Hostname: page-component-857557d7f7-nk9cn Total loading time: 0 Render date: 2025-12-07T06:08:02.219Z Has data issue: false hasContentIssue false

DUALITY FOR COHOMOLOGY OF CURVES WITH COEFFICIENTS IN ABELIAN VARIETIES

Published online by Cambridge University Press:  19 December 2018

TAKASHI SUZUKI*
Affiliation:
Department of Mathematics, Chuo University, 1-13-27 Kasuga, Bunkyo-ku, Tokyo 112-8551, Japan email tsuzuki@gug.math.chuo-u.ac.jp
Rights & Permissions [Opens in a new window]

Abstract

In this paper, we formulate and prove a duality for cohomology of curves over perfect fields of positive characteristic with coefficients in Néron models of abelian varieties. This is a global function field version of the author’s previous work on local duality and Grothendieck’s duality conjecture. It generalizes the perfectness of the Cassels–Tate pairing in the finite base field case. The proof uses the local duality mentioned above, Artin–Milne’s global finite flat duality, the nondegeneracy of the height pairing and finiteness of crystalline cohomology. All these ingredients are organized under the formalism of the rational étale site developed earlier.

Information

Type
Article
Copyright
© 2018 Foundation Nagoya Mathematical Journal

1 Introduction

1.1 Aim of the paper

Let $X$ be a proper smooth geometrically connected curve over a perfect field $k$ of characteristic $p>0$. Let $A$ be an abelian variety over the function field $K$ of $X$, with dual $A^{\vee }$. We denote the Néron models of $A$ and $A^{\vee }$ over $X$ by ${\mathcal{A}}$ and ${\mathcal{A}}^{\vee }$, respectively. Let ${\mathcal{A}}_{0}^{\vee }$ be the maximal open subgroup scheme of ${\mathcal{A}}^{\vee }$ with connected fibers. We view these group schemes over $X$ as representable sheaves on $X$ in the flat topology and hence in the étale topology. In this paper, we formulate and prove a duality that relates the two étale cohomology complexes

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}),\qquad R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\end{eqnarray}$$

to each other. ($R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})$ is a complex whose $n$th cohomology is $H^{n}(X,{\mathcal{A}})$.)

This is the global function field version of the author’s work [Reference SuzukiSuz14] on duality for cohomology of local fields with coefficients in abelian varieties that solved Grothendieck’s duality conjecture [Gro72, IX, Conjecture 1.3]. The complexes above are not viewed as complexes of mere abelian groups, but as complexes of sheaves on the ind-rational pro-étale site $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ defined in [Reference SuzukiSuz14]. In particular, this duality treats the structure of $H^{n}(X,{\mathcal{A}})$ as the perfection (inverse limit along Frobenii) of a smooth group scheme over the base field $k$, which is related to Artin–Milne’s duality [Reference Artin and Milne.AM76] for cohomology of $X$ with coefficients in finite flat group schemes. The object $H^{1}(X,{\mathcal{A}})$ has unipotent connected part (which is $p$-power torsion) and is the sheafified version of the Tate–Shafarevich group of $A/K$, which we call the Tate–Shafarevich scheme. See Proposition 3.2.10 and the preceding paragraph for a more precise (but a bit subtle) relationship to Tate–Shafarevich groups. The above duality generalizes the perfectness of the Cassels–Tate pairing in the finite base field case. The duality for the part $H^{0}(X,{\mathcal{A}})=A(K)$ includes the nondegeneracy of the height pairing. When showing the finite-dimensionality of the Tate–Shafarevich scheme, we will use the finiteness of crystalline cohomology of proper smooth surfaces over $k$.

Our duality extends a partial result of Milne [Reference MilneMil06, III, Theorem 11.6] on duality $T_{p}\unicode[STIX]{x1D6E4}(K,A^{\vee })\leftrightarrow H^{2}(X,{\mathcal{A}})$. He also pointed out that the part killed by $p$ (not the whole $p$-primary part) of the Tate–Shafarevich group can be infinite when $k$ is algebraically closed [Reference MilneMil06, III, Remark 9.9]. This phenomenon has been studied by Vvedenskiĭ ([Reference VvedenskiĭVve81] for example), which, in our formulation, can be explained by the connected part of the Tate–Shafarevich scheme. It appears from [Reference VvedenskiĭVve78, Remark 7] that Vvedenskiĭ at least once imagined a possibility to construct a duality theory similar to ours.

The most interesting part of this duality theory lies in $p$-torsion. The prime-to-$p$ part is essentially classical (cf. [Reference RaynaudRay95]). When $k$ has zero characteristic, which is out of scope of this paper, one should probably take (nonsemistable) degenerations of Hodge structures into account.

When the first version of this paper was uploaded to arXiv, Česnavičius pointed the author to the preprint [Reference Demarche and HarariDH18] by Demarche–Harari, which was written independently at almost the same time. They develop compact support flat cohomology, a key technical tool that we also develop, in a way very similar to us. The setting and the details are different. For more details, see Section 1.4.

1.2 Statement of the main theorem

Now we formulate our duality. Let $\operatorname{Ab}(X_{\text{fppf}})$ and $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ be the categories of sheaves of abelian groups on the fppf site $X_{\text{fppf}}$ and the ind-rational pro-étale site $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ [Reference SuzukiSuz14, Section 2.1], respectively. We define a left exact functor

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\;\cdot \;):\operatorname{Ab}(X_{\text{fppf}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

by sending an fppf sheaf $F$ on $X$ to the pro-étale sheafification of the presheaf $k^{\prime }\mapsto F(X\times _{k}k^{\prime })$ on $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$, where $k^{\prime }$ runs through ind-rational $k$-algebras. Denote its right derived functor by

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\;\cdot \;):D(X_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

and set $\mathbf{H}^{n}(X,\;\cdot \;)=H^{n}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\;\cdot \;)$. The complex of sheaves $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ on $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ is the main object of study. Denote the derived sheaf-Hom functor for $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ by $R\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}$. For $G\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, its Serre dual [Reference SuzukiSuz14, Section 2.4] is defined by

$$\begin{eqnarray}G^{\text{SD}}=R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}(G,\mathbb{Z}).\end{eqnarray}$$

The Poincaré biextension ${\mathcal{A}}_{0}^{\vee }\otimes ^{L}{\mathcal{A}}\rightarrow \mathbf{G}_{m}[1]$ as a morphism in $D(X_{\text{fppf}})$, the cup product and the degree map of the Picard scheme induce morphisms

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\mathbf{G}_{m})[1]\rightarrow \mathbf{H}^{1}(X,\mathbf{G}_{m})\rightarrow \mathbb{Z}\end{eqnarray}$$

in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. Hence we have a morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}}.\end{eqnarray}$$

Its Serre dual

(1.2.1)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}}\end{eqnarray}$$

is our duality morphism. Let

$$\begin{eqnarray}V=V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}=(T\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}})\otimes \mathbb{Q}\end{eqnarray}$$

be the rational Tate module of the maximal divisible subsheaf of $\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })$.

Theorem A. There exist canonical morphisms

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}}\rightarrow V\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}[1]\end{eqnarray}$$

such that the triangle

$$\begin{eqnarray}V[-1]\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}}\rightarrow V\end{eqnarray}$$

is distinguished in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$.

Both of the objects $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}}$ are concentrated in degrees $-1,0,1,2$, while $V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$ has no connected part, is uniquely divisible and concentrated in the single degree zero. Hence our duality morphism is “close” to be an isomorphism. The divisible part of the usual Tate–Shafarevich group, when $k$ is finite, is conjectured to be zero. But our Tate–Shafarevich scheme, $\mathbf{H}^{1}(X,{\mathcal{A}})$, has nonzero divisible part in general, which might be called the space of “transcendental cycles with coefficients in the Néron model.”

Concrete consequences of this theorem will be explained at Section 3.4. We will also give a version for cohomology of dense open subschemes of $X$ in Section 4.1 and explain the link of Theorem A to the known duality theory in the finite base field case in Section 4.2. A very small amount of explicit calculations is given in the course of Remark 3.4.2.

A small remark is that the first cohomology classifies torsors or principal bundles. The Tate–Shafarevich scheme $\mathbf{H}^{1}(X,{\mathcal{A}})$ might alternatively be called the moduli of $G$-bundles on $X$ and denoted by $\text{Bun}_{G}(X)$, where $G={\mathcal{A}}$. We do not pursue this viewpoint, merely mentioning that our crucial point is to evaluate functors on ind-rational $k$-algebras only.

1.3 Outline of proof

Here is a rough outline of the proof. The first deep input is Artin–Milne’s global duality [Reference Artin and Milne.AM76, Corollary (4.9)] for a finite flat group scheme $N$ over $X$:

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,N^{\text{CD}})\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,N),\end{eqnarray}$$

where $N^{\text{CD}}$ is the Cartier dual of $N$, the cohomology is taken in the fppf topology and we ignored the shift of degrees for simplicity. Artin–Milne uses the étale site of all perfect $k$-schemes and we will bring their result to our site $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ by restriction and pro-étale sheafification. At each closed point $x\in X$, we have Bester’s local finite flat duality [Reference BesterBes78, Theorem 3.1] in the form stated in [Reference SuzukiSuz14, Theorem (5.2.1.2)]:

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},N^{\text{CD}})\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},N),\end{eqnarray}$$

where $\hat{{\mathcal{O}}}_{x}$ is the completed local ring of $X$ at $x$, and the fppf cohomology functor $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$ and its version with support $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$ with values in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ are as defined in [Reference SuzukiSuz14, Section 3.3]. ([Reference SuzukiSuz14, Theorem (5.2.1.2)] fits better in the setting of the present paper as it is a derived categorical and sheaf-theoretic version of the duality isomorphism of profinite abelian groups in [Reference BesterBes78, Theorem 3.1]. Also [Reference SuzukiSuz14, Theorem (5.2.1.2)] corrects some inaccuracies in Bester’s paper; see [Reference SuzukiSuz14, Remark 5.2.1.5].) Hence we can pass from $X$ to a dense open subscheme $U$ of $X$:

(1.3.1)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,N^{\text{CD}})\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,N),\end{eqnarray}$$

where $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)$ is the fppf cohomology with compact support with values in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ that we define and study in this paper. The fppf cohomology with compact support, as usual abelian groups, is already defined in [Reference MilneMil06, III, Section 0]. We need some clarification about the definition given in [Reference MilneMil06, III, Section 0], more than just bringing it to $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. This point was simultaneously found by Demarche and Harari [Reference Demarche and HarariDH18]; see Section 1.4 for the details. The duality (1.3.1) implies, taking $U$ small enough so that $A$ has good reduction over $U$, a duality

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee }[n])\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}}[n])\end{eqnarray}$$

for any $n\geqslant 1$. At each $x\in X$, we have local duality for abelian varieties [Reference SuzukiSuz14, Remark (4.2.10)]

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0}^{\vee })\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}).\end{eqnarray}$$

Hence we can pass from $U$ to $X$ and take the limit in $n$:

(1.3.2)$$\begin{eqnarray}R\,\underset{n}{\varprojlim }(R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Q}/\mathbb{Z}.\end{eqnarray}$$

So far we have treated essentially torsion objects. To pass to the desired integral statement

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\end{eqnarray}$$

up to $V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$, we need to study the structure of $\mathbf{H}^{n}(X,{\mathcal{A}})$ for all $n$. (Of course we need the double dual $\text{SD}\text{SD}$ for the precise statement.) The Lang–Néron theorem (the arbitrary base field version of the Mordell–Weil theorem) shows that the group of rational points $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ is an étale group with finitely generated group of geometric points extended by an abelian variety over $k$, the $K/k$-trace of $A$. The part

$$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee }))/\text{torsion}\longleftrightarrow (\unicode[STIX]{x1D70B}_{0}\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))/\text{torsion}\end{eqnarray}$$

is given by the height pairing.

For higher cohomology, the duality theorems used so far and finiteness statements therein give some information. But this information does not exclude the possibility that $\mathbf{H}^{1}(X,{\mathcal{A}})$ has connected divisible unipotent ind-algebraic part (such as the direct limit of the groups of Witt vectors $\mathop{\varinjlim }\nolimits_{n}W_{n}$) since $\mathbf{H}_{x}^{2}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})$ does have such part. We need here a global finiteness result to eliminate this possibility. That is, we compare $\mathbf{H}^{1}(X,{\mathcal{A}})$ with the Brauer group of a proper smooth surface over $k$ up to some discrepancy of bounded torsion (i.e., discrepancy killed by multiplication by some positive integer) and hence with the part $F=p$ of the second rational crystalline cohomology of the surface. This part does not have connected part by finiteness of crystalline cohomology, which proves the desired property of $\mathbf{H}^{1}(X,{\mathcal{A}})$. Dualizing, this in turn shows that $\mathbf{H}^{2}(X,{\mathcal{A}})$ has no connected part.

Then we can pass to the rational statement

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\otimes \mathbb{Q},\end{eqnarray}$$

where the derived inverse limit is for multiplication by positive integers $n$. The nondegeneracy of the height pairing shows that this rational duality pairing is perfect up to $V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$. Then we can finally pass to the desired integral statement, proving the main theorem.

Two remarks are in order. The reader may have an impression from the above outline that the duality theorem of this paper is closely related to duality for crystalline cohomology and Milne’s flat duality for surfaces [Reference MilneMil76]. Probably it will likely be possible to reduce our theorem to those duality theories when $A$ is the Jacobian of a proper smooth curve over $K$ with a $K$-rational point or has semistable reduction everywhere. But there are differences between these special cases and the general case, coming from isogenies of $p$-power degrees and wild ramification. Without our formulation, it could be difficult to formulate a duality statement that is invariant under isogenies and/or admits Galois descent.

The local duality [Reference SuzukiSuz14] used above is where the site $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ plays a crucial role (see [Reference SuzukiSuz14, Section 1.2] for a little more details). This should be compared with Milne’s approach [Reference MilneMil06, III, Section 11] that uses the étale site of all perfect $k$-schemes.

1.4 Organization

Up to the point (1.3.2) above, we will know that $\mathbf{H}^{n}(X,{\mathcal{A}})$ for $n\geqslant 1$ is an ind-algebraic group of so-called “cofinite type.” This is a basic finiteness condition for ind-algebraic groups. Section 2.1 develops the notion of cofinite type objects in the setting of ind-categories of general artinian abelian categories. Section 2.2 studies cofinite type objects in the special case of the category of ind-algebraic groups. Section 2.3 briefly reviews the ind-rational pro-étale site and study some derived limits such as the one $R\,\mathop{\varprojlim }\nolimits_{n}$ mentioned above.

In Section 2.5 (Section 2.4 to be mentioned soon), we basically review the local duality [Reference SuzukiSuz14]. Additionally, we improve and simplify the formulation. We eliminate the relative fppf site ($\operatorname{Spec}K_{\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}}$ in the notation of [Reference SuzukiSuz14, Section 3.3]) and instead formulate the same duality result using the usual fppf site for local fields. We can still define the key notion of the structure morphism of a local field, which targets the ind-rational étale site of the residue field. This newer version is not, however, a morphism of sites, but only a “premorphism of sites,” which does not necessarily have an exact pullback functor. A premorphism of sites is not really new and nothing but a morphism of topologies in the terminology of [Reference ArtinArt62, Definition 2.4.2]. (We change the terminology since the direction of morphisms is somewhat confusing.) Its definition is recalled and some basic site-theoretic propositions are given in the preceding Section 2.4. The comparison between the two formulations is given in Appendix A. We do use the local duality results of [Reference SuzukiSuz14] and do not reprove them.

As explained above, we will need fppf cohomology of curves with compact support $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)$ as a complex of sheaves on $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$. We will develop this machinery in Section 2.7. There are actually two versions of compact support fppf cohomology currently known, as explained in [Reference MilneMil06, III, Remark 0.6(b)]. Their difference is whether we use Henselian local rings ${\mathcal{O}}_{x}^{h}$ or their completions $\hat{{\mathcal{O}}}_{x}$ for the local components. The latter version is what we need in the duality theory of this paper. But this latter version has a problem about covariant functoriality in $U$ coming from the difference between two versions of local cohomology with support $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$. We will see this problem in Remark 2.7.9. The difference between the two versions of local cohomology with support vanishes if the coefficient sheaf is representable by a smooth or finite flat group scheme, by a Greenberg approximation argument. However, we need various mapping cone constructions in order to establish basics about compact support cohomology including covariant functoriality, which forces us to work on the level of complexes and, consequently, to explicitly take injective resolutions everywhere. For nonrepresentable sheaves, the difference between the two versions of local cohomology with support is unavoidable. Hence we need to keep all track of this difference from the beginning to build the theory. Some part of this consideration is purely local, which is the content of the preceding Section 2.6. After these subsections, we can and will ignore the difference since we are interested in smooth or finite flat group schemes only.

Let us mention here that the same problem of compact support fppf cohomology is realized and solved by Demarche and Harari [Reference Demarche and HarariDH18] independently at almost the same time. Their setting is over a finite base field $k=\mathbb{F}_{q}$, cohomology is viewed as usual complexes of abelian groups, not sheaves, and coefficient sheaves are affine group schemes for the most part. Their method to establish covariant functoriality of compact support fppf cohomology is very similar to our method in Section 2.62.7.

Back to this paper, Section 3.1 proves (1.3.1). Section 3.2 begins by proving (1.3.2) (or the statement before limit in $n$). Various consequences and further structural results on $\mathbf{H}^{n}(X,{\mathcal{A}})$ are given: the structure of $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ (Proposition 3.2.3); the structure of $\mathbf{H}^{1}(X,{\mathcal{A}})$ and its relation to the Tate–Shafarevich group (Proposition 3.2.3, 3.2.10, 3.2.12); how the duality pairing interacts with Néron component groups at the closed points of $X$ (Proposition 3.2.9); and the relation to the height pairing (Proposition 3.2.13).

The properties of $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$, $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })$ and the pairing between them proven up to this point, together with the preliminaries in Section 2, allow us to deduce the main theorem as a result of formal calculations. This is done in Section 3.3. The outline of proof above basically explains that any choice of the mapping cone of the duality morphism (1.2.1) is concentrated in degree zero with cohomology isomorphic to $V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$. In addition, we heavily use derived limit arguments, the structure of $\mathbf{H}^{n}(X,{\mathcal{A}})$ and the nondegeneracy of the height pairing, to show that the mapping cone can be taken canonically.

Section 3.4 just collects all the results obtained so far to state them as a single theorem. This finishes the proof of Theorem A. A relation to Milne’s result [Reference MilneMil06, Theorem 11.6] is also explained. The intersection of the connected part and the divisible part of the Tate–Shafarevich scheme $\mathbf{H}^{1}(X,{\mathcal{A}})$ is an interesting but difficult finite étale $p$-group. We briefly explain this in Remark 3.4.2.

Section 4.1 gives a duality for open $U\subset X$:

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })\longleftrightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})\end{eqnarray}$$

up to the same obstruction $V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$. Again, it requires some work to ensure that the mapping cone of this duality morphism can be taken canonically.

Applying the derived global section $R\unicode[STIX]{x1D6E4}(k,\;\cdot \;)$ to the complex of sheaves $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ recovers the usual complex of abelian groups $R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})$. In a nonderived categorical language, this means that there is a spectral sequence

$$\begin{eqnarray}E_{2}^{ij}=H^{i}(k,\mathbf{H}^{j}(X,{\mathcal{A}}))\Longrightarrow H^{i+j}(X,{\mathcal{A}}).\end{eqnarray}$$

In Section 4.2, when the base field $k$ is finite, we apply $R\unicode[STIX]{x1D6E4}(k,\;\cdot \;)$ to translate our duality theorem into the classical duality theorems. We explain the relations to the finiteness of Tate–Shafarevich groups, Kato–Trihan’s arithmetic cohomology [Reference Kato and TrihanKT03], the Cassels–Tate pairing and the Weil-étale cohomology $R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}})$.

Notation. We fix two universes ${\mathcal{U}}_{0}$ and ${\mathcal{U}}$ such that $\mathbb{N}\in {\mathcal{U}}_{0}\in {\mathcal{U}}$. The categories of ${\mathcal{U}}$-small sets and ${\mathcal{U}}$-small abelian groups are denoted by $\operatorname{Set}$ and $\operatorname{Ab}$, respectively. For an abelian category ${\mathcal{A}}$, the category of complexes is denoted by $\operatorname{Ch}({\mathcal{A}})$. Its homotopy category is $K({\mathcal{A}})$ and derived category $D({\mathcal{A}})$. We denote the full subcategory of bounded, bounded below and bounded above complexes by $D^{b}({\mathcal{A}})$, $D^{+}({\mathcal{A}})$ and $D^{-}({\mathcal{A}})$, respectively. Similar notation applies to $\operatorname{Ch}({\mathcal{A}})$ and $K({\mathcal{A}})$. The mapping cone of a morphism $A\rightarrow B$ in $\operatorname{Ch}({\mathcal{A}})$ is denoted by $[A\rightarrow B]$. If $A\rightarrow B$ is a morphism in a triangulated category together with a certain canonical choice of a mapping cone, then this mapping cone is also denoted by $[A\rightarrow B]$ by abuse of notation. If we say $A\rightarrow B\rightarrow C$ is a distinguished triangle in a triangulated category, we implicitly assume that a morphism $C\rightarrow A[1]$ to the shift of $A$ is given, and the triangle $A\rightarrow B\rightarrow C\rightarrow A[1]$ is distinguished.

2 Site-theoretic foundations and local duality

2.1 Ind-objects of cofinite type

Let $p$ be a prime number. Let ${\mathcal{A}}$ be an artinian (${\mathcal{U}}_{0}$-)small abelian category such that any object of ${\mathcal{A}}$ is killed by some power of $p$. Note that no nonzero object of ${\mathcal{A}}$ is divisible, which will be used frequently later on. Denote its ind-category by $\text{I}{\mathcal{A}}$ (or more precisely, the ${\mathcal{U}}_{0}$-indcategory, where index sets are ${\mathcal{U}}_{0}$-small). The category $\text{I}{\mathcal{A}}$ is abelian by [Reference Kashiwara and SchapiraKS06, Theorem 8.6.5(i)]. Note that objects of ${\mathcal{A}}$ need not be artinian in $\text{I}{\mathcal{A}}$. For instance, the additive algebraic group scheme $\mathbf{G}_{a}$ over a field of characteristic $p$ contains the ind-finite-étale subgroup $\overline{\mathbb{F}_{p}}$. We say that an object $A\in \text{I}{\mathcal{A}}$ is of cofinite type if the $p^{n}$-torsion part $A[p^{n}]$ is in ${\mathcal{A}}$ for all $n\geqslant 0$. In this case, the equality $A=\varinjlim A[p^{n}]$ gives a presentation of $A$ as a filtered direct limit of objects of ${\mathcal{A}}$. Denote by $\text{I}^{f}{\mathcal{A}}$ the full subcategory of $\text{I}{\mathcal{A}}$ consisting of objects of cofinite type.

Proposition 2.1.1. For any $A\in \text{I}^{f}{\mathcal{A}}$, we have $A/p^{n}A\in {\mathcal{A}}$ for all$n\geqslant 0$.

Proof. Let $A\in \text{I}^{f}{\mathcal{A}}$. To show $A/p^{n}A\in {\mathcal{A}}$, it is enough to show this for $n=1$. The object $A/pA$ is the union of the increasing sequence of subobjects $A[p^{m}]/p(A[p^{m+1}])$ each in ${\mathcal{A}}$. It is enough to show that this sequence stabilizes, or $A[p^{m-1}]+p(A[p^{m+1}])=A[p^{m}]$ for large $m$. This is equivalent that the decreasing sequence

$$\begin{eqnarray}\cdots \overset{p}{{\hookrightarrow}}A[p^{m+1}]/A[p^{m}]\overset{p}{{\hookrightarrow}}A[p^{m}]/A[p^{m-1}]\overset{p}{{\hookrightarrow}}\cdots\end{eqnarray}$$

stabilizes. It indeed does since each object in the sequence belongs to the artinian category ${\mathcal{A}}$.◻

Proposition 2.1.2. The category $\text{I}^{f}{\mathcal{A}}$ is an abelian subcategory of $\text{I}{\mathcal{A}}$.

Proof. Let $f:A\rightarrow B$ be a morphism in $\text{I}^{f}{\mathcal{A}}$. Then $\operatorname{Ker}(f)[p^{n}]$ for any $n$ is the kernel of the restriction $A[p^{n}]\rightarrow B[p^{n}]$ of $f$. Hence $\operatorname{Ker}(f)[p^{n}]\in {\mathcal{A}}$, thus $\operatorname{Ker}(f)\in \text{I}^{f}{\mathcal{A}}$. Therefore in the exact sequence $0\rightarrow \operatorname{Ker}(f)\rightarrow A\rightarrow \operatorname{Im}(f)\rightarrow 0$, the kernel and cokernel of $p^{n}$ on the first two terms are in ${\mathcal{A}}$. Hence $\operatorname{Im}(f)\in \text{I}^{f}{\mathcal{A}}$. The same argument for $0\rightarrow \operatorname{Im}(f)\rightarrow B\rightarrow \operatorname{Coker}(f)\rightarrow 0$ implies that $\operatorname{Coker}(f)\in \text{I}^{f}{\mathcal{A}}$. Hence $\text{I}^{f}{\mathcal{A}}$ is an abelian subcategory of $\text{I}{\mathcal{A}}$.◻

Proposition 2.1.3. For any $A\in \text{I}^{f}{\mathcal{A}}$, the decreasing sequence

$$\begin{eqnarray}\cdots {\hookrightarrow}p^{2}A{\hookrightarrow}pA{\hookrightarrow}A\end{eqnarray}$$

stabilizes.

Proof. For any $n$, we have $p^{n}:A/(A[p^{n}]+pA)\overset{{\sim}}{\rightarrow }p^{n}A/p^{n+1}A$. The isomorphism

$$\begin{eqnarray}A/pA\cong \underset{n}{\varinjlim }\,(A[p^{n}]+pA)/pA\end{eqnarray}$$

in $\text{I}{\mathcal{A}}$ factors through $(A[p^{n}]+pA)/pA$ for some $n$ since $A/pA\in {\mathcal{A}}$. Hence $A/pA\cong (A[p^{n}]+pA)/pA$ for all large $n$. For such $n$, we have $A=A[p^{n}]+pA$ and $p^{n+1}A=p^{n}A$. Hence the sequence stabilizes.◻

For $A\in \text{I}^{f}{\mathcal{A}}$, we define

$$\begin{eqnarray}\displaystyle & \displaystyle A_{\text{div}}=\mathop{\bigcap }_{n}p^{n}A\quad (=p^{n}A\text{ for some }n), & \displaystyle \nonumber\\ \displaystyle & \displaystyle A_{/\text{div}}=A/A_{\text{div}}. & \displaystyle \nonumber\end{eqnarray}$$

Proposition 2.1.4. If $A\in \text{I}^{f}{\mathcal{A}}$, then $A_{\text{div}}\in \text{I}^{f}{\mathcal{A}}$ is divisible and$A_{/\text{div}}\in {\mathcal{A}}$. The sequence

$$\begin{eqnarray}0\rightarrow A_{\text{div}}\rightarrow A\rightarrow A_{/\text{div}}\rightarrow 0\end{eqnarray}$$

is exact. The object $A_{\text{div}}$ is the largest divisible subobject of $A$. The object $A_{/\text{div}}$ is the largest quotient of $A$ that belongs to ${\mathcal{A}}$.

Proof. Obvious. ◻

Proposition 2.1.5. The category $\text{I}^{f}{\mathcal{A}}$ is artinian.

Proof. By the previous proposition, an object of ${\mathcal{A}}$ is artinian in $\text{I}^{f}{\mathcal{A}}$ since an object of $\text{I}^{f}{\mathcal{A}}$ embeddable into an object of ${\mathcal{A}}$ does not have divisible part and hence itself is in ${\mathcal{A}}$. It is enough to show that a decreasing sequence $A_{0}\supset A_{1}\supset \cdots \,$ of divisible objects in $\text{I}^{f}{\mathcal{A}}$ stabilizes. Since $A_{n}[p]\in {\mathcal{A}}$ is artinian for any $n$, there is some $m$ such that $A_{m}[p]=A_{m+1}[p]=\cdots \,$. From the exact sequence

$$\begin{eqnarray}0\rightarrow A_{n}[p]\rightarrow A_{n}[p^{2}]\overset{p}{\rightarrow }A_{n}[p]\rightarrow 0,\end{eqnarray}$$

we know that $A_{m}[p^{2}]=A_{m+1}[p^{2}]=\cdots \,$ for the same $m$. Inductively, we have $A_{m}=A_{m+1}=\cdots \,$.◻

Denote the procategory of ${\mathcal{A}}$ by $\text{P}{\mathcal{A}}$ and the ind-category of $\text{P}{\mathcal{A}}$ by $\text{I}\text{P}{\mathcal{A}}$. For $A\in \text{I}^{f}{\mathcal{A}}$, we define

$$\begin{eqnarray}\displaystyle & \displaystyle TA=\underset{n}{\varprojlim }\,A[p^{n}]\in \text{P}{\mathcal{A}}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle VA=\underset{m}{\varinjlim }\,TA\in \text{I}\text{P}{\mathcal{A}}, & \displaystyle \nonumber\end{eqnarray}$$

where the direct limit in the second definition is for multiplication by $p^{m}$. We call $TA$ the Tate module of $A$ and $VA$ the rational Tate module of $A$. For example, if $A$ is $\mathbb{Q}_{p}/\mathbb{Z}_{p}$ in the category of torsion abelian groups, then $TA=\mathbb{Z}_{p}$ in the category of profinite abelian groups and $VA=\mathbb{Q}_{p}$ as an ind-object of profinite abelian groups. For general ${\mathcal{A}}$, if $A\in {\mathcal{A}}$, then the system $\{A[p^{n}]\}$ that defines $TA$ has essentially zero transition morphisms, hence $TA=0$. For each $m$, we consider the morphism $TA\rightarrow A$ in $\text{I}\text{P}{\mathcal{A}}$ given by $(a_{n})_{n}\mapsto a_{m}$. They form a morphism $VA\rightarrow A$ in $\text{I}\text{P}{\mathcal{A}}$.

Proposition 2.1.6. The functor $A\mapsto VA$ from $\text{I}^{f}{\mathcal{A}}$ to $\text{I}\text{P}{\mathcal{A}}$ is exact.

Proof. Let $0\rightarrow A\rightarrow B\rightarrow C\rightarrow 0$ be exact in $\text{I}^{f}{\mathcal{A}}$. For any $n$, we have an exact sequence

$$\begin{eqnarray}0\rightarrow A[p^{n}]\rightarrow B[p^{n}]\rightarrow C[p^{n}]\rightarrow A/p^{n}A\rightarrow B/p^{n}B\rightarrow C/p^{n}C\rightarrow 0\end{eqnarray}$$

in ${\mathcal{A}}$. The inverse limit in $n$ in $\text{P}{\mathcal{A}}$ gives an exact sequence

$$\begin{eqnarray}0\rightarrow TA\rightarrow TB\rightarrow TC\rightarrow A_{/\text{div}}\rightarrow B_{/\text{div}}\rightarrow C_{/\text{div}}\rightarrow 0\end{eqnarray}$$

in $\text{P}{\mathcal{A}}$. (Note that Mittag-Leffler conditions are not relevant here since filtered inverse limits in a procategory are exact by definition.) Since each $(\;\cdot \;)_{/\text{div}}\in {\mathcal{A}}$ is killed by multiplication by some power of $p$, the direct limit in multiplication by $m$ gives the desired exact sequence $0\rightarrow VA\rightarrow VB\rightarrow VC\rightarrow 0$ in $\text{I}\text{P}{\mathcal{A}}$.◻

Proposition 2.1.7. Let $A\in \text{I}^{f}{\mathcal{A}}$. Then we have an exact sequence

$$\begin{eqnarray}0\rightarrow TA\overset{p^{n}}{\rightarrow }TA\rightarrow A_{\text{div}}\cap A[p^{n}]\rightarrow 0\end{eqnarray}$$

in $\text{P}{\mathcal{A}}$ for any $n\geqslant 0$ and an exact sequence

$$\begin{eqnarray}0\rightarrow TA\rightarrow VA\rightarrow A_{\text{div}}\rightarrow 0\end{eqnarray}$$

in $\text{I}\text{P}{\mathcal{A}}$. We have $TA=T(A_{\text{div}})$ and $VA=V(A_{\text{div}})$.

Proof. The multiplication-by-$p^{n}$ map gives an exact sequence $0\rightarrow A[p^{n}]\rightarrow A\rightarrow p^{n}A\rightarrow 0$. We have $T(A[p^{n}])=0$ and $A[p^{n}]_{/\text{div}}=A[p^{n}]$ since $A[p^{n}]\in {\mathcal{A}}$ has no divisible part. Hence the long exact sequence in the proof of the previous proposition for this sequence is

$$\begin{eqnarray}0\rightarrow TA\rightarrow T(p^{n}A)\rightarrow A[p^{n}]\rightarrow A_{/\text{div}}\rightarrow (p^{n}A)_{/\text{div}}\rightarrow 0.\end{eqnarray}$$

in $\text{P}{\mathcal{A}}$. Since $A/p^{n}A\in {\mathcal{A}}$, we have $T(p^{n}A)=TA$. From this, we get the first exact sequence. The direct limit in $n$ gives the second exact sequence. We have $T(A_{/\text{div}})=0$ since $A_{/\text{div}}\in {\mathcal{A}}$. It follows that $TA=T(A_{\text{div}})$ and $VA=V(A_{\text{div}})$.◻

Next, let ${\mathcal{A}}$ be an artinian abelian category such that any object is killed by multiplication by some positive integer (not necessarily a prime power). For each prime $p$, denote by ${\mathcal{A}}_{p}$ the full subcategory of ${\mathcal{A}}$ of objects killed by a power of $p$. This is an artinian abelian category. Any object of ${\mathcal{A}}$ can be canonically decomposed as $A=\bigoplus _{p}A_{p}$, where each summand $A_{p}$ belongs to ${\mathcal{A}}_{p}$ and $A_{p}=0$ for almost all $p$. We call $A_{p}$ the $p$-primary part of $A$. Consider the ind-category $\text{I}{\mathcal{A}}$. Any object of $\text{I}{\mathcal{A}}$ can be canonically decomposed as $A=\bigoplus _{p}A_{p}$ (a filtered direct limit of finite partial sums), where each summand $A_{p}$ belongs to $\text{I}{\mathcal{A}}_{p}:=\text{I}({\mathcal{A}}_{p})$.

We say that an object $A\in \text{I}{\mathcal{A}}$ is of cofinite type if the $n$-torsion part $A[n]$ is in ${\mathcal{A}}$ for all $n\geqslant 1$. In this case, the equality $A=\varinjlim A[n]$ gives a presentation of $A$ as a filtered direct limit of objects of ${\mathcal{A}}$. Denote by $\text{I}^{f}{\mathcal{A}}$ the full subcategory of $\text{I}{\mathcal{A}}$ consisting of objects of cofinite type. An object $A\in \text{I}{\mathcal{A}}$ is of cofinite type if and only if $A_{p}$ is of cofinite type for all $p$. Note that $\text{I}^{f}{\mathcal{A}}$ is not necessarily artinian. For instance, the object $\bigoplus _{p}\mathbb{Z}/p\mathbb{Z}$ in the category of torsion abelian groups of cofinite type is not artinian. For $A\in \text{I}^{f}{\mathcal{A}}$, we define

$$\begin{eqnarray}A_{\text{div}}=\bigoplus _{p}(A_{p})_{\text{div}}\in \text{I}{\mathcal{A}}\qquad A_{/\text{div}}=A/A_{\text{div}}\in \text{I}{\mathcal{A}},\end{eqnarray}$$

where $(A_{p})_{\text{div}}\in \text{I}{\mathcal{A}}_{p}$ is previously defined. We also define

$$\begin{eqnarray}TA=\underset{n}{\varprojlim }\,A[n]\in \text{P}{\mathcal{A}},\qquad VA=\underset{m}{\varinjlim }\,TA\in \text{I}\text{P}{\mathcal{A}},\end{eqnarray}$$

where the inverse limit in the first definition is over multiplication by $n\geqslant 1$ and the direct limit in the second is over multiplication by $m\geqslant 1$. Also define $T_{p}A=T(A_{p})$, $V_{p}A=V(A_{p})$.

Proposition 2.1.8.
  1. (1) For any $A\in \text{I}^{f}{\mathcal{A}}$ and $n\geqslant 1$, we have $A/nA\in {\mathcal{A}}$.

  2. (2) The category $\text{I}^{f}{\mathcal{A}}$ is an abelian subcategory of $\text{I}{\mathcal{A}}$.

  3. (3) If $A\in \text{I}^{f}{\mathcal{A}}$, then $A_{\text{div}}$ is the largest divisible subobject of $A$, and $A_{/\text{div}}$ is the largest quotient of $A$ such that $(A_{/\text{div}})_{p}$ belongs to ${\mathcal{A}}_{p}$ for all $p$. The sequence

    $$\begin{eqnarray}0\rightarrow A_{\text{div}}\rightarrow A\rightarrow A_{/\text{div}}\rightarrow 0\end{eqnarray}$$
    is an exact sequence in $\text{I}^{f}{\mathcal{A}}$.
  4. (4) Let $A\in \text{I}^{f}{\mathcal{A}}$. Then we have exact sequences

    $$\begin{eqnarray}0\rightarrow TA\overset{n}{\rightarrow }TA\rightarrow A_{\text{div}}\cap A[n]\rightarrow 0\end{eqnarray}$$
    in $\text{P}{\mathcal{A}}$ for any $n\geqslant 1$ and
    $$\begin{eqnarray}0\rightarrow TA\rightarrow VA\rightarrow A_{\text{div}}\rightarrow 0\end{eqnarray}$$
    in $\text{I}\text{P}{\mathcal{A}}$. We have $TA=T(A_{\text{div}})=\prod _{p}T_{p}A$ and $VA=V(A_{\text{div}})$.

Proof. All the statements are reduced to the $p$-primary parts for primes $p$, which have already been proved.◻

Contrary to the $p$-primary case, $A_{/\text{div}}$ might not be in ${\mathcal{A}}$ for general $A\in \text{I}^{f}{\mathcal{A}}$. An example is $\bigoplus _{p}\mathbb{Z}/p\mathbb{Z}$. This is related to the Mittag-Leffler condition. Here is a general fact.

Proposition 2.1.9. Let $B$ be an object of any abelian category ${\mathcal{B}}$. Consider the inverse system in ${\mathcal{B}}$ given by multiplication maps by positive integers on $B$. This system is Mittag-Leffler if and only if $B$ has a divisible subobject $B^{\prime }$ such that $B/B^{\prime }$ is killed by multiplication by some positive integer. In this case, $B^{\prime }$ is the maximal divisible subobject of $B$, and we say that $B$ is $\text{divisibly ML}$.

Proof. Elementary. ◻

Proposition 2.1.10. Let $A\in \text{I}^{f}{\mathcal{A}}$. Then $A$ is divisibly ML if and only if $A_{/\text{div}}\in {\mathcal{A}}$ if and only if $A_{p}$ is divisible for almost all $p$.

Proof. We have

$$\begin{eqnarray}A_{/\text{div}}=\bigoplus _{p}A_{p}\bigg/\bigoplus _{p}(A_{p})_{\text{div}}=\bigoplus _{p}(A_{p})_{/\text{div}}.\end{eqnarray}$$

The result follows from this. ◻

Note that divisibly ML objects in $\text{I}^{f}{\mathcal{A}}$ do not form an abelian subcategory. For instance, the direct sum of the exact sequences $0\rightarrow \mathbb{Z}/p\mathbb{Z}\rightarrow \mathbb{Q}_{p}/\mathbb{Z}_{p}\rightarrow \mathbb{Q}_{p}/\mathbb{Z}_{p}\rightarrow 0$ for primes $p$ gives an exact sequence $0\rightarrow \bigoplus _{p}\mathbb{Z}/p\mathbb{Z}\rightarrow \mathbb{Q}/\mathbb{Z}\rightarrow \mathbb{Q}/\mathbb{Z}\rightarrow 0$, where the last two terms are divisible but the first term is not divisibly ML.

2.2 Ind-algebraic groups of cofinite type

Let $k$ ($\in {\mathcal{U}}_{0}$) be a perfect field of characteristic $p>0$. We quickly recall some notation about perfections of algebraic groups from [Reference SuzukiSuz14, Section 2.1]. Let $\text{Alg}/k$ be the category of quasi-algebraic groups over $k$ (commutative, as assumed throughout the paper) in the sense of Serre [Reference SerreSer60]. Recall that a quasi-algebraic group is the perfection (inverse limit along Frobenii) of a (not necessarily connected) algebraic group [Reference SerreSer60, Section 1.2, Définition 2; Section 1.4, Proposition 10]. For example, the perfection of the additive group $\operatorname{Spec}k[x]$ is $\operatorname{Spec}k[x,x^{1/p},x^{1/p^{2}},\ldots \,]$, which we simply call the additive group and denote by $\mathbf{G}_{a}$ by abuse of notation. We say that a quasi-algebraic group is a unipotent group, a torus or a (semi-)abelian variety if it is the perfection of such a group. We call an object of the procategory $\text{P}\text{Alg}/k$ a proalgebraic group, an object of the ind-category $\text{I}\text{Alg}/k$ an ind-algebraic group and an object of the ind-procategory $\text{I}\text{P}\text{Alg}/k=\text{I}(\text{P}\text{Alg}/k)$ an ind-proalgebraic group. Let $\text{Alg}_{\text{uc}}/k$ be the full subcategory of $\text{Alg}/k$ consisting of groups whose identity component is unipotent. Let $\text{L}\text{Alg}/k$ be the category of perfections of smooth group schemes over $k$. A finitely generated étale group is an étale group with a finitely generated group of geometric points. A lattice is a finitely generated étale group with no torsion. Let $\text{FEt}/k\subset \text{FGEt}/k\subset \text{Et}/k$ be the categories of finite étale groups, finitely generated étale groups, étale groups, respectively, over $k$. For any prime $l$ (possibly equal to $p$), we denote the full subcategory of $\text{FEt}/k$ of groups of $l$-power order by $\text{FEt}_{l}/k$. For $A\in \text{L}\text{Alg}/k$, we denote its identity component by $A_{0}\in \text{Alg}/k$ and set $\unicode[STIX]{x1D70B}_{0}(A)=A/A_{0}\in \text{Et}/k$ ([Reference Demazure and GabrielDG70, II, Section 5, Proposition 1.8] plus perfection, noting that perfection does not change the underlying topological space). The endofunctor $A\mapsto A_{0}$ on $\text{Alg}/k$ extends to endofunctors on $\text{P}\text{Alg}/k$, $\text{I}\text{Alg}/k$ and $\text{I}\text{P}\text{Alg}/k$, still denoted by $A\mapsto A_{0}$. We say that $A\in \text{I}\text{P}\text{Alg}/k$ is connected if $A=A_{0}$. For $A\in \text{I}\text{P}\text{Alg}/k$, we define $\unicode[STIX]{x1D70B}_{0}(A)=A/A_{0}\in \text{I}\text{P}\text{FEt}/k$.

The category $\text{Alg}_{\text{uc}}/k$ is an artinian abelian category such that any object is killed by multiplication by some positive integer. Hence we can apply the results and notation in the previous subsection to ${\mathcal{A}}=\text{Alg}_{\text{uc}}/k$.

Proposition 2.2.1. For any $A\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$, we have $A_{0},\unicode[STIX]{x1D70B}_{0}(A)\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$.

Proof. We may assume that $A$ is unipotent. Consider the sequence

$$\begin{eqnarray}\displaystyle 0 & \rightarrow & \displaystyle A_{0}[p^{n}]\rightarrow A[p^{n}]\rightarrow \unicode[STIX]{x1D70B}_{0}(A)[p^{n}]\nonumber\\ \displaystyle & \rightarrow & \displaystyle A_{0}/p^{n}A_{0}\rightarrow A/p^{n}A\rightarrow \unicode[STIX]{x1D70B}_{0}(A)/p^{n}\unicode[STIX]{x1D70B}_{0}(A)\rightarrow 0\nonumber\end{eqnarray}$$

for any $n\geqslant 0$. Since $A[p^{n}]$ is quasi-algebraic and $\unicode[STIX]{x1D70B}_{0}(A)[p^{n}]$ is étale, the image of $A[p^{n}]\rightarrow \unicode[STIX]{x1D70B}_{0}(A)[p^{n}]$ is finite and hence $A_{0}[p^{n}]$ is quasi-algebraic. Therefore $A_{0}\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$ and so $A_{0}/p^{n}A_{0}$ is quasi-algebraic by Proposition 2.1.1. This implies that $\unicode[STIX]{x1D70B}_{0}(A)[p^{n}]$ is quasi-algebraic (i.e., finite étale) and $\unicode[STIX]{x1D70B}_{0}(A)\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$.◻

We define $\text{L}^{f}\text{Alg}_{\text{uc}}/k$ to be the full subcategory of $\text{I}^{f}\text{Alg}_{\text{uc}}/k$ of objects whose identity component is quasi-algebraic (i.e., belongs to $\text{Alg}_{\text{uc}}/k$). Equivalently, an object $A\in \text{L}^{f}\text{Alg}_{\text{uc}}/k$ is an extension of a torsion étale group of cofinite type ($=\unicode[STIX]{x1D70B}_{0}(A)$) by a connected unipotent quasi-algebraic group ($=A_{0}$). Such objects can naturally be viewed as objects of $\text{L}\text{Alg}/k$; see [Reference SuzukiSuz13, Section 2.1, Footnote 3].

Proposition 2.2.2. Let $A\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$. Then $(A_{0})_{\text{div}}=(A_{\text{div}})_{0}$. We denote these isomorphic objects by $A_{0\text{div}}$.

Proof. We may assume that $A$ is unipotent. It is enough to show that $(A_{0})_{\text{div}}$ is connected and $(A_{\text{div}})_{0}$ is divisible. Take a power $p^{n}$ of $p$ that kills $(A_{0})_{/\text{div}}\in \text{Alg}_{\text{uc}}/k$. Then multiplication by $p^{n}$ gives a surjection $A_{0}{\twoheadrightarrow}(A_{0})_{\text{div}}$. Hence $(A_{0})_{\text{div}}$ is connected. On the other hand, the exact sequence $0\rightarrow (A_{\text{div}})_{0}\rightarrow A_{\text{div}}\rightarrow \unicode[STIX]{x1D70B}_{0}(A_{\text{div}})\rightarrow 0$ and the snake lemma gives a surjection $\unicode[STIX]{x1D70B}_{0}(A_{\text{div}})[p]{\twoheadrightarrow}(A_{\text{div}})_{0}/p((A_{\text{div}})_{0})$. This implies that $(A_{\text{div}})_{0}/p((A_{\text{div}})_{0})$ is étale and connected, hence zero. Thus $(A_{\text{div}})_{0}$ is divisible.◻

An example of a connected divisible group in $\text{I}^{f}\text{Alg}_{\text{uc}}/k$ is the direct limit $\mathop{\varinjlim }\nolimits_{n}W_{n}$ of (perfections of) groups of $p$-typical Witt vectors of finite length. We have $T\,\mathop{\varinjlim }\nolimits_{n}W_{n}=\mathop{\varprojlim }\nolimits_{n}W_{n}$, which is the group $W$ of $p$-typical Witt vectors of infinite length.

Proposition 2.2.3. Let $A\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$. Then the following are equivalent.

  1. (1) $A\in \text{L}^{f}\text{Alg}_{\text{uc}}/k$.

  2. (2) $A_{0\text{div}}=0$.

  3. (3) $V_{p}A\in \text{I}\text{P}\text{FEt}/k$. (Recall that $V_{p}A$ is defined in Section 2.1.)

  4. (4) $V_{p}A(k^{\prime })$ as a functor on algebraically closed fields $k^{\prime }$ over $k$ is constant.

Proof. We may assume that $A$ is unipotent. The equivalence between (1) and (2) follows from the previous proposition. We know that $V_{p}A=V_{p}(A_{\text{div}})$ surjects onto $A_{\text{div}}$. Hence $(V_{p}A)_{0}=(V_{p}(A_{\text{div}}))_{0}$ surjects onto $A_{0\text{div}}$. This shows that (3) is equivalent to (2). Obviously (3) implies (4).

We show that (4) implies (2). We may assume that $k=\overline{k}$ and $A$ is divisible. Then the assumption implies that $V_{p}A(k^{\prime })$ as a functor on arbitrary perfect fields $k^{\prime }$ over $k$ is constant. The surjection $V_{p}A{\twoheadrightarrow}A$ implies that $A(k^{\prime })$ as a functor on perfect fields $k^{\prime }$ over $k$ is constant. This implies that $A_{0}(k^{\prime })$ as a functor on perfect fields $k^{\prime }$ over $k$ is constant. Hence we may further assume that $A$ is connected. We then want to show that $A=0$ if $A(k^{\prime })$ as a functor on perfect fields $k^{\prime }$ over $k$ is constant. Let $A_{n}=(A[p^{n}])_{0}\in \text{Alg}_{\text{uc}}/k$. Then $A=\mathop{\varinjlim }\nolimits_{n}A_{n}$. It follows that the generic point of $A_{n}$ for any $n$ maps to a $k$-value point of $A$ and hence to a $k$-valued point of $A_{m}$ for some $m\geqslant n$. This means that the injective morphism $A_{n}{\hookrightarrow}A_{m}$ of quasi-algebraic groups is generically constant. Therefore $A_{n}=0$ for any $n$ and $A=0$.◻

Proposition 2.2.4. The category $\text{L}^{f}\text{Alg}_{\text{uc}}/k$ is an abelian subcategory of $\text{I}^{f}\text{Alg}_{\text{uc}}/k$.

Proof. This follows from the equivalence between (1) and (3) (or (4)) of the previous proposition. ◻

Proposition 2.2.5. Let $A\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$ and define $A_{0\cap \text{div}}:=A_{0}\cap A_{\text{div}}$. Then $A_{0\cap \text{div}}/A_{0\text{div}}\in \text{FEt}_{p}/k$. We have exact sequences

$$\begin{eqnarray}\displaystyle & \displaystyle 0\rightarrow A_{0\cap \text{div}}\rightarrow A_{0}\rightarrow (A_{/\text{div}})_{0}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow A_{0\cap \text{div}}\rightarrow A_{\text{div}}\rightarrow (\unicode[STIX]{x1D70B}_{0}A)_{\text{div}}\rightarrow 0. & \displaystyle \nonumber\end{eqnarray}$$

Proof. Since $A_{0}/A_{0\text{div}}=(A_{0})_{/\text{div}}$ is quasi-algebraic unipotent, it is killed by some power $p^{n}$ of $p$. Hence $A_{0\cap \text{div}}/A_{0\text{div}}$ is a subgroup of the $p^{n}$-torsion part of $A_{\text{div}}/A_{0\text{div}}=\unicode[STIX]{x1D70B}_{0}(A_{\text{div}})$. But $\unicode[STIX]{x1D70B}_{0}(A_{\text{div}})$ is a torsion étale group of cofinite type. Hence its $p^{n}$-torsion part is finite. Therefore $A_{0\cap \text{div}}/A_{0\text{div}}$ is finite. The exactness of the sequences is clear.◻

The group $A_{0\cap \text{div}}$ in the proposition is in general nonzero even if $A\in \text{L}^{f}\text{Alg}_{\text{uc}}/k$. It is $\mathbb{Z}/p\mathbb{Z}$ if $A$ is the cokernel of the diagonal embedding of $\mathbb{Z}/p\mathbb{Z}$ into $\mathbf{G}_{a}\oplus (\mathbb{Q}_{p}/\mathbb{Z}_{p})$. In particular, $A_{0\cap \text{div}}$ can be neither connected nor divisible. On the other hand, $A_{0\text{div}}$ is always connected and divisible.

Proposition 2.2.6. For any $A\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$, we have $\unicode[STIX]{x1D70B}_{0}(A)_{/\text{div}}=\unicode[STIX]{x1D70B}_{0}(A_{/\text{div}})$. Denote these isomorphic objects by $\unicode[STIX]{x1D70B}_{0}A_{/\text{div}}$. The kernel of $A{\twoheadrightarrow}\unicode[STIX]{x1D70B}_{0}A_{/\text{div}}$ is $A_{0+\text{div}}:=A_{0}+A_{\text{div}}$.

Proof. Obvious. ◻

For example, if $A=\mathbf{G}_{a}\oplus \mathbb{Q}/\mathbb{Z}\oplus \mathbb{Z}/p\mathbb{Z}$, then $A_{0+\text{div}}=\mathbf{G}_{a}\oplus \mathbb{Q}/\mathbb{Z}$ and $\unicode[STIX]{x1D70B}_{0}A_{\text{div}}=\mathbb{Z}/p\mathbb{Z}$. To see a more nontrivial example, one may work out the subgroup $A$ of $\mathop{\varinjlim }\nolimits_{n}W_{n}=\{(\ldots ,x_{-2},x_{-1},x_{0})\mid x_{n}=0\text{ for almost all }n\}$ defined by $(F-1)x_{n}=(F-1)^{2}x_{-1}=0$ for all $n\leqslant -2$, where $F$ is the Frobenius. Then $A_{0}\cong \mathbf{G}_{a}$ is defined by $x_{n}=0$ for all $n\leqslant -1$, $A_{\text{div}}\cong \mathbb{Q}_{p}/\mathbb{Z}_{p}$ by $(F-1)x_{n}=0$ for all $n\leqslant 0$, $A_{0+\text{div}}$ by $(F-1)x_{n}=0$ for all $n\leqslant -1$, and $\unicode[STIX]{x1D70B}_{0}A_{/\text{div}}\cong \mathbb{Z}/p\mathbb{Z}$.

Here is a connection diagram between the named subgroups of $A\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$:

All the arrows are inclusions. The label of an arrow means that the subquotient there is of that type. For example, $A_{0+\text{div}}/A_{\text{div}}$ is connected quasi-algebraic. The symbol $\bigoplus _{l}\text{FEt}_{l}$ means that the (sub)quotient there is a direct sum of finite étale $l$-primary groups for primes $l$ (with $l=p$ allowed), which is finite if and only if $A$ is divisibly ML. The upper and lower sides of the square have the same subquotient, and the left and right sides of the square have the same subquotient. The two step subquotient $A_{0}/A_{0\text{div}}$ is connected quasi-algebraic and, similarly, $A_{\text{div}}/A_{0\text{div}}$ is étale divisible. In particular, we have $\unicode[STIX]{x1D70B}_{0}(A_{0+\text{div}})=\unicode[STIX]{x1D70B}_{0}(A)_{\text{div}}$ and $(A_{0+\text{div}})_{/\text{div}}=(A_{/\text{div}})_{0}$. Later we will see in Proposition 3.2.3 that the cohomology object $G$ in any positive degree of the complex $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ mentioned in Introduction belongs to $\text{I}^{f}\text{Alg}_{\text{uc}}/k$. With some more effort, we will see in Theorem 3.4.1 that $G_{0\text{div}}=0$ (i.e., $G\in \text{L}^{f}\text{Alg}_{\text{uc}}/k$) and $G/G_{0+\text{div}}$ is finite (i.e., $G$ is divisibly ML).

2.3 The ind-rational pro-étale site and some derived limits

We quickly recall some definitions and notation about the ind-rational (pro-)étale site from [Reference SuzukiSuz14, Section 2.1]. We say that a $k$-algebra is rational if it is a finite direct product of perfections (direct limit along Frobenii) of finitely generated fields over $k$, and ind-rational if it is a filtered union of rational $k$-subalgebras. We denote the category of rational (resp. ind-rational) $k$-algebras with $k$-algebra homomorphisms by $k^{\text{rat}}$ (resp. $k^{\text{ind}\text{rat}}$). (More precisely, we only consider ${\mathcal{U}}_{0}$-small ind-rational $k$-algebras. Hence $k^{\text{ind}\text{rat}}$ is a ${\mathcal{U}}$-small ${\mathcal{U}}_{0}$-category.) We can endow $k^{\text{ind}\text{rat}}$ with the étale topology. That is, an étale covering of $k^{\prime }\in k^{\text{ind}\text{rat}}$ is a finite family $\{k_{i}^{\prime }\}$ of étale $k^{\prime }$-algebras such that the product $\prod k_{i}^{\prime }$ is faithfully flat over $k^{\prime }$. The resulting site is the ind-rational étale site $\operatorname{Spec}k_{\text{et}}^{\text{ind}\text{rat}}$. We can also endow $k^{\text{ind}\text{rat}}$ with the pro-étale topology of Bhatt and Scholze [Reference Bhatt and ScholzeBS15]. That is, a pro-étale covering of $k^{\prime }\in k^{\text{ind}\text{rat}}$ is a finite family $\{k_{i}^{\prime }\}$ of $k^{\prime }$-algebras such that each $k_{i}^{\prime }$ is a filtered direct limit of étale $k^{\prime }$-algebras and the product $\prod k_{i}^{\prime }$ is faithfully flat over $k^{\prime }$. The resulting site is the ind-rational pro-étale site $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$. The category of sheaves of sets (resp. abelian groups) on $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ is denoted by $\operatorname{Set}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ (resp. $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$). (Here, the target categories $\operatorname{Set}$ and $\operatorname{Ab}$ for sheaves are the categories of ${\mathcal{U}}$-small sets and abelian groups.) The category of complexes in $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is denoted by $\operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, its homotopy category by $K(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ and its derived category by $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. See [Reference Kashiwara and SchapiraKS06, Chapter 18] for details about unbounded derived categories of sheaves on sites. The cohomology of $k^{\prime }\in k^{\text{ind}\text{rat}}$ with coefficients in $G\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is denoted by $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}}^{\prime },G)$, with $n$th cohomology $H^{n}(k_{\text{pro}\text{et}}^{\prime },G)$ when $G\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. The sheaf-Hom, $n$th sheaf-Ext and derived sheaf-Hom functors are denoted by $\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}$, $\mathbf{Ext}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}^{n}$ and $R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}$, respectively. Similar notation applies to $\operatorname{Spec}k_{\text{et}}^{\text{ind}\text{rat}}$. For $G\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, we define its Serre dual [Reference SuzukiSuz14, Section 2.4] to be

$$\begin{eqnarray}G^{\text{SD}}=R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}(G,\mathbb{Z}).\end{eqnarray}$$

See the list of examples of Serre duals in the paragraph after the proof of [Reference SuzukiSuz14, Proposition (2.4.1)]. A key example there is $\mathbf{G}_{a}^{\text{SD}}\cong \mathbf{G}_{a}[-2]$, which comes from $\mathbf{G}_{a}\overset{{\sim}}{\rightarrow }\mathbf{Ext}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}^{1}(\mathbf{G}_{a},\mathbb{Z}/p\mathbb{Z})$ defined by sending $c\in \mathbf{G}_{a}$ to the pullback of the Artin–Schreier extension class $0\rightarrow \mathbb{Z}/p\mathbb{Z}\rightarrow \mathbf{G}_{a}\rightarrow \mathbf{G}_{a}\rightarrow 0$ by the multiplication-by-$c$ map $\mathbf{G}_{a}\rightarrow \mathbf{G}_{a}$ to the final term. In particular, $\mathbf{G}_{a}$ is Serre reflexive. We say that $G$ is Serre reflexive if the canonical morphism $G\rightarrow G^{\text{SD}\text{SD}}$ is an isomorphism. For $G\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, we denote its torsion part by $G_{\text{tor}}$ and set $G_{/\text{tor}}=G/G_{\text{tor}}$. We say that a sheaf $F\in \operatorname{Set}(k_{\text{et}}^{\text{ind}\text{rat}})$ is locally of finite presentation if it commutes with filtered direct limits as a functor $k^{\text{ind}\text{rat}}\rightarrow \operatorname{Set}$ [Reference SuzukiSuz14, Section 2.4]. In this case, $F$ is automatically a sheaf for the pro-étale topology.

The natural Yoneda functor $\text{Alg}/k\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ extends to an additive functor $\text{I}\text{P}\text{Alg}/k\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. (Here, as before, we consider the ${\mathcal{U}}_{0}$-procategory of $\text{Alg}/k$ and the ${\mathcal{U}}_{0}$-indcategory of the ${\mathcal{U}}_{0}$-category $\text{P}\text{Alg}/k$.) This functor is exact by [Reference SuzukiSuz14, Proposition (2.1.2)(e)], which is fully faithful and induces a fully faithful embedding $D^{b}(\text{I}\text{P}\text{Alg}/k){\hookrightarrow}D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ by [Reference SuzukiSuz14, Proposition (2.3.4)]. If $G\in \text{I}\text{P}\text{Alg}/k$ is an extension of an object of $\text{I}\text{Alg}_{\text{uc}}/k$ by an object of $\text{P}\text{Alg}/k$, then we denote

$$\begin{eqnarray}G^{\text{SD}^{\prime }}=\mathbf{Ext}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}^{1}(G,\mathbb{Q}/\mathbb{Z}).\end{eqnarray}$$

By [Reference SuzukiSuz14, Proposition (2.4.1)(b)], we have $G^{\text{SD}}=G^{\text{SD}^{\prime }}[-2]$. The same proposition shows that if $G$ is connected unipotent proalgebraic or ind-algebraic, then $G^{\text{SD}^{\prime }}$ is connected unipotent ind-algebraic or proalgebraic, respectively, and $G\overset{{\sim}}{\rightarrow }G^{\text{SD}^{\prime }\text{SD}^{\prime }}$. In particular, $G$ is Serre reflexive. By [Reference SuzukiSuz14, Proposition (2.4.1)(b)], if $G$ is semiabelian, then $G^{\text{SD}^{\prime }}$ is the Pontryagin dual of the Tate module $TG$. If $G$ is not necessarily connected, then we denote $G_{0}^{\text{SD}^{\prime }}=(G_{0})^{\text{SD}^{\prime }}$. If $G$ is an extension of a torsion étale group by a profinite-étale group, then its Pontryagin dual is denoted by $G^{\text{PD}}$, which can be given by $\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}(G,\mathbb{Q}/\mathbb{Z})$, or by $G^{\text{SD}}[1]$ by [Reference SuzukiSuz14, Proposition (2.4.1)(b)]. For example, if $G=\mathbb{Q}_{p}$, then $G^{\text{PD}}\cong \mathbb{Q}_{p}$. If the torsion part of $G\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is such a group, then we denote $G_{\text{tor}}^{\text{PD}}=(G_{\text{tor}})^{\text{PD}}$. If $G$ is a lattice over $k$, then its dual lattice is denoted by $G^{\text{LD}}$, which can be given by $\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}(G,\mathbb{Z})$. If the torsion-free quotient of $G\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is such a group, then we denote $G_{/\text{tor}}^{\text{LD}}=(G_{/\text{tor}})^{\text{LD}}$.

For an abelian (${\mathcal{U}}$-)category, we say that it has exact products if the product $\prod _{\unicode[STIX]{x1D706}\in \unicode[STIX]{x1D6EC}}A_{\unicode[STIX]{x1D706}}$ of any family of objects $\{A_{\unicode[STIX]{x1D706}}\}$ (with $\unicode[STIX]{x1D6EC}\in {\mathcal{U}}$) exists and for any family of surjections $\{A_{\unicode[STIX]{x1D706}}{\twoheadrightarrow}{B_{\unicode[STIX]{x1D706}}\}}_{\unicode[STIX]{x1D706}\in \unicode[STIX]{x1D6EC}}$ (with $\unicode[STIX]{x1D6EC}\in {\mathcal{U}}$), the morphism $\prod A_{\unicode[STIX]{x1D706}}\rightarrow \prod B_{\unicode[STIX]{x1D706}}$ is surjective.

Proposition 2.3.1. The category $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ has enough projectives and exact products. For every sequence

$$\begin{eqnarray}\cdots \overset{\unicode[STIX]{x1D711}_{3}}{\rightarrow }A_{3}\overset{\unicode[STIX]{x1D711}_{2}}{\rightarrow }A_{2}\overset{\unicode[STIX]{x1D711}_{1}}{\rightarrow }A_{1}\end{eqnarray}$$

in $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, its derived inverse limit is represented by the complex $\prod _{n}A_{n}\rightarrow \prod _{n}A_{n}$ in degrees $0$ and $1$, where the morphism sends $(a_{n})_{n}$ to $(a_{n}-\unicode[STIX]{x1D711}_{n}(a_{n+1}))_{n}$. If the sequence is Mittag-Leffler, then its derived inverse limit is zero in positive degrees.

Proof. For any $k^{\prime }\in k^{\text{ind}\text{rat}}$, there exists an ind-étale faithfully flat homomorphism $k^{\prime }\rightarrow k^{\prime \prime }$ to a w-contractible ring $k^{\prime \prime }$ [Reference Bhatt and ScholzeBS15, Definition 2.4.1] by [Reference Bhatt and ScholzeBS15, Lemma 2.4.9]. An ind-étale algebra over an ind-rational $k$-algebra is ind-rational over $k$ by [Reference SuzukiSuz13, Proposition 2.1.2], so $k^{\prime \prime }\in k^{\text{ind}\text{rat}}$. By the definition of w-contractibility, we know that the sheaf of free abelian groups $\mathbb{Z}[\operatorname{Spec}k^{\prime \prime }]$ generated by the representable sheaf of sets $\operatorname{Spec}k^{\prime \prime }$ is a projective object of $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. The natural morphism $\mathbb{Z}[\operatorname{Spec}k^{\prime \prime }]\rightarrow \mathbb{Z}[\operatorname{Spec}k^{\prime }]$ is surjective. Since any object of $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is a quotient of a direct sum (with index set $\in {\mathcal{U}}$) of objects of the form $\mathbb{Z}[\operatorname{Spec}k^{\prime }]$ with $k^{\prime }\in k^{\text{ind}\text{rat}}$, it follows that $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ has enough projectives.

The category $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ has products. Having enough projectives, it has exact products. The stated complex calculates the derived limit by [Reference NeemanNee01, Remark A.3.6] or [Reference RoosRoo06, Lemma 2.2]. The statement about Mittag-Leffler sequences follows from [Reference RoosRoo06] or [Reference NeemanNee01, Lemma A.3.15]. ◻

Let $\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ be the procategory of $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. (Here we consider the ${\mathcal{U}}$-procategory since $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is a ${\mathcal{U}}$-category.) A filtered inverse system $\{{A_{\unicode[STIX]{x1D706}}\}}_{\unicode[STIX]{x1D706}}$ in $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ as an object of $\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is denoted by . The category $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is a Grothendieck category and hence has enough injectives. Therefore the procategory $\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ also has enough injectives by [Reference Kashiwara and SchapiraKS06, Exercise 13.6(ii)]. As $\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ has exact products (and exact filtered inverse limits), it follows from [Reference Kashiwara and SchapiraKS06, Theorem 14.4.4] that any additive functor from $\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ to an abelian category admits a right derived functor on (unbounded) derived categories, which can be calculated by K-injective (or homotopically injective) complexes in $\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. In particular, the inverse limit functor $\varprojlim :\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ admits a right derived functor $R\varprojlim :D(\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}))\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. For a more detailed treatment of $R\varprojlim$, see [Reference Kashiwara and SchapiraKS06, Proposition 13.3.15, Corollary 13.3.16, Example 13.3.17(i)].Footnote 1

For a sheaf $A\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, the multiplication maps by integers $n\geqslant 1$ yield a pro-object  with index set $\mathbb{Z}_{{\geqslant}1}$ ordered by divisibility. Note that there is a cofinal map $\unicode[STIX]{x1D70E}$ from the set of positive integers with the usual ordering ${\leqslant}$ to the set of positive integers with the divisibility ordering. For example, we can take $\unicode[STIX]{x1D70E}(n)$ to be the $n$th power of the product of the first $n$ primes. Hence the pro-object  is isomorphic to the sequence $\cdots \rightarrow A_{\unicode[STIX]{x1D70E}(2)}\rightarrow A_{\unicode[STIX]{x1D70E}(1)}$. The assignment  defines an exact functor $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\rightarrow \text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ and hence a triangulated functor $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\rightarrow D(\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}))$. By composing it with $R\varprojlim$, we have a triangulated endofunctor on $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, which we denote by $A\mapsto R\,\mathop{\varprojlim }\nolimits_{n}A$. The objects  and $R\,\mathop{\varprojlim }\nolimits_{n}A$ are uniquely divisible. We have a projection  to the $n=1$ term in $D(\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}))$. This induces a morphism $R\,\mathop{\varprojlim }\nolimits_{n}A\rightarrow A$ in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. It is an isomorphism if and only if $A$ is uniquely divisible. We have $R\,\mathop{\varprojlim }\nolimits_{n}A=0$ if $A$ is killed by multiplication by some positive integer.

For $A\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, we define $A_{\text{div}}$ to be the image of the natural morphism $\mathop{\varprojlim }\nolimits_{n}A\rightarrow A$. It is the maximal divisible subsheaf of $A$. We denote $A_{/\text{div}}=A/A_{\text{div}}$. We define

$$\begin{eqnarray}TA=\underset{n}{\varprojlim }\,A[n],\qquad VA=(TA)\otimes \mathbb{Q},\qquad \hat{A}=\underset{n}{\varprojlim }\,(A\otimes \mathbb{Z}/n\mathbb{Z}),\end{eqnarray}$$

where $A[n]$ is the $n$-torsion part of $A$. If $A\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$, then Section 2.1 gives an object $TA\in \text{P}\text{Alg}_{\text{uc}}/k$, while the above definition gives an object$TA\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. They are compatible under the Yoneda functor $\text{P}\text{Alg}_{\text{uc}}/k{\hookrightarrow}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. A similar relation holds for $VA$. We denote $TA_{\text{div}}=T(A_{\text{div}})$, $VA_{\text{div}}=V(A_{\text{div}})$. If $A_{\text{div}}$ is a torsion étale group, we denote $A_{\text{div}}^{\text{PD}}=(A_{\text{div}})^{\text{PD}}$. The notation $A_{/\text{div}}^{\wedge }$ means $(A_{/\text{div}})^{\wedge }$. For $A\in \text{FGEt}/k$, we denote $A_{/\text{tor}}^{\text{LD}\wedge }=(A_{/\text{tor}}^{\text{LD}})^{\wedge }$. (The general rule here is that we read the subscript first and then the superscript, from the inside to the outside. Sticking to one such convention reduces excessive usage of parentheses. But we are not strictly consistent with this rule, as the ind-rational pro-étale site $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ or the open subscheme ${\mathcal{A}}_{0}^{\vee }\subset {\mathcal{A}}^{\vee }$ of the Néron model shows.)

Proposition 2.3.2. For any $A\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, we have $R^{m}\,\mathop{\varprojlim }\nolimits_{n}A=0$ for $m\geqslant 2$. If $A$ is divisibly ML (Proposition 2.1.9), then $R\,\mathop{\varprojlim }\nolimits_{n}A=\mathop{\varprojlim }\nolimits_{n}A$ and $A_{/\text{div}}\overset{{\sim}}{\rightarrow }\hat{A}$. If $A_{\text{div}}$ is torsion, then $VA\overset{{\sim}}{\rightarrow }\,\mathop{\varprojlim }\nolimits_{n}A$.

Proof. As above, represent the pro-object  as a sequence $\cdots \rightarrow A_{\unicode[STIX]{x1D70E}(2)}\rightarrow A_{\unicode[STIX]{x1D70E}(1)}$. Then the vanishing of the derived limit in degree ${\geqslant}2$ follows from the previous proposition. If $A$ is divisibly ML, then $R\,\mathop{\varprojlim }\nolimits_{n}A=\,\mathop{\varprojlim }\nolimits_{n}A$ by the previous proposition. We then have $\hat{A}=(A_{/\text{div}})^{\wedge }=A_{/\text{div}}$. For general $A$, we have an exact sequence $0\rightarrow TA\rightarrow \mathop{\varprojlim }\nolimits_{n}A\rightarrow A_{\text{div}}\rightarrow 0$. Tensoring by $\mathbb{Q}$, we have an exact sequence $0\rightarrow VA\rightarrow \mathop{\varprojlim }\nolimits_{n}A\rightarrow A_{\text{div}}\otimes \mathbb{Q}\rightarrow 0$. Hence $VA=\mathop{\varprojlim }\nolimits_{n}A$ if $A_{\text{div}}\otimes \mathbb{Q}=0$.◻

Proposition 2.3.3. Let $A\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$ be divisibly ML. Then the morphism $VA\rightarrow A$ induces an isomorphism $R\,\mathop{\varprojlim }\nolimits_{n}A=VA$ in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$.

Proof. Recall that any object of $\text{Alg}_{\text{uc}}/k$ is killed by multiplication by some positive integer. Hence $A_{\text{div}}$ is torsion as a sheaf over $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$. (Recall that this means $A(k^{\prime })$ is a torsion abelian group for any $k^{\prime }\in k^{\text{ind}\text{rat}}$. See [AGV73, IX, Définition 1.1, Proposition 1.2(iii)].) Now the result follows from the previous proposition.◻

We denote $\mathbb{A}^{\infty }=\hat{\mathbb{Z}}\otimes _{\mathbb{Z}}\mathbb{Q}\in \text{I}\text{P}\text{FEt}/k$ (which is the adele ring of $\mathbb{Q}$ without components of infinite places). We have $\mathbb{A}^{\infty }=V(\mathbb{Q}/\mathbb{Z})$. The assumption that $A$ is divisibly ML in the above proposition cannot be dropped. For instance, applying $R\,\mathop{\varprojlim }\nolimits_{n}$ to the exact sequence $0\rightarrow \bigoplus _{p}\mathbb{Z}/p\mathbb{Z}\rightarrow \mathbb{Q}/\mathbb{Z}\rightarrow \mathbb{Q}/\mathbb{Z}\rightarrow 0$, we know that $R\,\mathop{\varprojlim }\nolimits_{n}\bigoplus _{p}\mathbb{Z}/p\mathbb{Z}$ is isomorphic to $[\mathbb{A}^{\infty }\rightarrow \mathbb{A}^{\infty }][-1]$, or $[\hat{\mathbb{Z}}\rightarrow \hat{\mathbb{Z}}][-1]\otimes \mathbb{Q}$, or $(\prod _{p}\mathbb{Z}/p\mathbb{Z})/(\bigoplus _{p}\mathbb{Z}/p\mathbb{Z})[-1]$.

Proposition 2.3.4. Let $A\in \text{FGEt}/k$. Then $R\,\mathop{\varprojlim }\nolimits_{n}A=A\otimes _{\mathbb{Z}}\mathbb{A}^{\infty }/\mathbb{Q}[-1]$ in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$.

Proof. We may assume that $A$ is torsion-free. We have an exact sequence $0\rightarrow A\rightarrow A\otimes \mathbb{Q}\rightarrow A\otimes \mathbb{Q}/\mathbb{Z}\rightarrow 0$. Applying $R\,\mathop{\varprojlim }\nolimits_{n}$ and using the previous proposition, we obtain a distinguished triangle $R\,\mathop{\varprojlim }\nolimits_{n}A\rightarrow A\otimes \mathbb{Q}\rightarrow A\otimes \mathbb{A}^{\infty }$. From this, the result follows.◻

For $A\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, there is a canonical choice of a mapping cone of the above defined morphism . We denote this cone by . It is an object of $D(\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}))$. (This construction can also be understood more conceptually using the derived functor of the procategory extension of the two variable functor $\otimes$, at least when $A$ is assumed to be bounded above.)

Proposition 2.3.5. Let $A\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. The distinguished triangle

induces an exact sequence

in $\text{P}\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ for any $i\in \mathbb{Z}$.

Proof. Obvious. ◻

Proposition 2.3.6. If a morphism $A\rightarrow B$ in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ induces an isomorphism

$$\begin{eqnarray}A\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\overset{{\sim}}{\rightarrow }B\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\end{eqnarray}$$

for any $n\geqslant 1$, then it also induces an isomorphism

Proof. This follows from the previous proposition. ◻

If $A\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, we can apply $R\varprojlim$ to . We denote the result by

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}).\end{eqnarray}$$

Proposition 2.3.7. For $A\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, we have a distinguished triangle

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,A\rightarrow A\rightarrow R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z}).\end{eqnarray}$$

If a morphism $A\rightarrow B$ in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ induces an isomorphism

$$\begin{eqnarray}A\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\overset{{\sim}}{\rightarrow }B\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\end{eqnarray}$$

for any $n\geqslant 1$, then it also induces an isomorphism

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\overset{{\sim}}{\rightarrow }R\,\underset{n}{\varprojlim }\,(B\otimes ^{L}\mathbb{Z}/n\mathbb{Z}).\end{eqnarray}$$

Proof. This follows from the previous proposition. ◻

Proposition 2.3.8. Let $A\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. Assume that

$$\begin{eqnarray}R^{i}\,\underset{n}{\varprojlim }\,(H^{j}(A)\otimes \mathbb{Z}/n\mathbb{Z})=R^{i}\,\underset{n}{\varprojlim }\,(H^{j}(A)[n])=0\end{eqnarray}$$

for any $i\geqslant 1$ and $j\in \mathbb{Z}$. Then we have an exact sequence

$$\begin{eqnarray}0\rightarrow H^{i}(A)^{\wedge }\rightarrow H^{i}R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\rightarrow TH^{i+1}(A)\rightarrow 0.\end{eqnarray}$$

for all $i\in \mathbb{Z}$.

Proof. By Proposition 2.3.5, we know that applying $R^{i}\varprojlim$ to  results zero for any $i\geqslant 1$ and $j\in \mathbb{Z}$. Hence the $i$th cohomology of $R\,\mathop{\varprojlim }\nolimits_{n}(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z})$ is given by applying $\varprojlim$ to the $i$th cohomology of  for any $i\in \mathbb{Z}$. Applying $\varprojlim$ to the exact sequence in Proposition 2.3.5, we get the required exact sequence.◻

Proposition 2.3.9. For any $A\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, the natural morphism

$$\begin{eqnarray}\Bigl(R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\Bigr)\otimes \mathbb{Q}\rightarrow R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Q}/\mathbb{Z})\end{eqnarray}$$

is an isomorphism. Denote these isomorphic objects by $A\,\hat{\otimes }\,\mathbb{A}^{\infty }$. We have a natural distinguished triangle

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,A\rightarrow A\otimes \mathbb{Q}\rightarrow A\,\hat{\otimes }\,\mathbb{A}^{\infty }.\end{eqnarray}$$

The assignment $A\mapsto A\,\hat{\otimes }\,\mathbb{A}^{\infty }$ is a triangulated endofunctor on $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ that sends $\mathbb{Z}$ to $\mathbb{A}^{\infty }$.

Proof. We have a distinguished triangle

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,A\rightarrow A\rightarrow R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\end{eqnarray}$$

and hence a distinguished triangle

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,A\rightarrow A\otimes \mathbb{Q}\rightarrow \Bigl(R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\Bigr)\otimes \mathbb{Q}.\end{eqnarray}$$

Also we have a distinguished triangle

$$\begin{eqnarray}A\rightarrow A\otimes \mathbb{Q}\rightarrow A\otimes ^{L}\mathbb{Q}/\mathbb{Z}\end{eqnarray}$$

and hence a distinguished triangle

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,A\rightarrow A\otimes \mathbb{Q}\rightarrow R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Q}/\mathbb{Z}).\end{eqnarray}$$

Comparing the two triangles, we get the result. ◻

The boundedness conditions in the following two propositions might be unnecessary.

Proposition 2.3.10. For $A\in D^{+}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, there is a canonical isomorphism

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z})=R\,\mathbf{Hom}_{k}(\mathbb{Q}/\mathbb{Z},A)[1].\end{eqnarray}$$

Proof. We have

$$\begin{eqnarray}R\,\mathbf{Hom}_{k}(\mathbb{Q},A)=R\,\underset{n}{\varprojlim }\,R\,\mathbf{Hom}_{k}(\mathbb{Z},A)=R\,\underset{n}{\varprojlim }\,A\end{eqnarray}$$

by the proof of [Reference SuzukiSuz14, Proposition (2.2.3)] (which needs the bounded below condition). Comparing the two distinguished triangles

$$\begin{eqnarray}\displaystyle & \displaystyle R\,\underset{n}{\varprojlim }\,A\rightarrow A\rightarrow R\,\underset{n}{\varprojlim }\,(A\otimes ^{L}\mathbb{Z}/n\mathbb{Z}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\,\mathbf{Hom}_{k}(\mathbb{Q},A)\rightarrow R\,\mathbf{Hom}_{k}(\mathbb{Z},A)\rightarrow R\,\mathbf{Hom}_{k}(\mathbb{Q}/\mathbb{Z},A)[1], & \displaystyle \nonumber\end{eqnarray}$$

we get the result. ◻

Proposition 2.3.11. For $A\in D^{-}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ and $B\in D^{+}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, there is a canonical isomorphism

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }(R\,\mathbf{Hom}_{k}(A,B)\otimes ^{L}\mathbb{Z}/n\mathbb{Z})=R\,\mathbf{Hom}_{k}(A\otimes ^{L}\mathbb{Q}/\mathbb{Z},B)[1].\end{eqnarray}$$

Proof. We know that $R\,\mathbf{Hom}_{k}(A,B)$ is bounded below. By the previous proposition, the left-hand side can be written as

$$\begin{eqnarray}R\,\mathbf{Hom}_{k}(\mathbb{Q}/\mathbb{Z},R\,\mathbf{Hom}_{k}(A,B))[1].\end{eqnarray}$$

This is also the right-hand side by the derived tensor-Hom adjunction. ◻

2.4 Premorphisms of sites

For a site $S$, we denote its category of sheaves of sets (resp. abelian groups) by $\operatorname{Set}(S)$ (resp. $\operatorname{Ab}(S)$). The category of complexes in $\operatorname{Ab}(S)$ is $\operatorname{Ch}(S)$, its homotopy category $K(S)$ and derived category $D(S)$. We say $A\in \operatorname{Ch}(S)$ is K-limp if $R\unicode[STIX]{x1D6E4}(X,A)=\unicode[STIX]{x1D6E4}(X,A)$ for any object $X\in S$, where $\unicode[STIX]{x1D6E4}$ in the right-hand side is applied termwise. See also [Reference SpaltensteinSpa88, Corollary 5.17] and [Reference SchnürerSch17, Appendix A]. This is not a “K”-version of limp sheaf as defined in [Sta18, Tag 072Y] since $X$ is not allowed to be an arbitrary sheaf of sets but only a representable one. K-injectives are K-limp.

The following three propositions are essentially well known, at least for bounded below complexes of sheaves or just sheaves; see [Reference MilneMil80, III, Sections 1, 2].

Proposition 2.4.1. Let $S$ be a site and $A,J\in \operatorname{Ch}(S)$ complexes such that $J$ is K-injective. Then the total complex of the sheaf-Hom double complex $\mathbf{Hom}_{S}(A,J)$ is K-limp.

Proof. Note that $\mathbf{Hom}_{S}(A,J)=R\,\mathbf{Hom}_{S}(A,J)$. By [Reference Kashiwara and SchapiraKS06, Proposition 18.6.6], we have $R\unicode[STIX]{x1D6E4}(X,R\,\mathbf{Hom}_{S}(A,J))=R\operatorname{Hom}_{S/X}(A,J)$ for any $X\in S$, where $S/X$ is the localization of $S$ at $X$. Hence

$$\begin{eqnarray}\displaystyle & & \displaystyle R\unicode[STIX]{x1D6E4}(X,\mathbf{Hom}_{S}(A,J))=R\operatorname{Hom}_{S/X}(A,J)\nonumber\\ \displaystyle \hspace{74.39996pt} & & \displaystyle \quad =\operatorname{Hom}_{S/X}(A,J)=\unicode[STIX]{x1D6E4}(X,\mathbf{Hom}_{S}(A,J)).\hspace{48.0pt}\square\nonumber\end{eqnarray}$$

Let $f^{-1}$ be a functor from the underlying category of a site $S$ to the underlying category of another site $S^{\prime }$. The right composition with $f^{-1}$ defines a functor from the category of presheaves of sets on $S^{\prime }$ to the category of presheaves of sets on $S$, which is the pushforward functor for presheaves. If this functor sends sheaves to sheaves, $f^{-1}$ is called a continuous functor in the terminology of [AGV72, III, Definition 1.1]. We say that $f^{-1}$ then defines a continuous map of sites $f:S^{\prime }\rightarrow S$. The pushforward functor $\operatorname{Set}(S^{\prime })\rightarrow \operatorname{Set}(S)$ or $\operatorname{Ab}(S^{\prime })\rightarrow \operatorname{Ab}(S)$ is denoted by $f_{\ast }$. The left adjoint to $f_{\ast }:\operatorname{Set}(S^{\prime })\rightarrow \operatorname{Set}(S)$ exists by [AGV72, III, Proposition 1.2], which we denote by $f^{\ast \text{set}}:\operatorname{Set}(S)\rightarrow \operatorname{Set}(S^{\prime })$. The left adjoint to $f_{\ast }:\operatorname{Ab}(S^{\prime })\rightarrow \operatorname{Ab}(S)$ exists by [AGV72, III, Proposition 1.7], which we denote by $f^{\ast }:\operatorname{Set}(S)\rightarrow \operatorname{Set}(S^{\prime })$. They are called pullback functors. The continuous map $f:S^{\prime }\rightarrow S$ is called a morphism of sites if the pullback $f^{\ast \text{set}}:\operatorname{Set}(S)\rightarrow \operatorname{Set}(S^{\prime })$ for sheaves of sets is exact. In this case, the pullback for sheaves of abelian groups $f^{\ast }:\operatorname{Ab}(S)\rightarrow \operatorname{Ab}(S^{\prime })$ and $f^{\ast \text{set}}$ are compatible with forgetting group structures by [AGV72, III, Proposition 1.7, 4)], so we do not have to distinguish $f^{\ast }$ and $f^{\ast \text{set}}$. In general, we use $f^{\ast }$ to mean the pullback functor for sheaves of abelian groups.

Proposition 2.4.2. Let $f:S^{\prime }\rightarrow S$ be a continuous map of sites. For any K-limp complex $A^{\prime }$ in $\operatorname{Ab}(S^{\prime })$, we have $Rf_{\ast }A^{\prime }=f_{\ast }A^{\prime }$ in $D(S)$.

Proof. Let $A^{\prime }\overset{{\sim}}{\rightarrow }I^{\prime }$ be a K-injective replacement. Let $B^{\prime }$ be the mapping cone of $A^{\prime }\rightarrow I^{\prime }$, which is an exact complex and hence $Rf_{\ast }B^{\prime }=0$ in $D(S)$. For any $X^{\prime }\in S^{\prime }$, consider the distinguished triangle $\unicode[STIX]{x1D6E4}(X^{\prime },A^{\prime })\rightarrow \unicode[STIX]{x1D6E4}(X^{\prime },I^{\prime })\rightarrow \unicode[STIX]{x1D6E4}(X^{\prime },B^{\prime })$. In $D(\operatorname{Ab})$, we have $\unicode[STIX]{x1D6E4}(X^{\prime },A^{\prime })=R\unicode[STIX]{x1D6E4}(X^{\prime },A^{\prime })$ by assumption and $\unicode[STIX]{x1D6E4}(X^{\prime },I^{\prime })=R\unicode[STIX]{x1D6E4}(X^{\prime },I^{\prime })$ by the K-injectivity of $I^{\prime }$. Hence $\unicode[STIX]{x1D6E4}(X^{\prime },B^{\prime })=R\unicode[STIX]{x1D6E4}(X^{\prime },B^{\prime })$, which is zero in $D(\operatorname{Ab})$ since $B^{\prime }$ is exact. This means that $B^{\prime }$ is also exact as a complex of presheaves. Hence $f_{\ast }B^{\prime }$ is exact. Consider the distinguished triangle $f_{\ast }A^{\prime }\rightarrow f_{\ast }I^{\prime }\rightarrow f_{\ast }B^{\prime }$. We have $f_{\ast }I^{\prime }=Rf_{\ast }I^{\prime }$ and $f_{\ast }B^{\prime }=Rf_{\ast }B^{\prime }=0$, so $f_{\ast }A^{\prime }=Rf_{\ast }A^{\prime }$.◻

If $S$ and $S^{\prime }$ are sites defined by pretopologies, and if $f^{-1}$ is a functor from the underlying category of $S$ to that of $S^{\prime }$ that sends coverings to coverings and $f^{-1}(Y\times _{X}Z)=f^{-1}(Y)\times _{f^{-1}(X)}f^{-1}(Z)$ whenever $Y\rightarrow X$ appears in a covering family, then $f^{-1}$ is called a morphism of topologies in the terminology of [Reference ArtinArt62, Definition 2.4.2] and defines a continuous map $f:S^{\prime }\rightarrow S$. We call such a continuous map a premorphism of sites. In this case, $f_{\ast }$ sends acyclic sheaves (i.e., those $A^{\prime }$ with $H^{n}(X^{\prime },A^{\prime })=0$ for any object $X^{\prime }$ of $S^{\prime }$ and $n\geqslant 1$) to acyclic sheaves and hence induces the Leray spectral sequence $R\unicode[STIX]{x1D6E4}(X,Rf_{\ast }A^{\prime })=R\unicode[STIX]{x1D6E4}(f^{-1}(X),A^{\prime })$ for any $X\in S$ and $A^{\prime }\in \operatorname{Ab}(S^{\prime })$ [Reference ArtinArt62, Section II.4]. Here is a slight generalization.

Proposition 2.4.3. Let $f:S^{\prime }\rightarrow S$ be a premorphism of sites defined by pretopologies. Then $f_{\ast }$ sends K-limp complexes to K-limp complexes. For any $X\in S$ and $A^{\prime }\in D(S^{\prime })$, we have an isomorphism

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(X,Rf_{\ast }A^{\prime })=R\unicode[STIX]{x1D6E4}(f^{-1}X,A^{\prime }),\end{eqnarray}$$

where $f^{-1}:S\rightarrow S^{\prime }$ is the underlying functor of $f$.

Proof. By [Reference SuzukiSuz13, Lemma 3.7.2], we know that $f^{\ast }:\operatorname{Ab}(S)\rightarrow \operatorname{Ab}(S^{\prime })$ admits a left derived functor $Lf^{\ast }:D(S)\rightarrow D(S^{\prime })$, which is left adjoint to $Rf_{\ast }:D(S^{\prime })\rightarrow D(S)$ and sends the sheaf $\mathbb{Z}[X]$ of free abelian groups generated by any $X\in S$ to $\mathbb{Z}[f^{-1}X]$. For any $A^{\prime }\in \operatorname{Ch}(S^{\prime })$ and $X\in S$, we have

$$\begin{eqnarray}\displaystyle R\unicode[STIX]{x1D6E4}(X,Rf_{\ast }A^{\prime }) & = & \displaystyle R\operatorname{Hom}_{S}(\mathbb{Z}[X],Rf_{\ast }A^{\prime })=R\operatorname{Hom}_{S^{\prime }}(Lf^{\ast }\mathbb{Z}[X],A^{\prime })\nonumber\\ \displaystyle & = & \displaystyle R\operatorname{Hom}_{S^{\prime }}(\mathbb{Z}[f^{-1}X],A^{\prime })=R\unicode[STIX]{x1D6E4}(f^{-1}X,A^{\prime }).\nonumber\end{eqnarray}$$

If $A^{\prime }$ is K-limp, then $Rf_{\ast }A^{\prime }=f_{\ast }A^{\prime }$ by Proposition 2.4.2 and $R\unicode[STIX]{x1D6E4}(f^{-1}X,A^{\prime })=\unicode[STIX]{x1D6E4}(f^{-1}X,A^{\prime })$. Hence $R\unicode[STIX]{x1D6E4}(X,f_{\ast }A^{\prime })=\unicode[STIX]{x1D6E4}(f^{-1}X,A^{\prime })=\unicode[STIX]{x1D6E4}(X,f_{\ast }A^{\prime })$, and so $f_{\ast }A^{\prime }$ is K-limp.◻

2.5 Local duality without relative sites

Let $\hat{K}_{x}$ be a complete discrete valuation field with perfect residue field $k_{x}$ ($\in {\mathcal{U}}_{0}$) of characteristic $p>0$. The ring of integers is denoted by $\hat{{\mathcal{O}}}_{x}$ with maximal ideal $\hat{\mathfrak{p}}_{x}$. By the fppf site $\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}$ of $\hat{{\mathcal{O}}}_{x}$, we mean the category of (${\mathcal{U}}_{0}$-small) $\hat{{\mathcal{O}}}_{x}$-algebras endowed with the fppf topology. The same applies to the fppf site $\operatorname{Spec}\hat{K}_{x,\text{fppf}}$ of $\hat{K}_{x}$. Sheaves take values in $\operatorname{Set}$ or $\operatorname{Ab}$ (of ${\mathcal{U}}$-small sets or abelian groups). The Hom functor and the sheaf-Hom functor for $\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}$ are denoted by $\operatorname{Hom}_{\hat{{\mathcal{O}}}_{x}}$ and $\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}$, respectively. Their derived functors are denoted by $\operatorname{Ext}_{\hat{{\mathcal{O}}}_{x}}^{n}$, $\mathbf{Ext}_{\hat{{\mathcal{O}}}_{x}}^{n}$, $R\operatorname{Hom}_{\hat{{\mathcal{O}}}_{x}}$, $R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}$. Similar notation applies to $\operatorname{Spec}\hat{K}_{x,\text{fppf}}$. We denote $k_{x}^{\text{ind}\text{rat}}=(k_{x})^{\text{ind}\text{rat}}$, $\operatorname{Spec}k_{x,\text{et}}^{\text{ind}\text{rat}}=\operatorname{Spec}(k_{x})_{\text{et}}^{\text{ind}\text{rat}}$ and $\operatorname{Spec}k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}=\operatorname{Spec}(k_{x})_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$.

The ring $\hat{{\mathcal{O}}}_{x}$ has a natural structure of a $W(k_{x})$-algebra of profinite length, where $W$ denotes the ring of $p$-typical Witt vectors of infinite length. In equal characteristic, this structure factors through $k_{x}$ so that $\hat{{\mathcal{O}}}_{x}$ is profinite length over $k_{x}$. In mixed characteristic, it is finite free over $W(k_{x})$. For $k_{x}^{\prime }\in k_{x}^{\text{ind}\text{rat}}$, we define

$$\begin{eqnarray}\displaystyle & \displaystyle \hat{\mathbf{O}}_{x}(k_{x}^{\prime })=W(k_{x}^{\prime })\hat{\otimes }_{W(k_{x})}\hat{{\mathcal{O}}}_{x}=\underset{n}{\varprojlim }(W_{n}(k_{x}^{\prime })\otimes _{W_{n}(k_{x})}\hat{{\mathcal{O}}}_{x}/\hat{\mathfrak{p}}_{x}^{n}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle \hat{\mathbf{K}}_{x}(k_{x}^{\prime })=\hat{\mathbf{O}}_{x}(k_{x}^{\prime })\otimes _{\hat{{\mathcal{O}}}_{x}}\hat{K}_{x}. & \displaystyle \nonumber\end{eqnarray}$$

Proposition 2.5.1. The functors $k_{x}^{\prime }\mapsto \hat{\mathbf{O}}_{x}(k_{x}^{\prime }),\hat{\mathbf{K}}_{x}(k_{x}^{\prime })$ define premorphisms of sites

$$\begin{eqnarray}\displaystyle \unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}:\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}\rightarrow \operatorname{Spec}k_{x,\text{et}}^{\text{ind}\text{rat}}, & & \displaystyle \nonumber\\ \displaystyle \unicode[STIX]{x1D70B}_{\hat{K}_{x}}:\operatorname{Spec}\hat{K}_{x,\text{fppf}}\rightarrow \operatorname{Spec}k_{x,\text{et}}^{\text{ind}\text{rat}}. & & \displaystyle \nonumber\end{eqnarray}$$

Proof. From the proof of [Reference SuzukiSuz13, Proposition 2.3.1], we know that if $k_{x}^{\prime }\rightarrow k_{x}^{\prime \prime }$ is a (faithfully flat) étale homomorphism in $k_{x}^{\text{ind}\text{rat}}$, then $\hat{\mathbf{O}}_{x}(k_{x}^{\prime })\rightarrow \hat{\mathbf{O}}_{x}(k_{x}^{\prime \prime })$ is (faithfully flat) étale, and if $k_{x}^{\prime \prime \prime }\in k_{x}^{\text{ind}\text{rat}}$ is another object, then $\hat{\mathbf{O}}_{x}(k_{x}^{\prime \prime })\otimes _{\hat{\mathbf{O}}_{x}(k_{x}^{\prime })}\hat{\mathbf{O}}_{x}(k_{x}^{\prime \prime \prime })$ is isomorphic to $\hat{\mathbf{O}}_{x}(k_{x}^{\prime \prime }\otimes _{k_{x}^{\prime }}k_{x}^{\prime \prime \prime })$. Hence $\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}$ is a premorphism of sites. So is $\unicode[STIX]{x1D70B}_{\hat{K}_{x}}$.◻

The scheme morphism $j:\operatorname{Spec}\hat{K}_{x}{\hookrightarrow}\operatorname{Spec}\hat{{\mathcal{O}}}_{x}$ defines a morphism of sites

$$\begin{eqnarray}j:\operatorname{Spec}\hat{K}_{x,\text{fppf}}\rightarrow \operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}.\end{eqnarray}$$

Its pullback functor $j^{\ast }$ is the restriction functor for sheaves from $\hat{{\mathcal{O}}}_{x}$ to $\hat{K}_{x}$. This restriction functor is frequently omitted by abuse of notation, so a sheaf (or a complex of sheaves) $F$ on $\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}$ restricted to $\operatorname{Spec}\hat{K}_{x,\text{fppf}}$ will frequently be written by just $F$. More generally, if $f:Z\rightarrow Y$ is a morphism of schemes, then the restriction $f^{\ast }F$ of a sheaf (or a complex of sheaves) $F$ on $Y_{\text{fppf}}$ will frequently be written by just $F$. But the important point is that $f$ is a localization morphism [AGV72, III, Section 5]. Recall from [AGV72, III, Section 5] that for a site $S$ and its object $V$, the category $S/V$ of pairs $(W,g)$ ($W$ an object of $S$ and $g:W\rightarrow V$ a morphism in $S$) is equipped with the natural induced topology. The functor $j_{V}:(W,g)\mapsto W$ defines a continuous map of sites $S\rightarrow S/V$ whose pushforward functor $\operatorname{Set}(S)\rightarrow \operatorname{Set}(S/V)$ is the restriction functor. In our situation, by $f$ being a localization morphism, we mean that $f:Z\rightarrow Y$ belongs to the underlying category of $Y_{\text{fppf}}$ and $Z_{\text{fppf}}$ can be identified with $Y_{\text{fppf}}/Z$. Hence $f^{\ast }$ admits an exact left adjoint $f_{!}:\operatorname{Ab}(Z_{\text{fppf}})\rightarrow \operatorname{Ab}(Y_{\text{fppf}})$ by [AGV72, IV, Proposition 11.3.1]. Hence $f^{\ast }$ send (K-)injectives to (K-)injectives. This will be needed to pass from sheaf or complex level statements to derived categorical statements.

We have the pushforward functors

$$\begin{eqnarray}\displaystyle & \displaystyle (\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}})_{\ast }:\operatorname{Ab}(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow \operatorname{Ab}(k_{x,\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle (\unicode[STIX]{x1D70B}_{\hat{K}_{x}})_{\ast }:\operatorname{Ab}(\hat{K}_{x,\text{fppf}})\rightarrow \operatorname{Ab}(k_{x,\text{et}}^{\text{ind}\text{rat}}) & \displaystyle \nonumber\end{eqnarray}$$

by $\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}$ and $\unicode[STIX]{x1D70B}_{\hat{K}_{x}}$. Let $\operatorname{Ab}(k_{x,\text{et}}^{\text{ind}\text{rat}})\rightarrow \operatorname{Ab}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$ be the pro-étale sheafification functor. The composite functors of these functors are denoted by

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;):\operatorname{Ab}(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow \operatorname{Ab}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\;\cdot \;):\operatorname{Ab}(\hat{K}_{x,\text{fppf}})\rightarrow \operatorname{Ab}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}). & \displaystyle \nonumber\end{eqnarray}$$

We have the right derived functors

$$\begin{eqnarray}\displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;):D(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\;\cdot \;):D(\hat{K}_{x,\text{fppf}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}). & \displaystyle \nonumber\end{eqnarray}$$

Since sheafification is exact, $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$ is the composite of $R(\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}})_{\ast }$ and the pro-étale sheafification functor $D(k_{x,\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$. The same is true for $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\;\cdot \;)$. We denote $\mathbf{H}^{n}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)=H^{n}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$ and similarly $\mathbf{H}^{n}(\hat{K}_{x},\;\cdot \;)$. By Proposition A.1, these functors are compatible with the functors $R\tilde{\boldsymbol{\unicode[STIX]{x1D6E4}}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$ and $R\tilde{\boldsymbol{\unicode[STIX]{x1D6E4}}}(\hat{K}_{x},\;\cdot \;)$ in the notation of [Reference SuzukiSuz14, paragraph before Proposition (3.3.8)].

The pro-étale sheafification in general makes it hard to calculate the derived global section $R\unicode[STIX]{x1D6E4}(k_{x,\text{pro}\text{et}}^{\prime },\;\cdot \;)$ of the objects $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},G)$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},G)$ at each $k_{x}^{\prime }\in k_{x}^{\text{ind}\text{rat}}$ for a general complex $G$. In general, it is not clear whether the natural morphisms

$$\begin{eqnarray}\displaystyle & \displaystyle R\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},G)\rightarrow R\unicode[STIX]{x1D6E4}(k_{x,\text{pro}\text{et}},R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},G)) & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\unicode[STIX]{x1D6E4}(\hat{K}_{x},G)\rightarrow R\unicode[STIX]{x1D6E4}(k_{x,\text{pro}\text{et}},R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},G)) & \displaystyle \nonumber\end{eqnarray}$$

are isomorphisms or not. The first (resp. second) morphism is an isomorphism if $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},G)$ (resp. $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},G)$) is P-acyclic in the sense of [Reference SuzukiSuz14, Section 2.4]. This condition is satisfied for $G$ a smooth group scheme or a finite flat group scheme by [Reference SuzukiSuz14, Propositions (3.4.2), (3.4.3)]. This is why the notion of P-acyclicity is introduced in [Reference SuzukiSuz14, Section 2.4]. The following proposition shows we can eliminate P-acyclicity from [Reference SuzukiSuz14] to a certain extent.

Proposition 2.5.2. Let $G\in D(\hat{{\mathcal{O}}}_{x,\text{fppf}})$. For any w-contractible $k_{x}^{\prime }\in k_{x}^{\text{ind}\text{rat}}$ ([Reference Bhatt and ScholzeBS15, Definition 2.4.1], which includes any algebraically closed field over $k_{x}$), we have

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(k_{x,\text{pro}\text{et}}^{\prime },R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},G))=R\unicode[STIX]{x1D6E4}(\hat{\mathbf{O}}_{x}(k_{x}^{\prime }),G).\end{eqnarray}$$

A similar relation holds with $\hat{{\mathcal{O}}}_{x}$ replaced by $\hat{K}_{x}$.

Proof. We have

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(k_{x,\text{et}}^{\prime },R(\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}})_{\ast }G)=R\unicode[STIX]{x1D6E4}(\hat{\mathbf{O}}_{x}(k_{x}^{\prime }),G)\end{eqnarray}$$

by Proposition 2.4.3. Hence it is enough to show that $R\unicode[STIX]{x1D6E4}(k_{x,\text{pro}\text{et}}^{\prime },\tilde{F})=R\unicode[STIX]{x1D6E4}(k_{x,\text{et}}^{\prime },F)$ for any $F\in D(k_{\text{et}}^{\text{ind}\text{rat}})$, where $\tilde{F}$ is the pro-étale sheafification of $F$. The section functor $\unicode[STIX]{x1D6E4}(k_{x}^{\prime },\;\cdot \;)$ is exact on $\operatorname{Ab}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$ and $\operatorname{Ab}(k_{x,\text{et}}^{\text{ind}\text{rat}})$ since $k_{x}^{\prime }$ is w-contractible. Hence it is enough to show that $\unicode[STIX]{x1D6E4}(k_{x}^{\prime },\tilde{F})=\unicode[STIX]{x1D6E4}(k_{x}^{\prime },F)$ for any $F\in \operatorname{Ab}(k_{x,\text{et}}^{\text{ind}\text{rat}})$. Sheafification is given by applying the zeroth Čech cohomology presheaf functor $\text{}\underline{\check{H}}^{0}$ twice [Reference MilneMil80, III, Remark 2.2(c)]. Any pro-étale covering of a w-contractible scheme has a section by definition and the Čech complex for a covering with section is null-homotopic. Hence for any presheaf $F$ on $\operatorname{Spec}k_{x,\text{pro}\text{et}}^{\prime }$, we have

$$\begin{eqnarray}\hspace{72.60004pt}\unicode[STIX]{x1D6E4}(k_{x}^{\prime },\text{}\underline{\check{H}}^{0}\text{}\underline{\check{H}}^{0}(F))=\unicode[STIX]{x1D6E4}(k_{x}^{\prime },\text{}\underline{\check{H}}^{0}(F))=\unicode[STIX]{x1D6E4}(k_{x}^{\prime },F).\hspace{44.64001pt}\square\end{eqnarray}$$

We define a functor by the mapping fiber construction

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\;\cdot \;):=[\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},j^{\ast }\;\cdot \;)][-1]:\operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow \operatorname{Ch}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

on the category of complexes of sheaves. This is an additive functor that commutes with the translation functors, that is, a functor of additive categories with translation in the terminology of [Reference Kashiwara and SchapiraKS06, Definition 10.1.1]. It induces a functor on the homotopy categories by [Reference Kashiwara and SchapiraKS06, Proposition 11.2.9]. We have its right derived functor

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)=[R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},j^{\ast }\;\cdot \;)][-1]:D(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

by [Reference Kashiwara and SchapiraKS06, Theorem 14.3.1(vi)]. We set $\mathbf{H}_{x}^{n}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)=H^{n}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$. Note that $\mathbf{H}_{x}^{0}\neq \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}$ in this definition. (It can be shown that they define the same right derived functors by the same method as [Reference SuzukiSuz14, Proposition (3.3.3)].) By Proposition A.2, this is compatible with the functor denoted by $R\tilde{\boldsymbol{\unicode[STIX]{x1D6E4}}}_{x}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$ in [Reference SuzukiSuz14, paragraph before Proposition (3.3.8)]. We frequently omit the restriction functor $j^{\ast }$ from this type of formulas by abuse of notation.

Let $A,B\in \operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}})$. For any $k_{x}^{\prime }\in k_{x}^{\text{ind}\text{rat}}$, the functoriality of $\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$ gives a morphism

$$\begin{eqnarray}\operatorname{Hom}_{\hat{\mathbf{O}}_{x}(k_{x}^{\prime })}(A,B)\rightarrow \operatorname{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}/k_{x}^{\prime }}(\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},A),\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},B))\end{eqnarray}$$

in $\operatorname{Ch}(\operatorname{Ab})$, where $\operatorname{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}/k_{x}^{\prime }}$ is the Hom functor for the category of sheaves on the localization of $\operatorname{Spec}k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}$ at $k_{x}^{\prime }$. This is functorial on $k_{x}^{\prime }$, so we have a morphism

(2.5.1)$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}(A,B))\rightarrow \mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},A),\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},B))\end{eqnarray}$$

in $\operatorname{Ch}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$. The right derived functor of the left-hand side (as a functor on $A,B$) is

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}(A,B))\end{eqnarray}$$

by Propositions 2.4.1, 2.4.2 and the theorem on derived functors of composition [Reference Kashiwara and SchapiraKS06, Proposition 10.3.5(ii)]. Hence we have a morphism

(2.5.2)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}(A,B))\rightarrow R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},A),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},B))\end{eqnarray}$$

in $D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$, functorial on $A,B\in D(\hat{{\mathcal{O}}}_{x,\text{fppf}})$, by the universal property of the right derived functor. (Note that the right-hand side of this is not the right derived functor of the right-hand side of (2.5.1). The problem is that $\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}$ is only a premorphism of sites and hence its pushforward functor might not send (K-)injectives to (K-)injectives. Sheafification also might not send (K-)injectives to (K-)injectives.) Similarly, we have a morphism

(2.5.3)$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}(A,B))\rightarrow \mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},A),\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},B))\end{eqnarray}$$

in $\operatorname{Ch}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$, functorial on $A,B\in \operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}})$. Deriving, we have a morphism

(2.5.4)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}(A,B))\rightarrow R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},A),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},B))\end{eqnarray}$$

in $D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$, functorial on $A,B\in D(\hat{{\mathcal{O}}}_{x,\text{fppf}})$. Also we have a morphism

(2.5.5)$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\mathbf{Hom}_{\hat{K}_{x}}(A,B))\rightarrow \mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A),\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},B))\end{eqnarray}$$

in $\operatorname{Ch}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$, functorial on $A,B\in \operatorname{Ch}(\hat{K}_{x,\text{fppf}})$. Deriving, we have a morphism

(2.5.6)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},R\,\mathbf{Hom}_{\hat{K}_{x}}(A,B))\rightarrow R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A),R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},B))\end{eqnarray}$$

in $D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$, functorial on $A,B\in D(\hat{K}_{x,\text{fppf}})$. These morphisms are compatible with the morphisms in [Reference SuzukiSuz14, Proposition (3.3.8)] by Proposition A.3.

With the comparison results in Appendix A, we can translate the results of [Reference SuzukiSuz14] to our setting. In particular, by [Reference SuzukiSuz14, Propositions (3.4.2)(a), (3.4.3)(e); Section 4.1], any term in the localization triangle

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\mathbf{G}_{m})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\mathbf{G}_{m})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\mathbf{G}_{m})[1]\end{eqnarray}$$

is concentrated in degree zero, where we have an exact sequence

$$\begin{eqnarray}0\rightarrow \hat{\mathbf{O}}_{x}^{\times }\rightarrow \hat{\mathbf{K}}_{x}^{\times }\rightarrow \mathbb{Z}\rightarrow 0.\end{eqnarray}$$

We call the (iso)morphisms

(2.5.7)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\mathbf{G}_{m})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\mathbf{G}_{m})[1]=\mathbb{Z}\end{eqnarray}$$

the trace (iso)morphisms (at $x$).

Let $A$ be an abelian variety over $\hat{K}_{x}$ with dual $A^{\vee }$. By the Barsotti–Weil formula [Reference OortOor66, Chapter III, Theorem (18.1)], we have a canonical isomorphism $A^{\vee }\cong \mathbf{Ext}_{\hat{K}_{x}}^{1}(A,\mathbf{G}_{m})$ defined by the Poincaré bundle. With $\mathbf{Hom}_{\hat{K}_{x}}(A,\mathbf{G}_{m})=0$, we have a canonical morphism $A^{\vee }\rightarrow R\,\mathbf{Hom}_{\hat{K}_{x}}(A,\mathbf{G}_{m})[1]$ in $D(\hat{K}_{x,\text{fppf}})$. Hence we have a morphism

$$\begin{eqnarray}\displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A^{\vee }) & \rightarrow & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},R\,\mathbf{Hom}_{\hat{K}_{x}}(A,\mathbf{G}_{m}))[1]\nonumber\\ \displaystyle & \rightarrow & \displaystyle R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A),R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\mathbf{G}_{m}))[1]\nonumber\\ \displaystyle & \rightarrow & \displaystyle R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A),\mathbb{Z})[1]=R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A)^{\text{SD}_{x}}[1]\nonumber\end{eqnarray}$$

by (2.5.6) and (2.5.7), where $\text{SD}_{x}=R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(\;\cdot \;,\mathbb{Z})$. Its Serre dual, when $A$ is replaced by $A^{\vee }$, is denoted by

$$\begin{eqnarray}\unicode[STIX]{x1D703}_{A}:R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A^{\vee })^{\text{SD}_{x}\text{SD}_{x}}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A)^{\text{SD}_{x}}[1].\end{eqnarray}$$

These morphisms agree with the morphisms defined in [Reference SuzukiSuz14, Section 4.1]. Hence [Reference SuzukiSuz14, Theorem (4.1.2)] implies that $\unicode[STIX]{x1D703}_{A}$ is an isomorphism.

Proposition 2.5.3. Let $A$ be an abelian variety over $\hat{K}_{x}$ with Néron model ${\mathcal{A}}$. Let ${\mathcal{A}}_{0}$ be the maximal open subgroup scheme of ${\mathcal{A}}$ with connected fibers and ${\mathcal{A}}_{x}$ the special fiber of ${\mathcal{A}}$ over $x=\operatorname{Spec}k_{x}$. Then $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})$ is concentrated in degree $0$; $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0})$ in degree $0$; $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})$ in degree $2$; and $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0})$ in degrees $1,2$. We have

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})=\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A)\in \text{P}\text{Alg}/k_{x},\qquad \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0})=\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A)_{0}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle \mathbf{H}_{x}^{1}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0})=\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\in \text{FEt}/k_{x}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle \mathbf{H}_{x}^{2}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})=\mathbf{H}_{x}^{2}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0})=\mathbf{H}^{1}(\hat{K}_{x},A)\in \text{I}\text{Alg}_{\text{uc}}/k_{x}. & \displaystyle \nonumber\end{eqnarray}$$

The isomorphic groups in the third line are divisible.

Proof. The group $\mathbf{H}^{1}(\hat{K}_{x},A)$ is divisible since $\mathbf{H}^{2}(\hat{K}_{x},A[n])=0$ for any $n\geqslant 1$ by [Reference SuzukiSuz14, Proposition (3.4.3)(b)]. The statements about $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})$, $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0})$ follow from [Reference SuzukiSuz14, Propositions (3.4.2)(a), (3.4.3)(d)]. These propositions, the localization triangle

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A)\end{eqnarray}$$

and the similar triangle for ${\mathcal{A}}_{0}$ imply the rest of the statements.◻

We recall [Reference SuzukiSuz14, Remark (4.2.10)]. Let ${\mathcal{A}}$ and ${\mathcal{A}}^{\vee }$ be the Néron models of $A$ and $A^{\vee }$, respectively. Let ${\mathcal{A}}_{0}^{\vee }$ be the maximal open subgroup scheme of ${\mathcal{A}}^{\vee }$ with connected fibers. The Poincaré biextension $A^{\vee }\otimes ^{L}A\rightarrow \mathbf{G}_{m}[1]$ as a morphism in $D(\hat{K}_{x,\text{fppf}})$ canonically extends to a biextension ${\mathcal{A}}_{0}^{\vee }\otimes ^{L}{\mathcal{A}}\rightarrow \mathbf{G}_{m}[1]$ as a morphism in $D(\hat{{\mathcal{O}}}_{x,\text{fppf}})$ by [Gro72, IX, 1.4.3]. Hence we have a morphism ${\mathcal{A}}_{0}^{\vee }\rightarrow R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}({\mathcal{A}},\mathbf{G}_{m}[1])$. With the functoriality morphism (2.5.3) and the trace isomorphism (2.5.7), we have morphisms

$$\begin{eqnarray}\displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0}^{\vee }) & \rightarrow & \displaystyle R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},\mathbf{G}_{m}[1]))\nonumber\\ \displaystyle & = & \displaystyle R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}),\mathbb{Z})=R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})^{\text{SD}_{x}}.\nonumber\end{eqnarray}$$

Proposition 2.5.4. The diagram

is a morphism of distinguished triangles. Applying $\text{SD}_{x}\text{SD}_{x}$, the induced diagram

is an isomorphism of distinguished triangles.

Proof. By [Reference SuzukiSuz14, Proposition (3.3.8)], we know that the first diagram is a morphism of distinguished triangles. The terms in the first diagram can be identified with the terms in the first diagram of [Reference SuzukiSuz14, Proposition (4.2.3)]. The uniqueness stated in [Reference SuzukiSuz14, Proposition (4.2.3)] shows that the morphisms in the first diagram here and the morphisms in the first diagram of [Reference SuzukiSuz14, Proposition (4.2.3)] are equal. Hence the second diagram is an isomorphism of distinguished triangles by [Reference SuzukiSuz14, Proposition (4.2.7), Theorem (4.1.2)]. ◻

Assume that $k_{x}$ is a finite extension of another perfect field $k$. We have a finite étale morphism $f_{x}:\operatorname{Spec}k_{x}\rightarrow \operatorname{Spec}k$. This induces a morphism of sites

$$\begin{eqnarray}f_{x}:\operatorname{Spec}k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}\rightarrow \operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}\end{eqnarray}$$

Its pushforward functor is the Weil restriction functor $\operatorname{Res}_{k_{x}/k}$, which is exact [Reference Bhatt and ScholzeBS15, Lemma 6.1.17]. We denote the composites of $\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$, $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)$ and $(f_{x})_{\ast }=\operatorname{Res}_{k_{x}/k}$ by

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;):\operatorname{Ab}(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;):D(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\end{eqnarray}$$

respectively, with cohomology $\mathbf{H}^{n}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)=H^{n}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)$. Similar functors

$$\begin{eqnarray}\displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\;\cdot \;):\operatorname{Ab}(\hat{K}_{x,\text{fppf}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}), & & \displaystyle \nonumber\\ \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;):\operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow \operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}) & & \displaystyle \nonumber\end{eqnarray}$$

and their derived functor are defined. We have

$$\begin{eqnarray}(f_{x})_{\ast }R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(\;\cdot \;,\mathbb{Z})=R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}((f_{x})_{\ast }(\;\cdot \;),\mathbb{Z})\end{eqnarray}$$

by the duality for finite étale morphisms [Reference MilneMil80, V, Proposition 1.13] (which is for the étale topology; the same proof works for the pro-étale topology). In other words, we have

$$\begin{eqnarray}\operatorname{Res}_{k_{x}/k}\,\circ \,\text{SD}_{x}=\text{SD}\,\circ \,\operatorname{Res}_{k_{x}/k},\end{eqnarray}$$

where $\text{SD}=R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}(\;\cdot \;,\mathbb{Z})$ as before. Hence applying $\operatorname{Res}_{k_{x}/k}$ to the duality statements above over $k_{x}$ defines some new duality statements over $k$. Doing this for $\unicode[STIX]{x1D703}_{A}$ for instance defines a new isomorphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A^{\vee })^{\text{SD}\text{SD}}\overset{{\sim}}{\rightarrow }R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A)^{\text{SD}}\end{eqnarray}$$

in $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. (Here we are not using any specific property of our duality isomorphism. We are using nothing but the obvious fact that if a morphism $C\rightarrow D$ in $D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is an isomorphism, then the induced morphism $\operatorname{Res}_{k_{x}/k}C\rightarrow \operatorname{Res}_{k_{x}/k}D$ is an isomorphism.) More explicitly, this morphism comes from the morphisms

$$\begin{eqnarray}\displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A^{\vee }) & \rightarrow & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,R\,\mathbf{Hom}_{\hat{K}_{x}}(A,\mathbf{G}_{m}))[1]\nonumber\\ \displaystyle & \rightarrow & \displaystyle R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A),R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\mathbf{G}_{m}))[1]\nonumber\\ \displaystyle & \rightarrow & \displaystyle R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A),\mathbb{Z})[1]=R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A)^{\text{SD}}[1]\nonumber\end{eqnarray}$$

The trace morphism in this situation used here is

(2.5.8)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\mathbf{G}_{m})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\mathbf{G}_{m})[1]=\operatorname{Res}_{k_{x}/k}\hat{\mathbf{K}}_{x}^{\times }/\hat{\mathbf{O}}_{x}^{\times }=\operatorname{Res}_{k_{x}/k}\mathbb{Z}\rightarrow \mathbb{Z},\end{eqnarray}$$

where the last morphism is

$$\begin{eqnarray}\mathbb{Z}[\operatorname{Hom}_{k}(k_{x},\overline{k})]\cong \mathbb{Z}^{[k_{x}:k]}\overset{\text{sum}}{{\twoheadrightarrow}}\mathbb{Z}\end{eqnarray}$$

on geometric points.

2.6 Henselizations and completions

Let ${\mathcal{O}}_{x}^{h}$ be an excellent Henselian discrete valuation ring of equal characteristic $p>0$ with perfect residue field $k_{x}$ ($\in {\mathcal{U}}_{0}$). We denote the maximal ideal by $\mathfrak{p}_{x}^{h}$ and the fraction field by $K_{x}^{h}$. The corresponding objects after completion are denoted by $\hat{{\mathcal{O}}}_{x}$, $\hat{\mathfrak{p}}_{x}$ and $\hat{K}_{x}$.

We make a variant of the constructions in the previous subsection using Henselizations instead of completions. For $k_{x}^{\prime }\in k_{x}^{\text{ind}\text{rat}}$, the Henselization of the (nonlocal) ring $k_{x}^{\prime }\otimes _{k_{x}}{\mathcal{O}}_{x}^{h}$ at the ideal $k_{x}^{\prime }\otimes _{k_{x}}\mathfrak{p}_{x}^{h}$ [Reference RaynaudRay70b, Chapter XI, Section 2] is denoted by $(k_{x}^{\prime }\otimes _{k_{x}}{\mathcal{O}}_{x}^{h})^{h}$. We define

$$\begin{eqnarray}\mathbf{O}_{x}^{h}(k_{x}^{\prime })=(k_{x}^{\prime }\otimes _{k_{x}}{\mathcal{O}}_{x}^{h})^{h},\qquad \mathbf{K}_{x}^{h}(k_{x}^{\prime })=\mathbf{O}_{x}^{h}(k_{x}^{\prime })\otimes _{{\mathcal{O}}_{x}^{h}}K_{x}^{h}.\end{eqnarray}$$

The functors $\mathbf{O}_{x}^{h}$ and $\mathbf{K}_{x}^{h}$ define premorphisms of sites

$$\begin{eqnarray}\unicode[STIX]{x1D70B}_{{\mathcal{O}}_{x}^{h}}:\operatorname{Spec}{\mathcal{O}}_{x,\text{fppf}}^{h}\rightarrow \operatorname{Spec}k_{x,\text{et}}^{\text{ind}\text{rat}},\qquad \unicode[STIX]{x1D70B}_{K_{x}^{h}}:\operatorname{Spec}K_{x,\text{fppf}}^{h}\rightarrow \operatorname{Spec}k_{x,\text{et}}^{\text{ind}\text{rat}}.\end{eqnarray}$$

The pro-étale sheafifications of their pushforward functors are denoted by

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h},\;\cdot \;),:\operatorname{Ab}({\mathcal{O}}_{x,\text{fppf}}^{h})\rightarrow \operatorname{Ab}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x}^{h},\;\cdot \;):\operatorname{Ab}(K_{x,\text{fppf}}^{h})\rightarrow \operatorname{Ab}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}). & \displaystyle \nonumber\end{eqnarray}$$

We set

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)=[\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h},\;\cdot \;),\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x}^{h},j^{\ast }\;\cdot \;)][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle \operatorname{Ch}({\mathcal{O}}_{x,\text{fppf}}^{h})\rightarrow \operatorname{Ch}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\end{eqnarray}$$

where $j:\operatorname{Spec}K_{x}^{h}\rightarrow \operatorname{Spec}{\mathcal{O}}_{x}^{h}$ is the natural morphism inducing a morphism on the fppf sites. We have their right derived functors

$$\begin{eqnarray}\displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h},\;\cdot \;),:D({\mathcal{O}}_{x,\text{fppf}}^{h})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x}^{h},\;\cdot \;),:D(K_{x,\text{fppf}}^{h})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)=[R\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h},\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x}^{h},j^{\ast }\;\cdot \;)][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle D({\mathcal{O}}_{x,\text{fppf}}^{h})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}). & \displaystyle \nonumber\end{eqnarray}$$

Again, we frequently omit the restriction functor $j^{\ast }$ by abuse of notation.

We say that a sheaf $F\in \operatorname{Set}({\mathcal{O}}_{x,\text{fppf}}^{h})$ is locally of finite presentation if it commutes with filtered direct limits as a functor on the category of ${\mathcal{O}}_{x}^{h}$-algebras.

Proposition 2.6.1. For any sheaf $A\in \operatorname{Ab}({\mathcal{O}}_{x,\text{fppf}}^{h})$ locally of finite presentation and $n\geqslant 0$, the sheaf $R^{n}(\unicode[STIX]{x1D70B}_{{\mathcal{O}}_{x}^{h}})_{\ast }A\in \operatorname{Ab}(k_{x,\text{et}}^{\text{ind}\text{rat}})$ is locally of finite presentation. In particular, we have $\mathbf{H}^{n}({\mathcal{O}}_{x}^{h},A)=R^{n}(\unicode[STIX]{x1D70B}_{{\mathcal{O}}_{x}^{h}})_{\ast }A$. That is, $\mathbf{H}^{n}({\mathcal{O}}_{x}^{h},A)$ is the étale sheafification of the presheaf

$$\begin{eqnarray}k_{x}^{\prime }\mapsto H^{n}(\mathbf{O}_{x}^{h}(k_{x}^{\prime }),A).\end{eqnarray}$$

A similar statement holds when ${\mathcal{O}}_{x}^{h}$ is replaced by $K_{x}^{h}$.

Proof. The sheaf $\mathbf{O}_{x}^{h}$ is locally of finite presentation by the construction of Henselization. Let $A\in \operatorname{Ab}({\mathcal{O}}_{x,\text{fppf}}^{h})$ be locally of finite presentation. Let $k_{x}^{\prime }=\bigcup k_{x,\unicode[STIX]{x1D706}}^{\prime }\in k^{\text{ind}\text{rat}}$ with $k_{x,\unicode[STIX]{x1D706}}^{\prime }\in k_{x}^{\text{rat}}$. Then for any $n\geqslant 0$, we have

$$\begin{eqnarray}H^{n}(\mathbf{O}_{x}^{h}(k_{x}^{\prime }),A)=\underset{\unicode[STIX]{x1D706}}{\varinjlim }\,H^{n}(\mathbf{O}_{x}^{h}(k_{x,\unicode[STIX]{x1D706}}^{\prime }),A).\end{eqnarray}$$

By sheafification, we know that $R^{n}(\unicode[STIX]{x1D70B}_{{\mathcal{O}}_{x}^{h}})_{\ast }A$ is locally of finite presentation. The same proof works for $K_{x}^{h}$.◻

Define a functor $\operatorname{Ch}({\mathcal{O}}_{x,\text{fppf}}^{h})\rightarrow \operatorname{Ch}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$ of additive categories with translation by

$$\begin{eqnarray}o_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)=[\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},f^{\ast }\;\cdot \;)],\end{eqnarray}$$

where $f:\operatorname{Spec}\hat{{\mathcal{O}}}_{x}\rightarrow \operatorname{Spec}{\mathcal{O}}_{x}^{h}$ is the natural scheme morphism inducing a morphism on the fppf sites. We call its right derived functor

$$\begin{eqnarray}Ro_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)=[R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},f^{\ast }\;\cdot \;)]\end{eqnarray}$$

the obstruction for cohomological approximation (at $x$). Define $D({\mathcal{O}}_{x,\text{fppf}}^{h})_{\text{ca}}$ to be the kernel of the functor $Ro_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)$, that is, the full subcategory of $D({\mathcal{O}}_{x,\text{fppf}}^{h})$ consisting of objects $A$ with $Ro_{x}({\mathcal{O}}_{x}^{h},A)=0$, or

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h},A)=R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},f^{\ast }A).\end{eqnarray}$$

Such an object $A$ is said to satisfy cohomological approximation.

Proposition 2.6.2. Any smooth group scheme or finite flat group scheme over ${\mathcal{O}}_{x}^{h}$ satisfies cohomological approximation.

Proof. Any finite flat group scheme $N$ is a closed subgroup scheme of some smooth affine group scheme $G$ by [Reference BégueriBég81, Proposition 2.2.1]. The fppf quotient $H=G/N$ is a smooth affine group scheme by descent. We have a distinguished triangle $Ro_{x}({\mathcal{O}}_{x}^{h},N)\rightarrow Ro_{x}({\mathcal{O}}_{x}^{h},G)\rightarrow Ro_{x}({\mathcal{O}}_{x}^{h},H)$. If two of the terms are zero, then so is the other. Hence the finite flat case is reduced to the smooth case.

Assume that $A$ is a smooth group scheme over ${\mathcal{O}}_{x}^{h}$. Let $N\subset A$ be the schematic closure of the identity section of $A\times _{{\mathcal{O}}_{x}^{h}}K_{x}^{h}$. Then $N$ is an étale group scheme over ${\mathcal{O}}_{x}^{h}$ with trivial generic fiber and $A/N$ is a separated smooth group scheme over ${\mathcal{O}}_{x}^{h}$ by [Reference RaynaudRay70a, Proposition 3.3.5]. The both objects $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h},N)$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},N)$ are isomorphic to $N\times _{{\mathcal{O}}_{x}^{h}}k_{x}$.

Hence we may assume that $A$ is smooth separated. Then we have $\mathbf{H}_{x}^{0}(\hat{{\mathcal{O}}}_{x},A)=0$. Also $\mathbf{H}^{n}(\hat{{\mathcal{O}}}_{x},A)=0$ for $n\geqslant 1$ by [Reference SuzukiSuz14, Proposition (3.4.2)(a)]. Hence

$$\begin{eqnarray}\displaystyle & \displaystyle \mathbf{H}_{x}^{1}(\hat{{\mathcal{O}}}_{x},A)=\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A)/\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},A), & \displaystyle \nonumber\\ \displaystyle & \displaystyle \mathbf{H}_{x}^{n}(\hat{{\mathcal{O}}}_{x},A)=\mathbf{H}^{n-1}(\hat{K}_{x},A) & \displaystyle \nonumber\end{eqnarray}$$

for $n\geqslant 2$. These sheaves are locally of finite presentation (even before pro-étale sheafification) by the Greenberg approximation argument [Reference SuzukiSuz14, Propositions (3.2.8), (3.2.9)]. On the other hand, the sheaves $\mathbf{H}^{n}({\mathcal{O}}_{x}^{h},A)$, $\mathbf{H}^{n}(K_{x}^{h},A)$ and thus $\mathbf{H}_{x}^{n}({\mathcal{O}}_{x}^{h},A)$ are locally of finite presentation for $n\geqslant 0$ by the previous proposition. Therefore it is enough to show that the morphism

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}_{x}({\mathcal{O}}_{x}^{h},A)\rightarrow R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},A)\end{eqnarray}$$

is an isomorphism when $k$ is algebraically closed by [Reference SuzukiSuz14, the second paragraph after Proposition (2.4.1)]. The statements to prove are

$$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D6E4}(K_{x}^{h},A)/\unicode[STIX]{x1D6E4}({\mathcal{O}}_{x}^{h},A)=\unicode[STIX]{x1D6E4}(\hat{K}_{x},A)/\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},A), & \displaystyle \nonumber\\ \displaystyle & \displaystyle H^{n}(K_{x}^{h},A)=H^{n}(\hat{K}_{x},A),\quad n\geqslant 1. & \displaystyle \nonumber\end{eqnarray}$$

They can be proven in the same way as the Greenberg approximation argument [Reference SuzukiSuz14, Propositions (3.2.8), (3.2.9)] (or equivalently, by the same argument as [Reference MilneMil06, I, Remark 3.10] combined with [Reference ČesnavičiusČes15, Proposition 3.5(b)]). ◻

Perhaps any group scheme locally of finite type over ${\mathcal{O}}_{x}^{h}$ might satisfy cohomological approximation since the Greenberg approximation itself holds in this generality. We do not pursue this point. But see [Reference Demarche and HarariDH18, Lemma 2.6, Remark 2.7] for a related result.

If $k_{x}$ is a finite extension of another perfect field $k$ and $f:\operatorname{Spec}k_{x}\rightarrow \operatorname{Spec}k$ is the natural morphism, then the composite of $\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h},\;\cdot \;)$ and $f_{\ast }$ is denoted by

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;):\operatorname{Ch}({\mathcal{O}}_{x,\text{fppf}}^{h})\rightarrow \operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

with right derived functor

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;):D({\mathcal{O}}_{x,\text{fppf}}^{h})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}).\end{eqnarray}$$

Similar notation applies to other objects, defining $\boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x}^{h}/k,\;\cdot \;)$, $\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;)$, $o_{x}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;)$ and their derived functors.

2.7 The fppf site of a curve over the rational étale site of the base

Let $U$ be a smooth geometrically connected curve over a perfect field $k$ ($\in {\mathcal{U}}_{0}$) of characteristic $p>0$, with smooth compactification $X$ and function field $K$. By the fppf site $U_{\text{fppf}}$ of $U$, we mean the category of (${\mathcal{U}}_{0}$-small) $U$-schemes endowed with the fppf topology. For $k^{\prime }\in k^{\text{ind}\text{rat}}$, we denote $U_{k^{\prime }}=U\times _{k}k^{\prime }$. The functor sending $k^{\prime }\in k^{\text{ind}\text{rat}}$ to $U_{k^{\prime }}$ defines a premorphism of sites

$$\begin{eqnarray}\unicode[STIX]{x1D70B}_{U}:U_{\text{fppf}}\rightarrow \operatorname{Spec}k_{\text{et}}^{\text{ind}\text{rat}}.\end{eqnarray}$$

We define a left exact functor

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;):\operatorname{Ab}(U_{\text{fppf}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

by the composite of the pushforward functor $\unicode[STIX]{x1D70B}_{U\ast }$ and the pro-étale sheafification. We have its right derived functor

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;):D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}),\end{eqnarray}$$

with cohomologies $\mathbf{H}^{n}=H^{n}R\boldsymbol{\unicode[STIX]{x1D6E4}}$. For any $A\in \operatorname{Ab}(U_{\text{fppf}})$ and $n\geqslant 0$, the sheaf $\mathbf{H}^{n}(U,A)$ on $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ is the pro-étale sheafification of the presheaf

$$\begin{eqnarray}k^{\prime }\mapsto H^{n}(U_{k^{\prime }},A).\end{eqnarray}$$

Let $\operatorname{Spec}k_{\text{et}}^{\text{perf}}$ be the category of (${\mathcal{U}}_{0}$-small) perfect $k$-schemes endowed with the étale topology. The structure morphism $U\rightarrow \operatorname{Spec}k$ induces a morphism of sites

$$\begin{eqnarray}\unicode[STIX]{x1D70B}_{U}^{\text{perf}}:U_{\text{fppf}}\rightarrow \operatorname{Spec}k_{\text{et}}^{\text{perf}},\end{eqnarray}$$

which is defined by the functor that sends a perfect $k$-scheme $S$ to $U\times _{k}S$. In [Reference Artin and Milne.AM76, (3.1)], this is denoted by $\unicode[STIX]{x1D70B}^{f,p}$.

Proposition 2.7.1. Let

$$\begin{eqnarray}f:\operatorname{Spec}k_{\text{et}}^{\text{perf}}\rightarrow \operatorname{Spec}k_{\text{et}}^{\text{ind}\text{rat}}\end{eqnarray}$$

be the premorphism of sites defined by the identity functor. Then we have $\unicode[STIX]{x1D70B}_{U}=f\,\circ \,\unicode[STIX]{x1D70B}_{U}^{\text{perf}}$. The functor $f_{\ast }$ is exact. The functors $\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)$ are the pro-étale sheafifications of $f_{\ast }\unicode[STIX]{x1D70B}_{U\ast }^{\text{perf}}$ and $f_{\ast }R\unicode[STIX]{x1D70B}_{U\ast }^{\text{perf}}$, respectively.

Proof. Obvious. ◻

In this sense, our constructions are pro-étale sheafifications of restrictions (from all perfect schemes to only ind-rational algebras) of the constructions in [Reference Artin and Milne.AM76]. In the next section, we will translate results in [Reference Artin and Milne.AM76] to our setting in this way.

For a closed point $x\in X$, let $k_{x}$, ${\mathcal{O}}_{x}^{h}$, $\hat{{\mathcal{O}}}_{x}$, $K_{x}^{h}$, $\hat{K}_{x}$ be the residue field, the Henselian local ring, the completed local ring, their fraction fields, respectively, at $x$. The results and notation in the previous two subsections apply to $\hat{{\mathcal{O}}}_{x}$ and ${\mathcal{O}}_{x}^{h}$.

Assume $x\in U$. For any $k^{\prime }\in k^{\text{ind}\text{rat}}$, we have $\mathbf{O}_{x}^{h}(k^{\prime }\otimes _{k}k_{x})=(k^{\prime }\otimes _{k}{\mathcal{O}}_{x}^{h})^{h}.$ Hence the morphism $\operatorname{Spec}{\mathcal{O}}_{x}^{h}\rightarrow U$ induces a morphism $\operatorname{Spec}\mathbf{O}_{x}^{h}(k^{\prime }\otimes _{k}k_{x})\rightarrow U_{k^{\prime }}$. This induces a homomorphism

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}(U_{k^{\prime }},A)\rightarrow \unicode[STIX]{x1D6E4}(\mathbf{O}_{x}^{h}(k^{\prime }\otimes _{k}k_{x}),A)\end{eqnarray}$$

for any $A\in \operatorname{Ab}(U_{\text{fppf}})$. Thus we have a morphism

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;)\end{eqnarray}$$

of left exact functors $\operatorname{Ab}(U_{\text{fppf}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ and a morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;)\end{eqnarray}$$

of triangulated functors $D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$.

Assume $x\not \in U$. For any $k^{\prime }\in k^{\text{ind}\text{rat}}$, we similarly have a morphism $\operatorname{Spec}\mathbf{K}_{x}^{h}(k^{\prime }\otimes _{k}k_{x})\rightarrow U_{k^{\prime }}$. This induces a homomorphism

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}(U_{k^{\prime }},A)\rightarrow \unicode[STIX]{x1D6E4}(\mathbf{K}_{x}^{h}(k^{\prime }\otimes _{k}k_{x}),A)\end{eqnarray}$$

for any $A\in \operatorname{Ab}(U_{\text{fppf}})$. Thus we have a morphism

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x}^{h}/k,\;\cdot \;)\end{eqnarray}$$

of left exact functors $\operatorname{Ab}(U_{\text{fppf}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ and a morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x}^{h}/k,\;\cdot \;)\end{eqnarray}$$

of triangulated functors $D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$.

For a dense open subscheme $V\subset U$, we have a morphism

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(V,\;\cdot \;)\end{eqnarray}$$

of left exact functors $\operatorname{Ab}(U_{\text{fppf}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ and a morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(V,\;\cdot \;)\end{eqnarray}$$

of triangulated functors $D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. Let $Z=U\setminus V$, which is a finite set of closed points of $X$. We define a functor of additive categories with translation by

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{Z}(U,\;\cdot \;)=[\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(V,\;\cdot \;)][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle \operatorname{Ch}(U_{\text{fppf}})\rightarrow \operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}). & \displaystyle \nonumber\end{eqnarray}$$

We have its right derived functor

$$\begin{eqnarray}\displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{Z}(U,\;\cdot \;)=[R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(V,\;\cdot \;)][-1]: & & \displaystyle \nonumber\\ \displaystyle D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}). & & \displaystyle \nonumber\end{eqnarray}$$

When $Z=\{x\}$, these are also denoted by $\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(U,\;\cdot \;)$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(U,\;\cdot \;)$.

Proposition 2.7.2. Let $V\subset U$ be a dense open subscheme and set $Z=U\setminus V$. The diagram

of left exact functors $\operatorname{Ab}(U_{\text{fppf}})\rightarrow \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is commutative. The induced morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{Z}(U,\;\cdot \;)\rightarrow \bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;)\end{eqnarray}$$

of triangulated functors $D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is an isomorphism.

Proof. The first statement about the commutativity is obvious. For the second statement, it is enough to treat the case that $Z$ is a singleton $\{x\}$. We want to show that the morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(U,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;)\end{eqnarray}$$

is an isomorphism. Let $f_{x}:\operatorname{Spec}k_{x}\rightarrow \operatorname{Spec}k$ be the natural morphism. The above morphism is the pro-étale sheafification of

$$\begin{eqnarray}[R\unicode[STIX]{x1D70B}_{U\ast }\rightarrow R\unicode[STIX]{x1D70B}_{V\ast }][-1]\rightarrow f_{x\ast }[R(\unicode[STIX]{x1D70B}_{{\mathcal{O}}_{x}^{h}})_{\ast }\rightarrow R(\unicode[STIX]{x1D70B}_{K_{x}^{h}})_{\ast }][-1].\end{eqnarray}$$

Applying $R\unicode[STIX]{x1D6E4}(k^{\prime },\;\cdot \;)$ for $k^{\prime }\in k^{\text{ind}\text{rat}}$ to this morphism before sheafification, we have a morphism

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}_{x}(U_{k^{\prime }},\;\cdot \;)\rightarrow R\unicode[STIX]{x1D6E4}_{x}((k^{\prime }\otimes _{k}{\mathcal{O}}_{x}^{h})^{h},\;\cdot \;),\end{eqnarray}$$

where the left-hand (resp. right-hand) side is the fppf cohomology of $U_{k^{\prime }}=U\times _{k}k^{\prime }$ (resp. $(k^{\prime }\otimes _{k}{\mathcal{O}}_{x}^{h})^{h}$) with support on $x\times _{k}k^{\prime }$ (resp. the ideal generated by $k^{\prime }\otimes _{k}\mathfrak{p}_{x}^{h}$) [Reference MilneMil06, III, Proposition 0.3]. It is enough to show that this morphism is an isomorphism for any $k^{\prime }$. We may assume that $k^{\prime }\in k^{\text{rat}}$ since cohomology and henselization commute with filtered inverse limits of schemes. Then $k^{\prime }$ is a finite product of perfect fields. Hence we may assume that $k^{\prime }$ is a perfect field. Replacing $k$ by $k^{\prime }$, we may assume that $k^{\prime }=k$. Hence we are reduced to showing that

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}_{x}(U,\;\cdot \;)\overset{{\sim}}{\rightarrow }R\unicode[STIX]{x1D6E4}_{x}({\mathcal{O}}_{x}^{h},\;\cdot \;)\end{eqnarray}$$

on $D(U_{\text{fppf}})$. Since $\operatorname{Spec}{\mathcal{O}}_{x}^{h}$ is a filtered inverse limit of affine étale $U$-schemes and cohomology commutes with such limits, we can push them forward to the étale sites. The statement to prove is thus

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}_{x}(U_{\text{et}},\;\cdot \;)\overset{{\sim}}{\rightarrow }R\unicode[STIX]{x1D6E4}_{x}({\mathcal{O}}_{x,\text{et}}^{h},\;\cdot \;)\end{eqnarray}$$

on $D(U_{\text{et}})$. This is the excision isomorphism of étale cohomology [Reference MilneMil80, III, Corollary 1.28].◻

We define a functor of additive categories with translation by

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)=\Bigl[\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow \bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\;\cdot \;)\Bigr][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle \operatorname{Ch}(U_{\text{fppf}})\rightarrow \operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}), & \displaystyle \nonumber\end{eqnarray}$$

where the sum is over all $x\in X\setminus U$. We have its right derived functor

$$\begin{eqnarray}\displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)=\Bigl[R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow \bigoplus _{x\not \in U}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\;\cdot \;)\Bigr][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}). & \displaystyle \nonumber\end{eqnarray}$$

Here we are working with unbounded complexes, making the definition of compact support cohomology more involved than [Reference Demarche and HarariDH18]. This is important in view of the definition of the pairing (2.7.2) below (which uses derived tensor product $\otimes ^{L}$).

For another dense open subscheme $V\subset U$, unfortunately there is no obvious morphism from $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,\;\cdot \;)$ to $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)$, and no natural distinguished triangle

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,A)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,A)\rightarrow \bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,A),\end{eqnarray}$$

unless $A\in D(U_{\text{fppf}})$ satisfies cohomological approximation at all $x\in U\setminus V$ (see Propositions 2.7.3 and 2.7.4 below). We need to define a variant of $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,\;\cdot \;)$ that does admit a natural morphism to $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)$. We define a functor $\operatorname{Ch}(U_{\text{fppf}})\rightarrow \operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ of additive categories with translation by

$$\begin{eqnarray}\displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,U,\;\cdot \;) & = & \displaystyle \Bigl[\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)\rightarrow \bigoplus _{x\in U\setminus V}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\Bigr][-1]\nonumber\\ \displaystyle & = & \displaystyle \Bigl[\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow \bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\;\cdot \;)\oplus \bigoplus _{x\in U\setminus V}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\Bigr][-1].\nonumber\end{eqnarray}$$

Here we used the fact that for any two morphisms $A,B\rightarrow C$ of complexes in an additive category, we have natural isomorphisms of complexes

$$\begin{eqnarray}[A\oplus B\rightarrow C]\cong [A\rightarrow [B\rightarrow C]]\cong [B\rightarrow [A\rightarrow C]],\end{eqnarray}$$

or dually, for any two morphisms $A\rightarrow B,C$ of complexes in an additive category, we have natural isomorphisms of (mapping fiber) complexes

$$\begin{eqnarray}[A\rightarrow B\oplus C][-1]\cong [[A\rightarrow B][-1]\rightarrow C][-1]\cong [[A\rightarrow C][-1]\rightarrow B][-1].\end{eqnarray}$$

We have the right derived functor $D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$

$$\begin{eqnarray}\displaystyle & & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,U,\;\cdot \;)=\Bigl[R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)\rightarrow \bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\Bigr][-1]\nonumber\\ \displaystyle & & \displaystyle \quad =\Bigl[R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,\;\cdot \;)\rightarrow \bigoplus _{x\not \in U}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\;\cdot \;)\oplus \bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\Bigr][-1].\nonumber\end{eqnarray}$$

By definition, we have a distinguished triangle

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,U,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)\rightarrow \bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;).\end{eqnarray}$$

By this, we mean the values of these functors at any object form a distinguished triangle.

We also need to define a variant of $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)$. Define

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,\;\cdot \;)=\Bigl[\boldsymbol{\unicode[STIX]{x1D6E4}}(V,\;\cdot \;)\rightarrow \bigoplus _{x\notin U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\;\cdot \;)\oplus \bigoplus _{x\in U\setminus V}\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)[1]\Bigr][-1],\end{eqnarray}$$

where the morphism from $\boldsymbol{\unicode[STIX]{x1D6E4}}(V,\;\cdot \;)$ to $\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)[1]$ is the composite

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(V,\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x}^{h}/k,\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)[1].\end{eqnarray}$$

Its derived functor is

$$\begin{eqnarray}\displaystyle & & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,\;\cdot \;)\nonumber\\ \displaystyle & & \displaystyle \quad =\Bigl[R\boldsymbol{\unicode[STIX]{x1D6E4}}(V,\;\cdot \;)\rightarrow \bigoplus _{x\notin U}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,\;\cdot \;)\oplus \bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)[1]\Bigr][-1].\nonumber\end{eqnarray}$$

For any $x\in U\setminus V$, the inclusion into the second summand defines morphisms

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,\;\cdot \;)\end{eqnarray}$$

and

(2.7.1)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,\;\cdot \;).\end{eqnarray}$$

The last morphism will be used later to connect trace morphisms in the local and global situations.

Let $Z=U\setminus V$. Define

$$\begin{eqnarray}o_{Z}(U,\;\cdot \;)=\Bigl[\boldsymbol{\unicode[STIX]{x1D6E4}}_{Z}(U,\;\cdot \;)\rightarrow \bigoplus _{x\in U\setminus V}\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\Bigr],\end{eqnarray}$$

with derived functor

$$\begin{eqnarray}Ro_{Z}(U,\;\cdot \;)=\Bigl[R\boldsymbol{\unicode[STIX]{x1D6E4}}_{Z}(U,\;\cdot \;)\rightarrow \bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\Bigr].\end{eqnarray}$$

We explain the notation to be used in the next proposition. A commutative diagram of distinguished triangles in a triangulated category

means a commutative diagram all of whose rows and columns are distinguished triangles (where the right lower square is actually “anticommutative” [Reference Kashiwara and SchapiraKS06, Diagram (10.5.5)], but we largely ignore commutative vs. anticommutative issues, which is especially harmless if $A$ or $C^{\prime \prime }$ is zero for example). As usual, we will hide the shifted terms $B[1],A^{\prime \prime }[1],A[2]$ etc. for brevity and mention the remaining $3\times 3$ diagram as a commutative diagram of distinguished triangles. A commutative diagram of distinguished triangles of triangulated functors

means a commutative diagram of morphisms of triangulated functors whose values at any object form a commutative diagram of distinguished triangles in the above sense.

Proposition 2.7.3. Let $V\subset U$ be a dense open subscheme and set $Z=U\setminus V$. The natural morphisms form a commutative diagram

of distinguished triangles of triangulated functors $K(U_{\text{fppf}})\rightarrow K(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. With the isomorphism

$$\begin{eqnarray}Ro_{Z}(U,\;\cdot \;)\overset{{\sim}}{\rightarrow }\bigoplus _{x\in U\setminus V}Ro_{x}({\mathcal{O}}_{x}^{h}/k,\;\cdot \;)\end{eqnarray}$$

coming from the previous proposition, we have a canonical commutative diagram

of distinguished triangles of triangulated functors $D(U_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$.

Proof. For $A\in \operatorname{Ch}(U_{\text{fppf}})$, let

$$\begin{eqnarray}\displaystyle & \displaystyle C=\boldsymbol{\unicode[STIX]{x1D6E4}}(U,A),\qquad D=\boldsymbol{\unicode[STIX]{x1D6E4}}(V,A),\qquad E=\bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A), & \displaystyle \nonumber\\ \displaystyle & \displaystyle C^{\prime }=\bigoplus _{x\in U\setminus V}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,A),\qquad D^{\prime }=\bigoplus _{x\in U\setminus V}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A). & \displaystyle \nonumber\end{eqnarray}$$

These are objects of $\operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. We have a natural commutative diagram

in $\operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. The value at $A$ of the first diagram shifted by one can be written as

(which is actually a $4\times 4$ diagram as we are omitting the shifted terms). It is routine to check that this diagram is commutative up to homotopy. The second diagram results from the first.◻

We say that an object $A\in D(U_{\text{fppf}})$satisfies cohomological approximation if $Ro_{x}({\mathcal{O}}_{x}^{h},A)=0$ for any $x\in U$. We denote by $D(U_{\text{fppf}})_{\text{ca}}$ the full subcategory of $D(U_{\text{fppf}})$ consisting of objects satisfying cohomological approximation. It is a triangulated subcategory.

Proposition 2.7.4. On $D(U_{\text{fppf}})_{\text{ca}}$, we have isomorphisms

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,U,\;\cdot \;)=R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,\;\cdot \;),\qquad R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,\;\cdot \;)=R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;),\end{eqnarray}$$

a distinguished triangle

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)\rightarrow \bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\end{eqnarray}$$

and a morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\;\cdot \;)\end{eqnarray}$$

for any $x\in U$ compatible with (2.7.1).

Proof. Obvious from the previous proposition. ◻

Proposition 2.7.5. If $A$ is a smooth group scheme or a finite flat group scheme over $U$, then $A\in D(U_{\text{fppf}})_{\text{ca}}$.

Proof. This follows from Proposition 2.6.2. ◻

Let $B,C\in D(U_{\text{fppf}})$. To simplify the notation, we denote $R\,\mathbf{Hom}_{{\mathcal{O}}_{x}^{h}}$ for any $x\in U$ by $[\;\cdot \;,\;\cdot \;]_{{\mathcal{O}}_{x}^{h}}$. Denote similarly $R\,\mathbf{Hom}_{U}$ by $[\;\cdot \;,\;\cdot \;]_{U}$ and $R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}$ by $[\;\cdot \;,\;\cdot \;]_{k}$. A similar construction to (2.5.1) defines a morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,[B,C]_{U})\rightarrow [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,C)]_{k}.\end{eqnarray}$$

This is equivalent to a morphism

(2.7.2)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,A)\otimes ^{L}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,A\otimes ^{L}B)\end{eqnarray}$$

via the derived tensor-Hom adjunction [Reference Kashiwara and SchapiraKS06, Theorem 18.6.4(vii)] and the change of variables $[B,C]_{U}{\rightsquigarrow}A$ and $A\otimes ^{L}B{\rightsquigarrow}C$.

For each $x\in U\setminus V$, we have natural morphisms

$$\begin{eqnarray}\displaystyle & & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,[B,C]_{{\mathcal{O}}_{x}^{h}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,[B,C]_{{\mathcal{O}}_{x}^{h}})\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,[R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,C)]_{k}\rightarrow [R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,C)]_{k}\nonumber\end{eqnarray}$$

using the morphisms (2.5.3) and (2.7.1). Using the morphisms in Proposition 2.7.3, we have morphisms

$$\begin{eqnarray}\displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,[B,C]_{U}) & \rightarrow & \displaystyle [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,C)]_{k}\nonumber\\ \displaystyle & \rightarrow & \displaystyle [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,C)]_{k},\nonumber\\ \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(V,[B,C]_{V}) & \rightarrow & \displaystyle [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,C)]_{k}\nonumber\\ \displaystyle & \rightarrow & \displaystyle [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,U,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,C)]_{k}.\nonumber\end{eqnarray}$$

Proposition 2.7.6. The above morphisms give a morphism from the distinguished triangles

$$\begin{eqnarray}\bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,[B,C]_{{\mathcal{O}}_{x}^{h}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,[B,C]_{U})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(V,[B,C]_{V})\end{eqnarray}$$

to the distinguished triangle

$$\begin{eqnarray}\displaystyle \bigoplus _{x\in U\setminus V}[R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,C)]_{k} & \rightarrow & \displaystyle [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,C)]_{k}\nonumber\\ \displaystyle & \rightarrow & \displaystyle [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,U,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,C)]_{k}.\nonumber\end{eqnarray}$$

Proof. We only prove the commutativity of the square

for $x\in U\setminus V$. There are two more squares whose commutativity has to be proven. They can be treated similarly, so we omit their treatment. It is enough to show that the diagram

in $K(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ for $x\in U\setminus V$ and $B,C\in K(U_{\text{fppf}})$ is commutative, where $[\;\cdot \;,\;\cdot \;]_{k}^{c}$ is the sheaf-Hom complex functor $\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}$. By the tensor-Hom adjunction, it is enough to show that the morphism

(2.7.3)$$\begin{eqnarray}\displaystyle & & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,B)\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,A\otimes B)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,A\otimes B)\end{eqnarray}$$

and the morphism

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(U,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,A\otimes B)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,A\otimes B)\end{eqnarray}$$

in $\operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ are compatible up to homotopy via

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,A)\overset{{\sim}}{\leftarrow }\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(U,A)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(U,A)\end{eqnarray}$$

(where $\overset{{\sim}}{\leftarrow }$ is a quasi-isomorphism). It is routine to check that the morphism (2.7.3), the morphism

$$\begin{eqnarray}\displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(U,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B) & \rightarrow & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(U,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}(U,B)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(U,A\otimes B)\nonumber\\ \displaystyle & \rightarrow & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,A\otimes B)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,A\otimes B)\nonumber\end{eqnarray}$$

and the morphism

$$\begin{eqnarray}\displaystyle & & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}(U,B)\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,A\otimes B)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,V,A\otimes B)\nonumber\end{eqnarray}$$

are all compatible (without homotopy) via

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x}^{h}/k,A)\overset{{\sim}}{\leftarrow }\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(U,A)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,A).\end{eqnarray}$$

Hence we need to show that the diagram

in $\operatorname{Ch}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is commutative up to homotopy. The left upper term is the total complex of the three term complex in degrees $0,1,2$ of double complexes

$$\begin{eqnarray}\displaystyle & & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(U,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}(U,B)\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,\boldsymbol{\unicode[STIX]{x1D6E4}}(U,A)\otimes \bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,B)\oplus \bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A)\otimes \boldsymbol{\unicode[STIX]{x1D6E4}}(U,B)\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,\bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A)\otimes \bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,B).\nonumber\end{eqnarray}$$

The right lower term is the total complex of the two term complex in degrees $0,1$ of double complexes

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}(U,A\otimes B)\rightarrow \bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A\otimes B).\end{eqnarray}$$

The required homotopy is given by the projection to the diagonal

$$\begin{eqnarray}\bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A)\otimes \bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,B)\rightarrow \bigoplus _{x\not \in U}\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x}/k,A\otimes B)\end{eqnarray}$$

in degree $2$ and zero in other degrees.◻

Proposition 2.7.7. Let $A,B,C\in D(U_{\text{fppf}})_{\text{ca}}$. Let $A\rightarrow R\,\mathbf{Hom}_{U}(B,C)$, or equivalently, $A\otimes ^{L}B\rightarrow C$, be a morphism in $D(U_{\text{fppf}})$. Then the morphism in Proposition 2.7.6 induces a morphism of distinguished triangles from

$$\begin{eqnarray}\bigoplus _{x\in U\setminus V}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,A)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,A)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(V,A)\end{eqnarray}$$

to

$$\begin{eqnarray}\displaystyle \bigoplus _{x\in U\setminus V}[R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,C)]_{k} & \rightarrow & \displaystyle [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,C)]_{k}\nonumber\\ \displaystyle & \rightarrow & \displaystyle [R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(V,B),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,C)]_{k}.\nonumber\end{eqnarray}$$

Proof. This follows from Proposition 2.7.4. ◻

We say that a sheaf $F\in \operatorname{Set}(U_{\text{fppf}})$ is locally of finite presentation if $F(\mathop{\varprojlim }\nolimits_{\unicode[STIX]{x1D706}}U_{\unicode[STIX]{x1D706}})=\mathop{\varinjlim }\nolimits_{\unicode[STIX]{x1D706}}F(U_{\unicode[STIX]{x1D706}})$ for any filtered inverse system $\{U_{\unicode[STIX]{x1D706}}\}$ of quasi-compact quasi-separated $U$-schemes with affine transition morphisms.

Proposition 2.7.8. For any sheaf $A\in \operatorname{Ab}(U_{\text{fppf}})$ locally of finite presentation and $n\geqslant 0$, the sheaf $R^{n}\unicode[STIX]{x1D70B}_{U\ast }A\in \operatorname{Ab}(k_{\text{et}}^{\text{ind}\text{rat}})$ is locally of finite presentation. In particular, we have $\mathbf{H}^{n}(U,A)=R^{n}\unicode[STIX]{x1D70B}_{U\ast }A$. That is, $\mathbf{H}^{n}(U,A)$ is the étale sheafification of the presheaf

$$\begin{eqnarray}k^{\prime }\mapsto H^{n}(U_{k^{\prime }},A).\end{eqnarray}$$

Proof. The functor $k^{\prime }\mapsto U_{k^{\prime }}$ from the opposite category of $k^{\text{ind}\text{rat}}$ to the category of $U$-schemes commutes with filtered inverse limits. The same proof as Proposition 2.6.1 works.◻

In particular, in the situation of this proposition, we have isomorphisms

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,A))=R\unicode[STIX]{x1D6E4}(k_{\text{et}},R\unicode[STIX]{x1D70B}_{U,\ast }A)=R\unicode[STIX]{x1D6E4}(U,A)\end{eqnarray}$$

and a spectral sequence

$$\begin{eqnarray}E_{2}^{ij}=H^{i}(k_{\text{pro}\text{et}},\mathbf{H}^{j}(U,A))\Longrightarrow H^{i+j}(U,A).\end{eqnarray}$$

Remark 2.7.9. Proposition 2.6.2 may be false if $A$ is replaced by a general fppf sheaf over ${\mathcal{O}}_{x}^{h}$. In fact, whenever ${\mathcal{O}}_{x}^{h}$ is not complete, there exists an fppf sheaf $A$ over ${\mathcal{O}}_{x}^{h}$ such that the map $H_{x}^{0}({\mathcal{O}}_{x}^{h},A)\rightarrow H_{x}^{0}(\hat{{\mathcal{O}}}_{x},A)$ is not surjective, where $H_{x}^{0}({\mathcal{O}}_{x}^{h},A)$ is the kernel of $\unicode[STIX]{x1D6E4}({\mathcal{O}}_{x}^{h},A)\rightarrow \unicode[STIX]{x1D6E4}(K_{x}^{h},A)$ and $H_{x}^{0}(\hat{{\mathcal{O}}}_{x},A)$ is defined similarly. This shows that the distinguished triangle in Proposition 2.7.4 and a similar long exact sequence stated in [Reference MilneMil06, III, Remark 0.6(b)] do not exist for general sheaves. Coefficients in smooth group schemes and finite flat group schemes as stated in Proposition 2.7.5 are sufficient in this paper and in [Reference MilneMil06, III].

To give an example of such an fppf sheaf, consider the two ring homomorphisms ${\mathcal{O}}_{x}^{h}{\hookrightarrow}{\mathcal{O}}_{x}^{h}[t]{\twoheadrightarrow}k_{x}$, where the first one is the inclusion and the second is the map $t\mapsto 0$ followed by the reduction map. Denote the morphisms $\operatorname{Spec}k_{x,\text{fppf}}\rightarrow \operatorname{Spec}{\mathcal{O}}_{x}^{h}[t]_{\text{fppf}}\rightarrow \operatorname{Spec}{\mathcal{O}}_{x,\text{fppf}}^{h}$ induced on the fppf sites by $i$ and $f$, respectively. Then a desired counterexample is given by $A=f_{!}i_{\ast }\mathbb{Z}$, where $f_{!}$ is the left adjoint of $f^{\ast }$ [Reference MilneMil80, II, Remark 3.18].

Indeed, let $f_{!}^{\text{pre}}$ be the left adjoint of the presheaf pullback functor by $f$ and set $A^{\text{pre}}=f_{!}^{\text{pre}}i_{\ast }\mathbb{Z}$. Denote $\hat{{\mathcal{O}}}_{x}\otimes _{{\mathcal{O}}_{x}^{h}}\hat{{\mathcal{O}}}_{x}$ by $\hat{{\mathcal{O}}}_{x}^{\otimes 2}$. Then $\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},A^{\text{pre}})$ and $\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x}^{\otimes 2},A^{\text{pre}})$ are the free abelian groups generated by the sets $\hat{\mathfrak{p}}_{x}=\mathfrak{p}_{x}^{h}\hat{{\mathcal{O}}}_{x}$ and $\mathfrak{p}_{x}^{h}\hat{{\mathcal{O}}}_{x}^{\otimes 2}$, respectively. Consider the element of $\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},A^{\text{pre}})$ corresponding to any element $c$ of $\hat{\mathfrak{p}}_{x}$ not in ${\mathcal{O}}_{x}^{h}$. Since $c\otimes 1$ and $1\otimes c$ are distinct in $\hat{{\mathcal{O}}}_{x}^{\otimes 2}$, these elements do not glue in $\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x}^{\otimes 2},A^{\text{pre}})$. They do not even after replacing $\hat{{\mathcal{O}}}_{x}^{\otimes 2}$ by its fppf cover since $f_{!}^{\text{pre}}$ sends separated presheaves to separated presheaves by construction. Hence the corresponding element of $\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},A)=H_{x}^{0}(\hat{{\mathcal{O}}}_{x},A)$ does not come from an element of $\unicode[STIX]{x1D6E4}({\mathcal{O}}_{x}^{h},A)=H_{x}^{0}({\mathcal{O}}_{x}^{h},A)$.

3 Global duality and its proof

In the rest of this paper, let $X$ be a proper smooth geometrically connected curve over a perfect field $k$ ($\in {\mathcal{U}}_{0}$) of characteristic $p>0$ with function field $K$. We continue the notation in Section 2.7. We denote $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ by $\operatorname{Ab}(k)$ and $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ by $D(k)$. We denote $\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}$ by $\mathbf{Hom}_{k}$ and use the notation $\mathbf{Ext}_{k}^{n}$ and $R\,\mathbf{Hom}_{k}$ similarly. Exact sequences and distinguished triangles of objects over $k$ are always considered in $\operatorname{Ab}(k)$ or $D(k)$ unless otherwise noted.

Let $A$ be an abelian variety over $K$ with Néron model ${\mathcal{A}}$ over $X$. (Actually Sections 3.1 and 3.3 do not use $A$.) The maximal open subgroup scheme of ${\mathcal{A}}$ with connected fibers is denoted by ${\mathcal{A}}_{0}$. The fiber of ${\mathcal{A}}$ over any closed point $x=\operatorname{Spec}k_{x}$ of $X$ is denoted by ${\mathcal{A}}_{x}$. The dual of $A$ is denoted by $A^{\vee }$, with Néron model ${\mathcal{A}}^{\vee }$ and the corresponding subscheme ${\mathcal{A}}_{0}^{\vee }$ and the fibers ${\mathcal{A}}_{x}^{\vee }$.

3.1 Duality for finite flat coefficients over open curves

By Proposition 2.7.8, the sheaf $\mathbf{H}^{n}(X,\mathbf{G}_{m})\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ for any $n$ is locally of finite presentation and the étale sheafification of the presheaf

$$\begin{eqnarray}k^{\prime }\mapsto H^{n}(X_{k^{\prime }},\mathbf{G}_{m}).\end{eqnarray}$$

Let $\operatorname{Pic}_{X}=\operatorname{Pic}_{X/k}$ be the perfection of the Picard scheme of $X$ over $k$. This represents the sheaf $R^{1}\unicode[STIX]{x1D70B}_{X\ast }^{\text{perf}}\mathbf{G}_{m}$ on $\operatorname{Spec}k_{\text{et}}^{\text{perf}}$ and hence the sheaf $\mathbf{H}^{1}(X,\mathbf{G}_{m})$ on $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$.

Proposition 3.1.1. We have $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\mathbf{G}_{m})=\mathbf{G}_{m}$, $\mathbf{H}^{1}(X,\mathbf{G}_{m})=\operatorname{Pic}_{X}$ and $\mathbf{H}^{n}(X,\mathbf{G}_{m})=0$ for $n\geqslant 2$.

Proof. The statements for $n=0,1$ are obvious. For $n\geqslant 2$, it is a classical fact [Reference MilneMil80, III, Example 2.22, Case (d)] that $H^{n}(X_{k^{\prime }},\mathbf{G}_{m})=0$ for any algebraically closed fields $k^{\prime }$ over $k$. This implies $\mathbf{H}^{n}(X,\mathbf{G}_{m})=0$ since this sheaf is locally of finite presentation.◻

Let $U\subset X$ be a dense open subscheme. The smooth group scheme $\mathbf{G}_{m}\in \operatorname{Ab}(U_{\text{fppf}})$ satisfies cohomological approximation by Proposition 2.7.5. Hence by Proposition 2.7.4 and the above proposition, we have morphisms

(3.1.1)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\mathbf{G}_{m})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\mathbf{G}_{m})\rightarrow \operatorname{Pic}_{X}[-1]\overset{\deg }{\rightarrow }\mathbb{Z}[-1].\end{eqnarray}$$

We call the composite the (global) trace morphism.

Proposition 3.1.2. The global trace morphism $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\mathbf{G}_{m})\rightarrow \mathbb{Z}[-1]$ and the local trace morphism $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\mathbf{G}_{m})\rightarrow \mathbb{Z}[-1]$ at $x\in U$ in (2.5.8) are compatible under the morphism $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,\mathbf{G}_{m})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\mathbf{G}_{m})$ in Proposition 2.7.4.

Proof. We are comparing the degree morphism $\operatorname{Pic}_{X}\rightarrow \mathbb{Z}$ and the valuation morphism $\operatorname{Res}_{k_{x}/k}\hat{\mathbf{K}}_{x}^{\times }/\hat{\mathbf{O}}_{x}^{\times }\rightarrow \mathbb{Z}$. It is enough to compare them on $\overline{k}$-points since $\hat{\mathbf{K}}_{x}^{\times }/\hat{\mathbf{O}}_{x}^{\times }\cong \mathbb{Z}$ is étale. Then the comparison is between the abstract group homomorphisms $\operatorname{Pic}(X_{\overline{k}})\rightarrow \mathbb{Z}$ and $\bigoplus _{\overline{x}\rightarrow x}\hat{K}_{\overline{x}}^{\times }/\hat{{\mathcal{O}}}_{\overline{x}}^{\times }\rightarrow \mathbb{Z}$, where the sum is over all $\overline{k}$-points of $X_{\overline{k}}$ lying over $x$. This is obvious.◻

Tensoring $\mathbb{Z}/p\mathbb{Z}[-1]$, the global trace morphism induces a morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\unicode[STIX]{x1D707}_{p})\rightarrow \mathbb{Z}/p\mathbb{Z}[-2].\end{eqnarray}$$

Propositions 2.7.1 and 2.7.8 allow us to translate the results of [Reference Artin and Milne.AM76] to our setting. The kernel of the morphism

$$\begin{eqnarray}\mathbf{H}^{1}(X,\unicode[STIX]{x1D6FA}_{X}^{1})\overset{C-1}{\rightarrow }\mathbf{H}^{1}(X,\unicode[STIX]{x1D6FA}_{X}^{1}),\qquad \text{or}\qquad \mathbf{G}_{a}\overset{F-1}{\rightarrow }\mathbf{G}_{a}\end{eqnarray}$$

identifies $\mathbf{H}^{2}(X,\unicode[STIX]{x1D707}_{p})$ as $\mathbb{Z}/p\mathbb{Z}$ as explained in [Reference Artin and Milne.AM76, Introduction], where $C$ is the Cartier operator and $F$ is the Frobenius. This gives a morphism $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\unicode[STIX]{x1D707}_{p})\rightarrow \mathbb{Z}/p\mathbb{Z}$. This is equal to the above trace morphism since we have a commutative diagram

where $\operatorname{Res}$ denotes the residue map.

We need the following result of Milne [Reference MilneMil06, III, Theorem 11.1], which is the generalization for open curves of the corresponding result of Artin and Milne [Reference Artin and Milne.AM76]. Since [Reference SuzukiSuz14, Section 5.2] replaces Bester’s local finite flat duality, we give a proof based on [Reference SuzukiSuz14, Theorem (5.2.1.2)] for clarity.

Theorem 3.1.3. Let $U\subset X$ be a dense open subscheme and $N$ a finite flat group scheme over $U$. Then $R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,N)\in D^{b}(\text{I}\text{Alg}_{\text{uc}})$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,N)\in D^{b}(\text{P}\text{Alg}_{\text{uc}})$, which are both concentrated in degrees $0,1,2$. Consider the morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,N^{\text{CD}})\otimes ^{L}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,N)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\mathbf{G}_{m})\rightarrow \mathbb{Z}[-1]\end{eqnarray}$$

given by the cup product morphism (2.7.2) and the trace morphism (3.1.1). The induced morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,N^{\text{CD}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,N)^{\text{SD}}[-1]\end{eqnarray}$$

in $D(k)$ is an isomorphism.

Proof. The proof proceeds by dévissage.

Step 1: If $U=X$, then the theorem in this case is [Reference Artin and Milne.AM76, (0.3), (4.9)].

Step 2: Let $V\subset U$ be a dense open subscheme. The theorem is true for $N$ over $U$ if and only if so is for $N$ over $V$. Indeed, the groups $N$, $N^{\text{CD}}$ and $\mathbf{G}_{m}$ satisfy cohomological approximation by Proposition 2.7.5. The natural morphism $N^{\text{CD}}\rightarrow R\,\mathbf{Hom}_{U}(N,\mathbf{G}_{m})$, the trace morphism $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\mathbf{G}_{m})\rightarrow \mathbb{Z}[-1]$ and the morphism in Proposition 2.7.7 give a morphism of distinguished triangles

We know that $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,N^{\text{CD}})$ is in $D^{b}(\text{I}\text{Alg}_{\text{uc}}/k)$ concentrated in degree $2$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,N)$ is in $D^{b}(\text{P}\text{Alg}_{\text{uc}}/k)$ concentrated in degrees $0,1$ by [Reference SuzukiSuz14, Proposition (3.4.2)(b), Proposition (3.4.6)]. The left vertical morphism is an isomorphism by Bester’s duality [Reference SuzukiSuz14, Theorem (5.2.1.2)]. Hence the theorem for $N$ over $U$ and the theorem for $N$ over $V$ are equivalent.

Step 3: If $0\rightarrow N^{\prime }\rightarrow N\rightarrow N^{\prime \prime }\rightarrow 0$ is an exact sequence of finite flat group schemes over $U$ and the theorem is true for $N^{\prime }$ and $N^{\prime \prime }$, then so is for $N$.

Step 4: The general case. The group $N_{K}=N\times _{X}K$ over $K$ has a filtration by finite flat subgroup schemes whose each successive subquotient or its dual is of height one. A finite flat group scheme over $K$ of height one or with dual of height one extends to $X$ as a finite flat group scheme by [Reference MilneMil06, III, Propositions B.4, B.5]. By spreading out, we know that $N_{V}=N\times _{X}V$ over some dense open $V\subset U$ has a filtration by finite flat subgroup schemes whose successive subquotients are finite flat and extendable to $X$ as finite flat group schemes. Hence the previous three steps imply the theorem.◻

For how the above theorem is related to the corresponding duality statement [Reference MilneMil06, III, Theorem 8.2] in the finite base field case, see Remark 4.2.11 below.

3.2 Mod $n$ duality for Néron models and preliminary calculations

Let $A$ be as in the beginning of this section. The Poincaré biextension $A^{\vee }\otimes ^{L}A\rightarrow \mathbf{G}_{m}[1]$ as a morphism in $D(K_{\text{fppf}})$ canonically extends to a biextension ${\mathcal{A}}_{0}^{\vee }\otimes ^{L}{\mathcal{A}}\rightarrow \mathbf{G}_{m}[1]$ as a morphism in $D(X_{\text{fppf}})$ by [Gro72, IX, 1.4.3]. With this morphism, the cup product morphism (2.7.2) and the trace morphism (3.1.1), we have morphisms

(3.2.1)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\mathbf{G}_{m}[1])\rightarrow \mathbb{Z}.\end{eqnarray}$$

Proposition 3.2.1. Let $n\geqslant 1$. Then $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Z}/n\mathbb{Z}$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0})\otimes ^{L}\mathbb{Z}/n\mathbb{Z}$ are both in $D^{b}(\text{Alg}_{\text{uc}})$ concentrated in degrees $-1,0,1$. Consider the morphism (3.2.1) derived-tensored with $\mathbb{Z}/n\mathbb{Z}$:

$$\begin{eqnarray}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\otimes ^{L}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\rightarrow \mathbb{Z}/n\mathbb{Z}\rightarrow \mathbb{Z}[1],\end{eqnarray}$$

where the last morphism is the connecting morphism for the short exact sequence $0\rightarrow \mathbb{Z}\rightarrow \mathbb{Z}\rightarrow \mathbb{Z}/n\mathbb{Z}\rightarrow 0$. The induced morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\rightarrow (R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Z}/n\mathbb{Z})^{\text{SD}}[1]\end{eqnarray}$$

in $D(k)$ is an isomorphism.

Proof. We denote $(\;\cdot \;)\otimes ^{L}\mathbb{Z}/n\mathbb{Z}$ by $(\;\cdot \;)_{n}$ to simplify the notation. Take a dense open subscheme $U\subset X$ over which ${\mathcal{A}}$ is an abelian scheme. We have a morphism of distinguished triangles

by the same method as the proof of the previous theorem. The right vertical morphism is an isomorphism by the previous proposition since $({\mathcal{A}}^{\vee }\otimes ^{L}\mathbb{Z}/n\mathbb{Z})[-1]$ over $U$ is the $n$-torsion part of ${\mathcal{A}}$, which is finite flat over $U$. The left vertical morphism is an isomorphism by Proposition 2.5.4. Hence so is the middle vertical morphism. The object $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,{\mathcal{A}}_{0}^{\vee })_{n}$ is in $D^{b}(\text{I}\text{Alg}_{\text{uc}}/k)$ concentrated in degrees $0,1$ and the object $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,{\mathcal{A}})_{n}$ is in $D^{b}(\text{P}\text{Alg}_{\text{uc}}/k)$ concentrated in degrees $-1,0$ by Proposition 2.5.3. The objects $R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })_{n}$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})_{n}$ are in $D^{b}(\text{I}\text{Alg}_{\text{uc}}/k)$ and in $D^{b}(\text{P}\text{Alg}_{\text{uc}}/k)$, respectively, and concentrated in degrees $-1,0,1$ by the previous proposition. Hence the same is true for $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })_{n}$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})_{n}$. Therefore they are in both $D^{b}(\text{P}\text{Alg}_{\text{uc}}/k)$ and $D^{b}(\text{I}\text{Alg}_{\text{uc}}/k)$, hence in $D^{b}(\text{Alg}_{\text{uc}}/k)$.◻

In the next proposition, we consider not necessarily perfect group schemes over $k$. But we will soon apply perfection.

Proposition 3.2.2. Let $\operatorname{Res}_{X/k}{\mathcal{A}}$ be the Weil restriction of ${\mathcal{A}}$ as a functor on the category of (not necessarily perfect) $k$-schemes. Let $\operatorname{Tr}_{K/k}A$ be the $K/k$-trace of $A$ ([Reference LangLan83, VIII, Theorem 8], [Reference ConradCon06, Definition 6.1]), which is an abelian variety over $k$ with canonical $K$-morphism $(\operatorname{Tr}_{K/k}A)_{K}\rightarrow A$. Let $\operatorname{Tr}_{K/k}A\rightarrow \operatorname{Res}_{X/k}{\mathcal{A}}$ be the $k$-morphism induced by the $X$-morphism $(\operatorname{Tr}_{K/k}A)_{X}\rightarrow {\mathcal{A}}$, which itself is induced by $(\operatorname{Tr}_{K/k}A)_{K}\rightarrow A$.

Then $\operatorname{Res}_{X/k}{\mathcal{A}}$ is represented by a group scheme locally of finite type over $k$. The morphism $\operatorname{Tr}_{K/k}A\rightarrow (\operatorname{Res}_{X/k}{\mathcal{A}})_{0}$ to the identity component has finite infinitesimal kernel and cokernel. The étale $k$-group $\unicode[STIX]{x1D70B}_{0}\operatorname{Res}_{X/k}{\mathcal{A}}$ of connected components has finitely generated group of geometric points.

Proof. Since $X$ is proper over $k$ and ${\mathcal{A}}$ is quasi-projective over $X$ by [Reference Bosch, Lütkebohmert and RaynaudBLR90, 6.4/1], the result on existence of Hilbert schemes [Reference GrothendieckGro95, Section 4.c] shows that $\operatorname{Res}_{X/k}{\mathcal{A}}$ is a group scheme locally of finite type over $k$. The kernel of $(\operatorname{Tr}_{K/k}A)_{K}\rightarrow A$ is a finite infinitesimal $K$-group by [Reference ConradCon06, Theorem 6.12]. It follows that $\operatorname{Tr}_{K/k}A\rightarrow \operatorname{Res}_{K/k}{\mathcal{A}}$ is injective on geometric points. Hence its kernel is a finite infinitesimal $k$-group. The group of $\overline{k}$-points of the cokernel of $\operatorname{Tr}_{K/k}A\rightarrow \operatorname{Res}_{K/k}{\mathcal{A}}$ is ${\mathcal{A}}(X_{\overline{k}})/(\operatorname{Tr}_{K/k}A)(\overline{k})$, which is finitely generated by the Lang–Néron theorem [Reference ConradCon06, Theorem 7.1]. This implies that the connected group scheme $(\operatorname{Res}_{X/k}{\mathcal{A}})_{0}/\operatorname{Tr}_{K/k}A$ of finite type over $k$ has finitely generated group of geometric points. It follows that $(\operatorname{Res}_{X/k}{\mathcal{A}})_{0}/\operatorname{Tr}_{K/k}A$ has trivial reduced part and hence is finite infinitesimal. Hence $\unicode[STIX]{x1D70B}_{0}\operatorname{Res}_{X/k}{\mathcal{A}}$ has finitely generated group of geometric points.◻

Proposition 3.2.3. The sheaf $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ is an extension of a finitely generated étale group by an abelian variety over $k$. For $n\geqslant 1$, we have $\mathbf{H}^{n}(X,{\mathcal{A}})\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$. The group $\mathbf{H}^{2}(X,{\mathcal{A}})$ is divisible. For $n\geqslant 3$, we have $\mathbf{H}^{n}(X,{\mathcal{A}})=0$.

Proof. The statement about $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ follows from the previous proposition since it has the same values as $\operatorname{Res}_{X/k}{\mathcal{A}}$ on ind-rational $k$-algebras and hence is represented by the perfection of $\operatorname{Res}_{X/k}{\mathcal{A}}$.

We show that the sheaf $\mathbf{H}^{n}(X,{\mathcal{A}})$ for any $n\geqslant 1$ is torsion. By Proposition 2.7.8, it is an étale sheafification of the presheaf $k^{\prime }\mapsto H^{n}(X_{k^{\prime }},{\mathcal{A}})$ and locally of finite presentation. Hence it is enough to show that the abstract group $H^{n}(X_{k^{\prime }},{\mathcal{A}})$ for $n\geqslant 1$ is torsion for any algebraically closed field $k^{\prime }$ over $k$. Assume that $k=\overline{k}$. For any dense open $U\subset X$, by applying $R\unicode[STIX]{x1D6E4}(k,\;\cdot \;)$ to the isomorphism in Proposition 2.7.2, we have a distinguished triangle

$$\begin{eqnarray}\bigoplus _{x\not \in U}R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})\rightarrow R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow R\unicode[STIX]{x1D6E4}(U,{\mathcal{A}}).\end{eqnarray}$$

Taking the limit in smaller and smaller $U$, we have a distinguished triangle

$$\begin{eqnarray}\bigoplus _{x\in X}R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})\rightarrow R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow R\unicode[STIX]{x1D6E4}(K,A).\end{eqnarray}$$

We have $R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})=H^{1}(\hat{K}_{x},A)[-2]$ by Proposition 2.5.3. Since Galois cohomology groups are torsion in positive degrees, it follows that $H^{n}(X,{\mathcal{A}})$ is torsion for any $n\geqslant 1$. The same is true if $k$ is replaced by any algebraically closed field over $k$. Hence $\mathbf{H}^{n}(X,{\mathcal{A}})$ for $n\geqslant 1$ is torsion.

Let $m\geqslant 1$. Denote $C_{m}=R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Z}/m\mathbb{Z}$. For any $n\in \mathbb{Z}$, we have an exact sequence

$$\begin{eqnarray}0\rightarrow \mathbf{H}^{n-1}(X,{\mathcal{A}})\otimes \mathbb{Z}/m\mathbb{Z}\rightarrow H^{n-1}(C_{m})\rightarrow \mathbf{H}^{n}(X,{\mathcal{A}})[m]\rightarrow 0,\end{eqnarray}$$

where $[m]$ denotes the $m$-torsion part. We have $H^{n-1}(C_{m})\in \text{Alg}_{\text{uc}}/k$ by Proposition 3.2.1. We have $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\otimes \mathbb{Z}/m\mathbb{Z}\in \text{FEt}/k$ by what we have shown above about the structure of $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$. Therefore $\mathbf{H}^{1}(X,{\mathcal{A}})[m]\in \text{Alg}_{\text{uc}}/k$. Since $m$ is arbitrary, the torsionness shown above then shows that $\mathbf{H}^{1}(X,{\mathcal{A}})\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$. Hence $\mathbf{H}^{1}(X,{\mathcal{A}})\otimes \mathbb{Z}/m\mathbb{Z}\in \text{Alg}_{\text{uc}}/k$ by Proposition 2.1.1. Repeating the same argument, we have $\mathbf{H}^{2}(X,{\mathcal{A}})\in \text{I}^{f}\text{Alg}_{\text{uc}}/k$. Since $H^{n-1}(C_{m})=0$ for $n\geqslant 3$ by Proposition 3.2.1, we know that $\mathbf{H}^{2}(X,{\mathcal{A}})$ is divisible and $\mathbf{H}^{n}(X,{\mathcal{A}})=0$ for $n\geqslant 3$.◻

For each closed point $x\in X$, we regard the component group $\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})$ to be an étale group over $k_{x}$. Let $i_{x}:x{\hookrightarrow}X$ be the inclusion, where we identified $x=\operatorname{Spec}k_{x}$. We have an exact sequence

$$\begin{eqnarray}0\rightarrow {\mathcal{A}}_{0}\rightarrow {\mathcal{A}}\rightarrow \bigoplus _{x}i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\rightarrow 0\end{eqnarray}$$

in $\operatorname{Ab}(X_{\text{fppf}})$. The sheaf $i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})$ for any $x$ is an étale scheme over $X$ if the étale group $\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})$ over $k_{x}$ is constant and an étale algebraic space over $X$ in general by [Reference RaynaudRay70a, Proposition (3.3.6.1)].

Proposition 3.2.4. The above sequence induces a distinguished triangle

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\rightarrow \bigoplus _{x}\operatorname{Res}_{k_{x}/k}\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}),\end{eqnarray}$$

where $\operatorname{Res}_{k_{x}/k}$ denotes the Weil restriction functor. In particular, we have an exact sequence

$$\begin{eqnarray}\displaystyle 0 & \rightarrow & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0})\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\rightarrow \bigoplus _{x}\operatorname{Res}_{k_{x}/k}\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\nonumber\\ \displaystyle & \rightarrow & \displaystyle \mathbf{H}^{1}(X,{\mathcal{A}}_{0})\rightarrow \mathbf{H}^{1}(X,{\mathcal{A}})\rightarrow 0\nonumber\end{eqnarray}$$

and an isomorphism $\mathbf{H}^{2}(X,{\mathcal{A}}_{0})=\mathbf{H}^{2}(X,{\mathcal{A}})$. The group $\bigoplus _{x}\operatorname{Res}_{k_{x}/k}\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})$ is finite étale over $k$.

Proof. Obvious. ◻

Let ${\mathcal{G}}_{m}$ be the Néron (lft) model of $\mathbf{G}_{m}$ over $X$ [Reference Bosch, Lütkebohmert and RaynaudBLR90, 10.1/5]. It fits in the canonical exact sequence

$$\begin{eqnarray}0\rightarrow \mathbf{G}_{m}\rightarrow {\mathcal{G}}_{m}\rightarrow \bigoplus _{x}i_{x\ast }\mathbb{Z}\rightarrow 0\end{eqnarray}$$

of smooth group schemes over $X$, where the sum is over all closed points $x\in X$. At each $x$, we have Grothendieck’s pairing [Gro72, IX, 1.2.1]

$$\begin{eqnarray}\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}^{\vee })\times \unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\rightarrow \mathbb{Q}/\mathbb{Z}\end{eqnarray}$$

over $k_{x}$. Combining these two, in $D(X_{\text{fppf}})$, we have morphisms

$$\begin{eqnarray}\displaystyle & & \displaystyle \left(\bigoplus _{x}i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}^{\vee })\right)\otimes ^{L}\left(\bigoplus _{x}i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\right)\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,\bigoplus _{x}i_{x\ast }\mathbb{Q}/\mathbb{Z}\rightarrow \bigoplus _{x}i_{x\ast }\mathbb{Z}[1]\rightarrow \mathbf{G}_{m}[2].\nonumber\end{eqnarray}$$

Denote $\unicode[STIX]{x1D6F7}_{A,X}=\bigoplus _{x}i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})$ and $\unicode[STIX]{x1D6F7}_{A^{\vee },X}$ similarly. The above defines a morphism

$$\begin{eqnarray}\unicode[STIX]{x1D6F7}_{A^{\vee },X}\otimes ^{L}\unicode[STIX]{x1D6F7}_{A,X}\rightarrow \mathbf{G}_{m}[2].\end{eqnarray}$$

Recall again that we have morphisms

$$\begin{eqnarray}{\mathcal{A}}_{0}^{\vee }\otimes ^{L}{\mathcal{A}}\rightarrow \mathbf{G}_{m}[1],\qquad {\mathcal{A}}^{\vee }\otimes ^{L}{\mathcal{A}}_{0}\rightarrow \mathbf{G}_{m}[1]\end{eqnarray}$$

defined by the canonical extensions of the Poincaré biextension.

Proposition 3.2.5. The above morphisms

form a morphism of distinguished triangles in $D(X_{\text{fppf}})$, where the horizontal triangles are the natural ones.

To prove this, we need some notation and three lemmas. Let $X_{\text{sm}}$ be the smooth site of $X$, that is, the category of smooth schemes over $X$ with $X$-scheme morphisms endowed with the étale (or smooth) topology. Denote the sheaf-Hom functor for $X_{\text{sm}}$ by $\mathbf{Hom}_{X_{\text{sm}}}$ and $\mathbf{Ext}_{X_{\text{sm}}}^{n}$, $R\,\mathbf{Hom}_{X_{\text{sm}}}$ similarly (while $\mathbf{Hom}_{X}$ is still the sheaf-Hom for $X_{\text{fppf}}$).

Lemma 3.2.6. To prove Proposition 3.2.5, it is enough to show the modified statement in $D(X_{\text{sm}})$ where $R\,\mathbf{Hom}_{X}$ is replaced by $R\,\mathbf{Hom}_{X_{\text{sm}}}$.

Proof. Let $f:X_{\text{fppf}}\rightarrow X_{\text{sm}}$ be the premorphism of sites defined by the identity functor. By [Reference SuzukiSuz13, Lemma 3.7.2], the pullback functor $f^{\ast }:\operatorname{Ab}(X_{\text{sm}})\rightarrow \operatorname{Ab}(X_{\text{fppf}})$ admits a left derived functor $Lf^{\ast }:D(X_{\text{sm}})\rightarrow D(X_{\text{fppf}})$, which is left adjoint to $Rf_{\ast }:D(X_{\text{fppf}})\rightarrow D(X_{\text{sm}})$ and satisfies $L_{n}f^{\ast }\mathbb{Z}[Y]=0$ for any smooth $X$-scheme $Y$ and $n\geqslant 1$ (where $\mathbb{Z}[Y]$ is the sheaf of free abelian groups generated by the representable sheaf of sets $Y$). By [Reference SuzukiSuz18, Proposition 4.2], we have $Lf^{\ast }G=G$ for any smooth group algebraic space $G$ over $X$. Also $Rf_{\ast }G=G$ since the fppf cohomology with coefficients in a smooth group algebraic space agrees with the étale cohomology [Reference MilneMil80, III, Remark 3.11(b)]. Therefore if $H$ is another smooth group algebraic space over $X$, then

$$\begin{eqnarray}\displaystyle & & \displaystyle Rf_{\ast }R\,\mathbf{Hom}_{X}(G,H)=Rf_{\ast }R\,\mathbf{Hom}_{X}(Lf^{\ast }G,H)\nonumber\\ \displaystyle & & \displaystyle \quad =R\,\mathbf{Hom}_{X_{\text{sm}}}(G,Rf_{\ast }H)=R\,\mathbf{Hom}_{X_{\text{sm}}}(G,H)\nonumber\end{eqnarray}$$

by the derived tensor-Hom adjunction [Reference SuzukiSuz18, Proposition 3.1(1)], which is applicable to our situation since the category of smooth schemes over $X$ has finite products. Applying $Rf_{\ast }$ to the diagram in the statement, we know that the modified statement implies the original statement.◻

Lemma 3.2.7. Let $x\in X$ be a closed point and $N$ a finite étale group over $k_{x}$ with Pontryagin dual $N^{\text{PD}}$. Let $i_{x}:x_{\text{sm}}\rightarrow X_{\text{sm}}$ be the premorphism of sites defined by the inclusion $i_{x}:x{\hookrightarrow}X$. Then the truncation $\unicode[STIX]{x1D70F}_{{\leqslant}2}$ of $R\,\mathbf{Hom}_{X_{\text{sm}}}(i_{x\ast }N,\mathbf{G}_{m})$ in degrees ${\leqslant}2$ is canonically isomorphic to $i_{x\ast }N^{\text{PD}}[-2]$.

Proof. Let $j_{x}:U=X\setminus \{x\}{\hookrightarrow}X$ and denote its extension-by-zero functor by $j_{x!}:D(U_{\text{sm}})\rightarrow D(X_{\text{sm}})$. We have a distinguished triangle

$$\begin{eqnarray}R\,\mathbf{Hom}_{X_{\text{sm}}}(i_{x\ast }N,\mathbf{G}_{m})\rightarrow R\,\mathbf{Hom}_{X_{\text{sm}}}(N,\mathbf{G}_{m})\rightarrow R\,\mathbf{Hom}_{X_{\text{sm}}}(j_{x!}N,\mathbf{G}_{m})\end{eqnarray}$$

in $D(X_{\text{sm}})$ (where $N$ is base-changed to $X$). We have

$$\begin{eqnarray}R\,\mathbf{Hom}_{X_{\text{sm}}}(N,\mathbf{G}_{m})=N^{\text{PD}}\otimes ^{L}\mathbf{G}_{m}[-1],\end{eqnarray}$$
$$\begin{eqnarray}\displaystyle & & \displaystyle R\,\mathbf{Hom}_{X_{\text{sm}}}(j_{x!}N,\mathbf{G}_{m})=Rj_{x\ast }R\,\mathbf{Hom}_{U_{\text{sm}}}(N,\mathbf{G}_{m})\nonumber\\ \displaystyle & & \displaystyle \quad =Rj_{x\ast }(N^{\text{PD}}\otimes ^{L}\mathbf{G}_{m})[-1]=N^{\text{PD}}\otimes ^{L}Rj_{x\ast }\mathbf{G}_{m}[-1].\nonumber\end{eqnarray}$$

Hence

$$\begin{eqnarray}R\,\mathbf{Hom}_{X_{\text{sm}}}(i_{x\ast }N,\mathbf{G}_{m})=N^{\text{PD}}\otimes ^{L}[\mathbf{G}_{m}\rightarrow Rj_{x\ast }\mathbf{G}_{m}][-2].\end{eqnarray}$$

By definition, $j_{x\ast }\mathbf{G}_{m}$ is the Néron model over $X$ of $\mathbf{G}_{m}$ over $U$. Hence it fits in the exact sequence

$$\begin{eqnarray}0\rightarrow \mathbf{G}_{m}\rightarrow j_{x\ast }\mathbf{G}_{m}\rightarrow i_{x\ast }\mathbb{Z}\rightarrow 0.\end{eqnarray}$$

We have $R^{1}j_{x\ast }\mathbf{G}_{m}=0$ by the proof of [Reference MilneMil06, III, Lemma C.10]. Hence

$$\begin{eqnarray}\hspace{28.79993pt}\unicode[STIX]{x1D70F}_{{\leqslant}2}R\,\mathbf{Hom}_{X_{\text{sm}}}(i_{x\ast }N,\mathbf{G}_{m})=N^{\text{PD}}\otimes ^{L}i_{x\ast }\mathbb{Z}[-2]=i_{x\ast }N^{\text{PD}}[-2].\hspace{6.60004pt}\square\end{eqnarray}$$

Lemma 3.2.8. We have

$$\begin{eqnarray}\mathbf{Hom}_{X_{\text{sm}}}({\mathcal{A}},\mathbf{G}_{m})=\mathbf{Hom}_{X_{\text{sm}}}({\mathcal{A}}_{0},\mathbf{G}_{m})=0.\end{eqnarray}$$

Proof. Any morphism from ${\mathcal{A}}$ or ${\mathcal{A}}_{0}$ to $\mathbf{G}_{m}$ over any smooth scheme over $X$ is generically zero and hence zero. This implies the result.◻

Now we prove Proposition 3.2.5.

Proof of Proposition 3.2.5.

The commutativity of the left square in the diagram in the statement is easy to see. To see the commutativity of the right square, the above three lemmas show that it is enough to check the commutativity of the diagram

where we denoted

$$\begin{eqnarray}\unicode[STIX]{x1D6F7}_{A,X}^{\text{PD}}=\bigoplus _{x}i_{x\ast }(\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})^{\text{PD}})\end{eqnarray}$$

Any morphism ${\mathcal{A}}^{\vee }\rightarrow \unicode[STIX]{x1D6F7}_{A,X}^{\text{PD}}$ is determined by its values at ${\mathcal{O}}_{x}^{sh}$ for all closed points $x\in X$, where ${\mathcal{O}}_{x}^{sh}$ is the strict henselization of $X$ at $x$. Hence it is enough to show that the diagram

is commutative, where $\overline{k_{x}}$ is the algebraic closure of $k_{x}$. Let $K_{x}^{sh}$ be the fraction field of ${\mathcal{O}}_{x}^{sh}$. Let $\unicode[STIX]{x1D709}:0\rightarrow \mathbf{G}_{m}\rightarrow H\rightarrow A\rightarrow 0$ be an extension as an element of

$$\begin{eqnarray}{\mathcal{A}}^{\vee }({\mathcal{O}}_{x}^{sh})=A^{\vee }(K_{x}^{sh})=\operatorname{Ext}_{K_{x,\text{sm}}^{sh}}^{1}(A,\mathbf{G}_{m}),\end{eqnarray}$$

where the second isomorphism is the Barsotti–Weil formula [Reference OortOor66, Chapter III, Theorem (18.1)]. Let ${\mathcal{H}}$ be the Néron (lft) model over ${\mathcal{O}}_{x}^{sh}$ of $H$ and ${\mathcal{H}}_{0}$ the maximal open subgroup scheme with connected fibers. The image of $\unicode[STIX]{x1D709}$ under the left vertical morphism is the extension $0\rightarrow \mathbf{G}_{m}\rightarrow {\mathcal{H}}_{0}\rightarrow {\mathcal{A}}_{0}\rightarrow 0$ (which is exact since $R^{1}j_{x\ast }\mathbf{G}_{m}=0$ and ${\mathcal{H}}_{0}$ is of finite type). Its image under the lower vertical morphism is the extension

$$\begin{eqnarray}\unicode[STIX]{x1D702}_{1}:0\rightarrow \mathbf{G}_{m}\rightarrow {\mathcal{H}}_{0}\rightarrow {\mathcal{A}}\rightarrow i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\rightarrow 0\end{eqnarray}$$

given by composing it with $0\rightarrow {\mathcal{A}}_{0}\rightarrow {\mathcal{A}}\rightarrow i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\rightarrow 0$. On the other hand, we have an exact sequence

$$\begin{eqnarray}0\rightarrow i_{x\ast }\mathbb{Z}\rightarrow i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{H}}_{x})\rightarrow i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\rightarrow 0.\end{eqnarray}$$

The image of the extension $\unicode[STIX]{x1D709}$ under the upper horizontal morphism followed by the right vertical morphism is the extension

$$\begin{eqnarray}\unicode[STIX]{x1D702}_{2}:0\rightarrow \mathbf{G}_{m}\rightarrow {\mathcal{G}}_{m}\rightarrow i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{H}}_{x})\rightarrow i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\rightarrow 0\end{eqnarray}$$

given by the composite with $0\rightarrow \mathbf{G}_{m}\rightarrow {\mathcal{G}}_{m}\rightarrow i_{x\ast }\mathbb{Z}\rightarrow 0$. We need to show that $\unicode[STIX]{x1D702}_{1}$ and $\unicode[STIX]{x1D702}_{2}$ are equivalent. Denote by ${\mathcal{H}}^{\prime }$ the inverse image of ${\mathcal{A}}_{0}$ by ${\mathcal{H}}{\twoheadrightarrow}{\mathcal{A}}$. Consider the extension

$$\begin{eqnarray}\unicode[STIX]{x1D702}_{3}:0\rightarrow \mathbf{G}_{m}\rightarrow {\mathcal{H}}_{0}\times _{{\mathcal{A}}_{0}}{\mathcal{H}}^{\prime }\rightarrow {\mathcal{H}}\rightarrow i_{x\ast }\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})\rightarrow 0,\end{eqnarray}$$

where the first morphism (from $\mathbf{G}_{m}$ to ${\mathcal{H}}_{0}\times _{{\mathcal{A}}_{0}}{\mathcal{H}}^{\prime }$) is the inclusion into the first factor, the second the projection onto to the second factor and the third the natural morphism. We have a morphism $\unicode[STIX]{x1D702}_{3}\rightarrow \unicode[STIX]{x1D702}_{1}$ of extensions, where ${\mathcal{H}}_{0}\times _{{\mathcal{A}}_{0}}{\mathcal{H}}^{\prime }\rightarrow {\mathcal{H}}_{0}$ is the first projection. We also have a morphism $\unicode[STIX]{x1D702}_{3}\rightarrow \unicode[STIX]{x1D702}_{1}$, where ${\mathcal{H}}_{0}\times _{{\mathcal{A}}_{0}}{\mathcal{H}}^{\prime }\rightarrow {\mathcal{G}}_{m}$ is the subtraction map $(a,b)\mapsto a-b$. Therefore $\unicode[STIX]{x1D702}_{1}$ and $\unicode[STIX]{x1D702}_{2}$ are equivalent. This proves that the right square in the statement is commutative.

We finally show that the hidden square

is commutative. Interchanging the variables, this is equivalent to showing that the diagram

is commutative. This diagram is the same, up to replacing $A$ by $A^{\vee }$, as the diagram whose commutativity has just been proved.◻

Proposition 3.2.9. The diagram in the previous proposition, after applying $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\;\cdot \;)$, the cup product morphism (2.7.2) and the trace morphism (3.1.1), induce a morphism of distinguished triangles

in $D(k)$. The right vertical morphism is the sum over $x\in X$ of the Weil restrictions of Grothendieck’s pairings, which is an isomorphism [Reference SuzukiSuz14, Theorem C].

Proof. The existence of the stated morphism of distinguished triangle is self-explanatory. To show the description of the right vertical morphism, it is enough to show that the morphism $\bigoplus _{x}i_{x\ast }\mathbb{Z}\rightarrow \mathbf{G}_{m}[1]$ after applying $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\;\cdot \;)$ can be identified with the summation map $\bigoplus _{x}\operatorname{Res}_{k_{x}/k}\mathbb{Z}\rightarrow \mathbb{Z}$. The group $\bigoplus _{x}i_{x\ast }\mathbb{Z}$ can be identified with the sheaf of divisors on $X$ and the sequence $0\rightarrow \mathbf{G}_{m}\rightarrow {\mathcal{G}}_{m}\rightarrow \bigoplus _{x}i_{x\ast }\mathbb{Z}\rightarrow 0$ can be identified with the divisor exact sequence. Hence the composite

$$\begin{eqnarray}\boldsymbol{\unicode[STIX]{x1D6E4}}\left(X,\bigoplus _{x}\operatorname{Res}_{k_{x}/k}\mathbb{Z}\right)\rightarrow \mathbf{H}^{1}(X,\mathbf{G}_{m})\rightarrow \mathbb{Z}\end{eqnarray}$$

of the connecting morphism of the divisor exact sequence and the degree map is the summation map. ◻

If $k$ is algebraically closed or finite, we denote the Tate–Shafarevich group of $A$ over $K$ by $\unicode[STIX]{x0428}(A/K)$, which is the kernel of the natural homomorphism from $H^{1}(K,A)$ to the direct sum of $H^{1}(\hat{K}_{x},A)$ over the closed points $x\in X$. If $k$ is algebraically closed, we also call $\unicode[STIX]{x0428}(A/K)$ the Tate–Shafarevich group of ${\mathcal{A}}$ over $X$ and denote it by $\unicode[STIX]{x0428}({\mathcal{A}}/X)$. If $k$ is a general perfect field with algebraic closure $\overline{k}$, then the group $\unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k}}/X_{\overline{k}})$ has a natural action of $G_{k}=\operatorname{Gal}(\overline{k}/k)$. Let $\unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k}}/X_{\overline{k}})^{G_{k}}$ be the $G_{k}$-invariant part, which is independent of the choice of an algebraic closure $\overline{k}$. We use similar notation when $k$ is replaced by any perfect field $k^{\prime }$ over $k$. Consider the functor

$$\begin{eqnarray}k^{\prime }\mapsto \unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k^{\prime }}}/X_{\overline{k^{\prime }}})^{G_{k^{\prime }}}\end{eqnarray}$$

on perfect fields $k^{\prime }$ over $k$, which commutes with filtered direct limits. (Note that $\unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k^{\prime }}}/X_{\overline{k^{\prime }}})$ cannot be written as “$\unicode[STIX]{x0428}(A\times _{k}k^{\prime }/K\otimes _{k}k^{\prime })$”; the latter does not even make sense if $k^{\prime }$ is not algebraic over $k$ since the ring $K\otimes _{k}k^{\prime }$ in this case is not a field. The scheme $X_{\overline{k^{\prime }}}$ has much more closed points than $X_{\overline{k}}$, which significantly affect the definition of $\unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k^{\prime }}}/X_{\overline{k^{\prime }}})$.) The above functor uniquely extends to a functor $k^{\text{ind}\text{rat}}\rightarrow \operatorname{Ab}$ that commutes with finite products and filtered direct limits. We still denote this extended functor by $k^{\prime }\mapsto \unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k^{\prime }}}/X_{\overline{k^{\prime }}})^{G_{k^{\prime }}}$ by abuse of notation. It is obviously a sheaf for the étale topology. It is moreover a sheaf for the pro-étale topology since it commutes with filtered direct limits.

Proposition 3.2.10. The sheaf  $\mathbf{H}^{1}(X,{\mathcal{A}})$ on $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}$ is canonically isomorphic to the sheaf $k^{\prime }\mapsto \unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k^{\prime }}}/X_{\overline{k^{\prime }}})^{G_{k^{\prime }}}$.

Proof. The sheaf $\mathbf{H}^{1}(X,{\mathcal{A}})$ is locally of finite presentation as seen before. It is enough to show that the group of $k^{\prime }$-valued points of $\mathbf{H}^{1}(X,{\mathcal{A}})$ is canonically isomorphic to $\unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k^{\prime }}}/X_{\overline{k^{\prime }}})$ for any algebraically closed field $k^{\prime }$ over $k$. The former group is $H^{1}(X_{\overline{k^{\prime }}},{\mathcal{A}})$. That it is canonically isomorphic to the latter is [Reference MilneMil06, Lemma 11.5].◻

Proposition 3.2.11. The group $V_{p}H^{1}(X_{k^{\prime }},{\mathcal{A}})$ as a functor on algebraically closed fields $k^{\prime }$ over $k$ is constant.

Proof. We may assume that $k=\overline{k}$. By [Reference KatzKat99, Theorem 11], there exist a proper smooth geometrically connected curve $C$ over $K$ having a $K$-rational point and a surjective homomorphism $J{\twoheadrightarrow}A$ from the Jacobian $J$ of $C$ over $K$. By Poincaré complete reducibility, there exists a homomorphism $A\rightarrow J$ over $K$ such that the composite $A\rightarrow J{\twoheadrightarrow}A$ is multiplication by some positive integer $m$. Let ${\mathcal{J}}$ be the Néron model of $J$ over $X$. Then we have homomorphisms ${\mathcal{A}}\rightarrow {\mathcal{J}}\rightarrow {\mathcal{A}}$ over $X$ whose composite is multiplication by $m$. Therefore $V_{p}H^{1}(X,{\mathcal{A}})$ is a direct factor of $V_{p}H^{1}(X,{\mathcal{J}})$. Hence it is enough to show that $V_{p}H^{1}(X,{\mathcal{J}})$ does not depend on the algebraically closed base field $k$. Let ${\mathcal{C}}/X$ be a proper flat regular model of $C/K$ [Reference LipmanLip78]. By [Reference GrothendieckGro66, Section 4.6], there exists a canonical isomorphism $H^{1}(X,{\mathcal{J}})\cong H^{2}({\mathcal{C}},\mathbf{G}_{m})$. Hence it is enough to show that $V_{p}H^{2}({\mathcal{C}},\mathbf{G}_{m})$ does not depend on the algebraically closed base field $k$. Note that $H^{2}({\mathcal{C}},\mathbf{G}_{m})$ is the Brauer group of the proper smooth surface ${\mathcal{C}}$ over $k$.

By [Reference IllusieIll79, II, (5.8.5)], we have a canonical exact sequence

$$\begin{eqnarray}0\rightarrow \text{NS}({\mathcal{C}})\otimes \mathbb{Q}_{p}\rightarrow H^{2}({\mathcal{C}},\mathbb{Q}_{p}(1))\rightarrow V_{p}H^{2}({\mathcal{C}},\mathbf{G}_{m})\rightarrow 0,\end{eqnarray}$$

where $\text{NS}$ denotes the Néron–Severi group and the middle term is

$$\begin{eqnarray}(\underset{n}{\varprojlim }\,H^{2}({\mathcal{C}},\unicode[STIX]{x1D707}_{p^{n}}))\otimes \mathbb{Q}.\end{eqnarray}$$

The group $\text{NS}({\mathcal{C}})$ does not depend on $k$. By [Reference IllusieIll79, II, Theorem 5.5.3], there exists a canonical exact sequence

$$\begin{eqnarray}0\rightarrow H^{2}({\mathcal{C}},\mathbb{Q}_{p}(1))\rightarrow H_{\text{crys}}^{2}({\mathcal{C}}/W(k))\otimes \mathbb{Q}\overset{F-p}{\rightarrow }H_{\text{crys}}^{2}({\mathcal{C}}/W(k))\otimes \mathbb{Q}\rightarrow 0,\end{eqnarray}$$

where the last two groups are the rational crystalline cohomology. They are finite-dimensional over $W(k)[1/p]$. Hence $H^{2}({\mathcal{C}},\mathbb{Q}_{p}(1))$ is finite-dimensional over $\mathbb{Q}_{p}$ whose dimension does not depend on $k$. Therefore $V_{p}H^{2}({\mathcal{C}},\mathbf{G}_{m})$ does not depend on $k$. This finishes the proof.◻

Proposition 3.2.12. The group $\mathbf{H}^{1}(X,{\mathcal{A}})$ is in $\text{L}^{f}\text{Alg}_{\text{uc}}/k$.

Proof. This follows from the previous propositions, Proposition 3.2.3 and Proposition 2.2.3. ◻

Proposition 3.2.13. The morphism (3.2.1) induces a morphism

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\otimes \unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow \mathbb{Z}.\end{eqnarray}$$

This agrees with the height pairing [Reference Moret-BaillyMB85, III, Section 3].

Proof. This morphism is equal to the morphism

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\otimes \unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow H^{1}(X,\mathbf{G}_{m})\rightarrow \mathbb{Z}.\end{eqnarray}$$

Hence a more explicit description can be given as follows. Let $P$ be the extension of the Poincaré bundle to ${\mathcal{A}}_{0}^{\vee }\times _{X}{\mathcal{A}}$. Let $f:X\rightarrow {\mathcal{A}}_{0}^{\vee }$ and $g:X\rightarrow {\mathcal{A}}$ be sections over $X$. By pulling back $P$ by $f\times g:X\rightarrow {\mathcal{A}}_{0}^{\vee }\times _{X}{\mathcal{A}}$, we have a line bundle on $X$. Its degree is the value of the pairing at $(f,g)$. This pairing is equal to the height pairing by [Reference Moret-BaillyMB85, III, Section 3].◻

Proposition 3.2.14. Consider the morphism

$$\begin{eqnarray}\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee }))_{/\text{tor}}\times \unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))_{/\text{tor}}\rightarrow \mathbb{Z}\end{eqnarray}$$

coming from (3.2.1). The induced morphism

$$\begin{eqnarray}\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee }))_{/\text{tor}}\rightarrow \unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))_{/\text{tor}}^{\text{LD}}\end{eqnarray}$$

is injective with finite cokernel.

Proof. This follows from the previous proposition and the nondegeneracy of the height pairing [Reference ConradCon06, Theorem 9.15]. ◻

We summarize the results obtained so far.

Proposition 3.2.15. Let $C=R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ and $D=R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })$, which are objects of $D(k)$. The sheaf $H^{0}C$ is an extension of a finitely generated étale group by an abelian variety; $H^{1}C\in \text{L}^{f}\text{Alg}/k$; $H^{2}C\in \text{I}^{f}\text{Alg}/k$, which is divisible; and $H^{n}C=0$ for other values of $n$. The same are true for $D$. There is a canonical pairing $C\otimes ^{L}D\rightarrow \mathbb{Z}$ in $D(k)$. The induced morphism

$$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\end{eqnarray}$$

is injective with finite cokernel. For any $n\geqslant 1$, the induced morphism

$$\begin{eqnarray}D\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\rightarrow (C\otimes ^{L}\mathbb{Z}/n\mathbb{Z})^{\text{SD}}[1]\end{eqnarray}$$

is an isomorphism.

3.3 Formal steps toward duality for Néron models

Throughout this subsection, we fix two objects $C,D\in D(k)$ and a morphism $C\otimes ^{L}D\rightarrow \mathbb{Z}$, and assume the following:

  1. (1) The sheaf $H^{0}C$ is an extension of a finitely generated étale group by an abelian variety; $H^{1}C\in \text{L}^{f}\text{Alg}/k$; $H^{2}C\in \text{I}^{f}\text{Alg}/k$, which is divisible; and $H^{n}C=0$ for other values of $n$.

  2. (2) The same are true for $D$.

  3. (3) The morphism

    $$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\end{eqnarray}$$
    induced from $C\otimes ^{L}D\rightarrow \mathbb{Z}$ is injective and its cokernel $\unicode[STIX]{x1D6FF}_{\text{Height}}$ is finite.
  4. (4) For any $n\geqslant 1$, the induced morphism

    $$\begin{eqnarray}D\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\rightarrow (C\otimes ^{L}\mathbb{Z}/n\mathbb{Z})^{\text{SD}}[1]\end{eqnarray}$$
    is an isomorphism.

It follows from (1) that $C\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\in D^{b}(\text{Alg}_{\text{uc}}/k)$ for any $n\geqslant 1$. The same is true for $D$. Hence the isomorphism in (4) belongs to $D^{b}(\text{Alg}_{\text{uc}}/k)$.

The goal of this subsection is to prove that there exist canonical morphisms

$$\begin{eqnarray}C^{\text{SD}}\rightarrow VH^{1}D\rightarrow D^{\text{SD}\text{SD}}[1]\end{eqnarray}$$

such that the triangle

$$\begin{eqnarray}VH^{1}D[-1]\rightarrow D^{\text{SD}\text{SD}}\rightarrow C^{\text{SD}}\rightarrow VH^{1}D\end{eqnarray}$$

is distinguished (Proposition 3.3.19). In order to prove this, we first extract as much information as possible from the limit in $n$ of the isomorphism (4) (Proposition 3.3.1 to 3.3.7). Then we describe each cohomology object of $D^{\text{SD}\text{SD}}$ and $C^{\text{SD}}$ (Proposition 3.3.8). They are concentrated in degrees $-1,0,1,2$. With these two steps and (3), we can show that there exist canonical distinguished triangles

$$\begin{eqnarray}\displaystyle & \displaystyle D^{\text{SD}\text{SD}}\otimes \mathbb{Q}\rightarrow C^{\text{SD}}\otimes \mathbb{Q}\rightarrow VH^{1}D, & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\,\underset{n}{\varprojlim }\,D^{\text{SD}\text{SD}}\rightarrow R\,\underset{n}{\varprojlim }\,C^{\text{SD}}\rightarrow VH^{1}D & \displaystyle \nonumber\end{eqnarray}$$

(Proposition 3.3.11 to 3.3.13 for the first triangle and Propositions 3.3.15 to 3.3.17 for the second). These are easier to establish than the integral statement since $D^{\text{SD}\text{SD}}\otimes \mathbb{Q}$ and $C^{\text{SD}}\otimes \mathbb{Q}$ are concentrated in degrees $-1$, $0$, and $R\,\mathop{\varprojlim }\nolimits_{n}D^{\text{SD}\text{SD}}$ and $R\,\mathop{\varprojlim }\nolimits_{n}C^{\text{SD}}$ are concentrated in degrees $1$, $2$. Observe that these ranges of degrees have no intersection. This disjointness puts a strong restriction on possible choices of a mapping cone of $D^{\text{SD}\text{SD}}\rightarrow C^{\text{SD}}$. With this, we can get the desired canonical distinguished triangle.

We begin with taking limits:

Proposition 3.3.1. The morphism $D\rightarrow C^{\text{SD}}$ induces an isomorphism

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,(D\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\overset{{\sim}}{\rightarrow }R\,\underset{n}{\varprojlim }\,(C^{\text{SD}}\otimes ^{L}\mathbb{Z}/n\mathbb{Z}).\end{eqnarray}$$

Proof. For any $n\geqslant 1$, we have

$$\begin{eqnarray}C^{\text{SD}}\otimes ^{L}\mathbb{Z}/n\mathbb{Z}=(C\otimes ^{L}\mathbb{Z}/n\mathbb{Z})^{\text{SD}}[1].\end{eqnarray}$$

Hence Proposition 2.3.6 implies the result. ◻

To calculate the right-hand side, it is easier to write it with torsion objects only:

Proposition 3.3.2. We have

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }(R\,\mathbf{Hom}_{k}(C,\mathbb{Z})\otimes ^{L}\mathbb{Z}/n\mathbb{Z})=R\,\mathbf{Hom}_{k}(C\otimes ^{L}\mathbb{Q}/\mathbb{Z},\mathbb{Q}/\mathbb{Z}).\end{eqnarray}$$

Hence the isomorphism in the previous proposition can also be written as

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,(D\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\overset{{\sim}}{\rightarrow }R\,\mathbf{Hom}_{k}(C\otimes ^{L}\mathbb{Q}/\mathbb{Z},\mathbb{Q}/\mathbb{Z}).\end{eqnarray}$$

Proof. We have

$$\begin{eqnarray}R\,\mathbf{Hom}_{k}(C\otimes ^{L}\mathbb{Q}/\mathbb{Z},\mathbb{Q}/\mathbb{Z})=R\,\underset{n}{\varprojlim }\,R\,\mathbf{Hom}_{k}(C\otimes ^{L}\mathbb{Z}/n\mathbb{Z},\mathbb{Q}/\mathbb{Z})\end{eqnarray}$$

by [Reference SuzukiSuz14, Proposition (2.2.3)] (or its proof) in the notation therein. For any $n\geqslant 1$, we have

$$\begin{eqnarray}\displaystyle R\,\mathbf{Hom}_{k}(C\otimes ^{L}\mathbb{Z}/n\mathbb{Z},\mathbb{Q}/\mathbb{Z}) & = & \displaystyle R\,\mathbf{Hom}_{k}(C\otimes ^{L}\mathbb{Z}/n\mathbb{Z},\mathbb{Z}[1])\nonumber\\ \displaystyle & = & \displaystyle R\,\mathbf{Hom}_{k}(C,\mathbb{Z})\otimes ^{L}\mathbb{Z}/n\mathbb{Z}\nonumber\end{eqnarray}$$

since derived-tensoring $\mathbb{Z}/n\mathbb{Z}$ kills uniquely divisibles. The result follows by taking the limit.◻

Proposition 3.3.3. The groups $H^{2}C,H^{2}D\in \text{I}^{f}\text{Alg}/k$ are étale. The groups $H^{1}C,H^{1}D\in \text{L}^{f}\text{Alg}/k$ are divisibly ML. The isomorphism in Proposition 3.3.2 yields, on cohomology, the following duality pairings and morphisms:

  1. (1) Pontryagin duality between $T(H^{0}C)_{0}\in \text{P}\text{FEt}/k$ and $H^{2}D\in \text{I}\text{FEt}/k$.

  2. (2) Pontryagin duality between $T(H^{0}D)_{0}\in \text{P}\text{FEt}/k$ and $H^{2}C\in \text{I}\text{FEt}/k$.

  3. (3) Pontryagin duality between $(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{\text{tor}},\unicode[STIX]{x1D70B}_{0}(H^{1}D)_{/\text{div}}\in \text{FEt}/k$.

  4. (4) Pontryagin duality between $(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{\text{tor}},\unicode[STIX]{x1D70B}_{0}(H^{1}C)_{/\text{div}}\in \text{FEt}/k$.

  5. (5) An injection $(\unicode[STIX]{x1D70B}_{0}H^{1}C)_{\text{div}}^{\text{PD}}{\hookrightarrow}T(H^{1}D)_{\text{div}}$ in $\text{P}\text{FEt}/k$ whose cokernel $\unicode[STIX]{x1D6FF}_{\text{Tran}}$ is finite.

  6. (6) A surjection $(H^{1}C)_{0}^{\text{SD}^{\prime }}{\twoheadrightarrow}((H^{1}D)_{/\text{div}})_{0}$ of connected unipotent quasi-algebraic groups whose kernel $\unicode[STIX]{x1D6FF}_{\text{CT}}$ is finite.

  7. (7) An exact sequence $0\rightarrow \unicode[STIX]{x1D6FF}_{\text{Tran}}\rightarrow \unicode[STIX]{x1D6FF}_{\text{Height}}\rightarrow \unicode[STIX]{x1D6FF}_{\text{CT}}\rightarrow 0$ in $\text{FEt}/k$.

The suggestive subscripts $\text{CT}$ and $\text{Tran}$ will be explained after Theorem 3.4.1. The directions of the morphisms in (5) and (6) may look wrong, but they are indeed correct.

Proof. We set

$$\begin{eqnarray}X=C\otimes ^{L}\mathbb{Q}/\mathbb{Z},\qquad Y=R\,\mathbf{Hom}_{k}(X,\mathbb{Q}/\mathbb{Z}),\qquad Z=R\,\underset{n}{\varprojlim }\,(D\otimes ^{L}\mathbb{Z}/n\mathbb{Z}).\end{eqnarray}$$

We are going to write down the effects of the isomorphism $Z\overset{{\sim}}{\rightarrow }Y$ on their cohomology objects. For any $i$, we have

$$\begin{eqnarray}\displaystyle & \displaystyle 0\rightarrow (H^{i}C)\otimes \mathbb{Q}/\mathbb{Z}\rightarrow H^{i}X\rightarrow (H^{i+1}C)_{\text{tor}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (H^{-i+1}X)_{0}^{\text{SD}^{\prime }}\rightarrow H^{i}Y\rightarrow \unicode[STIX]{x1D70B}_{0}(H^{-i}X)^{\text{PD}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (H^{i}D)^{\wedge }\rightarrow H^{i}Z\rightarrow T(H^{i+1}D)\rightarrow 0, & \displaystyle \nonumber\end{eqnarray}$$

where the second line comes from [Reference SuzukiSuz14, Proposition (2.4.1)(b)] and [Reference MilneMil06, III, Theorem 0.14], and the third line comes from Proposition 2.3.8 and [Reference SuzukiSuz14, Proposition (2.1.2)(f)]. We have $H^{-1}X=(H^{0}C)_{\text{tor}}$. From the structure of $H^{0}C$ and the torsionness of $H^{1}C$ and $H^{2}C$, we have

$$\begin{eqnarray}\displaystyle & \displaystyle 0\rightarrow T(H^{0}C)_{0}\otimes \mathbb{Q}/\mathbb{Z}\rightarrow H^{-1}X\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{\text{tor}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow \unicode[STIX]{x1D70B}_{0}(H^{0}C)_{/\text{tor}}\otimes \mathbb{Q}/\mathbb{Z}\rightarrow H^{0}X\rightarrow H^{1}C\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle H^{1}X=H^{2}C, & \displaystyle \nonumber\end{eqnarray}$$

and $H^{i}X=0$ for other values of $i$. In particular, $H^{-1}X$ has trivial identity component. Using these, we have

$$\begin{eqnarray}\displaystyle & \displaystyle H^{-1}Y=\unicode[STIX]{x1D70B}_{0}(H^{2}C)^{\text{PD}}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (H^{2}C)_{0}^{\text{SD}}\rightarrow H^{0}Y\rightarrow \unicode[STIX]{x1D70B}_{0}(H^{0}X)^{\text{PD}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (H^{0}X)_{0}^{\text{SD}^{\prime }}\rightarrow H^{1}Y\rightarrow (H^{-1}X)^{\text{PD}}\rightarrow 0, & \displaystyle \nonumber\end{eqnarray}$$

and $H^{i}Y=0$ for other values of $i$. On the other hand, we know that $TH^{0}D=T(H^{0}D)_{0}$ and $(H^{0}D)^{\wedge }=\unicode[STIX]{x1D70B}_{0}(H^{0}D)^{\wedge }$ from the structure of $H^{0}D$ and that $(H^{2}D)^{\wedge }=0$ from the divisibility of $H^{2}D$. Hence

$$\begin{eqnarray}\displaystyle & \displaystyle H^{-1}Z=T(H^{0}D)_{0}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow \unicode[STIX]{x1D70B}_{0}(H^{0}D)^{\wedge }\rightarrow H^{0}Z\rightarrow T(H^{1}D)_{\text{div}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (H^{1}D)^{\wedge }\rightarrow H^{1}Z\rightarrow TH^{2}D\rightarrow 0, & \displaystyle \nonumber\end{eqnarray}$$

and $H^{i}Z=0$ for other values of $i$.

Since $H^{1}D\in \text{L}^{f}\text{Alg}_{\text{uc}}/k$, we know that $H^{0}Z\cong H^{0}Y$ has trivial identity component. Hence so does $H^{2}C$. Thus $H^{2}C$ is étale. Since $H^{1}C\in \text{L}^{f}\text{Alg}_{\text{uc}}/k$, we know that the identify component of $H^{0}X$ is quasi-algebraic. Hence so is $H^{1}Y\cong H^{1}Z$, and so is $TH^{2}D$. Therefore $H^{2}D$ is étale.

Hence $H^{i}Z\cong H^{i}Y$ has trivial identity component for $i\neq 1$. Comparing $H^{-1}$, $H^{0}$, $(H^{1})_{0}$, $\unicode[STIX]{x1D70B}_{0}H^{1}$ of $Y$ and $Z$, we have

$$\begin{eqnarray}\displaystyle & \displaystyle T(H^{0}D)_{0}\cong (H^{2}C)^{\text{PD}}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}D)^{\wedge }\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}X)^{\text{PD}}\rightarrow T(H^{1}D)_{\text{div}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle ((H^{1}D)^{\wedge })_{0}\cong (H^{0}X)_{0}^{\text{SD}^{\prime }}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow \unicode[STIX]{x1D70B}_{0}((H^{1}D)^{\wedge })\rightarrow (H^{-1}X)^{\text{PD}}\rightarrow TH^{2}D\rightarrow 0, & \displaystyle \nonumber\end{eqnarray}$$

respectively. We have (2) from the first line. Comparing the last line with the exact sequence for $H^{-1}X$ given above and using the finiteness of $(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{\text{tor}}$, we obtain (1), (3) and that $H^{1}D$ is divisibly ML. From the second line, we have $(\unicode[STIX]{x1D70B}_{0}H^{0}X)_{/\text{div}}=(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{\text{tor}}^{\text{PD}}$, which is finite. On the other hand, the exact sequence for $H^{0}X$ given above gives $(\unicode[STIX]{x1D70B}_{0}H^{0}X)_{/\text{div}}=(\unicode[STIX]{x1D70B}_{0}H^{1}C)_{/\text{div}}$. Hence we obtain (4) and that $H^{1}C$ is divisibly ML. By Proposition 2.3.2, we have $(H^{1}D)^{\wedge }=(H^{1}D)_{/\text{div}}$.

From the second line, we have

(3.3.1)$$\begin{eqnarray}0\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}^{\wedge }\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}X)_{\text{div}}^{\text{PD}}\rightarrow T(H^{1}D)_{\text{div}}\rightarrow 0,\end{eqnarray}$$

where $(\unicode[STIX]{x1D70B}_{0}H^{0}X)_{\text{div}}^{\text{PD}}=((\unicode[STIX]{x1D70B}_{0}H^{0}X)_{\text{div}})^{\text{PD}}$. Back to the relation between $X$ and $C$, we have

$$\begin{eqnarray}0\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}\otimes \mathbb{Q}/\mathbb{Z}\rightarrow (H^{0}X)_{0+\text{div}}\rightarrow (H^{1}C)_{0+\text{div}}\rightarrow 0.\end{eqnarray}$$

The long exact sequence of $\mathbf{Ext}_{k}^{\;\cdot \;}(\;\cdot \;,\mathbb{Q}/\mathbb{Z})$ for this short exact sequence and $((H^{1}D)_{/\text{div}})_{0}\cong (H^{0}X)_{0}^{\text{SD}^{\prime }}$ yields

(3.3.2)$$\begin{eqnarray}\displaystyle 0 & \rightarrow & \displaystyle (\unicode[STIX]{x1D70B}_{0}H^{1}C)_{\text{div}}^{\text{PD}}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}X)_{\text{div}}^{\text{PD}}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}\wedge }\nonumber\\ \displaystyle & \rightarrow & \displaystyle (H^{1}C)_{0}^{\text{SD}^{\prime }}\rightarrow ((H^{1}D)_{/\text{div}})_{0}\rightarrow 0.\end{eqnarray}$$

We will apply the lemma below to the two exact sequences (3.3.1) and (3.3.2). The composite

$$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}^{\wedge }{\hookrightarrow}(\unicode[STIX]{x1D70B}_{0}H^{0}X)_{\text{div}}^{\text{PD}}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}\wedge }\end{eqnarray}$$

is the completion of the injective morphism $(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}{\hookrightarrow}(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}$ with finite cokernel $\unicode[STIX]{x1D6FF}_{\text{Height}}$. Hence we can apply the lemma, yielding (5), (6) and (7).◻

Lemma 3.3.4. Let

$$\begin{eqnarray}\displaystyle & \displaystyle 0\rightarrow F\rightarrow W\rightarrow E^{\prime }\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow E\rightarrow W\rightarrow F^{\prime }\rightarrow G\rightarrow G^{\prime }\rightarrow 0 & \displaystyle \nonumber\end{eqnarray}$$

be exact sequences in an abelian category such that the composite $F\rightarrow W\rightarrow F^{\prime }$ is injective. Then the composite $E\rightarrow W\rightarrow E^{\prime }$ is injective, and we have an exact sequence

$$\begin{eqnarray}0\rightarrow \operatorname{Coker}(E{\hookrightarrow}E^{\prime })\rightarrow \operatorname{Coker}(F{\hookrightarrow}F^{\prime })\rightarrow \operatorname{Ker}(G{\twoheadrightarrow}G^{\prime })\rightarrow 0.\end{eqnarray}$$

Proof. Elementary. ◻

If the groups $(H^{1}C)_{0\cap \text{div}}$, $(H^{1}D)_{0\cap \text{div}}$ (which are in general finite étale $p$-groups by Proposition 2.2.5) are zero, then the morphisms in (5), (6) are simplified to more symmetric expressions

$$\begin{eqnarray}(H^{1}C)_{\text{div}}^{\text{PD}}{\hookrightarrow}T(H^{1}D)_{\text{div}},\qquad (H^{1}C)_{0}^{\text{SD}^{\prime }}{\twoheadrightarrow}(H^{1}D)_{0}.\end{eqnarray}$$

We do not assume these conditions. See Remark 3.4.2 for a little more details about this point.

Proposition 3.3.5. The cohomology objects $H^{n}$ of the complex $R\,\mathop{\varprojlim }\nolimits_{n}(D\otimes ^{L}\mathbb{Z}/n\mathbb{Z})$ are described as follows:

$$\begin{eqnarray}\displaystyle & \displaystyle H^{-1}=T(H^{0}D)_{0} & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}D)^{\wedge }\rightarrow H^{0}\rightarrow T(H^{1}D)_{\text{div}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (H^{1}D)_{/\text{div}}\rightarrow H^{1}\rightarrow TH^{2}D\rightarrow 0, & \displaystyle \nonumber\end{eqnarray}$$

and $H^{n}=0$ for $n\neq -1,0,1$. The cohomology objects $H^{n}$ of the complex $R\,\mathop{\varprojlim }\nolimits_{n}(C^{\text{SD}}\otimes ^{L}\mathbb{Z}/n\mathbb{Z})$ are described as follows:

$$\begin{eqnarray}\displaystyle & \displaystyle H^{-1}=(H^{2}C)^{\text{PD}}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{1}C)^{\text{PD}}\rightarrow H^{0}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}\wedge }\rightarrow \unicode[STIX]{x1D6FF}_{\text{CT}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow ((H^{1}D)_{/\text{div}})_{0}\rightarrow H^{1}\rightarrow (H^{0}C)_{\text{tor}}^{\text{PD}}\rightarrow 0, & \displaystyle \nonumber\end{eqnarray}$$

and $H^{n}=0$ for $n\neq -1,0,1$.

Proof. This is mostly given in the proof of Proposition 3.3.3. We show here only the second line for the second complex. Similar to (3.3.2), we have an exact sequence

$$\begin{eqnarray}\displaystyle 0 & \rightarrow & \displaystyle (\unicode[STIX]{x1D70B}_{0}H^{1}C)^{\text{PD}}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}X)^{\text{PD}}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}\wedge }\nonumber\\ \displaystyle & \rightarrow & \displaystyle (H^{1}C)_{0}^{\text{SD}^{\prime }}\rightarrow ((H^{1}D)_{/\text{div}})_{0}\rightarrow 0.\nonumber\end{eqnarray}$$

The $H^{0}$ of the second complex in the statement is $(\unicode[STIX]{x1D70B}_{0}H^{0}X)^{\text{PD}}$. The kernel of the last surjection is $\unicode[STIX]{x1D6FF}_{\text{CT}}$. Hence we get the desired exact sequence.◻

Tensoring $\mathbb{Q}$ to the isomorphism in Proposition 3.3.1, we have an isomorphism

(3.3.3)$$\begin{eqnarray}D\,\hat{\otimes }\,\mathbb{A}^{\infty }\overset{{\sim}}{\rightarrow }C^{\text{SD}}\,\hat{\otimes }\,\mathbb{A}^{\infty },\end{eqnarray}$$

where $\mathbb{A}^{\infty }=\hat{\mathbb{Z}}\otimes _{\mathbb{Z}}\mathbb{Q}\in \text{I}\text{P}\text{FEt}/k$ as in Section 2.3.

Proposition 3.3.6. The cohomology objects $H^{n}$ of the complex $D\,\hat{\otimes }\,\mathbb{A}^{\infty }$ are described as follows:

$$\begin{eqnarray}\displaystyle & \displaystyle H^{-1}=V(H^{0}D)_{0}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\otimes \mathbb{A}^{\infty }\rightarrow H^{0}\rightarrow V(H^{1}D)_{\text{div}}\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle H^{1}=VH^{2}D, & \displaystyle \nonumber\end{eqnarray}$$

and $H^{n}=0$ for other degrees. The cohomology objects $H^{n}$ of the complex $C^{\text{SD}}\,\hat{\otimes }\,\mathbb{A}^{\infty }$ are described as follows:

$$\begin{eqnarray}\displaystyle & \displaystyle H^{-1}=(VH^{2}C)^{\text{PD}}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle 0\rightarrow (V(H^{1}C)_{\text{div}})^{\text{PD}}\rightarrow H^{0}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\otimes \mathbb{A}^{\infty }\rightarrow 0, & \displaystyle \nonumber\\ \displaystyle & \displaystyle H^{1}=(V(H^{0}C)_{0})^{\text{PD}}, & \displaystyle \nonumber\end{eqnarray}$$

and $H^{n}=0$ for other degrees.

Proof. This follows from the previous proposition. ◻

Proposition 3.3.7. The isomorphism (3.3.3) induces Pontryagin duality between $V(H^{0}D)_{0}$ and $VH^{2}C$ and between $VH^{2}D$ and $V(H^{0}C)_{0}$. In the previous proposition, both of the sequence for $H^{0}$ of $D\,\hat{\otimes }\,\mathbb{A}^{\infty }$ and the sequence for $H^{0}$ of $C\,\hat{\otimes }\,\mathbb{A}^{\infty }$ canonically split. The parts $(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\otimes \mathbb{A}^{\infty }$ and $(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}\otimes \mathbb{A}^{\infty }$ are Pontryagin dual to each other. The parts $V(H^{1}D)_{\text{div}}$ and $V(H^{1}C)_{\text{div}}$ are Pontryagin dual to each other.

Proof. The induced morphism

$$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\otimes \mathbb{A}^{\infty }\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\otimes \mathbb{A}^{\infty }\end{eqnarray}$$

is an isomorphism. The rest follows from this. ◻

Proposition 3.3.8. The cohomology objects $H^{n}$ of the complex $D^{\text{SD}\text{SD}}$ are described as follows:

$$\begin{eqnarray}\displaystyle & \displaystyle H^{-1}=T(H^{0}D)_{0},\qquad H^{0}=\unicode[STIX]{x1D70B}_{0}H^{0}D, & \displaystyle \nonumber\\ \displaystyle & \displaystyle H^{1}=H^{1}D,\qquad H^{2}=H^{2}D, & \displaystyle \nonumber\end{eqnarray}$$

and $H^{n}=0$ for other values of $n$. The cohomology objects $H^{n}$ of the complex $C^{\text{SD}}$ are described as follows:

$$\begin{eqnarray}H^{-1}=(H^{2}C)^{\text{PD}},\end{eqnarray}$$
$$\begin{eqnarray}\displaystyle 0 & \rightarrow & \displaystyle (\unicode[STIX]{x1D70B}_{0}H^{1}C)^{\text{PD}}\rightarrow H^{0}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\nonumber\\ \displaystyle & \rightarrow & \displaystyle (H^{1}C)_{0}^{\text{SD}^{\prime }}\rightarrow H^{1}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{\text{tor}}^{\text{PD}}\rightarrow 0,\nonumber\end{eqnarray}$$
$$\begin{eqnarray}H^{2}=(T(H^{0}C)_{0})^{\text{PD}},\end{eqnarray}$$

and $H^{n}=0$ for other values of $n$.

Proof. First we treat $D^{\text{SD}\text{SD}}$. The cohomology objects of the mapping cone

$$\begin{eqnarray}[(H^{0}D)_{0}\rightarrow D]\end{eqnarray}$$

are in $\text{FGEt}/k$ in degree $0$, in $\text{I}\text{Alg}_{\text{uc}}/k$ in degrees $1$, $2$ and zero in other degrees. Each of them is Serre reflexive by [Reference SuzukiSuz14, Proposition (2.4.1)(b)]. Hence this mapping cone itself is Serre reflexive. On the other hand, the double Serre dual of the abelian variety $(H^{0}C)_{0}$ over $k$ is its Tate module placed in degree $-1$ by [Reference SuzukiSuz14, Proposition (2.4.1)(c)]. Therefore we have a distinguished triangle

$$\begin{eqnarray}T(H^{0}D)_{0}[1]\rightarrow D^{\text{SD}\text{SD}}\rightarrow [(H^{0}D)_{0}\rightarrow D].\end{eqnarray}$$

The desired description follows from this.

For $C^{\text{SD}}$, consider the hyperext spectral sequence

$$\begin{eqnarray}E_{2}^{ij}=\mathbf{Ext}_{k}^{i}(H^{-j}C,\mathbb{Z})\Longrightarrow H^{i+j}R\,\mathbf{Hom}_{k}(C,\mathbb{Z}).\end{eqnarray}$$

For any object $E$ of $\text{I}\text{Alg}/k$ or $\text{FGEt}/k$, we have

$$\begin{eqnarray}\displaystyle & \displaystyle \mathbf{Hom}_{k}(E,\mathbb{Z})=(\unicode[STIX]{x1D70B}_{0}E)_{/\text{tor}}^{\text{LD}},\qquad \mathbf{Ext}_{k}^{1}(E,\mathbb{Z})=(\unicode[STIX]{x1D70B}_{0}E)_{\text{tor}}^{\text{PD}}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle \mathbf{Ext}_{k}^{2}(E,\mathbb{Z})=E_{0}^{\text{SD}^{\prime }},\qquad \mathbf{Ext}_{k}^{{\geqslant}3}(E,\mathbb{Z})=0 & \displaystyle \nonumber\end{eqnarray}$$

by [Reference SuzukiSuz14, Proposition (2.4.1)(a), (b)]. We have

$$\begin{eqnarray}(H^{1}C)_{/\text{tor}}=(H^{2}C)_{/\text{tor}}=(H^{2}C)_{0}=0\end{eqnarray}$$

by Proposition 3.3.3. Hence the $E_{2}^{ij}$-term of the above spectral sequence is zero unless $(i,j)$ is $(-1,-2)$, $(-1,-1)$, $(0,0)$, $(2,-1)$ or $(2,0)$. We have

$$\begin{eqnarray}(H^{0}C)_{0}^{\text{SD}^{\prime }}=(T(H^{0}C)_{0})^{\text{PD}}\end{eqnarray}$$

by [Reference SuzukiSuz14, Proposition (2.4.1)(c)]. Therefore the $E_{2}$-sheet gives the desired description.◻

Proposition 3.3.9. The morphism

$$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\rightarrow (H^{1}C)_{0}^{\text{SD}^{\prime }}\end{eqnarray}$$

in the previous proposition and the morphism

$$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}\wedge }\rightarrow (H^{1}C)_{0}^{\text{SD}^{\prime }}\end{eqnarray}$$

in (the proof of) Proposition 3.3.5 are compatible under the natural morphism $(\;\cdot \;)\rightarrow (\;\cdot \;)^{\wedge }$.

Proof. The first morphism is the differential

(3.3.4)$$\begin{eqnarray}\mathbf{Hom}_{k}(H^{0}C,\mathbb{Z})\rightarrow \mathbf{Ext}_{k}^{2}(H^{1}C,\mathbb{Z})\end{eqnarray}$$

of the spectral sequence with $E_{2}^{ij}=\mathbf{Ext}_{k}^{i}(H^{-j}C,\mathbb{Z})$. The second is the connecting morphism

$$\begin{eqnarray}\mathbf{Hom}_{k}((H^{0}C)\otimes \mathbb{Q}/\mathbb{Z},\mathbb{Q}/\mathbb{Z})\rightarrow \mathbf{Ext}_{k}^{1}(H^{1}C,\mathbb{Q}/\mathbb{Z})\end{eqnarray}$$

for the short exact sequence

(3.3.5)$$\begin{eqnarray}0\rightarrow (H^{0}C)\otimes \mathbb{Q}/\mathbb{Z}\rightarrow H^{0}(C\otimes ^{L}\mathbb{Q}/\mathbb{Z})\rightarrow H^{1}C\rightarrow 0.\end{eqnarray}$$

We want to show that these morphisms are compatible under the natural morphism

$$\begin{eqnarray}\mathbf{Hom}_{k}(H^{0}C,\mathbb{Z})\rightarrow \mathbf{Hom}_{k}((H^{0}C)\otimes \mathbb{Q}/\mathbb{Z},\mathbb{Q}/\mathbb{Z})\end{eqnarray}$$

and the isomorphism

$$\begin{eqnarray}\mathbf{Ext}_{k}^{1}(H^{1}C,\mathbb{Q}/\mathbb{Z})=\mathbf{Ext}_{k}^{2}(H^{1}C,\mathbb{Z}).\end{eqnarray}$$

By general nonsense, (3.3.4) comes from the (shifted) connecting morphism $(H^{1}C)[-2]\rightarrow H^{0}C$ of the truncation distinguished triangle $H^{0}C\rightarrow \unicode[STIX]{x1D70F}_{{\leqslant}1}C\rightarrow (H^{1}C)[-1]$. This triangle induces a distinguished triangle

$$\begin{eqnarray}(H^{0}C)\otimes ^{L}\mathbb{Q}/\mathbb{Z}\rightarrow (\unicode[STIX]{x1D70F}_{{\leqslant}1}C)\otimes ^{L}\mathbb{Q}/\mathbb{Z}\rightarrow (H^{1}C)\otimes ^{L}\mathbb{Q}/\mathbb{Z}[-1].\end{eqnarray}$$

Taking $H^{0}$, we have an exact sequence

$$\begin{eqnarray}0\rightarrow (H^{0}C)\otimes \mathbb{Q}/\mathbb{Z}\rightarrow H^{0}((\unicode[STIX]{x1D70F}_{{\leqslant}1}C)\otimes ^{L}\mathbb{Q}/\mathbb{Z})\rightarrow H^{1}C\rightarrow 0,\end{eqnarray}$$

which recovers (3.3.5). From these observations, we can finish the comparison by applying $\mathbf{Ext}_{k}^{\;\cdot \;}(\;\cdot \;,\mathbb{Z})$.◻

Proposition 3.3.10. We have

$$\begin{eqnarray}\displaystyle & \displaystyle R\,\underset{n}{\varprojlim }\,(C\otimes ^{L}\mathbb{Z}/n\mathbb{Z})=R\,\underset{n}{\varprojlim }\,(C^{\text{SD}\text{SD}}\otimes ^{L}\mathbb{Z}/n\mathbb{Z}), & \displaystyle \nonumber\\ \displaystyle & \displaystyle C\,\hat{\otimes }\,\mathbb{A}^{\infty }=C^{\text{SD}\text{SD}}\,\hat{\otimes }\,\mathbb{A}^{\infty }. & \displaystyle \nonumber\end{eqnarray}$$

The same is true for $D$.

Proof. By Proposition 3.3.8, we have a distinguished triangle

$$\begin{eqnarray}V(H^{0}C)_{0}[1]\rightarrow C\rightarrow C^{\text{SD}\text{SD}}.\end{eqnarray}$$

We have

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }(V(H^{0}C)_{0}\otimes ^{L}\mathbb{Z}/n\mathbb{Z})=0\end{eqnarray}$$

since $V(H^{0}C)_{0}$ is uniquely divisible. This implies the result.◻

Proposition 3.3.11. The cohomology objects $H^{n}$ of the complex $D^{\text{SD}\text{SD}}\otimes \mathbb{Q}$ are described as follows:

$$\begin{eqnarray}H^{-1}=V(H^{0}D)_{0},\qquad H^{0}=(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\otimes \mathbb{Q},\end{eqnarray}$$

and $H^{n}=0$ for other values of $n$. The cohomology objects $H^{n}$ of the complex $C^{\text{SD}}\otimes \mathbb{Q}$ are described as follows:

$$\begin{eqnarray}H^{-1}=(VH^{2}C)^{\text{PD}},\qquad 0\rightarrow (VH^{1}C)^{\text{PD}}\rightarrow H^{0}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\otimes \mathbb{Q}\rightarrow 0,\end{eqnarray}$$

and $H^{n}=0$ for other values of $n$.

Proof. Tensor $\mathbb{Q}$ with the isomorphisms and exact sequences in Proposition 3.3.8.◻

Proposition 3.3.12. The morphism

$$\begin{eqnarray}D^{\text{SD}\text{SD}}\otimes \mathbb{Q}\rightarrow C^{\text{SD}}\otimes \mathbb{Q}\end{eqnarray}$$

induces an injection onto a direct summand in degree $0$ with cokernel $(VH^{1}C)^{\text{PD}}$ and isomorphisms on cohomology in other degrees.

Proof. From degree zero, we have a morphism

$$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\otimes \mathbb{Q}\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\otimes \mathbb{Q}.\end{eqnarray}$$

This is an isomorphism since $\unicode[STIX]{x1D6FF}_{\text{Height}}$ is finite. Together with the previous proposition, the statement about degree zero follows. For degree $-1$, consider the commutative diagram

The vertical morphisms induce isomorphisms in degree $-1$ by the previous proposition and Proposition 3.3.7. The last mentioned proposition also shows that the lower horizontal morphism induces an isomorphism in degree $-1$. Hence so is the upper morphism.◻

In particular, we have two canonical morphisms

$$\begin{eqnarray}D^{\text{SD}\text{SD}}\otimes \mathbb{Q}\rightarrow C^{\text{SD}}\otimes \mathbb{Q}\rightarrow (VH^{1}C)^{\text{PD}}.\end{eqnarray}$$

Proposition 3.3.13. There exists a unique morphism

$$\begin{eqnarray}(VH^{1}C)^{\text{PD}}[-1]\rightarrow D^{\text{SD}\text{SD}}\otimes \mathbb{Q}\end{eqnarray}$$

such that the resulting triangle

$$\begin{eqnarray}(VH^{1}C)^{\text{PD}}[-1]\rightarrow D^{\text{SD}\text{SD}}\otimes \mathbb{Q}\rightarrow C^{\text{SD}}\otimes \mathbb{Q}\rightarrow (VH^{1}C)^{\text{PD}}\end{eqnarray}$$

is distinguished.

Proof. The existence is clear from the previous proposition. The uniqueness follows from the following general lemma below. ◻

Lemma 3.3.14. If $E\rightarrow F\rightarrow G$ are two morphisms in the derived category of an abelian category such that $E$ (resp. $G$) is concentrated in nonpositive (resp. nonnegative) degrees and if there exists a morphism $G[-1]\rightarrow E$ that yields a distinguished triangle $G[-1]\rightarrow E\rightarrow F\rightarrow G$, then such a morphism $G[-1]\rightarrow E$ is unique.

Proof. Let $f,g:G[-1]\rightrightarrows E$ be two morphisms such that the two triangles $G[-1]\rightrightarrows E\rightarrow F\rightarrow G$ are both distinguished. By an axiom of triangulated category, there exists an automorphism $h$ on $G$ such that the diagram

is commutative. Hence the diagram

is commutative. Therefore there exists a morphism $r:E[1]\rightarrow G$ such that $h-1=r\,\circ \,f$. But the assumptions on $E$ and $G$ imply that $\operatorname{Hom}(E[1],G)=0$. Hence $r=0$, $h=1$ and thus $f=g$.◻

Proposition 3.3.15. The cohomology objects $H^{n}$ of the complex $R\,\mathop{\varprojlim }\nolimits_{n}D^{\text{SD}\text{SD}}$ are described as follows:

$$\begin{eqnarray}0\rightarrow (\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\otimes \mathbb{A}^{\infty }/\mathbb{Q}\rightarrow H^{1}\rightarrow VH^{1}D\rightarrow 0,\qquad H^{2}=VH^{2}D,\end{eqnarray}$$

and $H^{n}=0$ for other values of $n$. The cohomology objects $H^{n}$ of the complex $R\,\mathop{\varprojlim }\nolimits_{n}C^{\text{SD}}$ are described as follows:

$$\begin{eqnarray}H^{1}=(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\otimes \mathbb{A}^{\infty }/\mathbb{Q},\qquad H^{2}=(V(H^{0}C)_{0})^{\text{PD}},\end{eqnarray}$$

and $H^{n}=0$ for other values of $n$.

Proof. For $R\,\mathop{\varprojlim }\nolimits_{n}D^{\text{SD}\text{SD}}$, apply $R\,\mathop{\varprojlim }\nolimits_{n}$ to the groups in Proposition 3.3.8 and use Propositions 2.3.3 and 2.3.4.

For $R\,\mathop{\varprojlim }\nolimits_{n}C^{\text{SD}}$, we have

$$\begin{eqnarray}\displaystyle R\,\underset{n}{\varprojlim }\,R\,\mathbf{Hom}_{k}(C,\mathbb{Z}) & = & \displaystyle R\,\mathbf{Hom}_{k}(C\otimes \mathbb{Q},\mathbb{Z})\nonumber\\ \displaystyle & = & \displaystyle R\,\mathbf{Hom}_{k}(H^{0}C\otimes \mathbb{Q},\mathbb{Z})\nonumber\\ \displaystyle & = & \displaystyle R\,\underset{n}{\varprojlim }\,R\,\mathbf{Hom}_{k}(H^{0}C,\mathbb{Z})\nonumber\end{eqnarray}$$

by [Reference SuzukiSuz14, Proposition (2.3.3)(c)] and the torsionness result of higher cohomology in Proposition 3.3.3. The cohomology objects $H^{n}$ of the complex $R\,\mathbf{Hom}_{k}(H^{0}C,\mathbb{Z})$ are

$$\begin{eqnarray}H^{0}=(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}},\qquad H^{1}=(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{\text{tor}}^{\text{PD}},\qquad H^{2}=(T(H^{0}C)_{0})^{\text{PD}}\end{eqnarray}$$

by the same argument as the proof of Proposition 3.3.8. Applying $R\,\mathop{\varprojlim }\nolimits_{n}$ to these groups and using Propositions 2.3.3 and 2.3.4, we get the desired description.◻

Proposition 3.3.16. The morphism

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,D^{\text{SD}\text{SD}}\rightarrow R\,\underset{n}{\varprojlim }\,C^{\text{SD}}\end{eqnarray}$$

is a surjection onto a direct summand on cohomology in degree $1$ with kernel $VH^{1}D$ and an isomorphism on cohomology in any other degree.

Proof. The same argument as the proof of Proposition 3.3.12 works, this time using the isomorphism

$$\begin{eqnarray}(\unicode[STIX]{x1D70B}_{0}H^{0}D)_{/\text{tor}}\otimes \mathbb{A}^{\infty }/\mathbb{Q}\overset{{\sim}}{\rightarrow }(\unicode[STIX]{x1D70B}_{0}H^{0}C)_{/\text{tor}}^{\text{LD}}\otimes \mathbb{A}^{\infty }/\mathbb{Q}\end{eqnarray}$$

and the commutative diagram

In particular, we have two canonical morphisms

$$\begin{eqnarray}VH^{1}D[-1]\rightarrow R\,\underset{n}{\varprojlim }\,D^{\text{SD}\text{SD}}\rightarrow R\,\underset{n}{\varprojlim }\,C^{\text{SD}}\end{eqnarray}$$

Proposition 3.3.17. There exists a unique morphism

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }\,C^{\text{SD}}\rightarrow VH^{1}D\end{eqnarray}$$

such that the resulting triangle

$$\begin{eqnarray}VH^{1}D[-1]\rightarrow R\,\underset{n}{\varprojlim }\,D^{\text{SD}\text{SD}}\rightarrow R\,\underset{n}{\varprojlim }\,C^{\text{SD}}\rightarrow VH^{1}D\end{eqnarray}$$

is distinguished.

Proof. Similar to the proof of Proposition 3.3.13. ◻

Proposition 3.3.18. Consider the natural commutative diagram

where the rows are distinguished. The isomorphism $VH^{1}D\overset{{\sim}}{\rightarrow }(VH^{1}C)^{\text{PD}}$ in Proposition 3.3.7 is the unique morphism that completes the above diagram into a morphism of distinguished triangles.

Proof. A morphism $VH^{1}D\rightarrow (VH^{1}C)^{\text{PD}}$ with the required property exists by an axiom of triangulated categories. We need to show that such a morphism has to be the isomorphism of Proposition 3.3.7.

Consider the morphism of distinguished triangles

coming from Propositions 2.3.9, 3.3.7 and 3.3.10. Denote the upper triangle by $E\rightarrow F\rightarrow G$ and the lower by $E^{\prime }\rightarrow F^{\prime }\rightarrow G^{\prime }$. By Propositions 3.3.6, 3.3.11 and 3.3.15, we have a commutative diagram with exact rows

Hence any morphism $VH^{1}D\rightarrow (VH^{1}C)^{\text{PD}}$ with the required property has to be the connecting morphism

$$\begin{eqnarray}\operatorname{Ker}(H^{1}E\rightarrow H^{1}E^{\prime })\rightarrow \operatorname{Coker}(H^{0}F\rightarrow H^{0}F^{\prime })\end{eqnarray}$$

of the snake lemma for this diagram. Using Propositions 3.3.7, 3.3.12 and 3.3.16, we can see that this is indeed the isomorphism of Proposition 3.3.7. ◻

By Propositions 3.3.13 and 3.3.17, we have morphisms

(3.3.6)$$\begin{eqnarray}\displaystyle & \displaystyle C^{\text{SD}}\rightarrow C^{\text{SD}}\otimes \mathbb{Q}\rightarrow (VH^{1}C)^{\text{PD}}, & \displaystyle\end{eqnarray}$$
(3.3.7)$$\begin{eqnarray}\displaystyle & \displaystyle VH^{1}D[-1]\rightarrow R\,\underset{n}{\varprojlim }\,D^{\text{SD}\text{SD}}\rightarrow D^{\text{SD}\text{SD}}. & \displaystyle\end{eqnarray}$$

Together with the isomorphism $(VH^{1}C)^{\text{PD}}\cong VH^{1}D$ of Proposition 3.3.7, we have a triangle

$$\begin{eqnarray}VH^{1}D[-1]\rightarrow D^{\text{SD}\text{SD}}\rightarrow C^{\text{SD}}\rightarrow VH^{1}D\end{eqnarray}$$

Proposition 3.3.19. The above triangle is distinguished.

Proof. Let $W$ be any mapping cone of the morphism $D^{\text{SD}\text{SD}}\rightarrow C^{\text{SD}}$. Consider the commutative diagram of distinguished triangles

The left lower horizontal morphism is an isomorphism by Propositions 3.3.1 and 3.3.10. The (canonical choice of a) mapping fiber of the left upper horizontal morphism is $VH^{1}D[-1]$ by Proposition 3.3.17. Therefore $W$ can actually be taken as $VH^{1}D$, that is, there exists a morphism of distinguished triangle

The lower horizontal morphism in the left square has to be the morphism (3.3.7). For this choice of $W$, we show that the connecting morphism $C^{\text{SD}}\rightarrow VH^{1}D$ for the lower triangle is equal to the morphism (3.3.6) composed with the isomorphism $(VH^{1}C)^{\text{PD}}\cong VH^{1}D$ of Proposition 3.3.7. Consider the commutative diagram

By an axiom of triangulated categories, there exists a morphism $f:VH^{1}D\rightarrow (VH^{1}C)^{\text{PD}}$ that completes the diagram into a morphism of distinguished triangles. Hence we have a morphism of distinguished triangles

Applying Proposition 3.3.18 to this diagram, we know that $f$ has to be the isomorphism given by Proposition 3.3.7. Hence the morphism $C^{\text{SD}}\rightarrow (VH^{1}C)^{\text{PD}}$ has to come from the morphism (3.3.6).◻

Suppose that we have two other objects $C^{\prime },D^{\prime }\in D(k)$ and a morphism $C^{\prime }\otimes ^{L}D^{\prime }\rightarrow \mathbb{Z}$ satisfying the same assumptions (1)–(4) listed at the beginning of this subsection. Suppose also that we have a perfect pairing $C^{\prime \prime }\times D^{\prime \prime }\rightarrow \mathbb{Q}/\mathbb{Z}$ of finite étale groups over $k$. Suppose finally that we have two distinguished triangles $C^{\prime }\rightarrow C\rightarrow C^{\prime \prime }$ and $D\rightarrow D^{\prime }\rightarrow D^{\prime \prime }$ such that the morphisms

form a morphism of distinguished triangles.

Proposition 3.3.20. The morphism $D\rightarrow D^{\prime }$ induces an isomorphism $VH^{1}D\overset{{\sim}}{\rightarrow }VH^{1}D^{\prime }$. The diagram

is a commutative diagram of distinguished triangles.

Proof. Since $C^{\prime \prime }$ and $D^{\prime \prime }$ are finite, the morphism $H^{1}D\rightarrow H^{1}D^{\prime }$ is surjective with finite kernel. Hence $VH^{1}D\overset{{\sim}}{\rightarrow }VH^{1}D^{\prime }$. For the commutativity, the only part to check is hidden in the above diagram: the composite $VH^{1}D^{\prime }[-1]\rightarrow {D^{\prime }}^{\text{SD}\text{SD}}\rightarrow D^{\prime \prime }$ is zero; and the composite ${C^{\prime \prime }}^{\text{PD}}[-1]\rightarrow C^{\text{SD}}\rightarrow VH^{1}D$ is zero. These are obvious since there are no nonzero morphisms between finite groups and uniquely divisible groups with any shift.◻

3.4 Main theorem

Now we apply the results of the previous subsection to $C=R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$ and $D=R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })$. Statement (5) in the following theorem proves Theorem A in Introduction.

Theorem 3.4.1.
  1. (1) We have $\mathbf{H}^{n}(X,{\mathcal{A}})\in \text{L}\text{Alg}/k$ for any $n$ and $\mathbf{H}^{n}(X,{\mathcal{A}})=0$ for $n\neq 0,1,2$.

  2. (2)

    • $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})_{0}$ is an abelian variety.

    • $\unicode[STIX]{x1D70B}_{0}\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\in \text{FGEt}/k$.

    • $\mathbf{H}^{1}(X,{\mathcal{A}})_{0}$ is unipotent quasi-algebraic.

    • $\unicode[STIX]{x1D70B}_{0}\mathbf{H}^{1}(X,{\mathcal{A}})\in \text{Et}/k$ is torsion of cofinite type.

    • $\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}}\in \text{Et}/k$ is torsion of cofinite type.

    • $\mathbf{H}^{1}(X,{\mathcal{A}})_{/\text{div}}\in \text{Alg}_{\text{uc}}/k$.

    • $\mathbf{H}^{2}(X,{\mathcal{A}})\in \text{Et}/k$ is divisible torsion of cofinite type.

  3. (3) The same statements as (1) and (2) hold with ${\mathcal{A}}$ replaced by ${\mathcal{A}}_{0}^{\vee }$.

  4. (4) The group of $k^{\prime }$-valued points of $\mathbf{H}^{1}(X,{\mathcal{A}})$ for $k^{\prime }\in k^{\text{ind}\text{rat}}$ is given by $\unicode[STIX]{x0428}({\mathcal{A}}_{\overline{k^{\prime }}}/X_{\overline{k^{\prime }}})^{G_{k^{\prime }}}$ in the notation of the paragraph before Proposition 3.2.10.

  5. (5) There exists a canonical distinguished triangle

    $$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}}\rightarrow V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}.\end{eqnarray}$$
    The first morphism is induced from the morphism (3.2.1).
  6. (6) This distinguished triangle induces, on cohomology, the following duality pairings and morphisms:

    1. (a) Pontryagin duality between $T(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })_{0})$ and $\mathbf{H}^{2}(X,{\mathcal{A}})$.

    2. (b) Pontryagin duality between $T(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})_{0})$ and $\mathbf{H}^{2}(X,{\mathcal{A}}_{0}^{\vee })$.

    3. (c) Pontryagin duality between $(\unicode[STIX]{x1D70B}_{0}\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))_{\text{tor}}$ and $\unicode[STIX]{x1D70B}_{0}\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{/\text{div}}$.

    4. (d) Pontryagin duality between $(\unicode[STIX]{x1D70B}_{0}\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee }))_{\text{tor}}$ and $\unicode[STIX]{x1D70B}_{0}\mathbf{H}^{1}(X,{\mathcal{A}})_{/\text{div}}$.

    5. (e) An injection $\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee }))_{/\text{tor}}{\hookrightarrow}\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))_{/\text{tor}}^{\text{LD}}$ whose cokernel $\unicode[STIX]{x1D6FF}_{\text{Height}}$ is finite étale.

    6. (f) An injection $(\unicode[STIX]{x1D70B}_{0}\mathbf{H}^{1}(X,{\mathcal{A}}))_{\text{div}}^{\text{PD}}{\hookrightarrow}T\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$ whose cokernel $\unicode[STIX]{x1D6FF}_{\text{Tran}}$ is finite étale.

    7. (g) A surjection $\mathbf{H}^{1}(X,{\mathcal{A}})_{0}^{\text{SD}^{\prime }}{\twoheadrightarrow}(\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{/\text{div}})_{0}$ whose kernel $\unicode[STIX]{x1D6FF}_{\text{CT}}$ is finite étale.

    8. (h) An exact sequence $0\rightarrow \unicode[STIX]{x1D6FF}_{\text{Tran}}\rightarrow \unicode[STIX]{x1D6FF}_{\text{Height}}\rightarrow \unicode[STIX]{x1D6FF}_{\text{CT}}\rightarrow 0$.

    In particular, the morphism (6f) induces a Pontryagin duality between $V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}}$ and $V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$.

  7. (7) The morphism (6e) agrees with the height pairing.

  8. (8) The distinguished triangle (5) and the corresponding triangle with $A,A^{\vee }$ switched fit in the commutative diagram of distinguished triangles

    The left lower horizontal morphism is the sum over all closed points $x\in X$ of the Weil restrictions of Grothendieck’s pairings, which is an isomorphism [Reference SuzukiSuz14, Theorem C].

Proof. (1) and (2) follow from Propositions 3.2.15 and 3.3.3. (3) follows from Proposition 3.2.4. (4) is Proposition 3.2.10. (5) follows from Proposition 3.3.19. (6) follows from Propositions 3.3.3 and 3.2.15. (7) is Proposition 3.2.13. (8) follows from Propositions 3.2.9 and 3.3.20. ◻

Several comments are in order. We consider (6g) as a geometric analogue of the Cassels–Tate pairing in view of Proposition 4.2.8, whence the symbol $\unicode[STIX]{x1D6FF}_{\text{CT}}$. The divisible part of the Tate–Shafarevich group when $k$ is algebraically closed is of transcendental nature, whence the symbol $\unicode[STIX]{x1D6FF}_{\text{Tran}}$. Confusingly, the morphisms in (6f) and (6g) are from the duals, not to.

The exact sequence (6h) is mysterious. It came from applying the formal procedure of Lemma 3.3.4 to the two exact sequences (3.3.1) and (3.3.2) (where one should note that the central term $H^{0}X$ in the notation there is the “Selmer scheme” explained below). The morphism $\unicode[STIX]{x1D6FF}_{\text{Height}}{\twoheadrightarrow}\unicode[STIX]{x1D6FF}_{\text{CT}}$ is induced by the morphism

$$\begin{eqnarray}\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))_{/\text{tor}}^{\text{LD}}\rightarrow \mathbf{H}^{1}(X,{\mathcal{A}})_{0}^{\text{SD}^{\prime }}\end{eqnarray}$$

of Propositions 3.3.8 and 3.3.9 (i.e., the latter morphism factors through $\unicode[STIX]{x1D6FF}_{\text{Height}}{\twoheadrightarrow}\unicode[STIX]{x1D6FF}_{\text{CT}}$). This morphism is analogous to Artin’s period map [Reference ArtinArt74] for supersingular K3 surfaces.

Milne [Reference MilneMil06, III, paragraph before Theorem 11.6] made the hypothesis that the sheaf $R^{2}\unicode[STIX]{x1D70B}_{X}^{\text{perf}}{\mathcal{A}}$ on $\operatorname{Spec}k_{\text{et}}^{\text{perf}}$ has no connected part. This sheaf restricted to $\operatorname{Spec}k_{\text{et}}^{\text{ind}\text{rat}}$ is $\mathbf{H}^{2}(X,{\mathcal{A}})$ by Proposition 2.7.1. Hence the result $\mathbf{H}^{2}(X,{\mathcal{A}})\in \text{Et}/k$ in (2) above says that his hypothesis is true at least “birationally.”

Also [Reference MilneMil06, III, Theorem 11.6] (plus the fact that $H^{2}(K\otimes \overline{k},N)=0$ for finite flat $N$ and hence $H^{2}(K\otimes \overline{k},A)=0$) says that $\mathbf{H}^{2}(X,{\mathcal{A}})$ injects into the Pontryagin dual of $T(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })_{0})$. Thus (6a) above shows that this injection is actually bijective.

We have an exact sequence

$$\begin{eqnarray}0\rightarrow \unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))_{/\text{tor}}\otimes \mathbb{Q}/\mathbb{Z}\rightarrow H^{0}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Q}/\mathbb{Z})\rightarrow \mathbf{H}^{1}(X,{\mathcal{A}})\rightarrow 0\end{eqnarray}$$

in $\text{L}^{f}\text{Alg}_{\text{uc}}/k$. The first term is the Mordell–Weil group tensored with $\mathbb{Q}/\mathbb{Z}$. In Introduction, we called the last term the Tate–Shafarevich scheme. This terminology is justified by (4). Along this line, the middle term might be called the Selmer scheme.

Remark 3.4.2. As we saw after Lemma 3.3.4, if the finite étale $p$-groups $\mathbf{H}^{1}(X,{\mathcal{A}})_{0\cap \text{div}}$ and $\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{0\cap \text{div}}$ (intersection of connected part and divisible part) are zero, then the morphisms in (6f) and (6g) are simplified to more symmetric expressions

$$\begin{eqnarray}\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}}^{\text{PD}}{\hookrightarrow}T\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}\qquad \text{and}\qquad \mathbf{H}^{1}(X,{\mathcal{A}})_{0}^{\text{SD}^{\prime }}{\twoheadrightarrow}\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{0}.\end{eqnarray}$$

It is not clear whether these vanishing conditions are always satisfied or not. To see cases where the conditions are indeed satisfied, let $A$ be the Jacobian of a proper smooth geometrically connected curve $C$ over $K$ with a $K$-rational point (in particular, $A\cong A^{\vee }$) and ${\mathcal{C}}$ a proper flat regular model over $X$ of $C$. As we saw in the proof of Proposition 3.2.11, there exists a canonical isomorphism $H^{1}(X,{\mathcal{A}})\cong H^{2}({\mathcal{C}},\mathbf{G}_{m})$. Let $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})$ be the (pro-)étale sheafification of the presheaf $k^{\prime }\in k^{\text{ind}\text{rat}}\mapsto H^{2}({\mathcal{C}}_{k^{\prime }},\mathbf{G}_{m})$, which is locally of finite presentation. Then the above isomorphism extends to an isomorphism $\mathbf{H}^{1}(X,{\mathcal{A}})\cong \mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})$. Note that $\mathbf{H}^{1}(X,{\mathcal{A}}_{0})$ surjects onto $\mathbf{H}^{1}(X,{\mathcal{A}})$ with finite kernel by Proposition 3.2.4.

If ${\mathcal{C}}$ is an Artin supersingular K3 surface (hence an elliptic surface fibered over $X$), then [Reference ArtinArt74, (4.2), (4.4)] shows that ${\mathcal{C}}$ is Shioda supersingular (i.e., satisfies the Tate conjecture) and $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})\cong \mathbf{G}_{a}$. In particular, $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{0\cap \text{div}}=0$. If ${\mathcal{C}}$ is an Artin nonsupersingular K3 surface, then [Reference Milne and RamachandranMR15, Proposition 4.7, Lemma 2.1] applied to the de Rham–Witt complex of ${\mathcal{C}}$ with $r=1$, $j=3$ together with [Reference IllusieIll79, II, (5.7.6), Section 7.2(a)] (or [Reference YuiYui86, Proposition (4.4)]) shows that $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})$ is étale. Hence again $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{0\cap \text{div}}=0$. In this case, [Reference IllusieIll79, II, Proposition 5.2] shows that the dimension of $V_{p}\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})$ over $\mathbb{Q}_{p}$ is $22-2h-\unicode[STIX]{x1D70C}$, where $22$ is the second Betti number of ${\mathcal{C}}$, $h$ the height of the formal Brauer group of ${\mathcal{C}}$ and $\unicode[STIX]{x1D70C}$ the geometric Picard number of ${\mathcal{C}}$. As soon as the inequality $\unicode[STIX]{x1D70C}\leqslant 22-2h$ is strict, the group $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{\text{div}}$ is nonzero.

Similar arguments apply to the case where ${\mathcal{C}}$ is an abelian surface, showing that either $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{0}=0$ (nonsupersingular case) or $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{\text{div}}=0$ (supersingular case). Similarly, we know that $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{\text{div}}=0$ whenever the geometric Picard number of ${\mathcal{C}}$ is equal to the second Betti number of ${\mathcal{C}}$. This includes the cases of rational surfaces, ruled surface, Enriques surfaces ([Reference Bombieri and MumfordBM76, Theorem 4]; we drop the condition that $C$ has a $K$-rational point) and quasi-elliptic surfaces ([Reference Bombieri and MumfordBM76, Proposition 12]; we drop the condition that $C$ is smooth over $K$). If ${\mathcal{C}}$ is the product of $X$ with another curve $Y$ such that $X$ is a supersingular elliptic curve and $Y$ is a genus two curve whose Jacobian is absolutely simple with $p$-rank one (see [Reference O’Connor, McGuire, Naehrig and StrengHOMNS11, Section 8 Examples; Corollary 3] for the existence of such a curve; again we drop the condition that $C$ has a $K$-rational point), then a calculation similar to the above shows that $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{0}\cong \mathbf{G}_{a}$ and $T_{p}\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{\text{div}}$ has rank four. It is not clear whether $\mathbf{H}^{2}({\mathcal{C}},\mathbf{G}_{m})_{0\cap \text{div}}$ is trivial or not in this case.

If there is a case where $\mathbf{H}^{1}(X,{\mathcal{A}})_{0\cap \text{div}}$ is nonzero, it might mean that supersingularity and nonsupersingularity are somehow “mixed up” in ${\mathcal{A}}$ (or ${\mathcal{C}}$) in an interesting way. Perhaps it could be the Tate conjecture that does not allow such a mixing phenomenon.

4 Some more results

Let $X$ and $A$ be as in the beginning of the previous section.

4.1 Duality for Néron models over open curves

In this subsection, we set

$$\begin{eqnarray}V=V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}=(V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})^{\text{PD}},\end{eqnarray}$$

the last isomorphism coming from Theorem 3.4.1 (6). Hence (5) of the same theorem can be written as the canonical distinguished triangle

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}}\rightarrow V.\end{eqnarray}$$

Let $U\subset X$ be a dense open subscheme. We have a pairing

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,\mathbf{G}_{m})[1]\rightarrow \mathbb{Z},\end{eqnarray}$$

where the last morphism is the trace morphism (3.1.1). This induces a morphism

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})^{\text{SD}}.\end{eqnarray}$$

Proposition 4.1.1. We have a canonical morphism between canonical distinguished triangles

The left vertical morphism is an isomorphism.

Proof. This follows from the same method as the proof of Proposition 3.2.1. ◻

Proposition 4.1.2. The morphisms

$$\begin{eqnarray}\displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Q}/\mathbb{Z} & \rightarrow & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\otimes ^{L}\mathbb{Q}/\mathbb{Z}\nonumber\\ \displaystyle & \rightarrow & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})^{\text{SD}}\otimes ^{L}\mathbb{Q}/\mathbb{Z}\nonumber\end{eqnarray}$$

are isomorphisms.

Proof. We show that the first morphism is an isomorphism. We have

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Q}/\mathbb{Z}\overset{{\sim}}{\rightarrow }R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\otimes ^{L}\mathbb{Q}/\mathbb{Z}\end{eqnarray}$$

by the same proof as Proposition 3.3.10. We know that $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0}^{\vee })\in D^{b}(\text{I}\text{Alg}_{\text{uc}})$ is Serre reflexive for any $x$ by Proposition 2.5.3 and [Reference SuzukiSuz14, Proposition (2.4.1)(b)]. Hence

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Q}/\mathbb{Z}\overset{{\sim}}{\rightarrow }R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\otimes ^{L}\mathbb{Q}/\mathbb{Z}.\end{eqnarray}$$

Therefore the previous proposition gives the result.

The composite morphism has already been shown to be an isomorphism essentially in the proof of Proposition 3.2.1. ◻

Proposition 4.1.3. We have

$$\begin{eqnarray}\displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\otimes \mathbb{Q}\overset{{\sim}}{\rightarrow }R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\otimes \mathbb{Q}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}}\otimes \mathbb{Q}\overset{{\sim}}{\rightarrow }R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})^{\text{SD}}\otimes \mathbb{Q}. & \displaystyle \nonumber\end{eqnarray}$$

They induce a canonical isomorphism between distinguished triangles

Proof. An object of $\text{I}\text{Alg}_{\text{uc}}/k$ is torsion since a unipotent quasi-algebraic group in positive characteristic is torsion. Hence Propositions 2.5.3 and 2.5.4 show that $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x}/k,{\mathcal{A}}_{0}^{\vee })$ and $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x}/k,{\mathcal{A}})^{\text{SD}}$ for any $x$ are killed after tensored with $\mathbb{Q}$. This implies the result.◻

From this, we have canonical morphisms

$$\begin{eqnarray}\displaystyle & & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})^{\text{SD}}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})^{\text{SD}}\otimes \mathbb{Q}\rightarrow V\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}[1]\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}[1].\nonumber\end{eqnarray}$$

Here is the duality for ${\mathcal{A}}$ over $U$:

Proposition 4.1.4. Consider the triangle

$$\begin{eqnarray}V[-1]\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})^{\text{SD}}\rightarrow V\end{eqnarray}$$

coming from the above morphisms. This triangle is distinguished. The diagram

is a morphism of distinguished triangles.

Proof. Choose a mapping cone distinguished triangle

(4.1.1)$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})^{\text{SD}}\rightarrow E.\end{eqnarray}$$

Applying $(\;\cdot \;)\otimes \mathbb{Q}$ and $(\;\cdot \;)\otimes ^{L}\mathbb{Q}/\mathbb{Z}$, we get a commutative diagram of distinguished triangles

We have $E\otimes ^{L}\mathbb{Q}/\mathbb{Z}=0$ by Proposition 4.1.2. Hence $E\cong E\otimes \mathbb{Q}$ and $E$ is uniquely divisible. By an axiom of triangulated categories, the distinguished triangle in the middle row is isomorphic to the distinguished triangle

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(U,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\otimes \mathbb{Q}\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,{\mathcal{A}})^{\text{SD}}\otimes \mathbb{Q}\rightarrow V\end{eqnarray}$$

in the previous proposition. Therefore the distinguished triangle (4.1.1) can be chosen so that $E=V$ and the diagram

is a morphism of distinguished triangles. By an axiom of triangulated categories, the right half of the commutative diagram in Proposition 4.1.1 can be extended to a morphism of distinguished triangles

Tensoring $\mathbb{Q}$ has to recover the isomorphism of distinguished triangles in the previous proposition. Hence the right vertical morphism is the identity. Thus we have a morphism of distinguished triangles

With these diagrams, the result follows. ◻

4.2 Link to the finite base field case

In this section, we assume that the base field $k$ is the finite field $\mathbb{F}_{q}$ of $q$ elements. The main reference about the finite base field case of duality for abelian varieties is [Reference MilneMil06, II, Section 5; III, Section 9]. As in [Reference SuzukiSuz14, Section 10], let $k_{\text{pro}\text{zar}}$ be the full subcategory of the category of $k$-algebras consisting of filtered unions of finite products of copies of $k$. For $k^{\prime }\in k_{\text{pro}\text{zar}}$, we say that a finite family $\{k_{i}^{\prime }\}$ of $k^{\prime }$-algebras is a covering if $\prod k_{i}^{\prime }$ is faithfully flat over $k^{\prime }$. This defines a topology on $k_{\text{pro}\text{zar}}$. We denote the resulting site by $\operatorname{Spec}k_{\text{pro}\text{zar}}$. The identity functor defines a morphism of sites $\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}\rightarrow \operatorname{Spec}k_{\text{pro}\text{zar}}$. By abuse of notation, we denote its pushforward functor $D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{\text{pro}\text{zar}})$ by $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},\;\cdot \;)$, with cohomologies $H^{n}(k_{\text{pro}\text{et}},\;\cdot \;)$ and $H^{0}=\unicode[STIX]{x1D6E4}$. If $C\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is locally of finite presentation, then $H^{n}(k_{\text{pro}\text{et}},C)$ is a constant sheaf for any $n$.

We will first relate $R\,\mathbf{Hom}_{k}(\;\cdot \;,\mathbb{Q}/\mathbb{Z})$ ($=R\,\mathbf{Hom}_{k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}}(\;\cdot \;,\mathbb{Q}/\mathbb{Z})$) to $R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(\;\cdot \;,\mathbb{Q}/\mathbb{Z})$. For any $C\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$, we have morphisms

$$\begin{eqnarray}\displaystyle & & \displaystyle R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},R\,\mathbf{Hom}_{k}(C,\mathbb{Q}/\mathbb{Z}))\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},C),R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},\mathbb{Q}/\mathbb{Z}))\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},C),\mathbb{Q}/\mathbb{Z})[-1],\nonumber\end{eqnarray}$$

where the last morphism comes from $H^{1}(k,\mathbb{Q}/\mathbb{Z})=\mathbb{Q}/\mathbb{Z}$ and $H^{n}(k,\mathbb{Q}/\mathbb{Z})=0$ for $n\geqslant 2$.

Proposition 4.2.1. If $C\in \text{I}\text{P}\text{Alg}/k$, then the above morphism

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},R\,\mathbf{Hom}_{k}(C,\mathbb{Q}/\mathbb{Z}))\rightarrow R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},C),\mathbb{Q}/\mathbb{Z})[-1]\end{eqnarray}$$

is an isomorphism.

Proof. The same proof as [Reference SuzukiSuz14, Proposition (10.4)] works. ◻

We denote $(\;\cdot \;)^{\text{PD}}=R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(\;\cdot \;,\mathbb{Q}/\mathbb{Z})$. If $C\in \operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ is an extension of a torsion (constant) group by a profinite group, then $C^{\text{PD}}=\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(C,\mathbb{Q}/\mathbb{Z})$ is the usual Pontryagin dual. It follows that if $C=(\prod \mathbb{Z}/l\mathbb{Z})/(\bigoplus \mathbb{Z}/l\mathbb{Z})$, where $l$ ranges over all primes, then $C^{\text{PD}}\cong C[-1]$.

By abuse of notation, we denote the composite of

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\;\cdot \;):D(X_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

and

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},\;\cdot \;):D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{\text{pro}\text{zar}})\end{eqnarray}$$

by $R\unicode[STIX]{x1D6E4}(X,\;\cdot \;)$, with cohomologies $H^{n}(X,\;\cdot \;)$ and $H^{0}=\unicode[STIX]{x1D6E4}$. If $F\in \operatorname{Ab}(X_{\text{fppf}})$ is locally of finite presentation, then $H^{n}(X,F)$ is a constant sheaf for any $n$.

Now let $F$ be ${\mathcal{A}}$, the Néron model of an abelian variety $A$ as before. We will deduce the duality for $R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})$ from Proposition 3.3.2. The group $\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})$ is the Mordell–Weil group of $A$, which is finitely generated. We have an exact sequence

$$\begin{eqnarray}0\rightarrow \unicode[STIX]{x0428}(A/K)\rightarrow H^{1}(X,{\mathcal{A}})\rightarrow \bigoplus _{x}H^{1}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))\end{eqnarray}$$

by [Reference MilneMil06, III, Proposition 9.2]. In particular, $H^{1}(X,{\mathcal{A}})$ is a torsion group of cofinite type,

$$\begin{eqnarray}H^{1}(X,{\mathcal{A}})_{\text{div}}=\unicode[STIX]{x0428}(A/K)_{\text{div}},\end{eqnarray}$$

and the sequence

$$\begin{eqnarray}0\rightarrow \unicode[STIX]{x0428}(A/K)_{/\text{div}}\rightarrow H^{1}(X,{\mathcal{A}})_{/\text{div}}\rightarrow \bigoplus _{x}H^{1}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))\end{eqnarray}$$

is exact. It seems that the finiteness of $\unicode[STIX]{x0428}(A/K)_{/\text{div}}$ is not known unconditionally in general. Hence we cannot assume that $H^{1}(X,{\mathcal{A}})$ is divisibly ML. By [Reference MilneMil06, II, Proposition 5.1(a); III, Lemma 7.10(c)], the group $H^{2}(X,{\mathcal{A}})$ is torsion and $H^{n}(X,{\mathcal{A}})=0$ for $n\geqslant 3$.

Proposition 4.2.2. Applying $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},\;\cdot \;)$ to the second isomorphism in Proposition 3.3.2 gives an isomorphism

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }(R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\overset{{\sim}}{\rightarrow }(R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Q}/\mathbb{Z})^{\text{PD}}[-1].\end{eqnarray}$$

Proof. This follows from the previous proposition. ◻

The morphism in this proposition can alternatively be defined by the morphisms

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow R\unicode[STIX]{x1D6E4}(X,\mathbf{G}_{m})[1]\rightarrow H^{3}(X,\mathbf{G}_{m})[-2]\cong \mathbb{Q}/\mathbb{Z}[-2].\end{eqnarray}$$

The proposition says that this is a perfect pairing up to uniquely divisibles. This more or less recovers [Reference MilneMil06, III, Theorem 9.4]. We will state an integral version using Weil-étale cohomology at the end of this subsection.

We will express the conjectural finiteness of $\unicode[STIX]{x0428}(A/K)$ in terms of $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})$. Let $TH^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$ be the profinite Tate module of $H^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$. Let $VH^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}$ be $(TH^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}})\otimes \mathbb{Q}$. We need the following variant of [Reference MilneMil06, III, Theorem 9.4(b)] (cf. [Reference Kato and TrihanKT03, Proposition 2.4(3)]), which we deduce from the above proposition.

Proposition 4.2.3. In $H^{0}$, the isomorphism in Proposition 4.2.2 induces an exact sequence

$$\begin{eqnarray}0\rightarrow \unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })^{\wedge }\rightarrow H^{2}(X,{\mathcal{A}})^{\text{PD}}\rightarrow TH^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}\rightarrow 0.\end{eqnarray}$$

In particular, $H^{2}(X,{\mathcal{A}})$ is divisibly ML and of cofinite type.

Proof. Let $C$ be the isomorphic object in Proposition 4.2.2. We have a distinguished triangle

$$\begin{eqnarray}\displaystyle & & \displaystyle R\,\underset{n}{\varprojlim }\Bigl((\unicode[STIX]{x1D70F}_{{\leqslant}1}R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\Bigr)\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,C\rightarrow R\,\underset{n}{\varprojlim }(H^{2}(X,{\mathcal{A}}_{0}^{\vee })\otimes ^{L}\mathbb{Z}/n\mathbb{Z})[-2].\nonumber\end{eqnarray}$$

The third term is concentrated in degrees ${\geqslant}1$. Hence the first and second terms have the same $H^{0}$. Since $\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})$ is finitely generated and $H^{1}(X,{\mathcal{A}})$ is torsion of cofinite type, we can apply Proposition 2.3.8 (or its version for $\operatorname{Spec}k_{\text{pro}\text{zar}}$) to obtain an exact sequence

$$\begin{eqnarray}0\rightarrow \unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })^{\wedge }\rightarrow H^{0}C\rightarrow TH^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}}\rightarrow 0.\end{eqnarray}$$

On the other hand, the cohomology objects of $R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Q}/\mathbb{Z}$ are torsion (constant) groups, with $H^{1}$ given by $H^{2}(X,{\mathcal{A}})$. Hence $H^{0}C=H^{2}(X,{\mathcal{A}})^{\text{PD}}$. This gives the stated exact sequence. Since $\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})$ is finitely generated and $H^{1}(X,{\mathcal{A}})$ is of cofinite type, it follows that $H^{2}(X,{\mathcal{A}})$ is divisibly ML and of cofinite type.◻

We will describe $H^{n}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})$ for each $n$ in terms of $H^{1}(X,{\mathcal{A}})$. Note that for a finite étale group $C$ over $k$, the object $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},C)$ is the mapping fiber of the endomorphism $F-1$ on $C$, where $F$ is the $q$th power Frobenius morphism. Taking limits, we know that the same is true for an ind-profinite-étale group $C$. Since $\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}}$ is a torsion étale group of cofinite type by Theorem 3.4.1 (2), we know that $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})$ is the mapping fiber of $F-1$ on $V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}}$.

Proposition 4.2.4. The natural morphism

(4.2.1)$$\begin{eqnarray}VH^{1}(X,{\mathcal{A}})_{\text{div}}\rightarrow \unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})\end{eqnarray}$$

is an isomorphism. There exists a canonical exact sequence

(4.2.2)$$\begin{eqnarray}0\rightarrow \frac{H^{1}(X,{\mathcal{A}})_{/\text{div}}^{\wedge }}{H^{1}(X,{\mathcal{A}})_{/\text{div}}}\rightarrow H^{1}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})\rightarrow \!(VH^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}})^{\text{PD}}\rightarrow 0.\end{eqnarray}$$

A priori, $H^{1}(X,{\mathcal{A}})_{/\text{div}}$ might contain a subgroup isomorphic to $\bigoplus _{l}\mathbb{Z}/l\mathbb{Z}$ for example. In this case, the term $H^{1}(X,{\mathcal{A}})_{/\text{div}}^{\wedge }/H^{1}(X,{\mathcal{A}})_{/\text{div}}$ above would contain $(\prod \mathbb{Z}/l\mathbb{Z})/(\bigoplus \mathbb{Z}/l\mathbb{Z})$.

Proof. We have

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\,\hat{\otimes }\,\mathbb{A}^{\infty }=R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\,\hat{\otimes }\,\mathbb{A}^{\infty }).\end{eqnarray}$$

Denote these isomorphic objects by $C$. We first calculate $H^{n}C$ through the left-hand side. Recall that $\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})$ is finitely generated, $H^{1}(X,{\mathcal{A}})$ is torsion of cofinite type, and $H^{2}(X,{\mathcal{A}})$ is torsion, divisibly ML and of cofinite type (Proposition 4.2.3). Hence Proposition 2.3.8 gives exact sequences

(4.2.3)$$\begin{eqnarray}\begin{array}{@{}c@{}}\displaystyle 0\rightarrow \unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes \mathbb{A}^{\infty }\rightarrow H^{0}C\rightarrow VH^{1}(X,{\mathcal{A}})_{\text{div}}\rightarrow 0,\\ \displaystyle 0\rightarrow H^{1}(X,{\mathcal{A}})^{\wedge }\otimes \mathbb{Q}\rightarrow H^{1}C\rightarrow VH^{2}(X,{\mathcal{A}})_{\text{div}}\rightarrow 0\end{array}\end{eqnarray}$$

and $H^{n}C=0$ for $n\neq 0,1$. We have

$$\begin{eqnarray}H^{1}(X,{\mathcal{A}})^{\wedge }\otimes \mathbb{Q}=\frac{H^{1}(X,{\mathcal{A}})_{/\text{div}}^{\wedge }}{H^{1}(X,{\mathcal{A}})_{/\text{div}}}.\end{eqnarray}$$

Next we calculate $H^{n}C$ through the right-hand side. By Proposition 3.3.6 and Theorem 3.4.1 (2), (6a), the $H^{-1}$ (resp. $H^{1}$) of $R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\,\hat{\otimes }\,\mathbb{A}^{\infty }$ is the rational Tate module (resp. the Pontryagin dual of the rational Tate module) of an abelian variety over $k$. For any abelian variety $B$ over $k$, we have $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},B)=B(k)$ by Lang’s theorem, and $B(k)$ is finite. Hence $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},VB)=R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},(VB)^{\text{PD}})=0$. Therefore

$$\begin{eqnarray}C=R\unicode[STIX]{x1D6E4}\Bigl(k_{\text{pro}\text{et}},H^{0}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\,\hat{\otimes }\,\mathbb{A}^{\infty })\Bigr).\end{eqnarray}$$

By Propositions 3.3.6 and 3.3.7, we have a split exact sequence

$$\begin{eqnarray}0\rightarrow \unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))\otimes \mathbb{A}^{\infty }\rightarrow H^{0}(R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})\,\hat{\otimes }\,\mathbb{A}^{\infty })\rightarrow V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}}\rightarrow 0\end{eqnarray}$$

in $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. Since $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})_{0}$ is the perfection of an abelian variety over $k$, the group $H^{n}(k_{\text{pro}\text{et}},\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})_{0})$ is zero for $n\geqslant 1$ by Lang’s theorem and finite for $n=0$. Therefore $H^{n}(k_{\text{pro}\text{et}},\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))$ is $\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})$ if $n=0$ and $H^{n}(k_{\text{pro}\text{et}},\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})))$ if $n\geqslant 1$, which is zero if $n\geqslant 2$. Since $\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))\in \text{FGEt}/k$, the Frobenius action on $\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))\otimes \mathbb{Q}$ and hence on $\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))\otimes \mathbb{A}^{\infty }$ are semisimple. Therefore its invariant part and coinvariant part agree. Hence

$$\begin{eqnarray}H^{n}(k_{\text{pro}\text{et}},\unicode[STIX]{x1D70B}_{0}(\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))\otimes \mathbb{A}^{\infty })=\left\{\begin{array}{@{}ll@{}}\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes \mathbb{A}^{\infty }\quad & \text{if }n=0,1,\\ 0\quad & \text{otherwise}.\end{array}\right.\end{eqnarray}$$

Therefore we have split exact sequences

(4.2.4)$$\begin{eqnarray}\begin{array}{@{}c@{}}0\rightarrow \unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes \mathbb{A}^{\infty }\rightarrow H^{0}C\rightarrow \unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})\rightarrow 0,\\ 0\rightarrow \unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes \mathbb{A}^{\infty }\rightarrow H^{1}C\rightarrow H^{1}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})\rightarrow 0,\end{array}\end{eqnarray}$$

and $H^{n}C=0$ for $n\neq 0,1$.

Comparing the first exact sequences in (4.2.3) and (4.2.4), we know that (4.2.1) is an isomorphism. By Proposition 4.2.3, we have an exact sequence

$$\begin{eqnarray}0\rightarrow (VH^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}})^{\text{PD}}\rightarrow VH^{2}(X,{\mathcal{A}})_{\text{div}}\rightarrow (\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\otimes \mathbb{A}^{\infty })^{\text{PD}}\rightarrow 0.\end{eqnarray}$$

With this and the second exact sequence in (4.2.3), we obtain a surjection from $H^{1}C$ onto $(\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\otimes \mathbb{A}^{\infty })^{\text{PD}}$. Let $D$ be its kernel. Consider the diagram with exact rows

The composite map from $\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes \mathbb{A}^{\infty }$ to $(\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\otimes \mathbb{A}^{\infty })^{\text{PD}}$ is given by the height pairing tensored with $\mathbb{A}^{\infty }$, which is an isomorphism. Therefore the composite map from $D$ to $H^{1}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})$ is an isomorphism. This gives (4.2.2) since $D$ is an extension of $(VH^{1}(X,{\mathcal{A}}_{0}^{\vee })_{\text{div}})^{\text{PD}}$ by $H^{1}(X,{\mathcal{A}})_{/\text{div}}^{\wedge }/H^{1}(X,{\mathcal{A}})_{/\text{div}}$.◻

Proposition 4.2.5. The following are equivalent:

  • $\unicode[STIX]{x0428}(A/K)$ is finite.

  • $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})_{\text{div}})$ is zero.

Proof. The finiteness of $\unicode[STIX]{x0428}(A/K)$ is equivalent to the finiteness of $H^{1}(X,{\mathcal{A}})$, which itself is equivalent to

$$\begin{eqnarray}VH^{1}(X,{\mathcal{A}})_{\text{div}}=\frac{H^{1}(X,{\mathcal{A}})_{/\text{div}}^{\wedge }}{H^{1}(X,{\mathcal{A}})_{/\text{div}}}=0.\end{eqnarray}$$

It is also equivalent to the finiteness of $\unicode[STIX]{x0428}(A^{\vee }/K)$ by [Reference MilneMil06, I, Remark 6.14(c)]. Therefore the stated equivalence follows from the previous proposition.◻

Note that if $\unicode[STIX]{x0428}(A/K)$ is finite, then the full BSD conjecture for $A$ is true by the result of Kato–Trihan [Reference Kato and TrihanKT03].

Recall from [Reference Kato and TrihanKT03, Section 2.2] that a complex, called the arithmetic cohomology and denoted by $R\unicode[STIX]{x1D6E4}_{\mathit{ar}}$, is defined to be the mapping fiber of the morphism

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(U,{\mathcal{A}}_{\text{tor}})\oplus \bigoplus _{x\not \in U}(\unicode[STIX]{x1D6E4}(\hat{K}_{x},A)\otimes ^{L}\mathbb{Q}/\mathbb{Z})[-1]\rightarrow \bigoplus _{x\not \in U}R\unicode[STIX]{x1D6E4}(\hat{K}_{x},A_{\text{tor}}),\end{eqnarray}$$

where $U$ is an open subscheme of $X$ over which $A$ has good reduction. Here we understand the functor $R\unicode[STIX]{x1D6E4}(\hat{K}_{x},\;\cdot \;)$ as the composite of $R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\;\cdot \;)$ and $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{zar}},\;\cdot \;)$, and $\unicode[STIX]{x1D6E4}=H^{0}R\unicode[STIX]{x1D6E4}$. To be clear, we denote this complex by $R\unicode[STIX]{x1D6E4}_{\mathit{ar},A}$ and its cohomology objects by $H_{\mathit{ar},A}^{n}$. This mapping fiber is defined on the level of complexes, briefly explained in [Reference Kato and TrihanKT03, Section 2.1]. As they wrote, “how to use complexes is evident and so we do not explain it.” We follow the same style in the rest of this subsection.

We similarly have the mapping cone of the morphism

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Q}/\mathbb{Z}\rightarrow \left(\mathop{\prod }_{x\in X}\unicode[STIX]{x1D70F}_{{\geqslant}1}R\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})\right)\otimes ^{L}\mathbb{Q}/\mathbb{Z}.\end{eqnarray}$$

We have

$$\begin{eqnarray}\unicode[STIX]{x1D70F}_{{\geqslant}1}R\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})=H^{1}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))[-1].\end{eqnarray}$$

Hence we have the mapping cone

$$\begin{eqnarray}\left[R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Q}/\mathbb{Z}\rightarrow \bigoplus _{x}H^{1}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))\right].\end{eqnarray}$$

Also we have the mapping cone of the morphism

$$\begin{eqnarray}\left(\bigoplus _{x\in X}\unicode[STIX]{x1D70F}_{{\leqslant}1}R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0})\right)\otimes ^{L}\mathbb{Q}/\mathbb{Z}\rightarrow R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0})\otimes ^{L}\mathbb{Q}/\mathbb{Z}.\end{eqnarray}$$

We have

$$\begin{eqnarray}\unicode[STIX]{x1D70F}_{{\leqslant}1}R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0})=\unicode[STIX]{x1D6E4}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))[-1].\end{eqnarray}$$

Hence we have the mapping cone

$$\begin{eqnarray}\left[\bigoplus _{x}\unicode[STIX]{x1D6E4}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))\rightarrow R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0})\otimes ^{L}\mathbb{Q}/\mathbb{Z}\right].\end{eqnarray}$$

Proposition 4.2.6. There exist canonical isomorphisms

$$\begin{eqnarray}\displaystyle R\unicode[STIX]{x1D6E4}_{\mathit{ar},A} & \cong & \displaystyle \left[R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\otimes ^{L}\mathbb{Q}/\mathbb{Z}\rightarrow \bigoplus _{x}H^{1}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))\right][-2]\nonumber\\ \displaystyle & \cong & \displaystyle \left[\bigoplus _{x}\unicode[STIX]{x1D6E4}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))\rightarrow R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0})\otimes ^{L}\mathbb{Q}/\mathbb{Z}\right][-1].\nonumber\end{eqnarray}$$

Proof. We first show the first isomorphism. The mapping cone of

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}(\hat{K}_{x},A)\otimes ^{L}\mathbb{Q}/\mathbb{Z}[-1]\rightarrow R\unicode[STIX]{x1D6E4}(\hat{K}_{x},A_{\text{tor}})\end{eqnarray}$$

is $H^{1}(K_{x},A)\otimes ^{L}\mathbb{Q}/\mathbb{Z}[-2]$. Hence

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}_{\mathit{ar},A}=\left[R\unicode[STIX]{x1D6E4}(U,{\mathcal{A}})\otimes ^{L}\mathbb{Q}/\mathbb{Z}\rightarrow \bigoplus _{x\not \in U}H^{1}(K_{x},A)\otimes ^{L}\mathbb{Q}/\mathbb{Z}[-1]\right][-2]\end{eqnarray}$$

The morphism

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow R\unicode[STIX]{x1D6E4}(U,{\mathcal{A}})\end{eqnarray}$$

and the morphism

$$\begin{eqnarray}\bigoplus _{x\not \in U}H^{1}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})[-1]\rightarrow \bigoplus _{x\not \in U}H^{1}(\hat{K}_{x},A)[-1]\end{eqnarray}$$

have the same mapping cone $\bigoplus _{x\not \in U}R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})[1]$. This gives the desired result.

For the second isomorphism, use the distinguished triangle

$$\begin{eqnarray}\hspace{64.79993pt}R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0})\rightarrow R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow \bigoplus _{x}R\unicode[STIX]{x1D6E4}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})).\hspace{38.39996pt}\square\end{eqnarray}$$

Proposition 4.2.7. Consider the diagram

where the horizontal morphisms come from the morphisms in the previous proposition (with $R\,\mathop{\varprojlim }\nolimits_{n}(\;\cdot \;\otimes ^{L}\mathbb{Z}/n\mathbb{Z})$ applied), the left vertical one from Grothendieck’s pairing and the right vertical one from Proposition 4.2.2. This diagram is commutative and canonically induces an isomorphism

$$\begin{eqnarray}R\,\underset{n}{\varprojlim }(R\unicode[STIX]{x1D6E4}_{\mathit{ar},A^{\vee }}\otimes ^{L}\mathbb{Z}/n\mathbb{Z})\overset{{\sim}}{\rightarrow }(R\unicode[STIX]{x1D6E4}_{\mathit{ar},A})^{\text{PD}}[-2]\end{eqnarray}$$

on the mapping cones.

Proof. Consider the morphism

$$\begin{eqnarray}\bigoplus _{x}\unicode[STIX]{x1D70F}_{{\leqslant}1}R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0}^{\vee })\rightarrow R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee }).\end{eqnarray}$$

We have a natural morphism from this morphism to the morphism

$$\begin{eqnarray}\bigoplus _{x}\unicode[STIX]{x1D70F}_{{\leqslant}1}R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}({\mathcal{A}}^{\vee },\mathbf{G}_{m}))[1]\rightarrow R\unicode[STIX]{x1D6E4}(X,R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}({\mathcal{A}}^{\vee },\mathbf{G}_{m}))[1]\end{eqnarray}$$

(i.e., have a commutative square whose upper and lower sides are these morphisms). Using the functoriality/cup product morphisms similar to (2.5.4) and (2.7.2) and the morphism $R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},\mathbf{G}_{m})\rightarrow R\unicode[STIX]{x1D6E4}(X,\mathbf{G}_{m})$ we have a morphism from this morphism to the morphism

$$\begin{eqnarray}\displaystyle & & \displaystyle \bigoplus _{x}R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(R\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}),\unicode[STIX]{x1D70F}_{{\geqslant}3}R\unicode[STIX]{x1D6E4}(X,\mathbf{G}_{m}))[1]\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}),\unicode[STIX]{x1D70F}_{{\geqslant}3}R\unicode[STIX]{x1D6E4}(X,\mathbf{G}_{m}))[1].\nonumber\end{eqnarray}$$

We have $\unicode[STIX]{x1D70F}_{{\geqslant}3}R\unicode[STIX]{x1D6E4}(X,\mathbf{G}_{m})=\mathbb{Q}/\mathbb{Z}[-3]$. Hence the mapping cone of this morphism can be written as

$$\begin{eqnarray}R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}\Bigl([R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow \mathop{\prod }_{x}\unicode[STIX]{x1D70F}_{{\geqslant}1}R\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})],\mathbb{Q}/\mathbb{Z}\Bigr)[-1].\end{eqnarray}$$

These are functorial on the level of complexes. Hence we have a morphism

$$\begin{eqnarray}\displaystyle & & \displaystyle \Bigl[\bigoplus _{x}\unicode[STIX]{x1D70F}_{{\leqslant}1}R\unicode[STIX]{x1D6E4}_{x}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}}_{0}^{\vee })\rightarrow R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}}_{0}^{\vee })\Bigr]\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}\Bigl([R\unicode[STIX]{x1D6E4}(X,{\mathcal{A}})\rightarrow \mathop{\prod }_{x}\unicode[STIX]{x1D70F}_{{\geqslant}1}R\unicode[STIX]{x1D6E4}(\hat{{\mathcal{O}}}_{x},{\mathcal{A}})],\mathbb{Q}/\mathbb{Z}\Bigr)[-1].\nonumber\end{eqnarray}$$

Applying $R\,\mathop{\varprojlim }\nolimits_{n}(\;\cdot \;\otimes ^{L}\mathbb{Z}/n\mathbb{Z})$, we get the result.◻

This form of duality on $R\unicode[STIX]{x1D6E4}_{\mathit{ar},A}$ is also given in [Reference Trihan and VauclairTV17, Corollary 4.11, Remark 4.13].

For a torsion group $C$ of cofinite type, we call $(C_{/\text{div}})^{\text{PD}}$ the essentially torsion part of the profinite group $C^{\text{PD}}$, which is the product over all primes $l$ of the torsion part of the $l$-adic component of $C^{\text{PD}}$. For example, $\prod _{l}\mathbb{Z}/l\mathbb{Z}$ is the essentially torsion part of itself. By [Reference Kato and TrihanKT03, Proposition 2.4], we know that $H_{\mathit{ar},A}^{1}$ is isomorphic to the Selmer group of $A$. Hence the isomorphism in the above proposition in degree $1$ on the essentially torsion parts gives a perfect pairing

$$\begin{eqnarray}\unicode[STIX]{x0428}(A^{\vee }/K)_{/\text{div}}^{\wedge }\times \unicode[STIX]{x0428}(A/K)_{/\text{div}}\rightarrow \mathbb{Q}/\mathbb{Z}\end{eqnarray}$$

between a profinite group and a torsion group. For any prime $l$, this pairing on $l$-primary parts is a pairing between finite $l$-groups.

Proposition 4.2.8. This pairing agrees with the Cassels–Tate pairing.

Proof. The isomorphism in the proposition gives a Pontryagin duality between the exact sequences

$$\begin{eqnarray}\bigoplus _{x}\unicode[STIX]{x1D6E4}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}^{\vee }))\rightarrow H^{1}(X,{\mathcal{A}}_{0}^{\vee })_{/\text{div}}\rightarrow (H_{\mathit{ar},A^{\vee }}^{1})_{/\text{div}}\rightarrow 0\end{eqnarray}$$

and

$$\begin{eqnarray}0\rightarrow (H_{\mathit{ar},A}^{1})_{/\text{div}}^{\wedge }\rightarrow H^{1}(X,{\mathcal{A}})_{/\text{div}}^{\wedge }\rightarrow \bigoplus _{x}H^{1}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x})).\end{eqnarray}$$

Hence the proof of [Reference MilneMil06, III, Corollary 9.5] shows that the paring induced on the third term of the first exact sequence and the first term of the second exact sequence is the Cassels–Tate pairing. ◻

We briefly recall the triangulated functor

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(k_{W},\;\cdot \;):D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{\text{pro}\text{zar}})\end{eqnarray}$$

defined in [Reference SuzukiSuz14, Section 10]. Let $F$ be the $q$th power Frobenius morphism for any $k$-algebra, which induces an action on any object of $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$. The functor $k^{\prime }\mapsto k^{\prime }\otimes _{k}\overline{k}$ defines a premorphism of sites $f:\operatorname{Spec}k_{\text{pro}\text{et}}^{\text{ind}\text{rat}}\rightarrow \operatorname{Spec}k_{\text{pro}\text{zar}}$. Then

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(k_{W},C)=f_{\ast }[C\overset{F-1}{\rightarrow }C][-1]\end{eqnarray}$$

for $C\in D(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$.

We define

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(X_{W},\;\cdot \;)=R\unicode[STIX]{x1D6E4}(k_{W},R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,\;\cdot \;)):D(X_{\text{fppf}})\rightarrow D(k_{\text{pro}\text{zar}}).\end{eqnarray}$$

We denote $(\;\cdot \;)^{\text{LD}}=R\,\mathbf{Hom}_{k_{\text{pro}\text{zar}}}(\;\cdot \;,\mathbb{Z})$.

Proposition 4.2.9. We have

$$\begin{eqnarray}\displaystyle & \displaystyle R\unicode[STIX]{x1D6E4}(k_{W},R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}})=R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}}_{0}^{\vee }), & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\unicode[STIX]{x1D6E4}(k_{W},R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}})=R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}})^{\text{LD}}[-1], & \displaystyle \nonumber\\ \displaystyle & \displaystyle R\unicode[STIX]{x1D6E4}(k_{W},V\mathbf{H}^{1}(X,{\mathcal{A}}))=R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}})). & \displaystyle \nonumber\end{eqnarray}$$

Proof. For the first isomorphism, note that the mapping fiber of

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })^{\text{SD}\text{SD}}\end{eqnarray}$$

is concentrated in degree zero whose cohomology is $\mathop{\varprojlim }\nolimits_{n}\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })_{0}$. Since the endomorphism $F-1$ on the abelian variety $\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })_{0}$ is surjective with finite kernel by Lang’s theorem, we know that $\mathop{\varprojlim }\nolimits_{n}\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}_{0}^{\vee })_{0}$ is killed by applying $R\unicode[STIX]{x1D6E4}(k_{W},\;\cdot \;)$. This gives the desired result.

The second isomorphism follows from

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(k_{W},R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}})^{\text{SD}})=R\unicode[STIX]{x1D6E4}(k_{W},R\boldsymbol{\unicode[STIX]{x1D6E4}}(X,{\mathcal{A}}))^{\text{LD}}[-1],\end{eqnarray}$$

which is [Reference SuzukiSuz14, Proposition (10.4)]. For the third, we already saw in the proof of Proposition 4.2.5 that $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},V\mathbf{H}^{1}(X,{\mathcal{A}}))$ is the mapping fiber of $F-1$.◻

Here is a Weil-étale analogue of Proposition 4.2.2.

Proposition 4.2.10. Applying $R\unicode[STIX]{x1D6E4}(k_{W},\;\cdot \;)$ to the distinguished triangle in Theorem 3.4.1, we have a canonical distinguished triangle

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}}_{0}^{\vee })\rightarrow R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}})^{\text{LD}}[-1]\rightarrow R\unicode[STIX]{x1D6E4}(k_{W},V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })).\end{eqnarray}$$

The third term is zero if and only if $\unicode[STIX]{x0428}(A/K)$ is finite.

Proof. This follows from the previous proposition and Proposition 4.2.5. ◻

If $\unicode[STIX]{x0428}(A/K)$ is finite, then the cohomology objects of $R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}})$ and $R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}}_{0}^{\vee })$ are finitely generated abelian groups, and the above proposition gives a duality between these objects via $R\operatorname{Hom}(\;\cdot \;,\mathbb{Z})$.

We can also define a Weil-étale version $R\unicode[STIX]{x1D6E4}_{\mathit{ar},A,W}$ of $R\unicode[STIX]{x1D6E4}_{\mathit{ar},A}$ by

$$\begin{eqnarray}\displaystyle R\unicode[STIX]{x1D6E4}_{\mathit{ar},A,W} & = & \displaystyle \left[R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}})\rightarrow \bigoplus _{x}H^{1}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))[-1]\right][-1]\nonumber\\ \displaystyle & \cong & \displaystyle \left[\bigoplus _{x}\unicode[STIX]{x1D6E4}(k_{x},\unicode[STIX]{x1D70B}_{0}({\mathcal{A}}_{x}))[-1]\rightarrow R\unicode[STIX]{x1D6E4}(X_{W},{\mathcal{A}}_{0})\right]\nonumber\end{eqnarray}$$

and prove the existence of a canonical distinguished triangle

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}_{\mathit{ar},A^{\vee },W}\rightarrow R\unicode[STIX]{x1D6E4}_{\mathit{ar},A,W}^{\text{LD}}[-1]\rightarrow R\unicode[STIX]{x1D6E4}(k_{W},V\mathbf{H}^{1}(X,{\mathcal{A}}_{0}^{\vee })).\end{eqnarray}$$

Remark 4.2.11. We can also apply $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},\;\cdot \;)$ to Theorem 3.1.3 to recover the duality $R\unicode[STIX]{x1D6E4}(U,N)\leftrightarrow R\unicode[STIX]{x1D6E4}_{c}(U,N)$ for a finite flat group scheme $N$ over an open curve $U$ over $k=\mathbb{F}_{q}$ stated in [Reference MilneMil06, III, Theorem 8.2], including the topological group structures on the relevant cohomology groups. Define $R\unicode[STIX]{x1D6E4}_{c}(U,N)\in D(k_{\text{pro}\text{zar}})$ by $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,N))=R\unicode[STIX]{x1D6E4}(k_{W},R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,N))$. Since $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{c}(U,N)\in D^{b}(\text{P}\text{Alg}_{\text{uc}}/k)$ by Theorem 3.1.3, we know that $R\unicode[STIX]{x1D6E4}_{c}(U,N)\in D^{b}(\text{P}\text{Fin})$ by [Reference SuzukiSuz14, Proposition (10.3)(b)], where $\text{Fin}$ is the category of finite abelian groups. Therefore each cohomology object $H_{c}^{n}(U,N):=H^{n}R\unicode[STIX]{x1D6E4}_{c}(U,N)$ is a profinite group. Now applying $R\unicode[STIX]{x1D6E4}(k_{\text{pro}\text{et}},\;\cdot \;)$ to Theorem 3.1.3 and using Proposition 4.2.1, we obtain an isomorphism

$$\begin{eqnarray}R\unicode[STIX]{x1D6E4}(U,N)\overset{{\sim}}{\rightarrow }R\unicode[STIX]{x1D6E4}_{c}(U,N)^{\text{PD}}[-3]\end{eqnarray}$$

in $D(k_{\text{pro}\text{zar}})$. Hence we have a Pontryagin duality $H^{n}(U,N)\leftrightarrow H_{c}^{3-n}(U,N)$ for any $n$ between the torsion group and the profinite group, which recovers [Reference MilneMil06, III, Theorem 8.2].

A similar remark applies to Proposition 4.1.4 and [Reference MilneMil06, III, Theorem 9.4].

Acknowledgments

The author expresses his deep gratitude to Kazuya Kato. It was his suggestion to globalize the results of [Reference SuzukiSuz14] and study Tate–Shafarevich groups as sheaves. The author thanks Tetsuji Shioda for his encouragement to include explicit calculations, though the author’s calculations in Remark 3.4.2 are essentially no more than quotes of other researchers’ calculations. Also, an earlier version of this paper proved the finite-dimensionality of $\mathbf{H}^{1}(X,{\mathcal{A}})$ by reducing it to the semistable case, using Künnemann’s projective regular models [Reference KünnemannKün98] and relating $\mathbf{H}^{1}(X,{\mathcal{A}})$ to the Brauer group of a higher-dimensional variety. Michael Larsen’s suggestion to use Jacobians here made it possible to use surfaces only and reduce the length of the proof, which the author is grateful. Matthew Emerton’s idea (personal communication) to relate our duality with Brauer–Manin obstructions is regrettably not treated in this paper to save pages. This will be discussed elsewhere. The author is grateful to Kȩstutis Česnavičius for his information of the work [Reference Demarche and HarariDH18], to Cyril Demarche and David Harari for their encouragement to submit this paper independently, to Fabien Trihan for his information of the relation between Proposition 4.2.7 of this paper and [Reference Trihan and VauclairTV17, Corollary 4.11, Remark 4.13], to Thomas Geisser for discussions about finiteness of Tate–Shafarevich groups modulo divisible subgroups when $k$ is finite, and to the referees for their careful comments. This work was supported by JSPS KAKENHI Grant Number JP18J00415.

Appendix A. Comparison of local duality with and without relative sites

We continue the notation of Section 2.5. We recall the relative fppf site $\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}$ of $\hat{{\mathcal{O}}}_{x}$ over $k_{x}$ from [Reference SuzukiSuz13, Definition 2.3.2] and [Reference SuzukiSuz14, Section 3.3]. (Here, again, we abbreviate $(k_{x})^{\text{ind}\text{rat}}$ by $k_{x}^{\text{ind}\text{rat}}$ and $(\operatorname{Spec}k_{x})_{\text{et}}^{\text{ind}\text{rat}}$ by $\operatorname{Spec}k_{x,\text{et}}^{\text{ind}\text{rat}}$.) The category $\hat{{\mathcal{O}}}_{x}/k_{x}^{\text{ind}\text{rat}}$ is the category of pairs $(S,k_{S})$, where $k_{S}\in k_{x}^{\text{ind}\text{rat}}$ and $S$ a $\hat{\mathbf{O}}_{x}(k_{S})$-algebra. (In [Reference SuzukiSuz13, Section 2.3] and [Reference SuzukiSuz14, Section 3.3], $S$ is required to be finitely presented over $\hat{\mathbf{O}}_{x}(k^{\prime })$. The generalization to arbitrary algebras here does not change cohomology theory and Ext groups/sheaves as long as we treat sheaves representable by schemes locally of finite type over $\hat{{\mathcal{O}}}_{x}$.) A morphism $(S,k_{S})\rightarrow (S^{\prime },S_{S^{\prime }})$ consists of a $k_{x}$-algebra homomorphism $k_{S}\rightarrow k_{S^{\prime }}$ and an $\hat{\mathbf{O}}_{x}(k_{S})$-algebra homomorphism $S\rightarrow S^{\prime }$, with composition defined in the obvious way. A finite family of morphisms $\{(S,k_{S})\rightarrow (S_{i},k_{S_{i}})\}$ is called an fppf/étale covering if each $k_{S}\rightarrow k_{S_{i}}$ is étale and $S\rightarrow \prod _{i}S_{i}$ faithfully flat. This defines a site, which we denote by $\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}$. Its category of sheaves is denoted by $\operatorname{Ab}(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}})$. The categories $\operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}})$, $D(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}})$ and the functors $\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}}$, $R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}}$ are defined in a similar way as Section 2.5.

The functors

$$\begin{eqnarray}\displaystyle & \displaystyle k_{x}^{\text{ind}\text{rat}}\rightarrow \hat{{\mathcal{O}}}_{x}/k_{x}^{\text{ind}\text{rat}},\qquad k^{\prime }\mapsto (\hat{\mathbf{O}}_{x}(k^{\prime }),k^{\prime }), & \displaystyle \nonumber\\ \displaystyle & \displaystyle k_{x}^{\text{ind}\text{rat}}\rightarrow \hat{K}_{x}/k_{x}^{\text{ind}\text{rat}},\qquad k^{\prime }\mapsto (\hat{\mathbf{K}}_{x}(k^{\prime }),k^{\prime }), & \displaystyle \nonumber\\ \displaystyle & \displaystyle \hat{{\mathcal{O}}}_{x}/k_{x}^{\text{ind}\text{rat}}\rightarrow \hat{K}_{x}/k_{x}^{\text{ind}\text{rat}},\qquad (S,k_{S})\mapsto (S\otimes _{\hat{{\mathcal{O}}}_{x}}\hat{K}_{x},k_{S}) & \displaystyle \nonumber\end{eqnarray}$$

define morphisms of sites

$$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}^{\prime }:\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}\rightarrow \operatorname{Spec}k_{x,\text{et}}^{\text{ind}\text{rat}}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle \unicode[STIX]{x1D70B}_{\hat{K}_{x}}^{\prime }:\operatorname{Spec}\hat{K}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}\rightarrow \operatorname{Spec}k_{x,\text{et}}^{\text{ind}\text{rat}}, & \displaystyle \nonumber\\ \displaystyle & \displaystyle j^{\prime }:\operatorname{Spec}\hat{K}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}\rightarrow \operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}} & \displaystyle \nonumber\end{eqnarray}$$

such that $\unicode[STIX]{x1D70B}_{\hat{K}_{x}}^{\prime }=\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}^{\prime }\,\circ \,j^{\prime }$. We denote the composite of the derived pushforward

$$\begin{eqnarray}R(\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}^{\prime })_{\ast }:D(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{x,\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

and the pro-étale sheafification $D(k_{x,\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$ by

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(\hat{{\mathcal{O}}}_{x},\;\cdot \;):D(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}).\end{eqnarray}$$

Similarly, we denote the composite of the derived pushforward

$$\begin{eqnarray}R(\unicode[STIX]{x1D70B}_{\hat{K}_{x}}^{\prime })_{\ast }:D(\hat{K}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{x,\text{et}}^{\text{ind}\text{rat}})\end{eqnarray}$$

and the pro-étale sheafification $D(k_{x,\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}})$ by

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(\hat{K}_{x},\;\cdot \;):D(\hat{K}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}).\end{eqnarray}$$

In [Reference SuzukiSuz14, Section 3.3, paragraph before Proposition (3.3.8)], these functors were denoted by $R\tilde{\boldsymbol{\unicode[STIX]{x1D6E4}}}({\mathcal{O}}_{x},\;\cdot \;)$ and $R\tilde{\boldsymbol{\unicode[STIX]{x1D6E4}}}(K_{x},\;\cdot \;)$, respectively.

We compare the above morphisms of sites and the premorphisms of sites in Section 2.5. The functor sending $(S,k_{S})\in \hat{{\mathcal{O}}}_{x}/k^{\text{ind}\text{rat}}$ to $S$ defines a premorphism of sites

$$\begin{eqnarray}\bar{f}:\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}\rightarrow \operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}.\end{eqnarray}$$

Similarly, the functor sending $(S,k_{S})\in \hat{K}_{x}/k^{\text{ind}\text{rat}}$ to $S$ defines a premorphism of sites

$$\begin{eqnarray}f:\operatorname{Spec}\hat{K}_{x,\text{fppf}}\rightarrow \operatorname{Spec}\hat{K}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}.\end{eqnarray}$$

The pushforward functors $\bar{f}_{\ast }$ and $f_{\ast }$ are exact. They send the sheaf representable by a $\hat{{\mathcal{O}}}_{x}$-algebra (resp. $\hat{K}_{x}$-algebra) $S$ to the sheaf representable by $(S,k)$. Hence a scheme over $\hat{{\mathcal{O}}}_{x}$ (resp. $\hat{K}_{x}$) regarded as a sheaf on $\operatorname{Spec}\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}$ (resp. $\operatorname{Spec}\hat{K}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}$) mentioned in [Reference SuzukiSuz13, Section 2.3] is nothing but its image by $\bar{f}_{\ast }$ (resp. $f_{\ast }$).

Proposition A.1. We have $\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}=\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}^{\prime }\,\circ \,\bar{f}$ and $\unicode[STIX]{x1D70B}_{\hat{K}_{x}}=\unicode[STIX]{x1D70B}_{\hat{K}_{x}}^{\prime }\,\circ \,f$, and

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\;\cdot \;)=R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(\hat{{\mathcal{O}}}_{x},\bar{f}_{\ast }\;\cdot \;),\qquad R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},\;\cdot \;)=R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(\hat{{\mathcal{O}}}_{x},f_{\ast }\;\cdot \;).\end{eqnarray}$$

In particular, for a group scheme $A$ over $\hat{{\mathcal{O}}}_{x}$ or $\hat{K}_{x}$, we have

$$\begin{eqnarray}R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},A)=R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(\hat{{\mathcal{O}}}_{x},A),\qquad R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{K}_{x},A)=R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(\hat{{\mathcal{O}}}_{x},A).\end{eqnarray}$$

Proof. The composite $\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}^{\prime }\,\circ \,\bar{f}$ is defined by the composite of the functors $k_{x}^{\prime }\mapsto (\hat{\mathbf{O}}_{x}(k^{\prime }),k_{x}^{\prime })\mapsto \hat{\mathbf{O}}_{x}(k_{x}^{\prime })$ which defines $\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}$. Hence $\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}=\unicode[STIX]{x1D70B}_{\hat{{\mathcal{O}}}_{x}}^{\prime }\,\circ \,\bar{f}$. Similar for $\unicode[STIX]{x1D70B}_{\hat{K}_{x}}=\unicode[STIX]{x1D70B}_{\hat{K}_{x}}^{\prime }\,\circ \,f$. The equalities of the derived pushforwards follow from Propositions 2.4.2, 2.4.3 and the theorem on derived functors of composition [Reference Kashiwara and SchapiraKS06, Proposition 10.3.5(ii)].◻

Proposition A.2. The mapping fiber functor

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}^{\prime }({\mathcal{O}}_{x},\;\cdot \;):=[\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }({\mathcal{O}}_{x},\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(K_{x},{j^{\prime }}^{\ast }\;\cdot \;)][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle \operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}})\rightarrow \operatorname{Ch}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}) & \displaystyle \nonumber\end{eqnarray}$$

and the functor

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x},\;\cdot \;)=[\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x},\;\cdot \;)\rightarrow \boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x},j^{\ast }\;\cdot \;)][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle \operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow \operatorname{Ch}(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}) & \displaystyle \nonumber\end{eqnarray}$$

are compatible under the functor

$$\begin{eqnarray}\bar{f}_{\ast }:\operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow \operatorname{Ch}(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}).\end{eqnarray}$$

The right derived functors

$$\begin{eqnarray}\displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}^{\prime }({\mathcal{O}}_{x},\;\cdot \;)=[R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }({\mathcal{O}}_{x},\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(K_{x},{j^{\prime }}^{\ast }\;\cdot \;)][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle D(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}). & \displaystyle \nonumber\end{eqnarray}$$

and

$$\begin{eqnarray}\displaystyle & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}({\mathcal{O}}_{x},\;\cdot \;)=[R\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x},\;\cdot \;)\rightarrow R\boldsymbol{\unicode[STIX]{x1D6E4}}(K_{x},j^{\ast }\;\cdot \;)][-1]: & \displaystyle \nonumber\\ \displaystyle & \displaystyle D(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow D(k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}) & \displaystyle \nonumber\end{eqnarray}$$

are also compatible under the functor

$$\begin{eqnarray}\bar{f}_{\ast }:D(\hat{{\mathcal{O}}}_{x,\text{fppf}})\rightarrow D(\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{x,\text{et}}^{\text{ind}\text{rat}}).\end{eqnarray}$$

Proof. We have a commutative diagram of premorphisms of sites

Hence the first statement follows from Proposition A.1. We know that $j$ and $j^{\prime }$ are localization morphisms [AGV72, III, Section 5]. Hence their pullback functors admit exact left adjoints by [AGV72, IV, Proposition 11.3.1] and hence send K-injectives to K-injectives. With Proposition 2.4.2, we know that the first statement implies the second. ◻

By [Reference SuzukiSuz14, Proposition (3.3.8)], the functor $R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}^{\prime }({\mathcal{O}}_{x},\;\cdot \;)$ is isomorphic to the functor denoted by $R\tilde{\boldsymbol{\unicode[STIX]{x1D6E4}}}_{x}({\mathcal{O}}_{x},\;\cdot \;)$ in [Reference SuzukiSuz14, paragraph before Proposition (3.3.8)].

We can define the morphism of functoriality of $\bar{f}_{\ast }$

$$\begin{eqnarray}\bar{f}_{\ast }R\,\mathbf{Hom}_{{\mathcal{O}}_{x}}(A,B)\rightarrow R\,\mathbf{Hom}_{{\mathcal{O}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}}}(\bar{f}_{\ast }A,\bar{f}_{\ast }B)\end{eqnarray}$$

in $D({\mathcal{O}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}})$, functorial on $A,B\in D({\mathcal{O}}_{x,\text{fppf}})$, in a way similar to the definition of (2.5.2). Hence we have a morphism

$$\begin{eqnarray}\displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}({\mathcal{O}}_{x},R\,\mathbf{Hom}_{{\mathcal{O}}_{x}}(A,B)) & = & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }({\mathcal{O}}_{x},\bar{f}_{\ast }R\,\mathbf{Hom}_{{\mathcal{O}}_{x}}(A,B))\nonumber\\ \displaystyle & \rightarrow & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }({\mathcal{O}}_{x},R\,\mathbf{Hom}_{{\mathcal{O}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}}}(\bar{f}_{\ast }A,\bar{f}_{\ast }B)).\nonumber\end{eqnarray}$$

There is a similar morphism of functoriality of $f_{\ast }$.

Proposition A.3. Let $A,B\in D({\mathcal{O}}_{x,\text{fppf}})$. Under the above morphism, the morphism (2.5.2)

$$\begin{eqnarray}\displaystyle & & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}(A,B))\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},A),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},B))\nonumber\end{eqnarray}$$

and the morphism

$$\begin{eqnarray}\displaystyle & & \displaystyle R\boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(\hat{{\mathcal{O}}}_{x},R\,\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}}}(\bar{f}_{\ast }A,\bar{f}_{\ast }B))\nonumber\\ \displaystyle & & \displaystyle \quad \rightarrow \,R\,\mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}^{\prime }(\hat{{\mathcal{O}}}_{x},\bar{f}_{\ast }A),R\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}^{\prime }(\hat{{\mathcal{O}}}_{x},\bar{f}_{\ast }B))\nonumber\end{eqnarray}$$

in [Reference SuzukiSuz14, Proposition (3.3.8)] are compatible under the isomorphisms in the previous two propositions. Similar compatibilities hold for the morphisms (2.5.4), (2.5.6) and other morphisms in [Reference SuzukiSuz14, Proposition (3.3.8)].

Proof. This reduces to a corresponding nonderived statement about morphisms between

$$\begin{eqnarray}\displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}(\hat{{\mathcal{O}}}_{x},\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x}}(A,B)), & \displaystyle \nonumber\\ \displaystyle & \displaystyle \boldsymbol{\unicode[STIX]{x1D6E4}}^{\prime }(\hat{{\mathcal{O}}}_{x},\mathbf{Hom}_{\hat{{\mathcal{O}}}_{x,\text{fppf}}/k_{\text{et}}^{\text{ind}\text{rat}}}(\bar{f}_{\ast }A,\bar{f}_{\ast }B)), & \displaystyle \nonumber\\ \displaystyle & \displaystyle \mathbf{Hom}_{k_{x,\text{pro}\text{et}}^{\text{ind}\text{rat}}}(\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},A),\boldsymbol{\unicode[STIX]{x1D6E4}}_{x}(\hat{{\mathcal{O}}}_{x},B)) & \displaystyle \nonumber\end{eqnarray}$$

on the level of complexes of sheaves. Evaluate these complexes at each $k_{x}^{\prime }\in k_{x}^{\text{ind}\text{rat}}$ and then check the compatibility.◻

Footnotes

The author is a Research Fellow of Japan Society for the Promotion of Science.

1 For comparison of derived functors of inverse limits defined for pro-objects and for inverse systems, see [Reference ProsmansPro99, Corollary 7.3.7] for example. This reference assumes that the abelian category in question has exact products. This assumption is satisfied for our category $\operatorname{Ab}(k_{\text{pro}\text{et}}^{\text{ind}\text{rat}})$ by the proposition above.

References

Théorie des topos et cohomologie étale des schémas. Tome 1: Théorie des topos, Lecture Notes in Mathematics 269, Springer, Berlin, 1972; Séminaire de Géométrie Algébrique du Bois-Marie 1963–1964 (SGA 4), Dirigé par M. Artin, A. Grothendieck, et J. L. Verdier. Avec la collaboration de N. Bourbaki, P. Deligne et B. Saint-Donat.Google Scholar
Théorie des topos et cohomologie étale des schémas. Tome 3, Lecture Notes in Mathematics 305, Springer, New York, 1973; Séminaire de Géométrie Algébrique du Bois-Marie 1963–1964 (SGA 4), Dirigé par M. Artin, A. Grothendieck et J. L. Verdier. Avec la collaboration de P. Deligne et B. Saint-Donat.Google Scholar
Artin, M. and Milne., J. S., Duality in the flat cohomology of curves, Invent. Math. 35 (1976), 111129.10.1007/BF01390135Google Scholar
Artin, M., Grothendieck Topologies, Harvard University, Cambridge, Mass, 1962, 133pp.Google Scholar
Artin, M., Supersingular K3 surfaces., Ann. Sci. Éc. Norm. Supér. (4) 7 (1974), 543567.10.24033/asens.1279Google Scholar
Bégueri, L., Dualité sur un corps local à corps résiduel algébriquement clos, Mém. Soc. Math. Fr. (N.S.) No. 4 (1980/81), 121pp.Google Scholar
Bester, M., Local flat duality of abelian varieties, Math. Ann. 235(2) (1978), 149174.10.1007/BF01405011Google Scholar
Bosch, S., Lütkebohmert, W. and Raynaud, M., Néron Models, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)] 21, Springer, Berlin, 1990.10.1007/978-3-642-51438-8Google Scholar
Bombieri, E. and Mumford, D., Enriques’ classification of surfaces in char. p. III, Invent. Math. 35 (1976), 197232.10.1007/BF01390138Google Scholar
Bhatt, B. and Scholze, P., The pro-étale topology for schemes, Astérisque No. 369 (2015), 99201.Google Scholar
Česnavičius, K., Topology on cohomology of local fields, Forum Math. Sigma 3 (2015), e16 55.10.1017/fms.2015.18Google Scholar
Conrad, B., Chow’s K/k-image and K/k-trace, and the Lang–Néron theorem, Enseign. Math. (2) 52(1–2) (2006), 37108.Google Scholar
Demazure, M. and Gabriel, P., Groupes algébriques. Tome I: Géométrie algébrique, généralités, groupes commutatifs, North-Holland Publishing Co., Amsterdam, Masson & Cie, Éditeur, Paris, 1970, Avec un appendice Corps de classes local par Michiel Hazewinkel.Google Scholar
Demarche, C. and Harari, D., Artin–Mazur–Milne duality for fppf cohomology, preprint, 2018, arXiv:1804.03941v2.10.2140/ant.2019.13.2323Google Scholar
Grothendieck, A., Éléments de géométrie algébrique (rédigé avec la collaboration de Jean Dieudonné). IV. Étude locale des schémas et des morphismes de schémas. III, Publ. Math. Inst. Hautes Études Sci. No. 28 (1966), 255pp.Google Scholar
Groupes de monodromie en géométrie algébrique. I., Lecture Notes in Mathematics 288, Springer, New York, 1972, Séminaire de Géométrie Algébrique du Bois-Marie 1967–1969 (SGA 7 I), Dirigé par A. Grothendieck. Avec la collaboration de M. Raynaud et D. S. Rim.Google Scholar
Grothendieck, A., “Techniques de construction et théorèmes d’existence en géométrie algébrique. IV. Les schémas de Hilbert”, in Séminaire Bourbaki, Vol. 6, Mathematical Society of France, Paris, 1995, 249276, Exp. No. 221.Google Scholar
O’Connor, L. H., McGuire, G., Naehrig, M. and Streng, M., A CM construction for curves of genus 2 with p-rank 1, J. Number Theory 131(5) (2011), 920935.Google Scholar
Illusie, L., Complexe de de Rham-Witt et cohomologie cristalline, Ann. Sci. Éc. Norm. Supér. (4) 12(4) (1979), 501661.10.24033/asens.1374Google Scholar
Katz, N. M., Space filling curves over finite fields, Math. Res. Lett. 6(5–6) (1999), 613624.10.4310/MRL.1999.v6.n6.a2Google Scholar
Kashiwara, M. and Schapira, P., Categories and Sheaves, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] 332, Springer, New York, 2006.10.1007/3-540-27950-4Google Scholar
Kato, K. and Trihan, F., On the conjectures of Birch and Swinnerton–Dyer in characteristic p > 0, Invent. Math. 153(3) (2003), 537592.10.1007/s00222-003-0299-2+0,+Invent.+Math.+153(3)+(2003),+537–592.10.1007/s00222-003-0299-2>Google Scholar
Künnemann, K., Projective regular models for abelian varieties, semistable reduction, and the height pairing, Duke Math. J. 95(1) (1998), 161212.10.1215/S0012-7094-98-09505-9Google Scholar
Lang, S., Abelian Varieties, Springer, New York, 1983; Reprint of the 1959 original.10.1007/978-1-4419-8534-7Google Scholar
Lipman, J., Desingularization of two-dimensional schemes, Ann. Math. (2) 107(1) (1978), 151207.10.2307/1971141Google Scholar
Moret-Bailly, L., Pinceaux de variétés abéliennes, Astérisque No. 129 (1985), 266pp.Google Scholar
Milne, J. S., Duality in the flat cohomology of a surface, Ann. Sci. Éc. Norm. Supér. (4) 9(2) (1976), 171201.10.24033/asens.1309Google Scholar
Milne, J. S., Étale Cohomology, Princeton Mathematical Series 33, Princeton University Press, Princeton, NJ, 1980.Google Scholar
Milne, J. S., Arithmetic Duality Theorems, 2nd ed. BookSurge, LLC, Charleston, SC, 2006.Google Scholar
Milne, J. S. and Ramachandran, N., The p-cohomology of algebraic varieties and special values of zeta functions, J. Inst. Math. Jussieu 14(4) (2015), 801835.10.1017/S1474748014000176Google Scholar
Neeman, A., Triangulated Categories, Annals of Mathematics Studies 148, Princeton University Press, Princeton, NJ, 2001.10.1515/9781400837212Google Scholar
Oort, F., Commutative Group Schemes, Lecture Notes in Mathematics 15, Springer, New York, 1966.10.1007/BFb0097479Google Scholar
Prosmans, F., Derived limits in quasi-abelian categories, Bull. Soc. Roy. Sci. Liège 68(5–6) (1999), 335401.Google Scholar
Raynaud, M., Spécialisation du foncteur de Picard, Publ. Math. Inst. Hautes Études Sci. No. 38 (1970), 2776.10.1007/BF02684651Google Scholar
Raynaud, M., Anneaux locaux henséliens, Lecture Notes in Mathematics 169, Springer, New York, 1970.10.1007/BFb0069571Google Scholar
Raynaud, M., “Caractéristique d’Euler-Poincaré d’un faisceau et cohomologie des variétés abéliennes”, in Séminaire Bourbaki, Vol. 9, Soc. Math. France, Paris, 1995, 129147, Exp. No. 286.Google Scholar
Roos, J.-E., Derived functors of inverse limits revisited, J. Lond. Math. Soc. (2) 73(1) (2006), 6583.10.1112/S0024610705022416Google Scholar
Schnürer, O. M., Six operations on dg enhancements of derived categories of sheaves, preprint, 2017, arXiv:1507.08697v3.10.1007/s00029-018-0392-4Google Scholar
Serre, J.-P., Groupes proalgébriques, Publ. Math. Inst. Hautes Études Sci. 7 (1960), 67pp.10.1007/BF02699186Google Scholar
Spaltenstein, N., Resolutions of unbounded complexes, Compos. Math. 65(2) (1988), 121154.Google Scholar
The Stacks Project Authors. Stacks Project. http://stacks.math.columbia.edu, 2018.Google Scholar
Suzuki, T., Duality for local fields and sheaves on the category of fields, preprint, 2013, arXiv:1310.4941v5.Google Scholar
Suzuki, T., Grothendieck’s pairing on Néron component groups: Galois descent from the semistable case, accepted, preprint, 2014, arXiv:1410.3046v4.Google Scholar
Suzuki, T., Néron models of 1-motives and duality, preprint, 2018, arXiv:1806.07641v2.Google Scholar
Trihan, F. and Vauclair, D., On the non commutative Iwasawa main conjecture for abelian varieties over function fields, preprint, 2017, arXiv:1702.04620v1.Google Scholar
Vvedenskiĭ, O. N., Pairings in elliptic curves over global fields, Izv. Akad. Nauk SSSR Ser. Mat. 42 (1978), 237260; 469.Google Scholar
Vvedenskiĭ, O. N., The Artin effect in abelian varieties. II, Izv. Akad. Nauk SSSR Ser. Mat. 45 (1981), 2346; 239.Google Scholar
Yui, N., The arithmetic of the product of two algebraic curves over a finite field, J. Algebra 98(1) (1986), 102142.10.1016/0021-8693(86)90018-9Google Scholar