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

On patched completed homology and a conjecture of Venkatesh

Douglas Molin
Abstract

Let FF be a CM field and Π\Pi a regular algebraic cuspidal cohomological representation of 𝐆=PGL2/F\mathbf{G}=\operatorname{PGL}_{2}/F. A conjecture of Venkatesh describes the structure of the contribution of Π\Pi to the homology of the locally symmetric spaces associated to 𝐆\mathbf{G}. We investigate this conjecture in the setting of pp-adic homology with pp a totally split prime. Along the way, we elaborate on the relations between Venkatesh’s conjecture and completed homology, the Taylor-Wiles method and the pp-adic local Langlands correspondence. Our main result is a ‘big R=TR=T’ theorem in characteristic 0, from which we deduce a variant of the pp-adic realisation of Venkatesh’s conjecture, conditional on various natural conjectures and technical assumptions.

1 Introduction

Let 𝐆=PGL2/F\mathbf{G}=\operatorname{PGL}_{2}/F where FF is a CM field and suppose Π\Pi is a regular algebraic cuspidal automorphic representation of 𝐆(𝔸F)\mathbf{G}(\mathbb{A}_{F}) which is cohomological with respect to an algebraic representation of some weight λ\lambda. Then Π\Pi contributes to the homology of the locally symmetric spaces associated to 𝐆\mathbf{G} in the following sense.

Let l0=[F:]/2l_{0}=[F:\mathbb{Q}]/2 and fix a compact open subgroup U𝐆(𝔸F)U\subset\mathbf{G}(\mathbb{A}_{F}^{\infty}). There is an associated locally symmetric space XUX_{U} which – if UU is small enough – is a smooth manifold of dimension 3l03l_{0}, and we can consider its homology H(XU,λ)H_{\ast}(X_{U},\mathcal{L}_{\lambda}) with coefficients in a pp-adic local system λ\mathcal{L}_{\lambda} associated to the weight λ\lambda of Π\Pi. The homology is a finite-dimensional p\mathbb{Q}_{p}-vector space with an action of Hecke operators, and we define its ‘Π\Pi-part’ as the eigenspace

H(XU,λ)ΠH(XU,λ)H_{\ast}(X_{U},\mathcal{L}_{\lambda})_{\Pi}\subset H_{\ast}(X_{U},\mathcal{L}_{\lambda})

of homology classes with the same system of Hecke eigenvalues as Π\Pi. Let us assume for the purpose of this introduction that all eigenvalues of Π\Pi are rational and that Π\Pi is unramified at all places above pp. Using (𝔤,K)(\mathfrak{g},K)-cohomology, one can compute the dimensions of the graded pieces of the Π\Pi-part and see that, if d=dim(Π)Ud=\dim(\Pi^{\infty})^{U}, then

dimpHl0+i(XU,λ)Π={d(l0i) if i=0,,l00 otherwise.\dim_{\mathbb{Q}_{p}}H_{l_{0}+i}(X_{U},\mathcal{L}_{\lambda})_{\Pi}=\begin{cases}d\cdot\binom{l_{0}}{i}&\text{ if }i=0,\dots,l_{0}\\ 0&\text{ otherwise.}\end{cases}

To give an arithmetic explanation of this ‘spreading out’ in multiple degrees of Hecke eigenspaces, Venkatesh conjectures the existence of a vector space VΠV_{\Pi} and a natural action of the exterior algebra VΠ\wedge^{\ast}V_{\Pi} on H(XU,λ)ΠH_{\ast}(X_{U},\mathcal{L}_{\lambda})_{\Pi} such that the homology is a free graded module of rank dd over VΠ\wedge^{\ast}V_{\Pi}. In fact, Venkatesh’s conjecture is a more refined statement at the level of only conjecturally existing motives, but in the case of pp-adic homology one can make a concrete prediction. Letting SS denote the set of places of FF where Π\Pi ramifies, one can associate a Galois representation ρ:GF,SGL2(p)\rho:G_{F,S}\to\operatorname{GL}_{2}(\mathbb{Q}_{p}) to Π\Pi using the construction in [19] or [34]. Here, GF,SG_{F,S} is the Galois group of the maximal extension of FF unramified outside of SS. One expects VΠV_{\Pi} to be a Galois cohomology group defined in terms of ρ\rho, namely

VΠ=Hf1(GF,S,ad0ρ(1)),V_{\Pi}=H^{1}_{f}(G_{F,S},\operatorname{ad}^{0}\rho(1)),

the dual of the adjoint Bloch-Kato Selmer group (see Section 4). The Bloch-Kato conjecture predicts that the dimension of this vector space equals the order of vanishing at s=1s=1 of the adjoint LL-function L(s,ad0Π)L(s,\operatorname{ad}^{0}\Pi), which is known to be l0=[F:]/2l_{0}=[F:\mathbb{Q}]/2. In this way, one hopes to obtain an arithmetic proof of the dimension formula ()(\star).

In this article, we prove a version of the pp-adic conjecture under various conjectures and technical assumptions on ρ\rho, as well as the assumptions that p5p\geq 5 and pp is totally split in FF. Before stating our main result, let us describe in broad terms the strategy of proof. The main idea is to place the Π\Pi-part in a pp-adic family by relating it to completed homology. This is a pp-adic representation H~\widetilde{H} of G=vp𝐆(Fv)G=\prod_{v\mid p}\mathbf{G}(F_{v}) and it has an action of a ‘big’ Hecke algebra TT. The Π\Pi-part is related to completed homology by a spectral sequence. Using the Taylor-Wiles method for completed homology developed by Gee-Newton [16], we prove a ‘big R=TR=T’ theorem in characteristic 0 which identifies the spectrum of TT with the unrestricted deformation space of the Galois representation ρ\rho associated to Π\Pi.

The proof of the ‘R=TR=T’ theorem depends crucially on the condition that pp is totally split in FF and a type of local-global compatibility assumption at the places vpv\mid p. Under these assumptions, we are able to utilise the pp-adic local Langlands correspondence for GL2(p)\operatorname{GL}_{2}(\mathbb{Q}_{p}) as described in [32]. Paškūnas’ theory provides an equivalence between a category of GG-representations in which we find H~\widetilde{H} and a certain category of modules over a local deformation ring RlocR_{\operatorname{loc}}. This is the p\mathbb{Q}_{p}-algebra representing deformations of the tuple (ρv)vp(\rho_{v})_{v\mid p}, where ρv\rho_{v} denotes the restriction of ρ\rho to a decomposition group at vv. Under this equivalence, the image of H~\widetilde{H} is finitely generated, and thus we are able to carry out the depth estimates in the Taylor-Wiles method with this object. We learned of this strategy in [31].

Let us return to the Π\Pi-part H(XU,λ)ΠH_{\ast}(X_{U},\mathcal{L}_{\lambda})_{\Pi}. It is a finitely generated module over the big Hecke algebra R=TR=T, and we think of it as a coherent sheaf on the deformation space of ρ\rho. Its support is a closed subscheme consisting of deformations of ρ\rho satisfying local conditions coming from pp-adic Hodge theory. This subscheme can tautologically be described as the intersection between a space of deformations of ρ\rho with the space of deformations of the local factors ρv\rho_{v} satisfying the aforementioned conditions, where the intersection takes place within the space of all deformations of (ρv)vp(\rho_{v})_{v\mid p}. Under the Bloch-Kato conjecture, this intersection consists of a single point. In this way, one is able to view the Π\Pi-part as a module over the Tor\operatorname{Tor}-algebra representing the derived intersection, and the goal is to prove that it is free and moreover to identify the Tor\operatorname{Tor}-algebra with the exterior algebra VΠ\wedge^{\ast}V_{\Pi} mentioned above. Thus, the non-transverseness of the intersection of deformation spaces explains the spreading out into multiple degrees.

For the purpose of stating a simplified version of our main result, let RlocR_{\operatorname{loc}} be as above, i.e. the deformation ring parametrising deformations of the local representations ρv\rho_{v} for vpv\mid p. There exists a quotient RlocRloc(λ)R_{\operatorname{loc}}\twoheadrightarrow R_{\operatorname{loc}}(\lambda) corresponding to deformations satisfying the pp-adic Hodge theoretic conditions determined by the weight λ\lambda. Finally, let RglR_{\operatorname{gl}} denote the representing ring of deformations of ρ\rho unramified outside of SS. It turns out that RglR_{\operatorname{gl}} represents a closed subscheme of the formal spectrum of RlocR_{\operatorname{loc}}, and the space of global deformations satisfying the local conditions induced by λ\lambda is represented by the completed tensor product RglRlocRloc(λ)R_{\operatorname{gl}}\otimes_{R_{\operatorname{loc}}}R_{\operatorname{loc}}(\lambda).

We now present our main result (Theorem 7.12) without spelling out all of the assumptions.

Theorem 1.1.

Suppose p5p\geq 5 is totally split in FF and that ρ\rho has no global deformations satisfying the local conditions induced by λ\lambda. Then, under various additional assumptions on ρ\rho, the graded LL-vector space H(XU,λ)ΠH_{\ast}(X_{U},\mathcal{L}_{\lambda})_{\Pi} has a canonical structure of free module of rank dd over the Tor\operatorname{Tor}-algebra

TorRloc(Rgl,Rloc(λ)).\operatorname{Tor}_{\ast}^{R_{\operatorname{loc}}}\big{(}R_{\operatorname{gl}},R_{\operatorname{loc}}(\lambda)\big{)}.

If, in addition, RglR_{\operatorname{gl}} is formally smooth, there is a canonical isomorphism of graded-commutative rings

TorRloc(Rgl,Rloc(λ))LHf1(GF,S,ad0ρ(1)).\operatorname{Tor}_{\ast}^{R_{\operatorname{loc}}}\big{(}R_{\operatorname{gl}},R_{\operatorname{loc}}(\lambda)\big{)}\cong\wedge^{\ast}_{L}H^{1}_{f}\big{(}G_{F,S},\operatorname{ad}^{0}\rho(1)\big{)}.

The assumption that ρ\rho has no global deformations satisfying the conditions induced by λ\lambda is predicted to hold by the Bloch-Kato conjecture and is equivalent to the statement dimVΠ=l0\dim V_{\Pi}=l_{0}. Proving the final part of the theorem amounts to leveraging the smoothness assumption to compute the Tor\operatorname{Tor}-groups of the intersection RglRlocRloc(λ)R_{\operatorname{gl}}\otimes_{R_{\operatorname{loc}}}R_{\operatorname{loc}}(\lambda). This can be done by studying a natural short exact sequence of Galois cohomology groups coming from the Poitou-Tate sequence, by which the statement follows from an explicit computation with Koszul complexes.

A similar theorem for GLn/\operatorname{GL}_{n}/\mathbb{Q}, n2n\geq 2, has been proved using the theory of eigenvarieties and overconvergent cohomology by Hansen-Thorne [18, Theorem 1.1], and our application of the Poitou-Tate sequence is based on an analogous argument therein. Notably, our methods require no ‘small slope’ (or even finite slope) assumption on Π\Pi since we do not use overconvergent cohomology. However, with our methods we can at present only consider the case n=2n=2 and pp totally split.

Let us briefly outline the content of this article. In section 2, we recall some facts about the homology of locally symmetric spaces and their Hecke operators. In section 3, we discuss some notions from homological and commutative algebra which are used in later sections. Section 4 contains a brief discussion of Galois cohomology and Selmer groups, and Section 5 is devoted to deformation rings. In section 6, we turn to the representation-theoretic part of the story and introduce completed homology. The final section is devoted to the proof of our main result.

This article forms the basis of the author’s ‘mittseminarium’ given in February of 2024 at the University of Gothenburg.

2 Setup

Let p5p\geq 5 be a prime and L/pL/\mathbb{Q}_{p} a finite extension with ring of integers 𝒪\mathcal{O}, uniformiser ϖ\varpi and residue field k=𝒪/ϖk=\mathcal{O}/\varpi. At times, we tacitly assume LL is ‘large enough’ (e.g. to contain Hecke eigenvalues). We fix an isomorphism ι:¯p\iota:\overline{\mathbb{Q}}_{p}\simeq\mathbb{C} and a CM field FF in which pp is totally split.

Unless otherwise stated, completed tensor products are taken over 𝒪\mathcal{O}. When RR is an 𝒪\mathcal{O}-algebra and 𝔭R\mathfrak{p}\subset R is an ideal not containing ϖ\varpi, we use the same symbol to denote the ideal 𝔭\mathfrak{p} and the ideal it generates inside R[1/ϖ]R[1/\varpi]. We denote by 𝔸F\mathbb{A}^{\infty}_{F} the ring of finite adèles of FF.

2.1 Arithmetic locally symmetric spaces

We begin by recalling the construction of the locally symmetric spaces associated to 𝐆=PGL2/F\mathbf{G}=\operatorname{PGL}_{2}/F. A complete reference is [22, §6.1-2]. We introduce the following notation:

  • 𝐆=PGL2/F\mathbf{G}=\operatorname{PGL}_{2}/F,

  • G=𝐆(F)G_{\infty}=\mathbf{G}(F\otimes_{\mathbb{Q}}\mathbb{R}),

  • KGK_{\infty}\subset G_{\infty} a maximal compact connected subgroup,

  • D=G/KD_{\infty}=G_{\infty}/K_{\infty},

  • l0=rankGrankK=[F:]/2l_{0}=\operatorname{rank}G_{\infty}-\operatorname{rank}K_{\infty}=[F:\mathbb{Q}]/2,

  • q0=12(dimDl0)q_{0}=\frac{1}{2}(\dim D_{\infty}-l_{0}).

The integer l0l_{0} is called the defect of 𝐆\mathbf{G}, and turns up in many different settings. In fact, we have dimD=3l0\dim D_{\infty}=3l_{0} and hence q0=l0q_{0}=l_{0}. For groups other than PGL2/F\operatorname{PGL}_{2}/F, one usually has q0l0q_{0}\neq l_{0}, and we have opted to maintain the distinction in what follows.

Definition 2.1.

Let U𝐆(𝔸F)U\subset\mathbf{G}(\mathbb{A}^{\infty}_{F}) be an open compact subgroup. The locally symmetric space associated to UU is the double quotient

XU=𝐆(F)\(D×𝐆(𝔸F)/U),X_{U}=\mathbf{G}(F)\backslash(D_{\infty}\times\mathbf{G}(\mathbb{A}^{\infty}_{F})/U),

where the action of 𝐆(F)\mathbf{G}(F) is the diagonal action.

The space XUX_{U} decomposes as a finite disjoint union of subspaces of the form Γi\D\Gamma_{i}\backslash D_{\infty} where Γi=𝐆(F)giUgi1\Gamma_{i}=\mathbf{G}(F)\cap g_{i}Ug_{i}^{-1} for some g𝐆(𝔸F)g\in\mathbf{G}(\mathbb{A}_{F}^{\infty}).

Definition 2.2.

A good subgroup U𝐆(𝔸F)U\subset\mathbf{G}(\mathbb{A}^{\infty}_{F}) is an open compact subgroup of the form vUv\prod_{v}U_{v} such that:

  • (i)

    For every vv, UvPGL2(𝒪Fv)U_{v}\subseteq\operatorname{PGL}_{2}(\mathcal{O}_{F_{v}}).

  • (ii)

    For every g𝐆(𝔸F)g\in\mathbf{G}(\mathbb{A}^{\infty}_{F}) and every hgUg1𝐆(F),h\in gUg^{-1}\cap\mathbf{G}(F), the eigenvalues of hh generate a torsion-free subgroup of F¯\overline{F} (i.e., the subgroup gUg1gUg^{-1} is ‘neat’).

Proposition 2.3.

[22, Lemma 6.1] Let U𝐆(𝔸F)U\subset\mathbf{G}(\mathbb{A}^{\infty}_{F}) be a good subgroup. Then XUX_{U} is a smooth manifold of dimension 3l03l_{0} and homotopy equivalent to the geometric realisation of a finite simplicial complex. Moreover, if UUU^{\prime}\subset U is a normal compact open subgroup, then UU^{\prime} is also good and XUXUX_{U^{\prime}}\to X_{U} is a Galois cover of smooth manifolds with Galois group U/UU/U^{\prime}.

2.2 Homology of arithmetic locally symmetric spaces

Throughout this section, U=UpUp𝐆(𝔸F)U=U_{p}U^{p}\subset\mathbf{G}(\mathbb{A}^{\infty}_{F}) is a good subgroup. Here, Up=vpUvU_{p}=\prod_{v\mid p}U_{v} and Up=vpUvU^{p}=\prod_{v\nmid p}U_{v}. We fix a discrete left [Up]\mathbb{Z}[U_{p}]-module MM, viewed as a [U]\mathbb{Z}[U]-module by letting UU act via the projection UUpU\to U_{p}.

Definition 2.4.

The local system associated to MM is the sheaf M\mathcal{L}_{M} of continuous sections of the map

𝐆(F)\(D×𝐆(𝔸F)/U)×M/U𝐆(F)\(D×𝐆(𝔸F)/U),\mathbf{G}(F)\backslash(D_{\infty}\times\mathbf{G}(\mathbb{A}^{\infty}_{F})/U)\times M/U\to\mathbf{G}(F)\backslash(D_{\infty}\times\mathbf{G}(\mathbb{A}^{\infty}_{F})/U),

We denote by H(XU,M)H_{\ast}(X_{U},\mathcal{L}_{M}) the homology of XUX_{U} with local coefficients M\mathcal{L}_{M}.

There are two complexes commonly used to compute the homology with local coefficients displayed above. The first is the adèlic complex, which is useful for defining the Hecke action in a natural way. The drawback of the adèlic complex is its large size, and when finiteness properties are required one uses a Borel-Serre complex instead.

Definition 2.5.

The adélic complex of XUX_{U} with coefficients in MM is the chain complex

Cad(U,M)=Sing(G/K×𝐆(𝔸F))[𝐆(F)×U]M,C^{\operatorname{ad}}_{\bullet}(U,M)=\operatorname{Sing}_{\bullet}\big{(}G_{\infty}/K_{\infty}\times\mathbf{G}(\mathbb{A}^{\infty}_{F})\big{)}\otimes_{\mathbb{Z}[\mathbf{G}(F)\times U]}M,

where Sing(G/K×𝐆(𝔸F))\operatorname{Sing}_{\bullet}\big{(}G_{\infty}/K_{\infty}\times\mathbf{G}(\mathbb{A}^{\infty}_{F})\big{)} denotes the complex of singular chains with \mathbb{Z}-coefficients, viewed as a complex of right [𝐆(F)×U]\mathbb{Z}[\mathbf{G}(F)\times U]-modules.

To define our Borel-Serre complex, we consider the Borel-Serre bordification DDBSD_{\infty}\subset D_{\infty}^{\operatorname{BS}} (see [6, §7.1, Proposition 7.6]) and the principal UpU_{p}-bundle

𝐆(F)\(DBS×𝐆(𝔸F)/𝟏pUp)𝐆(F)\(DBS×𝐆(𝔸F)/UpUp)=XUpUpBS,\mathbf{G}(F)\backslash(D_{\infty}^{\operatorname{BS}}\times\mathbf{G}(\mathbb{A}^{\infty}_{F})/\mathbf{1}_{p}U^{p})\twoheadrightarrow\mathbf{G}(F)\backslash(D_{\infty}^{\operatorname{BS}}\times\mathbf{G}(\mathbb{A}^{\infty}_{F})/U_{p}U^{p})=X^{\operatorname{BS}}_{U_{p}U^{p}},

where 𝐆(𝔸F)\mathbf{G}(\mathbb{A}^{\infty}_{F}) is equipped with the discrete topology.

Fix a finite triangulation of XUpUpBSX^{\operatorname{BS}}_{U_{p}U^{p}}, and consider its associated complex of simplicial chains with \mathbb{Z}-coefficients. Pulling back the triangulation via the map above we obtain – possibly after finite refinement – a KpK_{p}-equivariant triangulation on the ‘infinite pp-level’ space 𝐆(F)\(DBS×𝐆(𝔸F)/𝟏pUp)\mathbf{G}(F)\backslash(D_{\infty}^{\operatorname{BS}}\times\mathbf{G}(\mathbb{A}^{\infty}_{F})/\mathbf{1}_{p}U^{p}). The associated complex of simplicial chains is then a bounded complex of finitely generated and free [Up]\mathbb{Z}[U_{p}]-modules which we denote by CBS(U)C^{\operatorname{BS}}_{\bullet}(U).

Definition 2.6.

The Borel-Serre complex with MM-coefficients is the complex

CBS(U,M)=CBS(U)[Up]M.C^{\operatorname{BS}}_{\bullet}(U,M)=C^{\operatorname{BS}}_{\bullet}(U)\otimes_{\mathbb{Z}[U_{p}]}M.
Proposition 2.7.

The complexes Cad(U,M)C^{\operatorname{ad}}_{\bullet}(U,M) and CBS(U,M)C^{\operatorname{BS}}_{\bullet}(U,M) are chain homotopic as [Up]\mathbb{Z}[U_{p}]-complexes, and

H(Cad(U,M))H(CBS(U,M))H(XU,M).H_{\ast}(C^{\operatorname{ad}}_{\bullet}(U,M))\cong H_{\ast}(C^{\operatorname{BS}}_{\bullet}(U,M))\cong H_{\ast}(X_{U},\mathcal{L}_{M}).

In light of this result, we fix once and for all a chain homotopy equivalence Cad(U,)CBS(U)C^{\operatorname{ad}}_{\bullet}(U,\mathbb{Z})\stackrel{{\scriptstyle\sim}}{{\to}}C^{\operatorname{BS}}_{\bullet}(U), and we define

H(XU,M):=H(XU,M),H_{\ast}(X_{U},M):=H_{\ast}(X_{U},\mathcal{L}_{M}),

2.3 Hecke operators

Let SS be a finite set of finite places of FF containing the set SpS_{p} of places above pp. We will only consider Hecke operators away from pp.

Definition 2.8.

The abstract Hecke algebra is the countably generated free 𝒪\mathcal{O}-algebra

𝕋S=𝒪[TvvS].\mathbb{T}^{S}=\mathcal{O}[T_{v}\mid v\notin S].

The variable TvT_{v} corresponds to the double coset operator given above vv by

[Kv(ϖv001)Kv],[K_{v}\begin{pmatrix}\varpi_{v}&0\\ 0&1\end{pmatrix}K_{v}],

where ϖv\varpi_{v} denotes an arbitrary choice of uniformiser of 𝒪Fv\mathcal{O}_{F_{v}} and Kv=PGL2(𝒪Fv)K_{v}=\operatorname{PGL}_{2}(\mathcal{O}_{F_{v}}). Given an automorphic representation Π\Pi (and a choice of ι:¯p\iota:\overline{\mathbb{Q}}_{p}\cong\mathbb{C}), we have a map

𝕋SEnd(vSΠvKv)\mathbb{T}^{S}\to\operatorname{End}_{\mathbb{C}}(\otimes^{\prime}_{v\notin S}\Pi_{v}^{K_{v}})

with kernel a maximal ideal.

Definition 2.9.

Let Π\Pi be an automorphic representation and suppose SS contains the ramified places of Π\Pi. Let Kv=PGL2(𝒪Fv)K_{v}=\operatorname{PGL}_{2}(\mathcal{O}_{F_{v}}) and

𝔑Π,ι=ker(𝕋SEnd(vSΠvKv)).\mathfrak{N}_{\Pi,\iota}=\ker\big{(}\mathbb{T}^{S}\to\operatorname{End}_{\mathbb{C}}(\otimes^{\prime}_{v\notin S}\Pi_{v}^{K_{v}})\big{)}.

The Π\Pi-part of the homology H(XU,M)H_{\ast}(X_{U},M) is defined as the localisation

H(XU,M)𝔑Π,ι.H_{\ast}(X_{U},M)_{\mathfrak{N}_{\Pi,\iota}}.

The dimensions of the graded pieces are given by the following formula.

Theorem 2.10.

Let 𝔭𝕋S[1/ϖ]\mathfrak{p}\subset\mathbb{T}^{S}[1/\varpi] be the maximal ideal associated to the cuspidal PGL2/F\operatorname{PGL}_{2}/F-representation Π\Pi contributing to homology with coefficients in σ\sigma, and set d=dim(Π)Upd=\dim_{\mathbb{C}}(\Pi^{\infty})^{U^{p}}. Then

dimLHi(XKUp,σ)𝔭={d(l0i) if i[q0,q0+l0],0 otherwise.\dim_{L}H_{i}(X_{KU^{p}},\sigma)_{\mathfrak{p}}=\begin{cases}d\cdot\binom{l_{0}}{i}&\text{ if }i\in[q_{0},q_{0}+l_{0}],\\ 0&\text{ otherwise.}\end{cases}
Proof.

The result follows from the proof of [18, Proposition 4.2]. ∎

We will consider the notion above when MM is an 𝒪[Up]\mathcal{O}[U_{p}]-module which is finitely generated over 𝒪\mathcal{O}. In particular, if s1s\geq 1 and M=𝒪/ϖsM=\mathcal{O}/\varpi^{s} with the trivial group action, we have for any open normal subgroup UpKU_{p}\subset K a homomorphism

𝕋SEndD((𝒪/ϖs)[K/Up])(Cad(UpUp,𝒪/ϖs)),\mathbb{T}^{S}\to\operatorname{End}_{D((\mathcal{O}/\varpi^{s})[K/U_{p}])}(C^{\operatorname{ad}}_{\bullet}(U_{p}U^{p},\mathcal{O}/\varpi^{s})),

where D()D(-) denotes the unbounded derived category. We define 𝕋S(UpUp,𝒪/ϖs)\mathbb{T}^{S}(U_{p}U^{p},\mathcal{O}/\varpi^{s}) as the image of the homomorphism above. Varying ss and UpU_{p}, one obtains a projective system.

Definition 2.11.

The big Hecke algebra is defined as the projective limit

𝕋S(Up)=lims,Up𝕋S(UpUp,𝒪/ϖs).\mathbb{T}^{S}(U^{p})=\varprojlim_{s,U_{p}}\mathbb{T}^{S}(U_{p}U^{p},\mathcal{O}/\varpi^{s}).

The big Hecke algebra is a semi-local profinite 𝒪\mathcal{O}-algebra, such that for any maximal ideal 𝔪𝕋S(Up)\mathfrak{m}\subset\mathbb{T}^{S}(U^{p}) the localisation 𝕋S(Up)𝔪\mathbb{T}^{S}(U^{p})_{\mathfrak{m}} is a local 𝔪\mathfrak{m}-adically complete 𝒪\mathcal{O}-algebra with residue field a finite extension of kk ([16, Lemma 2.1.14]).

3 Commutative and homological algebra

In this section we collect some general lemmas from commutative and homological algebra that will be used in later sections.

3.1 Projective limits

Lemma 3.1.

Let AA be a ring, (Mα)(M_{\alpha}) a projective system of finite length AA-modules and NN a finitely presented AA-module. Then there is a canonical isomorphism

(limMα)ANlimMαAN.(\varprojlim M_{\alpha})\otimes_{A}N\to\varprojlim M_{\alpha}\otimes_{A}N.
Proof.

See e.g. [1, Lemma 2.3.4]. ∎

Lemma 3.2.

([29, Proposition IV.2.7]) Let AA be a topological ring. Then the functor

lim:ProModcpt(A)Modcpt(A)\varprojlim:\operatorname{Pro}-\operatorname{Mod}^{\operatorname{cpt}}(A)\to\operatorname{Mod}^{\operatorname{cpt}}(A)

is exact, where ProModcpt(A)\operatorname{Pro}-\operatorname{Mod}^{\operatorname{cpt}}(A) is the category of projective systems of compact AA-modules.

3.2 Depth and regular sequences

Throughout this section, let (A,𝔪)(A,\mathfrak{m}) be a Noetherian local ring, IAI\subset A a proper ideal and MM a finitely generated AA-module. We recall some basic facts about regular sequences and the depth of modules, following [14, §18].

Definition 3.3.

An MM-regular II-sequence is a sequence a¯=a1,,arI\underline{a}=a_{1},\dots,a_{r}\in I such that for i=1,,ri=1,\dots,r, multiplication by aia_{i} is injective on M/(a1,,ai1)M/(a_{1},\dots,a_{i-1}).

We refer to AA-regular 𝔪\mathfrak{m}-sequences simply as regular sequences.

Definition 3.4.

The II-depth of MM is denoted dpI(M)\operatorname{dp}_{I}(M) and is the supremum of the lengths of MM-regular II-sequences.

When I=𝔪I=\mathfrak{m}, we denote the 𝔪\mathfrak{m}-depth of MM by dpA(M)\operatorname{dp}_{A}(M). It is clear from the definition that

dpIMdpAM.\operatorname{dp}_{I}M\leq\operatorname{dp}_{A}M.

The property of being MM-regular can be formulated in a homological way using Koszul complexes.

Definition 3.5.

Let a¯=a1,,ar\underline{a}=a_{1},\dots,a_{r} be an arbitrary sequence of elements of AA. The Koszul complex of a¯\underline{a} with coefficients in AA is the complex KA(a¯)K_{\bullet}^{A}(\underline{a}) of AA-modules which is non-zero only in degrees [0,r][0,r] where it is given by

KnA(a¯)\displaystyle K_{n}^{A}(\underline{a}) =1j1<<jnrA\displaystyle=\bigoplus_{1\leq j_{1}<\dots<j_{n}\leq r}A
d(b1,,bn)\displaystyle d(b_{1},\dots,b_{n}) =i=1r(1)i1ai(b1,,bi1,bi+1,,bn).\displaystyle=\sum_{i=1}^{r}(-1)^{i-1}a_{i}(b_{1},\dots,b_{i-1},b_{i+1},\dots,b_{n}).

The Koszul complex is exact precisely when the sequence a¯\underline{a} is regular on AA, in which case KnA(a¯)K_{n}^{A}(\underline{a}) is a free resolution of A/(a¯)A/(\underline{a}). More generally, one has the following homological characterisation of regular sequences.

Proposition 3.6.

Let a¯\underline{a} be an arbitrary sequence in AA. Then a¯\underline{a} is MM-regular if and only if, for every i1i\geq 1,

Hi(KA(a¯)AM)=0.H_{i}(K_{\bullet}^{A}(\underline{a})\otimes_{A}M)=0.

Note that if a¯\underline{a} is AA-regular, then H(KA(a¯)AM)Tor(A/(a¯),M)H_{\ast}(K_{\bullet}^{A}(\underline{a})\otimes_{A}M)\cong\operatorname{Tor}_{\ast}(A/(\underline{a}),M). Restated slightly, we have:

Corollary 3.7.

Suppose IAI\subset A can be generated by an AA-regular sequence. Then II can be generated by an MM-regular sequence if and only if

MA𝐋A/IM/IM,M\otimes^{\mathbf{L}}_{A}A/I\cong M/IM,

where we interpret M/IMM/IM as a complex concentrated in degree 0.

Lemma 3.8.

Let M,NM,N be finitely generated modules over AA. Suppose IAnnNI\subset\operatorname{Ann}N is an ideal generated by an AA-regular and MM-regular sequence. Then we have a natural isomorphism of complexes

MA𝐋NM/IMA/I𝐋NM\otimes^{\mathbf{L}}_{A}N\cong M/IM\otimes^{\mathbf{L}}_{A/I}N
Proof.

Since NN/INN\cong N/IN, we have

MA𝐋NMA𝐋A/IA/I𝐋NM/IMA/I𝐋N,\begin{split}M\otimes^{\mathbf{L}}_{A}N&\cong M\otimes^{\mathbf{L}}_{A}A/I\otimes^{\mathbf{L}}_{A/I}N\\ &\cong M/IM\otimes^{\mathbf{L}}_{A/I}N,\end{split}

using the assumptions and the homological criterion. ∎

The depth can also be defined in homological terms using the following result.

Proposition 3.9.

([14, Proposition 18.4])

dpI(M)=min{iExtAi(A/I,M)0}.\operatorname{dp}_{I}(M)=\min\{i\in\mathbb{N}\mid\operatorname{Ext}^{i}_{A}(A/I,M)\neq 0\}.
Lemma 3.10.

Suppose IAI\subset A can be generated by a regular sequence. Then

dpIM+dpIM=dpIA,\operatorname{dp}_{I}M+\operatorname{dp}^{\ast}_{I}M=\operatorname{dp}_{I}A,

where dpIM=max{iToriA(A/I,M)0}\operatorname{dp}^{\ast}_{I}M=\max\{i\mid\operatorname{Tor}_{i}^{A}(A/I,M)\neq 0\}.

Proof.

Let r=dpIAr=\operatorname{dp}_{I}A and a¯=a1,,ar\underline{a}=a_{1},\dots,a_{r} a regular sequence generating II. By [14, Proposition 17.15], there is a natural isomorphism of chain complexes

KA(a¯)AMHomA(KrA(a¯),M)K_{\bullet}^{A}(\underline{a})\otimes_{A}M\cong\operatorname{Hom}_{A}(K_{r-\bullet}^{A}(\underline{a}),M)

and hence for every ii an isomorphism

ToriA(A/I,M)ExtAri(A/I,M).\operatorname{Tor}_{i}^{A}(A/I,M)\cong\operatorname{Ext}^{r-i}_{A}(A/I,M).

The statement follows. ∎

Lemma 3.11.

Let A^\widehat{A} and M^\widehat{M} denote the 𝔪\mathfrak{m}-adic completions of AA and MM. Then

dpAM=dpA^M^.\operatorname{dp}_{A}M=\operatorname{dp}_{\widehat{A}}\widehat{M}.
Proof.

Since AA is Noetherian, A^\widehat{A} is a faithfully flat AA-module [25, Theorem 1.3.16(b)] and the maximal ideal of A^\widehat{A} is 𝔪A^\mathfrak{m}\widehat{A}. By flat base change, we have an isomorphism of AA-modules (see [39, Proposition 3.3.10])

ExtA^i(A^/𝔪A^,M^)ExtAi(A/𝔪,M)AA^.\operatorname{Ext}_{\widehat{A}}^{i}(\widehat{A}/\mathfrak{m}\widehat{A},\widehat{M})\cong\operatorname{Ext}_{A}^{i}(A/\mathfrak{m},M)\otimes_{A}\widehat{A}.

Since A^\widehat{A} is faithfully flat, the right hand side equals 0 if and only if ExtAi(A/𝔪,M)\operatorname{Ext}^{i}_{A}(A/\mathfrak{m},M) does, and the statement thus follows from Proposition 3.9. ∎

3.3 Cohen-Macaulay modules

We keep the notation from the previous section. One always has the inequalities

dpAMdimAMdimA.\operatorname{dp}_{A}M\leq\dim_{A}M\leq\dim A.
Definition 3.12.

We say MM is Cohen-Macaulay if dpAM=dimAM\operatorname{dp}_{A}M=\dim_{A}M, and maximal Cohen-Macaulay if dpAM=dimA\operatorname{dp}_{A}M=\dim A.

Lemma 3.13.

([36, Tag 00NT]) Suppose AA is regular and that MM is maximal Cohen-Macaulay over AA. Then MM is free.

Lemma 3.14.

Let AA be a complete local Noetherian 𝒪\mathcal{O}-algebra and MM a finitely generated maximal Cohen-Macaulay AA-module. Suppose 𝔭SpecA[1/ϖ]\mathfrak{p}\subset\operatorname{Spec}A[1/\varpi] is a regular closed point. Then M^𝔭\widehat{M}_{\mathfrak{p}} is free over A^𝔭\widehat{A}_{\mathfrak{p}}.

Proof.

We prove that M^𝔭\widehat{M}_{\mathfrak{p}} is maximal Cohen-Macaulay over A^𝔭\widehat{A}_{\mathfrak{p}} and use Proposition 3.13. It suffices to prove dimA𝔭dpA𝔭M𝔭\dim A_{\mathfrak{p}}\leq\operatorname{dp}_{A_{\mathfrak{p}}}M_{\mathfrak{p}}. By the Cohen structure theorem, AA is a quotient of a ring of formal power series over 𝒪\mathcal{O} and hence the quotient A/𝔭A/\mathfrak{p} is a finite extension of 𝒪\mathcal{O} and moreover,

dimA𝔭dimA1=dpAM1,\dim A_{\mathfrak{p}}\leq\dim A-1=\operatorname{dp}_{A}M-1,

where we in the second equality have used the maximal Cohen-Macaulay property of MM. Moving on, since 𝔪=𝔭+(ϖ)\mathfrak{m}=\mathfrak{p}+(\varpi), an application of [14, Lemma 18.3] gives the inequality

dpAM1=dp𝔭+(ϖ)M1dp𝔭M.\operatorname{dp}_{A}M-1=\operatorname{dp}_{\mathfrak{p}+(\varpi)}M-1\leq\operatorname{dp}_{\mathfrak{p}}M.

We have

dp𝔭MdpA𝔭M𝔭\operatorname{dp}_{\mathfrak{p}}M\leq\operatorname{dp}_{A_{\mathfrak{p}}}M_{\mathfrak{p}}

since the A𝔭A_{\mathfrak{p}}-depth can be calculated in terms of the group

ExtA𝔭(A𝔭/𝔭,M𝔭)ExtA(A/𝔭,M)AA𝔭.\operatorname{Ext}^{\ast}_{A_{\mathfrak{p}}}(A_{\mathfrak{p}}/\mathfrak{p},M_{\mathfrak{p}})\cong\operatorname{Ext}^{\ast}_{A}(A/\mathfrak{p},M)\otimes_{A}A_{\mathfrak{p}}.

Thus we have established dimA𝔭dpA𝔭M𝔭\dim A_{\mathfrak{p}}\leq\operatorname{dp}_{A_{\mathfrak{p}}}M_{\mathfrak{p}}, and the theorem follows. ∎

4 Galois cohomology

In this section, we discuss Galois cohomology and Bloch-Kato Selmer groups. For a complete reference, see [28], and for an introduction see [4].

Suppose ρ:GF,SGL2(L)\rho:G_{F,S}\to\operatorname{GL}_{2}(L) where SSpS\supseteq S_{p} is a finite set of finite places, and that ρ\rho is de Rham at all places above pp. First, we introduce the local Bloch-Kato Selmer groups. Let vSv\in S. If vpv\mid p, then

Hf1(GFv,ad0ρ)=ker(H1(GFv,ad0ρ)H1(GFv,ad0ρpBcris)),H^{1}_{f}(G_{F_{v}},\operatorname{ad}^{0}\rho)=\ker\big{(}H^{1}(G_{F_{v}},\operatorname{ad}^{0}\rho)\to H^{1}(G_{F_{v}},\operatorname{ad}^{0}\rho\otimes_{\mathbb{Q}_{p}}B_{\operatorname{cris}})\big{)},

where the map on the right is the natural one induced by the tensor product and BcrisB_{\operatorname{cris}} is Fontaine’s crystalline period ring. If vpv\nmid p, let IFvGFvI_{F_{v}}\subset G_{F_{v}} denote the inertia subgroup and set

Hf1(GFv,ad0ρ)=ker(H1(GF,S,ad0ρ)H1(IFv,ad0ρ)).H^{1}_{f}(G_{F_{v}},\operatorname{ad}^{0}\rho)=\ker\big{(}H^{1}(G_{F,S},\operatorname{ad}^{0}\rho)\to H^{1}(I_{F_{v}},\operatorname{ad}^{0}\rho)\big{)}.

where the map on the right is the one induced by the inclusion IFvGFvI_{F_{v}}\subset G_{F_{v}}.

For every place vv of FF there is a restriction map

resv:H1(GF,S,ad0ρ)H1(GFv,ad0ρ),\operatorname{res}_{v}:H^{1}(G_{F,S},\operatorname{ad}^{0}\rho)\to H^{1}(G_{F_{v}},\operatorname{ad}^{0}\rho),

and we define the global Bloch-Kato Selmer group as

Hf1(GF,S,ad0ρ)={cH1(GF,S,ad0ρ)vS:resv(c)Hf1(GFv,ad0ρ)},H^{1}_{f}(G_{F,S},\operatorname{ad}^{0}\rho)=\{c\in H^{1}(G_{F,S},\operatorname{ad}^{0}\rho)\mid\forall v\in S:\operatorname{res}_{v}(c)\in H^{1}_{f}(G_{F_{v}},\operatorname{ad}^{0}\rho)\},

or equivalently

ker(H1(GF,S,ad0ρ)vSH1(GFv,ad0ρ)Hf1(GFv,ad0ρ)).\ker\bigg{(}H^{1}(G_{F,S},\operatorname{ad}^{0}\rho)\to\prod_{v\in S}\frac{H^{1}(G_{F_{v}},\operatorname{ad}^{0}\rho)}{H^{1}_{f}(G_{F_{v}},\operatorname{ad}^{0}\rho)}\bigg{)}.

We use lowercase hh to denote dimension, e.g.

hf1(GF,S,V)=dimLHf1(GF,S,V).h^{1}_{f}(G_{F,S},V)=\dim_{L}H^{1}_{f}(G_{F,S},V).

If vSSpv\in S\setminus S_{p} and the Weil-Deligne representation WD(ρv)\operatorname{WD}(\rho_{v}) associated to ρv\rho_{v} is generic we have H0(GFv,ad0ρ(1))=0H^{0}(G_{F_{v}},\operatorname{ad}^{0}\rho(1))=0 by [2, Lemma 1.1.5]. Thus, by local Tate duality and the formula for the Euler-Poincaré characteristic,

h0(GFv,ad0ρ)=h1(GFv,ad0ρ).h^{0}(G_{F_{v}},\operatorname{ad}^{0}\rho)=h^{1}(G_{F_{v}},\operatorname{ad}^{0}\rho).
Proposition 4.1.

Suppose ρ:GF,SGL2(L)\rho:G_{F,S}\to\operatorname{GL}_{2}(L) is irreducible, where FF is a totally complex field in which pp is totally split, that ρ\rho is de Rham with distinct Hodge-Tate weights at all places above pp, and moreover that WD(ρv)\operatorname{WD}(\rho_{v}) is generic for all vSSpv\in S\setminus S_{p}. Then if l0=[F:]/2l_{0}=[F:\mathbb{Q}]/2 we have

hf1(GF,S,ad0ρ)=hf1(GF,S,ad0ρ(1))l0.h^{1}_{f}(G_{F,S},\operatorname{ad}^{0}\rho)=h^{1}_{f}(G_{F,S},\operatorname{ad}^{0}\rho(1))-l_{0}.
Proof.

By the Greenberg-Wiles formula, and the irreducibility of ρ\rho, we have

hf1(GF,S,ad0ρ)hf1(GF,S,ad0ρ(1))=vS(hf1(GFv,ad0ρ)h0(GFv,ad0ρ))vh0(GFv,ad0ρ).\begin{split}&h^{1}_{f}(G_{F,S},\operatorname{ad}^{0}\rho)-h^{1}_{f}(G_{F,S},\operatorname{ad}^{0}\rho(1))=\\ \sum_{v\in S}\big{(}&h^{1}_{f}(G_{F_{v}},\operatorname{ad}^{0}\rho)-h^{0}(G_{F_{v}},\operatorname{ad}^{0}\rho)\big{)}-\sum_{v\mid\infty}h^{0}(G_{F_{v}},\operatorname{ad}^{0}\rho).\\ \end{split}

Since WD(ρv)\operatorname{WD}(\rho_{v}) is generic for every vSSpv\in S\setminus S_{p}, the corresponding terms in the sum vanish. Furthermore, since FF is totally complex, the expression simplifies to

vp(hf1(GFv,ad0ρ)h0(GFv,ad0ρ))vdimad0ρ.\sum_{v\mid p}\big{(}h^{1}_{f}(G_{F_{v}},\operatorname{ad}^{0}\rho)-h^{0}(G_{F_{v}},\operatorname{ad}^{0}\rho)\big{)}-\sum_{v\mid\infty}\dim\operatorname{ad}^{0}\rho.

Now, pp is totally split in FF and the sum over vpv\mid p counts the total multiplicities of the negative Hodge-Tate weights of the ad0ρ|GFv\operatorname{ad}^{0}\rho|_{G_{F_{v}}}, (see [5, Corollary 3.8.4]). Since the Hodge-Tate weights of ρv\rho_{v} are assumed to be distinct, the total expression therefore equals [F:]/2=l0-[F:\mathbb{Q}]/2=-l_{0}. ∎

Theorem 4.2.

Suppose ρ:GF,SGL2(L)\rho:G_{F,S}\to\operatorname{GL}_{2}(L) is de Rham at all places above pp. Then there is an exact sequence of LL-vector spaces

0\displaystyle 0 Hf1(GF,S,ad0ρ)H1(GF,S,ad0ρ)vSH1(GFv,ad0ρ)Hf1(GFv,ad0ρ)\displaystyle\to H^{1}_{f}\big{(}G_{F,S},\operatorname{ad}^{0}\rho\big{)}\to H^{1}\big{(}G_{F,S},\operatorname{ad}^{0}\rho\big{)}\to\prod_{v\in S}\frac{H^{1}\big{(}G_{F_{v}},\operatorname{ad}^{0}\rho\big{)}}{H^{1}_{f}\big{(}G_{F_{v}},\operatorname{ad}^{0}\rho\big{)}}
Hf1(GF,S,ad0ρ(1))H2(GF,S,ad0ρ).\displaystyle\to H^{1}_{f}\big{(}G_{F,S},\operatorname{ad}^{0}\rho(1)\big{)}^{\vee}\to H^{2}(G_{F,S},\operatorname{ad}^{0}\rho).

Moreover, if the Weil-Deligne representation WD(ρv)\operatorname{WD}(\rho_{v}) associated to ρv\rho_{v} is generic for every vSSpv\in S\setminus S_{p}, the corresponding factors in the third term are 0.

Proof.

On [38, p.119], the analogous exact sequence with finite coefficients is derived from the Poitou-Tate sequence. We argue in the same way using the Poitou-Tate sequence for cohomology in characteristic 0, which is readily obtained from [38, Proposition 10] by identifying, for any LL-vector space VV with a continuous action of GF,SG_{F,S},

H1(GF,S,V)(limsH1(GF,S,Θ/ϖs))𝒪L,H^{1}(G_{F,S},V)\cong(\varprojlim_{s}H^{1}(G_{F,S},\Theta/\varpi^{s}))\otimes_{\mathcal{O}}L,

where Θ\Theta is an arbitrary choice of GF,SG_{F,S}-stable lattice in VV.

To prove the second part, recall that the genericity assumption at a place vSSpv\in S\setminus S_{p} implies the equality

h0(GFv,ad0ρ)=h1(GFv,ad0ρ).h^{0}(G_{F_{v}},\operatorname{ad}^{0}\rho)=h^{1}(G_{F_{v}},\operatorname{ad}^{0}\rho).

Thus, it suffices to prove that the left-hand side equals hur1(GFv,ad0ρ)h^{1}_{\operatorname{ur}}(G_{F_{v}},\operatorname{ad}^{0}\rho). We follow [4, Proposition 2.3(a)]. Let V=ad0ρV=\operatorname{ad}^{0}\rho. The inflation-restriction sequence yields an isomorphism

H1(GFv/Iv,VIv)Hf1(GFv,V),H^{1}(G_{F_{v}}/I_{v},V^{I_{v}})\cong H^{1}_{f}(G_{F_{v}},V),

and since GFv/Iv^G_{F_{v}}/I_{v}\cong\widehat{\mathbb{Z}} has cohomological dimension 11 ([17, Proposition 6.1.9]), the Euler-Poincaré characteristic formula implies

h1(GFv/Iv,VIv)=h0(GFv/Iv,VIv)=h0(GFv,V).h^{1}(G_{F_{v}}/I_{v},V^{I_{v}})=h^{0}(G_{F_{v}}/I_{v},V^{I_{v}})=h^{0}(G_{F_{v}},V).

Hence hf1(GFv,V)=h0(GFv,ad0ρ)h^{1}_{f}(G_{F_{v}},V)=h^{0}(G_{F_{v}},\operatorname{ad}^{0}\rho), as claimed. ∎

5 Galois deformation theory

In this section, we discuss the deformation theory of Galois representations, giving definitions and citing results from the literature that we shall need in the sequel.

5.1 Deformations of Galois representations

Let 𝔄𝒪\mathfrak{A}_{\mathcal{O}} be the category of discrete Artinian local 𝒪\mathcal{O}-algebras with residue field kk, and 𝔄^𝒪\widehat{\mathfrak{A}}_{\mathcal{O}} the category of complete Noetherian local 𝒪\mathcal{O}-algebras with residue field kk, with morphisms the continuous 𝒪\mathcal{O}-algebra homomorphisms. The latter is equivalent to a full subcategory of the category Pro𝔄𝒪\operatorname{Pro}-\mathfrak{A}_{\mathcal{O}}, and contains 𝔄𝒪\mathfrak{A}_{\mathcal{O}} as a full subcategory.

Let SS be a finite set of finite places of FF and let Γ\Gamma be one of the groups GF,SG_{F,S} and GFvG_{F_{v}}. Suppose ρ¯:ΓGL2(k)\bar{\rho}:\Gamma\to\operatorname{GL}_{2}(k) is a continuous representation and fix a continuous character χ:Γ𝒪×\chi:\Gamma\to\mathcal{O}^{\times} such that χdetρ¯modϖ.\chi\equiv\det\bar{\rho}\ \operatorname{mod}\varpi. For any A𝔄𝒪A\in\mathfrak{A}_{\mathcal{O}}, let redA\operatorname{red_{A}} be the reduction map GL2(A)GL2(k)\operatorname{GL}_{2}(A)\to\operatorname{GL}_{2}(k), and χA\chi_{A} the composition Γχ𝒪×A×\Gamma\stackrel{{\scriptstyle\chi}}{{\to}}\mathcal{O}^{\times}\to A^{\times}.

Definition 5.1.

Keeping the above notation, we make the following definitions.

  1. (i)

    A framed deformation of ρ¯\bar{\rho} to AA with determinant χ\chi is a continuous homomorphism ρA:ΓGL2(A)\rho_{A}:\Gamma\to\operatorname{GL}_{2}(A) fitting into a diagram

    GL2(A){{\operatorname{GL}_{2}(A)}}Γ{{\Gamma}}GL2(k){{\operatorname{GL}_{2}(k)}}redA\scriptstyle{\operatorname{red}_{A}}ρ¯\scriptstyle{\bar{\rho}}

    and satisfying detρA=χA\det\rho_{A}=\chi_{A}.

  2. (ii)

    Two framed deformations ρA,ρA\rho_{A},\rho_{A}^{\prime} are strictly equivalent if there is an element γker(redA)\gamma\in\ker(\operatorname{red}_{A}) such that

    γρAγ1=ρA.\gamma\rho_{A}\gamma^{-1}=\rho_{A}^{\prime}.
  3. (iii)

    A deformation of ρ¯\bar{\rho} to AA is an equivalence class of framed deformations to AA under the relation of strict equivalence.

All in all, we obtain functors

Dρ¯χ,Dρ¯vχ,Dρ¯vχ,:𝔄𝒪𝐒𝐞𝐭,D_{\bar{\rho}}^{\chi},D_{\bar{\rho}_{v}}^{\chi},D^{\chi,\square}_{\bar{\rho}_{v}}:\mathfrak{A}_{\mathcal{O}}\to\mathbf{Set},

where for example Dρ¯χ(A)D_{\bar{\rho}}^{\chi}(A) is the set of deformations of ρ¯\bar{\rho} to AA of determinant χ\chi. The superscript \square means ‘framed’. These functors extend in a canonical way to A^𝒪\widehat{A}_{\mathcal{O}}, and from now on we consider all functors of deformations as defined on the category A^𝒪\widehat{A}_{\mathcal{O}}.

Deformations of a continuous representation ρ:ΓGL2(L)\rho:\Gamma\to\operatorname{GL}_{2}(L) in characteristic 0 is defined in the same way. Let 𝔄^L\widehat{\mathfrak{A}}_{L} be the category of complete Noetherian local LL-algebras with residue field LL. As before, we define functors

Dρχ,Dρvχ,Dρvχ,:𝔄^L𝐒𝐞𝐭.D_{\rho}^{\chi},D_{\rho_{v}}^{\chi},D^{\chi,\square}_{\rho_{v}}:\widehat{\mathfrak{A}}_{L}\to\mathbf{Set}.

5.2 Deformations of pseudorepresentations

The notion of (non-framed) deformations is sometimes too restrictive in the sense that the corresponding functors can fail to be representable. To circumvent this issue, one introduces pseudorepresentations, which are roughly speaking functions that behave like traces of representations. Pseudodeformations can be defined in great generality, but for our purposes the following narrow definition will suffice.

Definition 5.2.

Let AA be a topological ring such that 2A×2\in A^{\times}. An AA-valued pseudorepresentation (of dimension 22) of GFvG_{F_{v}} is a continuous function τv:GFvA\tau_{v}:G_{F_{v}}\to A such that:

  • (i)

    τ(id)=2\tau(\operatorname{id})=2.

  • (ii)

    For every g1,g2G,τ(g1g2)=τ(g2g1).g_{1},g_{2}\in G,\ \tau(g_{1}g_{2})=\tau(g_{2}g_{1}).

  • (iii)

    For every g1,g2,g3Gg_{1},g_{2},g_{3}\in G,

    τ(g1g2g3)=τ(g1g2)τ(g3)+τ(g1g3)τ(g2)+τ(g2g3)τ(g1)τ(g1g3g2)τ(g1)τ(g2)τ(g3).\begin{split}\tau(g_{1}g_{2}g_{3})&=\tau(g_{1}g_{2})\tau(g_{3})+\tau(g_{1}g_{3})\tau(g_{2})+\tau(g_{2}g_{3})\tau(g_{1})\\ &-\tau(g_{1}g_{3}g_{2})-\tau(g_{1})\tau(g_{2})\tau(g_{3}).\end{split}
Definition 5.3.

Let τv\tau_{v} be an AA-valued pseudorepresentation of GFvG_{F_{v}}. The determinant of τv\tau_{v} is the function detτv:GFvA\det\tau_{v}:G_{F_{v}}\to A defined by

detτv(g)=12(τv(g)2τv(g2))\det\tau_{v}(g)=\frac{1}{2}\big{(}\tau_{v}(g)^{2}-\tau_{v}(g^{2})\big{)}

We will consider AA-valued pseudorepresentations where A𝔄^𝒪A\in\widehat{\mathfrak{A}}_{\mathcal{O}} is equipped with its profinite topology. Just as for representations, if τ¯v:GFvk\overline{\tau}_{v}:G_{F_{v}}\to k is a pseudorepresentation we define a functor

Dτ¯vps,χv:𝔄^𝒪𝐒𝐞𝐭D_{\overline{\tau}_{v}}^{\operatorname{ps},\chi_{v}}:\widehat{\mathfrak{A}}_{\mathcal{O}}\to\mathbf{Set}

taking an artinian 𝒪\mathcal{O}-algebra AA to the set of AA-valued pseudorepresentations with determinant χv\chi_{v} lifting τ¯v\overline{\tau}_{v}.

5.3 Deformation problems

In this brief section, we recall the definition of a deformation problem, following [22, §4].

Definition 5.4.

A local deformation problem for ρ¯v\bar{\rho}_{v} is a representable closed subfunctor DvDρ¯vχ,D_{v}\subseteq D^{\chi,\square}_{\bar{\rho}_{v}} such that Dv(A)D_{v}(A) is stable under conjugation by elements of kerredA\ker\operatorname{red}_{A}.

Definition 5.5.

A global deformation problem is a tuple 𝒮=(ρ¯,χ,S,{Dv}vS)\mathcal{S}=(\bar{\rho},\chi,S,\{D_{v}\}_{v\in S}) where

  • ρ¯:GFGL2(k)\bar{\rho}:G_{F}\to\operatorname{GL}_{2}(k) is an absolutely irreducible representation,

  • χ:GF𝒪×\chi:G_{F}\to\mathcal{O}^{\times} is a continuous character such that detρ¯=χ¯\det\bar{\rho}=\overline{\chi},

  • SS is a finite set of finite places of FF such that ρ¯\bar{\rho} and χ\chi are unramified outside SS (i.e. factor through GFGF,SG_{F}\twoheadrightarrow G_{F,S}),

  • DvD_{v} is a local deformation problem for every vSv\in S.

Definition 5.6.

Let 𝒮=(ρ¯,χ,S,{Dv}vS)\mathcal{S}=(\bar{\rho},\chi,S,\{D_{v}\}_{v\in S}) be a global deformation problem. A deformation [ρA]Dρ¯χ(A)[\rho_{A}]\in D_{\bar{\rho}}^{\chi}(A) is of type 𝒮\mathcal{S} if ρA\rho_{A} is unramified outside SS, satisfies detρA=χ\det\rho_{A}=\chi and for every vSv\in S,

ρA|GFvDv(A).\rho_{A}|_{G_{F_{v}}}\in D_{v}(A).

Implicit in this definition is the fact that the conditions are independent of the choice of representative ρA\rho_{A} of the strict equivalence class [ρA][\rho_{A}]. We define a functor

D𝒮:𝔄^𝒪𝐒𝐞𝐭,D_{\mathcal{S}}:\widehat{\mathfrak{A}}_{\mathcal{O}}\to\mathbf{Set},

by letting D𝒮(A)Dρ¯χ(A)D_{\mathcal{S}}(A)\subseteq D_{\bar{\rho}}^{\chi}(A) be the set of deformations of type 𝒮\mathcal{S}.

5.4 Universal deformation rings

When deformation functors are representable, many questions about deformations translate to questions about the representing objects, which are complete Noetherian 𝒪\mathcal{O}- or LL-algebras. As mentioned earlier, one reason to consider pseudorepresentations is that the corresponding functors are always representable. The same is true for framed deformations, as can be proved using Schlessinger’s criterion as in [27].

Theorem 5.7.

For any place vv, the functors Dρ¯vχ,D^{\chi,\square}_{\bar{\rho}_{v}} and Dτ¯vps,χvD_{\overline{\tau}_{v}}^{\operatorname{ps},\chi_{v}} are representable. We denote the representing objects by RvR_{v}^{\square} and RvpsR^{\operatorname{ps}}_{v}, respectively.

In contrast, functors of deformations are not representable in general, but the following class of representations have representable functors of deformations.

Definition 5.8.

A representation is Schur if it has only scalar endomorphisms.

Absolutely irreducible representations are always Schur. Recall that the definition of a global deformation problem includes the hypothesis that the residual representation is absolutely irreducible. The following results are deduced in the same way as Theorem 5.7.

Theorem 5.9.

Suppose ρ\rho (resp. ρv\rho_{v}, ρ¯,\bar{\rho}, ρ¯v\bar{\rho}_{v}) is Schur. Then DρχD_{\rho}^{\chi} (resp. Dρvχ,D_{\rho_{v}}^{\chi}, Dρ¯χD_{\bar{\rho}}^{\chi}, Dρ¯vχD_{\bar{\rho}_{v}}^{\chi}) is represented by an object RρR_{\rho} (resp. RρvR_{\rho_{v}}, RR, RvR_{v}) in 𝔄^𝒪\widehat{\mathfrak{A}}_{\mathcal{O}}.

Theorem 5.10.

Let 𝒮\mathcal{S} be a global deformation problem. Then D𝒮D_{\mathcal{S}} is represented by an object R𝒮R_{\mathcal{S}} in 𝔄^𝒪\widehat{\mathfrak{A}}_{\mathcal{O}}.

There are natural transformations

Dρvχ,DρvχtrDτ¯vps,χv,D^{\chi,\square}_{\rho_{v}}\to D_{\rho_{v}}^{\chi}\stackrel{{\scriptstyle\operatorname{tr}}}{{\to}}D_{\overline{\tau}_{v}}^{\operatorname{ps},\chi_{v}},

and thus RvR_{v}^{\square} is a RvpsR^{\operatorname{ps}}_{v}-algebra. Moreover, if RvR_{v} exists then we have homomorphisms

RvpsRvRv,R^{\operatorname{ps}}_{v}\to R_{v}\to R_{v}^{\square},

and RvR_{v}^{\square} is formally smooth of relative dimension 33 over RvR_{v}.

Suppose 𝒮\mathcal{S} is a deformation problem, with local conditions DvD_{v} represented by R¯v\overline{R}_{v}. By definition, R¯v\overline{R}_{v} is a quotient of RvR_{v}^{\square}. We define

R𝒮loc\displaystyle R_{\mathcal{S}}^{\operatorname{loc}} =^vSR¯v,\displaystyle=\widehat{\bigotimes}_{v\in S}\overline{R}_{v},
Rpps\displaystyle R^{\operatorname{ps}}_{p} =^vpRvps.\displaystyle=\widehat{\bigotimes}_{v\mid p}R^{\operatorname{ps}}_{v}.

These rings can also be interpreted as representing objects for certain functors, as the first item of the following lemma shows.

Lemma 5.11.

Let TT be a finite set, and suppose we have representable functors Dv:𝔄^𝒪𝐒𝐞𝐭D_{v}:\widehat{\mathfrak{A}}_{\mathcal{O}}\to\mathbf{Set} for every vTv\in T. Denote by RvR_{v} the representing object of DvD_{v}, and consider the functor DT=vTDv:𝔄^𝒪𝐒𝐞𝐭D_{T}=\prod_{v\in T}D_{v}:\widehat{\mathfrak{A}}_{\mathcal{O}}\to\mathbf{Set}. Then we have the following:

  • (i)

    DTD_{T} is represented by RT=^vTRvR_{T}=\widehat{\bigotimes}_{v\in T}R_{v}.

  • (ii)

    If 𝔭TRT[1/ϖ]\mathfrak{p}_{T}\subset R_{T}[1/\varpi] is a closed point generated by the joint image of closed points 𝔭vRv[1/ϖ]\mathfrak{p}_{v}\subset R_{v}[1/\varpi], we have a canonical isomorphism of LL-algebras

    (RT)𝔭T^vT,L(Rv)𝔭v.(R_{T})_{\mathfrak{p}_{T}}^{\wedge}\cong\widehat{\bigotimes}_{v\in T,L}(R_{v})_{\mathfrak{p}_{v}}^{\wedge}.
Proof.

(i) Let 𝔪v\mathfrak{m}_{v} be the maximal ideal of RvR_{v}, and A=(Ai)𝔄^𝒪A=(A_{i})\in\widehat{\mathfrak{A}}_{\mathcal{O}}. We have

(vTDv)(A)vTlimiHom𝒪(Rv,Ai)=vTlimilimr1Hom𝒪(Rv/𝔪vr,Ai)=limilimr1vTHom𝒪(Rv/𝔪vr,Ai)=limilimr1Hom𝒪(vTRv/𝔪vr,A)=Hom𝒪(^vTRv,A).\begin{split}(\prod_{v\in T}D_{v})(A)&\cong\prod_{v\in T}\varprojlim_{i}\operatorname{Hom}_{\mathcal{O}}(R_{v},A_{i})\\ &=\prod_{v\in T}\varprojlim_{i}\varinjlim_{r\geq 1}\operatorname{Hom}_{\mathcal{O}}(R_{v}/\mathfrak{m}_{v}^{r},A_{i})\\ &=\varprojlim_{i}\varinjlim_{r\geq 1}\prod_{v\in T}\operatorname{Hom}_{\mathcal{O}}(R_{v}/\mathfrak{m}_{v}^{r},A_{i})\\ &=\varprojlim_{i}\varinjlim_{r\geq 1}\operatorname{Hom}_{\mathcal{O}}(\bigotimes_{v\in T}R_{v}/\mathfrak{m}_{v}^{r},A)\\ &=\operatorname{Hom}_{\mathcal{O}}(\widehat{\bigotimes}_{v\in T}R_{v},A).\end{split}

Here, we use that AA is artinian which implies that lim\varprojlim commutes with Hom\operatorname{Hom} in the above way, and moreover that lim\varprojlim and lim\varinjlim commute with finite products.

(ii) Following [24, §2.3], the ring (RT)𝔭T(R_{T})_{\mathfrak{p}_{T}}^{\wedge} represents a functor

D(𝔭T):𝔄^L𝐒𝐞𝐭,D_{(\mathfrak{p}_{T})}:\widehat{\mathfrak{A}}_{L}\to\mathbf{Set},

constructed as a filtered colimit. Since filtered colimits commute with finite limits, it is not hard to see that

D(𝔭T)vpD(𝔭v),D_{(\mathfrak{p}_{T})}\cong\prod_{v\mid p}D_{(\mathfrak{p}_{v})},

and now the argument from (i) carried out in the category 𝔄^L\widehat{\mathfrak{A}}_{L} proves (ii). ∎

Theorem 5.12.

[2, Proposition 1.2.2] Let vSSpv\in S\setminus S_{p} and 𝔭vRv[1/ϖ]\mathfrak{p}_{v}\subset R_{v}^{\square}[1/\varpi] be the point corresponding to a representation ρv:GFvGL2(L)\rho_{v}:G_{F_{v}}\to\operatorname{GL}_{2}(L). Then the Weil-Deligne representation WD(ρv)\operatorname{WD}(\rho_{v}) is generic if and only if 𝔭v\mathfrak{p}_{v} is a regular point of SpecRv\operatorname{Spec}R_{v}^{\square}.

The dimensions of local deformation rings have been computed by Shotton [35]. Note that we are considering fixed-determinant deformations.

Theorem 5.13.

[35, Theorem 2.5] Let vSSpv\in S\setminus S_{p}. Then the local framed (fixed-determinant) deformation ring RvR_{v}^{\square} is an equidimensional reduced complete intersection ring, flat of relative dimension 33 over 𝒪\mathcal{O}.

Having discussed the local framed deformation rings at places of SSpS\setminus S_{p}, we move on to the subtler story of the places above pp. At these places, our local deformation problems will encode conditions from pp-adic Hodge theory. We denote the cyclotomic character by ε\varepsilon.

Definition 5.14.

Let vSpv\in S_{p}. A vv-adic Hodge type is a triple (𝐰v,τv,χv)(\mathbf{w}_{v},\tau_{v},\chi_{v}) where

  • 𝐰v=(av,bv)2\mathbf{w}_{v}=(a_{v},b_{v})\in\mathbb{Z}^{2} such that b>ab>a.

  • τv:IFvGL2(L)\tau_{v}:I_{F_{v}}\to\operatorname{GL}_{2}(L) is a representation with open kernel.

  • χv:GFv𝒪×\chi_{v}:G_{F_{v}}\to\mathcal{O}^{\times} is a continuous character such that χv|IFv=εav+bvdetτ\chi_{v}|_{I_{F_{v}}}=\varepsilon^{a_{v}+b_{v}}\det\tau.

Definition 5.15.

Let (𝐰v,τv,χv){(\mathbf{w}_{v},\tau_{v},\chi_{v})} be a vv-adic Hodge type. A continuous representation

ρv:GFvGL2(L)\rho_{v}:G_{F_{v}}\to\operatorname{GL}_{2}(L)

is of type (𝐰v,τv,χv){(\mathbf{w}_{v},\tau_{v},\chi_{v})} if it is potentially semi-stable with Hodge-Tate weights equal to 𝐰v\mathbf{w}_{v}, determinant equal to χv\chi_{v} and if WD(ρv)|IFvτv\operatorname{WD}(\rho_{v})|_{I_{F_{v}}}\cong\tau_{v}, where WD(ρv)\operatorname{WD}(\rho_{v}) is the Weil-Deligne representation associated to ρv\rho_{v}.

The representations of fixed vv-adic Hodge type form a Zariski closed subset of the generic fiber of the unrestricted deformation ring, as the following theorem shows.

Theorem 5.16.

[23, Theorem 2.7.6], [2, Theorem D] Let vSpv\in S_{p} and suppose ρ¯v:GFvGL2(k)\bar{\rho}_{v}:G_{F_{v}}\to\operatorname{GL}_{2}(k) is Schur with deformation ring RvR_{v}. For any vv-adic Hodge type (𝐰v,τv,χv){(\mathbf{w}_{v},\tau_{v},\chi_{v})} with detρ¯v=χ¯v\det\bar{\rho}_{v}=\overline{\chi}_{v}, there exists a reduced, 𝒪\mathcal{O}-flat quotient RvRv(𝐰v,τv)R_{v}\twoheadrightarrow R_{v}(\mathbf{w}_{v},\tau_{v}) such that for any closed point 𝔭vRv[1/ϖ]\mathfrak{p}_{v}\subset R_{v}[1/\varpi], the corresponding representation is of type (𝐰v,τv,χv){(\mathbf{w}_{v},\tau_{v},\chi_{v})} if and only if 𝔭vSpecRv(𝐰v,τv)[1/ϖ]\mathfrak{p}_{v}\in\operatorname{Spec}R_{v}(\mathbf{w}_{v},\tau_{v})[1/\varpi]. Moreover, if 𝔭v\mathfrak{p}_{v} corresponds to a representation ρv\rho_{v} such that WD(ρv)\operatorname{WD}(\rho_{v}) is generic, then 𝔭v\mathfrak{p}_{v} is a regular point of SpecRv(𝐰v,τv)[1/ϖ]\operatorname{Spec}R_{v}(\mathbf{w}_{v},\tau_{v})[1/\varpi]

The relation between deformation rings in characteristic 0 and pp is explicated in [24, Lemma 2.3.3, Proposition 2.3.5], from which one readily deduces the following two results.

Theorem 5.17.

Let ρ¯v:GFvGL2(k)\bar{\rho}_{v}:G_{F_{v}}\to\operatorname{GL}_{2}(k) be a representation with framed deformation ring Rρ¯vR_{\bar{\rho}_{v}}^{\square}. Suppose 𝔭vRρ¯v\mathfrak{p}_{v}\subset R_{\bar{\rho}_{v}}^{\square} is a closed point corresponding to ρv:GFvGL2(L)\rho_{v}:G_{F_{v}}\to\operatorname{GL}_{2}(L). Then the localisation and completion (Rρ¯v)𝔭v(R_{\bar{\rho}_{v}}^{\square})_{\mathfrak{p}_{v}}^{\wedge} represents DρvD_{\rho_{v}}^{\square}. If ρ¯v\bar{\rho}_{v} is Schur, the analogous statement holds for the deformation rings.

Theorem 5.18.

Let 𝒮=(ρ¯,χ,S,{Dρ¯vχ,}vS)\mathcal{S}=(\bar{\rho},\chi,S,\{D^{\chi,\square}_{\bar{\rho}_{v}}\}_{v\in S}) be the global deformation problem of deformations unramified outside of SS, represented by R𝒮R_{\mathcal{S}}. Suppose 𝔭R𝒮[1/ϖ]\mathfrak{p}\subset R_{\mathcal{S}}[1/\varpi] is a closed point corresponding to a representation ρ:GF,SGL2(L)\rho:G_{F,S}\to\operatorname{GL}_{2}(L). Then the localisation and completion R𝒮,ρ:=(R𝒮)𝔭R_{\mathcal{S},\rho}:={(R_{\mathcal{S}})}_{\mathfrak{p}}^{\wedge} represents the functor D𝒮,ρ:𝔄^L𝐒𝐞𝐭D_{\mathcal{S},\rho}:\widehat{\mathfrak{A}}_{L}\to\mathbf{Set} for which D𝒮,ρ(A)D_{\mathcal{S},\rho}(A) is the set of deformations of ρ\rho to AA of determinant χ\chi which are unramified outside of 𝒮\mathcal{S}.

When dealing with irreducible representations, passage from pseudodeformation rings to deformation rings is enabled by the following result.

Theorem 5.19.

Let ρv:GFvGL2(L)\rho_{v}:G_{F_{v}}\to\operatorname{GL}_{2}(L) be an absolutely irreducible characteristic 0 representation of GFvG_{F_{v}} and 𝔭vRtrρ¯vps,χv[1/ϖ]\mathfrak{p}_{v}\subset R^{\operatorname{ps},\chi_{v}}_{\operatorname{tr}\bar{\rho}_{v}}[1/\varpi] the closed point corresponding to trρv\operatorname{tr}\rho_{v}. Then

(Rtrρ¯vps,χv)𝔭vRρv,(R^{\operatorname{ps},\chi_{v}}_{\operatorname{tr}\bar{\rho}_{v}})_{\mathfrak{p}_{v}}^{\wedge}\cong R_{\rho_{v}},

the unrestricted deformation ring of ρv\rho_{v}.

Proof.

The ring (Rtrρ¯vps,χv)𝔭v(R^{\operatorname{ps},\chi_{v}}_{\operatorname{tr}\bar{\rho}_{v}})_{\mathfrak{p}_{v}}^{\wedge} is isomorphic to Rtrρvps,χvR^{\operatorname{ps},\chi_{v}}_{\operatorname{tr}\rho_{v}} by [23, Lemma 2.3.3, Proposition 2.3.5]. The latter is canonically isomorphic to RρvR_{\rho_{v}} since ρv\rho_{v} is absolutely irreducible (a consequence of the main theorem of [30]). ∎

The Galois cohomology groups discussed in Section are naturally isomorphic to tangent spaces of deformation rings, and local conditions on the cohomology classes corresponds to conditions on the restrictions ρ|GFv\rho|_{G_{F_{v}}}. For a ring RR, let T[R]\operatorname{T}[R] denote its Zariski tangent space, i.e. T(R)=(𝔪R/𝔪R2)\operatorname{T}(R)=(\mathfrak{m}_{R}/\mathfrak{m}_{R}^{2})^{\vee}.

Theorem 5.20.

We have the following natural isomorphisms of LL-vector spaces (when the deformation rings exist):

H1(GF,S,ad0ρ)\displaystyle H^{1}\big{(}G_{F,S},\operatorname{ad}^{0}\rho\big{)} T[R𝒮,ρ]\displaystyle\cong\operatorname{T}[R_{\mathcal{S},\rho}]
H1(GFv,ad0ρ)\displaystyle H^{1}\big{(}G_{F_{v}},\operatorname{ad}^{0}\rho\big{)} T[Rρv]\displaystyle\cong\operatorname{T}[R_{\rho_{v}}]
Hf1(GFv,ad0ρ)\displaystyle H^{1}_{f}\big{(}G_{F_{v}},\operatorname{ad}^{0}\rho\big{)} T[Rρv(σv)]\displaystyle\cong\operatorname{T}[R_{\rho_{v}}(\sigma_{v})]
Proof.

The first two follow from Theorems 5.17, 5.18 and a standard argument (see e.g. [1, Theorem 5.1.4]). The third item follows from the main theorem of [26] and [2, Proposition 1.3.12]. ∎

6 Representations and completed homology

In this section, we discuss the representation-theoretic input in our main result.

6.1 Iwasawa modules and categories of smooth representations

In this subsection, we recall some general facts about modules over Iwasawa algebras, following [15].

Definition 6.1.

The Iwasawa algebra (with 𝒪\mathcal{O}-coefficients) of a compact pp-adic analytic group KK is the profinite (possibly non-commutative) ring

𝒪[[K]]=limUpK𝒪[K/Up].\mathcal{O}[[K]]=\varprojlim_{\begin{subarray}{c}U_{p}\trianglelefteq K\end{subarray}}\mathcal{O}[K/U_{p}].

The inversion map gg1g\mapsto g^{-1} in KK induces an equivalence of categories between left- and right 𝒪[[K]]\mathcal{O}[[K]]-modules.

Suppose KK is an open compact subgroup of a possibly non-compact pp-adic analytic group GG. The Iwasawa algebra 𝒪[[K]]\mathcal{O}[[K]] contains the group algebra 𝒪[K]\mathcal{O}[K] as a subring, which in turn sits inside the full group algebra 𝒪[G]\mathcal{O}[G]. Of special interest to us are modules with an action of both 𝒪[[K]]\mathcal{O}[[K]] and 𝒪[G]\mathcal{O}[G] such that the restrictions to 𝒪[K]\mathcal{O}[K] coincide.

Definition 6.2.

Let GG be a pp-adic analytic group, KGK\subset G a compact subgroup and ζ:Z(G)𝒪×\zeta:Z(G)\to\mathcal{O}^{\times} a central character. The category of profinite augmented 𝒪[G]\mathcal{O}[G]-modules is the category ModG,ζpfa(𝒪)\operatorname{Mod}^{\operatorname{pfa}}_{G,\zeta}(\mathcal{O}) with objects the profinite topological 𝒪[[K]]\mathcal{O}[[K]]-modules admitting a neighbourhood basis of the identity given by 𝒪[[K]]\mathcal{O}[[K]]-submodules, with a compatible action of GG and with central character ζ\zeta. The morphisms are 𝒪[G]\mathcal{O}[G]-linear continuous maps.

The category ModG,ζpfa(𝒪)\operatorname{Mod}^{\operatorname{pfa}}_{G,\zeta}(\mathcal{O}) abelian and independent of the choice of KK. Pontryagin duality defines an anti-equivalence

ModG,ζsm(𝒪)\displaystyle\operatorname{Mod}^{\operatorname{sm}}_{G,\zeta}(\mathcal{O}) ModG,ζpfa(𝒪)\displaystyle\to\operatorname{Mod}^{\operatorname{pfa}}_{G,\zeta}(\mathcal{O})
V\displaystyle V V:=Hom𝒪(V,L/𝒪)\displaystyle\mapsto V^{\vee}:=\operatorname{Hom}_{\mathcal{O}}(V,L/\mathcal{O})

where ModG,ζsm(𝒪)\operatorname{Mod}^{\operatorname{sm}}_{G,\zeta}(\mathcal{O}) is the category of smooth GG-representations over 𝒪\mathcal{O} with central character ζ\zeta. The inverse of this functor is also given by ()(\cdot)^{\vee}. Here, a GG-representation VV is called smooth if

V=K,sVK[ϖs].V=\cup_{K,s}V^{K}[\varpi^{s}].

A stronger condition than smoothness is (local) admissibility. A smooth representation VV is said to be admissible if each term VK[ϖs]V^{K}[\varpi^{s}] in the union above is finitely generated over 𝒪\mathcal{O}. We say that VV is locally admissible if, for every vVv\in V, the subrepresentation of VV generated by vv is admissible.

The full subcategory of ModG,ζsm(𝒪)\operatorname{Mod}^{\operatorname{sm}}_{G,\zeta}(\mathcal{O}) consisting of locally admissible (resp. admissible) representations is denoted ModG,ζladm(𝒪)\operatorname{Mod}^{\operatorname{ladm}}_{G,\zeta}(\mathcal{O}) (resp. ModG,ζadm(𝒪)\operatorname{Mod}^{\operatorname{adm}}_{G,\zeta}(\mathcal{O})). These categories are also abelian, and the admissible representations are dual to the profinite augmented modules which are finitely generated as 𝒪[[K]]\mathcal{O}[[K]]-modules. We define

G,ζ(𝒪):=(ModG,ζladm),\displaystyle\mathfrak{C}_{G,\zeta}(\mathcal{O}):=(\operatorname{Mod}^{\operatorname{ladm}}_{G,\zeta})^{\vee},

so that the categories

ModG,ζadm(𝒪)ModG,ζladm(𝒪)ModG,ζsm(𝒪)\operatorname{Mod}^{\operatorname{adm}}_{G,\zeta}(\mathcal{O})\subset\operatorname{Mod}^{\operatorname{ladm}}_{G,\zeta}(\mathcal{O})\subset\operatorname{Mod}^{\operatorname{sm}}_{G,\zeta}(\mathcal{O})

are anti-equivalent under Pontryagin duality to

ModG,ζfga(𝒪)G,ζ(𝒪)ModG,ζpfa(𝒪).\operatorname{Mod}_{G,\zeta}^{\operatorname{fga}}(\mathcal{O})\subset\mathfrak{C}_{G,\zeta}(\mathcal{O})\subset\operatorname{Mod}^{\operatorname{pfa}}_{G,\zeta}(\mathcal{O}).

6.2 Representations of GL2(p)\operatorname{GL}_{2}(\mathbb{Q}_{p})

In this subsection, we cite some facts about GL2(p)\operatorname{GL}_{2}(\mathbb{Q}_{p})-representations from [32]. For now, we let G=GL2(p)G=\operatorname{GL}_{2}(\mathbb{Q}_{p}), BGB\subset G the subgroup of upper triangular matrices and K=GL2(p)K=\operatorname{GL}_{2}(\mathbb{Z}_{p}).

The category ModG,ζladm(𝒪)\operatorname{Mod}^{\operatorname{ladm}}_{G,\zeta}(\mathcal{O}) is canonically isomorphic to a direct product of subcategories

ModG,ζladm(𝒪)𝔅ModG,ζladm(𝒪)𝔅\operatorname{Mod}^{\operatorname{ladm}}_{G,\zeta}(\mathcal{O})\cong\prod_{\mathfrak{B}}\operatorname{Mod}^{\operatorname{ladm}}_{G,\zeta}(\mathcal{O})_{\mathfrak{B}}

([32, Proposition 5.34]). Here, ModG,ζladm(𝒪)𝔅\operatorname{Mod}^{\operatorname{ladm}}_{G,\zeta}(\mathcal{O})_{\mathfrak{B}} denotes the full subcategory consisting of representations with the property that all irreducible subquotients are isomorphic to one of a finite number of representations lying inside the block 𝔅\mathfrak{B}, which is a finite set of kk-representations. Pontryagin duality preserves this decomposition, and thus

G,ζ(𝒪)𝔅G,ζ(𝒪)𝔅.\mathfrak{C}_{G,\zeta}(\mathcal{O})\cong\prod_{\mathfrak{B}}\mathfrak{C}_{G,\zeta}(\mathcal{O})_{\mathfrak{B}}.

Since kk is not algebraically closed, there might irreducible representations which are not absolutely irreducible. After a finite base change, say from kk to kk^{\prime}, such a representation decomposes into a direct sum of absolutely irreducible kk^{\prime}-representations. Therefore, we will restrict our attention to blocks 𝔅\mathfrak{B} containing an absolutely irreducible representation and tacitly assume kk to be large enough throughout the rest of the paper.

Proposition 6.3.

([32, Proposition 5.42]) Suppose p5p\geq 5. The blocks 𝔅\mathfrak{B} containing an absolutely irreducible representation are as follows:

  1. (i)

    𝔅={π}\mathfrak{B}=\{\pi\}, where π\pi is supersingular.

  2. (ii)

    𝔅={IndBG(χ1χ2ε¯1),IndBG(χ2χ1ε¯1)}\mathfrak{B}=\{\operatorname{Ind}_{B}^{G}(\chi_{1}\otimes\chi_{2}\bar{\varepsilon}^{-1}),\operatorname{Ind}_{B}^{G}(\chi_{2}\otimes\chi_{1}\bar{\varepsilon}^{-1})\} with χ1/χ21,ε¯±1\chi_{1}/\chi_{2}\neq 1,\bar{\varepsilon}^{\pm 1}.

  3. (iii)

    𝔅={IndBG(χχε¯1)}\mathfrak{B}=\{\operatorname{Ind}_{B}^{G}(\chi\otimes\chi\bar{\varepsilon}^{-1})\}.

  4. (iv)

    𝔅={η,Spη,(IndBGα)η}\mathfrak{B}=\{\eta,\operatorname{Sp}\otimes\eta,(\operatorname{Ind}_{B}^{G}\alpha)\otimes\eta\}, where η=η0det:Gk×\eta=\eta_{0}\circ\det:G\to k^{\times} is a smooth character.

We will consider PGL2\operatorname{PGL}_{2}-representations, meaning we will have ζ\zeta equal to the trivial central character. As a consequence, case (iii) above will not feature later on.

For every π𝔅\pi\in\mathfrak{B}, we let PππP_{\pi}\twoheadrightarrow\pi^{\vee} be a projective envelope and define

P𝔅\displaystyle P_{\mathfrak{B}} =π𝔅Pππ𝔅π,\displaystyle=\bigoplus_{\pi\in\mathfrak{B}}P_{\pi}\twoheadrightarrow\bigoplus_{\pi\in\mathfrak{B}}\pi^{\vee},
E𝔅\displaystyle E_{\mathfrak{B}} =End𝔅(𝒪)(P𝔅).\displaystyle=\operatorname{End}_{\mathfrak{C}_{\mathfrak{B}}(\mathcal{O})}(P_{\mathfrak{B}}).

Then P𝔅P_{\mathfrak{B}} is a projective envelope of π𝔅π\bigoplus_{\pi\in\mathfrak{B}}\pi^{\vee} and is moreover a projective generator in the category G,ζ(𝒪)𝔅\mathfrak{C}_{G,\zeta}(\mathcal{O})_{\mathfrak{B}}. The ring E𝔅E_{\mathfrak{B}} is compact with respect to a natural topology and there is an equivalence of abelian categories

G,ζ(𝒪)𝔅\displaystyle\mathfrak{C}_{G,\zeta}(\mathcal{O})_{\mathfrak{B}} RModcpt(E𝔅)\displaystyle\to\operatorname{RMod}^{\operatorname{cpt}}(E_{\mathfrak{B}})
V\displaystyle V Hom𝔅(𝒪)(P𝔅,V).\displaystyle\mapsto\operatorname{Hom}_{\mathfrak{C}_{\mathfrak{B}}(\mathcal{O})}(P_{\mathfrak{B}},V).

The inverse of the functor Hom𝔅(𝒪)(P𝔅,)\operatorname{Hom}_{\mathfrak{C}_{\mathfrak{B}}(\mathcal{O})}(P_{\mathfrak{B}},-) is given by the completed tensor product, so that for any V𝔅(𝒪)V\in\mathfrak{C}_{\mathfrak{B}}(\mathcal{O}), there is a canonical isomorphism

VHom𝔅(𝒪)(P𝔅,V)^E𝔅P𝔅.V\cong\operatorname{Hom}_{\mathfrak{C}_{\mathfrak{B}}(\mathcal{O})}(P_{\mathfrak{B}},V)\widehat{\otimes}_{E_{\mathfrak{B}}}P_{\mathfrak{B}}.

So far we have only mentioned the ‘automorphic side’ of the pp-adic local Langlands correspondence. Let us now describe the ‘Galois side’ of the picture and relate E𝔅E_{\mathfrak{B}} to a pseudodeformation ring. To each block 𝔅\mathfrak{B} above, we associate a semisimple 2-dimensional kk-representation ρ¯𝔅\bar{\rho}_{\mathfrak{B}} of GpG_{\mathbb{Q}_{p}} by the following recipe:

  1. (i)

    ρ¯𝔅=𝐕(π)\bar{\rho}_{\mathfrak{B}}=\mathbf{V}(\pi) where 𝐕\mathbf{V} is Colmez’ Montreal functor (see [32, 5.7]).

  2. (ii)

    ρ¯𝔅=χ1χ2\bar{\rho}_{\mathfrak{B}}=\chi_{1}\oplus\chi_{2} (viewing χ1,χ2\chi_{1},\chi_{2} as Galois representations via local class field theory).

  3. (iii)

    ρ¯𝔅=χχ\bar{\rho}_{\mathfrak{B}}=\chi\oplus\chi.

  4. (iv)

    ρ¯𝔅=η0η0ε¯\bar{\rho}_{\mathfrak{B}}=\eta_{0}\oplus\eta_{0}\bar{\varepsilon}.

In each case, we have detρ¯𝔅=ζε¯\det\bar{\rho}_{\mathfrak{B}}=\zeta\bar{\varepsilon}.

Theorem 6.4.

Let Z𝔅=Z(E𝔅)Z_{\mathfrak{B}}=Z(E_{\mathfrak{B}}) denote the center of E𝔅E_{\mathfrak{B}}. Then we have the following:

  • (a)

    E𝔅E_{\mathfrak{B}} is a finitely generated module over Z𝔅Z_{\mathfrak{B}}.

  • (b)

    There is a canonical isomorphism of 𝒪\mathcal{O}-algebras

    Z𝔅Rtrρ¯𝔅ps,ζεZ_{\mathfrak{B}}\cong R^{\operatorname{ps},\zeta\varepsilon}_{\operatorname{tr}\bar{\rho}_{\mathfrak{B}}}

    where Rtrρ¯𝔅ps,ζεR^{\operatorname{ps},\zeta\varepsilon}_{\operatorname{tr}\bar{\rho}_{\mathfrak{B}}} is the pseudodeformation ring of 22-dimensional pseudorepresentations of GpG_{\mathbb{Q}_{p}} with determinant ζε\zeta\varepsilon lifting trρ¯𝔅\operatorname{tr}\bar{\rho}_{\mathfrak{B}}.

  • (c)

    Let 𝔅\mathfrak{B} be one of the blocks (i), (ii) or (iv), and suppose 𝔭vRtrρ¯𝔅ps,ζε[1/ϖ]\mathfrak{p}_{v}\subset R^{\operatorname{ps},\zeta\varepsilon}_{\operatorname{tr}\bar{\rho}_{\mathfrak{B}}}[1/\varpi] is a closed point corresponding to an irreducible GpG_{\mathbb{Q}_{p}}-representation ρv\rho_{v}, so that (Rtrρ¯𝔅ps,ζε)𝔭vRρv(R^{\operatorname{ps},\zeta\varepsilon}_{\operatorname{tr}\bar{\rho}_{\mathfrak{B}}})_{\mathfrak{p}_{v}}^{\wedge}\cong R_{\rho_{v}}. Then

    (E𝔅)𝔭vM|𝔅|(Rρv),(E_{\mathfrak{B}})_{\mathfrak{p}_{v}}^{\wedge}\cong M_{|\mathfrak{B}|}(R_{\rho_{v}}),

    the ring of |𝔅||\mathfrak{B}|-by-|𝔅||\mathfrak{B}| matrices with coefficients in RρvR_{\rho_{v}}, the unrestricted deformation ring of ρv\rho_{v}.

Proof.

(a,b) See [32, Theorem 1.5, Corollary 8.11, Corollary 9.25, Lemma 10.90].
(c) In the block (i), the natural map Z𝔅E𝔅Z_{\mathfrak{B}}\to E_{\mathfrak{B}} is an isomorphism ([32, Proposition 6.3]). The case (ii) is dealt with in [32, Corollary B.27], and we outline the proof here. We have a presentation

E𝔅=(HomG,ζ(𝒪)(Pi,Pj))1i,j2(Z𝔅𝟏Z𝔅b1Z𝔅b2Z𝔅𝟏)E_{\mathfrak{B}}=(\operatorname{Hom}_{\mathfrak{C}_{G,\zeta}(\mathcal{O})}(P_{i},P_{j}))_{1\leq i,j\leq 2}\cong\begin{pmatrix}Z_{\mathfrak{B}}\mathbf{1}&Z_{\mathfrak{B}}b_{1}\\ Z_{\mathfrak{B}}b_{2}&Z_{\mathfrak{B}}\mathbf{1}\end{pmatrix}

where b1b2=b2b1=cZ𝔅b_{1}\circ b_{2}=b_{2}\circ b_{1}=cZ_{\mathfrak{B}} for an element cZ𝔅c\in Z_{\mathfrak{B}} with the property that a point 𝔭vRtrρ¯𝔅ps,ζε[1/ϖ]\mathfrak{p}_{v}\subset R^{\operatorname{ps},\zeta\varepsilon}_{\operatorname{tr}\bar{\rho}_{\mathfrak{B}}}[1/\varpi] defines an irreducible representation if and only if c𝔭vc\notin\mathfrak{p}_{v} (put differently, the reducibility ideal is principal, generated by cc). Consequently the ring (E𝔅)𝔭v(E𝔅[1/c])𝔭v(E_{\mathfrak{B}})_{\mathfrak{p}_{v}}^{\wedge}\cong(E_{\mathfrak{B}}[1/c])_{\mathfrak{p}_{v}}^{\wedge} is a 22-by-22 matrix algebra over (Z𝔅)𝔭vRρv(Z_{\mathfrak{B}})_{\mathfrak{p}_{v}}^{\wedge}\cong R_{\rho_{v}}.

The block (iv) is dealt with similarly. We have a presentation of E𝔅E_{\mathfrak{B}} as a 33-by-33 matrix of homomorphisms PiPjP_{i}\to P_{j} ([32, p.134]). Here, the relevant relations are generated by two elements c0,c1Z𝔅c_{0},c_{1}\in Z_{\mathfrak{B}}, and as before, the point 𝔭v\mathfrak{p}_{v} defines an irreducible representation if (c0,c1)𝔭v(c_{0},c_{1})\not\subset\mathfrak{p}_{v}. Since ci1(c0,c1)=Z𝔅[1/ci]c_{i}^{-1}(c_{0},c_{1})=Z_{\mathfrak{B}}[1/c_{i}], the same argument as in the previous case proves the claim. ∎

6.3 Representations of PGL2(p)\prod{\operatorname{PGL}_{2}(\mathbb{Q}_{p})}

Let us now turn to our case of interest, namely

G\displaystyle G =vp𝐆(Fv)vpPGL2(p),\displaystyle=\prod_{v\mid p}\mathbf{G}(F_{v})\cong\prod_{v\mid p}\operatorname{PGL}_{2}(\mathbb{Q}_{p}),
K\displaystyle K =vpKvvpPGL2(p).\displaystyle=\prod_{v\mid p}K_{v}\cong\prod_{v\mid p}\operatorname{PGL}_{2}(\mathbb{Z}_{p}).

Our category of interest is G(𝒪)\mathfrak{C}_{G}(\mathcal{O}), the Pontryagin dual of the category of locally admissible GG-representations, and we make the identification

G(𝒪)(ModvpGL2(p),𝟏ladm(𝒪)).\mathfrak{C}_{G}(\mathcal{O})\cong\big{(}\operatorname{Mod}^{\operatorname{ladm}}_{\prod_{v\mid p}\operatorname{GL}_{2}(\mathbb{Q}_{p}),\mathbf{1}}(\mathcal{O})\big{)}^{\vee}.

The existence of a block decomposition relies only on general facts about locally finite categories and still hold for G(𝒪)\mathfrak{C}_{G}(\mathcal{O}). That is, there exists a set of blocks 𝔅\mathfrak{B}, projective generators P𝔅P_{\mathfrak{B}} and a decomposition

G(𝒪)𝔅𝔅(𝒪)𝔅RModcpt(E𝔅)\mathfrak{C}_{G}(\mathcal{O})\cong\prod_{\mathfrak{B}}\mathfrak{C}_{\mathfrak{B}}(\mathcal{O})\cong\prod_{\mathfrak{B}}\operatorname{RMod}^{\operatorname{cpt}}(E_{\mathfrak{B}})

where E𝔅E_{\mathfrak{B}} is the endomorphism ring of P𝔅P_{\mathfrak{B}}.

Lemma 6.5.

Let H𝔅(𝒪)H\in\mathfrak{C}_{\mathfrak{B}}(\mathcal{O}) and suppose HH is finitely generated as an 𝒪[[K]]\mathcal{O}[[K]]-module. Then HomG(𝒪)(P𝔅,H)\operatorname{Hom}_{\mathfrak{C}_{G}(\mathcal{O})}(P_{\mathfrak{B}},H) is a finitely generated E𝔅E_{\mathfrak{B}}-module.

Proof.

Recall that HH being finitely generated over 𝒪[[K]]\mathcal{O}[[K]] is equivalent to HH^{\vee} being admissible. Since E𝔅=EndG(𝒪)(P𝔅)E_{\mathfrak{B}}=\operatorname{End}_{\mathfrak{C}_{G}(\mathcal{O})}(P_{\mathfrak{B}}), it suffices to show that for some rr, there is a surjection

P𝔅rH.P_{\mathfrak{B}}^{\oplus r}\twoheadrightarrow H.

Indeed, applying HomG(𝒪)(P𝔅,)\operatorname{Hom}_{\mathfrak{C}_{G}(\mathcal{O})}(P_{\mathfrak{B}},-) to this diagram then proves the lemma.

We claim that the cosocle cosocH\operatorname{cosoc}H is a finite direct sum iπi\bigoplus_{i}\pi_{i}^{\vee} of irreducible objects. Indeed, every πiiπi=socH\pi_{i}\hookrightarrow\oplus_{i}\pi_{i}=\operatorname{soc}H^{\vee} has non-zero K0K_{0}-invariants for any pro-pp group K0K_{0}. Since the dual HH^{\vee} is admissible, we see that socH\operatorname{soc}H^{\vee} is a finite direct sum of irreducibles. The same then holds for cosocH=(socH).\operatorname{cosoc}H^{\vee}=(\operatorname{soc}H^{\vee})^{\vee}. Now, choose a surjection

P𝔅rcosocHP_{\mathfrak{B}}^{\oplus r}\twoheadrightarrow\operatorname{cosoc}H

for some rr. By the projectivity of P𝔅P_{\mathfrak{B}}, this map factors as

P𝔅rHcosocH,P_{\mathfrak{B}}^{\oplus r}\to H\to\operatorname{cosoc}H,

and the admissibility of VV implies that the second arrow is a superfluous surjection ([11, Lemma 4.6]). Thus the first map is also surjective, which completes the proof. ∎

Pan [31] has extended the results of the previous section to our setting.

Proposition 6.6.

Let (𝔅v)vp(\mathfrak{B}_{v})_{v\mid p} be a tuple of blocks of GL2(p),𝟏(𝒪)\mathfrak{C}_{\operatorname{GL}_{2}(\mathbb{Q}_{p}),\mathbf{1}}(\mathcal{O}), each containing an absolutely irreducible representation, and define

𝔅:=vp𝔅v:={vpπv:πv𝔅v}.\mathfrak{B}:=\otimes_{v\mid p}\mathfrak{B}_{v}:=\{\otimes_{v\mid p}\pi_{v}:\pi_{v}\in\mathfrak{B}_{v}\}.

Then 𝔅\mathfrak{B} is a block of G(𝒪)\mathfrak{C}_{G}(\mathcal{O}), and moreover:

  • (a)

    P𝔅:=^vpP𝔅vP_{\mathfrak{B}}:=\widehat{\bigotimes}_{v\mid p}P_{\mathfrak{B}_{v}} is a projective envelope of ^vpπv\widehat{\bigotimes}_{v\mid p}\pi_{v}^{\vee}.

  • (b)

    E𝔅=EndG(𝒪)(P𝔅)^vpE𝔅vE_{\mathfrak{B}}=\operatorname{End}_{\mathfrak{C}_{G}(\mathcal{O})}(P_{\mathfrak{B}})\cong\widehat{\bigotimes}_{v\mid p}E_{\mathfrak{B}_{v}}

  • (c)

    E𝔅E_{\mathfrak{B}} is a finitely generated module over RppsR^{\operatorname{ps}}_{p}.

  • (d)

    Let 𝔭v\mathfrak{p}_{v} be the ideal of Rvps=Rtrρ¯vpsR^{\operatorname{ps}}_{v}=R^{\operatorname{ps}}_{\operatorname{tr}\bar{\rho}_{v}} corresponding to the trace trρv\operatorname{tr}\rho_{v} of a irreducible representation ρv\rho_{v}, and let 𝔭\mathfrak{p} be the ideal of Rpps=^vpRvpsR^{\operatorname{ps}}_{p}=\widehat{\bigotimes}_{v\mid p}R^{\operatorname{ps}}_{v} generated by the joint image of the 𝔭v\mathfrak{p}_{v}’s. Then

    (E𝔅)𝔭M|𝔅|((Rpps)𝔭)(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}\cong M_{|\mathfrak{B}|}((R^{\operatorname{ps}}_{p})_{\mathfrak{p}}^{\wedge})

    as LL-algebras (possibly non-commutative).

Proof.

For (i-iii), see [31, §3.4]. For (iv), recall that for vpv\mid p, P𝔅vP_{\mathfrak{B}_{v}} is a direct sum iv=1|𝔅v|P𝔅v,iv\bigoplus_{i_{v}=1}^{|\mathfrak{B}_{v}|}P_{\mathfrak{B}_{v},i_{v}} and we have a matrix presentation of E𝔅E_{\mathfrak{B}} over RppsR^{\operatorname{ps}}_{p} given by

E𝔅=EndG(𝒪)(^vpP𝔅v)=EndG(𝒪)(vp^iv=1|𝔅v|P𝔅v,iv)=(^vpHomGL2(p),𝟏(𝒪)(P𝔅v,iv,P𝔅v,jv))(iv),(jv)\begin{split}E_{\mathfrak{B}}&=\operatorname{End}_{\mathfrak{C}_{G}(\mathcal{O})}(\widehat{\bigotimes}_{v\mid p}P_{\mathfrak{B}_{v}})\\ &=\operatorname{End}_{\mathfrak{C}_{G}(\mathcal{O})}\big{(}\widehat{\bigotimes_{v\mid p}}\bigoplus_{i_{v}=1}^{|\mathfrak{B}_{v}|}P_{\mathfrak{B}_{v},i_{v}}\big{)}\\ &=\big{(}\widehat{\bigotimes}_{v\mid p}\operatorname{Hom}_{\mathfrak{C}_{\operatorname{GL}_{2}(\mathbb{Q}_{p}),\mathbf{1}}(\mathcal{O})}(P_{\mathfrak{B}_{v},i_{v}},P_{\mathfrak{B}_{v},j_{v}})\big{)}_{(i_{v}),(j_{v})}\end{split}

where (iv)(i_{v}) and (jv)(j_{v}) run over sequences of length |𝔅v||\mathfrak{B}_{v}| such that 1iv,jv|𝔅v|1\leq i_{v},j_{v}\leq|\mathfrak{B}_{v}|. This defines a matrix presentation of E𝔅E_{\mathfrak{B}} over RppsR^{\operatorname{ps}}_{p} as in the proof of Theorem 6.4(iv), and there exists an element cRppsc\in R^{\operatorname{ps}}_{p} such that E𝔅[1/c]E_{\mathfrak{B}}[1/c] is a matrix algebra. Indeed, we may take cc equal to the product of cvRvpsc_{v}\in R^{\operatorname{ps}}_{v} chosen such that E𝔅v[1/cv]E_{\mathfrak{B}_{v}}[1/c_{v}] is a matrix algebra, as in the aforementioned proof. ∎

6.4 Functors of twisted coinvariants

In this section, we introduce the [Up]\mathbb{Z}[U_{p}]-modules whose associated local systems we will use as coefficients in homology and their associated functors of twisted coinvariants. They are indexed by vv-adic Hodge types (Definition 5.14).

Theorem 6.7.

([7, A.1.5.]) Let τv:IFvGL2(L)\tau_{v}:I_{F_{v}}\to\operatorname{GL}_{2}(L) be a representation with open kernel. There exists a unique smooth irreducible KvK_{v}-representation σ(τv)\sigma(\tau_{v}) on an LL-vector space characterised by the property

HomKv(πv,σ(τv))0LL(πv)|IFvτv,\operatorname{Hom}_{K_{v}}(\pi_{v},\sigma(\tau_{v}))\neq 0\iff\operatorname{LL}(\pi_{v})|_{I_{F_{v}}}\cong\tau_{v},

when πv\pi_{v} ranges over all smooth absolutely irreducible infinite-dimensional GvG_{v}-representation over LL and LL(πv)\operatorname{LL}(\pi_{v}) is the Weil-Deligne representation associated to πv\pi_{v} by the classical local Langlands correspondence (normalised as in [7]).

Definition 6.8.

Let (𝐰v,τv,χv){(\mathbf{w}_{v},\tau_{v},\chi_{v})} be a vv-adic Hodge type. The KvK_{v}-representation associated to (𝐰v,τv,χv){(\mathbf{w}_{v},\tau_{v},\chi_{v})} is the representation

σ(𝐰v,τv)=σ(τv)(Symbvav1L2)(det)av\sigma(\mathbf{w}_{v},\tau_{v})=\sigma(\tau_{v})\otimes(\operatorname{Sym}^{b_{v}-a_{v}-1}L^{2})\otimes(\det)^{a_{v}}

We write σ(𝐰v)=(Symbvav1L2)(det)av\sigma(\mathbf{w}_{v})=(\operatorname{Sym}^{b_{v}-a_{v}-1}L^{2})\otimes(\det)^{a_{v}} so that σ(𝐰v,τv)=σ(τv)σ(𝐰v)\sigma(\mathbf{w}_{v},\tau_{v})=\sigma(\tau_{v})\otimes\sigma(\mathbf{w}_{v}). Since KvK_{v} is compact, there is a KvK_{v}-stable 𝒪\mathcal{O}-lattice

σ(𝐰v,τv)=σ(τv)σ(𝐰v)σ(𝐰v,τv).\sigma^{\circ}(\mathbf{w}_{v},\tau_{v})={\sigma^{\circ}}(\tau_{v})\otimes{\sigma^{\circ}}(\mathbf{w}_{v})\subset\sigma(\mathbf{w}_{v},\tau_{v}).

Given a tuple (𝐰,τ,χ)=(𝐰v,τv,χv)vSp(\mathbf{w},\tau,\chi)={(\mathbf{w}_{v},\tau_{v},\chi_{v})}_{v\in S_{p}} of vv-adic Hodge types, we obtain a K=vpKvK=\prod_{v\mid p}K_{v}-representation upon forming the tensor product over vSpv\in S_{p},

σ(𝐰,τ)=vpσ(𝐰v,τv),\sigma(\mathbf{w},\tau)=\bigotimes_{v\mid p}\sigma(\mathbf{w}_{v},\tau_{v}),

containing the KK-stable 𝒪\mathcal{O}-lattice

σ(𝐰,τ)=vpσ(𝐰v,τv)σ(𝐰,τ),\sigma^{\circ}(\mathbf{w},\tau)=\bigotimes_{v\mid p}\sigma^{\circ}(\mathbf{w}_{v},\tau_{v})\subset\sigma(\mathbf{w},\tau),

which is a finitely generated 𝒪[[K]]\mathcal{O}[[K]]-module. Given such a σ=σ(𝐰,τ){\sigma^{\circ}}=\sigma^{\circ}(\mathbf{w},\tau) and a compact 𝒪[[K]]\mathcal{O}[[K]]-module NN, we have a natural isomorphism

N^𝒪[[K]]σN𝒪[[K]]σ,N\hat{\otimes}_{\mathcal{O}[[K]]}{\sigma^{\circ}}\cong N\otimes_{\mathcal{O}[[K]]}{\sigma^{\circ}},

and 𝒪[[K]]σ-\otimes_{\mathcal{O}[[K]]}{\sigma^{\circ}} defines a right exact functor RModcpt(𝒪[[K]])Modcpt(𝒪)\operatorname{RMod}^{\operatorname{cpt}}(\mathcal{O}[[K]])\to\operatorname{Mod}^{\operatorname{cpt}}(\mathcal{O}) ([8, Lemma 2.1]). If NN is of the form N=^vpNv,N=\widehat{\bigotimes}_{v\mid p}N_{v}, we have an isomorphism

N𝒪[[K]]σvp^(Nv𝒪[[Kv]]σv),N\otimes_{\mathcal{O}[[K]]}{\sigma^{\circ}}\cong\widehat{\bigotimes_{v\mid p}}(N_{v}\otimes_{\mathcal{O}[[K_{v}]]}\sigma^{\circ}_{v}),

where σv=σ(𝐰v,τv)\sigma^{\circ}_{v}=\sigma^{\circ}(\mathbf{w}_{v},\tau_{v}). We will use the notation

N(σ)\displaystyle N({\sigma^{\circ}}) :=N𝒪[[K]]σ,\displaystyle:=N\otimes_{\mathcal{O}[[K]]}{\sigma^{\circ}},
N(σ)\displaystyle N(\sigma) :=N𝒪[[K]]σ.\displaystyle:=N\otimes_{\mathcal{O}[[K]]}\sigma.

Note that N(σ)=N(σ)[1/ϖ]N(\sigma)=N({\sigma^{\circ}})[1/\varpi]. On projective objects, this functor has the following alternate description which occurs in the literature.

Proposition 6.9.

[16, Remark 5.1.7] Suppose PP is projective in RModcpt(𝒪[[K]])\operatorname{RMod}^{\operatorname{cpt}}(\mathcal{O}[[K]]), and σ=σ(𝐰,τ){\sigma^{\circ}}=\sigma^{\circ}(\mathbf{w},\tau). There is a natural isomorphism

P(σ)Hom𝒪[[K]]cts(P,(σ)d)dP({\sigma^{\circ}})\cong\operatorname{Hom}_{\mathcal{O}[[K]]}^{\operatorname{cts}}(P,({\sigma^{\circ}})^{\operatorname{d}})^{\operatorname{d}}

where ()d=Hom𝒪(,𝒪)(-)^{d}=\operatorname{Hom}_{\mathcal{O}}(-,\mathcal{O}) with the topology of pointwise convergence.

6.5 Description of PP

Let 𝔅=vp𝔅v\mathfrak{B}=\otimes_{v\mid p}\mathfrak{B}_{v} be a block of GG of the form considered in Theorem 6.6, and σ=σ(𝐰,τ)\sigma=\sigma(\mathbf{w},\tau). In this section, we will consider the module of twisted coinvariants P𝔅(σ)=P𝔅𝒪[[K]]σP_{\mathfrak{B}}(\sigma)=P_{\mathfrak{B}}\otimes_{\mathcal{O}[[K]]}\sigma of the projective generator P𝔅P_{\mathfrak{B}} of the category 𝔅(𝒪)\mathfrak{C}_{\mathfrak{B}}(\mathcal{O}), or rather its localisation and completion at a characteristic 0 point.

Let Rp,ρ=(Rpps)𝔭R_{p,\rho}=(R^{\operatorname{ps}}_{p})_{\mathfrak{p}}^{\wedge}. The localisation simplifies things greatly, since (E𝔅)𝔭(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge} is isomorphic to a matrix algebra over Rp,ρR_{p,\rho} when 𝔭\mathfrak{p} is the ideal corresponding to a tuple (ρv)(\rho_{v}) of irreducible representations (Proposition 6.6(d)), and thus P𝔅(σ)𝔭P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge} is a direct sum of |𝔅||\mathfrak{B}| pairwise isomorphic Rp,ρR_{p,\rho}-modules. For this reason, it will suffice to consider a single summand of P𝔅P_{\mathfrak{B}}.

For every vpv\mid p, let

πv=IndPG(χv,1χv,2ε1)𝔅v\pi_{v}=\operatorname{Ind}_{P}^{G}(\chi_{v,1}\otimes\chi_{v,2}\varepsilon^{-1})\in\mathfrak{B}_{v}

where χv,1/χv,21,ε\chi_{v,1}/\chi_{v,2}\neq 1,\varepsilon (but allowing χv,1ε=χv,2\chi_{v,1}\varepsilon=\chi_{v,2}) and let P=^vpPv^vpπvP=\widehat{\bigotimes}_{v\mid p}P_{v}\twoheadrightarrow\widehat{\bigotimes}_{v\mid p}\pi_{v}^{\vee} be the projective envelope as in Proposition 6.6(a). Then, as explained above,

P𝔅(σ)𝔭(P(σ)𝔭)|𝔅|P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge}\cong\big{(}P(\sigma)_{\mathfrak{p}}^{\wedge}\big{)}^{\oplus|\mathfrak{B}|}

where we view the elements of this module as vectors and the matrix algebra (E𝔅)𝔭(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge} acts by matrix multiplication. There is a natural forgetful functor

G(𝒪)RModcpt(𝒪[[K]]).\mathfrak{C}_{G}(\mathcal{O})\to\operatorname{RMod}^{\operatorname{cpt}}(\mathcal{O}[[K]]).
Proposition 6.10.

PP is projective in the category of compact right 𝒪[[K]]\mathcal{O}[[K]]-modules.

Proof.

We mimic the proof of [16, Lemma B.8]. We have P=^vpPvP=\widehat{\otimes}_{v\mid p}P_{v}, a tensor product of modules over

^vp𝒪[[Kv]]^vSp𝒪[[PGL2(p)]].\widehat{\bigotimes}_{v\mid p}\mathcal{O}[[K_{v}]]\simeq\widehat{\bigotimes}_{v\in S_{p}}\mathcal{O}[[\operatorname{PGL}_{2}(\mathbb{Z}_{p})]].

We proceed by induction on the size of the set SpS_{p}. By [33, Corollary 5.3], PvP_{v} is projective in RModcpt(𝒪[[Kv]])\operatorname{RMod}^{\operatorname{cpt}}(\mathcal{O}[[K_{v}]]). For the induction step, let wSpw\in S_{p} and define

Kw\displaystyle K^{w} =vSp{w}Kv,\displaystyle=\prod_{v\in S_{p}\setminus\{w\}}K_{v},
Pw\displaystyle P^{w} =^vSp{w}Pv,\displaystyle=\widehat{\otimes}_{v\in S_{p}\setminus\{w\}}P_{v},

so that K=Kw×KwK=K_{w}\times K^{w} and P=Pw^PwP=P_{w}\widehat{\otimes}P^{w}. By the induction hypothesis, PwP_{w} and PwP^{w} are projective over 𝒪[[Kw]]\mathcal{O}[[K_{w}]] and 𝒪[[Kw]]\mathcal{O}[[K^{w}]], respectively. The universal property of the completed tensor product implies that for any compact 𝒪[[K]]\mathcal{O}[[K]]-module NN,

Hom𝒪[[Kw×Kw]]cts(Pw^𝒪[[K]]Pw,N)Hom𝒪[[Kw]]cts(Pw,Hom𝒪[[Kw]]cts(Pw,N)).\operatorname{Hom}^{\operatorname{cts}}_{\mathcal{O}[[K_{w}\times K^{w}]]}(P_{w}\widehat{\otimes}_{\mathcal{O}[[K]]}P^{w},N)\cong\operatorname{Hom}^{\operatorname{cts}}_{\mathcal{O}[[K_{w}]]}(P_{w},\operatorname{Hom}^{\operatorname{cts}}_{\mathcal{O}[[K^{w}]]}(P^{w},N)).

Hence, the projectivity of PP follows from that of PwP_{w} and PwP^{w}. ∎

Given σ=σ(𝐰,τ)\sigma=\sigma(\mathbf{w},\tau) we write Rv(σv)R_{v}(\sigma_{v}) for the ring Rv(𝐰v,τv)R_{v}(\mathbf{w}_{v},\tau_{v}) introduced in Theorem 5.16, and Rp(σ)=^vpRv(σv)R_{p}({\sigma^{\circ}})=\widehat{\bigotimes}_{v\mid p}R_{v}(\sigma_{v}).

Theorem 6.11.

Let σ=σ(𝐰,τ)\sigma=\sigma(\mathbf{w},\tau). Then the action of RppsR^{\operatorname{ps}}_{p} on P(σ)P({\sigma^{\circ}}) factors through Rp(σ)R_{p}(\sigma) and P(σ)P(\sigma) is locally free of rank 11 over the regular locus of SpecRp(σ)[1/ϖ]\operatorname{Spec}R_{p}(\sigma)[1/\varpi].

Proof.

We have

P(σ)^vpPv(σv)P({\sigma^{\circ}})\cong\widehat{\bigotimes}_{v\mid p}P_{v}(\sigma^{\circ}_{v})

The modules Pv(σv)P_{v}(\sigma^{\circ}_{v}) are 𝒪\mathcal{O}-flat by [33, Lemma 2.10] and maximal Cohen-Macaulay over Rv(σv)R_{v}(\sigma_{v}) by [33, Corollary 6.4, 6.5] (here, we use Proposition 6.9). For every vpv\mid p, we fix a maximal regular sequence on Pv(σv)P_{v}(\sigma^{\circ}_{v}) of length dimRv(σv)\dim R_{v}(\sigma_{v}) containing ϖ\varpi (by extending ϖ\varpi to a maximal regular sequence). For a flat compact 𝒪\mathcal{O}-module MM, the functor ^𝒪M-\widehat{\otimes}_{\mathcal{O}}M is exact and thus the union of the regular sequences form a regular sequence of length

1+vp(dimRv(σv)1)=dimRp(σ),1+\sum_{v\mid p}(\dim R_{v}(\sigma_{v})-1)=\dim R_{p}(\sigma),

and thus dpRp(σ)P(σ)=dimRp(σ)\operatorname{dp}_{R_{p}(\sigma)}P({\sigma^{\circ}})=\dim R_{p}(\sigma). Thus, P(σ)P({\sigma^{\circ}}) is maximal Cohen-Macaulay over Rp(σ)R_{p}(\sigma). Therefore, if 𝔭Rp(σ)[1/ϖ]\mathfrak{p}\subset R_{p}(\sigma)[1/\varpi] is a regular closed point, (P(σ))𝔭(P({\sigma^{\circ}}))_{\mathfrak{p}} is free by Lemma 3.13. Finally, to compute the rank, let k(𝔭)=vpk(𝔭v)Lk(\mathfrak{p})=\otimes_{v\mid p}k(\mathfrak{p}_{v})\cong L be the residue field at 𝔭\mathfrak{p}. By [33, Prop. 2.22, 4.14], each Pv(σv)𝔭vP_{v}(\sigma_{v}^{\circ})_{\mathfrak{p}_{v}} is of rank 11 over Rv(σv)𝔭vR_{v}(\sigma_{v})_{\mathfrak{p}_{v}}, and hence

rkRp(σ)𝔭(P(σ)𝔭)=dimk(𝔭)(k(𝔭)Rp(σ)𝔭(P(σ))𝔭)=1.\operatorname{rk}_{R_{p}(\sigma)_{\mathfrak{p}}}\big{(}P({\sigma^{\circ}})_{\mathfrak{p}}\big{)}=\dim_{k(\mathfrak{p})}\big{(}k(\mathfrak{p})\otimes_{R_{p}(\sigma)_{\mathfrak{p}}}(P({\sigma^{\circ}}))_{\mathfrak{p}}\big{)}=1.

Corollary 6.12.

Let 𝔭Rpps[1/ϖ]\mathfrak{p}\subset R^{\operatorname{ps}}_{p}[1/\varpi] be the closed point corresponding to a tuple of absolutely irreducible representations, so that (E𝔅)𝔭M|𝔅|(Rp,ρ)(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}\cong M_{|\mathfrak{B}|}(R_{p,\rho}). Set Rp,ρ(σ)=Rp(σ)𝔭R_{p,\rho}(\sigma)=R_{p}(\sigma)_{\mathfrak{p}}^{\wedge}. Then P𝔅(σ)𝔭P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge} is isomorphic to Rp,ρ(σ)|𝔅|R_{p,\rho}(\sigma)^{\oplus|\mathfrak{B}|} as an (E𝔅)𝔭(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}-module, with (E𝔅)𝔭(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}-action given by matrix multiplication.

Proof.

This follows from Theorem 6.11 and the discussion at the beginning of this section. Indeed, we have

(P𝔅(σ)𝔭(P(σ)𝔭)|𝔅|Rp,ρ(σ)|𝔅|,(P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge}\cong\big{(}P({\sigma^{\circ}})_{\mathfrak{p}}^{\wedge}\big{)}^{\oplus|\mathfrak{B}|}\cong R_{p,\rho}(\sigma)^{\oplus|\mathfrak{B}|},

where the right-most module has the usual right action of M|𝔅|(Rp,ρ)M_{|\mathfrak{B}|}(R_{p,\rho}). ∎

6.6 Completed homology

In this section, we introduce the pp-adically completed homology of the group 𝐆\mathbf{G}. For a survey, see [9].

Fix a tame level Up𝐆(𝔸F,p)U^{p}\subset\mathbf{G}(\mathbb{A}^{\infty,p}_{F}), and let K1KK_{1}\subset K be chosen so that K1UpK_{1}U^{p} is good. Consider the tower of Galois covers

(XUpUp)Up,(X_{U_{p}U^{p}})_{U_{p}},

where UpU_{p} runs over a countable basis of neighbourhoods of the identity, normal in K1K_{1}. This defines a projective system of continuous K1K_{1}-spaces, and XUpUpX_{U_{p}U^{p}} is a K1/UpK_{1}/U_{p}-torsor over XK1UpX_{K_{1}U^{p}}. Since singular homology is a covariant functor, for any choice of coefficients one has a corresponding projective system of homology groups.

Definition 6.13.

We define the completed homology with tame level UpU^{p} as the projective limit

H~(XUp,𝒪)=limUpH(XUpUp,𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O})=\varprojlim_{U_{p}}H_{\ast}(X_{U_{p}U^{p}},\mathcal{O})

where UpU_{p} runs over a countable basis of neighbourhoods of the identity, normal in K1K_{1} for any choice of K1KK_{1}\subset K such that K1UpK_{1}U^{p} is a good subgroup.

Since H(XUpUp,𝒪)H_{\ast}(X_{U_{p}U^{p}},\mathcal{O}) is an 𝒪[K/Up]\mathcal{O}[K/U_{p}]-module, H~(XUp,𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O}) is an 𝒪[[K]]\mathcal{O}[[K]]-module.

Proposition 6.14.

Suppose K0KK_{0}\subset K is a subgroup such that K0UpK_{0}U^{p} is good. Then there is a canonical isomorphism

H~(XUp,𝒪)H(XK0Up,𝒪[[K0]]).\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O})\cong H_{\ast}(X_{K_{0}U^{p}},\mathcal{O}[[K_{0}]]).

Moreover, for any compact open K1KK_{1}\subset K, H~(XUp,𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O}) is a finitely generated 𝒪[[K1]]\mathcal{O}[[K_{1}]]-module.

Proof.

Let K0K_{0} be as in the statement and consider the natural Hecke-equivariant map

Cad(K0Up,𝒪[[K]])limUpCad(KUp,𝒪[K/Up]).C^{\operatorname{ad}}_{\bullet}(K_{0}U^{p},\mathcal{O}[[K]])\to\varprojlim_{U_{p}}C^{\operatorname{ad}}_{\bullet}(KU^{p},\mathcal{O}[K/U_{p}]).

Using our fixed choice of chain homotopy equivalence between the adèlic complex and the Borel-Serre complex, the map above corresponds to a map

CBS(KUp,𝒪[[K]])limUpCBS(KUp,𝒪[K/Up])C^{\operatorname{BS}}_{\bullet}(KU^{p},\mathcal{O}[[K]])\to\varprojlim_{U_{p}}C^{\operatorname{BS}}_{\bullet}(KU^{p},\mathcal{O}[K/U_{p}])

which is an isomorphism since CBS(K)C^{\operatorname{BS}}_{\bullet}(K) consists of free and finitely generated [K]\mathbb{Z}[K]-modules. Now,

CBS(KUp,𝒪[K/Up])limUpCBS(UpUp,𝒪)C^{\operatorname{BS}}_{\bullet}(KU^{p},\mathcal{O}[K/U_{p}])\cong\varprojlim_{U_{p}}C^{\operatorname{BS}}_{\bullet}(U_{p}U^{p},\mathcal{O})

and since CBS(UpUp,𝒪)C^{\operatorname{BS}}_{\bullet}(U_{p}U^{p},\mathcal{O}) consists of compact 𝒪\mathcal{O}-modules, Lemma 3.2 implies

H(limUpCBS(UpUp,𝒪))limUpH(CBS(UpUp,𝒪))H(XUp,𝒪),H_{\ast}\big{(}\varprojlim_{U_{p}}C^{\operatorname{BS}}_{\bullet}(U_{p}U^{p},\mathcal{O})\big{)}\cong\varprojlim_{U_{p}}H_{\ast}\big{(}C^{\operatorname{BS}}_{\bullet}(U_{p}U^{p},\mathcal{O}))\cong H_{\ast}(X_{U^{p}},\mathcal{O}\big{)},

as required. The finite generation of H~(XUp,𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O}) follows from the previous statement, together with the fact that if K1K0K_{1}\subset K_{0} is normal, then 𝒪[[K0]]\mathcal{O}[[K_{0}]] is finitely generated (in fact, free) over 𝒪[[K1]]\mathcal{O}[[K_{1}]]. ∎

Let G=vp𝐆(Fv)G=\prod_{v\mid p}\mathbf{G}(F_{v}). The action of KK on H~(XUp,𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O}) coming from the Iwasawa algebra 𝒪[[K]]\mathcal{O}[[K]] can be extended to an action of GG. For any gGg\in G and UpKU_{p}\subseteq K open compact, right multiplication by gg gives homomorphism

H(XUpUp,𝒪)H(X(gUpg1)Up,𝒪),H_{\ast}(X_{U_{p}U^{p}},\mathcal{O})\to H_{\ast}(X_{(gU_{p}g^{-1})U^{p}},\mathcal{O}),

which induces an endomorphism of the projective limit H~(XUp,𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O}), extending the 𝒪[K]\mathcal{O}[K]-action (see e.g. [16, Remark 3.4.13]). Since we have already established that H~(XUp,𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O}) is finitely generated over 𝒪[[K]]\mathcal{O}[[K]], we obtain the following.

Proposition 6.15.

[9, Theorem 1.1(1)] With GG-action defined as above, H~(XUp,𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O}) is a object of ModGfga(𝒪)\operatorname{Mod}_{G}^{\operatorname{fga}}(\mathcal{O}). In particular, H~(XUp,𝒪)G(𝒪)\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O})\in\mathfrak{C}_{G}(\mathcal{O}).

Almost by definition, the big Hecke algebra acts on completed homology. The action is equivariant for the GG-action; there exists a map

𝕋S(Up)𝔪EndG(𝒪)(H~(XUp,𝒪)𝔪).\mathbb{T}^{S}(U^{p})_{\mathfrak{m}}\to\operatorname{End}_{\mathfrak{C}_{G}(\mathcal{O})}(\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}).
Proposition 6.16.

Let σ=σ(𝐰,τ)=σ(𝐰)σ(τ)\sigma=\sigma(\mathbf{w},\tau)=\sigma(\mathbf{w})\otimes\sigma(\tau) and Kτ=ker(σ(τ))KK_{\tau}=\ker(\sigma(\tau))\subset K. Suppose KUpKU^{p} is good. There is a Hecke-equivariant homological first-quadrant spectral sequence

Ei,j2=Torj𝒪[[K]](H~i(XUp,𝒪),σ)Hi+j(XKτUp,σ(𝐰))[τ]E^{2}_{i,j}=\operatorname{Tor}_{j}^{\mathcal{O}[[K]]}\big{(}\widetilde{H}_{i}(X_{U^{p}},\mathcal{O}),\sigma)\implies H_{i+j}(X_{K_{\tau}U^{p}},\sigma(\mathbf{w})\big{)}[\tau^{\ast}]

where [τ][\tau^{\ast}] denotes the σ(τ)\sigma(\tau)^{\ast}-isotypic component.

Proof.

The quotient K/KτK/K_{\tau} is a finite group. By Proposition 6.14 and our choice of chain homotopy equivalence between the adèlic complex and the Borel-Serre complex, we have isomorphisms

H~(XUp,𝒪)H(Cad(KUp)𝒪[K]𝒪[[K]])H(CBS(KUp)𝒪[K]𝒪[[K]]).\begin{split}\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O})&\cong H_{\ast}\big{(}C^{\operatorname{ad}}_{\bullet}(KU^{p})\otimes_{\mathcal{O}[K]}\mathcal{O}[[K]]\big{)}\\ &\cong H_{\ast}\big{(}C^{\operatorname{BS}}_{\bullet}(KU^{p})\otimes_{\mathcal{O}[K]}\mathcal{O}[[K]]\big{)}.\end{split}

Let C~=CBS(KUp)𝒪[K]𝒪[[K]]\widetilde{C}_{\bullet}=C^{\operatorname{BS}}_{\bullet}(KU^{p})\otimes_{\mathcal{O}[K]}\mathcal{O}[[K]]. This is a complex of finitely generated projective 𝒪[[K]]\mathcal{O}[[K]]-modules, so there is a hyperhomology spectral sequence (see [39, Theorem 5.7.6])

Ei,j2=Torj𝒪[[K]](H~i(XUp,𝒪),σ)Hi+j(C~𝒪[[K]]σ(𝐰)Lσ(τ)).E^{2}_{i,j}=\operatorname{Tor}_{j}^{\mathcal{O}[[K]]}\big{(}\widetilde{H}_{i}(X_{U^{p}},\mathcal{O}),\sigma\big{)}\implies H_{i+j}\big{(}\widetilde{C}_{\bullet}\otimes_{\mathcal{O}[[K]]}\sigma(\mathbf{w})\otimes_{L}\sigma(\mathbf{\tau})\big{)}.

Note that

C~𝒪[[K]]σ(𝐰)Lσ(τ)(C~[𝒪[[Kτ]]]σ(𝐰)Lσ(τ))K/Kτ\widetilde{C}_{\bullet}\otimes_{\mathcal{O}[[K]]}\sigma(\mathbf{w})\otimes_{L}\sigma(\mathbf{\tau})\cong\big{(}\widetilde{C}_{\bullet}\otimes_{[}\mathcal{O}[[K_{\tau}]]]\sigma(\mathbf{w})\otimes_{L}\sigma(\mathbf{\tau})\big{)}_{K/K_{\tau}}

where ()K/Kτ(-)_{K/K_{\tau}} denotes taking coinvariants with respect to the action defined by

k(uu)=ukk1u,k(u\otimes u^{\prime})=uk\otimes k^{-1}u^{\prime},

where uC~nu\in\widetilde{C}_{n} and uσ(𝐰)Lσ(τ)u^{\prime}\in\sigma(\mathbf{w})\otimes_{L}\sigma(\mathbf{\tau}). In characteristic 0, taking coinvariants is an exact functor, and since KτK_{\tau} acts trivially on σ(τ)\sigma(\tau) we have

Hi+j(C~𝒪[[K]]σ(𝐰)Lσ(τ))(Hi+j(C~𝒪[[Kτ]]σ(𝐰))Lσ(τ))K/Kτ,H_{i+j}\big{(}\widetilde{C}_{\bullet}\otimes_{\mathcal{O}[[K]]}\sigma(\mathbf{w})\otimes_{L}\sigma(\mathbf{\tau})\big{)}\cong\big{(}H_{i+j}(\widetilde{C}_{\bullet}\otimes_{\mathcal{O}[[K_{\tau}]]}\sigma(\mathbf{w}))\otimes_{L}\sigma(\tau)\big{)}_{K/K\tau},

which by Schur’s lemma is precisely Hi+j(XKτUp,σ(𝐰))[τ]H_{i+j}(X_{K_{\tau}U^{p}},\sigma(\mathbf{w}))[\tau^{\ast}]. This completes the proof. ∎

In our setting, one expects the spectral sequence above to degenerate at the E2E^{2}-page after localisation at a ‘non-Eisenstein ideal’ 𝔪𝕋S(Up)\mathfrak{m}\subset\mathbb{T}^{S}(U^{p}). The condition that 𝔪\mathfrak{m} is non-Eisenstein means that the representation ρ¯𝔪\bar{\rho}_{\mathfrak{m}} introduced in the next section is absolutely irreducible.

Conjecture 6.17.

Let 𝔪𝕋S(Up)\mathfrak{m}\subset\mathbb{T}^{S}(U^{p}) be a non-Eisenstein maximal ideal. Then

H~(XUp,𝒪)𝔪=H~q0(XUp,𝒪)𝔪.\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}=\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}.
Remark 6.18.

As is shown in [16, Proposition 4.2.1(1)], this vanishing follows from conjectures of Calegari-Emerton [9, Conjecture 1.5] and Calegari-Geraghty [10, Conjecture B(4)(a)]. These conjectures are open in general, but known to hold when l0=1l_{0}=1, i.e. when FF is an imaginary quadratic field.

An immediate consequence of the conjecture and Proposition 6.16 is the isomorphism

Tori𝒪[[K]](H~q0(XUp,𝒪),σ)Hq0+i(XKτUp,σ(𝐰))[τ].\operatorname{Tor}_{i}^{\mathcal{O}[[K]]}\big{(}\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O}),\sigma\big{)}\cong H_{q_{0}+i}(X_{K_{\tau}U^{p}},\sigma(\mathbf{w}))[\tau^{\ast}].

6.7 Galois representations

Suppose that Π\Pi is a regular algebraic cuspidal automorphic representation which is cohomological with respect to an algebraic weight, and let SS be the set of ramified places of Π\Pi and the places above pp. The Hecke-equivariance of the spectral sequence of Proposition 6.16 implies that the action of the abstract Hecke algebra 𝕋S\mathbb{T}^{S} on the homology H(XUpUp,σ(𝐰))H_{\ast}(X_{U_{p}U^{p}},\sigma(\mathbf{w})) factors through the big Hecke algebra 𝕋S(Up)\mathbb{T}^{S}(U^{p}). The assumption that Π\Pi is cohomological means there is a σ(𝐰)\sigma(\mathbf{w}) such that the Π\Pi-part

H(XUpUp,σ(𝐰))𝔑Π,ιH_{\ast}(X_{U_{p}U^{p}},\sigma(\mathbf{w}))_{\mathfrak{N}_{\Pi,\iota}}

is non-zero in degrees [q0,q0+l0][q_{0},q_{0}+l_{0}]. By the observation above, if we let

𝔭=𝔭Π,ι=ker(𝕋S(Up)End(H(XUpUp,σ(𝐰))𝔑Π,ι))\mathfrak{p}=\mathfrak{p}_{\Pi,\iota}=\ker\big{(}\mathbb{T}^{S}(U^{p})\to\operatorname{End}(H_{\ast}(X_{U_{p}U^{p}},\sigma(\mathbf{w}))_{\mathfrak{N}_{\Pi,\iota}})\big{)}

then we have

H(XUpUp,σ(𝐰))𝔑Π,ι=H(XUpUp,σ(𝐰))𝔭.H_{\ast}(X_{U_{p}U^{p}},\sigma(\mathbf{w}))_{\mathfrak{N}_{\Pi,\iota}}=H_{\ast}(X_{U_{p}U^{p}},\sigma(\mathbf{w}))_{\mathfrak{p}}.

Moreover, we have a maximal ideal

𝔪=𝔪Π,ι=(𝔭,ϖ)𝕋S(Up),\mathfrak{m}=\mathfrak{m}_{\Pi,\iota}=(\mathfrak{p},\varpi)\subset\mathbb{T}^{S}(U^{p}),

to which we associate a residual representation

ρ¯𝔪:GF,SGL2(𝕋S(Up)/𝔪)\bar{\rho}_{\mathfrak{m}}:G_{F,S}\to\operatorname{GL}_{2}(\mathbb{T}^{S}(U^{p})/\mathfrak{m})

using the main result of [34]. By [12, Theorem 6.1.4], we may lift ρ¯𝔪\bar{\rho}_{\mathfrak{m}} to a representation

ρ𝔪:GF,SGL2(𝕋S(Up)𝔪)\rho_{\mathfrak{m}}:G_{F,S}\to\operatorname{GL}_{2}(\mathbb{T}^{S}(U^{p})_{\mathfrak{m}})

with determinant ε1\varepsilon^{-1}. This defines a surjective map

R𝒮𝕋S(Up)𝔪,R_{\mathcal{S}}\twoheadrightarrow\mathbb{T}^{S}(U^{p})_{\mathfrak{m}},

where 𝒮\mathcal{S} is the global deformation problem 𝒮=(ρ¯𝔪,S,ε,Dvε,)\mathcal{S}=(\bar{\rho}_{\mathfrak{m}},S,\varepsilon,D_{v}^{\varepsilon,\square}) corresponding to deformations of ρ¯𝔪\bar{\rho}_{\mathfrak{m}} that are unramified outside of SS. We denote by ρ=ρΠ,ι\rho=\rho_{\Pi,\iota} the composition

GF,Sρ𝔪GL2(𝕋S(Up)𝔪)GL2(L),G_{F,S}\stackrel{{\scriptstyle\rho_{\mathfrak{m}}}}{{\to}}\operatorname{GL}_{2}(\mathbb{T}^{S}(U^{p})_{\mathfrak{m}})\to\operatorname{GL}_{2}(L),

where LL is the residue field of 𝕋S(Up)𝔪\mathbb{T}^{S}(U^{p})_{\mathfrak{m}} at 𝔭\mathfrak{p} (if necessary, we replace LL by a finite extension). Thus, ρ\rho defines a maximal ideal 𝔭R𝒮[1/ϖ]\mathfrak{p}\subset R_{\mathcal{S}}[1/\varpi].

For every vSpv\in S_{p}, after twisting with ε¯\bar{\varepsilon} we obtain a pseudorepresentation tr(ρ¯𝔪,vε¯)\operatorname{tr}(\bar{\rho}_{\mathfrak{m},v}\otimes\bar{\varepsilon}) of trivial determinant which defines a block 𝔅𝔪,v\mathfrak{B}_{\mathfrak{m},v} of ModGL2(Fv),1pfa(𝒪)\operatorname{Mod}^{\operatorname{pfa}}_{\operatorname{GL}_{2}(F_{v}),1}(\mathcal{O}). We obtain a block of G(𝒪)\mathfrak{C}_{G}(\mathcal{O}) by forming the tensor product as in Proposition 6.6, i.e.

𝔅𝔪=vp𝔅𝔪,v.\mathfrak{B}_{\mathfrak{m}}=\otimes_{v\mid p}\mathfrak{B}_{\mathfrak{m},v}.

7 Main results

We keep the notation from the preceding section, and from now on make the following assumptions.

  • (i)

    ρ¯𝔪:GF,SGL2(k)\bar{\rho}_{\mathfrak{m}}:G_{F,S}\to\operatorname{GL}_{2}(k) is absolutely irreducible and the restriction ρ¯𝔪|GF(ζp)\bar{\rho}_{\mathfrak{m}}|_{G_{F(\zeta_{p})}} has adequate image (see [37, Definition 2.3]).

  • (ii)

    The characteristic 0 representations ρv\rho_{v} for all vSpv\in S_{p} are irreducible of vv-adic Hodge type (𝐰v,τv,χv){(\mathbf{w}_{v},\tau_{v},\chi_{v})}.

  • (iii)

    The local characteristic 0 representations ρv\rho_{v} have generic associated Weil-Deligne representations for all vSv\in S.

7.1 The patching argument

The goal of this subsection is to prove the following theorem. The notation R𝒮,ρR_{\mathcal{S},\rho} was introduced in Theorem 5.18.

Theorem 7.1.

With the same notation as in Section 6.7 and under Conjectures 6.17 and Conjecture 7.5 (see below), consider the E𝔅E_{\mathfrak{B}}-module

m0=Hom𝔅(𝒪)(P𝔅,H~q0(XUp,𝒪)𝔪).m_{0}=\operatorname{Hom}_{\mathfrak{C}_{\mathfrak{B}}(\mathcal{O})}\big{(}P_{\mathfrak{B}},\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}\big{)}.

Then (m0)𝔭(m_{0})_{\mathfrak{p}}^{\wedge} is free as an R𝒮,ρR_{\mathcal{S},\rho}-module, and hence isomorphic as an (E𝔅)𝔭(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}-module to a direct sum of modules of the form R𝒮,ρ|𝔅|R_{\mathcal{S},\rho}^{\oplus|\mathfrak{B}|} with (E𝔅)𝔭(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}-action defined by matrix multiplication.

This result is a characteristic 0 analogue of [16, Conjecture 5.1.2]. Note that our setting differs in that we have no ‘minimal level’ assumption ensuring the smoothness of the local framed deformation rings RvR_{v}^{\square} for vSSpv\in S\setminus S_{p}. For our purposes, it is sufficient to assume that the associated Weil-Deligne representations of ρv\rho_{v} are generic at all places vSSpv\in S\setminus S_{p}, as it ensures the restrictions ρv\rho_{v} define smooth points in the generic fibers of RvR_{v}^{\square} (Theorem 5.12).

Our proof of Theorem 7.1 is based on the construction of ‘patched completed homology’ in [16]. The strategy is to first prove an analogous result (Theorem 7.9) at ‘patched’ level and then ‘unpatch’ to deduce Theorem 7.1. Before we can state the patched analogue, we need to recall the construction of patched completed homology. We follow [16] with only slight adjustments.

To begin, we note that the assumption that ρ¯(GF(ζp))\bar{\rho}(G_{F(\zeta_{p})}) is adequate is equivalent to it being enormous ([16, Lemma 3.2.3]), and we let qq be an integer large enough to guarantee the existence of Taylor-Wiles primes as in [16, Lemma 3.3.1]. We let 𝒯\mathcal{T} be the power series ring over 𝒪\mathcal{O} in the SS-frame variables, i.e.

𝒯=𝒪[[{Xi,jvvS,i,j=1,2}]]/(X1,1v0),\mathcal{T}=\mathcal{O}[[\{X^{v}_{i,j}\mid v\in S,\ i,j=1,2\}]]/(X^{v_{0}}_{1,1}),

where v0v_{0} is an arbitrary element of SS. Then 𝒯\mathcal{T} is of relative dimension 4|S|14|S|-1 over 𝒪\mathcal{O}, and we define

𝒪=𝒯^𝒪[[(p×)q]],\mathcal{O}_{\infty}=\mathcal{T}\widehat{\otimes}\mathcal{O}[[(\mathbb{Z}_{p}^{\times})^{q}]],

a power series ring over 𝒪\mathcal{O} in 4|S|1+q4|S|-1+q variables. Let 𝐚=ker(𝒪𝒪)\mathbf{a}=\ker(\mathcal{O}_{\infty}\to\mathcal{O}) be the augmentation ideal of 𝒪\mathcal{O}_{\infty}, and set

S=Rpps^𝒪.S_{\infty}=R^{\operatorname{ps}}_{p}\widehat{\otimes}\mathcal{O}_{\infty}.

The construction features a second ring denoted RR_{\infty} which is as a power series ring over R𝒮S,loc=^vSRvR^{S,\operatorname{loc}}_{\mathcal{S}}=\widehat{\bigotimes}_{v\in S}R_{v}^{\square} in

|S|1+q[F:]l0|S|-1+q-[F:\mathbb{Q}]-l_{0}

variables, and it is constructed in such a way that there are maps

𝒪RR𝒮,\mathcal{O}_{\infty}\to R_{\infty}\twoheadrightarrow R_{\mathcal{S}},

where R𝒮R_{\mathcal{S}} is the ring representing the functor of type 𝒮\mathcal{S} deformations. Since RR_{\infty} is a RpR_{p}^{\square}-algebra, it is also an RppsR^{\operatorname{ps}}_{p}-algebra, and using this map we replace the diagram above with

S=Rpps^𝒪𝒪φRR𝒮𝕋S(Up).S_{\infty}=R^{\operatorname{ps}}_{p}\widehat{\otimes}_{\mathcal{O}}\mathcal{O}_{\infty}\stackrel{{\scriptstyle\varphi}}{{\to}}R_{\infty}\twoheadrightarrow R_{\mathcal{S}}\twoheadrightarrow\mathbb{T}^{S}(U^{p}).

The proof of Theorem 7.1 is based on analysing this diagram. One expects the final arrow to be an isomorphism, and we will confirm this expectation after localisation and completion at the point 𝔭R𝒮[1/ϖ]\mathfrak{p}\subset R_{\mathcal{S}}[1/\varpi] corresponding to ρ\rho.

Consider the diagram

SφRR𝒮LS_{\infty}\stackrel{{\scriptstyle\varphi}}{{\to}}R_{\infty}\twoheadrightarrow R_{\mathcal{S}}\to L

where the final map corresponds to ρ\rho. We define

𝔭\displaystyle\mathfrak{p}_{\infty} =ker(RL),\displaystyle=\ker(R_{\infty}\to L),
𝔮\displaystyle\mathfrak{q}_{\infty} =ker(SRL),\displaystyle=\ker(S_{\infty}\to R_{\infty}\to L),

so that 𝔮=φ1(𝔭)\mathfrak{q}_{\infty}=\varphi^{-1}(\mathfrak{p}_{\infty}) and we have a homomorphism (S)𝔮(R)𝔭(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}\to(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}.

Lemma 7.2.

The rings (S)𝔮(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge} and (R)𝔭(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} are regular and

dim(R)𝔭=dim(S)𝔮[F:]l0.\dim(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}=\dim(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}-[F:\mathbb{Q}]-l_{0}.
Proof.

For the purpose of this proof only, define the auxiliary ring

T=𝒪[[y1,,y|S|1+q[F:]l0]],T=\mathcal{O}[[y_{1},\dots,y_{|S|-1+q-[F:\mathbb{Q}]-l_{0}}]],

so that R=R𝒮S,loc^TR_{\infty}=R^{S,\operatorname{loc}}_{\mathcal{S}}\widehat{\otimes}T and S=Rpps^𝒪S_{\infty}=R^{\operatorname{ps}}_{p}\widehat{\otimes}\mathcal{O}_{\infty} are rings of formal power series with coefficients in R𝒮S,locR^{S,\operatorname{loc}}_{\mathcal{S}} and RppsR^{\operatorname{ps}}_{p}, respectively. The regularity of the points 𝔭,𝔮\mathfrak{p}_{\infty},\mathfrak{q}_{\infty} follows from that of the corresponding points of R𝒮S,locR^{S,\operatorname{loc}}_{\mathcal{S}} and RppsR^{\operatorname{ps}}_{p}. Using [3, Lemma 3.3] and Theorem 5.13, we find that the dimensions are

dim(R)𝔭=vSdim(Rv)𝔭v+|S|1+q[F:]l0]=3|SSp|+3|Sp|+vpdimRρv+|S|1+q[F:]l0=vpdimRρv+4|S|1+q[F:]l0,\begin{split}\dim(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}&=\sum_{v\in S}\dim(R_{v}^{\square})_{\mathfrak{p}_{v}}+|S|-1+q-[F:\mathbb{Q}]-l_{0}]\\ &=3|S\setminus S_{p}|+3|S_{p}|+\sum_{v\mid p}\dim R_{\rho_{v}}+|S|-1+q-[F:\mathbb{Q}]-l_{0}\\ &=\sum_{v\mid p}\dim R_{\rho_{v}}+4|S|-1+q-[F:\mathbb{Q}]-l_{0},\end{split}

and

dim(S)𝔮=vpdimRρv+dim𝒪1=vpdimRρv+4|S|1+q.\begin{split}\dim(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}&=\sum_{v\mid p}\dim R_{\rho_{v}}+\dim\mathcal{O}_{\infty}-1\\ &=\sum_{v\mid p}\dim R_{\rho_{v}}+4|S|-1+q.\end{split}

The result follows. ∎

So far, we have only mentioned the rings involved in the patching argument. Let us now recall the key features for us of the complex on which these rings act. Note that the ‘minimal level’ assumption present in [16] is not required to prove the results cited below.

Theorem 7.3.

There exists a perfect complex 𝒞~()\widetilde{\mathcal{C}}(\infty) of 𝒪[[K]]\mathcal{O}_{\infty}[[K]]-modules with a compatible GG-action such that the following hold.

  • (i)

    The action of 𝒪\mathcal{O}_{\infty} on 𝒞~()\widetilde{\mathcal{C}}(\infty) factors through the map 𝒪R\mathcal{O}_{\infty}\to R_{\infty}.

  • (ii)

    H(𝒞~())H_{\ast}(\widetilde{\mathcal{C}}(\infty)) lies in G(𝒪)\mathfrak{C}_{G}(\mathcal{O}).

  • (iii)

    Let 𝐚=ker(𝒪𝒪)\mathbf{a}=\ker(\mathcal{O}_{\infty}\to\mathcal{O}). There is a GG-equivariant isomorphism of 𝒪[[K]]\mathcal{O}_{\infty}[[K]]-modules.

    Hi(𝒞~()𝒪𝒪/𝐚)H~i(XUp,𝒪)𝔪.H_{i}(\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a})\cong\widetilde{H}_{i}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}.
  • (iv)

    Assume Conjecture 6.17. Then 𝒞~()\widetilde{\mathcal{C}}(\infty) has homology concentrated in degree q0q_{0}.

Proof.

The complex is constructed in [16, Section 3.4]. For (i), (iii) and (iv), see [16, Proposition 3.4.16(2), Remark 3.4.17, Proposition 3.4.19, Proposition 4.2.1]. To prove (ii), first note that since 𝒞~()\widetilde{\mathcal{C}}(\infty) is a perfect complex of 𝒪[[K]]\mathcal{O}_{\infty}[[K]]-modules,

H(𝒞~())limnH(𝒞~()𝒪𝒪/𝐚n).H_{\ast}(\widetilde{\mathcal{C}}(\infty))\cong\varprojlim_{n}H_{\ast}(\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a}^{n}).

The category G(𝒪)\mathfrak{C}_{G}(\mathcal{O}) has projective limits ([32, p.14]) and hence it is enough to prove that each term in the above limit is an element of G(𝒪)\mathfrak{C}_{G}(\mathcal{O}). We proceed by induction on nn, noting that the case n=1n=1 follows from (iii) and Proposition 6.15. For every n1n\geq 1, we have a short eact sequence of chain complexes

0𝒞~()𝒪𝐚n1/𝐚n𝒞~()𝒪𝒪/𝐚n𝒞~()𝒪𝒪/𝐚n100\to\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathbf{a}^{n-1}/\mathbf{a}^{n}\to\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a}^{n}\to\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a}^{n-1}\to 0

inducing a long exact sequence of homology groups

Hi(𝒞~()𝒪𝐚n1/𝐚n)Hi(𝒞~()𝒪𝒪/𝐚n)H(𝒞~()𝒪𝒪/𝐚n1),\dots\to H_{i}(\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathbf{a}^{n-1}/\mathbf{a}^{n})\to H_{i}(\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a}^{n})\to H_{\ast}(\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a}^{n-1})\to\dots,

which in turn decomposes into short exact sequences in the usual way. The category G(𝒪)\mathfrak{C}_{G}(\mathcal{O}) is closed under kernels, cokernels and extensions (in ModG,ζpfa(𝒪)\operatorname{Mod}^{\operatorname{pfa}}_{G,\zeta}(\mathcal{O})). Indeed, G(𝒪)\mathfrak{C}_{G}(\mathcal{O}) is abelian and moreover the inclusion ModGladm(𝒪)ModG,ζsm(𝒪)\operatorname{Mod}^{\operatorname{ladm}}_{G}(\mathcal{O})\hookrightarrow\operatorname{Mod}^{\operatorname{sm}}_{G,\zeta}(\mathcal{O}) preserves injectives (see [32, Corollary 5.18]), so that for any locally admissible V,WV,W we have a canonical isomorphism Extladm1(V,W)Extsm1(V,W)\operatorname{Ext}^{1}_{\operatorname{ladm}}(V,W)\cong\operatorname{Ext}^{1}_{\operatorname{sm}}(V,W). Viewing these groups as parametrising extensions, we obtain the claimed closedness of G(𝒪)\mathfrak{C}_{G}(\mathcal{O}) under extensions. Finally, by direct observation we see that

𝒞~()𝒪𝐚n1/𝐚(𝒞~()𝒪𝒪/𝐚)r\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathbf{a}^{n-1}/\mathbf{a}\cong(\widetilde{\mathcal{C}}(\infty)\otimes_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a})^{\oplus r}

for some rr depending on nn. The right-hand side has homology equal to a direct sum of copies of H~q0(XUp,𝒪)𝔪\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}, and the theorem now follows by induction on nn. ∎

Remark 7.4.

In light of item (ii), we think of 𝒞~()\widetilde{\mathcal{C}}(\infty) as a complex computing ‘patched completed homology’.

Neither H~q0(XUp,𝒪)\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O}) nor Hq0(𝒞~())H_{q_{0}}(\widetilde{\mathcal{C}}(\infty)) are finitely generated over the big Hecke algebra, so to carry out the depth estimate part of the Taylor-Wiles method, we work instead with the respective images in the category of E𝔅E_{\mathfrak{B}}-modules. This idea can only work under the following additional assumption,

Conjecture 7.5.

The two actions of RppsR^{\operatorname{ps}}_{p} on Hq0(𝒞~())H_{q_{0}}(\widetilde{\mathcal{C}}(\infty)) – one coming from the map RppsE𝔅R^{\operatorname{ps}}_{p}\to E_{\mathfrak{B}} and the other from the map Rpps𝕋S(Up)𝔪R^{\operatorname{ps}}_{p}\to\mathbb{T}^{S}(U^{p})_{\mathfrak{m}} – coincide.

Remark 7.6.

In [31, §3.5], a similar statement in a different setting is interpreted as a form of local-global compatibility at vpv\mid p. It seems a reasonable guess that under suitable assumptions the conjecture can be verified in our setting using recent results of Hevesi on local-global compatibility for completed homology [20].

From now on, we assume Conjecture 7.5. We define

m0\displaystyle m_{0} =HomG(𝒪)(P𝔅,H~q0(XUp,𝒪)𝔪),\displaystyle=\operatorname{Hom}_{\mathfrak{C}_{G}(\mathcal{O})}\big{(}P_{\mathfrak{B}},\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}\big{)},
m\displaystyle m_{\infty} =HomG(𝒪)(P𝔅,Hq0(𝒞~())).\displaystyle=\operatorname{Hom}_{\mathfrak{C}_{G}(\mathcal{O})}\big{(}P_{\mathfrak{B}},H_{q_{0}}(\widetilde{\mathcal{C}}(\infty))\big{)}.

Note that, as an RppsR^{\operatorname{ps}}_{p}-module,

m0π𝔅HomG(𝒪)(Pπ,H~q0(XUp,𝒪)𝔪)m_{0}\cong\bigoplus_{\pi\in\mathfrak{B}}\operatorname{Hom}_{\mathfrak{C}_{G}(\mathcal{O})}\big{(}P_{\pi},\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}\big{)}

and similarly for mm_{\infty}. By Theorem 7.3(i) and Conjecture 7.5, mm_{\infty} is an SS_{\infty}-module such that the action of SS_{\infty} factors through the map SφRS_{\infty}\stackrel{{\scriptstyle\varphi}}{{\to}}R_{\infty}.

Proposition 7.7.

With m0,mm_{0},m_{\infty} as above, we have:

  • (i)

    m𝒪𝐋𝒪/𝐚m0m_{\infty}\otimes^{\mathbf{L}}_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a}\cong m_{0}.

  • (ii)

    m0m_{0} is a finitely generated RppsR^{\operatorname{ps}}_{p}-module.

  • (iii)

    mm_{\infty} is finitely generated as an SS_{\infty}-module and as an RR_{\infty}-module.

Proof.

(i) follows from Theorem 7.3. Note that since H~q0(XUp,𝒪)𝔪\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}} is finitely generated over 𝒪[[K]]\mathcal{O}[[K]], m0m_{0} is finitely generated over E𝔅E_{\mathfrak{B}} by Lemma 6.5. Now, E𝔅E_{\mathfrak{B}} is finitely generated over RppsR^{\operatorname{ps}}_{p} (Proposition 6.6) and (ii) follows. Since the SS_{\infty}-action factors through SRS_{\infty}\to R_{\infty}, the first statement implies the second. From (i) we know that m𝒪𝒪m0m_{\infty}\otimes_{\mathcal{O}_{\infty}}\mathcal{O}\cong m_{0}, and therefore by Nakayama’s lemma for compact modules (see [8, Corollary 1.5]) it suffices to prove that mm_{\infty} is a compact module, which follows from that mm_{\infty} is an inverse limit of finitely generated modules. Thus, we have proved (iii). ∎

We obtain a corresponding diagram to what we had before:

SREndE𝔅(m).S_{\infty}\to R_{\infty}\to\operatorname{End}_{E_{\mathfrak{B}}}(m_{\infty}).

Localising and completing the diagram we obtain

(S)𝔮(R)𝔭End(E𝔅)𝔭((m)𝔭).(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}\to(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}\to\operatorname{End}_{(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}}((m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}).
Lemma 7.8.

The module (m)𝔭(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} is finitely generated over (R)𝔭(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} and (S)𝔮(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}.

Proof.

The first statement follows from the fact that mm_{\infty} is finitely generated over RR_{\infty}, which is Proposition 7.7(ii). For the second, note that since 𝔭\mathfrak{p}_{\infty} is maximal in R[1/ϖ]R_{\infty}[1/\varpi], we have

(m)𝔭limrm[1/ϖ]/𝔭r.(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}\cong\varprojlim_{r}m_{\infty}[1/\varpi]/\mathfrak{p}_{\infty}^{r}.

Now, by definition 𝔮=φ1(𝔭)\mathfrak{q}_{\infty}=\varphi^{-1}(\mathfrak{p}_{\infty}) and hence φ(𝔮r)𝔭r\varphi(\mathfrak{q}_{\infty}^{r})\subseteq\mathfrak{p}_{\infty}^{r}, so that the finitely generated (S)𝔮(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}-module (m)𝔮(m_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge} surjects onto (m)𝔭(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}. The statement follows. ∎

The following theorem is the patched counterpart of Theorem 7.1.

Theorem 7.9.

(m)𝔭(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} is a finitely generated and free (R)𝔭(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}-module.

Proof.

Let (𝐰,τ,ε)vSp(\mathbf{w},\tau,\varepsilon)_{v\in S_{p}} be the vv-adic Hodge types such that

σ=vpσ(𝐰v,τv,ε)\sigma=\otimes_{v\mid p}\ \sigma(\mathbf{w}_{v},\tau_{v},\varepsilon)

is the KK-module for which Π\Pi contributes (in degrees [q0,q0+l0][q_{0},q_{0}+l_{0}]) to H(XUpUp,σ)H_{\ast}(X_{U_{p}U^{p}},\sigma). We define, as before,

Rp,ρ(σ)=(^vpRv(σv,τv))𝔭R_{p,\rho}(\sigma)=\big{(}\widehat{\bigotimes}_{v\mid p}R_{v}(\sigma_{v},\tau_{v})\big{)}_{\mathfrak{p}}^{\wedge}

and set

𝐚σ,𝔭=ker((S)𝔮Rp,ρ(σ)).\mathbf{a}_{\sigma,\mathfrak{p}}=\ker\big{(}(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}\twoheadrightarrow R_{p,\rho}(\sigma)\big{)}.

By Theorem 5.16 and Lemma 7.2, the rings (S)𝔮(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge} and Rp,ρ(σ)R_{p,\rho}(\sigma) are regular. Thus 𝐚σ,𝔭\mathbf{a}_{\sigma,\mathfrak{p}} is generated by a regular sequence of length

dim(S)𝔮dimRp,ρ(σ)=dim(S)𝔮[F:].\dim(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}-\dim R_{p,\rho}(\sigma)=\dim(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}-[F:\mathbb{Q}].

We are going to combine Theorem 2.10 and Lemma 3.10 to prove that (m)𝔭(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} is maximal Cohen-Macaulay over the regular local ring (R)𝔭(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}.

Lemma 7.10.

We have a canonical isomorphism

((m0)𝔭(E𝔅)𝔭𝐋P𝔅(σ)𝔭))|𝔅|2(m0)𝔭𝐋Rp,ρP𝔅(σ)𝔭.\big{(}(m_{0})_{\mathfrak{p}}^{\wedge}\otimes^{\mathbf{L}}_{(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}}P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge})\big{)}^{\oplus|\mathfrak{B}|^{2}}\cong(m_{0})_{\mathfrak{p}}^{\wedge}\otimes^{\mathbf{L}}_{R_{p,\rho}}P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge}.
Proof.

Since P𝔅=π𝔅PπP_{\mathfrak{B}}=\oplus_{\pi\in\mathfrak{B}}P_{\pi}, we have

m0=π𝔅HomG(Pπ,H~q0(XUp,𝒪)𝔪).m_{0}=\oplus_{\pi\in\mathfrak{B}}\operatorname{Hom}_{G}(P_{\pi},\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}).

By Proposition 6.6, (E𝔅)𝔭(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge} is a rank |𝔅||\mathfrak{B}| matrix algebra over Rp,ρR_{p,\rho} and consequently the module (m0)𝔭(m_{0})_{\mathfrak{p}}^{\wedge} is a sum of |𝔅||\mathfrak{B}| pairwise isomorphic Rp,ρR_{p,\rho}-modules. The same is true for P𝔅(σ)𝔭P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge}. For any commutative ring RR, matrix algebra E=Mn(R)E=M_{n}(R), and RR-modules M,NM,N, one has

MR𝐋NMnE𝐋NnM\otimes^{\mathbf{L}}_{R}N\cong M^{\oplus n}\otimes^{\mathbf{L}}_{E}N^{\oplus n}

where Mn,NnM^{\oplus n},N^{\oplus n} are right and left EE-modules, respectively. Thus, we have

(m0)𝔭Rp,ρ𝐋P𝔅(σ)𝔭π1𝔅π2𝔅HomG(𝒪)(Pπ1,H~q0(XUp,𝒪)𝔪)𝔭Rp,ρ𝐋Pπ2(σ)𝔭π1𝔅π2𝔅(m0)𝔭(E𝔅)𝔭𝐋P𝔅(σ)𝔭((m0)𝔭(E𝔅)𝔭𝐋P𝔅(σ)𝔭))|𝔅|2,\begin{split}(m_{0})_{\mathfrak{p}}^{\wedge}\otimes^{\mathbf{L}}_{R_{p,\rho}}P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge}&\cong\bigoplus_{\pi_{1}\in\mathfrak{B}}\bigoplus_{\pi_{2}\in\mathfrak{B}}\operatorname{Hom}_{\mathfrak{C}_{G}(\mathcal{O})}\big{(}P_{\pi_{1}},\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}\big{)}_{\mathfrak{p}}^{\wedge}\otimes^{\mathbf{L}}_{R_{p,\rho}}P_{\pi_{2}}(\sigma)_{\mathfrak{p}}^{\wedge}\\ &\cong\bigoplus_{\pi_{1}\in\mathfrak{B}}\bigoplus_{\pi_{2}\in\mathfrak{B}}(m_{0})_{\mathfrak{p}}^{\wedge}\otimes^{\mathbf{L}}_{(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}}P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge}\\ &\cong((m_{0})_{\mathfrak{p}}^{\wedge}\otimes^{\mathbf{L}}_{(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}}P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge}))^{\oplus|\mathfrak{B}|^{2}},\end{split}

as claimed. ∎

Taking the derived quotient of (m)𝔭(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} over (S)𝔮(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge} with respect to 𝐚σ,𝔭\mathbf{a}_{\sigma,\mathfrak{p}} gives the following chain of isomorphisms:

(m)𝔭(S)𝔮𝐋(S)𝔮/𝐚σ,𝔭(m0)𝔭Rp,ρ𝐋Rp,ρ(σ)((m0E𝔅𝐋P𝔅𝒪[[K]]𝐋σ)𝔭)|𝔅|2((H~q0(XUp,𝒪)𝔪𝒪[[K]]𝐋σ)𝔭)|𝔅|2\begin{split}&(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}\otimes^{\mathbf{L}}_{(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}}(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}/\mathbf{a}_{\sigma,\mathfrak{p}}\\ \cong\ &(m_{0})_{\mathfrak{p}}^{\wedge}\otimes^{\mathbf{L}}_{R_{p,\rho}}R_{p,\rho}(\sigma)\\ \cong\ &\big{(}(m_{0}\otimes^{\mathbf{L}}_{E_{\mathfrak{B}}}P_{\mathfrak{B}}\otimes^{\mathbf{L}}_{\mathcal{O}[[K]]}{\sigma^{\circ}})_{\mathfrak{p}}^{\wedge}\big{)}^{\oplus|\mathfrak{B}|^{2}}\\ \cong\ &\big{(}(\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}\otimes^{\mathbf{L}}_{\mathcal{O}[[K]]}{\sigma^{\circ}})_{\mathfrak{p}}^{\wedge}\big{)}^{\oplus|\mathfrak{B}|^{2}}\end{split}

By Proposition 6.16 and Theorem 2.10, this complex has homology concentrated in degrees [q0,q0+l0][q_{0},q_{0}+l_{0}], and thus we have proved

Tori(S)𝔮((m)𝔭,Rp,ρ(σ))=0 for i[0,l0].\operatorname{Tor}_{i}^{(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}}\big{(}(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge},R_{p,\rho}(\sigma)\big{)}=0\text{ for }i\notin[0,l_{0}].

Applying Lemma 3.10, we obtain

dp(S)𝔮(m)𝔭dp𝐚σ,𝔭(m)𝔭=dim(S)𝔮[F:]l0=dim(R)𝔭.\begin{split}\operatorname{dp}_{(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}}(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}&\geq\operatorname{dp}_{\mathbf{a}_{\sigma,\mathfrak{p}}}(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}\\ &=\dim(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}-[F:\mathbb{Q}]-l_{0}\\ &=\dim(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}.\end{split}

Since the action of (S)𝔮(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge} on (m)𝔭(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} factors through (R)𝔭(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}, any regular sequence in (S)𝔮(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge} gives rise to a regular sequence in (R)𝔭(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}, and hence

dp(R)𝔭(m)𝔭dp(S)𝔮(m)𝔭dim(R)𝔭.\begin{split}\operatorname{dp}_{(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}}(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}&\geq\operatorname{dp}_{(S_{\infty})_{\mathfrak{q}_{\infty}}^{\wedge}}(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}\\ &\geq\dim(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}.\end{split}

Thus, (m)𝔭(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} is a maximal Cohen-Macaulay module over the regular local ring (R)𝔭(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}, and Theorem 7.9 now follows from 3.13. ∎

Having proved Theorem 7.9, we now deduce Theorem 7.1.

Corollary 7.11.

The natural surjections

(R/𝐚)𝔭R𝒮,ρ𝕋S(Up)𝔭(R_{\infty}/\mathbf{a})_{\mathfrak{p}_{\infty}}^{\wedge}\twoheadrightarrow R_{\mathcal{S},\rho}\twoheadrightarrow\mathbb{T}^{S}(U^{p})_{\mathfrak{p}}^{\wedge}

are isomorphisms. In particular, Theorem 7.1 holds.

Proof.

By Theorem 7.9, (m)𝔭(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge} is free over (R)𝔭(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}, and hence the unpatched module

(m0)𝔭=(m)𝔭𝒪𝒪/𝐚(m_{0})_{\mathfrak{p}}^{\wedge}=(m_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}\otimes_{\mathcal{O}_{\infty}}\mathcal{O}_{\infty}/\mathbf{a}

is free over (R)𝔭/𝐚(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}/\mathbf{a}. But the action of (R)𝔭/𝐚(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}/\mathbf{a} on (m0)𝔭(m_{0})_{\mathfrak{p}}^{\wedge} factors through the maps displayed in the statement of the theorem, which forces them to be injective. Thus both maps are bijections and (m0)𝔭(m_{0})_{\mathfrak{p}}^{\wedge} is free over (R)𝔭/𝐚R𝒮,ρ(R_{\infty})_{\mathfrak{p}_{\infty}}^{\wedge}/\mathbf{a}\cong R_{\mathcal{S},\rho}. ∎

7.2 The main theorem

In this section, we deduce our main theorem from Theorem 7.9.

Theorem 7.12.

Let FF be a CM field, suppose p5p\geq 5 is totally split in FF and let Π\Pi be a regular algebraic cuspidal automorphic representation of 𝐆=PGL2/F\mathbf{G}=\operatorname{PGL}_{2}/F which contributes to homology with weight σ(𝐰)\sigma(\mathbf{w}), and fix an isomorphism ι:¯p\iota:\overline{\mathbb{Q}}_{p}\to\mathbb{C}. Let τ\tau be the inertial type of Π\Pi determined by the local Langlands correspondence, set Kτ=ker(σ(τ))K_{\tau}=\ker(\sigma(\tau)) and suppose Up𝐆(𝔸F)U^{p}\subset\mathbf{G}(\mathbb{A}^{\infty}_{F}) is a good tame level subgroup. Let 𝔪=𝔪Π,ι𝕋S(Up)\mathfrak{m}=\mathfrak{m}_{\Pi,\iota}\subset\mathbb{T}^{S}(U^{p}) be the maximal ideal of the big Hecke algebra associated to (Π,ι)(\Pi,\iota) and its associated Galois representation by ρ𝔪:GF,SGL2(𝕋S(Up)𝔪)\rho_{\mathfrak{m}}:G_{F,S}\to\operatorname{GL}_{2}(\mathbb{T}^{S}(U^{p})_{\mathfrak{m}}). Let ρ=ρΠ,ι:GF,SGL2(L)\rho=\rho_{\Pi,\iota}:G_{F,S}\to\operatorname{GL}_{2}(L) be the characteristic 0 representation associated to (Π,ι)(\Pi,\iota).

Suppose the following statements hold.

  • (i)

    The residual representation ρ¯𝔪:GF,SGL2(k)\bar{\rho}_{\mathfrak{m}}:G_{F,S}\to\operatorname{GL}_{2}(k) is absolutely irreducible and the restriction ρ¯𝔪|GF(ζp)\bar{\rho}_{\mathfrak{m}}|_{G_{F(\zeta_{p})}} has adequate image.

  • (ii)

    The characteristic 0 representations ρv\rho_{v} are irreducible of vv-adic Hodge type (𝐰v,τv,χv){(\mathbf{w}_{v},\tau_{v},\chi_{v})}, for all vSpv\in S_{p}.

  • (iii)

    The local characteristic 0 representations ρv\rho_{v} have generic associated Weil-Deligne representations for all vSv\in S.

  • (iv)

    H~(XUp,𝒪)𝔪\widetilde{H}_{\ast}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}} vanishes outside degree q0q_{0} (Conjecture 6.17).

  • (v)

    Local-global compatibility in the sense of Conjecture 7.5 holds.

  • (vi)

    The adjoint Bloch-Kato Selmer group vanishes, i.e. Hf1(GF,ad0ρ)=0H^{1}_{f}(G_{F},\operatorname{ad}^{0}\rho)=0.

Then the graded LL-vector space

H(XKτUp,σ)𝔭[τ]H_{\ast}(X_{K_{\tau}U^{p}},\sigma)_{\mathfrak{p}}[\tau^{\ast}]

has a canonical structure of finitely generated and free graded module over the Tor\operatorname{Tor}-algebra

TorRp,ρ(R𝒮,ρ,Rp,ρ(σ)).\operatorname{Tor}_{\ast}^{R_{p,\rho}}\big{(}R_{\mathcal{S},\rho},R_{p,\rho}(\sigma)\big{)}.

Suppose in addition the following.

  • (vii)

    H2(GF,S,ad0ρ)=0H^{2}(G_{F,S},\operatorname{ad}^{0}\rho)=0, i.e. the ring R𝒮,ρR_{\mathcal{S},\rho} is smooth.

Then there is a canonical isomorphism of graded-commutative rings

TorRp,ρ(R𝒮,ρ,Rp,ρ(σ))Hf1(GF,S,ad0ρ(1)).\operatorname{Tor}_{\ast}^{R_{p,\rho}}\big{(}R_{\mathcal{S},\rho},R_{p,\rho}(\sigma)\big{)}\cong\wedge^{\ast}H^{1}_{f}\big{(}G_{F,S},\operatorname{ad}^{0}\rho(1)\big{)}.
Remark 7.13.

We have discussed assumptions (iv) and (v) where they appeared above. The assumption (vi) is known to hold in many cases, see [1]. Assumption (vii) is a special case of a conjecture of Jannsen [21].

Proof.

By Theorem 7.1, Proposition 6.10 and Theorem 6.12,

(H~q0(XUp,𝒪)𝔪𝒪[[K]]𝐋σ)Rpps𝐋(Rpps)𝔭((m0E𝔅𝐋P𝔅)𝒪[[K]]𝐋σ)Rpps𝐋(Rpps)𝔭(m0)𝔭(E𝔅)𝔭𝐋P𝔅(σ)𝔭R𝒮,ρmRp,ρ𝐋Rp,ρ(σ)\begin{split}(\widetilde{H}_{q_{0}}(X_{U^{p}},\mathcal{O})_{\mathfrak{m}}\otimes^{\mathbf{L}}_{\mathcal{O}[[K]]}\sigma)\otimes^{\mathbf{L}}_{R^{\operatorname{ps}}_{p}}(R^{\operatorname{ps}}_{p})_{\mathfrak{p}}^{\wedge}&\cong\big{(}(m_{0}\otimes^{\mathbf{L}}_{E_{\mathfrak{B}}}P_{\mathfrak{B}})\otimes^{\mathbf{L}}_{\mathcal{O}[[K]]}\sigma\big{)}\otimes^{\mathbf{L}}_{R^{\operatorname{ps}}_{p}}(R^{\operatorname{ps}}_{p})_{\mathfrak{p}}^{\wedge}\\ &\cong(m_{0})_{\mathfrak{p}}^{\wedge}\otimes^{\mathbf{L}}_{(E_{\mathfrak{B}})_{\mathfrak{p}}^{\wedge}}P_{\mathfrak{B}}(\sigma)_{\mathfrak{p}}^{\wedge}\\ &\cong R_{\mathcal{S},\rho}^{\oplus m}\otimes^{\mathbf{L}}_{R_{p,\rho}}R_{p,\rho}(\sigma)\end{split}

for some m1m\geq 1. By assumption (iv), the spectral sequence of Proposition 6.16 degenerates at the E2E^{2}-page, hence we obtain

Hq0+i(XKτUp,σ(𝐰))𝔭[τ]ToriRp,ρ(R𝒮,ρ,Rp,ρ(σ))m,\begin{split}H_{q_{0}+i}(X_{K_{\tau}U^{p}},\sigma(\mathbf{w}))_{\mathfrak{p}}[\tau^{\ast}]&\cong\operatorname{Tor}_{i}^{R_{p,\rho}}\big{(}R_{\mathcal{S},\rho},R_{p,\rho}(\sigma)\big{)}^{\oplus m},\end{split}

where Kτ=ker(σ(τ))K_{\tau}=\ker(\sigma(\tau)). This proves the first part of the theorem.

Under the additional assumption that H2(GF,ad0ρ)=0H^{2}(G_{F},\operatorname{ad}^{0}\rho)=0, all three of Rp,ρ,R𝒮,ρR_{p,\rho},R_{\mathcal{S},\rho} and Rp,ρ(σ)R_{p,\rho}(\sigma) are formally smooth LL-algebras. Moreover, the exact sequence of Theorem 4.2 simplifies to a short exact sequence

0H1(GF,S,ad0ρ)vpH1(GFv,ad0ρ)Hf1(GFv,ad0ρ)Hf1(GF,S,ad0ρ(1))0,0\to H^{1}\big{(}G_{F,S},\operatorname{ad}^{0}\rho\big{)}\to\prod_{v\mid p}\frac{H^{1}\big{(}G_{F_{v}},\operatorname{ad}^{0}\rho\big{)}}{H^{1}_{f}\big{(}G_{F_{v}},\operatorname{ad}^{0}\rho\big{)}}\to H^{1}_{f}\big{(}G_{F,S},\operatorname{ad}^{0}\rho(1)\big{)}^{\vee}\to 0,

and the final term has dimension hf1(GF,ad0ρ(1))=l0h^{1}_{f}(G_{F},\operatorname{ad}^{0}\rho(1))=l_{0} by Proposition 4.1.

Dualising the sequence and interpreting the cohomology groups as tangent spaces, we see that the closed subschemes of SpecRp,ρ\operatorname{Spec}R_{p,\rho} corresponding to R𝒮,ρR_{\mathcal{S},\rho} and Rp,ρ(σ)R_{p,\rho}(\sigma) share no tangent directions at 𝔭\mathfrak{p}, and thus we may express the maximal ideal of Rp,ρR_{p,\rho} as the sum

𝔪Rp,ρ=ker(Rp,ρR𝒮,ρ)+ker(Rp,ρRp,ρ(σ)).\mathfrak{m}_{R_{p,\rho}}=\ker(R_{p,\rho}\to R_{\mathcal{S},\rho})+\ker(R_{p,\rho}\to R_{p,\rho}(\sigma)).

Furthermore, the intersection of the summands is generated by l0l_{0} elements, corresponding to a basis of the dual Selmer group Hf1(GF,S,ad0ρ(1))H^{1}_{f}\big{(}G_{F,S},\operatorname{ad}^{0}\rho(1)\big{)}. A choice of generators y¯=y1,,yl0\underline{y}=y_{1},\dots,y_{l_{0}} thus determines an isomorphism (using Proposition 3.8)

TorRp,ρ(R𝒮,ρ,Rp,ρ(σ))TorL[[y¯]](L,L)Hf1(GF,S,ad0ρ(1)).\begin{split}\operatorname{Tor}_{\ast}^{R_{p,\rho}}\big{(}R_{\mathcal{S},\rho},R_{p,\rho}(\sigma)\big{)}&\cong\operatorname{Tor}_{\ast}^{L[[\underline{y}]]}(L,L)\\ &\cong\wedge^{\ast}H^{1}_{f}\big{(}G_{F,S},\operatorname{ad}^{0}\rho(1)\big{)}.\end{split}

By general properties of the Tor\operatorname{Tor}-product, these are isomorphisms of graded algebras (see [13, XI.§2]). Furthermore, the isomorphism does not depend on y¯\underline{y}, since any viable choice yields the same power series ring. ∎

References

  • [1] Lambert A’Campo. Rigidity of automorphic Galois representations over CM fields. Int. Math. Res. Not. IMRN, (6):4541–4623, 2024.
  • [2] Patrick B. Allen. Deformations of polarized automorphic Galois representations and adjoint Selmer groups. Duke Math. J., 165(13):2407–2460, 2016.
  • [3] Tom Barnet-Lamb, David Geraghty, Michael Harris, and Richard Taylor. A family of Calabi-Yau varieties and potential automorphy II. Publ. Res. Inst. Math. Sci., 47(1):29–98, 2011.
  • [4] Joël Bellaiche. An introduction to Bloch and Kato’s conjecture. 2009.
  • [5] Spencer Bloch and Kazuya Kato. LL-functions and Tamagawa numbers of motives. In The Grothendieck Festschrift, Vol. I, volume 86 of Progr. Math., pages 333–400. Birkhäuser Boston, Boston, MA, 1990.
  • [6] A. Borel and J.-P. Serre. Corners and arithmetic groups. Comment. Math. Helv., 48:436–491, 1973.
  • [7] Christophe Breuil and Ariane Mézard. Multiplicités modulaires et représentations de GL2(𝐙p){\rm GL}_{2}({\bf Z}_{p}) et de Gal(𝐐¯p/𝐐p){\rm Gal}(\overline{\bf Q}_{p}/{\bf Q}_{p}) en l=pl=p. Duke Math. J., 115(2):205–310, 2002. With an appendix by Guy Henniart.
  • [8] Armand Brumer. Pseudocompact algebras, profinite groups and class formations. J. Algebra, 4:442–470, 1966.
  • [9] Frank Calegari and Matthew Emerton. Completed cohomology—a survey. In Non-abelian fundamental groups and Iwasawa theory, volume 393 of London Math. Soc. Lecture Note Ser., pages 239–257. Cambridge Univ. Press, Cambridge, 2012.
  • [10] Frank Calegari and David Geraghty. Modularity lifting beyond the Taylor-Wiles method. Invent. Math., 211(1):297–433, 2018.
  • [11] Ana Caraiani, Matthew Emerton, Toby Gee, David Geraghty, Vytautas Paškūnas, and Sug Woo Shin. Patching and the pp-adic Langlands program for GL2(𝕡){\rm GL}_{2}(\mathbb{Q_{p}}). Compos. Math., 154(3):503–548, 2018.
  • [12] Ana Caraiani, Daniel R. Gulotta, Chi-Yun Hsu, Christian Johansson, Lucia Mocz, Emanuel Reinecke, and Sheng-Chi Shih. Shimura varieties at level Γ1(p)\Gamma_{1}(p^{\infty}) and Galois representations. Compos. Math., 156(6):1152–1230, 2020.
  • [13] Henri Cartan and Samuel Eilenberg. Homological algebra. Princeton Landmarks in Mathematics. Princeton University Press, Princeton, NJ, 1999. With an appendix by David A. Buchsbaum, Reprint of the 1956 original.
  • [14] David Eisenbud. Commutative Algebra: with a View Toward Algebraic Geometry. Graduate Texts in Mathematics. Springer New York, NY, 1995.
  • [15] Matthew Emerton. Ordinary parts of admissible representations of pp-adic reductive groups I. Definition and first properties. Astérisque, (331):355–402, 2010.
  • [16] Toby Gee and James Newton. Patching and the completed homology of locally symmetric spaces. J. Inst. Math. Jussieu, 21(2):395–458, 2022.
  • [17] Philippe Gille and Tamás Szamuely. Central simple algebras and Galois cohomology, volume 165 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, second edition, 2017.
  • [18] David Hansen and Jack A. Thorne. On the GLn\operatorname{GL}_{n}-eigenvariety and a conjecture of Venkatesh. Selecta Math. (N.S.), 23(2):1205–1234, 2017.
  • [19] Michael Harris, Kai-Wen Lan, Richard Taylor, and Jack Thorne. On the rigid cohomology of certain Shimura varieties. Res. Math. Sci., 3:Paper No. 37, 308, 2016.
  • [20] Bence Hevesi. Ordinary parts and local-global compatibility at =p\ell=p, 2023.
  • [21] Uwe Jannsen. Weights in arithmetic geometry. Jpn. J. Math., 5(1):73–102, 2010.
  • [22] Chandrashekhar B. Khare and Jack A. Thorne. Potential automorphy and the Leopoldt conjecture. Amer. J. Math., 139(5):1205–1273, 2017.
  • [23] Mark Kisin. Potentially semi-stable deformation rings. J. Amer. Math. Soc., 21(2):513–546, 2008.
  • [24] Mark Kisin. Moduli of finite flat group schemes, and modularity. Ann. of Math. (2), 170(3):1085–1180, 2009.
  • [25] Qing Liu. Algebraic geometry and arithmetic curves, volume 6 of Oxford Graduate Texts in Mathematics. Oxford University Press, Oxford, 2002. Translated from the French by Reinie Erné, Oxford Science Publications.
  • [26] Tong Liu. Torsion pp-adic Galois representations and a conjecture of Fontaine. Ann. Sci. École Norm. Sup. (4), 40(4):633–674, 2007.
  • [27] B. Mazur. Deforming Galois representations. In Galois groups over 𝐐{\bf Q} (Berkeley, CA, 1987), volume 16 of Math. Sci. Res. Inst. Publ., pages 385–437. Springer, New York, 1989.
  • [28] J. S. Milne. Arithmetic duality theorems, volume 1 of Perspectives in Mathematics. Academic Press, Inc., Boston, MA, 1986.
  • [29] Jürgen Neukirch. Algebraic number theory, volume 322 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999. Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder.
  • [30] Louise Nyssen. Pseudo-représentations. Math. Ann., 306(2):257–283, 1996.
  • [31] Lue Pan. The Fontaine-Mazur conjecture in the residually reducible case. J. Amer. Math. Soc., 35(4):1031–1169, 2022.
  • [32] Vytautas Paškūnas. The image of Colmez’s Montreal functor. Publ. Math. Inst. Hautes Études Sci., 118:1–191, 2013.
  • [33] Vytautas Paškūnas. On the Breuil-Mézard conjecture. Duke Math. J., 164(2):297–359, 2015.
  • [34] Peter Scholze. On torsion in the cohomology of locally symmetric varieties. Ann. of Math. (2), 182(3):945–1066, 2015.
  • [35] Jack Shotton. The Breuil-Mézard conjecture when lpl\neq p. Duke Math. J., 167(4):603–678, 2018.
  • [36] The Stacks project authors. The Stacks project. https://stacks.math.columbia.edu, 2023.
  • [37] Jack Thorne. On the automorphy of ll-adic Galois representations with small residual image. J. Inst. Math. Jussieu, 11(4):855–920, 2012. With an appendix by Robert Guralnick, Florian Herzig, Richard Taylor and Thorne.
  • [38] Larry C. Washington. Galois cohomology. In Modular forms and Fermat’s last theorem, pages 101–120. Springer, 2000.
  • [39] Charles A. Weibel. An introduction to homological algebra, volume 38 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1994.