Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-19T12:11:47.290Z Has data issue: false hasContentIssue false

Corrigendum: Around $\boldsymbol{\ell }$-independence

Published online by Cambridge University Press:  29 May 2020

Bruno Chiarellotto
Affiliation:
Dipartimento di Matematica ‘Tullio Levi-Civita’, Università Degli Studi di Padova, Via Trieste 63, 35121Padova, Italia email chiarbru@math.unipd.it
Christopher Lazda
Affiliation:
Warwick Mathematics Institute, Zeeman Building, University of Warwick, Coventry, CV4 7AL, UK email chris.lazda@warwick.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

We correct the proof of the main $\ell$-independence result of the above-mentioned paper by showing that for any smooth and proper variety over an equicharacteristic local field, there exists a globally defined such variety with the same ($p$-adic and $\ell$-adic) cohomology.

Type
Corrigendum
Copyright
© The Authors 2020

1 Introduction

It was pointed out to us by Zheng that the proof of [Reference Chiarellotto and LazdaCL18, Theorem 6.1] is invalid. The problem is in the final step of the proof on p. 237, where we showed that there was an exact sequence

$$\begin{eqnarray}0\rightarrow H_{\ell }^{i+n}(X)\rightarrow H_{\ell }^{i+n}(X_{0})\rightarrow H_{\ell }^{i+n}(X_{1})\rightarrow \cdots\end{eqnarray}$$

and claimed to deduce $\ell$-independence of $H_{\ell }^{i}(X)$ from $\ell$-independence of all the other terms $H_{\ell }^{i+n}(X_{n})$. Of course, this deduction does not work, since there might be infinitely many such other terms.

In their paper [Reference Lu and ZhengLZ19], Lu and Zheng provide (amongst other things) an alternative proof of this $\ell$-independence result, at least for $\ell \neq p$, see Theorem 1.4(2). In this corrigendum we will explain how to fix the proof of [Reference Chiarellotto and LazdaCL18, Theorem 6.1] by instead proving a stronger version of [Reference Chiarellotto and LazdaCL18, Corollary 5.5] where the semistable hypothesis is removed. In particular, this includes the case $\ell =p$.

Notation and conventions

We will use notation from [Reference Chiarellotto and LazdaCL18] freely.

2 Log structures

We begin with a general result on semistable reduction and log schemes. Let $R$ be a complete discrete valuation ring (DVR) with perfect residue field $k$, $\unicode[STIX]{x1D70B}$ a uniformiser for $R$, and let ${\mathcal{X}}\rightarrow \text{Spec}(R)$ be a strictly semistable scheme. That is, ${\mathcal{X}}$ is Zariski locally étale over $R[x_{1},\ldots ,x_{n}]/(x_{1}\cdots x_{r}-\unicode[STIX]{x1D70B})$ for some $n,r$. There is a natural log structure ${\mathcal{M}}_{{\mathcal{X}}}$ on ${\mathcal{X}}$ given by functions invertible outside the special fibre $X$, and we let ${\mathcal{M}}_{X}$ denote the pull-back of this log structure to $X$. We will also write $X_{i}$ for the reduction of ${\mathcal{X}}$ modulo $\unicode[STIX]{x1D70B}^{i+1}$, and $k^{\times }$ for $k$ equipped with the log structure pulled back from the canonical log structure $R^{\times }$ on $R$.

Proposition 2.1 (Illusie, Nakayama [Reference NakayamaNak98, Appendix A.4]).

If ${\mathcal{X}},{\mathcal{X}}^{\prime }$ are strictly semistable schemes over $R$, and $g:X_{1}\rightarrow X_{1}^{\prime }$ is an isomorphism between their mod $\unicode[STIX]{x1D70B}^{2}$-reductions, then $g$ induces a canonical isomorphism $g:(X,{\mathcal{M}}_{X})\overset{{\sim}}{\rightarrow }(X^{\prime },{\mathcal{M}}_{X^{\prime }})$ of log schemes over $k^{\times }$.

Sketch of proof.

Use $g$ to identify $X_{1}$ and $X_{1}^{\prime }$, and thus $X$ and $X^{\prime }$. Let ${\mathcal{M}}_{X}$ and ${\mathcal{M}}_{X}^{\prime }$ be the log structures on $X$ coming from ${\mathcal{X}}$ and ${\mathcal{X}}^{\prime }$ respectively.

Near a closed point of $X$ let $X^{(1)},\ldots ,X^{(r)}$ be the irreducible components of $X$, and pick $x_{1},\ldots ,x_{r}\in {\mathcal{O}}_{{\mathcal{X}}}$ such that $X^{(i)}=V(x_{i})$. Similarly pick $x_{1}^{\prime },\ldots ,x_{r}^{\prime }\in {\mathcal{O}}_{{\mathcal{X}}^{\prime }}$ such that $X^{(i)}=V(x_{i}^{\prime })$. Let $v\in {\mathcal{O}}_{{\mathcal{X}}}^{\ast }$ and $v^{\prime }\in {\mathcal{O}}_{{\mathcal{X}}^{\prime }}^{\ast }$ be such that $x_{1}\cdots x_{r}=v\unicode[STIX]{x1D70B}$ and $x_{1}^{\prime }\cdots x_{r}^{\prime }=v^{\prime }\unicode[STIX]{x1D70B}$. Then in a neighbourhood of $p$ the morphisms $({\mathcal{X}},{\mathcal{M}}_{{\mathcal{X}}})\rightarrow \text{Spec}(R^{\times })$ and $({\mathcal{X}}^{\prime },{\mathcal{M}}_{{\mathcal{X}}^{\prime }})\rightarrow \text{Spec}(R^{\times })$ can be described by the following diagrams:

Pulling back to $k$, we see that the morphisms $(X,{\mathcal{M}}_{X})\rightarrow \text{Spec}(k^{\times })$ and $(X,{\mathcal{M}}_{X}^{\prime })\rightarrow \text{Spec}(k^{\times })$ can be described by the diagrams

and

respectively, again in a neighbourhood of $p$. Since $V(x_{i})=V(x_{i}^{\prime })$ inside $X_{1}$, we must have $x_{i}=u_{i}x_{i}^{\prime }$ for some $u_{i}\in {\mathcal{O}}_{X_{1}}^{\ast }$, and so we can define an isomorphism

$$\begin{eqnarray}{\mathcal{M}}_{X}\overset{{\sim}}{\rightarrow }{\mathcal{M}}_{X}^{\prime }\end{eqnarray}$$

of log structures by mapping

$$\begin{eqnarray}(u,a_{1},\ldots ,a_{r})\mapsto (uu_{1}^{a_{1}}\cdots u_{r}^{a_{r}},a_{1},\ldots ,a_{r}).\end{eqnarray}$$

This is checked to be a morphism of log structures over $k^{\times }$ by using the above local descriptions. Note that any other choice $u_{i}^{\prime }$ must satisfy $(u_{i}-u_{i}^{\prime })x_{i}^{\prime }=0$ in ${\mathcal{O}}_{X_{1}}$, and hence we must have $u_{i}-u_{i}^{\prime }\in (\unicode[STIX]{x1D70B})$. In particular, the above isomorphism does not depend on the choice of $u_{i}$. By a similar argument, neither does it depend on the choice of $x_{i}$ and $x_{i}^{\prime }$, and so it glues to give a global isomorphism $(X,{\mathcal{M}}_{X})\cong (X,{\mathcal{M}}_{X}^{\prime })$ of log schemes over $k^{\times }$.◻

We will need to extend this result to cover morphisms between strictly semistable schemes over different bases. So suppose that $R\rightarrow S$ is a finite morphism of complete DVRs, with induced residue field extension $k\rightarrow k_{S}$. Let $\unicode[STIX]{x1D70B}_{S}$ be a uniformiser for $S$, and let $e=v_{\unicode[STIX]{x1D70B}_{S}}(\unicode[STIX]{x1D70B})$. We do not assume that the induced extension $Q(R)\rightarrow Q(S)$ of fraction fields is separable.

Suppose that we have strictly semistable schemes ${\mathcal{X}},{\mathcal{X}}^{\prime }$ over $R$ and ${\mathcal{Y}},{\mathcal{Y}}^{\prime }$ over $S$, and a pair of commutative diagrams

As before, let us write $Y_{j}$ for the reduction of ${\mathcal{Y}}$ modulo $\unicode[STIX]{x1D70B}_{S}^{j+1}$. Suppose that we have isomorphisms

$$\begin{eqnarray}g_{Y}:Y_{e}\overset{{\sim}}{\rightarrow }Y_{e}^{\prime },\quad g_{X}:X_{1}\overset{{\sim}}{\rightarrow }X_{1}^{\prime }\end{eqnarray}$$

of $S$- and $R$-schemes respectively such that the diagram

commutes. Then by Proposition 2.1 we obtain isomorphisms

$$\begin{eqnarray}g_{Y}:(Y,{\mathcal{M}}_{Y})\overset{{\sim}}{\rightarrow }(Y^{\prime },{\mathcal{M}}_{Y^{\prime }})\end{eqnarray}$$

of log schemes over $k_{S}^{\times }$, as well as

$$\begin{eqnarray}g_{X}:(X,{\mathcal{M}}_{X})\overset{{\sim}}{\rightarrow }(X^{\prime },{\mathcal{M}}_{X^{\prime }})\end{eqnarray}$$

of log schemes over $k^{\times }$. The above commutative diagrams of strictly semistable schemes induce commutative diagrams

of log schemes. Note that the morphism of punctured points along the bottom of each square is given by

$$\begin{eqnarray}\displaystyle k^{\ast }\oplus \mathbb{N} & \rightarrow & \displaystyle k_{S}^{\ast }\oplus \mathbb{N}\nonumber\\ \displaystyle (\unicode[STIX]{x1D706},a) & \mapsto & \displaystyle (\unicode[STIX]{x1D706}u^{a},ea),\nonumber\end{eqnarray}$$

where $u\in S^{\ast }$ is such that $\unicode[STIX]{x1D70B}=u\unicode[STIX]{x1D70B}_{S}^{e}$.

Proposition 2.2. The diagram

of log schemes commutes.

Proof. Let us use $g$ to identify $Y_{e}=Y_{e}^{\prime }$ and $Y=Y^{\prime }$, and let ${\mathcal{M}}_{Y}$ and ${\mathcal{M}}_{Y}^{\prime }$ be the log structures on $Y$ coming from ${\mathcal{Y}}$ and ${\mathcal{Y}}^{\prime }$ respectively. Similarly identify $X_{1}=X_{1}^{\prime }$ and $X=X^{\prime }$, and let ${\mathcal{M}}_{X}$ and ${\mathcal{M}}_{X}^{\prime }$ be the log structures on $X$ coming from ${\mathcal{X}}$ and ${\mathcal{X}}^{\prime }$ respectively.

Locally on $X$ and $Y$, choose functions $y_{1},\ldots ,y_{s}\in {\mathcal{O}}_{{\mathcal{Y}}}$, $y_{1}^{\prime },\ldots ,y_{s}^{\prime }\in {\mathcal{O}}_{{\mathcal{Y}}^{\prime }}$ cutting out the irreducible components of $Y$, and functions $x_{1},\ldots ,x_{r}\in {\mathcal{O}}_{{\mathcal{X}}}$ and $x_{1}^{\prime },\ldots ,x_{r}^{\prime }\in {\mathcal{O}}_{{\mathcal{X}}^{\prime }}$ cutting out the irreducible components of $X$. Write

$$\begin{eqnarray}f^{\ast }(x_{i})=\unicode[STIX]{x1D6FC}_{i}y_{1}^{d_{i1}}\cdots y_{s}^{d_{is}},\quad {f^{\prime }}^{\ast }(x_{i}^{\prime })=\unicode[STIX]{x1D6FC}_{i}^{\prime }{y_{1}^{\prime }}^{d_{i1}^{\prime }}\cdots {y_{s}^{\prime }}^{d_{is}^{\prime }},\end{eqnarray}$$

since both $d_{ij}$ and $d_{ij}^{\prime }$ are given by the multiplicity of the $j$th irreducible component of $Y$ in the scheme theoretic preimage of the $i$th irreducible component of $X$ inside $Y_{e}$, we must have $d_{ij}=d_{ij}^{\prime }$. Moreover, since $V(f^{\ast }(x_{i}))\subset V(\unicode[STIX]{x1D70B}_{S}^{e})=V(y_{1}^{e}\cdots y_{s}^{e})$ we must have $d_{ij}\leqslant e$ for all $i,j$.

Now choose $u_{i}\in {\mathcal{O}}_{X_{1}}^{\ast }$ such that $x_{i}=u_{i}x_{i}^{\prime }$, and $v_{j}\in {\mathcal{O}}_{Y_{e}}^{\ast }$ such that $y_{j}=v_{j}y_{j}^{\prime }$. Then the isomorphisms of log structures induced by $g_{Y}$ and $g_{X}$ are given by

$$\begin{eqnarray}\displaystyle {\mathcal{M}}_{Y}={\mathcal{O}}_{Y}^{\ast }\oplus \mathbb{N}^{s} & \rightarrow & \displaystyle {\mathcal{M}}_{Y}^{\prime }={\mathcal{O}}_{Y}^{\ast }\oplus \mathbb{N}^{s}\nonumber\\ \displaystyle (v,b_{1},\ldots ,b_{s}) & \mapsto & \displaystyle (vv_{1}^{b_{1}}\cdots v_{s}^{b_{s}},b_{1},\ldots ,b_{s})\nonumber\end{eqnarray}$$

and

$$\begin{eqnarray}\displaystyle {\mathcal{M}}_{X}={\mathcal{O}}_{X}^{\ast }\oplus \mathbb{N}^{r} & \rightarrow & \displaystyle {\mathcal{M}}_{X}^{\prime }={\mathcal{O}}_{X}^{\ast }\oplus \mathbb{N}^{r}\nonumber\\ \displaystyle (u,a_{1},\ldots ,a_{r}) & \mapsto & \displaystyle (uu_{1}^{a_{1}}\cdots u_{r}^{a_{r}},a_{1},\ldots ,a_{r})\nonumber\end{eqnarray}$$

respectively, and the morphisms ${\mathcal{M}}_{X}\rightarrow {\mathcal{M}}_{Y}$ and ${\mathcal{M}}_{X}^{\prime }\rightarrow {\mathcal{M}}_{Y}^{\prime }$ are defined by

$$\begin{eqnarray}(u,a_{1},\ldots ,a_{r})\mapsto \biggl(f^{\ast }(u)\unicode[STIX]{x1D6FC}_{1}^{a_{1}}\cdots \unicode[STIX]{x1D6FC}_{r}^{a_{r}},\mathop{\sum }_{i}d_{i1}a_{i},\ldots ,\mathop{\sum }_{i}d_{is}a_{i}\biggr)\end{eqnarray}$$

and

$$\begin{eqnarray}(u,a_{1},\ldots ,a_{r})\mapsto \biggl(f^{\ast }(u){\unicode[STIX]{x1D6FC}_{1}^{\prime }}^{a_{1}}\cdots {\unicode[STIX]{x1D6FC}_{r}^{\prime }}^{a_{r}},\mathop{\sum }_{i}d_{i1}a_{i},\ldots ,\mathop{\sum }_{i}d_{is}a_{i}\biggr)\end{eqnarray}$$

respectively. Hence in the diagram

the composite $f\circ g_{X}$ is given by

$$\begin{eqnarray}(u,a_{1},\ldots ,a_{r})\mapsto \biggl(f^{\ast }(u)(\unicode[STIX]{x1D6FC}_{1}^{\prime }f^{\ast }(u_{1}))^{a_{1}}\cdots (\unicode[STIX]{x1D6FC}_{r}^{\prime }f^{\ast }(u_{r}))^{a_{r}},\mathop{\sum }_{i}d_{i1}a_{i},\ldots ,\mathop{\sum }_{i}d_{is}a_{i}\biggr)\end{eqnarray}$$

and the composite $g_{Y}\circ f$ is given by

$$\begin{eqnarray}(u,a_{1},\ldots ,a_{r})\mapsto \biggl(f^{\ast }(u)(\unicode[STIX]{x1D6FC}_{1}v_{1}^{d_{11}}\cdots v_{s}^{d_{1s}})^{a_{1}}\cdots (\unicode[STIX]{x1D6FC}_{r}v_{1}^{d_{r1}}\cdots v_{s}^{d_{rs}})^{a_{r}},\mathop{\sum }_{i}d_{i1}a_{i},\ldots ,\mathop{\sum }_{i}d_{is}a_{i}\biggr).\end{eqnarray}$$

We thus need to show that $\unicode[STIX]{x1D6FC}_{i}^{\prime }f^{\ast }(u_{i})=\unicode[STIX]{x1D6FC}_{i}v_{1}^{d_{i1}}\cdots v_{s}^{d_{is}}$ in ${\mathcal{O}}_{Y}^{\ast }$ for all $i$. But now we write

$$\begin{eqnarray}\unicode[STIX]{x1D6FC}_{i}y_{1}^{d_{i1}}\cdots y_{s}^{d_{is}}=f^{\ast }(x_{i})=f^{\ast }(u_{i}x_{i}^{\prime })=f^{\ast }(u_{i})\unicode[STIX]{x1D6FC}_{i}^{\prime }{y_{1}^{\prime }}^{d_{i1}}\cdots {y_{s}^{\prime }}^{d_{is}}\end{eqnarray}$$

in ${\mathcal{O}}_{Y_{e}}$ and so deduce that

$$\begin{eqnarray}\unicode[STIX]{x1D6FC}_{i}v_{1}^{d_{i1}}\cdots v_{s}^{d_{is}}{y_{1}^{\prime }}^{d_{i1}}\cdots {y_{s}^{\prime }}^{d_{is}}=f^{\ast }(u_{i})\unicode[STIX]{x1D6FC}_{i}^{\prime }{y_{1}^{\prime }}^{d_{i1}}\cdots {y_{s}^{\prime }}^{d_{is}}.\end{eqnarray}$$

We deduce that the difference $\unicode[STIX]{x1D6FD}_{i}=\unicode[STIX]{x1D6FC}_{i}^{\prime }f^{\ast }(u_{i})-\unicode[STIX]{x1D6FC}_{i}v_{1}^{d_{i1}}\cdots v_{s}^{d_{is}}$ annihilates ${y_{1}^{\prime }}^{d_{i1}}\cdots {y_{s}^{\prime }}^{d_{is}}$ inside ${\mathcal{O}}_{Y_{e}}$, and since each $d_{ij}\leqslant e$ we deduce that in fact $\unicode[STIX]{x1D6FD}_{i}$ annihilates $\unicode[STIX]{x1D70B}_{S}^{e}$, and therefore must lie in $(\unicode[STIX]{x1D70B}_{S})$. Hence $\unicode[STIX]{x1D6FD}_{i}=0$ in ${\mathcal{O}}_{Y}$ and the proof is complete.◻

3 Functoriality of comparison isomorphisms

We will also need to know that the comparison isomorphisms [Reference Chiarellotto and LazdaCL18, Propositions 5.3, 5.4] are compatible with morphisms of semistable schemes over different bases. So let us suppose that we are again in the above set-up, where we have a commutative diagram

of strictly semistable schemes ${\mathcal{Y}}$ and ${\mathcal{X}}$ over $S$ and $R$ respectively, with $S$ the integral closure of $R$ in some finite extension of its fraction field. Let us assume that $R$, and hence $S$, is of equicharacteristic $p>0$, with fraction fields $F$ and $F_{S}$ respectively, whose absolute Galois groups we will denote by $G_{F}$ and $G_{F_{S}}$. Fix an embedding $F^{\text{sep}}{\hookrightarrow}F_{S}^{\text{sep}}$ of separable closures; note that this sends $F^{\text{tame}}$ into $F_{S}^{\text{tame}}$ and induces an injective homomorphism $G_{F_{S}}\rightarrow G_{F}$ with finite cokernel.

Let ${\mathcal{X}}^{\times }$ and ${\mathcal{Y}}^{\times }$ denote these semistable schemes endowed with their canonical log structures, and $X^{\times }$ and $Y^{\times }$ the corresponding log special fibres. We therefore have a commutative diagram

of log schemes. For every finite subextension $F\subset L\subset F^{\text{tame}}$, let $X_{L}^{\times }$ denote the corresponding base change of $X^{\times }$, and $X^{\times ,\text{tame}}$ the inverse limit of the étale topoi of all such $X_{L}^{\times }$; we have $Y^{\times ,\text{tame}}$ defined entirely similarly. Via the embedding $F^{\text{tame}}{\hookrightarrow}F_{S}^{\text{tame}}$ this induces a $G_{F_{S}}$-equivariant morphism of topoi

$$\begin{eqnarray}Y^{\times ,\text{tame}}\rightarrow X^{\times ,\text{tame}}\end{eqnarray}$$

and hence a $G_{F_{S}}$-equivariant morphism

$$\begin{eqnarray}H_{\acute{\text{e}}\text{t}}^{i}(X^{\times ,\text{tame}},\mathbb{Q}_{\ell })\rightarrow H_{\acute{\text{e}}\text{t}}^{i}(Y^{\times ,\text{tame}},\mathbb{Q}_{\ell })\end{eqnarray}$$

in cohomology, for any $\ell \neq p$. On the other hand we have a natural $G_{F_{S}}$-equivariant map

$$\begin{eqnarray}H_{\acute{\text{e}}\text{t}}^{i}({\mathcal{X}}\times _{R}F^{\text{sep}},\mathbb{Q}_{\ell })\rightarrow H_{\acute{\text{e}}\text{t}}^{i}({\mathcal{Y}}\times _{S}F_{S}^{\text{sep}},\mathbb{Q}_{\ell }),\end{eqnarray}$$

and by [Reference NakayamaNak98, Proposition 4.2] equivariant isomorphisms

$$\begin{eqnarray}\displaystyle H_{\acute{\text{e}}\text{t}}^{i}(X^{\times ,\text{tame}},\mathbb{Q}_{\ell }) & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\acute{\text{e}}\text{t}}^{i}({\mathcal{X}}\times _{R}F^{\text{sep}},\mathbb{Q}_{\ell }),\nonumber\\ \displaystyle H_{\acute{\text{e}}\text{t}}^{i}(Y^{\times ,\text{tame}},\mathbb{Q}_{\ell }) & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\acute{\text{e}}\text{t}}^{i}({\mathcal{Y}}\times _{S}F_{S}^{\text{sep}},\mathbb{Q}_{\ell }).\nonumber\end{eqnarray}$$

Proposition 3.1. The diagram

commutes.

Proof. Consider the commutative diagram

of topoi as in [Reference NakayamaNak98, §3], where ${\mathcal{Y}}^{\times ,\text{tame}}$ and ${\mathcal{X}}^{\times ,\text{tame}}$ are defined by ‘base change’ along $F_{S}\rightarrow F_{S}^{\text{tame}}$ and $F\rightarrow F^{\text{tame}}$ respectively. Then the isomorphism

$$\begin{eqnarray}H_{\acute{\text{e}}\text{t}}^{i}(Y^{\times ,\text{tame}},\mathbb{Q}_{\ell })\overset{{\sim}}{\rightarrow }H_{\acute{\text{e}}\text{t}}^{i}({\mathcal{Y}}\times _{S}F_{S}^{\text{sep}},\mathbb{Q}_{\ell })\end{eqnarray}$$

is given as the composite

$$\begin{eqnarray}H_{\acute{\text{e}}\text{t}}^{i}(Y^{\times ,\text{tame}},\mathbb{Q}_{\ell })\overset{{\sim}}{\leftarrow }H_{\acute{\text{e}}\text{t}}^{i}({\mathcal{Y}}^{\times ,\text{tame}},\mathbb{Q}_{\ell })\rightarrow H_{\acute{\text{e}}\text{t}}^{i}({\mathcal{Y}}\times _{S}F_{S}^{\text{sep}},\mathbb{Q}_{\ell })\end{eqnarray}$$

using the proper base change theorem in log-étale cohomology [Reference NakayamaNak97, Theorem 5.1], and there is a similar statement for ${\mathcal{X}}$. The claim then follows simply from commutativity of the above diagram of log schemes.◻

We will also need a version of this result for $p$-adic cohomology. Write $W=W(k)$, $W_{S}=W(k_{S})$, let $K=W[1/p]$, $K_{S}=W_{S}[1/p]$, and let ${\mathcal{R}}_{K}\supset {\mathcal{E}}_{K}^{\dagger }\subset {\mathcal{E}}_{K}$, and ${\mathcal{R}}_{K_{S}}\supset {\mathcal{E}}_{K_{S}}^{\dagger }\subset {\mathcal{E}}_{K_{S}}$ denote copies of the Robba ring, the bounded Robba ring and the Amice ring over $K$ and $K_{S}$ respectively. Lift the extension $F\rightarrow F_{S}$ to a finite flat morphism ${\mathcal{E}}_{K}^{\dagger }\rightarrow {\mathcal{E}}_{K_{S}}^{\dagger }$ which extends to finite flat morphisms ${\mathcal{R}}_{K}\rightarrow {\mathcal{R}}_{K_{S}}$ and ${\mathcal{E}}_{K}\rightarrow {\mathcal{E}}_{K_{S}}$. Then, as above, the morphism of log schemes $Y^{\times }\rightarrow X^{\times }$ induces a morphism

$$\begin{eqnarray}H_{\text{log}\text{-}\text{cris}}^{i}(X^{\times }/K^{\times })\rightarrow H_{\text{log}\text{-}\text{cris}}^{i}(Y^{\times }/K_{S}^{\times })\end{eqnarray}$$

in log crystalline cohomology, and the morphism ${\mathcal{Y}}_{F_{S}}\rightarrow {\mathcal{X}}_{F}$ induces a morphism

$$\begin{eqnarray}H_{\text{rig}}^{i}({\mathcal{X}}_{F}/{\mathcal{R}}_{K})\rightarrow H_{\text{rig}}^{i}({\mathcal{Y}}_{F_{S}}/{\mathcal{R}}_{K_{S}})\end{eqnarray}$$

in Robba-ring valued rigid cohomology. Then following [Reference Chiarellotto and LazdaCL18, Proposition 5.4] we can construct isomorphisms

$$\begin{eqnarray}\displaystyle H_{\text{log}\text{-}\text{cris}}^{i}(X^{\times }/K^{\times })\otimes _{K}{\mathcal{R}}_{K} & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\text{rig}}^{i}({\mathcal{X}}_{F}/{\mathcal{R}}_{K}),\nonumber\\ \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}(Y^{\times }/K_{S}^{\times })\otimes _{K_{S}}{\mathcal{R}}_{K_{S}} & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\text{rig}}^{i}({\mathcal{Y}}_{F_{S}}/{\mathcal{R}}_{K_{S}})\nonumber\end{eqnarray}$$

as follows. Let $t$ denote a co-ordinate on ${\mathcal{E}}_{K}^{\dagger }$ and $t_{S}$ a co-ordinate on ${\mathcal{E}}_{K_{S}}^{\dagger }$ such that $t\in W_{S}\unicode[STIX]{x27E6}t_{S}\unicode[STIX]{x27E7}$. Write $S_{K}=K\otimes W\unicode[STIX]{x27E6}t\unicode[STIX]{x27E7}$ and $S_{K_{S}}=K_{S}\otimes W_{S}\unicode[STIX]{x27E6}t_{S}\unicode[STIX]{x27E7}$. Equip $W\unicode[STIX]{x27E6}t\unicode[STIX]{x27E7}$ (respectively $W_{S}\unicode[STIX]{x27E6}t_{S}\unicode[STIX]{x27E7}$) with the log structure defined by the ideal $(t)\subset W\unicode[STIX]{x27E6}t\unicode[STIX]{x27E7}$ (respectively $(t_{S})\subset W\unicode[STIX]{x27E6}t_{S}\unicode[STIX]{x27E7}$) and define the log-crystalline cohomology groups

$$\begin{eqnarray}\displaystyle H_{\text{log}\text{-}\text{cris}}^{i}({\mathcal{X}}^{\times }/S_{K}) & := & \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}({\mathcal{X}}^{\times }/W\unicode[STIX]{x27E6}t\unicode[STIX]{x27E7})\otimes _{\mathbb{Z}}\mathbb{Q},\nonumber\\ \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}({\mathcal{Y}}^{\times }/S_{K_{S}}) & := & \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}({\mathcal{Y}}^{\times }/W_{S}\unicode[STIX]{x27E6}t_{S}\unicode[STIX]{x27E7})\otimes _{\mathbb{Z}}\mathbb{Q};\nonumber\end{eqnarray}$$

these are naturally endowed with the extra structure of $\text{log}\text{-}(\unicode[STIX]{x1D711},\unicode[STIX]{x1D6FB})$-modules over $S_{K}$ and $S_{K_{S}}$ respectively. Moreover, we have isomorphisms of $\unicode[STIX]{x1D711}$-modules

$$\begin{eqnarray}\displaystyle H_{\text{log}\text{-}\text{cris}}^{i}({\mathcal{X}}^{\times }/S_{K})\otimes _{S_{K},t\mapsto 0}K & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}(Y^{\times }/K_{S}^{\times }),\nonumber\\ \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}({\mathcal{Y}}^{\times }/S_{K_{S}})\otimes _{S_{K_{S}},t_{S}\mapsto 0}K_{S} & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}(Y^{\times }/K_{S}^{\times }),\nonumber\end{eqnarray}$$

by smooth and proper base change in log-crystalline cohomology, as well as isomorphisms of $(\unicode[STIX]{x1D711},\unicode[STIX]{x1D6FB})$-modules

$$\begin{eqnarray}\displaystyle H_{\text{log}\text{-}\text{cris}}^{i}({\mathcal{X}}^{\times }/S_{K})\otimes _{S_{K}}{\mathcal{R}}_{K} & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\text{rig}}^{i}({\mathcal{X}}_{F}/{\mathcal{R}}_{K}),\nonumber\\ \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}({\mathcal{Y}}^{\times }/S_{K_{S}})\otimes _{S_{K_{S}}}{\mathcal{R}}_{K_{S}} & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\text{rig}}^{i}({\mathcal{Y}}_{F_{S}}/{\mathcal{R}}_{K_{S}}),\nonumber\end{eqnarray}$$

by [Reference Lazda and PálLP16, Proposition 5.45]. It therefore follows from the logarithmic form of Dwork’s trick [Reference KedlayaKed10, Corollary 17.2.4] that the $(\unicode[STIX]{x1D711},\unicode[STIX]{x1D6FB})$-modules $H_{\text{rig}}^{i}({\mathcal{X}}_{F}/{\mathcal{R}}_{K})$ and $H_{\text{rig}}^{i}({\mathcal{Y}}_{F_{S}}/{\mathcal{R}}_{K_{S}})$ are unipotent, that there are isomorphisms

$$\begin{eqnarray}\displaystyle \left(H_{\text{rig}}^{i}({\mathcal{X}}_{F}/{\mathcal{R}}_{K})[\log t]\right)^{\unicode[STIX]{x1D6FB}=0} & \cong & \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}(X^{\times }/K^{\times }),\nonumber\\ \displaystyle \left(H_{\text{rig}}^{i}({\mathcal{Y}}_{F_{S}}/{\mathcal{R}}_{K_{S}})[\log t_{S}]\right)^{\unicode[STIX]{x1D6FB}=0} & \cong & \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}(Y^{\times }/K_{S}^{\times })\nonumber\end{eqnarray}$$

and moreover the connection $\unicode[STIX]{x1D6FB}$ on the rigid cohomology groups appearing on the left-hand side can be completely recovered from the monodromy operator $N$ on the right-hand side. This allows us to construct isomorphisms of $(\unicode[STIX]{x1D711},\unicode[STIX]{x1D6FB})$-modules

$$\begin{eqnarray}\displaystyle H_{\text{log}\text{-}\text{cris}}^{i}(X^{\times }/K^{\times })\otimes _{K}{\mathcal{R}}_{K} & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\text{rig}}^{i}({\mathcal{X}}_{F}/{\mathcal{R}}_{K}),\nonumber\\ \displaystyle H_{\text{log}\text{-}\text{cris}}^{i}(Y^{\times }/K_{S}^{\times })\otimes _{K_{S}}{\mathcal{R}}_{K_{S}} & \overset{{\sim}}{\rightarrow } & \displaystyle H_{\text{rig}}^{i}({\mathcal{Y}}_{F_{S}}/{\mathcal{R}}_{K_{S}})\nonumber\end{eqnarray}$$

where the left-hand side is endowed a natural connection coming from $N$; for more details see, for example, [Reference MarmoraMar08, §3.2].

Proposition 3.2. The diagram

commutes.

Proof. Given the construction of the horizontal isomorphisms outlined above, it suffices to show that the diagram

of log-crystalline cohomology groups commutes, which as in Proposition 3.1 simply follows from functoriality of log-crystalline cohomology. ◻

4 Cohomology and global approximations

Now suppose that $k$ is a finite field, $F=k(\!(t)\!)$, and $X/F$ is a smooth and proper variety.

Definition 4.1. We say that $X$ is globally defined if there exist a smooth curve $C/k$, a $k$-valued point $c\in C(k)$, a smooth and proper morphism $\mathbf{X}\rightarrow (C\setminus \{c\})$ and an isomorphism $F\cong \widehat{k(C)}_{c}$ such that $\mathbf{X}_{F}\cong X$.

We will prove the following strengthened version of [Reference Chiarellotto and LazdaCL18, Corollary 5.5].

Theorem 4.2. For any smooth and proper variety $X/F$ there exists a globally defined smooth and proper variety $Z/F$ such that

$$\begin{eqnarray}H_{\ell }^{i}(X)\cong H_{\ell }^{i}(Z)\end{eqnarray}$$

for all $\ell$ (including $\ell =p$).

Once we have shown this, the proof of [Reference Chiarellotto and LazdaCL18, Theorem 6.1] can then be completed using [Reference Chiarellotto and LazdaCL18, Proposition 5.8], exactly as in the proof of [Reference Chiarellotto and LazdaCL18, Theorem 5.1].

To prove Theorem 4.2, first of all choose a proper and flat model ${\mathcal{X}}$ for $X$ over the ring of integers ${\mathcal{O}}_{F}$. By [Reference de JongdJ96, Theorem 6.5] we may choose an alteration ${\mathcal{X}}_{0}\rightarrow {\mathcal{X}}$ and a finite extension $F_{0}/F$ such that ${\mathcal{X}}_{0}$ is strictly semistable over ${\mathcal{O}}_{F_{0}}$.

Next, we take the fibre product ${\mathcal{X}}_{0}\times _{{\mathcal{X}}}{\mathcal{X}}_{0}$, and let ${\mathcal{X}}_{1}^{\prime }$ denote the disjoint union of the reduced, irreducible components of ${\mathcal{X}}_{0}\times _{{\mathcal{X}}}{\mathcal{X}}_{0}$ which are flat over ${\mathcal{O}}_{F_{0}}$, or equivalently which map surjectively to $\text{Spec}({\mathcal{O}}_{F_{0}})$. Once more applying [Reference de JongdJ96, Theorem 6.5] to each of the connected components of ${\mathcal{X}}_{1}^{\prime }$ in turn enables us to produce:

  1. a 2-truncated augmented simplicial scheme

    $$\begin{eqnarray}{\mathcal{X}}_{1}\rightrightarrows {\mathcal{X}}_{0}\rightarrow {\mathcal{X}}\end{eqnarray}$$
    which is a proper hypercover after base changing to $F$;
  2. a collection $F_{1,1},\ldots ,F_{1,s}$ of finite field extensions of $F_{0}$

such that ${\mathcal{X}}_{1}$ is a disjoint union of schemes ${\mathcal{X}}_{1,j}$, for $1\leqslant j\leqslant s$, proper and strictly semistable over $\text{Spec}({\mathcal{O}}_{F_{1,j}})$.

Let $k_{0}$ denote the residue field of $F_{0}$, $k_{1,j}$ the residue field of $F_{1,j}$, and consider the intermediate extensions

$$\begin{eqnarray}F\subset F_{0}^{\text{un}}\subset F_{0}^{s}\subset F_{0}\subset F_{1,j}^{\text{un}}\subset F_{1,j}^{s}\subset F_{1,j},\end{eqnarray}$$

where $F_{0}^{\text{un}}/F$ and $F_{1,j}^{\text{un}}/F_{0}$ are separable and unramified, $F_{0}^{s}/F_{0}^{\text{un}}$ and $F_{1,j}^{s}/F_{1,j}^{\text{un}}$ are separable and totally ramified, and $F_{0}/F_{0}^{s}$ and $F_{1,j}/F_{1,j}^{s}$ are totally inseparable, of degree $p^{d_{0}}$ and $p^{d_{1,j}}$ respectively. Let $t$ denote a uniformiser for $F$, $t_{0}$ one for $F_{0}^{s}$, and let $P_{0}$ be the minimal polynomial of $t_{0}$ over $F_{0}^{\text{un}}$. Then $t_{0}^{\prime }:=t_{0}^{1/p^{d_{0}}}$ is a uniformiser for ${\mathcal{O}}_{F_{0}}$. Similarly, let $t_{1,j}$ be a uniformiser for $F_{1,j}^{s}$, and $P_{1,j}$ the minimal polynomial of $t_{1,j}$ over $F_{1,j}^{\text{un}}$. Then $t_{1,j}^{\prime }:=t_{1,j}^{1/p^{d_{1,j}}}$ is a uniformiser for ${\mathcal{O}}_{F_{1,j}}$.

Now choose a finitely generated sub-$k$-algebra $R\subset {\mathcal{O}}_{F}$, containing $t$, such that there exists a proper, flat scheme ${\mathcal{Y}}\rightarrow \text{Spec}(R)$ whose base change to ${\mathcal{O}}_{F}$ is exactly ${\mathcal{X}}$. By [Reference SpivakovskySpi99, Theorem 10.1], we may at any point increase $R$ to ensure that it is in fact smooth over $k$. Next, enlarge $R$ so that $R_{0}^{\text{un}}:=R\,\otimes _{k}\,k_{0}\subset {\mathcal{O}}_{F_{0}^{\text{un}}}$ contains all the coefficients of the minimal (Eisenstein) polynomial $P_{0}$ of $t_{0}$, and let $R_{0}^{s}$ denote the corresponding finite flat extension $R_{0}^{\text{un}}[x]/(P_{0})$ of $R_{0}^{\text{un}}$. We can thus consider $R_{0}^{s}\subset {\mathcal{O}}_{F_{0}^{s}}$ as a subring containing $t_{0}$, and we set $R_{0}=R_{0}^{s}[t_{0}^{\prime }]$. Hence we have $R_{0}\subset {\mathcal{O}}_{F_{0}}$ such that

$$\begin{eqnarray}R_{0}\otimes _{R}{\mathcal{O}}_{F}\overset{{\sim}}{\rightarrow }{\mathcal{O}}_{F_{0}}.\end{eqnarray}$$

Note also that $R_{0}$ is finite and flat over $R$; after localising $R$ within ${\mathcal{O}}_{F}$ we may in fact assume that $R_{0}$ is finite free over $R$.

Next we enlarge $R$ so that there exists a proper and flat morphism ${\mathcal{Y}}_{0}\rightarrow \text{Spec}(R_{0})$ whose base change to ${\mathcal{O}}_{F_{0}}$ is ${\mathcal{X}}_{0}$. Again, by further enlarging $R$ we may in addition assume that the map ${\mathcal{X}}_{0}\rightarrow {\mathcal{X}}$ arises from a proper surjective map

$$\begin{eqnarray}{\mathcal{Y}}_{0}\rightarrow {\mathcal{Y}}\end{eqnarray}$$

of $R$-schemes, and moreover that there exists an open cover of ${\mathcal{Y}}_{0}$ by schemes which are étale over $R_{0}[x_{1},\ldots ,x_{n}]/(x_{1}\cdots x_{r}-t_{0}^{\prime })$ for some $n,r$. In other words, ${\mathcal{Y}}_{0}$ is ‘strictly $t_{0}^{\prime }$-semistable’.

We now repeat this process to produce further finite free extensions $R_{0}\rightarrow R_{1,j}^{\text{un}}\rightarrow R_{1,j}^{s}\rightarrow R_{1,j}$ for all $j$, and an injection $R_{1,j}\subset {\mathcal{O}}_{F_{1,j}}$ containing the image of $t_{1,j}^{\prime }$ such that

$$\begin{eqnarray}R_{1,j}\otimes _{R}{\mathcal{O}}_{F}\overset{{\sim}}{\rightarrow }{\mathcal{O}}_{F_{1,j}}.\end{eqnarray}$$

We can also find proper, strictly $t_{1,j}^{\prime }$-semistable schemes ${\mathcal{Y}}_{1,j}\rightarrow \text{Spec}(R_{1,j})$ whose base change to ${\mathcal{O}}_{F_{1,j}}$ is ${\mathcal{X}}_{1,j}$, so that setting ${\mathcal{Y}}_{1}:=\coprod _{j}{\mathcal{Y}}_{1,j}$ (and again, possibly increasing $R$), we obtain a 2-truncated augmented simplicial scheme

$$\begin{eqnarray}{\mathcal{Y}}_{1}\rightrightarrows {\mathcal{Y}}_{0}\rightarrow {\mathcal{Y}}\end{eqnarray}$$

which becomes a proper hypercover over a dense open subscheme of $\text{Spec}(R)$, and whose base change to ${\mathcal{O}}_{F}$ is exactly our original 2-truncated augmented simplicial scheme

$$\begin{eqnarray}{\mathcal{X}}_{1}\rightrightarrows {\mathcal{X}}_{0}\rightarrow {\mathcal{X}}.\end{eqnarray}$$

Let $\unicode[STIX]{x1D704}:R{\hookrightarrow}{\mathcal{O}}_{F}$ denote the canonical inclusion, and $\unicode[STIX]{x1D704}^{\ast }:\text{Spec}({\mathcal{O}}_{F})\rightarrow \text{Spec}(R)$ the induced morphism of schemes. Note that since $\unicode[STIX]{x1D704}^{\ast }$ maps the generic point of $\text{Spec}({\mathcal{O}}_{F})$ to that of $\text{Spec}(R)$, the map ${\mathcal{Y}}\rightarrow \text{Spec}(R)$ is generically smooth. We may thus choose an open subset $U\subset \text{Spec}(R)$ such that ${\mathcal{Y}}_{U}\rightarrow U$ is smooth, and such that the base change of $[{\mathcal{Y}}_{1}\rightrightarrows {\mathcal{Y}}_{0}\rightarrow {\mathcal{Y}}]$ to $U$ is a proper hypercover.

Lemma 4.3. For any $n\geqslant 0$ there exists a smooth curve $C/k$, a rational point $c\in C(k)$, a uniformiser $t_{c}$ at $c$, and a locally closed immersion $C\rightarrow \text{Spec}(R)$ such that $C\setminus \{c\}\subset U$, and the induced map

$$\begin{eqnarray}\text{Spec}({\mathcal{O}}_{C,c}/\mathfrak{m}_{c}^{n})\rightarrow \text{Spec}(R)\end{eqnarray}$$

agrees with the modulo $t^{n}$-reduction of $\unicode[STIX]{x1D704}^{\ast }$ via the isomorphism

$$\begin{eqnarray}\widehat{{\mathcal{O}}}_{C,c}\overset{{\sim}}{\rightarrow }{\mathcal{O}}_{F}\end{eqnarray}$$

sending $t_{c}$ to $t$.

Proof. Since $R$ is smooth, we may choose étale co-ordinates around the image $\unicode[STIX]{x1D704}^{\ast }(s)$ of the closed point of $\text{Spec}({\mathcal{O}}_{F})$ under $\unicode[STIX]{x1D704}^{\ast }$. This induces an étale map $\text{Spec}(R)\rightarrow \mathbb{A}_{k}^{n}$ for some $n$, and it is a simple exercise to prove the corresponding claim for $\mathbb{A}_{k}^{n}$. We then just take the pull-back to $\text{Spec}(R)$.◻

The canonical inclusion $\unicode[STIX]{x1D704}$ induces similar inclusions

$$\begin{eqnarray}\unicode[STIX]{x1D704}_{0}^{\#}:R_{0}^{\#}{\hookrightarrow}R_{0}^{\#}\otimes _{R}{\mathcal{O}}_{F}={\mathcal{O}}_{F_{0}^{\#}}\end{eqnarray}$$

for $\#\in \{\text{un},s,\emptyset \}$, as well as

$$\begin{eqnarray}\unicode[STIX]{x1D704}_{1,j}^{\#}:R_{1,j}^{\#}{\hookrightarrow}R_{1,j}^{\#}\otimes _{R}{\mathcal{O}}_{F}={\mathcal{O}}_{F_{1,j}^{\#}}\end{eqnarray}$$

for all $j$, and again for $\#\in \{\text{un},s,\emptyset \}$. We will need the following form of Krasner’s lemma [Sta18, §0BU9].

Lemma 4.4. Let $K$ be a local field, with ring of integers ${\mathcal{O}}_{K}$, and let $P(x)$ be an Eisenstein polynomial over ${\mathcal{O}}_{K}$. Let $L$ be the corresponding finite totally ramified extension, and let $\unicode[STIX]{x1D6FC}$ be a root of $P$ in $L$. Then for any $m\geqslant 1$ there exists an $n\geqslant 2$ such that any $Q(x)\in {\mathcal{O}}_{K}[x]$ congruent to $P$ modulo $\mathfrak{m}_{K}^{n}$ is Eisenstein, and $L$ contains a root $\unicode[STIX]{x1D6FD}$ of $Q$ such that $L=K(\unicode[STIX]{x1D6FD})$ and $\unicode[STIX]{x1D6FC}\equiv \unicode[STIX]{x1D6FD}$ modulo $\mathfrak{m}_{L}^{m}$.

We will use this as follows: given $n_{1}\geqslant \max _{j}\{[F_{1,j}:F]\}$ Lemma 4.4 shows that there exists some $n_{0}\geqslant \max \left\{2,[F_{0}:F]\right\}$ such that any polynomial $Q_{1,j}$ with coefficients in ${\mathcal{O}}_{F_{1,j}^{\text{un}}}$ which agrees with the minimal polynomial $P_{1,j}$ of $t_{1,j}$ modulo $(t_{0}^{\prime })^{n_{0}}$ is Eisenstein, and has a root in ${\mathcal{O}}_{F_{1,j}^{s}}$ which agrees with $t_{1,j}$ modulo $t_{1,j}^{n_{1}}$. Applying the lemma again shows the existence of some $n\geqslant 2$ such that any polynomial $Q_{0}$ with coefficients in ${\mathcal{O}}_{F_{0}^{\text{un}}}$ which agrees with $P_{0}$ modulo $t^{n}$ is Eisenstein, and has a root in ${\mathcal{O}}_{F_{0}^{s}}$ which agrees with $t_{0}$ modulo $t_{0}^{n_{0}}$. Now choose a $k$-algebra homomorphism $\unicode[STIX]{x1D706}:R\rightarrow {\mathcal{O}}_{F}$ as provided by Lemma 4.3, that is, factoring through the local ring of some smooth point on a curve inside $\text{Spec}(R)$ and agreeing with $\unicode[STIX]{x1D704}$ modulo $t^{n}$.

Since $\unicode[STIX]{x1D706}$ is a $k$-algebra homomorphism, we have a canonical isomorphism $R_{0}^{\text{un}}\otimes _{R,\unicode[STIX]{x1D706}}{\mathcal{O}}_{F}\overset{{\sim}}{\rightarrow }{\mathcal{O}}_{F_{0}^{\text{un}}}$, which therefore induces a homomorphism

$$\begin{eqnarray}\unicode[STIX]{x1D706}_{0}^{\text{un}}:R_{0}^{\text{un}}\rightarrow {\mathcal{O}}_{F_{0}^{\text{un}}}\end{eqnarray}$$

extending $\unicode[STIX]{x1D706}$ and which agrees with $\unicode[STIX]{x1D704}_{0}^{\text{un}}$ modulo $t^{n}$. Now let $Q_{0}=\unicode[STIX]{x1D706}_{0}^{\text{un}}(P_{0})$ denote the image under $\unicode[STIX]{x1D706}_{0}^{\text{un}}$ of the minimal polynomial $P_{0}$ of $t_{0}$; this is therefore a monic polynomial with coefficients in ${\mathcal{O}}_{F_{0}^{\text{un}}}$, which agrees with $P_{0}$ modulo $t^{n}$. Thus it is also Eisenstein, and by the choice of $n$ we know that ${\mathcal{O}}_{F_{0}^{s}}$ contains a root of $\unicode[STIX]{x1D706}_{0}^{\text{un}}(P_{0})$ which is congruent to $t_{0}$ modulo $t_{0}^{n_{0}}$ and generates ${\mathcal{O}}_{F_{0}^{s}}$ as an ${\mathcal{O}}_{F_{0}^{\text{un}}}$-algebra. This then allows us to extend $\unicode[STIX]{x1D706}_{0}^{\text{un}}$ to a homomorphism

$$\begin{eqnarray}\unicode[STIX]{x1D706}_{0}^{s}:R_{0}^{s}\rightarrow {\mathcal{O}}_{F_{0}^{s}}\end{eqnarray}$$

which agrees with $\unicode[STIX]{x1D704}_{0}^{s}$ modulo $t_{0}^{n_{0}}$, and since $\unicode[STIX]{x1D706}_{0}^{s}(t_{0})$ generates ${\mathcal{O}}_{F_{0}^{s}}$ as an ${\mathcal{O}}_{F_{0}^{\text{un}}}$-algebra, we deduce that the diagram

is coCartesian. We can then extend this to a homomorphism

$$\begin{eqnarray}\unicode[STIX]{x1D706}_{0}:R_{0}\rightarrow {\mathcal{O}}_{F_{0}}\end{eqnarray}$$

agreeing with $\unicode[STIX]{x1D704}_{0}$ modulo $(t_{0}^{\prime })^{n_{0}}$, and forming a similar coCartesian diagram to $\unicode[STIX]{x1D706}_{0}^{s}$. We now play exactly the same game for all of the $R_{1,j}$, to produce $\unicode[STIX]{x1D706}_{1,j}:R_{1,j}\rightarrow {\mathcal{O}}_{F_{1,j}}$ extending all other $\unicode[STIX]{x1D706}_{0}$ and all previous $\unicode[STIX]{x1D706}_{1,j}^{\#}$, which agree with $\unicode[STIX]{x1D704}_{1,j}$ modulo $(t_{1,j}^{\prime })^{n_{1,j}}$, and which form coCartesian diagrams

Now let ${\mathcal{Z}}$ be the base change of ${\mathcal{Y}}$ to ${\mathcal{O}}_{F}$ via $\unicode[STIX]{x1D706}$; note that the generic fibre ${\mathcal{Z}}_{F}$ is globally defined by construction. Similarly define ${\mathcal{Z}}_{0}$ to be the base change of ${\mathcal{Y}}_{0}$ to ${\mathcal{O}}_{F_{0}}$ via $\unicode[STIX]{x1D706}_{0}$, ${\mathcal{Z}}_{1,j}$ the base change of ${\mathcal{Y}}_{1,j}$ to ${\mathcal{O}}_{F_{1,j}}$ via $\unicode[STIX]{x1D706}_{1,j}$, and ${\mathcal{Z}}_{1}:=\coprod _{j}{\mathcal{Z}}_{1,j}$, so we have a 2-truncated augmented simplicial scheme

$$\begin{eqnarray}{\mathcal{Z}}_{1}\rightrightarrows {\mathcal{Z}}_{0}\rightarrow {\mathcal{Z}}\end{eqnarray}$$

over ${\mathcal{O}}_{F}$, which gives a proper hypercover after base changing to $F$. For any $m\geqslant 2$ we can therefore take $n_{1}\geqslant m\max _{j}\{[F_{1,j}:F]\}$ to ensure:

  1. ${\mathcal{Z}}_{0}$ is a proper and strictly semistable scheme over ${\mathcal{O}}_{F_{0}}$, and each ${\mathcal{Z}}_{1,j}$ is a proper and strictly semistable scheme over ${\mathcal{O}}_{F_{1,j}}$;

  2. there is an isomorphism

    $$\begin{eqnarray}\left[{\mathcal{X}}_{1}\rightrightarrows {\mathcal{X}}_{0}\right]\otimes _{{\mathcal{O}}_{F}}{\mathcal{O}}_{F}/t^{m}\overset{{\sim}}{\rightarrow }\left[{\mathcal{Z}}_{1}\rightrightarrows {\mathcal{Z}}_{0}\right]\otimes _{{\mathcal{O}}_{F}}{\mathcal{O}}_{F}/t^{m}\end{eqnarray}$$
    of 2-truncated simplicial schemes, such that
    $$\begin{eqnarray}{\mathcal{X}}_{0}\otimes {\mathcal{O}}_{F}/t^{m}\overset{{\sim}}{\rightarrow }{\mathcal{Z}}_{0}\otimes {\mathcal{O}}_{F}/t^{m}\end{eqnarray}$$
    is in fact an isomorphism of ${\mathcal{O}}_{F_{0}}/(t^{m})$-schemes, and
    $$\begin{eqnarray}{\mathcal{X}}_{1}\otimes {\mathcal{O}}_{F}/t^{m}\overset{{\sim}}{\rightarrow }{\mathcal{Z}}_{1}\otimes {\mathcal{O}}_{F}/t^{m}\end{eqnarray}$$
    is obtained as a disjoint union of isomorphisms
    $$\begin{eqnarray}{\mathcal{X}}_{1,j}\otimes {\mathcal{O}}_{F}/t^{m}\overset{{\sim}}{\rightarrow }{\mathcal{Z}}_{1,j}\otimes {\mathcal{O}}_{F}/t^{m}\end{eqnarray}$$
    of ${\mathcal{O}}_{F_{1,j}}/(t^{m})$-schemes.

Thus if we let ${\mathcal{X}}_{0,s}^{\times }$ and ${\mathcal{Z}}_{0,s}^{\times }$ denote the log schemes over $k_{0}^{\times }$ given by the special fibres of ${\mathcal{X}}_{0}$ and ${\mathcal{Z}}_{0}$, and ${\mathcal{X}}_{1,s}^{\times }$ and ${\mathcal{Z}}_{1,s}^{\times }$ the log schemes over $\coprod _{j=1}^{s}\text{Spec}(k_{1,j}^{\times })$ given by the special fibres of ${\mathcal{X}}_{1}$ and ${\mathcal{Z}}_{1}$, then by Proposition 2.2 there is an isomorphism

$$\begin{eqnarray}[{\mathcal{Z}}_{1,s}^{\times }\rightrightarrows {\mathcal{Z}}_{0,s}^{\times }]\cong [{\mathcal{X}}_{1,s}^{\times }\rightrightarrows {\mathcal{X}}_{0,s}^{\times }]\end{eqnarray}$$

of 2-truncated simplicial log schemes over $k^{\times }$. Now by [Reference Chiarellotto and LazdaCL18, Propositions 5.3, 5.4] there are isomorphisms

$$\begin{eqnarray}\displaystyle H_{\ell }^{i}({\mathcal{X}}_{0,F_{0}}) & \cong & \displaystyle H_{\ell }^{i}({\mathcal{Z}}_{0,F_{0}}),\nonumber\\ \displaystyle H_{\ell }^{i}({\mathcal{X}}_{1,F_{1,j}}) & \cong & \displaystyle H_{\ell }^{i}({\mathcal{Z}}_{1,j,F_{1,j}})\nonumber\end{eqnarray}$$

between the cohomology of the generic fibres of ${\mathcal{X}}_{0},{\mathcal{X}}_{1,j}$ and ${\mathcal{Z}}_{0},{\mathcal{Z}}_{1,j}$, as Weil–Deligne representations over $F_{0}$ and $F_{1,j}$ respectively. If we define the category

$$\begin{eqnarray}\text{Rep}_{\mathbb{Q}_{\ell }^{\prime }}(\text{WD}_{F_{1}}):=\mathop{\prod }_{j=1}^{s}\text{Rep}_{\mathbb{ Q}_{\ell }^{\prime }}(\text{WD}_{F_{1,j}})\end{eqnarray}$$

of Weil–Deligne representations over $F_{1}:=\prod _{j}F_{1,j}$ to be the product of the categories of Weil–Deligne representations over each $F_{1,j}$, then by Propositions 3.1 and 3.2, the diagram

(with horizontal arrows given by the differences of the two pullback maps) commutes via the restriction functor from Weil–Deligne representations over $F_{0}$ to Weil–Deligne representations over $F_{1}$.

Let $\text{Ind}_{F_{i}}^{F}$ denote a right adjoint to the restriction functor from Weil–Deligne representations over $F$ to those over $F_{i}$: on the separable part this is the normal induction of representations, on the inseparable part it is a quasi-inverse to Frobenius pull-back, and $\text{Ind}_{F_{1}}^{F}=\bigoplus _{j}\text{Ind}_{F_{1,j}}^{F}$. We therefore have a commutative diagram

and, in particular, the kernels of both horizontal maps are isomorphic as Weil–Deligne representations over $F$. The proof of Theorem 4.2 now boils down to the following claim.

Proposition 4.5. Let $X_{1}\rightrightarrows X_{0}\rightarrow X$ be a 2-truncated semisimplicial proper hypercover of a smooth and proper $F$-variety $X$, such that there exist finite field extensions $F_{0}/F$ and $F_{1,j}/F_{0}$ for $1\leqslant j\leqslant s$, with $X_{0}$ smooth over $F_{0}$, and $X_{1}=\coprod _{j}X_{1,j}$ with $X_{1,j}$ smooth over $F_{1,j}$. If we set $F_{1}=\prod _{j=1}^{s}F_{1,j}$, then

$$\begin{eqnarray}H_{\ell }^{i}(X)\cong \text{ker}(\text{Ind}_{F_{0}}^{F}H_{\ell }^{i}(X_{0})\rightarrow \text{Ind}_{F_{1}}^{F}H_{\ell }^{i}(X_{1}))\end{eqnarray}$$

for all primes $\ell$.

Proof. By taking $\widetilde{F}_{1}/F$ a sufficiently large finite extension such that all of the $F_{1,j}$ embed into $\widetilde{F}_{1}$ and applying [Reference de JongdJ96, Theorem 4.1], we can extend $X_{1}\rightrightarrows X_{0}\rightarrow X$ to a full proper hypercover $X_{\bullet }\rightarrow X$ such that for $n\geqslant 2$ there exists a finite extension $F_{n}/\widetilde{F}_{1}$ with $X_{n}$ smooth over $F_{n}$. Now applying [Reference Chiarellotto and LazdaCL18, Lemma 6.4] we can see that the terms in $i$th column of the resulting spectral sequence have to be ‘quasi-pure’ of weight $i$. Therefore the spectral sequence degenerates exactly as in the proof of [Reference Chiarellotto and LazdaCL18, Theorem 6.1], and the proposition follows.◻

We now deduce from the proposition that $H_{\ell }^{i}(X)\cong H_{\ell }^{i}({\mathcal{Z}}_{F})$ as Weil–Deligne representations for all $i,\ell$, and by construction ${\mathcal{Z}}_{F}$ is globally defined. This completes the proof of Theorem 4.2

Remark 4.6. Note the use of the finite field hypothesis (via a weight argument) in the proof of Proposition 4.5. It might be possible to relax the assumption to $k$ perfect using a more sophisticated argument.

Acknowledgements

Both authors would like to thank W. Zheng for pointing out the error in [Reference Chiarellotto and LazdaCL18], as well as L. Illusie for useful discussions concerning log structures, in particular the proof of Proposition 2.1. We would also like to thank the anonymous referee for a careful reading of an earlier version of the manuscript, in particular for correcting a mistake in our use of alterations.

Footnotes

B. Chiarellotto was supported by the grants MIUR-PRIN 2015 ‘Number Theory and Arithmetic Geometry’ and Padova PRAT CDPA159224/15. C. Lazda was supported by the Netherlands Organisation for Scientific Research (NWO).

References

Chiarellotto, B. and Lazda, C., Around -independence, Compos. Math. 154 (2018), 223248.CrossRefGoogle Scholar
de Jong, A. J., Smoothness, semi-stability and alterations, Publ. Math. Inst. Hautes Études Sci. 83 (1996), 5193.Google Scholar
Kedlaya, K. S., p-adic differential equations, Cambridge Studies in Advanced Mathematics, vol. 125 (Cambridge University Press, Cambridge, 2010).CrossRefGoogle Scholar
Lazda, C. and Pál, A., Rigid cohomology over laurent series fields, Algebra and Applications, vol. 21 (Springer, 2016).10.1007/978-3-319-30951-4CrossRefGoogle Scholar
Lu, Q. and Zheng, W., Compatible systems and ramification, Compos. Math. 155 (2019), 23342353.CrossRefGoogle Scholar
Marmora, A., Facteurs epsilon p-adiques, Compos. Math. 144 (2008), 439483.CrossRefGoogle Scholar
Nakayama, C., Logarithmic étale cohomology, Math. Ann. 308 (1997), 365404.Google Scholar
Nakayama, C., Nearby cycles for log smooth families, Compos. Math. 112 (1998), 4575.10.1023/A:1000327225021CrossRefGoogle Scholar
Spivakovsky, M., A new proof of D. Popescu’s theorem on smoothing of ring homomorphisms, J. Amer. Math. Soc. 12 (1999), 381444.CrossRefGoogle Scholar
The Stacks project authors, The Stacks project (2018), https//stacks.math.columbia.edu.Google Scholar