1 Introduction
Let f be a cuspidal normalized newform of weight $2k$ and level $\Gamma _0(pN)$ , where p is a prime number and $N\in {\mathbf {Z}}_{>0}$ such that $p\nmid N$ . Consider the complex L-function attached to f
One way to study $L(f, s)$ is via p-adic method. That is, one can associate f with a p-adic L-function $L_p(f, s)$ , which p-adically interpolates the algebraic part of the special values $L(f, j)$ for $1\leq j\leq 2k-1$ . In particular, the interpolation property at $s= k$ is given by the formula
where $\Omega _f$ is the Deligne period of f at k ([Reference Deligne, Koblitz and Ogus6]).
Suppose moreover that $a_p = p^{k-1}$ ; the formula above shows that $L_p(f, s)$ vanishes at $s=k$ . In the case when $k=1$ , Mazur–Tate–Teitelbaum conjectured in [Reference Mazur, Tate and Teitelbaum15] that there exists an invariant ${\mathcal {L}}(f)$ such that
This conjecture is known as the trivial zero conjecture and has been proven by Greenberg–Stevens in [Reference Greenberg and Stevens9]. Moreover, for higher weights, various generalizations of the invariant ${\mathcal {L}}(f)$ has been proposed. The following is an incomplete list:
-
• In [Reference Greenberg8], R. Greenberg constructed the ${\mathcal {L}}$ -invariants for Galois representations that are ordinary at p and suggested a generalization of the trivial zero conjecture.
-
• In [Reference Mazur14], Fontaine–Mazur defined the ${\mathcal {L}}$ -invariant by studying the semistable module (à la Fontaine) associated with a p-adic representation.
-
• In [Reference Coleman5], R. Coleman proposed a construction of ${\mathcal {L}}$ -invariants as an application of his p-adic integration theory.
-
• In [Reference Teitelbaum18], J. Teitelbaum proposed a construction of ${\mathcal {L}}$ -invariants by applying the p-adic integration theory to p-adically uniformized Shimura curve.
All these ${\mathcal {L}}$ -invariants are known to be equal: Coleman–Iovita compared the second and the third in [Reference Coleman and iovita4]; Iovita–Spieß compared the second and the fourth in [Reference Iovita and Spieß13]; and the comparison between the first and the second is a special case of [Reference Benois2, Proposition 2.3.7].
It is a natural question to ask whether one can establish a similar philosophy for higher rank automorphic forms. Let us mention the following generalizations in our consideration:
-
• In [Reference Benois2], D. Benois generalized Greeberg’s construction to Galois representations of $\operatorname {\mathrm {Gal}}_{{\mathbf {Q}}}$ that satisfies some reasonable conditions. He also stated a trivial zero conjecture in such a generality ([op. cit., pp. 1579]).
-
• In [Reference Besser and de Shalit1], Besser–de Shalit generalized both the Fontaine–Mazur ${\mathcal {L}}$ -invariants and Coleman (or Teitelbaum) ${\mathcal {L}}$ -invariants by studying the p-adic cohomology groups of p-adically uniformized Shimura varieties. It is conjectured in loc. cit. that these two constructions give rise to the same ${\mathcal {L}}$ -invariants (or ${\mathcal {L}}$ -operators as called in loc. cit.). Authors of loc. cit. also speculated that the existence of a trivial zero conjecture for these two ${\mathcal {L}}$ -invariants. However, they were not able to provide an explicit statement.
This article concerns the comparison between Benois’s ${\mathcal {L}}$ -invariants and the Fontaine–Mazur type ${\mathcal {L}}$ -invariants of Besser–de Shalit. To explain our result, let us fix some notations: Let F be a number field such that for every prime ideal ${\mathfrak {p}}\subset {\mathcal {O}}_F$ sitting above p, the maximal unramified extension of ${\mathbf {Q}}_p$ in $F_{{\mathfrak {p}}}$ is ${\mathbf {Q}}_p$ itself; let E be a large enough value field that is a finite extension over ${\mathbf {Q}}_p$ . Suppose
is a Galois representation that is semistable at places above p. We further assume that $\rho $ satisfies the assumptions in §5.1. In particular, we assume the Frobenius eigenvalues on the associated semistable modules are given by $p^m, \ldots, p^{m-n+1}$ (for some suitable $m\in {\mathbf {Z}}$ that is independent of the prime ideals sitting above p) and the monodromy is maximal. We remark in the beginning that these assumptions are required so that we can perform the following two constructions:
-
• Following the suggestion in [Reference Rosso17] (see also [Reference Hida10]), one can consider the induction $\operatorname {\mathrm {Ind}}^{{\mathbf {Q}}}_F \rho $ . Part of the assumptions then allows us to attach the ${\mathcal {L}}$ -invariant in Benois’s style to $\operatorname {\mathrm {Ind}}_{F}^{{\mathbf {Q}}}\rho (m)$ . This resulting ${\mathcal {L}}$ -invariant is denoted by ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}(\rho (m))$ , where the subscript $\operatorname {\mathrm {GB}}$ stands for “Greenberg–Benois”. We refer the readers to §3 for the construction of ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}$ .
-
• We realized that the generalization of Fontaine–Mazur ${\mathcal {L}}$ -invariants suggested by Besser–de Shalit can be translated to the world of semistable modules of a local Galois representation. The other part of the assumptions in §5.1 then allow us to attach ${\mathcal {L}}$ -invariants of Fontaine–Mazur style to each local Galois representation $\rho _{{\mathfrak {p}}} = \rho |_{\operatorname {\mathrm {Gal}}_{F_{{\mathfrak {p}}}}}$ for every prime ${\mathfrak {p}}$ above p. We term such ${\mathcal {L}}$ -invariants ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}(\rho _{{\mathfrak {p}}})$ , where the subscript $\operatorname {\mathrm {FM}}$ stands for “Fontaine–Mazur”. We refer the readers to §4 for the construction of ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}$ .
Our main result reads as follows.
Theorem 5.4 We have an equality
where the index set runs through all prime ideals in ${\mathcal {O}}_F$ sitting above p
Since there is a well-stated trivial zero conjecture for ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}(\rho (m))$ in [Reference Benois2], our result immediately supplies an affirmative answer to Besser–de Shalit’s speculation of the relationship between their ${\mathcal {L}}$ -invariants and p-adic L-functions.
To close this introduction, let us mention that the generalization of ${\mathcal {L}}$ -invariants à la Coleman (or Teitelbaum) suggested by Besser–de Shalit replaces Coleman’s integration theory with Besser’s theory of finite polynomial cohomology. Although they only consider the case for the trivial coefficient (so that we can only see automorphic forms of weight associated with the differential 1-forms), one can hope a generalization for nontrivial coefficients by using finite polynomial cohomology with coefficients ([Reference Huang and J.-F.12]). We wish to come back to this in future projects and hopefully to compare this type of ${\mathcal {L}}$ -invariants with ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}(\rho _{{\mathfrak {p}}})$ as suggested in [Reference Besser and de Shalit1].
Notations
-
• Through out this article, we fix a prime number p.
-
• Given a field F, we fix a separable closure $\overline {F}$ and denote by $\operatorname {\mathrm {Gal}}_F = \operatorname {\mathrm {Gal}}(\overline {F}/F)$ its absolute Galois group.
2 Preliminaries on $(\varphi , \Gamma )$ -modules
2.1 General $(\varphi , \Gamma )$ -modules
Fix a compatible system of primitive p-power roots of unity $(\zeta _{p^n})_{n\in {\mathbf {Z}}_{\geq 0}}$ in $\overline {{\mathbf {Q}}}_p$ . Given a finite extension K of ${\mathbf {Q}}_p$ , consider $K(\zeta _{p^{\infty }}) = \bigcup _{n\in {\mathbf {Z}}_{\geq 0}} K(\zeta _{p^n})$ and denote by $\Gamma = \Gamma _K$ the Galois group $\operatorname {\mathrm {Gal}}(K(\zeta _{p^{\infty }})/K)$ . Moreover, for any $r\in [0, 1)$ , let
and
where $K^{\mathrm {unr}}$ is the maximal unramified extension of K in $\overline {{\mathbf {Q}}}_p$ and the infinite union is taken with respect to the inclusions ${\mathcal {R}}_K^{r} \hookrightarrow {\mathcal {R}}_K^{r'}$ for $r\leq r'<1$ . We call the ring ${\mathcal {R}}_K$ the Robba ring over K. It carries a $\varphi $ -action and a $\Gamma $ -action via the formula
where $\chi _{\operatorname {\mathrm {cyc}}}$ is the cyclotomic character.
In what follows, we shall consider a more generalized version of ${\mathcal {R}}_K$ . Let E be a finite extension of ${\mathbf {Q}}_p$ . We denote by ${\mathcal {R}}_{K, E} := {\mathcal {R}}_K \otimes _{{\mathbf {Q}}_p}E$ and call it the Robba ring over K with coefficients in E. We linearize the actions of $\varphi $ and $\Gamma $ on ${\mathcal {R}}_{K, E}$ via $\varphi \otimes \operatorname {\mathrm {id}}$ and $\gamma \otimes \operatorname {\mathrm {id}}$ , respectively. In what follows, we often assume E is large enough so that $K \subset E$ .
By a $(\varphi , \Gamma )$ -module over ${\mathcal {R}}_{K, E}$ , we mean a finite free ${\mathcal {R}}_{K, E}$ -module D together with a $\varphi $ -semilinear endomorphism $\varphi _D$ and a semilinear action by $\Gamma $ , which commute with each other, such that the induced map
is an isomorphism. We shall denote by ${\mathbf{\mathsf{Mod}}}_{{\mathcal {R}}_{K, E}}^{(\varphi , \Gamma )}$ the category of $(\varphi , \Gamma )$ -modules over ${\mathcal {R}}_{K, E}$ .
Let ${\mathbf{\mathsf{Rep}}}_{K}(E)$ the category of Galois representations of $\operatorname {\mathrm {Gal}}_K$ with coefficients in E. Then, by [Reference Benois2, Proposition 1.1.4], there is a fully faithful functor
Moreover, by letting ${\mathbf{\mathsf{Mod}}}_{K, E}^{(\varphi , N)}$ (resp., ${\mathbf{\mathsf{Mod}}}_{K, E}^{\varphi }$ ) the category of $(\varphi , N)$ -modules (resp., $\varphi $ -modules) over $K_0 = K \cap {\mathbf {Q}}_p^{\mathrm {unr}}$ with coefficients in E, there is a functor (see, for example, [Reference Benois2, §1.2.3])
such that if $\rho \in {\mathbf{\mathsf{Rep}}}_{K}(E)$ is semistable (resp., crystalline), then ([Reference Berger3, Théorème 0.2])
Here, ${\mathbf {D}}_{{\mathrm {st}}}$ (resp., ${\mathbf {D}}_{\operatorname {\mathrm {cris}}}$ ) is Fontaine’s semistable (resp., crystalline) functor ([Reference Fontaine7, Reference Berger3]), assigning a Galois representation in ${\mathbf{\mathsf{Rep}}}_{K}(E)$ a $(\varphi , N)$ -module (resp., $\varphi $ -module) over $K_0$ with coefficients in E.
Now, let D be a $(\varphi , \Gamma )$ -module over ${\mathcal {R}}_{K, E}$ . Recall the cohomology of D is defined by the cohomology of the Herr complex
where $\gamma $ is a (fixed) topological generator of $\Gamma $ . Note that, given $\alpha = (x, y)\in D \oplus D$ such that $(\gamma -1)x -(\varphi _D -1)y = 0$ , there is an extension
defined by
It turns out that such an assignment gives rise to an isomorphism
Furthermore, we write $H^1_{{\mathrm {st}}}(D)$ (resp., $H^1_f(D)$ ) the subspace of $H^1(D)$ , consisting of those semistable (resp., crystalline) extensions $D_{\alpha }$ (i.e., those satisfy $\operatorname {\mathrm {rank}}_{K_0\otimes _{{\mathbf {Q}}_p}E}{\mathcal {D}}_{{\mathrm {st}}}(D_{\alpha }) = \operatorname {\mathrm {rank}}_{K_0 \otimes _{{\mathbf {Q}}_p} E} {\mathcal {D}}_{{\mathrm {st}}}(D) +1$ (resp., $\operatorname {\mathrm {rank}}_{K_0\otimes _{{\mathbf {Q}}_p}E}{\mathcal {D}}_{\operatorname {\mathrm {cris}}}(D_{\alpha }) = \operatorname {\mathrm {rank}}_{K_0 \otimes _{{\mathbf {Q}}_p} E} {\mathcal {D}}_{\operatorname {\mathrm {cris}}}(D) +1$ )). According to [Reference Benois2, Proposition 1.4.2], if $\rho \in {\mathbf{\mathsf{Rep}}}_{K}(E)$ , then
where
Footnote 1 is the usual local Bloch–Kato Selmer group.
To conclude our discussion for general $(\varphi , \Gamma )$ -modules, we mention that if D is semistable,Footnote 2 then $H_{{\mathrm {st}}}^1(D)$ and $H_f^1(D)$ can be computed by complexes $C_{{\mathrm {st}}}^{\bullet }$ and $C_{\operatorname {\mathrm {cris}}}^{\bullet }$ , respectively ([Reference Benois2, Proposition 1.4.4]). Here,
and
2.2 $(\varphi , \Gamma )$ -modules of rank 1
Recall that $(\varphi , \Gamma )$ -modules of rank $1$ can be understood via continuous characters. More precisely, given a continuous character $\delta : K^\times \rightarrow E^\times $ and fixing a uniformizer $\varpi \in K$ , we can write $\delta = \delta ' \delta "$ with $\delta '|_{{\mathcal {O}}_{K}^{\times }} = \delta |_{{\mathcal {O}}_{K}^{\times }}$ , $\delta '(\varpi ) = 1$ and $\delta "(\varpi ) = \delta (\varpi )$ , $\delta "|_{{\mathcal {O}}_K^{\times }} = 1$ . By local class field theory, $\delta '$ defines a unique one-dimensional Galois representation $\chi _{\delta '}$ ; that is,
Footnote 3 which admits its associated $(\varphi , \Gamma )$ -module ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\chi _{\delta '})$ . However, we define ${\mathcal {R}}_{K, E}(\delta ") = {\mathcal {R}}_{K, E}e_{\delta "}$ such that $\varphi (e_{\delta "}) = \delta (\varpi )e_{\delta "}$ and $\gamma (e_{\delta "}) = e_{\delta "}$ . Then, the $(\varphi , \Gamma )$ -module associated with $\delta $ is defined to be
In particular, the cyclotomic character $\operatorname {\mathrm {Gal}}_{K} \rightarrow {\mathcal {O}}_{E}^\times $ has the associated $(\varphi , \Gamma )$ -module ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\chi _{\operatorname {\mathrm {cyc}}})$ . By [Reference Nakamura16, Lemma 2.13], we know that
where $\mathrm {Nm}^K_{{\mathbf {Q}}_p}$ is the norm function from K to ${\mathbf {Q}}_p$ .
Lemma 2.1. Let $\delta : K^\times \rightarrow E^\times $ be the character
such that all $m_{\sigma } \geq 1$ .
-
• If $\left ({\mathcal {D}}_{{\mathrm {st}}}({\mathcal {R}}_{K, E}(\delta ))^{\vee }(\chi _{\operatorname {\mathrm {cyc}}}) \right )^{\varphi =1}$ is nonzero, then the inclusion $H_{{\mathrm {st}}}^1({\mathcal {R}}_{K, E}(\delta )) \hookrightarrow H^1({\mathcal {R}}_{K, E}(\delta ))$ is an isomorphism.
-
• If $\left ({\mathcal {D}}_{{\mathrm {st}}}({\mathcal {R}}_{K, E}(\delta ))^{\vee }(\chi _{\operatorname {\mathrm {cyc}}}) \right )^{\varphi =1} =0 $ , then the inclusion $H_f^1({\mathcal {R}}_{K, E}(\delta )) \kern1.4pt{\hookrightarrow}\kern1.4pt H_{{\mathrm {st}}}^1({\mathcal {R}}_{K, E}(\delta )\kern-0.7pt)$ is an isomorphism.
Proof By [Reference Benois2, Corollary 1.4.5], we have the formula
Applying [Reference Rosso17, Proposition 2.1 & Lemma 2.3], we know that
The lemma then follows easily.
Suppose $\delta : K^\times \rightarrow E^\times $ is a continuous character as in Lemma 2.1. Suppose ${\mathcal {R}}_{K, E}(\delta )$ is semistable and so $\operatorname {\mathrm {rank}}_{K_0 \otimes _{{\mathbf {Q}}_p}E}{\mathcal {D}}_{{\mathrm {st}}}({\mathcal {R}}_{K, E}(\delta )) = 1$ . We fix a $K_0 \otimes _{{\mathbf {Q}}_p}E$ -basis $v_{\delta }$ for ${\mathcal {D}}_{{\mathrm {st}}}({\mathcal {R}}_{K, E}(\delta ))$ and define
Lemma 2.2. Suppose $\eta : K^\times \rightarrow E^\times $ is a continuous character of the form $\eta (z) = \prod _{\sigma : K \hookrightarrow \overline {{\mathbf {Q}}}_p} \sigma (z)^{n_{\sigma }}$ with all $n_{\sigma } \leq 0$ . Suppose
is a semistable extension (in the sense of §2.1). Then,
Moreover, there exists a unique ${\mathcal {L}}(D)\in E$ such that
Proof First of all, it follows from [Reference Benois2, Proposition 1.2.7] that ${\mathcal {R}}_{K, E}(\eta )$ is also semistable. Hence, by applying [op. cit., Proposition 1.4.4], we know that
Taking the cohomology of the short exact sequence in the lemma, we have a commutative diagram
which shows the first claim.
Since ${\mathcal {R}}_{K, E}(\eta )$ is semistable and it is of rank $1$ over ${\mathcal {R}}_{K,E}$ , it is crystalline and ${\mathcal {D}}_{{\mathrm {st}}}({\mathcal {R}}_{K, E}(\eta )) = {\mathcal {D}}_{\operatorname {\mathrm {cris}}}({\mathcal {R}}_{K, E}(\eta ))$ . This is because the monodromy operator is nilpotent. We consider the commutative diagram
induced by the short exact sequence in the lemma, where the rows are exact and the columns are the semistable complexes. Let $v_{\eta }$ be the element in ${\mathcal {D}}_{{\mathrm {st}}}({\mathcal {R}}_{K, E}(\eta ))$ that gives rise to the basis of $H^0({\mathcal {R}}_{K, E}(\eta ))$ as in [Reference Rosso17, Proposition 2.1]. In particular, $v_{\eta } \in \operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^0 {\mathcal {D}}_{{\mathrm {st}}}({\mathcal {R}}_{K, E}(\eta ))$ and $\varphi (v_{\eta }) = v_{\eta }$ . Using the relation $N \varphi = p \varphi N$ , one deduces that $1$ and $p^{-1}$ are Frobenius eigenvalues of ${\mathcal {D}}_{{\mathrm {st}}}(D)$ . We choose a lift $\widetilde {v}_{\eta }\in {\mathcal {D}}_{{\mathrm {st}}}(D)$ such that $\varphi (\widetilde {v}_{\eta }) = \widetilde {v}_{\eta }$ . This then implies that $N(\widetilde {v}_{\eta })$ has Frobenius eigenvalue $p^{-1}$ . The commutativity of the diagram then yields
Applying the exactness of the middle row, we see that
Since $N(\widetilde {v}_{\eta })$ is a basis for the Frobenius eigensubspace of ${\mathcal {D}}_{{\mathrm {st}}}(D)$ on which $\varphi $ acts via $p^{-1}$ , we see that b is invertible. We then conclude that
Therefore, ${\mathcal {L}}(D) := a/b$ .
3 Greenberg–Benois ${\mathcal {L}}$ -invariants
In this section, we first discuss the construction of Greenberg–Benois ${\mathcal {L}}$ -invariant over ${\mathbf {Q}}$ in §3.1 by summarizing Benois’s construction in [Reference Benois2, §2] (in particular [op. cit., (26)]). Then we follow the strategy in [Reference Rosso17], generalizing Benois’s construction to general number fields by considering inductions of Galois representations (§3.2).
3.1 Greenberg–Benois ${\mathcal {L}}$ -invariants over ${\mathbf {Q}}$
To define Greenberg–Benois ${\mathcal {L}}$ -invariants over ${\mathbf {Q}}$ , we start with a Galois representation
which is unramified outside a finite set of places. We denote by
and let ${\mathbf {Q}}_{S}$ be the maximal extension of ${\mathbf {Q}}$ that is unramified outside S.
Recall the Bloch–Kato Selmer group associated with $\rho $ : Given $v\in S$ , define the local Selmer groups
where $I_{\ell }$ stands for the inertia group at $\ell $ . Then, the Bloch-Kato Selmer group associated with $\rho $ is defined to be
Let $\rho _p := \rho |_{\operatorname {\mathrm {Gal}}_{{\mathbf {Q}}_p}}$ . We follow [Reference Benois2, §2.1.2, 2.1.4] and proceed with the following conditions:
-
(B1) The local representation $\rho _p$ is semistable with Hodge–Tate weights $k_1 \leq k_2 \leq \cdots \leq k_n$ , giving rise to the de Rham filtration $\operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^{\bullet } {\mathbf {D}}_{{\mathrm {st}}}(\rho )$ .
-
(B2) The Frobenius action on ${\mathbf {D}}_{{\mathrm {st}}}(\rho _p)$ is semisimple at $1$ and $p^{-1}$ .
-
(GB1) $H_f^1({\mathbf {Q}}, \rho ) = 0 = H_f^1({\mathbf {Q}}, \rho ^{\vee }(1))$ .Footnote 5
-
(GB2) $H^0(\operatorname {\mathrm {Gal}}_{{\mathbf {Q}}_S}, \rho ) = 0 = H^0(\operatorname {\mathrm {Gal}}_{{\mathbf {Q}}_S}, \rho ^{\vee }(1))$ .
-
(GB3) The associated $(\varphi , \Gamma )$ -module ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _p)$ has no saturated subquotientFootnote 6 isomorphic to $U_{k,m}$ with $k\geq 1$ and $m\geq 0$ ([Reference Benois2, §2.1.2]), where $U_{k,m}$ is the unique crystalline $(\varphi , \Gamma )$ -module sitting in a nonsplit short exact sequence
$$\begin{align*}0 \rightarrow {\mathcal{R}}_{{\mathbf{Q}}_p, E}(|z|z^k) \rightarrow U_{k,m} \rightarrow {\mathcal{R}}_{{\mathbf{Q}}_p, E}(z^{-m}) \rightarrow 0. \end{align*}$$
Given a regular submodule $D\subset {\mathbf {D}}_{{\mathrm {st}}}(\rho _p)$ (i.e., a $(\varphi , N)$ -submodule such that ${\mathbf {D}}_{{\mathrm {st}}}(\rho _p) = D \oplus \operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^0{\mathbf {D}}_{{\mathrm {st}}}(\rho _p)$ ), Benois defines a five-step filtration
Such a filtration then yields a filtration on ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _p)$ by
where $t = \log (1+T) \in {\mathcal {R}}_{{\mathbf {Q}}_p, E}$ .
Using this filtration, we define the exceptional subquotient
By [Reference Benois2, Proposition 2.1.7], we have
where M, $M_0$ , and $M_1$ sit inside a short exact sequence
Moreover, one has
and $\dim _E H^1(W)/H_f^1(W) = e_D = \operatorname {\mathrm {rank}} M_0 + \operatorname {\mathrm {rank}} W_0 +\operatorname {\mathrm {rank}} W_1$ ([Reference Benois2, §2.2.1]).
Under the assumption (GB1) and (GB2), one applies Poitou–Tate exact sequence and deduces an isomorphism
Note that the latter space contains an $e_D$ -dimensional subspace $\frac {H^1(W)}{H_f^1(W)} \cong \frac {H^1(\operatorname {\mathrm {Fil}}_1^{\operatorname {\mathrm {GB}}}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _p))}{H^1_f({\mathbf {Q}}_p, \rho )}$ . We then define $H^1(D, \rho )$ to be the image of $\frac {H^1(W)}{H_f^1(W)}$ in $H^1(\operatorname {\mathrm {Gal}}_{{\mathbf {Q}}_S}, \rho )$ .
To define the ${\mathcal {L}}$ -invariant, we further assume that
-
(GB4) $W_0 = 0$ and the Hodge–Tate weights for $\operatorname {\mathrm {Gr}}_1^{\operatorname {\mathrm {GB}}} {\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _p)$ are positive (see [Reference Benois2, Proposition 1.5.9]).
Benois shows that there is a decomposition ([Reference Benois2, §2.1.9] (see also the discussion in [Reference Harron and Jorza11, §1.2]))
and isomorphisms
There are natural morphisms $\varrho _{D, ?}: H^1(D, \rho ) \rightarrow {\mathcal {D}}_{\operatorname {\mathrm {cris}}}(\operatorname {\mathrm {Gr}}_1^{\operatorname {\mathrm {GB}}}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _p))$ (for $? \in \{f, c\}$ ) making the diagram
commutative. Under the assumption of (GB4), Benois shows that $\varrho _{D, c}$ is an isomorphism, and so one can define the Greenberg–Benois ${\mathcal {L}}$ - invariant attached to $\rho $ (with respect to D) as
3.2 Greenberg–Benois ${\mathcal {L}}$ -invariants over general number fields
To define the Greenberg–Benois ${\mathcal {L}}$ -invariants over general number fields, we follow the idea in [Reference Rosso17] (see also [Reference Hida10]) and consider the induction of a Galois representation. More precisely, let F be a number field and suppose we are given a Galois representation
where E is (again) a finite extension of ${\mathbf {Q}}_p$ . We shall consider the induction $\operatorname {\mathrm {Ind}}^{{\mathbf {Q}}}_F \rho $ and define S similarly as before.
Assume the following conditions hold for $\rho $ :
-
(B1) For each place ${\mathfrak {p}} |p$ in F, $\rho _{{\mathfrak {p}}} = \rho |_{\operatorname {\mathrm {Gal}}_{F_{{\mathfrak {p}}}}}$ is semistable with Hodge–Tate weights $k_{{\mathfrak {p}}, \sigma , 1} \leq k_{{\mathfrak {p}}, \sigma , 2}\leq \cdots \leq k_{{\mathfrak {p}}, \sigma , n}$ , where $\sigma : F_{{\mathfrak {p}}} \hookrightarrow \overline {{\mathbf {Q}}}_p$ .
-
(B2) For each place ${\mathfrak {p}} |p$ in F, the Frobenius action on ${\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}})$ is semistable at $1$ and $p^{-1}$ .
-
(GB1) $H_f^1({\mathbf {Q}}, \operatorname {\mathrm {Ind}}^{{\mathbf {Q}}}_{F} \rho ) = 0 = H_f^1({\mathbf {Q}}, \operatorname {\mathrm {Ind}}_F^{{\mathbf {Q}}} \rho ^{\vee }(1))$ .
-
(GB2) $H^0(\operatorname {\mathrm {Gal}}_{{\mathbf {Q}}_S}, \rho ) = 0 = H^0(\operatorname {\mathrm {Gal}}_{{\mathbf {Q}}_S}, \rho ^{\vee }(1))$ .
-
(GB3) The associated $(\varphi , \Gamma )$ -module ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }((\operatorname {\mathrm {Ind}}_F^{{\mathbf {Q}}}\rho )_p) = \bigoplus _{{\mathfrak {p}}|p} {\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _{{\mathfrak {p}}})$ has no saturated subquotient isomorphic to $U_{k,m}$ with $k\geq 1$ and $m\geq 0$ ([Reference Benois2, §2.1.2]).
For every ${\mathfrak {p}}|p$ , choose a regular subomdule $D_{{\mathfrak {p}}} \subset {\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}})$ . Then, $D := \bigoplus _{{\mathfrak {p}} |p} D_{{\mathfrak {p}} } \subset \bigoplus _{{\mathfrak {p}} | p}{\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}) = {\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }((\operatorname {\mathrm {Ind}}^{{\mathbf {Q}}}_{F}\rho )_p)$ is a regular submodule.Footnote 7 Moreover, if $W_0$ , $M_0$ , $M_1$ (resp., $W_{{\mathfrak {p}}, 0}$ , $M_{{\mathfrak {p}}, 0}$ , $M_{{\mathfrak {p}}, 1}$ ) are the corresponding subquotients of ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }((\operatorname {\mathrm {Ind}}^{{\mathbf {Q}}}_{F}\rho )_p)$ (resp., ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _{{\mathfrak {p}}})$ ) with respect to D (resp., $D_{{\mathfrak {p}}}$ ), then we have decompositions
Hence, by assuming
-
(GB4) $W_{{\mathfrak {p}}, 0} =0$ for every ${\mathfrak {p}} |p$ and the Hodge–Tate weights for $\operatorname {\mathrm {Gr}}_1^{\operatorname {\mathrm {GB}}} {\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }((\operatorname {\mathrm {Ind}}_F^{{\mathbf {Q}}} \rho )_p)$ are all positive,
we may then follow the same recipe and define the Greenberg–Benois ${\mathcal {L}}$ - invariant attached to $\rho $ (with respect to $\{D_{{\mathfrak {p}}}\}_{{\mathfrak {p}}|p}$ )
4 Fontaine–Mazur ${\mathcal {L}}$ -invariants
To define the Fontaine–Mazur ${\mathcal {L}}$ -invariants, we fix a finite extension K over ${\mathbf {Q}}_p$ . We shall be considering Galois representations
where E is (again) a finite extension of ${\mathbf {Q}}_p$ . In what follows, we consider the $(\varphi , N)$ -module ${\mathbf {D}}_{{\mathrm {st}}}(\rho )$ associated with $\rho $ . Note that if $K_0$ is the maximal unramified extension of ${\mathbf {Q}}_p$ in K, then ${\mathbf {D}}_{{\mathrm {st}}}(\rho )$ is a priori a $K_0$ -vector space. However, we shall linearize everything by base change to E.
Let q be the order of the residue field of K. We further assume $\rho $ enjoys the following properties:
-
(B1) The representation $\rho $ is semistable with Hodge–Tate weights $k_{\sigma , 1} \leq k_{\sigma , 2} \leq \cdots \leq k_{\sigma , n-1}\leq k_{\sigma , n}$ , where $\sigma : K \hookrightarrow \overline {{\mathbf {Q}}}_p$ . The Hodge–Tate weights give rise to the de Rham filtration $\operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^{\bullet } {\mathbf {D}}_{{\mathrm {st}}}(\rho ) = [\operatorname {\mathrm {Fil}}^{k_{\bullet , 1}}_{\operatorname {\mathrm {dR}}} {\mathbf {D}}_{{\mathrm {st}}}(\rho ) \supset \operatorname {\mathrm {Fil}}^{k_{\bullet , 2}}_{\operatorname {\mathrm {dR}}} {\mathbf {D}}_{{\mathrm {st}}}(\rho ) \supset \cdots \supset \operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^{k_{\bullet , n}}{\mathbf {D}}_{{\mathrm {st}}}(\rho )]$ .
-
(B2) The linearized Frobenius eigenvalues on ${\mathbf {D}}_{{\mathrm {st}}}(\rho )$ are $q^m, \ldots, q^{m-n+1}$ .
-
(FM1) Let $D_{(\varphi , N)}^{(i)}$ be the eigenspace in ${\mathbf {D}}_{{\mathrm {st}}}(\rho )$ on which the Frobenius acts via $q^{m-i}$ , and we assume that the induced monodromy operator N on $D_{(\varphi , N)}^{(i)}$ gives an isomorphism
$$\begin{align*}N : D_{(\varphi, N)}^{(i)} \rightarrow D_{(\varphi, N)}^{(i+1)}. \end{align*}$$ -
(FM2) Define Frobenius filtration $\operatorname {\mathrm {Fil}}_{\bullet }^{\varphi }{\mathbf {D}}_{{\mathrm {st}}}(\rho )$ by $\operatorname {\mathrm {Fil}}_j^{\varphi }{\mathbf {D}}_{{\mathrm {st}}}(\rho ) := \sum _{i>n-1-j} D_{(\varphi , N)}^{(i)}$ and assume the orthogonality
$$\begin{align*}{\mathbf{D}}_{{\mathrm{st}}}(\rho) = \operatorname{\mathrm{Fil}}_{\operatorname{\mathrm{dR}}}^{k_{\bullet, i}} {\mathbf{D}}_{{\mathrm{st}}}(\rho) \oplus \operatorname{\mathrm{Fil}}_i^{\varphi}{\mathbf{D}}_{{\mathrm{st}}}(\rho). \end{align*}$$
Lemma 4.1. Keep the notations and the assumptions as above. We abuse the notation and denote by $\operatorname {\mathrm {Gr}}_{\operatorname {\mathrm {dR}}}^{n-1}{\mathbf {D}}_{{\mathrm {st}}}(\rho ) := \operatorname {\mathrm {Fil}}^{k_{\bullet , n-1}}_{\operatorname {\mathrm {dR}}}{\mathbf {D}}_{{\mathrm {st}}}(\rho )/\operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^{k_{\bullet , n}} {\mathbf {D}}_{{\mathrm {st}}}(\rho )$ . Then, we have an inclusion
Proof Indeed, we have a sequence of identifications
where the third and the forth identifications follow from the orthogonality assumption.
Lemma 4.2. For every i, we have
Moreover, $m <k_{\sigma , n}$ for every $\sigma : K \hookrightarrow \overline {{\mathbf {Q}}}_p$ .
Proof Consider the twisted Galois representation $\rho (m)$ . One can similarly define the Frobenius filtration $\operatorname {\mathrm {Fil}}_{\bullet }^{\varphi }{\mathbf {D}}_{{\mathrm {st}}}(\rho (m))$ , and we denote by $D_{(\varphi , N)}^{(i)}(m)$ the graded pieces. Since each $\operatorname {\mathrm {Fil}}^{\varphi }_i {\mathbf {D}}_{{\mathrm {st}}}(\rho (m))$ is a $(\varphi , N)$ -module, [Reference Benois2, Proposition 1.2.7 (ii)] implies that we have an associated filtration $\operatorname {\mathrm {Fil}}_{\bullet }{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho (m))$ such that ${\mathcal {D}}_{{\mathrm {st}}}(\operatorname {\mathrm {Fil}}_{\bullet }{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho (m))) = \operatorname {\mathrm {Fil}}^{\varphi }_{i}{\mathbf {D}}_{{\mathrm {st}}}(\rho (m))$ .
Consider $\operatorname {\mathrm {Gr}}_{n}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho (m))$ . One sees that
on which the semistable Frobenius acts via $1$ . Hence, by [Reference Rosso17, Proposition 2.4], $\operatorname {\mathrm {Gr}}_{n}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho (m))$ is crystalline and
This shows that $\operatorname {\mathrm {rank}}_{K_0 \otimes _{{\mathbf {Q}}_p} E} D_{(\varphi , N)}^{(0)}(m) = 1$ . Using the formula in loc. cit., one also sees that $k_{\sigma , n}> m$ .
Since $\operatorname {\mathrm {rank}}_{K_0 \otimes _{{\mathbf {Q}}_p} E} D_{(\varphi , N)}^{(0)}(m) = 1$ , we see that $\operatorname {\mathrm {rank}}_{K_0 \otimes _{{\mathbf {Q}}_p} E} D_{(\varphi , N)}^{(0)} = 1$ . The result then can be concluded by applying (FM1).
Thanks to Lemma 4.1 and Lemma 4.2, we can now define the Fontaine–Mazur ${\mathcal {L}}$ -invariant. Let $v_0$ be a $K_0 \otimes _{{\mathbf {Q}}_p} E$ -basis for $D_{(\varphi , N)}^{(0)}$ and let $v_1 := Nv_0$ , which is a a $K_0 \otimes _{{\mathbf {Q}}_p} E$ -basis for $D_{(\varphi , N)}^{(1)}$ . The Fontaine–Mazur ${\mathcal {L}}$ - invariant attached to $\rho $ is then defined to be ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}(\rho ) \in K_0 \otimes _{{\mathbf {Q}}_p}E$ such that
Remark 4.3. In fact, if we write $\operatorname {\mathrm {Gr}}_{\operatorname {\mathrm {dR}}}^i {\mathbf {D}}_{{\mathrm {st}}}(\rho ) := \operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^{k_i}{\mathbf {D}}_{{\mathrm {st}}}(\rho )/\operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^{k_{i+1}} {\mathbf {D}}_{{\mathrm {st}}}(\rho )$ , then a similar argument as in Lemma 4.1 shows that
By using this inclusion, one can similarly define the i-th Fontaine–Mazur ${\mathcal {L}}$ -operator attached to $\rho $ to be ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}^{(i)}(\rho )\in K_0 \otimes _{{\mathbf {Q}}_p}E$ such that $v_{i-1} - {\mathcal {L}}_{\operatorname {\mathrm {FM}}}^{(i)}v_{i}\in \operatorname {\mathrm {Gr}}_{\operatorname {\mathrm {dR}}}^{n-i}{\mathbf {D}}_{{\mathrm {st}}}(\rho )$ , where $v_j = N^j v$ . Such a strategy was taken in [Reference Besser and de Shalit1]. However, it is believed that ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}^{(0)}(\rho ) = {\mathcal {L}}_{\operatorname {\mathrm {FM}}}(\rho )$ should determine all the other ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}^{(i)}(\rho )$ ’s (see, for example, [op. cit., §4.3.2]). Hence, we focus on ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}(\rho )$ . Moreover, one shall see, in what follows, that it is ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}(\rho )$ we can relate to Greenberg–Benois ${\mathcal {L}}$ -invariants.
5 Comparing the two ${\mathcal {L}}$ -invariants
The aim of this section is to prove the comparison theorem (Theorem 5.4). However, as aforementioned, to define ${\mathcal {L}}$ -invariants, there are some constraints one needs to put on the Galois representations. For reader’s convenience, we collect all the assumptions in §5.1 and briefly discuss a folklore about these assumptions.
5.1 Assumptions on the Galois representation
Let F be a number field and let E be a finite extension of ${\mathbf {Q}}_p$ such that, for every prime ideal ${\mathfrak {p}}$ in ${\mathcal {O}}_F$ sitting above p, $F_{{\mathfrak {p}}} \subset E$ . Write $F_{{\mathfrak {p}}, 0}$ for the maximal unramified extension of ${\mathbf {Q}}_p$ in $F_{{\mathfrak {p}}}$ ; we further assume that $F_{{\mathfrak {p}}, 0} = {\mathbf {Q}}_p$ for every ${\mathfrak {p}}$ . Suppose we are given a Galois representation
that is unramified outside a finite set of places. Let S be the set of places in F such that $\rho $ ramifies. We make the following assumptions:
-
(I) Basic assumptions:
-
(B1) For any prime ideal ${\mathfrak {p}} \subset {\mathcal {O}}_{F}$ sitting above p, $\rho _{{\mathfrak {p}}} := \rho |_{\operatorname {\mathrm {Gal}}_{F_{{\mathfrak {p}}}}}$ is semistable with Hodge–Tate weights $0\leq k_{{\mathfrak {p}}, \sigma , 1} \leq k_{{\mathfrak {p}}, \sigma , 2} \leq \cdots \leq k_{{\mathfrak {p}}, \sigma , n-1} \leq k_{{\mathfrak {p}}, n}$ , where $\sigma : F_{{\mathfrak {p}}} \hookrightarrow \overline {{\mathbf {Q}}}_p$ .
-
(B2) For any prime ideal ${\mathfrak {p}} \subset {\mathcal {O}}_{F}$ sitting above p, the Frobenius eigenvalues on ${\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}})$ are $p^m$ , …, $p^{m-n+1}$ such that $k_{{\mathfrak {p}}, \sigma , n}> m > k_{{\mathfrak {p}}, \sigma , n-1}$ , where the first inequality is always guaranteed by Lemma 4.2.Footnote 8
-
-
(II) Fontaine–Mazur assumptions:
-
(FM1) For any ${\mathfrak {p}}|p$ , let $D_{{\mathfrak {p}}, (\varphi , N)}^{(i)}$ be the eigenspace in ${\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}})$ on which the Frobenius acts via $p^{m-i}$ . We assume that the induced monodromy operator N on $D_{{\mathfrak {p}}, (\varphi , N)}^{(i)}$ gives an isomorphism
$$\begin{align*}N: D_{{\mathfrak{p}}, (\varphi, N)}^{(i)} \rightarrow D_{{\mathfrak{p}}, (\varphi, N)}^{(i+1)}. \end{align*}$$ -
(FM2) Define $\operatorname {\mathrm {Fil}}_j^{\varphi } {\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}) := \sum _{i>n-1-j} D_{{\mathfrak {p}}, (\varphi , N)}^{(i)}$ , and we call the ascending filtration $\operatorname {\mathrm {Fil}}_{\bullet }^{\varphi } {\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}})$ the Frobenius filtration on ${\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}})$ . We assume the orthogonality
$$\begin{align*}{\mathbf{D}}_{{\mathrm{st}}}(\rho_{{\mathfrak{p}}}) = \operatorname{\mathrm{Fil}}_{\operatorname{\mathrm{dR}}}^{k_{{\mathfrak{p}}, \bullet, i}}{\mathbf{D}}_{{\mathrm{st}}}(\rho_{{\mathfrak{p}}}) \oplus \operatorname{\mathrm{Fil}}_i^{\varphi} {\mathbf{D}}_{{\mathrm{st}}}(\rho_{{\mathfrak{p}}}). \end{align*}$$
-
-
(III) Greenberg–Benois assumptions:
-
(GB1) Vanishing of the Bloch–Kato Selmer groups
$$\begin{align*}H_f^1({\mathbf{Q}}, \operatorname{\mathrm{Ind}}^{{\mathbf{Q}}}_{F} \rho(m)) = H^1_f({\mathbf{Q}}, \operatorname{\mathrm{Ind}}^{{\mathbf{Q}}}_{F}\rho^{\vee}(1-m)) = 0. \end{align*}$$ -
(GB2) Vanishing of the zero-degree Galois cohomology
$$\begin{align*}H^0(\operatorname{\mathrm{Gal}}_{{\mathbf{Q}}_S}, \operatorname{\mathrm{Ind}}^{{\mathbf{Q}}}_{F} \rho(m)) = H^0(\operatorname{\mathrm{Gal}}_{{\mathbf{Q}}_S},\operatorname{\mathrm{Ind}}^{{\mathbf{Q}}}_F\rho^{\vee}(1-m)) =0. \end{align*}$$ -
(GB3) The associated $(\varphi , \Gamma )$ -module ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }((\operatorname {\mathrm {Ind}}_F^{{\mathbf {Q}}}\rho (m))_p) = \bigoplus _{{\mathfrak {p}}|p} {\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _{{\mathfrak {p}}} (m))$ does not admit a subquotient of the form $U_{k,r}$ with $k\geq 1$ and $r\geq 0$ ([Reference Benois2, §2.1.2]).
-
(GB4) For any ${\mathfrak {p}}|p$ , the space $W_{{\mathfrak {p}}, 0}$ for $\rho _{{\mathfrak {p}}}(m)$ vanishes (see [Reference Benois2, Proposition 2.1.7] or [Reference Rosso17, pp. 1238]).
-
Remark 5.1. For every ${\mathfrak {p}} | p$ , the Frobenius filtration $\operatorname {\mathrm {Fil}}_{\bullet }^{\varphi }{\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}})$ defines a filtration $\operatorname {\mathrm {Fil}}_{\bullet } {\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho )$ (similar as in the proof of Lemma 4.2). One observes that the graded pieces $\operatorname {\mathrm {Gr}}_{i}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho )$ of this filtration are all of rank $1$ over ${\mathcal {R}}_{F_{{\mathfrak {p}}}, E}$ (by [Reference Benois2, Proposition 1.2.7 (ii)]). In particular, $\operatorname {\mathrm {Fil}}_{\bullet }{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho )$ is a triangulation of ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho )$ . In fact, we have the following description for the graded pieces:
where
Remark 5.2. For every ${\mathfrak {p}} | p$ , since $k_{{\mathfrak {p}}, \sigma , n}> m > k_{{\mathfrak {p}},\sigma , n-1} $ , we see that $\operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^{k_{{\mathfrak {p}}, \bullet , n-1}-m} {\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}(m)) = \operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^0 {\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}(m))$ . Moreover, the orthogonality condition (FM2) implies that $\operatorname {\mathrm {Fil}}_{n-1}^{\varphi }{\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}(m))$ is a regular $(\varphi , N)$ -submodule of ${\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}(m))$ . Hence, in what follows, we naturally work with $D_{{\mathfrak {p}}} := \operatorname {\mathrm {Fil}}_{n-1}^{\varphi }{\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}(m)) \subset {\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}(m))$ and $D = \bigoplus _{{\mathfrak {p}} | p}D_{{\mathfrak {p}}} \subset \bigoplus _{{\mathfrak {p}} |p}{\mathbf {D}}_{{\mathrm {st}}}(\rho _{{\mathfrak {p}}}(m))$ . Moreover, in our situation, we shall see in the proof (e.g., (5.2)) that the corresponding $\operatorname {\mathrm {Gr}}_1^{\operatorname {\mathrm {GB}}} {\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }((\operatorname {\mathrm {Ind}}_F^{{\mathbf {Q}}}\rho )_p)$ has positive Hodge–Tate weights, and so we remove such assumption in (GB4).
Remark 5.3. We have many assumptions on our Galois representation $\rho $ . On the one hand, one sees that they are necessary in order to attach both ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}$ and ${\mathcal {L}}_{\operatorname {\mathrm {FM}}}$ to it. On the other hand, we remark that it is a folklore that they shall appear as Galois representations for automorphic forms of unitary groups whose corresponding Shimura varieties can be p-adically uniformized by Drinfeld’s upper-half spaces. For example, we are requiring maximal monodromy on our Galois representations. Such a phenomenon is expected to appear for the Galois representations attached to unitary automorphic representations that are Steinberg at p.
5.2 The main theorem
Theorem 5.4. Keep the notations and assumptions as above. We have an equality
Proof The proof of the theorem is similar to the proof of [Reference Benois2, Proposition 2.3.7], which relies on the following three steps:
Step 1. Fontaine–Mazur ${\mathcal {L}}$ -invariants and cohomology of $(\varphi , \Gamma )$ -modules.
Consider the triangulation $\operatorname {\mathrm {Fil}}_{\bullet }{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho )$ in Remark 5.1. We define
Hence, $\widetilde {W}_{{\mathfrak {p}}}$ sits inside the short exact sequence
for $\delta _{{\mathfrak {p}}, i} : F_{{\mathfrak {p}}}^\times \rightarrow E^\times $ described as in Remark 5.1.
As a result, $\widetilde {W}_{{\mathfrak {p}}}$ defines a class
However, by construction, we know that $\widetilde {W}_{{\mathfrak {p}}}$ is semistable (since ${\mathcal {D}}_{{\mathrm {st}}}(\widetilde {W}_{{\mathfrak {p}}}) = D_{{\mathfrak {p}}, (\varphi , N)}^{(0)} \oplus D_{{\mathfrak {p}}, (\varphi , N)}^{(1)}$ ), and so $\operatorname {\mathrm {cl}}(\widetilde {W}_{{\mathfrak {p}}}) \in H_{{\mathrm {st}}}^1({\mathcal {R}}_{F_{{\mathfrak {p}}}, E}(\delta _{{\mathfrak {p}}, 1}\delta _{{\mathfrak {p}}, 0}^{-1}))$ . Recall that $H_{{\mathrm {st}}}^1({\mathcal {R}}_{F_{{\mathfrak {p}}}, E} (\delta _{{\mathfrak {p}}, 1}\delta _{{\mathfrak {p}}, 0}^{-1}))$ can be computed via the complex $C_{{\mathrm {st}}}^{\bullet }({\mathcal {R}}_{F_{{\mathfrak {p}}}, E}(\delta _{{\mathfrak {p}}, 1}\delta _{{\mathfrak {p}}, 0}^{-1}))$ with
where the first map is given by $a \mapsto \big (a\ \mod \operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^0 {\mathcal {D}}_{{\mathrm {st}}}({\mathcal {R}}_{F_{{\mathfrak {p}}}, E}(\delta _{{\mathfrak {p}}, 1}\delta _{{\mathfrak {p}}, 0}^{-1})), (\varphi -1)a, Na \big )$ , while the second arrow is defined by $(a, b, c)\mapsto Nb-(p\varphi -1)c$ .
Now, choose a basis $v_{{\mathfrak {p}}, 0}\in D_{{\mathfrak {p}}, (\varphi , N)}^{(0)}$ over $F_{{\mathfrak {p}}, 0}\otimes _{{\mathbf {Q}}_p} E$ and let $v_{{\mathfrak {p}}, 1} := Nv_{{\mathfrak {p}}, 0}$ , which is a $F_{{\mathfrak {p}}, 0}\otimes _{{\mathbf {Q}}_p} E$ -basis for $D_{{\mathfrak {p}}, (\varphi , N)}^{(1)}$ . We again denote by $v_{{\mathfrak {p}}, i}$ for the image of $v_{{\mathfrak {p}}, i}$ in ${\mathcal {D}}_{{\mathrm {st}}}(\widetilde {W}_{{\mathfrak {p}}}(\delta _{{\mathfrak {p}}, 0}^{-1}))$ . By the proof of [Reference Benois2, Proposition 1.4.4 (ii)], we know that the class $\operatorname {\mathrm {cl}}(\widetilde {W}_{{\mathfrak {p}}})$ in $H^1(C_{{\mathrm {st}}}^{\bullet }({\mathcal {R}}_{F_{{\mathfrak {p}}}, E}(\delta _{{\mathfrak {p}}, 1}\delta _{{\mathfrak {p}}, 0}^{-1})))$ is given by
where $a\in {\mathcal {D}}_{{\mathrm {st}}}(\widetilde {W}_{{\mathfrak {p}}}(\delta _{{\mathfrak {p}}, 0}^{-1}))$ such that $v_{{\mathfrak {p}}, 0} + a \in \operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^0{\mathcal {D}}_{{\mathrm {st}}}(\widetilde {W}_{{\mathfrak {p}}}(\delta _{{\mathfrak {p}}, 0}^{-1}))$ . After untwisting, a defines an element, still denoted by $a\in {\mathcal {D}}_{{\mathrm {st}}}(\widetilde {W}_{{\mathfrak {p}}})$ such that $v_{{\mathfrak {p}}, 0} +a \in \operatorname {\mathrm {Fil}}_{\operatorname {\mathrm {dR}}}^{k_{{\mathfrak {p}}, \bullet , n}}{\mathcal {D}}_{{\mathrm {st}}}(\widetilde {W}_{{\mathfrak {p}}})$ . However, by construction,
Hence, we conclude that
Step 2. Computing ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}(\rho )$ .
Next, we would also like to compute the Greenberg–Benois ${\mathcal {L}}$ -invariant ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}(\rho )$ via cohomology of $(\varphi , \Gamma )$ -modules. As before, because of the decomposition ${\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }((\operatorname {\mathrm {Ind}}^{{\mathbf {Q}}}_{F}\rho )_p) = \bigoplus _{{\mathfrak {p}}|p}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\operatorname {\mathrm {Ind}}_{F_{{\mathfrak {p}}}}^{{\mathbf {Q}}_p}\rho _{{\mathfrak {p}}})$ , we can study each ${\mathfrak {p}}$ individually. Hence, fix ${\mathfrak {p}} | p$ . Computing the five-step filtration (3.1) explicitly, we have
which gives rise to a five-step filtration $\operatorname {\mathrm {Fil}}_{\bullet }^{\operatorname {\mathrm {GB}}}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _{{\mathfrak {p}}}(m))$ .
Let us simplify the notation and write
Similar as before, we see that $W_{{\mathfrak {p}}}$ sits inside the short exact sequence
where
By taking cohomology, we have the connecting homomorphism
where the equation follows from Lemma 2.1. Denoted by $\alpha _{\delta _{{\mathfrak {p}}, 1}'}^*$ and $\beta _{\delta _{{\mathfrak {p}}, 1}'}^*$ the two classes in $H^1({\mathcal {R}}_{F_{{\mathfrak {p}}, E}}(\delta _{{\mathfrak {p}}, 1}'))$ in Lemma 2.2. We know from loc. cit. that $\partial $ gives rise to a unique number ${\mathcal {L}}(W_{{\mathfrak {p}}})\in E$ such that
We claim that
Note that, in the definition of ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}(\rho (m))$ , one studies the cohomology of $\operatorname {\mathrm {Gr}}_1^{\operatorname {\mathrm {GB}}}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger }(\rho _{{\mathfrak {p}}}(m))$ . However, we are now having cohomology classes in $H^1(\operatorname {\mathrm {Gr}}_0^{\operatorname {\mathrm {GB}}}{\mathbf {D}}_{\operatorname {\mathrm {rig}}}^{\dagger } (\rho _{{\mathfrak {p}}}(m)))$ . To resolve this, we look at the short exact sequence
By [Reference Benois2, Proposition 2.2.4], the Greenberg–Benois ${\mathcal {L}}$ -invariant computed by this exact sequence (at each ${\mathfrak {p}}$ ) is the same as ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}(\rho (m))$ . Here,
and we want to compute ${\mathcal {L}}_{\operatorname {\mathrm {GB}}}(\rho (m))$ using the cohomology of ${\mathcal {R}}_{F_{{\mathfrak {p}}}, E}(\kappa _{{\mathfrak {p}}, 1})$ .
By (B2), we have $k_{{\mathfrak {p}}, \sigma , n-1}-m+1\leq 0$ and we let $u_{{\mathfrak {p}}} := \min _{\sigma }\{ k_{{\mathfrak {p}}, \sigma , n-1}-m+1\}$ . By [Reference Rosso17, (2.8)], there is an injection
where $x_{u_{{\mathfrak {p}}}}$ , $x_{k_{{\mathfrak {p}}, \bullet , n-1}-m+1}$ , $y_{u_{{\mathfrak {p}}}}$ , and $ y_{k_{{\mathfrak {p}}, \bullet , n-1}-m+1}$ are as defined in loc. cit. Footnote 9 ,Footnote 10 By the discussion on [Reference Rosso17, pp. 1238], we have a commutative diagram
where $\iota _c$ is an isomorphism. Moreover, [op. cit., Corollary 3.9] yields that
In particular, if ${\mathcal {L}}_{{\mathfrak {p}}}\in E$ such that
then
By definition, $H^1\left (\bigoplus _{{\mathfrak {p}}|p}D_{{\mathfrak {p}}}^{\vee }(1-m), \operatorname {\mathrm {Ind}}^{{\mathbf {Q}}}_{F}\rho ^{\vee }(1-m)\right ) \cong \bigoplus _{{\mathfrak {p}}|p}\frac {H^1(W_{{\mathfrak {p}}}^{\vee }(\chi _{\operatorname {\mathrm {cyc}}}))}{H_f^1(W_{{\mathfrak {p}}}^{\vee }(\chi _{\operatorname {\mathrm {cyc}}}))}$ . The vertical morphism in (5.4) is compatible with the natural morphism
induced from the short exact sequence (5.2). Note that the exact sequence
is dual to the exact sequence
We have $\partial ({\mathcal {L}}_{{\mathfrak {p}}} x_{k_{{\mathfrak {p}}, \bullet , n-1}-m+1} + y_{k_{{\mathfrak {p}}, \bullet , n-1}-m+1}) = 0 \in H^2({\mathcal {R}}_{F_{{\mathfrak {p}}}, E}(\kappa _{{\mathfrak {p}}, 0}))$ and $\beta _{\delta _{{\mathfrak {p}}, 1}'}^* + {\mathcal {L}}(W_{{\mathfrak {p}}}) \alpha _{\delta _{{\mathfrak {p}}, 1}'}^* = 0 \in H^1(W_{{\mathfrak {p}}})$ . Moreover, using the relation between $x_{k_{{\mathfrak {p}}, \bullet , n-1}-m+1}$ (resp., $y_{k_{{\mathfrak {p}}, \bullet , n-1}-m+1}$ ) and $\beta _{\delta _{{\mathfrak {p}}, 1}'}^*$ (resp., $-\alpha _{\delta _{{\mathfrak {p}}, 1}'}^*$ ), one sees that
which concludes our claim.
Step 3. Conclusion.
By construction, $W_{{\mathfrak {p}}}$ defines a class (see (5.2))
Note that, as classes in $H^1({\mathcal {R}}_{F_{{\mathfrak {p}}}, E}(\delta _{{\mathfrak {p}}, 1}\delta _{{\mathfrak {p}}, 0}^{-1}))$ , we have
Unwinding everything, we have
for some $c\in E$ . In particular, we conclude that
and so,
by (5.3).
Acknowledgements
This paper grew from a working group with Ting-Han Huang, Martí Roset Julià, and Giovanni Rosso, and I would like to thank them for interesting discussions. I especially thank Giovanni Rosso for valuable feedback regarding the early draft of this paper. I also thank David Loeffler for pointing out a mistake in an early version of this paper. I would also like to thank Muhammad Manji for interesting discussions on $(\varphi , \Gamma )$ -modules. Part of the work was done when I was visiting National Center for Theoretical Sciences in Taipei; I would like to thank the hospitality of the institute and Ming-Lun Hsieh. Finally, I thank the anonymous referees for their corrections and valuable suggestions, which helped to improve the exposition of the article. This work is supported by the ERC Consolidator grant “Shimura varieties and the BSD conjecture” and the Irish Research Council under grant number IRCLA/2023/849 (HighCritical).