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

\cellspacetoplimit

0.25cm \cellspacebottomlimit0.25cm \titlecontentssection [1.5em] \contentslabel1.5em \contentspage \titlecontentssubsection [2.3em] \contentslabel2.3em \contentspage

Crystalline condition for 𝑨𝐢𝐧𝐟\boldsymbol{A_{\mathrm{inf}}}–cohomology and ramification bounds

Pavel Čoupek Department of Mathematics, Michigan State University [email protected]
Abstract.

For a prime p>2p>2 and a smooth proper pp–adic formal scheme 𝒳\mathscr{X} over 𝒪K\mathcal{O}_{K} where KK is a pp–adic field, we study a series of conditions (Crs\mathrm{Cr}_{s}), s0s\geq 0 that partially control the GKG_{K}–action on the image of the associated Breuil–Kisin prismatic cohomology 𝖱ΓΔ(𝒳/𝔖)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S}) inside the AinfA_{\mathrm{inf}}–prismatic cohomology 𝖱ΓΔ(𝒳Ainf/Ainf)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}). The condition (Cr0\mathrm{Cr}_{0}) is a criterion for a Breuil–Kisin–Fargues GKG_{K}–module to induce a crystalline representation used by Gee and Liu in [EmertonGee2, Appendix F], and thus leads to a proof of crystallinity of Hiét(𝒳η¯,Qp)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Q}_{p}) that avoids the crystalline comparison. The higher conditions (Crs\mathrm{Cr}_{s}) are used to adapt the strategy of Caruso and Liu from [CarusoLiu] to establish ramification bounds for the mod pp representations Hiét(𝒳η¯,Z/pZ),\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}), for arbitrary ee and ii, which extend or improve existing bounds in various situations.

1.  Introduction

Let kk be a perfect field of characteristic p>2p>2 and K=W(k)[1/p]K^{\prime}=W(k)[1/p] the associated absolutely unramified field. Let K/KK/K^{\prime} be a totally ramified finite extension with ramification index ee, and denote by GKG_{K} its absolute Galois group. The goal of the present paper is to provide new bounds for ramification of the mod pp representations of GKG_{K} that arise as the étale cohomology groups Hiét(𝒳η¯,Z/pZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}) in terms of p,ip,i and ee, where 𝒳\mathscr{X} is a smooth proper pp–adic formal scheme over 𝒪K\mathcal{O}_{K} (and 𝒳η¯\mathscr{X}_{\overline{\eta}} is the geometric adic generic fiber). Concretely, let us denote by GKμG_{K}^{\mu} the μ\mu–th ramification group of GKG_{K} in the upper numbering (in the standard convention, e.g. [SerreLocalFields]) and GK(μ)=GKμ1G_{K}^{(\mu)}=G_{K}^{\mu-1}. The main result is as follows.

Theorem 1.1 (Theorem LABEL:thm:FinalRamification).

Set

α=logp(max{ipp1,(i1)ep1})+1,β=1pα(iepp11).\alpha=\left\lfloor\mathrm{log}_{p}\left(\mathrm{max}\left\{\frac{ip}{p-1},\frac{(i-1)e}{p-1}\right\}\right)\right\rfloor+1,\;\;\;\beta=\frac{1}{p^{\alpha}}\left(\frac{iep}{p-1}-1\right).

Then:

  1. (1)

    The group GK(μ)G_{K}^{(\mu)} acts trivially on Hiét(𝒳η¯,Z/pZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}) when μ>1+eα+max{β,ep1}.\mu>1+e\alpha+\mathrm{max}\left\{\beta,\frac{e}{p-1}\right\}.

  2. (2)

    Denote by LL the field K¯H\overline{K}^{H} where HH is the kernel of the GKG_{K}–representation ρ\rho given by Hiét(𝒳η¯,Z/pZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}). Then

    vK(𝒟L/K)<1+eα+β,v_{K}(\mathcal{D}_{L/K})<1+e\alpha+\beta,

    where 𝒟L/K\mathcal{D}_{L/K} denotes the different of the extension L/KL/K and vKv_{K} denotes the additive valuation on KK normalized so that vK(K×)=Z.v_{K}(K^{\times})=\mathbb{Z}.

In particular, unlike in previous results of this type (discussed below), there are no restrictions on the size of ee and ii with respect to pp.

Remark 1.2.

As the constants α,β\alpha,\beta appearing in Theorem 1.1 are quite complicated, let us draw some non–optimal, but more tractable consequences. The group GK(μ)G_{K}^{(\mu)} acts trivially on Hiét(𝒳η¯,Z/pZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}) when one of the following occurs:

  1. (1)

    epe\leq p and μ>1+e(logp(ipp1)+1)+e,\mu>1+e\left(\left\lfloor\mathrm{log}_{p}\left(\frac{ip}{p-1}\right)\right\rfloor+1\right)+e,

  2. (2)

    e>pe>p and μ>1+e(logp(iep1)+1)+p,\mu>1+e\left(\left\lfloor\mathrm{log}_{p}\left(\frac{ie}{p-1}\right)\right\rfloor+1\right)+p,111Strictly speaking, to obtain this precise form one has to replace (i1)e(i-1)e in α\alpha from Theorem 1.1 by ie,ie, and modify β\beta appropriately; one can show that such form of Theorem 1.1 is still valid.

  3. (3)

    i=1i=1 (e,pe,p are arbitrary) and μ>1+e(1+1p1).\mu>1+e\left(1+\frac{1}{p-1}\right).

Let us briefly summarize the history of related results. Questions of this type originate in Fontaine’s paper [Fontaine], where he proved that for a finite flat group scheme Γ\Gamma over 𝒪K\mathcal{O}_{K} that is annihilated by pnp^{n}, GK(μ)G_{K}^{(\mu)} acts trivially on Γ(K¯)\Gamma(\overline{K}) when μ>e(n+1/(p1))\mu>e(n+1/(p-1)); this is a key step in his proof that there are no non–trivial abelian schemes over Z\mathbb{Z}. In the same paper, Fontaine conjectured that general pnp^{n}–torsion cohomology would follow the same pattern: given a proper smooth variety XX over KK with good reduction, GK(μ)G_{K}^{(\mu)} should act trivially on Hiét(XK¯,Z/pnZ)\mathrm{H}^{i}_{\text{\'{e}t}}(X_{\overline{K}},\mathbb{Z}/p^{n}\mathbb{Z}) when μ>e(n+i/(p1))\mu>e(n+i/(p-1)).

This conjecture has been subsequently proved by Fontaine himself ([Fontaine2]) in the case when e=n=1,i<p1e=n=1,i<p-1 and by Abrashkin ([Abrashkin]; see also [Abrashkin2]) when e=1,i<p1e=1,i<p-1 and nn is arbitrary. This is achieved by using Fontaine–Laffaille modules (introduced in [FontaineLaffaille]), which parametrize quotients of pairs of GKG_{K}–stable lattices in crystalline representations with Hodge–Tate weights in [0,i][0,i] (such as Hiét(XK¯,Qp)\mathrm{H}^{i}_{\text{\'{e}t}}({X}_{\overline{K}},\mathbb{Q}_{p})^{\vee}). The (duals of the) representations Hiét(XK¯,Z/pnZ)\mathrm{H}^{i}_{\text{\'{e}t}}({X}_{\overline{K}},\mathbb{Z}/p^{n}\mathbb{Z}) are included among these thanks to a comparison theorem of Fontaine–Messing ([FontaineMessing]). Similarly to the orginal application, these ramification bounds lead to a scarcity result for existence of smooth proper Z\mathbb{Z}–schemes.

Various extensions to the semistable case subsequently followed. Under the asumption i<p1i<p-1 (and arbitrary ee), Hattori proved in [Hattori] a ramification bound for pnp^{n}–torsion quotients of lattices in semistable representations with Hodge–Tate weights in the range [0,i],[0,i], using (a variant of) Breuil’s filtered (ϕr,N)(\phi_{r},N)–modules. Thanks to a comparison result between log–crystalline and étale cohomology by Caruso ([CarusoLogCris]), this results in a ramification bound for Hiét(XK¯,Z/pnZ)\mathrm{H}^{i}_{\text{\'{e}t}}({X}_{\overline{K}},\mathbb{Z}/p^{n}\mathbb{Z}) when X{X} is proper with semistable reduction, assuming ie<p1ie<p-1 when n=1n=1 and (i+1)e<p1(i+1)e<p-1 when n2n\geq 2 222Recently, in [LiLiu] Li and Liu extended Caruso’s result to the range ie<p1ie<p-1 regardless of nn, for 𝒳/𝒪K\mathscr{X}/\mathcal{O}_{K} proper and smooth (formal) scheme. In view of this, results of [Hattori] should apply in these situations as well..

These results were further extended by Caruso and Liu in [CarusoLiu] for all pnp^{n}–torsion quotients of pairs of semistable lattices with Hodge–Tate weights in [0,i][0,i], without any restriction on ii or ee. The proof uses the theory of (φ,G^)(\varphi,\widehat{G})–modules, which are objects suitable for description of lattices in semistable representations. Roughly speaking, a (φ,G^)(\varphi,\widehat{G})–module consists of a Breuil–Kisin module MM and the datum of an action of G^=Gal(K(μp,π1/p)/K)\widehat{G}=\mathrm{Gal}(K(\mu_{p^{\infty}},\pi^{1/p^{\infty}})/K) on M^=M𝔖,φ^\widehat{M}=M\otimes_{\mathfrak{S},\varphi}\widehat{\mathcal{R}} where ^\widehat{\mathcal{R}} is a suitable subring of Fontaine’s period ring Ainf=W(𝒪CK)A_{\mathrm{inf}}=W(\mathcal{O}_{\mathbb{C}_{K}^{\flat}}) (and πK\pi\in K is a fixed choice of a uniformizer). An obstacle to applying the results of [CarusoLiu] to the torsion étale cohomology groups Hiét(XK¯,Z/pZ)\mathrm{H}^{i}_{\text{\'{e}t}}(X_{\overline{K}},\mathbb{Z}/p\mathbb{Z}) is that it is not quite clear when (duals of) such representations come as a quotient of two semistable lattices with Hodge–Tate weights in [0,i].[0,i]. This is indeed the case in the situation when e=1e=1, i<p1i<p-1 and XX has good reduction by the aforementioned Fontaine–Messing theorem, and it was also shown in the case i=1i=1 (no restriction on e,pe,p) for XX with semistable reduction by Emerton and Gee in [EmertonGee1], but in general the question seems open.

Nevertheless, the idea of the proof of Theorem 1.1 is to follow the general strategy of Caruso and Liu. While one does not necessarily have semistable lattices and the associated (φ,G^)(\varphi,\widehat{G})–modules to work with, a suitable replacement comes from the recently developed cohomology theories of Bhatt–Morrow–Scholze and Bhatt–Scholze ([BMS1, BMS2, BhattScholze]). Concretely, to a smooth pp–adic formal scheme 𝒳\mathscr{X} one can associate the “pnp^{n}-torsion prismatic cohomologies”

𝖱ΓΔ,n(𝒳/𝔖)=𝖱ΓΔ(𝒳/𝔖)𝖫Z/pnZ,𝖱ΓΔ,n(𝒳Ainf/Ainf)=𝖱ΓΔ(𝒳Ainf/Ainf)𝖫Z/pnZ\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/\mathfrak{S})=\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S})\stackrel{{\scriptstyle{\mathsf{L}}}}{{\otimes}}\mathbb{Z}/p^{n}\mathbb{Z},\;\;\;\;\;\;\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}},n}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}})=\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}})\stackrel{{\scriptstyle{\mathsf{L}}}}{{\otimes}}\mathbb{Z}/p^{n}\mathbb{Z}

where 𝖱ΓΔ(𝒳Ainf/Ainf),𝖱ΓΔ(𝒳/𝔖)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}),\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S}) are the prismatic avatars of the AinfA_{\mathrm{inf}}– and Breuil–Kisin cohomologies from [BMS1] and [BMS2], resp. Taking MBK=HiΔ,1(𝒳/𝔖)M_{\mathrm{BK}}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},1}(\mathscr{X}/\mathfrak{S}) and Minf=HiΔ,1(𝒳/Ainf),M_{\inf}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},1}(\mathscr{X}/A_{\mathrm{inf}}), Li and Liu showed in [LiLiu] that MBKM_{\mathrm{BK}} is a pp–torsion Breuil–Kisin module, MinfM_{\inf} is a pp–torsion Breuil–Kisin–Fargues GKG_{K}–module, and that these modules recover the étale cohomology group Hiét(𝒳η¯,Z/pZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}) essentially due to the étale comparison theorem for prismatic cohomology from [BhattScholze]. The pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) then serves as a suitable replacement of a (φ,G^)(\varphi,\widehat{G})–module in our context.

The most significant deviation from the strategy of [CarusoLiu] then stems from the fact that the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) obtained this way is “inherently pp–torsion”, that is, it does not come equipped with any apparent lift to analogous objects in characteristic 0. This is not the case in [CarusoLiu], where all torsion modules ultimately originate from a free (φ,G^)(\varphi,\widehat{G})–module (M,M^)(M,\widehat{M}). A key technical input in loc. cit. is to establish a partial control on the Galois action on MM inside M^,\widehat{M}, namely, a condition of the form

(1.1) gGK(π1/ps),xM:g(x)x(Jn,s+pnAinf)(M^^Ainf).\forall g\in G_{K(\pi^{1/p^{s}})},\;\;\forall x\in M:\;\;g(x)-x\in(J_{n,s}+p^{n}A_{\mathrm{inf}})(\widehat{M}\otimes_{\widehat{\mathcal{R}}}A_{\mathrm{inf}}).

Here Jn,sAinfJ_{n,s}\subseteq A_{\mathrm{inf}} are certain ideals (that are shrinking with growing ss). This is a “rational” fact, in the sense that this claim is a consequence of the description of the Galois action in terms of the monodromy operator on the associated Breuil module 𝒟(M^)\mathcal{D}(\widehat{M}) (cf. [Breuil], [LiuLatticesNew, §3.2]), a vector space over the characteristic 0 field KK^{\prime}.

As a starting point for replacing (1.1) in our context, we turn to a result by Gee and Liu in [EmertonGee2, Appendix F] (see also [Ozeki, Theorem 3.8]). Given a finite free Breuil–Kisin module MBKM_{\mathrm{BK}} (of finite height) and a compatible structure of Breuil–Kisin–Fargues GKG_{K}–module on Minf=MBK𝔖Ainf,M_{\inf}=M_{\mathrm{BK}}\otimes_{\mathfrak{S}}A_{\mathrm{inf}}, such that the image of MBKM_{\mathrm{BK}} under the natural map lands in (Minf)GK(π1/p)(M_{\inf})^{G_{K(\pi^{1/p^{\infty}})}}, the étale realization of MinfM_{\inf} is crystalline if and only if

(Cr0\mathrm{Cr}_{0}) gGK,xMBK:g(x)xφ1([ε¯]1)[π¯]Minf.\forall g\in G_{K},\;\;\forall x\in M_{\mathrm{BK}}:g(x)-x\in\varphi^{-1}([\underline{\varepsilon}]-1)[\underline{\pi}]M_{\inf}.

Here [][-] denotes the Teichmüller lift and ε¯,π¯\underline{\varepsilon},\underline{\pi} are the elements of 𝒪CK\mathcal{O}_{\mathbb{C}_{K}^{\flat}} given by a collection (ζpn)n(\zeta_{p^{n}})_{n} of (compatible) pnp^{n}–th roots of unity and a collection (π1/pn)n(\pi^{1/p^{n}})_{n} of pnp^{n}–th roots of the chosen uniformizer π\pi, resp. We call condition (Cr0\mathrm{Cr}_{0}) the crystalline condition. As the considered formal scheme 𝒳\mathscr{X} is assumed to be smooth over 𝒪K\mathcal{O}_{K}, it is reasonable to expect that the same condition applies to the pair MBK=HiΔ(𝒳/𝔖)M_{\mathrm{BK}}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S}) and Minf=HiΔ(𝒳Ainf/Ainf)M_{\inf}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}), despite the fact that the Breuil–Kisin and Breuil–Kisin–Fargues modules coming from prismatic cohomology are not necessarily free.

This is indeed the case and, moreover, it can be shown that the crystalline condition even applies to the embedding of the chain complexes 𝖱ΓΔ(𝒳/𝔖)𝖱ΓΔ(𝒳Ainf/Ainf)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S})\rightarrow\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}): to make sense of this claim, we model the cohomology theories by their associated Čech–Alexander complexes. These were introduced in [BhattScholze] in the case that 𝒳\mathscr{X} is affine, but can be extended to (at least) arbitrary separated smooth pp–adic formal schemes. We are then able to verify the condition termwise for this pair of complexes. More generally, we introduce a decreasing series of ideals IsI_{s}, s0s\geq 0 where I0=φ1([ε¯]1)[π¯]Ainf,I_{0}=\varphi^{-1}([\underline{\varepsilon}]-1)[\underline{\pi}]A_{\mathrm{inf}}, and then formulate and prove the analogue of (Cr0\mathrm{Cr}_{0}) for IsI_{s} and the action of GK(π1/ps).G_{K(\pi^{1/p^{s}})}. As a consequence, we obtain:

Theorem 1.3 (Theorem 4.1, Corollary 4.5, Proposition 4.8).

Let 𝒳\mathscr{X} be a smooth separated pp–adic formal scheme over 𝒪K\mathcal{O}_{K}.

  1. (1)

    Fix a compatible choice of Čech–Alexander complexes CˇBKCˇinf\check{C}_{\mathrm{BK}}^{\bullet}\subseteq\check{C}_{\inf}^{\bullet} that compute 𝖱ΓΔ(𝒳/𝔖)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S}) and 𝖱ΓΔ(𝒳Ainf/Ainf)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}), resp. Then for all s0s\geq 0, the pair (CˇBK,Cˇinf)(\check{C}_{\mathrm{BK}}^{\bullet},\check{C}_{\inf}^{\bullet}) satisfies (termwise) the condition

    (Crs\mathrm{Cr}_{s}) gGK(π1/ps),xCˇBK:g(x)xIsCˇinf.\forall g\in G_{K(\pi^{1/p^{s}})},\;\;\forall x\in\check{C}_{\mathrm{BK}}^{\bullet}:g(x)-x\in I_{s}\check{C}_{\inf}^{\bullet}.
  2. (2)

    The associated prismatic cohomology groups satisfy the crystalline condition, that is, the condition

    gGK,xHiΔ(𝒳/𝔖):g(x)xφ1([ε¯]1)[π¯]HiΔ(𝒳Ainf/Ainf).\forall g\in G_{K},\;\;\forall x\in\mathrm{H}^{i}_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S}):\;\;g(x)-x\in\varphi^{-1}([\underline{\varepsilon}]-1)[\underline{\pi}]\mathrm{H}^{i}_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}).
  3. (3)

    For all pairs of integers s,ns,n with s+1n1s+1\geq n\geq 1, the pnp^{n}–torsion prismatic cohomology groups satisfy the condition

    gGK(π1/ps),xHiΔ,n(𝒳/𝔖):g(x)xφ1([ε¯]1)[π¯]ps+1nHiΔ,n(𝒳Ainf/Ainf).\forall g\in G_{K(\pi^{1/p^{s}})},\;\;\forall x\in\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/\mathfrak{S}):\;\;g(x)-x\in\varphi^{-1}([\underline{\varepsilon}]-1)[\underline{\pi}]^{p^{s+1-n}}\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}).

Theorem 1.3 (3) specialized to n=1n=1 provides the desired analogue of the property (1.1) of (φ,G^)(\varphi,\widehat{G})–modules and allows us to carry out the proof of Theorem 1.1.

As a consequence of Theorem 1.3 (2), we obtain a proof of crystallinity of the cohomology groups Hiét(𝒳η¯,Qp)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Q}_{p}) in the proper case partially by means of “formal” pp–adic Hodge theory (Corollary 4.6). This fact is usually established via a a direct comparison between crystalline and étale cohomology, and in this generality is originally due to Bhatt, Morrow and Scholze ([BMS1]). Of course, since our setup relies on the machinery of prismatic cohomology and especially the étale comparison, the proof can be considered independent of the one from [BMS1] only in that it avoids the crystalline comparison theorem for (prismatic) AinfA_{\mathrm{inf}}–cohomology.

The bounds of Theorem 1.1 compare to the already known bounds as follows. Whenever the bounds of “semistable type” are known to apply to the situation of Hiét(𝒳η¯,Z/pZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}) (e.g. [CarusoLiu] when i=1i=1, [Hattori] when ie<p1ie<p-1 and 𝒳\mathscr{X} is a scheme), the bounds from Theorem 1.1 agree with those bounds. The bounds tailored to crystalline representations ([Fontaine2, Abrashkin]) are slightly better but their applicability is quite limited (e=1e=1 and i<p1i<p-1).

The fact that the cohomology groups Hiét(𝒳η¯,Z/pnZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p^{n}\mathbb{Z}) have an associated Breuil–Kisin module yields one more source of ramification estimates: in [Caruso], Caruso provides a very general bound for pnp^{n}–torsion GKG_{K}–modules based on their restriction to GK(π1/p)G_{K(\pi^{1/p^{\infty}})} via Fontaine’s theory of étale 𝒪\mathcal{O}_{\mathcal{E}}–modules. Using the Breuil–Kisin module HiΔ,n(𝒳/𝔖)\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/\mathfrak{S}) attached to Hiét(𝒳η¯,Z/pnZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p^{n}\mathbb{Z}), this bound becomes explicit (as discussed in more detail in Remark 5.6). Comparing this result to Theorem 1.1 is more ambiguous due to somewhat different shapes of the estimates, but roughly speaking, the estimate of Theorem 1.1 is approximately the same for epe\leq p, becomes worse when KK is absolutely tamely ramified with large ramification degree, and is expected to outperform Caruso’s bound in case of large wild absolute ramification (rel. to the tame part of the ramification).

In future work, we intend to extend the result of Theorem 1.1 to the case of arbitrary nn. This seems plausible thanks to the full statement of Theorem 1.3 (3). In a different direction, we plan to extend the results of the present paper to the case of semistable reduction, using the log–prismatic cohomology developed by Koshikawa in [Koshikawa]. An important facts in this regard are that the AinfA_{\mathrm{inf}}–log–prismatic cohomology groups are still Breuil–Kisin–Fargues GKG_{K}–modules by a result of Česnavičius and Koshikawa ([CesnaviciusKoshikawa]) and that by results of Gao, a variant of the condition (Cr0\mathrm{Cr}_{0}) might exist in the semistable case ([GaoBKGK]; see Remark 3.12 (3) below for details).

The outline of the paper is as follows. In §2 we establish some necessary technical results. Namely, we discuss non–zero divisors and regular sequences on derived complete and completely flat modules with respect to the weak topology of AinfA_{\mathrm{inf}}, and establish Čech–Alexander complexes in the case of a separated and smooth formal scheme. Next, §3 introduces the conditions (Crs\mathrm{Cr}_{s}), studies their basic algebraic properties and discusses in particular the crystalline condition (Cr0\mathrm{Cr}_{0}) in the case of Breuil–Kisin–Fargues GKG_{K}–modules. In §4 we prove the conditions (Crs\mathrm{Cr}_{s}) for the Alexander–Čech complexes of a separated smooth pp–adic scheme 𝒳\mathscr{X} over 𝔖\mathfrak{S} and AinfA_{\mathrm{inf}}, and draw some consequences for the inidividual cohomology groups (especially when 𝒳\mathscr{X} is proper), proving Theorem 1.3. Finally, in §5 we establish the ramification bounds for mod pp étale cohomology, proving Theorem 1.1. Subsequently, we discuss in detail how the bounds from Theorem 1.1 compare to the various known bounds from the literature.

Let us set up some basic notation used throughout the paper. We fix a perfect field kk of characteristic p>0p>0 and a finite totally ramified extension K/KK/K^{\prime} of degree ee where K=W(k)[1/p]K^{\prime}=W(k)[1/p]. We fix a uniformizer π𝒪K\pi\in\mathcal{O}_{K}, and a compatible system (πn)n(\pi_{n})_{n} of pnp^{n}–th roots of π\pi in CK\mathbb{C}_{K}, the completion of algebraic closure of KK. Setting 𝔖=W(k)[[u]]\mathfrak{S}=W(k)[[u]], the choice of π\pi determines a surjective map 𝔖𝒪K\mathfrak{S}\rightarrow\mathcal{O}_{K} via uπu\mapsto\pi; the kernel of this map is generated by an Eisenstein polynomial E(u)E(u) of degree ee. 𝔖\mathfrak{S} is endowed with a Frobenius lift (hence a δ\delta–structure) extending the one on W(k)W(k) by uupu\mapsto u^{p}.

Denote Ainf=Ainf(𝒪CK)=W(𝒪CK)A_{\mathrm{inf}}=\mathbb{A}_{\mathrm{inf}}(\mathcal{O}_{\mathbb{C}_{K}})=W(\mathcal{O}_{\mathbb{C}_{K}^{\flat}}) where W()W(-) denotes the Witt vectors construction and 𝒪CK=𝒪CK\mathcal{O}_{\mathbb{C}_{K}^{\flat}}=\mathcal{O}_{\mathbb{C}_{K}}^{\flat} is the tilt of 𝒪CK\mathcal{O}_{\mathbb{C}_{K}}, 𝒪CK=limxxp𝒪CK/p\mathcal{O}_{\mathbb{C}_{K}}^{\flat}=\varprojlim_{x\mapsto x^{p}}\mathcal{O}_{\mathbb{C}_{K}}/p. The choice of the system (πn)n(\pi_{n})_{n} describes an element π¯𝒪CKlimxxp𝒪CK\underline{\pi}\in\mathcal{O}_{\mathbb{C}_{K}^{\flat}}\simeq\varprojlim_{x\mapsto x^{p}}\mathcal{O}_{\mathbb{C}_{K}}, and hence an embedding of 𝔖\mathfrak{S} into AinfA_{\mathrm{inf}} via u[π¯]u\mapsto[\underline{\pi}] where [][-] denotes the Teichmüller lift. Under this embedding, E(u)E(u) is sent to a generator ξ\xi of the kernel of the canonical map θ:Ainf𝒪CK\theta:A_{\mathrm{inf}}\rightarrow\mathcal{O}_{\mathbb{C}_{K}} that lifts the canonical projection pr0:𝒪CK=limφ𝒪CK/p𝒪CK/p.\mathrm{pr}_{0}:\mathcal{O}_{\mathbb{C}_{K}}^{\flat}=\varprojlim_{\varphi}\mathcal{O}_{\mathbb{C}_{K}}/p\rightarrow\mathcal{O}_{\mathbb{C}_{K}}/p. Consequently, (𝔖,(E(u)))(Ainf,Kerθ)(\mathfrak{S},(E(u)))\rightarrow(A_{\mathrm{inf}},\mathrm{Ker}\,\theta) is a map of prisms. It is known that under such embedding, AinfA_{\mathrm{inf}} is faithfully flat over 𝔖\mathfrak{S} (see e.g. [EmertonGee2, Proposition 2.2.13]).

Similarly, we fix a choice of a compatible system of primitive pnp^{n}–th roots of unity (ζpn)n0(\zeta_{p^{n}})_{n\geq 0}. This defines an element ε¯\underline{\varepsilon} of 𝒪CK\mathcal{O}_{\mathbb{C}_{K}^{\flat}} in an analogous manner, and the embedding 𝔖Ainf\mathfrak{S}\hookrightarrow A_{\mathrm{inf}} extends to a map (actually still an embedding by [Caruso, Proposition 1.14]) W(k)[[u,v]]AinfW(k)[[u,v]]\rightarrow A_{\mathrm{inf}} by additionally setting v[ε¯]1v\mapsto[\underline{\varepsilon}]-1. Additionally, we denote by ω\omega the element ([ε¯]1)/([ε¯1/p]1)=[ε¯1/p]p1++[ε¯1/p]+1([\underline{\varepsilon}]-1)/([\underline{\varepsilon}^{1/p}]-1)=[\underline{\varepsilon}^{1/p}]^{p-1}+\dots+[\underline{\varepsilon}^{1/p}]+1. It is well–known that this is another generator of Kerθ\mathrm{Ker}\,\theta, therefore ω/ξ\omega/\xi is a unit in AinfA_{\mathrm{inf}}.

The choices of π,πn\pi,\pi_{n} and ζpn\zeta_{p^{n}}, hence also the maps 𝔖Ainf\mathfrak{S}\hookrightarrow A_{\mathrm{inf}} and W(k)[[u,v]]AinfW(k)[[u,v]]\hookrightarrow A_{\mathrm{inf}}, remain fixed throughout. For this reason, we often refer to [π¯][\underline{\pi}] as uu, [ε¯]1[\underline{\varepsilon}]-1 as vv, ξ\xi as E(u),E(u), etc.

Throughout the paper, we use freely the language of prisms and δ\delta–rings from [BhattScholze], and we adopt much of the related notation and conventions. In particular, a formal scheme 𝒳\mathscr{X} over a pp–adically complete ring AA always means a pp–adic formal scheme, and it is called smooth if it is locally of the form SpfR\mathrm{Spf}\,R for a (derived333As we will always consider the base AA to have bounded pp^{\infty}–torsion, there is no distinction between derived pp–completion and pp–adic completion in this case.) pp–completely smooth AA–algebra RR – that is, a pp–complete AA–algebra such that R/pR/p is a smooth A/pA/p–algebra and ToriA(R,A/p)=0\mathrm{Tor}_{i}^{A}(R,A/p)=0 for all i>0i>0. By the results of Elkik [Elkik] and the discussion in [BhattScholze, §1.2], RR is equivalently the pp–adic completion of a smooth AA-algebra.

Acknowledgements. I would like to express my gratitude to my Ph.D. advisor Tong Liu for suggesting the topic of this paper, his constant encouragement and many comments, suggestions and valuable insights. Many thanks go to Deepam Patel and Shuddhodan Kadattur Vasudevan for organizing the prismatic cohomology learning seminar at Purdue University in Fall 2019, and to Donu Arapura for a useful discussion of Čech theory. I would like to thank Xavier Caruso, Shin Hattori and Shizhang Li for reading an earlier version of this paper, and for providing me with useful comments and questions. The present paper is an adapted version of the author’s Ph.D. thesis at Purdue University. During the preparation of the paper, the author was partially supported by the Ross Fellowship and the Bilsland Fellowship of Purdue University, as well as Graduate School Summer Research Grants of Purdue University during summers 2019–2021.

2.  Preparations

2.1.  Regularity on (p,E(u))(p,E(u))–completely flat modules

The goal of this section is to prove that every (p,E(u))(p,E(u))–complete and (p,E(u))(p,E(u))–completely flat AinfA_{\mathrm{inf}}–module is torsion–free, and that any sequence p,xp,x with xAinf(Ainf×pAinf)x\in A_{\mathrm{inf}}\setminus(A_{\mathrm{inf}}^{\times}\cup pA_{\mathrm{inf}}) is regular on such modules.

Regarding completions and complete flatness, we adopt the terminology of [stacks, 091N], [BhattScholze], but since we apply these notions mostly to modules as opposed to objects of derived categories, our treatment is closer in spirit to [Positselski], [Rezk] and [YekutieliFlatness]. Given a ring AA and a finitely generated ideal I=(f1,f2,fn)I=(f_{1},f_{2},\dots f_{n}), the derived II–completion444That is, this is derived II–completion of MM as a module. This will be sufficient to consider for our purposes. of an AA–module MM is

(2.2) M^=M[[X1,X2,Xn]]/(X1f1,X2f2,,Xnfn)M[[X1,X2,Xn]].\widehat{M}=M[[X_{1},X_{2},\dots X_{n}]]/(X_{1}-f_{1},X_{2}-f_{2},\dots,X_{n}-f_{n})M[[X_{1},X_{2},\dots X_{n}]].

MM is said to be derived II–complete if the natural map MM^M\rightarrow\widehat{M} is an isomorphism. This is equivalent to the vanishing of ExtiA(Af,M)\mathrm{Ext}^{i}_{A}(A_{f},M) for i=0,1i=0,1 and all fIf\in I (equivalently, for f=fjf=f_{j} for all jj), and as a consequence, it can be shown that the category of derived II–complete modules forms a full abelian subcategory of the category of all AA–modules with exact inclusion functor (and the derived II–completion is its left adjoint; in particular, derived II–completion is right exact as a functor on AA–modules). Another consequence is that derived II–completeness is equivalent to derived JJ–completeness when I,JI,J are two finitely generated ideals and I=J\sqrt{I}=\sqrt{J}. There is always a natural surjection M^M^cl\widehat{M}\rightarrow{\widehat{M}}^{\mathrm{cl}} where ()^cl\widehat{(-)}^{\mathrm{cl}} stands for II–adic completion, which will be reffered to as classical II–completion for the rest of the paper. Just like for classsicaly II–complete modules, if MM is derived II–complete, then M/IM=0M/IM=0 implies M=0M=0 (this is referred to as derived Nakayama lemma).

A convenient consequence of the completion formula (2.2) is that in the case when M=RM=R is a derived II–complete AA–algebra, the isomorphism RR[[X1,Xn]]/(X1f1,,,Xnfn)R\rightarrow R[[X_{1},\dots X_{n}]]/(X_{1}-f_{1},\dots,...,X_{n}-f_{n}) picks a preferred representative in RR for the power series symbol j1,,jnaj1,,jnf1j1fmjn\sum_{j_{1},\dots,j_{n}}a_{j_{1},\dots,j_{n}}f_{1}^{j_{1}}\dots f_{m}^{j_{n}} as the preimage of the class represented by j1,,jnaj1,,jnX1j1Xnjn\sum_{j_{1},\dots,j_{n}}a_{j_{1},\dots,j_{n}}X_{1}^{j_{1}}\dots X_{n}^{j_{n}}. This gives an algebraically well–behaved notion of power series summation despite the fact that RR is not necessarily II–adically separated555This operation further leads to the notion of contramodules, discussed e.g. in [Positselski]..

An AA–module MM is said to be II–completely (faithfully) flat if ToriA(M,A/I)=0\mathrm{Tor}_{i}^{A}(M,A/I)=0 for all i>0i>0 and M/IMM/IM is a (faithfully) flat A/IA/I–module. Just like for derived completeness, II–complete flatness is equivalent to JJ–complete flatness when JJ is another finitely generated ideal with I=J\sqrt{I}=\sqrt{J} 666However, note that while (derived) II–completeness more generally implies (derived) II^{\prime}–completeness when II^{\prime} is a finitely generated ideal contained in I\sqrt{I}, the “opposite” works for flatness, i.e. II–complete flatness implies II^{\prime\prime}–complete flatness when when II^{\prime\prime} is a finitely generated ideal with III\subseteq\sqrt{I^{\prime\prime}}..

Let us start by a brief discussion of regular sequences on derived complete modules in general. For that purpose, given an AA–module MM and f¯=f1,,fnA\underline{f}=f_{1},\dots,f_{n}\in A, denote by Kos(M;f¯)\mathrm{Kos}(M;\underline{f}) the usual Koszul complex and let Hm(M;f¯)H_{m}(M;\underline{f}) denote the mm-th Koszul homology of MM with respect to f1,f2,,fnf_{1},f_{2},\dots,f_{n}.

The first lemma is a straightforward generalization of standard facts about Koszul homology (e.g. [Matsumura, Theorem 16.5]) and regularity on finitely generated modules.

Lemma 2.1.

Let AA be a ring, IAI\subseteq A a finitely generated ideal and let MM be a nonzero derived II-complete module. Let f¯=f1,f2,,fnI\underline{f}=f_{1},f_{2},\dots,f_{n}\in I. Then

  1. (1)

    f¯\underline{f} forms a regular sequence on MM if and only if Hm(M;f¯)=0H_{m}(M;\underline{f})=0 for all m1m\geq 1 if and only if H1(M;f¯)=0H_{1}(M;\underline{f})=0.

  2. (2)

    In this situation, any permutation of f1,f2,,fnf_{1},f_{2},\dots,f_{n} is also a regular sequence on MM.

Proof.

As Koszul homology is insensitive to the order of the elements f1,f2,,fnf_{1},f_{2},\dots,f_{n}, part (2) follows immediately from (1).

To prove (1), the forward implications are standard and hold in full generality (see e.g. [Matsumura, Theorem 16.5]). It remains to prove that the sequence f1,f2,fnf_{1},f_{2},\dots f_{n} is regular on MM if H1(M;f1,f2,,fn)=0H_{1}(M;f_{1},f_{2},\dots,f_{n})=0. We proceed by induction on nn. The case n=1n=1 is clear (H1(M;x)=M[x]H_{1}(M;x)=M[x] by definition, and M/xM0M/xM\neq 0 follows by derived Nakayama). Let n2n\geq 2, and denote f¯\underline{f}^{\prime} the truncated sequence f1,f2,,fn1f_{1},f_{2},\dots,f_{n-1}. Then we have Kos(M;f¯)Kos(M;f¯)Kos(A;fn),\mathrm{Kos}(M;\underline{f})\simeq\mathrm{Kos}(M;\underline{f^{\prime}})\otimes\mathrm{Kos}(A;f_{n}), which produces a short exact sequence

0Kos(M;f¯)Kos(M;f¯)Kos(M;f¯)[1]00\longrightarrow\mathrm{Kos}(M;\underline{f^{\prime}})\longrightarrow\mathrm{Kos}(M;\underline{f})\longrightarrow\mathrm{Kos}(M;\underline{f^{\prime}})[-1]\longrightarrow 0

of chain complexes. Taking homologies results in a long exact sequence

H1(M;f¯)±fnH1(M;f¯)H1(M;f¯)M/(f¯)M±fnM/(f¯)MM/(f¯)M0\cdots\rightarrow H_{1}(M;\underline{f}^{\prime})\stackrel{{\scriptstyle\pm f_{n}}}{{\longrightarrow}}H_{1}(M;\underline{f}^{\prime})\longrightarrow H_{1}(M;\underline{f})\longrightarrow M/(\underline{f}^{\prime})M\stackrel{{\scriptstyle\pm f_{n}}}{{\rightarrow}}M/(\underline{f}^{\prime})M\longrightarrow M/(\underline{f})M\rightarrow 0

(as in [Matsumura, Theorem 7.4]). By assumption, H1(M;f¯)=0H_{1}(M;\underline{f})=0 and thus, fnH1(M;f¯)=H1(M;f¯)f_{n}H_{1}(M;\underline{f}^{\prime})=H_{1}(M;\underline{f}^{\prime}) where fnIf_{n}\in I. Upon observing that H1(M;f¯)H_{1}(M;\underline{f}^{\prime}) is obtained from finite direct sum of copies of MM by repeatedly taking kernels and cokernels, it is derived II–complete. Thus, derived Nakayama implies that H1(M;f¯)=0H_{1}(M;\underline{f}^{\prime})=0 as well, and by induction hypothesis, f¯\underline{f}^{\prime} is a regular sequence on MM. Finally, the above exact sequence also implies that fnf_{n} is injective on M/(f¯)M,M/(\underline{f}^{\prime})M, and M/(f¯)M0M/(\underline{f})M\neq 0 is satisfied thanks to derived Nakayama again. This finishes the proof. ∎

Corollary 2.2.

Let AA be a derived II–complete ring for an ideal I=(f¯)I=(\underline{f}) where f¯=f1,f2,,fn\underline{f}=f_{1},f_{2},\dots,f_{n} is a regular sequence on AA, and let FF be a nonzero derived II–complete AA–module that is II–completely flat. Then f¯\underline{f} is a regular sequence on FF and consequently, each fif_{i} is a non–zero divisor on FF.

Proof.

By Lemma 2.1 (1), Hm(A;f¯)=0H_{m}(A;\underline{f})=0 for all m1m\geq 1, hence Kos(A;f¯)\mathrm{Kos}(A;\underline{f}) is a free resolution of A/IA/I. Thus, on one hand, the complex FAKos(A;f¯)F\otimes_{A}\mathrm{Kos}(A;\underline{f}) computes TorA(F,A/I)\mathrm{Tor}^{A}_{*}(F,A/I), hence is acyclic in positive degrees by II–complete flatness; on the other hand, this complex is by definition Kos(F;f¯)\mathrm{Kos}(F;\underline{f}). We may thus conclude that Hi(F;f¯)=0H_{i}(F;\underline{f})=0 for all i1i\geq 1. By Lemma 2.1, f¯\underline{f} is a regular sequence on FF, and it remains regular on FF after arbitrary permutation. This proves the claim. ∎

Now we specialize to the case at hand, that is, A=AinfA=A_{\mathrm{inf}}. Recall that this is a domain and so is Ainf/p=𝒪CKA_{\mathrm{inf}}/p=\mathcal{O}_{\mathbb{C}_{K}^{\flat}} (which is a rank 11 valuation ring).

Lemma 2.3.

For every element xAinf(Ainf×pAinf),x\in A_{\mathrm{inf}}\setminus(A_{\mathrm{inf}}^{\times}\cup pA_{\mathrm{inf}}), p,xp,x forms a regular sequence, and for all k,l,k,l, we have the equality pkAinfxlAinf=pkxlAinf,p^{k}A_{\mathrm{inf}}\cap x^{l}A_{\mathrm{inf}}=p^{k}x^{l}A_{\mathrm{inf}},. Furthermore, the ideal (p,x)\sqrt{(p,x)} is equal to (p,W(𝔪CK))(p,W(\mathfrak{m}_{\mathbb{C}_{K}^{\flat}})), the unique maximal ideal of AinfA_{\mathrm{inf}}. In particular, given two choices x,xx,x^{\prime} as above, we have (p,x)=(p,x)\sqrt{(p,x)}=\sqrt{(p,x^{\prime})}.

In particular, the equalities “(p,x)=(p,x)\sqrt{(p,x)}=\sqrt{(p,x^{\prime})}” imply that all the (p,x)(p,x)–adic topologies (for xx as above) are equivalent to each other; this is the so–called weak topology on AinfA_{\mathrm{inf}} (usually defined as (p,u)(p,u)–adic topology in our notation), and it is standard that AinfA_{\mathrm{inf}} is complete with respect to this topology.

Proof.

By assumption, the image x¯\overline{x} of xx in Ainf/p=𝒪CKA_{\mathrm{inf}}/p=\mathcal{O}_{\mathbb{C}_{K}^{\flat}} is non–zero and non–unit in Ainf/pA_{\mathrm{inf}}/p (non–unit since xAinf×x\notin A_{\mathrm{inf}}^{\times} and prad(Ainf)p\in\mathrm{rad}(A_{\mathrm{inf}})). Thus, xlx^{l} is a non–zero divisor both on AinfA_{\mathrm{inf}} and on Ainf/pA_{\mathrm{inf}}/p, hence the claim that pAinfxlAinf=pxlAinfpA_{\mathrm{inf}}\cap x^{l}A_{\mathrm{inf}}=px^{l}A_{\mathrm{inf}} follows for every ll. The element pp is itself non–zero divisor on AinfA_{\mathrm{inf}} and thus, p,xp,x is a regular sequence.

To obtain pkAinfxlAinf=pkxlAinfp^{k}A_{\mathrm{inf}}\cap x^{l}A_{\mathrm{inf}}=p^{k}x^{l}A_{\mathrm{inf}} for general kk, one can e.g. use induction on kk using the fact that pp is a non-zero divisor on AinfA_{\mathrm{inf}} (or simply note that one may replace elements in regular sequences by arbitrary positive powers).

To prove the second assertion, note that (x¯)=𝔪CK\sqrt{(\overline{x})}=\mathfrak{m}_{\mathbb{C}_{K}^{\flat}} since Ainf/p=𝒪CKA_{\mathrm{inf}}/p=\mathcal{O}_{\mathbb{C}_{K}^{\flat}} is a rank 11 valuation ring. It follows that (p,W(𝔪CK))(p,W(\mathfrak{m}_{\mathbb{C}_{K}^{\flat}})) is the unique maximal ideal of AinfA_{\mathrm{inf}} above (p)(p), hence the unique maximal ideal since prad(Ainf)p\in\mathrm{rad}(A_{\mathrm{inf}}), and that (p,x)\sqrt{(p,x)} is equal to this ideal. ∎

We are ready to prove the claim mentioned at the beginning of the section.

Corollary 2.4.

Let FF be a derived (p,E(u))(p,E(u))–complete and (p,E(u))(p,E(u))–completely flat AinfA_{\mathrm{inf}}–module, and let xAinf(Ainf×pAinf)x\in A_{\mathrm{inf}}\setminus(A_{\mathrm{inf}}^{\times}\cup pA_{\mathrm{inf}}). Then p,xp,x is a regular sequence on FF. In particular, for each k,l>0k,l>0, we have pkFxlF=pkxlFp^{k}F\cap x^{l}F=p^{k}x^{l}F. Consequently, FF is a torsion–free AinfA_{\mathrm{inf}}–module.

Proof.

By Lemma 2.3, AinfA_{\mathrm{inf}} and FF are derived (p,x)(p,x)–complete and FF is (p,x)(p,x)–completely flat over AinfA_{\mathrm{inf}}, and p,xp,x is a regular sequence on AinfA_{\mathrm{inf}}. Corollary 2.2 then proves the claim about regular sequence. The sequence pk,xlp^{k},x^{l} is then also regular on FF, and the claim pkFxlF=pkxlFp^{k}F\cap x^{l}F=p^{k}x^{l}F follows. To prove the “consequently” part, let yy be a non–zero and non–unit element of AinfA_{\mathrm{inf}}. Since AinfA_{\mathrm{inf}} is classically pp–complete, we have npnAinf=0\bigcap_{n}p^{n}A_{\mathrm{inf}}=0, and so there exist nn such that y=pnxy=p^{n}x with xpAinfx\notin pA_{\mathrm{inf}}. If xx is a unit, then yy is a non–zero divisor on FF since so is pnp^{n}. Otherwise xAinf(Ainf×pAinf)x\in A_{\mathrm{inf}}\setminus(A_{\mathrm{inf}}^{\times}\cup pA_{\mathrm{inf}}), so p,xp,x is a regular sequence on FF, and so is x,px,p (e.g. by Lemma 2.1). In particular p,xp,x are both non–zero divisors on FF, and hence so is y=pnxy=p^{n}x. ∎

Finally, we record the following consequence on flatness of (p,E(u))(p,E(u))–completely flat modules modulo powers of pp that seems interesting on its own.

Corollary 2.5.

Let xAinf(Ainf×pAinf)x\in A_{\mathrm{inf}}\setminus(A_{\mathrm{inf}}^{\times}\cup pA_{\mathrm{inf}}), and let FF be a derived (p,x)(p,x)–complete and (p,x)(p,x)–completely (faithfully) flat AinfA_{\mathrm{inf}}–module. Then FF is derived pp–complete and pp–completely (faithfully) flat. In particular, F/pnFF/p^{n}F is a flat Ainf/pnA_{\mathrm{inf}}/p^{n}–module for every n>0n>0.

Proof.

The fact that FF is derived pp–complete is clear since it is derived (p,x)(p,x)–complete. We need to show that F/pFF/pF is a flat Ainf/pA_{\mathrm{inf}}/p–module and that ToriAinf(F,Ainf/p)=0\mathrm{Tor}_{i}^{A_{\mathrm{inf}}}(F,A_{\mathrm{inf}}/p)=0 for all i>0i>0. The second claim is a consequence of the fact that pp is a non–zero divisor on both AinfA_{\mathrm{inf}} and FF by Corollary 2.4. For the first claim, note that Ainf/p=𝒪CKA_{\mathrm{inf}}/p=\mathcal{O}_{\mathbb{C}_{K}^{\flat}} is a valuation ring and therefore it is enough to show that F/pFF/pF is a torsion–free 𝒪CK\mathcal{O}_{\mathbb{C}_{K}^{\flat}}–module. This follows again by Corollary 2.4.

For the ‘faithful’ version, note that both the statements that F/pFF/pF is faithfully flat over Ainf/pA_{\mathrm{inf}}/p and that F/(p,x)FF/(p,x)F is faithfully flat over Ainf/(p,x)A_{\mathrm{inf}}/(p,x) are now equivalent to the statement F/𝔪F0F/\mathfrak{m}F\neq 0 where 𝔪=(p,W(𝔪CK))\mathfrak{m}=(p,W(\mathfrak{m}_{\mathbb{C}_{K}^{\flat}})) is the unique maximal ideal of AinfA_{\mathrm{inf}}. ∎

2.2.  Čech–Alexander complex

Next, we discuss the construction of Čech–Alexander complexes for computing prismatic cohomology, introduced in [BhattScholze] in the affine case, in a global situation. Throughout this section, let (A,I)(A,I) be a fixed bounded base prism, and let 𝒳\mathscr{X} be a smooth separated pp–adic formal scheme over A/I.A/I. Recall that (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} denotes the site whose underlying category is the opposite of the category of bounded prisms (B,IB)(B,IB) over (A,I)(A,I) together with a map of formal schemes Spf(B/IB)𝒳\mathrm{Spf}(B/IB)\rightarrow\mathscr{X} over A/IA/I. Covers in (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} are given by the opposites of faithfully flat maps (B,IB)(C,IC)(B,IB)\rightarrow(C,IC) of prisms, meaning that CC is (p,I)(p,I)–completely flat over (B,IB)(B,IB). The prismatic cohomology 𝖱ΓΔ(𝒳,A)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X},A) is then defined as the sheaf cohomology 𝖱Γ((𝒳/A)Δ,𝒪)\mathsf{R}\Gamma((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}},\mathcal{O})(=𝖱Γ((,𝒪)=\mathsf{R}\Gamma((*,\mathcal{O}) where * is the terminal sheaf) for the sheaf 𝒪=𝒪Δ\mathcal{O}=\mathcal{O}_{{{\mathbbl{\Delta}}}} on (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} defined by (B,IB)B(B,IB)\mapsto B.

Additionally, let us denote by Δ{{\mathbbl{\Delta}}} the site of all bounded prisms, i.e the opposite of the category of all bounded prisms and their maps, with topology given by faithfully flat maps of prisms.

In order to discuss the Čech–Alexander complex in a non-affine situation, a slight modification of the topology on (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} is convenient. The following proposition motivates the change.

Proposition 2.6.

Let (A,I)(A,I) be a bounded prism.

  1. (1)

    Given a collection of maps of (bounded) prisms (A,I)(Bi,IBi),(A,I)\rightarrow(B_{i},IB_{i}), i=1,2,,n,i=1,2,\dots,n, the canonical map (A,I)(C,IC)=(iBi,IiBi)(A,I)\rightarrow(C,IC)=\left(\prod_{i}B_{i},I\prod_{i}B_{i}\right) is a map of (bounded) prisms.

  2. (2)

    (C,IC)(C,IC) is flat over (A,I)(A,I) if and only if each (Bi,IBi)(B_{i},IB_{i}) is flat over (A,I)(A,I). In that situation, (C,IC)(C,IC) is faithfully flat prism over (A,I)(A,I) if and only if the family of maps of formal spectra Spf(Bi/IBi)Spf(A/I)\mathrm{Spf}(B_{i}/IB_{i})\rightarrow\mathrm{Spf}(A/I) is jointly surjective.

  3. (3)

    Let fAf\in A be an element. Then (Af^,IAf^)(\widehat{A_{f}},I\widehat{A_{f}}), where ()^\widehat{(-)} stands for the derived (equivalently, classical) (p,I)(p,I)–completion, is a bounded prism777We do consider the zero ring with its zero ideal a prism, hence allow the possibility of Af^=0\widehat{A_{f}}=0, which occurs e.g. when f(p,I).f\in(p,I). Whether the zero ring satisfies Definition 3.2 of [BhattScholze] depends on whether the inclusion of the empty scheme to itself is considered an effective Cartier divisor; following the usual definitions pedantically, it indeed seems to be the case. Also some related claims, such as [BhattScholze, Lemma 3.7 (3)] or [BhattNotes, Lecture 5, Corollary 5.2], suggest that the zero ring is allowed as a prism., and the map (A,I)(Af^,IAf^)(A,I)\rightarrow(\widehat{A_{f}},I\widehat{A_{f}}) is a flat map of prisms.

  4. (4)

    Let f1,,fnAf_{1},\dots,f_{n}\in A be a collection of elements generating the unit ideal. Then the canonical map (A,I)(iAfi^,IiAfi^)(A,I)\rightarrow\left(\prod_{i}\widehat{A_{f_{i}}},I\prod_{i}\widehat{A_{f_{i}}}\right) is a faithfully flat map of (bounded) prisms.

Proof.

The proof of (1) is more or less formal. The ring C=iBiC=\prod_{i}B_{i} has a unique AAδ\delta–algebra structure since the forgetful functor from δ\delta–rings to rings preserves limits, and CC is as product of (p,I)(p,I)–complete rings (p,I)(p,I)–complete. Clearly IC=i(IBi)IC=\prod_{i}(IB_{i}) is an invertible ideal since each IBiIB_{i} is. In particular, C[I]=0C[I]=0, hence a prism by [BhattScholze, Lemma 3.5]. Assuming that all (Bi,IBi)(B_{i},IB_{i}) are bounded, from C/IC=iBi/IBiC/IC=\prod_{i}B_{i}/IB_{i} we have C/IC[p]=C/IC[pk]C/IC[p^{\infty}]=C/IC[p^{k}] for kk big enough so that Bi/IBi[p]=Bi/IBi[pk]B_{i}/IB_{i}[p^{\infty}]=B_{i}/IB_{i}[p^{k}] for all ii, showing that (C,IC)(C,IC) is bounded.

The ((p,I)(p,I)–complete) flatness part of (2) is clear. For the faithful flatness statement, note that C/(p,I)C=iBi/(p,I)BiC/(p,I)C=\prod_{i}B_{i}/(p,I)B_{i}, hence A/(p,I)C/(p,I)CA/(p,I)\rightarrow C/(p,I)C is faithfully flat if and only if the map of spectra iSpec(Bi/(p,I)Bi)=Spec(C/(p,I)C)Spec(A/(p,I))\coprod_{i}\mathrm{Spec}({B_{i}/(p,I)B_{i}})=\mathrm{Spec}({C/(p,I)C})\rightarrow\mathrm{Spec}({A/(p,I)}) is surjective.

Let us prove (3). Since Af^\widehat{A_{f}} has prad(Af^),p\in\mathrm{rad}(\widehat{A_{f}}), the equality φn(fk)=fkpn+p()\varphi^{n}(f^{k})=f^{kp^{n}}+p(\dots) shows that φn(fk)\varphi^{n}(f^{k}) for each n,k0n,k\geq 0 is a unit in Af^\widehat{A_{f}}. Consequently, as in [BhattScholze, Remark 2.16], Af^=S1A^\widehat{A_{f}}=\widehat{S^{-1}A} for S={φn(fk)|n,k0}S=\{\varphi^{n}(f^{k})\;|\;n,k\geq 0\}, and the latter has a unique δ\delta–structure extending that of AA by [BhattScholze, Lemmas 2.15 and 2.17]. In particular, Af^\widehat{A_{f}} is a (p,I)(p,I)–completely flat AAδ\delta–algebra, hence (Af^,IAf^)(\widehat{A_{f}},I\widehat{A_{f}}) is flat prism over (A,I)(A,I) by [BhattScholze, Lemma 3.7 (3)].

Part (4) follows formally from parts (1)–(3). ∎

Construction 2.7.

Denote by (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg} the site whose underlying category is (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}. The covers on (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg} are given by the opposites of finite families {(B,IB)(Ci,ICi)}i\{(B,IB)\rightarrow(C_{i},IC_{i})\}_{i} of flat maps of prisms such that the associated maps {Spf(Ci/ICi)Spf(B/IB)}\{\mathrm{Spf}(C_{i}/IC_{i})\rightarrow\mathrm{Spf}(B/IB)\} are jointly surjective. Let us call these “faithfully flat families” for short. The covers of the initial object \varnothing 888That is, \varnothing corresponds to the zero ring, which we consider to be a prism as per the previous footnote. are the empty cover and the identity. We similarly extend Δ{{\mathbbl{\Delta}}} to Δ{{\mathbbl{\Delta}}}^{\amalg}, that is, we proclaim the identity cover and the empty cover to be covers of \varnothing, and generally proclaim (finite) faithfully flat families to be covers.

Clearly isomorphisms as well as composition of covers are covers in both cases. To check that (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg} and Δ{{\mathbbl{\Delta}}}^{\amalg} are sites, it thus remains to check the base change axiom. This is trivial for situations involving ,\varnothing, so it remains to check that given a faithfully flat family {(B,IB)(Ci,ICi)}i\{(B,IB)\rightarrow(C_{i},IC_{i})\}_{i} and a map of prisms (B,IB)(D,ID)(B,IB)\rightarrow(D,ID), the fibre products999Here we mean fibre products in the variance of the site, i.e. “pushouts of prisms”. We use the symbol \boxtimes to denote this operation. (Ci,ICi)(B,IB)(D,ID)(C_{i},IC_{i})\boxtimes_{(B,IB)}(D,ID) in Δ{{\mathbbl{\Delta}}}^{\amalg} exist and the collection {(D,ID)(Ci,ICi)(B,IB)(D,ID)}i\{(D,ID)\rightarrow(C_{i},IC_{i})\boxtimes_{(B,IB)}(D,ID)\}_{i} is a faithfully flat family; the existence and (p,I)(p,I)–complete flatness follows by the same proof as in [BhattScholze, Corollary 3.12], only with “(p,I)(p,I)–completely faithfully flat” replaced by “(p,I)(p,I)–completely flat” throughout, and the fact that the family is faithfully flat follows as well, since (i(Ci,ICi))(B,IB)(D,ID)=i((Ci,ICi)(B,IB)(D,ID))\left(\prod_{i}(C_{i},IC_{i})\right)\boxtimes_{(B,IB)}(D,ID)=\prod_{i}\left((C_{i},IC_{i})\boxtimes_{(B,IB)}(D,ID)\right) (and using Remark 2.8 (1) below).

Remark 2.8.
  1. (1)

    Note that for a finite family of objects (Ci,ICi)(C_{i},IC_{i}) in (𝒳/A)Δ,(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}, the structure map of the product (A,I)i(Ci,ICi)(A,I)\rightarrow\prod_{i}(C_{i},IC_{i}) together with the map of formal spectra (induced from the maps for individual ii’s)

    Spf(iCi/ICi)=iSpf(Ci/ICi)𝒳\mathrm{Spf}(\prod_{i}C_{i}/IC_{i})=\coprod_{i}\mathrm{Spf}(C_{i}/IC_{i})\rightarrow\mathscr{X}

    makes (iCi,IiCi)(\prod_{i}C_{i},I\prod_{i}C_{i}) into an object of (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} that is easily seen to be the coproduct of (Ci,ICi)(C_{i},IC_{i})’s. In view of Proposition 2.6 (2), one thus arrives at the equivalent formulation

    {YiZ}i is a (𝒳/A)Δ–cover iYiZ is a (𝒳/A)Δ–cover.\{Y_{i}\rightarrow Z\}_{i}\text{ is a }(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}\text{--cover }\Leftrightarrow\coprod_{i}Y_{i}\rightarrow Z\text{ is a }(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}\text{--cover.}

    That is, (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg} is the (finitely) superextensive site having covers of (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} as singleton covers. (Similar considerations apply to Δ{{\mathbbl{\Delta}}} and Δ{{\mathbbl{\Delta}}}^{\amalg}.)

  2. (2)

    The two sites are honestly different in that they define different categories of sheaves. Namely, for every finite coproduct Y=iYiY=\coprod_{i}Y_{i}, the collection of canonical maps {YiiYi}i\{Y_{i}\rightarrow\coprod_{i}Y_{i}\}_{i} forms a (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}–cover, and the sheaf axiom forces upon Shv((𝒳/A)Δ)\mathcal{F}\in\mathrm{Shv}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}) the identity (iYi)=i(Yi),\mathcal{F}\left(\coprod_{i}Y_{i}\right)=\prod_{i}\mathcal{F}(Y_{i}), which is not automatic101010For example, every constant presheaf is a sheaf for a topology given by singleton covers only, which is not the case for (𝒳/A)Δ.(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}.. In fact, Shv((𝒳/A)Δ)\mathrm{Shv}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}) can be identified with the full category of Shv((𝒳/A)Δ)\mathrm{Shv}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}) consisting of all sheaves compatible with finite disjoint unions in the sense above. In particular, the structure sheaf 𝒪=𝒪Δ:(B,IB)B\mathcal{O}=\mathcal{O}_{{{\mathbbl{\Delta}}}}:(B,IB)\mapsto B is a sheaf for the (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}–topology. (Again, the same is true for Δ{{\mathbbl{\Delta}}} and Δ{{\mathbbl{\Delta}}}^{\amalg}, including the fact that 𝒪:(B,IB)B\mathcal{O}:(B,IB)\mapsto B is a sheaf.)

Despite the above fine distinction, for the purposes of prismatic cohomology, the two topologies are interchangeable. This is a consequence of the following lemma.

Lemma 2.9.

Given an object (B,IB)(𝒳/A)Δ,(B,IB)\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}, one has Hi((B,IB),𝒪)=0\mathrm{H}^{i}((B,IB),\mathcal{O})=0 for i>0i>0.

Proof.

The sheaf 𝒪:(B,I)B\mathcal{O}:(B,I)\mapsto B on Δ{{\mathbbl{\Delta}}}^{\amalg} has vanishing positive Čech cohomology essentially by the proof of [BhattScholze, Corollary 3.12]: one needs to show acyclicity of the Čech complex for any Δ{{\mathbbl{\Delta}}}^{\amalg}–cover {(B,I)(Ci,ICi)}i,\{(B,I)\rightarrow(C_{i},IC_{i})\}_{i}, but the resulting Čech complex is identical to that for the Δ{{\mathbbl{\Delta}}}–cover (B,I)i(Ci,ICi)(B,I)\rightarrow\prod_{i}(C_{i},IC_{i}), for which the acyclicity is proved in [BhattScholze, Corollary 3.12]. By a general result (e.g. [stacks, 03F9]), this implies vanishing of HiΔ((B,I),𝒪)\mathrm{H}^{i}_{{{\mathbbl{\Delta}}}^{\amalg}}((B,I),\mathcal{O}) for all bounded prisms (B,I)(B,I) and all i>0i>0.

Now we make use of the fact that cohomology of an object can be computed as the cohomology of the corresponding slice site, [stacks, 03F3]. Let (B,IB)(𝒳/A)Δ.(B,IB)\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}. After forgetting structure, we may view (B,IB)(B,IB) as an object of Δ{{\mathbbl{\Delta}}}^{\amalg} as well, and then [stacks, 03F3] implies that for every i,i, we have the isomorphisms

Hi(𝒳/A)Δ((B,IB),𝒪)\displaystyle\mathrm{H}^{i}_{(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}}((B,IB),\mathcal{O}) Hi((𝒳/A)Δ/(B,IB),𝒪|(B,IB)),\displaystyle\simeq\mathrm{H}^{i}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}/(B,IB),\mathcal{O}|_{(B,IB)}),
HiΔ((B,IB),𝒪)\displaystyle\mathrm{H}^{i}_{{{\mathbbl{\Delta}}}^{\amalg}}((B,IB),\mathcal{O}) Hi((Δ/(B,IB),𝒪|(B,IB))\displaystyle\simeq\mathrm{H}^{i}(({{\mathbbl{\Delta}}}^{\amalg}/(B,IB),\mathcal{O}|_{(B,IB)})

(where 𝒞/c\mathcal{C}/c for a site 𝒞\mathcal{C} and c𝒞c\in\mathcal{C} denotes the slice site). Upon noting that the slice sites (𝒳/A)Δ/(B,IB),(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}/(B,IB), Δ/(B,IB){{\mathbbl{\Delta}}}^{\amalg}/(B,IB) are equivalent sites (in a manner that identifies the two versions of the sheaf 𝒪|(B,IB)\mathcal{O}|_{(B,IB)}), the claim follows. ∎

Corollary 2.10.

One has

𝖱Γ((𝒳/A)Δ,𝒪)=𝖱Γ((𝒳/A)Δ,𝒪).\mathsf{R}\Gamma((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}},\mathcal{O})=\mathsf{R}\Gamma((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg},\mathcal{O}).
Proof.

The coverings of (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg} contain the coverings of (𝒳/A)Δ,(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}, so we are in the situation of [stacks, 0EWK], namely, there is a morphism of sites ε:(𝒳/A)Δ(𝒳/A)Δ\varepsilon:(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}\rightarrow(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} given by the identity functor of the underlying categories, the pushforward functor ε:Shv((𝒳/A)Δ)Shv((𝒳/A)Δ)\varepsilon_{*}:\mathrm{Shv}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg})\rightarrow\mathrm{Shv}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}) being the natural inclusion and the (exact) inverse image functor ε1:Shv((𝒳/A)Δ)Shv((𝒳/A)Δ)\varepsilon^{-1}:\mathrm{Shv}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}})\rightarrow\mathrm{Shv}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}) is the sheafification with respect to the “{}^{\amalg}”-topology. One has

Γ((𝒳/A),)=Γ((𝒳/A),)ε\Gamma((\mathscr{X}/A)^{\amalg},-)=\Gamma((\mathscr{X}/A),-)\circ\varepsilon_{*}

(where ε\varepsilon_{*} denotes the inclusion of abelian sheaves in this context), hence

𝖱Γ((𝒳/A),𝒪)=𝖱Γ((𝒳/A),𝖱ε𝒪),\mathsf{R}\Gamma((\mathscr{X}/A)^{\amalg},\mathcal{O})=\mathsf{R}\Gamma((\mathscr{X}/A),\mathsf{R}\varepsilon_{*}\mathcal{O}),

and to conclude it is enough to show that 𝖱iε𝒪=0\mathsf{R}^{i}\varepsilon_{*}\mathcal{O}=0 i>0\forall i>0. But 𝖱iε𝒪\mathsf{R}^{i}\varepsilon_{*}\mathcal{O} is the sheafification of the presheaf given by (B,IB)Hi((B,IB),𝒪)(B,IB)\mapsto\mathrm{H}^{i}((B,IB),\mathcal{O}) ([stacks, 072W]), which is 0 by Lemma 2.9. Thus, 𝖱iε𝒪=0\mathsf{R}^{i}\varepsilon_{*}\mathcal{O}=0, which proves the claim. ∎

For an open pp–adic formal subscheme 𝒱𝒳\mathscr{V}\subseteq\mathscr{X}, denote by h𝒱h_{\mathscr{V}} the functor sending (B,IB)(𝒳/A)Δ(B,IB)\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} to the set of factorizations of the implicit map Spf(B/IB)𝒳\mathrm{Spf}(B/IB)\rightarrow\mathscr{X} through 𝒱𝒳;\mathscr{V}\hookrightarrow\mathscr{X}; that is,

h𝒱((B,IB))={ if the image of Spf(B/IB)𝒳 is contained in 𝒱, otherwise.h_{\mathscr{V}}((B,IB))=\begin{cases}*\;\;\text{ if the image of }\mathrm{Spf}(B/IB)\rightarrow\mathscr{X}\text{ is contained in }\mathscr{V},\\ \emptyset\;\;\text{ otherwise.}\end{cases}

Let (B,IB)(C,IC)(B,IB)\rightarrow(C,IC) correspond to a morphism in (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}. If Spf(B/IB)𝒳\mathrm{Spf}(B/IB)\rightarrow\mathscr{X} factors through 𝒱,\mathscr{V}, then so does Spf(C/IC)Spf(B/IB)𝒳\mathrm{Spf}(C/IC)\rightarrow\mathrm{Spf}(B/IB)\rightarrow\mathscr{X}. It follows that h𝒱h_{\mathscr{V}} forms a presheaf on (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} (with transition maps h𝒱((B,IB))h𝒱((C,IC))h_{\mathscr{V}}((B,IB))\rightarrow h_{\mathscr{V}}((C,IC)) given by *\mapsto* when h𝒱((B,IB))h_{\mathscr{V}}((B,IB))\neq\emptyset, and the empty map otherwise). Note that h𝒳h_{\mathscr{X}} is the terminal sheaf.

Proposition 2.11.

h𝒱h_{\mathscr{V}} is a sheaf on (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}.

Proof.

Consider a cover in (𝒳/A)Δ,(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}, which is given by a faithully flat family {(B,IB)(Ci,ICi)}i\{(B,IB)\rightarrow(C_{i},IC_{i})\}_{i}. One needs to check that the sequence

h𝒱((B,IB)))ih𝒱((Ci,ICi))i,jh𝒱((Ci,ICi)(B,IB)(Cj,ICj))h_{\mathscr{V}}((B,IB)))\rightarrow\prod_{i}h_{\mathscr{V}}((C_{i},IC_{i}))\rightrightarrows\prod_{i,j}h_{\mathscr{V}}((C_{i},IC_{i})\boxtimes_{(B,IB)}(C_{j},IC_{j}))

is an equalizer sequence. All the terms have at most one element; consequently, there are just two cases to consider, depending on whether the middle term is empty or not. In both cases, the pair of maps on the right necessarily agree, and so one needs to see that the map on the left is an isomorphism. This is clear in the case when the middle term is empty (since the only map into an empty set is an isomorphism). It remains to consider the case when the middle term is nonempty, which means that h𝒱((Ci,ICi))=h_{\mathscr{V}}((C_{i},IC_{i}))=* for all ii. In this case we need to show that h𝒱((B,IB))=h_{\mathscr{V}}((B,IB))=*. Since the maps Spf(Ci/ICi)Spf(B/IB)\mathrm{Spf}(C_{i}/IC_{i})\rightarrow\mathrm{Spf}(B/IB) are jointly surjective and each Spf(Ci/ICi)𝒳\mathrm{Spf}(C_{i}/IC_{i})\rightarrow\mathscr{X} lands in 𝒱\mathscr{V}, it follows that so does the map Spf(B/IB)𝒳\mathrm{Spf}(B/IB)\rightarrow\mathscr{X}. Thus, h𝒱((B,IB))=h_{\mathscr{V}}((B,IB))=*, which finishes the proof. ∎

When 𝒱\mathscr{V} is affine, one can cover the sheaf h𝒱h_{\mathscr{V}} by a representable sheaf. Note that the construction of the representing object is essentially equivalent to Construction 4.17 of [BhattScholze].

Construction 2.12 (Čech–Alexander cover of 𝒱\mathscr{V}).

Let us additionally assume that 𝒱=Spf(R)\mathscr{V}=\mathrm{Spf}(R) is affine. Choose a surjection P𝒱RP_{\mathscr{V}}\rightarrow R where P𝒱=A[X¯]^P_{\mathscr{V}}=\widehat{A[\underline{X}]} is a (p,I)(p,I)–completed free AA–algebra. Denote by J𝒱J_{\mathscr{V}} the kernel of the surjection. Then there is a commutative diagram with exact rows

0{0}J𝒱{J_{\mathscr{V}}}P𝒱{P_{\mathscr{V}}}R{R}0{0}J𝒱P𝒱δ^{\widehat{J_{\mathscr{V}}P_{\mathscr{V}}^{\delta}}}P𝒱δ^{\widehat{P_{\mathscr{V}}^{\delta}}}RP𝒱P𝒱δ^{\widehat{R\otimes_{P_{\mathscr{V}}}P_{\mathscr{V}}^{\delta}}}0,{0,}

where ()^\widehat{(-)} stands for derived (p,I)(p,I)–completion. Here for an AA–algebra SS, SδS^{\delta} denotes the “δ\delta–envelope” of SS, that is, the SS–algebra initial among SS–algebras endowed with an AAδ\delta–algebra structure. Note that P𝒱δ^=(P0𝒱)δ^\widehat{P_{\mathscr{V}}^{\delta}}=\widehat{(P^{0}_{\mathscr{V}})^{\delta}}, where P0𝒱=A[X¯]P^{0}_{\mathscr{V}}=A[\underline{X}] is the polynomial algebra before completion; in particular, since (P0𝒱)δ(P^{0}_{\mathscr{V}})^{\delta} is a flat P0𝒱P^{0}_{\mathscr{V}}–algebra (essentially by [BhattScholze, Lemma 2.11]), it follows that P𝒱δ^\widehat{P_{\mathscr{V}}^{\delta}} is (p,I)(p,I)–completely flat P𝒱P_{\mathscr{V}}–algebra. Consequently, the completions in the lower row of the diagram can be equivalently taken as classical (p,I)(p,I)-completions (cf. [BhattScholze, Lemma 3.7]).

Denote by J𝒱δ,P𝒱δ^J_{\mathscr{V}}^{\delta,\wedge}\subseteq\widehat{P_{\mathscr{V}}^{\delta}} the image of the map J𝒱P𝒱δ^P𝒱δ^,\widehat{J_{\mathscr{V}}P_{\mathscr{V}}^{\delta}}\rightarrow\widehat{P_{\mathscr{V}}^{\delta}}, i.e. the (p,I)(p,I)–complete ideal of P𝒱δ^\widehat{P_{\mathscr{V}}^{\delta}} topologically generated by J𝒱J_{\mathscr{V}}. Then we have a short exact sequence

0{0}J𝒱δ,{J_{\mathscr{V}}^{\delta,\wedge}}P𝒱δ^{\widehat{P_{\mathscr{V}}^{\delta}}}RP𝒱P𝒱δ^{\widehat{R\otimes_{P_{\mathscr{V}}}P_{\mathscr{V}}^{\delta}}}0.{0.}

Let (Cˇ𝒱,ICˇ𝒱)(\check{C}_{\mathscr{V}},I\check{C}_{\mathscr{V}}) be the prismatic envelope of (Pδ𝒱^,J𝒱δ,)(\widehat{P^{\delta}_{\mathscr{V}}},J_{\mathscr{V}}^{\delta,\wedge}). It follows from [BhattScholze, Proposition 3.13, Example 3.14] that (Cˇ𝒱,ICˇ𝒱)(\check{C}_{\mathscr{V}},I\check{C}_{\mathscr{V}}) exists and is given by a flat prism over (A,I)(A,I). The map

RRP𝒱Pδ𝒱^=Pδ𝒱^/J𝒱δ,Cˇ𝒱/ICˇ𝒱R\rightarrow\widehat{R\otimes_{P_{\mathscr{V}}}P^{\delta}_{\mathscr{V}}}=\widehat{P^{\delta}_{\mathscr{V}}}/J_{\mathscr{V}}^{\delta,\wedge}\rightarrow\check{C}_{\mathscr{V}}/I\check{C}_{\mathscr{V}}

of pp–complete rings corresponds to the map of formal schemes Spf(Cˇ𝒱/ICˇ𝒱)𝒱𝒳\mathrm{Spf}(\check{C}_{\mathscr{V}}/I\check{C}_{\mathscr{V}})\rightarrow\mathscr{V}\hookrightarrow\mathscr{X}. This defines an object of (𝒳/A)Δ,(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}, which we call a Čech–Alexander cover of 𝒱\mathscr{V}.

Remarks 2.13.


  1. (1)

    Note that (Cˇ𝒱,ICˇ𝒱)(\check{C}_{\mathscr{V}},I\check{C}_{\mathscr{V}}) is equivalently the prismatic envelope of (Pδ𝒱^,J𝒱Pδ𝒱^)(\widehat{P^{\delta}_{\mathscr{V}}},J_{\mathscr{V}}\widehat{P^{\delta}_{\mathscr{V}}}). Moreover, when the ideal J𝒱J_{\mathscr{V}} is finitely generated, one has the equality J𝒱δ,=J𝒱P𝒱δ^.J_{\mathscr{V}}^{\delta,\wedge}=J_{\mathscr{V}}\widehat{P_{\mathscr{V}}^{\delta}}.

  2. (2)

    Since the ring RR in Construction 2.12 is a pp-completely smooth A/IA/I–algebra, it is in particular a pp–completion of a finitely presented A/IA/I–algebra. It follows that the map P𝒱RP_{\mathscr{V}}\rightarrow R may be chosen so that P𝒱P_{\mathscr{V}} is the (derived) (p,I)(p,I)–completion of a polynomial AA–algebra of finite type, with the kernel J𝒱J_{\mathscr{V}} finitely generated. While such a choice may be preferable, we formulate the construction without imposing it, as it may be convenient to allow non–finite–type free algebras in the construction e.g. for the reasons of functoriality (see the remark at the end of [BhattScholze, Construction 4.17]).

Proposition 2.14.

Denote by hCˇ𝒱h_{\check{C}_{\mathscr{V}}} the sheaf represented by the object (Cˇ𝒱,ICˇ𝒱)(𝒳/A)Δ(\check{C}_{\mathscr{V}},I\check{C}_{\mathscr{V}})\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}. There exists a unique map of sheaves hCˇ𝒱h𝒱h_{\check{C}_{\mathscr{V}}}\rightarrow h_{\mathscr{V}}, and it is an epimorphism.

Proof.

If (B,IB)(𝒳/A)Δ(B,IB)\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} with hCˇ𝒱((B,IB)),h_{\check{C}_{\mathscr{V}}}((B,IB))\neq\emptyset, this means that Spf(B/IB)𝒳\mathrm{Spf}(B/IB)\rightarrow\mathscr{X} factors through 𝒱\mathscr{V} since it factors through Spf(Cˇ𝒱/ICˇ𝒱).\mathrm{Spf}(\check{C}_{\mathscr{V}}/I\check{C}_{\mathscr{V}}). Thus, we also have h𝒱((B,IB))=h_{\mathscr{V}}((B,IB))=*, and so the (necessarily unique) map hCˇ𝒱((B,IB))h𝒱((B,IB))h_{\check{C}_{\mathscr{V}}}((B,IB))\rightarrow h_{\mathscr{V}}((B,IB)) is defined. When hCˇ𝒱((B,IB))h_{\check{C}_{\mathscr{V}}}((B,IB)) is empty, the map hCˇ𝒱((B,IB))h𝒱((B,IB))h_{\check{C}_{\mathscr{V}}}((B,IB))\rightarrow h_{\mathscr{V}}((B,IB)) is still defined and unique, namely given by the empty map. Thus, the claimed morphism of sheaves exists and is unique.

We show that this map is an epimorphism. Let (B,IB)(𝒳/A)Δ(B,IB)\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} such that h𝒱((B,IB))=h_{\mathscr{V}}((B,IB))=*, i.e. Spf(B/IB)𝒳\mathrm{Spf}(B/IB)\rightarrow\mathscr{X} factors through 𝒱\mathscr{V}, and consider the map RB/IBR\rightarrow B/IB associated to the map Spf(B/IB)𝒱\mathrm{Spf}(B/IB)\rightarrow\mathscr{V}. Since P𝒱P_{\mathscr{V}} is a pp–completed free AA–algebra surjecting onto RR and BB is (p,I)(p,I)–complete, the map RB/IBR\rightarrow B/IB admits a lift P𝒱BP_{\mathscr{V}}\rightarrow B. This induces an AAδ\delta–algebra map P𝒱δ^B\widehat{P_{\mathscr{V}}^{\delta}}\rightarrow B which gives a morphism of δ\delta–pairs (P𝒱δ^,J𝒱P𝒱δ^)(B,IB)(\widehat{P_{\mathscr{V}}^{\delta}},J_{\mathscr{V}}\widehat{P_{\mathscr{V}}^{\delta}})\rightarrow(B,IB), and further the map of prisms (Cˇ𝒱,ICˇ𝒱)(B,IB)(\check{C}_{\mathscr{V}},I\check{C}_{\mathscr{V}})\rightarrow(B,IB) using the universal properties of objects involved. It is easy to see that this is indeed (the opposite of) a morphism in (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}. This shows that hCˇ𝒱((B,IB))h_{\check{C}_{\mathscr{V}}}((B,IB)) is nonempty whenever h𝒱((B,IB))h_{\mathscr{V}}((B,IB)) is. Thus, the map is an epimorphism. ∎

Let 𝔙={𝒱j}jJ\mathfrak{V}=\{\mathscr{V}_{j}\}_{j\in J} be an affine open cover of 𝒳\mathscr{X}. For n1n\geq 1 and a multi–index (j1,j2,,jn)Jn,(j_{1},j_{2},\dots,j_{n})\in J^{n}, denote by 𝒱j1,jn\mathscr{V}_{j_{1},\dots j_{n}} the intersection 𝒱j1𝒱jn\mathscr{V}_{j_{1}}\cap\dots\cap\mathscr{V}_{j_{n}}. As 𝒳\mathscr{X} is assumed to be separated, each 𝒱j1,jn\mathscr{V}_{j_{1},\dots j_{n}} is affine and we write 𝒱j1,jn=Spf(Rj1,,jn)\mathscr{V}_{j_{1},\dots j_{n}}=\mathrm{Spf}(R_{j_{1},\dots,j_{n}}).

Remark 2.15 (Binary products in (𝒳/A)Δ(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}).

For (B,IB),(C,IC)(𝒳/A)Δ(B,IB),(C,IC)\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}, let us denote their binary product by (B,IB)(C,IC)(B,IB)\boxtimes(C,IC). Let us describe it explicitly at least under the additional assumptions that

  1. (1)

    (B,IB),(C,IC)(B,IB),(C,IC) are flat prisms over (A,I),(A,I),

  2. (2)

    there are affine opens 𝒰,𝒱𝒳\mathscr{U},\mathscr{V}\subseteq\mathscr{X} such that h𝒰((B,IB))==h𝒱((C,IC))h_{\mathscr{U}}((B,IB))=*=h_{\mathscr{V}}((C,IC)).

Set 𝒲=𝒰𝒱\mathscr{W}=\mathscr{U}\cap\mathscr{V} and denote the rings corresponding to the affine open sets 𝒰,𝒱\mathscr{U},\mathscr{V} and 𝒲\mathscr{W} by R,SR,S and T,T, resp. Then any object (D,ID)(𝒳/A)Δ(D,ID)\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}} with maps both to (B,IB)(B,IB) and (C,IC)(C,IC) lives over 𝒲,\mathscr{W}, i.e. satisfies h𝒲((D,ID))=h_{\mathscr{W}}((D,ID))=*. This justifies the following construction. Consider the following commutative diagram, where \urcorner denotes the pushout of pp–complete commutative rings, i.e. taking the classically pp–completed tensor product ^\widehat{\otimes} (and B^ACB\widehat{\otimes}_{A}C is the derived, but equivalently classical, (p,I)(p,I)–completion of BACB\otimes_{A}C):

B^AC{B\widehat{\otimes}_{A}C}B{B}C{C}(B/IB^RT)^T(C/IC^ST){(B/IB\widehat{\otimes}_{R}T)\widehat{\otimes}_{T}(C/IC\widehat{\otimes}_{S}T)}B/IB{B/IB}B/IB^RT{B/IB\widehat{\otimes}_{R}T}A{{\color[rgb]{1,1,1}\definecolor[named]{pgfstrokecolor}{rgb}{1,1,1}\pgfsys@color@gray@stroke{1}\pgfsys@color@gray@fill{1}A}}C/IC^ST{C/IC\widehat{\otimes}_{S}T}C/IC{C/IC}R{R}T{T}T{T}S{S}

\llcorner

{\llcorner}{\lrcorner}

Let JB^ACJ\subseteq B\widehat{\otimes}_{A}C be the kernel of the map B^AC(B/IB^RT)^T(C/IC^ST).B\widehat{\otimes}_{A}C\rightarrow(B/IB\widehat{\otimes}_{R}T)\widehat{\otimes}_{T}(C/IC\widehat{\otimes}_{S}T). Then the product (B,IB)(C,IC)(B,IB)\boxtimes(C,IC) is given by the prismatic envelope of the δ\delta–pair (B^AC,J)(B\widehat{\otimes}_{A}C,J).

Proposition 2.16.

The Čech–Alexander covers can be chosen so that for all j1,,jnj_{1},\dots,j_{n} we have

(Cˇ𝒱j1,jn,ICˇ𝒱j1,jn)=(Cˇ𝒱j1,ICˇ𝒱j1)(Cˇ𝒱j2,ICˇ𝒱j2)(Cˇ𝒱jn,ICˇ𝒱jn).(\check{C}_{\mathscr{V}_{j_{1},\dots j_{n}}},I\check{C}_{\mathscr{V}_{j_{1},\dots j_{n}}})=(\check{C}_{\mathscr{V}_{j_{1}}},I\check{C}_{\mathscr{V}_{j_{1}}})\boxtimes(\check{C}_{\mathscr{V}_{j_{2}}},I\check{C}_{\mathscr{V}_{j_{2}}})\boxtimes\dots\boxtimes(\check{C}_{\mathscr{V}_{j_{n}}},I\check{C}_{\mathscr{V}_{j_{n}}}).
Proof.

Clearly it is enough to show the statement for binary products. More precisely, given two affine opens 𝒱1,𝒱2𝒳\mathscr{V}_{1},\mathscr{V}_{2}\subseteq\mathscr{X} and an arbitrary initial choice of (Cˇ𝒱1,ICˇ𝒱1)(\check{C}_{\mathscr{V}_{1}},I\check{C}_{\mathscr{V}_{1}}) and (Cˇ𝒱2,ICˇ𝒱2),(\check{C}_{\mathscr{V}_{2}},I\check{C}_{\mathscr{V}_{2}}), we show that P𝒱12R12P_{\mathscr{V}_{12}}\rightarrow R_{12} can be chosen so that the resulting Čech–Alexander cover (Cˇ𝒱12,ICˇ𝒱12)(\check{C}_{\mathscr{V}_{12}},I\check{C}_{\mathscr{V}_{12}}) of 𝒱12\mathscr{V}_{12} is equal to (Cˇ𝒱1,ICˇ𝒱1)(Cˇ𝒱2,ICˇ𝒱2)(\check{C}_{\mathscr{V}_{1}},I\check{C}_{\mathscr{V}_{1}})\boxtimes(\check{C}_{\mathscr{V}_{2}},I\check{C}_{\mathscr{V}_{2}}). For the purposes of this proof, let us refer to a prismatic envelope of a δ\delta–pair (S,J)(S,J) also as “the prismatic envelope of the arrow SS/JS\rightarrow S/J”.

Consider αi:P𝒱iRi,i=1,2\alpha_{i}:P_{\mathscr{V}_{i}}\twoheadrightarrow R_{i},\;i=1,2 as in Construction 2.12, and set P𝒱12=P𝒱1^AP𝒱2P_{\mathscr{V}_{12}}=P_{\mathscr{V}_{1}}\widehat{\otimes}_{A}P_{\mathscr{V}_{2}}. Then one has the induced surjection α1α2:P𝒱12R1^A/IR2\alpha_{1}\otimes\alpha_{2}:P_{\mathscr{V}_{12}}\rightarrow R_{1}\widehat{\otimes}_{A/I}R_{2}, which can be followed by the induced map R1^A/IR2R12R_{1}\widehat{\otimes}_{A/I}R_{2}\rightarrow R_{12}. This latter map is surjective as well since 𝒳\mathscr{X} is separated, and therefore the composition of these two maps α12:P𝒱12R12\alpha_{12}:P_{\mathscr{V}_{12}}\rightarrow R_{12} is surjective, with the kernel J𝒱12J_{\mathscr{V}_{12}} that contains (J𝒱1,J𝒱2)P𝒱12(J_{\mathscr{V}_{1}},J_{\mathscr{V}_{2}})P_{\mathscr{V}_{12}}. We may construct a diagram analogous to the one from Remark 2.15, which becomes the diagram

P𝒱12δ^{\widehat{P_{\mathscr{V}_{12}}^{\delta}}}P𝒱1δ^{\widehat{P_{\mathscr{V}_{1}}^{\delta}}}P𝒱2δ^{\widehat{P_{\mathscr{V}_{2}}^{\delta}}}R12^P𝒱12(P𝒱12δ^){R_{12}\widehat{\otimes}_{P_{\mathscr{V}_{12}}}(\widehat{P_{\mathscr{V}_{12}}^{\delta}})}R1^P𝒱1P𝒱1δ^{R_{1}\widehat{\otimes}_{P_{\mathscr{V}_{1}}}\widehat{P_{\mathscr{V}_{1}}^{\delta}}}R12^P𝒱1P𝒱1δ^{R_{12}\widehat{\otimes}_{P_{\mathscr{V}_{1}}}\widehat{P_{\mathscr{V}_{1}}^{\delta}}}A{{\color[rgb]{1,1,1}\definecolor[named]{pgfstrokecolor}{rgb}{1,1,1}\pgfsys@color@gray@stroke{1}\pgfsys@color@gray@fill{1}A}}R12^P𝒱2P𝒱2δ^{R_{12}\widehat{\otimes}_{P_{\mathscr{V}_{2}}}\widehat{P_{\mathscr{V}_{2}}^{\delta}}}R2^P𝒱2P𝒱2δ^{R_{2}\widehat{\otimes}_{P_{\mathscr{V}_{2}}}\widehat{P_{\mathscr{V}_{2}}^{\delta}}}R1{R_{1}}R12{R_{12}}R12{R_{12}}R2,{R_{2},}

\llcorner

{\llcorner}{\lrcorner}

where the expected arrow in the central column is replaced by an isomorphic one, namely the map obtained from the surjection P𝒱12R12P_{\mathscr{V}_{12}}\rightarrow R_{12} by the procedure as in Construction 2.12. Now (Cˇ𝒱12,ICˇ𝒱12)(\check{C}_{\mathscr{V}_{12}},I\check{C}_{\mathscr{V}_{12}}) is obtained as the prismatic envelope of this composed central arrow, while (Cˇ𝒱1,ICˇ𝒱1)(Cˇ𝒱2,ICˇ𝒱2)(\check{C}_{\mathscr{V}_{1}},I\check{C}_{\mathscr{V}_{1}})\boxtimes(\check{C}_{\mathscr{V}_{2}},I\check{C}_{\mathscr{V}_{2}}) is obtained the same way, but only after replacing the downward arrows on the left and right by their prismatic envelopes. Comparing universal properties, one easily sees that the resulting central prismatic envelope remains unchanged, proving the claim. ∎

Remark 2.17.

Suppose that for each jj, the initial choice of the map P𝒱jRjP_{\mathscr{V}_{j}}\rightarrow R_{j} has been made as in Remark 2.13 (2), that is, P𝒱jP_{\mathscr{V}_{j}} is the (p,I)(p,I)–completion of a finite type free AA–algebra and the ideal J𝒱jJ_{\mathscr{V}_{j}} is finitely generated. If now P𝒱j1,j2,jnP_{\mathscr{V}_{j_{1},j_{2},\dots j_{n}}} is the (p,I)(p,I)–completed free AA–algebra for 𝒱j1,j2,jn\mathscr{V}_{j_{1},j_{2},\dots j_{n}} obtained by iterating the procedure in the proof of Proposition 2.16, it is easy to see that in this case, the algebra P𝒱j1,j2,jnP_{\mathscr{V}_{j_{1},j_{2},\dots j_{n}}} is still the (p,I)(p,I)–completion of a finite type free AA–algebra, and it can be shown that the corresponding ideal J𝒱j1,j2,jnJ_{\mathscr{V}_{j_{1},j_{2},\dots j_{n}}} is finitely generated.

In more detail, given a ring BB and a finitely generated ideal JBJ\subseteq B, Let us call a BB–algebra CC JJ–completely finitely presented if CC is derived JJ–complete and there exists a map α:B[X¯]C\alpha:B[\underline{X}]\rightarrow C from the polynomial ring in finitely many variables X¯={X1,,Xn}\underline{X}=\{X_{1},\dots,X_{n}\} such that the derived JJ–completed map α^:B[X¯]^C\widehat{\alpha}:\widehat{B[\underline{X}]}\rightarrow C is surjective and with a finitely generated kernel. Then the algebra Rj1,j2,jnR_{j_{1},j_{2},\dots j_{n}} corresponding to 𝒱j1,j2,jn\mathscr{V}_{j_{1},j_{2},\dots j_{n}} is (p,I)(p,I)–completely finitely presented by Remark 2.13 (2), and since P𝒱j1,j2,jnP_{\mathscr{V}_{j_{1},j_{2},\dots j_{n}}} is the (p,I)(p,I)–completion of a finite type polynomial AA–algebra, the following lemma shows that J𝒱j1,j2,jnJ_{\mathscr{V}_{j_{1},j_{2},\dots j_{n}}} is finitely generated.

Lemma 2.18.

Let CC be a JJ–completely finitely presented BB–algebra, and consider a BB–algebra map β:B[Y¯]C\beta:B[\underline{Y}]\rightarrow C from a polynomial algebra in finitely many variables Y¯={Y1,,Ym}\underline{Y}=\{Y_{1},\dots,Y_{m}\} such that β^\widehat{\beta} is surjective. Then the kernel of β^\widehat{\beta} is finitely generated.

Proof.

The proof is an adaptation of the proof of [stacks, 00R2], which is a similar assertion about finitely presented algebras. Consider α\alpha as in Remark 2.17, and additionally let us fix a generating set (f1,f2,,fk)B[X¯]^(f_{1},f_{2},\dots,f_{k})\subseteq\widehat{B[\underline{X}]} of Kerα^\mathrm{Ker}\,\widehat{\alpha}.

For i=1,,m,i=1,\dots,m, let us choose giB[X¯]^g_{i}\in\widehat{B[\underline{X}]} such that α^(gi)=β(Yi)\widehat{\alpha}(g_{i})=\beta(Y_{i}). Then one can define a surjective map

θ0:B[X¯]^[Y¯]C,θ0B[X¯]^=α^,θ0(Yi)=β(Yi),\theta_{0}:\widehat{B[\underline{X}]}[\underline{Y}]\rightarrow C,\;\;\;\theta_{0}\mid_{\widehat{B[\underline{X}]}}=\widehat{\alpha},\;\;\theta_{0}(Y_{i})=\beta(Y_{i}),

and it is easy to see that Kerθ0=(f1,,fk,Y1g1,,Ymgm).\mathrm{Ker}\,\theta_{0}=(f_{1},\dots,f_{k},Y_{1}-g_{1},\dots,Y_{m}-g_{m}). That is, we have an exact sequence

(B[X¯]^[Y¯])k+mB[X¯]^[Y¯]θ0C0,(\widehat{B[\underline{X}]}[\underline{Y}])^{\oplus k+m}\rightarrow\widehat{B[\underline{X}]}[\underline{Y}]\stackrel{{\scriptstyle\theta_{0}}}{{\rightarrow}}C\rightarrow 0,

where the map on the left is a module map determined by the finite set of generators of Kerθ0\mathrm{Ker}\,\theta_{0}. After taking the derived JJ–completion, the sequence becomes the exact sequence

B[X¯,Y¯]^k+mB[X¯,Y¯]^θC0.\widehat{B[\underline{X},\underline{Y}]}^{\oplus k+m}\rightarrow\widehat{B[\underline{X},\underline{Y}]}\stackrel{{\scriptstyle\theta}}{{\rightarrow}}C\rightarrow 0.

That is, we have a surjective map θ:B[X¯,Y¯]^C\theta:\widehat{B[\underline{X},\underline{Y}]}\rightarrow C, which is determined on topological generators by θ(Xj)=α(Xj),θ(Yi)=β(Yi),\theta(X_{j})=\alpha(X_{j}),\theta(Y_{i})=\beta(Y_{i}), and the kernel of θ\theta is (f1,,fk,Y1g1,,Ymgm)(f_{1},\dots,f_{k},Y_{1}-g_{1},\dots,Y_{m}-g_{m}).

Next, we choose elements hjB[Y¯]^h_{j}\in\widehat{B[\underline{Y}]} such that β^(hj)=α(Xj)\widehat{\beta}(h_{j})=\alpha(X_{j}) for each jj. Then we have a surjective map ψ:B[X¯,Y¯]^B[Y¯]^\psi:\widehat{B[\underline{X},\underline{Y}]}\rightarrow\widehat{B[\underline{Y}]} given by XjhjX_{j}\mapsto h_{j} and YiYiY_{i}\mapsto Y_{i}, which has the property that β^ψ=θ\widehat{\beta}\circ\psi=\theta. That is,

Kerθ=Ker(β^ψ)=ψ1(Ker(β^)),\mathrm{Ker}\,\theta=\mathrm{Ker}\,(\widehat{\beta}\circ\psi)=\psi^{-1}(\mathrm{Ker}\,(\widehat{\beta})),

and therefore ψ(Kerθ)=Kerβ^\psi(\mathrm{Ker}\,\theta)=\mathrm{Ker}\,\widehat{\beta} since ψ\psi is surjective. But Kerθ\mathrm{Ker}\,\theta is finitely generated by the previous, and hence so is Kerβ^\mathrm{Ker}\,\widehat{\beta}.∎

Proposition 2.19.

The map jh𝒱jh𝒳=\coprod_{j}h_{\mathscr{V}_{j}}\rightarrow h_{\mathscr{X}}=* (where \coprod denotes the coproduct in Shv((𝒳/A)Δ)\mathrm{Shv}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg})) to the final object is an epimorphism, hence so is the map jhCˇ𝒱j\coprod_{j}h_{\check{C}_{\mathscr{V}_{j}}}\rightarrow*.

Proof.

It is enough to show that for a given object (B,IB)(𝒳/A)Δ,(B,IB)\in(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg}, there is a faithfully flat family (B,IB)(Ci,ICi)(B,IB)\rightarrow(C_{i},IC_{i}) in (𝒳/A)Δ,op(\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg,\mathrm{op}} such that prejh𝒱j((Ci,ICi))\coprod^{\mathrm{pre}}_{j}h_{\mathscr{V}_{j}}((C_{i},IC_{i}))\neq\emptyset for all ii where pre\coprod^{\mathrm{pre}} denotes the coproduct of presheaves.

With that aim, let us first consider the preimages 𝒲jSpf(B/IB)\mathscr{W}_{j}\subseteq\mathrm{Spf}(B/IB) of each 𝒱j\mathscr{V}_{j} under the map Spf(B/IB)𝒳\mathrm{Spf}(B/IB)\rightarrow\mathscr{X}. This is an open cover of Spf(B/IB)\mathrm{Spf}(B/IB) that corresponds to an open cover of SpecB/(p,I)B\mathrm{Spec}B/(p,I)B. One can then choose f1,f2,,fmf_{1},f_{2},\dots,f_{m} such that {Spec(B/(p,I)B)fi}i\{\mathrm{Spec}(B/(p,I)B)_{f_{i}}\}_{i} refines this cover, i.e. every Spec(B/(p,I)B)fi\mathrm{Spec}(B/(p,I)B)_{f_{i}} corresponds to an open subset of 𝒲j(i)\mathscr{W}_{j(i)} for some index j(i)j(i).

The elements f1,,fmf_{1},\dots,f_{m} generate the unit ideal of BB since they do so modulo (p,I)(p,I) which is contained in rad(B).\mathrm{rad}(B). Thus, the family

(B,IB)(Ci,ICi):=(Bfi^,IBfi^)i=1,2,,m(B,IB)\rightarrow(C_{i},IC_{i}):=(\widehat{B_{f_{i}}},I\widehat{B_{f_{i}}})\;\;i=1,2,\dots,m

is easily seen to give the desired faithfully flat family, with each prejh𝒱j((Ci,ICi))\coprod^{\mathrm{pre}}_{j}h_{\mathscr{V}_{j}}((C_{i},IC_{i})) nonempty, since each Spf(Ci/ICi)𝒳\mathrm{Spf}(C_{i}/IC_{i})\rightarrow\mathscr{X} factors through 𝒱j(i)\mathscr{V}_{j(i)} by construction. ∎

Remark 2.20.

The proof of Proposition 2.19 is the one step where we used the relaxation of the topology, namely the fact that the faithfully flat cover (B,IB)i(Ci,ICi)(B,IB)\rightarrow\prod_{i}(C_{i},IC_{i}) can be replaced by the family {(B,IB)(Ci,ICi)}i\{(B,IB)\rightarrow(C_{i},IC_{i})\}_{i}.

Finally, we obtain the Čech–Alexander complexes in the global case.

Proposition 2.21.

With the notation for 𝒱j1,j2,,jn\mathscr{V}_{j_{1},j_{2},\dots,j_{n}} as above and the choice of Čech–Alexander covers Cˇ𝒱j1,j2,,jn\check{C}_{\mathscr{V}_{j_{1},j_{2},\dots,j_{n}}} as in Proposition 2.16, 𝖱Γ((𝒳/A)Δ,𝒪)\mathsf{R}\Gamma((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}},\mathcal{O}) is modelled by the Čech–Alexander complex

(Cˇ𝔙\check{C}^{\bullet}_{\mathfrak{V}}) 0jCˇ𝒱jj1,j2Cˇ𝒱j1,j2j1,j2,j3Cˇ𝒱j1,j2,j30\longrightarrow\prod_{j}\check{C}_{\mathscr{V}_{j}}\longrightarrow\prod_{j_{1},j_{2}}\check{C}_{\mathscr{V}_{j_{1},j_{2}}}\longrightarrow\prod_{j_{1},j_{2},j_{3}}\check{C}_{\mathscr{V}_{j_{1},j_{2},j_{3}}}\longrightarrow\dots

(that is, the complex associated to the cosimplicial ring (j1,,jnCˇ𝒱j1,,jn)n(\prod_{j_{1},\dots,j_{n}}\check{C}_{\mathscr{V}_{j_{1},\dots,j_{n}}})_{n}).

Proof.

By [stacks, 079Z], the epimorphism of sheaves jhCˇ𝒱j\coprod_{j}h_{\check{C}_{\mathscr{V}_{j}}}\rightarrow* from Proposition 2.19 implies that there is a spectral sequence with E1E_{1}-page

E1p,q=Hq((jhCˇ𝒱j)×p,𝒪)=Hq(j1,j2,,jphCˇ𝒱j1,,jp,𝒪)=ji,,jpHq((Cˇ𝒱j1,,jp,ICˇ𝒱j1,,jp),𝒪)E_{1}^{p,q}=H^{q}\Big{(}\big{(}\coprod_{j}h_{\check{C}_{\mathscr{V}_{j}}}\big{)}^{\times p},\mathcal{O}\Big{)}=H^{q}\Big{(}\coprod_{j_{1},j_{2},\dots,j_{p}}h_{\check{C}_{\mathscr{V}_{j_{1},\dots,j_{p}}}},\mathcal{O}\Big{)}=\prod_{j_{i},\dots,j_{p}}H^{q}((\check{C}_{\mathscr{V}_{j_{1},\dots,j_{p}}},I\check{C}_{\mathscr{V}_{j_{1},\dots,j_{p}}}),\mathcal{O})

converging to Hp+q(,𝒪)=Hp+q((𝒳/A)Δ,𝒪)=Hp+q((𝒳/A)Δ,𝒪),H^{p+q}(*,\mathcal{O})=H^{p+q}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}}^{\amalg},\mathcal{O})=H^{p+q}((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}},\mathcal{O}), where we implicitly used Corollary 2.10 and the fact that hCˇ𝒱j1×hCˇ𝒱j2=hCˇ𝒱j1Cˇ𝒱j2=hCˇ𝒱j1,j2h_{\check{C}_{\mathscr{V}_{j_{1}}}}\times h_{\check{C}_{\mathscr{V}_{j_{2}}}}=h_{\check{C}_{\mathscr{V}_{j_{1}}}\boxtimes\check{C}_{\mathscr{V}_{j_{2}}}}=h_{\check{C}_{\mathscr{V}_{j_{1},j_{2}}}} as in Proposition 2.16, and similarly for higher multi–indices.

By Lemma 2.9, Hq((Cˇ𝒱j1,,jn,ICˇ𝒱j1,,jn),𝒪)=0H^{q}((\check{C}_{\mathscr{V}_{j_{1},\dots,j_{n}}},I\check{C}_{\mathscr{V}_{j_{1},\dots,j_{n}}}),\mathcal{O})=0 for every q>0q>0 and every multi–index j1,,jnj_{1},\dots,j_{n}. The first page is therefore concentrated in a single row of the form Cˇ𝔙\check{C}^{\bullet}_{\mathfrak{V}} and thus, the spectral sequence collapses on the second page. This proves that the cohomologies of 𝖱Γ((𝒳/A)Δ,𝒪)\mathsf{R}\Gamma((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}},\mathcal{O}) are computed as cohomologies of Cˇ𝔙\check{C}^{\bullet}_{\mathfrak{V}}, but in fact, this yields a quasi–isomorphism of the complexes themselves. (For example, analyzing the proof of [stacks, 079Z] via [stacks, 03OW], the double complex E0E_{0}^{\bullet\bullet} of the above spectral sequence comes with a natural map α:Cˇ𝔙Tot(E0),\alpha:\check{C}^{\bullet}_{\mathfrak{V}}\rightarrow\mathrm{Tot}(E_{0}^{\bullet\bullet}), and a natural quasi–isomorphism β:𝖱Γ((𝒳/A)Δ,𝒪)Tot(E0);\beta:\mathsf{R}\Gamma((\mathscr{X}/A)_{{{\mathbbl{\Delta}}}},\mathcal{O})\rightarrow\mathrm{Tot}(E_{0}^{\bullet\bullet}); when the spectral sequence collapses as above, α\alpha is also a quasi–isomorphism). ∎

Remarks 2.22.
  1. (1)

    Just as in the affine case, the formation of Čech–Alexander complexes is compatible with “termwise flat base–change” on the base prism essentially by [BhattScholze, Proposition 3.13]. That is, if (Cˇm,)m(\check{C}^{m},\partial)_{m} is a Čech–Alexander complex modelling 𝖱ΓΔ(𝒳/A)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/A) and (A,I)(B,IB)(A,I)\rightarrow(B,IB) is a flat map of prisms, then the complex (Cˇm^AB,1)m(\check{C}^{m}\widehat{\otimes}_{A}B,\partial\otimes 1)_{m} is a Čech–Alexander complex that computes 𝖱ΓΔ(𝒳B/B)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{B}/B).

  2. (2)

    Let now (A,I)(A,I) be the prism (Ainf,Kerθ)(A_{\mathrm{inf}},\mathrm{Ker}\,\theta) and let 𝒳\mathscr{X} be of the form 𝒳=𝒳0×𝒪K𝒪CK\mathscr{X}=\mathscr{X}^{0}\times_{\mathcal{O}_{K}}\mathcal{O}_{\mathbb{C}_{K}} where 𝒳0\mathscr{X}^{0} is a smooth separated formal 𝒪K\mathcal{O}_{K}–scheme. A convenient way to describe the GKG_{K}–action on 𝖱ΓΔ(𝒳/Ainf)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/A_{\mathrm{inf}}) is via base–change: given gGK,g\in G_{K}, the action of gg on AinfA_{\mathrm{inf}} gives a map of prisms g:(Ainf,(E(u)))(Ainf,(E(u)))g:(A_{\mathrm{inf}},(E(u)))\rightarrow(A_{\mathrm{inf}},(E(u))), and g𝒳=𝒳g^{*}\mathscr{X}=\mathscr{X} since 𝒳\mathscr{X} comes from 𝒪K\mathcal{O}_{K}. Base–change theorem for prismatic cohomology [BhattScholze, Theorem 1.8 (5)] then gives an AinfA_{\mathrm{inf}}–linear map g𝖱ΓΔ(𝒳/Ainf)𝖱ΓΔ(𝒳/Ainf);g^{*}\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/A_{\mathrm{inf}})\rightarrow\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/A_{\mathrm{inf}}); untwisting by gg on the left, this gives an AinfA_{\mathrm{inf}}gg–semilinear action map g:𝖱ΓΔ(𝒳/Ainf)𝖱ΓΔ(𝒳/Ainf)g:\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/A_{\mathrm{inf}})\rightarrow\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/A_{\mathrm{inf}}). The exact same procedure defines the GKG_{K}–action on the Čech–Alexander complexes modelling the cohomology theories since they are base–change compatible in the sense above.

3.  The conditions (𝐂𝐫𝒔\mathrm{Cr}_{s})

3.1.  Definition and basic properties

In order to describe the conditions (Crs\mathrm{Cr}_{s}), we need to fix more notation. For a natural number ss, denote by KsK_{s} the field K(πs)K(\pi_{s}) (where (πn)n(\pi_{n})_{n} is the compatible chain of pnp^{n}–th roots of π\pi chosen before, i.e. so that u=[(πn)n]u=[(\pi_{n})_{n}] in AinfA_{\mathrm{inf}}), and set K=sKsK_{\infty}=\bigcup_{s}K_{s}. Further set Kp=mK(ζpm)K_{p^{\infty}}=\bigcup_{m}K(\zeta_{p^{m}}) and for sN{}s\in\mathbb{N}\cup\{\infty\}, set Kp,s=KpKsK_{p^{\infty},s}=K_{p^{\infty}}K_{s}. Note that the field Kp,K_{p^{\infty},\infty} is the Galois closure of KK_{\infty}. Denote by G^\widehat{G} the Galois group Gal(Kp,/K)\mathrm{Gal}(K_{p^{\infty},\infty}/K) and by GsG_{s} the group Gal(K¯/Ks),\mathrm{Gal}(\overline{K}/K_{s}), for sN{}s\in\mathbb{N}\cup\{\infty\}.

The group G^\widehat{G} is generated by its two subgroups Gal(Kp,/Kp)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty}}) and Gal(Kp,/K)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{\infty}) (by [LiuBreuilConjecture, Lemma 5.1.2]). The subgroup Gal(Kp,/Kp)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty}}) is normal, and its element gg is uniquely determined by its action on the elements (πs)s(\pi_{s})_{s}, which takes the form g(πs)=ζpsasπsg(\pi_{s})=\zeta_{p^{s}}^{a_{s}}\pi_{s}, with the integers asa_{s} unique modulo psp^{s} and compatible with each other as ss increases. It follows that Gal(Kp,/Kp)Zp\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty}})\simeq\mathbb{Z}_{p}, with a topological generator τ\tau given by τ(πn)=ζpnπn\tau(\pi_{n})=\zeta_{p^{n}}\pi_{n} (where, again, ζpn\zeta_{p^{n}}’s are chosen as before, so that v=[(ζpn)n]1v=[(\zeta_{p^{n}})_{n}]-1).

Similarly, the image of GsG_{s} in G^\widehat{G} is the subgroup G^s=Gal(Kp,/Ks)\widehat{G}_{s}=\mathrm{Gal}(K_{p^{\infty},\infty}/K_{s}). Clearly G^s\widehat{G}_{s} contains Gal(Kp,/K)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{\infty}) and the intersection of G^s\widehat{G}_{s} with Gal(Kp,/Kp)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty}}) is Gal(Kp,/Kp,s).\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty},s}). Just as in the s=0s=0 case, G^s\widehat{G}_{s} is generated by these two subgroups, with the subgroup Gal(Kp,/Kp,s)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty},s}) normal and topologically generated by the element τps\tau^{p^{s}}.

There is a natural GKG_{K}–action on Ainf=W(𝒪CK),A_{\mathrm{inf}}=W(\mathcal{O}_{\mathbb{C}_{K}}^{\flat}), extended functorially from the natural action on 𝒪CK\mathcal{O}_{\mathbb{C}_{K}}^{\flat}. This action makes the map θ:Ainf𝒪CK\theta:A_{\mathrm{inf}}\rightarrow\mathcal{O}_{\mathbb{C}_{K}} GKG_{K}–equivariant, in particular, the kernel E(u)AinfE(u)A_{\mathrm{inf}} is GKG_{K}–stable. The GKG_{K}–action on the GKG_{K}–closure of 𝔖\mathfrak{S} in AinfA_{\mathrm{inf}} factors through G^\widehat{G}. Note that the subgroup Gal(Kp,/K)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{{\infty}}) of G^\widehat{G} acts trivially on elements of 𝔖\mathfrak{S}, and the action of the subgroup Gal(Kp,/Kp)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty}}) is determined by the equality τ(u)=(v+1)u\tau(u)=(v+1)u.

For an integer s0s\geq 0 and ii between 0 and ss, denote by ξs,i\xi_{s,i} the element

ξs,i=φs(v)ωφ(ω)φi(ω)=φ1(v)φi+1(ω)φi+2(ω)φs(ω)\xi_{s,i}=\frac{\varphi^{s}(v)}{\omega\varphi(\omega)\dots\varphi^{i}(\omega)}=\varphi^{-1}(v)\varphi^{i+1}(\omega)\varphi^{i+2}(\omega)\dots\varphi^{s}(\omega)

(recall that ω=v/φ1(v)\omega=v/\varphi^{-1}(v)), and set

Is=(ξs,0u,ξs,1up,,ξs,sups).I_{s}=\left(\xi_{s,0}u,\xi_{s,1}u^{p},\dots,\xi_{s,s}u^{p^{s}}\right).

For convenience of notation, we further set I=0I_{\infty}=0 and φ(v)u=0\varphi^{\infty}(v)u=0.

We are concerned with the following conditions.

Definition 3.1.

Let MinfM_{\inf} be an AinfA_{\mathrm{inf}}–module endowed with a GKG_{K}AinfA_{\mathrm{inf}}–semilinear action, let MBKM_{\mathrm{BK}} be an 𝔖\mathfrak{S}–module and let MBKMinfM_{\mathrm{BK}}\rightarrow M_{\inf} be an 𝔖\mathfrak{S}–linear map. Let s0s\geq 0 be an integer or \infty.

  1. (1)

    An element xMinfx\in M_{\inf} is called a (Crs\mathrm{Cr}_{s})–element if for every gGsg\in G_{s},

    g(x)xIsMinf.g(x)-x\in I_{s}M_{\inf}.
  2. (2)

    We say that the pair MBKMAinfM_{\mathrm{BK}}\rightarrow M_{A_{\mathrm{inf}}} satisfies the condition (Crs\mathrm{Cr}_{s}) if for every element xMBKx\in M_{\mathrm{BK}}, the image of xx in MinfM_{\inf} is (Crs\mathrm{Cr}_{s}).

  3. (3)

    An element xMinfx\in M_{\inf} is called a (Crs\mathrm{Cr}^{\prime}_{s})–element if for every gGsg\in G_{s}, there is an element yMinfy\in M_{\inf} such that

    g(x)x=φs(v)uy.g(x)-x=\varphi^{s}(v)uy.
  4. (4)

    We say that the pair MBKMAinfM_{\mathrm{BK}}\rightarrow M_{A_{\mathrm{inf}}} satisfies the condition (Crs\mathrm{Cr}^{\prime}_{s}) if for every element xMBKx\in M_{\mathrm{BK}}, the image of xx in MinfM_{\inf} is (Crs\mathrm{Cr}^{\prime}_{s}).

  5. (5)

    Aditionally, we call (Cr0\mathrm{Cr}_{0})–elements crystalline elements and we call the condition (Cr0\mathrm{Cr}_{0}) the crystalline condition.

Remarks 3.2.
  1. (1)

    Since I0=φ1(v)uAinf,I_{0}=\varphi^{-1}(v)uA_{\mathrm{inf}}, the crystalline condition equivalently states that for all gGKg\in G_{K} and all xx in the image of MBK,M_{\mathrm{BK}},

    g(x)xφ1(v)uMinf.g(x)-x\in\varphi^{-1}(v)uM_{\inf}.

    The reason for the extra terminology in the case s=0s=0 is that the condition is connected with a criterion for certain representations to be crystalline, as discussed in §3.2. The higher conditions (Crs\mathrm{Cr}_{s}) will on the other hand find application in computing bounds on ramification of pnp^{n}–torsion étale cohomology. The conditions (Crs\mathrm{Cr}^{\prime}_{s}) serve an auxillary purpose. Clearly (Crs\mathrm{Cr}^{\prime}_{s}) implies (Crs\mathrm{Cr}_{s}). The conditions (Cr\mathrm{Cr}_{\infty}), (Cr\mathrm{Cr}^{\prime}_{\infty}) are clearly both equivalent to the condition f(MBK)MinfGf(M_{\mathrm{BK}})\subseteq M_{\inf}^{G_{\infty}}.

  2. (2)

    Strictly speaking, one should talk about the crystalline condition (or (Crs\mathrm{Cr}_{s})) for the map ff, but we choose to talk about the the crystalline condition (or (Crs\mathrm{Cr}_{s})) for the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) instead, leaving the datum of the map ff implicit. This is because typically we consider the situation that MBKM_{\mathrm{BK}} is an 𝔖\mathfrak{S}–submodule of MinfGM_{\inf}^{G_{\infty}} and MBK𝔖AinfMAinfM_{\mathrm{BK}}\otimes_{\mathfrak{S}}A_{\mathrm{inf}}\simeq M_{A_{\mathrm{inf}}} via the natural map (or the derived (p,E(u))(p,E(u))–completed variant, MBK^𝔖AinfMAinfM_{\mathrm{BK}}\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}\simeq M_{A_{\mathrm{inf}}}). Also note that f:MBKMinff:M_{\mathrm{BK}}\rightarrow M_{\inf} satisfies the condition (Crs\mathrm{Cr}_{s}) if and only if f(MBK)Minff(M_{\mathrm{BK}})\subseteq M_{\inf} does.

Lemma 3.3.

For any integer ss, the ideals φs(v)uAinf\varphi^{s}(v)uA_{\mathrm{inf}} and IsI_{s} are GKG_{K}–stable.

Proof.

It is enough to prove that the ideals uAinfuA_{\mathrm{inf}} and vAinfvA_{\mathrm{inf}} are GKG_{K}–stable. Note that the GKG_{K}–stability of vAinfvA_{\mathrm{inf}} implies GKG_{K}–stability of φs(v)Ainf\varphi^{s}(v)A_{\mathrm{inf}} for any sZs\in\mathbb{Z} since φ\varphi is a GKG_{K}–equivariant automorphism of AinfA_{\mathrm{inf}}. Once we know this, we know that gφs(v)g\varphi^{s}(v) equals to φs(v)\varphi^{s}(v) times a unit for every gg and ss, the same is then true of φi(ω)=φi(v)/φi1(v)\varphi^{i}(\omega)=\varphi^{i}(v)/\varphi^{i-1}(v), hence also of all the elements ξi,s\xi_{i,s} and it follows that IsI_{s} is GKG_{K}–stable.

Given gGKg\in G_{K}, g(πn)=ζpnanπng(\pi_{n})=\zeta_{p^{n}}^{a_{n}}\pi_{n} for an ineger ana_{n} unique modulo pnp^{n} and such that an+1an(modpn)a_{n+1}\equiv a_{n}\pmod{p^{n}}. It follows that g(u)=[ε¯]aug(u)=[\underline{\varepsilon}]^{a}u for a pp–adic integer aa(=limnan=\lim_{n}a_{n}). (The Zp\mathbb{Z}_{p}–exponentiation used here is defined by [ε¯]a=limn[ε¯]an[\underline{\varepsilon}]^{a}=\lim_{n}[\underline{\varepsilon}]^{a_{n}} and the considered limit is with respect to the weak topology.) Thus, uAinfuA_{\mathrm{inf}} is GKG_{K}–stable.

Similarly, we have g(ζpn)=ζpnbn,g(\zeta_{p^{n}})=\zeta_{p^{n}}^{b_{n}}, for integers bnb_{n} coprime to pp, unique modulo pnp^{n} and compatible with each other as nn grows. It follows that g([ε¯])=[ε¯]bg([\underline{\varepsilon}])=[\underline{\varepsilon}]^{b} for b=limnbnb=\lim_{n}b_{n}, and so g(v)=(v+1)b1=limn((v+1)bn1).g(v)=(v+1)^{b}-1=\lim_{n}((v+1)^{b_{n}}-1). The resulting expression is still divisible by vv. To see that, fix the integers bnb_{n} to have all positive representatives. Then the claim follows from the formula

(v+1)bn1=v((v+1)bn1+(v+1)bn2++1),(v+1)^{b_{n}}-1=v((v+1)^{b_{n}-1}+(v+1)^{b_{n}-2}+\dots+1),

upon noting that the sequence of elements ((v+1)bn1+(v+1)bn2++1)=((v+1)bn1)/v((v+1)^{b_{n}-1}+(v+1)^{b_{n}-2}+\dots+1)=((v+1)^{b_{n}}-1)/v is still (p,v)(p,v)–adically (i.e. weakly) convergent thanks to Lemma 2.3. ∎

Let f:MBKMinff:M_{\mathrm{BK}}\rightarrow M_{\inf} be as in Definition 3.1. Lemma 3.3 shows that the modules Minf/IsMinfM_{\inf}/{I_{s}}M_{\inf} and Minf/φs(v)uMinfM_{\inf}/\varphi^{s}(v)uM_{\inf} have a well–defined GKG_{K}–action. Consequently, we get the following restatement of the conditions (Crs\mathrm{Cr}_{s}), (Crs\mathrm{Cr}^{\prime}_{s}).

Lemma 3.4.

Given f:MBKMinff:M_{\mathrm{BK}}\rightarrow M_{\inf} as in Definition 3.1, the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) satisfies the condition (Crs\mathrm{Cr}_{s}) ((Crs\mathrm{Cr}^{\prime}_{s}), resp.) if and only if the image of MBKM_{\mathrm{BK}} in Minf¯:=Minf/IsMinf\overline{M_{\inf}}:=M_{\inf}/I_{s}M_{\inf} (Minf¯:=Minf/φs(v)uMinf\overline{M_{\inf}}:=M_{\inf}/\varphi^{s}(v)uM_{\inf}, resp.) lands in Minf¯Gs\overline{M_{\inf}}^{G_{s}}.

In the case of the above–mentioned condition f(MBK)MinfG,f(M_{\mathrm{BK}})\subseteq M_{\inf}^{G_{\infty}}, the GKG_{K}–closure of f(MBK)f(M_{\mathrm{BK}}) in MinfM_{\inf} is contained in the GKG_{K}–submodule MGKp,M^{G_{K_{p^{\infty},\infty}}}, and thus, the GKG_{K}–action on it factors through G^\widehat{G}. Under mild assumtions on MinfM_{\inf}, the GsG_{s}–action on the elements of f(MBK)f(M_{\mathrm{BK}}) is ultimately determined by τps\tau^{p^{s}}, the topological generator of Gal(Kp,/Kp,s)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty},s}). Consequently, the conditions (Crs\mathrm{Cr}^{\prime}_{s}) are also determined by the action of this single element:

Lemma 3.5.

Let f:MBKMinff:M_{\mathrm{BK}}\rightarrow M_{\inf} be as in Definition 3.1. Additionally, assume that MinfM_{\inf} is classically (p,E(u))(p,E(u))–complete and (p,E(u))(p,E(u))–completely flat, and that the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) satisfies (Cr\mathrm{Cr}_{\infty}). Then the action of G^\widehat{G} on elements of f(MBK)f(M_{\mathrm{BK}}) makes sense, and (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) satisfies (Crs\mathrm{Cr}^{\prime}_{s}) if and only if

xf(MBK):τps(x)xφs(v)uMinf.\forall x\in f(M_{\mathrm{BK}}):\tau^{p^{s}}(x)-x\in\varphi^{s}(v)uM_{\inf}.
Proof.

Clearly the stated condition is necessary. To prove sufficiency, assume the above condition for τps\tau^{p^{s}}. By the fixed–point interpretation of the condition (Crs\mathrm{Cr}^{\prime}_{s}) as in Lemma 3.4, it is clear that the analogous condition holds for every element gτpsg\in\langle\tau^{p^{s}}\rangle.

Next, assume that gg is an element of Gal(Kp,/Kp,s)\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty},s}), the pp–adic closure of τps\langle\tau^{p^{s}}\rangle. Then gg can be written as limnτpsan\lim_{n}\tau^{p^{s}a_{n}}, with the sequence of integers (an)(a_{n}) pp–adically convergent. For xf(MBK),x\in f(M_{\mathrm{BK}}), by continuity we have g(x)x=limn(τpsan(x)x)g(x)-x=\lim_{n}(\tau^{p^{s}a_{n}}(x)-x), which is equal to limnφs(v)uyn\lim_{n}\varphi^{s}(v)uy_{n} with ynMinfy_{n}\in M_{\inf}. Since the sequence (yn)(y_{n}) is still convergent (using the fact that the (p,E(u))(p,E(u))–adic topology is the (p,φs(v)u)(p,\varphi^{s}(v)u)–adic topology, and that p,φs(v)up,\varphi^{s}(v)u is a regular sequence on MinfM_{\inf}), we have that g(x)x=φs(v)uyg(x)-x=\varphi^{s}(v)uy where y=limnyny=\lim_{n}y_{n}.

To conclude, note that a general element of G^s\widehat{G}_{s} is of the form g1g2g_{1}g_{2} where g1Gal(Kp,/Kp,s)g_{1}\in\mathrm{Gal}(K_{p^{\infty},\infty}/K_{p^{\infty},s}) and g2Gal(Kp,/K).g_{2}\in\mathrm{Gal}(K_{p^{\infty},\infty}/K_{\infty}). Then for xf(MBK)x\in f(M_{\mathrm{BK}}), by the assumption f(MBK)MinfGf(M_{\mathrm{BK}})\subseteq M_{\inf}^{G_{\infty}} we have g1g2(x)x=g1(x)xg_{1}g_{2}(x)-x=g_{1}(x)-x, and so the condition (Crs\mathrm{Cr}^{\prime}_{s}) is proved by the previous part. ∎

Let us now discuss some basic algebraic properties of the conditions (Crs\mathrm{Cr}_{s}) and (Crs\mathrm{Cr}^{\prime}_{s}). The basic situation when they are satisfied is the inclusion 𝔖Ainf\mathfrak{S}\hookrightarrow A_{\mathrm{inf}} itself.

Lemma 3.6.

The pair 𝔖Ainf\mathfrak{S}\hookrightarrow A_{\mathrm{inf}} satisfies the conditions (Crs\mathrm{Cr}^{\prime}_{s}) (hence also (Crs\mathrm{Cr}_{s})) for all s0s\geq 0.

Proof.

Note that 𝔖Ainf\mathfrak{S}\hookrightarrow A_{\mathrm{inf}} satisfies the assumptions of Lemma 3.5, so it is enough to consider the action of the element τpsG^s\tau^{p^{s}}\in\widehat{G}_{s}. For an element f=iaiui𝔖f=\sum_{i}a_{i}u^{i}\in\mathfrak{S} we have

τps(f)f=i0ai((v+1)psu)ii0aiui=i1ai((v+1)psi1)ui,\tau^{p^{s}}(f)-f=\sum_{i\geq 0}a_{i}((v+1)^{p^{s}}u)^{i}-\sum_{i\geq 0}a_{i}u^{i}=\sum_{i\geq 1}a_{i}((v+1)^{p^{s}i}-1)u^{i},

and thus,

τps(f)fφs(v)u=i1ai(v+1)psi1φs(v)ui1=i1ai(v+1)psi1(v+1)ps1ui1\frac{\tau^{p^{s}}(f)-f}{\varphi^{s}(v)u}=\sum_{i\geq 1}a_{i}\frac{(v+1)^{p^{s}i}-1}{\varphi^{s}(v)}u^{i-1}=\sum_{i\geq 1}a_{i}\frac{(v+1)^{p^{s}i}-1}{(v+1)^{p^{s}}-1}u^{i-1}

Since φs(v)=(v+1)ps1\varphi^{s}(v)=(v+1)^{p^{s}}-1 divides (v+1)psi1(v+1)^{p^{s}i}-1 for each ii, the obtained series has coefficients in AinfA_{\mathrm{inf}}, showing that τps(f)fφs(v)uAinf\tau^{p^{s}}(f)-f\in\varphi^{s}(v)uA_{\mathrm{inf}} as desired. ∎

The following lemma shows that in various contexts, it is often sufficient to verify the conditions (Crs\mathrm{Cr}_{s}), (Crs\mathrm{Cr}^{\prime}_{s}) on generators.

Lemma 3.7.

Fix an integer s0s\geq 0. Let (C) be either the condition (Crs\mathrm{Cr}_{s}) or (Crs\mathrm{Cr}^{\prime}_{s}).

  1. (1)

    Let MinfM_{\inf} be an AinfA_{\mathrm{inf}}–module with a GKG_{K}AinfA_{\mathrm{inf}}–semilinear action. The set of all (C)–elements forms an 𝔖\mathfrak{S}–submodule of MinfM_{\inf}.

  2. (2)

    Let CinfC_{\inf} be an AinfA_{\mathrm{inf}}–algebra endowed with a GKG_{K}–semilinear action. The set of (C)–elements of CinfC_{\inf} forms an 𝔖\mathfrak{S}–subalgebra of CinfC_{\inf}.

  3. (3)

    If the algebra CinfC_{\inf} from (2) is additionally AinfA_{\mathrm{inf}}δ\delta–algebra such that GKG_{K} acts by δ\delta–maps (i.e. δg=gδ\delta g=g\delta for all gGKg\in G_{K}) then the set of all (C)–elements forms a 𝔖\mathfrak{S}δ\delta–subalgebra of CinfC_{\inf}.

  4. (4)

    If the algebra CinfC_{\inf} as in (2) is additionally derived (p,E(u))(p,E(u))–complete and CBKCinfC_{\mathrm{BK}}\rightarrow C_{\inf} is a map of 𝔖\mathfrak{S}–algebras that satisfies the condition (C), then so does CBK^Cinf,\widehat{C_{\mathrm{BK}}}\rightarrow C_{\inf}, where CBK^\widehat{C_{\mathrm{BK}}} is the derived (p,E(u))(p,E(u))–completion of CBKC_{\mathrm{BK}}. In particular, the set of all (C)–elements in CinfC_{\inf} forms a derived (p,E(u))(p,E(u))–complete 𝔖\mathfrak{S}–subalgebra of CinfC_{\inf}.

Proof.

Let JJ be the ideal IsI_{s} if (C)=(Crs\mathrm{Cr}_{s}) and the ideal φs(v)uAinf\varphi^{s}(v)uA_{\mathrm{inf}} if (C)=(Crs\mathrm{Cr}^{\prime}_{s}). In view of Lemma 3.4, the sets described in (1),(2) are obtained as the preimages of (Minf/JMinf)Gs\left(M_{\inf}/JM_{\inf}\right)^{G_{s}} (ring (Cinf/JCinf)Gs\left(C_{\inf}/JC_{\inf}\right)^{G_{s}}, resp.) under the canonical projection MinfMinf/JMinfM_{\inf}\rightarrow M_{\inf}/JM_{\inf} (CinfCinf/JCinf,C_{\inf}\rightarrow C_{\inf}/JC_{\inf}, resp.). As these GsG_{s}–fixed points form an 𝔖\mathfrak{S}–module (𝔖\mathfrak{S}–algebra, resp.) by Lemma 3.6, this proves (1) and (2).

Similarly, to prove (3) we need to prove only that the ideal JCinfJC_{\inf} is a δ\delta–ideal and therefore the canonical projection CinfCinf/JCinfC_{\inf}\rightarrow C_{\inf}/JC_{\inf} is a map of δ\delta–rings.

Let us argue first in the case (Crs\mathrm{Cr}^{\prime}_{s}). As δ(u)=0\delta(u)=0, we have

δ(φs(v)u)=δ(φs(v))up=φ(φs(v))(φs(v))ppup=φs+1(v)(φs(v))ppup.\delta(\varphi^{s}(v)u)=\delta(\varphi^{s}(v))u^{p}=\frac{\varphi(\varphi^{s}(v))-(\varphi^{s}(v))^{p}}{p}u^{p}=\frac{\varphi^{s+1}(v)-(\varphi^{s}(v))^{p}}{p}u^{p}.

Recall that φs(v)=[ε¯]ps1\varphi^{s}(v)=[\underline{\varepsilon}]^{p^{s}}-1 divides φs+1(v)=([ε¯]ps)p1\varphi^{s+1}(v)=([\underline{\varepsilon}]^{p^{s}})^{p}-1. The numerator of the last fraction is thus divisible by φs(v)\varphi^{s}(v) and since φs(v)AinfpAinf=φs(v)pAinf\varphi^{s}(v)A_{\mathrm{inf}}\cap pA_{\mathrm{inf}}=\varphi^{s}(v)pA_{\mathrm{inf}} by Lemma 2.3, φs(v)\varphi^{s}(v) divides the whole fraction (φs+1(v)(φs(v))p)/p{(\varphi^{s+1}(v)-(\varphi^{s}(v))^{p})/p} in AinfA_{\mathrm{inf}}. (We note that this is true for every integer ss, in particular s=1s=-1, as well.)

Let us now prove that the ideal J=IsJ=I_{s} (hence also IsCinfI_{s}C_{\inf}) is a δ\delta–ideal. For any ii between 0 and s1s-1, we have

δ(ξs,i)=δ(φ1(v)φi+1(ω)φs(ω))=φ1(v)ωφi+2(ω)φs+1(ω)φ1(v)pφi+1(ω)pφs(ω)pp.\delta\left(\xi_{s,i}\right)=\delta(\varphi^{-1}(v)\varphi^{i+1}(\omega)\dots\varphi^{s}(\omega))=\frac{\varphi^{-1}(v)\omega\varphi^{i+2}(\omega)\dots\varphi^{s+1}(\omega)-\varphi^{-1}(v)^{p}\varphi^{i+1}(\omega)^{p}\dots\varphi^{s}(\omega)^{p}}{p}.

The numerator is divisible by ξs,i+1\xi_{s,i+1}, and so is the whole fraction thanks to Lemma 2.3. Thus, we have that δ(ξs,iupi)=δ(ξs,i)upi+1\delta(\xi_{s,i}u^{p^{i}})=\delta(\xi_{s,i})u^{p^{i+1}} is a multiple of ξs,i+1upi+1\xi_{s,i+1}u^{p^{i+1}}. Finally, when i=si=s, we have ξs,s=φ1(v)\xi_{s,s}=\varphi^{-1}(v), and δ(ξs,s)\delta(\xi_{s,s}) is thus a multiple of ξs,s\xi_{s,s} by the previous. Consequently, δ(ξs,sups)=δ(ξs,s)ups+1\delta(\xi_{s,s}u^{p^{s}})=\delta(\xi_{s,s})u^{p^{s+1}} is divisible by ξs,sups\xi_{s,s}u^{p^{s}}. This shows that IsI_{s} is a δ\delta–ideal.

Finally, let us prove (4). Note that E(u)ue(modp𝔖)E(u)\equiv u^{e}\pmod{p\mathfrak{S}}, hence (p,E(u))=(p,ue)=(p,u)\sqrt{(p,E(u))}=\sqrt{(p,u^{e})}=\sqrt{(p,u)} even as ideals of 𝔖\mathfrak{S}; consequently, the derived (p,E(u))(p,E(u))–completion agrees with the derived (p,u)(p,u)–completion both for 𝔖\mathfrak{S}– and AinfA_{\mathrm{inf}}–modules. We may therefore replace (p,E(u))(p,E(u))–completions with (p,u)(p,u)–completions throughout.

Since CinfC_{\inf} is derived (p,u)(p,u)–complete, any power series of the form

f=i,jci,jpiujf=\sum_{i,j}c_{i,j}p^{i}u^{j}

with ci,jCinfc_{i,j}\in C_{\inf} defines a unique111111Here we are using the preferred representatives of powers series as mentioned at the beginning of §2.1. element fCinff\in C_{\inf}, and ff comes from CBK^\widehat{C_{\mathrm{BK}}} if and only if the coefficients ci,jc_{i,j} may be chosen in the image of the map CBKCinfC_{\mathrm{BK}}\rightarrow C_{\inf}. Assuming this, for gGsg\in G_{s} we have

g(f)f=i,jg(ci,j)pi(γu)ji,jci,jpiuj=g(f)-f=\sum_{i,j}g(c_{i,j})p^{i}(\gamma u)^{j}-\sum_{i,j}c_{i,j}p^{i}u^{j}=
=i,j(g(ci,j)γjg(ci,j)+g(ci,j)ci,j)piuj,=\sum_{i,j}\left(g(c_{i,j})\gamma^{j}-g(c_{i,j})+g(c_{i,j})-c_{i,j}\right)p^{i}u^{j},

where γ\gamma is the AinfA_{\mathrm{inf}}–unit such that g(u)=γug(u)=\gamma u. Thus, it is clearly enough to show, upon assuming the condition (C) for (CBK,Cinf)(C_{\mathrm{BK}},C_{\inf}), that the terms (g(ci,j)γjg(ci,j))piuj\left(g(c_{i,j})\gamma^{j}-g(c_{i,j})\right)p^{i}u^{j} and (g(ci,j)ci,j)piuj\left(g(c_{i,j})-c_{i,j}\right)p^{i}u^{j} are in JCinfJC_{\inf} when gGsg\in G_{{s}}. (Note that an element d=i,jdi,jpiujd=\sum_{i,j}d_{i,j}p^{i}u^{j} with di,jJCinfd_{i,j}\in JC_{\inf} is itself in JCinfJC_{\inf}, since JJ is finitely generated.)

We have g(ci,j)ci,jJCinfg(c_{i,j})-c_{i,j}\in JC_{\inf} by assumption, so it remains to treat the term g(ci,j)(γj1)g(c_{i,j})(\gamma^{j}-1). Since (γj1)(\gamma^{j}-1) is divisible by γ1\gamma-1, it is also divisible by φs(v)\varphi^{s}(v) by Lemma 3.6. Thus, the terms g(ci,j)(γj1)piujg(c_{i,j})(\gamma^{j}-1)p^{i}u^{j} are divisible by φs(v)u\varphi^{s}(v)u when j1j\geq 1, and are 0 when j=0j=0; in either case, they are members of JCinfJC_{\inf}.

To prove the second assertion of (4), let now CBKCinfC_{\mathrm{BK}}\subseteq C_{\inf} be the 𝔖\mathfrak{S}–subalgebra of all crystalline elements. By the previous, the map CBK^Cinf\widehat{C_{\mathrm{BK}}}\rightarrow C_{\inf} satisfies (C), and hence the image CBK+C_{\mathrm{BK}}^{+} of this map consists of (C)–elements. Thus, we have CBKCBK+CBK,C_{\mathrm{BK}}\subseteq C_{\mathrm{BK}}^{+}\subseteq C_{\mathrm{BK}}, and hence, CBKC_{\mathrm{BK}} is derived (p,E(u))(p,E(u))–complete since so is CBK+C_{\mathrm{BK}}^{+}. ∎

Remark 3.8.

One consequence of Lemma 3.7 is that the 𝔖\mathfrak{S}–subalgebra \mathfrak{C} of AinfA_{\mathrm{inf}} formed by all crystalline elements (or even (Cr0\mathrm{Cr}^{\prime}_{0})–elements) forms a prism, with the distinguished invertible ideal I=E(u)I=E(u)\mathfrak{C}. As Lemma 3.6 works for any choice of Breuil–Kisin prism associated to K/KK/K^{\prime} in AinfA_{\mathrm{inf}}, \mathfrak{C} contains all of these (in particular, it contains all GKG_{K}–translates of 𝔖\mathfrak{S}).

For future use in applications to pnp^{n}–torsion modules, we consider the following approximation of the ideals IsI_{s} appearing in the conditions (Crs\mathrm{Cr}_{s}).

Lemma 3.9.

Consider a pair of integers n,sn,s with s0,n1s\geq 0,n\geq 1. Set t=max{0,s+1n}t=\mathrm{max}\left\{0,s+1-n\right\}. Then the image of the ideal IsI_{s} in the ring Wn(𝒪CK)=Ainf/pnW_{n}(\mathcal{O}_{\mathbb{C}_{K}^{\flat}})=A_{\mathrm{inf}}/p^{n} is contained in the ideal φ1(v)uptWn(𝒪CK)\varphi^{-1}(v)u^{p^{t}}W_{n}(\mathcal{O}_{\mathbb{C}_{K}^{\flat}}). That is, we have Is+pnAinfφ1(v)uptAinf+pnAinf.I_{s}+p^{n}A_{\mathrm{inf}}\subseteq\varphi^{-1}(v)u^{p^{t}}A_{\mathrm{inf}}+p^{n}A_{\mathrm{inf}}.

Proof.

When t=0t=0 there is nothing to prove, therefore we may assume that t=s+1n>0t=s+1-n>0. In the definition of IsI_{s}, we may replace the elements

ξs,i=φ1(v)φi+1(ω)φi+2(ω)φs(ω)\xi_{s,i}=\varphi^{-1}(v)\varphi^{i+1}(\omega)\varphi^{i+2}(\omega)\dots\varphi^{s}(\omega)

by the elements

ξs,i=φ1(v)φi+1(E(u))φi+2(E(u))φs(E(u)),\xi^{\prime}_{s,i}=\varphi^{-1}(v)\varphi^{i+1}(E(u))\varphi^{i+2}(E(u))\dots\varphi^{s}(E(u)),

since the quotients ξs,i/ξs,i\xi_{s,i}/\xi^{\prime}_{s,i} are AinfA_{\mathrm{inf}}–units.

It is thus enough to show that for every ii with 0is,0\leq i\leq s, the element

ϑs,i=ξs,iupiφ1(v)=φi+1(E(u))φi+2(E(u))φs(E(u))upi\vartheta_{s,i}=\frac{\xi^{\prime}_{s,i}u^{p^{i}}}{\varphi^{-1}(v)}=\varphi^{i+1}(E(u))\varphi^{i+2}(E(u))\dots\varphi^{s}(E(u))u^{p^{i}}

taken modulo pnp^{n} is divisible by ups+1nu^{p^{s+1-n}}.

This is clear when is+1ni\geq s+1-n, and so it remains to discuss the cases when isn.i\leq s-n. Write φj(E(u))=(ue)pj+pxj\varphi^{j}(E(u))=(u^{e})^{p^{j}}+px_{j} (with xj𝔖x_{j}\in\mathfrak{S}). Then it is enough to show that

(*) ϑs,iupi=((ue)pi+1+pxi+1)((ue)pi+2+pxi+2)((ue)ps+pxs)\frac{\vartheta_{s,i}}{u^{p^{i}}}=((u^{e})^{p^{i+1}}+px_{i+1})((u^{e})^{p^{i+2}}+px_{i+2})\dots((u^{e})^{p^{s}}+px_{s})

taken modulo pnp^{n} is divisible by

ups+1npi=upi(p1)(1+p++psni).u^{p^{s+1-n}-p^{i}}=u^{p^{i}(p-1)(1+p+\dots+p^{s-n-i})}.

Since we are interested in the product (*3.1) only modulo pnp^{n}, in expanding the brackets we may ignore the terms that use the expressions of the form pxjpx_{j} at least nn times. Each of the remaining terms contains the product of at least sin+1s-i-n+1 distinct terms from the following list:

(ue)pi+1,(ue)pi+2,,(ue)ps.(u^{e})^{p^{i+1}},(u^{e})^{p^{i+2}},\dots,(u^{e})^{p^{s}}.

Thus, each of the remaining terms is divisible by (at least)

(ue)pi+1+pi+2++psn+1=(ue)pi(p)(1+p++psni),(u^{e})^{p^{i+1}+p^{i+2}+\dots+p^{s-n+1}}=(u^{e})^{p^{i}\cdot(p)\cdot(1+p+\dots+p^{s-n-i})},

which is more than needed. This finishes the proof. ∎

3.2.  Crystalline condition for Breuil–Kisin–Fargues GKG_{K}–modules

The situation of central interest regarding the crystalline condition is the inclusion MBKMinfGM_{\mathrm{BK}}\rightarrow M_{\inf}^{G_{\infty}} such that Ainf𝔖MBKMinfA_{\mathrm{inf}}\otimes_{\mathfrak{S}}M_{\mathrm{BK}}\rightarrow M_{\inf} is an isomorphism, where MBKM_{\mathrm{BK}} is a Breuil–Kisin module and MinfM_{\inf} is a Breuil–Kisin–Fargues GKG_{K}–module. The version of these notions used in this paper is tailored to the context of prismatic cohomology. Namely, we have:

Definition 3.10.
  1. (1)

    A Breuil–Kisin module is a finitely generated 𝔖\mathfrak{S}–module MM together with a 𝔖[1/E(u)]\mathfrak{S}[1/E(u)]–linear isomorphism

    φ=φM[1/E]:(φ𝔖M)[1/E(u)]M[1/E(u)].\varphi=\varphi_{M[1/E]}:(\varphi_{\mathfrak{S}}^{*}M)[1/E(u)]\stackrel{{\scriptstyle\sim}}{{\rightarrow}}M[1/E(u)].

    For a positive integer ii, the Breuil–Kisin module MM is said to be of height i\leq i if φM[1/E]\varphi_{M[1/E]} is induced (by linearization and localization) by a φ𝔖\varphi_{\mathfrak{S}}–semilinear map φM:MM\varphi_{M}:M\rightarrow M such that, denoting φlin:φMM\varphi_{\mathrm{lin}}:\varphi^{*}M\rightarrow M its linearization, there exists an 𝔖\mathfrak{S}–linear map ψ:MφM\psi:M\rightarrow\varphi^{*}M such that both the compositions ψφlin\psi\circ\varphi_{\mathrm{lin}} and φlinψ\varphi_{\mathrm{lin}}\circ\psi are multiplication by E(u)iE(u)^{i}. A Breuil–Kisin module is of finite height if it is of height i\leq i for some ii.

  2. (2)

    A Breuil–Kisin–Fargues module is a finitely presented AinfA_{\mathrm{inf}}–module MM such that M[1/p]M[1/p] is a free Ainf[1/p]A_{\mathrm{inf}}[1/p]–module, together with an Ainf[1/E(u)]A_{\mathrm{inf}}[1/E(u)]–linear isomorphism

    φ=φM[1/E]:(φAinfM)[1/E(u)]M[1/E(u)].\varphi=\varphi_{M[1/E]}:(\varphi_{A_{\mathrm{inf}}}^{*}M)[1/E(u)]\stackrel{{\scriptstyle\sim}}{{\rightarrow}}M[1/E(u)].

    Similarly, the Breuil–Kisin–Fargues module is called of height i\leq i if φM[1/E]\varphi_{M[1/E]} comes from a φAinf\varphi_{A_{\mathrm{inf}}}–semilinear map φM:MM\varphi_{M}:M\rightarrow M admitting an AinfA_{\mathrm{inf}}–linear map ψ:MφM\psi:M\rightarrow\varphi^{*}M such that ψφlin\psi\circ\varphi_{\mathrm{lin}} and φlinψ\varphi_{\mathrm{lin}}\circ\psi are multiplication maps by E(u)iE(u)^{i}, where φlin\varphi_{\mathrm{lin}} is the inearization of φM\varphi_{M}. A Breuil–Kisin–Fargues module is of finite height if it is of height i\leq i for some ii.

  3. (3)

    A Breuil–Kisin–Fargues GKG_{K}–module (of height i\leq i, of finite height, resp.) is a Breuil–Kisin–Fargues module (of height i\leq i, of finite height, resp.) that is additionally endowed with an AinfA_{\mathrm{inf}}–semilinear GKG_{K}–action that makes φM[1/E]\varphi_{M[1/E]} GKG_{K}–equivariant (that makes also φM\varphi_{M} GKG_{K}–equivariant in the finite height cases).

That is, the definition of a Breuil–Kisin module agrees with the one in [BMS1], and MinfM_{\inf} is a Breuil–Kisin–Fargues module in the sense of the above definition if and only if φAinfMinf\varphi_{A_{\mathrm{inf}}}^{*}M_{\inf} is a Breuil–Kisin–Fargues module in the sense of [BMS1]121212This is to account for the fact that while Breuil–Kisin–Fargues modules in the sense of [BMS1] appear as AinfA_{\mathrm{inf}}–cohomology groups of smooth proper formal schemes, Breuil–Kisin–Fargues modules in the above sense appear as prismatic AinfA_{\mathrm{inf}}–cohomology groups of smooth proper formal schemes.. The notion of Breuil–Kisin module of height i\leq i agrees with what is called “(generalized) Kisin modules of height ii” in [LiLiu]. The above notion of finite height Breuil–Kisin–Fargues modules agrees with the one from [EmertonGee2, Appendix F] except that the modules are not assumed to be free. Also note that under these definitions, for a Breuil–Kisin module MBKM_{\mathrm{BK}} (of height i,\leq i, resp.), the AinfA_{\mathrm{inf}}–module Minf=Ainf𝔖MBKM_{\inf}=A_{\mathrm{inf}}\otimes_{\mathfrak{S}}M_{\mathrm{BK}} is a Breuil–Kisin–Fargues module (of height i,\leq i, resp.), without the need to twist the embedding 𝔖Ainf\mathfrak{S}\rightarrow A_{\mathrm{inf}} by φ\varphi.

The connection between Breuil–Kisin–, Breuil–Kisin–Fargues GKG_{K}–modules and the crystalline condition (justifying its name) is the following theorem.

Theorem 3.11 ([EmertonGee2, Appendix F], [GaoBKGK]).

Let MinfM_{\inf} be a free Breuil–Kisin–Fargues GKG_{K}–module which admits as an 𝔖\mathfrak{S}–submodule a free Breuil–Kisin module MBKMinfGM_{\mathrm{BK}}\subseteq M_{\inf}^{G_{\infty}} of finite height, such that Ainf𝔖MBKMinfA_{\mathrm{inf}}\otimes_{\mathfrak{S}}M_{\mathrm{BK}}\stackrel{{\scriptstyle\sim}}{{\rightarrow}}M_{\inf} (as Breuil–Kisin–Fargues modules) via the natural map, and such that the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) satisfies the crystalline condition. Then the étale realization of MinfM_{\inf},

V(Minf)=(W(CK)AinfMinf)φ=1[1p],V(M_{\inf})=\left(W(\mathbb{C}_{K}^{\flat})\otimes_{A_{\mathrm{inf}}}M_{\inf}\right)^{\varphi=1}\left[\frac{1}{p}\right],

is a crystalline representation.

Remarks 3.12.
  1. (1)

    Theorem 3.11 is actually an equivalence: If V(Minf)V(M_{\inf}) is crystalline, it can be shown that the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) satisfies the crystalline condition. We state the theorem in the one direction since this is the one that we use. However, the converse direction motivates why it is resonable to expect the crystalline condition for prismatic cohomology groups that is discussed in Section 4.

  2. (2)

    Strictly speaking, in [EmertonGee2, Appendix F] one assumes extra conditions on the pair MinfM_{\inf} (“satisfying all descents”); however, these extra assumptions are used only for a semistable version of the statement. Theorem 3.11 in its equivalence form is therefore only implicit in the proof of [EmertonGee2, Theorem F.11]. (See also [Ozeki, Theorem 3.8] for a closely related result.)

  3. (3)

    On the other hand, Theorem 3.11 in the one–sided form as above is a consequence of [GaoBKGK, Proposition 7.11] that essentially states that V(M)V(M) is crystalline if and only if the much weaker condition

    gGK:(g1)MBKφ1(v)W(𝔪𝒪CK)Minf\forall g\in G_{K}:\;\;(g-1)M_{\mathrm{BK}}\subseteq\varphi^{-1}(v)W(\mathfrak{m}_{\mathcal{O}_{\mathbb{C}_{K}^{\flat}}})M_{\inf}

    is satisfied. We note a related result of loc. cit.: V(M)V(M) is semistable if and only if

    gGK:(g1)MBKW(𝔪𝒪CK)Minf.\forall g\in G_{K}:\;\;(g-1)M_{\mathrm{BK}}\subseteq W(\mathfrak{m}_{\mathcal{O}_{\mathbb{C}_{K}^{\flat}}})M_{\inf}.

    This semistable criterion above might be a good starting point in generalizing the results of Sections 4 and 5 of the present paper to the case of semistable reduction, using the log–prismatic cohomology developed in [Koshikawa]. Thus, a natural question to ask is: Similarly to how the crystalline condition is a stronger version of the crystallinity criterion from [GaoBKGK], what is an analogous stronger (while still generally valid) version of the semistability criterion from [GaoBKGK]?

It will be convenient later to have version of Theorem 3.11 that applies to not necessarily free Breuil–Kisin and Breuil–Kisin–Fargues modules. Recall that, by [BMS1, Propostition 4.3], any Breuil–Kisin module MBKM_{\mathrm{BK}} is related to a free Breuil–Kisin module MBK,freeM_{\mathrm{BK},\mathrm{free}} by a functorial exact sequence

0{0}MBK,tor{M_{\mathrm{BK},\mathrm{tor}}}MBK{M_{\mathrm{BK}}}MBK,free{M_{\mathrm{BK},\mathrm{free}}}MBK¯{\overline{M_{\mathrm{BK}}}}0{0}

where MBK,torM_{\mathrm{BK},\mathrm{tor}} is a pnp^{n}-torsion module for some nn and MBK¯\overline{M_{\mathrm{BK}}} is supported at the maximal ideal (p,u)(p,u). Taking the base–change to Ainf,A_{\mathrm{inf}}, one obtains an analogous exact sequence

0{0}Minf,tor{M_{\inf,\mathrm{tor}}}Minf{M_{\inf}}Minf,free{M_{\inf,\mathrm{free}}}Minf¯{\overline{M_{\inf}}}0{0}

(also described by [BMS1, Proposition 4.13]) where Minf,freeM_{\inf,\mathrm{free}} is a free Breuil–Kisin–Fargues module. Clearly the maps MBKMfreeM_{\mathrm{BK}}\rightarrow M_{\mathrm{free}} and MinfMinf,freeM_{\inf}\rightarrow M_{\inf,\mathrm{free}} become isomorphisms after inverting pp.

Assume that MinfM_{\inf} is endowed with a GKG_{K}–action that makes it a Breuil–Kisin–Fargues GKG_{K}–module. The functoriality of the latter exact sequence implies that the GKG_{K}–action on MinfM_{\inf} induces a GKG_{K}–action on Minf,freeM_{\inf,\mathrm{free}}, endowing it with the structure of a free Breuil–Kisin–Faruges GKG_{K}–module. In more detail, given σGK\sigma\in G_{K}, the semilinear action map σ:MinfMinf\sigma:M_{\inf}\rightarrow M_{\inf} induces an AinfA_{\mathrm{inf}}–linear map σlin:σMinfMinf\sigma_{\mathrm{lin}}:\sigma^{*}M_{\inf}\rightarrow M_{\inf} where σM=Ainfσ,AinfM\sigma^{*}M=A_{\mathrm{inf}}\otimes_{\sigma,A_{\mathrm{inf}}}M. As σ\sigma is an isomorphism fixing pp, E(u)E(u) up to unit and the ideal (p,u)Ainf,(p,u)A_{\mathrm{inf}}, it is easy to see that σMinf\sigma^{*}M_{\inf} is itself a Breuil–Kisin–Fargues module, and the exact sequence from [BMS1, Proposition 4.13] for σMinf\sigma^{*}M_{\inf} can be identified with the upper row of the diagram

0{0}σMinf,tor{\sigma^{*}M_{\inf,\mathrm{tor}}}σMinf{\sigma^{*}M_{\inf}}σMinf,free{\sigma^{*}M_{\inf,\mathrm{free}}}σMinf¯{\sigma^{*}\overline{M_{\inf}}}0{0}0{0}Minf,tor{M_{\inf,\mathrm{tor}}}Minf{M_{\inf}}Minf,free{M_{\inf,\mathrm{free}}}Minf¯{\overline{M_{\inf}}}0,{0,}σlin\scriptstyle{\sigma_{\mathrm{lin}}}σlin\scriptstyle{\sigma_{\mathrm{lin}}}σlin\scriptstyle{\sigma_{\mathrm{lin}}}σlin\scriptstyle{\sigma_{\mathrm{lin}}}

where the second vertical map is the linearization of σ\sigma and the rest is induced by functoriality of the sequence. Finally, untwisting σMinf,free,\sigma^{*}M_{\inf,\mathrm{free}}, the third vertical map σlin\sigma_{\mathrm{lin}} induces a semilinear map σ:Minf,freeMinf,free\sigma:M_{\inf,\mathrm{free}}\rightarrow M_{\inf,\mathrm{free}}. Note that the module Minf[1/p]Minf,free[1/p]M_{\inf}[1/p]\simeq M_{\inf,\mathrm{free}}[1/p] inherits the GKG_{K}–action from MinfM_{\inf}; it is easy to see that the GKG_{K}-action on Minf,freeM_{\inf,\mathrm{free}} agrees with the one on Minf[1/p]M_{\inf}[1/p] when viewing Minf,freeM_{\inf,\mathrm{free}} as its submodule.

Proposition 3.13.

Assume that the pair MBKMinfM_{\mathrm{BK}}\hookrightarrow M_{\inf} satisfies the crystalline condition. Then so does the pair MBK,freeMinf,free.M_{\mathrm{BK},\mathrm{free}}\hookrightarrow M_{\inf,\mathrm{free}}.

Proof.

Notice that the crystalline condition is satisfied for MBK[1/p]Minf[1/p]M_{BK}[1/p]\rightarrow M_{\inf}[1/p] and by [BMS1, Propositions 4.3, 4.13], this map can be identified with MBK,free[1/p]Minf,free[1/p]M_{\mathrm{BK},\mathrm{free}}[1/p]\hookrightarrow M_{\inf,\mathrm{free}}[1/p]. Thus, the following lemma finishes the proof. ∎

Lemma 3.14.

Let FinfF_{\inf} be a free AinfA_{\mathrm{inf}}–module endowed with AinfA_{\mathrm{inf}}–semilinear GKG_{K}–action and let FBKFinfF_{BK}\subseteq F_{\inf} be a free 𝔖\mathfrak{S}–submodule such that FBK[1/p]Finf[1/p]F_{BK}[1/p]\hookrightarrow F_{\inf}[1/p] satisfies the crystalline condition. Then the pair FBKFinfF_{BK}\hookrightarrow F_{\inf} satisfies the crystalline condition.

Proof.

Fix an element aFBKa\in F_{BK} and gGKg\in G_{K}. The crystalline condition holds after inverting pp, and so we have the equality

b:=(g1)a=φ1(v)ucpkb:=(g-1)a=\varphi^{-1}(v)u\frac{c}{p^{k}}

with cFinfc\in F_{\inf}. In other words (using that pkp^{k} is a non-zero divisor on FinfF_{\inf}), we have

pkb=φ1(v)ucpkFinfφ1(v)uFinf=pkφ1(v)uFinf,p^{k}b=\varphi^{-1}(v)uc\in p^{k}F_{\inf}\cap\varphi^{-1}(v)uF_{\inf}=p^{k}\varphi^{-1}(v)uF_{\inf},

where the last equality follows by Lemma 2.3 since FinfF_{\inf} is a free module. In particular, we have

pkb=pkφ1(v)udp^{k}b=p^{k}\varphi^{-1}(v)ud

for yet another element dFinfd\in F_{\inf}. As pkp^{k} is a non–zero divisor on AinfA_{\mathrm{inf}}, hence on Finf,F_{\inf}, we may cancel out to conclude

(g1)a=b=φ1(v)udφ1(v)uFinf,(g-1)a=b=\varphi^{-1}(v)ud\in\varphi^{-1}(v)uF_{\inf},

as desired. ∎

Proposition 3.13 leads to the following strenghtening of Theorem 3.11.

Theorem 3.15.

The “free” assumption in Theorem 3.11 is superfluous. That is, given a Breuil–Kisin–Fargues GKG_{K}–module MinfM_{\inf} together with its Breuil–Kisin–𝔖\mathfrak{S}–submodule MBKMinfGM_{\mathrm{BK}}\subseteq M_{\inf}^{G_{\infty}} of finite height such that Ainf𝔖MBKMinfA_{\mathrm{inf}}\otimes_{\mathfrak{S}}M_{\mathrm{BK}}\stackrel{{\scriptstyle\sim}}{{\rightarrow}}M_{\inf} and such that the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) satisfies the crystalline condition, the representation

V(Minf)=(W(CK)AinfMinf)φ=1[1p]V(M_{\inf})=\left(W(\mathbb{C}_{K}^{\flat})\otimes_{A_{\mathrm{inf}}}M_{\inf}\right)^{\varphi=1}\left[\frac{1}{p}\right]

is crystalline.

Proof.

With the notation as above, upon realizing that V(Minf)V(M_{\inf}) and V(Minf,free)V(M_{\inf,\mathrm{free}}) agree, the result is a direct consequence of Proposition 3.13. ∎

4.  Conditions (𝐂𝐫𝒔\mathrm{Cr}_{s}) for cohomology

4.1.  (𝐂𝐫𝒔\mathrm{Cr}_{s}) for Čech–Alexander complexes

Let 𝒳\mathscr{X} be a smooth separated pp–adic formal scheme over 𝒪K\mathcal{O}_{K}. Denote by ČBK\v{C}_{\mathrm{BK}}^{\bullet} a Čech–Alexander complex that models 𝖱ΓΔ(𝒳/𝔖)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S}) and set Cˇinf=CˇBK^𝔖Ainf\check{C}_{\inf}^{\bullet}=\check{C}_{\mathrm{BK}}^{\bullet}\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}, computed termwise – by Remark 2.22, this is a Čech–Alexander complex modelling 𝖱ΓΔ(𝒳Ainf/Ainf)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}}}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}). We aim to prove the following.

Theorem 4.1.

For every m0m\geq 0 and sN{}s\in\mathbb{N}\cup\{\infty\}, the pair ČBKmČinfm\v{C}_{\mathrm{BK}}^{m}\rightarrow\v{C}_{\inf}^{m} satisfies the condition (Crs\mathrm{Cr}_{s}).

Let Spf(R)=𝒱𝒳\mathrm{Spf}(R)=\mathscr{V}\subseteq\mathscr{X} be an affine open formal subscheme. Then it is enough to prove the content of Theorem 4.1 for CˇBKCˇinf\check{C}_{\mathrm{BK}}\rightarrow\check{C}_{\inf} where CˇBK\check{C}_{\mathrm{BK}} and Cˇinf=CˇBK^𝔖Ainf\check{C}_{\inf}=\check{C}_{\mathrm{BK}}\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}} are the Čech–Alexander covers of 𝒱\mathscr{V} and 𝒱=𝒱×𝔖Ainf\mathscr{V}^{\prime}=\mathscr{V}\times_{\mathfrak{S}}A_{\mathrm{inf}} with respect to the base prism 𝔖\mathfrak{S} and AinfA_{\mathrm{inf}}, respectively, since the Čech–Alexander complexes termwise consist of products of such covers. Let R=R^𝒪K𝒪CK(=R^𝔖Ainf)R^{\prime}=R\widehat{\otimes}_{\mathcal{O}_{K}}\mathcal{O}_{\mathbb{C}_{K}}(=R\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}).

Let us fix a choice of the free 𝔖\mathfrak{S}–algebra P0=𝔖[{Xi}iI]P_{0}=\mathfrak{S}[\{X_{i}\}_{i\in I}] whose (p,E(u))(p,E(u))–completion is the algebra P=P𝒱P=P_{\mathscr{V}} as in Construction 2.12, with JJ being the kernel of the surjection PRP\rightarrow R. Then the corresponding choices at the AinfA_{\mathrm{inf}}–level are P0=P0𝔖AinfP_{0}^{\prime}=P_{0}\otimes_{\mathfrak{S}}A_{\mathrm{inf}} and P=P^𝔖Ainf,P^{\prime}=P\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}, and the associated (p,E(u))(p,E(u))–completed “δ\delta–envelopes” are also related by the completed base change; that is, we have a diagram with exact rows

(4.3) JPδ^{\widehat{JP^{\delta}}}Pδ^{\widehat{P^{\delta}}}RPPδ^{\widehat{R\otimes_{P}P^{\delta}}}0{0}J(P)δ^{\widehat{J(P^{\prime})^{\delta}}}(P)δ^{\widehat{(P^{\prime})^{\delta}}}RP(P)δ^{\widehat{R^{\prime}\otimes_{P^{\prime}}(P^{\prime})^{\delta}}}0.{0.}^𝔖Ainf{-\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}}α\scriptstyle{\alpha}

By Remark 2.17, we may and do assume that the set of variables {Xi}i={X1,,Xm}\{X_{i}\}_{i}=\{X_{1},\dots,X_{m}\} is finite, and that the ideal JJ is finitely generated. Consequently, after replacing the maps on the left by their respective images (and invoking Remark 2.13 (1)), diagram (4.3) becomes

(4.4) 0{0}JPδ^{J\widehat{P^{\delta}}}Pδ^{\widehat{P^{\delta}}}RPPδ^{\widehat{R\otimes_{P}P^{\delta}}}0{0}0{0}J(P)δ^{J\widehat{(P^{\prime})^{\delta}}}(P)δ^{\widehat{(P^{\prime})^{\delta}}}RP(P)δ^{\widehat{R^{\prime}\otimes_{P^{\prime}}(P^{\prime})^{\delta}}}0{0}^𝔖Ainf\scriptstyle{-\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}}^𝔖Ainf\scriptstyle{-\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}}α\scriptstyle{\alpha}

where the rows are exact. The prescription g(Xi)=Xig(X_{i})=X_{i} determines uniquely a continuous, AinfA_{\mathrm{inf}}–semilinear Galois action by δ\delta–maps on (P)δ^P0δ𝔖Ainf^\widehat{(P^{\prime})^{\delta}}\simeq\widehat{P_{0}^{\delta}\otimes_{\mathfrak{S}}A_{\mathrm{inf}}} (and, in particular, this action satisfies g(δj(Xi))=δj(Xi)g(\delta^{j}(X_{i}))=\delta^{j}(X_{i}) for all gGKg\in G_{K} and all i,ji,j). Similarly, the term RP(P)δ^RPPδ𝒪K𝒪CK^\widehat{R^{\prime}\otimes_{P^{\prime}}(P^{\prime})^{\delta}}\simeq\widehat{R\otimes_{P}P^{\delta}\otimes_{\mathcal{O}_{K}}\mathcal{O}_{\mathbb{C}_{K}}} is given the (linear) GKG_{K}–action prescribed by g(xa)=xg(a)g(x\otimes a)=x\otimes g(a) for every gGKg\in G_{K}, a𝒪CKa\in\mathcal{O}_{\mathbb{C}_{K}^{\flat}} and xx coming from the first row. This makes the map α\alpha GKG_{K}–equivariant, and therefore the kernel J(P)δ^\widehat{J(P^{\prime})^{\delta}} GKG_{K}–stable. As a consequence, the action extends to the prismatic envelope (Cˇinf,ICˇinf)(\check{C}_{\inf},I\check{C}_{\inf}) where the action obtained this way agrees with the one indicated in Remark 2.22. Upon taking the prismatic envelope (CˇBK,ICˇBK)(\check{C}_{\mathrm{BK}},I\check{C}_{\mathrm{BK}}) of the δ\delta–pair (P0δ^,J0P0δ^)(\widehat{P_{0}^{\delta}},J_{0}\widehat{P_{0}^{\delta}}), we arrive at the situation CˇBKCˇinf=CˇBK^𝔖Ainf\check{C}_{\mathrm{BK}}\hookrightarrow\check{C}_{\inf}=\check{C}_{\mathrm{BK}}\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}} for which we wish to verify the conditions (Crs\mathrm{Cr}_{s}).

With the goal of understanding the GKG_{K}–action on Cˇinf\check{C}_{\inf} even more explicitly, in similar spirit to the proof of [BhattScholze, Proposition 3.13] we employ the following approximation of the prismatic envelope.

Definition 4.2.

Let BB be a δ\delta–ring, JBJ\subseteq B an ideal with a fixed generating set x¯={xi}iΛ,\underline{x}=\{x_{i}\}_{i\in\Lambda}, and let bJb\in J be an element. Denote by 𝔟0\mathfrak{b}_{0} be the kernel of the BB–algebra map

B[T¯]=B[{Ti}iΛ]\displaystyle B[\underline{T}]=B[\{T_{i}\}_{i\in\Lambda}] B[1b]\displaystyle\longrightarrow B\left[\frac{1}{b}\right]
Ti\displaystyle T_{i} xib,\displaystyle\longmapsto\frac{x_{i}}{b},

and let 𝔟\mathfrak{b} be the δ\delta-ideal in B{T¯}B\{\underline{T}\} generated by 𝔟0\mathfrak{b}_{0}. Then we denote by B{x¯b}B\{\frac{\underline{x}}{b}\} the δ\delta–ring B{T¯}/𝔟B\{\underline{T}\}/\mathfrak{b}, and call it weak δ\delta–blowup algebra of x¯\underline{x} and bb.

That is, the above construction adjoins (in δ\delta–sense) the fractions xi/bx_{i}/b to BB together with all relations among them that exist in B[1/b]B[1/b], making it possible to naturally compute with fractions.

Note that if BCB\rightarrow C is a map of BBδ\delta–algebras such that JC=bCJC=bC and this ideal is invertible, the fact that the localization map CC[1b]C\rightarrow C[\frac{1}{b}] is injective shows that there is a unique map of BBδ\delta–algebras B{x¯b}CB\{\frac{\underline{x}}{b}\}\rightarrow C. (In fact, if bb happens to be a non–zero divisor on B{x¯b}B\{\frac{\underline{x}}{b}\}, then B{x¯b}B\{\frac{\underline{x}}{b}\} is initial among all such BBδ\delta–algebras; this justifies the name ’weak δ\delta–blowup algebra’.)

The purpose of the construction is the following.

Proposition 4.3.

Let (A,I)(A,I) be a bounded orientable prism with an orientation dId\in I. Consider a map of δ\delta–pairs (A,I)(B,J)(A,I)\rightarrow(B,J) and assume that (C,IC)(C,IC) is a prismatic envelope for (B,J)(B,J) that is classically (p,I)(p,I)–complete. Let x¯={xi}iΛ\underline{x}=\{x_{i}\}_{i\in\Lambda} be a system of generators of JJ. Then there is a surjective map of δ\delta–rings B{x¯d}^clC\widehat{B\{\frac{\underline{x}}{d}\}}^{\mathrm{cl}}\rightarrow C, where ()^cl\widehat{(-)}^{\mathrm{cl}} denotes the classical (p,I)(p,I)–completion.

Note that the assumptions apply to a Čech–Alexander cover in place of (C,IC)(C,IC) since it is (p,I)(p,I)–completely flat over the base prism, hence classically (p,I)(p,I)–complete by [BhattScholze, Proposition 3.7].

Proof.

Since JC=dCJC=dC and and dd is a non–zero divisor on CC, there is an induced map B{x¯d}CB\{\frac{\underline{x}}{d}\}\rightarrow C and hence a map of δ\delta–rings B{x¯d}^clC\widehat{B\{\frac{\underline{x}}{d}\}}^{\mathrm{cl}}\rightarrow C (using [BhattScholze, Lemma 2.17]).

To see that this map is surjective, let CC^{\prime} denote its image in CC, and denote by ι\iota the inclusion of CC^{\prime} into CC. Then CC^{\prime} is (derived, and, consequently, clasically) (p,I)(p,I)–complete AAδ\delta–algebra with C[d]=0C^{\prime}[d]=0. It follows that (C,IC)=(C,(d))(C^{\prime},IC^{\prime})=(C^{\prime},(d)) is a prism by [BhattScholze, Lemma 3.5] and thus, by the universal property of CC, there is a map of BBδ\delta–algebras r:CCr:C\rightarrow C^{\prime} which is easily seen to be right inverse to ι\iota. Hence, ι\iota is surjective, proving the claim. ∎

Finally, we are ready to prove the following proposition which, as noted above, proves Theorem 4.1.

Proposition 4.4.

The pair CˇBKCˇinf\check{C}_{\mathrm{BK}}\rightarrow\check{C}_{\inf} satisfies the conditions (Crs\mathrm{Cr}_{s}) for every sN{}s\in\mathbb{N}\cup\{\infty\}.

Proof.

Fix a generating set y1,y2,,yny_{1},y_{2},\dots,y_{n} of JJ, and set P1=Pδ^,P1=(P)δ^P_{1}=\widehat{P^{\delta}},P_{1}^{\prime}=\widehat{(P^{\prime})^{\delta}}. We obtain a commutative diagram

(4.5) P1{y¯E(u)}^{\widehat{P_{1}\{\frac{\underline{y}}{E(u)}\}}}P1{y¯E(u)}^{\widehat{P_{1}^{\prime}\{\frac{\underline{y}}{E(u)}\}}}CˇBK{\check{C}_{\mathrm{BK}}}Cˇinf,{\check{C}_{\inf},}

where the vertical maps are the surjective maps from Proposition 4.3, and the horizontal maps come from the (p,E(u))(p,E(u))–completed base change ^𝔖Ainf-\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}.

The GKG_{K}–action on P1P_{1}^{\prime} naturally extends to P1{y¯E(u)}P_{1}^{\prime}\{\frac{\underline{y}}{E(u)}\} by the rule on generators

g(yjE(u))=g(yj)g(E(u))=γ1g(yj)E(u)g\left(\frac{y_{j}}{E(u)}\right)=\frac{g(y_{j})}{g(E(u))}=\gamma^{-1}\frac{g(y_{j})}{E(u)}

where γ\gamma is the AinfA_{\mathrm{inf}}–unit such that g(E(u))=γE(u)g(E(u))=\gamma E(u) (note that the fraction on the right–hand side makes sense as g(yj)JP1g(y_{j})\in JP_{1}^{\prime}). The action can be again extended continuously to the (p,E(u))(p,E(u))–adic completion, and this action makes the right vertical map GKG_{K}–equivariant.

It is therefore enough to prove the validity of the conditions (Crs\mathrm{Cr}_{s}) for the pair (P1{y¯E(u)}^,P1{y¯E(u)}^)(\widehat{P_{1}\{\frac{\underline{y}}{E(u)}\}},\widehat{P_{1}^{\prime}\{\frac{\underline{y}}{E(u)}\}}). By Lemma 3.7 (3),(4), it is enough to check the conditions for the topological generators of P1{y¯E(u)}^\widehat{P_{1}\{\frac{\underline{y}}{E(u)}\}} as an 𝔖\mathfrak{S}δ\delta–algebra, which are X1,X2,,XmX_{1},X_{2},\dots,X_{m} and y1/E(u),y2/E(u),,yn/E(u).y_{1}/E(u),y_{2}/E(u),\dots,y_{n}/E(u).

Fix an integer sN{}s\in\mathbb{N}\cup\{\infty\}. Since the elements X1,X2,,XmX_{1},X_{2},\dots,X_{m} satisfy g(Xi)Xi=0g(X_{i})-X_{i}=0 for every gGsg\in G_{s}, by Lemma 3.7 the pair PPP\rightarrow P^{\prime} satisfies the stronger condition (Crs\mathrm{Cr}^{\prime}_{s}). In particular, (Crs\mathrm{Cr}^{\prime}_{s}) holds not only for the variables XiX_{i}, but also for y1,y2,,yny_{1},y_{2},\dots,y_{n} since they come from PP.

Let us now fix an index 1jn1\leq j\leq n and an element gGsg\in G_{s}. We may write

g(yj)yj=φs(v)uzj=ξs,0uE(u)z~jg(y_{j})-y_{j}=\varphi^{s}(v)uz_{j}=\xi_{s,0}uE(u)\tilde{z}_{j}

for some zj,zj~Pz_{j},\tilde{z_{j}}\in P^{\prime} (that are equal up to a multiplication by an AinfA_{\mathrm{inf}}–unit). Similarly, we have

g1(E(u))E(u)=(γ11)E(u)=φs(v)ua=ξs,0uE(u)a~g^{-1}(E(u))-E(u)=(\gamma^{-1}-1)E(u)=\varphi^{s}(v)ua=\xi_{s,0}uE(u)\tilde{a}

with a,a~Ainfa,\tilde{a}\in A_{\mathrm{inf}} (again equal up to a unit).

Thus, regarding the generator yj/E(u),y_{j}/E(u), we have that

g(yjE(u))yjE(u)=γ1g(yj)yjE(u)=γ1g(yj)γ1yj+γ1yjyjE(u)=g\left(\frac{y_{j}}{E(u)}\right)-\frac{y_{j}}{E(u)}=\frac{\gamma^{-1}g(y_{j})-y_{j}}{E(u)}=\frac{\gamma^{-1}g(y_{j})-\gamma^{-1}y_{j}+\gamma^{-1}y_{j}-y_{j}}{E(u)}=
=γ1g(yj)yjE(u)+(γ11)yjE(u)=γ1ξs,0uz~j+ξs,0ua~yjIs.=\gamma^{-1}\frac{g(y_{j})-y_{j}}{E(u)}+(\gamma^{-1}-1)\frac{y_{j}}{E(u)}=\gamma^{-1}\xi_{s,0}u\tilde{z}_{j}+\xi_{s,0}u\tilde{a}y_{j}\in I_{s}.

This shows that each of the generators yj/E(u)y_{j}/E(u) is a (Crs\mathrm{Cr}_{s})–element, which finishes the proof. ∎

4.2.  Consequences for cohomology groups

Let us now use Theorem 4.1 to draw some conclusions for individual cohomology groups. The first is the crystalline condition for the prismatic cohomology groups and its consequence for pp–adic étale cohomology. As before, let 𝒳\mathscr{X} be a separated smooth pp–adic formal scheme over 𝒪K\mathcal{O}_{K}. Denote by 𝒳Ainf\mathscr{X}_{A_{\mathrm{inf}}} the base change 𝒳×𝒪K𝒪CK=𝒳×𝔖Ainf\mathscr{X}\times_{\mathcal{O}_{K}}\mathcal{O}_{\mathbb{C}_{K}}=\mathscr{X}\times_{\mathfrak{S}}A_{\mathrm{inf}}, and by 𝒳η¯\mathscr{X}_{\overline{\eta}} the geometric adic generic fiber.

Corollary 4.5.

For every i0,i\geq 0, the pair HΔi(𝒳/𝔖)HΔi(𝒳Ainf/Ainf)H_{{{\mathbbl{\Delta}}}}^{i}(\mathscr{X}/\mathfrak{S})\rightarrow H_{{{\mathbbl{\Delta}}}}^{i}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}) satisfies the conditions (Cr0\mathrm{Cr}_{0}) and (Cr\mathrm{Cr}_{\infty}).

Proof.

By the results of Section 2.2, we may and do model the cohomology theories by the Čech–Alexander complexes

CˇBKCˇinf=CˇBK^𝔖Ainf,\check{C}_{\mathrm{BK}}^{\bullet}\rightarrow\check{C}_{\inf}^{\bullet}=\check{C}_{\mathrm{BK}}^{\bullet}\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}},

and by Theorem 4.1 the conditions (Cr0\mathrm{Cr}_{0}) and (Cr\mathrm{Cr}_{\infty}) termwise hold for this pair. The condition (Cr\mathrm{Cr}_{\infty}) for HiΔ(𝒳/𝔖)HiΔ(𝒳/Ainf)GH^{i}_{{{\mathbbl{\Delta}}}}(\mathscr{X}/\mathfrak{S})\subseteq H^{i}_{{{\mathbbl{\Delta}}}}(\mathscr{X}/A_{\mathrm{inf}})^{G_{\infty}} thus follows immediately, and it remains to verify the crystalline condition.

Each of the terms Cˇinfi\check{C}_{\inf}^{i} is (p,E(u))(p,E(u))–completely flat over AinfA_{\mathrm{inf}}, which means in particular that the terms Cˇinfi\check{C}_{\inf}^{i} are torsion–free by Corollary 2.4. Denote the differentials on CˇBK,Cˇinf\check{C}_{\mathrm{BK}}^{\bullet},\check{C}_{\inf}^{\bullet} by \partial and \partial^{\prime}, resp.

To prove the crystalline condition for cohomology groups, it is clearly enough to verify the condition at the level of cocycles. Given xZi(CˇBK),x\in Z^{i}(\check{C}_{\mathrm{BK}}^{\bullet}), denote by xx^{\prime} its image in Zi(Cˇinf)Z^{i}(\check{C}_{\inf}^{\bullet}). For gGKg\in G_{K} we have g(x)x=φ1(v)uyg(x^{\prime})-x^{\prime}=\varphi^{-1}(v)uy^{\prime} for some yCˇinfiy^{\prime}\in\check{C}_{\inf}^{i}. As g(x)xZi(Cˇinf),g(x^{\prime})-x^{\prime}\in Z^{i}(\check{C}_{\inf}^{\bullet}), we have

φ1(v)u(y)=(φ1(v)uy)=0,\varphi^{-1}(v)u\partial^{\prime}(y^{\prime})=\partial^{\prime}(\varphi^{-1}(v)uy^{\prime})=0,

and the torsion–freeness of Cˇinfi+1\check{C}_{\inf}^{i+1} implies that (y)=0\partial^{\prime}(y^{\prime})=0. Thus, yZi(Cˇinf)y^{\prime}\in Z^{i}(\check{C}_{\inf}) as well, showing that g(x)xφ1(v)uZi(Cˇinf),g(x^{\prime})-x^{\prime}\in\varphi^{-1}(v)uZ^{i}(\check{C}_{\inf}^{\bullet}), as desired. ∎

When 𝒳\mathscr{X} is proper over 𝒪K\mathcal{O}_{K}, we use Corollary 4.5 to reprove the result from [BMS1] that the étale cohomology groups Hiét(𝒳η¯,Qp)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Q}_{p}) are in this case crystalline representations.

Corollary 4.6.

Assume that 𝒳\mathscr{X} is additionally proper over 𝒪K.\mathcal{O}_{K}. Then for any i0,i\geq 0, the pp–adic étale cohomology Héti(𝒳η¯,Qp)\mathrm{H}_{\text{\'{e}t}}^{i}(\mathscr{X}_{\overline{\eta}},\mathbb{Q}_{p}) is a crystalline representation.

Proof.

It follows from [BhattScholze, Theorem 1.8] (and faithful flatness of Ainf/𝔖A_{\mathrm{inf}}/\mathfrak{S}) that MBK=HΔi(𝒳/𝔖)M_{\mathrm{BK}}=\mathrm{H}_{{{\mathbbl{\Delta}}}}^{i}(\mathscr{X}/\mathfrak{S}) and Minf=HΔi(𝒳Ainf/Ainf)M_{\inf}=\mathrm{H}_{{{\mathbbl{\Delta}}}}^{i}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}) are Breuil–Kisin and Breuil–Kisin–Fargues modules, resp., such that Minf=MBK𝔖AinfM_{\inf}=M_{\mathrm{BK}}\otimes_{\mathfrak{S}}A_{\mathrm{inf}}. Moreover, MinfM_{\inf} has the structure of a Breuil–Kisin–Fargues GKG_{K}–module with

V(Minf):=(W(CK)AinfMinf)φ=1[1p]Héti(𝒳η¯,Qp)V(M_{\inf}):=\left(W(\mathbb{C}_{K}^{\flat})\otimes_{A_{\mathrm{inf}}}M_{\inf}\right)^{\varphi=1}\left[\frac{1}{p}\right]\simeq\mathrm{H}_{\text{\'{e}t}}^{i}(\mathscr{X}_{\overline{\eta}},\mathbb{Q}_{p})

as GKG_{K}–representations. By Corollary 4.5, the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) satisfies all the assumptions of Theorem 3.15. The claim thus follows. ∎

For the purposes of obtaining a bound on ramification of pp–torsion étale cohomology in §5, let us recall the notion of torsion prismatic cohomology as defined in [LiLiu], and discuss the consequences of the conditions (Crs\mathrm{Cr}_{s}) in this context.

Definition 4.7.

Given a bounded prism (A,I)(A,I) and a smooth pp–adic formal scheme 𝒳\mathscr{X} over A/IA/I, the pnp^{n}–torsion prismatic cohomology of 𝒳\mathscr{X} is defined as

𝖱ΓΔ,n(𝒳/A)=𝖱ΓΔ(𝒳/A)𝖫ZZ/pnZ.\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/A)=\mathsf{R}\Gamma_{{\mathbbl{\Delta}}}(\mathscr{X}/A)\stackrel{{\scriptstyle{\mathsf{L}}}}{{\otimes}}_{\mathbb{Z}}\mathbb{Z}/p^{n}\mathbb{Z}.

We denote the cohomology groups of 𝖱ΓΔ,n(𝒳/A)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/A) by HiΔ,n(𝒳/A)\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/A) (and refer to them as pnp^{n}–torsion prismatic cohomology groups).

Proposition 4.8.

Let s,ns,n be a pair of integers satisfying s0,n1s\geq 0,n\geq 1. Set t=max{0,s+1n}.t=\mathrm{max}\left\{0,s+1-n\right\}. Then the torsion prismatic cohomology groups HiΔ,n(𝒳/𝔖)HiΔ,n(𝒳Ainf/Ainf)\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/\mathfrak{S})\rightarrow\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}) satisfy the following condition:

gGs:(g1)HiΔ,n(𝒳/𝔖)φ1(v)uptHiΔ,n(𝒳Ainf/Ainf).\forall g\in G_{s}:\;\;\;(g-1)\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/\mathfrak{S})\subseteq\varphi^{-1}(v)u^{p^{t}}\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}).
Proof.

The proof is a slightly refined variant of the proof of Corollary 4.5. Consider again the associated Čech–Alexander complexes over 𝔖\mathfrak{S} and AinfA_{\mathrm{inf}},

CˇBKCˇinf=CˇBK^𝔖Ainf.\check{C}_{\mathrm{BK}}^{\bullet}\rightarrow\check{C}_{\inf}^{\bullet}=\check{C}_{\mathrm{BK}}^{\bullet}\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}}.

Both of these complexes are given by torsion–free, hence Z\mathbb{Z}–flat, modules by Corollary 2.4. Consequently, 𝖱ΓΔ,n(𝒳/𝔖)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/\mathfrak{S}) is modelled by CˇBK,n:=CˇBK/pnCˇBK\check{C}_{\mathrm{BK},n}^{\bullet}:=\check{C}_{\mathrm{BK}}^{\bullet}/p^{n}\check{C}_{\mathrm{BK}}^{\bullet}, and similarly for 𝖱ΓΔ,n(𝒳Ainf/Ainf)\mathsf{R}\Gamma_{{{\mathbbl{\Delta}}},n}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}) and Cˇinf,n=Cˇinf/pnCˇinf\check{C}_{\inf,n}^{\bullet}=\check{C}_{\inf}^{\bullet}/p^{n}\check{C}_{\inf}^{\bullet}. That is, the considered maps between cohomology groups are obtained as the maps on cohomologies for the base–change map of chain complexes

CˇBK,nCˇinf,n=CˇBK,n^𝔖Ainf,\check{C}_{\mathrm{BK},n}^{\bullet}\rightarrow\check{C}_{\inf,n}^{\bullet}=\check{C}_{\mathrm{BK},n}^{\bullet}\widehat{\otimes}_{\mathfrak{S}}A_{\mathrm{inf}},

and as in the proof of Corolary 4.5, it is enough to establish the desired condition for the respective groups of cocycles.

Set α=φ1(v)upt\alpha=\varphi^{-1}(v)u^{p^{t}}. Note that by Lemma 3.9, the condition (Crs\mathrm{Cr}_{s}) for the pair of complexes CˇBK,nCˇinf,n\check{C}_{\mathrm{BK},n}^{\bullet}\rightarrow\check{C}_{\inf,n}^{\bullet} implies the condition

gGs:(g1)CˇBK,nαCˇinf,n\forall g\in G_{s}:\;\;\;(g-1)\check{C}_{\mathrm{BK},n}^{\bullet}\subseteq\alpha\check{C}_{\inf,n}^{\bullet}

(meant termwise as usual), and since the terms of the complex Cˇinf\check{C}_{\inf}^{\bullet} are (p,E(u))(p,E(u))–complete and (p,E(u))(p,E(u))–completely flat, α\alpha is a non–zero divisor on the terms of Cˇinf,n\check{C}_{\inf,n}^{\bullet} by Corollary 2.4.

So pick any element xZi(CˇBK,n)x\in Z^{i}(\check{C}_{\mathrm{BK},n}^{\bullet}). The image xx^{\prime} of xx in Cˇinf,ni\check{C}_{\inf,n}^{i} lies in Zi(Cˇinf,n)Z^{i}(\check{C}_{\inf,n}^{\bullet}) and for any chosen gGsg\in G_{s} we have g(x)x=αyg(x^{\prime})-x^{\prime}=\alpha y^{\prime} for some yCˇinf,niy^{\prime}\in\check{C}_{\inf,n}^{i}. Now g(x)xg(x^{\prime})-x^{\prime} lies in Zi(Cˇinf,n),Z^{i}(\check{C}_{\inf,n}^{\bullet}), so αy=g(x)x\alpha y^{\prime}=g(x^{\prime})-x^{\prime} satisfies

0=(αy)=α(y).0=\partial^{\prime}(\alpha y^{\prime})=\alpha\partial^{\prime}(y^{\prime}).

Since α\alpha is a non–zero divisor on Cˇinf,ni+1\check{C}_{\inf,n}^{i+1}, it follows that (y)=0\partial^{\prime}(y^{\prime})=0, that is, yy^{\prime} lies in Zi(Cˇinf,n).Z^{i}(\check{C}_{\inf,n}^{\bullet}). We thus infer that g(x)x=αyαZi(Cˇinf,n),g(x^{\prime})-x^{\prime}=\alpha y^{\prime}\in\alpha Z^{i}(\check{C}_{\inf,n}^{\bullet}), as desired. ∎

5.  Ramification bounds for mod pp étale cohomology

5.1.  Ramification bounds

We are ready to discuss the implications to the question of ramification bounds for mod pp étale cohomology groups Héti(𝒳η¯,Z/pZ)\mathrm{H}_{\text{\'{e}t}}^{i}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}) when 𝒳\mathscr{X} is smooth and proper pp–adic formal scheme over 𝒪K\mathcal{O}_{K}.

We define an additive valuation vv^{\flat} on 𝒪CK\mathcal{O}_{\mathbb{C}_{K}}^{\flat} by v(x)=v(x)v^{\flat}(x)=v(x^{\sharp}) where vv is the valuation on 𝒪CK\mathcal{O}_{\mathbb{C}_{K}} normalized so that v(π)=1v(\pi)=1, and ():𝒪CK𝒪CK(-)^{\sharp}:\mathcal{O}_{\mathbb{C}_{K}}^{\flat}\rightarrow\mathcal{O}_{\mathbb{C}_{K}} is the multiplicative lift of pr0:𝒪CK𝒪CK/p\mathrm{pr}_{0}:\mathcal{O}_{\mathbb{C}_{K}}^{\flat}\rightarrow\mathcal{O}_{\mathbb{C}_{K}}/p. This way, we have v(π¯)=1v^{\flat}(\underline{\pi})=1 and v(ε¯1)=pe/(p1)v^{\flat}(\underline{\varepsilon}-1)=pe/(p-1). For a real number c0c\geq 0, denote by 𝔞>c\mathfrak{a}^{>c} (𝔞c,\mathfrak{a}^{\geq c}, resp.) the ideal of 𝒪CK\mathcal{O}_{\mathbb{C}_{K}}^{\flat} formed by all elements xx with v(x)>cv^{\flat}(x)>c (v(x)cv^{\flat}(x)\geq c, resp.).

Similarly, we fix an additive valuation vKv_{K} of KK normalized by vK(π)=1v_{K}(\pi)=1. Then for an algebraic extension L/KL/K and a real number c0c\geq 0, we denote by 𝔞L>c\mathfrak{a}_{L}^{>c} the ideal consisting of all elements x𝒪Lx\in\mathcal{O}_{L} with vK(x)>cv_{K}(x)>c (and similarly, for ’\geq’ as well).

For a finite extensions M/F/KM/F/K and a real number m0m\geq 0, let us recall (a version of131313Fontaine’s original condition uses the ideals 𝔞Em\mathfrak{a}_{E}^{\geq m} instead. Up to changing some inequalities from ‘<<’ to ‘\leq’ and vice versa, the conditions are fairly equivalent.) Fontaine’s property (PmM/F)(P_{m}^{M/F}):

(PmM/F)For any algebraic extension E/F, the existence of an 𝒪F–algebra map𝒪M𝒪E/𝔞E>m implies the existence of an F–injection of fields ME.\begin{array}[]{cc}(P_{m}^{M/F})&\begin{array}[]{l}\text{For any algebraic extension }E/F,\text{ the existence of an }\mathcal{O}_{F}\text{--algebra map}\\ \mathcal{O}_{M}\rightarrow\mathcal{O}_{E}/\mathfrak{a}_{E}^{>m}\text{ implies the existence of an }F\text{--injection of fields }M\hookrightarrow E.\end{array}\end{array}

We also recall the upper ramification numbering in the convention used in [Fontaine, CarusoLiu]. For G=Gal(M/F)G=\mathrm{Gal}(M/F) and a non–negative real number λ,\lambda, set

G(λ)={gG|vM(g(x)x)λx𝒪M},G_{(\lambda)}=\{g\in G\;|\;v_{M}(g(x)-x)\geq\lambda\;\;\forall x\in\mathcal{O}_{M}\},

where vMv_{M} is again the additive valuation of MM normalized by vM(M×)=Zv_{M}(M^{\times})=\mathbb{Z}.

For t0,t\geq 0, set

ϕM/F(t)=0tdt[G(1):G(t)]\phi_{M/F}(t)=\int_{0}^{t}\frac{\mathrm{d}t}{[G_{(1)}:G_{(t)}]}

(which makes sense as G(t)G(1)G_{(t)}\subseteq G_{(1)} for all t>1t>1). Then ϕM/F\phi_{M/F} is a piecewise–linear increasing continuous concave function. Denote by ψM/F\psi_{M/F} its inverse, and set G(μ)=G(ψM/F(μ)).G^{(\mu)}=G_{(\psi_{M/F}(\mu))}.

Denote by λM/F\lambda_{M/F} the infimum of all λ0\lambda\geq 0 such that G(λ)={id},G_{(\lambda)}=\{\mathrm{id}\}, and by μM/F\mu_{M/F} the infimum of all μ0\mu\geq 0 such that G(μ)={id}.G^{(\mu)}=\{\mathrm{id}\}. Clearly one has μM/F=ϕM/F(λM/F).\mu_{M/F}=\phi_{M/F}(\lambda_{M/F}).

Remark 5.1.

Let us compare the indexing conventions with [SerreLocalFields] and [Fontaine], as the results therein are (implicitly or explicitly) used. If GS-(μ),GF-(μ)G^{\text{S-}(\mu)},G^{\text{F-}(\mu)} are the upper–index ramification groups in [SerreLocalFields] and [Fontaine], resp., and similarly we denote GS-(λ)G_{\text{S-}(\lambda)} and GF-(λ)G_{\text{F-}(\lambda)} the lower–index ramification groups, then we have

G(μ)=GS-(μ1)=GF-(μ),G(λ)=GS-(λ1)=GF-(λ/e~),G^{(\mu)}=G^{\text{S-}(\mu-1)}=G^{\text{F-}(\mu)},\;\;\;\;G_{(\lambda)}=G_{\text{S-}(\lambda-1)}=G_{\text{F-}(\lambda/\tilde{e})},

where e~=eM/F\tilde{e}=e_{M/F} is the ramification index of M/FM/F.

In particular, since the enumeration differs from the one in [SerreLocalFields] only by a shift by one, the upper indexing is still compatible with passing to quotients, and it make sense to set

GF(μ)=limM/FGal(M/F)(μ)G_{F}^{(\mu)}=\varprojlim_{M^{\prime}/F}\mathrm{Gal}(M^{\prime}/F)^{(\mu)}

where M/FM^{\prime}/F varies over finite Galois extensions M/FM^{\prime}/F contained in a fixed algebraic closure K¯\overline{K} of KK (and GF=limM/FGal(M/F)G_{F}=\varprojlim_{M^{\prime}/F}\mathrm{Gal}(M^{\prime}/F) is the absolute Galois group).

Regarding μ\mu, the following transitivity formula is useful.

Lemma 5.2 ([CarusoLiu, Lemma 4.3.1]).

Let N/M/FN/M/F be a pair of finite extensions with both N/FN/F and M/FM/F Galois. Then we have μN/F=max(μM/F,ϕM/F(μN/M)).\mu_{N/F}=\mathrm{max}(\mu_{M/F},\phi_{M/F}(\mu_{N/M})).

The property (PM/Fm)(P^{M/F}_{m}) is connected with the ramification of the field extension M/FM/F as follows.

Proposition 5.3.

Let M/F/KM/F/K be finite extensions of fields with M/FM/F Galois and let m>0m>0 be a real number. If the property (PM/Fm)(P^{M/F}_{m}) holds, then:

  1. (1)

    ([Yoshida, Proposition 3.3]) μM/FeF/Km.\mu_{M/F}\leq e_{F/K}m. In fact, μM/F/eF/K\mu_{M/F}/e_{F/K} is the infimum of all m>0m>0 such that (PM/Fm)(P^{M/F}_{m}) is valid.

  2. (2)

    ([CarusoLiu, Corollary 4.2.2]) vK(𝒟M/F)<m,v_{K}(\mathcal{D}_{M/F})<m, where 𝒟M/F\mathcal{D}_{M/F} denotes the different of the field extension M/FM/F.

Corollary 5.4.

Both the assumptions and the conclusions of Proposition 5.3 are insensitive to replacing FF by any unramified extension of FF contained in MM.

Proof.

Let F/FF^{\prime}/F be an unramified extension such that FMF^{\prime}\subseteq M. The fact that (PM/Fm)(P^{M/F}_{m}) is equivalent to (PM/Fm)(P^{M/F^{\prime}}_{m}) is proved in [Yoshida, Proposition 2.2]. To show that also the conclusions are the same for FF and FF^{\prime}, it is enough to observe that eF/K=eF/K,eM/F=eM/F,e_{F^{\prime}/K}=e_{F/K},e_{M/F^{\prime}}=e_{M/F}, vK(𝒟M/F)=vK(𝒟M/F)v_{K}(\mathcal{D}_{M/F^{\prime}})=v_{K}(\mathcal{D}_{M/F}) and μM/F=μM/F\mu_{M/F^{\prime}}=\mu_{M/F}. The first two equalities are clear since F/FF^{\prime}/F is unramified. The third equality follows from 𝒟M/F=𝒟M/F𝒟F/F\mathcal{D}_{M/F}=\mathcal{D}_{M/F^{\prime}}\mathcal{D}_{F^{\prime}/F} upon noting that 𝒟F/F\mathcal{D}_{F^{\prime}/F} is the unit ideal. Finally, by Lemma 5.2, we have μM/F=max(μF/F,ϕF/F(μM/F)).\mu_{M/F}=\mathrm{max}(\mu_{F^{\prime}/F},\phi_{F^{\prime}/F}(\mu_{M/F^{\prime}})). As F/FF^{\prime}/F is unramified, we have μF/F=0\mu_{F^{\prime}/F}=0 and ϕF/F(t)=t\phi_{F^{\prime}/F}(t)=t for all t0t\geq 0. The fourth equality thus follows as well. ∎

Let 𝒳\mathscr{X} be a proper and smooth pp–adic formal scheme over 𝒪K\mathcal{O}_{K}. Fix the integer ii, and denote by TT^{\prime} the Galois module Hiét(𝒳η¯,Z/pZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}). Let LL be the splitting field of TT^{\prime}, i.e. L=K¯KerρL=\overline{K}^{\mathrm{Ker}\,\rho} where ρ:GKAutFp(T)\rho:G_{K}\rightarrow\mathrm{Aut}_{\mathbb{F}_{p}}(T^{\prime}) is the associated representation. The goal is to provide an upper bound on vK(𝒟L/K)v_{K}(\mathcal{D}_{L/K}), and a constant μ0=μ0(e,i,p)\mu_{0}=\mu_{0}(e,i,p) such that GK(μ)G_{K}^{(\mu)} acts trivially on TT^{\prime} for all μ>μ0\mu>\mu_{0}.

To achieve this, we follow rather closely the strategy of [CarusoLiu]. The main difference is that the input of (φ,G^)(\varphi,\widehat{G})–modules attached to the discussed GKG_{K}–respresentations in [CarusoLiu] is in our situation replaced by a pp–torsion Breuil–Kisin module and a Breuil–Kisin–Fargues GKG_{K}–module that arise as the pp–torsion prismatic 𝔖\mathfrak{S}– and AinfA_{\mathrm{inf}}–cohomology, resp. Let us therefore lay out the strategy, referring to proofs in [CarusoLiu] whenever possible, and describe the needed modifications where necessary. To facilitate this approach further, the notation used will usually reflect the notation of [CarusoLiu], except for mostly omitting the index nn throughout (which in our situation is always equal to 11).

The relation of the above–mentioned pp–torsion prismatic cohomologies to the pp–torsion étale cohomology is as follows.

Proposition 5.5 ([LiLiu, Proposition 7.2, Corollary 7.4, Remark 7.5]).

Let 𝒳\mathscr{X} be a smooth and proper pp–adic formal scheme over 𝒪K\mathcal{O}_{K}. Then

  1. (1)

    MBK=HiΔ,n(𝒳/𝔖)M_{\mathrm{BK}}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/\mathfrak{S}) is a pnp^{n}–torsion Breuil–Kisin module of height i\leq i, and we have

    Héti(𝒳η¯,Z/pnZ)Tn(MBK):=(MBKWn(k)[[u]]Wn(CK))φ=1\mathrm{H}_{\text{\'{e}t}}^{i}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p^{n}\mathbb{Z})\simeq T_{n}(M_{\mathrm{BK}}):=\left(M_{\mathrm{BK}}\otimes_{W_{n}(k)[[u]]}W_{n}(\mathbb{C}_{K}^{\flat})\right)^{\varphi=1}

    as Z/pnZ[G]\mathbb{Z}/p^{n}\mathbb{Z}[G_{\infty}]–modules.

  2. (2)

    Minf=HiΔ,n(𝒳Ainf/Ainf)M_{\inf}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}) is a pnp^{n}–torsion Breuil–Kisin–Fargues GKG_{K}–module of height i\leq i, and we have

    Héti(𝒳η¯,Z/pnZ)Tn(Minf):=(MinfWn(𝒪CK)Wn(CK))φ=1\mathrm{H}_{\text{\'{e}t}}^{i}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p^{n}\mathbb{Z})\simeq T_{n}(M_{\inf}):=\left(M_{\inf}\otimes_{W_{n}(\mathcal{O}_{\mathbb{C}_{K}^{\flat}})}W_{n}(\mathbb{C}_{K}^{\flat})\right)^{\varphi=1}

    as Z/pnZ[GK]\mathbb{Z}/p^{n}\mathbb{Z}[G_{K}]–modules.

  3. (3)

    We have MBK𝔖Ainf=MBKWn(k)[[u]]Wn(𝒪CK)Minf,M_{\mathrm{BK}}\otimes_{\mathfrak{S}}A_{\mathrm{inf}}=M_{\mathrm{BK}}\otimes_{W_{n}(k)[[u]]}W_{n}(\mathcal{O}_{\mathbb{C}_{K}^{\flat}})\simeq M_{\inf}, and the natural map MBKMinfM_{\mathrm{BK}}\hookrightarrow M_{\inf} has the image contained in MinfGM_{\inf}^{G_{\infty}}.

So let M0BK=HiΔ,1(𝒳/𝔖)M^{0}_{\mathrm{BK}}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},1}(\mathscr{X}/\mathfrak{S}) and Minf0=HiΔ,1(𝒳Ainf/Ainf)M_{\inf}^{0}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},1}(\mathscr{X}_{A_{\mathrm{inf}}}/A_{\mathrm{inf}}), so that T1(MBK0)=T1inf(Minf0)=Héti(𝒳η¯,Z/pZ).T_{1}(M_{\mathrm{BK}}^{0})=T_{1}^{\inf}(M_{\inf}^{0})=\mathrm{H}_{\text{\'{e}t}}^{i}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z}). Observe further that, since uu is a unit of W1(CK)=CK,W_{1}(\mathbb{C}_{K}^{\flat})=\mathbb{C}_{K}^{\flat}, we have T1(MBK0)=T1(MBK)T_{1}(M_{\mathrm{BK}}^{0})=T_{1}(M_{\mathrm{BK}}) and T1inf(Minf0)=T1inf(Minf),T_{1}^{\inf}(M_{\inf}^{0})=T_{1}^{\inf}(M_{\inf}), where MBK=MBK0/MBK0[u]M_{\mathrm{BK}}=M_{\mathrm{BK}}^{0}/M_{\mathrm{BK}}^{0}[u^{\infty}] and Minf=Minf0/Minf0[u]M_{\inf}=M_{\inf}^{0}/M_{\inf}^{0}[u^{\infty}] are again a Breuil–Kisin module and a Breuil–Kisin–Fargues GKG_{K}–module, resp., of height i\leq i. Since 𝔖Ainf\mathfrak{S}\hookrightarrow A_{\mathrm{inf}} is faithfully flat, it is easy to see that the isomorphism MinfMBK𝔖AinfM_{\inf}\simeq M_{\mathrm{BK}}\otimes_{\mathfrak{S}}A_{\mathrm{inf}} remains true. Furthermore, the pair (MBK,Minf)(M_{\mathrm{BK}},M_{\inf}) satisfies the conditions

(5.6) gGsxMBK:g(x)xφ1(v)upsMinf\forall g\in G_{s}\;\;\forall x\in M_{\mathrm{BK}}:\;\;g(x)-x\in\varphi^{-1}(v)u^{p^{s}}M_{\inf}

for all s0s\geq 0, since the pair (MBK0,Minf0)(M_{\mathrm{BK}}^{0},M_{\inf}^{0}) satisfies the analogous conditions by Proposition 4.8. Finally, the module MBKM_{\mathrm{BK}} is finitely generated and uu–torsion–free k[[u]]k[[u]]–module, hence a finite free k[[u]]k[[u]]–module (and, consequently, MinfM_{\inf} is a finite free 𝒪CK\mathcal{O}_{\mathbb{C}_{K}^{\flat}}–module).

Instead of using T=T1(Minf)=Héti(𝒳η¯,Z/pnZ)T^{\prime}=T_{1}(M_{\inf})=\mathrm{H}_{\text{\'{e}t}}^{i}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p^{n}\mathbb{Z}) directly, we work with the dual module

T:=T,inf1(Minf)=HomAinf,φ(Minf,𝒪CK)Héti(𝒳η¯,Z/pZ)T:=T^{*,\inf}_{1}(M_{\inf})=\mathrm{Hom}_{A_{\mathrm{inf}},\varphi}(M_{\inf},\mathcal{O}_{\mathbb{C}_{K}^{\flat}})\simeq\mathrm{H}_{\text{\'{e}t}}^{i}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p\mathbb{Z})^{\vee}

instead; this is equivalent, as the splitting field of TT is still LL. Note that

TT1(MBK)=Hom𝔖,φ(MBK,𝒪CK)T\simeq T^{*}_{1}(M_{\mathrm{BK}})=\mathrm{Hom}_{\mathfrak{S},\varphi}(M_{\mathrm{BK}},\mathcal{O}_{\mathbb{C}_{K}^{\flat}})

as a Z/pZ[G]\mathbb{Z}/p\mathbb{Z}[G_{\infty}]–module.

Remark 5.6 (Ramification bounds of [Caruso]).

Similarly to the discussion above we may take, for any n1,n\geq 1, M0BK=HiΔ,n(𝒳/𝔖),M^{0}_{\mathrm{BK}}=\mathrm{H}^{i}_{{{\mathbbl{\Delta}}},n}(\mathscr{X}/\mathfrak{S}), and MBK=M0BK/M0BK[u]M_{\mathrm{BK}}=M^{0}_{\mathrm{BK}}/M^{0}_{\mathrm{BK}}[u^{\infty}]. Then the GG_{\infty}–module

T:=Tn(MBK)=Hom𝔖,φ(MBK,Wn(𝒪CK))T:=T^{*}_{n}(M_{\mathrm{BK}})=\mathrm{Hom}_{\mathfrak{S},\varphi}(M_{\mathrm{BK}},W_{n}(\mathcal{O}_{\mathbb{C}_{K}^{\flat}}))

is the restriction of Hiét(𝒳η¯,Z/pnZ)\mathrm{H}^{i}_{\text{\'{e}t}}(\mathscr{X}_{\overline{\eta}},\mathbb{Z}/p^{n}\mathbb{Z})^{\vee} to GG_{\infty}. Denoting by 𝒪\mathcal{O}_{\mathcal{E}} the pp–adic completion of 𝔖[1/u]\mathfrak{S}[1/u], M:=MBK𝔖𝒪M_{\mathcal{E}}:=M_{\mathrm{BK}}\otimes_{\mathfrak{S}}\mathcal{O}_{\mathcal{E}} then becomes an étale φ\varphi–module over 𝒪\mathcal{O}_{\mathcal{E}} in the sense of [Fontaine3, §A], with the natural map MBKMM_{\mathrm{BK}}\rightarrow M_{\mathcal{E}} injective; thus, in terminology of [Caruso], MBKM_{\mathrm{BK}} serves as a φ\varphi–lattice of height dividing E(u)iE(u)^{i}. Upon observing that TT is the GG_{\infty}–respresentation associated with MM_{\mathcal{E}} (see e.g. [Caruso, §2.1.3]), Theorem 2 of [Caruso] implies the ramification bound

μL/K1+c0(K)+e(s0(K)+logp(ip))+ep1.\mu_{L/K}\leq 1+c_{0}(K)+e\left(s_{0}(K)+\mathrm{log}_{p}(ip)\right)+\frac{e}{p-1}.

Here c0(K),s0(K)c_{0}(K),s_{0}(K) are constants that depend on the field KK and that generally grow with increasing ee. (Their precise meaning is described in § LABEL:subsec:Comparisons.)

We employ the following approximations of the functors T1T_{1}^{*} and T1,infT_{1}^{*,\inf}.

Notation 5.7.

For a real number c0c\geq 0, we define

Jc(MBK)=Hom𝔖,φ(MBK,𝒪CK/𝔞>c),J_{c}(M_{\mathrm{BK}})=\mathrm{Hom}_{\mathfrak{S},\varphi}(M_{\mathrm{BK}},\mathcal{O}_{\mathbb{C}_{K}^{\flat}}/\mathfrak{a}^{>c}),
Jinfc(Minf)=HomAinf,φ(Minf,𝒪CK/𝔞>c).J^{\inf}_{c}(M_{\inf})=\mathrm{Hom}_{A_{\mathrm{inf}},\varphi}(M_{\inf},\mathcal{O}_{\mathbb{C}_{K}^{\flat}}/\mathfrak{a}^{>c}).

We further set J(MBK)=T1(MBK)J_{\infty}(M_{\mathrm{BK}})=T_{1}^{*}(M_{\mathrm{BK}}) and Jinf(Minf)=T1,inf(Minf)J^{\inf}_{\infty}(M_{\inf})=T_{1}^{*,\inf}(M_{\inf}). Given c,dR0{}c,d\in\mathbb{R}^{\geq 0}\cup\{\infty\} with cd,c\geq d, we denote by ρc,d:Jc(MBK)Jd(MBK)\rho_{c,d}:J_{c}(M_{\mathrm{BK}})\rightarrow J_{d}(M_{\mathrm{BK}}) (ρinfc,d:Jinfc(Minf)Jinfimum