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

Finiteness of for some Elliptic Curves of Analytic Rank >1>1

Paul Hamacher, Ziquan Yang, and Xiaolei Zhao
Abstract

We prove the finiteness of (i.e., the BSD conjecture) for a class of elliptic curves over function fields. This is an application of a more general theorem that the Tate conjecture in codimension 11 is “generically true” for mod pp reductions of complex projective varieties with h2,0=1h^{2,0}=1, under a mild assumption on moduli. We also prove the Tate conjecture for a class of algebraic surfaces of general type with pg=1p_{g}=1.

1 Introduction

For an elliptic curve \mathcal{E} over a global field, it is conjectured that the Tate-Shafarevich group is finite. This is known in many cases, and in particular for function fields, when the analytic rank of \mathcal{E} is at most 11 ([Ulmer, Qiu, Kolyvagin]). However, there are very few examples when the conjecture is known when the analytic rank is >1>1. For function fields the only previously known examples correspond to elliptic surfaces which are rational, K3 ([ASD]), or dominated by products of curves (e.g., [UlmerLargeRank]).

One of the purposes of this paper is to prove this finiteness for a new class of examples in the function field case. Let kk be a finite field of characteristic p.p. Then we show

Theorem A.

Assume p5p\geq 5. Let π:XC\pi\colon X\to C be a minimal elliptic surface (with a zero section) over a smooth proper kk-curve CC of genus 11 and =Xη\mathcal{E}=X_{\eta} be the fiber over the generic point ηC\eta\in C. Suppose that

  1. (i)

    ht()=1\mathrm{ht}(\mathcal{E})=1;

  2. (ii)

    all geometric fibers of π\pi are irreducible.

Then ()\Sha(\mathcal{E}) is finite.

Here ht()\mathrm{ht}(\mathcal{E}) denotes the height of \mathcal{E}, i.e., the degree of the fundamental line bundle L(R1π𝒪X)L\coloneqq(R^{1}\pi_{*}\mathcal{O}_{X})^{\vee}. We remark that by the Shioda-Tate formula, condition (ii) is equivalent to rankNS(X)=rankMW()+2\mathrm{rank\,}\mathrm{NS}(X)=\mathrm{rank\,}\mathrm{MW}(\mathcal{E})+2, where NS(X)\mathrm{NS}(X) is the Néron-Severi group of XX and MW()\mathrm{MW}(\mathcal{E}) is the Mordell-Weil group of \mathcal{E}. Moreover, on the coarse moduli space of π:XC\pi\colon X\to C satisfying all conditions except possibly (ii) in the theorem, (ii) is generically satisfied, and there are infinitely many 99-dimensional families of \mathcal{E} such that rankMW()2\mathrm{rank\,}\mathrm{MW}(\mathcal{E})\geq 2 up to replacing kk by a finite extension (see (7.4.5)).

For function fields, it is known that the finiteness of is equivalent to the BSD conjecture ([KatoTrihan, MilneAT]), which is further equivalent to the Tate conjecture for the corresponding elliptic surfaces. In fact, Thm A is obtained by refining a special case of a more general theorem on the Tate conjecture. To explain this, we consider arithmetic families of the following type: Suppose that KK is a field of characteristic 0 and SS is a smooth connected KK-variety. We say that a smooth projective morphism f:𝒳Sf:\mathcal{X}\to S is a \heartsuit-family if for every geometric point sSs\to S, the fiber 𝒳s\mathcal{X}_{s} is connected, dimH0(Ω2𝒳s)=1\dim\mathrm{H}^{0}(\Omega^{2}_{\mathcal{X}_{s}})=1 (i.e., h2,0(𝒳s)=1h^{2,0}(\mathcal{X}_{s})=1), and the Kodaira-Spencer map

:TS/KHom¯(R1fΩ1𝒳/S,R2f𝒪𝒳)\nabla\colon T_{S/K}\to\underline{\mathrm{Hom}}(R^{1}f_{*}\Omega^{1}_{\mathcal{X}/S},R^{2}f_{*}\mathcal{O}_{\mathcal{X}}) (1)

is nontrivial. Our general theorem states:

Theorem B.

Let 𝖬\mathsf{M} be a connected and separated scheme over Spec()\mathrm{Spec}(\mathbb{Z}) which is smooth and of finite type and let f:𝒳𝖬f\colon\mathcal{X}\to\mathsf{M} be a smooth projective morphism with geometrically connected fibers. If the restriction of 𝒳\mathcal{X} to 𝖬\mathsf{M}_{\mathbb{Q}} is a \heartsuit-family, then for p0p\gg 0, every fiber of 𝒳\mathcal{X} over a point s𝖬𝔽ps\in\mathsf{M}_{\mathbb{F}_{p}} satisfies the Tate conjecture in codimension 11, i.e., for every prime p\ell\neq p, the Chern class map

c1:Pic(𝒳s)H2e´t(𝒳s¯,(1))Gal(s¯/s)c_{1}:\mathrm{Pic}\,(\mathcal{X}_{s}){\otimes}\mathbb{Q}_{\ell}\to\mathrm{H}^{2}_{\mathrm{{\acute{e}}t}}(\mathcal{X}_{\bar{s}},\mathbb{Q}_{\ell}(1))^{\mathrm{Gal}(\bar{s}/s)}

is surjective, where s¯\bar{s} is a geometric point over ss.111Over a finite field there is also a crystalline version of the Tate conjecture, which in codimension 11 is known to be equivalent to the \ell-adic version (see [Morrow, Prop. 4.1]).

The above theorem is roughly saying that under a mild assumption, the Tate conjecture in codimension 11 is “generically true” in an arithmetic family of varieties with h2,0=1h^{2,0}=1. Clearly, it is both a positive characteristic analogue of Moonen’s main result in [Moonen], and a generalization of the Tate conjecture for K3 surfaces, for which people made great progress in the past decade (e.g., [MPTate, Maulik, Charles]). The base scheme 𝖬\mathsf{M} above should be thought of as the moduli of varieties of a certain type. Note that we do not require in the theorem above that ss is a closed point—it is allowed to have positive transcendence degree over 𝔽p\mathbb{F}_{p}. Also, although the above theorem is non-effective in pp, for concretely given families we often can make it effective, as in the case of Thm A. To illustrate this, we also analyze a class of surfaces of general type:

Theorem C.

Assume that kk is a field finitely generated over 𝔽p\mathbb{F}_{p} for p5p\geq 5. Let XX be a minimal smooth projective geometrically connected surface over kk. Let KXK_{X} denote its canonical divisor, pgp_{g} the geometric genus h0(KX)h^{0}(K_{X}), and qq the irregularity h1(𝒪X)h^{1}(\mathcal{O}_{X}). If pg=KX2=1p_{g}=K_{X}^{2}=1, q=0q=0 and KXK_{X} is ample, then XX satisfies the Tate conjecture.

We remark that condition (ii) in Thm A is an analogue of the condition “KXK_{X} is ample” above. Over \mathbb{C}, surfaces with the above invariants were classified by Todorov and Catanese ([Todorov, Catanese0]). They are simply connected, have a coarse moduli space of dimension 1818 and were among the first examples of pg=1p_{g}=1 surfaces for which both the local and the global Torelli theorems fail ([Catanese]).

Sketch of Proofs

We first explain how to prove Theorem B. We build on the overall strategy of Madapusi-Pera [MPTate], which has two main steps. The first is to construct an integral period morphism ρ:𝖬𝒮\rho:\mathsf{M}\to\mathscr{S}, where 𝒮\mathscr{S} is the canonical integral model of a Shimura variety Sh(G)\mathrm{Sh}(G) defined by a suitable special orthogonal group GG.222For the exposition of ideas, we temporarily suppress Hermitian symmetric domains and level structures from the notation of Shimura varieties. Up to passing to its spinor cover, 𝒮\mathscr{S} is equipped with a family of abelian schemes 𝒜\mathcal{A}. The second is to construct, for each geometric point s𝖬s\to\mathsf{M} and t=ρ(s)t=\rho(s), a morphism θ:L(𝒜t)NS(𝒳s)\theta:\mathrm{L}(\mathcal{A}_{t})\to\mathrm{NS}(\mathcal{X}_{s}), where L(𝒜t)\mathrm{L}(\mathcal{A}_{t}) is a distinguished subspace of End(𝒜t)\mathrm{End}(\mathcal{A}_{t}). Then the Tate conjecture for NS(𝒳s)\mathrm{NS}(\mathcal{X}_{s}) follows from a variant of Tate’s theorem for L(𝒜t)\mathrm{L}(\mathcal{A}_{t}). As a crucial step, Madapusi-Pera proved that ρ\rho is étale. This boils down to the geometric fact that a K3 surface XX has unobstructed deformation and H1(ΩX)H1(TX)\mathrm{H}^{1}(\Omega_{X})\,{\cong}\,\mathrm{H}^{1}(T_{X}). Unfortunately, this is rarely true when XX is not a close relative of a hyperkähler variety.

The main contribution of our paper is a method to remove the dependence of the above strategy on any good local property of ρ\rho (not even flatness). Indeed, the condition on the Kodaira-Spencer map contained in the definition of a \heartsuit-family just amounts to asking dimim(ρ)>0\dim\mathrm{im}(\rho_{\mathbb{C}})>0, so even the situation dim𝖬<dim𝒮\dim\mathsf{M}<\dim\mathscr{S} is allowed in Thm B. Below we explain in more detail the difficulties in extending the two steps above and how to overcome them.

(1.0.1)   

It is not hard to construct a period morphism ρ:𝖬Sh(G)\rho_{\mathbb{C}}:\mathsf{M}_{\mathbb{C}}\to\mathrm{Sh}(G)_{\mathbb{C}} over \mathbb{C}, as the target is a moduli of variations of Hodge structures with some additional data. As in [MPTate], the idea to construct ρ\rho is to descend ρ\rho_{\mathbb{C}} to a morphism ρ\rho_{\mathbb{Q}} over \mathbb{Q}, and then appeal to the extension property of 𝒮\mathscr{S}. If for every s𝖬()s\in\mathsf{M}(\mathbb{C}), the motive 𝔥2(𝒳s)\mathfrak{h}^{2}(\mathcal{X}_{s}) defined by H2(𝒳s,)\mathrm{H}^{2}(\mathcal{X}_{s},\mathbb{Q}) is an abelian motive (e.g., in the K3 case), then one shows that the action of ρ\rho_{\mathbb{C}} on \mathbb{C}-points are Aut()\mathrm{Aut}(\mathbb{C})-equivariant, so that ρ\rho_{\mathbb{C}} descends to \mathbb{Q} just as in [MPTate].

To treat the general case, we absorb inputs from Moonen’s work [Moonen]. Let b𝖬()b\in\mathsf{M}(\mathbb{C}) be a point lying over the generic point of 𝖬\mathsf{M}, and let T(𝒳b)\mathrm{T}(\mathcal{X}_{b}) be the orthogonal complement of (0,0)(0,0)-classes in H2(𝒳b,(1))\mathrm{H}^{2}(\mathcal{X}_{b},\mathbb{Q}(1)). Let EE be the endomorphism algebra of the Hodge structure T(𝒳b)\mathrm{T}(\mathcal{X}_{b}), which is known to be either a totally real or a CM field. In the latter case, one can still show that 𝔥2(𝒳s)\mathfrak{h}^{2}(\mathcal{X}_{s}) is abelian for each ss (see (2.2.7)), so that the argument of [MPTate] still applies. In the former case, we need to consider an auxiliary Shimura subvariety Sh(𝒢)Sh(G)\mathrm{Sh}(\mathcal{G})_{\mathbb{C}}\subseteq\mathrm{Sh}(G)_{\mathbb{C}}, defined by the Weil restriction 𝒢\mathcal{G} of a special orthogonal group over EE.

Up to replacing 𝖬\mathsf{M} by a connected étale cover and bb by a lift, the restriction of ρ\rho_{\mathbb{C}} to 𝖬\mathsf{M}^{\circ} factors through a morphism ϱ:𝖬Sh(𝒢)\varrho_{\mathbb{C}}:\mathsf{M}^{\circ}\to\mathrm{Sh}(\mathcal{G})_{\mathbb{C}}, where 𝖬\mathsf{M}^{\circ} is the connected component of 𝖬\mathsf{M}_{\mathbb{C}} containing bb. We will show that the field of definition FF of 𝖬\mathsf{M}^{\circ} always contains EE. Interestingly, to descend ρ\rho_{\mathbb{C}} to \mathbb{Q} it suffices to descend ϱ\varrho_{\mathbb{C}} to FF. The trick is to consider the left adjoint to the base change functor from \mathbb{Q}-schemes to FF-schemes. To show that ϱ\varrho_{\mathbb{C}} descends to FF, consider the submotive 𝔱(𝒳b)\mathfrak{t}(\mathcal{X}_{b}) of 𝔥2(𝒳b)\mathfrak{h}^{2}(\mathcal{X}_{b}) defined by T(𝒳b)\mathrm{T}(\mathcal{X}_{b}). When dimET(𝒳b)\dim_{E}\mathrm{T}(\mathcal{X}_{b}) is odd, although we do not know that 𝔱(𝒳b)\mathfrak{t}(\mathcal{X}_{b}) (or equivalently 𝔥2(𝒳b)\mathfrak{h}^{2}(\mathcal{X}_{b})) is abelian, Moonen’s work [Moonen] tells us that (a slight variant of) “the E/E/\mathbb{Q}-norm” NmE/(𝔱(𝒳b))\mathrm{Nm}_{E/\mathbb{Q}}(\mathfrak{t}(\mathcal{X}_{b})) is abelian. This allows us to show that ϱ\varrho_{\mathbb{C}} descends to FF because considering NmE/(𝔱(𝒳b))\mathrm{Nm}_{E/\mathbb{Q}}(\mathfrak{t}(\mathcal{X}_{b})) amounts to considering a different faithful representation of 𝒢\mathcal{G}. As we document in a separate reference file [YangSystem], under some hypotheses the canonical models of Shimura varieties of abelian type over their reflex fields have a moduli interpretation attached to any faithful representation. When dimET(𝒳b)\dim_{E}\mathrm{T}(\mathcal{X}_{b}) is even, some adaptation is needed, which we omit here.

(1.0.2)   

Next, we explain how to construct θ:L(𝒜t)NS(𝒳s)\theta:\mathrm{L}(\mathcal{A}_{t})\to\mathrm{NS}(\mathcal{X}_{s}), where the main novelty of our method lies. As in [MPTate], the key to construct θ\theta is to find, for each ζL(𝒜t)\zeta\in\mathrm{L}(\mathcal{A}_{t}), a characteristic 0 point t~\widetilde{t} on 𝒮\mathscr{S} lifting tt, such that (i) ζ\zeta deforms to 𝒜t~\mathcal{A}_{\widetilde{t}} and (ii) t~\widetilde{t} comes from a lifting s~\widetilde{s} of ss via ρ\rho. The existence of t~\widetilde{t} which satisfies (i) was already shown in [CSpin]. When ρ\rho is étale (or at least smooth), (ii) is then automatically satisfied. This is where [MPTate] crucially relies on the étaleness of ρ\rho.

The challenge to generalize this, especially when dim𝖬<dim𝒮\dim\mathsf{M}<\dim\mathscr{S}, is that there is no general way to characterize locally the image of ρ\rho in 𝒮\mathscr{S}, so for any given t~\widetilde{t} one cannot decide directly whether it satisfies (ii) or not. Indeed, this is essentially a local Schottky problem, which is famously hard. To overcome this, we revisit Deligne’s insight for [Del02, Thm 1.6] but replace the local crystalline analysis by a global and topological argument.

Pretend for a moment that 𝖬\mathsf{M}_{\mathbb{C}} and 𝖬k\mathsf{M}_{k} are both connected, where kk is the (separably closed) field defining ss, and set p=charkp=\mathrm{char\,}k. Let η¯\bar{\eta} and η¯p\bar{\eta}_{p} be geometric generic points over 𝖬\mathsf{M}_{\mathbb{C}} and 𝖬k\mathsf{M}_{k} respectively. Pick a prime p\ell\neq p and restrict f:𝒳𝖬f:\mathcal{X}\to\mathsf{M} to (p)\mathbb{Z}_{(p)}. Suppose for the sake of contradiction that for some ζ\zeta, there is no lifting t~\widetilde{t} which satisfies (i) and (ii). Then using that the deformation of ζ\zeta is controlled by a single equation, we can show that ζ\zeta deforms along the formal completion of 𝖬k\mathsf{M}_{k} at ss, and hence gives rise to an element of L(𝒜ρ(η¯p))\mathrm{L}(\mathcal{A}_{\rho(\bar{\eta}_{p})}). This further induces an element in (R2f)η¯p(R^{2}f_{*}\mathbb{Q}_{\ell})_{\bar{\eta}_{p}} which is stabilized by an open subgroup of π1e´t(𝖬k,η¯p)\pi_{1}^{\mathrm{{\acute{e}}t}}(\mathsf{M}_{k},\bar{\eta}_{p}). On the other hand, we show using the theorem of the fixed part that all elements of (R2f)η¯(R^{2}f_{*}\mathbb{Q}_{\ell})_{\bar{\eta}} stabilized by an open subgroup of π1(𝖬,η¯)\pi_{1}(\mathsf{M}_{\mathbb{C}},\bar{\eta}) come from L(𝒜η¯)\mathrm{L}(\mathcal{A}_{\bar{\eta}}){\otimes}\mathbb{Q}_{\ell}. Therefore, to derive a constradiction it suffices to show that 𝖬k\mathsf{M}_{k} does not have more “π1\pi_{1}-invariants” than 𝖬\mathsf{M}_{\mathbb{C}}.

Using Hironaka’s resolution of singularities and a spreading out argument, we can find an open subscheme UU of Spec()\mathrm{Spec\,}(\mathbb{Z}) such that 𝖬U\mathsf{M}_{U} admits a compactification whose boundary is a relative normal crossing divisor. Then we apply Grothendieck’s specialization theorems for tame fundamental group and Abhyankar’s lemma to show that 𝖬k\mathsf{M}_{k} indeed cannot have more “π1\pi_{1}-invariants” than 𝖬\mathsf{M}_{\mathbb{C}}, when (p)U(p)\in U.333We thank Aaron Landesman for pointing out to us the applicability of Abhyankar’s lemma, which simplifies our original argument. This proves Thm B.

(1.0.3)   

To prove Theorem A and C, we need to avoid the spreading out argument above. Although compactifying moduli spaces is in general a hard geometric problem, one can find for 𝖬\mathsf{M} in question a partial compactification. That is, a morphism 𝖬B\mathsf{M}\to B and a smooth proper 𝒫\mathscr{P} over BB such that 𝖬\mathsf{M} is an open subscheme of 𝒫\mathscr{P}. Then we can find lots of smooth proper curves in 𝒫\mathscr{P}. If the boundary 𝔇=𝒫𝖬\mathfrak{D}=\mathscr{P}-\mathsf{M} is generically reduced modulo pp, then we can find a curve on 𝖬k\mathsf{M}_{k} which deforms to characteristic 0 such that by looking at the curve we can already prove that 𝖬k\mathsf{M}_{k} does not have more “π1\pi_{1}-invariants” than 𝖬\mathsf{M}_{\mathbb{C}}. Of course, such curve needs to be chosen wisely, and we do this by repeatedly applying the Baire category theorem. We will also need ρ\rho to satisfy a stronger condition, namely dimim(ρ)>dimB\dim\mathrm{im}(\rho_{\mathbb{C}})>\dim B_{\mathbb{C}}, as opposed to just dimim(ρ)>0\dim\mathrm{im}(\rho_{\mathbb{C}})>0. This is known to hold for the surfaces in question.

The boundary 𝔇\mathfrak{D} is essentially a discriminant scheme, i.e., 𝒳\mathcal{X} extends to a family over 𝒫\mathscr{P}, and 𝔇\mathfrak{D} is precisely the locus where the extension fails to be smooth. In general, it is possible for a discriminant scheme over \mathbb{Z} to be generically non-reduced modulo a certain prime (cf. [Saitodisc, Thm 4.2]), so the task is to determine an effective range of pp for which this does not happen. Drawing ideas from enumerative geometry, we show that this happens only when a general fiber over 𝔇𝔽¯p\mathfrak{D}_{\bar{\mathbb{F}}_{p}} is “more singular” than that over 𝔇\mathfrak{D}_{\mathbb{C}}. To exclude this possibility when p5p\geq 5 for the surfaces in Thm C, it suffices to adapt Katz’ results on Lefschetz pencils. Doing this for Thm A is much more involved. In particular, we need to develop some nonlinear Bertini theorems (see §7.2) tailored to handle Weierstrass equations, the key input being Kodaira’s classification for the singular fibers in an elliptic fibration.

Remark (1.0.4).

Recently Fu and Moonen proved the Tate conjecture for Gushel-Mukai varieties in characteristic p5p\geq 5. The middle cohomology of these varieties behaves like that of a K3 surface up to a Tate twist. An earlier version of the paper also discussed these varieties. For our method, these varieties can be treated in a similar way as the surfaces in Thm C. However, as Fu-Moonen [FuMoonen] gave much more thorough treatment of these varieties and proved that the relevant integral period morphism ρ\rho is indeed smooth, we removed the section from the current version. In general, it is hard to determine whether ρ\rho is smooth integrally even when ρ\rho_{\mathbb{C}} is smooth, as one can tell from [FuMoonen], but when this can be achieved, ρ\rho has many more potential applications than the divisorial Tate conjecture (e.g., CM lifting and the Tate conjecture for self-correspondences, as shown in [IIK]).

On the other hand, our main purpose is to deal with the situation when the smoothness of ρ\rho cannot be hoped for. In particular, Thm B implies that the characteristic pp counterparts of the surfaces in Moonen’s list [Moonen, Thm 9.4] satisfy the Tate conjecture for p0p\gg 0, and we expect that our methods for the refinements Thm A and C can be adapted to make Thm B effective for most classes of these surfaces.

Organization of Paper

In section 2, we review and mildly extend Moonen’s results in [Moonen] on the motives of fibers of \heartsuit-families. In particular, we recap the norm functors used in loc. cit. In section 3, we discuss the moduli interpretations of Shimura varieties of abelian type over their reflex fields, as documented in [YangSystem], and recap the integral models of orthogonal Shimura varieties from [CSpin]. In section 4, we construct the period morphisms for \heartsuit-families in characteristic 0, using results from sections 2 and 3. In section 5, we prove Thm B after giving a more effective version (see (5.3.3)). In section 6, we set up some basic tools to analyze deformation of curves on parameter spaces, including applications of the Baire category theorem. Finally, in section 7 and 8, we use tools from section 6 as well as (5.3.3) to prove Thm A and C respectively. In particular, in section 7 we study the geometry of natural parameter spaces of elliptic surfaces, which we hope to be of independent interest.

(1.0.5)   

Finally we introduce some notations and conventions.
(a) Let f:XSf:X\to S be a morphism between schemes. If TST\to S is another morphism, denote by XTX_{T} the base change X×STX\times_{S}T and by X(T)X(T) the set of morphisms TXT\to X as SS-schemes. By a geometric fiber of XX, we mean XsX_{s} for some geometric point sSs\to S. If xx is a point (resp. geometric point) on a scheme XX, we write k(x)k(x) for its residue field (resp. field of definition). By a variety over a field kk we mean a scheme which is reduced, separated and of finite type over kk.
(b) The letters pp and \ell will always denote some prime numbers and p\ell\neq p unless otherwise noted. We write ^\widehat{\mathbb{Z}} for the profinite completion of \mathbb{Z} and ^p\widehat{\mathbb{Z}}^{p} for its prime-to-pp part. Define 𝔸f:=^\mathbb{A}_{f}:=\widehat{\mathbb{Z}}{\otimes}\mathbb{Q} and 𝔸pf:=^p\mathbb{A}^{p}_{f}:=\widehat{\mathbb{Z}}^{p}{\otimes}\mathbb{Q}. If kk is a perfect field of characteristic pp, we write W(k)W(k) for its ring of Witt vectors.
(c) For a field kk of characteristic 2\neq 2, we consider a quadratic form qq on a finite dimensional kk-vector space VV simultaneously a symmetric bilinear pairing ,\langle-,-\rangle such that q(v)=v,vq(v)=\langle v,v\rangle for every vVv\in V.
(d) For a finite free module MM over a ring RR, we write MM^{\otimes} the direct sum of all the RR-modules which can be formed from MM by taking duals, tensor products, symmetric powers and exterior powers. We also use this notation for sheaves of modules on some Grothendieck topology whenever it makes an obvious sense.
(e) We use the following abbreviations: VHS for “variations of Hodge structures”, LHS (resp. RHS) for “left (resp. right) hand side”, and ODP for “ordinary double point”. Unless otherwise noted, the local system in a VHS is a local system of \mathbb{Q}-vector spaces. Moreover, we always assume that the VHS is pure, i.e., it is a direct sum of those pure of some weight (or the weight filtration is split).

2 Preliminaries

2.1   Motives and Norm Functors

(2.1.1)   

Let kk be a field of characteristic 0. We denote by 𝖬𝗈𝗍AH(k)\mathsf{Mot}_{\mathrm{AH}}(k) the neutral \mathbb{Q}-linear Tannakian category of motives over kk with morphisms defined by absolute Hodge correspondences (cf. [Pan94, §2] where it is denoted by k\mathscr{M}_{k}). We have the Tate objects 𝟏(n)\mathbf{1}(n) for every nn\in\mathbb{Z} in this category. For any object M𝖬𝗈𝗍AH(k)M\in\mathsf{Mot}_{\mathrm{AH}}(k), we write M(n)M(n) for M𝟏(n)M{\otimes}\mathbf{1}(n) and by an absolute Hodge cycle on MM we mean a morphism 𝟏M\mathbf{1}\to M.

Following [MPTate, §2], we denote by ω\omega_{\ell} the \ell-adic realization functor which sends 𝖬𝗈𝗍AH(k)\mathsf{Mot}_{\mathrm{AH}}(k) to the category of finite dimensional \mathbb{Q}_{\ell}-vector spaces with an action of Galk:=Gal(k¯/k)\mathrm{Gal}_{k}:=\mathrm{Gal}(\bar{k}/k), where k¯\bar{k} is some chosen algebraic closure of kk. Putting ω\omega_{\ell}’s together we obtain ωe´t\omega_{\mathrm{{\acute{e}}t}}, which takes values in the category of finite free 𝔸f\mathbb{A}_{f}-modules with a Galk\mathrm{Gal}_{k}-action. Let ωdR\omega_{\mathrm{dR}} denote the de Rham realization functor, which takes values in the category of filtered kk-vector spaces. If k=k=\mathbb{C}, we additionally consider the Betti realization ωB\omega_{B} (resp. the Hodge realization ωHdg\omega_{\mathrm{Hdg}}) which takes values in the category of \mathbb{Q}-vector spaces (resp. Hodge structures). For a smooth projective variety XX over kk, 𝔥i(X)\mathfrak{h}^{i}(X) denotes the object such that ω?(𝔥i(X))\omega_{?}(\mathfrak{h}^{i}(X)) is the iith ??-cohomology of XX, for ?=B,,dR?=B,\ell,\mathrm{dR} whenever applicable.

Let 𝖬𝗈𝗍𝖠𝖻(k)𝖬𝗈𝗍AH(k)\mathsf{Mot}_{\mathsf{Ab}}(k)\subseteq\mathsf{Mot}_{\mathrm{AH}}(k) be the full Tannakian sub-category generated by the Artin motives and the motives attached to abelian varieties. We will repeatedly make use of the following fact ([LNM900, Ch I], cf. [MPTate, Thm 2.3]):

Theorem (2.1.2).

The functor ωHdg\omega_{\mathrm{Hdg}} is fully faithful when restricted to 𝖬𝗈𝗍𝖠𝖻()\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}). In particular, for every M𝖬𝗈𝗍𝖠𝖻()M\in\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}), every element sωB(M)Fil0ωdR(M)s\in\omega_{B}(M)\cap\mathrm{Fil}^{0}\omega_{\mathrm{dR}}(M) is given by an absolute Hodge cycle.

We often refer to objects in 𝖬𝗈𝗍𝖠𝖻(k)\mathsf{Mot}_{\mathsf{Ab}}(k) as abelian motives.

(2.1.3)   

We will often consider the automorphism σ-{\otimes}_{\sigma}\mathbb{C} on 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}) defined by an element σAut()\sigma\in\mathrm{Aut}(\mathbb{C}) (cf. [LNM900, §II 6.7], see also [MPTate, Prop. 2.2]). For M𝖬𝗈𝗍AH()M\in\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}), we write MσM^{\sigma} for MσM{\otimes}_{\sigma}\mathbb{C}. Base change properties of étale (resp. de Rham) cohomology give us a 𝔸f\mathbb{A}_{f}-linear (resp. σ\sigma-linear) canonical isomorphism ωe´t(M)bcωe´t(Mσ)\omega_{\mathrm{{\acute{e}}t}}(M)\,{\cong}\,_{\mathrm{bc}}\omega_{\mathrm{{\acute{e}}t}}(M^{\sigma}) (resp. ωdR(M)bcωdR(Mσ)\omega_{\mathrm{dR}}(M)\,{\cong}\,_{\mathrm{bc}}\omega_{\mathrm{dR}}(M^{\sigma})). Here the subscript “bc” is short for “base change”. For an absolute Hodge class sωB(M)s\in\omega_{B}(M), we write sσs^{\sigma} for the class in ωB(Mσ)\omega_{B}(M^{\sigma}) which has the same étale and de Rham realizations as ss under the “bc\,{\cong}\,_{\mathrm{bc}}” isomorphisms.

Finally, we remark that for M𝖬𝗈𝗍AH()M\in\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}), it makes sense to say whether a tensor sωB(M)s\in\omega_{B}(M)^{\otimes} (see (1.0.5)(d)) is absolute Hodge, because ss has to lie in ωB\omega_{B} of a finite direct sum of tensorial constructions on MM. And when ss is absolute Hodge, we may form sσωB(Mσ)s^{\sigma}\in\omega_{B}(M^{\sigma})^{\otimes} for σAut()\sigma\in\mathrm{Aut}(\mathbb{C}), extending the notation in the previous paragraph.

(2.1.4)   

Next, we recall the basics of norm functors. The reader may refer to [Moonen, §3] for more details. Let kk be a field of characteristic 0 and EE be a finite field extension of kk. Let 𝒞\mathscr{C} be any Tannakian kk-linear category and 𝒞(E)\mathscr{C}_{(E)} be the category of EE-modules in 𝒞\mathscr{C}. For any object M𝒞(E)M\in\mathscr{C}_{(E)}, we write M(k)M_{(k)} for the underlying object in 𝒞\mathscr{C} when we forget the EE-linear structure. In [Ferrand], Ferrand gave a general construction of a norm functor NmE/k:𝒞(E)𝒞\mathrm{Nm}_{E/k}\colon\mathscr{C}_{(E)}\to\mathscr{C}, which was summarized in [Moonen, §3.6]. Note that we restricted to the case when EE is a field, as opposed to a general étale kk-algebra in loc. cit. This is good enough for our purposes.

We first consider the case when 𝒞\mathscr{C} is the category of kk-modules 𝖬𝗈𝖽k\mathsf{Mod}_{k}. For any M𝖬𝗈𝖽EM\in\mathsf{Mod}_{E}, there is a functorial polynomial map νM:MNmE/k(M)\nu_{M}\colon M\to\mathrm{Nm}_{E/k}(M) such that νM(em)=NormE/k(e)νM(m)\nu_{M}(em)=\mathrm{Norm}_{E/k}(e)\nu_{M}(m) for any eEe\in E and mMm\in M. The norm functor NmE/k\mathrm{Nm}_{E/k} is a {\otimes}-functor and is non-additive (unless E=kE=k). However, for any M1,M2𝖬𝗈𝖽EM_{1},M_{2}\in\mathsf{Mod}_{E}, there is an identification

NmE/k(HomE(M1,M2))=Homk(NmE/k(M1),NmE/k(M2)).\mathrm{Nm}_{E/k}(\mathrm{Hom}_{E}(M_{1},M_{2}))=\mathrm{Hom}_{k}(\mathrm{Nm}_{E/k}(M_{1}),\mathrm{Nm}_{E/k}(M_{2})).

The above identification gives us a structural map

η:ResE/kGL(𝒱)GL(NmE/k(𝒱))\eta:\mathrm{Res}_{E/k}\mathrm{GL}(\mathcal{V})\to\mathrm{GL}(\mathrm{Nm}_{E/k}(\mathcal{V}))

for any 𝒱𝖬𝗈𝖽E\mathcal{V}\in\mathsf{Mod}_{E}, which sends an EE-linear automorphism ff to NmE/k(f)\mathrm{Nm}_{E/k}(f). Let TET_{E} denote the torus ResE/k𝔾m,E\mathrm{Res}_{E/k}\mathbb{G}_{m,E} and TE1T_{E}^{1} denote the kernel of the norm map TETkT_{E}\to T_{k}. Viewing 𝔾m,E\mathbb{G}_{m,E} as the diagonal torus of GL(𝒱)\mathrm{GL}(\mathcal{V}), so that TE1T_{E}^{1} is a subgroup of ResE/kGL(𝒱)\mathrm{Res}_{E/k}\mathrm{GL}(\mathcal{V}), we have ker(η)=T1E\ker(\eta)=T^{1}_{E}.

Notation (2.1.5).

Take k=k=\mathbb{Q} and EE to be a totally real field. Let 𝒱\mathcal{V} be an EE-vector space equipped with a quadratic form ϕ~:𝒱E\widetilde{\phi}:\mathcal{V}\to E. We often drop ϕ~\widetilde{\phi} from the notation when it is assumed.

  1. (a)

    Write 𝒱()\mathcal{V}_{(\mathbb{Q})} for the underlying \mathbb{Q}-vector space of 𝒱\mathcal{V}, equipped with a quadratic form ϕ:𝒱()\phi:\mathcal{V}_{(\mathbb{Q})}\to\mathbb{Q} given by trE/ϕ~\mathrm{tr}_{E/\mathbb{Q}}\circ\widetilde{\phi}.

  2. (b)

    Write 𝒢(𝒱)\mathcal{G}(\mathcal{V}) for ResE/SO(𝒱)\mathrm{Res}_{E/\mathbb{Q}}\mathrm{SO}(\mathcal{V}), Z1E(𝒱)Z^{1}_{E}(\mathcal{V}) for T1E𝒢(𝒱)T^{1}_{E}\cap\mathcal{G}(\mathcal{V}), and (𝒱)\mathcal{H}(\mathcal{V}) for 𝒢(𝒱)/Z1E(𝒱)\mathcal{G}(\mathcal{V})/Z^{1}_{E}(\mathcal{V}), or simply 𝒢,Z1E\mathcal{G},Z^{1}_{E} and \mathcal{H} when 𝒱\mathcal{V} is understood.

  3. (c)

    Denote by Cl+(𝒱)\mathrm{Cl}^{+}(\mathcal{V}) the even Clifford algebra of 𝒱\mathcal{V} over EE and by Cl+E/(𝒱)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathcal{V}) its norm NmE/Cl+(𝒱)\mathrm{Nm}_{E/\mathbb{Q}}\mathrm{Cl}^{+}(\mathcal{V}).

  4. (d)

    Let \mathcal{E} denote the 11-dimensional quadratic form over EE given by equiping EE with the form ϕ~(α)=α2\widetilde{\phi}(\alpha)=\alpha^{2}.

Recall our convention (1.0.5)(c). The association 𝒱𝒱()\mathcal{V}\mapsto\mathcal{V}_{(\mathbb{Q})} in (a) above defines an equivalence of categories between quadratic forms over EE and quadratic forms over \mathbb{Q} with an self-adjoint EE-action ([Knus, Ch 1, Thm 7.4.1]). We call 𝒱()\mathcal{V}_{(\mathbb{Q})} the transfer of 𝒱\mathcal{V}, and 𝒱\mathcal{V} the EE-bilinear lift of 𝒱()\mathcal{V}_{(\mathbb{Q})}.

Lemma (2.1.6).

Let EE be a totally real field, and let 𝔴𝖬𝗈𝗍AH()(E)\mathfrak{w}\in\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C})_{(E)} be a motive equipped with an EE-action and a symmetric EE-bilinear form ϕ~:𝔴E𝔴𝟏E\widetilde{\phi}:\mathfrak{w}{\otimes}_{E}\mathfrak{w}\to\mathbf{1}_{E}. If dimEωB(𝔴)\dim_{E}\omega_{B}(\mathfrak{w}) is odd, then 𝔑(𝔴):=NmE/(𝔴)det(𝔴())\mathfrak{N}(\mathfrak{w}):=\mathrm{Nm}_{E/\mathbb{Q}}(\mathfrak{w}){\otimes}_{\mathbb{Q}}\det(\mathfrak{w}_{(\mathbb{Q})}) is (noncanonically) isomorphic to a submotive of Cl+E/(𝔴,ϕ~)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{w},\widetilde{\phi}).

The meaning of Cl+E/(𝔴,ϕ~)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{w},\widetilde{\phi}) is explained in the proof.

Proof.

This follows from the content in [Moonen, §5.4]. We give a sketch so that the reader can easily check the details from loc. cit. Set W:=ωB(𝔪)W:=\omega_{B}(\mathfrak{m}) and consider 𝔪\mathfrak{m} as a representation of its motivic Galois group Gmot(𝔴)G_{\mathrm{mot}}(\mathfrak{w}), which respects ϕ~\widetilde{\phi} and the EE-action. The motive Cl+E/(𝔴,ϕ~)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{w},\widetilde{\phi}) then corresponds to the Gmot(𝔴)G_{\mathrm{mot}}(\mathfrak{w})-representation Cl+E/(W,ϕ~)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(W,\widetilde{\phi}).

Let Σ\Sigma be the set of embeddings σ:E\sigma:E\hookrightarrow\mathbb{C}, 𝔖\mathfrak{S} be the symmetric group of Σ\Sigma, and QQ be the set of 𝔖\mathfrak{S}-orbits of Σ\mathbb{N}^{\Sigma}. Then under the assumption that dimEW\dim_{E}W is odd, there is an ascending filtration Fil\mathrm{Fil}_{\bullet} on Cl+E/(𝔴,ϕ~)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{w},\widetilde{\phi}) indexed by QQ such that for some q1,q2Qq_{1},q_{2}\in Q, Filq1/Filq2𝔑(𝔴)\mathrm{Fil}_{q_{1}}/\mathrm{Fil}_{q_{2}}\,{\cong}\,\mathfrak{N}(\mathfrak{w}). Therefore, 𝔑(𝔴)\mathfrak{N}(\mathfrak{w}) is a subquotient of Cl+E/(𝔴,ϕ~)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{w},\widetilde{\phi}). As 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}) is semisimple, 𝔑(𝔴)\mathfrak{N}(\mathfrak{w}) is in fact (non-canonically) a subobject. ∎

2.2   Motives of Varieties with h2,0=1h^{2,0}=1

Definition (2.2.1).

A polarized Hodge structure VV of weight 0 is said to be of K3-type if dimV(1,1)=dimV(1,1)=1\dim V_{\mathbb{C}}^{(-1,1)}=\dim V_{\mathbb{C}}^{(1,-1)}=1, and Vi,j=0V_{\mathbb{C}}^{i,j}=0 when |ij|>2|i-j|>2. The transcendental part T(V)T(V) of VV is the orthogonal complement of V(0,0):=VV(0,0)V^{(0,0)}:=V\cap V_{\mathbb{C}}^{(0,0)}.

We recall the following fundamental result of Zarhin:

Theorem (2.2.2) ([Zarhin, §2]).

Let VV be a Hodge structure of K3-type such that V(0,0)=0V^{(0,0)}=0 and let ϕ\phi be its polarization form. Then the endomorphism algebra EEndHdgVE\coloneqq\mathrm{End}_{\mathrm{Hdg}}V is either a totally real field or a CM field, and the adjoint map ee¯e\mapsto\bar{e} defined by ϕ(ex,y)=ϕ(x,e¯y)\phi(ex,y)=\phi(x,\bar{e}y) is the identity map when EE is totally real and is complex conjugation when EE is CM.

To discuss motives in families, we first give a definition:

Definition (2.2.3).

Let SS be a connected smooth \mathbb{C}-variety. For every \mathbb{Q}-local system 𝖵B\mathsf{V}_{B} over SS and bS()b\in S(\mathbb{C}), we write Mon(𝖵B,b)\mathrm{Mon}(\mathsf{V}_{B},b) for the Zariski closure of the image of π1(S,b)\pi_{1}(S,b) in GL(𝖵B,b)\mathrm{GL}(\mathsf{V}_{B,b}), and Mon(𝖵B,b)\mathrm{Mon}^{\circ}(\mathsf{V}_{B},b) for its identity component. When 𝖵=(𝖵B,𝖵dR)\mathsf{V}=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}}) is a polarizable VHS444This means that 𝖵B\mathsf{V}_{B} is the \mathbb{Q}-local system and 𝖵dR\mathsf{V}_{\mathrm{dR}} is the filtered flat vector bundle in the VHS. Similar conventions apply throughout the paper. over SS, we say that 𝖵\mathsf{V} has maximal monodromy if Mon(𝖵B,b)\mathrm{Mon}^{\circ}(\mathsf{V}_{B},b) is equal to the derived group of the Mumford-Tate group MT(𝖵b)\mathrm{MT}(\mathsf{V}_{b}), where bb is any Hodge-generic point.

For the terminology “Hodge-generic points”, see for example [Moonen-Fom, 31]. Note that [AndreMT, §5] says that in the above notation, Mon(𝖵B,b)\mathrm{Mon}^{\circ}(\mathsf{V}_{B},b) is always a normal subgroup of the derived group of MT(𝖵b)\mathrm{MT}(\mathsf{V}_{b}) (see also [Peters, Thm 16]).

(2.2.4)   

For the rest of (2.2), let SS be a connected smooth \mathbb{C}-variety and f:𝒳Sf:\mathcal{X}\to S be a \heartsuit-family of relative dimension dd. Let 𝖵=(𝖵B,𝖵dR)\mathsf{V}=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}}) be the VHS on SS defined by R2f(1)R^{2}f_{*}\mathbb{Q}(1). Note that by the definition of a \heartsuit-family, 𝖵\mathsf{V} satisfies condition (P) in [Moonen, Prop. 6.4]. Let 𝝃\boldsymbol{\xi} be a relatively ample line bundle on 𝒳/S\mathcal{X}/S, which defines a symmetric bilinear pairing ,\langle-,-\rangle on 𝖵\mathsf{V} such that α,β=αβc1(𝝃)d2\langle\alpha,\beta\rangle=\alpha\cup\beta\cup c_{1}(\boldsymbol{\xi})^{d-2} for local sections α,β\alpha,\beta. Choose a Hodge-generic base point bS()b\in S(\mathbb{C}).

Suppose for a moment that Mon(𝖵B,b)\mathrm{Mon}(\mathsf{V}_{B},b) is connected. Then for ρ:=dimNS(𝒳b)\rho:=\dim\mathrm{NS}(\mathcal{X}_{b})_{\mathbb{Q}}, 𝖵\mathsf{V} admits an orthogonal decomposition 𝟏Sρ𝖯\mathbf{1}_{S}^{\oplus\rho}\oplus\mathsf{P}, where 𝖯\mathsf{P} is a VHS polarized by the pairing induced by 𝝃\boldsymbol{\xi} and 𝟏S\mathbf{1}_{S} is the unit VHS. Note that the Hodge structure 𝖯b=(𝖯B,b,𝖯dR,b)\mathsf{P}_{b}=(\mathsf{P}_{B,b},\mathsf{P}_{\mathrm{dR},b}), together with its polarization, is of K3-type and satisfies the hypothesis of (2.2.2). Let EE be the endomorphism field of 𝖯b\mathsf{P}_{b}. As Mon(𝖵B,b)=Mon(𝖵B,b)MT(𝖵b)=MT(𝖯b)\mathrm{Mon}(\mathsf{V}_{B},b)=\mathrm{Mon}^{\circ}(\mathsf{V}_{B},b)\subseteq\mathrm{MT}(\mathsf{V}_{b})=\mathrm{MT}(\mathsf{P}_{b}) and MT(𝖯b)\mathrm{MT}(\mathsf{P}_{b}) commutes with EE, the EE-action on 𝖯B,b\mathsf{P}_{B,b} commutes with π1(S,b)\pi_{1}(S,b) and hence extends to an action on 𝖯\mathsf{P} (see e.g., [Peters, Cor. 12]). If 𝖵\mathsf{V} (or equivalently 𝖯\mathsf{P}) has maximal monodromy, we say that (𝒳/S,𝝃)(\mathcal{X}/S,\boldsymbol{\xi}) is in case (R+) if EE is totally real and (CM) if EE is CM; we further divide (R+) into case (R1) for dimE𝖯B,b\dim_{E}\mathsf{P}_{B,b} odd and case (R2) for dimE𝖯B,b\dim_{E}\mathsf{P}_{B,b} even. We say we are in case (R2’) if 𝖵\mathsf{V} has non-maximal monodromy, which can only happen when EE is totally real and dimE𝖯B,b=4\dim_{E}\mathsf{P}_{B,b}=4. See [Moonen, Prop. 6.4(iii)] and its proof.

If Mon(𝖵B,b)\mathrm{Mon}(\mathsf{V}_{B},b) is not connected, we say that (𝒳/S,𝝃)(\mathcal{X}/S,\boldsymbol{\xi}) is in case ?? for ?? = (R1), (R2), (CM) or (R2’) if it is in case ?? up to replacing SS by a connected étale cover SS^{\prime} and bb by a lift such that Mon(𝖵B,b)\mathrm{Mon}(\mathsf{V}_{B},b) becomes connected. The definition is clearly independent of these choices.

Proposition (2.2.5).

Suppose that 𝖵\mathsf{V} has non-maximal monodromy, or equivalently (𝒳/S,𝛏)(\mathcal{X}/S,\boldsymbol{\xi}) belongs to case (R2’). Then for a general sSs\in S and X:=𝒳sX:=\mathcal{X}_{s}, the Kodaira-Spencer map s:TsSHom(H1(Ω1X),H2(𝒪X))\nabla_{s}:T_{s}S\to\mathrm{Hom}(\mathrm{H}^{1}(\Omega^{1}_{X}),\mathrm{H}^{2}(\mathcal{O}_{X})) has rank 11.

Proof.

We may assume that Mon(𝖵B,b)\mathrm{Mon}(\mathsf{V}_{B},b) is conncted, so that there is a decomposition 𝟏Sρ𝖯\mathbf{1}_{S}^{\oplus\rho}\oplus\mathsf{P} as above. The rank of s\nabla_{s} achieves its maximum on an open dense USU\subseteq S. Choose some sUs\in U. By [Voisin, Thm 3.5] (cf. [Moonen, Prop. 6.4]), for some point zz in a small analytic neighborhood of ss, 𝖯B,z(0,0)\mathsf{P}_{B,z}^{(0,0)} contains a nonzero class ζ\zeta. Let ZSZ^{\prime}\subseteq S be the irreducible component of the Noether-Lefschetz loci defined by (z,ζ)(z,\zeta) and ZZ be its smooth locus. Up to moving zz along ZZ^{\prime} a little bit, we may assume that zz lies in ZZ, is Hodge-generic for the VHS 𝖯|Z\mathsf{P}|_{Z}, and ranks=rankz\mathrm{rank\,}\nabla_{s}=\mathrm{rank\,}\nabla_{z}. Let WW be the Hodge structure of K3-type defined by the fiber of 𝖯\mathsf{P} at zz and let TT be its transcendental part. Then dimET<dimEW=4\dim_{E}T<\dim_{E}W=4. Set F:=EndHdgTF:=\mathrm{End}_{\mathrm{Hdg}}T, which contains EE.

By assumption on a \heartsuit-family, ranks>0\mathrm{rank\,}\nabla_{s}>0. Suppose by way of contradiction that ranks>1\mathrm{rank\,}\nabla_{s}>1. Then as ZSZ\subseteq S has codimension 11 (cf. [Voisin, Lem. 3.1]), z\nabla_{z} does not vanish on TzZT_{z}Z. This implies that Mon(𝖯B|Z,z)1\mathrm{Mon}^{\circ}(\mathsf{P}_{B}|_{Z},z)\neq 1 and dimET=3\dim_{E}T=3 by [Moonen, Prop. 6.4] and its proof. Indeed, if FF is CM, then dimFT2\dim_{F}T\geq 2, so that dimET4\dim_{E}T\geq 4, which is impossible. Therefore, FF is totally real and dimFT3\dim_{F}T\geq 3. This forces E=FE=F and dimET=3\dim_{E}T=3. Now let 𝒯\mathcal{T} be the EE-bilinear lift of TT. Then by [Zarhin, Thm 2.2.1] MT(T)=ResE/SO(𝒯)\mathrm{MT}(T)=\mathrm{Res}_{E/\mathbb{Q}}\mathrm{SO}(\mathcal{T}). As this group is simple, Mon(𝖯B|Z,z)ResE/SO(𝒯)\mathrm{Mon}^{\circ}(\mathsf{P}_{B}|_{Z},z)\,{\cong}\,\mathrm{Res}_{E/\mathbb{Q}}\mathrm{SO}(\mathcal{T}). On the other hand, [Moonen, §8.1] tells us that Mon(𝖯B,s)ResE/L\mathrm{Mon}^{\circ}(\mathsf{P}_{B},s)\,{\cong}\,\mathrm{Res}_{E/\mathbb{Q}}L for some EE-form LL of SL2\mathrm{SL}_{2}. However, as SL2\mathrm{SL}_{2} and SO(3)\mathrm{SO}(3) are not even isomorphic over \mathbb{C}, ResE/L\mathrm{Res}_{E/\mathbb{Q}}L cannot have a subgroup isomorphic to ResE/SO(𝒯)\mathrm{Res}_{E/\mathbb{Q}}\mathrm{SO}(\mathcal{T}), which constradicts the fact that parallel transport (noncanonically) sends Mon(𝖯B|Z,z)\mathrm{Mon}^{\circ}(\mathsf{P}_{B}|_{Z},z) into Mon(𝖯B,s)\mathrm{Mon}^{\circ}(\mathsf{P}_{B},s). ∎

(2.2.6)   

The following statements are the key inputs we will use from Moonen’s paper [Moonen]. A little adaptation we make is that we will uniformly use the category of motives 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}) with absolute Hodge cycles, whereas Moonen used André’s category with motivated cycles ([AndreMot]). This adaptation makes a difference only for (2.2.7)(d) below. We use 𝖬𝗈𝗍()\mathsf{Mot}(\mathbb{C}) to denote André’s category (for base field \mathbb{C}) when explaining this difference, but otherwise all motives are considered in 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}). Note that motivated cycles are automatically absolute Hodge, so that 𝖬𝗈𝗍()\mathsf{Mot}(\mathbb{C}) is a subcategory of 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}).

Recall our notations in (2.1.4) and (2.2.4) for the theorem below:

Theorem (2.2.7).

Assume that Mon(𝖵B,b)\mathrm{Mon}(\mathsf{V}_{B},b) is connected and let 𝖯\mathsf{P} and EE be as in (2.2.4). Let sS()s\in S(\mathbb{C}) be any point and write 𝔭s𝖬𝗈𝗍AH()\mathfrak{p}_{s}\in\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}) for the submotive of 𝔥2(𝒳s)(1)\mathfrak{h}^{2}(\mathcal{X}_{s})(1) such that ωHdg(𝔭s)=𝖯s\omega_{\mathrm{Hdg}}(\mathfrak{p}_{s})=\mathsf{P}_{s}. Then the action of EE on 𝖯s\mathsf{P}_{s} is absolute Hodge, i.e., induced by an action of EE on 𝔭s\mathfrak{p}_{s}. Moreover:

  1. (a)

    In case (R1), NmE/(𝔭s)det(𝔭s,())\mathrm{Nm}_{E/\mathbb{Q}}(\mathfrak{p}_{s}){\otimes}_{\mathbb{Q}}\det(\mathfrak{p}_{s,(\mathbb{Q})}) is an object of 𝖬𝗈𝗍𝖠𝖻()\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}).

  2. (b)

    In case (R2), for 𝟏E:=𝟏E\mathbf{1}_{E}:=\mathbf{1}{\otimes}E and 𝔭s:=𝔭s𝟏E\mathfrak{p}_{s}^{\sharp}:=\mathfrak{p}_{s}\oplus\mathbf{1}_{E}, NmE/(𝔭s)det(𝔭s,())\mathrm{Nm}_{E/\mathbb{Q}}(\mathfrak{p}^{\sharp}_{s}){\otimes}_{\mathbb{Q}}\det(\mathfrak{p}^{\sharp}_{s,(\mathbb{Q})}) is an object of 𝖬𝗈𝗍𝖠𝖻()\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}).

  3. (c)

    In case (R2’), NmE/(𝔭s)\mathrm{Nm}_{E/\mathbb{Q}}(\mathfrak{p}_{s}) is an object of 𝖬𝗈𝗍𝖠𝖻()\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}).

  4. (d)

    In case (CM), 𝔭s\mathfrak{p}_{s} is an object of 𝖬𝗈𝗍𝖠𝖻()\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}).

Proof.

The statement that the action of EE on 𝖯s\mathsf{P}_{s} is absolute Hodge is implied by [Moonen, Prop. 6.6]. Note that our 𝔭s\mathfrak{p}_{s} is Moonen’s 𝑽s\boldsymbol{V}_{s}, and our 𝖯\mathsf{P} is Moonen’s 𝕍\mathbb{V}. Below we view 𝔭s\mathfrak{p}_{s} as an object of 𝖬𝗈𝗍AH()(E)\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C})_{(E)}. In this proof, End¯\underline{\mathrm{End}} (or Hom¯\underline{\mathrm{Hom}}) always mean internal End\mathrm{End} (or Hom\mathrm{Hom}) in the category 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}).

(a) We first treat the case (R+). Let 𝒱\mathcal{V} be the EE-bilinear lift of the quadratic form given by 𝖯B,b\mathsf{P}_{B,b} with its self-dual EE-action. By [Moonen, §6.9, 6.10] there is a family of abelian schemes 𝒜S\mathscr{A}\to S with multiplication by D:=Cl+E/(𝒱)D:=\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathcal{V}) (see (2.1.5)) such that there is an isomorphism

Cl+E/(𝔭s)End¯D(𝔥1(𝒜s))\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}_{s})\stackrel{{\scriptstyle\sim}}{{\to}}\underline{\mathrm{End}}_{D}(\mathfrak{h}^{1}(\mathscr{A}_{s})) (2)

of algebra objects in 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}). In case (R1), dimE𝒱\dim_{E}\mathcal{V} is odd, and by (2.1.6) NmE/(𝔭s)det(𝔭s,())\mathrm{Nm}_{E/\mathbb{Q}}(\mathfrak{p}_{s}){\otimes}\det(\mathfrak{p}_{s,(\mathbb{Q})}) is non-canonically a submotive of Cl+E/(𝔭s)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}_{s}), and hence is an object of 𝖬𝗈𝗍𝖠𝖻()\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}).

(b) In case (R2), (2) still holds, so that Cl+E/(𝔭s)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}_{s}) is still an object of 𝖬𝗈𝗍𝖠𝖻()\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}), but a further trick is needed to recover (a variant of) NmE/(𝔭s)\mathrm{Nm}_{E/\mathbb{Q}}(\mathfrak{p}_{s}) from Cl+E/(𝔭s)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}_{s}). Recall \mathcal{E} defined in (2.1.5)(d). As in [Moonen, §6.11, 6.12], we consider the VHS 𝖯:=𝖯(𝟏S())\mathsf{P}^{\sharp}:=\mathsf{P}\oplus(\mathbf{1}_{S}{\otimes}\mathcal{E}_{(\mathbb{Q})}), where 𝟏S\mathbf{1}_{S} standards for the unit VHS on SS. Let 𝒱:=𝒱\mathcal{V}^{\sharp}:=\mathcal{V}\oplus\mathcal{E} and D:=Cl+E/(𝒱)D^{\sharp}:=\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathcal{V}^{\sharp}). Then by [Moonen, §6.9, 6.10] again, there is an abelian scheme 𝒜S\mathscr{A}^{\sharp}\to S with multiplication by DD^{\sharp} such that

Cl+E/(𝖯)End¯D(𝖧)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathsf{P}^{\sharp})\,{\cong}\,\underline{\mathrm{End}}_{D^{\sharp}}(\mathsf{H}) (3)

where 𝖧\mathsf{H} is the VHS given by the first relative cohomology of 𝒜\mathscr{A}^{\sharp}. It is shown in loc. cit. that the fiber of the isomorphism (3) at every \mathbb{C}-point on SS is induced by an absolute Hodge cycle. Here is a summary of the argument in our notations: Choose a point s0S()s_{0}\in S(\mathbb{C}) such that 𝖯s0(0,0)0\mathsf{P}_{s_{0}}^{(0,0)}\neq 0. By the last paragraph of [Moonen, §6.12], there is an isomorphism

Cl+E/(𝔭s0)Cl+E/(𝔭s0)2[E:]\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}^{\sharp}_{s_{0}})\,{\cong}\,\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}_{s_{0}})^{\oplus 2^{[E:\mathbb{Q}]}}

of objects in 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}). As Cl+E/(𝔭s0)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}_{s_{0}}) is an object of 𝖬𝗈𝗍𝖠𝖻()\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}), so is Cl+E/(𝔭s0)\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}_{s_{0}}^{\sharp}). Therefore, by (2.1.2) the isomorphism (3) is absolute Hodge at s0s_{0}, and hence so at every other ss by Deligne’s Principle B. This implies Cl+E/(𝔭s)𝖬𝗈𝗍𝖠𝖻()\mathrm{Cl}^{+}_{E/\mathbb{Q}}(\mathfrak{p}^{\sharp}_{s})\in\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}), so that (b) follows from (2.1.6) again.

(c) This follows directly from [Moonen, Prop. 8.5].

(d) In case (CM), [Moonen, §7.4] tells us that there exists a motive 𝔲\mathfrak{u} (denoted by 𝑼\boldsymbol{U} therein), an abelian variety AA over \mathbb{C} and an abelian scheme S\mathscr{B}\to S, all equipped with multiplication by EE, such that there is an isomorphism

𝔭sE𝔲𝔥s:=Hom¯E(𝔥1(A),𝔥1(s)).\mathfrak{p}_{s}{\otimes}_{E}\mathfrak{u}\stackrel{{\scriptstyle\sim}}{{\to}}\mathfrak{h}_{s}:=\underline{\mathrm{Hom}}_{E}(\mathfrak{h}^{1}(A),\mathfrak{h}^{1}(\mathscr{B}_{s})).

This implies that 𝔭s𝔲𝖬𝗈𝗍𝖠𝖻()\mathfrak{p}_{s}{\otimes}\mathfrak{u}\in\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}). In order to show 𝔭s𝖬𝗈𝗍𝖠𝖻()\mathfrak{p}_{s}\in\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}), it suffices to argue that 𝔲 1E\mathfrak{u}\,{\cong}\,\mathbf{1}_{E} in 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}).

In loc. cit., 𝔲\mathfrak{u} is in fact constructed as an object 𝖬𝗈𝗍()\mathsf{Mot}(\mathbb{C}). Moonen remarked that conjecturally there should be an isomorphism 𝔲 1E\mathfrak{u}\,{\cong}\,\mathbf{1}_{E} in 𝖬𝗈𝗍()\mathsf{Mot}(\mathbb{C}) and proved that 𝔲\mathfrak{u} indeed has trivial Hodge and \ell-adic realizations. We note that his argument in [Moonen, Lem. 7.5] in fact implies that 𝔲 1E\mathfrak{u}\,{\cong}\,\mathbf{1}_{E} in 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}), i.e., ωB(𝔲)\omega_{B}(\mathfrak{u}) is spanned by absolute Hodge classes: The idea is to take advantage of the fact that 𝔲\mathfrak{u} is independent of ss, and that for some s0Ss_{0}\in S, the algebraic part ωB(𝔭s0)(0,0)\omega_{B}(\mathfrak{p}_{s_{0}})^{(0,0)} of ωB(𝔭s0)\omega_{B}(\mathfrak{p}_{s_{0}}) is nonempty. Since every class in ωB(𝔲)\omega_{B}(\mathfrak{u}) is of type (0,0)(0,0), we have ωB(𝔭s0)(0,0)EωB(𝔲)=ωB(𝔥s0)(0,0)\omega_{B}(\mathfrak{p}_{s_{0}})^{(0,0)}{\otimes}_{E}\omega_{B}(\mathfrak{u})=\omega_{B}(\mathfrak{h}_{s_{0}})^{(0,0)}. By Lefschetz (1,1)(1,1)-theorem, every class in ωB(𝔭s0)(0,0)\omega_{B}(\mathfrak{p}_{s_{0}})^{(0,0)} comes from a line bundle on 𝒳s\mathcal{X}_{s} and hence is absolute Hodge. As classes in ωB(𝔥s0)(0,0)\omega_{B}(\mathfrak{h}_{s_{0}})^{(0,0)} are also absolute Hodge, we may now conclude by (2.2.8) below. ∎

Proposition (2.2.8).

Let EE be a number field and let 𝔪,𝔫\mathfrak{m},\mathfrak{n} be objects of 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}) with EE-action. Let 𝔥𝔪E𝔫\mathfrak{h}\coloneqq\mathfrak{m}{\otimes}_{E}\mathfrak{n}. Let mωB(𝔪),nωB(𝔫)m\in\omega_{B}(\mathfrak{m}),n\in\omega_{B}(\mathfrak{n}) be nonzero Hodge cycles and define hωB(𝔥)h\in\omega_{B}(\mathfrak{h}) to be mEnm{\otimes}_{E}n. If hh and mm are both absolute Hodge, then so is nn.

Proof.

Recall notations in (2.1.3). We need to show that for every σAut()\sigma\in\mathrm{Aut}(\mathbb{C}), the image nn^{\prime} of n1n{\otimes}1 under the canonical isomorphisms

ωB(𝔫)(𝔸f)ωdR(𝔫)×ω𝔸f(𝔫)bcωdR(𝔫σ)×ω𝔸f(𝔫σ)ωB(𝔫σ)(𝔸f).\omega_{B}(\mathfrak{n})\otimes(\mathbb{C}\otimes\mathbb{A}_{f})\cong\omega_{\mathrm{dR}}(\mathfrak{n})\times\omega_{\mathbb{A}_{f}}(\mathfrak{n})\,{\cong}\,_{\mathrm{bc}}\omega_{\mathrm{dR}}(\mathfrak{n}^{\sigma})\times\omega_{\mathbb{A}_{f}}(\mathfrak{n}^{\sigma})\cong\omega_{B}(\mathfrak{n}^{\sigma})\otimes(\mathbb{C}\otimes\mathbb{A}_{f}).

is contained in ωB(𝔫σ)\omega_{B}(\mathfrak{n}^{\sigma}) (i.e., is of the form nσ1n^{\sigma}{\otimes}1 for some nσωB(𝔫σ)n^{\sigma}\in\omega_{B}(\mathfrak{n}^{\sigma})). Since mm and hh are absolute Hodge, we know that mσωB(𝔪σ)m^{\sigma}\in\omega_{B}(\mathfrak{m}^{\sigma}) and hσ=mσnωB(𝔥σ)h^{\sigma}=m^{\sigma}\otimes n^{\prime}\in\omega_{B}(\mathfrak{h}^{\sigma}). Then apply (2.2.9) with k=k=\mathbb{Q}, R=𝔸f×R=\mathbb{A}_{f}\times\mathbb{C}, M=ωB(𝔪σ)M=\omega_{B}(\mathfrak{m}^{\sigma}) and N=ωB(𝔫σ)N=\omega_{B}(\mathfrak{n}^{\sigma}) to deduce that nωB(𝔫σ)n^{\prime}\in\omega_{B}(\mathfrak{n}^{\sigma}). ∎

Lemma (2.2.9).

Let E/kE/k be a field extension, RR be any kk-algebra and M,NM,N be finite dimensional EE-vector spaces. Let HMENH\coloneqq M{\otimes}_{E}N and hHkRh\in H{\otimes}_{k}R be a nonzero element of the form mnm{\otimes}n under the canonical isomorphism

HkR(MkR)EkR(NkR),H{\otimes}_{k}R\,{\cong}\,(M{\otimes}_{k}R){\otimes}_{E{\otimes}_{k}R}(N{\otimes}_{k}R),

where mMkRm\in M{\otimes}_{k}R and nNkRn\in N{\otimes}_{k}R. If hHh\in H and mMm\in M, then nNn\in N.

Proof.

Let α=dimEM\alpha=\dim_{E}M and β=dimEN\beta=\dim_{E}N. By choosing bases of MM and NN, we may assume M=EαM=E^{\alpha} and N=EβN=E^{\beta}. Identify HH with the space of (α×β)(\alpha\times\beta)-matrices over EE and hh with mnTm\cdot n^{T}. We denote by (mi),(nj),(hi,j)(m_{i}),(n_{j}),(h_{i,j}) the respective EkRE{\otimes}_{k}R-coordinates of m,nm,n and hh and fix ii such that mi=0m_{i}\not=0. Then for every jj, hi,j=minjh_{i,j}=m_{i}\cdot n_{j} and hence nj=mi1hi,jEn_{j}=m_{i}^{-1}h_{i,j}\in E because mi,hi,jEEkRm_{i},h_{i,j}\in E\subseteq E{\otimes}_{k}R by assumption. ∎

3 Moduli Interpretations of Shimura Varieties

In this section, we first recall the moduli description for the canonical models of some Shimura varieties of abelian type from [YangSystem]. Then we give some prelimary results on those of orthogonal type over totally real fields and review their integral models when the reflex field is \mathbb{Q}.

3.1   Systems of realizations

Definition (3.1.1).

Let kk be a subfield of \mathbb{C} and SS be a smooth kk-variety. By a system of realizations we mean a tuple 𝖵=(𝖵B,𝖵dR,𝖵e´t,idR,ie´t)\mathsf{V}=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}},\mathsf{V}_{\mathrm{{\acute{e}}t}},i_{\mathrm{dR}},i_{\mathrm{{\acute{e}}t}}) where

  • 𝖵B\mathsf{V}_{B} is a \mathbb{Q}-local system over S:=SkS_{\mathbb{C}}:=S{\otimes}_{k}\mathbb{C};

  • 𝖵dR\mathsf{V}_{\mathrm{dR}} is a filtered flat vector bundle over SS;

  • 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}} is an étale local system of 𝔸f\mathbb{A}_{f}-coefficients over SS;

  • idR:(𝖵B𝒪anS,idd)(𝖵dR|S)ani_{\mathrm{dR}}:(\mathsf{V}_{B}{\otimes}\mathcal{O}^{\mathrm{an}}_{S_{\mathbb{C}}},\mathrm{id}{\otimes}d)\stackrel{{\scriptstyle\sim}}{{\to}}(\mathsf{V}_{\mathrm{dR}}|_{S_{\mathbb{C}}})^{\mathrm{an}} is an isomorphism of flat (holomorphic) vector bundles such that (𝖵B,𝖵dR|S)(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}}|_{S_{\mathbb{C}}}) is a polarizable VHS;

  • ie´t:𝖵B𝔸f𝖵e´t|Si_{\mathrm{{\acute{e}}t}}:\mathsf{V}_{B}{\otimes}\mathbb{A}_{f}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{V}_{\mathrm{{\acute{e}}t}}|_{S_{\mathbb{C}}} is an isomorphism between (the pro-étale sheaf associated to) 𝖵B𝔸f\mathsf{V}_{B}{\otimes}\mathbb{A}_{f} and 𝖵e´t|S\mathsf{V}_{\mathrm{{\acute{e}}t}}|_{S_{\mathbb{C}}}.

We may often omit (idR,ie´t)(i_{\mathrm{dR}},i_{\mathrm{{\acute{e}}t}}) in the notation and write 𝖵B𝔸f=𝖵e´t|S\mathsf{V}_{B}{\otimes}\mathbb{A}_{f}=\mathsf{V}_{\mathrm{{\acute{e}}t}}|_{S_{\mathbb{C}}} and 𝖵B𝒪San=𝖵dR|S\mathsf{V}_{B}{\otimes}\mathcal{O}_{S_{\mathbb{C}}}^{\mathrm{an}}=\mathsf{V}_{\mathrm{dR}}|_{S_{\mathbb{C}}} a little abusively. Let 𝖱(S)\mathsf{R}(S) denote the category of systems of realizations over SS, with morphisms defined in the obvious way. Then 𝖱(S)\mathsf{R}(S) is a naturally a Tannakian category. Let 𝟏S\mathbf{1}_{S} be its unit object and for each nn\in\mathbb{Z} let 𝟏S(n)\mathbf{1}_{S}(n) be the Tate object. For every 𝖵𝖱(S)\mathsf{V}\in\mathsf{R}(S), set 𝖵(n):=𝖵𝟏S(n)\mathsf{V}(n):=\mathsf{V}{\otimes}\mathbf{1}_{S}(n). Note that if k=k=\mathbb{C}, then 𝖱(S)\mathsf{R}(S) is naturally identified with the category of polarizable VHS over SS. We write H0(𝖵)\mathrm{H}^{0}(\mathsf{V}) for Hom(𝟏,𝖵)\mathrm{Hom}(\mathbf{1},\mathsf{V}).

Our definition is a little different from [FuMoonen, §6.1] because we fixed an embedding kk\hookrightarrow\mathbb{C}, but see [YangSystem, (3.1.3)] for a comparison. Also, recall our convention in (1.0.5)(e) for VHS. We remind the reader that since SS is defined over kk\subseteq\mathbb{C}, when we write sS()s\in S(\mathbb{C}), we mean a kk-linear morphism Spec()S\mathrm{Spec\,}(\mathbb{C})\to S, which we view simultaneously as a closed point on S=SkS_{\mathbb{C}}=S{\otimes}_{k}\mathbb{C}.

Next, we recall how to define a 𝖪\mathsf{K}-level structure in this context. Interestingly this definition can be leveraged to define additional structures on 𝖵\mathsf{V}, as (3.1.3) below shows.

Definition (3.1.2).

Let GG be a reductive group over \mathbb{Q}, let VRep(G)V\in\mathrm{Rep}(G) be a finite dimensional GG-representation and 𝖪G(𝔸f)\mathsf{K}\subseteq G(\mathbb{A}_{f}) be a compact open subgroup. For any subfield kk\subseteq\mathbb{C}, smooth kk-variety SS and any system of realizations 𝖵=(𝖵B,𝖵dR,𝖵e´t)\mathsf{V}=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}},\mathsf{V}_{\mathrm{{\acute{e}}t}}) on SS we make the following definitions.

  1. (a)

    A (G,V,𝖪)(G,V,\mathsf{K})-level structure on 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}}, or simply a 𝖪\mathsf{K}-level structure, is a global section [η][\eta] of 𝖪\Isom¯(V𝔸f,𝖵e´t)\mathsf{K}\backslash\underline{\mathrm{Isom}}(V{\otimes}\mathbb{A}_{f},\mathsf{V}_{\mathrm{{\acute{e}}t}}). Here Isom¯(V𝔸f,𝖵e´t)\underline{\mathrm{Isom}}(V{\otimes}\mathbb{A}_{f},\mathsf{V}_{\mathrm{{\acute{e}}t}}) denotes the pro-étale sheaf consisting of isomorphisms V𝔸f𝖵e´tV{\otimes}\mathbb{A}_{f}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{V}_{\mathrm{{\acute{e}}t}}, and 𝖪\mathsf{K} acts by pre-composition through its image in GL(V𝔸f)\mathrm{GL}(V{\otimes}\mathbb{A}_{f}). For each geometric point sSs\to S, write the 𝖪\mathsf{K}-orbit of isomorphisms V𝔸f𝖵e´t,sV{\otimes}\mathbb{A}_{f}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{V}_{\mathrm{{\acute{e}}t},s} determined by [η][\eta] as [η]s[\eta]_{s}.

  2. (b)

    If [η][\eta] is a (G,V,𝖪)(G,V,\mathsf{K})-level structure on 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}}, every element vI:=(V)Gv\in I:=(V^{\otimes})^{G} gives rise to a global section of 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}}, which we denote by η(v)e´t\eta(v)_{\mathrm{{\acute{e}}t}}, such that for every geometric point sSs\to S, every representative of [η]s[\eta]_{s} sends v1v{\otimes}1 to η(v)e´t,s\eta(v)_{\mathrm{{\acute{e}}t},s}. We say that [η][\eta] is 𝖵\mathsf{V}-rational if (i) for all vIv\in I, there exists a global section 𝐯=(𝐯B,𝐯dR,𝐯e´t)H0(𝖵)\mathbf{v}=(\mathbf{v}_{B},\mathbf{v}_{\mathrm{dR}},\mathbf{v}_{\mathrm{{\acute{e}}t}})\in\mathrm{H}^{0}(\mathsf{V}^{\otimes}) such that 𝐯e´t=η(v)e´t\mathbf{v}_{\mathrm{{\acute{e}}t}}=\eta(v)_{\mathrm{{\acute{e}}t}}, and (ii) for every sS()s\in S(\mathbb{C}), (𝖵B,s,{𝐯B,s}vI)(V,{v}vI)(\mathsf{V}_{B,s},\{\mathbf{v}_{B,s}\}_{v\in I})\,{\cong}\,(V,\{v\}_{v\in I}).

  3. (c)

    Let Ω\Omega be a G()G(\mathbb{R})-conjugacy class of morphisms 𝕊G\mathbb{S}\to G_{\mathbb{R}}. When VV is faithful, we say that a 𝖵\mathsf{V}-rational 𝖪\mathsf{K}-level structure is of type Ω\Omega if for every sS()s\in S(\mathbb{C}), under some (and hence every) isomorphism (𝖵B,s,{𝐯B,s}vI)(V,{v}vI)(\mathsf{V}_{B,s},\{\mathbf{v}_{B,s}\}_{v\in I})\,{\cong}\,(V,\{v\}_{v\in I}) the Hodge structure on 𝖵B,s\mathsf{V}_{B,s} is defined by an element of Ω\Omega.

  4. (d)

    If there is a morphism GGG^{\prime}\to G from another reductive subgroup GG^{\prime}, and 𝖪G(𝔸f)\mathsf{K}^{\prime}\subseteq G^{\prime}(\mathbb{A}_{f}) is a compact open subgroup whose image is contained in 𝖪\mathsf{K}, then we say that a 𝖪\mathsf{K}^{\prime}-level structure [η][\eta^{\prime}] on 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}} refines [η][\eta] if [η][\eta] is induced by [η][\eta^{\prime}] via the natural (forgetful) map 𝖪\Isom¯(V𝔸f,𝖵e´t)𝖪\Isom¯(V𝔸f,𝖵e´t)\mathsf{K}^{\prime}\backslash\underline{\mathrm{Isom}}(V{\otimes}\mathbb{A}_{f},\mathsf{V}_{\mathrm{{\acute{e}}t}})\to\mathsf{K}\backslash\underline{\mathrm{Isom}}(V{\otimes}\mathbb{A}_{f},\mathsf{V}_{\mathrm{{\acute{e}}t}}).

The definitions (b) and (c) above are used to simplify the formalism of the moduli interpretation of Shimura varieties. To explain this, we give a PEL-type example:

Example (3.1.3).

Let (B,,(V,ψ))(B,\ast,(V,\psi)) be a simple PEL-datum (i.e., BB is a simple \mathbb{Q}-algebra with positive involution \ast of type AA or CC and (V,ψ)(V,\psi) is a symplectic BB-module). We denote by GGSpB(V)G\coloneqq\mathrm{GSp}_{B}(V) the \mathbb{Q}-group of BB-linear similitudes and fix an open compact subgroup 𝖪G(𝔸f)\mathsf{K}\subset G(\mathbb{A}_{f}). Then there exists a unique G()G(\mathbb{R})-conjugacy class Ω\Omega such that (G,Ω)(G,\Omega) is a Shimura datum and Sh𝖪(G,Ω)()\mathrm{Sh}_{\mathsf{K}}(G,\Omega)(\mathbb{C}) is in canonical bijection with the set \mathcal{I} of isomorphism classes of tuples (𝖵,λ,i,[η])(\mathsf{V},\lambda,i,[\eta]) where

  • 𝖵\mathsf{V} is a \mathbb{Q}-Hodge structure (𝖵B,𝖵dR)(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}}) of type {(1,0),(0,1)}\{(1,0),(0,1)\},

  • i:BEnd(𝖵)i\colon B\hookrightarrow\mathrm{End}(\mathsf{V}) is an algebra morphism,

  • λ\lambda is a ×\mathbb{Q}^{\times}-equivalence class of BB-linear symplectic pairing of 𝖵\mathsf{V} and

  • [η][\eta] is a 𝖪\mathsf{K}-orbit of a BB-linear similitude V𝔸f𝖵e´tV\otimes\mathbb{A}_{f}\to\mathsf{V}_{\mathrm{{\acute{e}}t}}

such that ()(\spadesuit) there exists a BB-linear similitude V𝖵BV\,{\cong}\,\mathsf{V}_{B} through which the Hodge structure on 𝖵\mathsf{V} is defined by an element of Ω\Omega (see e.g. [MilIntro, Prop. 8.14, Thm. 8.17]). Here we implicitly applied the equivalence AH1(A,)A\mapsto\mathrm{H}^{1}(A,\mathbb{Q}) between the category of complex abelian varieties up to isogeny and that of polarizable Hodge structures of type {(1,0),(0,1)}\{(1,0),(0,1)\}.

Note that an object of 𝖱():=𝖱(Spec())\mathsf{R}(\mathbb{C}):=\mathsf{R}(\mathrm{Spec\,}(\mathbb{C})) is nothing but a polarizable Hodge structure. We can more concisely define \mathcal{I} as the isomorphism classes of pairs (𝖵,[η])(\mathsf{V},[\eta]) where

  • 𝖵𝖱()\mathsf{V}\in\mathsf{R}(\mathbb{C}) and

  • [η][\eta] is a 𝖵\mathsf{V}-rational (G,V,𝖪)(G,V,\mathsf{K})-level structure on 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}} of type Ω\Omega.

Indeed, if we view each bBb\in B as a tensor in End(V)=VV\mathrm{End}_{\mathbb{Q}}(V)=V^{\vee}\otimes V and c×ψc\coloneqq\mathbb{Q}^{\times}\cdot\psi as a tensor in V2(V)2V^{2}\otimes(V^{\vee})^{2} (cf. [Kim:RZ, Ex. 2.1.6]), we have i(b)𝔸f=η(b)e´ti(b){\otimes}\mathbb{A}_{f}=\eta(b)_{\mathrm{{\acute{e}}t}} and λ𝔸f=η(c)e´t\lambda{\otimes}\mathbb{A}_{f}=\eta(c)_{\mathrm{{\acute{e}}t}} in the notation of (3.1.2)(b). In particular, the datum of ii and λ\lambda is remembered by the condition that η(b)e´t\eta(b)_{\mathrm{{\acute{e}}t}} and η(c)e´t\eta(c)_{\mathrm{{\acute{e}}t}} come from global sections (i.e., elements) of 𝖵B\mathsf{V}_{B}^{\otimes} of Hodge type (0,0)(0,0). Since GG is the stabilizer of BB and cc in GL(V)\mathrm{GL}(V), we may equivalently say that this is true for all v(V)Gv\in(V^{\otimes})^{G}. Therefore, the existence of ii and λ\lambda such that there exists a BB-linear similitude V𝖵BV\,{\cong}\,\mathsf{V}_{B} is equivalent to [η][\eta] being 𝖵\mathsf{V}-rational. Then the condition ()(\spadesuit) can be simply stated as “[η][\eta] is of type Ω\Omega” in the sense of (3.1.2)(c).

Now we introduce some notation to keep track of Galois descent data in a system of realizations:

Notation (3.1.4).

Let SS be a smooth variety over a subfield kk of \mathbb{C}. Let sS()s\in S(\mathbb{C}) and σAut(/k)\sigma\in\mathrm{Aut}(\mathbb{C}/k) be any elements. Denote by σ(s)S()\sigma(s)\in S(\mathbb{C}) the point given by pre-composing the kk-linear morphism s:Spec()Ss:\mathrm{Spec\,}(\mathbb{C})\to S with Spec(σ)\mathrm{Spec\,}(\sigma). Given (𝖵B,𝖵e´t,𝖵dR)𝖱(S)(\mathsf{V}_{B},\mathsf{V}_{\mathrm{{\acute{e}}t}},\mathsf{V}_{\mathrm{dR}})\in\mathsf{R}(S), we write σ𝖵e´t,s:𝖵B,s𝔸f𝖵B,σ(s)𝔸f\sigma_{\mathsf{V}_{\mathrm{{\acute{e}}t}},s}:\mathsf{V}_{B,s}{\otimes}\mathbb{A}_{f}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{V}_{B,\sigma(s)}{\otimes}\mathbb{A}_{f} the natural isomorphism induced by 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}}, viewed as a descent of 𝖵B𝔸f\mathsf{V}_{B}{\otimes}\mathbb{A}_{f} from SS_{\mathbb{C}} to SS; similarly, we write σ𝖵dR,s:𝖵B,s𝖵B,σ(s)\sigma_{\mathsf{V}_{\mathrm{dR}},s}:\mathsf{V}_{B,s}{\otimes}\mathbb{C}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{V}_{B,\sigma(s)}{\otimes}\mathbb{C} for the natural σ\sigma-linear isomorphism induced by the descent 𝖵dR\mathsf{V}_{\mathrm{dR}} of (the algebraization of) 𝖵B𝒪San\mathsf{V}_{B}{\otimes}\mathcal{O}_{S_{\mathbb{C}}}^{\mathrm{an}}.

To clarify the meaning of σ𝖵e´t,s\sigma_{\mathsf{V}_{\mathrm{{\acute{e}}t}},s}, we remark that ss and σ(s)\sigma(s) are usually different closed points on SS_{\mathbb{C}}, and they are equal if and only if σ\sigma fixes the residue field k(s0)k(s_{0}), where s0Ss_{0}\in S is the image of ss. The collection of σ𝖵e´t,s\sigma_{\mathsf{V}_{\mathrm{{\acute{e}}t}},s} as σ\sigma runs through Aut(k(s)/k(s0))=Aut(/k(s0))\mathrm{Aut}(k(s)/k(s_{0}))=\mathrm{Aut}(\mathbb{C}/k(s_{0})) is nothing but the Galois action on the stalk 𝖵B,s𝔸f=𝖵e´t,s\mathsf{V}_{B,s}{\otimes}\mathbb{A}_{f}=\mathsf{V}_{\mathrm{{\acute{e}}t},s}.

Using the notations of (2.1.3) and (3.1.4), we define:

Definition (3.1.5).

We say that a system of realizations 𝖵\mathsf{V} is weakly abelian-motivic (weakly AM) if for every sS()s\in S(\mathbb{C}), there exists M𝖬𝗈𝗍𝖠𝖻()M\in\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}) such that ωHdg(M)(𝖵B,s,𝖵dR,s)\omega_{\mathrm{Hdg}}(M)\,{\cong}\,(\mathsf{V}_{B,s},\mathsf{V}_{\mathrm{dR},s}); moreover, for any such isomorphism γ:ωHdg(M)(𝖵B,s,𝖵dR,s)\gamma:\omega_{\mathrm{Hdg}}(M)\stackrel{{\scriptstyle\sim}}{{\to}}(\mathsf{V}_{B,s},\mathsf{V}_{\mathrm{dR},s}) and σAut(/k)\sigma\in\mathrm{Aut}(\mathbb{C}/k), there exists an isomorphism γσ:ωHdg(Mσ)(𝖵B,σ(s),𝖵dR,σ(s))\gamma^{\sigma}:\omega_{\mathrm{Hdg}}(M^{\sigma})\stackrel{{\scriptstyle\sim}}{{\to}}(\mathsf{V}_{B,\sigma(s)},\mathsf{V}_{\mathrm{dR},\sigma(s)}) such that the Betti components γB,γσB\gamma_{B},\gamma^{\sigma}_{B} of γ,γσ\gamma,\gamma^{\sigma} fit into commutative diagrams:

ωe´t(M){{\omega_{\mathrm{{\acute{e}}t}}(M)}}𝖵B,s𝔸f{{\mathsf{V}_{B,s}{\otimes}\mathbb{A}_{f}}}ωe´t(Mσ){{\omega_{\mathrm{{\acute{e}}t}}(M^{\sigma})}}𝖵B,σ(s)𝔸f{{\mathsf{V}_{B,\sigma(s)}{\otimes}\mathbb{A}_{f}}}bc\scriptstyle{\,{\cong}\,_{\mathrm{bc}}}σ𝖵e´t,s\scriptstyle{\sigma_{\mathsf{V}_{\mathrm{{\acute{e}}t}},s}}γσB𝔸f\scriptstyle{\gamma^{\sigma}_{B}{\otimes}\mathbb{A}_{f}}γB𝔸f\scriptstyle{\gamma_{B}{\otimes}\mathbb{A}_{f}} and ωdR(M){{\omega_{\mathrm{dR}}(M)}}𝖵B,s{{\mathsf{V}_{B,s}{\otimes}\mathbb{C}}}ωdR(Mσ){{\omega_{\mathrm{dR}}(M^{\sigma})}}𝖵B,σ(s){{\mathsf{V}_{B,\sigma(s)}{\otimes}\mathbb{C}}}bc\scriptstyle{\,{\cong}\,_{\mathrm{bc}}}σ𝖵dR,s\scriptstyle{\sigma_{\mathsf{V}_{\mathrm{dR}},s}}γσB\scriptstyle{\gamma^{\sigma}_{B}{\otimes}\mathbb{C}}γB\scriptstyle{\gamma_{B}{\otimes}\mathbb{C}} (4)

We note that γσ\gamma^{\sigma} is uniquely determined by γ\gamma provided that it exists. Denote the full subcategory of 𝖱(S)\mathsf{R}(S) given by these objects by 𝖱am(S)\mathsf{R}^{*}_{\mathrm{am}}(S). It is easy to check that if TST\to S is a morphism between smooth kk-varieties, the natural pullback functor 𝖱(S)𝖱(T)\mathsf{R}(S)\to\mathsf{R}(T) sends 𝖱am(S)\mathsf{R}^{*}_{\mathrm{am}}(S) to 𝖱am(T)\mathsf{R}^{*}_{\mathrm{am}}(T).

Lemma (3.1.6).

([YangSystem, (3.4.1)]) Let SS be a smooth variety over kk\subseteq\mathbb{C} and take 𝖵,𝖶𝖱am(S)\mathsf{V},\mathsf{W}\in\mathsf{R}_{\mathrm{am}}^{*}(S). Let φ\varphi_{\mathbb{C}} be a morphism 𝖵|S𝖶|S\mathsf{V}|_{S_{\mathbb{C}}}\to\mathsf{W}|_{S_{\mathbb{C}}}. Then φ\varphi_{\mathbb{C}} descends to a morphism 𝖵𝖶\mathsf{V}\to\mathsf{W} if and only if either the étale or the de Rham component of φ\varphi_{\mathbb{C}} descends to SS.

When applied to the case 𝖵=𝟏S\mathsf{V}=\mathbf{1}_{S}, the above lemma says that for 𝖶𝖱am(S)\mathsf{W}\in\mathsf{R}^{*}_{\mathrm{am}}(S), if the étale realization of a global section of 𝖶B|S\mathsf{W}_{B}|_{S_{\mathbb{C}}} which is everywhere of Hodge type (0,0)(0,0) descends to SS, then so does the de Rham realization, and vice versa. This is a global version of the following statement: Suppose that M𝖬𝗈𝗍AH(k)M\in\mathsf{Mot}_{\mathrm{AH}}(k) and vωB(M)v\in\omega_{B}(M_{\mathbb{C}}) is absolute Hodge. Then v𝔸fωe´t(M)v{\otimes}\mathbb{A}_{f}\in\omega_{\mathrm{{\acute{e}}t}}(M_{\mathbb{C}}) descends to kk (i.e., is Aut(/k)\mathrm{Aut}(\mathbb{C}/k)-invariant) if and only if vωdR(M)v{\otimes}\mathbb{C}\in\omega_{\mathrm{dR}}(M_{\mathbb{C}}) descends to ωdR(M)\omega_{\mathrm{dR}}(M). Note that if M𝖬𝗈𝗍𝖠𝖻(k)M\in\mathsf{Mot}_{\mathsf{Ab}}(k), then any Hodge cycle vωB(M)v\in\omega_{B}(M_{\mathbb{C}}) is automatically absolute Hodge.

For future reference we give a handy lemma on (b) and (c) in (3.1.2).

Lemma (3.1.7).

Suppose that in (3.1.2)(a), the representation VRep(G)V\in\mathrm{Rep}(G) is faithful, SS is geometrically connected as a kk-variety, and for some bS()b\in S(\mathbb{C}) there is an isomorphism ηb:V𝖵B,b\eta_{b}:V\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{V}_{B,b} such that [η]b=𝖪(ηb𝔸f)[\eta]_{b}=\mathsf{K}\cdot(\eta_{b}{\otimes}\mathbb{A}_{f}); moreover, for some Ω\Omega as in (3.1.2)(c), the Hodge structure on 𝖵B,b\mathsf{V}_{B,b} is defined by an element of Ω\Omega via ηb\eta_{b}.

Then, assuming either k=k=\mathbb{C} or 𝖵𝖱am(S)\mathsf{V}\in\mathsf{R}^{*}_{\mathrm{am}}(S), [η][\eta] is 𝖵\mathsf{V}-rational and of type Ω\Omega.

Proof.

Note that by assumption SS_{\mathbb{C}} is connected, and the fact that ηb𝔸f\eta_{b}{\otimes}\mathbb{A}_{f} represents a 𝖪\mathsf{K}-level structure at bb implies that its 𝖪\mathsf{K}-orbit is πe´t1(S,b)\pi^{\mathrm{{\acute{e}}t}}_{1}(S,b)-stable. In particular, for each v(V)Gv\in(V^{\otimes})^{G}, π1(S,b)\pi_{1}(S_{\mathbb{C}},b) fixes ηb(v)\eta_{b}(v), so there exists a 𝐯BH0(𝖵B)\mathbf{v}_{B}\in\mathrm{H}^{0}(\mathsf{V}_{B}^{\otimes}) such that 𝐯B,b=ηb(v)\mathbf{v}_{B,b}=\eta_{b}(v). Likewise, there exists 𝐯e´tH0(𝖵e´t)\mathbf{v}_{\mathrm{{\acute{e}}t}}\in\mathrm{H}^{0}(\mathsf{V}_{\mathrm{{\acute{e}}t}}^{\otimes}) such that 𝐯e´t,b=ηb(v)1\mathbf{v}_{\mathrm{{\acute{e}}t},b}=\eta_{b}(v){\otimes}1, and 𝐯e´t\mathbf{v}_{\mathrm{{\acute{e}}t}} is precisely the η(v)e´t\eta(v)_{\mathrm{{\acute{e}}t}} in (3.1.2)(b).

Note that 𝐯B\mathbf{v}_{B} is necessarily of Hodge type (0,0)(0,0) at bb, because the Mumford-Tate group of the Hodge structure (𝖵B,b,𝖵dR,b)(\mathsf{V}_{B,b},\mathsf{V}_{\mathrm{dR},b}) is contained in GG via ηb\eta_{b}. This implies that 𝐯B\mathbf{v}_{B} is of Hodge type (0,0)(0,0) everywhere by the theorem of the fixed part. Let 𝐯dR,\mathbf{v}_{\mathrm{dR},\mathbb{C}} be the global section in (𝖵B𝒪San)(\mathsf{V}_{B}{\otimes}\mathcal{O}_{S_{\mathbb{C}}}^{\mathrm{an}})^{\otimes} induced by 𝐯B\mathbf{v}_{B}. Then 𝐯dR,\mathbf{v}_{\mathrm{dR},\mathbb{C}} algebraizes to a global section of (𝖵dR|S)(\mathsf{V}_{\mathrm{dR}}|_{S_{\mathbb{C}}})^{\otimes} (cf. [DelVB, II Thm 5.9]).

If k=k=\mathbb{C} (so that S=SS=S_{\mathbb{C}}), then we have already shown that [η][\eta] is 𝖵\mathsf{V}-rational; moreover, one deduces from the connectedness of SS_{\mathbb{C}} and [DelVdShimura, (1.1.12)] that [η][\eta] remains of type Ω\Omega on SS_{\mathbb{C}}. If kk\subseteq\mathbb{C} is a general subfield, it remains to show that 𝐯dR,\mathbf{v}_{\mathrm{dR},\mathbb{C}} descends to SS under the additional assumption that 𝖵𝖱am(S)\mathsf{V}\in\mathsf{R}^{*}_{\mathrm{am}}(S). However, as the étale realization of 𝐯B\mathbf{v}_{B} descends to a global section of 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}}^{\otimes} (i.e., 𝐯e´t\mathbf{v}_{\mathrm{{\acute{e}}t}}), this follows from (3.1.6). ∎

3.2   Shimura varieties

Let (G,Ω)(G,\Omega) be a Shimura datum which satisfies the axioms in [Milne:CanonicalModels, II (2.1)]. Let E(G,Ω)E(G,\Omega) be the reflex field, ZZ be the center of GG and ZsZ_{s} be the maximal anisotropic subtorus of ZZ that is split over \mathbb{R}. In this paper, we always assume that

the weight is defined over  and Zs is trivial.\text{the weight is defined over $\mathbb{Q}$ and $Z_{s}$ is trivial}. (5)

Note that in particular the latter condition ensures that Z()Z(\mathbb{Q}) is discrete in Z(𝔸f)Z(\mathbb{A}_{f}) ([MilIntro, Rmk 5.27]). We will often drop the Hermitian symmetric domain Ω\Omega from the notation of Shimura varieties when no confusion would arise.

For any compact open subgroup 𝖪G(𝔸f)\mathsf{K}\subseteq G(\mathbb{A}_{f}), let Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}} denote the resulting Shimura variety with a complex uniformization G()\Ω×G(𝔸f)/𝖪G(\mathbb{Q})\backslash\Omega\times G(\mathbb{A}_{f})/\mathsf{K} and let Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) denote the canonical model over E(G,Ω)E(G,\Omega). Let Sh(G)\mathrm{Sh}(G) denote the inverse limit lim𝖪Sh𝖪(G)\varprojlim_{\mathsf{K}}\mathrm{Sh}_{\mathsf{K}}(G) as 𝖪\mathsf{K} runs through all compact open subgroups. Under our assumptions, Sh(G)()\mathrm{Sh}(G)(\mathbb{C}) is described by G()\Ω×G(𝔸f)G(\mathbb{Q})\backslash\Omega\times G(\mathbb{A}_{f}) ([MilIntro, (5.28)]). Note that Sh𝖪(G)=Sh(G)/𝖪\mathrm{Sh}_{\mathsf{K}}(G)=\mathrm{Sh}(G)/\mathsf{K}.

(3.2.1)   

Let GGL(V)G\to\mathrm{GL}(V) be a representation. For any neat compact open subgroup 𝖪G(𝔸f)\mathsf{K}\subseteq G(\mathbb{A}_{f}), we can attach to Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}} an automorphic VHS (𝖵B,𝖵dR,)(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR},\mathbb{C}}) (cf. [Milne:CanonicalModels, Ch. II 3.3], [Taelman2, §2.2]). In particular, 𝖵B\mathsf{V}_{B} is defined to be the contraction product

𝖵B:=V×G()[Ω×G(𝔸f)/𝖪]\mathsf{V}_{B}:=V\times^{G(\mathbb{Q})}[\Omega\times G(\mathbb{A}_{f})/\mathsf{K}] (6)

The filtration on 𝖵dR,\mathsf{V}_{\mathrm{dR},\mathbb{C}} is obtained by descending the filtration on the tautological VHS on V×ΩV\times\Omega. Analogously, the automorphic étale local system 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}} on the proétale site of Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) is defined as the contraction product

𝖵e´t:=V𝔸f×𝖪Sh(G)\mathsf{V}_{\mathrm{{\acute{e}}t}}:=V_{\mathbb{A}_{f}}\times^{\mathsf{K}}\mathrm{Sh}(G) (7)

which comes with a comparison isomorphism 𝖵B𝔸f𝖵e´t|Sh𝖪(G)\mathsf{V}_{B}\otimes\mathbb{A}_{f}\,{\cong}\,\mathsf{V}_{\mathrm{{\acute{e}}t}}|_{\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}}}. Moreover, by construction 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}} over Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) comes with a tautological 𝖪\mathsf{K}-level structure [ηV][\eta_{V}], i.e., a global section of 𝖪\Isom¯(V𝔸f,𝖵e´t)\mathsf{K}\backslash\underline{\mathrm{Isom}}(V{\otimes}\mathbb{A}_{f},\mathsf{V}_{\mathrm{{\acute{e}}t}}).

Define [ηanV][\eta^{\mathrm{an}}_{V}] to be the restriction of [ηV][\eta_{V}] to Sh𝖪(G){\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}}}. Using (𝖵B,𝖵dR,)(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR},\mathbb{C}}), we can already give a moduli interpretation of Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}}. Below is a reformulation of [Mil94, Prop. 3.10] in our terminology (cf. [YangSystem, (4.1.2)])555Technically, the example discussed in (3.1.3) may not satisfy (5), but it helps illustrate how to compare our formulation and Milne’s..

Theorem (3.2.2).

Assume that VV is faithful. For every smooth \mathbb{C}-variety TT, let V(T)\mathcal{M}_{V}(T) be the groupoid of tuples of the form (𝖶,[ξan])(\mathsf{W},[\xi^{\mathrm{an}}]) where 𝖶=(𝖶B,𝖶dR)\mathsf{W}=(\mathsf{W}_{B},\mathsf{W}_{\mathrm{dR}}) is a VHS over TT and [ξan][\xi^{\mathrm{an}}] is a 𝖪\mathsf{K}-level structure on 𝖶B𝔸f\mathsf{W}_{B}{\otimes}\mathbb{A}_{f} which is 𝖶\mathsf{W}-rational and is of type Ω\Omega.

Then (𝖵:=(𝖵B,𝖵dR,),[ηanV])(\mathsf{V}_{\mathbb{C}}:=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR},\mathbb{C}}),[\eta^{\mathrm{an}}_{V}]) is an object of V(Sh𝖪(G))\mathcal{M}_{V}(\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}}) and for every object (𝖶,[ξan])V(T)(\mathsf{W},[\xi^{\mathrm{an}}])\in\mathcal{M}_{V}(T) there exists a unique morphism ρ:TSh𝖪(G)\rho:T\to\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}} such that ρ(𝖵,[ηanV])(𝖶,[ξan])\rho^{*}(\mathsf{V}_{\mathbb{C}},[\eta^{\mathrm{an}}_{V}])\,{\cong}\,(\mathsf{W},[\xi^{\mathrm{an}}]).

We remark that as 𝖪\mathsf{K} is neat, assumption (5) ensures that the objects in V(T)\mathcal{M}_{V}(T) above have no nontrivial automorphisms (cf. [Mil94, Rmk 3.11]).

To describe the moduli problem for the canonical model, we restrict to the following subclass of Shimura data, which contains all cases we will consider in the following chapters.

Assumption (3.2.3).

There exists a morphism of Shimura data (G~,Ω~)(G,Ω)(\widetilde{G},\widetilde{\Omega})\to(G,\Omega) such that

  1. (i)

    (G~,Ω~)(\widetilde{G},\widetilde{\Omega}) is of Hodge type and also satisfies assumption (5);

  2. (ii)

    G~G\widetilde{G}\to G is surjective and the kernel lies in the center of G~\widetilde{G};

  3. (iii)

    the embedding of reflex fields E(G,Ω)E(G~,Ω~)E(G,\Omega)\subseteq E(\widetilde{G},\widetilde{\Omega}) is an equality.

Below we assume that (G,Ω)(G,\Omega) is a Shimura datum which satisfies the above assumptions.

(3.2.4)   

For any representation GGL(V)G\to\mathrm{GL}(V), and 𝖵B,𝖵dR,,𝖵e´t\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR},\mathbb{C}},\mathsf{V}_{\mathrm{{\acute{e}}t}} set up in (3.2.1), there exists a unique descent 𝖵dR\mathsf{V}_{\mathrm{dR}} of 𝖵dR,\mathsf{V}_{\mathrm{dR},\mathbb{C}} to Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) such that (𝖵B,𝖵dR,𝖵e´t)(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}},\mathsf{V}_{\mathrm{{\acute{e}}t}}) is a weakly AM system of realizations (see ([YangSystem, (4.2.2)])). We call 𝖵:=(𝖵B,𝖵dR,𝖵e´t)𝖱am(Sh𝖪(G))\mathsf{V}:=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}},\mathsf{V}_{\mathrm{{\acute{e}}t}})\in\mathsf{R}^{*}_{\mathrm{am}}(\mathrm{Sh}_{\mathsf{K}}(G)) the automorphic system (of realizations) on Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) attached to VRep(G)V\in\mathrm{Rep}(G). The tautological 𝖪\mathsf{K}-level structure [ηV][\eta_{V}] on 𝖵\mathsf{V} is 𝖵\mathsf{V}-rational, and is of type Ω\Omega when VV is faithful.

Theorem (3.2.5).

([YangSystem, (4.3.1)]) Assume that VV is faithful. Let EE^{\prime}\subseteq\mathbb{C} be a subfield which contains E=E(G,Ω)E=E(G,\Omega) and TT be a smooth EE^{\prime}-variety. Let V,E(T)\mathcal{M}_{V,E^{\prime}}(T) be the groupoid of pairs (𝖶,[ξ])(\mathsf{W},[\xi]) where 𝖶=(𝖶B,𝖶dR,𝖶e´t)𝖱am(T)\mathsf{W}=(\mathsf{W}_{B},\mathsf{W}_{\mathrm{dR}},\mathsf{W}_{\mathrm{{\acute{e}}t}})\in\mathsf{R}^{*}_{\mathrm{am}}(T), and [ξ][\xi] is a 𝖪\mathsf{K}-level structure on 𝖶e´t\mathsf{W}_{\mathrm{{\acute{e}}t}} such that [ξ][\xi] is 𝖶\mathsf{W}-rational and (𝖶,[ξ])(\mathsf{W},[\xi]) is of type Ω\Omega.

Then (𝖵=(𝖵B,𝖵dR,𝖵e´t),[ηV])(\mathsf{V}=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}},\mathsf{V}_{\mathrm{{\acute{e}}t}}),[\eta_{V}]) is an object of V,E(Sh𝖪(G))\mathcal{M}_{V,E}(\mathrm{Sh}_{\mathsf{K}}(G)), and for each (𝖶,[ξ])V,E(T)(\mathsf{W},[\xi])\in\mathcal{M}_{V,E^{\prime}}(T), there exists a unique ρ:TSh𝖪(G)E\rho:T\to\mathrm{Sh}_{\mathsf{K}}(G)_{E^{\prime}} such that (𝖶,[ξ])ρ(𝖵,[ηV])(\mathsf{W},[\xi])\,{\cong}\,\rho^{*}(\mathsf{V},[\eta_{V}]).

Note that any object in V,E(T)\mathcal{M}_{V,E^{\prime}}(T) above has no nontrivial automorphisms because this is already true when E=E^{\prime}=\mathbb{C}. This is a key fact that we shall use repeatedly throughout the paper. The above is proved by first constructing a morphism ρ:TSh𝖪(G)\rho_{\mathbb{C}}:T_{\mathbb{C}}\to\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}} such that (𝖶,[ξ])|Tρ(𝖵,[ηV])(\mathsf{W},[\xi])|_{T_{\mathbb{C}}}\,{\cong}\,\rho_{\mathbb{C}}^{*}(\mathsf{V},[\eta_{V}]) and then showing that the action of ρ\rho_{\mathbb{C}} on the \mathbb{C}-points are Aut(/E)\mathrm{Aut}(\mathbb{C}/E^{\prime})-equivariant. This is clearly inspired by the proof of [MPTate, Cor. 5.4]. However, unlike [MPTate, Prop. 5.6(1)], we show that (𝖶,[ξ])|Tρ(𝖵,[ηV])(\mathsf{W},[\xi])|_{T_{\mathbb{C}}}\,{\cong}\,\rho_{\mathbb{C}}^{*}(\mathsf{V},[\eta_{V}]) descends over TT using a rigidity lemma about weakly-AM systems of realizations (see [YangSystem, (3.4.4), (3.4.5) and (4.3.2)]).

3.3   Orthogonal Shimura Varieties over Totally Real Fields

(3.3.1)   

Let EE be a totally real number field and let 𝒱\mathcal{V} be a quadratic form over EE which has signature (2,dimE𝒱2)(2,\dim_{E}\mathcal{V}-2) at a unique real place τ\tau and is negative definitive at every other real place. We set 𝒢:=ResE/SO(𝒱)\mathcal{G}:=\mathrm{Res}_{E/\mathbb{Q}}\mathrm{SO}(\mathcal{V}) and

Ω𝒱{v(𝒱τ)v,v=0,v,v¯>0}.\Omega_{\mathcal{V}}\coloneqq\{v\in\mathbb{P}(\mathcal{V}{\otimes}_{\tau}\mathbb{C})\mid\langle v,v\rangle=0,\langle v,\bar{v}\rangle>0\}.

Let VV be the transfer 𝒱()\mathcal{V}_{(\mathbb{Q})} (recall (2.1.5)). Set G:=SO(V)G:=\mathrm{SO}(V) and define Ω:=ΩV\Omega:=\Omega_{V} as above (applied to the E=E=\mathbb{Q} case). Note that 𝒢\mathcal{G} is the identity component of the centeralizer of the EE-action in GG. We view Ω\Omega as a G()G(\mathbb{R})-conjugacy class of morphisms 𝕊G\mathbb{S}\to G_{\mathbb{R}} such that an element vΩv\in\Omega corresponds to the morphism hh which gives VV a Hodge structure of K3-type with v=V(1,1)v=V_{\mathbb{C}}^{(1,-1)} (cf. [CSpin, §3.1]). Likewise we view Ω𝒱\Omega_{\mathcal{V}} as a 𝒢()\mathcal{G}(\mathbb{R})-conjugacy class of morphisms 𝕊𝒢\mathbb{S}\to\mathcal{G}_{\mathbb{R}}, which are those whose composition with 𝒢G\mathcal{G}_{\mathbb{R}}\to G_{\mathbb{R}} defines a Hodge structure on VV which is preserved by the EE-action on V=𝒱()V=\mathcal{V}_{(\mathbb{Q})}. It is well-known that (𝒢,Ω𝒱)(\mathcal{G},\Omega_{\mathcal{V}}) is a Shimura datum with reflex field EE, viewed as a subfield of \mathbb{C} under τ\tau.

Let G~\widetilde{G} denote CSpin(V)\mathrm{CSpin}(V). Then (G~,Ω)(\widetilde{G},\Omega) is a Shimura datum of Hodge type with reflex field \mathbb{Q} which admits a natural morphism to (G,Ω)(G,\Omega).

Lemma (3.3.2).

(𝒢,Ω𝒱)(\mathcal{G},\Omega_{\mathcal{V}}) satisfies Assumption (3.2.3).

Proof.

Consider 𝒢~:=ResE/CSpin(𝒱)\widetilde{\mathcal{G}}^{\prime}:=\mathrm{Res}_{E/\mathbb{Q}}\mathrm{CSpin}(\mathcal{V}). It is well known that (𝒢~,Ω𝒱)(\widetilde{\mathcal{G}}^{\prime},\Omega_{\mathcal{V}}) is a Shimura datum with reflex field EE. Unfortunately, it does not satisfy condition (5) unless E=E=\mathbb{Q} as the center ZZ^{\prime} of 𝒢~\widetilde{\mathcal{G}}^{\prime} has identity component ResE/𝔾m,E\mathrm{Res}_{E/\mathbb{Q}}\mathbb{G}_{m,E}. Thus we modify this approach by dividing out the maximal anisotropic torus of ResE/𝔾m,E\mathrm{Res}_{E/\mathbb{Q}}\mathbb{G}_{m,E}; explicitly we consider (Nm=\mathrm{Nm}= the norm map)

𝒢~𝒢~/ker(Nm:ResE/𝔾m,E𝔾m)).\widetilde{\mathcal{G}}\coloneqq\widetilde{\mathcal{G}}^{\prime}/\ker(\mathrm{Nm}\colon\mathrm{Res}_{E/\mathbb{Q}}\mathbb{G}_{m,E}\to\mathbb{G}_{m})).

Note that ker(𝒢~𝒢)=ResE/𝔾m,E\ker(\widetilde{\mathcal{G}^{\prime}}\to\mathcal{G})=\mathrm{Res}_{E/\mathbb{Q}}\mathbb{G}_{m,E}, so we obtain a morphism (𝒢~,Ω𝒱)(𝒢,Ω𝒱)(\widetilde{\mathcal{G}},\Omega_{\mathcal{V}})\to(\mathcal{G},\Omega_{\mathcal{V}}) of Shimura data. It remains to check (i) and (iii) in (3.2.3). First, note that 𝒢~\widetilde{\mathcal{G}} can be canonically identified with the fiber product 𝒢×GG~\mathcal{G}\times_{G}\widetilde{G} (see [Moonen, §4.2, 4.3]), so that there is a diagram with exact rows

1{1}𝔾m{{\mathbb{G}_{m}}}𝒢~{{\widetilde{\mathcal{G}}}}𝒢{\mathcal{G}}1{1}1{1}𝔾m{{\mathbb{G}_{m}}}G~{{\widetilde{G}}}G{G}1.{1.}\scriptstyle{\lrcorner}

Now Zs(𝒢~)=1Z_{s}(\widetilde{\mathcal{G}})=1 is clear. As (G~,Ω)(\widetilde{G},\Omega) is of Hodge type and (𝒢~,Ω𝒱)(\widetilde{\mathcal{G}},\Omega_{\mathcal{V}}) embeds into (G~,Ω)(\widetilde{G},\Omega), (𝒢~,Ω𝒱)(\widetilde{\mathcal{G}},\Omega_{\mathcal{V}}) is also of Hodge type. As the reflex fields of (𝒢,Ω𝒱)(\mathcal{G},\Omega_{\mathcal{V}}) and (𝒢~,Ω𝒱)(\widetilde{\mathcal{G}}^{\prime},\Omega_{\mathcal{V}}) are both well known to be EE, the same must be true for (𝒢~,Ω𝒱)(\widetilde{\mathcal{G}},\Omega_{\mathcal{V}}). ∎

(3.3.3)   

We do some preparations to future reference. Below for any \mathbb{Q}-linear Tannakian category 𝒞\mathscr{C}, we write 𝔑:𝒞(E)𝒞\mathfrak{N}:\mathscr{C}_{(E)}\to\mathscr{C} for either one of the following two functors: MNmE/(M)M\mapsto\mathrm{Nm}_{E/\mathbb{Q}}(M) or MNmE/(M)det(M())M\mapsto\mathrm{Nm}_{E/\mathbb{Q}}(M){\otimes}_{\mathbb{Q}}\det(M_{(\mathbb{Q})}) (see notations in (2.1.4)). Any conclusion applies to both functors.

Recall the notations from (2.1.5), and write Z1E(𝒱)Z^{1}_{E}(\mathcal{V}) and (𝒱)\mathcal{H}(\mathcal{V}) simply as Z1EZ^{1}_{E} and \mathcal{H}. Then N:=𝔑(𝒱)N:=\mathfrak{N}(\mathcal{V}) is a faithful representation of \mathcal{H}, as the composite 𝒢ResE/GL(𝒱)GL(N)\mathcal{G}\to\mathrm{Res}_{E/\mathbb{Q}}\mathrm{GL}(\mathcal{V})\to\mathrm{GL}(N) has kernel exactly Z1EZ^{1}_{E}. Let 𝒦𝒢(𝔸f)\mathcal{K}\subseteq\mathcal{G}(\mathbb{A}_{f}) and 𝒞(𝔸f)\mathcal{C}\subseteq\mathcal{H}(\mathbb{A}_{f}) be compact open subgroups such that 𝒞\mathcal{C} contains the image of 𝒦\mathcal{K}. For each prime \ell, let 𝒦\mathcal{K}_{\ell} be the image of 𝒦\mathcal{K} under the projection 𝒢(𝔸f)𝒢()\mathcal{G}(\mathbb{A}_{f})\to\mathcal{G}(\mathbb{Q}_{\ell}).

Let Ω\Omega_{\mathcal{H}} be the ()\mathcal{H}(\mathbb{R})-conjugacy class of morphisms 𝕊\mathbb{S}\to\mathcal{H}_{\mathbb{R}} which contains the image of Ω𝒱\Omega_{\mathcal{V}}. Then (,Ω)(\mathcal{H},\Omega_{\mathcal{H}}) is a Shimura datum which admits a natural morphism (𝒢,Ω𝒱)(,Ω)(\mathcal{G},\Omega_{\mathcal{V}})\to(\mathcal{H},\Omega_{\mathcal{H}}). Since Z1EZ^{1}_{E} lies in the center of 𝒢\mathcal{G} and is discrete, by [DeligneTdShimura, Prop. 3.8] (,Ω)(\mathcal{H},\Omega_{\mathcal{H}}) has the same reflex field as (𝒢,Ω𝒱)(\mathcal{G},\Omega_{\mathcal{V}}). Moreover, one easily checks that (,Ω)(\mathcal{H},\Omega_{\mathcal{H}}) satisfies (3.2.3) using that (𝒢,Ω𝒱)(\mathcal{G},\Omega_{\mathcal{V}}) does.

Remark (3.3.4).

Note that if dimE𝒱\dim_{E}\mathcal{V} is odd, then Z1EZ^{1}_{E} is trivial; in this case, the reader should read the content below with (𝒢,Ω𝒱,𝒦)=(,Ω,𝒞)(\mathcal{G},\Omega_{\mathcal{V}},\mathcal{K})=(\mathcal{H},\Omega_{\mathcal{H}},\mathcal{C}) in mind. The case when dimE𝒱\dim_{E}\mathcal{V} is even will only be used for §4.3.

(3.3.5)   

Let kk\subseteq\mathbb{C} be a subfield and TT be a smooth kk-variety. Note that the identification V=𝒱()V=\mathcal{V}_{(\mathbb{Q})} gives VV an EE-action, which commutes with 𝒢\mathcal{G}. Suppose that 𝖶𝖱(T)\mathsf{W}\in\mathsf{R}(T) is a system equipped with a (𝒢,V,𝒦)(\mathcal{G},V,\mathcal{K})-level structure [μ][\mu]. Then [μ][\mu] transports the EE-action on V𝔸fV{\otimes}\mathbb{A}_{f} to one on 𝖶e´t\mathsf{W}_{\mathrm{{\acute{e}}t}}, through which we may view [μ][\mu] as a global section of 𝒦\Isom¯E(𝒱𝔸f,𝖶e´t)\mathcal{K}\backslash\underline{\mathrm{Isom}}_{E}(\mathcal{V}{\otimes}_{\mathbb{Q}}\mathbb{A}_{f},\mathsf{W}_{\mathrm{{\acute{e}}t}}). Moreover, there is a natural map

𝒦\Isom¯E(𝒱𝔸f,𝖶e´t)𝒞\Isom¯(𝔑(𝒱)𝔸f,𝔑(𝖶e´t))\mathcal{K}\backslash\underline{\mathrm{Isom}}_{E}(\mathcal{V}{\otimes}_{\mathbb{Q}}\mathbb{A}_{f},\mathsf{W}_{\mathrm{{\acute{e}}t}})\to\mathcal{C}\backslash\underline{\mathrm{Isom}}(\mathfrak{N}(\mathcal{V}){\otimes}\mathbb{A}_{f},\mathfrak{N}(\mathsf{W}_{\mathrm{{\acute{e}}t}}))

through which [μ][\mu] defines a (,N,𝒞)(\mathcal{H},N,\mathcal{C})-level structure on 𝔑(𝖶e´t)\mathfrak{N}(\mathsf{W}_{\mathrm{{\acute{e}}t}}), which we denote by 𝔑([μ])\mathfrak{N}([\mu]).

Lemma (3.3.6).

Let TT, 𝖶\mathsf{W} and [μ][\mu] be as above. Let 𝖶𝖱(T)\mathsf{W}^{\prime}\in\mathsf{R}(T) be another system with 𝒦\mathcal{K}-level structure [μ][\mu^{\prime}]. Suppose that there is an isomorphism γe´t:(𝖶e´t,[μ])|T(𝖶e´t,[μ])|T\gamma_{\mathrm{{\acute{e}}t}}:(\mathsf{W}_{\mathrm{{\acute{e}}t}},[\mu])|_{T_{\mathbb{C}}}\,{\cong}\,(\mathsf{W}_{\mathrm{{\acute{e}}t}}^{\prime},[\mu^{\prime}])|_{T_{\mathbb{C}}} such that 𝔑(γe´t)\mathfrak{N}(\gamma_{\mathrm{{\acute{e}}t}}) descends to an isomorphism (𝔑(𝖶e´t),𝔑([μ]))(𝔑(𝖶e´t),𝔑([μ]))(\mathfrak{N}(\mathsf{W}_{\mathrm{{\acute{e}}t}}),\mathfrak{N}([\mu]))\,{\cong}\,(\mathfrak{N}(\mathsf{W}_{\mathrm{{\acute{e}}t}}^{\prime}),\mathfrak{N}([\mu^{\prime}])) over TT.

If \ell is a prime such that 𝒦Z1E()=1\mathcal{K}_{\ell}\cap Z^{1}_{E}(\mathbb{Q}_{\ell})=1, then the \ell-adic component γ\gamma_{\ell} of γe´t\gamma_{\mathrm{{\acute{e}}t}} descends to an isomorphism 𝖶𝖶\mathsf{W}_{\ell}\,{\cong}\,\mathsf{W}_{\ell}^{\prime} over TT.

Proof.

We may assume that TT is connected. Choose a base point bT()b\in T(\mathbb{C}). Set γ:=γ,b\gamma:=\gamma_{\ell,b} and let gπ1e´t(T,b)g\in\pi_{1}^{\mathrm{{\acute{e}}t}}(T,b) be any element. Our goal is to show that δ:=g1γ1gγ=1\delta:=g^{-1}\gamma^{-1}g\gamma=1. Note that as 𝔑(γe´t)\mathfrak{N}(\gamma_{\mathrm{{\acute{e}}t}}) descends to TT, we already know that 𝔑(δ)=1\mathfrak{N}(\delta)=1. Let μ:𝒱𝔸f𝖶e´t,b\mu:\mathcal{V}{\otimes}_{\mathbb{Q}}\mathbb{A}_{f}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{W}_{\mathrm{{\acute{e}}t},b} be a representative of [μ]b[\mu]_{b}. Then μ1gμ𝒦\mu^{-1}g\mu\in\mathcal{K}_{\ell}, because the 𝒦\mathcal{K}-orbit [μ]b[\mu]_{b} is π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b)-stable. Set μ:=γμ\mu^{\prime}:=\gamma\mu. Then μ\mu^{\prime} represents [μ]b[\mu^{\prime}]_{b}, so that (μ)1gμ𝒦(\mu^{\prime})^{-1}g\mu^{\prime}\in\mathcal{K}_{\ell}. Now we have

μ1δμ=[μ1g1μ][(μ)1gμ]𝒦.\mu^{-1}\delta\mu=[\mu^{-1}g^{-1}\mu][(\mu^{\prime})^{-1}g\mu^{\prime}]\in\mathcal{K}_{\ell}.

Note that 𝔑(μ1δμ)=1GL(N)\mathfrak{N}(\mu^{-1}\delta\mu)=1\in\mathrm{GL}(N{\otimes}\mathbb{Q}_{\ell}). However, as the kernel of the 𝒦\mathcal{K}_{\ell}-action on NN{\otimes}\mathbb{Q}_{\ell} lies in Z1E()Z^{1}_{E}(\mathbb{Q}_{\ell}), we must have μ1δμ=1\mu^{-1}\delta\mu=1, i.e., δ=1\delta=1. ∎

(3.3.7)   

We will often consider the following diagram of Shimura data:

(𝒢,Ω𝒱){{(\mathcal{G},\Omega_{\mathcal{V}})}}(G,Ω){{(G,\Omega)}}(,Ω){{(\mathcal{H},\Omega_{\mathcal{H}})}}i\scriptstyle{i}π\scriptstyle{\pi}

Let 𝖪G(𝔸f),𝒦𝒢(𝔸f)\mathsf{K}\subseteq G(\mathbb{A}_{f}),\mathcal{K}\subseteq\mathcal{G}(\mathbb{A}_{f}) and 𝒞(𝔸f)\mathcal{C}\subseteq\mathcal{H}(\mathbb{A}_{f}) be neat compact open subgroups such that 𝒦𝖪\mathcal{K}\subseteq\mathsf{K} and 𝒞\mathcal{C} contains the image of 𝒦\mathcal{K}. Then we have Shimura morphisms i:Sh𝒦(𝒢)Sh𝖪(G)Ei\colon\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})\to\mathrm{Sh}_{\mathsf{K}}(G)_{E} and π:Sh𝒦(𝒢)Sh𝒞()\pi\colon\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})\to\mathrm{Sh}_{\mathcal{C}}(\mathcal{H}) defined over EE. Let 𝖵\mathsf{V} (resp. 𝖵~\widetilde{\mathsf{V}}) be the automorphic system on Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) (resp. Sh𝒦(𝒢)\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})) attached to the standard representation VV of GG (resp. 𝒢\mathcal{G}) and [ηV][\eta_{V}] (resp. [η𝒱][\eta_{\mathcal{V}}]) be the tautological 𝖪\mathsf{K}-level structure (resp. 𝒦\mathcal{K}-level structure), as defined in (3.2.4). Then there is a natural identification 𝖵~=i(𝖵)\widetilde{\mathsf{V}}=i^{*}(\mathsf{V}) and [η𝒱][\eta_{\mathcal{V}}] refines i([ηV])i^{*}([\eta_{V}]) in the sense of (3.1.2)(d).

Note that [η𝒱][\eta_{\mathcal{V}}] endows 𝖵~\widetilde{\mathsf{V}} with a canonical EE-action. A priori it only defines an EE-action on 𝖵~e´t\widetilde{\mathsf{V}}_{\mathrm{{\acute{e}}t}}, but since [η𝒱][\eta_{\mathcal{V}}] is 𝖵~\widetilde{\mathsf{V}}-rational, its restriction to Sh𝒦(𝒢)\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})_{\mathbb{C}} comes from an EE-action on 𝖵~B\widetilde{\mathsf{V}}_{B}. Then by (3.1.6), one deduces that the resulting EE-action on 𝖵~dR|Sh𝒦(𝒢)\widetilde{\mathsf{V}}_{\mathrm{dR}}|_{\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})_{\mathbb{C}}} via the Riemann-Hilbert correspondence descends to Sh𝒦(𝒢)\mathrm{Sh}_{\mathcal{K}}(\mathcal{G}). Therefore, it makes sense to form 𝔑(𝖵~)\mathfrak{N}(\widetilde{\mathsf{V}}). Since 𝖵~\widetilde{\mathsf{V}} is weakly AM in the sense of (3.1.5), so is 𝔑(𝖵~)\mathfrak{N}(\widetilde{\mathsf{V}}). This is simply because we may apply the functor 𝔑()\mathfrak{N}(-) to 𝖬𝗈𝗍𝖠𝖻()(E)\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C})_{(E)} and it commutes with the cohomological realizations.

Now let 𝖭\mathsf{N} be the automorphic system on Sh𝒞()\mathrm{Sh}_{\mathcal{C}}(\mathcal{H}) attached to NRep()N\in\mathrm{Rep}(\mathcal{H}) and let [ηN][\eta_{N}] be its tautological 𝒞\mathcal{C}-level structure. We claim that there is an (necessarily unique) isomorphism

π(𝖭,[ηN])(𝔑(𝖵~),𝔑([η𝒱])).\pi^{*}(\mathsf{N},[\eta_{N}])\,{\cong}\,(\mathfrak{N}({\widetilde{\mathsf{V}}}),\mathfrak{N}([\eta_{\mathcal{V}}])). (8)

Indeed, to verify this isomorphism one first checks its implications on the automorphic VHS and étale local systems, which follow from the explicit descriptions in (3.2.1), then applies (3.1.6).

3.4   Integral Model

Let VV be a quadratic form over \mathbb{Q} and suppose that there is a self-dual (p)\mathbb{Z}_{(p)}-lattice L(p)VL_{(p)}\subseteq V for a prime p>2p>2. Then G:=SO(V)G:=\mathrm{SO}(V) extends to the reductive (p)\mathbb{Z}_{(p)}-group SO(L(p))\mathrm{SO}(L_{(p)}), which we still write as GG by abuse of notation. Let 𝖪G(𝔸f)\mathsf{K}\subseteq G(\mathbb{A}_{f}) be a neat compact open subgroup of the form 𝖪p𝖪p\mathsf{K}_{p}\mathsf{K}^{p} with 𝖪p=G(p)\mathsf{K}_{p}=G(\mathbb{Z}_{p}) and 𝖪pG(𝔸pf)\mathsf{K}^{p}\subseteq G(\mathbb{A}^{p}_{f}). Then by [KisinInt] and [CSpin], Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) admits a canonical integral model 𝒮𝖪(G)\mathscr{S}_{\mathsf{K}}(G) over (p)\mathbb{Z}_{(p)}.

The Shimura variety 𝒮𝖪(G)\mathscr{S}_{\mathsf{K}}(G) is typically studied via the corresponding spinor Shimura variety, which is of Hodge type. Let G~CSpin(L(p))\widetilde{G}\coloneqq\mathrm{CSpin}(L_{(p)}). Set 𝕂p\mathbb{K}_{p} to be G~(p)\widetilde{G}(\mathbb{Z}_{p}), 𝕂pG~(𝔸pf)\mathbb{K}^{p}\subseteq\widetilde{G}(\mathbb{A}^{p}_{f}) to be a small enough compact open subgroup whose image in G(𝔸pf)G(\mathbb{A}^{p}_{f}) is contained in 𝖪p\mathsf{K}^{p}, and 𝕂\mathbb{K} to be the product 𝕂p𝕂p\mathbb{K}_{p}\mathbb{K}^{p}. The reflex field of (G~,Ω)(\widetilde{G},\Omega) is \mathbb{Q} and by [KisinInt] there is an canonical integral model 𝒮𝕂(G~)\mathscr{S}_{\mathbb{K}}(\widetilde{G}) over (p)\mathbb{Z}_{(p)}. There is a suitable sympletic space (H,ψ)(H,\psi) and a Siegel half space ±\mathcal{H}^{\pm} such that there is an embedding of Shimura data (G~,Ω)(GSp(ψ),±)(\widetilde{G},\Omega)\hookrightarrow(\mathrm{GSp}(\psi),\mathcal{H}^{\pm}) which eventually equips 𝒮𝕂(G~)\mathscr{S}_{\mathbb{K}}(\widetilde{G}) with a universal abelian scheme 𝒜\mathscr{A}.666Technically, in [KisinInt] and [CSpin], 𝒜\mathscr{A} is only defined as a sheaf of abelian schemes up to prime-to-pp quasi-isogeny. However, for 𝕂p\mathbb{K}^{p} sufficiently small, we can take 𝒜\mathscr{A} to be an actual abelian scheme (cf. [KisinInt, (2.1.5)]). Let a:𝒜𝒮𝕂(G~)a:\mathscr{A}\to\mathscr{S}_{\mathbb{K}}(\widetilde{G}) be the structural morphism. Define the sheaves 𝐇B:=R1a(p)\mathbf{H}_{B}:=R^{1}a_{\mathbb{C}*}\mathbb{Z}_{(p)}, 𝐇:=R1a¯\mathbf{H}_{\ell}:=R^{1}a_{*}\underline{\mathbb{Q}}_{\ell} (p\ell\neq p), 𝐇p:=R1a¯p\mathbf{H}_{p}:=R^{1}a_{\mathbb{Q}*}\underline{\mathbb{Z}}_{p}, 𝐇dR:=R1aΩ𝒜/𝒮𝕂(G~)\mathbf{H}_{\mathrm{dR}}:=R^{1}a_{*}\Omega^{\bullet}_{\mathscr{A}/\mathscr{S}_{\mathbb{K}}(\widetilde{G})} and 𝐇cris:=R1a¯cris𝒪𝒜𝔽p/p\mathbf{H}_{\mathrm{cris}}:=R^{1}\bar{a}_{\mathrm{cris}*}\mathcal{O}_{\mathscr{A}_{\mathbb{F}_{p}}/\mathbb{Z}_{p}} (a¯:=a𝔽p\bar{a}:=a{\otimes}\mathbb{F}_{p}). The abelian scheme 𝒜\mathscr{A} is equipped with a “CSpin-structure”: a /2\mathbb{Z}/2\mathbb{Z}-grading, Cl(L)\mathrm{Cl}(L)-action and an idempotent projector 𝝅?:End(𝐇?)End(𝐇?)\boldsymbol{\pi}_{?}:\mathrm{End}(\mathbf{H}_{?})\to\mathrm{End}(\mathbf{H}_{?}) for ?=B,,p,dR,cris?=B,\ell,p,\mathrm{dR},\mathrm{cris} on (various applicable fibers of) 𝒮𝕂(G~)\mathscr{S}_{\mathbb{K}}(\widetilde{G}). We use 𝐋?\mathbf{L}_{?} to denote the images of 𝝅?\boldsymbol{\pi}_{?}, and recall the definition of special endomorphisms ([CSpin, Def. 5.2, see also Lem. 5.4, Cor. 5.22]):

Definition (3.4.1).

For any 𝒮𝕂(G~)\mathscr{S}_{\mathbb{K}}(\widetilde{G})-scheme TT, fEnd(𝒜T)f\in\mathrm{End}(\mathscr{A}_{T}) is called a special endomorphism if for some (and hence all) 𝒪T×\ell\in\mathcal{O}_{T}^{\times}, the \ell-adic realization of ff lies in 𝐋|TEnd(𝐇|T)\mathbf{L}_{\ell}|_{T}\subseteq\mathrm{End}(\mathbf{H}_{\ell}|_{T}); if 𝒪T=k\mathcal{O}_{T}=k for a perfect field kk in characteristic pp, then equivalently ff is called a special endomorphism if the crystalline realization of ff lies in 𝐋cris,T\mathbf{L}_{\mathrm{cris},T}. We write the submodule of End(𝒜T)\mathrm{End}(\mathscr{A}_{T}) consisting of special endomorphisms as L(𝒜T)\mathrm{L}(\mathscr{A}_{T}).

(3.4.2)   

It is explained in [CSpin, §5.24] that the sheaves 𝐋?\mathbf{L}_{?} on (applicable fibers of) 𝒮𝕂(G~)\mathscr{S}_{\mathbb{K}}(\widetilde{G}) in fact descend to the corresponding fibers of 𝒮𝖪(G)\mathscr{S}_{\mathsf{K}}(G). We denote the descent of these sheaves by the same letters. It is not hard to see that 𝐇B[1/p]\mathbf{H}_{B}[1/p] together with the restrictions of p𝐇×𝐇p[1/p]\prod_{\ell\neq p}\mathbf{H}_{\ell}\times\mathbf{H}_{p}[1/p] and 𝐇dR\mathbf{H}_{\mathrm{dR}} to Sh𝕂(G~)\mathrm{Sh}_{\mathbb{K}}(\widetilde{G}) is nothing but the automorphic system attached to HRep(G~)H\in\mathrm{Rep}(\widetilde{G}), in our terminology (3.2.4). Similarly, 𝐋B[1/p]\mathbf{L}_{B}[1/p] together with the restrictions of p𝐋×𝐋p[1/p]\prod_{\ell\neq p}\mathbf{L}_{\ell}\times\mathbf{L}_{p}[1/p] and 𝐋dR\mathbf{L}_{\mathrm{dR}} to Sh𝕂(G~)\mathrm{Sh}_{\mathbb{K}}(\widetilde{G}) (or Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G)) is precisely the automorphic system attached to VRep(G~)V\in\mathrm{Rep}(\widetilde{G}) (or Rep(G)\mathrm{Rep}(G)). Readers who wish to check this can look at how the automorphic systems are constructed in [YangSystem, §4.2], which is essentially a generalization of [CSpin, §5.24]. The construction of the 𝐋?\mathbf{L}_{?}-sheaves are also summarized in more detail in [Yang, (3.1.3)]. In particular, there are natural identifications of 𝖵B\mathsf{V}_{B} with 𝐋B[1/p]\mathbf{L}_{B}[1/p], and 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}} (resp. 𝖵dR\mathsf{V}_{\mathrm{dR}}) with the restriction of p𝐋×𝐋p[1/p]\prod_{\ell\neq p}\mathbf{L}_{\ell}\times\mathbf{L}_{p}[1/p] (resp. 𝐋dR\mathbf{L}_{\mathrm{dR}}) to Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G), where (𝖵B,𝖵dR,𝖵e´t)(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}},\mathsf{V}_{\mathrm{{\acute{e}}t}}) was the notation we used in (3.3.7).

(3.4.3)   

Let Sh𝖪p(G)\mathrm{Sh}_{\mathsf{K}_{p}}(G) be the limit lim𝖪pSh𝖪p𝖪p(G)\varprojlim_{\mathsf{K}^{p}}\mathrm{Sh}_{\mathsf{K}_{p}\mathsf{K}^{\prime p}}(G) as 𝖪p\mathsf{K}^{\prime p} runs through the compact open subgroups of G(𝔸pf)G(\mathbb{A}^{p}_{f}) and define 𝒮𝖪p(G)\mathscr{S}_{\mathsf{K}_{p}}(G) similarly. The canonical extension property of 𝒮𝖪p(G)\mathscr{S}_{\mathsf{K}_{p}}(G) is that for any regular, formally smooth (p)\mathbb{Z}_{(p)}-scheme SS, every morphism S𝒮𝖪p(G)S_{\mathbb{Q}}\to\mathscr{S}_{\mathsf{K}_{p}}(G) extends to SS. We give an extension property for finite level, which is certainly well known to experts:

Theorem (3.4.4).

Let SS be a smooth (p)\mathbb{Z}_{(p)}-scheme which admits a morphism ρ:SSh𝖪(G)\rho:S_{\mathbb{Q}}\to\mathrm{Sh}_{\mathsf{K}}(G). If ρ𝖵\rho^{*}\mathsf{V}_{\ell} extends to a local system 𝖶\mathsf{W}_{\ell} over SS for every prime p\ell\neq p, then ρ\rho extends to a morphism S𝒮𝖪(G)S\to\mathscr{S}_{\mathsf{K}}(G), through which 𝖶\mathsf{W}_{\ell} is identified with the pullback of 𝐋\mathbf{L}_{\ell}.

Proof.

To simplify notation, let 𝖵e´t(p)p𝖵\mathsf{V}_{\mathrm{{\acute{e}}t}}^{(p)}\coloneqq\prod_{\ell\neq p}\mathsf{V}_{\ell} and 𝖶e´t(p):=p𝖶\mathsf{W}_{\mathrm{{\acute{e}}t}}^{(p)}:=\prod_{\ell\neq p}\mathsf{W}_{\ell}. Note that 𝖶e´t(p)|S=ρ(𝖵e´t(p))\mathsf{W}_{\mathrm{{\acute{e}}t}}^{(p)}|_{S_{\mathbb{Q}}}=\rho^{*}(\mathsf{V}_{\mathrm{{\acute{e}}t}}^{(p)}). Now the prime-to-pp part of the tautological level structure [ηV][\eta_{V}] gives us a section [ηV(p)]:Sh𝖪(G)𝖪p\Isom¯(V𝔸pf,𝖵e´t(p))[\eta_{V}^{(p)}]:\mathrm{Sh}_{\mathsf{K}}(G)\to\mathsf{K}^{p}\backslash\underline{\mathrm{Isom}}(V{\otimes}\mathbb{A}^{p}_{f},\mathsf{V}_{\mathrm{{\acute{e}}t}}^{(p)}), where the target is viewed as a pro-étale cover of Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G). By [stacks-project, 0BQM], ρ[η(p)V]\rho^{\ast}[\eta^{(p)}_{V}] extends to a morphism S𝖪p\Isom¯(V𝔸fp,𝖶e´t(p))S\to\mathsf{K}^{p}\backslash\underline{\mathrm{Isom}}(V\otimes\mathbb{A}_{f}^{p},\mathsf{W}_{\mathrm{{\acute{e}}t}}^{(p)}), where the target is a pro-étale cover of SS. We define S~\widetilde{S} as the pull-back

S~{{\widetilde{S}}}Isom¯(V𝔸f,𝖶e´t(p)){{\underline{\mathrm{Isom}}(V\otimes\mathbb{A}_{f},\mathsf{W}_{\mathrm{{\acute{e}}t}}^{(p)})}}S{S}𝖪\Isom¯(V𝔸f,𝖶e´t(p)){{\mathsf{K}\backslash\underline{\mathrm{Isom}}(V\otimes\mathbb{A}_{f},\mathsf{W}_{\mathrm{{\acute{e}}t}}^{(p)})}}\scriptstyle{\lrcorner}

Note that since S~S\widetilde{S}\to S is a 𝖪p\mathsf{K}^{p}-torsor for the proétale topology, S~\widetilde{S} is representable by a scheme. As Sh𝖪p(G)Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}_{p}}(G)\to\mathrm{Sh}_{\mathsf{K}}(G) can be defined by an analogous construction, ρ\rho lifts to a 𝖪p\mathsf{K}^{p}-equivariant morphism ρ~:S~Sh𝖪p(G)\widetilde{\rho}\colon\widetilde{S}_{\mathbb{Q}}\to\mathrm{Sh}_{\mathsf{K}_{p}}(G). The canonical extension property allows us to extend it to a (necessarily 𝖪p\mathsf{K}^{p}-equivariant) morphism S~𝒮𝖪p(G)\widetilde{S}\to\mathscr{S}_{\mathsf{K}_{p}}(G). Now the 𝖪p\mathsf{K}^{p}-action defines an étale descend datum which yields the desired morphism S𝒮𝖪(G)S\to\mathscr{S}_{\mathsf{K}}(G). ∎

4 Period Morphisms

4.1   The Basic Set-up

We first state a few basic definitions and results which will be needed to construct the period morphism.

Definition (4.1.1).

Let TT be a connected noetherian normal scheme with geometric point tt, \ell be a prime with 𝒪T×\ell\in\mathcal{O}_{T}^{\times} and 𝖶\mathsf{W}_{\ell} be an étale \mathbb{Q}_{\ell}-local system over TT. We denote by Mon(𝖶,t)\mathrm{Mon}(\mathsf{W}_{\ell},t) the Zariski closure of the image of π1e´t(T,t)\pi_{1}^{\mathrm{{\acute{e}}t}}(T,t) in GL(𝖶,t)\mathrm{GL}(\mathsf{W}_{\ell,t}), and denote by Mon(𝖶,t)\mathrm{Mon}^{\circ}(\mathsf{W}_{\ell},t) the identity component of Mon(𝖶,t)\mathrm{Mon}(\mathsf{W}_{\ell},t).

If 𝖶\mathsf{W}_{\ell}^{\prime} is another \mathbb{Q}_{\ell}-local system, then we say that 𝖶\mathsf{W}_{\ell} and 𝖶\mathsf{W}_{\ell}^{\prime} are étale locally isomorphic if they are isomorphic over some finite connected étale cover TT^{\prime} of TT; or equivalently, there is an isomorphism 𝖶,t𝖶,t\mathsf{W}_{\ell,t}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{W}^{\prime}_{\ell,t} which is equivariant under an open subgroup of πe´t1(T,t)\pi^{\mathrm{{\acute{e}}t}}_{1}(T,t).

Lemma (4.1.2).

Let TT be a noetherian integral normal scheme with generic point η\eta. Let f:𝒴Tf:\mathcal{Y}\to T be a smooth proper morphism.

  1. (a)

    The natural map Pic(𝒴)Pic(𝒴η)\mathrm{Pic}\,(\mathcal{Y})\to\mathrm{Pic}\,(\mathcal{Y}_{\eta}) is surjective with kernel Pic(T)\mathrm{Pic}\,(T).

  2. (b)

    If for some geometric point bb over η\eta and prime 𝒪T×\ell\in\mathcal{O}_{T}^{\times}, Mon(R2f,b)\mathrm{Mon}(R^{2}f_{*}\mathbb{Q}_{\ell},b) is connected, then the natural map NS(𝒴η)NS(𝒴b)\mathrm{NS}(\mathcal{Y}_{\eta})_{\mathbb{Q}}\to\mathrm{NS}(\mathcal{Y}_{b})_{\mathbb{Q}} is an isomorphism.

Proof.

(a) We may always extend a line bundle on 𝒴η\mathcal{Y}_{\eta} to 𝒴U\mathcal{Y}_{U} for some open dense subscheme UTU\subseteq T, and then to a line bundle on 𝒴\mathcal{Y} (use e.g., [Hartshorne, Prop. II.6.5]). But any two extensions to 𝒴\mathcal{Y} differ by an element of Pic(T)\mathrm{Pic}\,(T) by [EGAIV4, ErrIV Cor. 21.4.13].

(b) Let η¯\bar{\eta} be the geometric point over η\eta obtained by taking the separable closure of k(η)k(\eta) in k(b)k(b). Then every class of NS(𝒴b)\mathrm{NS}(\mathcal{Y}_{b})_{\mathbb{Q}} descends to η¯\bar{\eta}. As the natural morphism Gal(η¯/η)=πe´t1(η,b)πe´t1(T,b)\mathrm{Gal}(\bar{\eta}/\eta)=\pi^{\mathrm{{\acute{e}}t}}_{1}(\eta,b)\to\pi^{\mathrm{{\acute{e}}t}}_{1}(T,b) is surjective, and Gal(η¯/η)\mathrm{Gal}(\bar{\eta}/\eta) acts on NS(𝒴η¯)\mathrm{NS}(\mathcal{Y}_{\bar{\eta}})_{\mathbb{Q}} through a finite quotient, the connectedness assumption on Mon(R2f,b)\mathrm{Mon}(R^{2}f_{*}\mathbb{Q}_{\ell},b) implies that Gal(η¯/η)\mathrm{Gal}(\bar{\eta}/\eta) in fact acts trivially. This implies that every class in NS(𝒴η¯)\mathrm{NS}(\mathcal{Y}_{\bar{\eta}})_{\mathbb{Q}} descends to η\eta. ∎

Now we state the set-up we will work with for the entire section.

Set-up (4.1.3).

Let FF\subseteq\mathbb{C} be a subfield finitely generated over \mathbb{Q} and let SS be a connected smooth FF-variety with generic point η\eta. Let f:𝒳Sf\colon\mathcal{X}\to S be a \heartsuit-family with a relatively ample line bundle 𝝃\boldsymbol{\xi}. We fix a subspace ΛNS(𝒳η)\Lambda\subseteq\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}} containing the class of 𝝃η\boldsymbol{\xi}_{\eta}. Now choose an FF-linear embedding k(η)k(\eta)\hookrightarrow\mathbb{C} and let bS()b\in S(\mathbb{C}) be the resulting point, which we also view as a closed point on SS_{\mathbb{C}}. Let SSS^{\circ}\subset S_{\mathbb{C}} denote the connected component containing bb. We assume that for some (and hence every, see (4.1.4) below) prime \ell, Mon(R2f,b)\mathrm{Mon}(R^{2}f_{*}\mathbb{Q}_{\ell},b) is connected.

Define a pairing on H2dR(𝒳/S):=R2fΩ𝒳/S\mathrm{H}^{2}_{\mathrm{dR}}(\mathcal{X}/S):=R^{2}f_{*}\Omega^{\bullet}_{\mathcal{X}/S} by (x,y)xyc1(𝝃)d2(x,y)\mapsto x\cup y\cup c_{1}(\boldsymbol{\xi})^{d-2} for local sections x,yx,y and d=dim𝒳/Sd=\dim\mathcal{X}/S. By (4.1.2), every line bundle on 𝒳η\mathcal{X}_{\eta} extends to a relative line bundle on 𝒳/S\mathcal{X}/S. By taking Chern classes, we obtain a well defined embedding Λ¯H2dR(𝒳/S)\underline{\Lambda}\hookrightarrow\mathrm{H}^{2}_{\mathrm{dR}}(\mathcal{X}/S), where Λ¯\underline{\Lambda} is the constant sheaf with fiber Λ\Lambda. Define 𝖯dR\mathsf{P}_{\mathrm{dR}} to be the orthogonal complement of Λ¯\underline{\Lambda} in H2dR(𝒳/S)(1)\mathrm{H}^{2}_{\mathrm{dR}}(\mathcal{X}/S)(1). We define the primitive Betti cohomology 𝖯B\mathsf{P}_{B} over SS_{\mathbb{C}} and étale cohomology 𝖯e´t\mathsf{P}_{\mathrm{{\acute{e}}t}} over SS analogously. Then 𝖯:=(𝖯B,𝖯dR,𝖯e´t)𝖱(S)\mathsf{P}:=(\mathsf{P}_{B},\mathsf{P}_{\mathrm{dR}},\mathsf{P}_{\mathrm{{\acute{e}}t}})\in\mathsf{R}(S) is a system of realizations over SS in the sense of (3.1.1). Note that we applied a Tate twist so that the VHS 𝖯|S\mathsf{P}|_{S_{\mathbb{C}}} has weight 0. Since we assumed Mon(R2f,b)\mathrm{Mon}(R^{2}f_{*}\mathbb{Q}_{\ell},b) is connected for some \ell, NS(𝒳η)=NS(𝒳b)\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}}=\mathrm{NS}(\mathcal{X}_{b})_{\mathbb{Q}} by (4.1.2). This implies that 𝖯\mathsf{P} orthogonally decomposes into (Λ𝟏S)𝖯0(\Lambda^{\perp}{\otimes}\mathbf{1}_{S})\oplus\mathsf{P}_{0} where Λ\Lambda^{\perp} is the orthogonal complement of Λ\Lambda in NS(𝒳η)\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}} and 𝖯0=(𝖯0,B,𝖯0,dR,𝖯0,e´t)\mathsf{P}_{0}=(\mathsf{P}_{0,B},\mathsf{P}_{0,\mathrm{dR}},\mathsf{P}_{0,\mathrm{{\acute{e}}t}}) is another object in 𝖱(S)\mathsf{R}(S). Moreover, there are no nonzero (0,0)(0,0)-classes in 𝖯0,B,b\mathsf{P}_{0,B,b}. Note that 𝖯0\mathsf{P}_{0} is nothing but 𝖯\mathsf{P} when Λ=NS(𝒳η)\Lambda=\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}}.

Lemma (4.1.4).

In the above set-up, Mon(R2f,b)\mathrm{Mon}(R^{2}f_{*}\mathbb{Q}_{\ell},b) is connected for every prime \ell, and bb is a Hodge-generic point for the VHS 𝖯|S\mathsf{P}|_{S^{\circ}}.

Proof.

The first statement follows from the assumption that FF is finitely generated over \mathbb{Q} and [Larsen-Pink, Prop. 6.14], which implies that the étale group scheme of connected components of Mon(R2f,b)\mathrm{Mon}(R^{2}f_{*}\mathbb{Q}_{\ell},b) is independent of \ell. The second statement follows from the main theorem of [Moonen].777The main theorem of [Moonen] is an overkill for the purpose here and only used to avoid a case by case discussion of Mumford-Tate groups. Since the statement only concerns SS^{\circ}, we may replace SS by SS^{\circ} and FF by its field of definition and thus assume that SS is geometrically connected. Then we may make use of the notion of Galois-generic points ([Moonen-Fom, Def. 4.2.1]). As the Momford-Tate conjecture is known for H2H^{2} of the fibers of 𝒳/S\mathcal{X}/S and bb lies above η\eta, the fact that η\eta is Galois-generic implies that bb is Hodge-generic. ∎

(4.1.5)   

Let VV be a quadratic form over \mathbb{Q} which is isomorphic to 𝖯B,b\mathsf{P}_{B,b} and fix an isometry μb:V𝖯B,b\mu_{b}:V\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{B,b}. Let G:=SO(V)G:=\mathrm{SO}(V). Let V0:=μb1(𝖯0,B,b)V_{0}:=\mu_{b}^{-1}(\mathsf{P}_{0,B,b}). Note that via μb\mu_{b}, the monodromy representation of πe´t1(S,b)\pi^{\mathrm{{\acute{e}}t}}_{1}(S,b) takes values in G(𝔸f)G(\mathbb{A}_{f}). We say that 𝖪\mathsf{K} is admissible if the image of π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b) in GL(𝖯B,b𝔸f)\mathrm{GL}(\mathsf{P}_{B,b}{\otimes}\mathbb{A}_{f}) lies in 𝖪\mathsf{K} via μb\mu_{b}, or equivalently, μb\mu_{b} extends to a (G,V,𝖪)(G,V,\mathsf{K})-level structure [μ][\mu] on 𝖯e´t\mathsf{P}_{\mathrm{{\acute{e}}t}} with [μ]b=𝖪(μb𝔸f)[\mu]_{b}=\mathsf{K}\cdot(\mu_{b}{\otimes}\mathbb{A}_{f}). Define Ω:={v(V)v,v=0,v,v¯>0}\Omega:=\{v\in\mathbb{P}(V{\otimes}\mathbb{C})\mid\langle v,v\rangle=0,\langle v,\bar{v}\rangle>0\} as in (3.3.3). Then (G,Ω)(G,\Omega) is a Shimura datum of abelian type with reflex field \mathbb{Q}. Again let 𝖵=(𝖵B,𝖵dR,𝖵e´t)\mathsf{V}=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}},\mathsf{V}_{\mathrm{{\acute{e}}t}}) be the automorphic system on Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) attached to VRep(G)V\in\mathrm{Rep}(G), and let [ηV][\eta_{V}] be the tautological 𝖪\mathsf{K}-level structure on 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}} (see (3.2.4)).

(4.1.6)   

Suppose that (𝒳/S,𝝃)|S(\mathcal{X}/S,\boldsymbol{\xi})|_{S^{\circ}} belongs to case (R+) or (R2’) described in (2.2.4), i.e., the endomorphism field EE of the Hodge structure on 𝖯0,B,b\mathsf{P}_{0,B,b} is totally real. By (2.2.2), the EE-action on 𝖯0,B,b\mathsf{P}_{0,B,b} is self-adjoint. Let EE act on V0V_{0} through μb\mu_{b}. Recall the notations in (2.1.5). Let 𝒱\mathcal{V} be the EE-bilinear lift of V0V_{0}, i.e., V0=𝒱()V_{0}=\mathcal{V}_{(\mathbb{Q})} and set 𝒢:=ResE/SO(𝒱)\mathcal{G}:=\mathrm{Res}_{E/\mathbb{Q}}\mathrm{SO}(\mathcal{V}), :=𝒢/ZE1\mathcal{H}:=\mathcal{G}/Z_{E}^{1}. In addition, set 𝒱:=𝒱\mathcal{V}^{\sharp}:=\mathcal{V}\oplus\mathcal{E} and 𝒢:=ResE/SO(𝒱)\mathcal{G}^{\sharp}:=\mathrm{Res}_{E/\mathbb{Q}}\mathrm{SO}(\mathcal{V}^{\sharp}). Embed SO(𝒱)\mathrm{SO}(\mathcal{V}) into SO(𝒱)\mathrm{SO}(\mathcal{V}^{\sharp}) by acting trivially on \mathcal{E}. This induces an embedding 𝒢𝒢\mathcal{G}\hookrightarrow\mathcal{G}^{\sharp}.

We say that a neat compact open subgroup 𝖪G(𝔸f)\mathsf{K}\subset G(\mathbb{A}_{f}) satisfies condition ()(\sharp) depending on the particular case (𝒦:=𝖪𝒢(𝔸f)\mathcal{K}:=\mathsf{K}\cap\mathcal{G}(\mathbb{A}_{f}) below):

  • (R1)

    Always.

  • (R2)

    If the image of 𝒦\mathcal{K} in 𝒢(𝔸f)\mathcal{G}^{\sharp}(\mathbb{A}_{f}) lies in some neat compact open subgroup.

  • (R2’)

    If the image of 𝒦\mathcal{K} in (𝔸f)\mathcal{H}(\mathbb{A}_{f}) lies in some neat compact open subgroup.

In case (R2’), we say that for a prime 0\ell_{0}, 𝖪0\mathsf{K}_{\ell_{0}} is sufficiently small if 𝒦0ZE1(0)=1\mathcal{K}_{\ell_{0}}\cap Z_{E}^{1}(\mathbb{Q}_{\ell_{0}})=1. Here 𝖪0\mathsf{K}_{\ell_{0}} is the image of 𝖪\mathsf{K} under the projection G(𝔸f)G(0)G(\mathbb{A}_{f})\to G(\mathbb{Q}_{\ell_{0}}) and 𝒦0\mathcal{K}_{\ell_{0}} is defined similarly.

(4.1.7)   

For any subfield FF^{\prime}\subseteq\mathbb{C} which contains FF and sS(F)s\in S(F^{\prime}), we denote by 𝔭s𝖬𝗈𝗍AH(F)\mathfrak{p}_{s}\in\mathsf{Mot}_{\mathrm{AH}}(F^{\prime}) the submotive of 𝔥2(𝒳s)(1)\mathfrak{h}^{2}(\mathcal{X}_{s})(1) such that ω?(𝔭s)=𝖯?,s\omega_{?}(\mathfrak{p}_{s})=\mathsf{P}_{?,s} for ?=B?=B (when F=F^{\prime}=\mathbb{C}), or dR,e´t\mathrm{dR},\mathrm{{\acute{e}}t}. Note that for any σAut(/F)\sigma\in\mathrm{Aut}(\mathbb{C}/F) and sS()s\in S(\mathbb{C}), there is a natural isomorphism (𝔭s)σ𝔭σ(s)(\mathfrak{p}_{s})^{\sigma}\,{\cong}\,\mathfrak{p}_{\sigma(s)} in 𝖬𝗈𝗍AH()\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C}), which we shall label as σ𝔭,s\sigma_{\mathfrak{p},s}; moreover, the following diagrams tautologically commute (recall the notations in (2.1.3) and (3.1.4)):

ωe´t(𝔭s){{\omega_{\mathrm{{\acute{e}}t}}(\mathfrak{p}_{s})}}𝖯e´t,s{{\mathsf{P}_{\mathrm{{\acute{e}}t},s}}}𝖯dR,s{{\mathsf{P}_{\mathrm{dR},s}}}𝖯dR,s{{\mathsf{P}_{\mathrm{dR},s}}}ωe´t((𝔭s)σ){{\omega_{\mathrm{{\acute{e}}t}}((\mathfrak{p}_{s})^{\sigma})}}ωe´t(𝔭σ(s))=𝖯e´t,σ(s){{\omega_{\mathrm{{\acute{e}}t}}(\mathfrak{p}_{\sigma(s)})=\mathsf{P}_{\mathrm{{\acute{e}}t},\sigma(s)}}}ωdR((𝔭s)σ){{\omega_{\mathrm{dR}}((\mathfrak{p}_{s})^{\sigma})}}ωdR(𝔭σ(s))=𝖯dR,σ(s){{\omega_{\mathrm{dR}}(\mathfrak{p}_{\sigma(s)})=\mathsf{P}_{\mathrm{dR},\sigma(s)}}}bc\scriptstyle{\,{\cong}\,_{\mathrm{bc}}}σ𝔭,s\scriptstyle{\sigma_{\mathfrak{p},s}}σ𝖯e´t,s\scriptstyle{\sigma_{\mathsf{P}_{\mathrm{{\acute{e}}t}},s}}σ𝔭,s\scriptstyle{\sigma_{\mathfrak{p},s}}bc\scriptstyle{\,{\cong}\,_{\mathrm{bc}}}σ𝖯dR,s\scriptstyle{\sigma_{\mathsf{P}_{\mathrm{dR}},s}}

(4.1.8)   

For the rest of section 4, we will put a standing assumption that SS is geometically connected as an FF-variety, so that S=SS_{\mathbb{C}}=S^{\circ}.

Let 𝖪G(𝔸f)\mathsf{K}\subseteq G(\mathbb{A}_{f}) be a neat compact open subgroup and assume that it is also admissible, so that μb\mu_{b} extends to a 𝖪\mathsf{K}-level structure [μ][\mu] on 𝖯\mathsf{P} with [μ]b=𝖪(μb𝔸f)[\mu]_{b}=\mathsf{K}(\mu_{b}{\otimes}\mathbb{A}_{f}). Now (3.1.7) guarantees that the restriction of [μ][\mu] to SS_{\mathbb{C}} is (𝖯|S)(\mathsf{P}|_{S_{\mathbb{C}}})-rational and is of type Ω\Omega. Therefore, (3.2.2) gives us a unique morphism ρ:SSh𝖪(G)\rho_{\mathbb{C}}:S_{\mathbb{C}}\to\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}} such that

(ρ)(𝖵,[ηV])(𝖯,[μ])|S.(\rho_{\mathbb{C}})^{*}(\mathsf{V},[\eta_{V}])\,{\cong}\,(\mathsf{P},[\mu])|_{S_{\mathbb{C}}}. (9)

Note that as 𝖪\mathsf{K} is neat, the above isomorphism is unique; we denote its étale component by αe´t:ρ(𝖵e´t)𝖯e´t|S\alpha^{\circ}_{\mathrm{{\acute{e}}t}}:\rho_{\mathbb{C}}^{*}(\mathsf{V}_{\mathrm{{\acute{e}}t}})\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\mathrm{{\acute{e}}t}}|_{S_{\mathbb{C}}} and let α\alpha^{\circ}_{\ell} be its \ell-adic component for a prime \ell.

Theorem (4.1.9).

In the notations above, assume either

  1. (a)

    𝔭s𝖬𝗈𝗍𝖠𝖻()\mathfrak{p}_{s}\in\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}) for every sSs\in S_{\mathbb{C}} (e.g., when the family (𝒳/S,𝝃)|S(\mathcal{X}/S,\boldsymbol{\xi})|_{S_{\mathbb{C}}} belongs to case (CM)), or

  2. (b)

    the family (𝒳/S,𝝃)|S(\mathcal{X}/S,\boldsymbol{\xi})|_{S_{\mathbb{C}}} belongs to case (R+) = (R1) + (R2) and 𝖪\mathsf{K} satisfies condition ()(\sharp) as defined in (4.1.6).

Then ρ\rho_{\mathbb{C}} descends to a morphism ρ:SSh𝖪(G)F\rho:S\to\mathrm{Sh}_{\mathsf{K}}(G)_{F} over FF. Moreover, αe´t\alpha^{\circ}_{\mathrm{{\acute{e}}t}} descends to an isomorphism αe´t:ρ(𝖵e´t)𝖯e´t\alpha_{\mathrm{{\acute{e}}t}}:\rho^{*}(\mathsf{V}_{\mathrm{{\acute{e}}t}})\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\mathrm{{\acute{e}}t}} over SS.

Case (a) is easy: The diagrams in (4.1.7) readily imply that 𝖯\mathsf{P} over SS is weakly AM, so that the conclusion follows from (3.2.5). In fact, we have that the de Rham component of the isomorphism (9) also descends to SS. Note that case (a) in particular covers the case when (𝒳/S,𝝃)|S(\mathcal{X}/S,\boldsymbol{\xi})|_{S_{\mathbb{C}}} belongs to case (CM) by (2.2.7)(d). The proof of case (b) is the content of §4.2 below. If (𝒳/S,𝝃)|S(\mathcal{X}/S,\boldsymbol{\xi})|_{S_{\mathbb{C}}} belongs to case (R2’) and it is not known that 𝔭s𝖬𝗈𝗍𝖠𝖻()\mathfrak{p}_{s}\in\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}) for every sSs\in S_{\mathbb{C}}, a slightly weaker version of the above theorem holds:

Theorem (4.1.10).

Assume that the family (𝒳/S,𝛏)|S(\mathcal{X}/S,\boldsymbol{\xi})|_{S^{\circ}} belongs to case (R2’), Λ=Λ0\Lambda=\Lambda_{0}, 𝖪\mathsf{K} satisfies condition ()(\sharp), and for a prime 0\ell_{0}, 𝖪0\mathsf{K}_{\ell_{0}} is sufficiently small as defined in (4.1.6).

Then ρ:SSh𝖪(G)\rho_{\mathbb{C}}:S_{\mathbb{C}}\to\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}} descends to a morphism ρ:SFSh𝖪(G)F\rho:S_{F^{\prime}}\to\mathrm{Sh}_{\mathsf{K}}(G)_{F^{\prime}} for a finite extension F/FF^{\prime}/F in \mathbb{C}. Moreover, α0\alpha^{\circ}_{\ell_{0}} descends to an isomorphism α0:ρ𝖵0𝖯0|SF\alpha_{\ell_{0}}:\rho^{*}\mathsf{V}_{\ell_{0}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\ell_{0}}|_{S_{F^{\prime}}}, and ρ𝖵\rho^{*}\mathsf{V}_{\ell} is étale locally isomorphic to 𝖯|SF\mathsf{P}_{\ell}|_{S_{F^{\prime}}} for every other \ell in the sense of (4.1.1).

Remark (4.1.11).

In literature, to define period morphisms to orthogonal Shimura varieties, one usually keeps track of a trivialization of the determinants (cf. [MPTate, Prop. 4.3]). Such a trivialization (i.e., an isometry det(V)¯det(𝖯B)\underline{\det(V)}\stackrel{{\scriptstyle\sim}}{{\to}}\det(\mathsf{P}_{B}) in our notations) is implicit in the statement that (the restriction to SS_{\mathbb{C}} of) [μ][\mu] is (𝖯|S)(\mathsf{P}|_{S_{\mathbb{C}}})-rational as a (G,V,𝖪)(G,V,\mathsf{K})-level structure, because det(V)\det(V) is GG-invariant. More concretely, one obtains this trivialization by globalizing det(μb)\det(\mu_{b}), using that π1(S,b)\pi_{1}(S_{\mathbb{C}},b) fixes det(𝖯B,b)\det(\mathsf{P}_{B,b}) (cf. the proof of (3.1.7)). However, we remark that if SS were not geometrically connected, we would not be able to show that [μ][\mu] is (𝖯|S)(\mathsf{P}|_{S_{\mathbb{C}}})-rational over the connected components of SS_{\mathbb{C}} other than SS^{\circ}, unless we know det(𝖯B,b)\det(\mathsf{P}_{B,b}) is spanned by an absolute Hodge class (e.g., in case (a) of (4.1.9)).

On the other hand, for the purpose of putting level structures, the lack of absolute-Hodgeness of det(𝖯B,b)\det(\mathsf{P}_{B,b}) is partially remedied by independence-of-\ell type results on algebraic monodromy (e.g., [Saitodisc, Lem. 3.2], cf. [Taelman2, Cor. 5.9]) which implies that if for some \ell, det(𝖯B,b)\det(\mathsf{P}_{B,b}){\otimes}\mathbb{Q}_{\ell} is π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b)-invariant, then the same is true for all \ell. In (4.1.4) we used Larsen-Pink’ result [Larsen-Pink, Prop. 6.14] to achieve a similar effect. Later in (4.2.4) this is used to overcome a similar difficulty: We do not know that the tensors which cut out ResE/SO(𝒱)\mathrm{Res}_{E/\mathbb{Q}}\mathrm{SO}(\mathcal{V}) from ResE/O(𝒱)\mathrm{Res}_{E/\mathbb{Q}}\mathrm{O}(\mathcal{V}) are given by absolute-Hodge tensors on 𝖯B,b\mathsf{P}_{B,b} via μb\mu_{b}. As far as we are aware of, this cannot be deduced from (2.2.7). However, we can put 𝒦\mathcal{K}-level structures on 𝖯e´t\mathsf{P}_{\mathrm{{\acute{e}}t}} in question and proceed.

4.2   Case (R+): Maximal Monodromy

Lemma (4.2.1).

It suffices to prove (4.1.9) when Λ=NS(𝒳η)\Lambda=\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}}.

Proof.

Let Λ\Lambda^{\perp} be the orthogonal complement of Λ\Lambda in Λ0:=NS(𝒳η)\Lambda_{0}:=\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}} and set M=μb1(Λ)M=\mu_{b}^{-1}(\Lambda^{\perp}). Then V=V0MV=V_{0}\oplus M and we view G0G_{0} as the stabilizer of MM in GG. The image of π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b) in G(𝔸f)G(\mathbb{A}_{f}) via μb\mu_{b} actually lies in G0(𝔸f)G_{0}(\mathbb{A}_{f}). Therefore, 𝖪0=𝖪G0(𝔸f)\mathsf{K}_{0}=\mathsf{K}\cap G_{0}(\mathbb{A}_{f}) satisfies condition ()(\sharp) for Λ0\Lambda_{0}.

Let 𝖵0\mathsf{V}_{0} be the automorphic system on Sh0:=Sh𝖪0(G0)\mathrm{Sh}_{0}:=\mathrm{Sh}_{\mathsf{K}_{0}}(G_{0}) given by V0Rep(G0)V_{0}\in\mathrm{Rep}(G_{0}), [ηV0][\eta_{V_{0}}] be the 𝖪0\mathsf{K}_{0}-level structure on 𝖵0\mathsf{V}_{0}, and [μ0][\mu_{0}] be the 𝖪0\mathsf{K}_{0}-level structure on 𝖯0\mathsf{P}_{0} defined by μ0,b:=μb|V0\mu_{0,b}:=\mu_{b}|_{V_{0}}. As in the paragraph above (4.1.9), we obtain a morphism ρ0,:SSh0,\rho_{0,\mathbb{C}}:S_{\mathbb{C}}\to\mathrm{Sh}_{0,\mathbb{C}} such that (𝖵0,[ηV0])|S(𝖯0,[μ0])|S(\mathsf{V}_{0},[\eta_{V_{0}}])|_{S_{\mathbb{C}}}\,{\cong}\,(\mathsf{P}_{0},[\mu_{0}])|_{S_{\mathbb{C}}}. Define α0,e´t\alpha^{\circ}_{0,\mathrm{{\acute{e}}t}} accordingly.

Consider the Shimura morphism j:Sh0Sh:=Sh𝖪(G)j:\mathrm{Sh}_{0}\to\mathrm{Sh}:=\mathrm{Sh}_{\mathsf{K}}(G). Then we have natural identifications

j(𝖵)=𝖵0(M𝟏Sh0), and 𝖯=𝖯0(Λ𝟏S)j^{*}(\mathsf{V})=\mathsf{V}_{0}\oplus(M{\otimes}\mathbf{1}_{\mathrm{Sh}_{0}}),\text{ and }\mathsf{P}=\mathsf{P}_{0}\oplus(\Lambda^{\perp}{\otimes}\mathbf{1}_{S}) (10)

Moreover, the level structure [ηV0][\eta_{V_{0}}] (resp. [μ0][\mu_{0}]) refines j([ηV])j^{*}([\eta_{V}]) (resp. [μ][\mu]) in the sense of (3.1.2)(d). Therefore, one easily checks that

(jρ0,)(𝖵,[ηV])(𝖯,[μ])|S.(j_{\mathbb{C}}\circ\rho_{0,\mathbb{C}})^{*}(\mathsf{V},[\eta_{V}])\,{\cong}\,(\mathsf{P},[\mu])|_{S_{\mathbb{C}}}.

By the uniqueness statement in (3.2.2), this implies that ρ=jρ0,\rho_{\mathbb{C}}=j_{\mathbb{C}}\circ\rho_{0,\mathbb{C}}.

Assume now that ρ0,\rho_{0,\mathbb{C}} descends to ρ0\rho_{0} over FF, and α0,e´t\alpha^{\circ}_{0,\mathrm{{\acute{e}}t}} descends to α0,e´t\alpha_{0,\mathrm{{\acute{e}}t}} over SS. Then ρ\rho_{\mathbb{C}} descends to jFρ0j_{F}\circ\rho_{0}, and αe´t\alpha^{\circ}_{\mathrm{{\acute{e}}t}} descends to an isomorphism α0,e´t(idΛ𝔸¯f)\alpha_{0,\mathrm{{\acute{e}}t}}\oplus(\mathrm{id}_{\Lambda^{\perp}}{\otimes}\underline{\mathbb{A}}_{f}) over SS. ∎

(4.2.2)   

In this section we prove (4.1.9). By (4.2.1), we may assume that Λ=Λ0\Lambda=\Lambda_{0}, so that V=V0V=V_{0} and 𝖯=𝖯0\mathsf{P}=\mathsf{P}_{0}. Let E,𝒢,𝒱,𝒢,𝖪,𝒦E,\mathcal{G},\mathcal{V},\mathcal{G}^{\sharp},\mathsf{K},\mathcal{K} be as introduced in (4.1.6). In particular, EE is the endomorphism field of the Hodge structure 𝖯b\mathsf{P}_{b}, and 𝒱\mathcal{V} is the EE-bilinear lift of VV, which carries an EE-action via the isometry μb:V𝖯B,b\mu_{b}:V\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{B,b} fixed in (4.1.5).

Note that by (2.2.7) each eEe\in E, viewed as an element of End(𝖯B,b)=End(ωB(𝔭b))\mathrm{End}(\mathsf{P}_{B,b})=\mathrm{End}(\omega_{B}(\mathfrak{p}_{b})), is absolute Hodge, so its image in End(𝖯e´t,b)\mathrm{End}(\mathsf{P}_{\mathrm{{\acute{e}}t},b}) is stabilized by an open subgroup of π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b). Since we assumed that Mon(𝖯,b)\mathrm{Mon}(\mathsf{P}_{\ell},b) is connected for every \ell, the EE-action on 𝖯e´t,b\mathsf{P}_{\mathrm{{\acute{e}}t},b} must already be π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b)-equivariant. This has the following consequence:

Lemma (4.2.3).

The action of EE on 𝖯B,b\mathsf{P}_{B,b} extends to an action on 𝖯\mathsf{P}. If τ:E\tau:E\hookrightarrow\mathbb{C} is the embedding induced by the action of EE on Fil1𝖯dR,b\mathrm{Fil}^{1}\mathsf{P}_{\mathrm{dR},b}, or equivalently the unique indefinite real place of 𝒱\mathcal{V}, then τ(E)F\tau(E)\subseteq F.

Proof.

For the first statement, it is clear that the EE-action on 𝖯B,b\mathsf{P}_{B,b} (resp. 𝖯e´t,b\mathsf{P}_{\mathrm{{\acute{e}}t},b}) extends (necessarily uniquely) to 𝖯B\mathsf{P}_{B} and (resp. 𝖯e´t\mathsf{P}_{\mathrm{{\acute{e}}t}}) because it is π1(S,b)\pi_{1}(S_{\mathbb{C}},b) (resp. π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b))-equivariant. It remains to show that the EE-action on 𝖯dR|S\mathsf{P}_{\mathrm{dR}}|_{S_{\mathbb{C}}}, obtained via the Riemann-Hilbert correspondence, descends to an action on 𝖯dR\mathsf{P}_{\mathrm{dR}}. This follows from the fact that the de Rham realization of every eEe\in E, as an element of 𝖯dR,b\mathsf{P}_{\mathrm{dR},b}^{\otimes}, descends to 𝖯dR,η\mathsf{P}_{\mathrm{dR},\eta}^{\otimes}: Since ee is absolute Hodge and its étale component is πe´t1(η,b)\pi^{\mathrm{{\acute{e}}t}}_{1}(\eta,b)-invariant, its de Rham component descends to η\eta (cf. the argument for [KisinInt, (2.2.2)]).

We remind the reader that by S()S(\mathbb{C}) we mean the set of FF-linear morphisms Spec()S\mathrm{Spec\,}(\mathbb{C})\to S. The first statement implies that for every sS()s\in S(\mathbb{C}), 𝖯B,s\mathsf{P}_{B,s} carries an action of EE, which is self-adjoint by (2.2.2); moreover, for every σAut(/F)\sigma\in\mathrm{Aut}(\mathbb{C}/F), the σ\sigma-linear isomorphism σ𝖯dR,s:𝖯dR,s𝖯dR,σ(s)\sigma_{\mathsf{P}_{\mathrm{dR}},s}:\mathsf{P}_{\mathrm{dR},s}\,{\cong}\,\mathsf{P}_{\mathrm{dR},\sigma(s)} of filtered vector spaces is EE-equivariant (see (3.1.4) for this notation). Therefore, if we let τs:E\tau_{s}:E\to\mathbb{R} be the place through which EE acts on Fil1𝖯dR,s\mathrm{Fil}^{1}\mathsf{P}_{\mathrm{dR},s}, then τσ(s)=στs\tau_{\sigma(s)}=\sigma\circ\tau_{s}. Now we use that τs\tau_{s} can also be characterized as the unique real place of EE such that 𝖯B,sτs\mathsf{P}_{B,s}{\otimes}_{\tau_{s}}\mathbb{R} is indefinite. Parallel transport implies that τs\tau_{s} is constant on S()S(\mathbb{C}), which by assumption is connected. As τ=τb\tau=\tau_{b} and σ(b)S()\sigma(b)\in S(\mathbb{C}), στ=τ\sigma\circ\tau=\tau for every σAut(/F)\sigma\in\mathrm{Aut}(\mathbb{C}/F). This implies that τ(E)F\tau(E)\subseteq F. ∎

(4.2.4)   

Below we shall view EE as a subfield of FF (and hence of \mathbb{C}) via τ\tau as above and drop τ\tau from the notation. Now we recall the discussion in (3.3.7): Let Ω𝒱Ω\Omega_{\mathcal{V}}\subseteq\Omega be the Hermitian symmetric subdomain {w(𝒱E)w,w¯>0,w,w=0}\{w\in\mathbb{P}(\mathcal{V}{\otimes}_{E}\mathbb{C})\mid\langle w,\bar{w}\rangle>0,\langle w,w\rangle=0\}. Then (𝒢,Ω𝒱)(\mathcal{G},\Omega_{\mathcal{V}}) is a Shimura subdatum of (G,Ω)(G,\Omega) with reflex field EE. Therefore, there is a Shimura morphism i:Sh𝒦(𝒢)Sh𝖪(G)i_{\mathbb{C}}:\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})_{\mathbb{C}}\to\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}} which descends to i:Sh𝒦(𝒢)Sh𝖪(G)Ei:\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})\to\mathrm{Sh}_{\mathsf{K}}(G)_{E} over EE. Let 𝖵~\widetilde{\mathsf{V}} be the automorphic system on Sh𝒦(𝒢)\mathrm{Sh}_{\mathcal{K}}(\mathcal{G}) defined by VV and let [η𝒱][\eta_{\mathcal{V}}] be its tautological 𝒦\mathcal{K}-structure. Recall that 𝖵~\widetilde{\mathsf{V}} is identified with i𝖵i^{*}\mathsf{V}, and [η𝒱][\eta_{\mathcal{V}}] refines i([ηV])i^{*}([\eta_{V}]).

Consider the situation in (4.1.8). Recall that we assumed that Mon(𝖯,b)\mathrm{Mon}(\mathsf{P}_{\ell},b) is connected for every \ell. Since the centralizer of the EE-action in O(V)\mathrm{O}(V) can be identified with ResE/O(𝒱)\mathrm{Res}_{E/\mathbb{Q}}\mathrm{O}(\mathcal{V}), which contains 𝒢\mathcal{G} as the identity component, the monodromy action of π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b) must take values in 𝒦:=𝖪𝒢(𝔸f)\mathcal{K}:=\mathsf{K}\cap\mathcal{G}(\mathbb{A}_{f}) via μb\mu_{b}. Therefore, there exists a (𝒢,V,𝒦)(\mathcal{G},V,\mathcal{K})-level structure [μ~][\widetilde{\mu}] on 𝖯e´t\mathsf{P}_{\mathrm{{\acute{e}}t}} such that [μ~]b=𝒦(μb𝔸f)[\widetilde{\mu}]_{b}=\mathcal{K}\cdot(\mu_{b}{\otimes}\mathbb{A}_{f}). As the Hodge structure on 𝖯B,b\mathsf{P}_{B,b} is defined by a point on Ω𝒱\Omega_{\mathcal{V}} via μb\mu_{b}, by (3.1.7) the restriction of [μ~][\widetilde{\mu}] to SS_{\mathbb{C}} is (𝖯|S)(\mathsf{P}|_{S_{\mathbb{C}}})-rational and of type Ω𝒱\Omega_{\mathcal{V}}. One easily checks that:

Lemma (4.2.5).

Let ϱ:SSh𝒦(𝒢)\varrho_{\mathbb{C}}:S_{\mathbb{C}}\to\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})_{\mathbb{C}} be the unique morphism such that

(ϱ)(𝖵~,[η𝒱])(𝖯,[μ~])|S(\varrho_{\mathbb{C}})^{*}({\widetilde{\mathsf{V}}},[\eta_{\mathcal{V}}])\,{\cong}\,(\mathsf{P},[\widetilde{\mu}])|_{S_{\mathbb{C}}} (11)

given by (3.2.2). Then ρ=iϱ\rho_{\mathbb{C}}=i_{\mathbb{C}}\circ\varrho_{\mathbb{C}}.

Hence we reduce (4.1.9) to:

Theorem (4.2.6).

Under the hypothesis of (4.1.9) and notations above, the morphism ϱ\varrho_{\mathbb{C}} descends to an FF-morphism ϱ:SSh𝒦(𝒢)F\varrho:S\to\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})_{F}; moreover, the étale component αe´t:ϱ𝖵~e´t𝖯e´t|S\alpha^{\circ}_{\mathrm{{\acute{e}}t}}:\varrho_{\mathbb{C}}^{*}\widetilde{\mathsf{V}}_{\mathrm{{\acute{e}}t}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\mathrm{{\acute{e}}t}}|_{S_{\mathbb{C}}} of (11) descends to an isomorphism of αe´t:ϱ𝖵~e´t𝖯e´t\alpha_{\mathrm{{\acute{e}}t}}:\varrho^{*}{\widetilde{\mathsf{V}}}_{\mathrm{{\acute{e}}t}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\mathrm{{\acute{e}}t}} over SS.

Note that αe´t\alpha_{\mathrm{{\acute{e}}t}}^{\circ} above agrees with the one in (4.1.9) because 𝖵~e´t=i𝖵e´t{\widetilde{\mathsf{V}}}_{\mathrm{{\acute{e}}t}}=i^{*}\mathsf{V}_{\mathrm{{\acute{e}}t}}.

Below for any \mathbb{Q}-linear Tannakian category 𝒞\mathscr{C}, we write 𝔑()\mathfrak{N}(-) for the functor 𝒞(E)𝒞\mathscr{C}_{(E)}\to\mathscr{C} which sends every M𝒞(E)M\in\mathscr{C}_{(E)} to NmE/(M)det(M())\mathrm{Nm}_{E/\mathbb{Q}}(M){\otimes}_{\mathbb{Q}}\det(M_{(\mathbb{Q})}) (cf. (2.1.4)). Recall that in (3.3.5) we explained how to apply 𝔑()\mathfrak{N}(-) to a 𝒦\mathcal{K}-level structure on a system of realizations.

(4.2.7)   

We first treat the case when m:=dimE𝒱m:=\dim_{E}\mathcal{V} is odd. In this case, N:=𝔑(𝒱)𝖱𝖾𝗉(𝒢)N:=\mathfrak{N}(\mathcal{V})\in\mathsf{Rep}(\mathcal{G}) is faithful. Let 𝖭\mathsf{N} be the automorphic system on Sh𝖪(𝒢)\mathrm{Sh}_{\mathsf{K}}(\mathcal{G}) associated to NN. Let [ηN][\eta_{N}] be the tautological 𝒦\mathcal{K}-level structure on 𝖭\mathsf{N}. Then by (3.3.7), there is a unique isomorphism

(𝔑(𝖵~),𝔑(η𝒱))(𝖭,[ηN]).(\mathfrak{N}({\widetilde{\mathsf{V}}}),\mathfrak{N}(\eta_{\mathcal{V}}))\,{\cong}\,(\mathsf{N},[\eta_{N}]). (12)

As remarked in (3.3.4), one should read (3.3.7) with (,𝒞,Ω)=(𝒢,𝒦,Ω𝒱)(\mathcal{H},\mathcal{C},\Omega_{\mathcal{H}})=(\mathcal{G},\mathcal{K},\Omega_{\mathcal{V}}).

Lemma (4.2.8).

𝔑(𝖯)𝖱(S)\mathfrak{N}(\mathsf{P})\in\mathsf{R}(S) is weakly AM in the sense of (3.1.5).

Proof.

As the formation of 𝔑()\mathfrak{N}(-) on 𝖬𝗈𝗍AH()(E)\mathsf{Mot}_{\mathrm{AH}}(\mathbb{C})_{(E)} commutes with cohomological realizations, for every sS()s\in S(\mathbb{C}), 𝔑((𝖯B,s,𝖯dR,s))=ωHdg(𝔑(𝔭s))\mathfrak{N}((\mathsf{P}_{B,s},\mathsf{P}_{\mathrm{dR},s}))=\omega_{\mathrm{Hdg}}(\mathfrak{N}(\mathfrak{p}_{s})). By (2.2.7)(a), 𝔑(𝔭s)𝖬𝗈𝗍𝖠𝖻()\mathfrak{N}(\mathfrak{p}_{s})\in\mathsf{Mot}_{\mathsf{Ab}}(\mathbb{C}). Therefore, in (3.1.5) we may take M=𝔑(𝔭s)M=\mathfrak{N}(\mathfrak{p}_{s}), so that Mσ=(𝔑(𝔭s))σ=𝔑((𝔭s)σ)M^{\sigma}=(\mathfrak{N}(\mathfrak{p}_{s}))^{\sigma}=\mathfrak{N}((\mathfrak{p}_{s})^{\sigma}). By appling 𝔑()\mathfrak{N}(-) to the objects in the diagram (4.1.7), one checks that the diagrams in (3.1.5) commute for 𝔑(𝖯)\mathfrak{N}(\mathsf{P}). ∎

Lemma (4.2.9).

There exists a unique morphism ρN:SSh𝒦(𝒢)F\rho_{N}:S\to\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})_{F} such that

ρN(𝖭,[ηN])(𝔑(𝖯),𝔑([μ~])).\rho_{N}^{*}(\mathsf{N},[\eta_{N}])\,{\cong}\,(\mathfrak{N}(\mathsf{P}),\mathfrak{N}([\widetilde{\mu}])). (13)

Moreover, ϱ=ρN|S\varrho_{\mathbb{C}}=\rho_{N}|_{S_{\mathbb{C}}}.

Proof.

One easily checks from (3.3.5) that 𝔑([μ~])=𝒦(𝔑(μb)𝔸f)\mathfrak{N}([\widetilde{\mu}])=\mathcal{K}\cdot(\mathfrak{N}(\mu_{b}){\otimes}\mathbb{A}_{f}), so by (3.1.7) 𝔑([μ~])\mathfrak{N}([\widetilde{\mu}]) is 𝔑(𝖯)\mathfrak{N}(\mathsf{P})-rational. Moreover, as [μ~][\widetilde{\mu}] is of type Ω𝒱\Omega_{\mathcal{V}}, so is 𝔑([μ~])\mathfrak{N}([\widetilde{\mu}]). As 𝔑(𝖯)\mathfrak{N}(\mathsf{P}) is weakly AM, (3.2.5) gives the ρN\rho_{N} for which (13) holds. On the other hand, by appling 𝔑()\mathfrak{N}(-) to (11), we see that

(ϱ)(𝖭,[η𝔑])(𝔑(𝖯),𝔑([μ~]))|S.(\varrho_{\mathbb{C}})^{*}(\mathsf{N},[\eta_{\mathfrak{N}}])\,{\cong}\,(\mathfrak{N}(\mathsf{P}),\mathfrak{N}([\widetilde{\mu}]))|_{S_{\mathbb{C}}}. (14)

By the uniquness statement in (3.2.2), this implies that ϱ=ρN|S\varrho_{\mathbb{C}}=\rho_{N}|_{S_{\mathbb{C}}}. ∎

Proof of (4.2.6) for mm odd: To affirm the first statement, set ϱ:=ρN\varrho:=\rho_{N}. The second statement now follows from (3.3.6). Indeed, comparing (12), (13) and (14), we see that 𝔑(αe´t)\mathfrak{N}(\alpha^{\circ}_{\mathrm{{\acute{e}}t}}) descends to SS. However, as dimE𝒱\dim_{E}\mathcal{V} is odd, Z1E(𝒱)=1Z^{1}_{E}(\mathcal{V})=1 (see (2.1.5)). By (3.3.6), αe´t\alpha^{\circ}_{\mathrm{{\acute{e}}t}} descends to SS. ∎

(4.2.10)   

Recall that in (4.1.6) 𝒱=𝒱\mathcal{V}^{\sharp}=\mathcal{V}\oplus\mathcal{E} for \mathcal{E} defined in (2.1.5). If m=dimE𝒱m=\dim_{E}\mathcal{V} is even, we let 𝒱\mathcal{V}^{\sharp} play the role of 𝒱\mathcal{V} in the above proof. Recall that in (4.1.9) we assumed that 𝖪\mathsf{K} satisfies condition ()(\sharp) and we are currently in situation V=V0V=V_{0}, so that 𝒦=𝖪𝒢(𝔸f)𝒰\mathcal{K}=\mathsf{K}\cap\mathcal{G}(\mathbb{A}_{f})\subseteq\mathcal{U} for some neat compact open 𝒰𝒢(𝔸f)\mathcal{U}\subseteq\mathcal{G}^{\sharp}(\mathbb{A}_{f}). Define the Hermitian symmetric domain Ω𝒱\Omega_{\mathcal{V}^{\sharp}} with 𝒱\mathcal{V} replaced by 𝒱\mathcal{V}^{\sharp} in Ω𝒱\Omega_{\mathcal{V}}. Then we obtain an embedding of Shimura data (𝒢,Ω𝒱)(𝒢,Ω𝒱)(\mathcal{G},\Omega_{\mathcal{V}})\hookrightarrow(\mathcal{G}^{\sharp},\Omega_{\mathcal{V}^{\sharp}}). By [DeligneTdShimura, (1.15)], for some 𝒰𝒦\mathcal{U}^{\prime}\supseteq\mathcal{K}, the Shimura morphism Sh𝒦(𝒢)Sh𝒰(𝒢)\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})\to\mathrm{Sh}_{\mathcal{U}^{\prime}}(\mathcal{G}^{\sharp}) is an embedding. Replacing 𝒰\mathcal{U} by 𝒰𝒰\mathcal{U}\cap\mathcal{U}^{\prime} if necessary, we may assume that the Shimura morphism Sh𝒦(𝒢)Sh𝒰(𝒢)\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})\to\mathrm{Sh}_{\mathcal{U}}(\mathcal{G}^{\sharp}) is also an embedding. Below we write this embedding simply as j:Sh𝒦Sh𝒰j:\mathrm{Sh}_{\mathcal{K}}\to\mathrm{Sh}_{\mathcal{U}}. Let 𝖶\mathsf{W} be the automorphic system of realizations on Sh𝒰\mathrm{Sh}_{\mathcal{U}} given by W:=(𝒱)()𝖱𝖾𝗉(𝒢)W:=(\mathcal{V}^{\sharp})_{(\mathbb{Q})}\in\mathsf{Rep}(\mathcal{G}^{\sharp}) and let [ηW][\eta_{W}] be the tautological 𝒰\mathcal{U}-level structure on 𝖶\mathsf{W}.

The reader should now apply the discussion in (3.3.7) with (𝒢,Ω𝒱)(\mathcal{G},\Omega_{\mathcal{V}}) replaced by (𝒢,ΩV)(\mathcal{G}^{\sharp},\Omega_{V^{\sharp}}), 𝖵~\widetilde{\mathsf{V}} replaced by 𝖶\mathsf{W}, (,𝒞,Ω)=(𝒢,𝒰,Ω𝒱)(\mathcal{H},\mathcal{C},\Omega_{\mathcal{H}})=(\mathcal{G}^{\sharp},\mathcal{U},\Omega_{\mathcal{V}^{\sharp}}) and π=id\pi=\mathrm{id} (cf. (3.3.4)). In particular, 𝖶\mathsf{W} is equipped with a natural EE-action. This time we set N=𝔑(𝒱)Rep(𝒢)N=\mathfrak{N}(\mathcal{V}^{\sharp})\in\mathrm{Rep}(\mathcal{G}^{\sharp}). Let 𝖭\mathsf{N} be the automorphic system on Sh𝒰\mathrm{Sh}_{\mathcal{U}} given by NN and let [ηN][\eta_{N}] be the tautological 𝒰\mathcal{U}-level structure. Note that NRep(𝒢)N\in\mathrm{Rep}(\mathcal{G}^{\sharp}) is faithful as dimE𝒱\dim_{E}\mathcal{V}^{\sharp} is odd. Now (3.3.7) tells us that

(𝖭,[ηN])=(𝔑(𝖶),𝔑([ηW]))(\mathsf{N},[\eta_{N}])=(\mathfrak{N}(\mathsf{W}),\mathfrak{N}([\eta_{W}])) (15)

It is not hard to see that the restriction of 𝖶\mathsf{W} to Sh𝒦\mathrm{Sh}_{\mathcal{K}} is naturally identified with 𝖵~(𝟏Sh𝒦){\widetilde{\mathsf{V}}}\oplus(\mathcal{E}{\otimes}\mathbf{1}_{\mathrm{Sh}_{\mathcal{K}}}). Correspondingly, we set 𝖰:=𝖯(𝟏S)\mathsf{Q}:=\mathsf{P}\oplus(\mathcal{E}{\otimes}\mathbf{1}_{S}) and define νb:W𝖰B,b\nu_{b}:W\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{Q}_{B,b} by μbid\mu_{b}\oplus\mathrm{id}_{\mathcal{E}}. Then νb\nu_{b} defines a 𝒰\mathcal{U}-level structure [ν][\nu] with [ν]b=𝒰(νb𝔸f)[\nu]_{b}=\mathcal{U}\cdot(\nu_{b}{\otimes}\mathbb{A}_{f}). Define β:=jϱ\beta_{\mathbb{C}}:=j_{\mathbb{C}}\circ\varrho_{\mathbb{C}}. Then we have:

(β)(𝖶,[ηW])(𝖰,[ν])|S(\beta_{\mathbb{C}})^{*}(\mathsf{W},[\eta_{W}])\,{\cong}\,(\mathsf{Q},[\nu])|_{S_{\mathbb{C}}} (16)
Lemma (4.2.11).

There exists a unique morphism βN:S(Sh𝒰)F\beta_{N}:S\to(\mathrm{Sh}_{\mathcal{U}})_{F} such that

βN(𝖭,[ηN])(𝔑(𝖰),𝔑([ν])).\beta_{N}^{*}(\mathsf{N},[\eta_{N}])\,{\cong}\,(\mathfrak{N}(\mathsf{Q}),\mathfrak{N}([\nu])). (17)

Moreover, β=βN|S\beta_{\mathbb{C}}=\beta_{N}|_{S_{\mathbb{C}}}.

Proof.

By (2.2.7)(b) and a slight variant of the argument for (4.2.8), 𝔑(𝖰)\mathfrak{N}(\mathsf{Q}) is weakly AM. As 𝔑([ν])b=𝒰(𝔑(μb)𝔸f)\mathfrak{N}([\nu])_{b}=\mathcal{U}\cdot(\mathfrak{N}(\mu_{b}){\otimes}\mathbb{A}_{f}), (3.1.7) says that the (𝒢,N,𝒰)(\mathcal{G}^{\sharp},N,\mathcal{U})-level structure 𝔑([ν])\mathfrak{N}([\nu]) is 𝔑(𝖰)\mathfrak{N}(\mathsf{Q})-rational. As [μ~][\widetilde{\mu}] is of type Ω𝒱\Omega_{\mathcal{V}}, [ν][\nu] is of type Ω𝒱\Omega_{\mathcal{V}^{\sharp}}, so that 𝔑([ν])\mathfrak{N}([\nu]) is also of type Ω𝒱\Omega_{\mathcal{V}^{\sharp}}. Applying (3.2.5) to the faithful representation NN of 𝒢\mathcal{G}^{\sharp}, we obtain the desired map βN\beta_{N} such that (17) holds. By the uniqueness statement in (3.2.2), to show β=βN|S\beta_{\mathbb{C}}=\beta_{N}|_{S_{\mathbb{C}}} it suffices to observe that

β(𝔑(𝖶),𝔑([ηW]))(𝔑(𝖰),𝔑([ν]))|S.\beta_{\mathbb{C}}^{*}(\mathfrak{N}(\mathsf{W}),\mathfrak{N}([\eta_{W}]))\,{\cong}\,(\mathfrak{N}(\mathsf{Q}),\mathfrak{N}([\nu]))|_{S_{\mathbb{C}}}.

One checks this by applying 𝔑()\mathfrak{N}(-) to (16). ∎

Proof of (4.2.6) for mm even: The above implies that β\beta_{\mathbb{C}} descends to βN\beta_{N} over FF. Recall that β=jϱ\beta_{\mathbb{C}}=j_{\mathbb{C}}\circ\varrho_{\mathbb{C}} and jj_{\mathbb{C}} is an embedding. As the actions on \mathbb{C}-points of both β\beta_{\mathbb{C}} and jj_{\mathbb{C}} are Aut(/F)\mathrm{Aut}(\mathbb{C}/F)-equivariant, the same is true for ϱ\varrho_{\mathbb{C}}, so that ϱ\varrho_{\mathbb{C}} descends to a morphism ϱ\varrho over FF with βN=jFϱ\beta_{N}=j_{F}\circ\varrho.

Let λe´t:(𝖶e´t,[ηW])|S(𝖰e´t,[ν])|S\lambda^{\circ}_{\mathrm{{\acute{e}}t}}:(\mathsf{W}_{\mathrm{{\acute{e}}t}},[\eta_{W}])|_{S_{\mathbb{C}}}\,{\cong}\,(\mathsf{Q}_{\mathrm{{\acute{e}}t}},[\nu])|_{S_{\mathbb{C}}} be the étale component of (16). Then 𝔑(λe´t)\mathfrak{N}(\lambda^{\circ}_{\mathrm{{\acute{e}}t}}) is the étale component of (17) restricted to SS_{\mathbb{C}}. Therefore, 𝔑(λe´t)\mathfrak{N}(\lambda^{\circ}_{\mathrm{{\acute{e}}t}}) descends to SS. Applying (3.3.6) to 𝒱\mathcal{V}^{\sharp}, we have that λe´t\lambda^{\circ}_{\mathrm{{\acute{e}}t}} descends to an isomorphism λe´t:βN𝖶e´t𝖰e´t\lambda_{\mathrm{{\acute{e}}t}}:\beta_{N}^{*}\mathsf{W}_{\mathrm{{\acute{e}}t}}\,{\cong}\,\mathsf{Q}_{\mathrm{{\acute{e}}t}} over SS. Note the decompositions βN𝖶=ϱ𝖵~(𝟏S)\beta_{N}^{*}\mathsf{W}=\varrho^{*}{\widetilde{\mathsf{V}}}\oplus(\mathcal{E}{\otimes}\mathbf{1}_{S}) and 𝖰=𝖯(𝟏S)\mathsf{Q}=\mathsf{P}\oplus(\mathcal{E}{\otimes}\mathbf{1}_{S}). Since λe´t\lambda_{\mathrm{{\acute{e}}t}}^{\circ} respects these decompositions over SS_{\mathbb{C}} by construction, its descent λe´t\lambda_{\mathrm{{\acute{e}}t}} over SS must also respect these decompositions, which are defined over SS. Hence λe´t\lambda_{\mathrm{{\acute{e}}t}} restricts to the sought after αe´t:ϱ𝖵~e´t𝖯e´t\alpha_{\mathrm{{\acute{e}}t}}:\varrho^{*}{\widetilde{\mathsf{V}}}_{\mathrm{{\acute{e}}t}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\mathrm{{\acute{e}}t}} in (4.2.6). ∎

4.3   Case (R2’): non-maximal monodromy

By (2.2.5), we expect this case to be rare in practice. Readers who are not particularly interested in this case might skip to the next section.

Lemma (4.3.1).

Let kk\subseteq\mathbb{C} be a subfield and A,BA,B be kk-varieties with AA being geometrically connected. Suppose that there is a morphism f:ABf:A_{\mathbb{C}}\to B_{\mathbb{C}} over \mathbb{C}, and an étale morphism g:BCg:B\to C over kk such that for some h:ACh:A\to C, gf=hg_{\mathbb{C}}\circ f=h_{\mathbb{C}}. Then ff descends to a subfield kk^{\prime} of \mathbb{C} which is finite over kk such that gkf=hkg_{k^{\prime}}\circ f=h_{k^{\prime}}.

Proof.

We assume without loss of generality that kk is algebraically closed. The graph Γf:A(A×CB)\Gamma_{f}\colon A_{\mathbb{C}}\to(A\times_{C}B)_{\mathbb{C}} of ff (as a morphism between CC_{\mathbb{C}}-schemes) defines a section of the étale morphism (g×Ch):(A×CB)A(g\times_{C}h)_{\mathbb{C}}\colon(A\times_{C}B)_{\mathbb{C}}\to A_{\mathbb{C}}. Hence Γf\Gamma_{f} maps AA_{\mathbb{C}} isomorphically onto a connected component DD_{\mathbb{C}} of (A×CB)(A\times_{C}B)_{\mathbb{C}}. Since kk is algebraically closed, DD_{\mathbb{C}} comes from an extension of scalars of a connected component D(A×CB)D\subset(A\times_{C}B). As the natural projection DAD\to A is defined over kk and its base change to \mathbb{C} is the inverse of Γf\Gamma_{f}, we must have that Γf\Gamma_{f} is also defined over kk, and hence so is ff. ∎

Below for any \mathbb{Q}-linear Tannakian category 𝒞\mathscr{C}, we write 𝔑()\mathfrak{N}(-) for the functor NmE/:𝒞(E)𝒞\mathrm{Nm}_{E/\mathbb{Q}}:\mathscr{C}_{(E)}\to\mathscr{C} (cf. (2.1.4)).
Proof of (4.1.10). As before, set 𝒦:=𝖪𝒢(𝔸f)\mathcal{K}:=\mathsf{K}\cap\mathcal{G}(\mathbb{A}_{f}) and let i:Sh𝒦(𝒢)Sh𝖪(G)Ei:\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})\to\mathrm{Sh}_{\mathsf{K}}(G)_{E} be the Shimura morphism. Our discussions in (4.2.2) up to (4.2.5) apply without any change in the (R2’) case, so that ρ\rho_{\mathbb{C}} factors through a morphism ϱ:SSh𝒦(𝒢)\varrho_{\mathbb{C}}:S_{\mathbb{C}}\to\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})_{\mathbb{C}} such that

ϱ(𝖵~,[η𝒱])(𝖯,[μ~])|S,\varrho_{\mathbb{C}}^{*}({\widetilde{\mathsf{V}}},[\eta_{\mathcal{V}}])\,{\cong}\,(\mathsf{P},[\widetilde{\mu}])|_{S_{\mathbb{C}}}, (18)

and we still have EFE\subseteq F. We first show that ϱ\varrho_{\mathbb{C}} descends to a morphism ϱ:SFSh𝒦(𝒢)F\varrho:S_{F^{\prime}}\to\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})_{F^{\prime}} for some finite extension F/FF^{\prime}/F in \mathbb{C}.

Set N:=𝔑(𝒱)N:=\mathfrak{N}(\mathcal{V}). Then NN is a faithful representation of =(𝒱)\mathcal{H}=\mathcal{H}(\mathcal{V}) (2.1.5). Since in (4.1.10) we assumed that 𝖪\mathsf{K} satisfies condition ()(\sharp), there exists a neat 𝒞(𝔸f)\mathcal{C}\subseteq\mathcal{H}(\mathbb{A}_{f}) such that the image of 𝒦\mathcal{K} lies in 𝒞\mathcal{C}. Now recall our discussion in (3.3.7) and the notations therein. Let π:Sh𝒦(𝒢)Sh𝒞()\pi:\mathrm{Sh}_{\mathcal{K}}(\mathcal{G})\to\mathrm{Sh}_{\mathcal{C}}(\mathcal{H}) be the natural Shimura morphism over EE. Let 𝔑([μ~])\mathfrak{N}([\widetilde{\mu}]) denote the 𝒞\mathcal{C}-level structure on 𝔑(𝖯e´t)\mathfrak{N}(\mathsf{P}_{\mathrm{{\acute{e}}t}}) such that 𝔑([μ~])b=𝒞(𝔑(μ~b)𝔸f)\mathfrak{N}([\widetilde{\mu}])_{b}=\mathcal{C}\cdot(\mathfrak{N}(\widetilde{\mu}_{b}){\otimes}\mathbb{A}_{f}). By applying 𝔑()\mathfrak{N}(-) to the diagrams in (4.1.7), (2.2.7)(c) implies that 𝔑(𝖯)\mathfrak{N}(\mathsf{P}) is weakly AM. One checks using (3.1.7) that 𝔑(μ~)\mathfrak{N}(\widetilde{\mu}) is 𝔑(𝖯)\mathfrak{N}(\mathsf{P})-rational, and is of type Ω\Omega_{\mathcal{H}}. Then by (3.2.5), we obtain a morphism ρN:SSh𝒞()F\rho_{N}:S\to\mathrm{Sh}_{\mathcal{C}}(\mathcal{H})_{F} such that

(ρN)(𝖭,[ηN])(𝔑(𝖯),𝔑([μ~])).(\rho_{N})^{*}(\mathsf{N},[\eta_{N}])\,{\cong}\,(\mathfrak{N}(\mathsf{P}),\mathfrak{N}([\widetilde{\mu}])). (19)

By applying 𝔑()\mathfrak{N}(-) to (18) and comparing with (8) in (3.3.7), for β:=πϱ\beta_{\mathbb{C}}:=\pi_{\mathbb{C}}\circ\varrho_{\mathbb{C}} we obtain

β(𝖭,[ηN])=ϱπ(𝖭,[ηN])ϱ(𝔑(𝖵~),𝔑([η𝒱]))(𝔑(𝖯),𝔑([μ~]))|S.\beta_{\mathbb{C}}^{*}(\mathsf{N},[\eta_{N}])=\varrho_{\mathbb{C}}^{*}\pi_{\mathbb{C}}^{*}(\mathsf{N},[\eta_{N}])\,{\cong}\,\varrho_{\mathbb{C}}^{*}(\mathfrak{N}({\widetilde{\mathsf{V}}}),\mathfrak{N}([\eta_{\mathcal{V}}]))\,{\cong}\,(\mathfrak{N}(\mathsf{P}),\mathfrak{N}([\widetilde{\mu}]))|_{S_{\mathbb{C}}}.

Therefore, by the uniqueness statement in (3.2.2), β=ρN|S\beta_{\mathbb{C}}=\rho_{N}|_{S_{\mathbb{C}}}, i.e., β\beta_{\mathbb{C}} is defined over FF. As π\pi is étale and is defined over EFE\subseteq F. By (4.3.1), ϱ\varrho_{\mathbb{C}} descends to a morphism ϱ\varrho over some finite extension F/FF^{\prime}/F in \mathbb{C} such that (ρN)F=πFϱ(\rho_{N})_{F^{\prime}}=\pi_{F^{\prime}}\circ\varrho.

The above gives the first statement of (4.1.10) and we now turn to the second. Recall that we defined αe´t:ρ(𝖵e´t,[ηV])(𝖯,[μ])|S\alpha^{\circ}_{\mathrm{{\acute{e}}t}}:\rho_{\mathbb{C}}^{*}(\mathsf{V}_{\mathrm{{\acute{e}}t}},[\eta_{V}])\stackrel{{\scriptstyle\sim}}{{\to}}(\mathsf{P},[\mu])|_{S_{\mathbb{C}}} in (4.1.8). Since 𝖵~e´t=i𝖵e´t\widetilde{\mathsf{V}}_{\mathrm{{\acute{e}}t}}=i^{*}\mathsf{V}_{\mathrm{{\acute{e}}t}} and ρ=iϱ\rho_{\mathbb{C}}=i_{\mathbb{C}}\circ\varrho_{\mathbb{C}}, we may alternatively view αe´t\alpha^{\circ}_{\mathrm{{\acute{e}}t}} as the étale component of (18), i.e., an isomorphism ϱ𝖵~e´t𝖯e´t|S\varrho_{\mathbb{C}}^{*}\widetilde{\mathsf{V}}_{\mathrm{{\acute{e}}t}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\mathrm{{\acute{e}}t}}|_{S_{\mathbb{C}}}, which sends ϱ[η𝒱]\varrho_{\mathbb{C}}^{*}[\eta_{\mathcal{V}}] to [μ~][\widetilde{\mu}]. As (ρN)F=πFϱ(\rho_{N})_{F^{\prime}}=\pi_{F^{\prime}}\circ\varrho, (19) gives us an isomorphism

ϱ(𝔑(𝖵~e´t),𝔑([η𝒱]))=ϱπ(𝖭e´t,[ηN])(𝔑(𝖯e´t),𝔑([μ~]))|SF\varrho^{*}(\mathfrak{N}(\widetilde{\mathsf{V}}_{\mathrm{{\acute{e}}t}}),\mathfrak{N}([\eta_{\mathcal{V}}]))=\varrho^{*}\pi^{*}(\mathsf{N}_{\mathrm{{\acute{e}}t}},[\eta_{N}])\,{\cong}\,(\mathfrak{N}(\mathsf{P}_{\mathrm{{\acute{e}}t}}),\mathfrak{N}([\widetilde{\mu}]))|_{S_{F^{\prime}}}

whose restriction to SS_{\mathbb{C}} is 𝔑(αe´t)\mathfrak{N}(\alpha^{\circ}_{\mathrm{{\acute{e}}t}}). This implies that 𝔑(αe´t)\mathfrak{N}(\alpha^{\circ}_{\mathrm{{\acute{e}}t}}) descends to SFS_{F^{\prime}}. As we assumed that 𝒦0Z1E(0)=1\mathcal{K}_{\ell_{0}}\cap Z^{1}_{E}(\mathbb{Q}_{\ell_{0}})=1, (3.3.6) tells us that the 0\ell_{0}-adic component α0:𝖵~0|S𝖯0|S\alpha^{\circ}_{\ell_{0}}:{\widetilde{\mathsf{V}}}_{\ell_{0}}|_{S_{\mathbb{C}}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\ell_{0}}|_{S_{\mathbb{C}}} descends to SFS_{F^{\prime}}. For every other \ell, 𝒦\mathcal{K}_{\ell} still contains an open subgroup 𝒦\mathcal{K}^{\prime}_{\ell} such that 𝒦Z1E()=1\mathcal{K}^{\prime}_{\ell}\cap Z^{1}_{E}(\mathbb{Q}_{\ell})=1. Hence (3.3.6) implies that ϱ𝖵~\varrho^{*}{\widetilde{\mathsf{V}}}_{\ell} is étale-locally isomorphic to 𝖯|SF\mathsf{P}_{\ell}|_{S_{F^{\prime}}}. ∎

Remark (4.3.2).

We remark that (4.1.10) is slightly weaker than (4.1.9) (e.g., one cannot descend ρ\rho_{\mathbb{C}} to FF but only to some finite extension) fundamentally because the representation 𝒢GL(N)\mathcal{G}\to\mathrm{GL}(N) is not faithful, but has a finite kernel. In the (R2) case, this was avoided because we worked with 𝒱\mathcal{V}^{\sharp} instead.

5 Proof of Theorem B

5.1   A specialization lemma for monodromy

Definition (5.1.1).

Let SS be a noetherian integral normal scheme. Let 𝒪S×\ell\in\mathcal{O}_{S}^{\times} be a prime and 𝖶\mathsf{W}_{\ell} be an étale \mathbb{Q}_{\ell}-local system. We denote by λ(𝖶)\lambda(\mathsf{W}_{\ell}) the dimension dimlimU𝖶,sU\dim\varinjlim_{U}\mathsf{W}_{\ell,s}^{U} where ss is a geometric point on SS and UU runs through open subgroups of π1e´t(S,s)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,s).

It is clear that the definition is independent of the choice of ss.

Definition (5.1.2).

Let TT be a noetherian base scheme and SS be a smooth TT-scheme of finite type.

  1. (a)

    Let 𝒪T×\ell\in\mathcal{O}_{T}^{\times} be a prime and 𝖶\mathsf{W}_{\ell} be an étale \mathbb{Q}_{\ell}-local system. We say that 𝖶\mathsf{W}_{\ell} has constant λgeo\lambda^{\mathrm{geo}} if there exists a number λ\lambda such that for every geometric point tTt\to T and every connected component SS^{\circ} of StS_{t}, λ(𝖶|S)=λ\lambda(\mathsf{W}_{\ell}|_{S^{\circ}})=\lambda. When this condition is satisfied, write λgeo(𝖶)\lambda^{\mathrm{geo}}(\mathsf{W}_{\ell}) for λ\lambda.

  2. (b)

    We say that S¯\overline{S} is a good relative compactification of SS if S¯\overline{S} is a smooth proper TT-scheme and there exists a relative normal crossing divisor DD of S¯\overline{S} such that S=S¯DS=\overline{S}-D.

Lemma (5.1.3).

Let TT be a DVR with special point tt and generic point η\eta. Assume chark(η)=0\mathrm{char\,}k(\eta)=0 and 𝒪T×\ell\in\mathcal{O}_{T}^{\times}. Let STS\to T be a smooth morphism of finite type with SS being connected. Let 𝖶\mathsf{W}_{\ell} be a \mathbb{Q}_{\ell}-local system over SS. If SS admits a good relative compactification S¯\overline{S} over TT, then 𝖶\mathsf{W}_{\ell} has constant λgeo\lambda^{\mathrm{geo}} over TT.

Proof.

Let T~\widetilde{T} be the strict Henselianization of TT and let t~\widetilde{t} and η~\widetilde{\eta} be the special and generic point of T~\widetilde{T}. Let SS^{\prime} be a connected component of ST~S_{\widetilde{T}}. Then [stacks-project, 055J] tells us that Sη~S^{\prime}_{\widetilde{\eta}} is connected. As Sη~S^{\prime}_{\widetilde{\eta}} necessarily contains a k(η~)k(\widetilde{\eta})-rational point, Sη~S^{\prime}_{\widetilde{\eta}} is geometrically connected. Now by applying [stacks-project, 0E0N] to S¯\overline{S}, Sη~S_{\widetilde{\eta}} and St~S_{\widetilde{t}} have the same number of geometric connected components, so St~S^{\prime}_{\widetilde{t}} must be connected. To prove the lemma we may replace TT by T~\widetilde{T} and SS by SS^{\prime}, so that StS_{t} and SηS_{\eta} are both geometrically connected. Let η¯\bar{\eta} be the geometric point over η\eta defined by a chosen algebraic closure of k(η)k(\eta).

Choose a section σ:TS\sigma:T\to S and set a=σ(t),b=σ(η¯)a=\sigma(t),b=\sigma(\bar{\eta}). Note that σ\sigma provides an étale path between aa and bb, through which we identify 𝖶,a\mathsf{W}_{\ell,a} with 𝖶,b\mathsf{W}_{\ell,b} and πe´t1(S,a)\pi^{\mathrm{{\acute{e}}t}}_{1}(S,a) with πe´t1(S,b)\pi^{\mathrm{{\acute{e}}t}}_{1}(S,b). Let ρ:πe´t1(S,a)GL(𝖶,a)\rho:\pi^{\mathrm{{\acute{e}}t}}_{1}(S,a)\to\mathrm{GL}(\mathsf{W}_{\ell,a}) be the monodromy representation. For any group GG with a morphism Gπe´t1(S,a)G\to\pi^{\mathrm{{\acute{e}}t}}_{1}(S,a) implicitly understood, write ρ(G)\rho(G)^{\circ} for the identity component of the Zariski closure of the image of GG in GL(𝖶,a)\mathrm{GL}(\mathsf{W}_{\ell,a}). Clearly, ρ(G)\rho(G)^{\circ} remains unchanged if we replace GG by a finite index subgroup. It suffices to show that ρ(π1e´t(St,a))=ρ(π1e´t(Sη¯,b))\rho(\pi_{1}^{\mathrm{{\acute{e}}t}}(S_{t},a))^{\circ}=\rho(\pi_{1}^{\mathrm{{\acute{e}}t}}(S_{\bar{\eta}},b))^{\circ}, i.e., Mon(𝖶|St,a)=Mon(𝖶|Sη¯,b)\mathrm{Mon}^{\circ}(\mathsf{W}_{\ell}|_{S_{t}},a)=\mathrm{Mon}^{\circ}(\mathsf{W}_{\ell}|_{S_{\bar{\eta}}},b) in the notation in (4.1.1).

Take a sequence n\mathcal{F}_{n} of locally constant free /n\mathbb{Z}/\ell^{n}\mathbb{Z}-modules over SS such that 𝖶(limnn)\mathsf{W}_{\ell}\,{\cong}\,(\varprojlim_{n}\mathcal{F}_{n}){\otimes}\mathbb{Q}_{\ell}. As chark(η)=0\mathrm{char\,}k(\eta)=0, each n\mathcal{F}_{n} is tamely ramified over S¯\overline{S} by Abhyankar’s lemma [SGA1, XIII App. Prop. 5.5]888There is a typo in the statement: YY should be Supp(D)\mathrm{Supp}(D), not XSupp(D)X-\mathrm{Supp}(D).. Let us use a superscript “tt” to indicate tame fundamental group. Then we know that ρ(πe´t1(S?,a))=ρ(π1e´t(S?,a)t)\rho(\pi^{\mathrm{{\acute{e}}t}}_{1}(S_{?},a))^{\circ}=\rho(\pi_{1}^{\mathrm{{\acute{e}}t}}(S_{?},a)^{t})^{\circ} for ?=,t?=\emptyset,t. By [SGA1, XIII Ex. 2.10], the natural map π1e´t(St,a)tπ1e´t(S,a)t\pi_{1}^{\mathrm{{\acute{e}}t}}(S_{t},a)^{t}\to\pi_{1}^{\mathrm{{\acute{e}}t}}(S,a)^{t} is an isomorphism, so it remains to show that ρ(π1e´t(Sη¯,b))=ρ(π1e´t(S,b)t)\rho(\pi_{1}^{\mathrm{{\acute{e}}t}}(S_{\bar{\eta}},b))^{\circ}=\rho(\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b)^{t})^{\circ}.

The section σ(η)\sigma(\eta) induces an isomorphism πe´t1(S,b)=πe´t1(Sη¯,b)Galk(η)\pi^{\mathrm{{\acute{e}}t}}_{1}(S,b)=\pi^{\mathrm{{\acute{e}}t}}_{1}(S_{\bar{\eta}},b)\rtimes\mathrm{Gal}_{k(\eta)}. As σ(𝖶)\sigma^{*}(\mathsf{W}_{\ell}) is necessarily trivial, the subgroup Galk(η)πe´t1(S,b)\mathrm{Gal}_{k(\eta)}\subseteq\pi^{\mathrm{{\acute{e}}t}}_{1}(S,b) acts trivially on 𝖶,b\mathsf{W}_{\ell,b}. Therefore, ρ(πe´t1(Sη¯,b))=ρ(πe´t1(Sη,b))=ρ(πe´t1(S,b)t)\rho(\pi^{\mathrm{{\acute{e}}t}}_{1}(S_{\bar{\eta}},b))^{\circ}=\rho(\pi^{\mathrm{{\acute{e}}t}}_{1}(S_{\eta},b))^{\circ}=\rho(\pi^{\mathrm{{\acute{e}}t}}_{1}(S,b)^{t}) as desired. Note that the second equality follows from the simple fact that π1e´t(Sη,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S_{\eta},b) maps surjectively to π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b), and πe´t1(S,b)t\pi^{\mathrm{{\acute{e}}t}}_{1}(S,b)^{t} is a quotient of π1e´t(S,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(S,b). ∎

Proposition (5.1.4).

Let SS be a connected smooth \mathbb{C}-variety and f:𝒳Sf:\mathcal{X}\to S be a \heartsuit-family. Let ss be any Hodge-generic point on SS. Then for any prime \ell and 𝖵:=R2f\mathsf{V}_{\ell}:=R^{2}f_{*}\mathbb{Q}_{\ell}, dimNS(𝒳s)=λ(𝖵)\dim\mathrm{NS}(\mathcal{X}_{s})_{\mathbb{Q}}=\lambda(\mathsf{V}_{\ell}).

Proof.

Up to replacing SS by a connected étale cover, assume that M:=Mon(𝖵,s)M_{\ell}:=\mathrm{Mon}(\mathsf{V}_{\ell},s) is connected (see (4.1.1)). Set ρ:=NS(𝒳s)\rho:=\mathrm{NS}(\mathcal{X}_{s})_{\mathbb{Q}}. By (4.1.2), ρ=NS(𝒳η)\rho=\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}}. As 𝖬\mathsf{M}_{\ell} is unchanged if we replace SS by a further connected étale cover, dim𝖵,sM=λ(𝖵)\dim\mathsf{V}_{\ell,s}^{M_{\ell}}=\lambda(\mathsf{V}_{\ell}). Let 𝖵=(𝖵B,𝖵dR)\mathsf{V}=(\mathsf{V}_{B},\mathsf{V}_{\mathrm{dR}}) be the VHS R2f(1)R^{2}f_{*}\mathbb{Q}(1). Then 𝖵\mathsf{V} splits into 𝟏Sρ𝖯\mathbf{1}_{S}^{\oplus\rho}\oplus\mathsf{P} for some VHS 𝖯\mathsf{P} such that 𝖯B,s(0,0)=0\mathsf{P}_{B,s}^{(0,0)}=0. It suffices to argue that 𝖯,sM=0\mathsf{P}_{\ell,s}^{M_{\ell}}=0, where 𝖯=𝖯B\mathsf{P}_{\ell}=\mathsf{P}_{B}{\otimes}\mathbb{Q}_{\ell}.

Set M:=Mon(𝖵B,s)M:=\mathrm{Mon}(\mathsf{V}_{B},s) (see (2.2.3)). Then we have M=MM_{\ell}=M{\otimes}\mathbb{Q}_{\ell}. This implies that V,sM=VB,sMV_{\ell,s}^{M_{\ell}}=V_{B,s}^{M}{\otimes}\mathbb{Q}_{\ell}, so we reduce to showing that 𝖯B,sM=0\mathsf{P}_{B,s}^{M}=0. We recall that Deligne’s theorem of the fixed part [DelHdg, (4.1.2)] says that the subspace 𝖯B,sM\mathsf{P}_{B,s}^{M} has a Hodge structure which is respected by the embedding 𝖯B,sM𝖯B,s\mathsf{P}_{B,s}^{M}\hookrightarrow\mathsf{P}_{B,s}. Since the Hodge structure on 𝖯B,s\mathsf{P}_{B,s} is irreducible ([HuyK3Book, §3 Lem. 2.7]), 𝖯B,sM\mathsf{P}_{B,s}^{M} is either 0 or 𝖯B,s\mathsf{P}_{B,s}. But it cannot be 𝖯B,s\mathsf{P}_{B,s} because M1M\neq 1 by our assumption that 𝒳/S\mathcal{X}/S is a \heartsuit-family. ∎

5.2   An Effective Theorem

First, we extend the set-ups in (4.1.3) and (4.1.5) to a family over (p)\mathbb{Z}_{(p)} when F=F=\mathbb{Q}.

Set-up (5.2.1).

Let 𝖬\mathsf{M} be a connected separated scheme over (p)\mathbb{Z}_{(p)} which is smooth and of finite type for some prime p>2p>2. Let (f:𝒳𝖬)(f\colon\mathcal{X}\to\mathsf{M}) be a smooth projective morphism of relative dimension dd such that 𝒳|𝖬\mathcal{X}|_{\mathsf{M}_{\mathbb{Q}}} is a \heartsuit-family. Let η\eta be the generic point of 𝖬\mathsf{M}. Let 𝝃\boldsymbol{\xi} be a relatively ample line bundle on 𝒳/𝖬\mathcal{X}/\mathsf{M}, which endows R2f𝔸pf(1)R^{2}f_{*}\mathbb{A}^{p}_{f}(1) and R2fp(1)R^{2}f_{\mathbb{Q}*}\mathbb{Z}_{p}(1) a symmetric bilinear pairing. Let ΛΛ0:=NS(𝒳η)\Lambda\subseteq\Lambda_{0}:=\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}} be a subspace which contains the class of 𝝃η\boldsymbol{\xi}_{\eta}. Recall that by (4.1.2), Pic(𝒳η)=Pic(𝒳)/Pic(𝖬)\mathrm{Pic}\,(\mathcal{X}_{\eta})=\mathrm{Pic}\,(\mathcal{X})/\mathrm{Pic}\,(\mathsf{M}). By choosing a section Λ0Pic(𝒳η)\Lambda_{0}\hookrightarrow\mathrm{Pic}\,(\mathcal{X}_{\eta})_{\mathbb{Q}}, we obtain an embedding Λ¯0R2f𝔸pf(1)\underline{\Lambda}_{0}\to R^{2}f_{*}\mathbb{A}^{p}_{f}(1), and for every field kk and s𝖬(k)s\in\mathsf{M}(k), Λ0\Lambda_{0} (and hence Λ\Lambda) is naturally a subspace of NS(𝒳s)\mathrm{NS}(\mathcal{X}_{s})_{\mathbb{Q}}. We write PNS(𝒳s)\mathrm{PNS}(\mathcal{X}_{s})_{\mathbb{Q}} for the orthogonal complement of Λ\Lambda in NS(𝒳s)\mathrm{NS}(\mathcal{X}_{s})_{\mathbb{Q}}. Note that these definitions are independent of the section Λ0Pic(𝒳η)\Lambda_{0}\hookrightarrow\mathrm{Pic}\,(\mathcal{X}_{\eta})_{\mathbb{Q}} chosen. Choose a base point b𝖬()b\in\mathsf{M}(\mathbb{C}) lying above η\eta and let the connected component of 𝖬\mathsf{M}_{\mathbb{C}} which contains bb be 𝖬\mathsf{M}^{\circ}.

We assume that Mon(R2f2,b)\mathrm{Mon}(R^{2}f_{*}\mathbb{Q}_{2},b) is connected. Now apply the set-ups in (4.1.3) and (4.1.5) with S=𝖬S=\mathsf{M}_{\mathbb{Q}} and S=𝖬S^{\circ}=\mathsf{M}^{\circ} and define a system of realizations (𝖯B,𝖯dR,𝖯e´t)𝖱(𝖬)(\mathsf{P}_{B},\mathsf{P}_{\mathrm{dR}},\mathsf{P}_{\mathrm{{\acute{e}}t}})\in\mathsf{R}(\mathsf{M}_{\mathbb{Q}}); moreover, we fix an isometry μb:V𝖯B,b\mu_{b}:V\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{B,b} and define the Shimura datum (G,Ω)(G,\Omega). Let (R2fp)tf(R^{2}f_{\mathbb{Q}*}\mathbb{Z}_{p})_{\mathrm{tf}} be the image of R2fpR^{2}f_{\mathbb{Q}*}\mathbb{Z}_{p} in R2fpR^{2}f_{\mathbb{Q}*}\mathbb{Q}_{p} and define (R2f(p))tf(R^{2}f_{\mathbb{C}*}\mathbb{Z}_{(p)})_{\mathrm{tf}} similarly (tf\mathrm{tf} is short for “torsion-free”). Let 𝐏B:=𝖯B(R2f(p)(1))tf\mathbf{P}_{B}:=\mathsf{P}_{B}\cap(R^{2}f_{\mathbb{C}*}\mathbb{Z}_{(p)}(1))_{\mathrm{tf}} and 𝐏p:=𝖯p(R2fp(1))tf\mathbf{P}_{p}:=\mathsf{P}_{p}\cap(R^{2}f_{\mathbb{Q}*}\mathbb{Z}_{p}(1))_{\mathrm{tf}}. For every p\ell\neq p, let 𝐏\mathbf{P}_{\ell} be the orthogonal complement of Λ\Lambda in R2f(1)R^{2}f_{*}\mathbb{Q}_{\ell}(1), so that 𝖯\mathsf{P}_{\ell} over 𝖬\mathsf{M}_{\mathbb{Q}} extends to 𝐏\mathbf{P}_{\ell} over 𝖬\mathsf{M}. If kk is a perfect field of characteristic pp and W:=W(k)W:=W(k), for every point t𝖬(k)t\in\mathsf{M}(k), 𝝃t\boldsymbol{\xi}_{t} defines a pairing on the F-isocrystal H2cris(𝒳t/W)[1/p]\mathrm{H}^{2}_{\mathrm{cris}}(\mathcal{X}_{t}/W)[1/p] and we write 𝐏cris,t[1/p]\mathbf{P}_{\mathrm{cris},t}[1/p] for the orthogonal complement of the classes in ΛNS(𝒳t)\Lambda\subseteq\mathrm{NS}(\mathcal{X}_{t})_{\mathbb{Q}}. Assume that the (p)\mathbb{Z}_{(p)}-pairing on L(p):=μb1(𝐏B,b)L_{(p)}:=\mu_{b}^{-1}(\mathbf{P}_{B,b}) is self-dual. We abusively write the reductive (p)\mathbb{Z}_{(p)}-group SO(L(p))\mathrm{SO}(L_{(p)}) also as GG.

Under the above set-up, we define:

Definition (5.2.2).

Let 𝖪G(𝔸f)\mathsf{K}\subseteq G(\mathbb{A}_{f}) be a neat compact open subgroup of the form 𝖪p𝖪p\mathsf{K}_{p}\mathsf{K}^{p} for 𝖪p=G(p)\mathsf{K}_{p}=G(\mathbb{Z}_{p}) and 𝖪pG(𝔸pf)\mathsf{K}^{p}\subseteq G(\mathbb{A}^{p}_{f}). Let 𝒮𝖪(G)\mathscr{S}_{\mathsf{K}}(G) denote the integral model of Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) over (p)\mathbb{Z}_{(p)}. Let 0p\ell_{0}\neq p be a prime. We say that a morphism ρ:𝖬𝒮𝖪(G)\rho:\mathsf{M}\to\mathscr{S}_{\mathsf{K}}(G) is an 0\ell_{0}-admissible period morphism if (recall the notations in (3.4.2))

  1. (a)

    there exists an isometry αB:ρ𝐋B|𝖬𝐏B|𝖬\alpha_{B}^{\circ}:\rho_{\mathbb{C}}^{*}\mathbf{L}_{B}|_{\mathsf{M}^{\circ}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathbf{P}_{B}|_{\mathsf{M}^{\circ}} compatible with the Hodge filtrations (i.e., induces an isomorphism (ρ𝖵)|𝖬𝖯|𝖬(\rho_{\mathbb{C}}^{*}\mathsf{V})|_{\mathsf{M}^{\circ}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}|_{\mathsf{M}^{\circ}} of VHS over 𝖬\mathsf{M}^{\circ});

  2. (b)

    there is an isometry α0:ρ𝐋0𝐏0\alpha_{\ell_{0}}:\rho^{*}\mathbf{L}_{\ell_{0}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathbf{P}_{\ell_{0}} whose restriction to 𝖬\mathsf{M}^{\circ} agrees with αB0\alpha_{B}^{\circ}{\otimes}\mathbb{Q}_{\ell_{0}};

  3. (c)

    for every p\ell\neq p, ρ𝐋\rho^{*}\mathbf{L}_{\ell} is étale-locally isomorphic to 𝐏\mathbf{P}_{\ell} over 𝖬\mathsf{M};

  4. (d)

    ρ𝐋p\rho_{\mathbb{Q}}^{*}\mathbf{L}_{p} is étale locally isomorphic to 𝐏p\mathbf{P}_{p} over 𝖬\mathsf{M}_{\mathbb{Q}}.

Note that the isomorphism α0\alpha_{\ell_{0}} is unique if it exists. If (b) is satisfied for every prime p\ell\neq p, then we simply say that ρ\rho is admissible.

Theorem (5.2.3).

Consider the set-up in (5.2.1). Assume that for some prime 0p\ell_{0}\neq p

  1. (a)

    𝐏0\mathbf{P}_{\ell_{0}} has constant λgeo\lambda^{\mathrm{geo}} over (p)\mathbb{Z}_{(p)} as defined in (5.1.2), and

  2. (b)

    there is an 0\ell_{0}-admissible period morphism ρ:𝖬𝒮𝖪(G)\rho:\mathsf{M}\to\mathscr{S}_{\mathsf{K}}(G) for some 𝖪\mathsf{K} as in (5.2.2).

Then for every kk which is finitely generated over 𝔽p\mathbb{F}_{p} and t:Spec(k)𝖬t:\mathrm{Spec\,}(k)\to\mathsf{M}, the fiber 𝒳t\mathcal{X}_{t} satisfies the Tate conjecture in codimension 11.

Proof.

Define G~:=CSpin(L)\widetilde{G}:=\mathrm{CSpin}(L) and 𝕂p:=G~(p)\mathbb{K}_{p}:=\widetilde{G}(\mathbb{Z}_{p}) as in §3.4. Choose a compact open subgroup 𝕂pG~(𝔸pf)\mathbb{K}^{p}\subseteq\widetilde{G}(\mathbb{A}^{p}_{f}) whose image is contained in 𝖪p\mathsf{K}^{p} such that 𝕂:=𝕂p𝕂p\mathbb{K}:=\mathbb{K}_{p}\mathbb{K}^{p} is neat. Up to replacing 𝖬\mathsf{M} by a further connected étale cover, let us assume that ρ\rho can be lifted to a morphism 𝖬𝒮:=𝒮𝕂(G~)\mathsf{M}\to\mathscr{S}:=\mathscr{S}_{\mathbb{K}}(\widetilde{G}). Below we shall use ρ\rho to denote this lift. Recall the definition of special endomorphisms in (3.4.1). Under these preparations, we have the following proposition:

Proposition (5.2.4).

For every algebraically closed field κ\kappa and geometric point s:Spec(κ)𝖬s:\mathrm{Spec\,}(\kappa)\to\mathsf{M}, there is an isomorphism θs:L(𝒜ρ(s))PNS(𝒳s)\theta_{s}:\mathrm{L}(\mathscr{A}_{\rho(s)})_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathrm{PNS}(\mathcal{X}_{s})_{\mathbb{Q}} such that the diagram

L(𝒜ρ(s)){{\mathrm{L}(\mathscr{A}_{\rho(s)})_{\mathbb{Q}}}}PNS(𝒳s){{\mathrm{PNS}(\mathcal{X}_{s})_{\mathbb{Q}}}}𝐋0,ρ(s){{\mathbf{L}_{\ell_{0},{\rho(s)}}}}𝐏0,s{{\mathbf{P}_{\ell_{0},s}}}θs\scriptstyle{\theta_{s}}α0,s\scriptstyle{\alpha_{\ell_{0},s}} (20)

commutes, where the vertical arrows are cycle class maps.

Now we prove (5.2.3) assuming the proposition above. First, we introduce a pp-adic analogue of λ()\lambda(-) over a closed point: Let pr:=W(𝔽pr)\mathbb{Z}_{p^{r}}:=W(\mathbb{F}_{p^{r}}) and let pr\mathbb{Q}_{p^{r}} be its fraction field. Let σ\sigma denote the Frobenius action on pr\mathbb{Q}_{p^{r}}. Let HH be an FF-isocrystal over pr\mathbb{Q}_{p^{r}}, i.e., a finite dimensional pr\mathbb{Q}_{p^{r}}-vector space equipped with an isomorphism F:HσprHF:H{\otimes}_{\sigma}\mathbb{Q}_{p^{r}}\stackrel{{\scriptstyle\sim}}{{\to}}H. The p\mathbb{Q}_{p}-vector space HF=piH^{F=p^{i}} is defined to be {hH:F(h1)=pih}\{h\in H:F(h{\otimes}1)=p^{i}h\}. For every rmr\mid m, one naturally defines an F-isocrystal structure on HprpmH{\otimes}_{\mathbb{Q}_{p^{r}}}\mathbb{Q}_{p^{m}}. Now we set

λ(H):=limrmdimp(Hpm)F=p.\lambda(H):=\varinjlim_{r\mid m}\dim_{\mathbb{Q}_{p}}(H{\otimes}\mathbb{Q}_{p^{m}})^{F=p}.

Choose a geometric point t¯\bar{t} over tt. Below any 𝖬\mathsf{M}-scheme TT is automatically viewed also as an 𝒮\mathscr{S}-scheme via ρ\rho.

Let us first assume that the field kk in (5.2.3) is finite. Then we have

dimH2e´t(𝒳t¯,(1))Gal(t¯/t)=dimp(H2cris(𝒳t/W(k))[1/p])F=p,\dim_{\mathbb{Q}_{\ell}}\mathrm{H}^{2}_{\mathrm{{\acute{e}}t}}(\mathcal{X}_{\bar{t}},\mathbb{Q}_{\ell}(1))^{\mathrm{Gal}(\bar{t}/t)}=\dim_{\mathbb{Q}_{p}}(\mathrm{H}^{2}_{\mathrm{cris}}(\mathcal{X}_{t}/W(k))[1/p])^{F=p},

where FF is the natural Frobenius action on crystalline cohomology. Indeed, both sides are equal to the order of the pole of the zeta function of 𝒳t\mathcal{X}_{t} at 11 ([Morrow, Prop. 4.1]). This implies that λ(t𝐏)=λ(𝐏cris,t[1/p])\lambda(t^{*}\mathbf{P}_{\ell})=\lambda(\mathbf{P}_{\mathrm{cris},t}[1/p]) for every prime p\ell\neq p. By assumption (5.2.2)(c), we also have λ(t𝐏)=λ(t𝐋)\lambda(t^{*}\mathbf{P}_{\ell})=\lambda(t^{*}\mathbf{L}_{\ell}). Similarly, we argue that

λ(𝐋cris,t[1/p])=λ(𝐏cris,t[1/p]).\lambda(\mathbf{L}_{\mathrm{cris},t}[1/p])=\lambda(\mathbf{P}_{\mathrm{cris},t}[1/p]). (21)

Indeed, suppose that k=𝔽prk=\mathbb{F}_{p^{r}}. Lift tt to some t~𝖬(pr)\widetilde{t}\in\mathsf{M}(\mathbb{Z}_{p^{r}}). Take vv to be the generic point of t~\widetilde{t}, and let v¯\bar{v} be a geometric point over vv. Using the pp-adic comparison isomorphism and the compatibility with cycle class maps (cf. [BMS, Thm 14.3] and [IIK, Cor. 11.6]), as well as [CSpin, Prop. 4.7], we see that

𝐏cris,t[1/p](𝐏p,v¯pBcris)Galpr and 𝐋cris,t[1/p](𝐋p,v¯pBcris)Galpr.\mathbf{P}_{\mathrm{cris},t}[1/p]\,{\cong}\,(\mathbf{P}_{p,\bar{v}}{\otimes}_{\mathbb{Q}_{p}}B_{\mathrm{cris}})^{\mathrm{Gal}_{\mathbb{Q}_{p^{r}}}}\text{ and }\mathbf{L}_{\mathrm{cris},t}[1/p]\,{\cong}\,(\mathbf{L}_{p,\bar{v}}{\otimes}_{\mathbb{Q}_{p}}B_{\mathrm{cris}})^{\mathrm{Gal}_{\mathbb{Q}_{p^{r}}}}.

Now (21) follows because by assumption (5.2.2)(d) there exists an isomorphism 𝐏p,v¯𝐋p,v¯\mathbf{P}_{p,\bar{v}}\,{\cong}\,\mathbf{L}_{p,\bar{v}} which is equivariant under an open subgroup of Galpr\mathrm{Gal}_{\mathbb{Q}_{p^{r}}}, so that the F-isocrystals above are isomorphic up to base changing to pm\mathbb{Q}_{p^{m}} for some rmr\mid m.

Now for any kk (not necessarily finite) and tt in (5.2.3), we claim that

L(𝒜t)=𝐋,t¯Gal(t¯/t)\mathrm{L}(\mathscr{A}_{t})_{\mathbb{Q}}{\otimes}\mathbb{Q}_{\ell}=\mathbf{L}_{\ell,\bar{t}}^{\mathrm{Gal}(\bar{t}/t)} (22)

The above paragraph implies that if kk is finite, then every prime p\ell\neq p, λ(t𝐋)=λ(𝐋cris,t[1/p])\lambda(t^{*}\mathbf{L}_{\ell})=\lambda(\mathbf{L}_{\mathrm{cris},t}[1/p]); that is, assumption [MPTate, (6.2)] is satisfied for ρ(t)\rho(t), so that (22) is given by [MPTate, Thm 6.4]. When kk is not finite, the claim follows from the proof of [MPTate, Cor. 6.11]. Indeed, we may assume that kk is the fraction field of some smooth and geometrically connected variety TT over a finite extension of 𝔽p\mathbb{F}_{p}, and tt extends to a morphism T𝖬T\to\mathsf{M}. Let us choose a closed point t0t_{0} on TT, which we also view as a k(t0)k(t_{0})-valued point on 𝖬\mathsf{M}. Choose a geometric point t¯0\bar{t}_{0} over t0t_{0}, and let γ\gamma be an étale path connecting t¯\bar{t} and t¯0\bar{t}_{0}. By [FC90, I Prop. 2.7], End(𝒜T)=End(𝒜t)\mathrm{End}(\mathscr{A}_{T})=\mathrm{End}(\mathscr{A}_{t}). This gives us a specialization morphism L(𝒜t)L(𝒜t0)\mathrm{L}(\mathscr{A}_{t})\to\mathrm{L}(\mathscr{A}_{t_{0}}) which fits into a commutative diagram below

L(𝒜t){{\mathrm{L}(\mathscr{A}_{t})}}𝐋,t¯{{\mathbf{L}_{\ell,\bar{t}}}}L(𝒜t0){{\mathrm{L}(\mathscr{A}_{t_{0}})}}𝐋,t¯0{{\mathbf{L}_{\ell,\bar{t}_{0}}}}End(𝒜t){{\mathrm{End}(\mathscr{A}_{t})}}End(𝐇,t¯){{\mathrm{End}(\mathbf{H}_{\ell,\bar{t}})}}End(𝒜t0){{\mathrm{End}(\mathscr{A}_{t_{0}})}}End(𝐇,t¯0){{\mathrm{End}(\mathbf{H}_{\ell,\bar{t}_{0}})}}γ\scriptstyle{\gamma}γ\scriptstyle{\gamma}

(recall the notations in §3.4). We claim that all vertical squares are Cartesian. For the squares on the left and right, this is clear by the definition (3.4.1). For the square at the front, this is obtained by applying (22) to t0t_{0}. Hence the remaining diagram at the back must also be Cartesian. Therefore, the surjectivity of End(𝒜t)End(𝐇,t¯)Gal(t¯/t)\mathrm{End}(\mathscr{A}_{t}){\otimes}\mathbb{Q}_{\ell}\to\mathrm{End}(\mathbf{H}_{\ell,\bar{t}})^{\mathrm{Gal}(\bar{t}/t)} ([Zarhin2]) implies the surjectivity of L(𝒜t)𝐋,t¯Gal(t¯/t)\mathrm{L}(\mathscr{A}_{t}){\otimes}\mathbb{Q}_{\ell}\to\mathbf{L}_{\ell,\bar{t}}^{\mathrm{Gal}(\bar{t}/t)}. Hence we have affirmed (22) for tt.

Finally, combining (20), (22), and assumption (5.2.2)(c), we have

dimPNS(𝒳t¯)=λ(t𝐏0)=λ(t𝐋0)=λ(t𝐋)=λ(t𝐏)\dim\mathrm{PNS}(\mathcal{X}_{\bar{t}})_{\mathbb{Q}}=\lambda(t^{*}\mathbf{P}_{\ell_{0}})=\lambda(t^{*}\mathbf{L}_{\ell_{0}})=\lambda(t^{*}\mathbf{L}_{\ell})=\lambda(t^{*}\mathbf{P}_{\ell})

This implies the Tate conjecture in codimension 11 for 𝒳t\mathcal{X}_{t}. ∎

Now we prove (5.2.4), which is the key geometric input to (5.2.3).

Proof.

We first prove the statement when charκ=0\mathrm{char\,}\kappa=0. Without loss of generality, we may assume that κ\kappa can be embedded to \mathbb{C}; moreover, as Aut()\mathrm{Aut}(\mathbb{C}) acts transitively on the set of connected components of 𝖬\mathsf{M}_{\mathbb{C}}, we may choose an embedding such that the resulting \mathbb{C}-point lies on the distinguished component 𝖬\mathsf{M}^{\circ} (defined in (5.2.1)) of ρ\rho. To prove the statement we may replace ss by this \mathbb{C}-point. Then the statement follows from Hodge theory. Indeed, we obtain a commutative diagram999Note that we are considering every 𝖬\mathsf{M}-scheme also as an 𝒮\mathscr{S}-scheme via ρ\rho, so 𝒜s\mathscr{A}_{s} (resp. 𝐋B,s\mathbf{L}_{B,s}) is the same as 𝒜ρ(s)\mathscr{A}_{\rho(s)} (resp. 𝐋B,ρ(s)\mathbf{L}_{B,\rho(s)}). Similar conventions apply below.:

L(𝒜s){{\mathrm{L}(\mathscr{A}_{s})_{\mathbb{Q}}}}PNS(𝒳s){{\mathrm{PNS}(\mathcal{X}_{s})_{\mathbb{Q}}}}𝐋B,s[1/p](0,0){{\mathbf{L}_{B,s}[1/p]^{(0,0)}}}𝐏B,s[1/p](0,0){{\mathbf{P}_{B,s}[1/p]^{(0,0)}}}θs\scriptstyle{\theta_{s}}\scriptstyle{\,{\cong}\,}\scriptstyle{\,{\cong}\,}αB,s\scriptstyle{\alpha^{\circ}_{B,s}} (23)

The vertical maps are again cycle class maps, but this time they are isomorphisms. For the arrow on the right, we are applying the Lefschetz (1,1)(1,1)-theorem. Since α0,s=αB,s0\alpha_{\ell_{0},s}=\alpha^{\circ}_{B,s}{\otimes}\mathbb{Q}_{\ell_{0}} by assumption, we obtain (20).

Now we assume that charκ=p\mathrm{char\,}\kappa=p. Set W=W(κ)W=W(\kappa). We shall construct θs\theta_{s} by considering characteristic 0 liftings of ss. Let 𝒮^s\widehat{\mathscr{S}}_{s} be the formal completion of 𝒮𝕂(G~)W\mathscr{S}_{\mathbb{K}}(\widetilde{G})_{W} at ρ(s)\rho(s). For a special endomorphism ζL(𝒜s)\zeta\in\mathrm{L}(\mathscr{A}_{s}) consider the following functor:

Def𝒮(ζ,s):R{s~𝒮^s(R)ζ deforms to L(𝒜s~)}\mathrm{Def}_{\mathscr{S}}(\zeta,s):R\mapsto\{\widetilde{s}\in\widehat{\mathscr{S}}_{s}(R)\mid\zeta\textit{ deforms to }\mathrm{L}(\mathscr{A}_{\widetilde{s}})\} (24)

where RR runs through all Artin WW-algebras. By [CSpin, §5.14], Def𝒮(ζ,s)\mathrm{Def}_{\mathscr{S}}(\zeta,s) is represented by a closed formal subscheme of 𝒮^s\widehat{\mathscr{S}}_{s} cut out by a single formal power series fζ𝒪𝒮^sf_{\zeta}\in\mathcal{O}_{\widehat{\mathscr{S}}_{s}}. Similarly, let 𝖬^s\widehat{\mathsf{M}}_{s} be the formal completion of 𝖬W\mathsf{M}_{W} at ss. Then ρ\rho restricts to a morphism 𝖬^s𝒮^s\widehat{\mathsf{M}}_{s}\to\widehat{\mathscr{S}}_{s}. Consider the pullback of 𝒜\mathscr{A} to 𝖬\mathsf{M} and define Def𝖬(ζ,s)\mathrm{Def}_{\mathsf{M}}(\zeta,s) to be the functor defined by (24) with 𝒮^s\widehat{\mathscr{S}}_{s} replaced by 𝖬^s\widehat{\mathsf{M}}_{s}. Then we have a fiber diagram:

Def𝖬(ζ,s){\mathrm{Def}_{\mathsf{M}}(\zeta,s)}Def𝒮(ζ,s){\mathrm{Def}_{\mathscr{S}}(\zeta,s)}𝖬^s{\widehat{\mathsf{M}}_{s}}𝒮^s{\widehat{\mathscr{S}}_{s}}ρ\scriptstyle{\rho}\scriptstyle{\lrcorner} (25)

In particular, Def𝖬(ζ,s)\mathrm{Def}_{\mathsf{M}}(\zeta,s) is a closed formal subscheme of 𝖬^s\widehat{\mathsf{M}}_{s} cut out by the pullback ρ(fζ)\rho^{*}(f_{\zeta}). Now we prove the key intermediate lemma:

Lemma (5.2.5).

Up to replacing ζ\zeta by a power, Def𝖬(ζ,s)\mathrm{Def}_{\mathsf{M}}(\zeta,s) is flat over WW.

Proof.

It suffices show that if ζ\zeta does not lie in the image of L(𝒜𝖬)L(𝒜s)\mathrm{L}(\mathscr{A}_{\mathsf{M}})_{\mathbb{Q}}\to\mathrm{L}(\mathscr{A}_{s})_{\mathbb{Q}}, then Def𝖬(ζ,s)\mathrm{Def}_{\mathsf{M}}(\zeta,s) is flat over WW, because otherwise up to replacing ζ\zeta by a power, Def𝖬(ζ,s)=𝖬^s\mathrm{Def}_{\mathsf{M}}(\zeta,s)=\widehat{\mathsf{M}}_{s}. Suppose by way of contradiction that Def𝖬(ζ,s)\mathrm{Def}_{\mathsf{M}}(\zeta,s) is not flat over WW, which is equivalent to saying that ρ(fζ)\rho^{*}(f_{\zeta}) vanishes on the entire mod pp disk 𝖬^κ,s:=𝖬^sWκ\widehat{\mathsf{M}}_{\kappa,s}:=\widehat{\mathsf{M}}_{s}{\otimes}_{W}\kappa, i.e., the formal completion of 𝖬κ\mathsf{M}_{\kappa} at ss. Let 𝜻\boldsymbol{\zeta} be the deformation of ζ\zeta over 𝖬^κ,s\widehat{\mathsf{M}}_{\kappa,s}. Let use write uu for the generic point of 𝖬^κ,s\widehat{\mathsf{M}}_{\kappa,s}. Let SS be the connected component of 𝖬κ\mathsf{M}_{\kappa} which contains ss and let vv be its generic point. Since 𝖬^κ,s\widehat{\mathsf{M}}_{\kappa,s} is also the completion of SS at ss, there is natural embedding k(v)k(u)k(v)\hookrightarrow k(u) of residue fields. Let u¯\bar{u} be the geometric point over uu defined by a chosen algebraic closure of k(u)k(u). We view it also as a geometric point over vv..

Recall that we assumed that Mon(R2fe´t2,b)\mathrm{Mon}(R^{2}f_{\mathrm{{\acute{e}}t}*}\mathbb{Q}_{2},b) is connected in (5.2.1), so that NS(𝒳η)NS(𝒳b)\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}}\hookrightarrow\mathrm{NS}(\mathcal{X}_{b})_{\mathbb{Q}} is an isomorphism by (4.1.2). Therefore, π1e´t(𝖬,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(\mathsf{M},b) acts trivially on NS(𝒳b)0\mathrm{NS}(\mathcal{X}_{b}){\otimes}\mathbb{Q}_{\ell_{0}}. Since α0,b\alpha_{\ell_{0},b} is π1e´t(𝖬,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(\mathsf{M},b)-equivariant by assumption (5.2.2)(b), π1e´t(𝖬,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(\mathsf{M},b) acts trivially on L(𝒳b)0\mathrm{L}(\mathcal{X}_{b}){\otimes}\mathbb{Q}_{\ell_{0}} as well. Hence every element of L(𝒜b)\mathrm{L}(\mathscr{A}_{b})_{\mathbb{Q}} is defined over η\eta. It follows from [FC90, I Prop. 2.7] that L(𝒜η)=L(𝒜𝖬)\mathrm{L}(\mathscr{A}_{\eta})=\mathrm{L}(\mathscr{A}_{\mathsf{M}}). By (5.1.4), dimPNS(𝒳b)=λgeo(𝐏0)\dim\mathrm{PNS}(\mathcal{X}_{b})_{\mathbb{Q}}=\lambda^{\mathrm{geo}}(\mathbf{P}_{\ell_{0}}). Assumption (5.2.2)(b) implies that 𝐋0|𝖬\mathbf{L}_{\ell_{0}}|_{\mathsf{M}} also has constant λgeo\lambda^{\mathrm{geo}} and λgeo(𝐋0|𝖬)=λgeo(𝐏0)\lambda^{\mathrm{geo}}(\mathbf{L}_{\ell_{0}}|_{\mathsf{M}})=\lambda^{\mathrm{geo}}(\mathbf{P}_{\ell_{0}}). Combining these observations, we must have dimL(𝒜𝖬)=λgeo(𝐋0|𝖬)\dim\mathrm{L}(\mathscr{A}_{\mathsf{M}})_{\mathbb{Q}}=\lambda^{\mathrm{geo}}(\mathbf{L}_{\ell_{0}}|_{\mathsf{M}}).

Now consider the subspace I:=limU𝐋0,u¯UI:=\varinjlim_{U}\mathbf{L}_{\ell_{0},\bar{u}}^{U}, as UU runs through open subgroups of πe´t1(v,u¯)\pi^{\mathrm{{\acute{e}}t}}_{1}(v,\bar{u}), so that dimI=λ(𝐋0|S)=λgeo(𝐋0|𝖬)\dim I=\lambda(\mathbf{L}_{\ell_{0}}|_{S})=\lambda^{\mathrm{geo}}(\mathbf{L}_{\ell_{0}}|_{\mathsf{M}}) (recall definition (5.1.2)). As the endomorphism scheme of an abelian scheme is representable and unramified over the base, every element in L(𝒜u¯)\mathrm{L}(\mathscr{A}_{\bar{u}}) is defined over some finite separable extension of k(v)k(v) inside k(u¯)k(\bar{u}). Therefore, we obtain a well defined map L(𝒜u¯)0I\mathrm{L}(\mathscr{A}_{\bar{u}}){\otimes}\mathbb{Q}_{\ell_{0}}\hookrightarrow I. However, we note that the composite

L(𝒜𝖬)0L(𝒜u¯)0I\mathrm{L}(\mathscr{A}_{\mathsf{M}}){\otimes}\mathbb{Q}_{\ell_{0}}\to\mathrm{L}(\mathscr{A}_{\bar{u}}){\otimes}\mathbb{Q}_{\ell_{0}}\hookrightarrow I

has to be an isomorphism because dimL(𝒜𝖬)=dimI\dim\mathrm{L}(\mathscr{A}_{\mathsf{M}})_{\mathbb{Q}}=\dim I. This forces the natural map L(𝒜𝖬)L(𝒜u¯)\mathrm{L}(\mathscr{A}_{\mathsf{M}})_{\mathbb{Q}}\to\mathrm{L}(\mathscr{A}_{\bar{u}})_{\mathbb{Q}} to be an isomorphism. Now note that 𝜻u¯L(𝒜u¯)\boldsymbol{\zeta}_{\bar{u}}\in\mathrm{L}(\mathscr{A}_{\bar{u}})_{\mathbb{Q}}. But as we assumed that ζ\zeta does not come from L(𝒜𝖬)\mathrm{L}(\mathscr{A}_{\mathsf{M}})_{\mathbb{Q}}, the same has to be true for 𝜻u¯\boldsymbol{\zeta}_{\bar{u}}. This gives the desired contradiction. ∎

Suppose we have replaced ζ\zeta by a power so that the above lemma holds. Following Deligne’s argument for [Del02, Cor. 1.7], we show that there exists a DVR VV finite flat over WW such that Def𝖬(ζ,s)\mathrm{Def}_{\mathsf{M}}(\zeta,s) admits a VV-valued point. Indeed, suppose that Def𝖬(ζ,s)=Spf(R)\mathrm{Def}_{\mathsf{M}}(\zeta,s)=\mathrm{Spf}(R) for some complete local WW-algebra RR. As pp is not a zero divisor in RR, one may complete pp into a system of parameters (p,x1,,xd)(p,x_{1},\cdots,x_{d}) for RR, where d=dimR1d=\dim R-1 ([stacks-project, 0BWY]). Let R:=R/(x1,,xd)R^{\prime}:=R/(x_{1},\cdots,x_{d}). By [stacks-project, 00LJ], pp is not a zero divisor in RR^{\prime}, so that RR^{\prime} remains flat over WW. In particular it has zero-dimensional fibres over WW and hence is quasi-finite. By [stacks-project, 04GG(13)] we can write R=A×BR^{\prime}=A\times B where AA is a finite WW-algebra and B/pB=0B/pB=0. Since SpecR\mathrm{Spec\,}R^{\prime} is closed in SpecR\mathrm{Spec\,}R we must have B=0B=0, i.e., RR^{\prime} is finite over WW. Now, by flatness RR^{\prime} admits a WW-algebra morphism RKR^{\prime}\to K for some finite extension KK of W[1/p]W[1/p], but then by finiteness RR^{\prime} necessarily maps into the integral closure of WW in KK, which is the desired VV.101010We spell out the commutative algebra details because it is a little counterintuitive that this argument does not require RR to be reasonably non-singular (cf. (5.2.6)).

Let s~𝖬W(V)\widetilde{s}\in\mathsf{M}_{W}(V) denote the point defined by the above morphism RVR\to V. Choose an algebraic clsoure K¯\bar{K}. As VV is a strictly Henselian DVR, it defines an étale path γs~\gamma_{\widetilde{s}} connecting w:=s~K¯w:=\widetilde{s}_{\bar{K}} and ss. There is a compatible specialization map NS(𝒳w)NS(𝒳s)\mathrm{NS}(\mathcal{X}_{w})\to\mathrm{NS}(\mathcal{X}_{s}) along VV (cf. [MPJumping, Prop. 3.6] and its proof), which we denote by sp(s~)\mathrm{sp}(\widetilde{s}). There is a similar specialization map of special endomorphisms. Now we have the following diagram:

L(𝒜w){{\mathrm{L}(\mathscr{A}_{w})_{\mathbb{Q}}}}PNS(𝒳w){{\mathrm{PNS}(\mathcal{X}_{w})_{\mathbb{Q}}}}L(𝒜s){{\mathrm{L}(\mathscr{A}_{s})_{\mathbb{Q}}}}PNS(𝒳s){{\mathrm{PNS}(\mathcal{X}_{s})_{\mathbb{Q}}}}𝐋0,w{{\mathbf{L}_{\ell_{0},w}}}𝐏0,w{{\mathbf{P}_{\ell_{0},w}}}𝐋0,s{{\mathbf{L}_{\ell_{0},s}}}𝐋0,s{{\mathbf{L}_{\ell_{0},s}}}sp(s~)\scriptstyle{\mathrm{sp}(\widetilde{s})}θw\scriptstyle{\theta_{w}}sp(s~)\scriptstyle{\mathrm{sp}(\widetilde{s})}α0,s\scriptstyle{\alpha_{\ell_{0},s}}γs~\scriptstyle{\gamma_{\widetilde{s}}}α0,w\scriptstyle{\alpha_{\ell_{0},w}}γs~\scriptstyle{\gamma_{\widetilde{s}}}θs\scriptstyle{\theta_{s}} (26)

Note that charK¯=0\mathrm{char\,}\bar{K}=0, so we have shown that θw\theta_{w} exists. The map θs\theta_{s} does not exist yet. But by construction of s~\widetilde{s}, ζ\zeta lifts to some (necessarily unique) ζ~\widetilde{\zeta} over ww, and we can define θs(ζ)\theta_{s}(\zeta) to be sp(s~)(θw(ζ~))\mathrm{sp}(\widetilde{s})(\theta_{w}(\widetilde{\zeta})). Note that θs(ζ)\theta_{s}(\zeta) does not depend on the choice of s~\widetilde{s} because its class in 𝐏0,s\mathbf{P}_{\ell_{0},s} is completely determined by the class of ζ\zeta in 𝐋0,s\mathbf{L}_{\ell_{0},s} via α0,s\alpha_{\ell_{0},s}. Repeating this construction for every ζL(𝒜s)\zeta\in\mathrm{L}(\mathscr{A}_{s}), we obtain the desired map θs\theta_{s}. ∎

Remark (5.2.6).

In [Del02], Deligne deduced Cor. 1.7 directly from Thm 1.6—there is no control on how singular the formal scheme Def(X0,L0)\mathrm{Def}(X_{0},L_{0}) in Thm 1.6 might be (here X0X_{0} is a K3 surface over an algebraically closed field kk of characteristic p>0p>0 and Def(X0,L0)\mathrm{Def}(X_{0},L_{0}) is the universal deformation space of the pair (X0,L0)(X_{0},L_{0})). From later work [OgusCrystals, (2.2)], we do know that if L0L_{0} is not a pp-th power of another line bundle and p>2p>2, then Def(X0,L0)\mathrm{Def}(X_{0},L_{0}) in Deligne’s Thm 1.6 is indeed always regular. However, as Deligne’s argument shows, this is unnecessary for one to deduce that Def(X0,L0)\mathrm{Def}(X_{0},L_{0}) admits a point valued over some finite flat extension of W(k)W(k).

5.3   Proof of Theorem B

(5.3.1)   

The reader may have noticed that (4.1.9) and (4.1.10) apply to geometrically connected bases. To make use of these results, we consider the following simple-minded functor, which sometimes goes by the name “Grothendieck restriction”: If kkk\subseteq k^{\prime} is a field extension, and TT is a kk^{\prime}-scheme, we write T(k)T_{(k)} for the kk-scheme obtained by composing the structure morphism of TT with the projection Spec(k)Spec(k)\mathrm{Spec\,}(k^{\prime})\to\mathrm{Spec\,}(k). Then TT(k)T\mapsto T_{(k)} is the left adjoint to the base change functor from kk-schemes to kk^{\prime}-schemes.

We will only apply this functor when k=k=\mathbb{Q}. Let SS be a connected smooth \mathbb{Q}-variety. Then its scheme of connected components π0(S)\pi_{0}(S) is the spectrum of some number field FF. The natural morphism Sπ0(S)S\to\pi_{0}(S) endows SS with the structure of an FF-variety. We denote this FF-variety by FS{}^{F}S. Note that now S=(FS)()S=({}^{F}S)_{(\mathbb{Q})} and FS{}^{F}S is geometrically connected as an FF-variety.

(5.3.2)   

Since SS and FS{}^{F}S have the same underlying scheme, étale sheaves or filtered flat vector bundles on SS are naturally identified with the corresponding structures on FS{}^{F}S and vice versa and we do not distinguish them notationally. If YY is a \mathbb{Q}-variety, then there is a canonical bijection between the sets of morphisms ϵ:MorF(FS,YF)Mor(S,Y)\epsilon:\mathrm{Mor}_{F}({}^{F}S,Y_{F})\stackrel{{\scriptstyle\sim}}{{\to}}\mathrm{Mor}_{\mathbb{Q}}(S,Y). Suppose now that ρ:FSYF\rho:{}^{F}S\to Y_{F} is a morphism, and MM is an étale sheaf or a filtered flat vector bundle on YY, then ρ(M|YF)\rho^{*}(M|_{Y_{F}}) is canonically identified with ϵ(ρ)M\epsilon(\rho)^{*}M because ϵ(ρ)=(YFY)ρ\epsilon(\rho)=(Y_{F}\to Y)\circ\rho as morphisms of schemes.

Proposition (5.3.3).

Let USpec([21])U\subseteq\mathrm{Spec}(\mathbb{Z}[2^{-1}]) be an open subscheme. Let 𝖬\mathsf{M} be a connected separated smooth UU-scheme of finite type with generic point η\eta. Let 𝒳S\mathcal{X}\to S be a smooth projective morphism such that 𝒳|𝖬\mathcal{X}|_{\mathsf{M}_{\mathbb{Q}}} is a \heartsuit-family.

Let 𝛏\boldsymbol{\xi} be a relatively ample line bundle of 𝒳/𝖬\mathcal{X}/\mathsf{M} and let ΛNS(𝒳η)\Lambda\subseteq\mathrm{NS}(\mathcal{X}_{\eta})_{\mathbb{Q}} be a subspace containing the class of 𝛏η\boldsymbol{\xi}_{\eta}. Let b𝖬()b\in\mathsf{M}(\mathbb{C}) be a point lying above η\eta and 𝖬\mathsf{M}^{\circ} be the connected component of 𝖬\mathsf{M}_{\mathbb{C}} containing bb. Assume that R2f2R^{2}f_{*}\mathbb{Q}_{2} has constant λgeo\lambda^{\mathrm{geo}}, and for every pUp\in U, the pairing on PH2(𝒳b,(p))tf\mathrm{PH}^{2}(\mathcal{X}_{b},\mathbb{Z}_{(p)})_{\mathrm{tf}} is self-dual. Then we have:

  1. (a)

    If (𝒳/𝖬,𝝃)|𝖬(\mathcal{X}/\mathsf{M},\boldsymbol{\xi})|_{\mathsf{M}^{\circ}} has maximal monodromy (see (2.2.3)), then for every s𝖬s\in\mathsf{M}, 𝒳s\mathcal{X}_{s} satisfies the Tate conjecture in codimension 11.

  2. (b)

    If (𝒳/𝖬,𝝃)|𝖬(\mathcal{X}/\mathsf{M},\boldsymbol{\xi})|_{\mathsf{M}^{\circ}} belongs to case (R2’) (see (2.2.4)), then up to replacing UU by a nonempty open subscheme, the above is true.

Here PH2(𝒳b,(p))tf\mathrm{PH}^{2}(\mathcal{X}_{b},\mathbb{Z}_{(p)})_{\mathrm{tf}} denotes the submodule of H2(𝒳b,(p))tf\mathrm{H}^{2}(\mathcal{X}_{b},\mathbb{Z}_{(p)})_{\mathrm{tf}} consisting of elements orthogonal to the image of Λ\Lambda in NS(𝒳b)\mathrm{NS}(\mathcal{X}_{b})_{\mathbb{Q}} under the pairing on H2(𝒳b,)\mathrm{H}^{2}(\mathcal{X}_{b},\mathbb{Q}) induced by 𝝃b\boldsymbol{\xi}_{b}. Recall that the big monodromy case contains (R+) = (R1) + (R2) and (CM).

Proof.

The hypothesis remains unchanged if we replace 𝖬\mathsf{M} by a connected étale cover (and bb by a lift). Therefore, we may assume that Mon(R2f2,b)\mathrm{Mon}(R^{2}f_{*}\mathbb{Q}_{2},b) is connected. Moreover, to prove (b), we may assume that Λ=NS(𝒳b)\Lambda=\mathrm{NS}(\mathcal{X}_{b})_{\mathbb{Q}}. Indeed, replacing Λ\Lambda by NS(𝒳b)\mathrm{NS}(\mathcal{X}_{b})_{\mathbb{Q}} might make PH2(𝒳b,(p))tf\mathrm{PH}^{2}(\mathcal{X}_{b},\mathbb{Z}_{(p)})_{\mathrm{tf}} no longer self-dual only for finitely many pp, but we are allowed to shrink UU.

Apply the set-ups in (4.1.3) to S=𝖬S=\mathsf{M}_{\mathbb{Q}} and S=𝖬S^{\circ}=\mathsf{M}^{\circ}. Let V,G=SO(V)V,G=\mathrm{SO}(V) and μb:V𝖯B,b\mu_{b}:V\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{B,b} be as defined in (4.1.6). Let LVL\subseteq V be the \mathbb{Z}-lattice defined by μb1(𝖯B,bH2(𝒳b,)tf)\mu_{b}^{-1}(\mathsf{P}_{B,b}\cap\mathrm{H}^{2}(\mathcal{X}_{b},\mathbb{Z})_{\mathrm{tf}}) and choose a N0N\gg 0 such that 𝖪:=G(𝔸f)ker(GL(L^)GL(L/2NL))\mathsf{K}:=G(\mathbb{A}_{f})\cap\ker(\mathrm{GL}(L{\otimes}\widehat{\mathbb{Z}})\to\mathrm{GL}(L/2^{N}L)) satisfies condition ()(\sharp); in case (b), we additionally require 𝖪0=2\mathsf{K}_{\ell_{0}=2} to be sufficiently small (see (4.1.6) for these notions). Note that 𝖪\mathsf{K} is of the form q𝖪q\prod_{q}\mathsf{K}_{q}, where qq runs through all primes and 𝖪q\mathsf{K}_{q} is a compact open subgroup of G(q)G(\mathbb{Q}_{q}). By (4.1.4), up to replacing 𝖬\mathsf{M} by a further finite connected étale cover, we assume that 𝖪\mathsf{K} is admissible, i.e., the image of π1e´t(𝖬,b)\pi_{1}^{\mathrm{{\acute{e}}t}}(\mathsf{M},b) in GL(𝖯e´t,b)\mathrm{GL}(\mathsf{P}_{\mathrm{{\acute{e}}t},b}) lies in 𝖪\mathsf{K} via the chosen isometry μb\mu_{b}. Now we set FF to be the field such that π0(S)=Spec(F)\pi_{0}(S)=\mathrm{Spec\,}(F) and recall that the base point bb induces an embedding FF\hookrightarrow\mathbb{C} through which S=(FS)S^{\circ}=({}^{F}S)_{\mathbb{C}}.

For (a), we may now apply (4.1.9) to FS{}^{F}S to conclude that there is a period morphism ρ:FSSh𝖪(G)F\rho:{}^{F}S\to\mathrm{Sh}_{\mathsf{K}}(G)_{F} together with an isomorphism αB:ρ(𝖵B)|S𝖯B|S\alpha^{\circ}_{B}:\rho_{\mathbb{C}}^{*}(\mathsf{V}_{B})|_{S^{\circ}}\to\mathsf{P}_{B}|_{S^{\circ}} which preserves the Hodge filtrations such that αe´t:=αB𝔸f\alpha^{\circ}_{\mathrm{{\acute{e}}t}}:=\alpha^{\circ}_{B}{\otimes}\mathbb{A}_{f} descends to an isomorphism ρ𝖵e´t𝖯e´t\rho^{*}\mathsf{V}_{\mathrm{{\acute{e}}t}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\mathrm{{\acute{e}}t}} over FS{}^{F}S. By (5.3.2) we may view ρ\rho as a morphism SSh𝖪(G)S\to\mathrm{Sh}_{\mathsf{K}}(G) over \mathbb{Q} and αe´t\alpha_{\mathrm{{\acute{e}}t}} as an isomorphism of étale sheaves over SS. Let p2p\neq 2 be a prime in UU. Restrict 𝖬\mathsf{M} to (p)\mathbb{Z}_{(p)} and apply the set-ups in (5.2.1). Note that by construction, 𝖪p=SO(Lp)\mathsf{K}_{p}=\mathrm{SO}(L{\otimes}\mathbb{Z}_{p}) is hyperspecial. Let 𝒮\mathscr{S} be the integral model of Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) over (p)\mathbb{Z}_{(p)}. Recall that 𝖵e´t\mathsf{V}_{\mathrm{{\acute{e}}t}} (resp. 𝖯e´t\mathsf{P}_{\mathrm{{\acute{e}}t}}) is the restriction of p𝐋×𝐋p[1/p]\prod_{\ell\neq p}\mathbf{L}_{\ell}\times\mathbf{L}_{p}[1/p] to Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) (resp. p𝐏×𝐏p[1/p]\prod_{\ell\neq p}\mathbf{P}_{\ell}\times\mathbf{P}_{p}[1/p] to S=𝖬S=\mathsf{M}_{\mathbb{Q}}) (see (3.4.2) and (5.2.1)). The existence of αe´t\alpha_{\mathrm{{\acute{e}}t}} allows us to apply (3.4.4) to extend ρ\rho to a morphism 𝖬(p)𝒮\mathsf{M}_{\mathbb{Z}_{(p)}}\to\mathscr{S}, which by construction is admissible in the sense of (5.2.2). Now we conclude by (5.2.3).

For (b) a minor adaptation is needed. Take 0=2\ell_{0}=2. By (4.1.10) we still have a period morphism ρ:SSh𝖪(G)\rho_{\mathbb{C}}:S^{\circ}\to\mathrm{Sh}_{\mathsf{K}}(G)_{\mathbb{C}} equipped with an isomorphism αB:ρ𝖵B𝖯B|S\alpha^{\circ}_{B}:\rho_{\mathbb{C}}^{*}\mathsf{V}_{B}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{B}|_{S^{\circ}} which preserves the Hodge filtrations, but ρ\rho_{\mathbb{C}} does not descend all the way to FF. Instead, we only have that for some finite extension F/FF^{\prime}/F in \mathbb{C}, and S~:=(FS)FF\widetilde{S}:=({}^{F}S){\otimes}_{F}F^{\prime}, ρ\rho_{\mathbb{C}} descends to a morphism ρ:S~Sh𝖪(G)F\rho:\widetilde{S}\to\mathrm{Sh}_{\mathsf{K}}(G)_{F^{\prime}} such that

  1. (i)

    α0:=αB0\alpha^{\circ}_{\ell_{0}}:=\alpha^{\circ}_{B}{\otimes}\mathbb{Q}_{\ell_{0}} descends to an isomorphism ρ𝖵0𝖯0|S~\rho^{*}\mathsf{V}_{\ell_{0}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathsf{P}_{\ell_{0}}|_{\widetilde{S}};

  2. (ii)

    and ρ𝖵\rho^{*}\mathsf{V}_{\ell} is étale locally isomorphic to 𝖯|S~\mathsf{P}_{\ell}|_{\widetilde{S}} for every other \ell.

Set S:=S~()S^{\prime}:=\widetilde{S}_{(\mathbb{Q})}. Then by (5.3.2) again, we may view ρ\rho as a morphism SSh𝖪(G)S^{\prime}\to\mathrm{Sh}_{\mathsf{K}}(G) and (i), (ii) above as statements about étale sheaves over SS^{\prime}. Note that SS^{\prime} is naturally a connected étale over of SS. As LL is self-dual at every pUp\in U, there exists a UU-scheme 𝒮\mathscr{S} such that 𝒮(p)\mathscr{S}{\otimes}\mathbb{Z}_{(p)} is the integral canonical model of Sh𝖪(G)\mathrm{Sh}_{\mathsf{K}}(G) over (p)\mathbb{Z}_{(p)} (cf. the first theorem of [Lovering]). By a standard spreading out argument, up to further shrinking UU, we may assume that SS^{\prime} is the \mathbb{Q}-fiber of a smooth UU-scheme 𝖬\mathsf{M}^{\prime}; moreover, SSS^{\prime}\to S extends to an étale covering map 𝖬𝖬\mathsf{M}^{\prime}\to\mathsf{M} and ρ:SSh𝖪(G)\rho:S^{\prime}\to\mathrm{Sh}_{\mathsf{K}}(G) extends to a morphism ρU:𝖬𝒮\rho_{U}:\mathsf{M}^{\prime}\to\mathscr{S}. The reader readily checks using (i) and (ii) above that for each pUp\in U the localization ρU(p)\rho_{U}{\otimes}\mathbb{Z}_{(p)} is an 0\ell_{0}-admissible period morphism in the sense of (5.2.2). Therefore, the conclusion follows from (5.2.3). ∎

We are now ready to prove Theorem B.

Proof.

Note that the conclusion of Theorem B is for p0p\gg 0. By (5.3.3) it suffices to show that there are exists an open subscheme USpec([21])U\subseteq\mathrm{Spec\,}(\mathbb{Z}[2^{-1}]) such that after we restrict 𝖬\mathsf{M} to UU, R2f2R^{2}f_{*}\mathbb{Q}_{2} has constant λgeo\lambda^{\mathrm{geo}}. By combining Nagata’s compactification and Hironaka’s resolution of singularities in characteristic zero, we can find a compactification 𝖬¯\bar{\mathsf{M}} of the generic fiber 𝖬\mathsf{M}_{\mathbb{Q}} such that the boundary 𝔇𝖬¯𝖬\mathfrak{D}\coloneqq\bar{\mathsf{M}}-\mathsf{M}_{\mathbb{Q}}, equipped with the reduced scheme structure, is a normal crossing divisor. For some open subscheme UU of Spec([21])\mathrm{Spec}(\mathbb{Z}[2^{-1}]), 𝖬¯\bar{\mathsf{M}} and 𝔇\mathfrak{D} are defined over UU, and 𝔇\mathfrak{D} becomes a relative normal crossing divisor over UU. This implies that for (p)U(p)\in U, 𝖬(p)\mathsf{M}_{\mathbb{Z}_{(p)}} admits a good compactification relative to (p)\mathbb{Z}_{(p)} in the sense of (5.1.2). Now we conclude by (5.1.3). ∎

6 Deforming Curves on Parameter Spaces

6.1   Families of Curves which homogeneously dominate a Variety

Let kk be an algebraically closed field. For a morphism f:XYf\colon X\to Y betweem kk-varieties, we say that ff has equi-dimensional fibers if for every two points y,yYy,y^{\prime}\in Y, dimXy=dimXy\dim X_{y}=\dim X_{y^{\prime}}. If f:XYf:X\to Y is a smooth morphism between kk-varieties, denote by 𝒯(X/Y)\mathcal{T}(X/Y) the relative tangent bundle, i.e., the dual of Ω1X/Y\Omega^{1}_{X/Y}. If Y=Spec(k)Y=\mathrm{Spec\,}(k), then we simply write 𝒯X\mathcal{T}X for 𝒯(X/Y)\mathcal{T}(X/Y), and for a kk-point xXx\in X, we write 𝒯xX\mathcal{T}_{x}X for the tangent space TxXT_{x}X to emphasize that it is a fiber of 𝒯X\mathcal{T}X.

Definition (6.1.1).

Let SS and TT be two smooth irreducible kk-varieties, g:𝒞Tg\colon\mathscr{C}\to T be a smooth family of connected curves, and φ:𝒞S\varphi\colon\mathscr{C}\to S be a morphism with equi-dimensional fibers. Let UU be the maximal open subvariety of 𝒞\mathscr{C} on which the composition 𝒯(𝒞/T)𝒯𝒞φ(𝒯S)\mathcal{T}(\mathscr{C}/T)\hookrightarrow\mathcal{T}\mathscr{C}\to\varphi^{*}(\mathcal{T}S) does not vanish. Suppose that

  1. (a)

    the induced morphism 𝒞T×S\mathscr{C}\to T\times S is quasi-finite,

  2. (b)

    for every kk-point sSs\in S, Us:=Uφ1(s)U_{s}:=U\cap\varphi^{-1}(s) is dense in φ1(s)\varphi^{-1}(s);

  3. (c)

    the morphism U(𝒯S)U\to\mathbb{P}(\mathcal{T}S) has equi-dimensional fibers.

Then we say that the family of curves 𝒞/T\mathscr{C}/T homogeneously dominates SS (via the morphism φ\varphi). If there exists an open dense subvariety TTT^{\prime}\subseteq T such that the restriction 𝒞|T\mathscr{C}|_{T^{\prime}} homogeneously dominates SS, then we say that 𝒞/T\mathscr{C}/T strongly dominates SS.

The natural morphism U(𝒯S)U\to\mathbb{P}(\mathcal{T}S) is induced by the identification 𝒞(𝒯(𝒞/T))\mathscr{C}\,{\cong}\,\mathbb{P}(\mathcal{T}(\mathscr{C}/T)). Roughly speaking, the family 𝒞/T\mathscr{C}/T homogeneously dominates SS if there are curves passing through every given point on SS in any given direction, and the sub-family of such curves has a fixed dimension.

The notion “𝒞/T\mathscr{C}/T strongly dominates SS” is only defined for convenience, as sometimes the natural families of curves have some bad locus on TT of smaller dimension which does not affect applications.

Lemma (6.1.2).

Let 𝒫S\mathscr{P}\to S be a smooth morphism between smooth kk-varieties. Let 𝒳𝒫\mathcal{X}\subseteq\mathscr{P} be a relative effective Cartier divisor whose total space is smooth. If sS(k)s\in S(k) is a point such that 𝒳s\mathcal{X}_{s} has isolated singularities, then there exists an open dense subvariety U(𝒯sS)U\subseteq\mathbb{P}(\mathcal{T}_{s}S) with the following property: For every unramified morphism φ:CS\varphi\colon C\to S from a smooth curve CC which sends a point cCc\in C to ss, the total space of the pullback family 𝒳|C\mathcal{X}|_{C} has no singularity on 𝒳s\mathcal{X}_{s} if dφ((𝒯cC))Ud\varphi(\mathbb{P}(\mathcal{T}_{c}C))\in U.

Proof.

Since the question is étale-local in nature, we might as well assume that S=𝔸mkS=\mathbb{A}^{m}_{k} for m=dimSm=\dim S, s=0s=0, 𝒳s\mathcal{X}_{s} has a single isolated singularity at a kk-point P𝒳s𝒫sP\in\mathcal{X}_{s}\subseteq\mathscr{P}_{s}, and 𝒫\mathscr{P} is isomorphic to 𝔸nS=S×𝔸nk\mathbb{A}^{n}_{S}=S\times\mathbb{A}^{n}_{k} near PP. Let xix_{i}’s and sjs_{j}’s be the coordinates on 𝔸nk\mathbb{A}^{n}_{k} and SkS_{k} respectively. Suppose that 𝒳\mathcal{X} is locally cut out by an equation F(x1,,xn,s1,,sm)F(x_{1},\cdots,x_{n},s_{1},\cdots,s_{m}) near PP. That PP is a singularity of the fiber 𝒳s\mathcal{X}_{s} but not of the total space 𝒳\mathcal{X} implies that F/sj0\partial F/\partial s_{j}\neq 0 at PP for some jj. One may simply take UU to be the open subscheme of (𝒯sS)m1\mathbb{P}(\mathcal{T}_{s}S)\,{\cong}\,\mathbb{P}^{m-1} where the coordinate of sj\partial_{s_{j}} is nonzero. ∎

Definition (6.1.3).

Let f:XSf\colon X\to S be a finite type morphism between schemes.

  1. (a)

    Let the singular locus Sing(f)\mathrm{Sing}(f) be the reduced closed subscheme of XX whose support consists of all points where ff fails to be smooth.111111Note that this definition is different from the one given in [stacks-project, Tag 0C3H].

  2. (b)

    If ff is in addition proper and flat, we say that the scheme-theoretic image of Sing(f)\mathrm{Sing}(f) is the (generalized) discriminant locus of ff, and denote it by Disc(f)\mathrm{Disc}(f).

  3. (c)

    In the above situation, we say that Disc(f)\mathrm{Disc}(f) is mild if it has codimension at least 11 in SS and there exists a dense open subscheme VDisc(f)V\subseteq\mathrm{Disc}(f) such that for every geometric point ss on VV, the fiber XsX_{s} has only isolated singularities.

Remark (6.1.4).

Note that the properness assumption on ff implies that Disc(f)\mathrm{Disc}(f) is closed in SS. Moreover, since it is defined to be the scheme-theoretic image of a reduced scheme, it is also reduced. Its formation commutes with flat base change but not arbitrary base change: For any morphism TST\to S, Disc(fT)\mathrm{Disc}(f_{T}) is always the reduced subscheme of Disc(f)T\mathrm{Disc}(f)_{T}, so they are equal if and only if the latter is reduced.

Proposition (6.1.5).

Let 𝒳\mathcal{X} and SS be as in (6.1.2). Suppose that

  • the generalized discriminant variety DDisc(f)SD\coloneqq\mathrm{Disc}(f)\subseteq S is mild;

  • there is a family of smooth curves g:𝒞Tg:\mathscr{C}\to T which homogeneously dominates SS through a morphism φ:𝒞S\varphi:\mathscr{C}\to S.

Then for a general kk-point tTkt\in T_{k}, the total space of the pullback family 𝒳|𝒞t\mathcal{X}|_{\mathscr{C}_{t}} is smooth.

Proof.

By assumption we have a diagram

𝒞{\mathscr{C}}𝒳{\mathcal{X}}T{T}S{S}g\scriptstyle{g}φ\scriptstyle{\varphi}f\scriptstyle{f}

Let 𝒵S×T\mathcal{Z}\subseteq S\times T be the subset of points (s,t)(s,t) such that 𝒞t\mathscr{C}_{t} passes through ss and the total space of 𝒳|𝒞t\mathcal{X}|_{\mathscr{C}_{t}} has a singularity lying above φ1t(s)\varphi^{-1}_{t}(s). It is easy to see that 𝒵\mathcal{Z} is constructible: Let 𝒳|𝒞\mathcal{X}|_{\mathscr{C}} be the pullback of 𝒳\mathcal{X} along φ\varphi. Then we have a natural morphism 𝒳|𝒞T×kS\mathcal{X}|_{\mathscr{C}}\to T\times_{k}S and 𝒵\mathcal{Z} is the set-theoretic image of Sing(𝒳|𝒞T)\mathrm{Sing}(\mathcal{X}|_{\mathscr{C}}\to T). Endow 𝒵\mathcal{Z} with the structure of a reduced scheme.

It suffices to show that dim𝒵<dimT\dim\mathcal{Z}<\dim T. Let VV be as in (6.1.3)(c) and for sV(k)s\in V(k), let Usφ1(s)U_{s}\subseteq\varphi^{-1}(s) be as in (6.1.1)(b). Let t=g(s)t=g(s). By (6.1.2), there exists a proper closed subvariety Us,badUU_{s,\mathrm{bad}}\subseteq U such that the total space of 𝒳|𝒞t\mathcal{X}|_{\mathscr{C}_{t}} is not smooth near the fiber 𝒳u\mathcal{X}_{u} only if uUs,badu\in U_{s,\mathrm{bad}}. Let 𝒵~sφ1(s)\widetilde{\mathcal{Z}}_{s}\subseteq\varphi^{-1}(s) be the union of the complement of UsU_{s} and the Zariski closure of Us,badU_{s,\mathrm{bad}}. Then 𝒵~s\widetilde{\mathcal{Z}}_{s} is a proper closed subvariety of φ1(s)\varphi^{-1}(s). Since the morphism φ1(s)T\varphi^{-1}(s)\to T is quasi-finite and the fiber 𝒵sT\mathcal{Z}_{s}\subseteq T is contained in the image of 𝒵~s\widetilde{\mathcal{Z}}_{s}, we have

dim𝒵s<dimφ1(s)=dim𝒞dimS.\dim\mathcal{Z}_{s}<\dim\varphi^{-1}(s)=\dim\mathscr{C}-\dim S.

Since the image of 𝒵\mathcal{Z} in SS is contained in DD, and ss runs over an open dense subvariety of DD, we have that dim𝒵dim𝒞2=dimT1\dim\mathcal{Z}\leq\dim\mathscr{C}-2=\dim T-1 as desired. ∎

6.2   Applications of the Baire category theorem

Let kk be an algebraically closed field of characteristic p>0p>0. Set W:=W(k)W:=W(k) and K:=W[1/p]K:=W[1/p]. Choose an algebraic closure K¯\bar{K} of KK.

Lemma (6.2.1).

Suppose that SBS\to B is a flat morphism between irreducible smooth WW-schemes of finite type. Let NN be a countable union of closed proper subschemes of SK¯S_{\bar{K}}. Let bB(k)b\in B(k) be any point and B^b\widehat{B}_{b} be the formal completion of BB at bb. Then the subset of points b~B^b(W)\widetilde{b}\in\widehat{B}_{b}(W) such that supp(Sb~K¯)\mathrm{supp}(S_{\widetilde{b}}{\otimes}\bar{K}) is not contained in NN is analytically dense.

Proof.

Let U:=B^b(W)U:=\widehat{B}_{b}(W). By taking the union of NN with all its Galois conjugates, we may assume that NN is defined over KK. Let N1,N2,N_{1},N_{2},\cdots be the irreducible components of NN. By flatness the morphism SBS\to B is open, so for each ii there exists a proper closed subscheme ZiBZ_{i}\subseteq B such that every zB(K)z\in B(K) satisfies supp(Sz)Ni\mathrm{supp\,}(S_{z})\subseteq N_{i} only if zZiz\in Z_{i}. Indeed, one may simply take ZiZ_{i} to be the complement of the image of SNiS-N_{i}. Since each UZi(W)U-Z_{i}(W) is open dense in analytic topology, we may conclude by the Baire category theorem for complete metric spaces that Ui=1Zi(W)U-\cup_{i=1}^{\infty}Z_{i}(W) is analytically dense. ∎

Lemma (6.2.2).

Let SS and TT be smooth irreducible WW-schemes of finite type and NN be a countable union of closed proper subschemes of SK¯S_{\bar{K}}. Let tTkt\in T_{k} be a closed point and T^t\widehat{T}_{t} be the formal completion of TT at tt.

Suppose that 𝒞\mathscr{C} is a smooth family of geometrically connected curves over TT, and there is a morphism φ:𝒞S\varphi\colon\mathscr{C}\to S such that the family (𝒞/T)K¯(\mathscr{C}/T)_{\bar{K}} over K¯\bar{K} strongly dominates SK¯S_{\bar{K}}. Then the subset of points t~T^t(W)\widetilde{t}\in\widehat{T}_{t}(W) such that φ(supp(𝒞t~K¯))\varphi(\mathrm{supp\,}(\mathscr{C}_{\widetilde{t}}{\otimes}\bar{K})) is not contained NN is analytically dense.

Proof.

Again by Galois descent we may assume that N=i=1NiN=\cup_{i=1}^{\infty}N_{i} for irreducible closed subschemes NiN_{i} of SKS_{K}. Let Mi:=𝒞K×φNiM_{i}:=\mathscr{C}_{K}\times_{\varphi}N_{i} and M:=i=1MiM:=\cup_{i=1}^{\infty}M_{i}. The assumption that (𝒞/T)K¯(\mathscr{C}/T)_{\bar{K}} strongly dominates SK¯S_{\bar{K}} implies that each MiM_{i} is a proper closed subscheme. Now we apply the above lemma with (SB,N)(S\to B,N) replaced by (𝒞T,M)(\mathscr{C}\to T,M). ∎

7 Elliptic Surfaces with pg=q=1p_{g}=q=1

7.1   Generalities on Elliptic Surfaces

In this section we recall some basic facts about elliptic surfaces and describe their moduli. Let kk be an algebraically closed field of characteristic 2,3\neq 2,3. Let CC be a smooth projective curve over kk and π:XC\pi\colon X\to C be an elliptic surface over CC with a zero section σ:CX\sigma\colon C\to X through which we also view CC as a curve on XX. The fundamental line bundle LL of X/CX/C is defined to be the dual of the normal bundle NC/XN_{C/X}, or equivalently that of R1π𝒪XR^{1}\pi_{*}\mathcal{O}_{X}. The degree of LL is defined to be the height of XX, which we denote by ht(X)\mathrm{ht}(X). Set Vr=H0(Lr)V_{r}=\mathrm{H}^{0}(L^{r}). There exists a pair (a4,a6)V4×V6{0}(a_{4},a_{6})\in V_{4}\times V_{6}-\{0\}, which is unique up to the action of λk×\lambda\in k^{\times} by λ(a4,a6)=(λ4a4,λ6a6)\lambda\cdot(a_{4},a_{6})=(\lambda^{4}a_{4},\lambda^{6}a_{6}), such that XX is the minimal resolution of the hypersurface X(L2L3𝒪C)X^{\prime}\subseteq\mathbb{P}(L^{2}\oplus L^{3}\oplus\mathcal{O}_{C}) defined by the Weierstrass equation ([Kas, Thm 1])

y2z=x3a4xz2a6z3,y^{2}z=x^{3}-a_{4}xz^{2}-a_{6}z^{3}, (27)

where x,y,zx,y,z as homogenenous coordinates on L2,L3,𝒪CL^{2},L^{3},\mathcal{O}_{C} respectively. The hypersurface XX^{\prime} has at most rational double point singularities and is called the Weierstrass normal form of the original surface XX. If XX^{\prime} is smooth, then of course X=XX=X^{\prime}. In this paper, we only consider XX with ht(X)>0\mathrm{ht}(X)>0.

Next, we recall that Kodaira classified all the possible singular fibers in the elliptic fibration π:XC\pi\colon X\to C when k=k=\mathbb{C} in [Kodaira], and his classification is well known to hold verbatim in characteristic 2,3\neq 2,3 as well. We refer the reader to [SchSh, §4] for a summary. Set Δ4a4327a62\Delta\coloneqq 4a_{4}^{3}-27a_{6}^{2}. Let cCc\in C be a point and denote by valc\mathrm{val}_{c} the valuation defined by a uniformizer at cc. The only facts we shall need from loc. cit. are the following:

Proposition (7.1.1).
  1. (a)

    The fiber XcX_{c} is singular if and only if Δ\Delta vanishes at cc, i.e., valc(Δ)1\mathrm{val}_{c}(\Delta)\geq 1.

  2. (b)

    XcX_{c} is of In\mathrm{I}_{n} type (n>0)(n>0) if and only if valc(a4)=valc(a6)=0\mathrm{val}_{c}(a_{4})=\mathrm{val}_{c}(a_{6})=0, and n=valc(Δ)n=\mathrm{val}_{c}(\Delta).

  3. (c)

    XcX_{c} is of II\mathrm{II}-type if and only if valc(a4)1\mathrm{val}_{c}(a_{4})\geq 1 and valc(a6)=1\mathrm{val}_{c}(a_{6})=1.

  4. (d)

    If XcX_{c} is a singular fiber of any other type, valc(Δ)3\mathrm{val}_{c}(\Delta)\geq 3.

  5. (e)

    If XcX_{c} is of I1\mathrm{I}_{1}-type or II\mathrm{II}-type, then Xc=XcX_{c}=X^{\prime}_{c}. In other words, the singularity on XcX^{\prime}_{c} is not a surface singularity.

  6. (f)

    If XcX_{c} is of I2\mathrm{I}_{2}-type, then XcX^{\prime}_{c} has a unique ODP singularity given by contracting the irreducible component not meeting the zero section.

The degree of the discriminant Δ\Delta is 12χ(𝒪X)=e(X)12\chi(\mathcal{O}_{X})=e(X). Recall that the genus g(C)g(C) is equal to the irregularity q(X)q(X) and we have pg(X)=χ(𝒪X)1+g(C)p_{g}(X)=\chi(\mathcal{O}_{X})-1+g(C). Therefore, elliptic surfaces with pg=1p_{g}=1 fall into two types:

  • χ(𝒪X)=2\chi(\mathcal{O}_{X})=2 and g(C)=0g(C)=0. These are elliptic K3 surfaces.

  • χ(𝒪X)=g(C)=1\chi(\mathcal{O}_{X})=g(C)=1. These surfaces have Kodaira dimension 11.

We are interested in the latter class. Note that although these surfaces are elliptic fibrations over genus 11 curves, one should not confuse them with bielliptic surfaces, which are of Kodaira dimension 0.

(7.1.2)   

For future reference we introduce some notation. Let BB be a base scheme and 𝒱\mathcal{V} be a vector bundle over BB. We denote by 𝔸(𝒱)\mathbb{A}(\mathcal{V}) the relative affine space over BB defined by 𝒱\mathcal{V} and 𝔸(𝒱)\mathbb{A}(\mathcal{V})^{*} the open part of 𝔸(𝒱)\mathbb{A}(\mathcal{V}) minus the zero section. Given a sequence of numbers q=(q0,,qm)q=(q_{0},\cdots,q_{m}) such that qiq_{i}’s are invertible in 𝒪B\mathcal{O}_{B} and vector bundles 𝒱0,,𝒱m\mathcal{V}_{0},\cdots,\mathcal{V}_{m} such that 𝒱=i=0m𝒱i\mathcal{V}=\oplus_{i=0}^{m}\mathcal{V}_{i}, we denote by 𝒫q(𝒱)\mathcal{P}_{q}(\mathcal{V}) the resulting weighted projective stack, i.e., the quotient stack of 𝔾m\mathbb{G}_{m}-action on 𝒱\mathcal{V} given by

λ:(v0,,vm)(λq0v0,,λqmvm) for λ𝔾m,\lambda\colon(v_{0},\cdots,v_{m})\mapsto(\lambda^{q_{0}}v_{0},\cdots,\lambda^{q_{m}}v_{m})\text{ for }\lambda\in\mathbb{G}_{m},

and by q(𝒱)\mathbb{P}_{q}(\mathcal{V}) the coarse moduli space of 𝒫q(𝒱)\mathcal{P}_{q}(\mathcal{V}). It is well known that this coarse moduli space can be constructed explicitly by applying the relative Proj functor to a sheaf of graded algebras over BB. We omit the details. If qq is not specified, then it is assumed to be (1,,1)(1,\cdots,1).

Set-up (7.1.3).

Let BB be a [1/6]\mathbb{Z}[1/6]-scheme, ϖ:𝒞B\varpi\colon\mathcal{C}\to B be a family of smooth projective curves over BB of genus gg and \mathcal{L} be a relative line bundle on 𝒞\mathcal{C} of degree hh. Assume that 4h2g14h\geq 2g-1 and h1h\geq 1. Let 𝒱r\mathcal{V}_{r} denote the vector bundle ϖr\varpi_{*}\mathcal{L}^{r} for r4r\geq 4. Let 𝒳~\widetilde{\mathcal{X}} be the subscheme of 𝔸(𝒱4𝒱6)×B(23𝒪𝒞)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}\times_{B}\mathbb{P}(\mathcal{L}^{2}\oplus\mathcal{L}^{3}\oplus\mathcal{O}_{\mathcal{C}}) defined by the Weierstrass equation (27) in the obvious way. Let μ\mu be the 𝔾m\mathbb{G}_{m}-action on 𝔸(𝒱4𝒱6)×B(23𝒪𝒞)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}\times_{B}\mathbb{P}(\mathcal{L}^{2}\oplus\mathcal{L}^{3}\oplus\mathcal{O}_{\mathcal{C}}) defined by

λ((a4,a6),[x:y:z])=((λ4a4,λ6a6),[λ2x:λ3y:z]), for λ𝔾m.\lambda\cdot((a_{4},a_{6}),[x:y:z])=((\lambda^{4}a_{4},\lambda^{6}a_{6}),[\lambda^{2}x:\lambda^{3}y:z]),\text{ for }\lambda\in\mathbb{G}_{m}. (28)

Let 𝒬(μ)\mathcal{Q}(\mu) denote the quotient stack of the 𝔾m\mathbb{G}_{m}-action μ\mu. Then 𝒳~\widetilde{\mathcal{X}} descends to an algebraic substack 𝒳\mathcal{X} of 𝒬(μ)\mathcal{Q}(\mu). Note that 𝒬(μ)\mathcal{Q}(\mu), and hence 𝒳\mathcal{X}, admit natural morphisms to 𝒫(4,6)(𝒱4𝒱6)\mathcal{P}_{(4,6)}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}). Set 𝔇:=Disc(𝒳~/𝔸(𝒱4𝒱6))\mathfrak{D}:=\mathrm{Disc}(\widetilde{\mathcal{X}}/\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}) (see (6.1.3)). As a quasi-cone, it defines a closed substack 𝔇¯\overline{\mathfrak{D}} in 𝒫(4,6)(𝒱4𝒱6)\mathcal{P}_{(4,6)}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}). Let 𝕌\mathbb{U} denote the open complement of 𝔇\mathfrak{D} in 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}.

Remark (7.1.4).

Note that with fixed (𝒞,)(\mathcal{C},\mathcal{L}), the formation of 𝒱4,𝒱6,𝒳~,𝒳\mathcal{V}_{4},\mathcal{V}_{6},\widetilde{\mathcal{X}},\mathcal{X} and 𝕌\mathbb{U} naturally commutes with base change among BB-schemes. We will implicitly use this for the rest of Sec. 7. However, a priori 𝔇\mathfrak{D} and 𝔇¯\overline{\mathfrak{D}} might not commute with non-flat base change as they can become non-reduced (cf. (6.1.4)). Much of Sec. 7 is devouted to giving conditions to exclude this possibility. The key intermediate result is (7.3.4), which will play an important role in the proof of Theorem A.

We remark that 𝒳\mathcal{X} is only “stacky” because of the base.

Lemma (7.1.5).

Let TT be an BB-scheme and ν:T𝒫(4,6)(𝒱4𝒱6)\nu\colon T\to\mathcal{P}_{(4,6)}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}) be a morphism. Then the pullback ν𝒬(μ)\nu^{*}\mathcal{Q}(\mu), and hence ν𝒳\nu^{*}\mathcal{X}, are flat projective schemes over TT.

Proof.

The reader can check that 𝒬(μ)\mathcal{Q}(\mu) is in fact a 2\mathbb{P}^{2}-bundle over 𝒫(4,6)(𝒱4𝒱6)×B𝒞\mathcal{P}_{(4,6)}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})\times_{B}\mathcal{C}. Therefore, the pullback ν𝒬(μ)\nu^{*}\mathcal{Q}(\mu) is a 2\mathbb{P}^{2}-bundle over the scheme T×B𝒞T\times_{B}\mathcal{C}. Being a closed substack of the scheme ν𝒬(μ)\nu^{*}\mathcal{Q}(\mu), ν𝒳\nu^{*}\mathcal{X} has to be a projective scheme. The flatness of ν𝒬(μ)\nu^{*}\mathcal{Q}(\mu) is clear, and one deduces the flatness of ν𝒳\nu^{*}\mathcal{X} using that it is locally cut out by a single equation, and its geometric fibers are of codimension 11 (cf. [stacks-project, 00MF]). ∎

Proposition (7.1.6).

The morphism 𝒳~B\widetilde{\mathcal{X}}\to B is smooth.

Proof.

As 𝒳~\widetilde{\mathcal{X}} is flat over BB, it suffices to check smoothness of geometric fibers. Hence we may assume that B=Spec(k)B=\mathrm{Spec}(k), where kk is an algebraically closed field of characteristic 2,3\neq 2,3. Let us simply write 𝔸\mathbb{A}^{*} for 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}. Choose a point u𝔸(k)u\in\mathbb{A}^{*}(k) and c𝒞(k)c\in\mathcal{C}(k). Choose a uniformizer tt of CC at cc and bases {σi},{θj}\{\sigma_{i}\},\{\theta_{j}\} for 𝒱4,𝒱6\mathcal{V}_{4},\mathcal{V}_{6} respectively. Then the formal complection of 𝔸×C\mathbb{A}^{*}\times C at (u,c)(u,c) can be identified with Spf(R)\mathrm{Spf}(R), where R=k[[t,α0,,α4hg,β0,,β6hg]]R=k[\![t,\alpha_{0},\cdots,\alpha_{4h-g},\beta_{0},\cdots,\beta_{6h-g}]\!].

By choosing a local 𝒪C\mathcal{O}_{C}-generator of LL at cc, we turn σi\sigma_{i}’s and θj\theta_{j}’s into elements in k[[t]]k[\![t]\!]. Let ({ai},{bj})k10h2g+2(\{a_{i}\},\{b_{j}\})\in k^{10h-2g+2} be the affine coordinates of uu in 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}). Then the restriction of 𝒳~\widetilde{\mathcal{X}} to Spf(R)\mathrm{Spf}(R) can be identified with the subscheme of ProjR[x,y,z]\mathrm{Proj\,}R[x,y,z] defined by the equation

Wy2zx3+(i=04hg(ai+αi)σi(t))xz2+(j=06hg(bj+βj)θj(t))z3=0.W\coloneqq y^{2}z-x^{3}+(\sum_{i=0}^{4h-g}(a_{i}+\alpha_{i})\sigma_{i}(t))xz^{2}+(\sum_{j=0}^{6h-g}(b_{j}+\beta_{j})\theta_{j}(t))z^{3}=0. (29)

Let rr be the special point of Spf(R)\mathrm{Spf}(R). The singularity of the (generalized) elliptic curve 𝒳~t\widetilde{\mathcal{X}}_{t} defined by the above equation when tt, αi\alpha_{i}’s and βj\beta_{j}’s all vanish cannot appear on the z=0z=0 chart. So we may set z=1z=1 in the above equation and consider the resulting scheme in SpecR[x,y]\mathrm{Spec}R[x,y]. Since θj(0)0\theta_{j}(0)\neq 0 for some jj, we deduce that the partial derivative W/βj\partial W/\partial\beta_{j} remains nonzero on the special fiber. This implies that the total space of the restriction of 𝒳~\widetilde{\mathcal{X}} to Spf(R)\mathrm{Spf}(R) is smooth. But the choice of (u,c)(u,c) is arbitrary, so 𝒳~\widetilde{\mathcal{X}} is smooth. ∎

7.2   Nonlinear Bertini Theorems for Families of Elliptic Surfaces

In this section, kk remains an algebraically closed field of characteristic 2,3\neq 2,3.

Proposition (7.2.1).

Let CC be a smooth projective curve of genus gg over kk and let LL be a line bundle on CC with degree hh. Set VrH0(Lr)V_{r}\coloneqq\mathrm{H}^{0}(L^{r}) for every rr\in\mathbb{N}. For every dd\in\mathbb{N}, consider the closed subset of 𝔸(V4V6)×C\mathbb{A}(V_{4}\oplus V_{6})\times C defined by (ΔV12\Delta\in V_{12} is defined by 4a4327a624a_{4}^{3}-27a_{6}^{2} as before)

𝒦d{(a4,a6,c)V4×V6×Cvalc(Δ)d}\mathcal{K}_{d}\coloneqq\{(a_{4},a_{6},c)\in V_{4}\times V_{6}\times C\mid\mathrm{val}_{c}(\Delta)\geq d\}

and endow it with the reduced subscheme structure. Likewise, let DC×CD\subseteq C\times C be the diagonal and define a closed subscheme in 𝔸(V4V6)×(C×CD)\mathbb{A}(V_{4}\oplus V_{6})\times(C\times C-D) by

𝒦2+{(a4,a6,c,c)V4×V6×(C×CD)valc(Δ) and valc(Δ) are both 2}.\mathcal{K}_{2}^{+}\coloneqq\{(a_{4},a_{6},c,c^{\prime})\in V_{4}\times V_{6}\times(C\times C-D)\mid\mathrm{val}_{c}(\Delta)\text{ and }\mathrm{val}_{c^{\prime}}(\Delta)\text{ are both }\geq 2\}.

If 2hg+12h\geq g+1, then we have the following:

  1. (a)

    𝒦d\mathcal{K}_{d} has codimension dd for d3d\leq 3.

  2. (b)

    𝒦2\mathcal{K}_{2} has two irreducible components 𝒦2(I2)\mathcal{K}_{2}(\mathrm{I}_{2}) and 𝒦2(II)\mathcal{K}_{2}(\mathrm{II}) characterized by conditions valc(a6)=0\mathrm{val}_{c}(a_{6})=0 and valc(a6)1\mathrm{val}_{c}(a_{6})\geq 1 respectively.

  3. (c)

    𝒦2+\mathcal{K}_{2}^{+} has codimension 44.

Proof.

Recall that by the Riemann-Roch theorem, for any line bundle MM on CC, if deg(M)2g1\mathrm{deg}(M)\geq 2g-1, then h0(M)=deg(M)g+1h^{0}(M)=\mathrm{deg}(M)-g+1; if deg(M)=2g2\mathrm{deg}(M)=2g-2, then h0(M)=deg(M)g+1h^{0}(M)=\mathrm{deg}(M)-g+1 unless MωCM\,{\cong}\,\omega_{C}.

Fix any point cCc\in C and consider the projection 𝒦dC\mathcal{K}_{d}\to C. It suffices to show that the fiber 𝒦d,c\mathcal{K}_{d,c} over cc, viewed naturally as a closed subscheme of 𝔸(V4V6)\mathbb{A}(V_{4}\oplus V_{6}), has codimension dd. We identify the completion of CC along cc with Spf(k[[t]])\mathrm{Spf}(k[\![t]\!]) by choosing a uniformizer tt. After choosing a local generator of LL, we may consider the Taylor series of any σH0(Lr)\sigma\in\mathrm{H}^{0}(L^{r}), which is a power series σ(t)k[[t]]\sigma(t)\in k[\![t]\!]. By the first paragraph, for r4r\geq 4 and d3d\leq 3, we have

h0(Lr((1d)c))=h0(Lr(dc))+1.h^{0}(L^{r}((1-d)c))=h^{0}(L^{r}(-dc))+1.

Therefore, we may choose a basis σ0,,σ4hg\sigma_{0},\cdots,\sigma_{4h-g} for V4V_{4} such that valc(σi)=0\mathrm{val}_{c}(\sigma_{i})=0 for i=0,1,2i=0,1,2 and {σi}3i4hg\{\sigma_{i}\}_{3\leq i\leq 4h-g} forms a basis for H0(L4(3c))\mathrm{H}^{0}(L^{4}(-3c)). We may assume that σ0(t)1,σ1(t)t\sigma_{0}(t)\equiv 1,\sigma_{1}(t)\equiv t and σ2(t)t2\sigma_{2}(t)\equiv t^{2} modulo t3t^{3}. We choose a basis {θ0,,θ6hg}\{\theta_{0},\cdots,\theta_{6h-g}\} in an entirely similar way.

With the given choices of bases, we use {αi}\{\alpha_{i}\} and {βj}\{\beta_{j}\} for the coordinates of V4V_{4} and V6V_{6} respectively, so that Δ\Delta can be expressed as

Δ=4(i=04hgαiσi)327(j=06hgβjθj)2.\Delta=4\left(\sum_{i=0}^{4h-g}\alpha_{i}\sigma_{i}\right)^{3}-27\left(\sum_{j=0}^{6h-g}\beta_{j}\theta_{j}\right)^{2}. (30)

Then the fiber 𝒦d,c\mathcal{K}_{d,c} (d3d\leq 3) is cut out in 𝔸(V4V6)\mathbb{A}(V_{4}\oplus V_{6}) by the first dd equations from below:

{Δ(0)=4α0327β02Δ(0)=3(4α02α118β0β1)Δ(0)=24(α0α12+α02α2)54(β12+2β0β2)\displaystyle\begin{cases}\Delta(0)&=4\alpha_{0}^{3}-27\beta_{0}^{2}\\ \Delta^{\prime}(0)&=3(4\alpha_{0}^{2}\alpha_{1}-18\beta_{0}\beta_{1})\\ \Delta^{\prime\prime}(0)&=24(\alpha_{0}\alpha_{1}^{2}+\alpha_{0}^{2}\alpha_{2})-54(\beta_{1}^{2}+2\beta_{0}\beta_{2})\end{cases} (31)

The statement (a) is clear for d=0,1d=0,1. For d=2d=2, it is clear that 𝒦2\mathcal{K}_{2} contains the following subscheme

𝒦2(II){(a4,a6,c)V4×V6×Cvalc(a4)1,valc(a6)1},\mathcal{K}_{2}(\mathrm{II})\coloneqq\{(a_{4},a_{6},c)\in V_{4}\times V_{6}\times C\mid\mathrm{val}_{c}(a_{4})\geq 1,\mathrm{val}_{c}(a_{6})\geq 1\},

such that the fiber of 𝒦2(II)\mathcal{K}_{2}(\mathrm{II}) over cc is simply cut out by α0=β0=0\alpha_{0}=\beta_{0}=0. Let 𝔸2𝔸(kσ0,θ0)\mathbb{A}^{2}\coloneqq\mathbb{A}(k\langle\sigma_{0},\theta_{0}\rangle) be the affine space with coordinates (α0,β0)(\alpha_{0},\beta_{0}) and C𝔸2C^{\prime}\subseteq\mathbb{A}^{2} be the cuspidal curve defined by Δ(0)=0\Delta(0)=0. Then the fiber of 𝒦2𝒦2(II)\mathcal{K}_{2}-\mathcal{K}_{2}(\mathrm{II}) over a point in C{(0,0)}C^{\prime}-\{(0,0)\} is given by a codimension 11 hyperplane in 𝔸(kσi,βji,j1)\mathbb{A}(k\langle\sigma_{i},\beta_{j}\rangle_{i,j\geq 1}). This implies that 𝒦2𝒦2(II)\mathcal{K}_{2}-\mathcal{K}_{2}(\mathrm{II}) is irreducible of codimension 22 in 𝔸(V4V6)\mathbb{A}(V_{4}\oplus V_{6}), and we denote this component by 𝒦2(I2)\mathcal{K}_{2}(\mathrm{I}_{2}). Note that this implies (b). To see the d=3d=3 case for (a), just note that Δ(0)\Delta^{\prime\prime}(0) does not vanish identically on both 𝒦2(I2)\mathcal{K}_{2}(\mathrm{I}_{2}) and 𝒦2(II)\mathcal{K}_{2}(\mathrm{II}).

Finally we treat (c). We consider the projection Φ:𝒦2+(C×CD)\Phi\colon\mathcal{K}_{2}^{+}\to(C\times C-D) and take a point (c,c)(C×CD)(c,c^{\prime})\in(C\times C-D). Denote the fiber of Φ\Phi over (c,c)(c,c^{\prime}) by Φ(c,c)\Phi_{(c,c^{\prime})}. We assume first that L4ωC(2c+2c)L^{4}\not\cong\omega_{C}(2c+2c^{\prime}). This condition is automatically satisfied when 2h>g+12h>g+1 and ensures that h0(Lr(2c2c))=rh+g5h^{0}(L^{r}(-2c-2c^{\prime}))=rh+g-5 for r4r\geq 4. Then we may choose σ0,,σ3V4\sigma_{0},\cdots,\sigma_{3}\in V_{4} with the following vanishing orders:

σ0\sigma_{0} σ1\sigma_{1} σ2\sigma_{2} σ3\sigma_{3}
valc\mathrm{val}_{c} 0 11 2\geq 2 2\geq 2
valc\mathrm{val}_{c^{\prime}} 2\geq 2 2\geq 2 0 11
(32)

Then we complete {σ0,,σ3}\{\sigma_{0},\cdots,\sigma_{3}\} to a basis {σi}\{\sigma_{i}\} of V4V_{4} by adjoining a basis for H0(L4(2c2c))\mathrm{H}^{0}(L^{4}(-2c-2c^{\prime})). Let t,st,s be uniformizers of the completions of CC along cc and cc^{\prime} respectively. After choosing local generators of LL, we may consider Taylor series σi(t)k[[t]]\sigma_{i}(t)\in k[\![t]\!] and σi(s)k[[s]]\sigma_{i}(s)\in k[\![s]\!], and assume that σ0(t)1,σ1(t)tmodt2\sigma_{0}(t)\equiv 1,\sigma_{1}(t)\equiv t\mod t^{2} and σ2(s)1,σ3(s)smods2\sigma_{2}(s)\equiv 1,\sigma_{3}(s)\equiv s\mod s^{2}. Choose an entirely similar basis {θj}\{\theta_{j}\} for V6V_{6} and express Δ\Delta again as in (30). Then the defining equations for Φ(c,c)\Phi_{(c,c^{\prime})} in 𝔸(V4V6)\mathbb{A}(V_{4}\oplus V_{6}) are

{4α0327β02=4α2327β22=03(4α02α118β0β1)=3(4α22α318β2β3)=0\displaystyle\begin{cases}&4\alpha_{0}^{3}-27\beta_{0}^{2}=4\alpha_{2}^{3}-27\beta_{2}^{2}=0\\ &3(4\alpha_{0}^{2}\alpha_{1}-18\beta_{0}\beta_{1})=3(4\alpha_{2}^{2}\alpha_{3}-18\beta_{2}\beta_{3})=0\end{cases} (33)

By the same argument for the d=2d=2 case in (a), the above equations define a codimension 44 subscheme. The point is that the variables with indices 0,10,1 do not interfere with those with 2,32,3.

It remains to deal with the case when 2h=g+12h=g+1 and L4ωC(2c+2c)L^{4}\,{\cong}\,\omega_{C}(2c+2c^{\prime}). Note that in this case g1g\geq 1, so the condition L4ωC(2c+2c)L^{4}\,{\cong}\,\omega_{C}(2c+2c^{\prime}) defines a closed subscheme of (C×CD)(C\times C-D) of codimension at least 11. Therefore, it is enough to show that the codimension of Φ(c,c)\Phi_{(c,c^{\prime})} is at least 33. Note that we are able to choose a basis {θj}\{\theta_{j}\} just as before, but this time choose {σ0,σ1,σ2}\{\sigma_{0},\sigma_{1},\sigma_{2}\} with the following vanishing orders:

σ0\sigma_{0} σ1\sigma_{1} σ2\sigma_{2}
valc\mathrm{val}_{c} 0 11 22
valc\mathrm{val}_{c^{\prime}} 22 11 0
(34)

and complete it to a basis of V4V_{4} by adjoining a basis of H0(L4(2c2c))H^{0}(L^{4}(-2c-2c^{\prime})). Assume that σ0(t)=1,σ1(t)=tmodt3\sigma_{0}(t)=1,\sigma_{1}(t)=t\mod t^{3} and σ2(s)=1mods\sigma_{2}(s)=1\mod s. Then the conditions valc(Δ)2\mathrm{val}_{c}(\Delta)\geq 2 and valc(Δ)1\mathrm{val}_{c^{\prime}}(\Delta)\geq 1 give us 33 equations which are necessarily satisfied by Φ(c,c)\Phi_{(c,c^{\prime})}:

{4α0327β02=4α2327β22=03(4α02α118β0β1)=0\displaystyle\begin{cases}&4\alpha_{0}^{3}-27\beta_{0}^{2}=4\alpha_{2}^{3}-27\beta_{2}^{2}=0\\ &3(4\alpha_{0}^{2}\alpha_{1}-18\beta_{0}\beta_{1})=0\end{cases} (35)

It is clear that these indeed cut out a subscheme of codimension 33. ∎

Proposition (7.2.2).

Assume 2hg+12h\geq g+1. Apply set-up(7.1.3) to BSpec(k)B\coloneqq\mathrm{Spec\,}(k). The resulting discriminant 𝔇\mathfrak{D} is a proper subvariety of 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}. If codim𝔇=1\mathrm{codim\,}\mathfrak{D}=1, then 𝔇\mathfrak{D} has a unique irreducible component 𝔇0\mathfrak{D}_{0} of maximal dimension; moreover, for a general point aa on 𝔇0\mathfrak{D}_{0}, 𝒳~a\widetilde{\mathcal{X}}_{a} is smooth away from a single ODP.

Proof.

As B=Spec(k)B=\mathrm{Spec\,}(k), (𝒱4,𝒱6,𝒞)(\mathcal{V}_{4},\mathcal{V}_{6},\mathcal{C}) above is the same as (V4,V6,C)(V_{4},V_{6},C) in (7.2.1), and we use the notations from (7.2.1) and the results in (7.1.1) throughout the proof below.

Note that for a=(a4,a6)𝔸(V4V6)a=(a_{4},a_{6})\in\mathbb{A}(V_{4}\oplus V_{6})^{*} such that Δa=4a4327a63H0(L12)\Delta_{a}=4a_{4}^{3}-27a_{6}^{3}\in\mathrm{H}^{0}(L^{12}) does not vanish identically on CC, 𝒳~a\widetilde{\mathcal{X}}_{a} is singular if and only if its elliptic fibration has a reducible singular fiber. It is clear that 𝔇\mathfrak{D} is contained in the image of 𝒦2\mathcal{K}_{2} in 𝔸(V4V6)\mathbb{A}(V_{4}\oplus V_{6}), and hence has codimension at least 11. If codim𝔇=1\mathrm{codim\,}\mathfrak{D}=1, then by (7.2.1)(c), there exists an open dense subset U𝔇U\subseteq\mathfrak{D} such that if aUa\in U, 𝒳~a\widetilde{\mathcal{X}}_{a} has at most one singular fiber not of I1\mathrm{I}_{1}-type. If moreover this singular fiber is of II\mathrm{II}-type, then 𝒳~a\widetilde{\mathcal{X}}_{a} is smooth and a𝔇a\not\in\mathfrak{D}. Therefore, the only possible irreducible component of maximal dimension in 𝔇0\mathfrak{D}_{0} is the Zariski closure of the image of 𝒦2(I2)\mathcal{K}_{2}(\mathrm{I}_{2}). This implies the second statement in the proposition. ∎

7.3   Mod pp Behavior of Discriminants

Set-up (7.3.1).

Suppose that in (7.1.3) 𝒪B\mathcal{O}_{B} is a local ring, so that the vector bundles 𝒱4\mathcal{V}_{4} and 𝒱6\mathcal{V}_{6} are trivial 𝒪B\mathcal{O}_{B}-modules. By choosing 𝒪B\mathcal{O}_{B}-generators for 𝒱4\mathcal{V}_{4} and 𝒱6\mathcal{V}_{6}, we identify 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}) with 𝔸d1𝔸d2\mathbb{A}^{d_{1}}\oplus\mathbb{A}^{d_{2}}, where d1=4hg+1d_{1}=4h-g+1 and d2=6hg+1d_{2}=6h-g+1. Assume that 2hg+12h\geq g+1 as in the results in §7.2. Consider 1=Proj𝒪B[u,v]\mathbb{P}^{1}=\mathrm{Proj\,}\mathcal{O}_{B}[u,v]. Let 𝔸1=Spec(𝒪B[u])\mathbb{A}^{1}=\mathrm{Spec\,}(\mathcal{O}_{B}[u]) be the v=1v=1 chart on 1\mathbb{P}^{1}, and let :B1\infty:B\to\mathbb{P}^{1} denote the section defined by v=0v=0. Let 𝒲r\mathcal{W}_{r} be the 𝒪B\mathcal{O}_{B}-module of degree rr homogenous polynomials in 𝒪B[u,v]\mathcal{O}_{B}[u,v] or equivalently the module of degree r\leq r polynomials in 𝒪B[u]\mathcal{O}_{B}[u]. Consider the open subscheme T𝔸(𝒲4d1𝒲6d2)T\subseteq\mathbb{A}(\mathcal{W}_{4}^{d_{1}}\oplus\mathcal{W}_{6}^{d_{2}}) consisting of the points of the form

{(f1,,fd1,g1,,gd2)the common vanishing locus V({fi,gj})=}.\{(f_{1},\cdots,f_{d_{1}},g_{1},\cdots,g_{d_{2}})\mid\text{the common vanishing locus\,}V(\{f_{i},g_{j}\})=\emptyset\}.

Then it is clear that there is a natural morphism 1B×BT=1T𝒫(4,6)(𝒱4𝒱6)\mathbb{P}^{1}_{B}\times_{B}T=\mathbb{P}^{1}_{T}\to\mathcal{P}_{(4,6)}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}). By setting v=1v=1 in the polynomials fif_{i}’s and gjg_{j}’s, we also obtain an BB-morphism 𝔸1B×BT=𝔸1T𝔸(𝒱4𝒱6)\mathbb{A}^{1}_{B}\times_{B}T=\mathbb{A}^{1}_{T}\to\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}. Recall that 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*} denotes the affine space 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}) minus the zero section. This morphism fits into a commutative diagram

𝔸1T𝔸(𝒱4𝒱6)1T𝒫(4,6)(𝒱4𝒱6)φφ¯.\leavevmode\hbox to105.99pt{\vbox to56.18pt{\pgfpicture\makeatletter\hbox{\hskip 52.99724pt\lower-28.13954pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-52.99724pt}{-28.0397pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 9.45555pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-5.15001pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\mathbb{A}^{1}_{T}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 9.45555pt\hfil&\hfil\hskip 51.53334pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-23.22783pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 27.53337pt\hfil\cr\vskip 18.00005pt\cr\hfil\hskip 9.10832pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.80278pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\mathbb{P}^{1}_{T}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 9.10832pt\hfil&\hfil\hskip 55.54167pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-27.23616pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\mathcal{P}_{(4,6)}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 31.5417pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-33.88614pt}{10.91753pt}\pgfsys@lineto{-6.6778pt}{10.91753pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-6.47781pt}{10.91753pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-22.37157pt}{14.63138pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\varphi}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{}{{}}\pgfsys@moveto{-43.54169pt}{1.20457pt}\pgfsys@lineto{-43.54169pt}{-15.14227pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{1.0}{1.0}{0.0}{-43.54169pt}{1.20457pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-43.54169pt}{-15.34225pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-41.18892pt}{-6.44887pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{21.45554pt}{2.05782pt}\pgfsys@lineto{21.45554pt}{-16.28003pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{21.45554pt}{-16.48001pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{23.8083pt}{-7.31108pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-34.23337pt}{-25.5397pt}\pgfsys@lineto{-10.68613pt}{-25.5397pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-10.48615pt}{-25.5397pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-24.75977pt}{-23.18694pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\overline{\varphi}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}. (36)

For the content below, recall (6.1.1) and definition of 𝒳~,𝒳,𝔇,𝔇¯\widetilde{\mathcal{X}},\mathcal{X},\mathfrak{D},\overline{\mathfrak{D}} and 𝕌\mathbb{U} in (7.1.3). Assume that kk is an algebraically closed field of characteristic 2,3\neq 2,3.

Proposition (7.3.2).

Suppose that B=Spec(k)B=\mathrm{Spec\,}(k). Then the family 𝔸1T/T\mathbb{A}^{1}_{T}/T strongly dominates 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*} via φ\varphi. Moreover, for a general point tTt\in T, φ¯t()𝔇¯\overline{\varphi}_{t}(\infty)\not\in\overline{\mathfrak{D}}.

Proof.

The first statement is an exercise of dimension counting, so we omit the details. For the second statement, it suffices to exhibit a single such tt as the condition is open. Note that the automorphism group of 1\mathbb{P}^{1} naturally acts on TT. We start with any point tTt^{\prime}\in T be such that for some point w𝔸1w\in\mathbb{A}^{1}, φt(w)𝔇\varphi_{t}(w)\not\in\mathfrak{D}. Then we can always apply an automorphism of 1\mathbb{P}^{1} to switch ww and \infty. This gives us a point tTt\in T and by construction φ¯t()𝔇¯\overline{\varphi}_{t}(\infty)\not\in\overline{\mathfrak{D}}. ∎

Remark (7.3.3).

We expect that 𝔇\mathfrak{D} to always be of codimension exactly 11 in 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}, but not smaller. We make a simple observation for a case when this is easily guaranteed: Suppose that k=k=\mathbb{C} in (7.3.2). If for a general point tt on TT, the H2H^{2} of the restriction of 𝒳~\widetilde{\mathcal{X}} to φt(𝕌)\varphi^{*}_{t}(\mathbb{U}) defines a non-isotrivial VHS, then codim𝔇=1\mathrm{codim\,}\mathfrak{D}=1. Indeed, otherwise for a general tt, the family φ¯t(𝒳)\overline{\varphi}_{t}^{*}(\mathcal{X}) is a smooth family over 1\mathbb{P}^{1}_{\mathbb{C}}, which cannot support a non-isotrivial VHS.

Lemma (7.3.4).

Suppose that chark=p5\mathrm{char\,}k=p\geq 5 and in set-up(7.3.1) BB is taken to be Spec(W)\mathrm{Spec\,}(W) for W:=W(k)W:=W(k). Assume that 𝔇K\mathfrak{D}_{K} has codimension 11 in 𝔸(𝒱4𝒱6)K\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}_{K} over K:=W[1/p]K:=W[1/p]. Then for a general kk-point tTkt\in T_{k}, φ¯t\overline{\varphi}_{t} has the following properties:

  1. (a)

    φ¯t()𝔇¯k\overline{\varphi}_{t}(\infty)\not\in\overline{\mathfrak{D}}_{k}, the total space of the family φ¯t(𝒳)1k\overline{\varphi}^{*}_{t}(\mathcal{X})\to\mathbb{P}^{1}_{k} is smooth and every fiber has at most a single ODP singularity.

  2. (b)

    φ¯t(𝔇¯k)\overline{\varphi}^{*}_{t}(\overline{\mathfrak{D}}_{k}), or equivalently φt(𝔇k)\varphi_{t}^{*}(\mathfrak{D}_{k}), is reduced.

  3. (c)

    For every t~T(W)\widetilde{t}\in T(W) lifting tt, φt~(𝔇)\varphi_{\widetilde{t}}^{*}(\mathfrak{D}) is étale over WW, so that the open subcurve φt~(𝕌)𝔸1t~\varphi_{\widetilde{t}}^{*}(\mathbb{U})\subseteq\mathbb{A}^{1}_{\widetilde{t}} has a good compactification relative to WW.

Proof.

Since 𝔇¯\overline{\mathfrak{D}} is proper over WW, it is not hard to see that 𝔇k\mathfrak{D}_{k} is also of codimension 11 in 𝔸(𝒱4𝒱6)k\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}_{k}. We break into 3 steps.
Step 1: By (7.2.2) and (6.1.4), the irreducible components of maximal dimension of 𝔇K\mathfrak{D}_{K} and (𝔇k)red(\mathfrak{D}_{k})_{\mathrm{red}} are unique. Let us denote them by 𝔇K\mathfrak{D}_{K}^{\circ} and 𝔇k\mathfrak{D}_{k}^{\circ} respectively. Moreover, by (7.2.2), there exists an open dense subscheme V𝔇kV\subseteq\mathfrak{D}^{\circ}_{k}, such that for every vV(k)v\in V(k), 𝒳v\mathcal{X}_{v} has at most a single ODP singularity. In particular, by (7.3.2), as well as (6.1.5) and its proof, for a general point tT(k)t\in T(k), φ¯t()𝔇¯k\overline{\varphi}_{t}(\infty)\not\in\overline{\mathfrak{D}}_{k}, the total space of the pullback family φt(𝒳)\varphi_{t}^{*}(\mathcal{X}) is smooth, and the image of φt\varphi_{t} only intersects 𝔇k\mathfrak{D}^{\circ}_{k} on VV, so that every singular fiber of φt(𝒳)\varphi_{t}^{*}(\mathcal{X}) over 𝔸1t\mathbb{A}^{1}_{t} has a single ODP; moreover, as φ¯t()𝔇¯k\overline{\varphi}_{t}(\infty)\not\in\overline{\mathfrak{D}}_{k}, the total space of φ¯t(𝒳)\overline{\varphi}_{t}^{*}(\mathcal{X}) is smooth. Hence we may conclude (a).
Step 2: Next, we show the following claim ()(\star): If tt is a general point, for any t~T(W)\widetilde{t}\in T(W) lifting tt, Z:=φt~(𝔇)Z:=\varphi_{\widetilde{t}}^{*}(\mathfrak{D}) is flat over WW. Let 𝔇0,,𝔇r\mathfrak{D}_{0},\cdots,\mathfrak{D}_{r} be the irreducible components of 𝔇\mathfrak{D} such that 𝔇0\mathfrak{D}_{0} is the component which contains 𝔇K\mathfrak{D}_{K}^{\circ}. We claim that 𝔇0\mathfrak{D}_{0} contains 𝔇k\mathfrak{D}^{\circ}_{k} as well. To simplify notation, let us write 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}) as 𝒜\mathscr{A}. Then 𝔇0,K𝒜K\mathfrak{D}_{0,K}\subseteq\mathscr{A}_{K}^{*} is cut out by a single polynomial FF in the coordinates of the affine space 𝒜K\mathscr{A}_{K}. By minimally clearing denominators, we may assume that the coefficients of FF are defined in WW and generate WW. Using the fact that 𝒜\mathscr{A} is affine and 𝒪𝒜\mathcal{O}_{\mathscr{A}} is a UFD, one checks that the Zariski closure of 𝔇0,K\mathfrak{D}_{0,K} in 𝒜\mathscr{A} contains the vanishing locus V(F)V(F) of F𝒪𝒜F\in\mathcal{O}_{\mathscr{A}}. Note that FF is weighted-homogeneous, so V(F)k𝒜kV(F)_{k}\subseteq\mathscr{A}_{k} at least contains the origin. In paricular, V(F)BV(F)\to B is surjective. By [stacks-project, Tag 0B2J], dimV(F)k=dimV(F)K\dim V(F)_{k}=\dim V(F)_{K}. This implies that dim𝔇0,k=dim𝔇0,K\dim\mathfrak{D}_{0,k}=\dim\mathfrak{D}_{0,K}. By the uniqueness of 𝔇k\mathfrak{D}_{k}^{\circ} as an irreducible component of maximal dimension, we conclude that 𝔇k𝔇0\mathfrak{D}_{k}^{\circ}\subseteq\mathfrak{D}_{0}. By applying [stacks-project, Tag 0B2J] again, we also conclude that for any i>0i>0, 𝔇i,k\mathfrak{D}_{i,k} has codimension 2\geq 2 in 𝒜k\mathscr{A}^{*}_{k}.

Set U𝔇0U\subseteq\mathfrak{D}_{0} be the complement of the closed subscheme i>1(𝔇0𝔇i)\cup_{i>1}(\mathfrak{D}_{0}\cap\mathfrak{D}_{i}). Then (Uk)red(U_{k})_{\mathrm{red}} is dense in 𝔇k\mathfrak{D}_{k}^{\circ}. As tt is general and the family 𝔸1Tk\mathbb{A}^{1}_{T_{k}} strongly dominates 𝔸(𝒱4𝒱6)k\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})_{k}, we may assume that the intersection im(φt)(𝔇k)red\mathrm{im}(\varphi_{t})\cap(\mathfrak{D}_{k})_{\mathrm{red}} is transverse and lies in UkU_{k}. Now we can prove the claim ()(\star). Indeed, note that 𝔇0\mathfrak{D}_{0} is a Weil divisor, and hence also a Cartier divisor of 𝒜\mathscr{A}^{*}, as 𝒜\mathscr{A}^{*} is regular. This implies that Z=φt~(𝔇0)Z=\varphi^{*}_{\widetilde{t}}(\mathfrak{D}_{0}) is everywhere locally cut out in 1W\mathbb{P}^{1}_{W} by a single equation. Since Zk1kZ_{k}\subseteq\mathbb{P}^{1}_{k} is of codimension 11, ZZ is flat over WW by [stacks-project, Tag 00MF].
Step 3: Finally, we show (b) and (c) simultaneously. Note that if we show (c) for some t~\widetilde{t}, then we can already conclude (b), which conversely implies (c) for all t~\widetilde{t}. Indeed, 𝒪Z\mathcal{O}_{Z} is isomorphic to W[x]/(f(x))W[x]/(f(x)) for some f(x)pW[x]f(x)\not\in pW[x]. If Zk=φ(𝔇k)Z_{k}=\varphi^{*}(\mathfrak{D}_{k}) is reduced, then by Hensel’s lemma ZZ has to be ad disjoint union of several copies of WW. Hence it suffices to show (c) for some t~\widetilde{t}, for which we may assume that the generic fiber φt~K\varphi_{\widetilde{t}}{\otimes}K intersects 𝔇K\mathfrak{D}_{K} transversely. What we are using here is that 𝔸1TK¯\mathbb{A}^{1}_{T}{\otimes}\bar{K} strongly dominates 𝔸(𝒱4𝒱6)K¯\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}_{\bar{K}} and the points on T(W)T(W) which lift tt are Zariski dense on TKT_{K}.

Since ZZ is finite and flat, degKZK=degkZk\mathrm{deg}_{K}\,Z_{K}=\mathrm{deg}_{k}\,Z_{k}. As ZKZ_{K} is reduced, degZK\mathrm{deg}Z_{K} is the number of closed points on ZK¯Z_{\bar{K}}. Therefore, to show that ZkZ_{k} is reduced, it suffices to show that ZkZ_{k} has the same number of closed points, i.e., |Zk(k)|=|ZK¯(K¯)||Z_{k}(k)|=|Z_{\bar{K}}(\bar{K})|. Now we compute |Zk(k)||Z_{k}(k)| by topology. Choose a prime p\ell\neq p. By combining the Leray spectral sequence and the Grothendieck-Ogg-Shafarevich formula, we have

χ(φ¯t(𝒳))=χ(𝒳η¯t)χ(1k)+zZk(k)(χ(𝒳z)χ(𝒳η¯t)swz(He´t(𝒳η¯t,)))\chi(\overline{\varphi}_{t}^{*}(\mathcal{X}))=\chi(\mathcal{X}_{\overline{\eta}_{t}})\chi(\mathbb{P}^{1}_{k})+\sum_{z\in Z_{k}(k)}\left(\chi(\mathcal{X}_{z})-\chi(\mathcal{X}_{\overline{\eta}_{t}})-\mathrm{sw}_{z}(\mathrm{H}_{\mathrm{{\acute{e}}t}}^{*}(\mathcal{X}_{\overline{\eta}_{t}},\mathbb{Q}_{\ell}))\right)

where η¯t\overline{\eta}_{t} is a geometric generic point over 1t\mathbb{P}^{1}_{t}, and swz(He´t(𝒳η¯t,))\mathrm{sw}_{z}(\mathrm{H}_{\mathrm{{\acute{e}}t}}^{*}(\mathcal{X}_{\overline{\eta}_{t}},\mathbb{Q}_{\ell})) denotes the alternating sum of Swan conductors

swz(He´t(𝒳η¯t,))i=04(1)iswz(He´ti(𝒳η¯t,)).\mathrm{sw}_{z}(\mathrm{H}_{\mathrm{{\acute{e}}t}}^{*}(\mathcal{X}_{\overline{\eta}_{t}},\mathbb{Q}_{\ell}))\coloneqq\sum_{i=0}^{4}(-1)^{i}\mathrm{sw}_{z}(\mathrm{H}_{\mathrm{{\acute{e}}t}}^{i}(\mathcal{X}_{\overline{\eta}_{t}},\mathbb{Q}_{\ell})).

Since for every zZk(k)z\in Z_{k}(k), 𝒳z\mathcal{X}_{z} is smooth away from an ODP, [WeilI, §4.2] tells us that the local monodromy action on He´t(𝒳η¯t,)\mathrm{H}_{\mathrm{{\acute{e}}t}}^{*}(\mathcal{X}_{\overline{\eta}_{t}},\mathbb{Q}_{\ell}) factors through /2\mathbb{Z}/2\mathbb{Z} and hence is tamely ramified as p2p\neq 2. This implies that the Swan conductors all vanish. Moreover, by loc. cit. we also know that χ(𝒳z)χ(𝒳η¯t)=1\chi(\mathcal{X}_{z})-\chi(\mathcal{X}_{\overline{\eta}_{t}})=-1, so

|Zk(k)|=χ(𝒳η¯t)χ(1k)χ(φ¯t(𝒳)).|Z_{k}(k)|=\chi(\mathcal{X}_{\overline{\eta}_{t}})\chi(\mathbb{P}^{1}_{k})-\chi(\overline{\varphi}_{t}^{*}(\mathcal{X})). (37)

By considering how singularities might degenerate, we easily see that the fiber of 𝒳\mathcal{X} over each point in ZK¯(K¯)Z_{\bar{K}}(\bar{K}) has at most an ODP singularity. Therefore, by the same computation as above, for u:=t~K¯u:=\widetilde{t}_{\bar{K}} and a geometric generic point η¯u\overline{\eta}_{u} of 1u\mathbb{P}^{1}_{u}, we have

|ZK¯(K¯)|=χ(𝒳η¯u)χ(1K¯)χ(φ¯u(𝒳)).|Z_{\bar{K}}(\bar{K})|=\chi(\mathcal{X}_{\overline{\eta}_{u}})\chi(\mathbb{P}^{1}_{\bar{K}})-\chi(\overline{\varphi}_{u}^{*}(\mathcal{X})). (38)

Using the smooth and proper base change theorem for étale cohomology, it is not hard to see that χ(φ¯u(𝒳))=χ(φ¯t(𝒳))\chi(\overline{\varphi}_{u}^{*}(\mathcal{X}))=\chi(\overline{\varphi}_{t}^{*}(\mathcal{X})) and χ(𝒳η¯u)=χ(𝒳η¯t)\chi(\mathcal{X}_{\overline{\eta}_{u}})=\chi(\mathcal{X}_{\overline{\eta}_{t}}). Hence we conclude that |Zk(k)|=|ZK¯(K¯)||Z_{k}(k)|=|Z_{\bar{K}}(\bar{K})| as desired. ∎

Note that the discussion of Swan conductors fundamentally uses the p2p\neq 2 assumption. Of course it is irrelevant here because we are working with the p5p\geq 5, but we remark that in [Saitodisc, Thm 4.2] the non-reducedness of the relevant discriminant scheme modulo 22 can indeed be explained by the fact that ordinary quadratic singularities behave differently in characteristic 22.121212This was explained in the appendix in the earlier arXiv version of the paper, which will be published separately. We also remark that for results in §7.2 crucially uses the 2hg+12h\geq g+1 assumption, which is indeed satisfied by the case we care about (g=h=1g=h=1).

7.4   Proof of Theorem A

In this section we work with the following set-up:

Set-up (7.4.1).

Let 1,1\mathscr{M}_{1,1} be the moduli stack of the pair of a genus 11 curve together with a degree 11 line bundle (i.e., an elliptic curve). Let BB be the [1/6]\mathbb{Z}[1/6]-scheme defined by

{(a,b)Spec([16][a,b])4a327b20}.\{(a,b)\in\mathrm{Spec}(\mathbb{Z}[\frac{1}{6}][a,b])\mid 4a^{3}-27b^{2}\neq 0\}.

Then the Weierstrass equation equips BB with a surjective morphism B1,1B\to\mathscr{M}_{1,1}. Let (ϖ:𝒞B,)(\varpi:\mathcal{C}\to B,\mathcal{L}) be the restriction of the universal family over 1,1\mathscr{M}_{1,1} to BB. Apply the constructions in (7.1.3) with this triple (B,𝒞,)(B,\mathcal{C},\mathcal{L}) and define the objects 𝒱r=ϖr\mathcal{V}_{r}=\varpi_{*}\mathcal{L}^{r} (r=4,6r=4,6), 𝒳~𝔸(𝒱4𝒱6)\widetilde{\mathcal{X}}\to\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*} and 𝔇=Disc(𝒳~/𝔸(𝒱4𝒱6))\mathfrak{D}=\mathrm{Disc}(\widetilde{\mathcal{X}}/\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*}) accordingly. Below we write 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*} as 𝖬~\widetilde{\mathsf{M}}, the open subscheme 𝖬~𝔇\widetilde{\mathsf{M}}-\mathfrak{D} as 𝖬\mathsf{M}, and the restriction of 𝒳~\widetilde{\mathcal{X}} to 𝖬\mathsf{M} as 𝒳\mathcal{X}^{\circ}.

Remark (7.4.2).

For any algebraically closed field kk of characteristic p2,3p\neq 2,3 and elliptic surface XX over kk with pg=q=1p_{g}=q=1, there exists a point z𝖬~(k)z\in\widetilde{\mathsf{M}}(k) such that XX is the minimal model of 𝒳~z\widetilde{\mathcal{X}}_{z}. Moreover, there are no reducible fibers in the elliptic fibration of XX if and only if z𝖬(k)z\in\mathsf{M}(k). The choice of zz is unique up to the 𝔾m\mathbb{G}_{m}-action on 𝖬\mathsf{M} given by λ(a4,a6)=(λ4a4,λ6a6)\lambda\cdot(a_{4},a_{6})=(\lambda^{4}a_{4},\lambda^{6}a_{6}) where (a4,a6)(a_{4},a_{6}) is the relative coordinate on 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6}).

We need a lower bound on the rank of the Kodaira-Spencer map, or equivalently the image of the period morphism over \mathbb{C}.131313This bound is now obsolete due to a more recent preprint [EGW], but we keep it here to illustrate that the proof of (7.4.4) below works under much weaker inputs, in case the reader wishes to study other varieties (cf. (1.0.4)). Note also that the period map over \mathbb{C} in [EGW] is not flat, despite being generically étale (Cor. 5.1 ibid), and its mod pp properties are unclear, so computations in §7.2, 7.3 remain necessary for the current method.

Lemma (7.4.3).

For a general z𝖬z\in\mathsf{M}_{\mathbb{C}} and X𝒳zX\coloneqq\mathcal{X}_{z}, the Kodaira-Spencer map

Tz𝖬Hom(H1(Ω1X),H2(𝒪X))T_{z}\mathsf{M}_{\mathbb{C}}\to\mathrm{Hom}(\mathrm{H}^{1}(\Omega^{1}_{X}),\mathrm{H}^{2}(\mathcal{O}_{X}))

has rank at least 33.

Proof.

This follows a construction of Ikeda and the Artin-Brieskorn resolution. In [Ikeda], Ikeda constructed a subfamily of elliptic surfaces over \mathbb{C} with the given invariants pg=q=1p_{g}=q=1 using bielliptic curves of genus 33. Let C~\widetilde{C} be a bielliptic curve of genus 33, equipped with an involution σ\sigma such that C~/σ\widetilde{C}/\sigma is a smooth genus 11 curve CC. On the symmetric square C~(2)\widetilde{C}^{(2)} of C~\widetilde{C}, σ\sigma also lifts to an involution σ(2)\sigma^{(2)}. Consider the surface Y=C~(2)/σ(2)Y^{\prime}={\widetilde{C}}^{(2)}/\sigma^{(2)}, which is shown to be a projective surface of Kodaira dimension 11 with 66 ODPs. Its minimal resolution YY is an elliptic surface with pg=q=1p_{g}=q=1. By Prop. 2.9 in loc. cit., the morphism C~C\widetilde{C}\to C can be recovered from YY. Note however the Weierstrass model of YY is singular, so YY is not given by a point on 𝖬\mathsf{M}.

Note that for every \mathbb{C}-point s𝖬~s\in\widetilde{\mathsf{M}}_{\mathbb{C}} with image tt in BB_{\mathbb{C}}, ss is given by a pair (a4,a6)H0(t4)×H0(t6)(a_{4},a_{6})\in\mathrm{H}^{0}(\mathcal{L}_{t}^{4})\times\mathrm{H}^{0}(\mathcal{L}_{t}^{6}). Let 𝖬𝖬~\mathsf{M}^{\prime}_{\mathbb{C}}\subseteq\widetilde{\mathsf{M}}_{\mathbb{C}} be the open subscheme which consists of those ss such that Δs=4a4327a62\Delta_{s}=4a_{4}^{3}-27a_{6}^{2} does not vanish identically on the base curve 𝒞t\mathcal{C}_{t}. Then 𝒳~s\widetilde{\mathcal{X}}_{s} is the Weierstrass normal form of an elliptic surface and by [Kas, Thm 1] has at most rational double point singularities. By applying the Artin-Brieskorn resolution [Artin-Res] to 𝒳~|𝖬\widetilde{\mathcal{X}}|_{\mathsf{M}^{\prime}_{\mathbb{C}}}, we obtain a smooth and proper algebraic space 𝒳𝖬\mathcal{X}_{\mathbb{C}}^{\sharp}\to\mathsf{M}_{\mathbb{C}}^{\sharp}, where 𝖬\mathsf{M}_{\mathbb{C}}^{\sharp} is an algebraic space which admits a morphism to 𝖬\mathsf{M}^{\prime}_{\mathbb{C}} bijective on geometric points. Moreover, [Artin-Res, Thm 2] tells us that for any \mathbb{C}-point zz of 𝖬\mathsf{M}^{\sharp}_{\mathbb{C}} which maps to a point ss of 𝖬\mathsf{M}^{\prime}_{\mathbb{C}}, the Henselianization of 𝖬\mathsf{M}^{\sharp}_{\mathbb{C}} at zz maps surjectively to that of 𝖬\mathsf{M}^{\prime}_{\mathbb{C}} at ss. Let 𝖬+\mathsf{M}^{+}_{\mathbb{C}} be a resolution of singularities of 𝖬\mathsf{M}^{\sharp}_{\mathbb{C}} and pullback the family 𝒳|𝖬\mathcal{X}^{\sharp}|_{\mathsf{M}^{\prime}_{\mathbb{C}}} to 𝖬+\mathsf{M}^{+}_{\mathbb{C}}.

Note that all elliptic surfaces which can be constructed as in the first paragraph can be found as fibers of this family over 𝖬+\mathsf{M}^{+}_{\mathbb{C}}. Let Ω\Omega be the period domain parametrizing Hodge structures of K3-type on the integral lattice Λ\Lambda given by the Betti cohomology of any complex elliptic surface with pg=q=1p_{g}=q=1. Let 𝖬~+\widetilde{\mathsf{M}}^{+}_{\mathbb{C}} be the universal cover of 𝖬+\mathsf{M}^{+}_{\mathbb{C}}. Then up to an action of O(Λ)\mathrm{O}(\Lambda) there is a well defined period map 𝖬~+Ω\widetilde{\mathsf{M}}^{+}_{\mathbb{C}}\to\Omega. The moduli space of bielliptic curves of genus 33 over \mathbb{C} is 44-dimensional, so [Ikeda, Thm 1.1(1)] implies that the period image of 𝖬~+\widetilde{\mathsf{M}}^{+}_{\mathbb{C}} is of dimension 3\geq 3. Since 𝖬𝖬\mathsf{M}_{\mathbb{C}}\subseteq\mathsf{M}^{\prime}_{\mathbb{C}} is open and dense, the discussion on the Henselization of 𝖬\mathsf{M}^{\sharp}_{\mathbb{C}} at \mathbb{C}-points in the preceeding paragraph implies that the preimage of 𝖬𝖬\mathsf{M}_{\mathbb{C}}\subseteq\mathsf{M}^{\prime}_{\mathbb{C}} in 𝖬+\mathsf{M}^{+}_{\mathbb{C}} is also open and dense. This implies that the preimage of 𝖬\mathsf{M}_{\mathbb{C}} in 𝖬~+\widetilde{\mathsf{M}}^{+}_{\mathbb{C}} also has period image of dim3\dim\geq 3 by a continuity argument. ∎

We are now ready to prove (a more general form of) Theorem A:

Theorem (7.4.4).

Assume that kk is a field finitely generated over 𝔽p\mathbb{F}_{p} for p5p\geq 5 and let k¯\bar{k} be a separable closure. Let XX be an elliptic surface over kk with pg=q=1p_{g}=q=1. If all fibers in the elliptic fibration of Xk¯X_{\bar{k}} are irreducible, then the Tate conjecture holds for XX.

Proof.

Let η\eta be the generic point of 𝖬\mathsf{M}. Let 𝔣\mathfrak{f} and 𝔰\mathfrak{s} be the classes in NS(𝒳η)\mathrm{NS}(\mathcal{X}^{\circ}_{\eta}) such that 𝔣\mathfrak{f} (resp. 𝔰\mathfrak{s}) is given by a smooth fiber (resp. zero section) in the elliptic fibration of 𝒳η\mathcal{X}^{\circ}_{\eta}. We may polarize the family 𝒳/𝖬\mathcal{X}^{\circ}/\mathsf{M} with the subspace 𝔣,𝔰\mathbb{Q}\langle\mathfrak{f},\mathfrak{s}\rangle, as it contains the class of 𝝃η\boldsymbol{\xi}_{\eta} for some relatively ample line bundle 𝝃\boldsymbol{\xi}. Since 𝔣,𝔣=0,𝔣,𝔰=1\langle\mathfrak{f},\mathfrak{f}\rangle=0,\langle\mathfrak{f},\mathfrak{s}\rangle=1 and 𝔰,𝔰=1\langle\mathfrak{s},\mathfrak{s}\rangle=-1, one deduces that for any s𝖬()s\in\mathsf{M}(\mathbb{C}), the lattice PH2(𝒳s,(p))tf\mathrm{PH}^{2}(\mathcal{X}^{\circ}_{s},\mathbb{Z}_{(p)})_{\mathrm{tf}} is self-dual for every p5p\geq 5. Note that 𝖬\mathsf{M}_{\mathbb{C}} is clearly connected, and by (7.4.3) and (2.2.5), the family (𝒳/𝖬,𝝃)|𝖬(\mathcal{X}^{\circ}/\mathsf{M},\boldsymbol{\xi})|_{\mathsf{M}_{\mathbb{C}}} has maximal monodromy, as defined in (2.2.3). By (7.4.2), it suffices to apply (5.3.3)(a) to the family f:𝒳𝖬f:\mathcal{X}^{\circ}\to\mathsf{M}, for which we only need to prove that 𝕃:=R2f2\mathbb{L}:=R^{2}f_{*}\mathbb{Q}_{2} has constant λgeo\lambda^{\mathrm{geo}} over Spec([1/6])\mathrm{Spec\,}(\mathbb{Z}[1/6]). We do so in two steps:

Step 1: Fix a prime p5p\geq 5, set W:=W(𝔽¯p)W:=W(\bar{\mathbb{F}}_{p}), K:=W[1/p]K:=W[1/p] and choose an isomorphism K¯\bar{K}\,{\cong}\,\mathbb{C}. It suffices to show that there exists a smooth connected WW-curve CC with a morphism to 𝖬\mathsf{M} such that (recall (5.1.1) and (5.1.2))

  1. (i)

    λ(𝕃|C)=λ(𝕃|𝖬)\lambda(\mathbb{L}|_{C_{\mathbb{C}}})=\lambda(\mathbb{L}|_{\mathsf{M}_{\mathbb{C}}}), and

  2. (ii)

    CC has a good compactification over WW.

Assume the existence of such curves for the moment, we first show how the theorem follows. Note that 𝖬𝔽¯p\mathsf{M}_{\bar{\mathbb{F}}_{p}} and 𝖬\mathsf{M}_{\mathbb{C}} are both irreducible. Choose a base point s𝖬()s\in\mathsf{M}(\mathbb{C}) which lies over the generic point η\eta of 𝖬\mathsf{M}. Let sps_{p} be a geometric point over the generic point ηp\eta_{p} of 𝖬𝔽¯p\mathsf{M}_{\bar{\mathbb{F}}_{p}}. Since ss specializes to sps_{p}, dimNS(𝒳sp)dimNS(𝒳s)\dim\mathrm{NS}(\mathcal{X}^{\circ}_{s_{p}})_{\mathbb{Q}}\geq\dim\mathrm{NS}(\mathcal{X}^{\circ}_{s})_{\mathbb{Q}}; moreover, as every element in NS(𝒳sp)\mathrm{NS}(\mathcal{X}^{\circ}_{s_{p}})_{\mathbb{Q}} is stabilized by an open subgroup of π1e´t(𝖬𝔽¯p,sp)\pi_{1}^{\mathrm{{\acute{e}}t}}(\mathsf{M}_{\bar{\mathbb{F}}_{p}},s_{p}), λ(𝕃|𝖬𝔽¯p)dimNS(𝒳sp)\lambda(\mathbb{L}|_{\mathsf{M}_{\bar{\mathbb{F}}_{p}}})\geq\dim\mathrm{NS}(\mathcal{X}_{s_{p}})_{\mathbb{Q}}. On the other hand, as π1\pi_{1} is a functor, we have λ(𝕃|𝖬𝔽¯p)λ(𝕃|C𝔽¯p)\lambda(\mathbb{L}|_{\mathsf{M}_{\bar{\mathbb{F}}_{p}}})\leq\lambda(\mathbb{L}|_{C_{\bar{\mathbb{F}}_{p}}}) by default. But the curve CC has constant λgeo\lambda^{\mathrm{geo}} because it has a good relative compactification (5.1.3), so we have

dimNS(𝒳s)=(5.1.4)λ(𝕃|𝖬)=λ(𝕃|C)=λ(𝕃|C𝔽¯p)λ(𝕃|𝖬𝔽¯p).\dim\mathrm{NS}(\mathcal{X}^{\circ}_{s})_{\mathbb{Q}}\stackrel{{\scriptstyle\ref{prop: thm of fixed part}}}{{=}}\lambda(\mathbb{L}|_{\mathsf{M}_{\mathbb{C}}})=\lambda(\mathbb{L}|_{C_{\mathbb{C}}})=\lambda(\mathbb{L}|_{C_{\bar{\mathbb{F}}_{p}}})\geq\lambda(\mathbb{L}|_{\mathsf{M}_{\bar{\mathbb{F}}_{p}}}).

This implies that λ(𝕃|𝖬)=λ(𝕃|𝖬𝔽¯p)\lambda(\mathbb{L}|_{\mathsf{M}_{\mathbb{C}}})=\lambda(\mathbb{L}|_{\mathsf{M}_{\bar{\mathbb{F}}_{p}}}), i.e., 𝖬W\mathsf{M}_{W} has constant λgeo\lambda^{\mathrm{geo}}.

Step 2: It remains to construct the curve CC which satisfies (i) and (ii). Let 𝖵\mathsf{V} be the VHS on 𝖬\mathsf{M}_{\mathbb{C}} given by R2f(1)R^{2}f_{\mathbb{C}*}\mathbb{Q}(1). We say that an irreducible smooth \mathbb{C}-variety ZZ admitting an understood morphism to 𝖬\mathsf{M}_{\mathbb{C}} admissible if its image is not contained in the Noether-Lefschetz loci of 𝖵\mathsf{V} and the restriction 𝖵|Z\mathsf{V}|_{Z} is non-isotrivial. This implies, by (5.1.4), that λ(𝕃|Z)=λ(𝕃|𝖬)\lambda(\mathbb{L}|_{Z})=\lambda(\mathbb{L}|_{\mathsf{M}_{\mathbb{C}}}).

We construct CC in two steps. First, let bB(k)b\in B(k) be any point and B^b\widehat{B}_{b} be the formal completion of BWB_{W} at bb. We claim that for some b~B^b(W)\widetilde{b}\in\widehat{B}_{b}(W), 𝖬b~\mathsf{M}_{\widetilde{b}}{\otimes}\mathbb{C} is admissible. Indeed, we first observe that since dimB=2\dim B_{\mathbb{C}}=2, (7.4.3) implies that for a general point zBz\in B_{\mathbb{C}}, 𝖵|𝖬z\mathsf{V}|_{\mathsf{M}_{z}} is non-isotrivial. Note that the KK-points given by an analytically dense subset of B^b(W)\widehat{B}_{b}(W) are Zariski dense on BKB_{K} (and hence also on BK¯B_{\bar{K}}). As the Noether-Lefschetz (NL) loci of 𝖵\mathsf{V} is a countable union of proper closed subvarieties, we may now apply (6.2.1) to the case SS being 𝖬\mathsf{M} and NN being the NL loci to conclude that the desired lifting b~\widetilde{b} exists.

Next, we take the base BB in the context of (7.3.1) to be b~\widetilde{b} above (and hence 𝔸(𝒱4𝒱6)\mathbb{A}(\mathcal{V}_{4}\oplus\mathcal{V}_{6})^{*} and 𝕌\mathbb{U} in (7.3.1) are 𝖬~b~\widetilde{\mathsf{M}}_{\widetilde{b}} and 𝖬b~\mathsf{M}_{\widetilde{b}} respectively in the current context). Then we obtain an irreducible smooth WW-scheme TT and a morphism φ:𝔸1T𝖬~b~\varphi:\mathbb{A}^{1}_{T}\to\widetilde{\mathsf{M}}_{\widetilde{b}} such that φ\varphi_{\mathbb{C}} defines a strongly dominating families of curves on 𝖬~b~\widetilde{\mathsf{M}}_{\widetilde{b}}{\otimes}\mathbb{C} by (7.3.2). Let tt be a general kk-point on TkT_{k} such that the conclusion of (7.3.4) holds. Since 𝖵|𝖬b~\mathsf{V}|_{\mathsf{M}_{\widetilde{b}}{\otimes}\mathbb{C}} is non-isotrivial, for a general zTz\in T_{\mathbb{C}}, the restriction of 𝖵\mathsf{V} to φz(𝖬b~)\varphi_{z}^{*}(\mathsf{M}_{\widetilde{b}}{\otimes}\mathbb{C}) is also non-isotrivial. Similarly, applying (6.2.2) to the NL loci on 𝖬b~\mathsf{M}_{\widetilde{b}}{\otimes}\mathbb{C} again, we conclude that for some t~T(W)\widetilde{t}\in T(W) lifting tt, φt~(𝖬b~)\varphi_{\widetilde{t}{\otimes}\mathbb{C}}^{*}(\mathsf{M}_{\widetilde{b}}{\otimes}\mathbb{C}) is admissible. Therefore, if we set CC to be φt~(𝖬b~)\varphi^{*}_{\widetilde{t}}(\mathsf{M}_{\widetilde{b}}), then it indeed satisfies condition (i), and (7.3.4)(c) guarantees that this CC also satisfies (ii), as desired. ∎

Remark (7.4.5).

Let ZZ_{\mathbb{C}} be an irreducible component of the Noether-Lefschetz (NL) loci on 𝖬\mathsf{M}_{\mathbb{C}} and take ZZ to be the Zariski closure of ZZ_{\mathbb{C}} in 𝖬\mathsf{M}. Then by specialization of line bundles we know that every geometric fiber of 𝒳\mathcal{X}^{\circ} over ZZ has Picard rank 3\geq 3. By the Shioda-Tate formula141414See Landesman’s note people.math.harvard.edu/~landesman/assets/shioda-tate.pdf for a proof of the formula over finite fields., this implies that, if kk is a finite field, sZ(k)s\in Z(k) and \mathcal{E} is the generic fiber of the elliptic fibration on 𝒳s\mathcal{X}^{\circ}_{s}, then up to replacing kk by a finite extension rankMW()>0\mathrm{rank\,}\mathrm{MW}(\mathcal{E})>0. Moreover, as 𝒳s\mathcal{X}^{\circ}_{s} has even second Betti number, the Weil conjecture implies that rankNS(𝒳sk¯)\mathrm{rank\,}\mathrm{NS}(\mathcal{X}^{\circ}_{s}{\otimes}\bar{k}) is also even, so up to extending kk again we must in fact have rankMW()2\mathrm{rank\,}\mathrm{MW}(\mathcal{E})\geq 2.

Note that ZZ is of codimension 11 in 𝖬\mathsf{M}, so that ZZ has relative dimension 1010 over [1/6]\mathbb{Z}[1/6]. The 𝔾m\mathbb{G}_{m}-action on 𝖬\mathsf{M} stabilizes ZZ and the quotient Z/𝔾mZ/\mathbb{G}_{m} gives rise to a 99-dimensional subfamily in the coarse moduli of elliptic surfaces with pg=q=1p_{g}=q=1 over [1/6]\mathbb{Z}[1/6] (cf. (7.4.2)). By [Moonen, Prop. 6.4], the NL loci has infinitely many components, so there are lots of these examples.

8 Surfaces with pg=K2=1p_{g}=K^{2}=1 and q=0q=0

Recall our notations for weighted projective spaces in (7.1.2) and set Q:=(1,2,2,3,3)Q:=(1,2,2,3,3). Throughout this section, kk denotes an arbitrary algebraically closed field of characteristic 2,3\neq 2,3 unless otherwise specified.

Theorem (8.0.1).

Let XX be a minimal surface over kk with pg=KX2=1p_{g}=K_{X}^{2}=1 and q=0q=0. Then the canonical model XX^{\prime} of XX only has rational double point singularities. Moreover, if we let (x0,x1,x2,x3,x4)(x_{0},x_{1},x_{2},x_{3},x_{4}) be the coordinates of Q(k5)\mathbb{P}_{Q}(k^{\oplus 5}), then for j=1,2j=1,2