1. Main theorem
Let $p$ be a prime number. Rapoport–Zink spaces [Reference Rapoport and ZinkRZ96] are deformation spaces of
$p$-divisible groups equipped with some extra structure. This article concerns the geometry of Rapoport–Zink spaces of EL type (endomorphisms
$+$ level structure). In particular we consider the infinite-level spaces
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$, which are preperfectoid spaces [Reference Scholze and WeinsteinSW13]. An example is the space
${{\mathscr M}}_{H,\infty }$, where
$H/\overline {{{\mathbf {F}}}}_p$ is a
$p$-divisible group of height
$n$. The points of
${{\mathscr M}}_{H,\infty }$ over a nonarchimedean field
$K$ containing
$W(\overline {{{\mathbf {F}}}}_p)$ are in correspondence with isogeny classes of
$p$-divisible groups
$G/\mathcal {O}_K$ equipped with a quasi-isogeny
$G\otimes _{\mathcal {O}_K} \mathcal {O}_K/p \to H\otimes _{\overline {{{\mathbf {F}}}}_p} \mathcal {O}_K/p$ and an isomorphism
$\mathbf {Q}_p^{n}\cong VG$ (where
$VG$ is the rational Tate module).
The infinite-level space ${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ appears as the limit of finite-level spaces, each of which is a smooth rigid-analytic space. We would like to investigate the question of smoothness for the space
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ itself, which is quite a different matter. We need the notion of cohomological smoothness [Reference ScholzeSch17], which makes sense for general morphisms of analytic adic spaces, and which is reviewed in § 4. Roughly speaking, an adic space is cohomologically smooth over
$C$ (where
$C/\mathbf {Q}_p$ is complete and algebraically closed) if it satisfies local Verdier duality. In particular, if
$U$ is a quasi-compact adic space which is cohomologically smooth over
$\operatorname {Spa} (C,\mathcal {O}_C)$, then the cohomology group
$H^{i}(U,{{\mathbf {F}}}_\ell )$ is finite for all
$i$ and all primes
$\ell \neq p$.
Our main theorem shows that each connected component of the geometric fiber of ${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ has a dense open subset which is cohomologically smooth.
Theorem 1.0.1 Let ${{\mathcal {D}}}$ be a basic EL datum (cf. § 2). Let
$C$ be a complete algebraically closed extension of the field of scalars of
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$, and let
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\circ }$ be a connected component of the base change
${{\mathscr M}}_{{{\mathcal {D}}},\infty ,C}$. Let
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\circ ,\operatorname {non-sp}}\subset {{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\circ }$ be the nonspecial locus (cf. § 3.5), corresponding to
$p$-divisible groups without extra endomorphisms. Then
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\circ ,\operatorname {non-sp}}$ is cohomologically smooth over
$C$.
We remark that outside of trivial cases, $\pi _0({{\mathscr M}}_{{{\mathcal {D}}},\infty ,C})$ has no isolated points, which implies that no open subset of
${{\mathscr M}}_{{{\mathcal {D}}},\infty ,C}$ can be cohomologically smooth. (Indeed, the
$H^{0}$ of any quasi-compact open fails to be finitely generated.) Therefore it really is necessary to work with individual connected components of the geometric fiber of
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$.
Theorem 1.0.1 is an application of the perfectoid version of the Jacobian criterion for smoothness, due to Fargues and Scholze [Reference Fargues and ScholzeFS]; cf. Theorem 4.2.1. The latter theorem involves the Fargues–Fontaine curve $X_C$ (reviewed in § 3). It asserts that a functor
${{\mathscr M}}$ on perfectoid spaces over
$\operatorname {Spa}(C,\mathcal {O}_C)$ is cohomologically smooth, when
${{\mathscr M}}$ can be interpreted as global sections of a smooth morphism
$Z\to X_C$, subject to a certain condition on the tangent bundle
$\operatorname {Tan}_{Z/X_C}$.
In our application to Rapoport–Zink spaces, we construct a smooth morphism $Z\to X_C$, whose moduli space of global sections is isomorphic to
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\circ }$ (Lemma 5.2.1). Next, we show that a geometric point
$x \in {{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\circ }(C)$ lies in
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\circ ,\textrm {non-sp}}(C)$ if and only if the corresponding section
$s \colon X_C \rightarrow Z$ satisfies the condition that all slopes of the vector bundle
$s^{\ast }\operatorname {Tan}_{Z/X_C}$ on
$X_C$ are positive (Theorem 5.5.1). This is exactly the condition on
$\operatorname {Tan}_{Z/X_C}$ required by Theorem 4.2.1, so we can conclude that
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\circ }$ is cohomologically smooth.
The geometry of Rapoport–Zink spaces is related to the geometry of Shimura varieties. As an example, consider the tower of classical modular curves $X(p^{\infty })$, considered as rigid spaces over
$C$. There is a perfectoid space
$X(p^{\infty })$ over
$C$ for which
$X(p^{\infty })\sim \varprojlim _n X(p^{n})$, and a Hodge–Tate period map
$\pi _{HT}\colon X(p^{\infty })\to {{\mathbf {P}}}^{1}_C$ [Reference ScholzeSch15], which is
$\operatorname {GL}_2(\mathbf {Q}_p)$-equivariant. Let
$X(p^{\infty })^{\circ }\subset X(p^{\infty })$ be a connected component.
Corollary 1.0.2 The following are equivalent for a $C$-point
$x$ of
$X(p^{\infty })^{\circ }$.
(i) The point
$x$ corresponds to an elliptic curve
$E$, such that the
$p$-divisible group
$E[p^{\infty }]$ has
$\operatorname {End} E[p^{\infty }]=\mathbf {Z}_p$.
(ii) The stabilizer of
$\pi _{HT}(x)$ in
$\operatorname {PGL}_2(\mathbf {Q}_p)$ is trivial.
(iii) There is a neighborhood of
$x$ in
$X(p^{\infty })^{\circ }$ which is cohomologically smooth over
$C$.
2. Review of Rapoport–Zink spaces at infinite level
2.1 The infinite-level Rapoport–Zink space
${{\mathscr M}}_{H,\infty }$
Let $k$ be a perfect field of characteristic
$p$, and let
$H$ be a
$p$-divisible group of height
$n$ and dimension
$d$ over
$k$. We review here the definition of the infinite-level Rapoport–Zink space associated with
$H$.
First there is the formal scheme ${{\mathscr M}}_H$ over
$\operatorname {Spf} W(k)$ parametrizing deformations of
$H$ up to isogeny, as in [Reference Rapoport and ZinkRZ96]. For a
$W(k)$-algebra
$R$ in which
$p$ is nilpotent,
${{\mathscr M}}_H(R)$ is the set of isomorphism classes of pairs
$(G,\rho )$, where
$G/R$ is a
$p$-divisible group and
$\rho \colon H\otimes _k R/p \to G\otimes _R R/p$ is a quasi-isogeny.
The formal scheme ${{\mathscr M}}_H$ locally admits a finitely generated ideal of definition. Therefore it makes sense to pass to its adic space
${{\mathscr M}}_H^{\operatorname{ad}}$, which has generic fiber
$({{\mathscr M}}_H^{\operatorname{ad}})_\eta$, a rigid-analytic space over
$\operatorname {Spa}(W(k)[1/p],W(k))$. Then
$({{\mathscr M}}_H^{\operatorname{ad}})_{\eta }$ has the following moduli interpretation: it is the sheafification of the functor assigning to a complete affinoid
$(W(k)[1/p],W(k))$-algebra
$(R,R^{+})$ the set of pairs
$(G,\rho )$, where
$G$ is a
$p$-divisible group defined over an open and bounded subring
$R_0\subset R^{+}$, and
$\rho \colon H\otimes _k R_0/p\to G\otimes _{R_0} R_0/p$ is a quasi-isogeny. There is an action of
$\operatorname {Aut} H$ on
${{\mathscr M}}_H^{\operatorname{ad}}$ obtained by composition with
$\rho$.
Given such a pair $(G,\rho )$, Grothendieck–Messing theory produces a surjection
$M(H)\otimes _{W(k)} R \to \operatorname {Lie} G[1/p]$ of locally free
$R$-modules, where
$M(H)$ is the covariant Dieudonné module. There is a Grothendieck–Messing period map
$\pi _{GM}\colon ({{\mathscr M}}_H^{\operatorname{ad}})_{\eta }\to {{\mathcal {F}}}\ell$, where
${{\mathcal {F}}}\ell$ is the rigid-analytic space parametrizing rank
$d$ locally free quotients of
$M(H)[1/p]$. The morphism
$\pi _{GM}$ is equivariant for the action of
$\operatorname {Aut} H$. It has open image
${{\mathcal {F}}}\ell ^{a}$ (the admissible locus).
We obtain a tower of rigid-analytic spaces over $({{\mathscr M}}_H^{\operatorname{ad}})_{\eta }$ by adding level structures. For a complete affinoid
$(W(k)[1/p],W(k))$-algebra
$(R,R^{+})$, and an element of
$({{\mathscr M}}_H^{\operatorname{ad}})_{\eta }(R,R^{+})$ represented locally on
$\operatorname {Spa}(R,R^{+})$ by a pair
$(G,\rho )$ as above, we have the Tate module
$TG=\varprojlim _m G[p^{m}]$, considered as an adic space over
$\operatorname {Spa}(R,R^{+})$ with the structure of a
$\mathbf {Z}_p$-module [Reference Scholze and WeinsteinSW13, (3.3)]. Finite-level spaces
${{\mathscr M}}_{H,m}$ are obtained by trivializing the
$G[p^{m}]$; these are finite étale covers of
$({{\mathscr M}}_H^{\operatorname{ad}})_{\eta }$. The infinite-level space is obtained by trivializing all of
$TG$ at once, as in the following definition.
Definition 2.1.1 [SW13, Definition 6.3.3] Let ${{\mathscr M}}_{H,\infty }$ be the functor which sends a complete affinoid
$(W(k)[1/p],W(k))$-algebra
$(R,R^{+})$ to the set of triples
$(G,\rho ,\alpha )$, where
$(G,\rho )$ is an element of
$({{\mathscr M}}_H)^{\operatorname{ad}}_{\eta }(R,R^{+})$, and
$\alpha \colon \mathbf {Z}_p^{n} \to TG$ is a
$\mathbf {Z}_p$-linear map which is an isomorphism pointwise on
$\operatorname {Spa}(R,R^{+})$.
There is an equivalent definition in terms of isogeny classes of triples $(G,\rho ,\alpha )$, where this time
$\alpha \colon \mathbf {Q}_p^{n}\to VG$ is a trivialization of the rational Tate module. Using this definition, it becomes clear that
${{\mathscr M}}_{H,\infty }$ admits an action of the product
$\operatorname {GL}_n(\mathbf {Q}_p)\times \operatorname {Aut}^{0} H$, where
$\operatorname {Aut}^{0}$ means automorphisms in the isogeny category. Then the period map
$\pi _{GM}\colon {{\mathscr M}}_{H,\infty }\to {{\mathcal {F}}}\ell$ is equivariant for
$\operatorname {GL}_n(\mathbf {Q}_p)\times \operatorname {Aut}^{0} H$, where
$\operatorname {GL}_n(\mathbf {Q}_p)$ acts trivially on
${{\mathcal {F}}}\ell$.
We remark that ${{\mathscr M}}_{H,\infty }\sim \varprojlim _m {{\mathscr M}}_{H,m}$ in the sense of [Reference Scholze and WeinsteinSW13, Definition 2.4.1].
One of the main theorems of [Reference Scholze and WeinsteinSW13] is the following.
Theorem 2.1.2 The adic space ${{\mathscr M}}_{H,\infty }$ is a preperfectoid space.
This means that for any perfectoid field $K$ containing
$W(k)$, the base change
${{\mathscr M}}_{H,\infty }\times _{\operatorname {Spa}(W(k)[1/p],W(k))} \operatorname {Spa}(K,\mathcal {O}_K)$ becomes perfectoid after
$p$-adically completing.
We sketch here the proof of Theorem 2.1.2. Consider the ‘universal cover’ $\tilde {H}=\varprojlim _p H$ as a sheaf of
$\mathbf {Q}_p$-vector spaces on the category of
$k$-algebras. This has a canonical lift to the category of
$W(k)$-algebras [Reference Scholze and WeinsteinSW13, Proposition 3.1.3(ii)], which we continue to call
$\tilde {H}$. The adic generic fiber
$\tilde {H}^{\operatorname{ad}}_{\eta }$ is a preperfectoid space, as can be checked ‘by hand’: it is a product of the
$d$-dimensional preperfectoid open ball
$(\operatorname {Spa} W(k) [\kern-1pt[ {T_1^{1/p^{\infty }},\ldots ,T_d^{1/p^{\infty }}} ]\kern-1pt])_{\eta }$ by the constant adic space
$VH^{\mathop e\limits^{{\prime}}t}$, where
$H^{\mathop e\limits{{\prime}}t}$ is the étale part of
$H$. Given a triple
$(G,\rho ,\alpha )$ representing an element of
${{\mathscr M}}_{H,\infty }(R,R^{+})$, the quasi-isogeny
$\rho$ induces an isomorphism
$\tilde {H}^{\operatorname{ad}}_{\eta }\times _{\operatorname {Spa}(W(k)[1/p],W(k))} \operatorname {Spa}(R,R^{+})\to \tilde {G}^{\operatorname{ad}}_{\eta }$; composing this with
$\alpha$ gives a morphism
$\mathbf {Q}_p^{n}\to \tilde {H}^{\operatorname{ad}}_{\eta }(R,R^{+})$. We have therefore described a morphism
${{\mathscr M}}_{H,\infty } \to (\tilde {H}^{\operatorname{ad}}_{\eta })^{n}$.
Theorem 2.1.2 follows from the fact that the morphism ${{\mathscr M}}_{H,\infty }\to (\tilde {H}^{\operatorname{ad}})_{\eta }^{n}$ presents
${{\mathscr M}}_{H,\infty }$ as an open subset of a Zariski closed subset of
$(\tilde {H}^{\operatorname{ad}})_{\eta }^{n}$. We conclude this subsection by spelling out how this is done. We have a quasi-logarithm map
$\operatorname {qlog}_H\colon \tilde {H}^{\operatorname{ad}}_{\eta } \to M(H)[1/p] \otimes _{W(k)[1/p]} {{\mathbf {G}}}_a$ [Reference Scholze and WeinsteinSW13, Definition 3.2.3], a
$\mathbf {Q}_p$-linear morphism of adic spaces over
$\operatorname {Spa}(W(k)[1/p],W(k))$.
Now suppose $(G,\rho )$ is a deformation of
$H$ to
$(R,R^{+})$. The logarithm map on
$G$ fits into an exact sequence of
$\mathbf {Z}_p$-modules:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU1.png?pub-status=live)
After taking projective limits along multiplication-by-$p$, this turns into an exact sequence of
$\mathbf {Q}_p$-vector spaces,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU2.png?pub-status=live)
On the other hand, we have a commutative diagram.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU3.png?pub-status=live)
The lower horizontal map $M(H)\otimes _{W(k)} R\to \operatorname {Lie} G[1/p]$ is the quotient by the
$R$-submodule of
$M(H)\otimes _{W(k)} R$ generated by the image of
$VG(R,R^{+}) \to \tilde {G}_{\eta }^{\operatorname{ad}}(R,R^{+})\cong \tilde {H}_{\eta }^{\operatorname{ad}}(R,R^{+})\to M(H)\otimes _{W(k)} R$.
Now suppose we have a point of ${{\mathscr M}}_{H,\infty }(R,R^{+})$ represented by a triple
$(G,\rho ,\alpha )$. Then we have a
$\mathbf {Q}_p$-linear map
$\mathbf {Q}_p^{n}\to \tilde {H}_{\eta }^{\operatorname{ad}}(R,R^{+})\to M(H)\otimes _{W(k)} R$. The cokernel of its
$R$-extension
$R^{n}\to M(H)\otimes _{W(k)} R$ is a projective
$R$-module of rank
$d$, namely
$\operatorname {Lie} G[1/p]$. This condition on the cokernel allows us to formulate an alternate description of
${{\mathscr M}}_{H,\infty }$ which is independent of deformations.
Proposition 2.1.3 The adic space ${{\mathscr M}}_{H,\infty }$ is isomorphic to the functor which assigns to a complete affinoid
$(W(k)[1/p],W(k))$-algebra
$(R,R^{+})$ the set of
$n$-tuples
$(s_1,\ldots ,s_n)\in \tilde {H}^{\operatorname{ad}}_{\eta }(R,R^{+})^{n}$ such that the following conditions are satisfied.
(i) The quotient of
$M(H)\otimes _{W(k)} R$ by the
$R$-span of the
$\operatorname {qlog}(s_i)$ is a projective
$R$-module
$W$ of rank
$d$.
(ii) For all geometric points
$\operatorname {Spa}(C,\mathcal {O}_C)\to \operatorname {Spa}(R,R^{+})$, the sequence
is exact.\[ 0 \to \mathbf{Q}_p^{n}\stackrel{(s_1,\ldots,s_n)}{\to} \tilde{H}^{\operatorname{ad}}_{\eta}(C,\mathcal{O}_C)\to W\otimes_R C \to 0 \]
2.2 Infinite-level Rapoport–Zink spaces of EL type
This article treats the more general class of Rapoport–Zink spaces of EL type. We review these here.
Definition 2.2.1 Let $k$ be an algebraically closed field of characteristic
$p$. A rational EL datum is a quadruple
${{\mathcal {D}}}=(B,V, H,\mu )$, where:
–
$B$ is a semisimple
$\mathbf {Q}_p$-algebra;
–
$V$ is a finite
$B$-module;
–
$H$ is an object of the isogeny category of
$p$-divisible groups over
$k$, equipped with an action
$B\to \operatorname {End} H$;
–
$\mu$ is a conjugacy class of
$\overline {\mathbf {Q}}_p$-rational cocharacters
$\mathbf {G}_m\to {{\mathbf {G}}}$, where
${{\mathbf {G}}}/\mathbf {Q}_p$ is the algebraic group
$\operatorname {GL}_B(V)$.
These are subject to the following conditions.
– If
$M(H)$ is the (rational) Dieudonné module of
$H$, then there exists an isomorphism
$M(H)\cong V\otimes _{\mathbf {Q}_p} W(k)[1/p]$ of
$B\otimes _{\mathbf {Q}_p} W(k)[1/p]$-modules. In particular
$\dim V=\operatorname {ht} H$.
– In the weight decomposition of
$V\otimes _{\mathbf {Q}_p} \overline {\mathbf {Q}}_p\cong \bigoplus _{i\in \mathbf {Z}} V_i$ determined by
$\mu$, only weights 0 and 1 appear, and
$\dim V_0=\dim H$.
The reflex field $E$ of
${{\mathcal {D}}}$ is the field of definition of the conjugacy class
$\mu$. We remark that the weight filtration (but not necessarily the weight decomposition) of
$V\otimes _{\mathbf {Q}_p}\overline {\mathbf {Q}}_p$ may be descended to
$E$, and so we will be viewing
$V_0$ and
$V_1$ as
$B\otimes _{\mathbf {Q}_p} E$-modules.
The infinite-level Rapoport–Zink space ${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ is defined in [Reference Scholze and WeinsteinSW13] in terms of moduli of deformations of the
$p$-divisible group
$H$ along with its
$B$-action. It admits an alternate description along the lines of Proposition 2.1.3.
Proposition 2.2.2 [SW13, Theorem 6.5.4] Let ${{\mathcal {D}}}=(B,V,H,\mu )$ be a rational EL datum. Let
$\breve {E}=E\cdot W(k)$. Then
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ is isomorphic to the functor which inputs a complete affinoid
$(\breve {E},\mathcal {O}_{\breve {E}})$-algebra
$(R,R^{+})$ and outputs the set of
$B$-linear maps
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU5.png?pub-status=live)
subject to the following conditions.
– Let
$W$ be the quotient
Then\[ V\otimes_{\mathbf{Q}_p} R\stackrel{\operatorname{qlog}_H\circ s}{\longrightarrow} M(H)\otimes_{W(k)} R \to W \to 0. \]
$W$ is a finite projective
$R$-module, which locally on
$R$ is isomorphic to
$V_0\otimes _{E} R$ as a
$B\otimes _{\mathbf {Q}_p} R$-module.
– For any geometric point
$x=\operatorname {Spa}(C,\mathcal {O}_C)\to \operatorname {Spa}(R,R^{+})$, the sequence of
$B$-modules
is exact.\[ 0\to V \to \tilde{H}(\mathcal{O}_C) \to W\otimes_R C \to 0 \]
If ${{\mathcal {D}}} = (\mathbf {Q}_p, \mathbf {Q}_p^{n}, H, \mu )$, where
$H$ has height
$n$ and dimension
$d$ and
$\mu (t)=(t^{\oplus d},1^{\oplus (n-d)})$, then
$E=\mathbf {Q}_p$ and
${{\mathscr M}}_{{{\mathcal {D}}},\infty } = {{\mathscr M}}_{H,\infty }$.
In general, we call $\breve E$ the field of scalars of
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$, and for a complete algebraically closed extension
$C$ of
$\breve E$, we write
${{\mathscr M}}_{{{\mathcal {D}}},\infty ,C} = {{\mathscr M}}_{{{\mathcal {D}}},\infty } \times _{\operatorname {Spa}(\breve E,\mathcal {O}_{\breve E})} \operatorname {Spa}(C,\mathcal {O}_C)$ for the corresponding geometric fiber of
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$.
The space ${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ admits an action by the product group
${{\mathbf {G}}}(\mathbf {Q}_p)\times J(\mathbf {Q}_p)$, where
$J/\mathbf {Q}_p$ is the algebraic group
$\operatorname {Aut}_B^{\circ }(H)$. A pair
$(\alpha ,\alpha ')\in {{\mathbf {G}}}(\mathbf {Q}_p)\times J(\mathbf {Q}_p)$ sends
$s$ to
$\alpha '\circ s\circ \alpha ^{-1}$.
There is once again a Grothendieck–Messing period map $\pi _{GM}\colon {{\mathscr M}}_{{{\mathcal {D}}},\infty }\to {{\mathcal {F}}}\ell _\mu$ onto the rigid-analytic variety whose
$(R,R^{+})$-points parametrize
$B\otimes _{\mathbf {Q}_p} R$-module quotients of
$M(H)\otimes _{W(k)} R$ which are projective over
$R$, and which are of type
$\mu$ in the sense that they are (locally on
$R$) isomorphic to
$V_0\otimes _E R$. The morphism
$\pi _{GM}$ sends an
$(R,R^{+})$-point of
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ to the quotient
$W$ of
$M(H)\otimes _{W(k)} R$ as above. It is equivariant for the action of
${{\mathbf {G}}}(\mathbf {Q}_p)\times J(\mathbf {Q}_p)$, where
${{\mathbf {G}}}(\mathbf {Q}_p)$ acts trivially on
${{\mathcal {F}}}\ell _\mu$. In terms of deformations of the
$p$-divisible group
$H$, the period map
$\pi _{GM}$ sends a deformation
$G$ to
$\operatorname {Lie} G$.
There is also a Hodge–Tate period map $\pi _{HT}\colon {{\mathscr M}}_{{{\mathcal {D}}},\infty }\to {{\mathcal {F}}}\ell _{\mu }'$, where
${{\mathcal {F}}}\ell _\mu '(R,R^{+})$ parametrizes
$B\otimes _{\mathbf {Q}_p} R$-module quotients of
$V\otimes _{\mathbf {Q}_p} R$ which are projective over
$R$, and which are (locally on
$R$) isomorphic to
$V_1\otimes _E R$. The morphism
$\pi _{HT}$ sends an
$(R,R^{+})$-point of
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ to the image of
$V\otimes _{\mathbf {Q}_p} R\to M(H)\otimes _{W(k)} R$. It is equivariant for the action of
${{\mathbf {G}}}(\mathbf {Q}_p)\times J(\mathbf {Q}_p)$, where this time
$J(\mathbf {Q}_p)$ acts trivially on
${{\mathcal {F}}}\ell _\mu '(R,R^{+})$. In terms of deformations of the
$p$-divisible group
$H$, the period map
$\pi _{HT}$ sends a deformation
$G$ to
$(\operatorname {Lie} G^{\vee })^{\vee }$.
3. The Fargues–Fontaine curve
3.1 Review of the curve
We briefly review here some constructions and results from [Reference Fargues and FontaineFF18]. First we review the absolute curve, and then we cover the version of the curve which works in families.
Fix a perfectoid field $F$ of characteristic
$p$, with
$F^{\circ }\subset F$ its ring of integral elements. Let
$\varpi \in F^{\circ }$ be a pseudo-uniformizer for
$F$, and let
$k$ be the residue field of
$F$. Let
$W(F^{\circ })$ be the ring of Witt vectors, which we equip with the
$(p,[\varpi ])$-adic topology. Let
${{\mathcal {Y}}}_F=\operatorname {Spa}(W(F^{\circ }),W(F^{\circ }))\backslash \{ {\left \lvert {p[\varpi ]} \right \rvert =0} \}$. Then
${{\mathcal {Y}}}_F$ is an analytic adic space over
$\mathbf {Q}_p$. The Frobenius automorphism of
$F$ induces an automorphism
$\phi$ of
${{\mathcal {Y}}}_F$. Let
$B_F=H^{0}({{\mathcal {Y}}}_F,\mathcal {O}_{{{\mathcal {Y}}}_F})$, a
$\mathbf {Q}_p$-algebra endowed with an action of
$\phi$. Let
$P_F$ be the graded ring
$P_F=\bigoplus _{n\geq 0} B_F^{\phi =p^{n}}$. Finally, the Fargues–Fontaine curve is
$X_F=\operatorname {Proj} P_F$. It is shown in [Reference Fargues and FontaineFF18] that
$X_F$ is the union of spectra of Dedekind rings, which justifies the use of the word ‘curve’ to describe
$X_F$. Note however that there is no ‘structure morphism’
$X_F\to \operatorname {Spec} F$.
If $x\in X_F$ is a closed point, then the residue field of
$x$ is a perfectoid field
$F_x$ containing
$\mathbf {Q}_p$ which comes equipped with an inclusion
$i\colon F\hookrightarrow F_x^{\flat }$, which presents
$F_x^{\flat }$ as a finite extension of
$F$. Such a pair
$(F_x,i)$ is called an untilt of
$F$. Then
$x\mapsto (F_x,i)$ is a bijection between closed points of
$X_F$ and isomorphism classes of untilts of
$F$, modulo the action of Frobenius on
$i$. Thus if
$F=E^{\flat }$ is the tilt of a given perfectoid field
$E/\mathbf {Q}_p$, then
$X_{E^{\flat }}$ has a canonical closed point
$\infty$, corresponding to the untilt
$E$ of
$E^{\flat }$.
An important result in [Reference Fargues and FontaineFF18] is the classification of vector bundles on $X_F$. (By a vector bundle on
$X_F$ we are referring to a locally free
$\mathcal {O}_{X_F}$-module
${{\mathcal {E}}}$ of finite rank. We will use the notation
$V({{\mathcal {E}}})$ to mean the corresponding geometric vector bundle over
$X_F$, whose sections correspond to sections of
${{\mathcal {E}}}$.) Recall that an isocrystal over
$k$ is a finite-dimensional vector space
$N$ over
$W(k)[1/p]$ together with a Frobenius semilinear automorphism
$\phi$ of
$N$. Given
$N$, we have the graded
$P_F$-module
$\bigoplus _{n\geq 0} (N\otimes _{W(k)[1/p]} B_F)^{\phi =p^{n}}$, which corresponds to a vector bundle
${{\mathcal {E}}}_F(N)$ on
$X_F$. Then the Harder–Narasimhan slopes of
${{\mathcal {E}}}_F(N)$ are negative to those of
$N$. If
$F$ is algebraically closed, then every vector bundle on
$X_F$ is isomorphic to
${{\mathcal {E}}}_F(N)$ for some
$N$.
It is straightforward to ‘relativize’ the above constructions. If $S=\operatorname {Spa}(R,R^{+})$ is an affinoid perfectoid space over
$k$, one can construct the adic space
${{\mathcal {Y}}}_S$, the ring
$B_S$, the scheme
$X_S$, and the vector bundles
${{\mathcal {E}}}_S(N)$ as above. Frobenius-equivalences classes of untilts of
$S$ correspond to effective Cartier divisors of
$X_S$ of degree 1.
In our applications, we will start with an affinoid perfectoid space $S$ over
$\mathbf {Q}_p$. We will write
$X_S=X_{S^{\flat }}$, and we will use
$\infty$ to refer to the canonical Cartier divisor of
$X_{S}$ corresponding to the untilt
$S$ of
$S^{\flat }$. Thus if
$N$ is an isocrystal over
$k$, and
$S=\operatorname {Spa}(R,R^{+})$ is an affinoid perfectoid space over
$W(k)[1/p]$, then the fiber of
${{\mathcal {E}}}_{S}(N)$ over
$\infty$ is
$N\otimes _{W(k)[1/p]} R$.
Let $S = \operatorname {Spa}(R,R^{+})$ be as above and let
$\infty$ be the corresponding Cartier divisor. We denote the completion of the ring of functions on
$\mathcal {Y}_S$ along
$\infty$ by
$B_{\operatorname {dR}}^{+}(R)$. It comes equipped with a surjective homomorphism
$\theta \colon B_{\operatorname {dR}}^{+}(R) \rightarrow R$, whose kernel is a principal ideal
$\ker (\theta ) = (\xi )$.
3.2 Relation to
$p$-divisible groups
Here we recall the relationships between $p$-divisible groups and global sections of vector bundles on the Fargues–Fontaine curve. Let us fix a perfect field
$k$ of characteristic
$p$, and write
$\operatorname {Perf}_{W(k)[1/p]}$ for the category of perfectoid spaces over
$W(k)[1/p]$. Given a
$p$-divisible group
$H$ over
$k$ with covariant isocrystal
$N$, if
$H$ has slopes
$s_1,\ldots ,s_k \in \mathbb {Q}$, then
$N$ has the slopes
$1-s_1, \ldots ,1-s_k$. For an object
$S$ in
$\operatorname {Perf}_{W(k)[1/p]}$ we define the vector bundle
${{\mathcal {E}}}_S(H)$ on
$X_S$ by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU8.png?pub-status=live)
Under this normalization, the Harder–Narasimhan slopes of ${{\mathcal {E}}}_S(H)$ are (pointwise on
$S$) the same as the slopes of
$H$.
Let us write $H^{0}({{\mathcal {E}}}(H))$ for the sheafification of the functor on
$\operatorname {Perf}_{W(k)[1/p]}$, which sends
$S$ to
$H^{0}(X_S,{{\mathcal {E}}}_S(H))$.
Proposition 3.2.1 Let $H$ be a
$p$-divisible group over a perfect field
$k$ of characteristic
$p$, with isocrystal
$N$. There is an isomorphism
$\tilde {H}^{\operatorname{ad}}_{\eta }\cong H^{0}({{\mathcal {E}}}(H))$ of sheaves on
$\operatorname {Perf}_{W(k)[1/p]}$ making the following diagram commute,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU9.png?pub-status=live)
where the morphism $H^{0}({{\mathcal {E}}}(H))\to N\otimes _{W(k)[1/p]} {{\mathbf {G}}}_a$ sends a global section of
${{\mathcal {E}}}(H)$ to its fiber at
$\infty$.
Proof. Let $S=\operatorname {Spa}(R,R^{+})$ be an affinoid perfectoid space over
$W(k)[1/p]$. Then
$\tilde {H}^{\operatorname{ad}}_{\eta }(R,R^{+})\cong \tilde {H}(R^{\circ })\cong \tilde {H}(R^{\circ }/p)$. Observe that
$\tilde {H}(R^{\circ }/p)=\operatorname {Hom}_{R^{\circ }/p}(\mathbf {Q}_p/\mathbf {Z}_p,H)[1/p]$, where the Hom is taken in the category of
$p$-divisible groups over
$R^{\circ }/p$. Recall the crystalline Dieudonné functor
$G\mapsto M(G)$ from
$p$-divisible groups to Dieudonné crystals [Reference MessingMes72]. Since the base ring
$R^{\circ }/p$ is semiperfect, the latter category is equivalent to the category of finite projective modules over Fontaine's period ring
$A_\textrm {cris}(R^{\circ }/p)=A_\textrm {cris}(R^{\circ })$, equipped with Frobenius and Verschiebung.
Now we apply [Reference Scholze and WeinsteinSW13, Theorem A]: since $R^{\circ }/p$ is f-semiperfect, the crystalline Dieudonné functor is fully faithful up to isogeny. Thus
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU10.png?pub-status=live)
where the latter Hom is in the category of modules over $A_\textrm {cris}(R^{\circ })$ equipped with Frobenius. Recall that
$B_\textrm {cris}^{+}(R^{\circ })=A_\textrm {cris}(R^{\circ })[1/p]$. Since
$H$ arises via base change from
$k$, we have
$M(H)[1/p]=B_\textrm {cris}^{+}(R^{\circ }) \otimes _{W(k)[1/p]} N$. For its part,
$M(\mathbf {Q}_p/\mathbf {Z}_p)[1/p]=B_\textrm {cris}^{+}(R^{\circ })e$, for a basis element
$e$ on which Frobenius acts as
$p$. Therefore
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU11.png?pub-status=live)
On the Fargues–Fontaine curve side, we have by definition $H^{0}(X_S,{{\mathcal {E}}}_S(H))=(B_S\otimes _{W(k)[1/p]} N)^{\phi =p}$. The isomorphism between
$(B_S\otimes _{W(k)[1/p]} N)^{\phi =p}$ and
$(B_\textrm {cris}^{+}(R^{\circ }) \otimes _{W(k)[1/p]} N )^{\phi =p}$ is discussed in [Reference Le BrasLB18, Remarque 6.6].
The commutativity of the diagram in the proposition is [Reference Scholze and WeinsteinSW13, Proposition 5.1.6(ii)], at least in the case that $S$ is a geometric point, but this suffices to prove the general case.
With Proposition 3.2.1 we can reinterpret the infinite-level Rapoport Zink spaces as moduli spaces of modifications of vector bundles on the Fargues–Fontaine curve. First we do this for ${{\mathscr M}}_{H,\infty }$. In the following, we consider
${{\mathscr M}}_{H,\infty }$ as a sheaf on the category of perfectoid spaces over
$W(k)[1/p]$.
Proposition 3.2.2 Let $H$ be a
$p$-divisible group of height
$n$ and dimension
$d$ over a perfect field
$k$. Let
$N$ be the associated isocrystal over
$k$. Then
${{\mathscr M}}_{H,\infty }$ is isomorphic to the functor which inputs an affinoid perfectoid space
$S=\operatorname {Spa}(R,R^{+})$ over
$W(k)[1/p]$ and outputs the set of exact sequences
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn1.png?pub-status=live)
where $i_\infty \colon \operatorname {Spec} R\to X_S$ is the inclusion, and
$W$ is a projective
$\mathcal {O}_S$-module quotient of
$N\otimes _{W(k)[1/p]} \mathcal {O}_S$ of rank
$d$.
Proof. We briefly describe this isomorphism on the level of points over $S=\operatorname {Spa}(R,R^{+})$. Suppose that we are given a point of
${{\mathscr M}}_{H,\infty }(S)$, corresponding to a
$p$-divisible group
$G$ over
$R^{\circ }$, together with a quasi-isogeny
$\iota \colon H\otimes _{k} R^{\circ }/p \to G\otimes _{R^{\circ }} R^{\circ }/p$ and an isomorphism
$\alpha \colon \mathbf {Q}_p^{n}\to VG$ of sheaves of
$\mathbf {Q}_p$-vector spaces on
$S$. The logarithm map on
$G$ fits into an exact sequence of sheaves of
$\mathbf {Z}_p$-modules on
$S$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU12.png?pub-status=live)
After taking projective limits along multiplication-by-$p$, this turns into an exact sequence of sheaves of
$\mathbf {Q}_p$-vector spaces on
$S$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU13.png?pub-status=live)
The quasi-isogeny induces an isomorphism $\tilde {H}_{\eta }^{\operatorname{ad}} \times _{\operatorname {Spa} W(k)[1/p]} S \cong \tilde {G}_{\eta }^{\operatorname{ad}}$; composing this with the level structure gives an injective map
$\mathbf {Q}_p^{n}\to \tilde {H}^{\operatorname{ad}}_{\eta }(S)$, whose cokernel
$W$ is isomorphic to the projective
$R$-module
$\operatorname {Lie} G$ of rank
$d$. In light of Theorem 3.2.1, the map
$\mathbf {Q}_p^{n}\to \tilde {H}^{\operatorname{ad}}_{\eta }(S)$ corresponds to an
$\mathcal {O}_{X_{S}}$-linear map
$s\colon \mathcal {O}_{X_S}^{n}\to {{\mathcal {E}}}_S(H)$, which fits into the exact sequence in (3.2.1).
Similarly, we have a description of ${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ in terms of modifications.
Proposition 3.2.3 Let ${{\mathcal {D}}}=(B,V,H,\mu )$ be a rational EL datum. Then
${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ is isomorphic to the functor which inputs an affinoid perfectoid space
$S$ over
$\breve {E}$ and outputs the set of exact sequences of
$B \otimes _{\mathbf {Q}_p} \mathcal {O}_{X_S}$-modules
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU14.png?pub-status=live)
where $W$ is a finite projective
$\mathcal {O}_S$-module, which is locally isomorphic to
$V_0\otimes _{\mathbf {Q}_p} \mathcal {O}_S$ as a
$B\otimes _{\mathbf {Q}_p} \mathcal {O}_S$-module (using notation from Definition 2.2.1).
3.3 The determinant morphism, and connected components
If we are given a rational EL datum ${{\mathcal {D}}}$, there is a determinant morphism
$\det \colon {{\mathscr M}}_{{{\mathcal {D}}},\infty }\to {{\mathscr M}}_{\det {{\mathcal {D}}},\infty }$, which we review below. For an algebraically closed perfectoid field
$C$ containing
$W(k)[1/p]$, the base change
${{\mathscr M}}_{\det {{\mathcal {D}}},\infty ,C}$ is a locally profinite set of copies of
$\operatorname {Spa} C$. For a point
$\tau \in {{\mathscr M}}_{\det {{\mathcal {D}}},\infty }(C)$, let
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\tau }$ be the fiber of
${{\mathscr M}}_{{{\mathcal {D}}},\infty }\to {{\mathscr M}}_{\det {{\mathcal {D}}},\infty }$ over
$\tau$. We will prove in § 5 that each
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\tau ,\operatorname {non-sp}}$ is cohomologically smooth if
${{\mathcal {D}}}$ is basic. This implies that
$\pi _0({{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\tau ,\operatorname {non-sp}})$ is discrete, so that cohomogical smoothness of
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\tau ,\operatorname {non-sp}}$ is inherited by each of its connected components. This is Theorem 1.0.1. In certain cases (for example Lubin–Tate space) it is known that
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\tau }$ is already connected [Reference ChenChe14].
We first review the determinant morphism for the space ${{\mathscr M}}_{H,\infty }$, where
$H$ is a
$p$-divisible group of height
$n$ and dimension
$d$ over a perfect field
$k$ of characteristic
$p$. Let
$\breve {E}=W(k)[1/p]$. For a perfectoid space
$S=\operatorname {Spa}(R,R^{+})$ over
$\breve {E}$, we have the vector bundle
${{\mathcal {E}}}_S(H)$ and its determinant
$\det {{\mathcal {E}}}_S(H)$, a line bundle of degree
$d$. (This does not correspond to a
$p$-divisible group ‘
$\det H$’ unless
$d\leq 1$.) We define
${{\mathscr M}}_{\det H,\infty }(S)$ to be the set of morphisms
$s\colon \mathcal {O}_{X_S}\to \det {{\mathcal {E}}}_S(H)$, such that the cokernel of
$s$ is a projective
$B_{\operatorname {dR}}^{+}(R)/(\xi )^{d}$-module of rank 1, where
$(\xi )$ is the kernel of
$B_{\operatorname {dR}}^{+}(R)\to R$. The morphism
$\det \colon {{\mathscr M}}_{H,\infty }\to {{\mathscr M}}_{\det H,\infty }$ is simply
$s\mapsto \det s$.
Regarding the structure of ${{\mathscr M}}_{\det H,\infty }$: we claim that for an algebraically closed perfectoid field
$C/\breve {E}$, the set
${{\mathscr M}}_{\det H,\infty }(C)$ is a
$\mathbf {Q}_p^{\times }$-torsor. Indeed, since the vector bundle
${{\mathcal {E}}}_C(H)$ has degree
$d$, so does the line bundle
$\det {{\mathcal {E}}}_C(H)$, so that
$\det {{\mathcal {E}}}_C(H)\cong \mathcal {O}_{X_C}(d)$. A
$C$-point of
${{\mathscr M}}_{\det H,\infty }$ is therefore a global section of
$\mathcal {O}_{X_C}(d)$ with a zero of order
$d$ at
$\infty$. In other words, it is a nonzero element of
$\operatorname {Fil}^{0}B_C^{\phi =p^{d}}\cong \mathbf {Q}_p(d)$.
For the general case, let ${{\mathcal {D}}}=(B,V,H,\mu )$ be a rational EL datum. Let
$F=Z(B)$ be the center of
$B$. Then
$F$ is a semisimple commutative
$\mathbf {Q}_p$-algebra; i.e., it is a product of fields. The idea is now to construct the determinant datum
$\det {{\mathcal {D}}}=(F,\det _F V,\det _F H,\det _F\circ \mu )$, noting once again that there may not be a
$p$-divisible group ‘
$\det _F H$’. The determinant
$\det _F V$ is a free
$F$-module of rank 1. For a perfectoid space
$S=\operatorname {Spa}(R,R^{+})$ over
$\breve {E}$, we have the
$F\otimes _{\mathbf {Q}_p}{{\mathcal {O}}}_{X_S}$-module
${{\mathcal {E}}}_S(H)$ and its determinant
$\det _F {{\mathcal {E}}}_S(H)$; the latter is a locally free
$F\otimes _{\mathbf {Q}_p}{{\mathcal {O}}}_{X_S}$-module of rank 1. Let
$d$ be the degree of
$\det _F{{\mathcal {E}}}_S(H)$, considered as a function on
$\operatorname {Spec} F$. We define
${{\mathscr M}}_{\det {{\mathcal {D}}},\infty }(S)$ to be the set of
$F$-linear morphisms
$s\colon \det _F V\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_S} \to \det _F {{\mathcal {E}}}_S(H)$, such that the cokernel of
$s$ is (locally on
$\operatorname {Spec} F$) a projective
$B_{\operatorname {dR}}^{+}(R)/(\xi )^{d}$-module of rank 1. (We remark here that the
$\det _F$ in
$\det _F\circ \mu$ means the morphism from
${{\mathbf {G}}}=\operatorname {Aut}_B(V)$ to
${{\mathbf {G}}}^{\text {ab}}=\operatorname {Aut}_F(\det _F V)=\mathop { \rm Res}_{F/\mathbf {Q}_p}{{\mathbf {G}}}_m$. If
$\det _F \mu$ is a minuscule cocharacter, meaning that it is a vector of only 0s and 1s in the character group
$X_*({{\mathbf {G}}}^{\text {ab}})\cong \mathbf {Z}^{[F:\mathbf {Q}_p]})$, then
$\det {{\mathcal {D}}}$ is an honest rational EL datum.) The morphism
${{\mathscr M}}_{{{\mathcal {D}}},\infty }\to {{\mathscr M}}_{\det {{\mathcal {D}}},\infty }$ sends a
$B\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_S}$-linear map
$s\colon V\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_S} \to {{\mathcal {E}}}_S(H)$ to the
$F\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_S}$-linear map
$\det s\colon \det _F V\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_S} \to \det _F {{\mathcal {E}}}_S(H)$.
An argument similar to the above shows that for an algebraically closed perfectoid field $C/\breve {E}$, the set
${{\mathscr M}}_{\det {{\mathcal {D}}},\infty }(C)$ is an
$F^{\times }$-torsor, equal to the set of
$F$-bases for
$F(d)$. Here the Tate twist is interpreted (locally on
$\operatorname {Spec} F$) as the
$d$th tensor power of the rational Tate module of the Lubin–Tate module for
$F$.
3.4 Basic Rapoport–Zink spaces
The main theorem of this article concerns basic Rapoport–Zink spaces, so we recall some facts about these here.
Let $H$ be a
$p$-divisible group over a perfect field
$k$ of characteristic
$p$. The space
${{\mathcal {M}}}_{H,\infty }$ is said to be basic when the
$p$-divisible group
$H$ (or rather, its Dieudonné module
$M(H)$) is isoclinic. This is equivalent to saying that the natural map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU15.png?pub-status=live)
is an isomorphism, where on the right the endomorphisms are not required to commute with Frobenius.
More generally we have a notion of basicness for a rational EL datum $(B,H,V,\mu )$, referring to the following equivalent conditions.
– The
${{\mathbf {G}}}$-isocrystal (
${{\mathbf {G}}}=\operatorname {Aut}_B V$) associated to
$H$ is basic in the sense of Kottwitz [Reference KottwitzKot85].
– The natural map
is an isomorphism.\[ \operatorname{End}^{\circ}_B(H) \otimes_{\mathbf{Q}_p} W(k)[1/p] \to \operatorname{End}_{B\otimes_{\mathbf{Q}_p} W(k)[1/p]} M(H)[1/p] \]
– Considered as an algebraic group over
$\mathbf {Q}_p$, the automorphism group
$J=\operatorname {Aut}^{\circ }_B H$ is an inner form of
${{\mathbf {G}}}$.
– Let
$D'=\operatorname {End}^{\circ }_B H$. For any algebraically closed perfectoid field
$C$ containing
$W(k)$, the map
is an isomorphism.\[ D'\otimes_{\mathbf{Q}_p} \mathcal{O}_{X_C} \to \operatorname{\mathscr{E}{\mathcal{nd}}\,}_{(B\otimes_{\mathbf{Q}_p} \mathcal{O}_{X_C})} {{\mathcal{E}}}_C(H) \]
In brief, the duality theorem from [Reference Scholze and WeinsteinSW13] says the following. Given a basic EL datum ${{\mathcal {D}}}$, there is a dual datum
$\check {{{\mathcal {D}}}}$, for which the roles of the groups
${{\mathbf {G}}}$ and
$J$ are reversed. There is a
${{\mathbf {G}}}(\mathbf {Q}_p)\times J(\mathbf {Q}_p)$-equivariant isomorphism
${{\mathscr M}}_{{{\mathcal {D}}},\infty }\cong {{\mathscr M}}_{\check {{{\mathcal {D}}}},\infty }$ which exchanges the roles of
$\pi _{GM}$ and
$\pi _{HT}$.
3.5 The special locus
Let ${{\mathcal {D}}}=(B,V,H,\mu )$ be a basic rational EL datum relative to a perfect field
$k$ of characteristic
$p$, with reflex field
$E$. Let
$F$ be the center of
$B$. Define
$F$-algebras
$D$ and
$D'$ by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU18.png?pub-status=live)
Finally, let ${{\mathbf {G}}}=\operatorname {Aut}_B V$ and
$J=\operatorname {Aut}_B H$, considered as algebraic groups over
$\mathbf {Q}_p$. Then
${{\mathbf {G}}}$ and
$J$ both contain
$\mathop { \rm Res}_{F/\mathbf {Q}_p}{{\mathbf {G}}}_m$.
Let $C$ be an algebraically closed perfectoid field containing
$\breve {E}$, and let
$x\in {{\mathscr M}}_{{{\mathcal {D}}},\infty }(C)$. Then
$x$ corresponds to a
$p$-divisible group
$G$ over
$\mathcal {O}_C$ with endomorphisms by
$B$, and also it corresponds to a
$B\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_C}$-linear map
$s\colon V\otimes _{\mathbf {Q}_p}\mathcal {O}_X\to {{\mathcal {E}}}_C(N)$ as in Proposition 3.2.3. Define
$A_x=\operatorname {End}_B G$ (endomorphisms in the isogeny category). Then
$A_x$ is a semisimple
$F$-algebra. In light of Proposition 3.2.3, an element of
$A_x$ is a pair
$(\alpha ,\alpha ')$, where
$\alpha \in \operatorname {End}_{B\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_C}} V\otimes \mathcal {O}_{X_C}=\operatorname {End}_B V=D$ and
$\alpha '\in \operatorname {End}_{B\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_C}} {{\mathscr E}}_C(H)=D'$ (the last equality is due to basicness), such that
$s\circ \alpha =\alpha '\circ s$. Thus we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU19.png?pub-status=live)
Lemma 3.5.1 The following are equivalent.
(i) The
$F$-algebra
$A_x$ strictly contains
$F$.
(ii) The stabilizer of
$\pi _{GM}(x)\in \mathcal {F}\ell _\mu (C)$ in
$J(\mathbf {Q}_p)$ strictly contains
$F^{\times }$.
(iii) The stabilizer of
$\pi _{HT}(x)\in \mathcal {F}\ell _\mu '(C)$ in
${{\mathbf {G}}}(\mathbf {Q}_p)$ strictly contains
$F^{\times }$.
Proof. As in Proposition 3.2.3, let $s \colon V\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_S} \stackrel {s}{\to } {{\mathcal {E}}}_S(H)$ be the modification corresponding to
$x$.
Note that the condition (1) is equivalent to the existence of an invertible element $(\alpha ,\alpha ') \in A_x$ not contained in (the diagonally embedded)
$F$. Also note that if one of
$\alpha ,\alpha '$ lies in
$F$, then so does the other, in which case they are equal.
Suppose $(\alpha ,\alpha ')\in A_x$ is invertible. The point
$\pi _{GM}(x)\in {{\mathcal {F}}}\ell _\mu$ corresponds to the cokernel of the fiber of
$s$ at
$\infty$. Since
$\alpha '\circ s=s\circ \alpha$, the cokernels of
$\alpha '\circ s$ and
$s$ are the same, which means exactly that
$\alpha '\in J(\mathbf {Q}_p)$ stabilizes
$\pi _{GM}(x)$. Thus (1) implies (2). Conversely, if there exists
$\alpha '\in J(\mathbf {Q}_p)\backslash F^{\times }$ which stabilizes
$\pi _{GM}(x)$, it means that the
$B\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_C}$-linear maps
$s$ and
$\alpha '\circ s$ have the same cokernel, and therefore there exists
$\alpha \in \operatorname {End}_{B\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_C}} V\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_C}=D$ such that
$s\circ \alpha = \alpha '\circ s$, and then
$(\alpha ,\alpha ')\in A_x\backslash F^{\times }$. This shows that (2) implies (1).
The equivalence between (1) and (3) is proved similarly.
Definition 3.5.2 The special locus in ${{\mathscr M}}_{{{\mathcal {D}}},\infty }$ is the subset
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\operatorname {sp}}$ defined by the condition
$A_x\neq F$. The nonspecial locus
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\operatorname {non-sp}}$ is the complement of the special locus.
The special locus is built out of ‘smaller’ Rapoport–Zink spaces, in the following sense. Let $A$ be a semisimple
$F$-algebra, equipped with two
$F$-embeddings
$A\to D$ and
$A\to D'$, so that
$A\otimes _F B$ acts on
$V$ and
$H$. Also assume that a cocharacter in the conjugacy class
$\mu$ factors through a cocharacter
$\mu _0\colon {{\mathbf {G}}}_m\to \operatorname {Aut}_{A\otimes _F B} V$. Let
${{\mathcal {D}}}_0=(A\otimes _F B,V,H,\mu _0)$. Then there is an evident morphism
${{\mathscr M}}_{{{\mathcal {D}}}_0,\infty }\to {{\mathscr M}}_{{{\mathcal {D}}},\infty }$. The special locus
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\operatorname {sp}}$ is the union of the images of all the
${{\mathscr M}}_{{{\mathcal {D}}}_0,\infty }$, as
$A$ ranges through all semisimple
$F$-subalgebras of
$D\times D'$ strictly containing
$F$.
4. Cohomological smoothness
Let $\operatorname {Perf}$ be the category of perfectoid spaces in characteristic
$p$, with its pro-étale topology [Reference ScholzeSch17, Definition 8.1]. For a prime
$\ell \neq p$, there is a notion of
$\ell$-cohomological smoothness [Reference ScholzeSch17, Definition 23.8]. We only need the notion for morphisms
$f\colon Y'\to Y$ between sheaves on
$\operatorname {Perf}$ which are separated and representable in locally spatial diamonds. If such an
$f$ is
$\ell$-cohomologically smooth, and
$\Lambda$ is an
$\ell$-power torsion ring, then the relative dualizing complex
$Rf^{!}\Lambda$ is an invertible object in
$D_{\mathop e\limits{{\prime}}t}(Y',\Lambda )$ (thus, it is v-locally isomorphic to
$\Lambda [n]$ for some
$n\in {{\mathbf {Z}}}$), and the natural transformation
$Rf^{!}\Lambda \otimes f^{*}\to Rf^{!}$ of functors
$D_{\mathop e\limits{{\prime}}t}(Y,\Lambda )\to D_{\mathop e\limits{{\prime}}t}(Y',\Lambda )$ is an equivalence [Reference ScholzeSch17, Proposition 23.12]. In particular, if
$f$ is projection onto a point, and
$Rf^{!}\Lambda \cong \Lambda [n]$, one derives a statement of Poincaré duality for
$Y'$:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU20.png?pub-status=live)
We will say that $f$ is cohomologically smooth if it is
$\ell$-cohomologically smooth for all
$\ell \neq p$. As an example, if
$f\colon Y'\to Y$ is a separated smooth morphism of rigid-analytic spaces over
$\mathbf {Q}_p$, then the associated morphism of diamonds
$f^{\diamond }\colon (Y')^{\diamond }\to Y^{\diamond }$ is cohomologically smooth [Reference ScholzeSch17, Proposition 24.3]. There are other examples where
$f$ does not arise from a finite-type map of adic spaces. For instance, if
$\tilde {B}_C=\operatorname {Spa} C \langle T^{1/p^{\infty }} \rangle$ is the perfectoid closed ball over an algebraically closed perfectoid field
$C$, then
$\tilde {B}_C$ is cohomologically smooth over
$C$.
If $Y$ is a perfectoid space over an algebraically closed perfectoid field
$C$, it seems quite difficult to detect whether
$Y$ is cohomologically smooth over
$C$. We will review in § 4.2 a ‘Jacobian criterion’ from [Reference Fargues and ScholzeFS] which applies to certain kinds of
$Y$. But first we give a classical analogue of this criterion in the context of schemes.
4.1 The Jacobian criterion: classical setting
Proposition 4.1.1 Let $X$ be a smooth projective curve over an algebraically closed field
$k$. Let
$Z\to X$ be a smooth morphism. Define
${{\mathscr M}}_Z$ to be the functor which inputs a
$k$-scheme
$T$ and outputs the set of sections of
$Z\to X$ over
$X_T$, that is, the set of morphisms
$s$ making
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU21.png?pub-status=live)
commute, subject to the condition that, fiberwise on $T$, the vector bundle
$s^{*}\operatorname {Tan}_{Z/X}$ has vanishing
$H^{1}$. Then
${{\mathscr M}}_Z \to \operatorname {Spec} k$ is formally smooth.
Here $\operatorname {Tan}_{Z/X}$ is the tangent bundle, equal to the
$\mathcal {O}_Z$-linear dual of the sheaf of differentials
$\Omega _{Z/X}$, which is locally free of finite rank. Let
$\pi \colon X\times _k T \to T$ be the projection. For
$t\in T$, let
$X_t$ be the fiber of
$\pi$ over
$t$, and let
$s_t\colon X_t\to Z$ be the restriction of
$s$ to
$X_t$. By proper base change, the fiber of
$R^{1}\pi _*s^{*}\operatorname {Tan}_{Z/X}$ at
$t\in T$ is
$H^{1}(X_t,s_t^{*}\operatorname {Tan}_{Z/X})$. The condition about the vanishing of
$H^{1}$ in the proposition is equivalent to
$H^{1}(X_t,s_t^{*}\operatorname {Tan}_{Z/X})=0$ for each
$t\in T$. By Nakayama's lemma, this condition is equivalent to
$R^{1}\pi _*s^{*}\operatorname {Tan}_{Z/X}=0$.
Proof. Suppose we are given a commutative diagram,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn2.png?pub-status=live)
where $T_0\to T$ is a first-order thickening of affine schemes; thus
$T_0$ is the vanishing locus of a square-zero ideal sheaf
$I\subset \mathcal {O}_T$. Note that
$I$ becomes an
$\mathcal {O}_{T_0}$-module.
The morphism $T_0\to {{\mathscr M}}_Z$ in (4.1.1) corresponds to a section of
$Z\to X$ over
$T_0$. Thus there is a solid diagram.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn3.png?pub-status=live)
We claim that there exists a dotted arrow making the diagram commute. Since $Z\to X$ is smooth, it is formally smooth, and therefore this arrow exists Zariski-locally on
$X$. Let
$\pi \colon X\times _k T\to T$ and
$\pi _0\colon X\times _k T_0\to T_0$ be the projections. Then
$X\times _k T_0$ is the vanishing locus of the ideal sheaf
$\pi ^{*}I\subset \mathcal {O}_{X\times _k T}$. Note that sheaves of sets on
$X\times _k T$ are equivalent to sheaves of sets on
$X\times _k T_0$; under this equivalence,
$\pi ^{*}I$ and
$\pi _0^{*}I$ correspond. By [Sta14, Remark 36.9.6], the set of such morphisms form a (Zariski) sheaf of sets on
$X\times _k T$, which when viewed as a sheaf on
$X\times _k T_0$ is a torsor for
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU22.png?pub-status=live)
This torsor corresponds to class in
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU23.png?pub-status=live)
This $H^{1}$ is the limit of a spectral sequence with terms
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU24.png?pub-status=live)
But since $T_0$ is affine and
$R^{q}\pi _{0*}(s_0^{*}\operatorname {Tan}_{Z/X}\otimes \ \pi _0^{*}I)$ is quasi-coherent, the above terms vanish for all
$p>0$, and therefore
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU25.png?pub-status=live)
Since $s_0^{*}\operatorname {Tan}_{Z/X}$ is locally free, we have
$s_0^{*}\operatorname {Tan}_{Z/X}\otimes \ \pi _0^{*}I \cong s_0^{*}\operatorname {Tan}_{Z/X}\otimes ^{{{\mathbf {L}}}} \ \pi _{0*}I$, and we may apply the projection formula [Sta14, Lemma 35.21.1] to obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU26.png?pub-status=live)
Now we apply the hypothesis about vanishing of $H^{1}$, which implies that
$R\pi _{0*}s_0^{*}\operatorname {Tan}_{Z/X}$ is quasi-isomorphic to the locally free sheaf
$\pi _{0*}s_0^{*}\operatorname {Tan}_{Z/X}$ in degree 0. Therefore the complex displayed above has
$H^{1}=0$.
Thus our torsor is trivial, and so a morphism $s\colon X\times _k T\to Z$ exists filling in (4.1.2). The final thing to check is that
$s$ corresponds to a morphism
$T\to {{\mathscr M}}_Z$, i.e., that it satisfies the fiberwise
$H^{1}=0$ condition. But this is automatic, since
$T_0$ and
$T$ have the same schematic points.
In the setup of Proposition 4.1.1, let $s\colon X \times _k {{\mathscr M}}_Z \to Z$ be the universal section. That is, the pull-back of
$s$ along a morphism
$T \to {{\mathscr M}}_Z$ is the section
$X\times _k T\to Z$ to which this morphism corresponds. Let
$\pi \colon X\times _k {{\mathscr M}}_Z\to {{\mathscr M}}_Z$ be the projection. By Proposition 4.1.1
${{\mathscr M}}_Z\to \operatorname {Spec} k$ is formally smooth. There is an isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU27.png?pub-status=live)
Indeed, the proof of Proposition 4.1.1 shows that $\pi _*s^{*}\operatorname {Tan}_{Z/X}$ has the same universal property with respect to first order deformations as
$\operatorname {Tan}_{{{\mathscr M}}_Z/\operatorname {Spec} k}$.
The following example is of similar spirit as our main application of the perfectoid Jacobian criterion below.
Example 4.1.2 Let $X = {{\mathbf {P}}}^{1}$ over the algebraically closed field
$k$. For
$d \in {{\mathbf {Z}}}$, let
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU27a.png?pub-status=live)
be the geometric vector bundle over $X$ whose global sections are
$\Gamma (X,{{\mathcal {O}}}(d))$. Fix integers
$n,d, \delta > 0$ and let
$P$ be a homogeneous polynomial over
$k$ of degree
$\delta$ in
$n$ variables. Then
$P$ defines a morphism
$P \colon \prod _{i=1}^{n} V_d \rightarrow V_{d\delta }$, by sending sections
$(s_i)_{i=1}^{n}$ of
$V_d$ to the section
$P(s_1,\ldots ,s_n)$ of
$V_{d\delta }$. Fix a global section
$f \colon X \rightarrow V_{d\delta }$ to the projection morphism and consider the pull-back of
$P$ along
$f$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU28.png?pub-status=live)
Moreover, let $Z$ be the smooth locus of
$P^{-1}(f)$ over
$X$. It is an open subset. The derivatives
${\partial P}/{\partial x_i}$ of
$P$ are homogeneous polynomials of degree
$\delta - 1$ in
$n$ variables, hence can be regarded as functions
$\prod _{i=1}^{n} V_d \rightarrow V_{d(\delta - 1)}$. A point
$y \in P^{-1}(f)$ lies in
$Z$ if and only if
$({\partial P}/{\partial x_i})(y)$,
$i=1,\ldots ,n$ are not all zero. We wish to apply Proposition 4.1.1 to
$Z/X$. Let
${{\mathscr M}}'_Z$ denote the space of global sections of
$Z$ over
$X$, that is for a
$k$-scheme
$T$,
${{\mathscr M}}_Z'(T)$ is the set of morphisms
$s \colon X \times _k T \rightarrow Z$ as in the proposition (without any further conditions). A
$k$-point
$g \in {{\mathscr M}}'_Z(k)$ corresponds to a section
$g \colon X \rightarrow \prod _{i=1}^{n} V_d$, satisfying
$P \circ g = f$. In general, for a (geometric) vector bundle
$V$ on
$X$ with corresponding locally free
${{\mathcal {O}}}_X$-module
${{\mathscr E}}$, the pull-back of the tangent space
$\mathrm {Tan}_{V/X}$ along a section
$s \colon X \rightarrow V$ is canonically isomorphic to
${{\mathscr E}}$. Hence in our situation (using that
$Z \subseteq P^{-1}(f)$ is open) the tangent space
$g^{\ast } \mathrm {Tan}_{Z/X}$ can be computed from the short exact sequence,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU29.png?pub-status=live)
where $D_g P$ is the derivative of
$P$ at
$g$. It is the
${{\mathcal {O}}}_X$-linear map given by
$(t_i)_{i=1}^{n} \mapsto \sum _{i = 1}^{n} ({\partial P}/{\partial x_i})(g)t_i$ (note that
$({\partial P}/{\partial x_i})(g)$ are global sections of
${{\mathcal {O}}}(d(\delta - 1))$). Note that
$D_g P$ is surjective: by Nakayama, it suffices to check this fiberwise, where it is true by the condition defining
$Z$.
The space ${{\mathscr M}}_Z$ is the subfunctor of
${{\mathscr M}}'_Z$ consisting of all
$g$ such that (fiberwise)
$g^{\ast }\mathrm {Tan}_{Z/X} = \mathrm {ker}(D_g P)$ has vanishing
$H^{1}$. Writing
$\mathrm {ker}(D_g P) = \bigoplus _{i=1}^{r} {{\mathcal {O}}}(m_i)$ (
$m_i \in {{\mathbf {Z}}}$), this is equivalent to
$m_i \geq -1$. By the Proposition 4.1.1 we conclude that
${{\mathscr M}}_Z$ is formally smooth over
$k$.
Consider now a numerical example. Let $n = 3$,
$d = 1$ and
$\delta = 4$ and let
$g \in {{\mathscr M}}'_Z(k)$. Then
$D_g P \in \mathrm {Hom}_{{{\mathcal {O}}}_X}({{\mathcal {O}}}(1)^{\oplus 3}, {{\mathcal {O}}}(4)) = \Gamma (X,{{\mathcal {O}}}(3)^{\oplus 3})$, a
$12$-dimensional
$k$-vector space, and moreover,
$D_g P$ lies in the open subspace of surjective maps. We have the short exact sequence of
${{\mathcal {O}}}_X$-modules
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn4.png?pub-status=live)
This shows that $g^{\ast } \mathrm {Tan}_{Z/X}$ has rank
$2$ and degree
$-1$. Moreover, being a subbundle of
${{\mathcal {O}}}(1)^{\oplus 3}$ it only can have slopes less than or equal to
$1$. There are only two options, either
$g^{\ast } \mathrm {Tan}_{Z/X} \cong {{\mathcal {O}}}(-1) \oplus {{\mathcal {O}}}$ or
$g^{\ast } \mathrm {Tan}_{Z/X} \cong {{\mathcal {O}}}(-2) \oplus {{\mathcal {O}}}(1)$. The point
$g$ lies in
${{\mathscr M}}_Z$ if and only if the first option occurs for
$g$. Which option occurs can be seen from the long exact cohomology sequence associated to (4.1.3):
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU30.png?pub-status=live)
It is clear that $\Gamma (X, g^{\ast } \mathrm {Tan}_{Z/X})$ is
$1$-dimensional if and only if
$g^{\ast } \mathrm {Tan}_{Z/X} \cong {{\mathcal {O}}}(-1) \oplus {{\mathcal {O}}}$ and
$2$-dimensional otherwise. The first option is generic, i.e.,
${{\mathscr M}}_Z$ is an open subscheme of
${{\mathscr M}}'_Z$.
4.2 The Jacobian criterion: perfectoid setting
We present here the perfectoid version of Proposition 4.1.1
Theorem 4.2.1 (Fargues and Scholze [FS]) Let $S=\operatorname {Spa}(R,R^{+})$ be an affinoid perfectoid space in characteristic
$p$. Let
$Z\to X_S$ be a smooth morphism of schemes. Let
${{\mathscr M}}_Z^{>0}$ be the functor which inputs a perfectoid space
$T\to S$ and outputs the set of sections of
$Z\to X_S$ over
$T$, that is, the set of morphisms
$s$ making
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU31.png?pub-status=live)
commute, subject to the condition that, fiberwise on $T$, all Harder–Narasimhan slopes of the vector bundle
$s^{*}\operatorname {Tan}_{Z/X_S}$ are positive. Then
${{\mathscr M}}_Z^{>0}\to S$ is a cohomologically smooth morphism of locally spatial diamonds.
Example 4.2.2 Let $S=\eta =\operatorname {Spa}(C,\mathcal {O}_C)$, where
$C$ is an algebraically closed perfectoid field of characteristic
$0$, and let
$Z={{\mathbf {V}}}({{\mathcal {E}}}_S(H))\to X_S$ be the geometric vector bundle attached to
${{\mathcal {E}}}_S(H)$, where
$H$ is a
$p$-divisible group over the residue field of
$C$. Then
${{\mathscr M}}_Z=H^{0}({{\mathcal {E}}}_S(H))$ is isomorphic to
$\tilde {H}_\eta ^{\operatorname{ad}}$ by Proposition 3.2.1. Let
$s\colon X_{{{\mathscr M}}_Z}\to Z$ be the universal morphism; then
$s^{*}\operatorname {Tan}_{Z/X_S}$ is the constant Banach–Colmez space associated to
$H$ (i.e., the pull-back of
${{\mathcal {E}}}_S(H)$ along
$X_{{{\mathscr M}}_Z} \rightarrow X_S$). This has vanishing
$H^{1}$ if and only if
$H$ has no étale part. This is true if and only if
${{\mathscr M}}_Z^{>0}$ is isomorphic to a perfectoid open ball. The perfectoid open ball is cohomologically smooth, in accord with Theorem 4.2.1. In contrast, if the étale quotient
$H^{\mathop e\limits{{\prime}}t}$ has height
$d > 0$, then
$\pi _0(\tilde {H}_{\eta }^{\operatorname{ad}})\cong \mathbf {Q}_p^{d}$ implies that
$\tilde {H}_{\eta }^{\operatorname{ad}}$ is not cohomologically smooth.
In the setup of Theorem 4.2.1, suppose that $x= \operatorname {Spa}(C,\mathcal {O}_C)\to S$ is a geometric point, and that
$x\to {{\mathscr M}}_Z^{>0}$ is an
$S$-morphism, corresponding to a section
$s\colon X_C\to Z$. Then
$s^{*}\operatorname {Tan}_{Z/X_S}$ is a vector bundle on
$X_C$. In light of the discussion in the previous section, we are tempted to interpret
$H^{0}(X_C,s^{*}\operatorname {Tan}_{Z/X_S})$ as the ‘tangent space of
${{\mathscr M}}_Z^{>0}\to S$ at
$x$’. At points
$x$ where
$s^{*}\operatorname {Tan}_{Z/X_S}$ has only positive Harder–Narasimhan slopes, this tangent space is a perfectoid open ball.
5. Proof of the main theorem
5.1 Dilatations and modifications
As preparation for the proof of Theorem 1.0.1, we review the notion of a dilatation of a scheme at a locally closed subscheme [Reference Bosch, Lütkebohmert and RaynaudBLR90, § 3.2].
Throughout this subsection, we fix some data. Let $X$ be a curve, meaning that
$X$ is a scheme which is locally the spectrum of a Dedekind ring. Let
$\infty \in X$ be a closed point with residue field
$C$. Let
$i_\infty \colon \operatorname {Spec} C \to X$ be the embedding, and let
$\xi \in \mathcal {O}_{X,\infty }$ be a local uniformizer at
$\infty$.
Proposition 5.1.1 Let $V\to X$ be a morphism of finite type, and let
$Y\subset V_\infty$ be a locally closed subscheme of the fiber of
$V$ at
$\infty$.
There exists a morphism of $X$-schemes
$V'\to V$ which is universal for the following property:
$V'\to X$ is flat at
$\infty$, and
$V_\infty '\to V_\infty$ factors through
$Y\subset V_\infty$.
The $X$-scheme
$V'$ is the dilatation of
$V$ at
$Y$. We review here its construction.
First suppose that $Y\subset V_\infty$ is closed. Let
${{\mathscr I}}\subset \mathcal {O}_V$ be the ideal sheaf which cuts out
$Y$. Let
$B\to V$ be the blow-up of
$V$ along
$Y$. Then
${{\mathscr I}}\cdot \mathcal {O}_B$ is a locally principal ideal sheaf. The dilatation
$V'$ of
$V$ at
$Y$ is the open subscheme of
$B$ obtained by imposing the condition that the ideal
$({{\mathscr I}}\cdot \mathcal {O}_B)_x\subset \mathcal {O}_{B,x}$ is generated by
$\xi$ at all
$x\in B$ lying over
$\infty$.
We give here an explicit local description of the dilatation $V'$. Let
$\operatorname {Spec} A$ be an affine neighborhood of
$\infty$, such that
$\xi \in A$, and let
$\operatorname {Spec} R\subset V$ be an open subset lying over
$\operatorname {Spec} A$. Let
$I=(f_1,\ldots ,f_n)$ be the restriction of
${{\mathscr I}}$ to
$\operatorname {Spec} R$, so that
$I$ cuts out
$Y\cap \operatorname {Spec} A$. Then the restriction of
$V'\to V$ to
$\operatorname {Spec} R$ is
$\operatorname {Spec} R'$, where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU32.png?pub-status=live)
Now suppose $Y\subset V_\infty$ is only locally closed, so that
$Y$ is open in its closure
$\overline {Y}$. Then the dilatation of
$V$ at
$Y$ is the dilatation of
$V\backslash (\overline {Y}\backslash Y)$ at
$Y$.
Note that a dilatation $V'\to V$ is an isomorphism away from
$\infty$, and that it is affine.
Example 5.1.2 Let
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU33.png?pub-status=live)
be an exact sequence of $\mathcal {O}_X$-modules, where
${{\mathcal {E}}}$ (and thus
${{\mathcal {E}}}'$) is locally free, and
$W$ is a
$C$-vector space. (This is an elementary modification of the vector bundle
${{\mathcal {E}}}$.) Let
$K=\ker ({{\mathcal {E}}}_\infty \to W)$.
Let ${{\mathbf {V}}}({{\mathcal {E}}})\to X$ be the geometric vector bundle corresponding to
${{\mathcal {E}}}$. Similarly, we have
${{\mathbf {V}}}({{\mathcal {E}}}')\to X$, and an
$X$-morphism
${{\mathbf {V}}}({{\mathcal {E}}}')\to {{\mathbf {V}}}({{\mathcal {E}}})$. Let
${{\mathbf {V}}}(K)\subset {{\mathbf {V}}}({{\mathcal {E}}})_\infty$ be the affine space associated to
$K\subset {{\mathcal {E}}}_\infty$. We claim that
${{\mathbf {V}}}({{\mathcal {E}}}')$ is isomorphic to the dilatation
${{\mathbf {V}}}({{\mathcal {E}}})'$ of
${{\mathbf {V}}}({{\mathcal {E}}})$ at
${{\mathbf {V}}}(K)$. Indeed, by the universal property of dilatations, there is a morphism
${{\mathbf {V}}}({{\mathcal {E}}}')\to {{\mathbf {V}}}({{\mathcal {E}}})'$, which is an isomorphism away from
$\infty$.
To see that ${{\mathbf {V}}}({{\mathcal {E}}}')\to {{\mathbf {V}}}({{\mathcal {E}}})'$ is an isomorphism, it suffices to work over
$\mathcal {O}_{X,\infty }$. Over this base, we may give a basis
$f_1,\ldots ,f_n$ of global sections of
${{\mathcal {E}}}$, with
$f_1,\ldots ,f_k$ lifting a basis for
$K\subset {{\mathcal {E}}}_\infty$. Then the localization of
${{\mathbf {V}}}({{\mathcal {E}}})'\to {{\mathbf {V}}}({{\mathcal {E}}})$ at
$\infty$ is isomorphic to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU34.png?pub-status=live)
This agrees with the localization of ${{\mathbf {V}}}({{\mathcal {E}}}')\to {{\mathbf {V}}}({{\mathcal {E}}})$ at
$\infty$.
Lemma 5.1.3 Let $V\to X$ be a smooth morphism, let
$Y\subset V_\infty$ be a smooth locally closed subscheme, and let
$\pi \colon V'\to V$ be the dilatation of
$V$ at
$Y$. Then
$V'\to X$ is smooth, and
$\operatorname {Tan}_{V'/X}$ lies in an exact sequence of
$\mathcal {O}_{V'}$-modules
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn5.png?pub-status=live)
where $N_{Y/V_\infty }$ is the normal bundle of
$Y\subset V_\infty$, and
$j\colon Y\to V$ is the inclusion.
Finally, let $T\to X$ be a morphism which is flat at
$\infty$, and let
$s\colon T\to V$ be a morphism of
$X$-schemes, such that
$s_\infty$ factors through
$Y$. By the universal property of dilatations,
$s$ factors through a morphism
$s'\colon T\to V'$. Then we have an exact sequence of
$\mathcal {O}_V$-modules
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn6.png?pub-status=live)
Proof. One reduces to the case that $Y$ is closed in
$V_\infty$. The smoothness of
$V'\to X$ is [Reference Bosch, Lütkebohmert and RaynaudBLR90, § 3.2, Proposition 3]. We turn to the exact sequence (5.1.1). The morphism
$\operatorname {Tan}_{V'/X}\to \pi ^{*}\operatorname {Tan}_{V/X}$ comes from functoriality of the tangent bundle. To construct the morphism
$\pi ^{*}\operatorname {Tan}_{V/X}\to \pi ^{*}j_*N_{Y/V_\infty }$, we consider the diagram
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU35.png?pub-status=live)
in which the outer rectangle is cartesian. For its part, the normal bundle $N_{Y/V_\infty }$ sits in an exact sequence of
$\mathcal {O}_Y$-modules
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU36.png?pub-status=live)
The composite
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU37.png?pub-status=live)
induces by adjunction a morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU38.png?pub-status=live)
where the last step is justified because $j$ is a closed immersion.
We check that (5.1.1) is exact using our explicit description of $V'$. The sequence is clearly exact away from the preimage of
$Y$ in
$V'$, since on the complement of this locus, the morphism
$\pi$ is an isomorphism, and
$\pi ^{*}j_*=0$. Therefore we let
$y\in Y$ and check exactness after localization at
$y$. Let
${{\mathcal {I}}}\subset \mathcal {O}_V$ be the ideal sheaf which cuts out
$Y$, and let
$I\subset \mathcal {O}_{V,y}$ be the localization of
${{\mathcal {I}}}$ at
$y$. Then
$\mathcal {O}_{V_\infty ,y}=\mathcal {O}_{V,y}/\xi$. Since
$Y\subset V_\infty$ are both smooth at
$y$, we can find a system of local coordinates
$\overline {f}_1,\ldots ,\overline {f}_n\in \mathcal {O}_{V_\infty ,y}$ (meaning that the differentials
$d\overline {f}_i$ form a basis for
$\Omega ^{1}_{V_{\infty }/C,y}$), such that
$\overline {f}_{k+1},\ldots ,\overline {f}_n$ generate
$I/\xi$. If
$\partial /\partial \overline {f}_i$ are the dual basis, then the stalk of
$N_{Y/V_\infty }$ at
$y$ is the free
$\mathcal {O}_{Y,y}$-module with basis
$\partial /\partial \overline {f}_{k+1},\ldots ,\partial /\partial \overline {f}_n$.
Choose lifts $f_i\in \mathcal {O}_{V,y}$ of the
$\overline {f}_i$. Then
$I$ is generated by
$\xi ,f_k,\ldots ,f_n$. The localization of
$V'\to V$ over
$y$ is
$\operatorname {Spec} \mathcal {O}_{V',y}$, where
$\mathcal {O}_{V',y}=\mathcal {O}_{V,y}[g_{k+1},\ldots ,g_n]/(\xi \text {-torsion})$, where
$\xi g_i=f_i$ for
$i=k+1,\ldots ,n$. Then the stalk of
$\operatorname {Tan}_{V'/X}$ at
$y$ is the free
$\mathcal {O}_{V',y}$-module with basis
$\partial /\partial f_1,\ldots ,\partial /\partial f_k,\partial /\partial g_{k+1},\ldots ,\partial /\partial g_n$, whereas the stalk of
$\pi ^{*}\operatorname {Tan}_{V/X}$ at
$y$ is the free
$\mathcal {O}_{V',y}$-module with basis
$\partial /\partial f_1,\ldots ,\partial / \partial f_n$. The quotient between these stalks is evidently the free module over
$\mathcal {O}_{V',y}/\xi$ with basis
$\partial /\partial f_{k+1},\ldots ,\partial /\partial f_n$, and this agrees with the stalk of
$\pi ^{*}j_*N_{Y/V_\infty }$.
Given a morphism of $X$-schemes
$s\colon T\to V$ as in the lemma, we apply
$(s')^{*}$ to (5.1.1); this is exact because
$s'$ is flat. The term on the right is
$s^{*}j_*N_{Y/V_\infty }\cong i_{T_\infty *}s_\infty ^{*} N_{Y/V_\infty }$ (once again, this is valid because
$j$ is a closed immersion).
5.2 The space
${{\mathscr M}}_{H,\infty }$ as global sections of a scheme over
$X_C$
We will prove Theorem 1.0.1 for the Rapoport–Zink spaces of the form ${{\mathscr M}}_{H,\infty }$ before proceeding to the general case. Let
$H$ be a
$p$-divisible group of height
$n$ and dimension
$d$ over a perfect field
$k$. In this context,
$\breve {E}=W(k)[1/p]$. Let
${{\mathcal {E}}}={{\mathcal {E}}}_C(H)$. Throughout, we will be interpreting
${{\mathscr M}}_{H,\infty }$ as a functor on
$\operatorname {Perf}_{\breve {E}}$ as in Proposition 3.2.2.
We have a determinant morphism $\det \colon {{\mathscr M}}_{H,\infty }\to {{\mathscr M}}_{\det H, \infty }$. Let
$\tau \in {{\mathscr M}}_{\det H,\infty }(C)$ be a geometric point of
${{\mathscr M}}_{\det H, \infty }$. This point corresponds to a section
$\tau$ of
${{\mathbf {V}}}(\det {{\mathcal {E}}})\to X_C$, which we also call
$\tau$. Let
${{\mathscr M}}_{H,\infty }^{\tau }$ be the fiber of
$\det$ over
$\tau$.
Our first order of business is to express ${{\mathscr M}}_{H,\infty }^{\tau }$ as the space of global sections of a smooth morphism
$Z\to X_C$, defined as follows. We have the geometric vector bundle
${{\mathbf {V}}}({{\mathcal {E}}}^{n})\to X$, whose global sections parametrize morphisms
$s\colon \mathcal {O}_{X_C}^{n}\to {{\mathcal {E}}}$. Let
$U_{n-d}$ be the locally closed subscheme of the fiber of
${{\mathbf {V}}}({{\mathcal {E}}}^{n})$ over
$\infty$, which parametrizes all morphisms of rank
$n-d$. We consider the dilatation
${{\mathbf {V}}}({{\mathcal {E}}}^{n})^{\operatorname {rk}_\infty = n-d} \rightarrow {{\mathbf {V}}}({{\mathcal {E}}}^{n})$ of
${{\mathbf {V}}}({{\mathcal {E}}}^{n})$ along
$U_{n-d}$. For any flat
$X_C$-scheme
$T$,
${{\mathbf {V}}}({{\mathcal {E}}}^{n})^{\operatorname {rk}_\infty = n-d}(T)$ is the set of all
$s \colon {{\mathcal {O}}}_T^{n} \rightarrow {{\mathcal {E}}}_T$ such that
$\textrm {cok}(s) \otimes C$ is projective
${{\mathcal {O}}}_T \otimes C$-module of rank
$d$. Define
$Z$ as the following cartesian product.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn7.png?pub-status=live)
Lemma 5.2.1 Let ${{\mathscr M}}_Z$ be the functor which inputs a perfectoid space
$T/C$ and outputs the set of sections of
$Z\to X_C$ over
$X_T$. Then
${{\mathscr M}}_Z$ is isomorphic to
${{\mathscr M}}_{H,\infty }^{\tau }$.
Proof. Let $T=\operatorname {Spa}(R,R^{+})$ be an affinoid perfectoid space over
$C$. The morphism
$X_T\to X_C$ is flat. (This can be checked locally:
$B_{\operatorname {dR}}^{+}(R)$ is torsion-free over the discrete valuation ring
$B_{\operatorname {dR}}^{+}(C)$, and so it is flat.) By the description in (5.2.1), an
$X_T$-point of
${{\mathscr M}}_Z$ corresponds to a morphism
$\sigma \colon \mathcal {O}_{X_T}^{n}\to {{\mathcal {E}}}_T(H)$ which has the following properties.
(1) The cokernel of
$\sigma _\infty$ is a projective
$R$-module quotient of
${{\mathcal {E}}}_T(H)_\infty$ of rank
$d$.
(2) The determinant of
$\sigma$ equals
$\tau$.
On the other hand, by Proposition 3.2.2, ${{\mathscr M}}_{H,\infty }(T)$ is the set of morphisms
$\sigma \colon \mathcal {O}_{X_T}^{n}\to {{\mathcal {E}}}_T(H)$ satisfying the following.
(
$1'$) The cokernel of
$\sigma$ is
$i_{\infty *} W$, for a projective
$R$-module quotient
$W$ of
${{\mathcal {E}}}_T(H)_\infty$ of rank
$d$.
(
$2$) The determinant of
$\sigma$ equals
$\tau$.
We claim the two sets of conditions are equivalent for a morphism $\sigma \colon \mathcal {O}_{X_T}^{n}\to {{\mathcal {E}}}_T(H)$. Clearly (
$1'$) implies (1), so that (
$1'$) and (
$2$) together imply (1) and (2) together. Conversely, suppose (1) and (2) hold. Since
$\tau$ represents a point of
${{\mathscr M}}_{\det H,\infty }$, it is an isomorphism outside of
$\infty$, and therefore so is
$\sigma$. This means that
$\operatorname {cok} \sigma$ is supported at
$\infty$. Thus
$\operatorname {cok} \sigma$ is a
$B_{\operatorname {dR}}^{+}(R)$-module. For degree reasons, the length of
$(\operatorname {cok} \sigma )\otimes _{B_{\operatorname {dR}}^{+}(R)} B_{\operatorname {dR}}^{+}(C')$ has length
$d$ for every geometric point
$\operatorname {Spa}(C',(C')^{+})\to T$. Whereas condition (1) says that
$(\operatorname {cok} \sigma )\otimes _{B_{\operatorname {dR}}^{+}(R)} R$ is a projective
$R$-module of rank
$d$. This shows that
$(\operatorname {cok} \sigma )$ is already a projective
$R$-module of rank
$d$, which is condition (
$1'$).
Lemma 5.2.2 The morphism $Z\to X_C$ is smooth.
Proof. Let $\infty '\in X_C$ be a closed point, with residue field
$C'$. It suffices to show that the stalk of
$Z$ at
$\infty '$ is smooth over
$\operatorname {Spec} B_{\operatorname {dR}}^{+}(C')$.
If $\infty '\neq \infty$, then this stalk is isomorphic to the variety
$({{\mathbf {A}}}^{n^{2}})^{\det = \tau }$ consisting of
$n\times n$ matrices with fixed determinant
$\tau$. Since
$\tau$ is invertible in
$B_{\operatorname {dR}}^{+}(C')$, this variety is smooth.
Now suppose $\infty '=\infty$. Let
$\xi$ be a generator for the kernel of
$B_{\operatorname {dR}}^{+}(C)\to C$. Then the stalk of
$Z$ at
$\infty$ is isomorphic to the flat
$B_{\operatorname {dR}}^{+}(C)$-scheme
$Y$, whose
$T$-points for a flat
$B_{\operatorname {dR}}^{+}(C)$-scheme
$T$ are
$n\times n$ matrices with coefficients in
$\Gamma (T,\mathcal {O}_T)$, which are rank
$n-d$ modulo
$\xi$, and which have fixed determinant
$\tau$ (which must equal
$u\xi ^{d}$ for a unit
$u\in B_{\operatorname {dR}}^{+}(C)$). Consider the open subset
$Y_0\subset Y$ consisting of matrices
$M$ where the first
$(n-d)$ columns have rank
$(n-d)$. Then the final
$d$ columns of
$M$ are congruent modulo
$\xi$ to a linear combination of the first
$(n-d)$ columns. After row reduction operations only depending on those first
$(n-d)$ columns,
$M$ becomes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU39.png?pub-status=live)
with $\det Q=w$ for a unit
$w\in B_{\operatorname {dR}}^{+}(C)$ which only depends on the first
$(n-d)$ columns of
$M$. We therefore have a fibration
$Y_0\to {{\mathbf {A}}}^{n(n-d)}$, namely projection onto the first
$(n-d)$ columns, whose fibers are
${{\mathbf {A}}}^{d(n-d)} \times ({{\mathbf {A}}}^{d^{2}})^{\det = w}$, which is smooth. Therefore
$Y_0$ is smooth. The variety
$Y$ is covered by opens isomorphic to
$Y_0$, and so it is smooth.
We intend to apply Theorem 4.2.1 to the morphism $Z\to X$, and so we need some preparations regarding the relative tangent space of
${{\mathbf {V}}}({{\mathcal {E}}}^{n})^{\operatorname {rk}_\infty = n-d}\to X_C$.
5.3 A linear algebra lemma
Let $f\colon V'\to V$ be a rank
$r$ linear map between
$n$-dimensional vector spaces over a field
$C$. Thus there is an exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU40.png?pub-status=live)
with $\dim W=\dim W'=n-r$.
Consider the minor map $\Lambda \colon \operatorname {Hom}(V',V)\to \operatorname {Hom}(\bigwedge ^{r+1}V',\bigwedge ^{r+1}V)$ given by
$\sigma \mapsto \bigwedge ^{r+1} \sigma$. This is a polynomial map, whose derivative at
$f$ is a linear map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU41.png?pub-status=live)
Explicitly, this map is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn8.png?pub-status=live)
Lemma 5.3.1 Let
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU42.png?pub-status=live)
be the kernel of the map $\sigma \mapsto q\circ (\sigma \vert _{W'})$. Then
$\ker D_f\Lambda = K$.
Proof. Suppose $\sigma \in K$. Since
$f$ has rank
$r$, the exterior power
$\bigwedge ^{r+1} V'$ is spanned over
$C$ by elements of the form
$v_1\wedge \cdots \wedge v_{r+1}$, where
$v_{r+1}\in \ker f = W'$. Since
$f(v_{r+1})=0$, the sum in (5.3.1) reduces to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU43.png?pub-status=live)
Since $\sigma \in K$ and
$v_{r+1}\in W'$ we have
$\sigma (v_{r+1})\in \ker q=f(V')$, which means that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU44a.png?pub-status=live)
Thus $\sigma \in \ker D_f\Lambda$.
Now suppose $\sigma \in \ker D_f\Lambda$. Let
$w\in W'$. We wish to show that
$\sigma (w)\in f(V')$. Let
$v_1,\ldots ,v_r\in V'$ be vectors for which
$f(v_1),\ldots ,f(v_r)$ is a basis for
$f(V')$. Since
$\sigma \in \ker D_f\Lambda$, we have
$D_f\Lambda (\sigma )(v_1\wedge \cdots \wedge v_r\wedge w)=0$. On the other hand,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU44.png?pub-status=live)
because all other terms in the sum in (5.3.1) are 0, owing to $f(w)=0$. Since the wedge product above is 0, and the
$f(v_i)$ are a basis for
$f(V')$, we must have
$\sigma (w)\in f(V')$. Thus
$\sigma \in K$.
We interpret Lemma 5.3.1 as the calculation of a certain normal bundle. Let $Y={{\mathbf {V}}}(\operatorname {Hom}(V',V))$ be the affine space over
$C$ representing morphisms
$V'\to V$ over a
$C$-scheme, and let
$j\colon Y^{\operatorname {rk} =r}\to Y$ be the locally closed subscheme representing morphisms which are everywhere of rank
$r$. Thus,
$Y^{\operatorname {rk} = r}$ is an open subset of the fiber over 0 of (the geometric version of) the minor map
$\Lambda$. It is well known that
$Y^{\operatorname {rk} = r}/C$ is smooth of codimension
$(n-r)^{2}$ in
$Y/C$, and so the normal bundle
$N_{Y^{\operatorname {rk}= r}/Y}$ is locally free of this rank.
We have a universal morphism of $\mathcal {O}_{Y^{\operatorname {rk}=r}}$-modules
$\sigma \colon \mathcal {O}_{Y^{\operatorname {rk} = r}}\otimes _C V'\to \mathcal {O}_{Y^{\operatorname {rk} = r}}\otimes _C V$. Let
${{\mathcal {W}}}'=\ker \sigma$ and
${{\mathcal {W}}}=\operatorname {cok} \sigma$, so that
${{\mathcal {W}}}'$ and
${{\mathcal {W}}}$ are locally free
$\mathcal {O}_{Y^{\operatorname {rk} = r}}$-modules of rank
$n-r$. We also have the
$\mathcal {O}_{Y^{\operatorname {rk} = r}}$-linear morphism
$D\Lambda \colon \mathcal {O}_{Y^{\operatorname {rk}=r}} \otimes _C \operatorname {Hom}(V',V) \rightarrow \mathcal {O}_{Y^{\operatorname {rk}=r}} \otimes _C \operatorname {Hom}(\Lambda ^{r+1}V',\Lambda ^{r+1}V)$, whose kernel is precisely
$\operatorname {Tan}_{Y^{\operatorname {rk}=r}/C}$. The geometric interpretation of Lemma 5.3.1 is a commutative diagram with short exact rows.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn9.png?pub-status=live)
5.4 Moduli of morphisms of vector bundles with fixed rank at
$\infty$
We return to the setup of § 5.1. We have a curve $X$ and a closed point
$\infty \in X$, with inclusion map
$i_\infty$ and residue field
$C$.
Let ${{\mathcal {E}}}$ and
${{\mathcal {E}}}'$ be rank
$n$ vector bundles over
$X$, with fibers
$V={{\mathcal {E}}}_\infty$ and
$V'={{\mathcal {E}}}'_\infty$. We have the geometric vector bundle
${{\mathbf {V}}}(\operatorname {\mathscr {H} {\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))\to X$. If
$f\colon T\to X$ is a morphism, then
$T$-points of
${{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))$ classify
$\mathcal {O}_T$-linear maps
$f^{*}{{\mathcal {E}}}'\to f^{*}{{\mathcal {E}}}$.
Let ${{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))^{\operatorname {rk}_\infty = r}$ be the dilatation of
${{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))$ at the locally closed subscheme
${{\mathbf {V}}}(\operatorname {Hom}(V',V))^{\operatorname {rk} =r}$ of the fiber
${{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))_\infty ={{\mathbf {V}}}(\operatorname {Hom}(V',V))$. This has the following property, for a flat morphism
$f\colon T\to X$: the
$X$-morphisms
$s\colon T\to {{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))^{\operatorname {rk}_\infty = r}$ parametrize those
$\mathcal {O}_T$-linear maps
$\sigma \colon f^{*}{{\mathcal {E}}}'\to f^{*}{{\mathcal {E}}}$, for which the fiber
$\sigma _\infty \colon f_\infty ^{*}V'\to f_\infty ^{*}V$ has rank
$r$ everywhere on
$T_\infty$.
Given a morphism $s$ as above, corresponding to a morphism
$\sigma \colon f^{*}{{\mathcal {E}}}'\to f^{*}{{\mathcal {E}}}$, we let
${{\mathcal {W}}}'$ and
${{\mathcal {W}}}$ denote the kernel and cokernel of
$\sigma _\infty$. Then
${{\mathcal {W}}}'$ and
${{\mathcal {W}}}$ are locally free
$\mathcal {O}_{T_\infty }$-modules of rank
$r$. Let
$i_{T_\infty }\colon T_\infty \to T$ denote the pull-back of
$i_\infty$ through
$f$.
We intend to use Lemma 5.1.3 to compute $s^{*}\operatorname {Tan}_{{{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))^{\operatorname {rk}_\infty =r}/X}$. The tangent bundle
$\operatorname {Tan}_{{{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))/X}$ is isomorphic to the pull-back
$f^{*}\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}})$. Also, we have identified the normal bundle
$N_{{{\mathbf {V}}}(\operatorname {Hom}(V',V))^{\operatorname {rk} =r}/{{\mathbf {V}}}(\operatorname {Hom}(V',V)}$ in (5.3.2). So when we apply the lemma to this situation, we obtain an exact sequence of
$\mathcal {O}_T$-modules
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn10.png?pub-status=live)
where the third arrow is adjoint to the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU45.png?pub-status=live)
which sends $\sigma \in \operatorname {\mathscr {H}{\mathcal{om}}\,}(f_\infty ^{*}V',f_\infty ^{*}V)$ to the composite
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU46.png?pub-status=live)
The short exact sequence in (5.4.1) identifies the $\mathcal {O}_T$-module
$s^{*}\mathrm {Tan}_{{{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}}))^{\operatorname {rk}_\infty = r}/X}$ as a modification of
$f^{*}\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}}',{{\mathcal {E}}})$ at the divisor
$T_\infty$. We can say a little more in the case that
$\sigma$ itself is a modification. Let us assume that
$\sigma$ fits into an exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU47.png?pub-status=live)
Dualizing gives another exact sequence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU48.png?pub-status=live)
Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU49.png?pub-status=live)
The kernel of $\alpha \otimes \alpha '$ can be computed in terms of
$\ker \alpha =f^{\ast }{{\mathcal {E}}}'$ and
$\ker \alpha '=f^{\ast }({{\mathcal {E}}}^{\vee })$, see Lemma 5.4.1 below. It sits in a diagram.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn11.png?pub-status=live)
Lemma 5.4.1 Let ${{\mathcal {A}}}$ be an abelian
$\otimes$-category. Let
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU50.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU51.png?pub-status=live)
be two exact sequences in ${{\mathcal {A}}}$, with
$A,A',K,K'$ projective. The homology of the complex
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU52.png?pub-status=live)
is given by $H_2=0$,
$H_1\cong \operatorname {Tor}_1(B,B')$, and
$H_0\cong B\otimes B'$. Thus,
$K''=\ker (f\otimes f'\colon A\otimes A' \to B\otimes B')$ appears in a diagram
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU53.png?pub-status=live)
where both sequences are exact.
Proof. Let $C_{\bullet }$ be the complex
$K\to A$, and let
$C'_{\bullet }$ be the complex
$K'\to A'$. Since
$C'_{\bullet }$ is a projective resolution of
$B'$, we have a Tor spectral sequence [Sta14, Tag 061Z]
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU54.png?pub-status=live)
We have $E^{2}_{0,0}=B\otimes B'$ and
$E^{2}_{0,1}=\operatorname {Tor}_1(B,B')$, and
$E^{2}_{i,j}=0$ for all other
$(i,j)$. Therefore
$H_0(C_{\bullet } \otimes C'_{\bullet }) \cong B\otimes B'$ and
$H_1(C_{\bullet } \otimes C'_{\bullet }) \cong \operatorname {Tor}_1(B,B')$, which is the lemma.
5.5 A tangent space calculation
We return to the setup of § 5.2. Thus we have fixed a $p$-divisible group
$H$ over a perfect field
$k$, and an algebraically closed perfectoid field
$C$ containing
$W(k)[1/p]$. But now we specialize to the case that
$H$ is isoclinic. Therefore
$D=\operatorname {End} H$ (up to isogeny) is a central simple
$\mathbf {Q}_p$-algebra. Let
${{\mathcal {E}}}={{\mathcal {E}}}_C(H)$; we have
$\operatorname {\mathscr {H}{\mathcal{om}}\,}({{\mathcal {E}}},{{\mathcal {E}}})\cong D\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_C}$.
Recall the scheme $Z\to X_C$, defined as a fiber product in (5.2.1). Let
$s\colon X_C\to Z$ be a section. This corresponds to a morphism
$\sigma \colon \mathcal {O}_{X_C}^{n}\to {{\mathcal {E}}}$. Let
$W'$ and
$W$ be the cokernel of
$\sigma _\infty$; these are
$C$-vector spaces.
We are interested in the vector bundle $s^{*}\mathrm {Tan}_{Z/X_C}$. This is the kernel of the derivative of the determinant map:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU55.png?pub-status=live)
We apply (5.4.2) to give a description of $s^{*}\mathrm {Tan}_{{{\mathbf {V}}}({{\mathcal {E}}}^{n})^{\operatorname {rk}_\infty = n-d}/X_C}$. We get a diagram of
$\mathcal {O}_{X_C}$-modules.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn12.png?pub-status=live)
On the other hand, the horizontal exact sequence fits into a diagram.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn13.png?pub-status=live)
The arrow labeled ${\textrm {tr}}$ is induced from the
$\mathbf {Q}_p$-linear map
$M_n(\mathbf {Q}_p)\times D\to \mathbf {Q}_p$ carrying
$(\alpha ',\alpha )$ to
${\textrm {tr}}(\alpha ')-{\textrm {tr}}(\alpha )$ (reduced trace on
$D$). The commutativity of the lower right square boils down to the identity, valid for sections
$s_1,\ldots ,s_n\in H^{0}(X_C,{{\mathcal {E}}})$ and
$\alpha \in D$:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU56.png?pub-status=live)
(There is a similar identity for $\alpha '\in M_n(\mathbf {Q}_p)$.) Because the arrow labeled
$\tau$ is injective, we can combine (5.5.1) and (5.5.2) to arrive at a description of
$s^{*}\operatorname {Tan}_{Z/X_C}$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn14.png?pub-status=live)
We pass to duals to obtain the following.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn15.png?pub-status=live)
The dotted arrow is induced from the map $(M_n(\mathbf {Q}_p)\times D)\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_C}\to {{\mathcal {E}}}^{n}$ sending
$(\alpha ',\alpha )\otimes 1$ to
$\alpha \circ \sigma - \sigma \circ \alpha '$.
Theorem 5.5.1 If $s$ is a section to
$Z \rightarrow X_C$ corresponding, under the isomorphism of Lemma 5.2.1, to a point
$x \in {{\mathcal {M}}}_{H,\infty }^{\tau }(C)$, then the following are equivalent.
(i) The vector bundle
$s^{*}\mathrm {Tan}_{Z/X_C}$ has a Harder–Narasimhan slope which is less than or equal to
$0$.
(ii) The point
$x$ lies in the special locus
${{\mathscr M}}_{H,\infty }^{\tau ,\operatorname {sp}}$.
Proof. Let $\sigma \colon \mathcal {O}_{X_C}^{n}\to {{\mathcal {E}}}$ denote the homomorphism corresponding to
$x$. Condition (1) is true if and only if
$H^{0}(X_C,s^{*}\mathrm {Tan}_{Z/X_C}^{\vee })\neq 0$. We now take
$H^{0}$ of (5.5.4), noting that
$H^{0}(X_C,{{\mathcal {F}}}^{\vee })\to H^{0}(X_C,{{\mathcal {E}}}^{n})$ is injective. We find that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU57.png?pub-status=live)
This is nonzero exactly when $x$ lies in the special locus.
Combining Theorem 5.5.1 with the criterion for cohomological smoothness in Theorem 4.2.1 proves Theorem 1.0.1 for the space ${{\mathscr M}}_{H,\infty }$.
Naturally we wonder whether it is possible to give a complete description of $s^{*}\operatorname {Tan}_{Z/X_C}$, as this is the ‘tangent space’ of
${{\mathscr M}}_{H,\infty }^{\tau }$ at the point
$x$. Note that
$s^{*}\operatorname {Tan}_{Z/X_C}$ can only have nonnegative slopes, since it is a quotient of a trivial bundle. Therefore Theorem 5.5.1 says that 0 appears as a slope of
$s^{*}\operatorname {Tan}_{Z/X_C}$ if and only if
$s$ corresponds to a special point of
${{\mathscr M}}_{H,\infty }^{\tau }$.
Example 5.5.2 Consider the case that $H$ has dimension 1 and height
$n$, so that
${{\mathscr M}}_{H,\infty }$ is an infinite-level Lubin–Tate space. Suppose that
$x\in {{\mathscr M}}_{H,\infty }(C)$ corresponds to a section
$s\colon X_C\to Z$. Then
$s^{*}\operatorname {Tan}_{Z/X_C}$ is a vector bundle of rank
$n^{2}-1$ and degree
$n-1$, with slopes lying in
$[0,1/n]$; this already limits the possibilities for the slopes to a finite list.
If $n=2$ there are only two possibilities for the slopes appearing in
$s^{*}\operatorname {Tan}_{Z/X_C}$:
$\{ {1/3} \}$ and
$\{ {0,1/2} \}$. These correspond exactly to the nonspecial and special loci, respectively.
If $n=3$, there are a priori five possibilities for the slopes appearing in
$s^{*}\operatorname {Tan}_{Z/X_C}$:
$\{ {1/4,1/4} \}$,
$\{ {1/3,1/5} \}$,
$\{ {1/3,1/3,0,0} \}$,
$\{ {2/7,0} \}$, and
$\{ {1/3,1/4,0} \}$. But in fact the final two cases cannot occur: if 0 appears as a slope, then
$x$ lies in the special locus, so that
$A_x\neq \mathbf {Q}_p$. But as
$A_x$ is isomorphic to a subalgebra of
$\operatorname {End}^{\circ } H$, the division algebra of invariant 1/3, it must be the case that
$\dim _{\mathbf {Q}_p} A_x = 3$, which forces 0 to appear as a slope with multiplicity
$\dim _{\mathbf {Q}_p} A_x/\mathbf {Q}_p = 2$. On the nonspecial locus, we suspect that the generic (semistable) case
$\{ {1/4,1/4} \}$ always occurs, as otherwise there would be some unexpected stratification of
${{\mathscr M}}_{H,\infty }^{\circ ,\operatorname {non-sp}}$. But currently we do not know how to rule out the case
$\{ {1/3,1/5} \}$.
5.6 The general case
Let ${{\mathcal {D}}} = (B,V,H,\mu )$ be a rational EL datum over
$k$, with reflex field
$E$. Let
$F$ be the center of
$B$. As in § 3.5, let
$D=\operatorname {End}_{B} V$ and
$D'=\operatorname {End}_{B} H$, so that
$D$ and
$D'$ are both
$F$-algebras.
Let $C$ be a perfectoid field containing
$\breve {E}$, and let
$\tau \in {{\mathscr M}}_{\det {{\mathcal {D}}},\infty }(C)$. Let
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\tau }$ be the fiber of the determinant map over
$\tau$. We will sketch the proof that
${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\tau }\to \operatorname {Spa} C$ is cohomologically smooth. It is along the same lines as the proof for
${{\mathscr M}}_{H,\infty }$, but with some extra linear algebra added.
The space ${{\mathscr M}}_{{{\mathcal {D}}},\infty }^{\tau }$ may be expressed as the space of global sections of a smooth morphism
$Z\to X_C$, defined as follows. We have the geometric vector bundle
${{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}_B(V\otimes _{\mathbf {Q}_p}\mathcal {O}_X, {{\mathcal {E}}}_C(H)))$. In its fiber over
$\infty$, we have the locally closed subscheme whose
$R$-points for a
$C$-algebra
$R$ are morphisms, whose cokernel is as a
$B\otimes _{{{\mathbf {Q}}}_p} R$-module isomorphic to
$V_0 \otimes _{\breve E} R$, where
$V_0$ is the weight 0 subspace of
$V\otimes _{\mathbf {Q}_p} \breve {E}$ determined by
$\mu$. We then have the dilatation
${{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}_B(V\otimes _{\mathbf {Q}_p}\mathcal {O}_{X_C}, {{\mathcal {E}}}_C(H)))^{\mu }$ of
${{\mathbf {V}}}(\operatorname {\mathscr {H}{\mathcal{om}}\,}_B(V\otimes _{\mathbf {Q}_p}\mathcal {O}_X, {{\mathcal {E}}}_C(H)))$ at this locally closed subscheme. Its points over
$S=\operatorname {Spa}(R,R^{+})$ parametrize
$B$-linear morphisms
$s\colon V\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_S}\to {{\mathcal {E}}}_S(H)$, such that (locally on
$S$) the cokernel of the fiber
$s_\infty$ is isomorphic as a
$(B\otimes _{\mathbf {Q}_p} R)$-module to
$V_0\otimes _{\breve {E}} R$. Finally, the morphism
$Z\to X_C$ is defined by the following cartesian diagram.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU58.png?pub-status=live)
Let $x\in {{\mathscr M}}_{{{\mathcal {D}}},\infty }(C)$ correspond to a
$B$-linear morphism
$s\colon V\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_C} \to {{\mathcal {E}}}_C(H)$ and a section of
$Z\to X_C$ which we also call
$s$. Define
$B\otimes _{\mathbf {Q}_p} C$-modules
$W'$ and
$W$ by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU59.png?pub-status=live)
The analogue of (5.5.4) is a diagram which computes the dual of $s^{*}\operatorname {Tan}_{Z/X_C}$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqn16.png?pub-status=live)
This time, the dotted arrow is induced from the map $(D'\times D)\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_C}\to \operatorname {\mathscr {H}{\mathcal{om}}\,}(V\otimes _{\mathbf {Q}_p} \mathcal {O}_{X_C},{{\mathcal {E}}}_C(H))$ sending
$(\alpha ',\alpha )\otimes 1$ to
$\alpha \circ s- s\circ \alpha '$. Taking
$H^{0}$ in (
) shows that $H^{0}(X_C,s^{*}\operatorname {Tan}_{Z/X_C}^{\vee }) = A_x / F$, and this is nonzero exactly when
$x$ lies in the special locus.
5.7 Proof of Corollary 1.0.2
We conclude with a discussion of the infinite-level modular curve $X(p^{\infty })$. Recall from [Reference ScholzeSch15] the following facts about the Hodge–Tate period map
$\pi _{HT}\colon X(p^{\infty })\to {{\mathbf {P}}}^{1}$. The ordinary locus in
$X(p^{\infty })$ is sent to
${{\mathbf {P}}}^{1}(\mathbf {Q}_p)$. The supersingular locus is isomorphic to finitely many copies of
${{\mathscr M}}_{H,\infty ,C}$, where
$H$ is a connected
$p$-divisible group of height 2 and dimension 1 over the residue field of
$C$; the restriction of
$\pi _{HT}$ to this locus agrees with the
$\pi _{HT}$ we had already defined on each
${{\mathscr M}}_{H,\infty ,C}$.
We claim that the following are equivalent for a $C$-point
$x$ of
$X(p^{\infty })^{\circ }$.
(i) The point
$x$ corresponds to an elliptic curve
$E/\mathcal {O}_C$, such that the
$p$-divisible group
$E[p^{\infty }]$ has
$\operatorname {End} E[p^{\infty }]=\mathbf {Z}_p$.
(ii) The stabilizer of
$\pi _{HT}(x)$ in
$\operatorname {PGL}_2(\mathbf {Q}_p)$ is trivial.
(iii) There is a neighborhood of
$x$ in
$X(p^{\infty })^{\circ }$ which is cohomologically smooth over
$C$.
First we discuss the equivalence of (1) and (2). If $E$ is ordinary, then
$E[p^{\infty }]\cong \mathbf {Q}_p/\mathbf {Z}_p \times \mu _{p^{\infty }}$ certainly has endomorphism ring larger than
$\mathbf {Z}_p$, so that (1) is false. Meanwhile, the stabilizer of
$\pi _{HT}(x)$ in
$\operatorname {PGL}_2(\mathbf {Q}_p)$ is a Borel subgroup, so that (2) is false as well. The equivalence between (1) and (2) in the supersingular case is a special case of the equivalence discussed in § 3.5.
Theorem 1.0.1 tells us that ${{\mathscr M}}_{H,\infty }^{\circ ,\operatorname {non-sp}}$ is cohomologically smooth, which implies that shows that (2) implies (3). We therefore are left with showing that if (2) is false for a point
$x\in X(p^{\infty })^{\circ }$, then no neighborhood of
$x$ is cohomologically smooth.
First suppose that $x$ lies in the ordinary locus. This locus is fibered over
${{\mathbf {P}}}^{1}(\mathbf {Q}_p)$. Suppose
$U$ is a sufficiently small neighborhood of
$x$. Then
$U$ is contained in the ordinary locus, and so
$\pi _0(U)$ is nondiscrete. This implies that
$H^{0}(U,{{\mathbf {F}}}_\ell )$ is infinite, and so
$U$ cannot be cohomologically smooth.
Now suppose that $x$ lies in the supersingular locus, and that
$\pi _{HT}(x)$ has nontrivial stabilizer in
$\operatorname {PGL}_2(\mathbf {Q}_p)$. We can identify
$x$ with a point in
${{\mathscr M}}_{H,\infty }^{\circ ,\operatorname {sp}}(C)$. We intend to show that every neighborhood of
$x$ in
${{\mathscr M}}_{H,\infty }^{\circ }$ fails to be cohomologically smooth.
Not knowing a direct method, we appeal to the calculations in [Reference WeinsteinWei16], which constructed semistable formal models for each ${{\mathscr M}}_{H,m}^{\circ }$. The main result we need is Theorem 5.1.2, which uses the term ‘CM points’ for what we have called special points. There exists a decreasing basis of neighborhoods
$Z_{x,0}\supset Z_{x,1}\supset \cdots$ of
$x$ in
${{\mathscr M}}_{H,\infty }^{\circ }$. For each affinoid
$Z=\operatorname {Spa}(R,R^{+})$, let
$\overline {Z}=\operatorname {Spec} R^{+}\otimes _{\mathcal {O}_C} \kappa$, where
$\kappa$ is the residue field of
$C$. For each
$m\geq 0$, there exists a nonconstant morphism
$\overline {Z}_{x,m}\to C_{x,m}$, where
$C_{x,m}$ is an explicit nonsingular affine curve over
$\kappa$. This morphism is equivariant for the action of the stabilizer of
$Z_{x,m}$ in
$\operatorname {SL}_2(\mathbf {Q}_p)$. For infinitely many
$m$, the completion
$C_{x,m}^\textrm {{cl}}$ of
$C_{x,m}$ is a projective curve with positive genus.
Let $U\subset {{\mathscr M}}_{H,\infty }^{\circ }$ be an affinoid neighborhood of
$x$. Then there exists
$N\geq 0$ such that
$Z_{x,m}\subset U$ for all
$m\geq N$. Let
$K\subset \operatorname {SL}_2(\mathbf {Q}_p)$ be a compact open subgroup which stabilizes
$U$, so that
$U/K$ is an affinoid subset of the rigid-analytic curve
${{\mathscr M}}_{H,\infty }^{\circ }/K$. For each
$m\geq N$, let
$K_m\subset K$ be the stabilizer of
$Z_{x,m}$, so that
$K_m$ acts on
$C_{x,m}$.
There exists an integral model of $U/K$ whose special fiber contains as a component the completion of each
$\overline {Z}_{x,m}/K_m$ which has positive genus. Since there is a nonconstant morphism
$\overline {Z}_{x,m}/K_m\to C_{x,m}/K_m$, we must have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU60.png?pub-status=live)
Now we take a limit as $K$ shrinks. Since
$U\sim \varprojlim U/K$, we have
$H^{1}(U,{{\mathbf {F}}}_\ell )\cong \varinjlim H^{1}(U/K,{{\mathbf {F}}}_\ell )$. Also, for each
$m$, the action of
$K_m$ on
$C_{x,m}$ is trivial for all sufficiently small
$K$. Therefore
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20201123180221180-0641:S0010437X20007332:S0010437X20007332_eqnU61.png?pub-status=live)
This shows that $U$ is not cohomologically smooth.
Acknowledgements
The authors want to thank Peter Scholze for his help and his interest in their work. Also they thank Andreas Mihatsch for pointing out a mistake in a previous version of the manuscript. The first named author was supported by Peter Scholze's Leibniz Preis. The second author was supported by NSF Grant No. DMS-1440140 while in residence at the Mathematical Sciences Research Institute in Berkeley, California.