1. Introduction
In this article we want to study problems related to the finiteness of the $p$-primary torsion of the Brauer group of abelian varieties in positive characteristic
$p$. If
$k$ is a finite field and
$A$ is an abelian variety over
$k$, it is well-known that the Brauer group of
$A$, defined as
$\mathrm {Br}(A):=H^2_\mathrm {\acute {e}t}(A,\mathbb {G}_m)$, is a finite group [Reference TateTat94, Proposition 4.3]. The main input for this result is the Tate conjecture for divisors, proved by Tate in [Reference TateTat66]. If
$k$ is replaced by a finitely generated field extension of
$\mathbb {F}_p$ one can no longer expect
$\mathrm {Br}(A)$ to be finite (see [Reference Skorobogatov and ZarhinSZ08, § 1]). On the other hand, if
$\mathrm {Br}(A_{{k_s}}\!)^{k}$ is the transcendental Brauer group of
$A$, namely the image of
$\mathrm {Br}(A)\to \mathrm {Br}(A_{{k_s}}\!)$ where
${k_s}$ is a separable closure of
$k$, the group
$\mathrm {Br}(A_{{k_s}}\!)^{k}[\frac {1}{p}]$ is finite by [Reference Colliot-Thélène and SkorobogatovCS21, Theorem 16.2.3]. In [Reference Skorobogatov and ZarhinSZ08, Question 1], Skorobogatov and Zarhin asked whether the
$p$-primary torsion of
$\mathrm {Br}(A_{{k_s}}\!)^{k}$ is also finite. This question has a negative answer already for abelian surfaces, as we show in Proposition 5.4. Nonetheless, we prove the following alternative finiteness result. Write
${\bar {k}}$ for an algebraic closure of
$k_s$.
Theorem 1.1 (Theorem 5.2)
Let $A$ be an abelian variety over a finitely generated field
$k$ of characteristic
$p>0$. The transcendental Brauer group
$\mathrm {Br}(A_{{k_s}}\!)^{k}$ is a direct sum of a finite group and a finite exponent
$p$-group. In addition, if the Witt vector cohomology group
$H^2(A_{{\bar {k}}},W\mathcal {O}_{A_{\bar {k}}}\!)$ is a finite
$W({\bar {k}})$-module, then
$\mathrm {Br}(A_{{k_s}}\!)^{k}$ is finite.
The condition on $H^2(A_{{\bar {k}}},W\mathcal {O}_{A_{\bar {k}}}\!)$ is necessary to remove the ‘supersingular pathologies’ as the one of our counterexample and it is satisfied, for example, when the
$p$-rank of
$A$ is
$g$ or
$g-1$, where
$g$ is the dimension of
$A$ (see [Reference IllusieIll83, Corollary 6.3.16]). Note that if the formal Brauer group of
$A_{\bar {k}}$, denoted by
$\hat {\mathrm {Br}}(A_{\bar {k}})$, is a formal Lie group, then by [Reference Artin and MazurAM77, Corollary II.4.4] the cohomology group
$H^2(A_{{\bar {k}}},W\mathcal {O}_{A_{\bar {k}}}\!)$ is a finite
$W({\bar {k}})$-module if and only if
$\hat {\mathrm {Br}}(A_{\bar {k}})$ has finite height. Note also that the formal Brauer group of abelian surfaces is always a formal Lie group by [Reference Artin and MazurAM77, Corollary II.2.12]. As a consequence of Theorem 1.1, we deduce that the subgroup of Galois-fixed points of
$\mathrm {Br}(A_{{k_s}}\!)$, denoted by
$\mathrm {Br}(A_{{k_s}}\!)^{\Gamma _k}$, has finite exponent (Corollary 5.3). This is a variant of [Reference Skorobogatov and ZarhinSZ08, Question 2] for abelian varieties.
In this article, we also study the Galois-fixed points of $\mathrm {Br}(A_{{\bar {k}}})$. Ulmer in [Reference UlmerUlm14, § 7.3.1] conjectured that
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))^{\Gamma _k}=0$ where
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))$ is the
$p$-adic Tate module of
$\mathrm {Br}(A_{{\bar {k}}})$. Even in this case, we provide a counterexample to this conjecture. We use the following result.
Proposition 1.2 (Proposition 6.6)
Let $B$ be an abelian variety over a finitely generated field
$k$ of characteristic
$p>0$. Write
$A$ for
$B\times _k B$ and
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))$ for the
$p$-adic Tate module of
$\mathrm {Br}(A_{\bar {k}})$. There is a natural exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU1.png?pub-status=live)
where $\operatorname {Hom}(B,B^\vee )$ denotes the group of homomorphisms
$B\to B^\vee$ as abelian varieties over
$k$.
The proposition implies, for example, that when $\operatorname {End}(B)=\mathbb {Z}$ the
$\Gamma _k$-module
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))$ admits non-zero Galois-fixed points (Corollary 6.7). In this case,
$\mathrm {Br}(A_{{\bar {k}}})^{\Gamma _k}$ has infinite exponent since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU2.png?pub-status=live)
Note that if we replace $\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))$ with the
$\ell$-adic Tate module
$\mathrm {T_\ell }(\mathrm {Br}(A_{\bar {k}}))$, where
$\ell$ is a prime different from
$p$, then
$\mathrm {T_\ell }(\mathrm {Br}(A_{{k_s}}\!))=\mathrm {T_\ell }(\mathrm {Br}(A_{\bar {k}}))$ has no non-trivial Galois-fixed points.
These ‘exceptional classes’ in $\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))^{\Gamma _k}$ are naturally related to specialisation morphisms of Néron–Severi groups. We recall the following theorem, which was proved in [Reference AndréAnd96, Theorem 5.2] in characteristic
$0$ (see also [Reference Maulik and PoonenMP12]) and in [Reference AmbrosiAmb23] and [Reference ChristensenChr18] in positive characteristic.
Theorem 1.3 (André, Ambrosi, Christensen)
Let $K$ be an algebraically closed field which is not an algebraic extension of a finite field,
$X$ a finite-type
$K$-scheme, and
$\mathcal {Y}\to X$ a smooth proper morphism. For every geometric point
$\bar {\eta }$ of
$X$ there is an
$x\in X(K)$ such that
$\operatorname {rk}_\mathbb {Z}(\mathrm {NS}(\mathcal {Y}_{\bar {\eta }}))=\operatorname {rk}_\mathbb {Z}(\mathrm {NS}(\mathcal {Y}_x))$.Footnote 1
As it is well-known, the theorem is false when $K=\bar {\mathbb {F}}_p$ (see [Reference Maulik and PoonenMP12, Remark 1.12]). What we prove is that, in the known counterexamples, the elements in
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))^{\Gamma _k}$ explain the failure of Theorem 1.3. More precisely, we prove the following result.
Theorem 1.4 (Theorem 6.2)
Let $X$ be a connected normal scheme of finite type over
$\mathbb {F}_p$ with generic point
$\eta =\operatorname {Spec}(k)$ and let
$f:\mathcal {A}\to X$ be an abelian scheme over
$X$ with constant Newton polygon.Footnote 2 For every closed point
$x=\operatorname {Spec}(\kappa )$ of
$X$ we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU3.png?pub-status=live)
Note that in the inequality the left term is ‘motivic’, whereas the right term comes from some $p$-adic object which, as far as we know, has no
$\ell$-adic analogue. Note also that
$\mathrm {T}_p(\mathrm {Br}(\mathcal {A}_{\bar {x}}))^{\Gamma _\kappa }=0$ by Corollary 5.3 since
$\kappa$ is a perfect field.
To prove Theorem 1.1 we use a flat variant of the Tate conjecture. For every $n$, let
$H^2_\mathrm {fppf}(A_{\bar {k}},{\mu _{p^n}}\!)^k$ be the image of the extension of scalars morphism
$H^2_\mathrm {fppf}(A,{\mu _{p^n}}\!)\to H^2_\mathrm {fppf}(A_{\bar {k}},{\mu _{p^n}}\!)$.
Theorem 1.5 (Theorem 5.1)
After possibly replacing $k$ with a finite separable extension, the cycle class map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU4.png?pub-status=live)
becomes an isomorphism.Footnote 3
We obtain this result by using the crystalline Tate conjecture for abelian varieties, proved by de Jong in [Reference de JongdeJ98, Theorem 2.6]. The main issue that we have to overcome is the lack of a good comparison between crystalline and fppf cohomology of ${\mathbb {Z}_{p}}(1)$ over imperfect fields. To avoid this problem, we exploit the fact that we are working with abelian varieties. In this special case, the comparison is constructed using the
$p$-divisible group of
$A$ (and its dual).
The technical issue that we have to solve using the groups $H^2_{\mathrm {fppf}}(A_{\bar {k}},{\mu _{p^n}}\!)^k$ is that it is not clear a priori whether
$H^2_{\mathrm {fppf}}(A,{\mathbb {Z}_{p}}(1))\to \varprojlim _n H^2_{\mathrm {fppf}}(A_{\bar {k}},{\mu _{p^n}}\!)^k$ is surjective. This is done (after inverting
$p$) in Proposition 3.9, where we reduce to the case when
$A$ is the Jacobian of a curve. This idea was inspired by the proof of [Reference Colliot-Thélène and SkorobogatovCS13, Theorem 2.1].
1.6 Outline of the article
In § 3 we prove some general results on the cohomology of fppf sheaves. In particular, we prove Corollary 3.4, which is a first result on the relation between the Brauer group of a scheme over ${k_s}$ and
${\bar {k}}$. In this section, we also prove in Proposition 3.8 the exactness of some fundamental sequences for the groups
$H^2_\mathrm {fppf}(X_{\bar {k}},{\mu _{p^n}}\!)^k$. In § 4, we construct a morphism which relates
$H^2_\mathrm {fppf}(A,{\mathbb {Z}_{p}}(1))$ with
$\operatorname {Hom}(A[p^\infty ],A^\vee [p^\infty ])$ and we prove basic properties of this morphism as Propositions 4.5 and 4.6. In § 5, we prove the flat variant of the Tate conjecture (Theorem 5.1) and the finiteness result for the transcendental Brauer group (Theorem 5.2). Finally, in § 6, we look at the relation of our results with the theory of specialisation of Néron–Severi groups. In particular, we prove Theorem 6.2.
2. Notation
If $k$ is a field, we write
${\bar {k}}$ for a fixed algebraic closure of
$k$ and
${k_s}$ (respectively,
${k_i}$) for the separable (respectively, purely inseparable) closure of
$k$ in
${\bar {k}}$. We denote by
$\Gamma _k$ the absolute Galois group of
$k$. If
$x$ is a
$k$-point of a scheme, we denote by
$\bar {x}$ the induced
${\bar {k}}$-point. For an abelian group
$M$, we write
$\mathrm {T}_p(M)$ for the
$p$-adic Tate module of
$M$, which is the projective limit
$\varprojlim _{n}M[p^n]$, we write
$\mathrm {V}_p(M)$ for
$\mathrm {T}_p(M)[\frac {1}{p}]$, and we write
$M^\wedge$ for the
$p$-adic completion of
$M$. If
$M$ is endowed with a
$\Gamma _k$-action, we denote by
$M^{\Gamma _k}$ the subgroup of fixed points. For a scheme
$X$ and an fppf sheaf
$\mathcal {F}$, we denote by
$H^{\bullet }(X,\mathcal {F})$ the fppf cohomology groups and when
$X=\operatorname {Spec}(k)$ we simply write
$H^{\bullet }(k,\mathcal {F})$. If
$f:X\to Y$ is a morphism of schemes, we denote by
$R^{\bullet }f_*\mathcal {F}$ the fppf higher direct image functors over
$(\mathbf {Sch}/Y)_\mathrm {fppf}$. Finally, if
$X$ is a scheme over
$\mathbb {F}_p$, we write
$X^{\mathrm {perf}}$ for the projective limit
$\varprojlim (\cdots \xrightarrow {F}X\xrightarrow {F}X\xrightarrow {F}X)$ where
$F$ is the absolute Frobenius of
$X$.
3. Preliminary results
In this section we start by proving some results that we will use later on. We work over a field $k$ of arbitrary characteristic and we consider a scheme
$X$ over
$k$ with structural morphism
$q$.
Lemma 3.1 Let $\mathcal {F}$ be a sheaf over
$(\mathbf {Sch}/k)_\mathrm {fppf}$ such that
$q_*\mathcal {F}_{X}=\mathcal {F}$ and suppose that
$X$ has a
$k$-rational point. The group
$H^0(k,R^1q_*\mathcal {F}_{X}\!)$ is canonically isomorphic to
$H^1(X,\mathcal {F}_X\!)/H^1(k,\mathcal {F})$. In addition, the natural morphism
$H^2(X,\mathcal {F}_X\!)\to H^0(k,R^2q_*\mathcal {F}_{X}\!)$ sits in an exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU5.png?pub-status=live)
where $K$ is an extension of
$H^1(k,R^1q_*\mathcal {F}_{X}\!)$ by
$H^2(k,\mathcal {F})$.
Proof. We consider the Leray spectral sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU6.png?pub-status=live)
The morphisms $E^{i,0}_2=H^i(k,q_*\mathcal {F}_X\!)=H^i(k,\mathcal {F})\to H^i(X,\mathcal {F}_X\!)$ are injective since
$X$ admits a
$k$-rational point. We deduce that
$E^{1,1}_2=E^{1,1}_\infty$ and
$E^{2,0}_2=E^{2,0}_\infty$. This implies that the kernel of
$H^2(X,\mathcal {F}_X\!)\to E^{0,2}_\infty$ is an extension of
$E^{1,1}_2$ by
$E^{2,0}_2$, as we wanted. The obstruction for the map
$H^2(X,\mathcal {F}_X\!)\to E^{0,2}_2=H^0(k,R^2q_*\mathcal {F}_{X}\!)$ to be surjective lies in
$E^{2,1}_2=H^2(k,R^1q_*\mathcal {F}_X\!)$. This concludes the proof.
Definition 3.2 We say that a presheaf $\mathcal {F}$ on
$(\mathbf {Sch}^{\mathrm {qcqs}}/k)_{\mathrm {fppf}}$ is finitary if for every inverse system
$\{T^{(\ell )}\}_{\ell \in L}$ of quasi-compact quasi-separated
$k$-schemes with affine transition maps, the natural morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU7.png?pub-status=live)
is an isomorphism.
Lemma 3.3 Let $G$ be a commutative finite-type group scheme over
$k$. If
$X$ is quasi-compact quasi-separated, then
$R^iq_*G_X$ is finitary for
$i\geq 0$. In addition, the natural morphism
$H^0(k,R^i q_*G_X\!)\to H^{i}(X_{\bar {k}},G_{X_{\bar {k}}}\!)$ is injective.
Proof. Let ${\mathcal {H}}^i(q,G_{X}\!)$ be the higher presheaf pushforward of
$G_{X}$ on
$X$ with respect to
$q$. We first want to prove that
${\mathcal {H}}^i(q,G_{X}\!)$ is finitary for
$i\geq 0$. In other words, we want to prove that for every inverse system
$\{T^{(\ell )}\}_{\ell \in L}$ of quasi-compact quasi-separated
$k$-schemes, the natural morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU8.png?pub-status=live)
is an isomorphism, where $X^{(\ell )}:=X\times _k T^{(\ell )}$ and
$X^{(\infty )}:=\lim _{\ell \in L}X^{(\ell )}$. By [Sta23, Tag 01H0],
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU9.png?pub-status=live)
for every $\ell \in L \coprod \{\infty \}$, where
$\mathrm {HC}(X^{(\ell )})$ is the category of fppf hypercoverings of
$X^{(\ell )}$. Since each
$X^{(\ell )}$ is quasi-compact quasi-separated, by [Sta23, Tag 021P] we can replace the category
$\mathrm {HC}(X^{(\ell )})$ in the colimit with the subcategory
$\mathrm {HC}(X^{(\ell )})^{\mathrm {qcqs}}$, consisting of those hypercoverings such that
$U^{(\ell )}_n$ is quasi-compact quasi-separated for every
$n\geq 0$. By [Sta23, Lemma 01ZM], for
$U^{(\infty )}_\bullet \in \mathrm {HC}(X^{(\infty )})^{\mathrm {qcqs}}$ and
$n\geq 0$ there exists an
$\ell \in L$ and
$U_\bullet ^{(\ell,n)}\in \mathrm {HC}(X^{(\ell )})^{\mathrm {qcqs}}$ such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU10.png?pub-status=live)
where $\mathrm {tr}_n(-)$ denotes the
$n$th truncation of simplicial schemes and
$T^{(\infty )}:=\lim _{\ell \in L}T^{(\ell )}$. This implies that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU11.png?pub-status=live)
We are reduced to proving that for every $\ell \in L$ and
$U_\bullet ^{(\ell )}\in \mathrm {HC}(X^{(\ell )})^{\mathrm {qcqs}}$ we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU12.png?pub-status=live)
Since $G$ is of finite type over
$k$, this follows from [Sta23, Lemma 01ZM] and the exactness of filtered colimits.
Knowing that ${\mathcal {H}}^i(q,G_{X}\!)$ is finitary, in order to prove that
$R^iq_*G_X\!$ is finitary as well it is enough to prove that for every finitary presheaf
$\mathcal {F}$ on
$(\mathbf {Sch}^{\mathrm {qcqs}}/k)_{\mathrm {fppf}}$, the ‘partial’ sheafification
$\mathcal {F}^+$ (defined as in [Sta23, § 00W1]) is finitary. Similarly to the previous paragraph, the proof of this fact follows from the observation that each finite quasi-compact quasi-separated fppf covering of
$X^{(\infty )}$ descends to a covering of
$X^{(\ell )}$ for some
$\ell \in L$ and Čech cohomology commutes with filtered colimits (
$\check {H}^0$ is enough in this case).
For the second part, we note that for every presheaf $\mathcal {F}$ on
$(\mathbf {Sch}/k)_{\mathrm {fppf}}$ with sheafification
$\mathcal {F}^\sharp$, the natural morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU13.png?pub-status=live)
is an isomorphism because every fppf covering of $\operatorname {Spec}({\bar {k}})$ admits a section. This implies that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU14.png?pub-status=live)
Thanks to the previous part, we deduce that the composition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU15.png?pub-status=live)
is injective, where the colimit runs over all finite field extensions of $k$. This ends the proof.
With the previous results we can prove [Reference GrothendieckGro68, Proposition 5.6], which was stated by Grothendieck without a complete proof.Footnote 4
Corollary 3.4 If $k$ is separably closed and
$X$ is a proper
$k$-scheme, then there is a natural exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU16.png?pub-status=live)
In particular, if $\mathrm {Pic}_{X/k}$ is smooth, then the natural morphism
$\mathrm {Br}(X)\to \mathrm {Br}(X_{{\bar {k}}})$ is injective.
Proof. As in Lemma 3.1, we consider the Leray spectral sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU17.png?pub-status=live)
Since $X$ is proper over
$k$, by [Sta23, Tag 0BUG] we deduce that
$A:=H^0(X,\mathcal {O}_X\!)$ is a finite
$k$-algebra. This implies that
$q_*\mathbb {G}_m$ is represented by a smooth group scheme over
$k$. Thanks to [Reference GrothendieckGro68, Theorem 11.7], we deduce that
$E^{i,j}_2=0$ for
$i>0$ and
$j=0$, so that
$E^{1,1}_2=E^{1,1}_\infty$. The Leray spectral sequence produces then the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU18.png?pub-status=live)
To get the first part of the statement it is then enough to apply Lemma 3.3. For the second part, we note that when $\mathrm {Pic}_{X/{k}}$ is smooth, thanks to [Reference GrothendieckGro68, Theorem 11.7], the group
$H^1(k,\mathrm {Pic}_{X/{k}})$ vanishes.
Definition 3.5 For a scheme $X$ over
$k$ and a prime
$p$, we define
$H^2(X,\mathbb {Z}_p(1))$ as the projective limit
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU19.png?pub-status=live)
Remark 3.6 Note that we are defining $H^2(X,\mathbb {Z}_p(1))$ without taking into account higher inverse limits. Nonetheless, if
$k$ is algebraically closed of characteristic
$p$ and
$X$ is smooth and proper over
$k$, then
$R^1\varprojlim _n H^1(X,{\mu _{p^n}}\!)=R^1\varprojlim _n \mathrm {Pic}(X)[p^n]=0$ since
$\mathrm {Pic}(X)[p^\infty ]$ is a direct sum of a
$p$-divisible group and a finite group and
$R^1\varprojlim _n H^2(X,{\mu _{p^n}}\!)=0$ by [Reference IllusieIll79, Chapter II, Proposition 5.9].
Construction 3.7 The Kummer exact sequences for $X$ and
$X_{{\bar {k}}}$ (for the fppf topology) induce the following commutative diagram with exact rows.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqn1.png?pub-status=live)
We write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU20.png?pub-status=live)
for the complex obtained by taking images of the vertical arrows. Note that a priori $(\mathrm {Br}(X_{{\bar {k}}})[p^n])^{k}$ is smaller than
$\mathrm {Br}(X_{\bar {k}})^k[p^n]$, where
$\mathrm {Br}(X_{\bar {k}})^k:=\mathrm {im}(\mathrm {Br}(X)\to \mathrm {Br}(X_{\bar {k}}))$.
Since both $R^1\varprojlim _{n}\mathrm {Pic}(X)/p^n$ and
$R^1\varprojlim _{n}\mathrm {Pic}(X_{\bar {k}})/p^n$ vanish, we can also consider the following commutative diagram with exact rows:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU21.png?pub-status=live)
obtained by taking the projective limit of the diagrams (3.7.1) for various $n$. We denote by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU22.png?pub-status=live)
the complex obtained by taking images of the vertical arrows.
Proposition 3.8 If $\mathrm {char}(k)=p$ and
$A$ is an abelian variety over
$k$ such that the morphism
$\mathrm {Pic}(A)\to \mathrm {NS}(A_{{\bar {k}}})$ is surjective, then the complexes
$C_n(A)$ and
$\hat {C}(A)$ are acyclic.
Proof. If $K_{1,n}$ is the kernel of
$H^2_{}(A,{\mu _{p^n}}\!)\to H^2(A_{\bar {k}},{\mu _{p^n}}\!)$ and
$K_{2}$ is the kernel of
$\mathrm {Br}(A)\to \mathrm {Br}(A_{\bar {k}})$, in order to prove that
$C_n(A)$ is acyclic we have to show that
$K_{1,n}\to K_{2}[p^n]$ is surjective. Combining Lemmas 3.1 and 3.3, we deduce the following commutative diagram with exact rows.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU23.png?pub-status=live)
The morphism of exact sequences factors through the complex
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU24.png?pub-status=live)
which is acyclic because $\mathrm {Br}(k)$ is
$p$-divisible by [Reference Colliot-Thélène and SkorobogatovCS21, Theorem 1.3.7]. The image of
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU25.png?pub-status=live)
is $H^1_{}(k,\mathrm {Pic}_{A/k}^\circ )[p^n]$, thus we are reduced to prove that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU26.png?pub-status=live)
is surjective. Since $\mathrm {Pic}(A)\to \mathrm {NS}(A_{\bar {k}})$ is surjective, we know that
$\pi _0(\mathrm {Pic}_{A/k})$ is a constant finitely generated torsion-free group over
$k$ such that
$\mathrm {Pic}_{A/k}(k)\to \pi _0(\mathrm {Pic}_{A/k})(k)$ is surjective. Looking at the cohomology long exact sequence associated to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU27.png?pub-status=live)
we then deduce that $H^1_{}(k,\mathrm {Pic}_{A/k}^\circ )\xrightarrow {\sim } H^1_{}(k,\mathrm {Pic}_{A/k})$, which yields the desired result.
We now prove that $\hat {C}(A)$ is acyclic. The kernel of
$H^2(A,{\mathbb {Z}_{p}}(1))\to H^2(A_{\bar {k}},{\mathbb {Z}_{p}}(1))$ is
$\varprojlim _n K_{1,n}$ and the kernel of
$\mathrm {T}_p(\mathrm {Br}(A))\to \mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))$ is
$\mathrm {T}_p(K_2)$. Thus, again, we have to prove that
$\varprojlim _n K_{1,n}\to \mathrm {T}_p(K_2)$ is surjective. Combining the previous discussion and the fact that
$\mathrm {Br}(k)$ is
$p$-divisible, we deduce that the two groups sit in the following diagram with exact rows.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU28.png?pub-status=live)
For every $n>0$, the kernel of
$H^1_{}(k,\mathrm {Pic}_{A/k}[p^n]) \to H^1_{}(k,\mathrm {Pic}_{A/k})[p^n]$ is
$\mathrm {Pic}(A)/p^n$ and the groups
$(\mathrm {Pic}(A)/p^n)_{n>0}$ form a Mittag–Leffler system. We deduce that the morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU29.png?pub-status=live)
is surjective. This implies that $\hat {C}(A)$ is acyclic, as we wanted.
The proof of the following proposition was inspired by [Reference Colliot-Thélène and SkorobogatovCS13, Theorem 2.1].
Proposition 3.9 If $\mathrm {char}(k)=p$ and
$A$ is an abelian variety over
$k$, we have
$H^2(A_{\bar {k}},{\mathbb {Z}_{p}}(1))^{k}[\frac {1}{p}]=(\varprojlim _{n}H^2(A_{\bar {k}},{\mu _{p^n}}\!)^{k})[\frac {1}{p}]$ and
$\mathrm {T}_p(\mathrm {Br}(A_{{\bar {k}}}))^{k}[\frac {1}{p}]=\mathrm {V}_p(\mathrm {Br}(A_{{\bar {k}}})^{k}).$
Proof. We first note that the four ${\mathbb {Q}_{p}}$-vector spaces are invariant under isogenies of
$A$ and finite separable extension of
$k$. Indeed, for every isogeny
$\varphi :B\to A$ there exists an isogeny
$\psi : A\to B$ such that the composition
$\varphi \circ \psi$ is the multiplication by some positive integer
$n$. Since
$n$ is invertible in
${\mathbb {Q}_{p}}$, we deduce that
$\varphi ^*$ is an isomorphism at the level of cohomology groups. Similarly, if
$k'/k$ is a finite separable extension, then the pullback morphisms with respect to
$A_{k'}\to A$ admit as inverse the morphisms
$ ({1}/{[k':k]})\mathrm {Tr}_{A_{k'}/A}$.
Next, thanks to [Reference KatzKat99, Theorem 11], we note that there exists a proper smooth connected curve $C$ with a rational point and a morphism
$C\to A$ such that
$B:=\mathrm {Jac}(C)$ maps surjectively to
$A$. By Poincaré's complete reducibility theorem,
$B$ is isogenous to a product
$A\times _k A'$ with
$A'$ an abelian variety over
$k$. Since
$H^2(A_{\bar {k}},{\mathbb {Z}_{p}}(1))^{k}$ (respectively,
$\mathrm {Br}(A_{{\bar {k}}})^k$) is a direct summand of
$H^2(A_{\bar {k}}\times _{\bar {k}} A'_{\bar {k}},{\mathbb {Z}_{p}}(1))^{k}$ (respectively,
$\mathrm {Br}(A_{\bar {k}}\times _{\bar {k}} A'_{{\bar {k}}})^k$) and the property we want to prove is invariant by isogenies, it is then enough to prove the result for
$B$. In addition, since in the statement it is harmless to extend
$k$ to a finite separable extension, we may assume that
$\mathrm {Pic}(B)\to \mathrm {NS}(B_{{\bar {k}}})$ is surjective, so that
$H^1(k,\mathrm {Pic}^0_{B/k})=H^1(k,\mathrm {Pic}_{B/k})$.
Let $K_{1,n}$ be the kernel of the morphism
$H^2(B,{\mu _{p^n}}\!)\to H^2(B_{\bar {k}},{\mu _{p^n}}\!)$. By Lemmas 3.1 and 3.3, the group
$K_{1,n}$ is an extension of
$H^1(k,\mathrm {Pic}_{B/k}[p^n])$ by
$\mathrm {Br}(k)[p^n]$ and by the assumption
$\mathrm {Pic}_{C/k}[p^n]=\mathrm {Pic}_{B/k}[p^n]$. We deduce that
$K_{1,n}=\ker (H^2(C,{\mu _{p^n}}\!)\to H^2(C_{\bar {k}},{\mu _{p^n}}\!))$. By [Reference GrothendieckGro68, Rmq. 2.5.b], the group
$\mathrm {Br}(C_{\bar {k}})$ vanishes, thus
$H^2(C_{{\bar {k}}},{\mathbb {Z}_{p}}(1))={\mathbb {Z}_{p}}$ and the morphism
$H^2(C,{\mathbb {Z}_{p}}(1))\to H^2(C_{{\bar {k}}},{\mathbb {Z}_{p}}(1))$ is surjective because
$C$ has a rational point. This implies that
$R^1 \varprojlim _{n}K_{1,n}\to R^1 \varprojlim _{n}H^2(C,{\mu _{p^n}}\!)$ is injective. Since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU30.png?pub-status=live)
factors through $R^1 \varprojlim _{n}H^2(B,{\mu _{p^n}}\!),$ we deduce that
$R^1 \varprojlim _{n}K_{1,n}\to R^1 \varprojlim _{n}H^2(B,{\mu _{p^n}}\!)$ is injective as well. Therefore, the morphism
$H^2(B,{\mathbb {Z}_{p}}(1))\to \varprojlim _{n}H^2(B_{\bar {k}},{\mu _{p^n}}\!)^{k}$ is surjective. Thanks to Proposition 3.8, for every
$n$ the morphism
$H^2(B_{\bar {k}},{\mu _{p^n}}\!)^{k}\to (\mathrm {Br}(B_{\bar {k}})[p^n])^k$ is surjective with finite kernel, so that
$\varprojlim _{n}H^2(B_{\bar {k}},{\mu _{p^n}}\!)^{k}\to \varprojlim _n(\mathrm {Br}(B_{\bar {k}})[p^n])^k$ is surjective as well. This implies that
$\mathrm {T}_p(\mathrm {Br}(B_{{\bar {k}}}))^{k}\to \varprojlim _n(\mathrm {Br}(B_{\bar {k}})[p^n])^k$ is surjective.
It remains to prove that for every $n$ we have
$(\mathrm {Br}(B_{\bar {k}})[p^n])^k=\mathrm {Br}(B_{\bar {k}})^k[p^n]$. Consider the natural morphism
$K_3\to \mathrm {Br}(C)$ where
$K_3$ is the kernel of
$\mathrm {Br}(B)\to \mathrm {Br}(B_{{\bar {k}}})$. Thanks to Lemmas 3.1 and 3.3 and using the fact that
$\mathrm {Br}(C_{\bar {k}})=0$, this morphism sits in the following commutative diagram with exact rows.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU31.png?pub-status=live)
Since $\mathrm {Pic}(B)\to \mathrm {NS}(B_{\bar {k}})$ is surjective and
$C$ is a curve, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU32.png?pub-status=live)
We deduce that $K_3\to \mathrm {Br}(C)$ is an isomorphism, thus
$\mathrm {Br}(B)\to \mathrm {Br}(C)\xrightarrow {\sim } K_3$ provides a splitting of the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU33.png?pub-status=live)
This implies that $\mathrm {Br}(B)[p^n]\to \mathrm {Br}(B_{{\bar {k}}})^k[p^n]$ is surjective and this yields the desired result.
4. Constructing a morphism
Let $A$ be an abelian variety over a field
$k$. For a line bundle
$\mathcal {L}$ of
$A$ we write
$\varphi _{\mathcal {L}}:A\to A^\vee$ for the morphism which sends
$x\mapsto t_x^*\mathcal {L}\otimes \mathcal {L}^{-1}$, where
$t_x$ is the translation by
$x$. In this section we want to complete the following solid square.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU34.png?pub-status=live)
If $k$ is an algebraically closed field of characteristic
$0$ such a commutative diagram is constructed in [Reference Orr, Skorobogatov and ZarhinOSZ21, Lemma 2.6] using an analytic method. We propose instead an algebraic construction which works for any field.
4.1
Consider the morphism $h_1:H^2(A,{\mu _{p^n}}\!)\to H^2(A\times _k A,{\mu _{p^n}}\!)$ which sends a class
$\alpha$ to
$m^*(\alpha )-\pi _1^*(\alpha )-\pi _2^*(\alpha )$, where
$\pi _1$ and
$\pi _2$ are the two projections of
$A\times _k A$. This morphism has the property that the first Chern class
$c_1(\mathcal {L})\in H^2(A,{\mu _{p^n}}\!)$ of a line bundle
$\mathcal {L}$ is sent to
$c_1(\Lambda (\mathcal {L}))$, the first Chern class of the associated Mumford bundle
$\Lambda (\mathcal {L}):=m^*\mathcal {L}\otimes \pi _1^*\mathcal {L}^{-1}\otimes \pi _2^*\mathcal {L}^{-1}$. The Leray spectral sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqn2.png?pub-status=live)
induces a filtration $0\subseteq F^2 H^{2}(A\times _k A,{\mu _{p^n}}\!) \subseteq F^1 H^{2}(A\times _k A,{\mu _{p^n}}\!)\subseteq H^{2}(A\times _k A,{\mu _{p^n}}\!)$.
Lemma 4.2 The image of $h_1$ lies in
$F^1H^{2}(A\times _k A,{\mu _{p^n}}\!)$.
Proof. The spectral sequence (4.1.1) gives the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU35.png?pub-status=live)
Therefore, it is enough to check that the composition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU36.png?pub-status=live)
is the $0$-morphism. By [Reference Bragg and OlssonBO21, Corollary 1.4], there exists a commutative linear algebraic groupFootnote 5
$G$ over
$k$ which represents
$R^2q_{*}{\mu _{p^n}}$ on the big fppf site
$(\mathbf {Sch}/k)_{\mathrm {fppf}}$. Since
$R^2\pi _{2*}{\mu _{p^n}}$ is the restriction of
$R^2q_{*}{\mu _{p^n}}$ from
$(\mathbf {Sch}/k)_{\mathrm {fppf}}$ to
$(\mathbf {Sch}/A)_{\mathrm {fppf}}$, this implies that
$H^0(A,R^2\pi _{2*}{\mu _{p^n}}\!)$ can be computed as
$\mathrm {Mor}_{\mathbf {Sch}/k}(A,G)$. Thanks to the fact that
$G$ is affine, every morphism
$A\to G$ contracts
$A$ to a point. We deduce that
$\mathrm {Mor}_{\mathbf {Sch}/k}(A,G)=\mathrm {Mor}_{\mathbf {Sch}/k}(0_A,G)=H^0(k,R^2q_{*}{\mu _{p^n}}\!)$. By Lemma 3.3, the group
$H^0(k,R^2q_{*}{\mu _{p^n}}\!)$ is naturally a subgroup of
$H^2(A_{\bar {k}},{\mu _{p^n}}\!)$ and the induced morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU37.png?pub-status=live)
is given by the pullback via $i_1:A=A\times _k 0_A \hookrightarrow A\times _k A$ followed by the extension of scalars to
${\bar {k}}$. By construction, we have that
$i_1^*\circ h_1=i_1^*\circ m^*-i_1^*\circ \pi _1^*=0$. This concludes the proof.
Lemma 4.3 Let $G$ be a finite commutative group scheme killed by a positive integer
$n$. There is a natural injective morphism
$f_n:\operatorname {Hom}(A[n],G)\to H^1(A,G)$ which admits a retraction
$g_n$.
Proof. Write $P_{n}$ for the
$A[n]$-torsor over
$A$ given by the multiplication by
$n$. The morphism
$f_n$ is then defined by
$f_n(\sigma ):=\sigma _*P_{n}$ for every
$\sigma \in \operatorname {Hom}(A[n],G)$. We want to define now
$g_n$ which sends a
$G$-torsor
$P$ over
$A$ to an homomorphism
$g_n(P): A[n]\to G$. By Cartier duality, this is the same as defining a morphism
$g_n(P)^\vee :G^\vee \to (A[n])^\vee =A^\vee [n]$. For a scheme
$T$ over
$k$ and a
$T$-point of
$G^\vee$ corresponding to a morphism
$\tau :G_T\to \mathbb {G}_{m,T}$ we define
$g_n(P)^\vee (\tau )$ as
$\tau _{*}P_T\in H^1(A_T,\mathbb {G}_{m,T})[n]=A^\vee [n](T)$. To prove that
$g_n\circ f_n=\operatorname {id}$ it is enough to note that for every
$\sigma \in \operatorname {Hom}(A[n],G)$, every scheme
$T$ over
$k$, and every
$\tau \in \operatorname {Hom}(G_T,\mathbb {G}_{m,T})$ we have that
$g_n(f_n(\sigma ))^\vee (\tau )=\tau _*(\sigma _*P_n)_T=(\tau \circ \sigma _T)_*P_{n,T}$ is the line bundle over
$A_T$ associated to
$\tau \circ \sigma _T\in (A[n])^\vee (T)$ under the identification
$(A[n])^\vee =A^\vee [n]$.
4.4
Thanks to Lemma 4.2, we can define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU38.png?pub-status=live)
as the composition of $h_1$ and the natural morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU39.png?pub-status=live)
In addition, by Lemma 4.3 applied to $G=A^\vee [p^n]$, we get a morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU40.png?pub-status=live)
We write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU41.png?pub-status=live)
for the composition ${h_2}\circ \bar {h}_1$ and we denote with the same letter the induced morphism
$H^2(A,{\mathbb {Z}_{p}}(1))\to \operatorname {Hom}(A[p^\infty ],A^\vee [p^\infty ])$.
Proposition 4.5 The square
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqn3.png?pub-status=live)
is commutative.
Proof. We have to show that for every line bundle $\mathcal {L}\in \mathrm {Pic}(A)$ we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU42.png?pub-status=live)
Consider the Leray spectral sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqn4.png?pub-status=live)
The morphism $A\times _k A\xrightarrow {\operatorname {id}_A\times \varphi _{\mathcal {L}}} A\times _k A^\vee$ induces via pullback a morphism from (4.5.2) to (4.1.1). This produces the commutative diagram
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU43.png?pub-status=live)
where the composition of the lower horizontal arrows is $h$. If
$\mathcal {P}\in \mathrm {Pic}({A\times _k A^\vee })$ is the Poincaré bundle of
$A$, we have that
$\Lambda (\mathcal {L})=(\operatorname {id}_A\times \varphi _{\mathcal {L}})^* \mathcal {P}$. This implies that
$h_1(c_1(\mathcal {L}))=c_1(\Lambda (\mathcal {L}))=(\operatorname {id}_A\times \varphi _{\mathcal {L}})^* c_1(\mathcal {P})$. In addition, by direct inspection, we note that
$h_2(\varphi _{\mathcal {L}}^*([P]))=\varphi _\mathcal {L}|_{A[p^n]}$, where
$[P]\in H^1(A^\vee,A^\vee [p^n])$ is the class of the torsor
$A^\vee \xrightarrow {\cdot p^n} A^\vee$. It remains to prove that the morphism
$F^1H^2(A\times _k A^\vee,{\mu _{p^n}}\!)\to H^1(A^\vee,A^\vee [p^n])$ sends
$c_1(\mathcal {P})$ to
$[P]$. For this purpose, we introduce the Leray spectral sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqn5.png?pub-status=live)
The morphism $\delta :\mathbb {G}_m[1]\to {\mu _{p^n}}$ associated to the Kummer exact sequence induces a morphism from (4.5.3) to (4.5.2) which we denote with the same symbol. In turn, this produces the following commutative diagram.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU44.png?pub-status=live)
The upper horizontal arrow sends the line bundle $\mathcal {P}$ to
$\operatorname {id}_{A^\vee }\in H^0(A^\vee,A^\vee )$, while
$\delta$ sends
$\operatorname {id}_{A^\vee }$ to
$[P]$. This yields the desired result.
Proposition 4.6 The morphism $h:H^2(A_{\bar {k}},{\mathbb {Z}_{p}}(1))\to \operatorname {Hom}_{}(A_{\bar {k}}[p^\infty ],A^\vee _{\bar {k}}[p^\infty ])$ is an injective morphism with image
$\operatorname {Hom}^{\mathrm {sym}}_{}(A_{\bar {k}}[p^\infty ],A^\vee _{\bar {k}}[p^\infty ])$, the group of homomorphisms which are fixed by the involution
$\tau \mapsto \tau ^\vee$.
Proof. Suppose $\mathrm {char}(k)=p$ and write
$W$ for the ring of Witt vectors of
${\bar {k}}$. The crystalline cohomology groups of an abelian variety are torsion free by [Reference Berthelot, Breen and MessingBBM82, Corollary 2.5.5]. Therefore, thanks to the Künneth formula, [Reference BerthelotBer74, Theorem V.4.2.1], we have that
${H^*_{{\mathrm {crys}}}(A_{\bar {k}}\times _{\bar {k}} A_{\bar {k}}/W)=H^*_{{\mathrm {crys}}}(A_{\bar {k}}/W)\otimes H^*_{{\mathrm {crys}}}(A_{\bar {k}}/W)}$ so that
$m:A\times _k A\to A$ induces a morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU45.png?pub-status=live)
In degree $2$ we get a morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU46.png?pub-status=live)
which, in turn, induces a morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU47.png?pub-status=live)
Write $\sigma : \bigwedge ^2 H^1_{{\mathrm {crys}}}(A_{\bar {k}}/W)\xrightarrow {\sim } H^2_{{\mathrm {crys}}}(A_{\bar {k}}/W)$ for the natural isomorphism induced by the cup product, as in [Reference Berthelot, Breen and MessingBBM82, Corollary 2.5.5]. For every
$v\in H^1_{{\mathrm {crys}}}(A_{\bar {k}}/W)$, the pullback
$m^*(v)$ is equal to
$\pi _1^*(v)+\pi _2^*(v)$. Therefore, the composition
$(m^*-\pi _1^*-\pi _2^*)\circ \sigma : \bigwedge ^2 H^1_{{\mathrm {crys}}}(A_{\bar {k}}/W)\hookrightarrow H^1_{{\mathrm {crys}}}(A_{\bar {k}}/W)^{\otimes 2}$ is equal to the natural embedding
$v\wedge w\mapsto v\otimes w- w\otimes v$. By [Reference Berthelot, Breen and MessingBBM82, Theorem 5.1.8], we have that the
$F$-crystal
$H^1_{{\mathrm {crys}}}(A_{\bar {k}}/W)$ over
${\bar {k}}$ is canonically isomorphic to
$H^1_{{\mathrm {crys}}}(A_{\bar {k}}^\vee /W)^\vee$ with
$F$-structure defined as the dual of the
$F$-structure of
$H^1_{{\mathrm {crys}}}(A_{\bar {k}}^\vee /W)$ multiplied by
$p$. Thus, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU48.png?pub-status=live)
where $\mathbf {F\textbf {-}Crys}({\bar {k}})$ is the category of
$F$-crystals over
${\bar {k}}$. By [Reference IllusieIll79, Rmq. II.3.11.2], the
$F$-crystals
$H^1_{{\mathrm {crys}}}(A_{\bar {k}}/W)$ and
$H^1_{{\mathrm {crys}}}(A_{\bar {k}}^\vee /W)$ are the contravariant crystalline Dieudonné modules of the
$p$-divisible groups
$A_{\bar {k}}[p^\infty ]$ and
$A^\vee _{\bar {k}}[p^\infty ]$, thus we get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU49.png?pub-status=live)
On the other hand, by [Reference IllusieIll79, Theorem II.5.14], there is a canonical isomorphism $H^2(A_{\bar {k}},{\mathbb {Z}_{p}}(1))=H^2_{{\mathrm {crys}}}(A_{\bar {k}}/W)^{F=p}$. This concludes the case when
$\mathrm {char}(k)=p$. If
$p$ is invertible in
$k$ one can replace crystalline cohomology with
$p$-adic étale cohomology.
5. Main results
We are now ready to prove our main result, which is a flat version of the Tate conjecture for divisors of abelian varieties.
Theorem 5.1 If $A$ is an abelian variety over a finitely generated field
$k$ of characteristic
$p>0$, then
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}})^k)=0$. Moreover, after possibly replacing
$k$ with a finite separable extension, the cycle class map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU50.png?pub-status=live)
becomes an isomorphism.
Proof. To prove the statement we may assume that $\mathrm {Pic}(A)\to \mathrm {NS}(A_{{\bar {k}}})$ is surjective by extending
$k$. The
${\mathbb {Z}_{p}}$-module
$\operatorname {Hom}(A[p^\infty ],A^\vee [p^\infty ])$ embeds into
$\operatorname {Hom}(A_{\bar {k}}[p^\infty ],A^\vee _{\bar {k}}[p^\infty ])$, therefore the morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU51.png?pub-status=live)
induces a morphism $\tilde {h}:H^2(A_{\bar {k}},\mathbb {Z}_p(1))^{k}\to \operatorname {Hom}(A[p^\infty ],A^\vee [p^\infty ])$. By Proposition 4.6, we know that
$\tilde {h}$ is injective and
$\mathrm {im}(\tilde {h})$ is contained in
$\operatorname {Hom}^{\mathrm {sym}}(A[p^\infty ],A^\vee [p^\infty ])$. In addition, by Proposition 4.5, we have the following commutative square.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU52.png?pub-status=live)
The lower arrow is an isomorphism by [Reference de JongdeJ98, Theorem 2.6], and since $\mathrm {NS}(A)=\operatorname {Hom}^{\mathrm {sym}}(A,A^\vee )$, we deduce that
$\mathrm {im}(\tilde {h}\circ c_1)=\operatorname {Hom}^{\mathrm {sym}}(A[p^\infty ],A^\vee [p^\infty ])$. This implies that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU53.png?pub-status=live)
is surjective, thus by Proposition 3.9 we get that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU54.png?pub-status=live)
is surjective. Combining this with Proposition 3.8, we deduce that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU55.png?pub-status=live)
is surjective. For the result about the Brauer group, we just note that by the previous argument and Proposition 3.8, the ${\mathbb {Z}_{p}}$-module
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))^k$ vanishes. Therefore, thanks to Proposition 3.9, we deduce that
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}})^k)=\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}})^k)[\frac {1}{p}]=0$.
Theorem 5.2 Let $A$ be an abelian variety over a finitely generated field
$k$ of characteristic
$p>0$. The transcendental Brauer group
$\mathrm {Br}(A_{{k_s}}\!)^{k}$ is a direct sum of a finite group and a finite exponent
$p$-group. In addition, if the Witt vector cohomology group
$H^2(A_{{\bar {k}}},W\mathcal {O}_{A_{\bar {k}}}\!)$ is a finite
$W({\bar {k}})$-module, then
$\mathrm {Br}(A_{{k_s}}\!)^{k}$ is finite.
Proof. By [Reference Colliot-Thélène and SkorobogatovCS21, Theorem 16.2.3], the group $\mathrm {Br}(A_{{k_s}}\!)^{k}[\frac {1}{p}]$ is finite. Moreover, thanks to Corollary 3.4, the morphism
$\mathrm {Br}(A_{{k_s}}\!)\to \mathrm {Br}(A_{{\bar {k}}})$ is injective, which implies that the transcendental Brauer group is the same as
$\mathrm {Br}(A_{{\bar {k}}})^{k}$. Write
$\mathbf {Ab}_p^\star \subseteq \mathbf {Ab}$ for the full subcategory of the category of abstract abelian groups with objects those (possibly infinite)
$p$-groups isomorphic to
$({\mathbb {Q}_{p}}/{\mathbb {Z}_{p}})^{\oplus a} \oplus M$ for some
$a\geq 0$ and
$M$ a finite exponent
$p$-group. Equivalently,
$\mathbf {Ab}_p^\star$ is the subcategory of those
$p$-groups
$M$ such that
$M[p^{n+1}]/M[p^n]$ is finite for
$n$ big enough. Note that this subcategory is closed under the operation of taking subobjects, quotients, and finite direct sums. We first want to prove that
$H:=\varinjlim _n H^2(A_{\bar {k}},{\mu _{p^n}}\!)\in \mathbf {Ab}_p^\star$ and when
$H^2(A_{{\bar {k}}},W\mathcal {O}_{A_{\bar {k}}}\!)$ is a finite
$W({\bar {k}})$-module then, in addition,
$H[p]$ is finite (so that
$H[p^n]$ is also finite for every
$n\geq 0$). Thanks to the Kummer exact sequence this implies the same result for
$\mathrm {Br}(A_{{\bar {k}}})^{k}[p^\infty ]$.
Write $q: A_{\bar {k}} \to \operatorname {Spec}({\bar {k}})$ for the structural morphism. By [Reference Bragg and OlssonBO21, Corollary 1.4], for every
$n$ there exists a commutative linear algebraic group
$G_n$ representing
$R^2q_{*}\mu _{p^n}$.Footnote 6 Write
$U_n$ for the unipotent radical of
$G_n$ and
$D_n$ for the reductive quotient
$G_n/U_n$. Since
$G_n$ is commutative, there is a canonical Levi decomposition
$G_n=U_n\times D_n$. In particular, we have that
$H=U\times D$, where
$U:=\varinjlim _nU_n({\bar {k}})$ and
$D:=\varinjlim _nD_n({\bar {k}})$. For every
$n>0$, the group scheme
$D_n$ is finite, because it is a reductive group killed by
$p^n$. In addition, by [Reference Bragg and OlssonBO21, Proposition 10.7], there is a canonical isomorphism of formal groups
$\varinjlim _n \hat {G}_n=\varinjlim _n \hat {U}_n=\Phi ^2_{\mathrm {fl}}(A_{\bar {k}},\mathbb {G}_m)$, where
$\hat {G}_n$ and
$\hat {U}_n$ are the formal completions at the identity of
$G_n$ and
$U_n$ and
$\Phi ^2_{\mathrm {fl}}(-,-)$ is as in [Reference Bragg and OlssonBO21, § 10.6].
Applying $Rq_*$ to the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqn6.png?pub-status=live)
and using the fact that $\varinjlim _n R^1q_{*}\mu _{p^n}=\mathrm {Pic}_{A_{{\bar {k}}}/{\bar {k}}}[p^\infty ]$ is a
$p$-divisible group, we get the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU56.png?pub-status=live)
As a first consequence, we deduce that for every $n>0$ the group scheme
$D_n$ is the same as
$D_{n+1}[p^n]$, thus
$D[p]=D_1({\bar {k}})$ is finite. In particular, the abstract group
$D$ is in
$\mathbf {Ab}_p^\star$. To bound
$U$, we note that by [Reference MilneMil86, Proposition 3.1] the dimension of the chain of algebraic groups
$G_1\subseteq G_2\subseteq \cdots$ is eventually constant. Therefore, there exists
$N>0$ such that for every
$n\geq N$, the morphism
$(U_n)_{\mathrm {red}}\to (U_{n+1})_{\mathrm {red}}$ is an isomorphism. This shows that
$U$ is a finite exponent
$p$-group.
If $H^2(A_{\bar {k}},W\mathcal {O}_{A_{\bar {k}}}\!)$ is a finite
$W({\bar {k}})$-module, then the formal group
$\Phi ^2_{\mathrm {fl}}(A_{\bar {k}},\mathbb {G}_m)$ does not contain any copy of
$\hat {\mathbb {G}}_a$. Indeed, by [Reference Bragg and OlssonBO21, Corollary 12.5], the group
$H^2(A_{\bar {k}},W\mathcal {O}_{A_{\bar {k}}}\!)$ is the Cartier module of
$\Phi ^2_{\mathrm {fl}}(A_{\bar {k}},\mathbb {G}_m)$ and, by the assumption, it cannot contain
${\bar {k}}[[V]]$, the Cartier module of
$\hat {\mathbb {G}}_a$. Therefore, in this case, we have that each group
$U_n({\bar {k}})$ is trivial, so that
$H[p]=D[p]$ is finite.
We can finally prove that $\mathrm {Br}(A_{{\bar {k}}})^{k}[p^\infty ]$ has finite exponent. Suppose by contradiction that this is not the case. Since
$\mathrm {Br}(A_{{\bar {k}}})^{k}[p^\infty ]\in \mathbf {Ab}_p^\star$, we deduce that it contains a copy of
${\mathbb {Q}_{p}}/{\mathbb {Z}_{p}}$. On the other hand, by Theorem 5.1, the group
$\mathrm {T}_p(\mathrm {Br}(A_{{\bar {k}}})^{k})$ vanishes, which leads to a contradiction.
Corollary 5.3 The group $\mathrm {Br}(A_{{k_s}}\!)^{\Gamma _k}$ has finite exponent.
Proof. This follows from Theorem 5.2 thanks to [Reference Colliot-Thélène and SkorobogatovCS21, Theorem 5.4.12].
We end this section with some examples of abelian varieties over finitely generated fields with infinite transcendental Brauer group. Let $E$ be a supersingular elliptic curve over an infinite finitely generated field
$k$ and let
$A$ be the product
$E\times _k E$.
Proposition 5.4 After possibly extending $k$ to a finite separable extension, the transcendental Brauer group
$\mathrm {Br}(A_{{k_s}}\!)^k$ becomes infinite.
Proof. Even in this case we use that, thanks to Corollary 3.4, the transcendental Brauer group is the same as $\mathrm {Br}(A_{{\bar {k}}})^{k}$. Moreover, after extending the scalars we may assume that the morphism
$\mathrm {Pic}(A)\to \mathrm {NS}(A_{{\bar {k}}})$ is surjective. Combining Proposition 3.8 and the fact that
$\mathrm {NS}(A_{{\bar {k}}})/p$ is finite we deduce that it is enough to show that
$H^2(A_{{\bar {k}}},\mu _p)^k$ is infinite. We look at the Leray spectral sequence with respect to the second projection
$\pi _2:A=E\times _k E\to E$ (both over
$k$ and over
${\bar {k}}$). In the second page, we have that the boundary morphism
$H^1(E,R^1\pi _{2*} \mu _p) \to H^3 (E,\pi _{2*}\mu _p)$ vanishes because
$H^3 (E,\pi _{2*}\mu _p)\to H^3(A,\mu _p)$ admits a retraction induced by the zero section of
$\pi _2$. Since
$H^0(E_{\bar {k}},R^2\pi _{2*}\mu _p)=H^2(E_{\bar {k}},\mu _p)=\mathbb {Z}/p$, it is then enough to show that the image of
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU57.png?pub-status=live)
is infinite. By Lemma 4.3, we have that $\operatorname {End}(E[p])$ (respectively,
$\operatorname {End}(E_{{\bar {k}}}[p])$) admits a natural embedding in
$H^1(E,E[p])$ (respectively,
$H^1(E_{{\bar {k}}},E_{{\bar {k}}}[p])$). Since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU58.png?pub-status=live)
by the assumption that $E$ is supersingular, we deduce the desired result.
6. Specialisation of Néron–Severi groups
6.1
We want to start this section with an explicative example. Let $\mathcal {E}\to X$ be a non-isotrivial family of ordinary elliptic curves, where
$X$ is a connected normal scheme of finite type over
$\mathbb {F}_p$. Let
$\mathcal {A}$ be the fibred product
$\mathcal {E}\times _X \mathcal {E}$. We denote by
$E$ and
$A$ the generic fibres over the generic point
$\operatorname {Spec}(k)\hookrightarrow X$. The Kummer exact sequence induces the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU59.png?pub-status=live)
The group $\mathrm {NS}(A_{\bar {k}})_{\mathbb {Z}_p}$ is of rank
$2+\operatorname {rk}_\mathbb {Z}(\operatorname {End}(E_{\bar {k}}))=3$, whereas, by Proposition 4.6, the rank of
$H^2_{}(A_{\bar {k}},\mathbb {Z}_p(1))$ is
$2+\operatorname {rk}_{\mathbb {Z}_{p}}(\operatorname {End}(E_{\bar {k}}[p^\infty ]))=4$. This shows that
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))$ is of rank
$1$. The endomorphisms of
$E_{{\bar {k}}}[p^\infty ]$ are all defined over
${k_i}$, which implies that the action of
$\Gamma _k$ on
$H^2_{}(A_{\bar {k}},\mathbb {Z}_p(1))$ is trivial. In particular, the morphism
$\mathrm {NS}(A)_{\mathbb {Z}_p} \to H^2_{}(A_{\bar {k}},\mathbb {Z}_p(1))^{\Gamma _k}$ is not surjective and the cokernel
$\mathrm {T}_p(\mathrm {Br}(A_{\bar {k}}))^{\Gamma _k}$ is isomorphic to
${\mathbb {Z}_{p}}$.
In this case, the Galois action on $H^2_{}(A_{\bar {k}},\mathbb {Z}_p(1))$ is not enough to detect what classes are
${\mathbb {Z}_{p}}$-linear combinations of algebraic cycles. There is an additional obstruction to descend cohomology classes through the purely inseparable extension
${k_i}/k$. This extra purely inseparable obstruction gives an explanation of the failure of surjectivity of specialisation morphisms of Néron–Severi groups. In the example, if
$\operatorname {Spec}(\kappa )\hookrightarrow X$ is a closed point, we have that
$\mathrm {NS}(\mathcal {A}_{\kappa })=\mathrm {NS}(\mathcal {A}_{\bar {\kappa }})$ is of rank
$4$ because
$\operatorname {End}(\mathcal {E}_\kappa )=\operatorname {End}(\mathcal {E}_{\bar \kappa })$ is of rank
$2$ (there is an extra Frobenius endomorphism). Thus, the specialisation map
$\mathrm {NS}(A)\hookrightarrow \mathrm {NS}(\mathcal {A}_\kappa )$ is never surjective even if the rank of
$H^2_{}(A_{\bar {\kappa }},\mathbb {Z}_p(1))$ is
$4$ as the generic geometric fibre and the Galois action is trivial in both cases. One can interpret this failure by saying that the extra obstruction on
$H^2_{}(A_{\bar {k}},\mathbb {Z}_p(1))$ coming from the purely inseparable extension
${k_i}/k$ is trivial on
$H^2_{}(A_{\bar {\kappa }},\mathbb {Z}_p(1))$ since
$\kappa$ is perfect. In general, we prove the following theorem.
Theorem 6.2 Let $X$ be a connected normal scheme of finite type over
$\mathbb {F}_p$ with generic point
$\eta =\operatorname {Spec}(k)$ and let
$f:\mathcal {A}\to X$ be an abelian scheme over
$X$ with constant Newton polygon. For every closed point
$x=\operatorname {Spec}(\kappa )$ of
$X$ we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU60.png?pub-status=live)
Remark 6.3 Note that after replacing $X$ with a finite étale cover the action of
$\Gamma _k$ on
$\mathrm {NS}(\mathcal {A}_{\bar {\eta }})$ is trivial. Thus, we also get an inequality before taking Galois-fixed points.
To prove Theorem 6.2 we first need the following result.
Proposition 6.4 Under the assumptions of Theorem 6.2, the functor $\mathcal {F}$ which sendsFootnote 7
$T\in (X^{\mathrm {perf}})_{\mathrm {pro}\mathrm {\acute {e}t}}$ to
$\operatorname {Hom}^{{\mathrm {sym}}}(\mathcal {A}_T[p^\infty ],\mathcal {A}^\vee _T[p^\infty ])[\frac {1}{p}]$ is a semi-simple finite-rank
${\mathbb {Q}_{p}}$-local system such that for every
$\bar {x}\in X(\bar {\mathbb {F}}_p)$ we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU61.png?pub-status=live)
Proof. Let $\widetilde {\mathcal {F}}$ be the functor which sends
$T\in (X^{\mathrm {perf}})_{\mathrm {pro}\mathrm {\acute {e}t}}$ to
$\operatorname {Hom}(\mathcal {A}_T[p^\infty ],\mathcal {A}^\vee _T[p^\infty ])[\frac {1}{p}]$. We first note that to prove the result we can replace
$\mathcal {F}$ with
$\widetilde {\mathcal {F}}$ since
$\mathcal {F}$ is the kernel of the
${\mathbb {Q}_{p}}$-linear endomorphism
$\alpha -\operatorname {id}_{\widetilde {\mathcal {F}}}:\widetilde {\mathcal {F}}\to \widetilde {\mathcal {F}}$ where
$\alpha$ sends
$\tau \in \operatorname {Hom}(\mathcal {A}_T[p^\infty ],\mathcal {A}^\vee _T[p^\infty ])[\frac {1}{p}]$ to
$\tau ^\vee$. Write
$\mathbf {F\textrm {-}Crys}(X)$ for the category of
$F$-crystals over the absolute crystalline site of
$X$ and let
$\mathcal {M}_1,\mathcal {M}_2\in \mathbf {F\textrm {-}Crys}(X)$ be the contravariant crystalline Dieudonné modules of
$\mathcal {A}[p^\infty ]$ and
$\mathcal {A}^\vee [p^\infty ]$ over
$X$ constructed in [Reference Berthelot, Breen and MessingBBM82, Déf. 3.3.6]. By [Reference Berthelot, Breen and MessingBBM82, Theorem 5.1.8], we have that
$\mathcal {M}_1= \mathcal {M}_2^\vee (-1)$ where
$\mathcal {M}_2^\vee (-1)$ is the
$F$-crystal
$\mathcal {H}om(\mathcal {M}_2,\mathcal {O}_{X,\mathrm {cris}})$ endowed with the dual of the
$F$-structure of
$\mathcal {M}_2$ multiplied by
$p$. Thus
$\mathcal {M}_1^{\otimes 2}$ is equal to
$\mathcal {H}om(\mathcal {M}_2,\mathcal {M}_1)$ endowed with the natural
$F$-structure multiplied by
$p$. By [Reference LauLau13, Theorem D], for every perfect scheme
$T\to X$ we have canonical isomorphisms
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU62.png?pub-status=live)
where $\mathcal {M}_{1,T}$ and
$\mathcal {M}_{2,T}$ are the inverse images of
$\mathcal {M}_1$ and
$\mathcal {M}_2$ to
$T$. These isomorphisms are equivariant with respect to the action of the abstract group
$\operatorname {Aut}(T/X)$.
By [Reference KatzKat79, Theorem 2.5.1], the slope filtration of the $F$-crystal
$\mathcal {M}_{1,X^{\textrm {perf}}}^{\otimes 2}$ (which exists since
$\mathcal {A}\to X$ has constant Newton polygon) splits uniquely up to isogeny. We denote by
$\mathcal {N}^{[1]}_{X^\textrm {perf}}$ the slope
$1$ subobject of
$\mathcal {M}_{1,X^{\textrm {perf}}}^{\otimes 2}$, defined up to isogeny. Note that for every
$T\to X^\textrm {perf}$ we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU63.png?pub-status=live)
By construction, the $F$-crystal
$\mathcal {N}^{[1]}_{X^{\textrm {perf}}}(1)$ is unit-root. Therefore, by [Reference KatzKat73, Proposition 4.1.1], we deduce that
$\widetilde {\mathcal {F}}$ is a
${\mathbb {Q}_{p}}$-local system. In addition, by [Reference KatzKat73, Lemma 4.3.15], for every
$S=\operatorname {Spec}(R)\to X$ with
$R$ strictly henselian perfect ring we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU64.png?pub-status=live)
where $s$ is the closed point of
$S$. This implies that for every
$\bar {x}\in X(\bar {\mathbb {F}}_p)$ we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU65.png?pub-status=live)
For the semi-simplicity, since $X$ is normal, we can shrink
$X$ and assume it smooth. Write
$\mathcal {N}$ for the
$F$-isocrystal
$(R^1f_{{\mathrm {crys}}*}\mathcal {O}_{\mathcal {A},{\mathrm {crys}}})^{\otimes 2}$ and
$\mathcal {N}^{[1]}$ for the quotient
$\mathcal {N}^{\leq 1}/\mathcal {N}^{<1}$, where
$\mathcal {N}^{\leq 1}$ (respectively,
$\mathcal {N}^{<1}$) is the subobject of
$\mathcal {N}$ of slopes
$\leq 1$ (respectively,
$<1$). Note that by [Reference Berthelot, Breen and MessingBBM82, Theorem 2.5.6(ii)], the pullback of
$\mathcal {N}^{[1]}$ to
$X^{\mathrm {perf}}$ is isomorphic as an
$F$-isocrystal with
$\mathcal {N}^{[1]}_{X^\textrm {perf}}$ over
$X^{\textrm {perf}}$ defined above. Thanks to [Reference D'AddezioD'Ad23, Theorem 1.1.2], we have that
$\mathcal {N}^{[1]}$ is semi-simple as an
$F$-isocrystal.
By [Reference CrewCre87, Theorem 2.1], there is an equivalence between unit-root $F$-isocrystals over
$X$ and finite-rank
${\mathbb {Q}_{p}}$-local systems. By construction, Crew's and Katz's correspondences are compatible, in the sense that they agree after pulling back the objects through
$X^\mathrm {perf}\to X$. Since the étale fundamental groups of
$X$ and
$X^{\mathrm {perf}}$ are canonically isomorphic, we deduce that
$\widetilde {\mathcal {F}}$ is semi-simple as well. This yields the desired result.
6.5
Proof of Theorem 6.2 We look at the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU66.png?pub-status=live)
Thanks to Proposition 6.4, the operation of taking Galois-fixed points is exact. We get the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU67.png?pub-status=live)
Looking at the ranks we deduce the following equality
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqn7.png?pub-status=live)
By Proposition 6.4, the action of $\Gamma _k$ on
$H^2(A_{{\bar {k}}},{\mathbb {Q}_{p}}(1))=\operatorname {Hom}^{\mathrm {sym}}(A_{\bar {k}}[p^\infty ],A^\vee _{\bar {k}}[p^\infty ])[\frac {1}{p}]$ factors through the étale fundamental group of
$X$ associated to
$\bar {\eta }$, denoted by
$\pi _1^{\mathrm {\acute {e}t}}(X,\bar {\eta })$. In addition, if
$\kappa$ is the residue field of
$x$, the inclusion
$x\hookrightarrow X$ induces then an action of
$\Gamma _\kappa$ on
$H^2(A_{{\bar {k}}},{\mathbb {Q}_{p}}(1))$ which corresponds, up to conjugation, to the action of
$\Gamma _\kappa$ on
$H^2(A_{\bar {\kappa }},{\mathbb {Q}_{p}}(1))$. Therefore, by the Tate conjecture over finite fields (or Corollary 5.3), we get
$\mathrm {NS}(\mathcal {A}_{\bar {x}})^{\Gamma _\kappa }_{\mathbb {Q}_{p}}=H^2(A_{{\bar {k}}},{\mathbb {Q}_{p}}(1))^{\Gamma _\kappa }$. Since
$H^2(A_{{\bar {k}}},{\mathbb {Q}_{p}}(1))^{\Gamma _k}$ is a subspace of
$H^2(A_{{\bar {k}}},{\mathbb {Q}_{p}}(1))^{\Gamma _\kappa }$ we deduce that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqn8.png?pub-status=live)
We want to conclude this section with other examples of abelian varieties such that $\mathrm {T}_p(\mathrm {Br}(A_{{\bar {k}}}))^{\Gamma _k}\neq 0$. These are variants of the abelian surface of § 6.1 and they all provide counterexamples to the conjecture in [Reference UlmerUlm14, § 7.3.1] when
$\ell =p$.
Proposition 6.6 Let $A$ be an abelian variety which splits as a product
$B\times _k B$ with
$B$ an abelian variety over
$k$. There is a natural exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU68.png?pub-status=live)
Proof. We consider the exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU69.png?pub-status=live)
Arguing as in the proof of Proposition 4.6, the ${\mathbb {Z}_{p}}$-module
$\operatorname {Hom}(B_{\bar {k}}[p^\infty ],B_{\bar {k}}^\vee [p^\infty ])^{\Gamma _k}$ is naturally a direct summand of
$H^2(A_{{\bar {k}}}, {\mathbb {Z}_{p}}(1))^{\Gamma _k}$. Its preimage in
$\mathrm {NS}(A_{{\bar {k}}})_{\mathbb {Z}_{p}}^{\Gamma _k}$ corresponds to the
${\mathbb {Z}_{p}}$-module
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240119031445472-0754:S0010437X23007558:S0010437X23007558_eqnU70.png?pub-status=live)
This concludes the proof.
Corollary 6.7 If $\operatorname {End}(B)=\mathbb {Z}$, then
$\mathrm {T}_p(\mathrm {Br}(A_{{\bar {k}}}))^{\Gamma _k}\neq 0$.
Proof. By the assumption, $\operatorname {Hom}(B,B^\vee )_{\mathbb {Z}_{p}}$ is a
${\mathbb {Z}_{p}}$-module of rank
$1$. Therefore, by Proposition 6.6, it is enough to prove that the rank of
$\operatorname {Hom}(B_{\bar {k}}[p^\infty ],B_{\bar {k}}^\vee [p^\infty ])^{\Gamma _k}$ is greater than
$1$. Since
$\operatorname {End}(B)=\mathbb {Z}$, the abelian variety
$B$ is not supersingular, so that the
$p$-divisible group
$B[p^\infty ]$ admits at least two slopes. By the Dieudonné–Manin classification, this implies that
$B_{{k_i}}[p^\infty ]$ is isogenous to a direct sum
$\mathcal {G}_1\oplus \mathcal {G}_2$ of non-zero
$p$-divisible groups over
${k_i}$. Since
$\operatorname {End}(\mathcal {G}_1)[\frac {1}{p}]\oplus \operatorname {End}(\mathcal {G}_2)[\frac {1}{p}]$ embeds into
$\operatorname {End}(B_{{k_i}}[p^\infty ])[\frac {1}{p}]\simeq \operatorname {Hom}(B_{\bar {k}}[p^\infty ],B_{\bar {k}}^\vee [p^\infty ])[\frac {1}{p}]^{\Gamma _k}$, we deduce that
$\operatorname {rk}_{\mathbb {Z}_{p}}(\operatorname {Hom}(B_{\bar {k}}[p^\infty ],B_{\bar {k}}^\vee [p^\infty ])^{\Gamma _k})>1$, as we wanted.
Acknowledgements
I thank Emiliano Ambrosi for the discussions we had during the writing of [Reference Ambrosi and D'AddezioAD22], which inspired this article, Matthew Morrow and Kay Rülling for many enlightening conversations about the cohomology of $\mathbb {Z}_p(1)$, and Ofer Gabber, Luc Illusie, Peter Scholze, and Takashi Suzuki for answering some questions on the fppf site. I also thank Jean-Louis Colliot-Thélène, Bruno Kahn, Alexei Skorobogatov, and Takashi Suzuki for very useful comments on a first draft of this article. Finally, I thank the anonymous referees for their careful reading of the article and for the corrections they suggested.
The author was funded by the Deutsche Forschungsgemeinschaft (EXC-2046/1, project ID 390685689 and DA-2534/1-1, project ID 461915680) and by the Max-Planck Institute for Mathematics.
Conflicts of Interest
None.