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

Fontaine-Laffaille Theory over Power Series Rings

Christian Hokaj
Abstract.

Let kk be a perfect field of characteristic p>2p>2. We extend the equivalence of categories between Fontaine-Laffaille modules and Zp\mathbb{Z}_{p} lattices inside crystalline representations with Hodge-Tate weights at most p2p-2 of [FL82] to the situation where the base ring is the power series ring over the Witt vectors RW(k)t1,,tdR\coloneqq W(k)\llbracket t_{1},\cdots,t_{d}\rrbracket and TT a pp-adically complete ring that is étale over the Tate Algebra W(k)t1±1,,td±1W(k)\langle t_{1}^{\pm 1},\cdots,t_{d}^{\pm 1}\rangle.

1. Introduction

Let kk be a perfect field of characteristic p>2p>2, W(k)W(k) the ring of Witt-vectors over kk with φ\varphi its Frobenius automorphism, KK its fraction field, and RW(k)t1,,tdR\coloneqq W(k)\llbracket t_{1},\cdots,t_{d}\rrbracket the ring of power series over W(k)W(k) in dd variables. Let TT be a “small base ring,” a ring that is pp-adically completed étale over the Tate algebra, W(k)t1±1,,td±1W(k)\langle t_{1}^{\pm 1},\cdots,t_{d}^{\pm 1}\rangle (i.e. the ring of restricted Laurent series over W(k)W(k)), for some dd. Let rr be an integer such that 0rp2.0\leq r\leq p-2. When the base ring is W(k)W(k), Fontaine and Laffaille introduced in [FL82] strongly divisible W(k)W(k)-lattices, which we now call Fontaine-Laffaille modules, to study Zp\mathbb{Z}_{p}-lattices in crystalline representations of GKGal(K¯/K)G_{K}\coloneqq\mathrm{Gal}(\overline{K}/K) with Hodge-Tate weights in [0,r].[0,r]. Denote this category by RepZp,[0,r]cris(GK)\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{K}). We first recall the definition of a Fontaine-Laffaille module in the classical setting:

Definition 1.

A finite free Fontaine-Laffaille module over W(k)W(k) is a triple (M,Fili(M),φi)(M,\mathrm{Fil}^{i}(M),\varphi_{i}) where:

  • MM is a finite free module over W(k)W(k)

  • {Fili(M)}i=0\{\mathrm{Fil}^{i}(M)\}_{i=0}^{\infty} is a decreasing filtration of MM such that Fili+1(M)\mathrm{Fil}^{i+1}(M) is a direct summand of Fili(M)\mathrm{Fil}^{i}(M) for all ii, Fil0(M)=M\mathrm{Fil}^{0}(M)=M, Filr+1(M)=0.\mathrm{Fil}^{r+1}(M)=0.

  • φi\varphi_{i} are φ\varphi-semilinear maps Fili(M)M\mathrm{Fil}^{i}(M)\to M whose restriction to Fili+1(M)\mathrm{Fil}^{i+1}(M) is pφi+1.p\varphi_{i+1}.

  • iφi(FiliM)=M.\sum_{i}\varphi_{i}(\mathrm{Fil}^{i}M)=M.

Let MFff,[0,r](W(k))\mathrm{MF}^{\mathrm{ff},[0,r]}(W(k)) denote the category of finite free Fontaine-Laffaille modules over W(k)W(k). These modules classify lattices within crystalline representations as in the following theorem:

Theorem 2 ([FL82]).

When 0rp2,0\leq r\leq p-2, the functor given by

Tcris(M)(Filr(AcrisW(k)M))φr=1T_{\mathrm{cris}}(M)\coloneqq(\mathrm{Fil}^{r}(A_{\mathrm{cris}}\otimes_{W(k)}M))^{\varphi_{r}=1}

is an equivalence of categories

MFff,[0,r](W(k))RepZp,[0,r]cris(GK).\mathrm{MF}^{\mathrm{ff},[0,r]}(W(k))\leftrightarrow\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{K}).

The work of Faltings in [Fal89] extended this theory to more general base rings. Our goal is to extend the above theorem to the setting where the base ring is RR and to the setting of a small base ring TT.

Let us first recall the representation category in the relative case. Let GRG_{R} denote Gal(R¯[p1]/R[p1])=π1e´t(SpecR[1p],η)\mathrm{Gal}(\overline{R}[p^{-1}]/R[p^{-1}])=\pi_{1}^{\mathrm{\acute{e}t}}(\mathrm{Spec}R\left[\frac{1}{p}\right],\eta) where η\eta is a fixed geometric point and R¯\overline{R} is the union of finite RR-subalgebras RR^{\prime} of a fixed algebraic closure of Frac(R)\mathrm{Frac}(R) such that R[p1]R^{\prime}[p^{-1}] is étale over R[p1].R[p^{-1}]. Let RepZp,[0,r]cris(GR)\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{R}) denote the category of crystalline Zp\mathbb{Z}_{p} representations of GRG_{R} with Hodge-Tate weights in [0,r][0,r] and we make the analogous definitions over the base ring TT.

Using recent results of Du, Liu, Moon, and Shimizu in [Du+24] and the category of finite, free Fontaine-Laffaille modules in the relative case first introduced by Faltings in [Fal89] and more recently studied by [LMP23], which we will denote MFff,[0,r](R)\mathrm{MF}^{\mathrm{ff},[0,r]}_{\nabla}(R), we obtain the following result (see Definition 9 for the definition of Fontaine-Laffaille modules in the relative case, and see Section 2.3 for a brief review of the period rings and crystalline representations in the relative case):

Theorem 3.

Let rr be an integer satisfying 0rp20\leq r\leq p-2 and let 𝒯\mathcal{T} denote either the base ring RR or TT. The functor

Tcris:MFff,[0,r](𝒯)RepZp,[0,r]cris(G𝒯)T_{\mathrm{cris}}:\mathrm{MF}^{\mathrm{ff},[0,r]}_{\nabla}(\mathcal{T})\to\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{\mathcal{T}})

given by

Tcris(M)(Filr(Acris(𝒯)𝒯M))φr=1T_{\mathrm{cris}}(M)\coloneqq(\mathrm{Fil}^{r}(A_{\mathrm{cris}}(\mathcal{T})\otimes_{\mathcal{T}}M))^{\varphi_{r}=1}

is an equivalence of categories.

Showing this functor is fully faithful is routine, and the difficulty lies in showing essential surjectivity and specifically associating a Fontaine-Laffaille module to a given lattice inside a crystalline representation.

The key input of [Du+24] is the existence of a Kisin module 𝔐\mathfrak{M} over the ring 𝔖𝒯u\mathfrak{S}\coloneqq\mathcal{T}\llbracket u\rrbracket associated to an object of RepZp,[0,r]cris(G𝒯)\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{\mathcal{T}}). The ring 𝔖\mathfrak{S} comes equipped with a Frobenius map φ𝔖\varphi_{\mathfrak{S}} extending the Frobenius on 𝒯\mathcal{T}. We can then define the associated Fontaine-Laffaille module as 𝔐/u𝔐𝒯,φ𝔖𝒯\mathfrak{M}/u\mathfrak{M}\otimes_{\mathcal{T},\varphi_{\mathfrak{S}}}\mathcal{T}, but the Kisin module alone is not enough to obtain the full data of a Fontaine-Laffaille module over 𝒯\mathcal{T}.

In the classical case, Breuil in [Bre02] studied Zp\mathbb{Z}_{p} lattices in crystalline representations of GKG_{K} using strongly divisible SS-modules. To obtain the full data of our Fontaine-Laffaille module, we define SS in the relative setting to be the pp-adic completion of the divided power envelope of 𝔖\mathfrak{S} taken with respect to (E(u))(E(u)) with E(u)=upE(u)=u-p or u+pu+p. The ring SS comes with a canonical divided power filtration and additional structure which will descend to our proposed Fontaine-Laffaille module.

To show the main theorem holds over RR, we will verify that the data properly descends through the use of two base change maps relating our RR-module back to the classical theory. Let kgk_{g} be the perfection of Frac(R/pR)\mathrm{Frac}(R/pR) and RgW(kg)R_{g}\coloneqq W(k_{g}), which gives rise to a flat embedding RRgR\hookrightarrow R_{g}. Let b¯:RW(k)\overline{b}:R\to W(k) denote reduction moduo (t1,,td)(t_{1},\cdots,t_{d}). A result of [LMP23] allows us to verify that MM is an object of MFff,[0,r](R)\mathrm{MF}^{\mathrm{ff},[0,r]}_{\nabla}(R) by checking that MRRgM\otimes_{R}R_{g} and MW(k),b¯W(k)M\otimes_{W(k),\overline{b}}W(k) are Fontaine-Laffaille modules over W(kg)W(k_{g}) and W(k)W(k) in the classical sense.

To show the main theorem holds over TT, we obtain a result analogous to that of [LMP23] where we can check that MM is Fontaine-Laffaille at the pp-adic completion of localizations at maximal ideals. This enables us to use a base change TRT\to R as described in Section 2.2.3 and reduce to the setting where the base ring is RR.

Recently Würthen in [Wür23] proved this result in the setting of a smooth pp-adic formal scheme over a mixed characteristic complete discrete valuation ring with perfect residue field. Our work uses different methods to focus on the case where the base ring is more limited, and it is routine to extend the results from a small base ring to accommodate the setting of a smooth, pp-adic formal scheme.

The work of Imai, Kato, and Youcis in [IKY24],developed simultaneously, uses different methods to obtain a different fully faithful functor between lattices inside crystalline representations and Fontaine-Laffaille modules in a relative setting. Their functor has the advantage of being extendable beyond the Fontaine-Laffaille range, and it yields objects in a nicer category. Our functor has the advantage of being essentially surjective. They also place additional assumptions on the prismatic FF-crystal, and their technique has not yet been applied to the case where the base ring is a power series ring.

Acknowledgements: The author would like to thank his Ph.D advisor Tong Liu for suggesting the work in this paper, and for his extensive and invaluable guidance, comments, and conversation throughout the production of this paper and its earlier drafts. This paper includes and generalizes the content of the author’s Ph.D. thesis at Purdue University when he was partially supported by the Ross Fellowship of Purdue University and a Summer Research Grant during the summer of 2021.

2. Preliminary Results on some Categories and Functors in the Relative Setting

2.1. Base Ring Conventions

Much of integral pp-adic Hodge Theory has been established in the relative case in far more generality than Fontaine-Laffaille theory. We present the most general setting we will consider here.

Recall kk is a perfect field of characteristic p>2.p>2. Let KtrK^{tr} be a totally ramified extension of K=Frac(W(k))K=\mathrm{Frac}(W(k)) with ring of integers 𝒪Ktr\mathcal{O}_{K^{tr}}. Fix π\pi a uniformizer of KtrK^{tr} and let E(u)W(k)[u]E(u)\in W(k)[u] denote the monic minimal polynomial of π.\pi. Write W(k)T1±1,,Td±1W(k)\langle T_{1}^{\pm 1},\cdots,T_{d}^{\pm 1}\rangle for the pp-adic completion of the Laurent polynomial ring W(k)[T1±1,,Td±1]W(k)\left[T_{1}^{\pm 1},\cdots,T_{d}^{\pm 1}\right].

Convention 4.

In order to reference results with maximal generality, we will say \mathcal{R} is a general base ring if \mathcal{R} is a pp-adically complete 𝒪Ktr\mathcal{O}_{K^{tr}}-algebra which is of the form =𝒪KtrW(k)0\mathcal{R}=\mathcal{O}_{K^{tr}}\otimes_{W(k)}\mathcal{R}_{0} where 0\mathcal{R}_{0} is an integral domain obtained from W(k)T1±1,,Td±1W(k)\langle T_{1}^{\pm 1},\cdots,T_{d}^{\pm 1}\rangle by a finite number of iterations of the operations:

  • pp-adic completion of an étale extension

  • pp-adic completion of a localization;

  • completion with respect to an ideal containing pp.

Our new results will hold for a subset of these general base rings which we describe next:

Convention 5.

We will say a general base ring TT is small if it is pp-adically completed étale over W(k)T1±1,,Td±1W(k)\langle T_{1}^{\pm 1},\cdots,T_{d}^{\pm 1}\rangle for some dd.

The key strategy for showing our results over small base rings will be to reduce to the case of the power series ring R=W(k)t1,,tdR=W(k)\llbracket t_{1},\cdots,t_{d}\rrbracket due to the local structure of small base rings described in the following lemma:

Lemma 6.

Let TT be a small base ring that is pp-adically compaeted étale over W(k)T1±1,,Td±1W(k)\langle T_{1}^{\pm 1},\cdots,T_{d}^{\pm 1}\rangle and let 𝔪\mathfrak{m} be a maximal ideal (note that p𝔪p\in\mathfrak{m}). Then T𝔪^\widehat{T_{\mathfrak{m}}}, the pp-adic completion of the localization of TT at 𝔪\mathfrak{m}, is isomorphic to W(k)t1,,tdW(k)\llbracket t_{1},\cdots,t_{d}\rrbracket.

Proof.

Let 𝔪\mathfrak{m} be a maximal ideal of TT, and let T𝔪^\widehat{T_{\mathfrak{m}}} denote the pp-adic completion of the localization of TT at 𝔪.\mathfrak{m}. The ring TT is a regular Noetherian ring, and thus T𝔪^W(k)X1,,XdR\widehat{T_{\mathfrak{m}}}\cong W(k)\llbracket X_{1},\cdots,X_{d}\rrbracket\cong R by Cohen’s Structure Theorem. ∎

We will try to reserve the letter RR in usual font for referring to such a power series ring, the letter TT in usual font for a small base ring, and the stylized \mathcal{R} for a general base ring. We will use the stylized 𝒯\mathcal{T} to concisely state results that hold for a small base ring or the power series ring RR.

When we refer to the “classical” theory we will be referring to the setting where the base ring is KK: the historically typical setting of pp-adic Hodge Theory.

We will now describe the Frobenius structure and the category of Galois representations for a general base ring. We will let φ\varphi_{\mathcal{R}} denote the lift of Frobenius on /W(k)\mathcal{R}/W(k) uniquely determined by φ(Ti)=Tip\varphi(T_{i})=T_{i}^{p}. Let ¯\overline{\mathcal{R}} denote the union of finite \mathcal{R}-subalgebras \mathcal{R}^{\prime} of a fixed algebraic closure of Frac()\mathrm{Frac}(\mathcal{R}) such that [p1]\mathcal{R}^{\prime}[p^{-1}] is étale over [p1]\mathcal{R}[p^{-1}]. Set G=Gal(¯/[p1])G_{\mathcal{R}}=\mathrm{Gal}(\overline{\mathcal{R}}/\mathcal{R}[p^{-1}]). Let RepQp(G)\mathrm{Rep}_{\mathbb{Q}_{p}}(G_{\mathcal{R}}) denote the category of finite dimensional Qp\mathbb{Q}_{p} vector spaces with continuous GG_{\mathcal{R}} action, and let RepZpff(G)\mathrm{Rep}_{\mathbb{Z}_{p}}^{\mathrm{ff}}(G_{\mathcal{R}}) denote the full category of finite free Zp\mathbb{Z}_{p} modules equipped with a continuous action of 𝒢\mathcal{G}_{\mathcal{R}}.

Remark 7.

This describes the Frobenius structure for a general base ring, but we will actually take φR(ti)=(ti+1)p1\varphi_{R}(t_{i})=(t_{i}+1)^{p}-1 for our base ring RR in order to use the results of [LMP23] as stated. This is not significant, as we will show the relevant category of Fontaine-Laffaille modules defined below is independent of the choice of Frobenius.

2.2. Important Base Change Maps

Our main technique for developing the Fontaine-Laffaille Theory for new base rings will be to base change to rings where the Fontaine-Laffaille Theory is already known. We describe the important base change maps here.

2.2.1. Shilov Point Base Changes

Let kgk_{g} be the perfection of Frac(/p)\mathrm{Frac}(\mathcal{R}/p\mathcal{R}) and let gW(kg)\mathcal{R}_{g}\coloneqq W(k_{g}). We have a φ\varphi-compatiable, flat embedding g\mathcal{R}\hookrightarrow\mathcal{R}_{g}. Set 𝔖,gW(kg)u\mathfrak{S}_{\mathcal{R},g}\coloneqq W(k_{g})\llbracket u\rrbracket, and let S,gS_{\mathcal{R},g} be the pp-adically completed divided power envelope of 𝔖,g\mathfrak{S}_{\mathcal{R},g} with respect to E(u)E(u) which comes equipped with a φ\varphi structure and PD-filtration.

As the Shilov point above forgets the data of the connection \nabla, we will also need to consider a second Shilov point with imperfect residue field. Let L\mathcal{R}_{L} denote the pp-adic completion of the localization (p)\mathcal{R}_{(p)}. Then L\mathcal{R}_{L} is a complete discrete valuation ring with imperfect residue field. We define 𝔖L=Lu\mathfrak{S}_{L}=\mathcal{R}_{L}\llbracket u\rrbracket and SLS_{L} to be the pp-adically completed divided power envelope of 𝔖L\mathfrak{S}_{L} with respect to (E(u)).(E(u)).

2.2.2. Closed Fiber Base Change of a Power Series Ring

In the case of our base ring RR, we also have a natural projection π:RW(k)\pi:R\to W(k) given by ti0t_{i}\mapsto 0 for every ii. Showing that the Fontaine-Laffaille Theory holds for RR will amount to checking it is compatible with the Shilov point base change and the closed fiber base change.

2.2.3. Small Base Ring Base Changes

We will now discuss a base change map from a small base ring TT to RR suggested by Lemma 6. Let 𝔪=(p,f1,,fd)\mathfrak{m}=(p,f_{1},\cdots,f_{d}) be a maximal ideal of TT (recall TT is regular), and let T𝔪^\widehat{T_{\mathfrak{m}}} denote the pp-adic completion of the localization of TT at 𝔪\mathfrak{m} which is isomorphic to T𝔪^W(k)X1,,XdR\widehat{T_{\mathfrak{m}}}\cong W(k)\llbracket X_{1},\cdots,X_{d}\rrbracket\cong R by Lemma 6.

This gives us a flat base change ring homomorphism b𝔪:TT𝔪^Rb_{\mathfrak{m}}:T\to\widehat{T_{\mathfrak{m}}}\cong R. In order to make this map Frobenius equivariant, though, we must consider φR\varphi_{R} to be the lift of the Frobenius on W(k)W(k) which acts on the XiX_{i} as φR(X1)=b𝔪(φT(fi))\varphi_{R}(X_{1})=b_{\mathfrak{m}}(\varphi_{T}(f_{i})). For example, if 𝔪=(p,t11,,td1)\mathfrak{m}=(p,t_{1}-1,\cdots,t_{d}-1), then we must take φR(Xi)=(Xi+1)p1.\varphi_{R}(X_{i})=(X_{i}+1)^{p}-1.

By Lemma 12 below, there is a canonical equivalence between the categories of finite, free Fontaine-Laffaille modules over RR with different Frobenii, so this choice will not impact our results.

2.3. Review of relative period rings and crystalline representations

In this section we briefly review the crystalline period rings and the functor DcrisD_{\mathrm{cris}} in the relative case. For full details of their constructions, see [Bri08, Chapter 6] for a treatment of the period rings and [Bri08, Chapter 8] for a treatment of relevant functors. Let \mathcal{R} be a general base ring as in Convention 4. Let ¯^\widehat{\overline{\mathcal{R}}} be the pp-adic completion of ¯\overline{\mathcal{R}}. Set ¯=limφ¯/p¯\overline{\mathcal{R}}^{\flat}=\displaystyle\varprojlim_{\varphi}\overline{\mathcal{R}}/p\overline{\mathcal{R}}. Then we define Ainf()=W(¯)A_{\mathrm{inf}}(\mathcal{R})=W(\overline{\mathcal{R}}^{\flat}) and write [π][\pi^{\flat}] for the Teichmüler lift of π\pi^{\flat} where π=(π1/pn)n0\pi^{\flat}=(\pi^{1/p^{n}})_{n\geq 0} is a compatible sequence of pp-power roots of π\pi. Let θ:Ainf()¯^\theta:A_{\mathrm{inf}}(\mathcal{R})\to\widehat{\overline{\mathcal{R}}} be the unique surjective W(k)W(k)-algebra homomorphism lifting the first projection.

Let Acris()A_{\mathrm{cris}}(\mathcal{R}) be the pp-adic completion of the divided power envelope of Ainf()A_{\mathrm{inf}}(\mathcal{R}) with respect to the kernel of θ\theta. Let ϵn¯\epsilon_{n}\in\overline{\mathcal{R}} denote a (non-trivial) sequence of compatible pp-power roots of unity (i.e. such that ϵ0=1),ϵ11,ϵn=ϵn+1p\epsilon_{0}=1),\epsilon_{1}\neq 1,\epsilon_{n}=\epsilon_{n+1}^{p}. Let ϵ=(ϵn)n¯\epsilon=(\epsilon_{n})_{n}\in\overline{\mathcal{R}}^{\flat} and t=log[ϵ]Acris().t=\log[\epsilon]\in A_{\mathrm{cris}}(\mathcal{R}). Let Bcris()Acris()[p1,t1].B_{\mathrm{cris}}(\mathcal{R})\coloneqq A_{\mathrm{cris}}(\mathcal{R})[p^{-1},t^{-1}].

Let θ0\theta_{0} be the extension of θ\theta to 0W(k)Ainf()¯^\mathcal{R}_{0}\otimes_{W(k)}A_{\mathrm{inf}}(\mathcal{R})\to\widehat{\overline{\mathcal{R}}} and set OAcris()\mathrm{OA}_{\mathrm{cris}}(\mathcal{R}) to be the pp-adic completion of the divided power envelope of 0W(k)Ainf()\mathcal{R}_{0}\otimes_{W(k)}A_{\mathrm{inf}}(\mathcal{R}) with respect to the kernel of θ0\theta_{0}. Let OBcris()OAcris()[p1,t1]\mathrm{OB}_{\mathrm{cris}}(\mathcal{R})\coloneqq\mathrm{OA}_{\mathrm{cris}}(\mathcal{R})[p^{-1},t^{-1}].

For VRepQp(G)V\in\mathrm{Rep}_{\mathbb{Q}_{p}}(G_{\mathcal{R}}), we have a functor

Dcris(V)(OBcris()QpV)GD_{\mathrm{cris}}(V)\coloneqq(\mathrm{OB}_{\mathrm{cris}}(\mathcal{R})\otimes_{\mathbb{Q}_{p}}V)^{G_{\mathcal{R}}}

which is a finite projetive [p1]\mathcal{R}[p^{-1}] module of rank at most dimQpV\mathrm{dim}_{\mathbb{Q}_{p}}V equipped with a φ\varphi and an integrable connection \nabla induced by that on OBcris().\mathrm{OB}_{\mathrm{cris}}(\mathcal{R}). We say that VV is crystalline if the natural map

OBcris()[p1]Dcris(V)OBcris()QpV\mathrm{OB}_{\mathrm{cris}}(\mathcal{R})\otimes_{\mathcal{R}[p^{-1}]}D_{\mathrm{cris}}(V)\to\mathrm{OB_{\mathrm{cris}}}(\mathcal{R})\otimes_{\mathbb{Q}_{p}}V

is an isomorphism. An object NRepZpff(G)N\in\mathrm{Rep}_{\mathbb{Z}_{p}}^{\mathrm{ff}}(G_{\mathcal{R}}) is said to be crystalline if NZpQpN\otimes_{\mathbb{Z}_{p}}\mathbb{Q}_{p} is crystalline. There is also a contravariant version of DcrisD_{\mathrm{cris}} defined to be

Dcris(V)HomG(V,OBcris())D_{\mathrm{cris}}^{\vee}(V)\coloneqq\mathrm{Hom}_{G_{\mathcal{R}}}(V,\mathrm{OB}_{\mathrm{cris}}(\mathcal{R}))

dual to the above covariant version.

2.4. Finite free Fontaine-Laffaille modules

Here we introduce the various categories of Fontaine-Laffaille modules in the relative case, following the work of [Fal89] and [LMP23]. Here we will need to distinguish between the case of the base ring RR and a small base ring TT for the statement of some theorems.

Definition 8.

Let MFbigf(R)\mathrm{MF}_{\mathrm{big}}^{\mathrm{f}}(R) be the category whose objects consist of a an RR-module MM, a sequence of RR-modules Fi(M)\mathrm{F}^{i}(M), and sequences of RR-linear maps ιi:Fi(M)Fi1(M),πi:Fi(M)M,\iota_{i}:\mathrm{F}^{i}(M)\to\mathrm{F}^{i-1}(M),\pi_{i}:\mathrm{F}^{i}(M)\to M, and RR-semi-linear maps φi:Fi(M)M\varphi_{i}:\mathrm{F}^{i}(M)\to M, satisfying the following conditions:
(1) The composition Fi(M)ιiFi1(M)πi1M\mathrm{F}^{i}(M)\xrightarrow{\iota_{i}}{}\mathrm{F}^{i-1}(M)\xrightarrow{\pi_{i-1}}{}M is the map πi:Fi(M)M\pi_{i}:\mathrm{F}^{i}(M)\to M
(2) The map πi:Fi(M)M\pi_{i}:\mathrm{F}^{i}(M)\to M is an isomorphism for i0.i\ll 0.
(3) The composition of φi1\varphi_{i-1} with ιi:Fi(M)Fi1(M)\iota_{i}:\mathrm{F}^{i}(M)\to\mathrm{F}^{i-1}(M) is pφi.p\varphi_{i}.
Morphism between these objects are compatible collections of RR-linear maps between MM’s and Fi(M)\mathrm{F}^{i}(M)’s.

We write MF,bigf(R)\mathrm{MF}_{\nabla,\mathrm{big}}^{\mathrm{f}}(R) when we consider the category of such modules equipped with an integrable connection :MMRΩ^R\nabla:M\to M\otimes_{R}\widehat{\Omega}_{R} for which the φi\varphi_{i} are parallel and that satisfies Griffiths transversality. Specifically, when the following conditions hold:

  • Griffiths-transversality holds. i.e. the following diagram commutes: Fi(M){\mathrm{F}^{i}(M)}Fi1M{\mathrm{F}^{i-1}M}Fi1MRΩ^R{\mathrm{F}^{i-1}M\otimes_{R}\widehat{\Omega}_{R}}Fi2(M)RΩ^R{\mathrm{F}^{i-2}(M)\otimes_{R}\widehat{\Omega}_{R}}i\scriptstyle{\nabla_{i}}ιi\scriptstyle{\iota_{i}}i1\scriptstyle{\nabla_{i-1}}ιi1RΩ^R\scriptstyle{\iota_{i-1}\otimes_{R}\widehat{\Omega}_{R}}

    Note that this simplifies to (Fi(M))Fi1(M)RΩ^R\nabla(\mathrm{F}^{i}(M))\subset\mathrm{F}^{i-1}(M)\otimes_{R}\widehat{\Omega}_{R} when the ιi\iota_{i} are injective.

  • The φi\varphi_{i} are parallel: φi=(φi1Rdφ1)\nabla\circ\varphi_{i}=(\varphi_{i-1}\otimes_{R}d\varphi_{1})\circ\nabla as a map from Fi(M)\mathrm{F}^{i}(M) to MRΩ^RM\otimes_{R}\widehat{\Omega}_{R} where dφ1:Ω^RR,φRΩ^Rd\varphi_{1}:\widehat{\Omega}_{R}\otimes_{R,\varphi}R\to\widehat{\Omega}_{R} is dφ/pd\varphi/p.

We write MF,bigf,[a,b](R)\mathrm{MF}_{\nabla,\mathrm{big}}^{\mathrm{f},[a,b]}(R) when Fa(M)=M\mathrm{F}^{a}(M)=M and Fb+1(M)={0}.\mathrm{F}^{b+1}(M)=\{0\}.

We will replace “f” with “tor” when working in the setting of pp-power torsion modules as in [LMP23]. We define these categories analgously when working over small base rings.

Definition 9.

Let MFff(R)\mathrm{MF}^{\mathrm{ff}}(R) denote the full subcategory of MFf(R)\mathrm{MF}^{\mathrm{f}}(R) such that MM and Fi(M)\mathrm{F}^{i}(M) are finitely generated and free, each Fi+1(M)\mathrm{F}^{i+1}(M) is a direct summand of Fi(M)\mathrm{F}^{i}(M), Fi(M)={0}\mathrm{F}^{i}(M)=\{0\} for i0i\gg 0, and iφi(Fi(M))\sum_{i}\varphi_{i}(\mathrm{F}^{i}(M)) generate M.M. We call an object in this category a “finite free Fontaine-Laffaille module (over RR)” or sometimes just a “Fontaine-Laffaille module (over RR)” for convenience. The tags \nabla, [a,b][a,b], and “tor” are used in the same way as for the big category, and we define the categories analagously for small base rings.

We next recall a main theorem of [LMP23] which we will refine to our finite free setting over the base ring RR, and we will develop an analogue for small base rings. Recall the Shilov point RgR_{g} from Section 2.2.1 and that here is a φ\varphi-compatible, flat embedding bg:RRgb_{g}:R\hookrightarrow R_{g}. Recall from Section 2.2.2 that there is a natural projection b¯:RW(k)\overline{b}:R\to W(k) determined by ti0t_{i}\mapsto 0 for each ii. A main theorem of [LMP23] is as follows:

Theorem 10.

[LMP23, Proposition 2.2.5]
Suppose MMFbigtor,[0,r](R)M\in\mathrm{MF}_{\mathrm{big}}^{\mathrm{tor},[0,r]}(R) such that each Fili(M)\mathrm{\mathrm{Fil}^{i}(M)} is finite as an RR-module. Then MM is in MFff(R)\mathrm{MF}^{\mathrm{ff}}(R) if and only if:
(1) Both M0MR,b¯W(k)M_{0}\coloneqq M\otimes_{R,\overline{b}}W(k) and MgMRRgM_{g}\coloneqq M\otimes_{R}R_{g} are objects in MFtor,[0,r](W(kg))\mathrm{MF}^{\mathrm{tor},[0,r]}(W(k_{g})) and MFtor,[0,r](Rg),\mathrm{MF}^{\mathrm{tor},[0,r]}(R_{g}), respectively.
(2) MgM0W(k)RgM_{g}\cong M_{0}\otimes_{W(k)}R_{g} as RgR_{g}-modules.

Lt MFbigff,[0,r](R)\mathrm{MF}_{\mathrm{big}}^{\mathrm{ff},[0,r]}(R) denote the full subcategory of MFbigf,[0,r](R)\mathrm{MF}_{\mathrm{big}}^{\mathrm{f},[0,r]}(R) where MM and each Fili(M)\mathrm{Fil}^{i}(M) are finite, free RR-modules. Then we have:

Proposition 11.

Theorem 10, holds if each “tor\mathrm{tor}” is replaced with “ff.\mathrm{ff}.

Proof.

If MMFbigff,[0,r](R)M\in\mathrm{MF}_{\mathrm{big}}^{\mathrm{ff},[0,r]}(R), by tensoring with W(k)W(k) and RgR_{g} and noting that MM and Fili(M)\mathrm{Fil}^{i}(M) are finite free RR-modules, it is clear that (1) and (2) hold.

We can show the reverse direction by reducing to the torsion case. Let nn be a positive integer. It is clear that M/pnMR,b¯W(k)MRR/pnRR,b¯W(k)M0/pnM0M/p^{n}M\otimes_{R,\overline{b}}W(k)\cong M\otimes_{R}R/p^{n}R\otimes_{R,\overline{b}}W(k)\cong M_{0}/p^{n}M_{0} and M/pnMRRgMg/pnMgM/p^{n}M\otimes_{R}R_{g}\cong M_{g}/p^{n}M_{g} are elements of MFtor,[0,r](W(k))\mathrm{MF}^{\mathrm{tor},[0,r]}(W(k)) and MFtor,[0,r](Rg)\mathrm{MF}^{\mathrm{tor},[0,r]}(R_{g}), respectively.

By tensoring MgM0W(k)RgM_{g}\cong M_{0}\otimes_{W(k)}R_{g} with R/pnRR/p^{n}R, we also see that M0/pnM0M_{0}/p^{n}M_{0} and Mg/pnMgM_{g}/p^{n}M_{g} are of the same type. Then by Theorem 10, we conclude that M/pnMM/p^{n}M is an element of MFtor,[0,r](R).\mathrm{MF}^{\mathrm{tor},[0,r]}(R). Since M=limM/pnMM=\varprojlim M/p^{n}M and Fi(M)=limFi(M)/pnFi(M)\mathrm{F}^{i}(M)=\varprojlim\mathrm{F}^{i}(M)/p^{n}\mathrm{F}^{i}(M) and each is free, we deduce that each is finite free and we have the direct summand condition. Similarly, as φi(Fi(M/pnM))\varphi_{i}(\mathrm{F}^{i}(M/p^{n}M)) generates M/pnMM/p^{n}M, the same condition lifts to MM. ∎

We will now build to a theorem analogous to Theorem 11 but which holds for small base rings. First we will show that the category MFff(R)\mathrm{MF}_{\nabla}^{\mathrm{ff}}(R) is independent of the lift of Frobenius.

Lemma 12 ([Fal89]).

Let a,ba,b be such that 0bap10\leq b-a\leq p-1. Let φ,ψ\varphi,\psi be two lifts of the natural Frobenius on W(k)W(k) to RR which coincide modulo pp. We write RφR_{\varphi} (resp. RψR_{\psi}) when we are considering RR with the Frobenius lift φ\varphi (resp. ψ\psi). Then there is an equivalence between the corresponding categories MFff,[a,b](Rφ)\mathrm{MF}_{\nabla}^{\mathrm{ff},[a,b]}(R_{\varphi}) and MFff,[a,b](Rψ)\mathrm{MF}_{\nabla}^{\mathrm{ff},[a,b]}(R_{\psi}), and up to canonical isomorphism, MFff,[a,b](R)\mathrm{MF}_{\nabla}^{\mathrm{ff},[a,b]}(R) is independent of the lift of Frobenius.

Proof.

We show the result for the torsion categories, and the result for the finite free categories follows via projective limits.

The proof of [Fal89] applies almost identically in the torsion case, but we do not have an étale map from W(k)[t1,,td]RW(k)[t_{1},\cdots,t_{d}]\to R, but we still have that {i}i=1d{/Xi}i=1d\{\partial_{i}\}_{i=1}^{d}\coloneqq\{\partial/\partial X_{i}\}_{i=1}^{d} gives the dual basis of RR-derivations. From there we proceed identically to Faltings. By shifting if necessary, we can assume that a=0a=0, b=p1.b=p-1. The /Xi\partial/\partial X_{i} act on MM via .\nabla. Given a multi-index I=(i1,,id)I=(i_{1},\cdots,i_{d}), we get an endomorphism ()I\nabla(\partial)^{I} of MM. Let (φ(X)ψ(X))I(\varphi(X)-\psi(X))^{I} denote j=1d(φ(Xj)ψ(Xj))ij\prod_{j=1}^{d}(\varphi(X_{j})-\psi(X_{j}))^{i_{j}}, and |I|=j=1dij|I|=\sum_{j=1}^{d}i_{j}, and I!=j=1dij!I!=\prod_{j=1}^{d}i_{j}!. Comparing divisibility by pp, we can see this gives a well defined element.

Note that an object MMFbigtor,[a,b](R)M\in\mathrm{MF}_{\mathrm{big}}^{\mathrm{tor},[a,b]}(R) is in MFtor,[a,b](R)\mathrm{MF}^{\mathrm{tor},[a,b]}(R) if and only if the Frobenius ϕ\phi on MM induces an isomorphism M~R,φM\widetilde{M}\otimes_{R,\varphi}\to M where M~\widetilde{M} is the right exact functor defined explicitly as the cokernel of the map

θM:i=a+1bFi(M)i=abFi(M),\theta_{M}:\bigoplus_{i=a+1}^{b}\mathrm{F}^{i}(M)\to\bigoplus_{i=a}^{b}\mathrm{F}^{i}(M),

where θM\theta_{M} is given by

θM((xa+1,,xb))=(ιa+1(xa+1),pxa+1+ιa+2(xa+2),,pxb1+ιb(xb),pxb).\theta_{M}((x_{a+1},\cdots,x_{b}))=(\iota_{a+1}(x_{a+1}),-px_{a+1}+\iota_{a+2}(x_{a+2}),\cdots,-px_{b-1}+\iota_{b}(x_{b}),-px_{b}).

Equivalently, M~\widetilde{M} can be defined as the colimit of the following diagram:

Fi+1(M)Fi+1(M)Fi(M)Fi(M)Fi1(M)\cdots\rightarrow\mathrm{F}^{i+1}(M)\leftarrow\mathrm{F}^{i+1}(M)\rightarrow\mathrm{F}^{i}(M)\leftarrow\mathrm{F}^{i}(M)\rightarrow\mathrm{F}^{i-1}(M)\leftarrow\cdots

where the right arrows denote the ι\iota maps and the left arrows denote multiplication by pp.

Now we define a map α:M~R,φRM~R,ψR\alpha:\widetilde{M}\otimes_{R,\varphi}R\to\widetilde{M}\otimes_{R,\psi}R. Let mFi(M)m\in F^{i}(M), which defines an element of M~\widetilde{M}, and it suffices to show α\alpha is an isomorphism. We define:

α(m1)=I()I(m)(φ(X)ψ(X))II!pmin(|I|,i).\alpha(m\otimes 1)=\sum_{I}\nabla(\partial)^{I}(m)\otimes\frac{(\varphi(X)-\psi(X))^{I}}{I!\cdot p^{\min(|I|,i)}}.

Noting that p|I|p^{|I|} divides (φ(X)ψ(x))I(\varphi(X)-\psi(x))^{I}, we can confirm that (φ(X)ψ(X))II!pmin(|I|,i)\frac{(\varphi(X)-\psi(X))^{I}}{I!\cdot p^{\min(|I|,i)}} gives a well-defined element of RR and that this sum converges to 0 in the pp-adic topology. If |I|i|I|\leq i, the fraction is obviously well-defined as it forces |I|p1|I|\leq p-1 and vp(I!)=0.v_{p}(I!)=0. If |I|>i,|I|>i, this can be done by writing each iji_{j} in its base pp representation as ij=k=0aj,kpki_{j}=\sum_{k=0}^{\infty}a_{j,k}p^{k} and using the well-known identity

vp(ij!)=1p1k=0aj,k(pk1).v_{p}(i_{j}!)=\frac{1}{p-1}\sum_{k=0}^{\infty}a_{j,k}(p^{k}-1).

Then the pp-adic valuation of the fraction becomes at least

|I|min(|I|,i)j=1d1p1k=0aj,k(pk1)\displaystyle|I|-\mathrm{min}(|I|,i)-\sum_{j=1}^{d}\frac{1}{p-1}\sum_{k=0}^{\infty}a_{j,k}(p^{k}-1)
=i+j=1dk=1aj,k[pkpk1p1]\displaystyle=-i+\sum_{j=1}^{d}\sum_{k=1}^{\infty}a_{j,k}\left[p^{k}-\frac{p^{k}-1}{p-1}\right]
=i+j=1dk=1aj,k[pk(p2)+1p1]\displaystyle=-i+\sum_{j=1}^{d}\sum_{k=1}^{\infty}a_{j,k}\left[\frac{p^{k}(p-2)+1}{p-1}\right]

We can then see that for p>2p>2, pk(p2)+1>(p1)kp^{k}(p-2)+1>(p-1)^{k}. Since ip1<|I|i\leq p-1<|I|, the pp-adic valuation is nonnegative and thus the original fraction is well-defined, and we can see that |I|vp(I!)|I|-v_{p}(I!) grows without bound as |I||I| does.

As in Faltings, Taylor’s formula φ(r)=Iψ(()I(r))(φ(X)ψ(X))I/I!\varphi(r)=\sum_{I}\psi(\nabla(\partial)^{I}(r))\otimes(\varphi(X)-\psi(X))^{I}/I! shows us that α\alpha gives a well-defined map. The α\alpha’s satisfy transitivity for three different Frobenius lifts by the binomial formula. Applying this to φ,ψ,φ\varphi,\psi,\varphi gives us that the α\alpha are isomorphisms. It is easy to verify that α\alpha is parallel for the connections. ∎

We are now ready to state the analogue of Theorem 11 for a small base ring TT:

Theorem 13.

Suppose MM is in MFbigf(T)\mathrm{MF}_{\mathrm{big}}^{\mathrm{f}}(T) such that each Fi(M)\mathrm{F}^{i}(M) is finite as a TT-module. Then MM is in MFff(T)\mathrm{MF}^{\mathrm{ff}}(T) if and only if (M𝔪,Fi(M𝔪))(Mb𝔪,TT𝔪^,Fi(M)b𝔪,TT𝔪^)MFff(T𝔪^)(M_{\mathfrak{m}},\mathrm{F}^{i}(M_{\mathfrak{m}}))\coloneqq(M\otimes_{b_{\mathfrak{m}},T}\widehat{T_{\mathfrak{m}}},\mathrm{F}^{i}(M)\otimes_{b_{\mathfrak{m}},T}\widehat{T_{\mathfrak{m}}})\in\mathrm{MF}^{\mathrm{ff}}(\widehat{T_{\mathfrak{m}}}) for all maximal ideals 𝔪\mathfrak{m} of TT.

Proof.

Again we show the result in the torsion case and the result in the finite free case follows from taking projective limits.

For the forward direction, it is clear that M𝔪,Fi(M𝔪)M_{\mathfrak{m}},F^{i}(M_{\mathfrak{m}}) are finitely generated pp-power torsion T𝔪^\widehat{T_{\mathfrak{m}}}-modules with Fi(M𝔪)={0}\mathrm{F}^{i}(M_{\mathfrak{m}})=\{0\} for i0i\gg 0 since MM and Fi(M)\mathrm{F}^{i}(M) have these properties over TT. The axioms of MFbig(T𝔪^)\mathrm{MF}_{\mathrm{big}}(\widehat{T_{\mathfrak{m}}}) are verified directly by properties of the tensor product and the fact that b𝔪b_{\mathfrak{m}} is φ\varphi-equivariant with our choice of Frobenius. Recalling that M~\widetilde{M} is a cokernel and again using that b𝔪b_{\mathfrak{m}} is φ\varphi-equivariant, we obtain an isomorphism Mb𝔪,TT𝔪^~M~b𝔪,TT𝔪^.\widetilde{M\otimes_{b_{\mathfrak{m}},T}\widehat{T_{\mathfrak{m}}}}\cong\widetilde{M}\otimes_{b_{\mathfrak{m}},T}\widehat{T_{\mathfrak{m}}}. Thus φM𝔪:M𝔪~T𝔪^,φ𝔪T𝔪^\varphi_{M_{\mathfrak{m}}}:\widetilde{M_{\mathfrak{m}}}\otimes_{\widehat{T_{\mathfrak{m}}},\varphi_{\mathfrak{m}}}\widehat{T_{\mathfrak{m}}} induces an isomorphism, and M𝔪M_{\mathfrak{m}} is an object in MF(T𝔪^).\mathrm{MF}(\widehat{T_{\mathfrak{m}}}).

For the reverse direction, we are assuming that each Fi(M)\mathrm{F}^{i}(M) is a finite TT module, so we need only check that φM:M~T,φTM\varphi_{M}:\widetilde{M}\otimes_{T,\varphi}T\to M is an isomorphism. It suffices to check that this is an isomorphism locally at maximal ideals. Since T𝔪^\widehat{T_{\mathfrak{m}}} is a Noetherian local ring and (p)(p) is an ideal contained in 𝔪\mathfrak{m}, the ring map T𝔪T𝔪^T_{\mathfrak{m}}\to\widehat{T_{\mathfrak{m}}} is faithfully flat. Thus it suffices to check that T𝔪^φM\widehat{T_{\mathfrak{m}}}\otimes\varphi_{M} is an isomorphism for every maximal ideal 𝔪\mathfrak{m}, which is true by assumption since, as in the previous paragraph, M𝔪~Mb𝔪,TT𝔪^~M~b𝔪,TT𝔪^.\widetilde{M_{\mathfrak{m}}}\cong\widetilde{M\otimes_{b_{\mathfrak{m}},T}\widehat{T_{\mathfrak{m}}}}\cong\widetilde{M}\otimes_{b_{\mathfrak{m}},T}\widehat{T_{\mathfrak{m}}}.

2.5. Étale (φ𝒪,𝒪)(\varphi_{\mathcal{O}_{\mathcal{E}}},\mathcal{O}_{\mathcal{E}})-modules

In this section and the next we will work over a general base ring \mathcal{R} as in Convetion 4. Here we review the theory of étale (φ𝒪,𝒪)(\varphi_{\mathcal{O}_{\mathcal{E}}},\mathcal{O}_{\mathcal{E}})-modules in the relative case.

Let 𝔖u\mathfrak{S}_{\mathcal{R}}\coloneqq\mathcal{R}\llbracket u\rrbracket with Frobenius φ𝔖\varphi_{\mathfrak{S}_{\mathcal{R}}} obtained by setting φ𝔖(u)=up.\varphi_{\mathfrak{S}_{\mathcal{R}}}(u)=u^{p}. Let 𝒪,=𝒪\mathcal{O}_{\mathcal{E},\mathcal{R}}=\mathcal{O}_{\mathcal{E}} be the pp-adic completion of 𝔖[u1]\mathfrak{S}_{\mathcal{R}}[u^{-1}] with Frobenius φ𝒪\varphi_{\mathcal{O}_{\mathcal{E}}} extending that on 𝔖.\mathfrak{S}_{\mathcal{R}}.

Definition 14.

An étale (φ𝒪\varphi_{\mathcal{O}_{\mathcal{E}}}, 𝒪\mathcal{O}_{\mathcal{E}})-module is a pair (,φ)(\mathcal{M},\varphi_{\mathcal{M}}) where \mathcal{M} is a finitely generated 𝒪\mathcal{O}_{\mathcal{E}}-module and φ:\varphi_{\mathcal{M}}:\mathcal{M}\to\mathcal{M} is a φ\varphi-semilinear endomorphism such that the linearization 1φ:φ1\otimes\varphi_{\mathcal{M}}:\varphi^{*}\mathcal{M}\to\mathcal{M} (i.e. the map cmcφ(m)c\otimes m\mapsto c\varphi_{\mathcal{M}}(m) for cm𝒪φ𝒪,𝒪c\otimes m\in\mathcal{O}_{\mathcal{E}}\otimes_{\varphi_{\mathcal{O}_{\mathcal{E}}},\mathcal{O}_{\mathcal{E}}}\mathcal{M}) is an isomorphism.

Let Mod𝒪pr\mathrm{Mod}^{pr}_{\mathcal{O}_{\mathcal{E}}} denote the category of projective étale (φ𝒪\varphi_{\mathcal{O}_{\mathcal{E}}}, 𝒪\mathcal{O}_{\mathcal{E}})-modules whose morphisms are φ𝒪\varphi_{\mathcal{O}_{\mathcal{E}}}-compatible 𝒪\mathcal{O}_{\mathcal{E}}-linear maps and Mod𝒪ff\mathrm{Mod}^{ff}_{\mathcal{O}_{\mathcal{E}}} the category with the projective condition replaced by finite free. We now recall how étale (φ𝒪\varphi_{\mathcal{O}_{\mathcal{E}}}, 𝒪\mathcal{O}_{\mathcal{E}})-modules relate to Galois representations.

Recall that πnKtr¯\pi_{n}\in\overline{K^{tr}} is chosen compatibly so that π0=π\pi_{0}=\pi and πn+1p=πn\pi^{p}_{n+1}=\pi_{n}. Set KK_{\infty} to be the pp-adic completion of n=0Ktr(πn)\cup_{n=0}^{\infty}K^{tr}(\pi_{n}) and KK_{\infty}^{\flat} its tilt. Let E+=𝔖/p𝔖E_{\mathcal{R}_{\infty}}^{+}=\mathfrak{S}/p\mathfrak{S} and E~+\tilde{E}_{\mathcal{R}_{\infty}}^{+} denote the uu-adic completion of limφER+\varinjlim_{\varphi}E_{R_{\infty}}^{+}. Write ~\tilde{\mathcal{R}}_{\infty} for W(E~+)W(K)𝒪KW(\tilde{E}_{\mathcal{R}_{\infty}}^{+})\otimes_{W(K_{\infty}^{\flat})}\mathcal{O}_{K_{\infty}}.

Remark 15.

While we have described ~\tilde{\mathcal{R}}_{\infty} in full generality for completeness, for our base ring RR we will only need to use R~\tilde{R}_{\infty} which has a much more explicit description as

R~=n1,1idR(πn,1+tipn)\tilde{R}_{\infty}=\bigcup_{n\geq 1,1\leq i\leq d}R\left(\pi_{n},\sqrt[p^{n}]{1+t_{i}}\right)

for a fixed choice of compatibly chosen pnp^{n}th roots of 1+ti1+t_{i}.

Then we have the following relationship:

Proposition 16.

[Kim15, Prop. 7.7], [Du+24, Prop 2.16] There is a functor \mathcal{M} from the category RepZpff(G~)\mathrm{Rep}_{\mathbb{Z}_{p}}^{\mathrm{ff}}(G_{\tilde{\mathcal{R}}_{\infty}}) of finite free Zp\mathbb{Z}_{p}-modules with continuous G~G_{\tilde{\mathcal{R}}_{\infty}} action to the category Mod𝒪pr\mathrm{Mod}^{pr}_{\mathcal{O}_{\mathcal{E}}} which is an exact equivalence of categories. The inverse of \mathcal{M} is given by

T𝒪()(𝒪^ur𝒪)φ=1.T_{\mathcal{O}_{\mathcal{E}}}(\mathcal{M})\coloneqq(\widehat{\mathcal{O}_{\mathcal{E}}}^{\mathrm{ur}}\otimes_{\mathcal{O}_{\mathcal{E}}}\mathcal{M})^{\varphi=1}.

This functor behaves well with respect to base change in the following sense: Let \mathcal{R}^{\prime} be another base ring equipped with Frobenius satisfying the same conditions as \mathcal{R} with a φ\varphi-equivariant map 00\mathcal{R}_{0}\rightarrow\mathcal{R}_{0}^{\prime}. If TRepZpff(GR~)T\in\mathrm{Rep}_{\mathbb{Z}_{p}}^{\mathrm{ff}}(G_{\tilde{R}_{\infty}}), then TT can be considered as a G~G_{\tilde{\mathcal{R}}_{\infty}^{\prime}}-representation via the map G~G~G_{\tilde{\mathcal{R}}_{\infty}^{\prime}}\rightarrow G_{\tilde{\mathcal{R}}_{\infty}} and we have the isomorphism

𝒪,𝒪(T)~(T)\mathcal{O}_{\mathcal{E},\mathcal{R^{\prime}}}\otimes_{\mathcal{O}_{\mathcal{E}}}\mathcal{M}(T)\cong\mathcal{M}_{\tilde{\mathcal{R}}^{\prime}}(T)

as étale (φ𝒪,\varphi_{\mathcal{O}_{\mathcal{E},\mathcal{R}^{\prime}}}, 𝒪,\mathcal{O}_{\mathcal{E},\mathcal{R}^{\prime}})-modules.

2.6. Kisin Modules

We next review Kisin modules in the relative case and the main result of [Du+24] which provides the key input to the proof of our main theorem.

First we recall the the ring 𝔖(1)\mathfrak{S}^{(1)} considered in [Du+24]. Let Δ\mathcal{R}_{{\mathbbl{\Delta}}} denote the absolute prismatic site of \mathcal{R}. In full generality, 𝔖\mathfrak{S} is the ring 0u\mathcal{R}_{0}\llbracket u\rrbracket and EE is an choice of minimal Eisenstein polynomial of π\pi (in our setting, E(u)=upE(u)=u-p or u+pu+p). We write (𝔖,(E))(\mathfrak{S},(E)) for the Breuil-Kisin prism. Then (𝔖(1),(E))(\mathfrak{S}^{(1)},(E)) is defined to be the self-product of (𝔖,(E))(\mathfrak{S},(E)) in Δ\mathcal{R}_{{\mathbbl{\Delta}}} and that it satisfies a universal property as follows:

If (B,I)Δ(B,I)\in\mathcal{R}_{{\mathbbl{\Delta}}} and if we are given maps f1,f2:(𝔖,(E))(B,I)f_{1},f_{2}:(\mathfrak{S},(E))\to(B,I) such that the maps B/I\mathcal{R}\to B/I induced by f1,f2f_{1},f_{2} agree, then there is a map (𝔖(1),(E))(B,I)(\mathfrak{S}^{(1)},(E))\to(B,I) uniquely determined by f1,f2.f_{1},f_{2}. We will write p1,p2p_{1},p_{2} for the maps (𝔖,(E))(𝔖(1),(E))(\mathfrak{S},(E))\to(\mathfrak{S}^{(1)},(E)). We will also utilize the triple-self product of (𝔖,(E))(\mathfrak{S},(E)) denoted (𝔖(2),(E))(\mathfrak{S}^{(2)},(E)).

The details of the construction of (𝔖(1),(E))(\mathfrak{S}^{(1)},(E)) and a justification that this self-product exists can be found in [Du+24, Example 3.4]. We now introduce the category of free Kisin modules with descent data.

Definition 17 ([Du+24] Definition 3.24).

Let DD𝔖\mathrm{DD}_{\mathfrak{S}} denote the category consisting of triples (𝔐,φ𝔐,f)(\mathfrak{M},\varphi_{\mathfrak{M}},f) where

  • 𝔐\mathfrak{M} is a finite 𝔖\mathfrak{S}-module that is projective away from (p,E)(p,E) and saturated.

  • φ𝔐:𝔐𝔐\varphi_{\mathfrak{M}}:\mathfrak{M}\to\mathfrak{M} is a φ\varphi-semi-linear endomorphism such that (𝔐,φ𝔐)(\mathfrak{M},\varphi_{\mathfrak{M}}) has finite EE-height.

  • f:𝔖(1)p1,𝔖𝔐𝔖(1)p2,𝔖𝔐f:\mathfrak{S}^{(1)}\otimes_{p_{1},\mathfrak{S}}\mathfrak{M}\to\mathfrak{S}^{(1)}\otimes_{p_{2},\mathfrak{S}}\mathfrak{M} is an isomorphism of 𝔖(1)\mathfrak{S}^{(1)}-modules that is compatible with Frobenii and satisfies the cocycle condition over 𝔖(2).\mathfrak{S}^{(2)}.

An object of this category is called an integral Kisin descent datum. If we replace the third piece of data with a map f:𝔖(1)[p1]p1,𝔖𝔐𝔖(1)[p1]p2,𝔖𝔐f:\mathfrak{S}^{(1)}[p^{-1}]\otimes_{p_{1},\mathfrak{S}}\mathfrak{M}\to\mathfrak{S}^{(1)}[p^{-1}]\otimes_{p_{2},\mathfrak{S}}\mathfrak{M} that is an isomorphism of 𝔖(1)[p1]\mathfrak{S}^{(1)}[p^{-1}]-modules that is compatible with Frobenii and satisfies the cocycle condition over 𝔖(2)[p1]\mathfrak{S}^{(2)}[p^{-1}], we call such an object a rational Kisin descent datum. Let DD𝔖,[0,r]ff\mathrm{DD}_{\mathfrak{S},[0,r]}^{\mathrm{ff}} denote the full subcateogry consisting of objects with EE-height r\leq r and which are finite free over 𝔖\mathfrak{S}.

Remark 18.

The category DD𝔖\mathrm{DD}_{\mathfrak{S}} is equivalent to the category of completed prismatic FF-crystals on \mathcal{R} introduced in [Du+24].

The main theorem of [Du+24] is the following:

Theorem 19 ([Du+24] Prop. 3.25, Theorem 3.28).

There is an equivalence of categories between DD𝔖,[0,r]\mathrm{DD}_{\mathfrak{S},[0,r]} and RepZp,[0,r]cris\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}.

A main result of [Du+24] is the association of a Kisin module to a crystalline representation:

Theorem 20.

[Du+24, Theorem 4.19] Let VV be a crystalline Qp\mathbb{Q}_{p} representation of GG_{\mathcal{R}} with Hodge-Tate weights in [0,r][0,r], and let Λ\Lambda be a finite free Zp\mathbb{Z}_{p} lattice of VV stable under the GG_{\mathcal{R}} action. Let \mathcal{M} be the étale (φ𝒪\varphi_{\mathcal{O}_{\mathcal{E}}}, 𝒪\mathcal{O}_{\mathcal{E}})-module associated to Λ\Lambda as in Proposition 16. Then there exists a 𝔖\mathfrak{S} submodule 𝔐\mathfrak{M} of \mathcal{M} stable under Frobenius such that:

  • 𝔐\mathfrak{M} with the induced Frobenius is a Kisin module over 𝔖\mathfrak{S} of EE-height r\leq r.

  • 𝒪𝔖𝔐=\mathcal{O}_{\mathcal{E}}\otimes_{\mathfrak{S}}\mathfrak{M}=\mathcal{M}

  • 𝔐\mathfrak{M} is equipped with the data of 𝔐\nabla_{\mathfrak{M}}, a map

    𝔐:(0φ,0𝔐/u𝔐)[p1](0φ,0𝔐/u𝔐)[p1]0Ω^0\nabla_{\mathfrak{M}}:(\mathcal{R}_{0}\otimes_{\varphi,\mathcal{R}_{0}}\mathfrak{M}/u\mathfrak{M})[p^{-1}]\to(\mathcal{R}_{0}\otimes_{\varphi,\mathcal{R}_{0}}\mathfrak{M}/u\mathfrak{M})[p^{-1}]\otimes_{\mathcal{R}_{0}}\widehat{\Omega}_{\mathcal{R}_{0}}

    which is a topologically quasi-nilpotent connection commuting with Frobenius and satisfying rational SS-Griffiths transversality. (See [Du+24, Def 4.1] and Definition 29 below.)

If 𝔐,,\mathfrak{M},\mathcal{M}, and Λ\Lambda are as above, we call 𝔐\mathfrak{M} a Kisin module associated to Λ\Lambda. On ocassion we will want to forget the \nabla structure on the Kisin module.

Remark 21.

In the setting of our base ring RR, [Du+24, Remark 4.23] shows that the Kisin module associated to a lattice in a crystalline representation is projective. As RR is a local ring, the Kisin module is thus a free 𝔖\mathfrak{S}-module. Furthermore, in the setting of a small base ring TT, the Kisin module associated to a lattice in a crystalline representation is projective (but not necessarily free).

2.7. Base Change of the Kisin Module

Here we will again work over a general base ring \mathcal{R} to state the base change theorems in full generality.

We recall the following standard fact from commutative algebra:

Lemma 22.

[Du+24, Lemma 3.1] Let AA be a ring, QQ a flat AA-module, and N1,N2N_{1},N_{2} submodules of an AA-module NN. Then as submodules of QANQ\otimes_{A}N, we have

QA(N1N2)=(QAN1)(QAN2).Q\otimes_{A}(N_{1}\cap N_{2})=(Q\otimes_{A}N_{1})\cap(Q\otimes_{A}N_{2}).

2.7.1. Shilov point base change of the Kisin module

We will now recall the base change properties of the Kisin module. Let \mathcal{R}^{\prime} be another another general base ring for which there exists a map \mathcal{R}\rightarrow\mathcal{R}^{\prime} compatible with Frobenius which makes \mathcal{R^{\prime}} into an \mathcal{R}-module.

Theorem 23.

Let (𝔐,φ𝔐)(\mathfrak{M},\varphi_{\mathfrak{M}}) be the module and φ𝔐\varphi_{\mathfrak{M}} data of the Kisin module associated to the lattice Λ\Lambda inside the crystalline representation VV as in Theorem 20 with \nabla structure forgotten. Let (𝔐,φ𝔐)(\mathfrak{M}^{\prime},\varphi_{\mathfrak{M}^{\prime}}) be the module and φ𝔐\varphi_{\mathfrak{M}^{\prime}} data of the Kisin module associated to Λ|G\Lambda|_{G_{\mathcal{R}^{\prime}}} with \nabla structure forgotten, and assume 𝔐\mathfrak{M} and 𝔐\mathfrak{M}^{\prime} are finite free. Then 𝔐\mathfrak{M}\otimes_{\mathcal{R}}\mathcal{R}^{\prime} gives the module and φ𝔐\varphi_{\mathfrak{M}^{\prime}} data of the Kisin module associated to Λ|G\Lambda|_{G_{\mathcal{R}^{\prime}}}, i.e. 𝔐𝔐\mathfrak{M}\otimes_{\mathcal{R}}\mathcal{R^{\prime}}\cong\mathfrak{M}^{\prime}.

Proof.

Let \mathcal{M} denote the étale φ\varphi-module associated to Λ\Lambda as in Theorem 20. Let 𝔐\mathfrak{M}^{\prime} denote the Kisin module associated to Λ|G\Lambda|_{G_{\mathcal{R}^{\prime}}} and \mathcal{M}^{\prime} its associated étale φ\varphi-module. We have injections 𝔐\mathfrak{M}\hookrightarrow\mathcal{M} and 𝔐.\mathfrak{M}^{\prime}\hookrightarrow\mathcal{M}^{\prime}. By Proposition 16, we know that 𝒪,𝒪,\mathcal{M}^{\prime}\cong\mathcal{O}_{\mathcal{E},\mathcal{R}^{\prime}}\otimes_{\mathcal{O}_{\mathcal{E},\mathcal{R}}}\mathcal{M} and so we also have an injection

𝔐𝔖𝔖.\mathfrak{M}\otimes_{\mathfrak{S}_{\mathcal{R}}}\mathfrak{S}_{\mathcal{R^{\prime}}}\hookrightarrow\mathcal{M}^{\prime}.

Then 𝔐𝔖𝔖\mathfrak{M}\otimes_{\mathfrak{S}_{\mathcal{R}}}\mathfrak{S}_{\mathcal{R^{\prime}}} and 𝔐\mathfrak{M}^{\prime} are both the module data of Kisin modules lying inside of \mathcal{M^{\prime}}. We claim there is unique such module data. By [Du+24, Lemma 4.18] and Proposition 16 there is a unique Kisin module 𝔐L\mathfrak{M}_{L} associated to the étale φ𝒪,L\varphi_{\mathcal{O}_{\mathcal{E},L}}-module L=𝒪,𝒪,L\mathcal{M}_{L}=\mathcal{M}^{\prime}\otimes_{\mathcal{O}_{\mathcal{E},\mathcal{R}^{\prime}}}\mathcal{O}_{\mathcal{E},\mathcal{R}_{L}^{\prime}} and thus

𝔐𝔖𝔖𝔖L𝔖L=𝔐𝔖𝔖L\mathfrak{M}\otimes_{\mathfrak{S}_{\mathcal{R}}}\mathfrak{S}_{\mathcal{R^{\prime}}}\otimes_{\mathfrak{S}_{\mathcal{R}_{L}^{\prime}}}\mathfrak{S}_{\mathcal{R}^{\prime}_{L}}=\mathfrak{M}^{\prime}\otimes_{\mathfrak{S}_{\mathcal{R}^{\prime}}}\mathfrak{S}_{\mathcal{R}_{L}^{\prime}}

as submodules of L.\mathcal{M}_{L}. We also know that

𝔐𝔖𝔖𝔖𝒪,𝔐𝔖𝒪,.\mathfrak{M}\otimes_{\mathfrak{S}_{\mathcal{R}}}\mathfrak{S}_{\mathcal{R^{\prime}}}\otimes_{\mathfrak{S}_{\mathcal{R^{\prime}}}}\mathcal{O}_{\mathcal{E},\mathcal{R}^{\prime}}\cong\mathcal{M}^{\prime}\cong\mathfrak{M}^{\prime}\otimes_{\mathfrak{S}_{\mathcal{R}^{\prime}}}\mathcal{O}_{\mathcal{E},\mathcal{R}^{\prime}}.

Thus we have 𝔐𝔖𝔖\mathfrak{M}\otimes_{\mathfrak{S}_{\mathcal{R}}}\mathfrak{S}_{\mathcal{R}^{\prime}} contained in the intersection of 𝔐𝔖𝔖L\mathfrak{M}^{\prime}\otimes_{\mathfrak{S}_{\mathcal{R}^{\prime}}}\mathfrak{S}_{L}^{\prime} and 𝔐𝔖𝒪,\mathfrak{M}^{\prime}\otimes_{\mathfrak{S}_{\mathcal{R}^{\prime}}}\mathcal{O}_{\mathcal{E},\mathcal{R}^{\prime}}. By Lemma 22 and the fact that 𝔖L𝒪,=𝔖\mathfrak{S}_{\mathcal{R}_{L}^{\prime}}\cap\mathcal{O}_{\mathcal{E},\mathcal{R}^{\prime}}=\mathfrak{S}_{\mathcal{R}^{\prime}}, this intersection is 𝔐\mathfrak{M}^{\prime}, concluding the proof as 𝔐\mathfrak{M}^{\prime} and 𝔐\mathfrak{M} have the same rank and our maps are all compatible with the φ\varphi structures.

Working in the setting of RR, set 𝔖0W(k)u\mathfrak{S}_{0}\coloneqq W(k)\llbracket u\rrbracket and let S0S_{0} be the pp-adically completed divided power envelope of 𝔖0\mathfrak{S}_{0} with respect to E(u)E(u). We have induced maps 𝔖R𝔖0\mathfrak{S}_{R}\to\mathfrak{S}_{0} and 𝒪,R𝒪,W(k)\mathcal{O}_{\mathcal{E},R}\to\mathcal{O}_{\mathcal{E},W(k)} which confirm that if \mathcal{M} is an étale (φ𝒪,R,𝒪,R)(\mathcal{\varphi}_{\mathcal{O}_{\mathcal{E},R}},\mathcal{O}_{\mathcal{E},R})-module corresponding to a Zp\mathbb{Z}_{p}-stable lattice inside a crystalline representation of GRG_{R} then 0𝒪,W(k)\mathcal{M}_{0}\coloneqq\mathcal{M}\otimes\mathcal{O}_{\mathcal{E},W(k)} is an étale (φ𝒪,W(k),𝒪,W(k))(\mathcal{\varphi}_{\mathcal{O}_{\mathcal{E},W(k)}},\mathcal{O}_{\mathcal{E},W(k)})-module corresponding to a Zp\mathbb{Z}_{p}-stable lattice inside a crystalline representation of GK.G_{K}. Then by [Kis06, Lemma 2.1.15] or as a corollary of Theorem 23, we have

Corollary 23.1.

If 𝔐\mathfrak{M} is a Kisin module associated to a Zp\mathbb{Z}_{p}-stable lattice in a crystalline representation of GRG_{R}, then 𝔐𝔖𝔖0\mathfrak{M}\otimes_{\mathfrak{S}}\mathfrak{S}_{0} is, in the classical setting of [Kis06], a Kisin module associated to a Zp\mathbb{Z}_{p} stable lattice in a crystalline representation of GKG_{K}.

Remark 24.

Here we are using that 𝔐\mathfrak{M} is free by Remark 21 and that r[0,p2]r\in[0,p-2], which is why the corollary has only been stated for the base ring RR.

2.8. Breuil Modules

In this section we will work over 𝒯\mathcal{T} which is either a small base ring or the power series ring RR. We define S𝒯S_{\mathcal{T}} to be the pp-adically completed divided power envelope of 𝔖𝒯𝒯u\mathfrak{S}_{\mathcal{T}}\coloneqq\mathcal{T}\llbracket u\rrbracket with respect to E(u)E(u) a choice of minimal polynomial of pp. The ring S𝒯S_{\mathcal{T}} comes equipped with a Frobenius map φS𝒯\varphi_{S_{\mathcal{T}}} that extends φ𝔖𝒯.\varphi_{\mathfrak{S}_{\mathcal{T}}}. It also comes with its divided power filtration that we will denote FiliS𝒯.\mathrm{Fil}^{i}S_{\mathcal{T}}. Note that for 1ip11\leq i\leq p-1, φS𝒯(FiliS𝒯)piS\varphi_{S_{\mathcal{T}}}(\mathrm{Fil}^{i}S_{\mathcal{T}})\subset p^{i}S (just consider the action of φS𝒯\varphi_{S_{\mathcal{T}}} on E(u)ii!\frac{E(u)^{i}}{i!}), so we can set φS𝒯,i=φS𝒯pi:FiliS𝒯S𝒯.\varphi_{S_{\mathcal{T}},i}=\frac{\varphi_{S_{\mathcal{T}}}}{p^{i}}:\mathrm{Fil}^{i}S_{\mathcal{T}}\to S_{\mathcal{T}}. The ring S𝒯S_{\mathcal{T}} also comes equipped with an integrable connection S𝒯:S𝒯S𝒯𝒯Ω^𝒯.\nabla_{S_{\mathcal{T}}}:S_{\mathcal{T}}\to S_{\mathcal{T}}\otimes_{\mathcal{T}}\widehat{\Omega}_{\mathcal{T}}.

Definition 25.

For rp1r\leq p-1 a positive integer, let ModS,ff,r(𝒯)\mathrm{Mod}_{{S},\nabla}^{\mathrm{ff},\mathrm{r}}(\mathcal{T}) be the category whose objects are quadruples (,Fili(),,φ,i)(\mathscr{M},\mathrm{Fil}^{i}(\mathscr{M}),\nabla_{\mathscr{M}},\varphi_{\mathscr{M},i}) where:

  • \mathscr{M} is a finite, projective S𝒯S_{\mathcal{T}}-module.

  • Fili()\mathrm{Fil}^{i}(\mathscr{M}) is a decreasing filtration of \mathscr{M} with Fil0()=\mathrm{Fil}^{0}(\mathscr{M})=\mathscr{M}, Filr+1()(Fil1S𝒯)\mathrm{Fil}^{r+1}(\mathscr{M})\subset(\mathrm{Fil}^{1}S_{\mathcal{T}})\mathscr{M}, and (FiliS𝒯)FiliFili+1.(\mathrm{Fil}^{i}S_{\mathcal{T}})\mathrm{Fil}^{i}\mathscr{M}\subset\mathrm{Fil}^{i+1}\mathscr{M}.

  • φ,i\varphi_{\mathscr{M},i} are φS\varphi_{S}-semilinear maps φ,i:Fili()\varphi_{\mathscr{M},i}:\mathrm{Fil}^{i}(\mathscr{M})\to\mathscr{M} so that the composite Fili()Fili1()φ,i1\mathrm{Fil}^{i}(\mathscr{M})\to\mathrm{Fil}^{i-1}(\mathscr{M})\xrightarrow{\varphi_{\mathscr{M},i-1}}\mathscr{M} is pφ,i.p\varphi_{\mathscr{M},i}. Also, φr(Filr)\varphi_{r}(\mathrm{Fil}^{r}\mathscr{M}) generates .\mathscr{M}.

  • \nabla_{\mathscr{M}} is a topologically quasi-nilpotent integrable connection which satisfies SS-Griffiths transversality (see Definition 29 below) and commutes with each φ,i\varphi_{\mathscr{M},i} as before.

We refer to objects of this category as Breuil modules.

Given a Kisin module 𝔐\mathfrak{M}, we define the associated Breuil module to be

(𝔐)𝔐𝔖𝒯,φ𝔖𝒯S𝒯.\mathscr{M}(\mathfrak{M})\coloneqq\mathfrak{M}\otimes_{\mathfrak{S}_{\mathcal{T}},\varphi_{\mathfrak{S}_{\mathcal{T}}}}S_{\mathcal{T}}.

If such a Kisin module exists for a given Breuil module \mathscr{M}, we will say that “\mathscr{M} arises from a Kisin module.” In the next few sections, we show how the full data of a Breuil module can be constructed on 𝔖𝒯,φ𝔖𝒯S𝒯.\mathscr{M}\otimes_{\mathfrak{S}_{\mathcal{T}},\varphi_{\mathfrak{S}_{\mathcal{T}}}}S_{\mathcal{T}}.

2.8.1. The filtration from the Kisin module

Definition 26.

Let \mathscr{M} be a Breuil module that arises from a Kisin module 𝔐\mathfrak{M} which arises from a crystalline representation. We define a decreasing filtration on [p1]\mathscr{M}[p^{-1}] via:

Fili([p1]){m[p1]|(1φ𝔐)(m)Fili(S𝒯[p1])𝔖𝒯𝔐}\mathrm{Fil}^{i}(\mathscr{M}[p^{-1}])\coloneqq\{m\in\mathscr{M}[p^{-1}]\leavevmode\nobreak\ |\leavevmode\nobreak\ (1\otimes\varphi_{\mathfrak{M}})(m)\in\mathrm{Fil}^{i}(S_{\mathcal{T}}[p^{-1}])\otimes_{\mathfrak{S}_{\mathcal{T}}}\mathfrak{M}\}

and a filtration on \mathscr{M} via

Fili()Fili([p1]).\mathrm{Fil}^{i}(\mathscr{M})\coloneqq\mathscr{M}\cap\mathrm{Fil}^{i}(\mathscr{M}[p^{-1}]).

We have the following lemma which shows that the filtration structure descends integrally:

Lemma 27.

Let \mathscr{M} be a Breuil module that arises from a Kisin module 𝔐\mathfrak{M} and let 0i<p10\leq i<p-1. We can express the filtration on \mathscr{M} as in Definition 26 as

Fili()={m|(1φ𝔐)(m)Fili(S𝒯)𝔖𝔐}.\mathrm{Fil}^{i}(\mathscr{M})=\{m\in\mathscr{M}|(1\otimes\varphi_{\mathfrak{M}})(m)\in\mathrm{Fil}^{i}(S_{\mathcal{T}})\otimes_{\mathfrak{S}}\mathfrak{M}\}.
Proof.

Observe that Fili()Fili([p1]),\mathrm{Fil}^{i}(\mathscr{M})\subset\mathrm{Fil}^{i}(\mathscr{M}[p^{-1}]), so we can realize an element of Fili()\mathrm{Fil}^{i}(\mathscr{M}) as an element mm\in\mathscr{M} such that (1φ𝔐)(m)(1\otimes\varphi_{\mathfrak{M}})(m) is contained in

(Fili(S[p1])𝔖𝒯𝔐)(S𝒯𝔖𝒯𝔐)=(Fili(S𝒯[p1])S)𝔖𝔐(\mathrm{Fil}^{i}(S[p^{-1}])\otimes_{\mathfrak{S}_{\mathcal{T}}}\mathfrak{M})\cap(S_{\mathcal{T}}\otimes_{\mathfrak{S}_{\mathcal{T}}}\mathfrak{M})=(\mathrm{Fil}^{i}(S_{\mathcal{T}}[p^{-1}])\cap S)\otimes_{\mathfrak{S}}\mathfrak{M}

where we are relying on Lemma 22 (note 𝔐\mathfrak{M} is projective and hence flat). By comparing elements term-by-term we can show Fili(S𝒯[p1])S𝒯=FiliS𝒯\mathrm{Fil}^{i}(S_{\mathcal{T}}[p^{-1}])\cap S_{\mathcal{T}}=\mathrm{Fil}^{i}S_{\mathcal{T}}, as needed. Specifically, an element in S𝒯[p1]S_{\mathcal{T}}[p^{-1}] can be written as

x=k=0E(u)kk!j=0deg(E)1akjujx=\sum_{k=0}^{\infty}\frac{E(u)^{k}}{k!}\sum_{j=0}^{\mathrm{deg}(E)-1}a_{kj}u^{j}

with akj𝒯[p1].a_{kj}\in\mathcal{T}[p^{-1}]. For this element to be in Fili(S𝒯[p1])\mathrm{Fil}^{i}(S_{\mathcal{T}}[p^{-1}]) we must have akj=0a_{kj}=0 whenever 0k<i0\leq k<i. But xS𝒯x\in S_{\mathcal{T}} with each akj=0a_{kj}=0 whenever 0k<i0\leq k<i, so xFili(S𝒯)x\in\mathrm{Fil}^{i}(S_{\mathcal{T}}), as well. ∎

2.8.2. The φi\varphi_{i} structure

Definition 28.

Let \mathscr{M} be a Breuil module that arises from a Kisin module 𝔐\mathfrak{M}. We can also define a φ,i\varphi_{\mathscr{M},i} structure by setting φ,i\varphi_{\mathscr{M},i} to be the composition:

φ,i:Fili()1φ𝔐Fili(S𝒯)𝔖𝔐φS𝒯,i1.\varphi_{\mathscr{M},i}:\mathrm{Fil}^{i}(\mathscr{M})\xrightarrow{1\otimes\varphi_{\mathfrak{M}}}\mathrm{Fil}^{i}(S_{\mathcal{T}})\otimes_{\mathfrak{S}}\mathfrak{M}\xrightarrow{\varphi_{S_{\mathcal{T}},i}\otimes 1}\mathscr{M}.

Note the first arrow is well-defined by the definition of Fili().\mathrm{Fil}^{i}(\mathscr{M}). It is then clear that the composite Fili()Fili1()φ,i1\mathrm{Fil}^{i}(\mathscr{M})\to\mathrm{Fil}^{i-1}(\mathscr{M})\xrightarrow{\varphi_{\mathscr{M},i-1}}\mathscr{M} is pφ,ip\varphi_{\mathscr{M},i} because φS𝒯,i=φS𝒯pi.\varphi_{S_{\mathcal{T}},i}=\frac{\varphi_{S_{\mathcal{T}}}}{p^{i}}.

2.8.3. The integrable connection

Here we discuss where the connection

[p1]:[p1][p1]RΩ^R\nabla_{\mathscr{M}[p^{-1}]}:\mathscr{M}[p^{-1}]\to\mathscr{M}[p^{-1}]\otimes_{R}\widehat{\Omega}_{R}

arises from, and then we will show it descends to the integral setting.

Let \mathscr{M} be a Breuil module that arises from a Kisin module with descent datum 𝔐.\mathfrak{M}. By Theorem 19 there exists Λ\Lambda, a lattice inside a crystalline representation VΛZpQpV\coloneqq\Lambda\otimes_{\mathbb{Z}_{p}}\mathbb{Q}_{p} corresponding to 𝔐\mathfrak{M} under the equivalence of categories. Then we consider Dcris(V)D_{\mathrm{cris}}^{\vee}(V) which is a finite projective 𝒯[p1]\mathcal{T}[p^{-1}]-module equipped with an integrable connection D\nabla_{D}, a φ\varphi, and filtration structure which we will denote Fili(Dcris(V))\mathrm{Fil}^{i}(D_{\mathrm{cris}}^{\vee}(V)). Set M=𝒯φ,𝒯𝔐/u𝔐M=\mathcal{T}\otimes_{\varphi,\mathcal{T}}\mathfrak{M}/u\mathfrak{M}. Then by [Du+24, Theorem 4.28] there is a φ\varphi-compatible isomorphism M[p1]Dcris(V)M[p^{-1}]\to D_{\mathrm{cris}}^{\vee}(V). By [Du+24, Theorem 4.2] we can identify 𝒟S𝒯[p1]𝒯M\mathscr{D}\coloneqq S_{\mathcal{T}}[p^{-1}]\otimes_{\mathcal{T}}M with [p1]\mathscr{M}[p^{-1}] as modules. Also see Lemma 43.

The connection D\nabla_{D} can be further extended to an integrable connection on 𝒟Dcris𝒯[p1]S𝒯[p1]\mathscr{D}\coloneqq D_{\mathrm{cris}}^{\vee}\otimes_{\mathcal{T}[p^{-1}]}S_{\mathcal{T}}[p^{-1}] defined as [p1]D1+1S𝒯[p1]\nabla_{\mathscr{M}[p^{-1}]}\coloneqq\nabla_{D}\otimes 1+1\otimes\nabla_{S_{\mathcal{T}}[p^{-1}]}, and 𝒟\mathscr{D} comes equipped with a filtration defined inductively by setting Fil0𝒟=𝒟\mathrm{Fil}^{0}\mathscr{D}=\mathscr{D} and:

Fili𝒟{x𝒟|Nu(x)Fili1𝒟,q(x)Fili(Dcris(V))},\mathrm{Fil}^{i}\mathscr{D}\coloneqq\{x\in\mathscr{D}|N_{u}(x)\in\mathrm{Fil}^{i-1}\mathscr{D},q(x)\in\mathrm{Fil}^{i}(D_{\mathrm{cris}}^{\vee}(V))\},

where Nu:𝒟𝒟N_{u}:\mathscr{D}\to\mathscr{D} is the 𝒯\mathcal{T}-linear derivation Nu,S𝒯1N_{u,S_{\mathcal{T}}}\otimes 1 and q(x)q(x) is the projection M\mathscr{M}\to M as in [Du+24, Theorem 4.2]. By [Du+24, Lemma 4.31], the filtration Fili([p1])\mathrm{Fil}^{i}(\mathscr{M}[p^{-1}]) defined in Definition 26 using the φ\varphi-structure on 𝔐\mathfrak{M} and the filtration Fili(𝒟)\mathrm{Fil}^{i}(\mathscr{D}) defined above agree and, consequently, [p1]\nabla_{\mathscr{M}[p^{-1}]} satisfies rational S𝒯S_{\mathcal{T}}-Griffiths transversality as we define below:

Definition 29.

We say the connection [p1]\nabla_{\mathscr{M}[p^{-1}]} satisfies rational S𝒯S_{\mathcal{T}}-Griffiths transversality if for every ii

u(Fili+1([p1])Fili([p1])\partial_{u}(\mathrm{Fil}^{i+1}(\mathscr{M}[p^{-1}])\subset\mathrm{Fil}^{i}(\mathscr{M}[p^{-1}])

and

[p1](Fili+1([p1])Fili([p1])RΩ^R\nabla_{\mathscr{M}[p^{-1}]}(\mathrm{Fil}^{i+1}(\mathscr{M}[p^{-1}])\subset\mathrm{Fil}^{i}(\mathscr{M}[p^{-1}])\otimes_{R}\widehat{\Omega}_{R}

where u:[p1][p1]\partial_{u}:\mathscr{M}[p^{-1}]\to\mathscr{M}[p^{-1}] is the derivation given by u,S𝒯[p1]1.\partial_{u,S_{\mathcal{T}}[p^{-1}]}\otimes 1.

If [p1]\nabla_{\mathscr{M}[p^{-1}]} restricts to a connection :𝒯Ω^𝒯\nabla_{\mathscr{M}}:\mathscr{M}\to\mathscr{M}\otimes_{\mathcal{T}}\widehat{\Omega}_{\mathcal{T}} where the same condition on the filtration holds with Fil([p1])\mathrm{Fil}^{\bullet}(\mathscr{M}[p^{-1}]) replaced by Fil()\mathrm{Fil}^{\bullet}(\mathscr{M}), we say that \nabla_{\mathscr{M}} satisfies integral S𝒯S_{\mathcal{T}}-Griffiths transversality.

We now show this connection descends to the integral setting.

Proposition 30.

Let \mathscr{M} be a Breuil module that arises from a Kisin module 𝔐\ \mathfrak{M}. Then \mathscr{M} is stable under \nabla_{\mathscr{M}} and satisfies integral SS-Griffiths transversality.

Proof.

We consider the base change LSS𝒯,L\mathscr{M}_{L}\coloneqq\mathscr{M}\otimes_{S}S_{\mathcal{T},L} to the Shilov point with imperfect residue field (defined in Section 2.2.1. Let VV be the crystalline representation corresponding to the étale (φ𝒪,𝒪)(\varphi_{\mathcal{O}_{\mathcal{E}}},\mathcal{O}_{\mathcal{E}})-module 𝔐𝔖𝒯𝒪\mathfrak{M}\otimes_{\mathfrak{S}_{\mathcal{T}}}\mathcal{O}_{\mathcal{E}}. By [Moo24, Theorem 4.20] there is a Breuil module L\mathscr{M}_{L}^{\prime} associated to V|GLV|_{G_{L}}.

By [Moo24, p. 10] there exists a Kisin module 𝔐L\mathfrak{M}_{L}^{\prime} such that L=𝔐Lφ𝔖𝒯,L,𝔖𝒯,LS𝒯,L.\mathscr{M}_{L}^{\prime}=\mathfrak{M}_{L}^{\prime}\otimes_{\varphi_{\mathfrak{S}_{\mathcal{T},L}},\mathfrak{S}_{\mathcal{T},L}}S_{\mathcal{T},L}. By its uniqueness and Theorem 23 we must have 𝔐L𝔐𝔖𝒯𝔖𝒯,L\mathfrak{M}_{L}^{\prime}\cong\mathfrak{M}\otimes_{\mathfrak{S}_{\mathcal{T}}}\mathfrak{S}_{\mathcal{T},L} and thus L𝔐φ,𝔖𝒯S𝒯,LL.\mathscr{M}_{L}^{\prime}\cong\mathfrak{M}\otimes_{\varphi,\mathfrak{S}_{\mathcal{T}}}S_{\mathcal{T},L}\cong\mathscr{M}_{L}.

Thus L\mathscr{M}_{L} is stable under its integrable connection L=1+1SL.\nabla_{\mathscr{M}_{L}}=\nabla_{\mathscr{M}}\otimes 1+1\otimes\nabla_{S_{L}}. Thus the image of \mathscr{M} under [p1]\nabla_{\mathscr{M}[p^{-1}]} is contained in [p1]L.\mathscr{M}[p^{-1}]\cap\mathscr{M}_{L}. By Lemma 22:

[p1]L\displaystyle\mathscr{M}[p^{-1}]\cap\mathscr{M}_{L} =(SS𝒯[p1])(S𝒯S𝒯,L)\displaystyle=(\mathscr{M}\otimes_{S}S_{\mathcal{T}}[p^{-1}])\cap(\mathscr{M}\otimes_{S_{\mathcal{T}}}S_{\mathcal{T},L})
=S𝒯(S𝒯[p1]S𝒯,L)\displaystyle=\mathscr{M}\otimes_{S_{\mathcal{T}}}(S_{\mathcal{T}}[p^{-1}]\cap S_{\mathcal{T},L})
=\displaystyle=\mathscr{M}

where S𝒯[p1]S𝒯,L=S𝒯S_{\mathcal{T}}[p^{-1}]\cap S_{\mathcal{T},L}=S_{\mathcal{T}} can be checked term-by-term as in the proof of Lemma 43 below.

For integral SS-Griffiths Transversality, we know (Fili+1())FiliRΩ^R\nabla_{\mathscr{M}}(\mathrm{Fil}^{i+1}(\mathscr{M}))\subset\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}\widehat{\Omega}_{R} as it satisfies rational SS-Griffiths transversality and that (Fili+1())RΩ^R\nabla_{\mathscr{M}}(\mathrm{Fil}^{i+1}(\mathscr{M}))\subset\mathscr{M}\otimes_{R}\widehat{\Omega}_{R} because \mathscr{M} is stable under \nabla_{\mathscr{M}}. We conclude that \nabla_{\mathscr{M}} satisfies integral SS-Griffiths Transversality from Definition 26. ∎

2.8.4. The functor from Fontaine-Laffaille modules to Breuil modules

In the classical setting, it has already been established as the main result of [Gao19, Theorem 1.1] that there is a “direct” equivalence of categories between Fontaine-Laffaille modules and Breuil modules. We rely on this equivalence when reducing the relative setting to the classical setting. It is also well-known that the same story holds in the classical case as in the above section. There is an equivalence of categories between Kisin modules and lattices inside of crystalline representations. If Λ\Lambda is a lattice associated to Kisin module 𝔐\mathfrak{M}, there is a functor DcrisD_{\mathrm{cris}} such that Dcris(ΛZpQp)(𝔐/u𝔐)[p1]D_{\mathrm{cris}}(\Lambda\otimes_{\mathbb{Z}_{p}}\mathbb{Q}_{p})\cong(\mathfrak{M}/u\mathfrak{M})[p^{-1}] where M𝔐/u𝔐M\coloneqq\mathfrak{M}/u\mathfrak{M} is a Fontaine-Laffaille module in the classical sense of [FL82].

In the relative setting there is an analogous “obvious” functor from Fontaine-Laffaille modules to Breuil modules.

Definition 31.

Let MMF,[0,r]ffM\in\mathrm{MF}_{\nabla,[0,r]}^{\mathrm{ff}}. Then we can define the functor (M)=M^𝒯S𝒯\mathscr{M}\coloneqq\mathscr{M}(M)=M\widehat{\otimes}_{\mathcal{T}}S_{\mathcal{T}}, the pp-adic completion of M𝒯S.M\otimes_{\mathcal{T}}S. The additional structure is defined in the natural way as follows:

  • φφMφS𝒯\varphi_{\mathscr{M}}\coloneqq\varphi_{M}\otimes\varphi_{S_{\mathcal{T}}}

  • M1+1S𝒯\nabla_{\mathscr{M}}\coloneqq\nabla_{M}\otimes 1+1\otimes\nabla_{S_{\mathcal{T}}}

  • Filr()i=0rFili(M)^RFilri(S𝒯)\mathrm{Fil}^{r}(\mathscr{M})\coloneqq\sum_{i=0}^{r}\mathrm{Fil}^{i}(M)\widehat{\otimes}_{R}\mathrm{Fil}^{r-i}(S_{\mathcal{T}})

  • φri=0rφi,Mφri,S𝒯\varphi_{r}\coloneqq\sum_{i=0}^{r}\varphi_{i,M}\otimes\varphi_{r-i,S_{\mathcal{T}}}

where we’ve added an “MM” tag to a structure if it arises from that appropriate structure on MM and an “SS” tag if it arises from the appropriate structure on S𝒯S_{\mathcal{T}}. Much of our focus will be on trying to reverse this functor: obtaining Fontaine-Laffaille module data from Breuil module data.

In the classical setting, this functor is well-known to be an equivalence of categories and Hui Gao established a direct quasi-inverse in [Gao19, Theorem 1.1]

Remark 32.

Our results will imply that this functor M(M)M\mapsto\mathscr{M}(M) is an equivalence of categories in the setting with base ring 𝒯\mathcal{T}, as well.

2.9. Functors to Galois Representations

2.9.1. Definition of TcrisT_{\mathrm{cris}}

Let MMFff,[0,r](R).M\in\mathrm{MF}^{\mathrm{ff},[0,r]}_{\nabla}(R). Now we recall how the Galois action on Acris(R)RMA_{\mathrm{cris}}(R)\otimes_{R}M and hence

Tcris(M)(Filr(AcrisRM))φr=1T_{\mathrm{cris}}(M)\coloneqq(\mathrm{Fil}^{r}(A_{\mathrm{cris}}\otimes_{R}M))^{\varphi_{r}=1}

is defined. This is discussed in [LMP23, Section 2.3]. Set φ\varphi on RR to extend the natural φ\varphi on W(k)W(k) with φ(ti)=(1+ti)p1.\varphi(t_{i})=(1+t_{i})^{p}-1. Choose 1+ti~R¯\widetilde{1+t_{i}}\in\overline{R}^{\flat} to be a fixed sequence of compatible pip^{i}th roots of 1+ti1+t_{i} and denote by [1+ti~][\widetilde{1+t_{i}}] its Teichmüller lift. For any gGRg\in G_{R} we define βi(g)=g([1+ti~])[1+ti~]FiliAcris(R)\beta_{i}(g)=g([\widetilde{1+t_{i}}])-[\widetilde{1+t_{i}}]\in\mathrm{Fil}^{i}A_{\mathrm{cris}}(R), and we have for axAcris(R)RMa\otimes x\in A_{\mathrm{cris}}(R)\otimes_{R}M:

g(ax)=Ig(a)γI(β(g))()I(x)g(a\otimes x)=\sum_{I}g(a)\gamma_{I}(\beta(g))\otimes\nabla(\partial)^{I}(x)

where γI(β(g))\gamma_{I}(\beta(g)) is defined as β(g)ii!\frac{\prod_{\ell}\beta_{\ell}(g)^{i_{\ell}}}{\prod_{\ell}i_{\ell}!} and I=(i1,,id)I=(i_{1},\cdots,i_{d}) is a multi-index with ik0i_{k}\geq 0 for all kk and ()I=[(ti)]i1[(td)]id\nabla(\partial)^{I}=\left[\nabla(\frac{\partial}{\partial t_{i}})\right]^{i_{1}}\cdots\left[\nabla(\frac{\partial}{\partial t_{d}})\right]^{i_{d}}. It follows from [LMP23, Lemma 2.20] that this formula gives a well-defined Acris(R)A_{\mathrm{cris}}(R)-semilinear action of GRG_{R} which is compatible with the filtration structure (i.e. g(Acris(R)RFilrM)Filr(Acris(R)RM)g(A_{\mathrm{cris}}(R)\otimes_{R}\mathrm{Fil}^{r}M)\subset\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{R}M)) and commutes with the Frobenius structure. This ensures Tcris(M)T_{\mathrm{cris}}(M) is well-defined and that Tcris(M)Repcris,Zp[0,r](GR)T_{\mathrm{cris}}(M)\in\mathrm{Rep}_{\mathrm{cris},\mathbb{Z}_{p}}^{[0,r]}(G_{R}).

Tcris(M)T_{\mathrm{cris}}(M) can be defined analogously for MMFff,[0,r](T),M\in\mathrm{MF}^{\mathrm{ff},[0,r]}_{\nabla}(T), but φ(ti)\varphi(t_{i}) is set to be tipt_{i}^{p} and we use ti~T¯\widetilde{t_{i}}\in\overline{T}^{\flat} instead of 1+ti~\widetilde{1+t_{i}}.

2.9.2. The definition of T𝔖(𝔐)T_{\mathfrak{S}}(\mathfrak{M})

Write 𝔖\mathfrak{S} for 𝔖R\mathfrak{S}_{R}. Let (𝔐,φ𝔐,f)(\mathfrak{M},\varphi_{\mathfrak{M}},f) be a Kisin module with descent data associated to a crystalline representation. Here we will work over our base ring RR and explain how the map ff as in Definition 17 is used to construct a Galois action on 𝔐𝔖Ainf(R)\mathfrak{M}\otimes_{\mathfrak{S}}A_{\mathrm{inf}}(R) as in [DL22, 4.3]. First recall that we have an embedding RAinf(R)R\hookrightarrow A_{\mathrm{inf}}(R) via ti[ti+1~]1t_{i}\mapsto[\widetilde{t_{i}+1}]-1.

Let f1:(𝔖,(E))(Ainf(R),(E))f_{1}:(\mathfrak{S},(E))\to(A_{\mathrm{inf}}(R),(E)) be the map determined by u[π]u\mapsto[\pi^{\flat}] which extends the above embedding of RR into Ainf(R)A_{\mathrm{inf}}(R). Let gGRg\in G_{R} and let f2:(𝔖,(E))(Ainf(R),(E))f_{2}:(\mathfrak{S},(E))\to(A_{\mathrm{inf}}(R),(E)) be the map determined by ug([π])u\mapsto g([\pi^{\flat}]) and tig([ti+1~]1).t_{i}\mapsto g([\widetilde{t_{i}+1}]-1). Then the universal property described in Section 2.6 gives a unique map fg:(𝔖(1),(E))(Ainf(R),(E))f_{g}:(\mathfrak{S}^{(1)},(E))\to(A_{\mathrm{inf}}(R),(E)) and since ff satisfies the cocycle condition there exists a map dd making the following diagram commute:

𝔖(1)p1,𝔖𝔐{\mathfrak{S}^{(1)}\otimes_{p_{1},\mathfrak{S}}\mathfrak{M}}𝔖(1)p2,𝔖𝔐{\mathfrak{S}^{(1)}\otimes_{p_{2},\mathfrak{S}}\mathfrak{M}}Ainf(R)𝔖𝔐{A_{\mathrm{inf}}(R)\otimes_{\mathfrak{S}}\mathfrak{M}}Ainf(R)g,𝔖𝔐{A_{\mathrm{inf}}(R)\otimes_{g,\mathfrak{S}}\mathfrak{M}}f\scriptstyle{f}f1\scriptstyle{f_{1}}fg\scriptstyle{f_{g}}dg\scriptstyle{d_{g}}

and from this we have a GRG_{R} action on Ainf(R)𝔖𝔐.A_{\mathrm{inf}}(R)\otimes_{\mathfrak{S}}\mathfrak{M}. Alsom the GRG_{R} action commutes with the φ\varphi-structure.

We will make use of this GRG_{R} action on two separate functors. We define

T𝔖(𝔐)(𝔐𝔖W(R¯[1/u]))φ=1T_{\mathfrak{S}}(\mathfrak{M})\coloneqq(\mathfrak{M}\otimes_{\mathfrak{S}}W(\overline{R}^{\flat}\left[1/u\right])\leavevmode\nobreak\ )^{\varphi=1}

and

T𝔖r(𝔐)(Filrφ𝔐𝔖Ainf(R))φr=1T^{r}_{\mathfrak{S}}(\mathfrak{M})\coloneqq(\mathrm{Fil}^{r}\varphi^{*}\mathfrak{M}\otimes_{\mathfrak{S}}A_{\mathrm{inf}}(R))^{\varphi_{r}=1}

where we write φ𝔐=𝔖φ,𝔖𝔐\varphi^{*}\mathfrak{M}=\mathfrak{S}\otimes_{\varphi,\mathfrak{S}}\mathfrak{M} and we set Filrφ𝔐={xφ𝔐|(1φ)(x)E(u)r𝔐}\mathrm{Fil}^{r}\varphi^{*}\mathfrak{M}=\{x\in\varphi^{*}\mathfrak{M}|(1\otimes\varphi)(x)\in E(u)^{r}\mathfrak{M}\} and φr:Filrφ𝔐φ𝔐\varphi_{r}:\mathrm{Fil}^{r}\varphi^{*}\mathfrak{M}\to\varphi^{*}\mathfrak{M} defined by

φr(x)φ(x)φ(E(u)r).\varphi_{r}(x)\coloneqq\frac{\varphi(x)}{\varphi(E(u)^{r})}.

These functors are related via

T𝔖r(𝔐)T𝔖(𝔐)(r)T^{r}_{\mathfrak{S}}(\mathfrak{M})\cong T_{\mathfrak{S}}(\mathfrak{M})(r)

as in [LL23, Lemma 6.11], where (r)(r) denotes the rrth Tate twist. Their proof only establishes the isomorphism in the classical case, but the proof works mutatis mutandis in the relative case.

Remark 33.

We are using the covariant versions of these functors rather than the contravariant versions used in [Du+24] (defined below) which are dual to the covariant versions. With this in mind, if we take a lattice ΛRepcris,Zp[0,r](GR)\Lambda\in\mathrm{Rep}_{\mathrm{cris},\mathbb{Z}_{p}}^{[0,r]}(G_{R}), we should first consider Λ(r)Repcris,Zp[0,r](GR)\Lambda^{\vee}(r)\in\mathrm{Rep}_{\mathrm{cris},\mathbb{Z}_{p}}^{[0,r]}(G_{R}), the dual with the rrth Tate twist applied. Then using Theorem 20 as stated for the contravariant versions of the functors, there exists a Kisin module with descent data (𝔐,φ𝔐,f)(\mathfrak{M},\varphi_{\mathfrak{M}},f) such that T𝔖(𝔐)=Λ(r)T_{\mathfrak{S}}^{\vee}(\mathfrak{M})=\Lambda^{\vee}(r) Then taking duals and applying the rrth Tate twist, we obtain T𝔖r(𝔐)=ΛT_{\mathfrak{S}}^{r}(\mathfrak{M})=\Lambda, so we do have a Kisin module with descent data (𝔐,φ𝔐,f)(\mathfrak{M},\varphi_{\mathfrak{M}},f) with T𝔖r(𝔐)=ΛT_{\mathfrak{S}}^{r}(\mathfrak{M})=\Lambda for our covariant version, too.

Remark 34.

We can define T𝔖T_{\mathfrak{S}} and T𝔖rT_{\mathfrak{S}}^{r} analogously when working over a small base ring TT using the analogous embedding of TAinf(T)T\to A_{\mathrm{inf}}(T) via ti[ti~]t_{i}\mapsto[\widetilde{t_{i}}]. Note that the embedding TT𝔪^T\to\widehat{T_{\mathfrak{m}}} induces an embedding Ainf(T)Ainf(T𝔪^)A_{\mathrm{inf}}(T)\to A_{\mathrm{inf}}(\widehat{T_{\mathfrak{m}}}) which gives an isomorphism 𝒯𝔖T(𝔐)T𝔖T𝔪^(𝔐TT𝔪^)\mathcal{T}_{\mathfrak{S}_{T}}(\mathfrak{M})\to T_{\mathfrak{S}_{\widehat{T_{\mathfrak{m}}}}}(\mathfrak{M}\otimes_{T}\widehat{T_{\mathfrak{m}}}) when 𝔐\mathfrak{M} is a Kisin module over 𝔖T.\mathfrak{S}_{T}.

2.9.3. Equivalence between T𝔖T_{\mathfrak{S}} and T𝒪T_{\mathcal{O}_{\mathcal{E}}}

Again let 𝔐\mathfrak{M} be a Kisin module associated to a crystalline representation. By [Du+24, Lemma 4.27], the natural GR~G_{\tilde{R}_{\infty}}-equivariant map

T𝔖(𝔐)Hom𝔖,φ(𝔐,𝔖^ur)Hom𝒪,φ(,𝒪^ur)=T𝒪()T_{\mathfrak{S}}^{\vee}(\mathfrak{M})\coloneqq\mathrm{Hom}_{\mathfrak{S},\varphi}(\mathfrak{M},\widehat{\mathfrak{S}}^{\mathrm{ur}})\to\mathrm{Hom}_{\mathcal{O}_{\mathcal{E}},\varphi}(\mathcal{M},\widehat{\mathcal{O}_{\mathcal{E}}}^{\mathrm{ur}})=T_{\mathcal{O}_{\mathcal{E}}}^{\vee}(\mathcal{M})

is an isomorphism, though we need to make slight adjustments because we are using the covariant versions of these functors as defined above rather than the contravariant versions sometimes used in [Du+24].

Lemma 35.

The natural map

T𝒪()T𝔖(𝔐)T_{\mathcal{O}_{\mathcal{E}}}(\mathcal{M})\to T_{\mathfrak{S}}(\mathfrak{M})

is an isomorphism of GR~G_{\tilde{R}_{\infty}} representations.

Proof.

Recall =𝔐𝔖𝒪.\mathcal{M}=\mathfrak{M}\otimes_{\mathfrak{S}}\mathcal{O}_{\mathcal{E}}.

The embedding f1:𝔖Ainf(R)=W(R¯)f_{1}:\mathfrak{S}\to A_{\mathrm{inf}}(R)=W(\overline{R}^{\flat}) above extends to an embedding 𝒪^urW(R¯[1/u])\widehat{\mathcal{O}_{\mathcal{E}}}^{\mathrm{ur}}\to W(\overline{R}^{\flat}[1/u]) which gives us a natural embedding

T𝒪()=(𝒪𝒪^ur)φ=1=(𝔐𝔖𝒪^ur)φ=1(𝔐𝔖W(R¯[1/u]))φ=1=T𝔖(𝔐).T_{\mathcal{O}_{\mathcal{E}}}(\mathcal{M})=(\mathcal{M}\otimes_{\mathcal{O}_{\mathcal{E}}}\widehat{\mathcal{O}_{\mathcal{E}}}^{\mathrm{ur}})^{\varphi=1}=(\mathfrak{M}\otimes_{\mathfrak{S}}\widehat{\mathcal{O}_{\mathcal{E}}}^{\mathrm{ur}})^{\varphi=1}\hookrightarrow(\mathfrak{M}\otimes_{\mathfrak{S}}W(\overline{R}^{\flat}[1/u]))^{\varphi=1}=T_{\mathfrak{S}}(\mathfrak{M}).

The last embedding is known to be an isomorphism of GR~G_{\tilde{R}_{\infty}} representations as a consequence of [Du+24, Lemma 2.15], completing the proof. ∎

2.9.4. The definition of TS()T_{S}(\mathscr{M})

Write SS For SR.S_{R}. Let \mathscr{M} be a Breuil module. We can define a Galois action on Acris(R)SA_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M} is defined using \nabla in the same way as the Galois action on a Fontaine-Laffaille module was defined above. First embed SS into Acris(R)A_{\mathrm{cris}}(R) where uu maps to [π][\pi^{\flat}], the Teichmuller lift of a compatible sequence of pp-power roots of π\pi. If we let gGRg\in G_{R} and axAcris(R)Sa\otimes x\in A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M} we set

g(ax)=g(a)uj0t1j1tdjd(x)γj0(g[π][π])i=1dγji(g([ti])[ti])()g(a\otimes x)=\sum g(a)\partial_{u}^{j_{0}}\partial_{t_{1}}^{j_{1}}\cdots\partial_{t_{d}}^{j_{d}}(x)\cdot\gamma_{j_{0}}(g[\pi^{\flat}]-[\pi^{\flat}])\prod_{i=1}^{d}\gamma_{j_{i}}(g([t_{i}^{\flat}])-[t_{i}^{\flat}])\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ (*)

with notation as above in the previous section and the sum over multi-indices (j0,,jd)(j_{0},\cdots,j_{d}) of nonnegative integers. Similarly to Section 2.9.1, this gives us an Acris(R)A_{\mathrm{cris}}(R)-semilinear action of GRG_{R} which preserves the filtration structure (g(Filr(Acris(R)S)Filr(Acris(R)Sg(\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M})\subset\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M})) and commutes with Frobenius. It is also clear that if =MRS\mathscr{M}=M\otimes_{R}S, then

TS()Filr(Acris(R)S)φr=1=Filr(Acris(R)S(MRS))φr=1=Tcris(M).T_{S}(\mathscr{M})\coloneqq\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M})^{\varphi_{r}=1}=\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{S}(M\otimes_{R}S))^{\varphi_{r}=1}=T_{\mathrm{cris}}(M).

We can also view this Galois action on Acris(R)SA_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M} determined by \nabla as one arising from the descent data. We first recall that there are rings S(1)S^{(1)} and S(2)S^{(2)} analogous to 𝔖(1)\mathfrak{S}^{(1)} and 𝔖(2)\mathfrak{S}^{(2)}. The details of the construction can be found in [Du+24, Example 3.9]. We will write p1p_{1} and p2p_{2} for the maps from SS to S(1)S^{(1)} analogous to the p1p_{1} and p2p_{2} defined in Section 2.9.2

Given (𝔐,φ𝔐,f)(\mathfrak{M},\varphi_{\mathfrak{M}},f) a Kisin module associated to a lattice in a crystalline representation as in Theorem 20 , [Du+24, Construction 4.3] constructs a descent datum fSf_{S} on [p1]=S[p1]RM\mathscr{M}[p^{-1}]=S[p^{-1}]\otimes_{R}M as follows:

Let u:S[p1]RMS[p1]RM\partial_{u}:S[p^{-1}]\otimes_{R}M\to S[p^{-1}]\otimes_{R}M be the derivation given by u,S1\partial_{u,S}\otimes 1 and let ti\partial_{t_{i}} be the derivation given by ti,S1+1ti,M\partial_{t_{i},S}\otimes 1+1\otimes\partial_{ti,M}. Then we can define fS:S(2)p1,S[p1]S(2)p2,S[p1]f_{S}:S^{(2)}\otimes_{p_{1},S}\mathscr{M}[p^{-1}]\to S^{(2)}\otimes_{p_{2},S}\mathscr{M}[p^{-1}] as

fS(x)=Juj0t1j1tdjd(x)γj0(p2(u)p1(u))i=1dγji(p2(ti)p1(ti)),()f_{S}(x)=\sum_{J}\partial_{u}^{j_{0}}\partial_{t_{1}}^{j_{1}}\cdots\partial_{t_{d}}^{j_{d}}(x)\gamma_{j_{0}}(p_{2}(u)-p_{1}(u))\prod_{i=1}^{d}\gamma_{j_{i}}(p_{2}(t_{i})-p_{1}(t_{i})),\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ (**)

where the sum is over multi-indices J=(j0,,jd)J=(j_{0},\cdots,j_{d}) of nonnegative integers.

This descent datum determines a Galois action on Acris(R)A_{\mathrm{cris}}(R)\otimes\mathscr{M} in the same way as the descent datum ff determined a Galois action on Ainf(R)𝔖φ𝔐A_{\mathrm{inf}}(R)\otimes_{\mathfrak{S}}\varphi^{*}\mathfrak{M} just with 𝔖(1)\mathfrak{S}^{(1)} and Ainf(R)A_{\mathrm{inf}}(R) replaced by S(2)S^{(2)} and Acris(R)A_{\mathrm{cris}}(R). Specifically, let gGR.g\in G_{R}. The maps f1f_{1} and f2f_{2} from Section 2.9.2 extend to embeddings SAcris(R)S\to A_{\mathrm{cris}}(R) and by the analogous universal property as discussed at the beginning of Section 2.6, there exists a unique map fS,g:(S(1),(E))(Acris(R),(E))f_{S,g}:(S^{(1)},(E))\to(A_{\mathrm{cris}}(R),(E)) and since fSf_{S} satisfues the cocycle condition, there exists a map dd, depending on gg, making the diagram commute:

S(1)p1,S{S^{(1)}\otimes_{p_{1},S}\mathscr{M}}S(1)p2,S{S^{(1)}\otimes_{p_{2},S}\mathscr{M}}Acris(R)S{A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M}}Acris(R)g,S{A_{\mathrm{cris}}(R)\otimes_{g,S}\mathscr{M}}fS\scriptstyle{f_{S}}f1\scriptstyle{f_{1}}fS,g\scriptstyle{f_{S,g}}dg\scriptstyle{d_{g}}

which establishes the desired Galois action on Acris(R)SA_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M}.

So fSf_{S} determines a Galois action of GRG_{R} on Acris(R)SA_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M} and hence on

TS()Filr(Acris(R)S)φr=1.T_{S}(\mathscr{M})\coloneqq\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M})^{\varphi_{r}=1}.

We have seen two ways of defining a Galois action on TS()T_{S}(\mathscr{M}): the first via the descent datum fSf_{S} and the second via the connection \nabla. Via the explicit formulas ()(*) and ()(**) above, we see that the two ways of defining the Galois action are the same.

Remark 36.

We define TST_{S} analogously when working over a small base ring TT instead of RR. Writing SS for STS_{T} and letting \mathscr{M} be a Breuil module over SS, we have TS()Filr(Acris(T)S)φr=1.T_{S}(\mathscr{M})\coloneqq\mathrm{Fil}^{r}(A_{\mathrm{cris}}(T)\otimes_{S}\mathscr{M})^{\varphi_{r}=1}.

2.9.5. Equivalence between T𝔖rT_{\mathfrak{S}}^{r} and TST_{S}

We will now explore the relationship between T𝔖r(𝔐)T_{\mathfrak{S}}^{r}(\mathfrak{M}) and TS()T_{S}(\mathscr{M}) defined in the preceding sections. Continue to assume that (𝔐,φ𝔐,f)(\mathfrak{M},\varphi_{\mathfrak{M}},f) is a Kisin module arising from a Zp\mathbb{Z}_{p}-crystalline representation of GRG_{R}. Let \mathscr{M} be a Breuil module arising from (𝔐,φ𝔐,f)(\mathfrak{M},\varphi_{\mathfrak{M}},f) as in Section 2.8 (recall this means =(𝔐)𝔐𝔖,φ𝔖S\mathscr{M}=\mathscr{M}(\mathfrak{M})\coloneqq\mathfrak{M}\otimes_{\mathfrak{S},\varphi_{\mathfrak{S}}}S).

Now consider the natural map

α:Ainf(R)𝔖φ𝔐Acris(R)S.\alpha:A_{\mathrm{inf}}(R)\otimes_{\mathfrak{S}}\varphi^{*}\mathfrak{M}\to A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M}.

By [Du+24, Prop. 4.6], the map fSf_{S} as defined in the previous section (originally constructed in [Du+24, Construction 4.3]) satisfies

fS=S(1)φ,𝔖(1)f.f_{S}=S^{(1)}\otimes_{\varphi,\mathfrak{S}^{(1)}}f.

Thus we see that fSf_{S} and S(1)φ,𝔖(1)fS^{(1)}\otimes_{\varphi,\mathfrak{S}^{(1)}}f define the same Galois action on Acris(R)SA_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M}, as we discussed how ff defines a Galois action on Ainf(R)φ𝔐A_{\mathrm{inf}}(R)\otimes\varphi^{*}\mathfrak{M} using the following diagram:

𝔖(1)p1,𝔖𝔐{\mathfrak{S}^{(1)}\otimes_{p_{1},\mathfrak{S}}\mathfrak{M}}𝔖(1)p2,𝔖𝔐{\mathfrak{S}^{(1)}\otimes_{p_{2},\mathfrak{S}}\mathfrak{M}}Ainf(R)𝔖𝔐{A_{\mathrm{inf}}(R)\otimes_{\mathfrak{S}}\mathfrak{M}}Ainf(R)g,𝔖𝔐{A_{\mathrm{inf}}(R)\otimes_{g,\mathfrak{S}}\mathfrak{M}}f\scriptstyle{f}f1\scriptstyle{f_{1}}fg\scriptstyle{f_{g}}dg\scriptstyle{d_{g}}

and fSf_{S} defines a Galois action on Acris(R)A_{\mathrm{cris}}(R)\otimes\mathscr{M} via the same diagram mutatis mutandi:

S(1)p1,S{S^{(1)}\otimes_{p_{1},S}\mathscr{M}}S(1)p2,S{S^{(1)}\otimes_{p_{2},S}\mathscr{M}}Acris(R)S{A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M}}Acris(R)g,S{A_{\mathrm{cris}}(R)\otimes_{g,S}\mathscr{M}}fS\scriptstyle{f_{S}}f1\scriptstyle{f_{1}}fS,g\scriptstyle{f_{S,g}}dg\scriptstyle{d_{g}}

Thus we see that the Galois action established on T𝔖r(𝔐)T_{\mathfrak{S}}^{r}(\mathfrak{M}) is compatible with the Galois action on TS()T_{S}(\mathscr{M}) via the natural map α\alpha.

As =(𝔐)\mathscr{M}=\mathscr{M}(\mathfrak{M}), it is standard to see from the definitions of the filtration structure on 𝔐\mathfrak{M} and \mathscr{M} that the map α\alpha respects the filtration and φr\varphi_{r} structures. This gives us a natural injective map

T𝔖r(𝔐)TS()T_{\mathfrak{S}}^{r}(\mathfrak{M})\to T_{S}(\mathscr{M})

compatible with the GRG_{R}-action, filtration, and φi\varphi_{i}, structure.

We will later prove the following theorem:

Theorem 37.

The natural injection

T𝔖r(𝔐)TS()Filr(Acris(R)S)φr=1T_{\mathfrak{S}}^{r}(\mathfrak{M})\to T_{S}(\mathscr{M})\coloneqq\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M})^{\varphi_{r}=1}

induced by the embedding Ainf(R)Acris(R)A_{\mathrm{inf}}(R)\to A_{\mathrm{cris}}(R) is an isomorphism of GRG_{R} representations.

but we will need information about the Fontaine-Laffaille module associated to \mathscr{M} which won’t be established until the next section. Nevertheless, we state the result here in our section on Galois representations for completeness.

2.10. A summary of categories

Here we include a brief diagram summarizing the relationship between the many categories introduced above in the setting of our base ring RR:

  RepZp,[0,r]cris(GR){\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{R})}MF,[0,r]ff{\mathrm{MF}_{\nabla,[0,r]}^{\mathrm{ff}}}Mod𝒪ff{\leavevmode\nobreak\ \mathrm{Mod}_{\mathcal{O}_{\mathcal{E}}}^{\mathrm{ff}}}DD𝔖,[0,r]ff{\mathrm{DD}_{\mathfrak{S},[0,r]}^{\mathrm{ff}}}ModS,ff,r{\mathrm{Mod}_{S,\nabla}^{\mathrm{ff},\leavevmode\nobreak\ r}}Tcris\scriptstyle{T_{\mathrm{cris}}}^RS\scriptstyle{-\leavevmode\nobreak\ \widehat{\otimes}_{R}S}T𝒪\scriptstyle{T_{\mathcal{O}_{\mathcal{E}}}}𝒪𝔖\scriptstyle{\mathcal{O}_{\mathcal{E}}\otimes_{\mathfrak{S}}\leavevmode\nobreak\ -}𝔖,φS\scriptstyle{-\leavevmode\nobreak\ \otimes_{\mathfrak{S},\varphi}S}T𝔖r\scriptstyle{T_{\mathfrak{S}}^{r}}\scriptstyle{\cong}TS\scriptstyle{T_{S}}

2.11. TcrisT_{\mathrm{cris}} is fully faithful

To conclude Section 2, we will show that TcrisT_{\mathrm{cris}} is fully faithful. First we work over the base ring RR.

Let M1M_{1} and M2M_{2} be Fontaine-Laffaille modules over RR so that there exists a morphism

ψ:Tcris(M1)Tcris(M2)\psi:T_{\mathrm{cris}}(M_{1})\to T_{\mathrm{cris}}(M_{2})

of GRG_{R} representations. By Proposition 16 there is an equivalence of categories between Zp\mathbb{Z}_{p} stable lattices in finite free GRG_{R_{\infty}} representations and étale (φ𝒪,𝒪)(\varphi_{\mathcal{O}_{\mathcal{E}}},\mathcal{O}_{\mathcal{E}})-modules and thus exists a unique morphism ψ:12\psi_{\mathcal{M}}:\mathcal{M}_{1}\to\mathcal{M}_{2} where i\mathcal{M}_{i} is the étale (𝒪,φ𝒪)(\mathcal{O}_{\mathcal{E}},\varphi_{\mathcal{O}_{\mathcal{E}}})-module associated to the G𝒯G_{\mathcal{T}} representation Tcris(Mi).T_{\mathrm{cris}}(M_{i}).

Moreover, there is a unique morphism ψ𝔐:𝔐1𝔐2\psi_{\mathfrak{M}}:\mathfrak{M}_{1}\to\mathfrak{M}_{2} between the associated Kisin modules with descent data by [Du+24, Proposition 3.25, Theorem 3.28]. Proposition 3.25 establishes an equivalence of categories between DD𝔖f\mathrm{DD}^{f}_{\mathfrak{S}} and a category of prismatic FF-crystals, and Theorem 3.28 establishes that this category of prismatic FF-crystals is equivalent to RepZp,[0,r]cris(G𝒯)\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{\mathcal{T}}). So it suffices to show that ψ𝔐\psi_{\mathfrak{M}} induces a map ψM:M1M2\psi_{M}:M_{1}\to M_{2} such that Tcris(ψM)T_{\mathrm{cris}}(\psi_{M}) agrees with T𝔖r(ψ𝔐)T_{\mathfrak{S}}^{r}(\psi_{\mathfrak{M}}) on Tcris(M1)T𝔖r(𝔐1)T_{\mathrm{cris}}(M_{1})\cong T_{\mathfrak{S}}^{r}(\mathfrak{M}_{1}), which are isomorphic as established above.

Set ψ=ψ𝔐𝔖,φ𝔖S:12\psi_{\mathscr{M}}=\psi_{\mathfrak{M}}\otimes_{\mathfrak{S},\varphi_{\mathfrak{S}}}S:\mathscr{M}_{1}\to\mathscr{M}_{2} which is a morphism of the associated Breuil modules. We can then obtain a morphism on the associated Fontaine-Laffaille modules using the unique sections guaranteed by Theorem 43. Let s1s_{1} be the section M11M_{1}\to\mathscr{M}_{1} and s2s_{2} the section M22M_{2}\to\mathscr{M}_{2} as in the following diagram:

1{\mathscr{M}_{1}}2{\mathscr{M}_{2}}M1{M_{1}}M2{M_{2}}ψ\scriptstyle{\psi_{\mathscr{M}}}ψM\scriptstyle{\psi_{M}}s1\scriptstyle{s_{1}}s2\scriptstyle{s_{2}}

where the bottom map ψM\psi_{M} is defined to be the natural projection 2M2\mathscr{M}_{2}\to M_{2} composed with ψs1\psi_{\mathscr{M}}\circ s_{1}. Tensoring the above diagram with Acris,A_{\mathrm{cris}}, we see that Tcris(ψM)T_{\mathrm{cris}}(\psi_{M}) agrees with TS(ψ).T_{S}(\psi_{\mathscr{M}}). If we let gi:𝔐iig_{i}:\mathfrak{M}_{i}\to\mathscr{M}_{i} denote the map 𝔐i𝔐i𝔖,φ𝔖S\mathfrak{M}_{i}\to\mathfrak{M}_{i}\otimes_{\mathfrak{S},\varphi_{\mathfrak{S}}}S, we obtain a similar diagram

𝔐1{\mathfrak{M}_{1}}𝔐2{\mathfrak{M}_{2}}1{\mathscr{M}_{1}}2{\mathscr{M}_{2}}ψ𝔐\scriptstyle{\psi_{\mathfrak{M}}}g1\scriptstyle{g_{1}}g2\scriptstyle{g_{2}}ψ\scriptstyle{\psi_{\mathscr{M}}}

which shows that T𝔖r(ψ𝔐)T_{\mathfrak{S}}^{r}(\psi_{\mathfrak{M}}) will agree with TS(ψ).T_{S}(\psi_{\mathscr{M}}).

To complete the proof of fully-faithfulness, we need to justify the uniqueness of ψM\psi_{M}. This follows from the established rational theory. Such a morphism ψM:M1M2\psi_{M}:M_{1}\to M_{2} induces a morphism ψM[p1]:M1[p1]M2[p1]\psi_{M[p^{-1}]}:M_{1}[p^{-1}]\to M_{2}[p^{-1}] which is known to be unique as [Du+24, Proposition 4.28] establishes an isomorphism Mi[p1]Dcris(Tcris(Mi)ZpQp)M_{i}[p^{-1}]\to D_{\mathrm{cris}}^{\vee}(T_{\mathrm{cris}}(M_{i})\otimes_{\mathrm{Z}_{p}}\mathbb{Q}_{p}) and the induced morphism on DcrisD_{\mathrm{cris}}^{\vee} is known to be unique by the original theory of Brinon in [Bri08].

Thus we have established:

Theorem 38.

The functor

Tcris:MF,[0,r]ff(R)RepZp,[0,r]cris(GR)T_{\mathrm{cris}}:\mathrm{MF}^{\mathrm{ff}}_{\nabla,[0,r]}(R)\to\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{R})

given by

Tcris(M)(Filr(Acris(R)RM))φr=1T_{\mathrm{cris}}(M)\coloneqq(\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{R}M))^{\varphi_{r}=1}

is fully faithful.

We can also establish fully-faithfulness for small base rings via base change.

We also have the analogous result for the Tate Algebra:

Theorem 39.

The functor

Tcris:MF,[0,r]ff(T)RepZp,[0,r]cris(GT)T_{\mathrm{cris}}:\mathrm{MF}^{\mathrm{ff}}_{\nabla,[0,r]}(T)\to\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{T})

given by

Tcris(M)(Filr(Acris(T)TM))φr=1T_{\mathrm{cris}}(M)\coloneqq(\mathrm{Fil}^{r}(A_{\mathrm{cris}}(T)\otimes_{T}M))^{\varphi_{r}=1}

is fully faithful.

Proof.

Let M1M_{1} and M2M_{2} be Fontaine-Laffaille modules over TT so that there exists a morphism

ψ:Tcris(M1)Tcris(M2)\psi:T_{\mathrm{cris}}(M_{1})\to T_{\mathrm{cris}}(M_{2})

of GTG_{T} representations. Write M^\widehat{M} for MTT𝔪^.M\otimes_{T}\widehat{T_{\mathfrak{m}}}. We get an induced map Tcris(M^1)Tcris(M^2)T_{\mathrm{cris}}(\widehat{M}_{1})\to T_{\mathrm{cris}}(\widehat{M}_{2}), for which there exists a unique morhphism M^1M^2\widehat{M}_{1}\to\widehat{M}_{2} as TcrisT_{\mathrm{cris}} is fully faithful with the base ring RR. We also have a unique induced map D1=M1[p1]M2[p1]=D2D_{1}=M_{1}[p^{-1}]\to M_{2}[p^{-1}]=D_{2}, setting up the following diagram: M^1{\widehat{M}_{1}}M^2{\widehat{M}_{2}}M1{M_{1}}M2{M_{2}}φ\scriptstyle{\varphi} where the vertical maps are injective. But the image of MM under φ\varphi is contained in M^2M2[p1]=M2\widehat{M}_{2}\cap M_{2}[p^{-1}]=M_{2}, completing the proof of fully faithfulness. ∎

3. Essential Surjectivity of TcrisT_{\mathrm{cris}} over Power Series Rings

In this section we aim to prove the following two theorems, working over the base ring RR. Let I0SI_{0}\subset S denote the kernel of the projection SRS\to R given by u0.u\mapsto 0.

Theorem 40.

Let \mathscr{M} be a Breuil module which arises from a Kisin module which arises from a Λ,\Lambda, an object in RepZpcris[0,r](GR).\mathrm{Rep}_{\mathbb{Z}_{p}}^{\mathrm{cris}[0,r]}(G_{R}). Set

M/I0𝔐/u𝔐R,φ𝔖R.M\coloneqq\mathscr{M}/I_{0}\mathscr{M}\cong\mathfrak{M}/u\mathfrak{M}\otimes_{R,\varphi_{\mathfrak{S}}}R.

Then MM carries the data of a Fontaine-Laffaille module and satisfies MRSM\otimes_{R}S\cong\mathscr{M} as in Definition 31 We call it the “Fontaine-Laffaille module associated to Λ\Lambda.”

Theorem 41.

Let MM be the Fontaine-Laffaille module associated to Λ.\Lambda. Then

Tcris(M)Λ.T_{\mathrm{cris}}(M)\cong\Lambda.

Combined, these theorems complete the proof of essential surjectivity of Tcris.T_{cris}. For the remainder of Section 3, we will assume \mathscr{M} satisfies the hypotheses of the above theorem.

3.1. Fontaile-Laffaille data on MM

We will now define the Fontaine-Laffaille data on MM and show that MM is indeed a Fontaine-Laffaille module with our key tool being Theorem 11. As modules, we have from Theorem 23 and Proposition 23.1 that MgMRRgM_{g}\coloneqq M\otimes_{R}R_{g} and M0MRW(k)M_{0}\coloneqq M\otimes_{R}W(k) are isomorphic, as modules, to Fontaile-Laffaille modules arising from Zp\mathbb{Z}_{p} stable lattices in crystalline representations, so they come equipped with the data of a classical Fontaine-Laffaille module as in Definition 1. We do not yet know, though, whether this data, particularly the filtration data, is compatible with the base change maps. Showing this will be a major focus of this section.

3.1.1. The Filtration and φi\varphi_{i} structure on MM

We first develop the filtration and φi\varphi_{i} structure on M/I0𝔐/u𝔐R,φRM\coloneqq\mathscr{M}/I_{0}\mathscr{M}\cong\mathfrak{M}/u\mathfrak{M}\otimes_{R,\varphi}R. The key is to show that the following Lemma from [Du+24] descends to the integral situation:

Lemma 42 ([Du+24] Lemma 4.2).

Consider the projection q:Mq:\mathscr{M}\to M induced by the φ\varphi-compatible projection SRS\to R, u0.u\mapsto 0. Then qq admits a unique φ\varphi-compatible section s:M[p1][p1].s:M[p^{-1}]\to\mathscr{M}[p^{-1}]. Furthermore, 1s:S[p1]R[p1]M[p1][p1]1\otimes s:S[p^{-1}]\otimes_{R[p^{-1}]}M[p^{-1}]\to\mathscr{M}[p^{-1}] is an isomorphism.

We first prove the following:

Lemma 43.

The projection qq as in Lemma 42 admits a unique φ\varphi-compatible section

s:M.s:M\to\mathscr{M}.
Proof.

To attain this section we base change to the Shilov point obtaining the following commutative diagram:

M{M}M[1p]{M[\frac{1}{p}]\ }[1p]{\mathscr{M}[\frac{1}{p}]}Mg{M_{g}}Mg[1p]{M_{g}[\frac{1}{p}]}g[1p]{\mathscr{M}_{g}[\frac{1}{p}]}s\scriptstyle{s}sRg\scriptstyle{s\otimes R_{g}}

Note that sRgs\otimes R_{g} determines a section Mg[1p]g[1p]M_{g}[\frac{1}{p}]\to\mathscr{M}_{g}[\frac{1}{p}]. By [Gao19, Prop 3.2.3] this section is unique and it takes MgM_{g}, and thus MM, into g\mathscr{M}_{g}. Tracing MM through the top of the diagram, we see that the image of MM in g[1p]\mathscr{M}_{g}[\frac{1}{p}] is contained in [1p].\mathscr{M}[\frac{1}{p}].

From Lemma 22, we obtain:

g[1p]\displaystyle\mathscr{M}_{g}\cap\mathscr{M}\left[\frac{1}{p}\right] =(SgS[1p])\displaystyle=\mathscr{M}\otimes\left(S_{g}\cap S\left[\frac{1}{p}\right]\right)
=S.\displaystyle=\mathscr{M}\otimes S.

We can verify SgS[1p]=SS_{g}\cap S\left[\frac{1}{p}\right]=S term by term. If we take xx in the intersection, we can write it as

x=E(u)ii!aijujx=\sum\frac{E(u)^{i}}{i!}\sum a_{ij}u^{j}

with aijW(kg).a_{ij}\in W(k_{g}). We can also write it as

x=E(u)ii!bijujx=\sum\frac{E(u)^{i}}{i!}\sum b_{ij}u^{j}

with bijW(k)[1p].b_{ij}\in W(k)[\frac{1}{p}]. Comparing term-by-term we can see that aij=bij.a_{ij}=b_{ij}.

Lemma 44.

The map 1s:SRM1\otimes s:S\otimes_{R}M\to\mathscr{M} is an isomorphism.

Proof.

Since SRMS\otimes_{R}M and \mathscr{M} are finite, free SS-modules of the same rank, it suffices to show that 1s1\otimes s is surjective. Moreover, SS is a local ring, so by Nakayama’s Lemma it suffices to show that 1s1\otimes s is surjective after reducing modulo the maximal ideal of SS which is the ideal generated by p,t1,,tdp,t_{1},\cdots,t_{d}, uu, and the divided powers E(u)ii!\frac{E(u)^{i}}{i!}. After reducing modulo the maximal ideal, 1s1\otimes s becomes the identity map which is obviously surjective. ∎

We now define the filtration and φi\varphi_{i} structure on MM.

Definition 45.

Realizing MM as a submodule of \mathscr{M} via the section ss, we can define

Fili(M)FiliM.\mathrm{Fil}^{i}(M)\coloneqq\mathrm{Fil}^{i}\mathscr{M}\cap M.
Definition 46.

We set φM,i=φ,i|M\varphi_{M,i}=\varphi_{\mathscr{M},i}\big{|}_{M}.

Note that the filtration on \mathscr{M} induces a filtration Fili(M)\mathrm{Fil}^{i}(M) on MM for which the composite Fili(M)Fili1(M)φM,i1M\mathrm{Fil}^{i}(M)\to\mathrm{Fil}^{i-1}(M)\xrightarrow{\varphi_{M,i-1}}M is pφM,ip\varphi_{M,i}. By the corresponding properties on Fili\mathrm{Fil}^{i}\mathscr{M} (see Definition 25), it is easy to see that Fil0M=M\mathrm{Fil}^{0}M=M and Filr+1M={0}.\mathrm{Fil}^{r+1}M=\{0\}.

Thus MM has the structure of an object in MFbigf(R)\mathrm{MF}_{\mathrm{big}}^{\mathrm{f}}(R). We rely on Proposition 11 to show that we have an object of MFff(R)\mathrm{MF}^{\mathrm{ff}}(R) in the following section.

3.2. MRRgMgM\otimes_{R}R_{g}\cong M_{g} as Fontaine-Laffaille modules

We need to show that the filtration data on the tensor product MRRgM\otimes_{R}R_{g} is compatible with the filtration on MgM_{g}. The difficulty in doing this is establishing a base change theorem for the Breuil module \mathscr{M}. The ring SgS_{g} is very large, so treating \mathscr{M} as an RR-module and tensoring with RgR_{g} over RR will not get us back g\mathscr{M}_{g}: we would need to pp-adically complete RRg.\mathscr{M}\otimes_{R}R_{g}. The fact SS is non-Noetherian makes pp-adic completion difficult to work with.

Instead, we can work over the ring S/Fili(S)S/\mathrm{Fil}^{i}(S) which is Noetherian when i<pi<p. This will allow us to decompose Fili(g)\mathrm{Fil}^{i}(\mathscr{M}_{g}) and show that Fili(Mg)\mathrm{Fil}^{i}(M_{g}) is contained in the nicer piece.

Lemma 47.

Let ii be a nonnegative integer less than pp. Then we can decompose Fili(g)\mathrm{Fil}^{i}(\mathscr{M}_{g}) as follows:

Fili(g)=FiliRRg+SFili(Sg).\mathrm{Fil}^{i}(\mathscr{M}_{g})=\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g}+\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}(S_{g}).
Proof.

Another way to express the definition of Fili\mathrm{Fil}^{i}\mathscr{M} is as a kernel in the exact sequence

0Fili(S/FiliS)𝔖𝔐.0\to\mathrm{Fil}^{i}\mathscr{M}\to\mathscr{M}\to\left(S/\mathrm{Fil}^{i}S\right)\otimes_{\mathfrak{S}}\mathfrak{M}.

It follows quickly from the definition of Fili\mathrm{Fil}^{i}\mathscr{M} that SFili(S)\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}(S) is contained in the kernel, so we get an exact sequence

0Fili/(SFili(S))/(SFili(S))(S/FiliS)𝔖𝔐.0\to\mathrm{Fil}^{i}\mathscr{M}/\left(\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}(S)\right)\to\mathscr{M}/\left(\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}(S)\right)\to\left(S/\mathrm{Fil}^{i}S\right)\otimes_{\mathfrak{S}}\mathfrak{M}.

We can now tensor with RgR_{g} which maintains the exactness since RgR_{g} is a flat RR-module. As a further conesequence of the flatness of RgR_{g}, the tensor product commutes with quotients and we obtain the exact sequence:

0FiliRRgS(Fili(S)RRg)S(Fili(S))RRg((S/FiliS)𝔖𝔐)RRg0\to\frac{\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g}}{\mathscr{M}\otimes_{S}(\mathrm{Fil}^{i}(S)\otimes_{R}R_{g})}\to\frac{\mathscr{M}}{\mathscr{M}\otimes_{S}(\mathrm{Fil}^{i}(S))}\otimes_{R}R_{g}\to\left(\left(S/\mathrm{Fil}^{i}S\right)\otimes_{\mathfrak{S}}\mathfrak{M}\right)\otimes_{R}R_{g}

We also have on the Shilov point the exact sequence

0Filigg(Sg/FiliSg)𝔖g𝔐g0\to\mathrm{Fil}^{i}\mathscr{M}_{g}\to\mathscr{M}_{g}\to(S_{g}/\mathrm{Fil}^{i}S_{g})\otimes_{\mathfrak{S}_{g}}\mathfrak{M}_{g}

where the rightmost map restricts to the rightmost map of exact sequence (1)(1) via the embedding g\mathscr{M}\hookrightarrow\mathscr{M}_{g}. We can also see that SFiliSg\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}S_{g} is in the kernel, so this gives an exact sequence

0Filig(SFiliSg)g(SFiliSg)(Sg/FiliSg)𝔖g𝔐g.0\to\frac{\mathrm{Fil}^{i}\mathscr{M}_{g}}{\left(\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}S_{g}\right)}\to\frac{\mathscr{M}_{g}}{\left(\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}S_{g}\right)}\to(S_{g}/\mathrm{Fil}^{i}S_{g})\otimes_{\mathfrak{S}_{g}}\mathfrak{M}_{g}.

Now note that

gSFiliSgSSgSFiliSgSSgFiliSg\frac{\mathscr{M}_{g}}{\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}S_{g}}\cong\frac{\mathscr{M}\otimes_{S}S_{g}}{\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}S_{g}}\cong\mathscr{M}\otimes_{S}\frac{S_{g}}{\mathrm{Fil}^{i}S_{g}}

since \mathscr{M} is a finite, free SS-module. Also,

SFiliSRRgSFili(S)RRg.\frac{\mathscr{M}}{\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}S}\otimes_{R}R_{g}\cong\mathscr{M}\otimes\frac{S}{\mathrm{Fil}^{i}(S)}\otimes_{R}R_{g}.

Since S/FiliSS/\mathrm{Fil}^{i}S is a finite, free RR-module and Sg/FiliSgS_{g}/\mathrm{Fil}^{i}S_{g} is a finite, free-RgR_{g} module of the same rank, the rightmost maps of (2)(2) and (3)(3) are the same with the same domain. Thus the kernels must be the same.

Since FiliSRRgFili(Sg)\mathrm{Fil}^{i}S\otimes_{R}R_{g}\subset\mathrm{Fil}^{i}(S_{g}), we obtain that an element in Filig\mathrm{Fil}^{i}\mathscr{M}_{g} is determined by an element of FiliRRg\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g} up to an element in SFiliSg\mathscr{M}\otimes_{S}\mathrm{Fil}^{i}S_{g}, completing the proof of the lemma. ∎

We now want to use this decomposition to show that elements of Fili(Mg)\mathrm{Fil}^{i}(M_{g}) cannot arise containing any Fili(Sg)S\mathrm{Fil}^{i}(S_{g})\otimes_{S}\mathscr{M} components. Formally, we have the following lemma:

Lemma 48.
Fili(Mg)FiliRRg\mathrm{Fil}^{i}(M_{g})\subset\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g}
Proof.

Let xFili(Mg)x\in\mathrm{Fil}^{i}(M_{g}), which we know to be a subset of Fili(g).\mathrm{Fil}^{i}(\mathscr{M}_{g}). This means by Lemma 47 we can write x=y+zx=y+z with yFili(Sg)Sy\in\mathrm{Fil}^{i}(S_{g})\otimes_{S}\mathscr{M} and zFiliRRgz\in\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g}. Since Fili(Mg)MgMRRg\mathrm{Fil}^{i}(M_{g})\subset M_{g}\cong M\otimes_{R}R_{g}, we have that y=zxRRg.y=z-x\in\mathscr{M}\otimes_{R}R_{g}. We claim y(Fili(S)S)RRgFiliRRg.y\in(\mathrm{Fil}^{i}(S)\otimes_{S}\mathscr{M})\otimes_{R}R_{g}\subset\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g}.

Let e1,e2,,ene_{1},e_{2},\cdots,e_{n} be an SS-basis of .\mathscr{M}. Since yFili(Sg)Sy\in\mathrm{Fil}^{i}(S_{g})\otimes_{S}\mathscr{M}, we can write

y=m=1nsmemy=\sum_{m=1}^{n}s_{m}e_{m}

with smFili(Sg).s_{m}\in\mathrm{Fil}^{i}(S_{g}). Since yRRgy\in\mathscr{M}\otimes_{R}R_{g}, we can write

y\displaystyle y =j=1[m=1n(k=1ak,m,jE(u)kk!)em]rj\displaystyle=\sum_{j=1}^{\ell}\left[\sum_{m=1}^{n}\left(\sum_{k=1}^{\infty}a_{k,m,j}\frac{E(u)^{k}}{k!}\right)e_{m}\right]\otimes r_{j}
=m=1n[j=1k=1((ak,m,jrj)E(u)kk!)]em\displaystyle=\sum_{m=1}^{n}\left[\sum_{j=1}^{\ell}\sum_{k=1}^{\infty}\left((a_{k,m,j}\otimes r_{j})\frac{E(u)^{k}}{k!}\right)\right]e_{m}
=m=1n[k=1(j=1(ak,m,jrj))E(u)kk!]em\displaystyle=\sum_{m=1}^{n}\left[\sum_{k=1}^{\infty}\left(\sum_{j=1}^{\ell}(a_{k,m,j}\otimes r_{j})\right)\frac{E(u)^{k}}{k!}\right]e_{m}

with ak,m,jR[u]a_{k,m,j}\in R[u] and rjRg.r_{j}\in R_{g}. Thus,

sm=k=1(j=1(ak,m,jrj))E(u)kk!.s_{m}=\sum_{k=1}^{\infty}\left(\sum_{j=1}^{\ell}(a_{k,m,j}\otimes r_{j})\right)\frac{E(u)^{k}}{k!}.

Then since smFili(Sg)s_{m}\in\mathrm{Fil}^{i}(S_{g}), we must have

(j=1(ak,m,jrj))=0\left(\sum_{j=1}^{\ell}(a_{k,m,j}\otimes r_{j})\right)=0

for k<i,k<i, and then it is clear that y(Fili(S)S)RRgFiliRRgy\in(\mathrm{Fil}^{i}(S)\otimes_{S}\mathscr{M})\otimes_{R}R_{g}\subset\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g} and the lemma is proven. ∎

Theorem 49.

Fili(Mg)Fili(M)RRg.\mathrm{Fil}^{i}(M_{g})\cong\mathrm{Fil}^{i}(M)\otimes_{R}R_{g}.

Proof.

We know that Fili(Mg)MRRgMg\mathrm{Fil}^{i}(M_{g})\subset M\otimes_{R}R_{g}\cong M_{g} and, by Lemma 48, Fili(Mg)FiliRRg.\mathrm{Fil}^{i}(M_{g})\subset\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g}. By Lemma 22, we have:

Fili(Mg)\displaystyle\mathrm{Fil}^{i}(M_{g}) (FiliRRg)(MRRg)\displaystyle\subset\left(\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}R_{g}\right)\cap\left(M\otimes_{R}R_{g}\right)
=(FiliM)RRg\displaystyle=\left(\mathrm{Fil}^{i}\mathscr{M}\cap M\right)\otimes_{R}R_{g}
=Fili(M)RRg.\displaystyle=\mathrm{Fil}^{i}(M)\otimes_{R}R_{g}.

The inclusion in the other direction is easy to establish. Clearly Fili(M)RRg\mathrm{Fil}^{i}(M)\otimes_{R}R_{g} is inside MgM_{g} and Fili()RRg\mathrm{Fil}^{i}(\mathscr{M})\otimes_{R}R_{g} since Fili(M)\mathrm{Fil}^{i}(M) is a subset of Fili()\mathrm{Fil}^{i}(\mathscr{M}) by definition, Fili(M)M\mathrm{Fil}^{i}(M)\subset M and RgR_{g} is a flat RR-module. Thus we have:

Fili(M)RRg(Fili()RRg)Mg.\displaystyle\mathrm{Fil}^{i}(M)\otimes_{R}R_{g}\subset(\mathrm{Fil}^{i}(\mathscr{M})\otimes_{R}R_{g})\cap M_{g}.

But Fili()RRg\mathrm{Fil}^{i}(\mathscr{M})\otimes_{R}R_{g} as a subset of g\mathscr{M}_{g} is in the kernel of 1φ:gSg/FiliSg𝔖g𝔐g1\otimes\varphi:\mathscr{M}_{g}\to S_{g}/\mathrm{Fil}^{i}S_{g}\otimes_{\mathfrak{S}_{g}}\mathfrak{M}_{g}, so we have

Fili(M)RRg(Fili(g))Mg=Fili(Mg).\mathrm{Fil}^{i}(M)\otimes_{R}R_{g}\subset(\mathrm{Fil}^{i}(\mathscr{M}_{g}))\cap M_{g}=\mathrm{Fil}^{i}(M_{g}).

Now that we have shown the compatibility of the filtrations, MRRgM\otimes_{R}R_{g} has the full Fontaine-Laffaille module structure of MgM_{g}. We thus conclude:

Theorem 50.

MRRgM\otimes_{R}R_{g} is a Fontaine-Laffaille module over RgR_{g}, and MRRgMgM\otimes_{R}R_{g}\cong M_{g} as Fontaine-Laffaille modules.

Proof.

We established in Theorem 23 an isomorphism MRRgMgM\otimes_{R}R_{g}\cong M_{g} compatible with Frobenius. We established in Theorem 49 that Fili(M)RRgFili(Mg).\mathrm{Fil}^{i}(M)\otimes_{R}R_{g}\cong\mathrm{Fil}^{i}(M_{g}). Since MgM_{g} carries the structure of a Fontaine-Laffaile module, we have completed the proof of this theorem. ∎

3.3. MRW(k)M0M\otimes_{R}W(k)\cong M_{0} as Fontaine-Laffaille Modules

The strategy in this section is similar to the strategy in the above section where we analyze the Breuil module filtration to understand the Fontaine-Laffaille module filtration, but we have to deal with the additional concern that W(k)W(k) is not a flat RR-module, which means analyzing various terms of the form Tor1R(,W(k))\mathrm{Tor}_{1}^{R}(-,W(k)). These Tor terms aren’t too concerning because the barrier to W(k)W(k) being a flat RR-module is, effectively, tit_{i}-torsion, which does not show up in most of our modules.

Lemma 51.

We have (for i<pi<p):
(a) Tor1R(S/R,W(k))=0\mathrm{Tor}_{1}^{R}(S/R,W(k))=0
(b) Tor1R(/M,W(k))=0\mathrm{Tor}_{1}^{R}(\mathscr{M}/M,W(k))=0
(c) Tor1R(S/FiliS𝔖𝔐,W(k))=0\mathrm{Tor}_{1}^{R}(S/\mathrm{Fil}^{i}S\otimes_{\mathfrak{S}}\mathfrak{M},W(k))=0
(d) Tor1R(/Fili(),W(k))=0\mathrm{Tor}_{1}^{R}(\mathscr{M}/\mathrm{Fil}^{i}(\mathscr{M}),W(k))=0

Proof.

For (a), we compute Tor1R(S/R,W(k))=0\mathrm{Tor}^{R}_{1}(S/R,W(k))=0 by recognizing that

0RSS/R00\rightarrow R\rightarrow S\rightarrow S/R\rightarrow 0

is a free resolution of S/RS/R since SS is a free RR-module. When tensoring with W(k)W(k), it is clear that the first map remains injective, so indeed Tor1R(S/R,W(k))=0\mathrm{Tor}^{R}_{1}(S/R,W(k))=0.

For (b), by Lemma 44, we have that /M\mathscr{M}/M is a direct sum of finitely many copies of S/RS/R, so Tor1R(/M,W(k))=0\mathrm{Tor}^{R}_{1}(\mathscr{M}/M,W(k))=0 by part (a).

For part (c), we observe that S/FiliSS/\mathrm{Fil}^{i}S is (for i<pi<p) a free RR-module and hence a free W(k)W(k)-module. We also know 𝔐\mathfrak{M} is a free 𝔖\mathfrak{S}-module, giving us our conclusion.

For Part (d), we have an exact sequence

0Fili()S/Fili(S)𝔖𝔐.0\rightarrow\mathrm{Fil}^{i}(\mathscr{M})\hookrightarrow\mathscr{M}\rightarrow S/\mathrm{Fil}^{i}(S)\otimes_{\mathfrak{S}}\mathfrak{M}.

Below in Lemma 52, we justify that the cokernel of this map
S/Fili(S)𝔖𝔐\mathscr{M}\to S/\mathrm{Fil}^{i}(S)\otimes_{\mathfrak{S}}\mathfrak{M} is a finite free RR-module, hence a free W(k)W(k)-module.

Tensoring the induced sequence

0/Fili()S/Fili(S)𝔖𝔐0\to\mathscr{M}/\mathrm{Fil}^{i}(\mathscr{M})\to S/\mathrm{Fil}^{i}(S)\otimes_{\mathfrak{S}}\mathfrak{M}

with W(k)W(k) then applying Lemma 52 and part (c) gives us the result.

The following lemma completes the proof of (c):

Lemma 52.

Let f:S/FiliS𝔖𝔐f:\mathscr{M}\to S/\mathrm{Fil}^{i}S\otimes_{\mathfrak{S}}\mathfrak{M} be the map induced by 1φ𝔐:=Sφ,𝔖𝔐S𝔖𝔐.1\otimes\varphi_{\mathfrak{M}}:\mathscr{M}=S\otimes_{\varphi,\mathfrak{S}}\mathfrak{M}\to S\otimes_{\mathfrak{S}}\mathfrak{M}. Then coker(f)\mathrm{coker}(f) is a finite, free RR-module.

Proof.

Write φ𝔐\varphi^{*}\mathfrak{M} for 𝔖φ,𝔖𝔐\mathfrak{S}\otimes_{\varphi,\mathfrak{S}}\mathfrak{M}. We have an exact sequence

0φ𝔐𝔐𝔐/φ𝔐00\to\varphi^{*}\mathfrak{M}\to\mathfrak{M}\to\mathfrak{M}/\varphi^{*}\mathfrak{M}\to 0

Tensoring with SS, we obtain

Sφ,𝔖𝔐S𝔖𝔐S𝔖𝔐/φ𝔐0S\otimes_{\varphi,\mathfrak{S}}\mathfrak{M}\to S\otimes_{\mathfrak{S}}\mathfrak{M}\to S\otimes_{\mathfrak{S}}\mathfrak{M}/\varphi^{*}\mathfrak{M}\to 0

where the first map is 1φ𝔐.1\otimes\varphi_{\mathfrak{M}}. Then we obtain the following diagram projecting to the S/Fili(S)S/\mathrm{Fil}^{i}(S) level:

Sφ,𝔖𝔐{S\otimes_{\varphi,\mathfrak{S}}\mathfrak{M}}S𝔖𝔐{S\otimes_{\mathfrak{S}}\mathfrak{M}}S𝔖𝔐/φ𝔐{S\otimes_{\mathfrak{S}}\mathfrak{M}/\varphi^{*}\mathfrak{M}}0{0}S/FiliS𝔖𝔐{S/\mathrm{Fil}^{i}S\otimes_{\mathfrak{S}}\mathfrak{M}}S/FiliS𝔖𝔐/φ𝔐{S/\mathrm{Fil}^{i}S\otimes_{\mathfrak{S}}\mathfrak{M}/\varphi^{*}\mathfrak{M}}0{0}0{0}1φ𝔐\scriptstyle{1\otimes\varphi_{\mathfrak{M}}}f\scriptstyle{f}()\scriptstyle{(*)}

from which we conclude that ()(*) is surjective and that coker(f)=S/FiliS𝔖𝔐/φ𝔐.\mathrm{coker}(f)=S/\mathrm{Fil}^{i}S\otimes_{\mathfrak{S}}\mathfrak{M}/\varphi^{*}\mathfrak{M}. By Lemma 23.1 and the proof of Lemma 51(c) which justifies that Tor1R(S/FiliS,W(k))=0\mathrm{Tor}_{1}^{R}(S/\mathrm{Fil}^{i}S,W(k))=0 and allows us to commute the base change with the quotient, coker(f)RW(k)S0/FiliS0𝔖0𝔐0/φ𝔐0.\mathrm{coker}(f)\otimes_{R}W(k)\cong S_{0}/\mathrm{Fil}^{i}S_{0}\otimes_{\mathfrak{S}_{0}}\mathfrak{M}_{0}/\varphi^{*}\mathfrak{M}_{0}.

By tensoring 0φ𝔐0𝔐0𝔐0/φ𝔐00\to\varphi^{*}\mathfrak{M}_{0}\to\mathfrak{M}_{0}\to\mathfrak{M}_{0}/\varphi^{*}\mathfrak{M}\to 0 with S0S_{0}, we find that S0/FiliS0𝔖0𝔐0/φ𝔐0S_{0}/\mathrm{Fil}^{i}S_{0}\otimes_{\mathfrak{S}_{0}}\mathfrak{M}_{0}/\varphi^{*}\mathfrak{M}_{0} is the cokernel of f0:0S0/FiliS0𝔖𝔐f_{0}:\mathscr{M}_{0}\to S_{0}/\mathrm{Fil}^{i}S_{0}\otimes_{\mathfrak{S}}\mathfrak{M} induced by 1φ𝔐01\otimes\varphi_{\mathfrak{M}_{0}}. If i<pi<p, by [GLS14, Theorem 4.20] we know that S0/FiliS0=W(k)[u]/E(u)iS_{0}/\mathrm{Fil}^{i}S_{0}=W(k)[u]/E(u)^{i} and 𝔐0/φ𝔐0=iW(k)[u]/E(u)hi\mathfrak{M}_{0}/\varphi^{*}\mathfrak{M}_{0}=\displaystyle\oplus_{i}W(k)[u]/E(u)^{h_{i}} where hih_{i} are the Hodge-Tate weights from which it is clear that S0/FiliS0𝔖0𝔐/φ𝔐=coker(f0)S_{0}/\mathrm{Fil}^{i}S_{0}\otimes_{\mathfrak{S}_{0}}\mathfrak{M}/\varphi^{*}\mathfrak{M}=\mathrm{coker}(f_{0}) is a finite free W(k)W(k)-module of rank, say, mm.

Let e1,,eme_{1},\cdots,e_{m} be a basis of coker(f0)\mathrm{coker}(f_{0}) as a W(k)W(k)-module. By Nakayama’s Lemma, we can lift it to e1~,,em~\widetilde{e_{1}},\cdots,\widetilde{e_{m}} which generate coker(f)\mathrm{coker}(f) as an RR-module. Analogously, we can see that coker(f)RRgSg/Fili(Sg)𝔖g𝔐g/φ𝔐g\mathrm{coker}(f)\otimes_{R}R_{g}\cong S_{g}/\mathrm{Fil}^{i}(S_{g})\otimes_{\mathfrak{S}_{g}}\mathfrak{M}_{g}/\varphi^{*}\mathfrak{M}_{g} which, by the same reasoning replacing W(k)W(k) with W(kg)W(k_{g}), is a free W(kg)W(k_{g})-module of rank mm. The rank is the same since \mathscr{M} arises from a crystalline representation. A crystalline representation in the relative case is Hodge-Tate and will have the same Hodge-Tate weights on all of its fibers, so the decomposition of 𝔐0/φ𝔐0\mathfrak{M}_{0}/\varphi^{*}\mathfrak{M}_{0} will use the same Hodge-Tate weights hih_{i} as the analogous decomposition of 𝔐g/φ𝔐g.\mathfrak{M}_{g}/\varphi^{*}\mathfrak{M}_{g}. Then the images of e1~,,em~\widetilde{e_{1}},\cdots,\widetilde{e_{m}} in coker(f)RRg\mathrm{coker}(f)\otimes_{R}R_{g} must generate Sg/Fili(Sg)𝔖g𝔐g/φ𝔐gS_{g}/\mathrm{Fil}^{i}(S_{g})\otimes_{\mathfrak{S}_{g}}\mathfrak{M}_{g}/\varphi^{*}\mathfrak{M}_{g}. Therefore they are linearly independent over W(kg)W(k_{g}) and, hence, over RR, as well. Thus coker(f)\mathrm{coker}(f) is a finite, free RR-module. ∎

On the Shilov point we had to be concerned that SgS_{g} is very large: it is not true that RRg=g\mathscr{M}\otimes_{R}R_{g}=\mathscr{M}_{g}, we would need an additional pp-adic completion which does not behave nicely over non-Noetherian rings. On the closed fiber we have a little more freedom since S0S_{0} is smaller than SS and this concern is not present on the closed fiber. We can thus obtain a stronger version of Lemma 47 more easily:

Lemma 53.

We have

Fili(0)=FiliRW(k).\mathrm{Fil}^{i}(\mathscr{M}_{0})=\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}W(k).
Proof.

We have the exact sequence

0Fili𝑓(S/FiliS)𝔖𝔐coker(f)00\rightarrow\mathrm{Fil}^{i}\mathscr{M}\rightarrow\mathscr{M}\xrightarrow{f}(S/\mathrm{Fil}^{i}S)\otimes_{\mathfrak{S}}\mathfrak{M}\to\mathrm{coker}(f)\to 0

from the definition of Fili\mathrm{Fil}^{i}\mathscr{M}, and let ff be the labelled map. Factoring through the kernel, we obtain the short exact sequence

0/Fili(S/FiliS)𝔖𝔐coker(f)0.0\to\mathscr{M}/\mathrm{Fil}^{i}\mathscr{M}\to(S/\mathrm{Fil}^{i}S)\otimes_{\mathfrak{S}}\mathfrak{M}\to\mathrm{coker}(f)\to 0.

We now base change to W(k)W(k), noting that coker(f)\mathrm{coker}(f) is a finite, free RR-module by Lemma 52 and the tensor produdct commutes with the quotients by Lemma 51. We thus obtain the exact sequence

00/(FiliRW(k))(S0/FiliS0)𝔖0𝔐0coker(f)RW(k)0.0\to\mathscr{M}_{0}/(\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}W(k))\to(S_{0}/\mathrm{Fil}^{i}S_{0})\otimes_{\mathfrak{S}_{0}}\mathfrak{M}_{0}\to\mathrm{coker}(f)\otimes_{R}W(k)\to 0.

As Fili0\mathrm{Fil}^{i}\mathscr{M}_{0} is exactly the kernel of 0(S0/FiliS0)𝔖0𝔐0\mathscr{M}_{0}\to(S_{0}/\mathrm{Fil}^{i}S_{0})\otimes_{\mathfrak{S}_{0}}\mathfrak{M}_{0} inside of 0\mathscr{M}_{0} and FiliRW(k)\mathrm{Fil}^{i}\mathscr{M}\otimes_{R}W(k) is contained in this kernel, the proof of the lemma is complete. ∎

Now that we can reduce the Breuil module filtration to the classical case, we will introduce a second filtration on MM which behaves a little more nicely under base change. In the classical case, the two filtrations agree, which is what we will show in our power series ring setting.

Definition 54.

Let

fp:Mf_{p}:\mathscr{M}\to M

be the map determined by upu\mapsto p. Note this map sends E(u)ii!\frac{E(u)^{i}}{i!} to 0. Define FiM\mathrm{F}^{i}M to be the image of Fili\mathrm{Fil}^{i}\mathscr{M} under fpf_{p}.

Lemma 55.
FiliM=FiM\mathrm{Fil}^{i}M=\mathrm{F}^{i}M
Proof.

We first show that FiliMFiM\mathrm{Fil}^{i}M\subset\mathrm{F}^{i}M. Recall that we set 𝒟=[p1]\mathscr{D}=\mathscr{M}[p^{-1}] and D=M[p1]D=M[p^{-1}] upon which we can utilize the rational theory. Set FiliDFili𝒟D.\mathrm{Fil}^{i}D\coloneqq\mathrm{Fil}^{i}\mathscr{D}\cap D. We have the exact sequence

0Fili1𝒟(FilpS)𝒟EFili𝒟(FilpS)𝒟fpFiliD0.0\to\frac{\mathrm{Fil}^{i-1}\mathscr{D}}{(\mathrm{Fil}^{p}S)\mathscr{D}}\xrightarrow{\cdot E}\frac{\mathrm{Fil}^{i}\mathscr{D}}{(\mathrm{Fil}^{p}S)\mathscr{D}}\xrightarrow{f_{p}}\mathrm{Fil}^{i}D\to 0.

The kernel of fp:Fili𝒟FiliDf_{p}:\mathrm{Fil}^{i}\mathscr{D}\to\mathrm{Fil}^{i}D is (Fil1S)Fili1𝒟(\mathrm{Fil}^{1}S)\mathrm{Fil}^{i-1}\mathscr{D} by [Du+24, Lemma 4.31], and removing the FilpS\mathrm{Fil}^{p}S part ensures that Fili1𝒟\mathrm{Fil}^{i-1}\mathscr{D} will give the entire kernel. Recalling from the rational theory that

Fili=Fili𝒟,\mathrm{Fil}^{i}\mathscr{M}=\mathrm{Fil}^{i}\mathscr{D}\cap\mathscr{M},

we obtain the exact sequence

0Fili1(FilpS)Fili(FilpS)fpFiM0.\displaystyle 0\to\frac{\mathrm{Fil}^{i-1}\mathscr{M}}{(\mathrm{Fil}^{p}S)\mathscr{M}}\to\frac{\mathrm{Fil}^{i}\mathscr{M}}{(\mathrm{Fil}^{p}S)\mathscr{M}}\xrightarrow{f_{p}}\mathrm{F}^{i}M\to 0. (1)

Since FiliMFiliMFili𝒟D\mathrm{Fil}^{i}M\subset\mathrm{Fil}^{i}\mathscr{M}\cap M\subset\mathrm{Fil}^{i}\mathscr{D}\cap D, we have an injective map

FiliMFiliDFili𝒟(FilpS)𝒟fpFiliD.\mathrm{Fil}^{i}M\hookrightarrow\mathrm{Fil}^{i}D\hookrightarrow\frac{\mathrm{Fil}^{i}\mathscr{D}}{(\mathrm{Fil}^{p}S)\mathscr{D}}\xrightarrow{f_{p}}\mathrm{Fil}^{i}D.

On the other hand, this agrees with the map

FiliMFili(FilpS)fpFiMFiliD\mathrm{Fil}^{i}M\hookrightarrow\frac{\mathrm{Fil}^{i}\mathscr{M}}{(\mathrm{Fil}^{p}S)\mathscr{M}}\xrightarrow{f_{p}}\mathrm{F}^{i}M\hookrightarrow\mathrm{Fil}^{i}D

giving us that

FiliMFiM.\mathrm{Fil}^{i}M\subset\mathrm{F}^{i}M.

Tensoring (1)(1) with W(k)W(k) and applying Lemma 53 we obtain the exact

Fili10(FilpS0)0Fili0(FilpS0)0fpFiMRW(k)0.\frac{\mathrm{Fil}^{i-1}\mathscr{M}_{0}}{(\mathrm{Fil}^{p}S_{0})\mathscr{M}_{0}}\to\frac{\mathrm{Fil}^{i}\mathscr{M}_{0}}{(\mathrm{Fil}^{p}S_{0})\mathscr{M}_{0}}\xrightarrow{f_{p}}\mathrm{F}^{i}M\otimes_{R}W(k)\to 0.

But we also know from the classical theory that Fili(M0)\mathrm{Fil}^{i}(M_{0}) is the image of Fili0\mathrm{Fil}^{i}\mathscr{M}_{0} under fpf_{p}, so we obtain that FiliM0=FiMRW(k)\mathrm{Fil}^{i}M_{0}=\mathrm{F}^{i}M\otimes_{R}W(k) and hence FiMRW(k)\mathrm{F}^{i}M\otimes_{R}W(k) is a finite, free W(k)W(k)-module. Let e1¯,,ea¯\overline{e_{1}},\cdots,\overline{e_{a}} be a basis and by Nakayama’s Lemma we can lift it to a generating set e1,,eae_{1},\cdots,e_{a} of FiM\mathrm{F}^{i}M. Since FiM\mathrm{F}^{i}M is an image after upu\mapsto p, it is clear that we can safely mod out (1)(1) by FilpS\mathrm{Fil}^{p}S to obtain the exact

0Fili1(FilpS)Fili(FilpS)fpFiM0.0\to\frac{\mathrm{Fil}^{i-1}\mathscr{M}}{(\mathrm{Fil}^{p}S)\mathscr{M}}\to\frac{\mathrm{Fil}^{i}\mathscr{M}}{(\mathrm{Fil}^{p}S)\mathscr{M}}\xrightarrow{f_{p}}\mathrm{F}^{i}M\to 0.

Now we can safely base change to the Shilov point to obtain

0Fili1g(FilpSg)gFilig(FilpSg)gfpFiMRRg0.0\to\frac{\mathrm{Fil}^{i-1}\mathscr{M}_{g}}{(\mathrm{Fil}^{p}S_{g})\mathscr{M}_{g}}\to\frac{\mathrm{Fil}^{i}\mathscr{M}_{g}}{(\mathrm{Fil}^{p}S_{g})\mathscr{M}_{g}}\xrightarrow{f_{p}}\mathrm{F}^{i}M\otimes_{R}R_{g}\to 0.

the base change in the first two terms is as written due to Theorem 50 and the classical functor from Fontaine-Laffaille modules to Breuil modules described in Definition 31. Applying the classical theory to the Shilov point, we know FiMRRg=FiliMg\mathrm{F}^{i}M\otimes_{R}R_{g}=\mathrm{Fil}^{i}M_{g} and hence is a finite, free RgR_{g} module and is of the same rank as FiliM0\mathrm{Fil}^{i}M_{0}. Note that the rank is the same again because of the Hodge-Tate property of crystalline representations. The places where the filtrations of M0M_{0} and MgM_{g} increase in rank and the amount they increase in rank by are determined completely by the Hodge-Tate weights of the crystalline representation associated to MM. These are the same on each fiber and, consequently, the ranks of FiliM0\mathrm{Fil}^{i}M_{0} and FiliMg\mathrm{Fil}^{i}M_{g} are the same.

Thus with our embedding FiMFiliMRRg\mathrm{F}^{i}M\hookrightarrow\mathrm{Fil}^{i}M\otimes_{R}R_{g}, we see that e1,,eae_{1},\cdots,e_{a} is a basis of FiM\mathrm{F}^{i}M and hence FiM\mathrm{F}^{i}M is a finite, free RR-module.

Now we know that FiM\mathrm{F}^{i}M is a finite, free RR-module contained in both FiliMg\mathrm{Fil}^{i}M_{g} and Fili𝒟\mathrm{Fil}^{i}\mathscr{D} We conclude that

FiM\displaystyle\mathrm{F}^{i}M (FiMRRg)Fili𝒟\displaystyle\subset(\mathrm{F}^{i}M\otimes_{R}R_{g})\cap\mathrm{Fil}^{i}\mathscr{D}
=FiliMgFili𝒟\displaystyle=\mathrm{Fil}^{i}M_{g}\cap\mathrm{Fil}^{i}\mathscr{D}
=FiliM\displaystyle=\mathrm{Fil}^{i}M

Theorem 56.

Fili(M0)Fili(M)RW(k).\mathrm{Fil}^{i}(M_{0})\cong\mathrm{Fil}^{i}(M)\otimes_{R}W(k).

Proof.

In the proof of the previous lemma we established that FiMRW(k)=FiliM0\mathrm{F}^{i}M\otimes_{R}W(k)=\mathrm{Fil}^{i}M_{0}. The result of the previous lemma is that FiM=FiliM\mathrm{F}^{i}M=\mathrm{Fil}^{i}M, completing the proof. ∎

This completes the proof of the following theorem:

Theorem 57.

MRW(k)M\otimes_{R}W(k) is a Fontaine-Laffaille module over W(k)W(k), and MRW(k)M0M\otimes_{R}W(k)\cong M_{0} as Fontaine-Laffaille modules.

3.4. MM is an object of MFff\mathrm{MF}_{\nabla}^{\mathrm{ff}}

We now complete the proof of Theorem 40:

Proof.

Having established Theorem 50 and Theorem 57, we have shown Condition (1) of Proposition 11. Since MM is a finite, free RR-module, condition (2) holds, and thus MM is a Fontaine-Laffaille module over RR and belongs to the category MFff(R)\mathrm{MF}^{\mathrm{ff}}(R). ∎

Theorem 58.

MM is an object of the category MFff\mathrm{MF}^{\mathrm{ff}}_{\nabla}.

Proof.

By Theorem 57 and Theorem 20, we only need to show that MM is stable under 𝔐\nabla_{\mathfrak{M}} and that it satisfies integral Griffiths Transversality.

Using Lemma 42 and Lemma 22 we see that M=M[p1]M=M[p^{-1}]\cap\mathscr{M} as submodules of [p1].\mathscr{M}[p^{-1}]. We know M[p1]M[p^{-1}] is stable under \nabla_{\mathscr{M}} (which restricts to 𝔐\nabla_{\mathfrak{M}} on M[p1]M[p^{-1}]) and, by Proposition 30, so is \mathscr{M}, showing that MM is stable under 𝔐.\nabla_{\mathfrak{M}}.

Because Fili(M)=Fili()M\mathrm{Fil}^{i}(M)=\mathrm{Fil}^{i}(\mathscr{M})\cap M, MM is stable under \nabla_{\mathscr{M}}, and \nabla_{\mathscr{M}} satisfies integral SS-Griffiths transversality by lemma 30 we conclude 𝔐\nabla_{\mathfrak{M}} satisfies Griffiths transversality, as needed. ∎

3.5. Proof of Essential Surjectivity

We now complete the proof that TcrisT_{\mathrm{cris}} is essentially surjective. Let Λ\Lambda be an object of RepZp,[0,r]cris(GR)\mathrm{Rep}_{\mathbb{Z}_{p},[0,r]}^{\mathrm{cris}}(G_{R}). Let (𝔐,φ𝔐,f)(\mathfrak{M},\varphi_{\mathfrak{M}},f) be the Kisin module with descent data associated to Λ\Lambda guaranteed by Theorem 20 (note Remark 33). Note also this theorem associates a Kisin module with descent data to an étale (φ𝒪,𝒪)(\varphi_{\mathcal{O}_{\mathcal{E}}},\mathcal{O}_{\mathcal{E}})-module \mathcal{M} with T𝒪()(r)ΛT_{\mathcal{O}_{\mathcal{E}}}(\mathcal{M})(r)\cong\Lambda, and then T𝔖r(𝔐)ΛT_{\mathfrak{S}}^{r}(\mathfrak{M})\cong\Lambda by Lemma 35.

Let 𝔐φ,𝔖S\mathscr{M}\cong\mathfrak{M}\otimes_{\varphi,\mathfrak{S}}S be the Breuil module associated with 𝔐.\mathfrak{M}. We first prove the following stepping stone to Theorem 37.

Lemma 59.

The natural injection

T𝔖r(𝔐)TS()Filr(Acris(R)S)φr=1T_{\mathfrak{S}}^{r}(\mathfrak{M})\to T_{S}(\mathscr{M})\coloneqq\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R)\otimes_{S}\mathscr{M})^{\varphi_{r}=1}

induced by the embedding Ainf(R)Acris(R)A_{\mathrm{inf}}(R)\to A_{\mathrm{cris}}(R) is an isomorphism of GR~G_{\tilde{R}_{\infty}} representations.

Proof.

We prove the lemma by reducing to the Shilov point where we argue as in [Moo24, Lemma 4.7]. First note that using Lemma 35,

ΛT𝒪()(r)T𝔖r(𝔐)(Filrφ𝔐𝔖Ainf(R))φr=1(𝒪,gur^𝒪,gg)φ=1(r)Λ\Lambda\cong T_{\mathcal{O}_{\mathcal{E}}}(\mathcal{M})(r)\cong T^{r}_{\mathfrak{S}}(\mathfrak{M})\to(\mathrm{Fil}^{r}\varphi^{*}\mathfrak{M}\otimes_{\mathfrak{S}}A_{\mathrm{inf}}(R))^{\varphi_{r}=1}\cong(\widehat{\mathcal{O}_{\mathcal{E},g}^{\mathrm{ur}}}\otimes_{\mathcal{O}_{\mathcal{E},g}}\mathcal{M}_{g})^{\varphi=1}(r)\cong\Lambda

is an isomorphism of GR~g,G_{\tilde{R}_{g,{\infty}}} representations.

Then we obtain the diagram:

T𝔖r(𝔐){T_{\mathfrak{S}}^{r}(\mathfrak{M})}TS(){T_{S}(\mathscr{M})}(Filrφ𝔐𝔖Ainf)φr=1{(\mathrm{Fil}^{r}\varphi^{*}\mathfrak{M}\otimes_{\mathfrak{S}}A_{\mathrm{inf}})^{\varphi_{r}=1}}Filr(Acris(Rg)Sg)φr=1{\mathrm{Fil}^{r}(A_{\mathrm{cris}}(R_{g})\otimes_{S}\mathscr{M}_{g})^{\varphi_{r}=1}}\scriptstyle{\cong}\scriptstyle{\cong} where the bottom map is an isomorphism by [LL23, Prop 6.12] and AinfA_{\mathrm{inf}} represents the usual AinfA_{\mathrm{inf}} from the classical theory. We do not know whether the right vertical map is an injection since the natural map Acris(R)Acris(Rg)A_{\mathrm{cris}}(R)\to A_{\mathrm{cris}}(R_{g}) requires a pp-adic completion. We can, however, extend the diagram to the right because if MM is the Fontaine-Laffaille module associated to \mathscr{M}, then the unique section of Lemma 42 and the discussion of Section 2.9.4 induces an isomorphism TS()Tcris(M)T_{S}(\mathscr{M})\cong T_{\mathrm{cris}}(M), and we get a similar isomorphism on the Shilov point. This gives us the diagram:

T𝔖r(𝔐){T_{\mathfrak{S}}^{r}(\mathfrak{M})}TS(){T_{S}(\mathscr{M})}Tcris(M){T_{\mathrm{cris}}(M)}(Filrφ𝔐𝔖Ainf)φr=1{(\mathrm{Fil}^{r}\varphi^{*}\mathfrak{M}\otimes_{\mathfrak{S}}A_{\mathrm{inf}})^{\varphi_{r}=1}}Filr(gSAcris(Rg))φr=1{\mathrm{Fil}^{r}(\mathscr{M}_{g}\otimes_{S}A_{\mathrm{cris}}(R_{g}))^{\varphi_{r}=1}}Tcris(Mg){T_{\mathrm{cris}}(M_{g})}\scriptstyle{\cong}\scriptstyle{\cong}\scriptstyle{\cong}\scriptstyle{\cong}\scriptstyle{\cong} where the rightmost vertical arrow is an isomorphism by [LMP23, Cor 2.3.5]. Since T𝔖r(𝔐)TS()T_{\mathfrak{S}}^{r}(\mathfrak{M})\to T_{S}(\mathscr{M}) is an injection, it must then be an isomorphism, completing the proof of the lemma. ∎

The last diagram in the proof of this lemma then establishes that Tcris(M)ΛT_{\mathrm{cris}}(M)\cong\Lambda as GRG_{R_{\infty}} representations. We now are able to prove Theorem 37 by showing this is actually an isomorphism of GRG_{R} representations which comes from the fact that the above maps are compatible with the relevant Galois actions.

The isomorphism TS()Tcris(M)T_{S}(\mathscr{M})\to T_{\mathrm{cris}}(M) is determined by the unique \nabla-compatible section MM\to\mathscr{M}, and it is clear that the Galois action on Tcris(M)T_{\mathrm{cris}}(M) determined by \nabla on MM is compatible with the Galois action on TS()T_{S}(\mathscr{M}) determined by \nabla on \mathscr{M}, confirming that this map is actually an isomorphism of GRG_{R} representations.

To see that the map T𝔖r(𝔐)TS()T_{\mathfrak{S}}^{r}(\mathfrak{M})\to T_{S}(\mathscr{M}) is compatible with the GRG_{R} action, [Du+24, Proposition 4.6] confirms that there is a unique descent datum ff on 𝔐\mathfrak{M} such that fS=idSφ,𝔖ff_{S}=\mathrm{id}_{S}\otimes_{\varphi,\mathfrak{S}}f. As the descent data is used to construct the full GRG_{R} action in the same way as described in Section 2.9.4 and Section 2.9.2 and this is the same as the Galois action defined via the \nabla structure, the Galois action established on T𝔖r(𝔐)T_{\mathfrak{S}}^{r}(\mathfrak{M}) is compatible with the Galois action on TS()T_{S}(\mathscr{M}) and thus the map T𝔖r(𝔐)TS()T_{\mathfrak{S}}^{r}(\mathfrak{M})\to T_{S}(\mathscr{M}) which was previously shown to be an isomorphism of GRG_{R_{\infty}} representations is actually an isomorphism of GRG_{R} representations, completing the proof of Theorem 37.

From the diagram in the above proof, we have also shown that Tcris(M)T𝔖r(𝔐)ΛT_{\mathrm{cris}}(M)\cong T_{\mathfrak{S}}^{r}(\mathfrak{M})\cong\Lambda as GRG_{R} representations, which completes the proof of essential surjectivity over RR.

4. Essential Surjectivity of TcrisT_{\mathrm{cris}} over small base rings

We will now prove theorems analaogous to those in the last section but in more generality over small base rings. While the results are analogous, the proof is easier because we are able to leverage the theory from the previous section to define a φi\varphi_{i} structure via the rational theory, which would be difficult to do for the power series ring. Throughout this section, fix TT a small base ring. Write 𝔖=𝔖T=Tu\mathfrak{S}=\mathfrak{S}_{T}=T\llbracket u\rrbracket and S=STS=S_{T}. Let I0SI_{0}\subset S denote the kernel of the projection STS\to T given by u0u\mapsto 0.

4.1. Statement of Main Theorems

Theorem 60.

Let \mathscr{M} be a Breuil module which arises from a Kisin module which arises from a Λ,\Lambda, an object in RepZpcris[0,r](GT).\mathrm{Rep}_{\mathbb{Z}_{p}}^{\mathrm{cris}[0,r]}(G_{T}). Set

M/I0𝔐/u𝔐T,φ𝔖T.M\coloneqq\mathscr{M}/I_{0}\mathscr{M}\cong\mathfrak{M}/u\mathfrak{M}\otimes_{T,\varphi_{\mathfrak{S}}}T.

Then MM carries the data of a Fontaine-Laffaille module and satisfies MRSM\otimes_{R}S\cong\mathscr{M} as in Definition 31 We call it the “Fontaine-Laffaille module associated to Λ\Lambda.”

Theorem 61.

Let MM be the Fontaine-Laffaille module associated to Λ.\Lambda. Then

Tcris(M)Λ.T_{\mathrm{cris}}(M)\cong\Lambda.

4.2. Preliminaries

Similar to the previous section, our strategy will be to show the Fontaine-Laffaille data is compatible with the base change to T𝔪^R\widehat{T_{\mathfrak{m}}}\cong R discussed in Section 2.2.3 and then to apply Theorem 13. This time we will be able to rely on the rational theory combined with the theory we have established for a power series ring.

Recall that associated to VΛZpQpV\coloneqq\Lambda\otimes_{\mathbb{Z}_{p}}\otimes\mathbb{Q}_{p} is Dcris(V)D_{\mathrm{cris}}^{\vee}(V) (which we will denote with DD in this section), a finite, projective T[p1]T[p^{-1}] module equipped with a filtration structure FiliD\mathrm{Fil}^{i}D, a Frobenius structure φD\varphi_{D}, and a D\nabla_{D} structure. We will transfer the data on DD to MM in a similar way as how in the last section we transferred the data on \mathscr{M} to MM.

By Theorem 23, we can see that MTT𝔪^M\otimes_{T}\widehat{T_{\mathfrak{m}}} and Tcris(Λ|T𝔪^)T_{\mathrm{cris}}(\Lambda|_{\widehat{T_{\mathfrak{m}}}}), the Fontaine-Laffaille over T𝔪^\widehat{T_{\mathfrak{m}}} associated to Λ|T𝔪^\Lambda|_{\widehat{T_{\mathfrak{m}}}} are isomorphic as modules (note that T𝔪^\widehat{T_{\mathfrak{m}}} is a local ring, and a projective module over a local ring is free). Write M^\widehat{M} for Tcris(Λ|T𝔪^)T_{\mathrm{cris}}(\Lambda|_{\widehat{T_{\mathfrak{m}}}}) through the remainder of this section.

Lemma 62.

Consider the projection q:Mq:\mathscr{M}\to M induced by the φ\varphi-compatible projection STS\to T, u0u\mapsto 0. Then qq admits a unique φ\varphi-compatible section s:Ms:M\to\mathscr{M}.

Proof.

The proof of Lemma 43 works identically when the base ring is TT. ∎

This section gives us an injective map M𝒟[p1]M\hookrightarrow\mathscr{M}\hookrightarrow\mathscr{D}\coloneqq\mathscr{M}[p^{-1}] and the projection u0u\mapsto 0 takes 𝒟\mathscr{D} to DD, and thus we have an injective map MD.M\hookrightarrow D.

4.3. The Filtration Structure

We first define the filtration structure on MM via the filtration on DD from the rational theory.

Definition 63.

Set

FiliMFiliDM.\mathrm{Fil}^{i}M\coloneqq\mathrm{Fil}^{i}D\cap M.

It remains to show this filtration behaves well under base change.

Lemma 64.

FiliMTT𝔪^FiliM^\mathrm{Fil}^{i}M\otimes_{T}\widehat{T_{\mathfrak{m}}}\cong\mathrm{Fil}^{i}\widehat{M}

Proof.

Note that [Du+24, Lemma 2.13] describes the compatibility of DD under base change. Since M^\widehat{M} is a Fontaine-Laffaille module we know Dcris(Λ|T𝔪^)M^[p1]D_{\mathrm{cris}}^{\vee}(\Lambda|_{\widehat{T_{\mathfrak{m}}}})\cong\widehat{M}[p^{-1}] and FiliM^=FiliDcris(Λ|T𝔪^)M.\mathrm{Fil}^{i}\widehat{M}=\mathrm{Fil}^{i}D_{\mathrm{cris}}^{\vee}(\Lambda|_{\widehat{T_{\mathfrak{m}}}})\cap M. Using Lemma 22 we have

Fil(M)TT𝔪^\displaystyle\mathrm{Fil}(M)\otimes_{T}\widehat{T_{\mathfrak{m}}} =(FiliDM)TT𝔪^\displaystyle=(\mathrm{Fil}^{i}D\cap M)\otimes_{T}\widehat{T_{\mathfrak{m}}}
=(FiliDTT𝔪^)(MTT𝔪^)\displaystyle=(\mathrm{Fil}^{i}D\otimes_{T}\widehat{T_{\mathfrak{m}}})\cap(M\otimes_{T}\widehat{T_{\mathfrak{m}}})
=Fili(Dcris(Λ|T𝔪^))M^\displaystyle=\mathrm{Fil}^{i}(D_{\mathrm{cris}}^{\vee}(\Lambda|_{\widehat{T_{\mathfrak{m}}}}))\cap\widehat{M}
=FiliM^.\displaystyle=\mathrm{Fil}^{i}\widehat{M}.

4.4. The φi\varphi_{i} structure

We now define the φi\varphi_{i} structure on MM. Fix 𝔪\mathfrak{m} a maximal ideal of TT and again write M^\widehat{M} For MT𝔪^M\otimes\widehat{T_{\mathfrak{m}}} As M^\widehat{M} is a Fontaine-Laffaille module, there is a φi\varphi_{i} structure on M^\widehat{M} which we will denote φi^\widehat{\varphi_{i}}. Via our φ\varphi-compatible embedding MM^M\hookrightarrow\widehat{M}, we can define a φi\varphi_{i} structure on MM as

φiφ^i|FiliM.\varphi_{i}\coloneqq\widehat{\varphi}_{i}\bigg{|}_{\mathrm{Fil}^{i}M}.

Let D^M^[p1]=Dcris(Λ|T𝔪^)\widehat{D}\coloneqq\widehat{M}[p^{-1}]=D_{\mathrm{cris}}^{\vee}(\Lambda|_{\widehat{T_{\mathfrak{m}}}}) Note that φi^=φD^pi\widehat{\varphi_{i}}=\frac{\varphi_{\widehat{D}}}{p^{i}}, and the compatibility of DcrisD_{\mathrm{cris}}^{\vee} with base change in [Du+24, Lemma 2.13] gives us that φDpi\frac{\varphi_{D}}{p^{i}} acts the same on elements of FiliMD\mathrm{Fil}^{i}M\subset D and sends them into DD. Thus φi\varphi_{i} sends FiliM\mathrm{Fil}^{i}M into

DM^=MTT[p1]MTT𝔪^=MD\cap\widehat{M}=M\otimes_{T}T[p^{-1}]\cap M\otimes_{T}\widehat{T_{\mathfrak{m}}}=M

by Lemma 22. Since the φ^i\widehat{\varphi}_{i} satisfies φi1^=pφ^i\widehat{\varphi_{i-1}}=p\widehat{\varphi}_{i}, the same holds true for our new φi\varphi_{i} structure.

4.5. The \nabla Structure

As DD comes equipped with a D\nabla_{D} structure, set

MD|M.\nabla_{M}\coloneqq\nabla_{D}\big{|}_{M}.

The proof that MM is stable under M\nabla_{M} is similar to the proof that MM is stable under the φi\varphi_{i} structure.

We can embed MM into M^\widehat{M} which is stable under its M^\nabla_{\widehat{M}} structure, and we can also embed MM into DD which is stable under its D\nabla_{D} structure. As DM^=MD\cap\widehat{M}=M as above, we see that MM is stable under M\nabla_{M}.

4.6. TcrisT_{\mathrm{cris}} is essentially surjective over TT.

Combining the last few subsections, we have equipped MM with the data of an object in MFbig,(T)\mathrm{MF}_{\mathrm{big},\nabla}(T) which is compatible with each base change to T𝔪^\widehat{T_{\mathfrak{m}}}, completing the proof of Theorem 60 by utilizing Theorem 13. It only remains to show that

Theorem 65.

Tcris(M)Λ.T_{\mathrm{cris}}(M)\cong\Lambda.

Proof.

We define a map Tcris(M)Tcris(M^)T_{\mathrm{cris}}(M)\to T_{\mathrm{cris}}(\widehat{M}) via the functors to Galois representations from Kisin modules. As Tcris(M)=TS()T_{\mathrm{cris}}(M)=T_{S}(\mathscr{M}), we obtain an injective map

T𝔖r(𝔐)=(Filr𝔐Ainf(T))φr=1(FilrAcris(T))φr=1=Tcris(M)T_{\mathfrak{S}}^{r}(\mathfrak{M})=(\mathrm{Fil}^{r}\mathfrak{M}\otimes A_{\mathrm{inf}}(T))^{\varphi_{r}=1}\hookrightarrow(\mathrm{Fil}^{r}\mathscr{M}\otimes A_{\mathrm{cris}}(T))^{\varphi_{r}=1}=T_{\mathrm{cris}}(M)

and analogously for 𝒯𝔖r(𝔐𝔖T𝔖R)Tcris(M^)\mathcal{T}_{\mathfrak{S}}^{r}(\mathfrak{M}\otimes_{\mathfrak{S}_{T}}\mathfrak{S}_{R})\to T_{\mathrm{cris}}(\widehat{M}). We obtain the diagram

Tcris(M){T_{\mathrm{cris}}(M)}Tcris(M^){T_{\mathrm{cris}}(\widehat{M})}T𝔖r(𝔐){T_{\mathfrak{S}}^{r}(\mathfrak{M})}𝒯𝔖r(𝔐𝔖T𝔖T𝔪^){\mathcal{T}_{\mathfrak{S}}^{r}(\mathfrak{M}\otimes_{\mathfrak{S}_{T}}\mathfrak{S}_{\widehat{T_{\mathfrak{m}}}})}\scriptstyle{\cong}\scriptstyle{\cong} as we also have a map Tcris(M)Tcris(M^)T_{\mathrm{cris}}(M)\to T_{\mathrm{cris}}(\widehat{M}) since Tcris(M)Tcris(D)T_{\mathrm{cris}}(M)\subset T_{\mathrm{cris}}(D) and Tcris(M^)Tcris(D^)T_{\mathrm{cris}}(\widehat{M})\subset T_{\mathrm{cris}}(\widehat{D}) and Tcris(D)=Tcris(D^).T_{\mathrm{cris}}(D)=T_{\mathrm{cris}}(\widehat{D}).

The right vertical arrow in the diagram above is an isomorphism by the proof of Lemma 59 and the bottom horizontal arrow is an isomorphism by Remark 34 The left vertical arrow is an injection. Thus the injective Tcris(M)Tcris(M^)T_{\mathrm{cris}}(M)\to T_{\mathrm{cris}}(\widehat{M}) is forced to be an isomorphism, which quickly completes the proof. ∎

References

  • [Bre02] Christophe Breuil “Integral pp-adic Hodge Theory” In Advanced Studies in Pure Mathematics, 2002, pp. 51–80
  • [Bri08] Olivier Brinon “Représentations pp-adiques cristallines et de de Rham dans le cas relatif”, Mémoires de la Société Mathématique de France 112 Société mathématique de France, 2008 DOI: 10.24033/msmf.424
  • [DL22] Heng Du and Tong Liu “A prismatic Approach to (ϕ,G^)(\phi,\hat{G})-modules and FF-Crystals.”, 2022 arXiv:2107.12240v2 [math.NT]
  • [Du+24] Heng Du, Tong Liu, Yong Suk Moon and Koji Shimizu “Completed prismatic F-crystals and crystalline Zp-local systems” In Compositio Mathematica 160.5, 2024, pp. 1101–1166 DOI: 10.1112/S0010437X24007097
  • [Fal89] Gerd Faltings “Algebraic analysis, geometry, and number theory” Baltimore, MD: Johns Hopkins University Press, 1989, pp. 25–80
  • [FL82] Jean-Marc Fontaine and Guy Laffaille “Construction de représentations pp-adiques” In Annales scientifiques de l’École Normale Supérieure 15.4, 1982, pp. 547–608
  • [Gao19] Hui Gao “Fontaine-Laffaille modules and strongly divisible modules” In Annales mathématiques du Québec 43, 2019, pp. 145–159 DOI: https://doi.org/10.1007/s40316-017-0090-1
  • [GLS14] Toby Gee, Tong Liu and David Savitt “The Buzzard-Diamond-Jarvis Conjecture For Unitary Groups” In Journal of the Americal Mathematical Society 27.2, 2014, pp. 389–435
  • [IKY24] Naoki Imai, Hiroki Kato and Alex Youcis “The prismatic realization functor for Shimura varieties of abelian type”, 2024 arXiv:2310.08472 [math.NT]
  • [Kim15] Wansu Kim “The relative Breuil-Kisin classification of p-divisible groups and finite flat group schemes” In Int. Math. Res. Not., 2015, pp. 8152–8232
  • [Kis06] Mark Kisin “Crystalline representations and F-crystals” In Algebraic geometry and number theory 253, 2006, pp. 459–496
  • [LL23] Shizahang Li and Tong Liu “Comparison of Prismatic Cohomology and Derived de Rham Cohomology” In Journal of the European Mathematical Society, 2023 DOI: DOI 10.4171/JEMS/1377
  • [LMP23] Tong Liu, Yong Suk Moon and Deepam Patel “Relative Fontaine–Messing Theory Over Power Series Rings” In International Mathematics Research Notices 2024.5 Oxford University Press (OUP), 2023, pp. 4384–4455 DOI: 10.1093/imrn/rnad247
  • [Moo24] Yong Suk Moon “Strongly Divisible Lattices and Crystalline Cohomology in the Imperfect Residue Field Case” In Selecta Mathematica 30.12, 2024 DOI: https://doi.org/10.1007/s00029-023-00899-y
  • [Wür23] Matti Würthen “Divided prismatic Frobenius crystals of small height and the category \mathcal{MF} arXiv, 2023 URL: https://arxiv.org/pdf/2305.06081.pdf