Hostname: page-component-745bb68f8f-cphqk Total loading time: 0 Render date: 2025-02-06T01:46:28.528Z Has data issue: false hasContentIssue false

Big principal series, $p$-adic families and $\mathcal {L}$-invariants

Published online by Cambridge University Press:  25 April 2022

Lennart Gehrmann
Affiliation:
Fakultät für Mathematik, Universität Duisburg-Essen, Thea-Leymann-Straße 9, 45127 Essen, Germanylennart.gehrmann@uni-due.de
Giovanni Rosso
Affiliation:
Departments of Mathematics and Statistics, Concordia University, 1455 De Maisonneuve Blvd. W., Montréal, Québec, Canada H3G 1M8 giovanni.rosso@concordia.ca
Rights & Permissions [Opens in a new window]

Abstract

In earlier work, the first named author generalized the construction of Darmon-style $\mathcal {L}$-invariants to cuspidal automorphic representations of semisimple groups of higher rank, which are cohomological with respect to the trivial coefficient system and Steinberg at a fixed prime. In this paper, assuming that the Archimedean component of the group has discrete series we show that these automorphic $\mathcal {L}$-invariants can be computed in terms of derivatives of Hecke eigenvalues in $p$-adic families. Our proof is novel even in the case of modular forms, which was established by Bertolini, Darmon and Iovita. The main new technical ingredient is the Koszul resolution of locally analytic principal series representations by Kohlhaase and Schraen. As an application of our results we settle a conjecture of Spieß: we show that automorphic $\mathcal {L}$-invariants of Hilbert modular forms of parallel weight $2$ are independent of the sign character used to define them. Moreover, we show that they are invariant under Jacquet–Langlands transfer and, in fact, equal to the Fontaine–Mazur $\mathcal {L}$-invariant of the associated Galois representation. Under mild assumptions, we also prove the equality of automorphic and Fontaine–Mazur $\mathcal {L}$-invariants for representations of definite unitary groups of arbitrary rank. Finally, we study the case of Bianchi modular forms to show how our methods, given precise results on eigenvarieties, can also work in the absence of discrete series representations.

Type
Research Article
Copyright
© 2022 The Author(s). The publishing rights in this article are licensed to Foundation Compositio Mathematica under an exclusive licence

Introduction

Let $f=\sum _{n=1}^{\infty } a_n q^{n}$ be a normalized newform of weight $2$ and level $\Gamma _0(M)$ such that $M=pN$ with $p$ prime, $p\nmid N$ and $a_p=1$. Inspired by Teitelbaum's work (cf. [Reference TeitelbaumTei90]) on $\mathcal {L}$-invariants for automorphic forms on definite quaternion groups, Darmon in [Reference DarmonDar01] constructed the automorphic $\mathcal {L}$-invariant $\mathcal {L}(f)^{\pm }$ of $f$, depending a priori on a choice of sign at infinity. Let $a_p(k)$ be the $U_p$-eigenvalue of the $p$-adic family passing through $f$. In [Reference Bertolini, Darmon and IovitaBDI10] Bertolini, Darmon and Iovita (see also [Reference DasguptaDas05]) prove the following formula:

(0.1)\begin{equation} \mathcal{L}^{\pm}(f)=-2\frac{d a_p}{dk}\biggr|_{k=2}. \end{equation}

In particular, the automorphic $\mathcal {L}$-invariant is independent of the sign at infinity. They also prove a similar formula for Orton's $\mathcal {L}$-invariant (cf. [Reference OrtonOrt04]) in higher weight. Moreover, if $f$ admits a Jacquet–Langlands transfer $\operatorname {JL}(f)$ to a definite quaternion group, which is split at $p$, they show the analogous formula

\[ \mathcal{L}_T(\operatorname{JL}(f))=-2\frac{d a_p}{dk}\biggr|_{k=2} \]

for Teitelbaum's $\mathcal {L}$-invariant, thus proving that automorphic $\mathcal {L}$-invariants are preserved under Jacquet–Langlands transfers. This result was extended to Jacquet–Langlands transfers to indefinite quaternion groups over the rationals, which are split at $p$, by Dasgupta and Greenberg [Reference Dasgupta and GreenbergDG12], Longo, Rotger and Vigni [Reference Longo, Rotger and VigniLRV12] and Seveso [Reference SevesoSev13].

Over the last few years the construction of automorphic $\mathcal {L}$-invariants was generalized to various settings, e.g. to Hilbert modular cusp forms of parallel weight $2$ by Spieß (see [Reference SpießSpi14]) and to Bianchi modular cusp forms of even weight by Barrera and Williams (see [Reference Barrera Salazar and WilliamsBW19]). These generalizations have been defined as automorphic $\mathcal {L}$-invariants appear naturally in many context, from proving exceptional zero conjecture formulae to constructing and studying Stark–Heegner points.

Most recently, the first named author defined automorphic $\mathcal {L}$-invariants for certain cuspidal automorphic representations of higher rank semi-simple groups over a number field $F$, which are split at a fixed prime $ {\mathfrak {p}}$ of $F$ (see [Reference GehrmannGeh21]). The most crucial assumptions on the representation $\pi$ are that it is cohomological with respect to the trivial coefficient system and that the local factor $\pi _ {\mathfrak {p}}$ of $\pi$ at $ {\mathfrak {p}}$ is the Steinberg representation. Our main aim is to prove the analogue of equation (0.1) for these $\mathcal {L}$-invariants. Previous works used explicit computations with cocycles and it seems unlikely that one can generalize these to higher rank groups; instead, we give a new, more conceptual approach, which is novel even in the already known cases.

In § 1 we recall the definition of automorphic $\mathcal {L}$-invariants. Just as in the case of modular forms these $\mathcal {L}$-invariants depend on a choice of sign character at infinity. They also depend on a choice of degree of cohomology, in which the representation occurs. For this introduction, we suppress it because we are mostly interested in the case that there is only one interesting degree. Given a simple root $i$ of the group and a sign character $\epsilon$ the space of $\mathcal {L}$-invariants

\[ \mathcal{L}_{i}(\pi,{\mathfrak{p}})^{\epsilon} \subset \operatorname{Hom}^{\operatorname{ct}}(F_{{\mathfrak{p}}}^{\ast},{E}) \]

is a subspace of codimension at least one. Here ${E}$ denotes a large enough $p$-adic field. If strong multiplicity one holds, its codimension is exactly one. Whether a character of $F_ {\mathfrak {p}}^{\ast }$ belongs to the space of $\mathcal {L}$-invariants is decided by certain maps between cohomology groups of $ {\mathfrak {p}}$-arithmetic subgroups with values in duals of (locally analytic) generalized Steinberg representations. These maps are induced by cup products with one-extensions of the smooth generalized Steinberg representation corresponding to the simple root $i$ with the locally analytic Steinberg representations (see § 1.1 for a description of these extensions due to Ding, cf. [Reference DingDin19]).

As a first step, in § 2.2 we show that one can replace generalized Steinberg representations by locally analytic principal series representations and the extension classes by infinitesimal deformations of these principal series. As a consequence, we prove an automorphic analogue of the Colmez–Greenberg–Stevens formula (see Proposition 2.4). Let us remind ourselves that the Colmez–Greenberg–Stevens formula (see, for example, Theorem 3.4 of [Reference DingDin19]) states that one can compute Fontaine–Mazur $\mathcal {L}$-invariants of Galois representations by deforming them in $p$-adic families. Our analogue states that one can compute automorphic $\mathcal {L}$-invariants from the cohomology of $ {\mathfrak {p}}$-arithmetic groups with values in duals of big principal series representations, i.e. parabolic inductions of characters with values in units of affinoid algebras.

Thus, we reduce the problem to producing classes in these big cohomology groups. Here is where we need to impose further restrictions. First, we assume that the group under consideration is adjoint. Under this assumption Kohlhaase and Schraen constructed a Koszul resolution of locally analytic principal series representations (cf. [Reference Kohlhaase and SchraenKS12]), which we recall in § 3.1. Using that resolution we can lift overconvergent cohomology classes which are common eigenvectors for all $U_ {\mathfrak {p}}$-operators to big cohomology classes. Second, in order to have a nice enough theory of families of overconvergent cohomology classes (in the spirit of [Reference Ash and StevensAS08]) we consider only groups whose Archimedean component fulfils the Harish-Chandra condition. This implies that the automorphic representation we study only shows up in the middle degree cohomology of the associated locally symmetric space. We further assume that the map from the eigenvariety to the weight space $ {\mathcal {W}}$ is étale at the point corresponding to the automorphic representation $\pi$. Étaleness is implied by a suitable strong multiplicity one assumption and, thus, holds for example for Hilbert modular forms. Under these hypotheses we can show that we can lift the cohomology class corresponding to $\pi$ to a big cohomology class valued in functions on an open affinoid neighbourhood of the trivial character in weight space (see Theorem 3.12). This allows us give the generalization of (0.1) in Theorem 3.16. In particular, we see that automorphic $\mathcal {L}$-invariants are codimension one subspaces under our étaleness assumption.

As a first application, in § 4.1 we prove a conjecture of Spieß (cf. [Reference SpießSpi14, Conjecture 6.4]): we show that the $\mathcal {L}$-invariants of Hilbert modular forms of parallel weight $2$ are independent of the sign character used to define them. We further show that in this situation automorphic $\mathcal {L}$-invariants are invariant under Jacquet–Langlands transfers to quaternion groups which are split at $ {\mathfrak {p}}$. In fact, we show that all these $\mathcal {L}$-invariants agree with the Fontaine–Mazur $\mathcal {L}$-invariant of the associated Galois representation.Footnote 1 In particular, one can remove the assumptions in the main theorem of [Reference SpießSpi14] (see Theorem 6.10(b) there). Furthermore, the equality of automorphic and Fontaine–Mazur $\mathcal {L}$-invariants makes the construction of Stark–Heegner points for modular elliptic curves over totally real fields unconditional. Similarly, we prove the equality of automorphic and Fontaine–Mazur $\mathcal {L}$-invariants for definite unitary groups under mild assumptions in § 4.2, i.e. we give an alternative proof of the main result of [Reference DingDin19] for our global situation.

We end the paper by considering the easiest case of a group that does not fulfil the Harish-Chandra condition, i.e. we study Bianchi modular forms. As we cannot deform the automorphic representation over the whole weight space, in general one cannot compute the $\mathcal {L}$-invariant completely in terms of derivatives of Hecke-eigenvalues. But at least in case the Bianchi forms is the base change of a modular form, we overcome this problem; we show that the $\mathcal {L}$-invariants of the base change equal the $\mathcal {L}$-invariant of the modular form.

The assumption that the coefficient system is trivial is not necessary for the arguments of this article. However, the construction of automorphic $\mathcal {L}$-invariants relies on the existence of a well-behaved lattice in the Steinberg representation, which is not known for twists of the Steinberg by an algebraic representation in general. In case the existence of such a lattice is known, e.g. if $G$ is a form of $\operatorname {PGL}_2$ by a result of Vignéras (see [Reference VignérasVig08]), one can easily generalize our results to arbitrary weights and non-critical slopes.

Notation

If $X$ and $Y$ are topological spaces, we write $C(X,Y)$ for the set of continuous functions from $X$ to $Y$. All rings are assumed to be commutative and unital. The group of invertible elements of a ring $R$ will be denoted by $R^{\ast }$. If $M$ is an $R$-module we denote the $q$th exterior power of $M$ by $\Lambda ^{q}_R M$. If $R$ is a ring and $G$ a group, we denote the group ring of $G$ over $R$ by $R[G]$. Given topological groups $H$ and $G$ we write $\operatorname {Hom}^{\operatorname {ct}}(H,G)$ for the space of continuous homomorphism from $H$ to $G$. Let $\chi \colon G\to R^{\ast }$ be a character. We write $R[\chi ]$ for the $G$-representation, which underlying $R$-module is $R$ itself and on which $G$ acts via the character $\chi$. The trivial character of $G$ will be denoted by ${{\mathbb 1}}_G$. Let $H$ be an open subgroup of a locally profinite group $G$ and $M$ an $R$-linear representation $M$ of $H$. The compact induction $\operatorname {c-ind}^{G}_{H}M$ of $M$ from $H$ to $G$ is the space of all functions $f\colon G\to M$ which have finite support modulo $H$ and satisfy $f(gh)=h^{-1}.f(g)$ for all $h\in H,\ g\in G$.

The setup

We fix an algebraic number field $F$. In addition, we fix a finite place $ {\mathfrak {p}}$ of $F$ lying above the rational prime $p$ and choose embeddings

\[ \overline{{\mathbb{Q}}_p} \hookleftarrow \overline{{\mathbb{Q}}}\hookrightarrow {\mathbb{C}}. \]

If $v$ is a place of $F$, we denote by $F_{v}$ the completion of $F$ at $v$. If $v$ is a finite place, we let $\mathcal {O}_{v}$ denote the valuation ring of $F_{v}$ and $\operatorname {ord}_{v}$ the additive valuation such that $\operatorname {ord}_{v}(\varpi )=1$ for any local uniformizer $\varpi \in \mathcal {O}_{v}$.

Let $ {\mathbb {A}}$ be the adele ring of $F$, i.e. the restricted product over all completions $F_{v}$ of $F$. We write $ {\mathbb {A}}^{\infty }$ (respectively, $ {\mathbb {A}} {{}^{ {\mathfrak {p}},\infty }}$) for the restricted product over all completions of $F$ at finite places (respectively, finite places different from $ {\mathfrak {p}}$). More generally, if $S$ is a finite set of places of $F$ we denote by $ {\mathbb {A}}^{S}$ the restricted product of all completions $F_v$ with $v\notin S$.

If $H$ is an algebraic group over $F$ and $v$ is a place of $F$, we write $H_v=H(F_v)$. We put $H_\infty =\prod _{v\mid \infty }H_v$.

Throughout the article we fix a connected, adjoint, semi-simple algebraic group $G$ over $F$. We assume that the base change $G_{F_ {\mathfrak {p}}}$ of $G$ to $F_ {\mathfrak {p}}$ is split. Let $K_\infty \subseteq G_\infty$ denote a fixed maximal compact subgroup. The integers $\delta$ and $q$ are defined via

\[ \delta=\operatorname{rk} G_\infty - \operatorname{rk} K_\infty \]

and

\[ 2 q+\delta=\dim G_\infty - \dim K_\infty. \]

Finally, we fix a cuspidal automorphic representation $\pi =\otimes _v \pi _v$ of $G( {\mathbb {A}})$ with the following properties:

  1. (i) $\pi$ is cohomological with respect to the trivial coefficient system;

  2. (ii) $\pi$ is tempered at $\infty$; and

  3. (iii) $\pi _ {\mathfrak {p}}$ is the (smooth) Steinberg representation $\operatorname {St}^{\infty }_{G_{ {\mathfrak {p}}}}( {\mathbb {C}})$ of $G_ {\mathfrak {p}}$.

We denote by $ {\mathbb {Q}}_\pi \subseteq \overline { {\mathbb {Q}}}$ a fixed finite extension of $ {\mathbb {Q}}$ over which $\pi ^{\infty }$ has a model (see Theorem C of [Reference JanuszewskiJan18] for the existence of such an extension).

Hypothesis (SMO)

We assume that the following strong multiplicity one hypothesis on $\pi$ holds. If $\pi ^{\prime }$ is an automorphic representation of $G$ such that:

  1. (i) $\pi _v^{\prime }\cong \pi _v$ for all finite places $v\neq {\mathfrak {p}}$;

  2. (ii) $\pi _ {\mathfrak {p}}^{\prime }$ has an Iwahori-invariant vector; and

  3. (iii) $\pi ^{\prime }_\infty$ has non-vanishing $(\mathfrak {g},K_{\infty }^{\circ })$-cohomology;

then $\pi$ is cuspidal, $\pi ^{\prime }_ {\mathfrak {p}}\cong \pi _ {\mathfrak {p}}$ and $\pi _\infty$ is tempered.

It is known that this holds for cuspidal representations of $\mathrm {GL}_n$ by work of Jacquet and Shalika [Reference Jacquet and ShalikaJS81b, Reference Jacquet and ShalikaJS81a] and Piateski-Shapiro [Reference Piatetski-ShapiroPia79] and thus, in particular, for representations of $G=\mathrm {PGL}_n$.

For general groups strong multiplicity one fails, see, for example, [Reference Howe and Piatetski-ShapiroHP83]. Hence, it is harder to find explicit results about representations $\pi$ for which the hypothesis SMO holds, but nevertheless it is expected that SMO should hold in many cases; for example, for generic representation of $\mathrm {GSp}_4$ [Reference SoudrySou87]. The same strategy of [Reference SoudrySou87] could apply as long as we have an injective transfer from (a subclass of) representations of $G$ to representation of $\mathrm {GL}_n$, e.g. tempered representations of unitary groups that at each prime are not endoscopic.

1. Automorphic $\mathcal {L}$-invariants

In the following, we briefly sketch the construction of automorphic $\mathcal {L}$-invariants. For more details, see [Reference GehrmannGeh21].

1.1 Extensions

In this section we recall the computation of certain $\operatorname {Ext}^{1}$-groups of (locally analytic) generalized Steinberg representations due to Ding (see [Reference DingDin19]). We fix a finite extension ${E}$ of $ {\mathbb {Q}}_p$. If $V$ and $W$ are admissible locally $ {\mathbb {Q}}_p$-analytic ${E}$-representations of $G_ {\mathfrak {p}}$, we write $\operatorname {Ext}^{1}_{\operatorname {an}}(V,W)$ for the group of locally analytic extensions of $V$ by $W$.

Given an algebraic subgroup $H\subseteq G_{F_ {\mathfrak {p}}}$, we denote the group of $F_ {\mathfrak {p}}$-valued points of $H$ also by $H$. We fix a Borel subgroup $B$ of the split group $G_{F_ {\mathfrak {p}}}$ and a maximal split torus $T\subseteq B$ and denote by $\Delta$ the associated basis of simple roots. For a subset $I\subseteq \Delta$ we let $P_I\supseteq B$ be the corresponding parabolic containing $B$.

Suppose $M$ is a smooth representation of $P_I$ over a ring $R$; we define its smooth induction to $G_ {\mathfrak {p}}$ as

\[ i_{P_I}^{\infty}(M)=\{f\colon G_{\mathfrak{p}}\to M \mbox{ locally constant}\mid f(pg)=p.f(g)\ \forall p\in P_I,\ g\in G_{\mathfrak{p}} \}. \]

The generalized $R$-valued (smooth) Steinberg representation associated with $I\subseteq \Delta$ is given by the quotient

\[ v^{\infty}_{P_I}(R)= i^{\infty}_{P_I}(R) \bigg/ \sum_{I\subset J\subset \Delta, I\neq J} i^{\infty}_{P_J}(R). \]

Likewise, if $V$ is a locally $ {\mathbb {Q}}_ {\mathfrak {p}}$-analytic ${E}$-representation of $P_I$, we define its locally analytic induction to $G_ {\mathfrak {p}}$ as the space of functions

\[ {\mathbb{I}}_{P_I}^{\operatorname{an}}(V)=\{f\colon G_{\mathfrak{p}}\to \tau \mbox{ locally } {\mathbb{Q}}_p\mbox{-analytic}\mid f(pg)=p.f(g)\ \forall p\in P_I,\ g\in G_{\mathfrak{p}} \}. \]

We define the locally analytic generalized Steinberg representation with respect to $I$ as the quotient

\[ {\mathbb{V}}^{\operatorname{an}}_{P_I}({E})={\mathbb{I}}_{P_I}^{\operatorname{an}}({E})\bigg/\sum_{I\subset J\subset \Delta, I\neq J} {\mathbb{I}}^{\operatorname{an}}_{P_J}({E}). \]

We put $\operatorname {St}_{G_ {\mathfrak {p}}}^{\infty }(R)=v^{\infty }_B(R)$ and $\operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {an}}({E})= {\mathbb {V}}^{\operatorname {an}}_B({E})$. Similarly, replacing locally analyticity with continuity we define the continuous Steinberg representation $\operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {ct}}({E})$ and, more generally, $ {\mathbb {V}}^{\operatorname {ct}}_{P_I}({E})$. It is easy to see that $ {\mathbb {V}}^{\operatorname {ct}}_{P_I}({E})$ is the universal unitary completion of both $v^{\infty }_{P_I}({E})$ and $ {\mathbb {V}}^{\operatorname {an}}_{P_I}({E})$.

Let $i\in \Delta$ be a simple root and $\lambda \in \operatorname {Hom}^{\operatorname {ct}}(B,{E})$ a continuous homomorphism. Note that $\lambda$ is automatically locally analytic and is trivial on the unipotent radical of $B$. Thus, it can be identified with a character on $T$. We write $\tau _\lambda$ for the two-dimensional representation of $B$ given by

(1.1)\begin{equation} \tau_\lambda(b)=\begin{pmatrix} 1 & \lambda(b)\\ 0 & 1\end{pmatrix}. \end{equation}

By the exactness of the parabolic induction functor for locally analytic extensions (see [Reference KohlhaaseKoh11, Proposition 5.1 and Remark 5.4]) we have a short exact sequence of the form

\[ 0\longrightarrow {\mathbb{I}}_{B}^{\operatorname{an}}({E})\longrightarrow {\mathbb{I}}_{B}^{\operatorname{an}}(\tau_\lambda)\longrightarrow {\mathbb{I}}_{B}^{\operatorname{an}}({E})\longrightarrow 0. \]

We write

(1.2)\begin{equation} \widehat{\mathcal{E}}^{\operatorname{an}}(\lambda)\in \operatorname{Ext}^{1}_{\operatorname{an}}({\mathbb{I}}^{\operatorname{an}}_{B}({E}),{\mathbb{I}}_{B}^{\operatorname{an}}({E})) \end{equation}

for the associated extension class. Further, we define the class

(1.3)\begin{equation} \widetilde{\mathcal{E}}^{\operatorname{an}}_{i}(\lambda)\in \operatorname{Ext}^{1}_{\operatorname{an}}(i^{\infty}_{P_i}({E}),{\mathbb{I}}_{B}^{\operatorname{an}}({E})) \end{equation}

as the pullback of this extension along $i^{\infty }_{P_i}({E})\to {\mathbb {I}}^{\operatorname {an}}_{B}({E})$. Finally, taking pushforward along $ {\mathbb {I}}_{B}^{\operatorname {an}}({E})\to \operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {an}}({E})$ yields the extension class

(1.4)\begin{equation} \mathcal{E}^{\operatorname{an}}_{i}(\lambda)\in \operatorname{Ext}^{1}_{\operatorname{an}}(i^{\infty}_{P_i}({E}),\operatorname{St}_{G_{\mathfrak{p}}}^{\operatorname{an}}). \end{equation}

By an easy calculation we see that the map

\[ \operatorname{Hom}^{\operatorname{ct}}(B,{E})\longrightarrow \operatorname{Ext}^{1}_{\operatorname{an}}(i^{\infty}_{P_i}({E}),\operatorname{St}_{G_{\mathfrak{p}}}^{\operatorname{an}}({E})),\quad \lambda\longmapsto \mathcal{E}^{\operatorname{an}}_{i}(\lambda) \]

defines a homomorphism. The inclusion $B\hookrightarrow P_i$ induces an injection

\[ \operatorname{Hom}^{\operatorname{ct}}(P_i,{E})\longrightarrow \operatorname{Hom}^{\operatorname{ct}}(B,{E}). \]

The quotient can be identified with the space $\operatorname {Hom}^{\operatorname {ct}}(F_ {\mathfrak {p}}^{\ast },{E})$ via the map

(1.5)\begin{equation} \operatorname{Hom}^{\operatorname{ct}}(F_{\mathfrak{p}}^{\ast},{E})\longrightarrow \operatorname{Hom}^{\operatorname{ct}}(B,{E})/\operatorname{Hom}^{\operatorname{ct}}(P_i,{E}),\quad \lambda\longmapsto \lambda\circ i. \end{equation}

Alternatively, let $i^{\vee }$ denote the coroot associated with $i$. Then, the kernel of the map

(1.6)\begin{equation} \operatorname{Hom}^{\operatorname{ct}}(B,{E})\longrightarrow \operatorname{Hom}^{\operatorname{ct}}(F_{\mathfrak{p}}^{\ast},{E}),\quad \lambda\longmapsto \lambda_{i}=\lambda\circ i^{\vee} \end{equation}

is equal to $\operatorname {Hom}^{\operatorname {ct}}(P_i,{E})$ and, hence, the map induces an isomorphism

\[ \operatorname{Hom}^{\operatorname{ct}}(B,{E})/\operatorname{Hom}^{\operatorname{ct}}(P_i,{E})\longrightarrow \operatorname{Hom}^{\operatorname{ct}}(F_{\mathfrak{p}}^{\ast},{E}), \]

which is the inverse of the isomorphism above up to multiplication by two.

Theorem 1.1 (Ding)

The following hold.

  1. (i) The map $\operatorname {Hom}^{\operatorname {ct}}(B,{E})\to \operatorname {Ext}^{1}_{\operatorname {an}}(i^{\infty }_{P_i}({E}),\operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {an}}({E}))$, $\lambda \mapsto \mathcal {E}^{\operatorname {an}}_{i}(\lambda )$ is surjective with kernel $\operatorname {Hom}^{\operatorname {ct}}(P_i,{E})\subseteq \operatorname {Hom}^{\operatorname {ct}}(B,{E})$.

  2. (ii) The canonical map $\operatorname {Ext}^{1}_{\operatorname {an}}(v^{\infty }_{P_i}({E}),\operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {an}}({E}))\to \operatorname {Ext}^{1}_{\operatorname {an}}(i^{\infty }_{P_i}({E}),\operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {an}}({E}))$ is an isomorphism.

  3. (iii) The induced map $\operatorname {Hom}^{\operatorname {ct}}(F_ {\mathfrak {p}}^{\ast },{E})\longrightarrow \operatorname {Ext}^{1}_{\operatorname {an}}(v^{\infty }_{P_i}({E}),\operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {an}}({E}))$, $\lambda \mapsto \mathcal {E}^{\operatorname {an}}_{i}(\lambda \circ i)$ is an isomorphism.

Proof. The third claim is a direct consequence of the first two. For the proof of the first two claims in the case $G=\operatorname {PGL}_n$, see § 2.2 of [Reference DingDin19]. The general case is proven in § 2.4 of [Reference GehrmannGeh21].

1.2 Flawless lattices

We recall the notion of flawless smooth representations over a ring $R$.

Definition 1.2 A smooth $R$-representation $M$ of $G_ {\mathfrak {p}}$ is called flawless if:

  1. (i) $M$ is projective as an $R$-module; and

  2. (ii) there exists a finite length exact resolution

    \[ 0\longrightarrow C_m\longrightarrow\cdots\longrightarrow C_0\longrightarrow M \longrightarrow 0 \]
    of $M$ by smooth $R$-representations $C_i$ of $G_ {\mathfrak {p}}$, where each $C_i$ is a finite direct sum of modules of the form
    \[ \operatorname{c-ind}_{K_{\mathfrak{p}}}^{G_{\mathfrak{p}}}(L) \]
    with $K_ {\mathfrak {p}}\subseteq G_ {\mathfrak {p}}$ a compact, open subgroup and $L$ a smooth representation of $K_ {\mathfrak {p}}$ that is finitely generated projective over $R$.

The following is our main example.

Theorem 1.3 (Borel–Serre)

The Steinberg representation $\operatorname {St}_{G_ {\mathfrak {p}}}^{\infty }(R)$ is flawless for every ring $R$.

Proof. It is enough to prove that $\operatorname {St}_{G_ {\mathfrak {p}}}^{\infty }( {\mathbb {Z}})$ is flawless. By [Reference Borel and SerreBS76, Theorem 5.6], $\operatorname {St}_{G_ {\mathfrak {p}}}^{\infty }( {\mathbb {Z}})$ can be identified with the cohomology with compact supports of the Bruhat–Tits building of $G_ {\mathfrak {p}}$. Thus, its simplicial complex gives a flawless resolution of $\operatorname {St}_{G_ {\mathfrak {p}}}^{\infty }( {\mathbb {Z}})$.

1.3 Cohomology of $ {\mathfrak {p}}$-arithmetic groups

Let $A$ be a $ {\mathbb {Q}}_p$-affinoid algebra in the sense of Tate. Given a compact, open subgroup $K^{{\mathfrak {p}}}\subseteq G( {\mathbb {A}} {{}^{ {\mathfrak {p}},\infty }})$, an $A[G_ {\mathfrak {p}}]$-module $V$ and an $A[G(F)]$-module $W$, which is free and of finite rank over $A$, we define $ {\mathcal {C}}_A(K^{ {\mathfrak {p}}},V;W)$ as the space of all $A$-linear maps $\Phi \colon G( {\mathbb {A}} {{}^{ {\mathfrak {p}},\infty }})/K^{{\mathfrak {p}}} \times V\to W$. The $A$-module $ {\mathcal {C}}_{E}(K^{{\mathfrak {p}}},V;W)$ carries a natural $G(F)$-action given by

\[ (\gamma.\Phi)(g,v)=\gamma.(\Phi(\gamma^{-1}g,\gamma^{-1}.v)). \]

Suppose $V$ is a topological $A$-module equipped with a continuous $A$-linear $G_ {\mathfrak {p}}$-action we put $ {\mathcal {C}}^{\operatorname {ct}}_A(K^{{\mathfrak {p}}},V;W)=C(G( {\mathbb {A}} {{}^{ {\mathfrak {p}},\infty }})/K^{ {\mathfrak {p}}},\operatorname {Hom}_{A}^{\operatorname {ct}}(V,W))$. Here $W$ is endowed with its canonical topology as a finitely generated free $A$-module.

Let ${E}$ be a finite extension of $ {\mathbb {Q}}_p$ with ring of integers $\mathcal {O}_{E}$. Suppose that $V$ is a smooth ${E}$-representation of $G_ {\mathfrak {p}}$ that admits a flawless $G_ {\mathfrak {p}}$-stable $\mathcal {O}_{E}$-lattice $M$. As $M$ is finitely generated as an $\mathcal {O}_{E}[G_ {\mathfrak {p}}]$-module, the completion of $V$ with respect to $M$ is the universal unitary completion $V^{\operatorname {un}}$ of $V$. The following automatic continuity statement holds (see [Reference GehrmannGeh21, Proposition 3.11]).

Proposition 1.4 Suppose that $V$ is a smooth ${E}$-representation of $G_ {\mathfrak {p}}$ that admits a flawless $G_ {\mathfrak {p}}$-stable $\mathcal {O}_{E}$-lattice $M$. Then the canonical map

\[ \operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},V^{\operatorname{un}};{E}(\epsilon)))\longrightarrow \operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},V;{E}(\epsilon))) \]

is an isomorphism for every character $\epsilon \colon \pi _0(G_\infty )\to \{\pm 1\}$, every compact, open subgroup $K^{{\mathfrak {p}}}\subseteq G( {\mathbb {A}} {{}^{ {\mathfrak {p}},\infty }})$ and every degree $d\geqslant 0$.

This proposition combined with Theorem 1.3 implies the following.

Corollary 1.5 The canonical map

\[ \operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\operatorname{ct}}({E});{E}(\epsilon))) \longrightarrow \operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\infty}({E});{E}(\epsilon))) \]

is an isomorphism.

For a compact, open subset $K_ {\mathfrak {p}}\subseteq G_ {\mathfrak {p}}$ and a character $\epsilon \colon \pi _0(G_\infty ) \to \{\pm 1\}$ we put

\[ \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times K_{\mathfrak{p}}},W)^{\epsilon}=\operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},{E}[G_{\mathfrak{p}}/K_{\mathfrak{p}}];W(\epsilon))). \]

If the level $K^{{\mathfrak {p}}}\times K_ {\mathfrak {p}}$ is neat, this group is naturally isomorphic to (the epsilon component of) the singular cohomology with coefficients in $W$ of the locally symmetric space of level $K^{{\mathfrak {p}}}\times K_ {\mathfrak {p}}$ associated with $G$. More generally, let $V$ be a topological $A$-module, which is locally convex as a $ {\mathbb {Q}}_p$-vector space, with a continuous $A$-linear $K_ {\mathfrak {p}}$-action. We consider $\operatorname {c-ind}_{K_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}V$ with the locally convex inductive limit topology. This is a continuous $A$-module with a continuous $G_p$-action. We put

\[ \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times K_{\mathfrak{p}}},\operatorname{Hom}_{A}^{\operatorname{ct}}(V,W))^{\epsilon}=\operatorname{H}^{d}(G(F),{\mathcal{C}}_{A}^{\operatorname{ct}}(K^{{\mathfrak {p}}},\operatorname{c-ind}_{K_{\mathfrak{p}}}^{G_{\mathfrak{p}}}V;W(\epsilon))). \]

Again, this space can be identified with the cohomology of the corresponding locally symmetric space with values in the sheaf associated with $\operatorname {Hom}_{A}^{\operatorname {ct}}(V,A)$.

1.4 Automorphic $\mathcal {L}$-invariants

For the remainder of the article we fix a finite extension ${E}$ of $ {\mathbb {Q}}_p$ that contains $ {\mathbb {Q}}_\pi$ and a compact, open subgroup $K^{{\mathfrak {p}}}=\prod _{v\nmid {\mathfrak {p}} \infty }K_v\subseteq G( {\mathbb {A}} {{}^{ {\mathfrak {p}},\infty }})$ such that $(\pi {{}^{ {\mathfrak {p}},\infty }})^{K^{{\mathfrak {p}}}}\neq 0$. We may assume that $K_v$ is hyperspecial for every finite place $v$ such that $\pi _v$ is spherical.

As $v_{P_I}^{\infty }({E})=v_I^{\infty }( {\mathbb {Z}})\otimes {E}$ we have a canonical isomorphism

\[ {\mathcal{C}}_{\mathbb{Z}}(K^{{\mathfrak {p}}},v^{\infty}_{P_I}({\mathbb{Z}});W)\cong {\mathcal{C}}_{{E}}(K^{{\mathfrak {p}}},v^{\infty}_{P_I}({E});W) \]

for any ${E}$-vector space $W$. Hence, we abbreviate this space by $ {\mathcal {C}}(K^{{\mathfrak {p}}},v^{\infty }_{P_I};W)$ (and similarly for $\operatorname {St}_{G_ {\mathfrak {p}}}^{\infty }$ in place of $v^{\infty }_{P_I}$).

Let $I_ {\mathfrak {p}}\subseteq G_ {\mathfrak {p}}$ be an Iwahori subgroup. By Frobenius reciprocity the choice of an Iwahori-fixed vector yields a $G_ {\mathfrak {p}}$-equivariant homomorphism

(1.7)\begin{equation} \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}{E}\longrightarrow \operatorname{St}_{G_{\mathfrak{p}}}^{\infty}, \end{equation}

which, in turn, induces a Hecke-equivariant map

(1.8)\begin{equation} \operatorname{ev}^{(d)}\colon \operatorname{H}^{d}(G(F),{\mathcal{C}}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\infty};{E}(\epsilon)))\longrightarrow \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E})^{\epsilon}. \end{equation}

Let

\[ {\mathbb{T}}={\mathbb{T}}(K^{{\mathfrak {p}}}\times I_{\mathfrak{p}})_{E}=C_c(K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}\backslash G({\mathbb{A}}^{\infty})/K^{{\mathfrak {p}}}\times I_{\mathfrak{p}},{E}) \]

be the ${E}$-valued Hecke algebra of level $K^{{\mathfrak {p}}}\times I_ {\mathfrak {p}}$. By abuse of notation, we denote the model of $\pi ^{\infty }$ over ${E}$ also by $\pi ^{\infty }$. If $V$ is a $\mathbb { {\mathbb {T}}}$-module, we put

\[ V[\pi]=\sum_f \operatorname{im}(f) \]

where we sum over all $\mathbb {T}$-homomorphisms $f\colon (\pi ^{\infty })^{K^{{\mathfrak {p}}}\times I_ {\mathfrak {p}}}\to V.$ Similarly, we define

\[ {\mathbb{T}}^{{\mathfrak{p}}}={\mathbb{T}}(K^{{\mathfrak {p}}})_{E}=C_c(K^{{\mathfrak {p}}}\backslash G({\mathbb{A}}{{}^{{\mathfrak{p}},\infty}})/K^{{\mathfrak {p}}},{E}) \]

to be the Hecke algebra away from $ {\mathfrak {p}}$ and, given a $ {\mathbb {T}}^{ {\mathfrak {p}}}$-module $V$, we put

\[ V[\pi^{{\mathfrak{p}}}]=\sum_f \operatorname{im}(f) \]

where we sum over all $ {\mathbb {T}}^{ {\mathfrak {p}}}$-homomorphisms $f\colon (\pi {{}^{ {\mathfrak {p}},\infty }})^{K^{{\mathfrak {p}}}}\to V.$

For the proof of the following proposition that crucially relies on the hypothesis (SMO) we refer to [Reference GehrmannGeh21, Proposition 3.6].

Proposition 1.6 The map $\operatorname {ev}^{(d)}$ induces an isomorphism

\[ \operatorname{H}^{d}(G(F),{\mathcal{C}}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\infty};{E}(\epsilon)))[\pi^{{\mathfrak{p}}}]\xrightarrow{\cong} \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E})^{\epsilon}[\pi]. \]

on isotypic components. There exists an integer $m_\pi \geqslant 0$ such that

\[ \dim_{E} \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E})^{\epsilon}[\pi]= m_\pi\cdot \dim(\pi^{\infty})^{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}} \cdot\binom{\delta}{d-q}\ \forall d\geqslant 0 \]

for each sign character $\epsilon$.

By Corollary 1.5, we have a canonical isomorphism

\[ \operatorname{H}^{d}(G(F),{\mathcal{C}}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\infty};{E}(\epsilon))) \cong \operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}^{\operatorname{ct}}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\operatorname{ct}};{E}(\epsilon))). \]

Composing with the homomorphism coming from dualizing the continuous inclusion $\operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {an}}\hookrightarrow \operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {ct}}$ yields the map

\[ \operatorname{H}^{d}(G(F),{\mathcal{C}}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\infty};{E}(\epsilon)))\longrightarrow \operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}^{\operatorname{ct}}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\operatorname{an}};{E}(\epsilon))) \]

in cohomology. Thus, for every $i\in \Delta$ we get a well-defined cup-product pairing

\begin{align*} &\operatorname{H}^{d}(G(F),{\mathcal{C}}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\infty};{E}(\epsilon))) \times \operatorname{Ext}^{1}_{\operatorname{an}}(v_{P_i}^{\infty}({E}),\operatorname{St}_{G_{\mathfrak{p}}}^{\operatorname{an}}({E}))\\ &\quad \longrightarrow \operatorname{H}^{d+1}(G(F),{\mathcal{C}}(K^{{\mathfrak {p}}},v^{\infty}_{P_i};{E}(\epsilon))), \end{align*}

which commutes with the action of the Hecke algebra $ {\mathbb {T}}^{ {\mathfrak {p}}}$. By Theorem 1.1 we have a canonical isomorphism $\operatorname {Hom}^{\operatorname {ct}}(F_ {\mathfrak {p}}^{\ast },{E})\cong \operatorname {Ext}^{1}_{\operatorname {an}}(v^{\infty }_{P_i}({E}),\operatorname {St}_{G_ {\mathfrak {p}}}^{\operatorname {an}}({E}))$. Hence, taking cup product with the extension $\mathcal {E}^{\operatorname {an}}_{i}(\lambda \circ i)$ associated with a homomorphism $\lambda \in \operatorname {Hom}^{\operatorname {ct}}(F_ {\mathfrak {p}}^{\ast },{E})$ in (1.4) yields a map

\[ c^{(d)}_{i}(\lambda)^{\epsilon}\colon \operatorname{H}^{d}(G(F),{\mathcal{C}}(K^{{\mathfrak {p}}},\operatorname{St}_{G_{\mathfrak{p}}}^{\infty};{E}(\epsilon)))[\pi^{{\mathfrak{p}}}]\longrightarrow \operatorname{H}^{d+1}(G(F),{\mathcal{C}}(K^{{\mathfrak {p}}},v^{\infty}_{P_i};{E}(\epsilon)))[\pi^{{\mathfrak{p}}}] \]

on $\pi {{}^{ {\mathfrak {p}},\infty }}$-isotypic parts.

Definition 1.7 Given a character $\epsilon \colon \pi _0(G_\infty ) \to \{\pm 1\}$, an integer $d\in {\mathbb {Z}}$ with $0\leqslant d\leqslant \delta$ and a root $i\in \Delta$ we define

\[ \mathcal{L}_{i}^{(d)}(\pi,{\mathfrak{p}})^{\epsilon}\subseteq \operatorname{Hom}^{\operatorname{ct}}(F_{\mathfrak{p}}^{\ast},{E}) \]

as the kernel of the map $\lambda \mapsto c^{(q+d)}_{i}(\lambda )^{\epsilon }$.

Alternatively, we can consider cup products with extensions associated with homomorphisms $\lambda \colon B \to {E}$ and define the $\mathcal {L}$-invariant as a subspace of

\[ \operatorname{Hom}^{\operatorname{ct}}(B,{E})/\operatorname{Hom}^{\operatorname{ct}}(P_i,{E}). \]

This subspace is mapped to the $\mathcal {L}$-invariant defined above via the map (1.6) induced by the coroot $i^{\vee }$ associated with $i$.

Proposition 1.8 For all sign characters $\epsilon$, every degree $d\in [0,\delta ]\cap {\mathbb {Z}}$ and every root $i \in \Delta$ the $\mathcal {L}$-invariant

\[ \mathcal{L}_{i}^{(d)}(\pi,{\mathfrak{p}})^{\epsilon}\subseteq \operatorname{Hom}^{\operatorname{ct}}(F_{\mathfrak{p}}^{\ast},{E}) \]

is a subspace of codimension at least one, which does not contain the space of smooth homomorphisms.

Suppose $m_\pi =1$. Then, in the extremal cases $d=0$ and $d=\delta$ the codimension is exactly one.

Proof. This is Proposition 3.14 of [Reference GehrmannGeh21].

Remark 1.9 Let $\log _p\colon {E}^{\ast }\to {E}$ denote the branch of the $p$-adic logarithm such that $\log _p(p)=0.$ Let us assume that ${E}$ contains the images of all embeddings $\sigma \colon F_ {\mathfrak {p}}\to \overline { {\mathbb {Q}}_p}$. We put $\log _{p,\sigma }=\log _p \circ \sigma \colon F_ {\mathfrak {p}}^{\ast }\to {E}$. The set

\[ \{{\rm ord}_{\mathfrak{p}}\} \cup \{{\rm log}_{p,\sigma}\mid \sigma\colon F_{\mathfrak{p}} \to {E} \} \]

is a basis of $\operatorname {Hom}^{\operatorname {ct}}(F_ {\mathfrak {p}}^{\ast },{E})$. Suppose $\mathcal {L}\subseteq \operatorname {Hom}^{\operatorname {ct}}(F_ {\mathfrak {p}}^{\ast },{E})$ is a subspace of codimension one that does not contain $\operatorname {ord}_ {\mathfrak {p}}$. Then for each embedding $\sigma$ there exists a unique element $\mathcal {L}^{\sigma }\in {E}$ such that $\log _{p,\sigma }- \mathcal {L}^{\sigma }\operatorname {ord}_ {\mathfrak {p}} \in \mathcal {L}$ and these elements clearly form a basis of $\mathcal {L}$.

2. $\mathcal {L}$-invariants and big principal series

By the Colmez–Greenberg–Stevens formula (see [Reference DingDin19, Theorem 3.4]) one can calculate Fontaine–Mazur $\mathcal {L}$-invariants of certain local Galois representations by deforming them (or rather the attached $(\varphi,\Gamma )$-modules) in rigid analytic families. The goal of this section is to prove an automorphic version of that formula.

This implies that $\mathcal {L}$-invariants can be related to the existence of cohomology classes with values in (duals of) families of principal series.

2.1 Linear algebra over the dual numbers

Let $R$ be a ring and $R[\varepsilon ]=R[X]/X^{2}$ the ring of dual numbers over $R$. Let $M$ be an $R[\varepsilon ]$-module. We define several maps on dual spaces associated with $M$. First, multiplication with $\varepsilon$ induces the map $\mu \colon M/\varepsilon M\to \varepsilon M\hookrightarrow M$. We denote by

\[ \mu_\varepsilon^{\ast}\colon \operatorname{Hom}_R(M,R)\longrightarrow \operatorname{Hom}_R(M/\varepsilon M,R) \]

its $R$-dual. Second, reducing modulo $\varepsilon$ yields the map

\[ \operatorname{red}\colon \operatorname{Hom}_{R[\varepsilon]}(M,R[\varepsilon])\longrightarrow \operatorname{Hom}_{R}(M/\varepsilon M,R). \]

Finally, the map $\operatorname {add}\colon R[\varepsilon ]\to R, a+b\varepsilon \mapsto a+b$ induces the map

\[ \operatorname{add}_\ast\colon\operatorname{Hom}_{R[\varepsilon]}(M,R[\varepsilon])\longrightarrow \operatorname{Hom}_{R}(M,R). \]

The following easy computation is left to the reader.

Lemma 2.1 The maps $\mu _\varepsilon ^{\ast }$, $\operatorname {red}$ and $\operatorname {add}$ are functorial in $M$ and the following equality holds: $\operatorname {red}=\mu _\varepsilon ^{\ast }\circ \operatorname {add}_\ast.$

2.2 Infinitesimal deformations of principal series

Let $T\subseteq B\subset G_{F_ {\mathfrak {p}}}$ be the maximal torus respectively the Borel subgroup chosen in § 1.1. The map

\[ \tau\colon\operatorname{Hom}^{\operatorname{ct}}(B,{E})\longrightarrow\operatorname{Hom}^{\operatorname{ct}}(B,{E}[\varepsilon]^{\ast}),\quad \lambda \longmapsto \left[x\mapsto 1+\lambda(x)\epsilon\right] \]

defines an injective group homomorphism. Its image is the set of all continuous characters $\chi \colon B\to {E}[\varepsilon ]^{\ast }$ such that $\chi \equiv 1 \bmod \varepsilon$. The underlying ${E}$-representation of $\tau (\lambda )$ is the two-dimensional representation $\tau _\lambda$ defined in (1.1). Thus, we can view $ {\mathbb {I}}_{B}^{\operatorname {an}}(\tau _\lambda )= {\mathbb {I}}_{B}^{\operatorname {an}}(\tau (\lambda ))$ as an ${E}[\varepsilon ]$-representation of $G_ {\mathfrak {p}}$. Reducing modulo $\varepsilon$ induces the map

\[ \operatorname{red}_\lambda^{d,\epsilon}\colon \operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{{E}[\varepsilon]}(K^{{\mathfrak {p}}},{\mathbb{I}}_{B}^{\operatorname{an}}(\tau_\lambda);{E}[\varepsilon](\epsilon)))\longrightarrow\operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}_{B}^{\operatorname{an}}({E});{E}(\epsilon))) \]

in cohomology.

Let $i^{\vee }$ be the coroot associated with $i$. Given a character $\lambda \colon B\to {E}$ we put $\lambda _i=\lambda \circ i^{\vee }$.

Lemma 2.2 Let $\lambda \colon B\to {E}$ be a continuous character. If the isotypic component

\[ \operatorname{H}{}^{q+d}(G(F),{\mathcal{C}}{}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}_{B}^{\operatorname{an}}({E});{E}(\epsilon)))[\pi^{{\mathfrak{p}}}] \]

is contained in the image of $\operatorname {red}_\lambda ^{q+d,\epsilon }$, then the homomorphism $\lambda _i$ belongs to $\mathcal {L}_{i}^{(d)}(\pi, {\mathfrak {p}})^{\epsilon }.$

Proof. The inclusion $ {\mathbb {I}}_{B}^{\operatorname {an}}({E})\hookrightarrow {\mathbb {I}}_{B}^{\operatorname {an}}(\tau _\lambda )$ induces the map

\[ \operatorname{H}^{q+d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{{E}}(K^{{\mathfrak {p}}},{\mathbb{I}}_{B}^{\operatorname{an}}(\tau_\lambda);{E}(\epsilon)))\longrightarrow\operatorname{H}^{q+d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}_{B}^{\operatorname{an}}({E});{E}(\epsilon))) \]

in cohomology. Its image is the kernel of the cup product with the extension class $\widehat {\mathcal {E}}^{\operatorname {an}}(\lambda )$ defined in (1.2).

Thus, by Lemma 2.1 our assumption implies that the $\pi$-isotypic component $\operatorname {H}^{q+d}(G(F), {\mathcal {C}}^{\operatorname {ct}}_{E}(K^{{\mathfrak {p}}}, {\mathbb {I}}_{B}^{\operatorname {an}}({E});{E}(\epsilon )))[\pi ^{ {\mathfrak {p}}}]$ is contained in the kernel of the cup product with $\widehat {\mathcal {E}}^{\operatorname {an}}(\lambda )$. We have the following commutative diagram.

The claim now follows from the first part of the next lemma and a simple diagram chase.

Lemma 2.3 Let $J\subseteq \Delta$ be a subset.

  1. (i) The canonical map

    \[ \operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},v^{\infty}_{P_J}({E});{E}(\epsilon)))[\pi^{{\mathfrak{p}}}]\longrightarrow\operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},i^{\infty}_{P_J}({E});{E}(\epsilon)))[\pi^{{\mathfrak{p}}}] \]
    is injective for all $d$.
  2. (ii) It is an isomorphism in degree $d=q+|J|$.

Proof. The Jordan–Hölder decomposition of $i^{\infty }_{P_J}({E})$ consists of all generalized Steinberg representations $v^{\infty }_{P_I}({E})$ with $J\subseteq I$, each occurring with multiplicity one. Thus, the second claim follows from [Reference GehrmannGeh21, Proposition 3.9].

Via the well-known resolution (see, for example, [Reference OrlikOrl05, Proposition 11])

\[ 0 \to i^{\infty}_{P_\Delta}({E})\to \bigoplus_{\substack{I\subseteq J\subseteq \Delta\\ |\Delta\setminus I|=1}}i^{\infty}_{P_I}({E}) \to \cdots \to \bigoplus_{\substack{J\subseteq I\subseteq \Delta\\ |I\setminus J|=1}}i^{\infty}_{P_I}({E}) \to i^{\infty}_{P_J}({E}) \longrightarrow v^{\infty}_{P_J}({E}) \to 0 \]

one can reduce the first claim to the following statement: Let $\mathcal {E}_{J,I}$ be any smooth extension of $v^{\infty }_{P_J}({E})$ by $v^{\infty }_{P_I}({E})$, where $J\subseteq I \subseteq \Delta$ with $|I|=|J|+1$. Then the map

\[ \operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},v^{\infty}_{P_J}({E});{E}(\epsilon)))[\pi^{{\mathfrak{p}}}]\longrightarrow\operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},\mathcal{E}_{J,I};{E}(\epsilon)))[\pi^{{\mathfrak{p}}}] \]

is injective for all $d$. Equivalently, it is enough to show that the cup product

\[ \operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},v^{\infty}_{P_I}({E});{E}(\epsilon)))[\pi^{{\mathfrak{p}}}]\xrightarrow{\cup \mathcal{E}_{J,I}}\operatorname{H}^{d+1}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},v^{\infty}_{P_J}({E});{E}(\epsilon)))[\pi^{{\mathfrak{p}}}] \]

is the zero map for all $d$. Let $\mathcal {E}_{I,J}$ be the unique up to scalar non-split smooth extension of $v^{\infty }_{P_I}({E})$ by $v^{\infty }_{P_J}({E})$. By [Reference GehrmannGeh21, Corollary 3.8], the cup product

\[ \operatorname{H}^{d+1}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},v^{\infty}_{P_J}({E});{E}(\epsilon)))[\pi^{{\mathfrak{p}}}]\xrightarrow{\cup \mathcal{E}_{I,J}}\operatorname{H}^{d+2}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},v^{\infty}_{P_I}({E});{E}(\epsilon)))[\pi^{{\mathfrak{p}}}] \]

is an isomorphism. Therefore, it is enough to prove that taking the cup product with $\mathcal {E}_{J,I} \cup \mathcal {E}_{I,J}$ induces the zero map on $\operatorname {H}^{\bullet }(G(F), {\mathcal {C}}_{E}(K^{{\mathfrak {p}}},v^{\infty }_{P_I}({E});{E}(\epsilon )))[\pi ^{ {\mathfrak {p}}}]$. This is true because $\mathcal {E}_{J,I} \cup \mathcal {E}_{I,J}$ is a smooth $2$-extension of $v^{\infty }_{P_I}({E})$ by itself and the space of all such extensions is zero by [Reference OrlikOrl05, Theorem 1].

2.3 The automorphic Colmez–Greenberg–Stevens formula

Let $A$ be an ${E}$-affinoid algebra and $\chi \colon B\to A^{\ast }$ a locally analytic character. The parabolic induction $ {\mathbb {I}}_B^{\operatorname {an}}(\chi )$ is naturally an $A[G_ {\mathfrak {p}}]$-module. Given an ideal $ {\mathfrak {m}}\subseteq A$ we let $\chi _ {\mathfrak {m}}\colon B\to (A/ {\mathfrak {m}})^{\ast }$, $x\longmapsto \chi (x) \bmod {\mathfrak {m}}$ denote the reduction of $\chi$ modulo $ {\mathfrak {m}}$. Similar to Proposition 2.2.1 of [Reference HansenHan17], one can prove that

\[ \operatorname{Hom}_{A}^{\operatorname{ct}}({\mathbb{I}}_B^{\operatorname{an}}(\chi),A)\otimes_A A/{\mathfrak{m}} \cong \operatorname{Hom}_{A/{\mathfrak{m}}}^{\operatorname{ct}}({\mathbb{I}}_B^{\operatorname{an}}(\chi_{\mathfrak{m}}),A/{\mathfrak{m}}). \]

Let $ {\mathfrak {m}}\in \operatorname {Spm} A$ be an ${E}$-rational point such that $\chi _ {\mathfrak {m}}={{\mathbb 1}}_{G {\mathfrak {p}}}$. Thus, the reduction map induces the maps

\[ \operatorname{red}_\chi^{d,\epsilon}\colon \operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{A}(K^{{\mathfrak {p}}},{\mathbb{I}}_{B}^{\operatorname{an}}(\chi);A(\epsilon)))\longrightarrow\operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}_{B}^{\operatorname{an}}({E});{E}(\epsilon))) \]

in cohomology.

Let $i^{\vee }$ be the coroot associated with $i$. We put $\chi _i=\chi \circ i^{\vee } \in \operatorname {Hom}(F_{ {\mathfrak {p}}}^{\ast },A)$. Suppose $v\colon \operatorname {Spec} {E}[\varepsilon ]\to \operatorname {Spm} A$ is an element of the tangent space of $\operatorname {Spm} A$ at $ {\mathfrak {m}}$. The pullback $\chi _{i,v}$ of $\chi _i$ along $v$ is of the form $\chi _{i,v}=1+ ({\partial }/{\partial v})\chi _{i}\cdot \varepsilon$ for a unique homomorphism $ ({\partial }/{\partial v})\chi _{i}\colon F_ {\mathfrak {p}}^{\ast }\to {E}.$

Lemma 2.2 immediately implies the following.

Proposition 2.4 Suppose that the image of $\operatorname {red}_\chi ^{q+d,\epsilon }$ contains the $\pi {{}^{ {\mathfrak {p}},\infty }}$-isotypic component of $\operatorname {H}^{q+d}(G(F), {\mathcal {C}}^{\operatorname {ct}}_{E}(K^{{\mathfrak {p}}}, {\mathbb {I}}_{B}^{\operatorname {an}}({E});{E}(\epsilon )))$. Then the homomorphism $ ({\partial }/{\partial v})\chi _{i}$ belongs to $\mathcal {L}_{i}^{(d)}(\pi, {\mathfrak {p}})^{\epsilon }$ for every element $v$ of the tangent space of $\operatorname {Spm} A$ at $ {\mathfrak {m}}$.

3. Overconvergent cohomology

After giving a brief overview on Kohlhaase and Schraen's Koszul resolution of locally analytic principal series (see [Reference Kohlhaase and SchraenKS12]) we recall the control theorem relating overconvergent cohomology to classical cohomology as proven by Ash–Stevens, Urban and Hansen (see [Reference Ash and StevensAS08, Reference UrbanUrb11, Reference HansenHan17]). Combining the two results allows us to construct classes in the cohomology of (duals of) principal series. If $G_\infty$ fulfils the Harish-Chandra condition, i.e. if $\delta =0$, we can lift the construction to families of principal series. This implies our main theorem.

3.1 Koszul complexes

In [Reference Kohlhaase and SchraenKS12], Kohlhaase and Schraen construct a resolution of locally analytic principal series representations via a Koszul complex. We recall their construction in a slightly more general setup: instead of restricting to $p$-adic fields as coefficient rings we allow affinoid algebras. The proofs of [Reference Kohlhaase and SchraenKS12] carry over verbatim to this more general framework.

Let us fix some notation: we denote the Borel opposite of $B\subseteq {G_{F_ {\mathfrak {p}}}}$ by $\bar {B}$. Let $\bar {N}\subseteq \bar {B}$ be its unipotent radical. The chosen torus $T\subseteq B\subseteq G_{F_ {\mathfrak {p}}}$ determines an apartment in the Bruhat–Tits building of $G_ {\mathfrak {p}}$. We chose a chamber $C$ of that apartment and a special vertex $v$ of $C$ as in § 3.5. of [Reference CartierCar79]. The stabilizer $G_{ {\mathfrak {p}},0}\subseteq G_ {\mathfrak {p}}$ of $v$ is a maximal compact subgroup of $G_ {\mathfrak {p}}$ and the stabilizer $I_ {\mathfrak {p}}\subseteq G_{ {\mathfrak {p}},0}$ of $C$ is an Iwahori subgroup. Let $\mathfrak {G}_0$ be the Bruhat–Tits group scheme over $\mathcal {O}_ {\mathfrak {p}}$ associated with $G_{ {\mathfrak {p}},0}$. We define

\[ I_{\mathfrak{p}}^{n}=\operatorname{ker}(I_{\mathfrak{p}}\longrightarrow \mathfrak{G}_0(\mathcal{O}_{\mathfrak{p}}/{\mathfrak{p}}^{n})). \]

The open normal subgroups $I_ {\mathfrak {p}}^{n}\subseteq I_ {\mathfrak {p}}$ form a system of neighbourhoods of the identity in $I_ {\mathfrak {p}}$. The subgroup $T_0=T\cap G_{ {\mathfrak {p}},0}$ is maximal compact subgroup of $T$.

Let $X^{\ast }(T)$ (respectively, $X_{\ast }(T)$) denote the group of $F_ {\mathfrak {p}}$-rational characters (respectively, cocharacters) of $T$. The natural pairing

\[ \langle \cdot,\cdot \rangle\colon X^{\ast}(T) \times X_\ast(T)\longrightarrow {\mathbb{Z}} \]

is a perfect pairing. There is a natural isomorphism $T/T_0\cong X_{\ast }(T)$ characterized by

\[ \langle \chi,t\rangle=\operatorname{ord}_{{\mathfrak{p}}}(\chi(t)). \]

We denote by $\Phi ^{+}$ the set of positive roots with respect to $B$ and put

\[ T^{-}=\{t\in T\mid \operatorname{ord}_{{\mathfrak{p}}}(\alpha(t))\leqslant 0\ \forall \alpha\in\Phi^{+}\}. \]

Let $A$ be a $ {\mathbb {Q}}_p$-affinoid algebra. Restricting to $T_0$ gives a bijection between locally analytic characters $\chi \colon B\cap I_ {\mathfrak {p}}\to A^{\ast }$ and locally analytic characters $\chi \colon T_0\to A^{\ast }$. Given such a character $\chi$ we write

\[ {\mathcal{A}}_{\chi}=\{f\colon I_{\mathfrak{p}}\longrightarrow A\ \mbox{locally analytic}\mid f(bk)=\chi(b)f(k)\ \forall b\in B\cap I_{\mathfrak{p}},\ k\in I_p\} \]

for the locally analytic induction of $\chi$ to $I_ {\mathfrak {p}}$. It is naturally an $A[I_ {\mathfrak {p}}]$-module. Restricting a function $f\in {\mathcal {A}}_{\chi }$ to the intersection $I_ {\mathfrak {p}}\cap \bar {N}$ induces an isomorphism of $ {\mathcal {A}}_\chi$ with the space of all locally analytic functions from $I_ {\mathfrak {p}}\cap \bar {N}$ to $A$. There exists a minimal integer $n_\chi \geqslant 1$ such that $\chi$ restricted to $B\cap I_ {\mathfrak {p}}^{n_\chi }$ is rigid analytic. For any $n\geqslant n_\chi$ we define the $A[I_ {\mathfrak {p}}]$-submodule

\[ {\mathcal{A}}_{\chi}^{n}=\{f\in {\mathcal{A}}_{\chi} \mid f \mbox{ is rigid analytic on any coset in } I_{\mathfrak{p}}/I_{\mathfrak{p}}^{n}\}. \]

For later purposes, we define the dual spaces

\[ {\mathcal{D}}^{n}_{\chi}=\operatorname{Hom}_A^{\operatorname{ct}}({\mathcal{A}}^{n}_{\chi}, A). \]

By Frobenius reciprocity we can identify $\operatorname {End}_{A[ G_ {\mathfrak {p}}]}(\operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}} ( {\mathcal {A}}^{n}_{\chi }))$ with the space of all functions $\Psi \colon G_ {\mathfrak {p}}\to \operatorname {End}_{A}( {\mathcal {A}}_{\chi }^{n})$ such that:

  1. (i) $\Psi$ is $I_ {\mathfrak {p}}$-biequivariant, i.e. $\Psi (k_1 g k_2)=k_1 \Psi (g)k_2$ for all $k_1 ,k_2\in I_ {\mathfrak {p}}$, $g\in G_ {\mathfrak {p}}$; and

  2. (ii) for every $f\in {\mathcal {A}}_{\chi }^{n}$ the function $G_ {\mathfrak {p}}\to {\mathcal {A}}_{\chi }^{n}$, $g\mapsto \Psi (g)(f)$ is compactly supported.

Let $t$ be an element of $T^{-}$ and $f\in {\mathcal {A}}_{\chi }^{n}$. The function $I_ {\mathfrak {p}}\to A$, $u\mapsto f(tut^{-1})$ defines an element of $ {\mathcal {A}}_{\chi }^{n}$.

Lemma 3.1 For every element $t\in T^{-}$ there exists a unique $I_ {\mathfrak {p}}$-biequivariant function $\Psi _t\colon G_ {\mathfrak {p}} \to \operatorname {End}_A( {\mathcal {A}}_{\chi }^{n})$ such that:

  1. (i) $\operatorname {supp}(\Psi _t)= I_ {\mathfrak {p}} t^{-1} I_ {\mathfrak {p}}$; and

  2. (ii) $\Psi _t(t^{-1})(f)(u)=f(tut^{-1})$ for any $f\in {\mathcal {A}}_{\chi }^{n}$ and $u\in I_ {\mathfrak {p}}\cap \bar {N}$.

Proof. This is a minor generalization of [Reference Kohlhaase and SchraenKS12, Lemma 2.2].

For $t\in T^{-}$ we denote by $U_t$ the endomorphism of $\operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}} ( {\mathcal {A}}^{n}_{\chi })$ corresponding to $\Psi _t$. The following is a straightforward generalization of [Reference Kohlhaase and SchraenKS12, Lemma 2.3].

Lemma 3.2 We have $U_t U_{\tilde {t}}= U_{t\tilde {t}}$ for all $t,\tilde {t}\in T^{-}$.

Now let us fix a character $\chi \colon T\to A^{\ast }$ and let $\chi _0$ be its restriction to $T_0$. Given an open subset $C\subseteq G_ {\mathfrak {p}}$, which is stable under multiplication with $B$ from the left, we denote by

\[ {\mathbb{I}}^{\operatorname{an}}_B(\chi) (C)\subseteq {\mathbb{I}}^{\operatorname{an}}_B(\chi) \]

the subset of all functions with support in $C$. Restricting functions to $I_ {\mathfrak {p}}$ gives an $I_ {\mathfrak {p}}$-equivariant $A$-linear isomorphism

\[ {\mathbb{I}}^{\operatorname{an}}_B(\chi)(BI_{\mathfrak{p}})\xrightarrow{\cong} {\mathcal{A}}_{\chi_0}. \]

Thus, by Frobenius reciprocity its inverse induces a $G_ {\mathfrak {p}}$-equivariant $A$-linear map

(3.1)\begin{equation} \operatorname{aug}_\chi\colon \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}({\mathcal{A}}^{n}_{\chi_0})\longrightarrow\operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}({\mathcal{A}}_{\chi_0})\longrightarrow {\mathbb{I}}^{\operatorname{an}}_B(\chi) \end{equation}

for any integer $n\geqslant n_{\chi _0}$.

As the group $G_{F_ {\mathfrak {p}}}$ is adjoint, there exist elements $t_i\in T^{-}$, $i\in \Delta$, such that

\[ t_i\in \bigcap_{j\in\Delta\setminus\{i\}}\operatorname{ker}(j) \]

and

\[ \operatorname{ord}_{{\mathfrak{p}}}(i(t_i))=-1. \]

The element $t_i$ is uniquely determined by the value $i(t_i)^{-1}\in F_ {\mathfrak {p}}^{\ast }$, which is a uniformizer. Every element $t\in T$ can uniquely be written as $t=t_0\prod _{i\in \Delta }t_{i}^{n_i}$ with $t\in T_0$ and integers $n_{i}\in {\mathbb {Z}}$. Let us fix a choice of $t_i$, $i\in \Delta$, and put

\begin{align*} y_i=U_{t_i}-\chi(t_i). \end{align*}

Proposition 3.3 The $G_ {\mathfrak {p}}$-equivariant $A$-linear map $\operatorname {aug}_\chi$ is surjective with kernel $\sum _{i\in \Delta }\operatorname {im}(y_i).$

Proof. The same proof as for [Reference Kohlhaase and SchraenKS12, Proposition 2.4] works here.

By Lemma 3.1 the $G_ {\mathfrak {p}}$-representation $\operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}( {\mathcal {A}}^{n}_{\chi _0})$ is a module over the polynomial algebra $A[X_i\mid i\in \Delta ]$, where $X_i$ acts through the operator $T_i$. The Koszul complex of $\operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}( {\mathcal {A}}^{n}_{\chi _0})$ with respect to the endomorphisms $(y_i)_{i\in \Delta }$ is the complex $\Lambda ^{\bullet }_A(A^{\Delta }) \otimes \operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}( {\mathcal {A}}_{\chi _0})$ with boundary maps

(3.2)\begin{equation} d_k(e_{i_1}\wedge \cdots \wedge e_{i_k}\otimes f)=\sum_{l=1}^{k}(-1)^{l+1} e_{i_1}\wedge \cdots \wedge \widehat{e_{i_l}}\wedge \cdots \wedge e_{i_k}\otimes y_{i_l}(f). \end{equation}

The following is the main technical theorem of Kohlhaase–Schraen (cf. [Reference Kohlhaase and SchraenKS12, Theorem 2.5]) generalized to affinoid coefficient rings.

Theorem 3.4 (Kohlhaase–Schraen)

For any $n\geqslant n_{\chi _0}$ the augmented Koszul complex

\[ \Lambda^{\bullet}_A(A^{\Delta}) \otimes_A \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}({\mathcal{A}}^{n}_{\chi_0})\longrightarrow {\mathbb{I}}^{\operatorname{an}}_B(\chi)\longrightarrow 0 \]

with boundary maps (3.2) and augmentation map (3.1) is exact.

Remark 3.5 All of the results above remain valid if one replaces $ {\mathcal {A}}_{\chi _0}^{n}$ by $ {\mathcal {A}}_{\chi _0}$.

Example 3.6 Suppose $G_ {\mathfrak {p}}=\operatorname {PGL}_n(F_ {\mathfrak {p}})$, $T$ is the torus of diagonal matrices and $B$ is the Borel subgroup of upper triangular matrices. The simple roots of $T$ with respect to $B$ are given by

\[ i(\operatorname{diag}(x_1,\ldots,x_n))=x_i x_{i+1}^{-1} \]

for $1\leqslant i \leqslant d-1$. For each simple root $1\leqslant i \leqslant d-1$, we might choose for $t_i$ the image of the diagonal matrix

\[ \operatorname{diag}(1,\ldots,1,\pi,\ldots,\pi), \]

where $\pi$ is a uniformizer and exactly the first $i$ entries are equal to one. For the Iwahori subgroup $I_ {\mathfrak {p}}$, we may choose the image in $G_ {\mathfrak {p}}$ of all matrices in $\operatorname {GL}_n(\mathcal {O}_ {\mathfrak {p}})$, which are upper triangular modulo $ {\mathfrak {p}}$. Then $I_ {\mathfrak {p}}^{n}$ consists of all matrices in $I_ {\mathfrak {p}}$ which are congruent to the identity modulo $ {\mathfrak {p}}^{n}$.

There is also a smooth variant of the above result, which is probably well-known. As we could not find a reference in the literature, we give a proof of said variant in the following.

Let $\Omega$ be field of characteristic zero. With the same formula as before, we can define commuting Hecke operators $U_t\in \operatorname {End}_{G_ {\mathfrak {p}}}(\operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}({E}))$ for $t\in T^{-}$. Let ${{\mathbb 1}}\colon T_0\to \Omega ^{\times }$ denote the constant character on $T_0$. In case $\Omega ={E}$ we can identify ${E}$ with the subspace of constant functions in $ {\mathcal {A}}^{n}_{{{\mathbb 1}}}$ and the induced embedding

(3.3)\begin{equation} \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}({E})\hookrightarrow \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}({\mathcal{A}}^{n}_{{{\mathbb 1}}}) \end{equation}

is equivariant with respect to the operators $U_t$, $t\in T^{-}$.

The operators $U_t\in \operatorname {End}_{G_ {\mathfrak {p}}}(\operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}(\Omega ))$ are invertible. Moreover, by the Bernstein decomposition the map

\[ \Omega[X_i^{\pm 1} | i\in \Delta]\longrightarrow \operatorname{End}_{G_{\mathfrak{p}}}(\operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}(\Omega)),\ X_i\longmapsto U_{t_i} \]

is injective and $\operatorname {End}_{G_ {\mathfrak {p}}}(\operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}(\Omega ))$ is a free module of finite rank (and, therefore, flat) over $\Omega [X_i^{\pm 1} | i\in \Delta ]$ (see, for example, [Reference ReederRee92] for more details). As $\operatorname {c-ind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}(\Omega )$ is a flat module over its endomorphism algebra by a theorem of Borel (see [Reference BorelBor76, Theorem 4.10]), it is thus also flat as a $\Omega [X_i^{\pm 1} | i\in \Delta ]$-module.

Therefore, for any choice of elements $a_i\in \Omega ^{\times }$ the Koszul complex

\[ \Lambda^{\bullet}_\Omega(\Omega^{\Delta}) \otimes_\Omega \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}(\Omega) \]

associated with the regular sequence $y_i=U_{t_i}-a_i$, $i\in \Delta$, is a resolution of the $G_ {\mathfrak {p}}$-representation

\[ M_{\underline{a}}=\operatorname{coker} \Bigl(\bigoplus_{i \in \Delta} \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}(\Omega) \xrightarrow{(y_{i})_{i\in \Delta}}\operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}(\Omega)\Bigr). \]

Let $\chi _{\underline {a}}\colon T\to \Omega ^{\ast }$ be the unique smooth unramified character such that $\chi _{\underline {a}}(t_i)=a_i$. As before, we extend $\chi _{\underline {a}}$ to a character of $B$. Let $\phi \in i^{\infty }_B(\chi _{\underline {a}})$ be the unique element such that:

  1. (i) $\phi$ is invariant under $I_ {\mathfrak {p}}$;

  2. (ii) the support of $\phi$ is $BI_ {\mathfrak {p}}$; and

  3. (iii) $\phi (1)=1$.

Note that in case $\chi _{\underline {a}}={{\mathbb 1}}$ is the trivial character the image of $\phi$ under the quotient map $i^{\infty }_B(\Omega )\twoheadrightarrow \operatorname {St}^{\infty }(\Omega )$ is non-zero and, thus, generates the space of Iwahori invariants of the Steinberg representation.

By Frobenius reciprocity $\phi$ induces a $G_ {\mathfrak {p}}$-equivariant homomorphism

(3.4)\begin{equation} \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}\Omega \longrightarrow i^{\infty}_B(\chi_{\underline{a}}). \end{equation}

One can argue as in the proof of [Reference OllivierOll14, Proposition 4.4] that the map (3.4) induces an isomorphism

\[ M_{\underline{a}}\longrightarrow i^{\infty}_B(\chi_{\underline{a}}). \]

In conclusion, we see that the augmented Koszul complex

(3.5)\begin{equation} \Lambda^{\bullet}_A(A^{\Delta}) \otimes_A \operatorname{c-ind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}(\Omega)\longrightarrow i^{\infty}_B(\chi_{\underline{a}})\longrightarrow 0 \end{equation}

is exact.

3.2 Overconvergent cohomology

We show that the result from the previous section together with the theory of overconvergent cohomology allows us to construct cohomology classes with values in duals of locally analytic representations.

Before sticking to the case which is most relevant for our applications let us consider the general case: let $A$ be a $ {\mathbb {Q}}_p$-affinoid algebra and $\chi \colon T\to A$ a locally analytic character. Denote by $\chi _0$ its restriction to $T_0$ and fix elements $t_i\in T^{-}$, $i\in \Delta$, as before. For $n\geqslant n_{\chi _0}$ the continuous dual of the augmentation map (3.1) of the previous section yields the map

\[ \operatorname{aug}_\chi^{\ast}\colon\operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}^{\operatorname{an}}_{B}(\chi);A(\epsilon)))\longrightarrow \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{\mathcal{D}}^{n}_{\chi_0})^{\epsilon} \]

for every sign character $\epsilon$ and every integer $d\geqslant 0$. We denote the operator induced by $U_{t_i}$, $i\in \Delta$, on the right-hand side also by $U_{t_i}$ and similar for $U_{\tilde {t}}$. Then, Proposition 3.3 implies the following.

Corollary 3.7 We have

\[ \operatorname{im} (\operatorname{aug}_\chi^{\ast})\subseteq \bigcap_{i \in \Delta}\operatorname{ker} (U_{t_i}-\chi(t_i)). \]

For the remainder of the section we study the case of the trivial character ${{\mathbb 1}}={{\mathbb 1}}_{T}\colon T\to {E}^{\ast }$ of $T$ with values in the units of the $p$-adic field ${E}$. We denote its restriction to $T_0$ also by ${{\mathbb 1}}$. Let $n\geqslant 1$ be any integer. The map (3.3) induces a homomorphism

(3.6)\begin{equation} \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{\mathcal{D}}^{n}_{{{\mathbb 1}}})^{\epsilon} \longrightarrow \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E})^{\epsilon} \end{equation}

in cohomology, that is, equivariant with respect to the commuting actions of the Hecke algebra $ {\mathbb {T}}^{ {\mathfrak {p}}}$ and the operators $U_t$, $t \in T^{-}$.

In the following, we study the finite slope parts of the above cohomology groups (see, for example, [Reference HansenHan17, § 2.3], for definitions and notation).

Theorem 3.8 (Ash–Stevens, Urban and Hansen)

Let $\tilde {t}=\prod _{i\in \Delta }t_i.$

  1. (i) The space $\operatorname {H}^{d}(X_{K^{{\mathfrak {p}}}\times I_ {\mathfrak {p}}}, {\mathcal {D}}^{n}_{{{\mathbb 1}}})^{\epsilon }$ admits a slope decomposition with respect to $U_{\tilde {t}}$ and every rational number $h$.

  2. (ii) If $h$ is small enough with respect to $\tilde {t}$ and the trivial character ${{\mathbb 1}}$ (see, for example, the first formula on page 1690 of [Reference UrbanUrb11] or equation $(21)$ of [Reference Ash and StevensAS08]), then the map

    \[ \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{\mathcal{D}}^{n}_{{{\mathbb 1}}})^{\epsilon,\leqslant h} \longrightarrow \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E})^{\epsilon,\leqslant h}. \]
    is an isomorphism. In particular, this is an isomorphism for $h=0$.

Proof. For the first claim see § 2.3 of [Reference HansenHan17] and for the second see [Reference HansenHan17, Theorem 3.2.5]. The third claim is [Reference UrbanUrb11, Proposition 4.3.10]. Note that in all cases the authors consider all primes lying above $p$ at once. The same proofs work in our partial $ {\mathfrak {p}}$-adic setup.

Composing the dual of the augmentation map

(3.7)\begin{equation} \operatorname{aug}_{{\mathbb 1}}^{\ast}\colon \operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}^{\operatorname{an}}_{B}({E});{E}(\epsilon)))\longrightarrow \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{\mathcal{D}}^{n}_{{{\mathbb 1}}})^{\epsilon} \end{equation}

with the map (3.6) yields the map

(3.8)\begin{equation} \operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}^{\operatorname{an}}_{B}({E});{E}(\epsilon)))\longrightarrow \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E})^{\epsilon}. \end{equation}

As the diagram

is commutative up to multiplication with a non-zero constant, which comes from the choice of a Iwahori-fixed vector in (1.7), so is the induced diagram in cohomology as follows.

Proposition 3.9 Let $a_i\in E^{\times }$, $i\in \Delta$, be elements with $p$-adic valuation equal to $0$ and $\chi _{\underline {a}}\colon T\to E^{\times }$ the unique smooth unramified character such that $\chi _{\underline {a}}(t_i)=a_i$. The natural inclusion $i^{\infty }_{B}(\chi _{\underline {a}})\hookrightarrow {\mathbb {I}}^{\operatorname {an}}_{B}(\chi _{\underline {a}})$ induces an isomorphism

\[ \operatorname{H}^{d}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}^{\operatorname{an}}_{B}(\chi_{\underline{a}});{E}(\epsilon)))\xrightarrow{\cong} \operatorname{H}^{d}(G(F),{\mathcal{C}}_{E}(K^{{\mathfrak {p}}},i^{\infty}_{B}(\chi_{\underline{a}});{E}(\epsilon))) \]

for all $d\geqslant 0$. In particular, the map (3.8) induces an isomorphism

\[ \operatorname{H}^{q}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{E}(K^{{\mathfrak {p}}},{\mathbb{I}}^{\operatorname{an}}_{B}({E});{E}(\epsilon)))[\pi^{{\mathfrak {p}}}]\xrightarrow{\cong} \operatorname{H}^{q}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E})^{\epsilon}[\pi] \]

on $\pi$-isotypic components in degree $q$.

Proof. The second claim is a direct consequence of the first claim and Lemma 2.3.

Given an $I_ {\mathfrak {p}}$-representation $M$ we denote by $\operatorname {Coind}_{I_ {\mathfrak {p}}}^{G_ {\mathfrak {p}}}M$ the coinduction of $M$ to $G_ {\mathfrak {p}}$, i.e. the module of all functions $f\colon G_ {\mathfrak {p}}\to M$ such that $f(kg)=kf(g)$ for all $k\in I_ {\mathfrak {p}}$, $g\in G_ {\mathfrak {p}}$. By taking continuous duals the Koszul complex of Theorem 3.4 with $y_i=U_{t_i}-a_i$ yields a quasi-isomorphism

\[ \operatorname{Hom}_E^{\operatorname{ct}}({\mathbb{I}}^{\operatorname{an}}_{B}(\chi_{\underline{a}}),E) \xrightarrow{\cong} \Lambda^{\bullet}_{E}({E}^{\Delta}) \otimes_{E} \operatorname{Coind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}({\mathcal{D}}^{n}_{{{\mathbb 1}}}) \]

of complexes. Note that taking continuous duals in this case is exact by the Hahn–Banach theorem. Similarly by the smooth Koszul resolution (3.5) we have a quasi-isomorphism

\[ \operatorname{Hom}_E(i^{\infty}_{B}(\chi_{\underline{a}}),E) \xrightarrow{\cong} \Lambda^{\bullet}_{E}({E}^{\Delta}) \otimes_{E} \operatorname{Coind}_{I_{\mathfrak{p}}}^{G_{\mathfrak{p}}}({E}) \]

and the canonical diagram

is commutative.

From the associated spectral sequences for double complex, we deduce that it is enough to prove that the map of complexes

(3.9)\begin{equation} \Lambda^{\bullet}_{E}({E}^{\Delta})\otimes_{E} \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{\mathcal{D}}^{n}_{{{\mathbb 1}}}) \longrightarrow \Lambda^{\bullet}_{E}({E}^{\Delta})\otimes_{E} \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E}) \end{equation}

is a quasi-isomorphism for all $d\geqslant 0.$ The modules on the left-hand side admit a slope decomposition for $U_{\tilde {t}}$ by Theorem 3.8(i), whereas the modules on the right-hand side admit a slope decomposition because they are finite-dimensional ${E}$-vector spaces.

As the operators $y_i$ commute with $U_{\tilde {t}}$ they respect the slope decomposition on both sides (see [Reference UrbanUrb11, Lemma 2.3.2]). On the slope less than or equal to zero part the map (3.9) is even an isomorphism of complexes by Theorem 3.8(ii).

Thus, we are reduced to proving that

\[ \Lambda^{\bullet}_{E}({E}^{\Delta})\otimes_{E} \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{\mathcal{D}}^{n}_{{{\mathbb 1}}})^{>h} \longrightarrow \Lambda^{\bullet}_{E}({E}^{\Delta})\otimes_{E} \operatorname{H}^{d}(X_{K^{{\mathfrak {p}}}\times I_{\mathfrak{p}}},{E})^{>h} \]

is a quasi-isomorphism for all $d\geqslant 0$.

In fact, both sides are acyclic: standard properties of Koszul complexes imply that the operators $y_i$ act as multiplication by zero on the cohomology of the complexes. By the definition of a slope decomposition the operator $U_{\tilde {t}}-\prod _{i\in \Delta } a_i$ acts via isomorphisms on both complexes and, thus, on their cohomology. However, because $U_{\tilde {t}}-\prod _{i\in \Delta } a_i$ lies in the ideal generated by the operators $y_i$, it also acts via multiplication by zero, which proves the claim.

Remark 3.10 In order to keep the article short and avoid unnecessary notation, we decided to stick to the special case above. However, Proposition 3.9 also holds in a much more general situation, e.g. one can allow arbitrary coefficient systems and arbitrary non-critical principal series representations, given that the slope is small enough with respect to $\tilde {t}$.

3.3 Overconvergent families

Let $ {\mathcal {W}}$ be the weight space of $T$, i.e. it is the rigid space over ${E}$ such that for every ${E}$-affinoid algebra $A$ its $A$-points are given by

\[ {\mathcal{W}}(A)=\operatorname{Hom}^{\operatorname{ct}}(T,A^{\ast}). \]

It is smaller than the usual weight space as we only consider characters of the torus in $G_{F_ {\mathfrak {p}}}$. For any open affinoid $\mathcal {U} \subset {\mathcal {W}}$ and we let $\chi _{\mathcal {U}}$ be the corresponding universal weight. As in § 3.1 we define the space of $n$-analytic functions and distributions $ {\mathcal {A}}_{\chi _{\mathcal {U}}}^{n}$ and $ {\mathcal {D}}^{n}_{\chi _{\mathcal {U}}}$ for $n\gg 0$.

We suppose now that the group $G_\infty$ has discrete series or, equivalently, that $\delta =0$ (see [Reference KnappKna01, Theorem 12.20] for the equivalence of these two properties). Thus, by Proposition 1.6, the representation $\pi ^{\infty }$ appears only in the middle degree cohomology $\operatorname {H}^{q}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {E})$. Hence, we put $\mathcal {L}_{i}(\pi, {\mathfrak {p}})^{\epsilon }=\mathcal {L}_{i}^{(q)}(\pi, {\mathfrak {p}})^{\epsilon }$.

Let $ {\mathbb {T}}_{\operatorname {sph}}^{ {\mathfrak {p}}}\subseteq {\mathbb {T}}^{ {\mathfrak {p}}}$ be the spherical Hecke algebra of level $K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}$ over ${E}$, i.e. the commutative Hecke algebra generated by all Hecke operators at finite places $v$ such that $K_v$ is hyperspecial. We define $ {\mathbb {T}}_{\operatorname {sph}}$ to be the commutative algebra generated by $ {\mathbb {T}}_{\operatorname {sph}}$ and all $U_t$-operators for $t\in T^{-}$. Let $ {\mathfrak {m}}_\pi \subseteq {\mathbb {T}}_{\operatorname {sph}}$ (respectively, $ {\mathfrak {m}}_\pi ^{ {\mathfrak {p}}}\subseteq {\mathbb {T}}_{\operatorname {sph}}^{ {\mathfrak {p}}}$) the maximal ideal associated with $\pi$. We assume the following weak non-Eisenstein assumption on the maximal ideal $ {\mathfrak {m}}_\pi$ throughout this section.

Hypothesis (NE)

We have

\begin{align*} \operatorname{H}^{d}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {E})_{{\mathfrak{m}}_\pi}=0 \end{align*}

unless $d=q$.

Example 3.11 If the group $G$ is definite, the hypothesis is automatically true. By strong multiplicity one and the Jacquet–Langlands correspondence, it is also true for inner forms of $\operatorname {PGL}_2$ over totally real number fields.

Let $\mathcal {U}$ be an open affinoid of $ {\mathcal {W}}$ and define $\mathcal {O}_{\mathcal {U},{{\mathbb 1}}}$ to be the rigid localization of $\mathcal {O}_{ {\mathcal {W}}}(\mathcal {U})$ at the weight ${{\mathbb 1}}$ (which is the cohomological weight of $\pi$), i.e. 

\[ \mathcal{O}_{\mathcal{U},{{\mathbb 1}}}= \varinjlim_{{{\mathbb 1}} \in \mathcal{U}' \subset \mathcal{U}} \mathcal{O}_{{\mathcal{W}}}(\mathcal{U}'). \]

The following theorem is instrumental in calculating $\mathcal {L}$-invariants.

Theorem 3.12 After localization at the ideal $ {\mathfrak {m}}_{\pi }$ of $ {\mathbb {T}}_{\operatorname {sph}}$ and restricting to a small enough open affinoid $\mathcal {U}$ containing ${{\mathbb 1}}$, the canonical reduction map

\[ \operatorname{H}^{q} (X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}^{n}_{\chi_{\mathcal{U}}} )^{\epsilon}_{{\mathfrak{m}}_{\pi}}\rightarrow \operatorname{H}^{q}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {E})^{\epsilon}_{{\mathfrak{m}}_{\pi}} \]

is surjective. Moreover, $\operatorname {H}^{q} (X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}^{n}_{\chi _{\mathcal {U}}} )^{\epsilon }_{ {\mathfrak {m}}_\pi }$ is a free $\mathcal {O}_{\mathcal {U},{{\mathbb 1}}}$-module of rank equal to the dimension of $\operatorname {H}^{q}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {E})^{\epsilon }_{ {\mathfrak {m}}_{\pi }}$.

In addition, we have $\operatorname {H}^{d} (X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}^{n}_{\chi _{\mathcal {U}}} )^{\epsilon }_{ {\mathfrak {m}}_{\pi }}=0$ for all $d \neq q$.

Proof. The case of $\operatorname {PGL}_2$ over a totally real field is proven in detail in [Reference Barrera Salazar, Dimitrov and JorzaBDJ21, Theorem 2.14]. The main ingredient in their proof is the vanishing of the cohomology outside middle degree. Thus, the same proof works in our more general setup.

Definition 3.13 We say that $ {\mathfrak {m}}_\pi$ is $ {\mathfrak {p}}$-étale (with respect to $\epsilon$) if every Hecke operator $h\in {\mathbb {T}}_{\operatorname {sph}}$ acts on $\operatorname {H}^{q} (X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}^{n}_{\chi _{\mathcal {U}}} )^{\epsilon }_{ {\mathfrak {m}}_{\pi }}$ as multiplication by an element $\alpha _h\in \mathcal {O}_{\mathcal {U},{{\mathbb 1}}}^{\ast }.$

Theorem 3.12 immediately implies the following.

Corollary 3.14 Assume that $\dim _{E} \operatorname {H}^{q}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {E})^{\epsilon }_{ {\mathfrak {m}}_{\pi }}=1$. Then $ {\mathfrak {m}}_\pi$ is $ {\mathfrak {p}}$-étale.

Example 3.15 Suppose $G$ is an inner form of $\operatorname {PGL}_2$ over a totally real number field. Then the above corollary together with strong multiplicity one implies that $ {\mathfrak {m}}_\pi$ is $ {\mathfrak {p}}$-étale.

Let $t_i\in T^{-}$, $i\in \Delta$, be a choice of elements as in § 3.1. Suppose $ {\mathfrak {m}}_\pi$ is $ {\mathfrak {p}}$-étale (with respect to $\epsilon$) with associated eigenvalues $\alpha _{U_{t_i}}\in \mathcal {O}_{\mathcal {U},{{\mathbb 1}}}^{\ast }$, $i \in \Delta$. By possibly shrinking $\mathcal {U}$, we may assume that $\alpha _{U_{t_i}}\in \mathcal {O}_{\mathcal {U}}^{\ast }$. Every element $t\in T$ can be written uniquely as a product $t=t_0 \prod _{i\in \Delta } t_i^{n_i}$ with $t_0 \in T_0$, $n_i\in {\mathbb {Z}}$. Hence, there exists a unique character $\chi _\alpha \colon T\to \mathcal {O}_{\mathcal {U}}^{\ast }$ such that

\[ \left.\chi_{\alpha}\right|_{T_0}=\chi_{U} \]

and

\begin{align*} \chi_\alpha(t_i)=\alpha_{U_{t_i}}. \end{align*}

Theorem 3.16 Suppose that $ {\mathfrak {m}}_\pi$ is $ {\mathfrak {p}}$-étale (with respect to $\epsilon$) with associated character $\chi _\alpha \colon T \to \mathcal {O}_{\mathcal {U}}^{\ast }$. Then for every tangent vector $v$ of $\mathcal {U}$ at ${{\mathbb 1}}$ we have:

\[ \frac{\partial}{\partial v}\chi_{\alpha,i}\in \mathcal{L}_{i}(\pi,{\mathfrak{p}})^{\epsilon} \]

where $\chi _{\alpha,i}$ denotes the composition $\chi _{\alpha }\circ i^{\vee }.$ Moreover, the subspace $\mathcal {L}_{i}(\pi, {\mathfrak {p}})^{\epsilon }\subseteq \operatorname {Hom}^{\operatorname {ct}}(F_ {\mathfrak {p}}^{\ast },{E})$ has codimension one.

Proof. Let $ {\mathfrak {m}}_{{\mathbb 1}}\subseteq \mathcal {O}_{\mathcal {U}}$ be the maximal ideal corresponding to the trivial character. We put $A=\mathcal {O}_{\mathcal {U}}/ {\mathfrak {m}}_{{{\mathbb 1}}}^{2}$ and $\chi _0 = \chi _{\mathcal {U}} \bmod {\mathfrak {m}}_{{{\mathbb 1}}}^{2}$. Let us write $\widetilde {\chi }_\alpha = \chi _\alpha \bmod {\mathfrak {m}}_{{{\mathbb 1}}}^{2}\colon T \to A$. Using Theorem 3.12 and arguments with Koszul complexes as in the proof of Proposition 3.9 one shows that the image of the map

\[ \operatorname{H}^{q}(G(F),{\mathcal{C}}^{\operatorname{ct}}_{A}(K^{{\mathfrak {p}}},{\mathbb{I}}_{B}^{\operatorname{an}}(\widetilde{\chi}_\alpha);A(\epsilon)))_{{\mathfrak{m}}_{\pi}^{{\mathfrak{p}}}}\longrightarrow \operatorname{H}^{q} (X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, E )^{\epsilon}_{{\mathfrak{m}}_{\pi}^{{\mathfrak{p}}}} \]

is the intersection of the kernels of the homomorphisms $U_{t_i}-1$. Therefore, the first claim follows from Proposition 2.4.

We may assume that ${E}$ is large enough. We explain how to choose appropriate tangent vectors so that $\mathcal {L}_{i}(\pi, {\mathfrak {p}})^{\epsilon }$ contains elements of the form $\log _{p,\sigma }-\mathcal {L}^{\sigma }\operatorname {ord}_ {\mathfrak {p}}$ for every embedding $\sigma \colon F_ {\mathfrak {p}} \to {E}$. This will show that the $\mathcal {L}$-invariant has codimension at most one.

Let ${E} \langle k_\sigma \rangle$ be the ring of power series in the $k_\sigma$ and let $\mathcal {O}_{ {\mathfrak {p}},1}^{\ast }$ denote the free part of $\mathcal {O}_{ {\mathfrak {p}}}^{\ast }$. By smoothness of the weight space, we can embed $\mathcal {O}_{\mathcal {U}}^{\ast }$ in ${E} \langle k_\sigma \rangle$ and the structural morphism of ${E} [\kern-1pt[ \mathcal {O}_{ {\mathfrak {p}},1}^{\ast } ]\kern-1pt] $ into $\mathcal {O}_{\mathcal {U}}^{\ast }$ can be described via the map

\[ \mathcal{O}_{\mathfrak{p}}^{\ast} \ni \langle u \rangle\mapsto \prod_{\sigma } \sigma(u)^{k_{\sigma}}. \]

We can, hence, identify $\chi _{\alpha,i}$ with a map from $F_{ {\mathfrak {p}}}^{\ast }$ to $\mathcal {O}_{\mathcal {U}}^{\ast }$. For all $n \in \mathbb {Z}$ and $u \in \mathcal {O}_{ {\mathfrak {p}},1}^{\ast }$ we have

\[ \chi_{\alpha,i}(\pi^{n} u)=\alpha_{U_{t_i}}^{n} \prod_{\sigma } \sigma(u)^{k_{\sigma}}. \]

Note that

\[ \frac{{d}}{{d}k_{\sigma'}} (\sigma(u)^{k_{\sigma}} )=\delta_{\sigma,\sigma'} \log_p(\sigma(u))\sigma(u)^{k_{\sigma}}, \]

where

\[ \delta_{\sigma,\sigma'} = \begin{cases} 1 & \mbox{if}\ \sigma=\sigma' \\ 0 & \mbox{otherwise.}\end{cases} \]

Take for $v$ the direction where only $k_{\sigma }$ varies, derive $\chi _{\alpha,i}$ along $v$, and evaluate at $k_{\sigma '}=0$ for all $\sigma '$ (corresponding to the weight ${{\mathbb 1}}$) to obtain

\[ \frac{\partial}{\partial v}\chi_{\alpha,i}(\pi^{n} u)= \frac{{d}}{ {d}k_{\sigma}}\alpha_{U_{t_i}}({{\mathbb 1}})\operatorname{ord}_{\mathfrak{p}}(\pi^{n})+\alpha_{U_{t_i}}({{\mathbb 1}})\log_p(\sigma(u)). \]

By the Steinberg hypothesis, $\alpha _{U_{t_i}}({{\mathbb 1}})$ is not vanishing and we are done.

By Proposition 1.8, the $\mathcal {L}$-invariant has codimension at least one and, hence, the second claim follows.

4. Applications

4.1 Hilbert modular forms

We want to study the case of inner forms of $\operatorname {PGL}_2$, which are split at $ {\mathfrak {p}}$, over a totally real number field $F$ in detail. As there is only one simple root we drop it from the notation. If $G$ is equal to $\operatorname {PGL}_2$, one can attach $2^{[F: {\mathbb {Q}}]}$ a priori different $\mathcal {L}$-invariants $\mathcal {L}(\pi,p)^{\epsilon }$ to $\pi$; in this case, a conjecture of Spieß (cf. [Reference SpießSpi14, Conjecture 6.4]) states that the definition does not depend on $\epsilon$, i.e.

\[ \mathcal{L}(\pi,{\mathfrak{p}})^{\epsilon}=\mathcal{L}(\pi,{\mathfrak{p}})^{\epsilon'} \]

for all choices of sign characters $\epsilon$ and $\epsilon '$. In the same paper, Remark 6.6(b), Spieß also states that ‘an interesting and difficult problem’ is to show that the $\mathcal {L}$-invariant is invariant under Jacquet–Langlands transfers. In this section, thanks to Theorem 3.16, we settle both Spieß’ conjecture and his question.

Moreover, let $\rho _{\pi }\colon \operatorname {Gal}(\overline { {\mathbb {Q}}}/F)\to \operatorname {GL}_2({E})$ be the Galois representation associated with $\pi$ (or rather associated with the Jacquet–Langlands transfer $\operatorname {JL}(\pi )$ of $\pi$ to $\operatorname {PGL}_2$), which exists by work of Taylor (see [Reference TaylorTay89]). As $\pi$ is $ {\mathfrak {p}}$-ordinary, by Theorem 2 of [Reference WilesWil88] we know that the restriction $\rho _{\pi, {\mathfrak {p}}}\colon \operatorname {Gal}(\overline { {\mathbb {Q}}_p}/F_ {\mathfrak {p}})$ of $\rho _{\pi }$ to a decomposition group at $ {\mathfrak {p}}$ is ordinary. Moreover, because $\pi _ {\mathfrak {p}}$ is Steinberg, a result of Saito (see [Reference SaitoSai09]) implies that this restriction is a non-split extension of the trivial character by the cyclotomic character. Therefore, it gives a class $\langle \rho _{\pi, {\mathfrak {p}}}\rangle \in \operatorname {H}^{1}(\operatorname {Gal}(\overline { {\mathbb {Q}}_p}/F_ {\mathfrak {p}}),{E}(1))$, where ${E}(1)$ denotes the Tate twist of ${E}$. The Fontaine–Mazur $\mathcal {L}$-invariant $\mathcal {L}^{\operatorname {FM}}(\rho _{\pi, {\mathfrak {p}}})$ of $\rho _{\pi, {\mathfrak {p}}}$ is the orthogonal complement of $\langle \rho _{\pi, {\mathfrak {p}}}\rangle$ with respect to the local Tate pairing

\[ \operatorname{H}^{1}(\operatorname{Gal}(\overline{{\mathbb{Q}}_p}/F_{\mathfrak{p}}),{E}(1)) \times \operatorname{H}^{1}(\operatorname{Gal}(\overline{{\mathbb{Q}}_p}/F_{\mathfrak{p}}),{E}) \longrightarrow {E}. \]

By local class field theory we have a canonical isomorphism

\[ \operatorname{H}^{1}(\operatorname{Gal}(\overline{{\mathbb{Q}}_p}/F_{\mathfrak{p}}),{E})\cong \operatorname{Hom}^{\operatorname{ct}}(F_{\mathfrak{p}}^{\ast},{E}) \]

and, thus, we consider $\mathcal {L}^{\operatorname {FM}}(\rho _{\pi, {\mathfrak {p}}})$ as a codimension-one subspace of $\operatorname {Hom}^{\operatorname {ct}}(F_ {\mathfrak {p}}^{\ast },{E})$. We show that automorphic $\mathcal {L}$-invariants equal the Fontaine–Mazur $\mathcal {L}$-invariant of the associated Galois representation.

Theorem 4.1 Suppose $G$ is an inner form of $\operatorname {PGL}_2$ over a totally real number field $F$, which is split at the prime $ {\mathfrak {p}}$ of $F$. Let $\pi$ be a cuspidal automorphic representation of parallel weight $2$ of $G$ that is Steinberg at $ {\mathfrak {p}}$.

  1. (i) The automorphic $\mathcal {L}$-invariant of $\pi$ is independent of $\epsilon$, i.e. 

    \[ \mathcal{L}(\pi,{\mathfrak{p}})^{\epsilon} = \mathcal{L}(\pi,{\mathfrak{p}})^{\epsilon'} \]
    for all choices of $\epsilon$ and $\epsilon '$. We therefore put $\mathcal {L}(\pi, {\mathfrak {p}})=\mathcal {L}(\pi, {\mathfrak {p}})^{\epsilon }$ for any choice of sign character $\epsilon$.
  2. (ii) Let $\operatorname {JL}(\pi )$ be the Jacquet–Langlands transfer of $\pi$ to $\operatorname {PGL}_2$. The equality

    \[ \mathcal{L}(\pi,{\mathfrak{p}}) = \mathcal{L}(\operatorname{JL}(\pi),{\mathfrak{p}}) \]
    holds.
  3. (iii) Let $\rho _\pi$ the Galois representation associated with $\pi$. Then

    \[ \mathcal{L}(\pi,{\mathfrak{p}})=\mathcal{L}^{\operatorname{FM}}(\rho_{\pi,{\mathfrak{p}}}). \]

Proof. We actually prove that $\mathcal {L}(\pi, {\mathfrak {p}})^{\epsilon }=\mathcal {L}^{\operatorname {FM}}(\rho _{\pi, {\mathfrak {p}}})$ for every $\epsilon$, which implies all other claims. As $ {\mathfrak {m}}_\pi$ is $ {\mathfrak {p}}$-étale (with respect to $\epsilon$) we can deform the Hecke eigenvalues of $\pi$ to a family over an open affinoid subspace of weight space in the following sense: there exists an open affinoid $\mathcal {U}\subseteq {\mathcal {W}}$ containing the trivial character such that the eigenvalue $\alpha _h$ of each Hecke operator $h\in {\mathbb {T}}_{\operatorname {sph}}$ acting on $\operatorname {H}^{q} (X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}^{n}_{\chi _{\mathcal {U}}} )^{\epsilon }$ is an element of $\mathcal {O}_{ {\mathcal {W}}}(\mathcal {U})$. As explained in [Reference Barrera Salazar, Dimitrov and JorzaBDJ21, Theorem 2.14(ii)], we may shrink $\mathcal {U}$ further such that the common eigenspace associated to the eigenvalues $\alpha _h$ is a free $\mathcal {O}_{ {\mathcal {W}}}(\mathcal {U})$-submodule of $\operatorname {H}^{q} (X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}^{n}_{\chi _{\mathcal {U}}} )^{\epsilon }$ of rank equal to the dimension of $\operatorname {H}^{q}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {E})^{\epsilon }_{ {\mathfrak {m}}_{\pi }}$. By shrinking $\mathcal {U}$ even further we may assume that $\alpha _{t_i}\in \mathcal {O}_{ {\mathcal {W}}}(\mathcal {U})^{\times }$ has absolute value one for each $i\in \Delta$. Then, by the classicality of small slope overconvergent eigenforms, the specialization of these eigenvalues at a classical point $\lambda \in \mathcal {U}$ are the eigenvalues associated to a cuspidal representation $\pi _\lambda$ of weight $\lambda$.

Note that the Galois representation $\rho _{\pi }$ is irreducible, as the Hilbert modular form associated with $\pi$ is cuspidal. Thus, by standard arguments (see, for example, Theorem B of [Reference ChenevierChe14]) there exists (after possibly shrinking $\mathcal {U}$ again) a family of Galois representations $\rho _{\pi,\mathcal {U}}$ over $\mathcal {U}$ passing through $\rho _{\pi }$, i.e. the specialization of $\rho _{\pi,\mathcal {U}}$ at each classical point $\lambda \in \mathcal {U}$ is isomorphic to the Galois representation attached to $\pi _\lambda$. It follows from the results of Saito (see Theorem 1 of [Reference SaitoSai09]) that for every classical point $z\in \mathcal {U}$ the restriction of the Galois representation $\rho _{\pi,z}$ to a decomposition group is determined by the $U_ {\mathfrak {p}}$-eigenvalue. Moreover, the Weil–Deligne module of $\rho _{\pi }$ restricted at the decomposition group at $ {\mathfrak {p}}$ is the Weil–Deligne module associated with the Steinberg representation via the local Langlands correspondence. In particular, our representation (with the filtration induced by the ordinarity of the Galois representation) is non-critical special, as defined in § 3.1 of [Reference DingDin19]; indeed we know that the corresponding $(\varphi,\Gamma )$-module is an extension of special characters, so we just have to check whether the extensions are split or not. If one extension were split, then the Weil–Deligne representation would also be split, contradicting the fact the monodromy for the Steinberg is maximal.

Comparing the Galois theoretic Colmez–Greenberg–Stevens formula (see [Reference DingDin19, Theorem 3.4]) with its automorphic counterpart of Theorem 3.16 yields the result.

Remark 4.2 (i) In the Hilbert modular form case, the constructions simplify substantially and, thus, one can check that our methods also work for higher weights assuming that the representation is non-critical.

(ii) One can prove the first two claims of Theorem 4.1 without passing to the Galois side. Namely, using methods as in § 5 of [Reference HansenHan17] one can show that the eigenvalue of the overconvergent $ {\mathfrak {p}}$-adic family passing through $\pi$ is independent of the sign character and stable under the Jacquet–Langlands transfer. Thus, by Theorem 3.16, we can conclude the argument.

(iii) By definition, Fontaine–Mazur $\mathcal {L}$-invariants are stable under restricting the Galois representation to the absolute Galois group of finite extensions. Thus, we can deduce from Theorem 4.1 that automorphic $\mathcal {L}$-invariants of Hilbert modular forms are stable under abelian base change with respect to totally real extensions.

(iv) In the case of modular elliptic curves over totally real fields of class number one the equality of automorphic and Fontaine–Mazur $\mathcal {L}$-invariants was conjectured by Greenberg (see [Reference GreenbergGre09, Conjecture 2]). Theorem 4.1 thus implies that the construction of Stark–Heegner points over totally real fields is unconditional (see [Reference Guitart, Masdeu and SengünGMS15] for a detailed discussion of Stark–Heegner points).

4.2 Unitary groups

Let $\tilde {F}$ be a CM field with totally real subfield $F$. We assume that the prime $ {\mathfrak {p}}$ of $F$ is split in $\tilde {F}$. Let $U$ be the unitary group attached to a positive definite hermitian space over $\tilde {F}$ and $G$ the associated adjoint group. By construction, $G_ {\mathfrak {p}}$ is isomorphic to $\operatorname {PGL}_n(F_ {\mathfrak {p}})$. We can identify the simple roots with respect to the upper triangular Borel with the set $\{1,\ldots, n-1\}$. Let $\pi$ be an automorphic representation of $G$ such that $\pi _\infty = {\mathbb {C}}$ and $\pi _ {\mathfrak {p}}$ is the Steinberg representation of $\operatorname {PGL}_n(F_ {\mathfrak {p}})$. In this case the only sign character is the trivial character and, therefore, we drop it from the notation.

By Shin's appendix to [Reference GoldringGol14] the base change $\operatorname {BC}(\pi )$ of $\pi$ to $\operatorname {PGL}_n$ over $\tilde {F}$ exists and it is Steinberg at both primes of $\tilde {F}$ lying above $ {\mathfrak {p}}$. By the work of many people (see Theorem 2.1.1 of [Reference Barnet-Lamb, Gee, Geraghty and TaylorBGGT14] for a detailed discussion) we can attach a $p$-adic Galois representation

\[ \rho_{\pi}=\rho_{\operatorname{BC}(\pi)}\colon \operatorname{Gal}(\overline{{\mathbb{Q}}}/\tilde{F})\longrightarrow \operatorname{GL}_n({E}) \]

to $\operatorname {BC}(\pi )$. As we consider the trivial coefficient system the Steinberg representation is ordinary (cf. [Reference GeraghtyGer19, Lemma 5.6]). Therefore, as shown in [Reference ThorneTho15, Theorem 2.4], one deduces from the local–global compatibility theorem of Caraiani (see [Reference CaraianiCar14]) that the restriction $\rho _{\pi, {\mathfrak {p}}}$ of $\rho _{\pi }$ to the decomposition group of a prime above $ {\mathfrak {p}}$ can be brought in the following form: it is upper triangular and the $i$th diagonal entry is the $(1-i)$-power of the cyclotomic character.

Therefore, we have $n-1$ canonical two-dimensional subquotients $\rho _{\pi, {\mathfrak {p}},i}$ which are extensions of ${E}(-n+i)$ by ${E}(-n+i+1)$. We can consider the associated Fontaine–Mazur $\mathcal {L}$-invariant $\mathcal {L}^{\operatorname {FM}}_{i}(\rho _{\pi, {\mathfrak {p}}})=\mathcal {L}^{\operatorname {FM}}(\rho _{\pi,\pi,i}(n-i))$. Recall that the Weil–Deligne representation associated with the Steinberg has always maximal monodromy so the corresponding $(\varphi,\Gamma )$-module is non-critical split à la Ding, see § 3.1 of [Reference DingDin19]. Replacing the local–global compatibility results of Saito by those of Caraiani (see [Reference CaraianiCar14]), we can apply Theorem 3.4 of [Reference DingDin19] to the $(\varphi,\Gamma )$-module associated with the Galois representation $\rho _{\pi }$ and then the same proof as that of Theorem 4.1 yields the following result.

Theorem 4.3 Let $F$ be a totally real number field and $G$ the adjoint of a unitary group over $F$ compact at infinity and split at $ {\mathfrak {p}}$. Let $\pi$ be an automorphic representation of $G$ such that $\pi _\infty = {\mathbb {C}}$ and $\pi _{ {\mathfrak {p}}}$ is Steinberg. Suppose $\pi$ satisfies (SMO), that we can choose the tame level $K^{ {\mathfrak {p}}}$ such that $ {\mathfrak {m}}_{\pi }$ is $ {\mathfrak {p}}$-étale and that the Galois representation $\rho _{\pi }$ attached to $\pi$ is irreducible. Then we have

\[ \mathcal{L}^{\operatorname{FM}}_{i}(\rho_{\pi}) = \mathcal{L}_{i}(\pi,{\mathfrak{p}}) \]

for every $i=1,\ldots,n-1$.

Remark 4.4 Something can be said also when $F =\mathbb {Q}$ and $G=\operatorname {Sp}_{2g}$. In this case the Galois representations have been constructed by Scholze [Reference ScholzeSch15] but local–global compatibility is not known. Still, if one supposes that the $2g+1$-dimensional Galois representation $\operatorname {Std}(\rho _{\pi })$ associated with $\pi$ is semistable with maximal monodromy, then the Greenberg–Benois $\mathcal {L}(\operatorname {Std}(\rho _{\pi }))$-invariant has been calculated in [Reference RossoRos15, Theorem 1.3].

In this case the root system of $\operatorname {Sp}_{2g}$ can be identified with the set $\{1,\ldots,g\}$ via the identification with the root system of $\operatorname {GL}_g$, which is embedded in $\operatorname {Sp}_{2g}$ by

\[ \operatorname{GL}_g\longrightarrow \operatorname{Sp}_{2g},\quad A\longmapsto \biggl(\!\!\!\!\begin{array}{cc} \phantom{e}{}^{\rm t} A^{-1} & 0\\ 0 & A \end{array}\!\!\biggr), \]

and we see that the automorphic $\mathcal {L}$-invariant $\mathcal {L}_{1}(\pi,p)$ coincides with that of [Reference RossoRos15].

The same calculation in § 4.2 of [Reference RossoRos15] gives us that $\mathcal {L}_{i}(\pi,p)$ is the Greenberg–Benois $\mathcal {L}$-invariant for $\operatorname {Std}(\rho _{\pi })(i-1)$. (This case has not been treated in [Reference RossoRos15] as the $L$-values are not Deligne-critical, but Benois’ definition applies also in this case, see formula (96) in [Reference BenoisBen21].)

If $F_ {\mathfrak {p}} \neq {\mathbb {Q}}_p$, then the comparison is more subtle, as there is only just one Greenberg–Benois $\mathcal {L}$-invariant per $p$-adic place, and from the Galois side one needs to consider Galois invariant characters of $F^{*}_ {\mathfrak {p}}$.

5. Beyond discrete series: the Bianchi case

When $G_\infty$ does not fulfil the Harish-Chandra condition, the representation $\pi$ contributes to several degrees of cohomology of the associated locally symmetric space and the techniques used in Theorem 3.12 break down.

There are two tools to tackle the problem: first, one can use Hansen's Tor-spectral sequence (see Theorem 3.3.1 of [Reference HansenHan17])

\[ \operatorname{Tor}_{i}^{\mathcal{O}_{{\mathcal{W}}}} (\operatorname{H}^{q+i}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{U}}})^{\leqslant h}, {E} ) \Longrightarrow \operatorname{H}^{q}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\Sigma}})^{\leqslant h}, \]

where $\mathcal {U}\subseteq {\mathcal {W}}$ is an open affinoid and $\Sigma \subseteq \mathcal {U}$ is Zariski-closed, to analyse the overconvergent cohomology groups in question; second, one can use cases of Langlands functoriality in $p$-adic families to reduce to groups, which fulfil the Harish-Chandra condition.

In good situations it should be possible to calculate at least one of the $\mathcal {L}$-invariants $\mathcal {L}^{(d)}_{i}(\pi, {\mathfrak {p}})$ for $0 \leqslant d\leqslant \delta$, using Proposition 2.4. The main difficulty is that, in general, most classes do not lift to a big cohomology class as classes in $\operatorname {H}^{q+1}$ give lifting obstructions. However, as soon as one can show that at least a class lifts to a family in $\operatorname {H}^{q+d}$ (combined with some results on the tangent directions in the eigenvariety) Proposition 2.4 lets us calculate $\mathcal {L}^{(d)}_{i}(\pi, {\mathfrak {p}})$. If Venkatesh's conjecture (stating that the $\pi$-isotypic component of the cohomology is generated by the minimal degree cohomology as a module over the derived Hecke algebra) holds, then these $\mathcal {L}$-invariants $\mathcal {L}^{(d)}_{i}(\pi, {\mathfrak {p}})$, for varying $d$, are essentially all the same by the main result of [Reference GehrmannGeh19a].

Using unpublished work of Hansen (see also [Reference Barrera Salazar and WilliamsBW21]) we study the case of Bianchi modular forms. We now fix $F$ to be a quadratic imaginary field where $p$ is unramified and $\pi$ a cuspidal representation of $\operatorname {PGL}_{2,F}$ of parallel weight $2$ such that $\pi _ {\mathfrak {q}}$ is Steinberg for all primes $ {\mathfrak {q}}$ lying above $p$. We put $G=\operatorname {PGL}_{2,F}$ if $p$ is inert and, if $p$ is split, we define $G$ to be the Weil restriction $\operatorname {Res}_{F/ {\mathbb {Q}}}\operatorname {PGL}_{2,F}$. In the first case there is only one simple root and, thus, we drop it from the notation. In the second case, the simple roots can be identified with the two primes above $ {\mathfrak {p}}$. In both cases, the only sign character is the trivial one and we shall drop it from the notation as well. (The following theorem indicates that in the case of a split prime the partial eigenvarieties we considered before may not be big enough. The case that $p$ is split and $\pi$ is only Steinberg at one of the primes above $p$ could be handled similarly but one would have to introduce new notation. For the sake of brevity, we do not discuss it further.)

We recall some result on the eigenvariety for Bianchi modular forms due to Hansen and Barrera–Williams (see [Reference Barrera Salazar and WilliamsBW21, Lemma 4.4] and the proof of Theorem 4.5).

Theorem 5.1 (Hansen, Barrera–Williams)

Let $\mathcal {U}\subseteq {\mathcal {W}}$ be an open affinoid neighbourhood of the trivial character.

  1. (i) The system of eigenvalues associated with $\pi$ appears in $\operatorname {H}^{d}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}_{\chi _{\mathcal {U}}})$ if and only if $d=2$.

  2. (ii) There is at least one curve $\mathcal {S} \subseteq \mathcal {U}$ passing through ${{\mathbb 1}}$ such that

    \[ \operatorname{H}^{1}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{S}}})_{{\mathfrak{m}}_{\pi}}\neq 0. \]
    If such a curve $\mathcal {S}$ is smooth at ${{\mathbb 1}}$, the space is free of rank one over $\mathcal {O}_{\mathcal {S}, {{\mathbb 1}}}$ and the canonical map
    \[ \operatorname{H}^{1}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{S}}})_{{\mathfrak{m}}_{\pi}}\longrightarrow \operatorname{H}^{1}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {E})_{{\mathfrak{m}}_{\pi}} \]
    is surjective.

The following proposition completes the picture by taking into account the cohomology in degree two.

Proposition 5.2 For every curve $\mathcal {S}$ as in Theorem 5.1(ii) that is smooth at ${{\mathbb 1}}$ we have

\[ \operatorname{H}^{2}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{S}}})_{{\mathfrak{m}}_{\pi}}\neq 0. \]

More precisely, the space is free of rank one over $\mathcal {O}_{\mathcal {S}, {{\mathbb 1}}}$ and the canonical map

\[ \operatorname{H}^{2}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{S}}})_{{\mathfrak{m}}_{\pi}}\otimes_{\mathcal{O}_\mathcal{S}(\mathcal{S})}{E}\longrightarrow \operatorname{H}^{2}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {E})_{{\mathfrak{m}}_{\pi}} \]

is an isomorphism. Moreover, $\operatorname {H}^{2}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}_{\chi _{\mathcal {S}}})_{ {\mathfrak {m}}_{\pi }}$ and $\operatorname {H}^{1}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}_{\chi _{\mathcal {S}}})_{ {\mathfrak {m}}_{\pi }}$ are isomorphic as modules over the Hecke algebra.

Proof. Let $m$ be a generator of the maximal ideal of the localization. From the short exact sequence

\[ 0 \rightarrow {\mathcal{D}}_{\chi_{\mathcal{S}}} \stackrel{\times m}{\longrightarrow} {\mathcal{D}}_{\chi_{\mathcal{S}}} \rightarrow {\mathcal{D}}_{{{\mathbb 1}}} \longrightarrow 0 \]

we obtain a long exact sequence in cohomology

\begin{align*} & \xrightarrow{\times m}\operatorname{H}^{1}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{S}}})_{{\mathfrak{m}}_{\pi}} \longrightarrow \operatorname{H}^{1}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {E})_{{\mathfrak{m}}_{\pi}} \xrightarrow{0} \operatorname{H}^{2}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{S}}})_{{\mathfrak{m}}_{\pi}} \\ & \xrightarrow{\times m} \operatorname{H}^{2}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{S}}})_{{\mathfrak{m}}_{\pi}} \longrightarrow \operatorname{H}^{2}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {E})_{{\mathfrak{m}}_{\pi}} \longrightarrow 0. \end{align*}

Here the second map is the zero map as the first map is surjective by the second part of Theorem 5.1, and the last term is zero because the system of eigenvalues for $\pi$ does not appear in degrees greater by Theorem 5.1(i).

We then get that multiplication by $m$ is injective on $\operatorname {H}^{2}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}_{\chi _{\mathcal {S}}})_{ {\mathfrak {m}}_{\pi }}$ and, therefore, the map

\[ \operatorname{H}^{2}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {\mathcal{D}}_{\chi_{\mathcal{S}}})_{{\mathfrak{m}}_{\pi}}\otimes_{\mathcal{O}_{\mathcal{S}, {{\mathbb 1}}}}{E}\longrightarrow \operatorname{H}^{2}(X_{K^{{\mathfrak{p}}}\times I_{{\mathfrak{p}}}}, {E})_{{\mathfrak{m}}_{\pi}} \]

is an isomorphism. By hypothesis, $\operatorname {H}^{2}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {E})_{ {\mathfrak {m}}_{\pi }}$ is one-dimensional. Therefore, the $\mathcal {O}_{\mathcal {S}, {{\mathbb 1}}}$-module $\operatorname {H}^{2}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}_{\chi _{\mathcal {S}}})_{ {\mathfrak {m}}_{\pi }}$ is cyclic by Nakayama's Lemma. If it were torsion, then multiplication by $m$ would not be injective on it, which is a contradiction; so it is free.

Let $r$ be a generator of the ideal of $\mathcal {O}_{\mathcal {U},{{\mathbb 1}}}$ corresponding to $\mathcal {S}$. The modules $\operatorname {H}^{1}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}_{\chi _{\mathcal {S}}})_{ {\mathfrak {m}}_{\pi }}$ and $\operatorname {H}^{2}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}_{\chi _{\mathcal {S}}})_{ {\mathfrak {m}}_{\pi }}$ are kernel and cokernel, respectively, of the multiplication by $r$ map on $\operatorname {H}^{2}(X_{K^{ {\mathfrak {p}}}\times I_{ {\mathfrak {p}}}}, {\mathcal {D}}_{\chi _{\mathcal {U}}})_{ {\mathfrak {m}}_\pi }$. Multiplication by the appropriate power of $r$ yields the sought-after isomorphism.

We would like to apply Theorem 3.16 to conclude that $\mathcal {L}$-invariants are independent of the cohomological degree. There are, however, several problems. First, it is not clear whether one can find a curve $\mathcal {S}$ as in the theorem that is smooth at ${{\mathbb 1}}$. Second, if one finds such a curve, the automorphic Colmez–Greenberg–Stevens formula in the inert case only produces one element in the intersection of $\mathcal {L}^{(0)}(f,p)$ and $\mathcal {L}^{(1)}(f,p)$. As these spaces are two-dimensional by Proposition 1.8 this gives us no new information. In the split case the situation is better: one can at least compute the $\mathcal {L}$-invariant of one of the primes lying above $p$ and, if the tangent space of $\mathcal {S}$ at ${{\mathbb 1}}$ is generic enough, one can compute both.

Corollary 5.3 Let $F$ be an imaginary quadratic field unramified at $p$ and let $\pi$ be a cuspidal Bianchi newform of parallel weight $2$. Suppose that $p$ splits in $F$ as $p= {\mathfrak {p}} \overline { {\mathfrak {p}}}$ and $f$ is Steinberg at both, $ {\mathfrak {p}}$ and $\overline { {\mathfrak {p}}}$. Then at least one of the equalities

\[ \mathcal{L}^{(0)}(\pi,{\mathfrak{p}})=\mathcal{L}^{(1)}(\pi,{\mathfrak{p}}) \]

or

\[ \mathcal{L}^{(0)}(\pi,\overline{{\mathfrak{p}}})=\mathcal{L}^{(1)}(\pi,\overline{{\mathfrak{p}}}) \]

holds.

Remark 5.4 It was shown in [Reference GehrmannGeh19a] that the theorem above would also follow from Venkatesh's conjectures on the action of derived Hecke algebras.

The situation simplifies substantially if $\pi$ is the base change of a modular form $f$.

Theorem 5.5 Let $F$ be an imaginary quadratic field unramified at $p$. Let $f$ be a newform of weight $2$ that is Steinberg at $p$ and $\operatorname {BC}(f)$ its base change to $F$. The following hold.

  1. (i) If $p= {\mathfrak {p}} \overline { {\mathfrak {p}}}$ is split in $F$, then

    \[ \mathcal{L}^{(0)}(\operatorname{BC}(f),{\mathfrak{p}})=\mathcal{L}^{(1)}(\operatorname{BC}(f),{\mathfrak{p}})=\mathcal{L}^{(0)}(\operatorname{BC}(f),\overline{{\mathfrak{p}}})=\mathcal{L}^{(1)}(\operatorname{BC}(f),\overline{{\mathfrak{p}}})=\mathcal{L}(f,p). \]
  2. (ii) If $p$ is inert, we have

    \[ \mathcal{L}^{(0)}(\operatorname{BC}(f),p)\cap \operatorname{Hom}^{\operatorname{ct}}({\mathbb{Q}}_p^{\ast},{E})=\mathcal{L}^{(1)}(\operatorname{BC}(f),p) \cap \operatorname{Hom}^{\operatorname{ct}}({\mathbb{Q}}_p^{\ast},{E})= \mathcal{L}(f,p). \]
  3. (iii) If $p$ is inert, we have

    \[ \mathcal{L}^{(0)}(\operatorname{BC}(f),p)=\mathcal{L}^{(1)}(\operatorname{BC}(f),p). \]

Proof. In § 5 of [Reference Barrera Salazar and WilliamsBW19] it is shown that one may take $\mathcal {S}\subseteq {\mathcal {W}}$ to be the parallel weight curve (see § 5 of [Reference Barrera Salazar and WilliamsBW19]), which is clearly smooth at ${{\mathbb 1}}$. Using $p$-adic Langlands functoriality as explained in [Reference Barrera Salazar and WilliamsBW19], the first two claims follow by applying Theorem 3.16 (respectively, its analogue in the Bianchi setting) for the modular form $f$ and its base change to $F$. The last claim is a consequence of the second claim and the Galois invariance of automorphic $\mathcal {L}$-invariants attached to base change representations (see Lemma 3.1 of [Reference GehrmannGeh19b]).

Remark 5.6 In the inert case the equality $\mathcal {L}^{(0)}(\operatorname {BC}(f),p)\cap \operatorname {Hom}^{\operatorname {ct}}( {\mathbb {Q}}_p^{\ast },{E})=\mathcal {L}(f,p)$ was proven in [Reference GehrmannGeh19b, Lemma 3.3], using Artin formalism for $p$-adic $L$-functions and the exceptional zero formula. An analogous result for higher weight forms was proven by Barrera-Salazar and Williams in [Reference Barrera Salazar and WilliamsBW19, Proposition 10.2].

Our approach can be adapted to higher weights making the construction of Stark–Heegner cycles of [Reference Venkat and WilliamsVW21] unconditional in the base change case.

Footnotes

While working on this paper the first named author was visiting McGill University, supported by Deutsche Forschungsgemeinschaft, and he would like to thank these institutions. The second named author was partly funded by FRQNT grant 2019-NC-254031 and NSERC grant RGPIN-2018-04392. In addition, the authors would like to thank Henri Darmon, Michael Lipnowski, Vytautas Pas̆kūnas and Chris Williams for several intense and stimulating discussions. We also thank the referee for their careful reading of the paper and many suggestions.

1 The same result has been obtained by Spieß by different methods; see [Reference SpießSpi20].

References

Ash, A. and Stevens, G., $p$-adic deformations of arithmetic cohomology, Preprint (2008), https://math.bu.edu/people/ghs/preprints/Ash-Stevens-02-08.pdf.Google Scholar
Barnet-Lamb, T., Gee, T., Geraghty, D. and Taylor, R., Potential automorphy and change of weight, Ann. of Math. (2) 179 (2014), 501609.CrossRefGoogle Scholar
Barrera Salazar, D., Dimitrov, M. and Jorza, A., $p$-adic $L$-functions of Hilbert cusp forms and the trivial zero conjecture, J. Eur. Math. Soc. (2021), doi:10.4171/JEMS/1165.Google Scholar
Barrera Salazar, D. and Williams, C., Exceptional zeros and $\mathcal {L}$-invariants of Bianchi modular forms, Trans. Amer. Math. Soc. 372 (2019), 134.CrossRefGoogle Scholar
Barrera Salazar, D. and Williams, C., Families of Bianchi modular symbols: critical base-change $p$-adic $L$-functions and $p$-adic Artin formalism, Selecta Math. (N.S.) 27 (2021), 82.CrossRefGoogle Scholar
Benois, D., $p$-adic heights and $p$-adic Hodge theory, Mém. Soc. Math. Fr. (N.S.) 167 (2021).Google Scholar
Bertolini, M., Darmon, H. and Iovita, A., Families of automorphic forms on definite quaternion algebras and Teitelbaum's conjecture, Astérisque 331 (2010), 2964.Google Scholar
Borel, A., Admissible representations of a semi-simple group over a local field with vector fixed under an Iwahori subgroup, Invent. Math. 35 (1976), 233260.CrossRefGoogle Scholar
Borel, A. and Serre, J.-P., Cohomologie d'immeubles et de groupes S-arithmétiques, Topology 15 (1976), 211232.CrossRefGoogle Scholar
Caraiani, A., Monodromy and local-global compatibility for $l=p$, Algebra Number Theory 8 (2014), 15971646.CrossRefGoogle Scholar
Cartier, P., Representations of p-adic groups: a survey, in Automorphic forms, representations and L-functions (Proc. Sympos. Pure Math., Oregon State Univ., Corvallis, Ore., 1977), Part 1, Proceedings of Symposia in Pure Mathematics, vol. XXXIII (American Mathematical Society, Providence, RI, 1979), 111155.Google Scholar
Chenevier, G., The p-adic analytic space of pseudocharacters of a profinite group and pseudorepresentations over arbitrary rings, in Automorphic forms and Galois representations. Vol. 1, London Mathematical Society Lecture Note Series, vol. 414 (Cambridge University Press, Cambridge, 2014), 221285.CrossRefGoogle Scholar
Darmon, H., Integration on $H_{ p}\times H$ and arithmetic applications, Ann. of Math. (2) 154 (2001), 589639.CrossRefGoogle Scholar
Dasgupta, S., Stark–Heegner points on modular jacobians, Ann. Sci. Éc. Norm. Supér (4) 38 (2005), 427469.CrossRefGoogle Scholar
Dasgupta, S. and Greenberg, M., L-invariants and Shimura curves, Algebra Number Theory 6 (2012), 455485.CrossRefGoogle Scholar
Ding, Y., Simple $\mathcal {L}$-invariants for $\mathrm {GL}_n$, Trans. Amer. Math. Soc. 372 (2019), 79938042.CrossRefGoogle Scholar
Gehrmann, L., Derived Hecke algebra and automorphic L-invariants, Trans. Amer. Math. Soc. 372 (2019), 77677784.CrossRefGoogle Scholar
Gehrmann, L., Functoriality of automorphic $L$-invariants and applications, Doc. Math. 24 (2019), 12251243.Google Scholar
Gehrmann, L., Automorphic $\mathcal {L}$-invariants for reductive groups, J. Reine Angew. Math. 779 (2021), 57103.CrossRefGoogle Scholar
Geraghty, D., Modularity lifting theorems for ordinary Galois representations, Math. Ann. 373 (2019), 13411427.CrossRefGoogle Scholar
Goldring, W., Galois representations associated to holomorphic limits of discrete series, Compos. Math. 150 (2014), 191228.CrossRefGoogle Scholar
Greenberg, M., Stark–Heegner points and the cohomology of quaternionic Shimura varieties, Duke Math. J. 147 (2009), 541575.CrossRefGoogle Scholar
Guitart, X., Masdeu, M. and Sengün, M. H., Darmon points on elliptic curves over number fields of arbitrary signature, Proc. Lond. Math. Soc. (3) 111 (2015), 484518.CrossRefGoogle Scholar
Hansen, D., Universal eigenvarieties, trianguline Galois representations, and $p$-adic Langlands functoriality, J. Reine Angew. Math. 730 (2017), 164, with an appendix by James Newton.CrossRefGoogle Scholar
Howe, R. and Piatetski-Shapiro, I. I., Some examples of automorphic forms on ${\rm Sp}_{4}$, Duke Math. J. 50 (1983), 55106.CrossRefGoogle Scholar
Jacquet, H. and Shalika, J. A., On Euler products and the classification of automorphic forms. II, Amer. J. Math. 103 (1981), 777815.CrossRefGoogle Scholar
Jacquet, H. and Shalika, J. A., On Euler products and the classification of automorphic representations. I, Amer. J. Math. 103 (1981), 499558.CrossRefGoogle Scholar
Januszewski, F., Rational structures on automorphic representations, Math. Ann. 370 (2018), 18051881.CrossRefGoogle Scholar
Knapp, A. W., Representation theory of semisimple groups, Princeton Landmarks in Mathematics (Princeton University Press, Princeton, NJ, 2001). An overview based on examples, reprint of the 1986 original.Google Scholar
Kohlhaase, J., The cohomology of locally analytic representations, J. Reine Angew. Math. 651 (2011), 187240.Google Scholar
Kohlhaase, J. and Schraen, B., Homological vanishing theorems for locally analytic representations, Math. Ann. 353 (2012), 219258.CrossRefGoogle Scholar
Longo, M., Rotger, V. and Vigni, S., On rigid analytic uniformization of Jacobians of Shimura curves, Amer. J. Math. 134 (2012), 11971246.CrossRefGoogle Scholar
Ollivier, R., Resolutions for principal series representations of $p$-adic ${\rm GL}_n$, Münster J. Math. 7 (2014), 225240.Google Scholar
Orlik, S., On extensions of generalized Steinberg representations, J. Algebra 293 (2005), 611630.CrossRefGoogle Scholar
Orton, L., An elementary proof of a weak exceptional zero conjecture, Can. J. Math. 56 (2004), 373405.CrossRefGoogle Scholar
Piatetski-Shapiro, I. I., Multiplicity one theorems, in Automorphic forms, representations and L-functions (Proc. Sympos. Pure Math., Oregon State Univ., Corvallis, Ore., 1977), Part 1, Proceedings of Symposia in Pure Mathematics, vol. XXXIII (American Mathematical Society, Providence, RI, 1979), 209212.Google Scholar
Reeder, M., On certain Iwahori invariants in the unramified principal series, Pacific J. Math. 153 (1992), 313342.CrossRefGoogle Scholar
Rosso, G., $\mathcal {L}$-invariant for Siegel-Hilbert forms, Doc. Math. 20 (2015), 12271253.Google Scholar
Saito, T., Hilbert modular forms and $p$-adic Hodge theory, Compos. Math. 145 (2009), 10811113.CrossRefGoogle Scholar
Scholze, P., On torsion in the cohomology of locally symmetric varieties, Ann. of Math. (2) 182 (2015), 9451066.10.4007/annals.2015.182.3.3CrossRefGoogle Scholar
Seveso, M. A., The Teitelbaum conjecture in the indefinite setting, Amer. J. Math. 135 (2013), 15251557.CrossRefGoogle Scholar
Soudry, D., A uniqueness theorem for representations of ${\rm GSO}(6)$ and the strong multiplicity one theorem for generic representations of ${\rm GSp}(4)$, Israel J. Math. 58 (1987), 257287.CrossRefGoogle Scholar
Spieß, M., On special zeros of $p$-adic $L$-functions of Hilbert modular forms, Invent. Math. 196 (2014), 69138.CrossRefGoogle Scholar
Spieß, M., On $\mathcal {L}$-invariants associated to Hilbert modular forms, Preprint (2020), arXiv:2005.11892.Google Scholar
Taylor, R., On Galois representations associated to Hilbert modular forms, Invent. Math. 98 (1989), 265280.CrossRefGoogle Scholar
Teitelbaum, J. T., Values of $p$-adic $L$-functions and a $p$-adic Poisson kernel, Invent. Math. 101 (1990), 395410.CrossRefGoogle Scholar
Thorne, J. A., Automorphy lifting for residually reducible $l$-adic Galois representations, J. Amer. Math. Soc. 28 (2015), 785870.CrossRefGoogle Scholar
Urban, E., Eigenvarieties for reductive groups, Ann. of Math. (2) 174 (2011), 16851784.CrossRefGoogle Scholar
Venkat, G. and Williams, C., Stark–Heegner cycles attached to Bianchi modular forms, J. Lond. Math. Soc. 104 (2021), 394422.CrossRefGoogle Scholar
Vignéras, M.-F., A criterion for integral structures and coefficient systems on the tree of ${\rm PGL}(2, F)$, Pure Appl. Math. Q. 4 (2008), 12911316.CrossRefGoogle Scholar
Wiles, A., On ordinary $\lambda$-adic representations associated to modular forms, Invent. Math. 94 (1988), 529573.CrossRefGoogle Scholar