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

Absolute Hodge and \ell-adic Monodromy

David Urbanik
Abstract

Let 𝕍\mathbb{V} be a motivic variation of Hodge structure on a KK-variety SS, let \mathcal{H} be the associated KK-algebraic Hodge bundle, and let σAut(/K)\sigma\in\textrm{Aut}(\mathbb{C}/K) be an automorphism. The absolute Hodge conjecture predicts that given a Hodge vector v,sv\in\mathcal{H}_{\mathbb{C},s} above sS()s\in S(\mathbb{C}) which lies inside 𝕍s\mathbb{V}_{s}, the conjugate vector vσ,sσv_{\sigma}\in\mathcal{H}_{\mathbb{C},s_{\sigma}} is Hodge and lies inside 𝕍sσ\mathbb{V}_{s_{\sigma}}. We study this problem in the situation where we have an algebraic subvariety ZSZ\subset S_{\mathbb{C}} containing ss whose algebraic monodromy group HZ\textbf{H}_{Z} fixes vv. Using relationships between HZ\textbf{H}_{Z} and HZσ\textbf{H}_{Z_{\sigma}} coming from the theories of complex and \ell-adic local systems, we establish a criterion that implies the absolute Hodge conjecture for vv subject to a group-theoretic condition on HZ\textbf{H}_{Z}. We then use our criterion to establish new cases of the absolute Hodge conjecture.

1 Introduction

Let f:XSf:X\to S be a smooth projective111We follow the convention in [Har77], so this means that there exists an embedding ι:XSn\iota:X\hookrightarrow\mathbb{P}^{n}_{S} over SS. morphism of KK-varieties, with SS smooth and quasi-projective, and KK\subset\mathbb{C} a subfield. In this setting, the local system 𝕍=R2kfan(k)\mathbb{V}=R^{2k}f^{\textrm{an}}_{*}\mathbb{Q}(k) underlies a polarizable variation of Hodge structure, which has the property that the vector bundle 𝕍𝒪San\mathbb{V}\otimes\mathcal{O}_{S^{\textrm{an}}} admits a KK-algebraic model =R2kfΩX/S\mathcal{H}=R^{2k}f_{*}\Omega^{\bullet}_{X/S}, where ΩX/S\Omega^{\bullet}_{X/S} is the complex of relative differentials. Moreover, the underlying integral local system 𝕍𝕍\mathbb{V}_{\mathbb{Z}}\subset\mathbb{V} admits a comparison with the \ell-adic local system 𝕍=R2kfét(k)\mathbb{V}_{\ell}=R^{2k}f^{\textrm{\'{e}t}}_{*}\mathbb{Z}_{\ell}(k) on the étale site of SS_{\mathbb{C}}.

Notation.

For sS()s\in S(\mathbb{C}), we denote by csdR:𝕍,s,sc^{\textrm{dR}}_{s}:\mathbb{V}_{\mathbb{C},s}\xrightarrow{\sim}\mathcal{H}_{\mathbb{C},s} and cs:𝕍,s𝕍,,sc^{\ell}_{s}:\mathbb{V}_{\mathbb{Q}_{\ell},s}\xrightarrow{\sim}\mathbb{V}_{\ell,\mathbb{Q}_{\ell},s} the natural comparison isomorphisms.

The fact that \mathcal{H} and 𝕍\mathbb{V}_{\ell} come from a morphism f:XSf:X\to S of KK-varieties means that for each automorphism σAut(/K)\sigma\in\textrm{Aut}(\mathbb{C}/K) and point sS()s\in S(\mathbb{C}), there are isomorphisms ()σ:,s,sσ(-)_{\sigma}:\mathcal{H}_{\mathbb{C},s}\xrightarrow{\sim}\mathcal{H}_{\mathbb{C},s_{\sigma}} and ()σ:𝕍,,s𝕍,,sσ(-)_{\sigma}:\mathbb{V}_{\ell,\mathbb{Q}_{\ell},s}\xrightarrow{\sim}\mathbb{V}_{\ell,\mathbb{Q}_{\ell},s_{\sigma}} induced by conjugation. Combining these isomorphisms with the isomorphisms csdRc^{\textrm{dR}}_{s} and csc^{\ell}_{s}, we may consider several possible ways in which the rational structure of 𝕍\mathbb{V} may be preserved under conjugation:

Definition 1.1.

Let sS()s\in S(\mathbb{C}) be a point, v𝕍sv\in\mathbb{V}_{s} be a vector, W𝕍sW\subset\mathbb{V}_{s} a subspace, and σAut(/K)\sigma\in\textrm{Aut}(\mathbb{C}/K). Then we say:

  • (i)

    That vv has rational conjugates if csdR(v)σc_{s}^{\textrm{dR}}(v)_{\sigma} (resp. cs(v)σc_{s}^{\ell}(v)_{\sigma} for all \ell) are rational vectors of 𝕍\mathbb{V}. (They lie in the image of 𝕍sσ\mathbb{V}_{s_{\sigma}} under the comparison isomorphisms csσdRc_{s_{\sigma}}^{\textrm{dR}} and csσc^{\ell}_{s_{\sigma}}.)

  • (ii)

    That vv has a canonical rational conjugate if there exists vσ𝕍sσv_{\sigma}\in\mathbb{V}_{s_{\sigma}} such that csdR(v)σ=csσdR(vσ)c_{s}^{\textrm{dR}}(v)_{\sigma}=c_{s_{\sigma}}^{\textrm{dR}}(v_{\sigma}) and cs(v)σ=csσ(vσ)c_{s}^{\ell}(v)_{\sigma}=c_{s_{\sigma}}^{\ell}(v_{\sigma}) for all \ell. (The vector vσv_{\sigma} works for all comparison isomorphisms at once.)

  • (iii)

    That WW has rational conjugates if each element of WW does.

  • (iv)

    That WW has a canonical rational conjugate if each element of WW does.

Remark.

It is clear that if ww and ww^{\prime} have (canonical) rational conjugates then so do w+ww+w^{\prime} and λw\lambda w for any λ\lambda\in\mathbb{Q}, so we lose nothing by considering subspaces with (canonical) rational conjugates.

In the case K=K=\mathbb{Q}, the property that a Hodge vector vv has rational conjugates is equivalent to the absolute Hodge conjecture for vv formulated in [CS14].222Although the formulation in [Del82a], notably, is equivalent to the condition that vv has a canonical rational conjugate; see also the discussion in section 2.3. The property that vv has a canonical rational conjugate is a priori stronger, and would follow for Hodge vectors from the Hodge conjecture in the geometric setting. These properties are expected to hold for Hodge vectors more generally in the motivic setting. Loosely speaking, we will say that a variation 𝕍\mathbb{V} is motivic if it is generated from the geometric situation by functorial constructions; we give a precise definition suitable for our purposes in section 2.1.

Our central technical result, from which new cases of the absolute Hodge conjecture will follow, gives a new way of deducing that a subspace W𝕍sW\subset\mathbb{V}_{s} (or a subspace of the various tensor powers of 𝕍s\mathbb{V}_{s}) has rational conjugates by studying the monodromy of complex subvarieties ZSZ\subset S_{\mathbb{C}}. To formulate the statement we introduce some notation.

Notation.

Given an RR-module VV, we let Vm,n=(V)mVnV^{m,n}=(V^{*})^{\otimes m}\otimes V^{\otimes n}, with the analogous definition for RR-local systems 𝕍\mathbb{V}.

Notation.

Given an RR-module VV, we let 𝒯(V)=m,n0Vn,m\mathcal{T}(V)=\bigoplus_{m,n\geq 0}V^{n,m}, with the analogous definition for RR-local systems 𝕍\mathbb{V}.

Notation.

Let VV be a vector space over RR, and let T𝒯(V)T\subset\mathcal{T}(V) be a vector subspace. We denote by 𝐆TAut(V)\mathbf{G}_{T}\subset\textrm{Aut}(V) the RR-algebraic subgroup defined by the property that it fixes each element of TT.

Notation.

Let ZSZ\subset S_{\mathbb{C}} be an algebraic subvariety, and 𝕍\mathbb{V} be a local system. We denote by HZ\textbf{H}_{Z} the algebraic monodromy of ZZ, which is the identity component of the Zariski closure of the monodromy representation on 𝕍|Znor{\left.\kern-1.2pt\mathbb{V}\vphantom{\big{|}}\right|_{Z^{\textrm{nor}}}}, where ZnorZZ^{\textrm{nor}}\to Z is the normalization of ZZ.

Remark.

Given a subvariety ZSZ\subset S_{\mathbb{C}}, we may regard HZ\textbf{H}_{Z} as an algebraic subgroup of Aut(𝕍s)\textrm{Aut}(\mathbb{V}_{s}) for any point sZ()s\in Z(\mathbb{C}). We will use the notation HZ\textbf{H}_{Z} for each of these subgroups, even though the point ss may vary, as the context will leave no ambiguity about which algebraic group we mean.

Our main technical result is then the following:

Definition 1.2.

If GAut(𝕍s)G\subset\textrm{Aut}(\mathbb{V}_{s}) is an algebraic group, and W𝒯(𝕍s)W\subset\mathcal{T}(\mathbb{V}_{s}) is a subspace, we will say that GG fixes WW as a \mathbb{Q}-subspace if for each gG()g\in G(\mathbb{C}) we have gW=WgW=W.333We wish to stress that this condition is different from fixing WW pointwise, or fixing WW_{\mathbb{C}} as a subspace.

Theorem 1.3.

Let 𝕍\mathbb{V} be a motivic variation of Hodge structure on SS, and let sS()s\in S(\mathbb{C}). Suppose that T𝒯(𝕍s)T\subset\mathcal{T}(\mathbb{V}_{s}) is a subspace with canonical rational conjugates, that ZSZ\subset S_{\mathbb{C}} is a complex subvariety containing ss satisfying HZ𝐆T\textbf{H}_{Z}\subset\mathbf{G}_{T}, and that W𝒯(𝕍s)W\subset\mathcal{T}(\mathbb{V}_{s}) is a subspace containing TT. Suppose that the normalizer NN of HZ\textbf{H}_{Z} inside 𝐆T\mathbf{G}_{T} fixes WW as a \mathbb{Q}-subspace. Then WW has rational conjugates.

A simple application of 1.3 is then:

Corollary 1.4.

Suppose 𝕍\mathbb{V} is a motivic variation of Hodge structure on SS, let sS()s\in S(\mathbb{C}) and suppose that ZSZ\subset S_{\mathbb{C}} contains ss. Then if the only (m,n)(m,n) tensors fixed by HZ\textbf{H}_{Z} are of the form w\mathbb{Q}w for some w𝕍sm,nw\in\mathbb{V}_{s}^{m,n}, then the subspace w\mathbb{Q}w has rational conjugates.

Proof.

We show in section 2.2 that any motivic variation admits a polarization Q:𝕍𝕍Q:\mathbb{V}\otimes\mathbb{V}\to\mathbb{Q} such that the class of QsQ_{s} inside 𝕍s2,0\mathbb{V}^{2,0}_{s} has a canonical rational conjugate. Take TT to be the span of QsQ_{s}. From the fact that NN preserves QQ we learn it acts on ww by ±1\pm 1. Taking WW to be the span of QsQ_{s} and ww, we learn that WW has rational conjugates. ∎

As we explain in section 5.1, a form of the argument in 1.4 appears in Voisin’s paper [Voi07]. There it is used to prove a “weakly absolute Hodge” statement for Hodge classes coming from a family of fourfolds for which neither the Hodge nor absolute Hodge conjecture are known. Our 1.3 gives the full absolute conjecture in this case; this is 1.8.

Let Isom(𝕍s,𝕍sσ)\textrm{Isom}(\mathbb{V}_{s},\mathbb{V}_{s_{\sigma}}) be the \mathbb{Q}-algebraic variety consisting of all isomorphisms between 𝕍s\mathbb{V}_{s} and 𝕍sσ\mathbb{V}_{s_{\sigma}}. Our proof of 1.3 proceeds through the study of the \mathbb{Q}-subvariety

T(s,Z,σ)Isom(𝕍s,𝕍sσ),\mathcal{I}_{T}(s,Z,\sigma)\subset\textrm{Isom}(\mathbb{V}_{s},\mathbb{V}_{s_{\sigma}}),

defined by the property that rT(s,Z,σ)()r\in\mathcal{I}_{T}(s,Z,\sigma)(\mathbb{C}) if and only if rHZ,r1=HZσ,r\circ\textbf{H}_{Z,\mathbb{C}}\circ r^{-1}=\textbf{H}_{Z_{\sigma},\mathbb{C}} and rr sends each tTt\in T to its canonical rational conjugate tσTσt_{\sigma}\in T_{\sigma}. We observe that T(s,Z,σ)\mathcal{I}_{T}(s,Z,\sigma) is a torsor under the normalizer NN of HZ\textbf{H}_{Z} inside 𝐆T\mathbf{G}_{T}. 1.3 follows immediately from the following two facts about T(s,Z,σ)\mathcal{I}_{T}(s,Z,\sigma), established in section 3:

Proposition 1.5.
  • (i)

    The map

    rσdR:𝕍,scsdR,s𝜎,sσ(csσdR)1𝕍,sσr^{\textrm{dR}}_{\sigma}:\mathbb{V}_{\mathbb{C},s}\xrightarrow{c^{\textrm{dR}}_{s}}\mathcal{H}_{\mathbb{C},s}\xrightarrow{\sigma}\mathcal{H}_{\mathbb{C},s_{\sigma}}\xrightarrow{(c^{\textrm{dR}}_{s_{\sigma}})^{-1}}\mathbb{V}_{\mathbb{C},s_{\sigma}}

    induced by σ\sigma is a \mathbb{C}-point of T(s,Z,σ)\mathcal{I}_{T}(s,Z,\sigma).

  • (ii)

    For each \ell, the map

    rσ:𝕍,scs𝕍,,s𝜎𝕍,,sσ(csσ)1𝕍,sσr^{\ell}_{\sigma}:\mathbb{V}_{\mathbb{Q}_{\ell},s}\xrightarrow{c^{\ell}_{s}}\mathbb{V}_{\ell,\mathbb{Q}_{\ell},s}\xrightarrow{\sigma}\mathbb{V}_{\ell,\mathbb{Q}_{\ell},s_{\sigma}}\xrightarrow{(c^{\ell}_{s_{\sigma}})^{-1}}\mathbb{V}_{\mathbb{Q}_{\ell},s_{\sigma}}

    induced by σ\sigma is a \mathbb{Q}_{\ell}-point of T(s,Z,σ)\mathcal{I}_{T}(s,Z,\sigma).

Proof of 1.3 (assuming 1.5):.

We let =T(s,Z,σ)\mathcal{I}=\mathcal{I}_{T}(s,Z,\sigma) for ease of notation. That \mathcal{I} is naturally a torsor under NN implies that the subspace r(W)r(W) is independent of the choice of r()r\in\mathcal{I}(\mathbb{C}): any two choices rr and rr^{\prime} are related by r=rnr^{\prime}=rn with nN()n\in N(\mathbb{C}), and NN fixes WW by assumption. As \mathcal{I} is a \mathbb{Q}-variety, it has a ¯\overline{\mathbb{Q}}-point rr, hence Wσ:=r(W)W_{\sigma}:=r(W) is defined over ¯\overline{\mathbb{Q}}. If we can show that WσW_{\sigma} is in fact a \mathbb{Q}-subspace of 𝕍sσ\mathbb{V}_{s_{\sigma}}, then the proof will be complete as 1.5 then implies both that csdR(W)σ=csσdR(Wσ)c^{\textrm{dR}}_{s}(W)_{\sigma}=c^{\textrm{dR}}_{s_{\sigma}}(W_{\sigma}) and cs(W)σ=csσ(Wσ)c^{\ell}_{s}(W)_{\sigma}=c^{\ell}_{s_{\sigma}}(W_{\sigma}) for each \ell.

By 1.5(ii) the variety \mathcal{I} has \mathbb{Q}_{\ell}-points for every \ell, hence it follows that under some choice of isomorphism ¯\mathbb{C}\cong\overline{\mathbb{Q}_{\ell}}, every element of WσW_{\sigma} is defined over \mathbb{Q}_{\ell}. The proof is completed by the following:

Claim 1.6.

Let v𝔸n(¯)v\in\mathbb{A}^{n}(\overline{\mathbb{Q}}) be a ¯\overline{\mathbb{Q}}-point in affine space, and suppose that for each \ell there exists an isomorphism ¯\mathbb{C}\cong\overline{\mathbb{Q}_{\ell}} under which vv is identified with a \mathbb{Q}_{\ell}-point. Then vv is defined over \mathbb{Q}.

Proof.

As 𝔸n=(𝔸1)n\mathbb{A}^{n}=(\mathbb{A}^{1})^{n}, it suffices to consider the case n=1n=1. Let ff be the minimal polynomial of vv. The statement is unchanged by replacing vv with λv\lambda v for λ\lambda\in\mathbb{Q}, so we may assume that vv is integral and that ff is monic with integral coefficients. The hypothesis shows (using integrality) that ff has a root over \mathbb{\mathbb{Z}}_{\ell} for every \ell, hence a root modulo \ell for every \ell. As ff is irreducible, this implies that ff is linear by [CF67, p. 362, Ex. 6.2]. ∎

Remark.

The proof of 1.3 does not invoke the properties of the filtration FF^{\bullet} in any way, so it applies to motivic mixed variations as well. We focus exclusively on the pure case in this paper.

1.3 lets us deduce the following result on the absolute Hodge conjecture, as we show in section 2.3:

Theorem 1.7.

Let 𝕍\mathbb{V}, SS, ss, TT, and ZZ be as in 1.3, and suppose that WW is a subspace of Hodge vectors. Then the KK-absolute Hodge conjecture444See section 2.3 for a definition; the usual absolute Hodge conjecture is the case K=K=\mathbb{Q}. holds for each element of WW.

We give two applications of our results. The first is to resolve the absolute Hodge conjecture for classes coming from a family of fourfolds considered by Voisin in her paper [Voi07, Example 3.4]. More specifically, we show the following:

Theorem 1.8.

Let f:XSf:X\to S be the family of degree 66 hypersurfaces in 5\mathbb{P}^{5} stable under the involution ι(X0,,X5)=(X0,X1,X2,,X5)\iota(X_{0},\ldots,X_{5})=(-X_{0},-X_{1},X_{2},\ldots,X_{5}), let 𝕍\mathbb{V} be the geometric variation on middle cohomology of degree 2k2k, and let 𝕍\mathbb{V}_{-} be the subvariation obtained by taking anti-invariants under ι\iota. Then if k,k\mathcal{H}_{-}^{k,k} is the quotient Fk/Fk+1F_{-}^{k}/F_{-}^{k+1}, the rational Hodge classes v𝕍v\in\mathbb{V}_{-} for which the absolute Hodge conjecture holds are topologically dense in the real part of k,k\mathcal{H}_{-}^{k,k}.

Our method of establishing 1.8 is similar to the argument employed in Voisin’s paper to establish that such classes are “weakly absolute Hodge”. We are able to obtain a stronger conclusion due to the presence of \ell-adic input in our argument.

As a second application, we let 𝕍\mathbb{V} be a motivic variation on SS, let ZS×SZ\subset S\times S be the algebraic locus of points (x,y)(x,y) such that 𝕍x𝕍y\mathbb{V}_{x}\cong\mathbb{V}_{y} as integral polarized Hodge structures, and we let Z0ZZ_{0}\subset Z be an irreducible component containing the diagonal ΔS×S\Delta\subset S\times S. The variety Z0Z_{0} is a component of the Hodge locus for the variation 𝕍𝕍\mathbb{V}\boxtimes\mathbb{V}^{*} on S×SS\times S (see 5.2). We call an isomorphism 𝕍x𝕍y\mathbb{V}_{x}\cong\mathbb{V}_{y} with (x,y)Z0(x,y)\in Z_{0} generic if its class is Hodge at a generic point of Z0Z_{0}. We have the following result:

Theorem 1.9.

If 𝕍\mathbb{V} is a motivic variation of Hodge structure on the smooth KK-variety SS such that HS=Aut(𝕍s,Qs)\textbf{H}_{S}=\textrm{Aut}(\mathbb{V}_{s},Q_{s}) (monodromy is Zariski dense), then rational tensors fixed by HZ0\textbf{H}_{Z_{0}} have rational conjugates, and hence generic isomorphisms between 𝕍x\mathbb{V}_{x} and 𝕍y\mathbb{V}_{y} are KK-absolutely Hodge.

1.1 Structure of the Paper

In section 2 we discuss motivic variations and the Hodge and absolute Hodge conjectures for such variations. We review some properties of motivic variations in section 2.1, and list all the properties that we will be needed for the remainder of the paper. The properties in section 2.1 relating to polarizations require further justification, so we carry out the necessary constructions in section 2.2. As mentioned, section 2.3 discusses the relationship between our notion of rational conjugates and the Absolute Hodge conjecture.

In section 3 we prove 1.5. The proof of 1.5(i) is already implicit in the paper [KOU20], so we give a brief summary of the argument in section 3.1. To tackle the \ell-adic case in section 3.2 we reduce the necessary statement to a purely algebraic statement involving étale fundamental groups, after which the result follows by showing that σAut(/K)\sigma\in\textrm{Aut}(\mathbb{C}/K) essentially conjugates the representation of π1ét(S,s)\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s) on the fibre 𝕍,s\mathbb{V}_{\ell,s} to the representation of π1ét(S,sσ)\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s_{\sigma}) on the fibre 𝕍,sσ\mathbb{V}_{\ell,s_{\sigma}}.

The same ideas that give 1.3 and 1.7 in fact give several other variants of 1.3 and 1.7, as we discuss in section 4. We then consider the example of Voisin in section 5.1, and the application to generic isomorphisms of Hodge structures in section 5.2.

2 Motivic Variations

2.1 Properties of Motivic Variations

Recall that our central example of a variation of Hodge structures is the variation 𝕍\mathbb{V} obtained from a smooth, projective family f:XSf:X\to S of KK-varieties, with SS smooth. Let us briefly state some properties of such variations:

  • (i)

    The Hodge bundle 𝕍𝒪San\mathbb{V}\otimes\mathcal{O}_{S^{\textrm{an}}} admits a canonical model as a KK-algebraic vector bundle =R2kfΩX/S\mathcal{H}=R^{2k}f_{*}\Omega^{\bullet}_{X/S}, where ΩX/S\Omega^{\bullet}_{X/S} is the complex of relative algebraic differentials.

  • (ii)

    The filtration FF^{\bullet} on \mathcal{H} admits a canonical KK-algebraic model, agreeing with the filtration in the analytic setting. This is obtained from a filtration on ΩX/S\Omega^{\bullet}_{X/S} and an appropriate spectral sequence.

  • (iii)

    The connection \nabla whose flat sections are given by 𝕍\mathbb{V}_{\mathbb{C}} is KK-algebraic. This is due to Katz and Oda [KO68].

  • (iv)

    The \mathbb{Q}-local system 𝕍\mathbb{V} admits a canonical integral subsystem 𝕍𝕍\mathbb{V}_{\mathbb{Z}}\subset\mathbb{V}, such that 𝕍\mathbb{V}_{\mathbb{Z}}\otimes\mathbb{Z}_{\ell} admits a comparison isomorphism with 𝕍an\mathbb{V}_{\ell}^{\textrm{an}}, where 𝕍\mathbb{V}_{\ell} is an \ell-adic local system defined on the étale site of SS_{\mathbb{C}}. This is the relative version of the comparison between Betti and étale cohomology.

The following additional properties relate to the polarization on 𝕍\mathbb{V}, and will be established in the next section:

  • (v)

    The variation 𝕍\mathbb{V} admits a polarization Q:𝕍𝕍Q:\mathbb{V}\otimes\mathbb{V}\to\mathbb{Q} which is KK-algebraic, in the sense that we have a morphism 𝒬:𝒪S\mathcal{Q}:\mathcal{H}\otimes\mathcal{H}\to\mathcal{O}_{S} of KK-algebraic vector bundles whose analytification is identified with QQ after applying the isomorphism 𝕍𝒪Sanan\mathbb{V}\otimes\mathcal{O}_{S^{\textrm{an}}}\simeq\mathcal{H}^{\textrm{an}}.

  • (vi)

    There exists a bilinear form Q:𝕍𝕍Q_{\ell}:\mathbb{V}_{\ell}\otimes\mathbb{V}_{\ell}\to\mathbb{Z}_{\ell}, compatible with the polarization QQ in (v).

  • (vii)

    Let σAut(/K)\sigma\in\textrm{Aut}(\mathbb{C}/K), and denote by τσ:SétSét\tau_{\sigma}:S_{\textrm{\'{e}t}}\to S_{\textrm{\'{e}t}} the induced automorphism of the étale site of SS_{\mathbb{C}}. Then there exists a canonical isomorphism τσ𝕍𝕍\tau_{\sigma}^{*}\mathbb{V}_{\ell}\simeq\mathbb{V}_{\ell}, compatible with the polarization QQ_{\ell} in (vi).

The properties (i) through (vii) are preserved under any sufficiently functorial construction; for instance, the properties (i) through (vii) are preserved under duals, direct sums and tensor products. The theorems we prove will only ever use the above listed properties, and we will refer to such variations as motivic. Note that it is not immediately clear that this notion of motivic variation is the same as other similar notions that appear in the literature, but it is the definition which will be most useful for us.

We note the following elementary consequence of (v), (vi) and (vii) which will be useful later:

Lemma 2.1.

For motivic variations, the σ\sigma-conjugate of the class Qs𝕍s2,0Q_{s}\in\mathbb{V}^{2,0}_{s} is rational for 𝕍sσ2,0\mathbb{V}^{2,0}_{s_{\sigma}} (and given by QsσQ_{s_{\sigma}}) in both the de Rham and \ell-adic setting, where σAut(/K)\sigma\in\textrm{Aut}(\mathbb{C}/K). ∎

2.2 Construction of KK-algebraic Polarizations

The material in this section should be well-known to experts, but we cannot find an appropriate reference.

We consider the geometric situation with 𝕍=R2kfan,\mathbb{V}=R^{2k}f_{\textrm{an},*}\mathbb{Z}, =R2kfΩX/S\mathcal{H}=R^{2k}f_{*}\Omega^{\bullet}_{X/S} and 𝕍=R2kfét,\mathbb{V}_{\ell}=R^{2k}f_{\textrm{\'{e}t},*}\mathbb{Z}_{\ell}. We will define the primitive cohomology subsystems 𝕍prim𝕍\mathbb{V}_{\textrm{prim}}\subset\mathbb{V} and 𝕍,prim𝕍\mathbb{V}_{\ell,\textrm{prim}}\subset\mathbb{V}_{\ell}, and vector subbundle prim\mathcal{H}_{\textrm{prim}}\subset\mathcal{H}, as well as polarizations Q,QQ,Q_{\ell} and 𝒬\mathcal{Q} on the primitive cohomology satisfying the properties (v), (vi) and (vii). That properties (v), (vi) and (vii) hold for the original local systems and vector bundles follows from the usual procedure of obtaining a polarization on all of cohomology from a polarization on the primitive part.

In the case where S=SpecKS=\textrm{Spec}\hskip 1.49994ptK is a point, our definition of QQ and 𝒬\mathcal{Q} essentially appears (among other places) in [Del71]. Our goal is to generalize this to the relative setting, where we regard the scheme XX as an SS-scheme via the map ff. The first task is to define the relative Chern classes c1,SdR()c^{\textrm{dR}}_{1,S}(\mathcal{L}) and c1,Sét()c^{\textrm{\'{e}t}}_{1,S}(\mathcal{L}) of a line bundle \mathcal{L} on XX.

Lemma 2.2.

Let \mathcal{L} be a line bundle on XX. Then there exists a (necessarily unique) global section c1,SdR()c^{\textrm{dR}}_{1,S}(\mathcal{L}) of R2fΩX/SR^{2}f_{*}\Omega^{\bullet}_{X/S} whose restriction to the fibre at sSs\in S is the Chern class c1dR(|Xs)c^{\textrm{dR}}_{1}\left({\left.\kern-1.2pt\mathcal{L}\vphantom{\big{|}}\right|_{X_{s}}}\right).555In the sense that the equality holds after identifying (R2fΩX/S)s(R^{2}f_{*}\Omega^{\bullet}_{X/S})_{s} with the algebraic de Rham cohomology of the fibre XsX_{s} using proper base change.

Proof.

By [Sta20, Section 0FLE] we obtain a class c1dR()HdR2(X/S)c_{1}^{\textrm{dR}}(\mathcal{L})\in H^{2}_{\textrm{dR}}(X/S), where HdR(X/S)H^{\bullet}_{\textrm{dR}}(X/S) is the cohomology of the complex RΓΩX/SR\,\Gamma\,\Omega^{\bullet}_{X/S}. Choose an injective resolution ΩX/S\Omega^{\bullet}_{X/S}\to\mathcal{I}^{\bullet}. We may compute R2fΩX/SR^{2}f_{*}\Omega^{\bullet}_{X/S} as the cohomology sheaf of ff_{*}\mathcal{I}^{\bullet} in degree two. From the equality Γ(S,f)=Γ(X,)\Gamma(S,f_{*}\mathcal{I}^{\bullet})=\Gamma(X,\mathcal{I}^{\bullet}) we therefore obtain a morphism g:HdR2(X/S)Γ(S,R2fΩX/S)g:H^{2}_{\textrm{dR}}(X/S)\to\Gamma(S,R^{2}f_{*}\Omega^{\bullet}_{X/S}), and we may define c1,SdR()c^{\textrm{dR}}_{1,S}(\mathcal{L}) as the image of c1dR()c_{1}^{\textrm{dR}}(\mathcal{L}) under gg.

Let sSs\in S be a point, and r:Γ(S,R2fΩX/S)(R2fΩX/S)sr:\Gamma(S,R^{2}f_{*}\Omega^{\bullet}_{X/S})\to(R^{2}f_{*}\Omega^{\bullet}_{X/S})_{s} be the restriction of a global section to the fibre at ss. Let b:(R2fΩX/S)sHdR2(Xs/κ(s))b:(R^{2}f_{*}\Omega^{\bullet}_{X/S})_{s}\to H^{2}_{\textrm{dR}}(X_{s}/\kappa(s)) be the canonical base change morphism. As Chern classes are functorial, the map h:HdR2(X/S)HdR2(Xs/κ(s))h:H^{2}_{\textrm{dR}}(X/S)\to H^{2}_{\textrm{dR}}(X_{s}/\kappa(s)) induced by the inclusion ι:XsX\iota:X_{s}\hookrightarrow X sends c1dR()c^{dR}_{1}(\mathcal{L}) to c1dR(|Xs)c^{\textrm{dR}}_{1}\left({\left.\kern-1.2pt\mathcal{L}\vphantom{\big{|}}\right|_{X_{s}}}\right). It therefore suffices to check that h=brgh=b\circ r\circ g.

To make this verification, we recall the construction of the base change map at the level of complexes, following [Sta20, Section 0735]. We may identify ΩXs\Omega^{\bullet}_{X_{s}} with the pullback ιΩX/S\iota^{*}\Omega^{\bullet}_{X/S}, and find an injective resolution ιΩX/S𝒥\iota^{*}\Omega^{\bullet}_{X/S}\to\mathcal{J}^{\bullet}. We then obtain a commuting diagram

ιιΩX/S{\iota_{*}\iota^{*}\Omega^{\bullet}_{X/S}}ι𝒥{\iota_{*}\mathcal{J}^{\bullet}}ΩX/S{\Omega^{\bullet}_{X/S}}{\mathcal{I}^{\bullet}}adj.β\scriptstyle{\beta}

,

where the map β\beta is unique up to homotopy. The map bb (resp. the map hh) is constructed from β\beta by applying ff_{*} (resp. applying Γ\Gamma) and taking the induced map on cohomology. The required equality then follows from the fact that (fι)=Γ(f\circ\iota)_{*}=\Gamma. ∎

Lemma 2.3.

Let K¯\overline{K} be an algebraic closure of KK inside \mathbb{C}, and let \mathcal{L} be a line bundle on XK¯X_{\overline{K}}. Then there exists a (necessarily unique) global section c1,SK¯ét()c^{\textrm{\'{e}t}}_{1,S_{\overline{K}}}(\mathcal{L}) of R2fét,(1)R^{2}f_{\textrm{\'{e}t},*}\mathbb{Z}_{\ell}(1) whose restriction to the fibre at sS(K¯)s\in S(\overline{K}) is the Chern class c1ét(|Xs)c^{\textrm{\'{e}t}}_{1}\left({\left.\kern-1.2pt\mathcal{L}\vphantom{\big{|}}\right|_{X_{s}}}\right).666In the same sense as in 2.2.

Proof.

We argue analogously to 2.2. For ease of notation, we assume K=K¯K=\overline{K} and so replace XK¯X_{\overline{K}} by XX and SK¯S_{\overline{K}} by SS. By [Mil80, VI, §10] we obtain a class c1ét()Hét2(X,(1))c^{\textrm{\'{e}t}}_{1}(\mathcal{L})\in H^{2}_{\textrm{\'{e}t}}(X,\mathbb{Z}_{\ell}(1)). Choose an injective resolution \mathbb{Z}_{\ell}\to\mathcal{I}^{\bullet}.777Here we really mean a compatible system of resolutions /kk\mathbb{Z}/\ell^{k}\mathbb{Z}\to\mathcal{I}_{k}, where we regard \ell-adic sheaves as systems of /k\mathbb{Z}/\ell^{k}\mathbb{Z}-sheaves in the usual way. Proceeding as before, we may compute Hét2(X,(1))H^{2}_{\textrm{\'{e}t}}(X,\mathbb{Z}_{\ell}(1)) from the degree two cohomology of \mathcal{I}^{\bullet}, and 𝕍2\mathbb{V}^{2}_{\ell} as the degree two cohomology sheaf of fét,f_{\textrm{\'{e}t},*}\mathcal{I}^{\bullet}. From the equality Γ(S,fét,)=Γ(X,)\Gamma(S,f_{\textrm{\'{e}t},*}\mathcal{I}^{\bullet})=\Gamma(X,\mathcal{I}^{\bullet}) we therefore obtain a map g:Hét2(X,(1))Γ(S,R2fét,(1))g:H^{2}_{\textrm{\'{e}t}}(X,\mathbb{Z}_{\ell}(1))\to\Gamma(S,R^{2}f_{\textrm{\'{e}t},*}\mathbb{Z}_{\ell}(1)), and define c1,Sét()c^{\textrm{\'{e}t}}_{1,S}(\mathcal{L}) as the image of c1ét()c^{\textrm{\'{e}t}}_{1}(\mathcal{L}) under gg.

Letting sSs\in S be a (geometric) point, one defines r,br,b and hh analogously to 2.2, and similarly checks that h=brgh=b\circ r\circ g. ∎

Definition 2.4.

Let f:XSf:X\to S be a smooth projective morphism of KK-varieties, with SS smooth and fibres of dimension nn. Let 2k=R2kfΩX/S\mathcal{H}^{2k}=R^{2k}f_{*}\Omega^{\bullet}_{X/S}, 𝕍2k=R2kfét(k)\mathbb{V}^{2k}_{\ell}=R^{2k}f^{\textrm{\'{e}t}}_{*}\mathbb{Z}_{\ell}(k) and 𝕍2k=R2kfan(k)\mathbb{V}^{2k}=R^{2k}f^{\textrm{an}}_{*}\mathbb{Z}(k). Let \mathcal{L} be a very ample bundle over SS, and let ωdR=c1,SdR()\omega^{\textrm{dR}}=c^{\textrm{dR}}_{1,S}(\mathcal{L}), ωét=c1,SK¯ét()\omega^{\textrm{\'{e}t}}=c^{\textrm{\'{e}t}}_{1,S_{\overline{K}}}(\mathcal{L}) and ω\omega be the analytification of ωdR\omega^{\textrm{dR}}. We define:

  • (i)

    the operators

    LdR\displaystyle L^{\textrm{dR}} :2k2k+2\displaystyle:\mathcal{H}^{2k}\to\mathcal{H}^{2k+2} β\displaystyle\hskip 20.00003pt\beta βωdR\displaystyle\mapsto\beta\wedge\omega^{\textrm{dR}}
    Lét\displaystyle L^{\textrm{\'{e}t}}_{\ell} :𝕍2k𝕍2k+2\displaystyle:\mathbb{V}^{2k}_{\ell}\to\mathbb{V}^{2k+2}_{\ell} β\displaystyle\hskip 20.00003pt\beta βωét\displaystyle\mapsto\beta\wedge\omega^{\textrm{\'{e}t}}
    L\displaystyle L :𝕍2k𝕍2k+2\displaystyle:\mathbb{V}^{2k}\to\mathbb{V}^{2k+2} β\displaystyle\hskip 20.00003pt\beta βω;\displaystyle\mapsto\beta\wedge\omega;
  • (ii)

    and the subbundle and subsystems

    prim2k\displaystyle\mathcal{H}^{2k}_{\textrm{prim}} =kerLdR,n2k+1\displaystyle=\textrm{ker}\,L^{dR,n-2k+1}
    𝕍,prim2k\displaystyle\mathbb{V}^{2k}_{\ell,\textrm{prim}} =kerLét,n2k+1\displaystyle=\textrm{ker}\,L^{\textrm{\'{e}t},n-2k+1}_{\ell}
    𝕍prim2k\displaystyle\mathbb{V}^{2k}_{\textrm{prim}} =kerLn2k+1.\displaystyle=\textrm{ker}\,L^{n-2k+1}.

Now fix isomorphisms

ξdR\displaystyle\xi^{\textrm{dR}} :R2nfΩX/S𝒪S\displaystyle:R^{2n}f_{*}\Omega^{\bullet}_{X/S}\xrightarrow{\sim}\mathcal{O}_{S}
ξét\displaystyle\xi^{\textrm{\'{e}t}}_{\ell} :R2nfét,(n)\displaystyle:R^{2n}f_{\textrm{\'{e}t},*}\mathbb{Z}_{\ell}(n)\xrightarrow{\sim}\mathbb{Z}_{\ell}
ξ\displaystyle\xi :R2nf(n),\displaystyle:R^{2n}f_{*}\mathbb{Z}(n)\xrightarrow{\sim}\mathbb{Z},

compatible with the comparisons coming from analytification; for instance, using the relative version of the trace isomorphism (see [Har75] and [Con]). We then further define

  • (iii)

    the polarizations

    𝒬2k\displaystyle\mathcal{Q}^{2k} :prim2kprim2k𝒪S\displaystyle:\mathcal{H}^{2k}_{\textrm{prim}}\otimes\mathcal{H}^{2k}_{\textrm{prim}}\to\mathcal{O}_{S} ξdR[(α,β)\displaystyle\hskip 20.00003pt\xi^{\textrm{dR}}\circ[(\alpha,\beta) αβ(ωdR)n2k]\displaystyle\mapsto\alpha\wedge\beta\wedge(\omega^{\textrm{dR}})^{n-2k}]
    Q2k\displaystyle Q^{2k}_{\ell} :𝕍,prim2k𝕍,prim2k\displaystyle:\mathbb{V}^{2k}_{\ell,\textrm{prim}}\otimes\mathbb{V}^{2k}_{\ell,\textrm{prim}}\to\mathbb{Z}_{\ell} ξét[(α,β)\displaystyle\xi^{\textrm{\'{e}t}}_{\ell}\circ[(\alpha,\beta) αβ(ωét)n2k]\displaystyle\mapsto\alpha\wedge\beta\wedge(\omega^{\textrm{\'{e}t}})^{n-2k}]
    Q2k\displaystyle Q^{2k} :𝕍prim2k𝕍prim2k\displaystyle:\mathbb{V}^{2k}_{\textrm{prim}}\otimes\mathbb{V}^{2k}_{\textrm{prim}}\to\mathbb{Z} ξ[(α,β)\displaystyle\xi\circ[(\alpha,\beta) αβωn2k].\displaystyle\mapsto\alpha\cup\beta\cup\omega^{n-2k}].
Proposition 2.5.

The polarizations 𝒬2k,Q2k\mathcal{Q}^{2k},Q^{2k}_{\ell} and Q2kQ^{2k} of 2.4(iii) satisfy the properties (v), (vi) and (vii) of section 2.1.

Proof.

For properties (v) and (vi), this follows from the compatibility of the cup product with the comparison isomorphisms, as well as the compatibility of the maps ξdR,ξét\xi^{\textrm{dR}},\xi^{\textrm{\'{e}t}}_{\ell} and ξ\xi. For property (vii) this follows from the compatibility of the cup product with conjugation by Aut(/K)\textrm{Aut}(\mathbb{C}/K), as well as the fact that the section ω\omega is defined over KK. ∎

2.3 The Absolute Hodge Conjecture

Let YY be a smooth complex projective variety, and kk a non-negative integer. We define by H2k(Y)H^{2k}(Y) the intersection H2k(Yan,(k))Hk,k(Yan)H^{2k}(Y^{\textrm{an}},\mathbb{Q}(k))\cap H^{k,k}(Y^{\textrm{an}}), where H2k(Yan,(k))p+q=2kHp,q(Yan)H^{2k}(Y^{\textrm{an}},\mathbb{C}(k))\cong\bigoplus_{p+q=2k}H^{p,q}(Y^{\textrm{an}}) is the Hodge decomposition. According to [CS14], the absolute Hodge conjecture says the following:

Conjecture 2.6 (Absolute Hodge).

Let σAut(/)\sigma\in\textrm{Aut}(\mathbb{C}/\mathbb{Q}), and let cYdR:H2k(Yan,(k))HdR2k(Y/)c^{\textrm{dR}}_{Y}:H^{2k}(Y^{\textrm{an}},\mathbb{C}(k))\xrightarrow{\sim}H^{2k}_{\textrm{dR}}(Y/\mathbb{C}) and cY:H2k(Yan,(k))Hét2k(Y,(k))c^{\ell}_{Y}:H^{2k}(Y^{\textrm{an}},\mathbb{Q}_{\ell}(k))\xrightarrow{\sim}H^{2k}_{\textrm{\'{e}t}}(Y,\mathbb{Q}_{\ell}(k)) be the comparison isomorphisms. Then

  • (i)

    if vdRHdR2k(Y/)v^{\textrm{dR}}\in H^{2k}_{\textrm{dR}}(Y/\mathbb{C}) is the image of vH2k(Y)v\in H^{2k}(Y) under cYdRc^{\textrm{dR}}_{Y} then vσdRHdR2k(Yσ/)v^{\textrm{dR}}_{\sigma}\in H^{2k}_{\textrm{dR}}(Y_{\sigma}/\mathbb{C}) is the image of some vσH2k(Yσ)v_{\sigma}\in H^{2k}(Y_{\sigma}) under cYσdRc^{\textrm{dR}}_{Y_{\sigma}};

  • (ii)

    if vHét2k(Y,(k))v^{\ell}\in H^{2k}_{\textrm{\'{e}t}}(Y,\mathbb{Q}_{\ell}(k)) is the image of vH2k(Y)v\in H^{2k}(Y) under cYc^{\ell}_{Y} then vσHét2k(Yσ,(k))v^{\ell}_{\sigma}\in H^{2k}_{\textrm{\'{e}t}}(Y_{\sigma},\mathbb{Q}_{\ell}(k)) is the image of some vσH2k(Yσ)v_{\sigma}\in H^{2k}(Y_{\sigma}) under cYσc^{\ell}_{Y_{\sigma}}.

Remark.

Deligne in [Del82a] requires the class vσv_{\sigma} to be canonical, in the sense that it is the same for all comparison isomorphisms.

Remark.

2.6 generalizes to all Hodge tensors in the obvious way.

Definition 2.7.

More generally, we call the statement of 2.6 with Aut(/)\textrm{Aut}(\mathbb{C}/\mathbb{Q}) replaced by Aut(/K)\textrm{Aut}(\mathbb{C}/K) the KK-absolute Hodge conjecture.

Definition 2.8.

By the (KK-)absolute Hodge conjecture for vv, we mean the statement of 2.6 for some fixed vector vH2k(Y)v\in H^{2k}(Y).

Let us now deduce 1.7 from 1.3.

Proposition 2.9.

In the situation of 1.3, if each element of WW is Hodge, then the KK-absolute Hodge conjecture holds for each element of WW.

Proof.

What needs to be checked is that if wσw_{\sigma} is a rational conjugate to wWw\in W, then wσw_{\sigma} is automatically Hodge. We recall that a rational vector inside 𝕍s\mathbb{V}_{s} is Hodge if and only if it lies inside F0F^{0}, where FF^{\bullet} is the Hodge filtration. As FF^{\bullet} is a filtration by KK-algebraic bundles on \mathcal{H}, this shows the result in the de Rham case. In the \ell-adic case, we use the fact that Wσ=r(W)W_{\sigma}=r(W) is independent of the point rr of T(s,Z,σ)\mathcal{I}_{T}(s,Z,\sigma), hence taking r=rσr=r^{\ell}_{\sigma} we see that the result holds \ell-adically as well. ∎

3 Conjugation Isomorphisms and Monodromy

In this section we establish 1.5. The variety T(s,Z,σ)Isom(𝕍s,𝕍sσ)\mathcal{I}_{T}(s,Z,\sigma)\subset\textrm{Isom}(\mathbb{V}_{s},\mathbb{V}_{s_{\sigma}}) is defined by the conditions conditions rHZr1=HZσr\circ\textbf{H}_{Z}\circ r^{-1}=\textbf{H}_{Z_{\sigma}} and r(t)=tσr(t)=t_{\sigma} for each tTt\in T, where rr is a point of Isom(𝕍s,𝕍sσ)\textrm{Isom}(\mathbb{V}_{s},\mathbb{V}_{s_{\sigma}}) and the first equality is in the sense of \mathbb{Q}-subschemes. That rσdRr^{\textrm{dR}}_{\sigma} and rσr^{\ell}_{\sigma} satisfy the second condition is just the assumption that each tTt\in T has a canonical rational conjugate tσt_{\sigma}, so 1.5 reduces to the following statement:

Proposition 3.1.
  • (i)

    The map

    rσdR:𝕍,scsdR,s𝜎,sσ(csσdR)1𝕍,sσr^{\textrm{dR}}_{\sigma}:\mathbb{V}_{\mathbb{C},s}\xrightarrow{c^{\textrm{dR}}_{s}}\mathcal{H}_{\mathbb{C},s}\xrightarrow{\sigma}\mathcal{H}_{\mathbb{C},s_{\sigma}}\xrightarrow{(c^{\textrm{dR}}_{s_{\sigma}})^{-1}}\mathbb{V}_{\mathbb{C},s_{\sigma}}

    induced by σ\sigma is satisfies the property that rσdRHZ,(rσdR)1=HZσ,r^{\textrm{dR}}_{\sigma}\circ\textbf{H}_{Z,\mathbb{C}}\circ\left(r^{\textrm{dR}}_{\sigma}\right)^{-1}=\textbf{H}_{Z_{\sigma},\mathbb{C}}.

  • (ii)

    For each \ell, the map

    rσ:𝕍,scs𝕍,,s𝜎𝕍,,sσ(csσ)1𝕍,sσr^{\ell}_{\sigma}:\mathbb{V}_{\mathbb{Q}_{\ell},s}\xrightarrow{c^{\ell}_{s}}\mathbb{V}_{\ell,\mathbb{Q}_{\ell},s}\xrightarrow{\sigma}\mathbb{V}_{\ell,\mathbb{Q}_{\ell},s_{\sigma}}\xrightarrow{(c^{\ell}_{s_{\sigma}})^{-1}}\mathbb{V}_{\mathbb{Q}_{\ell},s_{\sigma}}

    induced by σ\sigma satisfies the property that induced by σ\sigma is satisfies the property that rσHZ,(rσ)1=HZσ,r^{\ell}_{\sigma}\circ\textbf{H}_{Z,\mathbb{Q}_{\ell}}\circ\left(r^{\ell}_{\sigma}\right)^{-1}=\textbf{H}_{Z_{\sigma},\mathbb{Q}_{\ell}}.

In both sections that follow, we denote the normalization of an algebraic variety ZZ by ZnZ^{n}. If ZZ is a subvariety of SS_{\mathbb{C}} and sS()s\in S(\mathbb{C}) is a point lying in Z()Z(\mathbb{C}), we will denote by sns^{n} a lift of ss to ZnZ^{n}.

3.1 The de Rham Case

The required statement is implicit in [KOU20]; we summarise the argument for expository purposes.

Let us temporarily denote by 𝕍\mathbb{V} the associated complex local system. We may argue as in the proof of [KOU20, Proposition 3.1] to obtain an equivalence of neutral Tannakian categories

𝕍|Znτ𝕍|Zσn,\langle{\left.\kern-1.2pt\mathbb{V}\vphantom{\big{|}}\right|_{Z^{n}}}\rangle^{\otimes}\simeq_{\tau}\langle{\left.\kern-1.2pt\mathbb{V}\vphantom{\big{|}}\right|_{Z^{n}_{\sigma}}}\rangle^{\otimes},

where the notation \langle-\rangle^{\otimes} denotes the neutral Tannakian category generated by the enclosed object (inside the appropriate category of local systems). Note that in the notation of [KOU20] we have 𝕍σ=𝕍\mathbb{V}^{\sigma}=\mathbb{V} from the fact that the connection \nabla is defined over KK. If FibZn,sn\textrm{Fib}_{Z^{n},s^{n}} and FibZσn,sσn\textrm{Fib}_{Z^{n}_{\sigma},s^{n}_{\sigma}} are the obvious fibre functors, then the Tannakian groups associated to FibZn,sn\textrm{Fib}_{Z^{n},s^{n}} and FibZσn,sσn\textrm{Fib}_{Z^{n}_{\sigma},s^{n}_{\sigma}} have natural faithful representations on 𝕍s\mathbb{V}_{s} and 𝕍sσ\mathbb{V}_{s_{\sigma}}, as follows from the fact that an automorphism of FibZn,sn\textrm{Fib}_{Z^{n},s^{n}} (resp. an automorphism of FibZσn,sσn\textrm{Fib}_{Z_{\sigma}^{n},s^{n}_{\sigma}}) is determined by its induced automorphism of 𝕍sn\mathbb{V}_{s^{n}} (resp. 𝕍sσn\mathbb{V}_{s^{n}_{\sigma}}).

Under these representations, the Tannakian groups associated to FibZn,sn\textrm{Fib}_{Z^{n},s^{n}} and FibZσn,sσn\textrm{Fib}_{Z^{n}_{\sigma},s^{n}_{\sigma}} agree with the complex algebraic monodromy groups of ZZ and ZσZ_{\sigma}. It follows from the fact that τ\tau is an equivalence that the induced map rsdR:𝕍,s𝕍,sσr^{\textrm{dR}}_{s}:\mathbb{V}_{\mathbb{C},s}\xrightarrow{\sim}\mathbb{V}_{\mathbb{C},s_{\sigma}} (the notation is in agreement with our previous definition of rsdRr^{\textrm{dR}}_{s}) gives an isomorphism of representations, and hence conjugates the (complex) algebraic monodromy groups.

3.2 The \ell-adic Case

We will reduce the problem to a statement about \ell-adic local systems and the étale fundamental group, which can then be solved entirely algebraically.

Definition 3.2.

Let 𝕍\mathbb{V}_{\ell} be an \ell-adic local system on the complex algebraic variety ZZ. We define H,Z\textbf{H}_{\ell,Z} to be the identity component of the Zariski closure of π1ét(Zn,sn)\pi_{1}^{\textrm{\'{e}t}}(Z^{n},s^{n}) inside Aut(𝕍,s)\textrm{Aut}(\mathbb{V}_{\ell,s})_{\mathbb{Q}_{\ell}}.

Lemma 3.3.

Let 𝕍\mathbb{V} be a \mathbb{Z}-local system on ZanZ^{\textrm{an}}, let 𝕍\mathbb{V}_{\ell} be an \ell-adic local system on ZZ, and suppose that we have an isomorphism 𝕍an𝕍\mathbb{V}^{\textrm{an}}_{\ell}\simeq\mathbb{V}\otimes\mathbb{Z}_{\ell}. Then the induced isomorphism on fibres identifies the groups HZ,\textbf{H}_{Z,\mathbb{Q}_{\ell}} and H,Z\textbf{H}_{\ell,Z}.

Proof.

Using the analytic (resp. \ell-adic) equivalence between local systems and monodromy representations, and the canonical identification of π1ét(Zn,sn)\pi_{1}^{\textrm{\'{e}t}}(Z^{n},s^{n}) with the profinite completion π1(Zn,sn)^\widehat{\pi_{1}(Z^{n},s^{n})}, the statement amounts to the following claim: given the commuting diagram

π1(Zn,sn){\pi_{1}(Z^{n},s^{n})}Aut(𝕍s){\textrm{Aut}(\mathbb{V}_{s}\otimes\mathbb{Z}_{\ell})}π1(Zn,sn)^{\widehat{\pi_{1}(Z^{n},s^{n})}}Aut(𝕍,s),{\textrm{Aut}(\mathbb{V}_{\ell,s}),}

the Zariski closures MM and M^\widehat{M} of the images im(π1(Zn,sn))\textrm{im}(\pi_{1}(Z^{n},s^{n})) and im(π1(Zn,sn)^)\textrm{im}(\widehat{\pi_{1}(Z^{n},s^{n})}) inside Aut(𝕍,s)\textrm{Aut}(\mathbb{V}_{\ell,s})_{\mathbb{Q}_{\ell}} coincide. We clearly have MM^M\subset\widehat{M} from the inclusion im(π1(Zn,sn))im(π1(Zn,sn)^)\textrm{im}(\pi_{1}(Z^{n},s^{n}))\subset\textrm{im}(\widehat{\pi_{1}(Z^{n},s^{n})}), so it suffices to show the reverse inclusion.

It follows from the fact that a group is dense in its profinite completion that if π1(Zn,sn)F\pi_{1}(Z^{n},s^{n})\to F is a morphism to a finite group FF, then the induced morphism π1(Zn,sn)^F\widehat{\pi_{1}(Z^{n},s^{n})}\to F has the same image. As a consequence, the image of π1(Zn,sn)^\widehat{\pi_{1}(Z^{n},s^{n})} inside Aut(𝕍,s)limkGL(/k)\textrm{Aut}(\mathbb{V}_{\ell,s})\cong\varprojlim_{k}\textrm{GL}(\mathbb{Z}/\ell^{k}\mathbb{Z}) consists of compatible sequences (αk)k1(\alpha_{k})_{k\geq 1} where each αk\alpha_{k} is a reduction modulo k\ell^{k} of an element in im(π1(Zn,sn))\textrm{im}(\pi_{1}(Z^{n},s^{n})). Letting ff be a polynomial function vanishing on im(π1(Zn,sn))\textrm{im}(\pi_{1}(Z^{n},s^{n})) with coefficients in \mathbb{Q}_{\ell}, it now suffices to show that ff vanishes on such compatible sequences (αk)k1(\alpha_{k})_{k\geq 1}. Scaling if necessary, we may assume that ff has coefficients in \mathbb{Z}_{\ell}. But then f(αk)=0f(\alpha_{k})=0 modulo k\ell^{k} holds for all kk, so the result follows. ∎

Applying 3.3, the proof of 3.1(ii) is reduced to showing that the conjugation isomorphism 𝕍,s𝕍,sσ\mathbb{V}_{\ell,s}\xrightarrow{\sim}\mathbb{V}_{\ell,s_{\sigma}} sends the image of π1ét(Zn,sn)\pi_{1}^{\textrm{\'{e}t}}(Z^{n},s^{n}) inside Aut(𝕍,s)\textrm{Aut}(\mathbb{V}_{\ell,s}) to the image of π1ét(Zσn,sσn)\pi_{1}^{\textrm{\'{e}t}}(Z^{n}_{\sigma},s^{n}_{\sigma}) inside Aut(𝕍,sσ)\textrm{Aut}(\mathbb{V}_{\ell,s_{\sigma}}). Applying the isomorphism 𝕍τσ𝕍\mathbb{V}_{\ell}\simeq\tau_{\sigma}^{*}\mathbb{V}_{\ell}, it suffices to show that there exists an isomorphism j:π1ét(S,s)π1ét(S,sσ)j:\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s)\xrightarrow{\sim}\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s_{\sigma}) sending im(π1ét(Zn,sn))π1ét(S,s)\textrm{im}(\pi_{1}^{\textrm{\'{e}t}}(Z^{n},s^{n}))\subset\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s) to im(π1ét(Zσn,sσn))π1ét(S,sσ)\textrm{im}(\pi_{1}^{\textrm{\'{e}t}}(Z^{n}_{\sigma},s^{n}_{\sigma}))\subset\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s_{\sigma}) and a commuting diagram of the following form:

π1ét(S,s){\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s)}Aut(𝕍,s){\textrm{Aut}(\mathbb{V}_{\ell,s})}π1ét(S,sσ){\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s_{\sigma})}Aut((τσ𝕍)sσ),{\textrm{Aut}((\tau_{\sigma}^{*}\mathbb{V}_{\ell})_{s_{\sigma}}),}j\scriptstyle{j} (1)

where the horizontal arrows are the natural representations and the vertical arrow on the right comes from the natural isomorphism 𝕍,s(τσ𝕍)sσ\mathbb{V}_{\ell,s}\simeq(\tau_{\sigma}^{*}\mathbb{V}_{\ell})_{s_{\sigma}} of étale stalks.

Conjugation Isomorphisms:

To describe the map jj, we begin with a more general construction. We let XX be a complex variety, and σAut(/K)\sigma\in\textrm{Aut}(\mathbb{C}/K). Then σ\sigma defines a categorical equivalence τσ:FÉT(X)FÉT(Xσ)\tau_{\sigma}:\textrm{F\'{E}T}(X)\to\textrm{F\'{E}T}(X_{\sigma}). Denote by FibX,x\textrm{Fib}_{X,x} the fibre functor at xXx\in X. Given απ1ét(X,x)\alpha\in\pi_{1}^{\textrm{\'{e}t}}(X,x), we may define an automorphism ασ\alpha_{\sigma} of FibXσ,xσ\textrm{Fib}_{X_{\sigma},x_{\sigma}} as follows: for each cover of XσX_{\sigma}, choose an isomorphic cover XσXσX^{\prime}_{\sigma}\to X_{\sigma} in the essential image of τσ\tau_{\sigma}; this is the conjugate of a cover XXX^{\prime}\to X of XX. Then define ασ(xσ)=α(x)σ\alpha_{\sigma}(x^{\prime}_{\sigma})=\alpha(x^{\prime})_{\sigma}. One checks that ασ\alpha_{\sigma} extends uniquely to a well-defined automorphism ασ\alpha_{\sigma} of FibXσ,sσ\textrm{Fib}_{X_{\sigma},s_{\sigma}}, and that the map π1ét(X,x)π1ét(Xσ,xσ)\pi_{1}^{\textrm{\'{e}t}}(X,x)\to\pi_{1}^{\textrm{\'{e}t}}(X_{\sigma},x_{\sigma}) defined by αασ\alpha\mapsto\alpha_{\sigma} is a group homomorphism. We define the map jj to be the case with X=SX=S_{\mathbb{C}} and x=sx=s, and the map jZnj_{Z^{n}} to be the case X=ZnX=Z^{n} and x=snx=s^{n}.

Completing the Proof:

Let iZn:π1ét(Zn,sn)π1ét(S,s)i_{Z^{n}}:\pi_{1}^{\textrm{\'{e}t}}(Z^{n},s^{n})\to\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s) and iZσn:π1ét(Zσn,sn)π1ét(S,s)i_{Z^{n}_{\sigma}}:\pi_{1}^{\textrm{\'{e}t}}(Z^{n}_{\sigma},s^{n})\to\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s) be the natural maps. It now suffices check that jιZn=ιZσnjZnj\circ\iota_{Z^{n}}=\iota_{Z^{n}_{\sigma}}\circ j_{Z^{n}} and that diagram (1) commutes. In the first case, one immediately checks that if απ1ét(S,s)\alpha\in\pi_{1}^{\textrm{\'{e}t}}(S_{\mathbb{C}},s) acts on fibres of SSS^{\prime}\to S_{\mathbb{C}} above ss by base changing to a cover of ZnZ^{n} and acting via π1ét(Zn,sn)\pi_{1}^{\textrm{\'{e}t}}(Z^{n},s^{n}), then the same is true for ασ\alpha_{\sigma} with respect to ZσnZ^{n}_{\sigma} and π1ét(Zσn,sσn)\pi_{1}^{\textrm{\'{e}t}}(Z^{n}_{\sigma},s^{n}_{\sigma}). In the second case, it is immediate from the explicit description of the isomorphism t:𝕍,s(τσ𝕍)sσt:\mathbb{V}_{\ell,s}\xrightarrow{\sim}(\tau_{\sigma}^{*}\mathbb{V}_{\ell})_{s_{\sigma}}, that together with the map jj, the map tt gives a map of representations (i.e., acting by α\alpha then applying tt is the same as applying tt and then acting by j(α)j(\alpha)). But tt giving a map of representations is equivalent to the commutativity of (1).

4 Variants of the Main Theorem

Although our applications will use 1.3 and 1.7, we wish to briefly explain how 1.5 may be used to prove variants of 1.3 and 1.7 which may be of independent interest. Let us first generalize the notation T(s,Z,σ)\mathcal{I}_{T}(s,Z,\sigma) established in the introduction.

Definition 4.1.

Let sS()s\in S(\mathbb{C}) be a point, {Zi}iI\{Z_{i}\}_{i\in I} a collection of subvarieties of SS_{\mathbb{C}} containing ss, σAut(/K)\sigma\in\textrm{Aut}(\mathbb{C}/K) an automorphism, and T𝒯(𝕍s)T\subset\mathcal{T}(\mathbb{V}_{s}) a collection of tensors with canonical rational conjugates. Then we define

=T(s,{Zi}iI,σ)Isom(𝕍s,𝕍sσ)\mathcal{I}=\mathcal{I}_{T}(s,\{Z_{i}\}_{i\in I},\sigma)\subset\textrm{Isom}(\mathbb{V}_{s},\mathbb{V}_{s_{\sigma}})

by the property that r()r\in\mathcal{I}(\mathbb{C}) if and only if rHZi,r1=Hi,σ,r\circ\textbf{H}_{Z_{i},\mathbb{C}}\circ r^{-1}=\textbf{H}_{\mathbb{Z}_{i,\sigma},\mathbb{C}} for all ii, and rr sends each tTt\in T to its canonical rational conjugate tσTσt_{\sigma}\in T_{\sigma}.

We observe that the normalizer NN in the statement of 1.3 is equal to T(s,Z,id)\mathcal{I}_{T}(s,Z,\textrm{id}), so we may view T(s,{Z}iI,id)\mathcal{I}_{T}(s,\{Z\}_{i\in I},\textrm{id}) as its natural generalization. The following generalization of 1.3 is then immediate from 3.1:

Theorem 4.2.

Let 𝕍\mathbb{V} be a motivic variation of Hodge structure on SS, and let sS()s\in S(\mathbb{C}). Suppose that T𝒯(𝕍s)T\subset\mathcal{T}(\mathbb{V}_{s}) is a subspace with canonical rational conjugates, that ZiSZ_{i}\subset S_{\mathbb{C}} for iIi\in I is a collection of complex subvarieties containing ss, and that W𝒯(𝕍s)W\subset\mathcal{T}(\mathbb{V}_{s}) is a subspace containing TT such that either

  • (i)

    T(s,{Zi}iI,id)\mathcal{I}_{T}(s,\{Z_{i}\}_{i\in I},\textrm{id}) fixes WW as a \mathbb{Q}-subspace; or

  • (ii)

    T(s,{Zi}iI,id)\mathcal{I}_{T}(s,\{Z_{i}\}_{i\in I},\textrm{id}) fixes WW pointwise.

Then

  • (i)

    each element ww of WW has rational conjugates; or

  • (ii)

    each element ww of WW has a canonical rational conjugate,

in cases (i) and (ii) respectively.

Proof.

As in the proof of 1.3, we obtain from 3.1 that rσdRr^{\textrm{dR}}_{\sigma} and rσr^{\ell}_{\sigma} for all \ell are points of =T(s,{Zi}iI,σ)\mathcal{I}=\mathcal{I}_{T}(s,\{Z_{i}\}_{i\in I},\sigma). It follows as in 1.3 that the points of WW are defined over ¯\overline{\mathbb{Q}} and (under an appropriate isomorphism ¯\mathbb{C}\cong\overline{\mathbb{Q}_{\ell}}) over ¯\overline{\mathbb{Q}}_{\ell} for every \ell. By 1.6 we conclude in that r(Wσ)r(W_{\sigma}) is a \mathbb{Q}-subspace, where r()r\in\mathcal{I}(\mathbb{C}). In Case (ii) we additionally know that r(w)=wσr(w)=w_{\sigma} is independent of the choice of rr, making the conjugates canonical. ∎

Finally let us give a similar, but simpler argument which establishes cases of a kind of “¯\overline{\mathbb{Q}}-absolute Hodge” conjecture. We note that a “¯\overline{\mathbb{Q}}-absolute Hodge” conjecture may in fact be enough for certain applications of absolute Hodge to the algebraicity of periods. The following requires no \ell-adic input.

Theorem 4.3.

Let 𝕍\mathbb{V} be a variation of Hodge structure on SS, suppose that SS, \nabla and \mathcal{H} are defined over KK, and choose sS()s\in S(\mathbb{C}). Let T𝒯(𝕍¯,s)T\subset\mathcal{T}(\mathbb{V}_{\overline{\mathbb{Q}},s}) be a ¯\overline{\mathbb{Q}}-subspace with canonical rational conjugates, let ZiSZ_{i}\subset S_{\mathbb{C}} for iIi\in I be a collection of complex subvarieties containing ss, and suppose that W𝒯(𝕍¯,s)W\subset\mathcal{T}(\mathbb{V}_{\overline{\mathbb{Q}},s}) is a subspace containing TT such that the orbit of [W][W] under =T(s,{Zi}iI,id)\mathcal{I}=\mathcal{I}_{T}(s,\{Z_{i}\}_{i\in I},\textrm{id}) inside the appropriate Grassmanian is finite. Then WW has ¯\overline{\mathbb{Q}}-rational conjugates.

Proof.

The proof follows immediately, as 3.1 ensures that rσdRr^{\textrm{dR}}_{\sigma} (and rσr^{\ell}_{\sigma}, if an appropriate comparison to an \ell-adic local system exists) lies inside \mathcal{I} and the assumptions ensure that r(W)r(W) is defined over ¯\overline{\mathbb{Q}} for any point rr of \mathcal{I}. ∎

Finally, we note that there is an additional step involved in translating 4.2 and 4.3 to absolute Hodge statements (like 1.7), but this follows exactly as in section 2.3 and we leave this to the reader.

5 Applications to Absolute Hodge

5.1 Voisin’s Example

In this section we prove 1.8, following an approach laid out by Voisin in Section 3 of [Voi07]. Let us first revisit the proof of our 1.4 in the context of 1.7.

We let 𝕍\mathbb{V} be a motivic variation on the smooth KK-variety SS, and T𝒯(𝕍s)T\subset\mathcal{T}(\mathbb{V}_{s}) be the \mathbb{Q}-span of the polarization QsQ_{s}; by 2.1 the subspace TT has canonical rational conjugates. Suppose that there exists a subvariety ZSZ\subset S_{\mathbb{C}} containing ss such that the fixed locus of HZ\textbf{H}_{Z} inside 𝕍sm,n\mathbb{V}^{m,n}_{s} is the line spanned by the Hodge class w𝕍sm,nw\in\mathbb{V}^{m,n}_{s}. Then if nn is a complex point of the normalizer of HZ\textbf{H}_{Z} inside 𝐆T\mathbf{G}_{T}, we have that nw=λwnw=\lambda w for some λ\lambda\in\mathbb{C}, and from the fact that Q(nw,nw)=Q(w,w)Q(nw,nw)=Q(w,w) we learn that λ=±1\lambda=\pm 1. 1.7 therefore applies, and the tensor ww is absolutely Hodge.

A very similar argument is given by Voisin in her paper (see Remark 1.2, Theorem 0.5(1) in [Voi07]). The language used is slightly different: Voisin only considers Hodge vectors vv (the (m,n)=(0,1)(m,n)=(0,1) case); the variety ZZ is taken to be a special subvariety (irreducible component of the Hodge locus); and the condition that the fixed locus of HZ\textbf{H}_{Z} consist of exactly the line spanned by ww takes the form of the condition that the restricted variation 𝕍|Z{\left.\kern-1.2pt\mathbb{V}\vphantom{\big{|}}\right|_{Z}} has w\mathbb{Q}w as its constant subvariation. Voisin also only obtains the weaker conclusion that the conjugate vσv_{\sigma} is a ¯\overline{\mathbb{Q}}-scalar multiple of a Hodge class; the essential difference, it seems to us, is the presence of \ell-adic input in our argument.

To study the example of 1.8, Voisin states a criterion [Voi07, Theorem 3.1], whose proof gives the following:

Proposition 5.1 (Voisin).

Let 𝕍\mathbb{V} be a polarizable variation of Hodge structure of weight 2k2k. Denote by ¯p,q\overline{\nabla}^{p,q} the 𝒪S\mathcal{O}_{S}-linear map p,qp1,q+1ΩS\mathcal{H}^{p,q}\to\mathcal{H}^{p-1,q+1}\otimes\Omega_{S} induced by the connection, let α𝕍s\alpha\in\mathbb{V}_{s} be a Hodge class with Hodge locus ZαZ_{\alpha}, and let λ\lambda be its projection to k,k\mathcal{H}^{k,k}. Suppose that

  • (i)

    the map μλ(v)=¯v(λ)\mu_{\lambda}(v)=\overline{\nabla}_{v}(\lambda) is surjective;

  • (ii)

    for p>k,p+q=2kp>k,p+q=2k the restriction of ¯p,q\overline{\nabla}^{p,q} to the the tangent space TsZαT_{s}Z_{\alpha} is injective;

  • (iii)

    and the restriction of ¯k,k\overline{\nabla}^{k,k} to the tangent space TsZαT_{s}Z_{\alpha} has kernel equal to the span of λ\lambda.

Then the fixed locus in 𝕍s\mathbb{V}_{s} of HZα\textbf{H}_{Z_{\alpha}} is exactly the line spanned by α\alpha.

Proof of 1.8:.

We observe that the variation 𝕍\mathbb{V}_{-} is motivic: taking anti-invariants defines an appropriate subbundle \mathcal{H}_{-}\subset\mathcal{H} and subsystem 𝕍,𝕍\mathbb{V}_{\ell,-}\subset\mathbb{V}_{\ell}, compatibly with the comparison isomorphisms, and the required properties are simply obtained by restriction. Arguing as Voisin does in [Voi07, Example 3.4], we verify that the hypotheses of 5.1 hold for classes α\alpha whose projections λ\lambda lie in a certain topologically dense subset of the underlying real subbundle of k,k\mathcal{H}^{k,k}_{-}. The result then follows from 1.7. ∎

5.2 Absolutely Hodge Isomorphisms

In this section we study the question of when an isomorphism of Hodge structures between two fibres in a motivic family is absolute Hodge. Related questions are considered in [Voi07, Section 3].

Definition 5.2.

Let S1S_{1} and S2S_{2} be algebraic varieties, and let 𝕍1\mathbb{V}_{1} and 𝕍2\mathbb{V}_{2} be variations of Hodge structure on S1S_{1} and S2S_{2}, respectively. We define 𝕍1𝕍2\mathbb{V}_{1}\boxtimes\mathbb{V}_{2} to be the variation p1𝕍1p2𝕍2p^{*}_{1}\mathbb{V}_{1}\otimes p^{*}_{2}\mathbb{V}_{2} on S1×S2S_{1}\times S_{2}, where pi:S1×S2Sip_{i}:S_{1}\times S_{2}\to S_{i} is the projection.

Suppose that 𝕍\mathbb{V} is an integral variation of Hodge structure on SS, and let ZS×SZ\subset S\times S be the algebraic subvariety whose points are pairs (x,y)(x,y) such that 𝕍x\mathbb{V}_{x} is isomorphic to 𝕍y\mathbb{V}_{y} as an integral polarized Hodge structure. These isomorphisms are Hodge tensors of the variation 𝕍𝕍\mathbb{V}\boxtimes\mathbb{V}^{*}. When a Torelli theorem is avaliable, the locus ZZ is simply equal to the diagonal Δ\Delta, and its algebraic monodromy is easily determined to be equal to the image of HS\textbf{H}_{S} under the diagonal action. In the general case, let Z0Z_{0} be an irreducible component of ZZ containing the diagonal ΔS×S\Delta\subset S\times S. We determine HZ0\textbf{H}_{Z_{0}} by reducing to the Torelli case using a recent result of Bakker, Brunebarbe and Tsimerman [BBT18].

To show our main result we first prove some lemmas on algebraic monodromy:

Lemma 5.3.

Let f:STf:S\to T be a proper, dominant morphism of irreducible complex algebraic varieties, and let 𝕍\mathbb{V} be a local system on TT. Then for any point sSs\in S, the isomorphism (f𝕍)s𝕍f(s)(f^{*}\mathbb{V})_{s}\xrightarrow{\sim}\mathbb{V}_{f(s)} induces an isomorphism between HS\textbf{H}_{S} and HT\textbf{H}_{T}.

Proof.

That f:STf:S\to T is proper means that it has a Stein factorization f=ghf=g\circ h, where h:SVh:S\to V is proper with connected fibres and gg is finite. We are thus reduced to showing the theorem in two cases:

  • (i)

    when ff additionally has connected fibres;

  • (ii)

    and when ff is additionally finite.

Let us first assume that both SS and TT are normal. If XX is a normal variety, then the fundamental group of any open subvariety UU surjects onto the fundamental group of XX, so we may replace both SS and TT by Zariski open sets, restricting ff appropriately. In case (i) this lets us assume that ff is surjective and flat, and hence by [Sta20, Lemma 01UA] universally open. Kollár has shown in [Kol19] that a universally open, surjective morphism of complex varieties with connected fibres satisfies the two-point path lifting property (see [Kol19, Definition 30]), and therefore induces a surjection of fundamental groups, from which the result follows. In case (ii) we may assume that ff is étale, from which the result follows as the monodromy of SS will be finite index in the monodromy of TT, and hence have the same Zariski closure.

Working now in the general case, it follows from the definition of algebraic monodromy that the algebraic monodromy of a local system 𝕍\mathbb{V} on XX agrees with the algebraic monodromy of the restriction of 𝕍\mathbb{V} to the normal locus of XX. The result then follows by restricting ff. ∎

Lemma 5.4.

Let 𝕍\mathbb{V} be a polarizable integral variation of Hodge structure on the smooth complex algebraic variety SS, and let Z0Z_{0} and Δ\Delta be as above. Then HZ0=HΔ\textbf{H}_{Z_{0}}=\textbf{H}_{\Delta}.

Proof.

The statement is unchanged under replacing SS with a finite étale cover, so we may assume that SS has unipotent monodromy at infinity. Let φ:SΓ\D\varphi:S\to\Gamma\backslash D be the period map on SS; here DD is the full period domain of polarized integral Hodge structures on a fixed lattice VV, and Γ=G()\Gamma=G(\mathbb{Z}) where G=Aut(V,Q)G=\textrm{Aut}(V,Q). Then Γ\D\Gamma\backslash D may be interpreted as the moduli space for polarized integral Hodge structures of the same type as the fibres of 𝕍\mathbb{V}, and the map φ\varphi sends ss to the isomorphism class of 𝕍s\mathbb{V}_{s}. Arguing as in [CPMS03, Corollary 13.7.6] we may complete the variety SS by adding finitely many points so that φ\varphi is proper. Applying the main theorem of [BBT18], we find that there exists a factorization φ=φg\varphi=\varphi^{\prime}\circ g where g:STg:S\to T is a dominant proper map of algebraic varieties and φ:TΓ\D\varphi:T\hookrightarrow\Gamma\backslash D is a closed embedding.

The variety TT admits a variation 𝕍\mathbb{V}^{\prime} such that 𝕍\mathbb{V} may be identified with g𝕍g^{*}\mathbb{V}^{\prime}, and φ\varphi^{\prime} is the period map of 𝕍\mathbb{V}^{\prime}. The variety ZZ is then the inverse image of the diagonal under the product morphism g×g:S×ST×Tg\times g:S\times S\to T\times T. If ΔSS×S\Delta_{S}\subset S\times S and ΔTT×T\Delta_{T}\subset T\times T are the respective diagonals, the maps Z0ΔTZ_{0}\to\Delta_{T} and ΔSΔT\Delta_{S}\to\Delta_{T} are both dominant and proper, so the result follows by 5.3. ∎

Let us note that it is a consequence of Deligne’s Principle A (see [Del82b, Theorem 3.8]) that if the tensor QsQ_{s} has canonical rational conjugates, then so does every rational tensor fixed by Aut(𝕍s,Qs)\textrm{Aut}(\mathbb{V}_{s},Q_{s}). 1.9 now follows from the more general:

Theorem 5.5.

If 𝕍\mathbb{V} is a motivic variation of Hodge structure on the smooth KK-variety SS such that rational tensors fixed by HS\textbf{H}_{S} have canonical rational conjugates and the points of the center 𝒵(HS)\mathcal{Z}(\textbf{H}_{S}) are defined over \mathbb{Q}, then tensors fixed by HZ0\textbf{H}_{Z_{0}} have rational conjugates, and generic isomorphisms between 𝕍x\mathbb{V}_{x} and 𝕍y\mathbb{V}_{y} are KK-absolutely Hodge.

Proof.

Let T𝒯(𝕍s)T\subset\mathcal{T}(\mathbb{V}_{s}) be the subspace of \mathbb{Q}-tensors fixed by HS\textbf{H}_{S}, and let TT^{*} be its dual. As the subspace TT (resp. the subspace TT^{*}) is fixed under algebraic monodromy, parallel transport gives a path-independent translate of TT (resp. TT^{*}) to any fibre of 𝕍\mathbb{V} (resp. 𝕍\mathbb{V}^{*}). It follows that we may unambiguously refer to the subspace R=TTR=T\otimes T^{*} inside 𝒯(𝕍x𝕍y)\mathcal{T}(\mathbb{V}_{x}\otimes\mathbb{V}^{*}_{y}) for any pair (x,y)(x,y). The fact that TT has canonical rational conjugates implies the same fact for RR. We then have:

Claim 5.6.

The groups 𝐆R\mathbf{G}_{R} and 𝐆T×𝐆T\mathbf{G}_{T}\times\mathbf{G}_{T^{*}} are equal as subgroups of Aut(𝕍x𝕍y)\textrm{Aut}(\mathbb{V}_{x}\otimes\mathbb{V}^{*}_{y}).

Proof.

Let U𝒯(𝕍x𝕍y)U\subset\mathcal{T}(\mathbb{V}_{x}\otimes\mathbb{V}^{*}_{y}) be the subspace of tensors fixed by 𝐆T×𝐆T\mathbf{G}_{T}\times\mathbf{G}_{T^{*}}. It suffices to show that U=TTU=T\otimes T^{*}. Let us first argue that the space UU is spanned by pure tensors; i.e., it has a basis of the form uuu\otimes u^{\prime} where u𝒯(𝕍x)u\in\mathcal{T}(\mathbb{V}_{x}) and u𝒯(𝕍y)u^{\prime}\in\mathcal{T}(\mathbb{V}^{*}_{y}). Indeed, suppose that 𝐆T×𝐆T\mathbf{G}_{T}\times\mathbf{G}_{T^{*}} fixes k=1mvkvk\sum_{k=1}^{m}v_{k}\otimes v^{\prime}_{k}, where {vk}k=1m\{v_{k}\}_{k=1}^{m} and {vk}k=1m\{v^{\prime}_{k}\}_{k=1}^{m} form linearly independent sets. Let (g,g)𝐆T()×𝐆T()(g,g^{\prime})\in\mathbf{G}_{T}(\mathbb{C})\times\mathbf{G}_{T^{*}}(\mathbb{C}) be a point. Then from the fact that (g,1)(g,1) fixes k=1mvkvk\sum_{k=1}^{m}v_{k}\otimes v^{\prime}_{k} we learn that gg fixes each vkv_{k}, and analogously we learn that (1,g)(1,g^{\prime}) fixes each vkv^{\prime}_{k}. It follows that (g,g)(g,g^{\prime}) fixes each term in the sum.

We are reduced to arguing that any element of the form vvv\otimes v^{\prime} with v𝒯(𝕍x)v\in\mathcal{T}(\mathbb{V}_{x}) and v𝒯(𝕍y)v^{\prime}\in\mathcal{T}(\mathbb{V}^{*}_{y}) fixed by 𝐆T×𝐆T\mathbf{G}_{T}\times\mathbf{G}_{T^{*}} must satisfy vTv\in T and vTv^{\prime}\in T^{*}. From the fact that 𝐆T×𝐆T\mathbf{G}_{T}\times\mathbf{G}_{T^{*}} acts on vv and vv^{\prime} separately we learn immediately that 𝐆T\mathbf{G}_{T} (resp. 𝐆T\mathbf{G}_{T^{*}}) acts on vv (resp. vv^{\prime}) through a character. As 𝐆T\mathbf{G}_{T} (resp. 𝐆T\mathbf{G}_{T^{*}}) is a connected, semisimple group (isomorphic to HS\textbf{H}_{S}), we learn that 𝐆T\mathbf{G}_{T} (resp. 𝐆T\mathbf{G}_{T^{*}}) fixes vv (resp. vv^{\prime}). ∎

We may now complete the proof by computing the normalizer NN of HZ0\textbf{H}_{Z_{0}} inside HS×HS\textbf{H}_{S}\times\textbf{H}_{S}. It follows from 5.4 that HZ0\textbf{H}_{Z_{0}} is the diagonal subgroup DD of HS×HS\textbf{H}_{S}\times\textbf{H}_{S}. The normalizer NN of DD is the group generated by DD and the product of centers 𝒵(HS)×𝒵(HS)\mathcal{Z}(\textbf{H}_{S})\times\mathcal{Z}(\textbf{H}_{S}). It follows that NN preserves the subspace WW of all rational tensors fixed by HZ0\textbf{H}_{Z_{0}}, and so the result follows by 1.3 and 1.7. ∎

References

  • [BBT18] Benjamin Bakker, Yohan Brunebarbe, and Jacob Tsimerman. o-minimal GAGA and a conjecture of Griffiths. arXiv e-prints, page arXiv:1811.12230, November 2018.
  • [CF67] John William Scott Cassels and Albrecht Frölich. Algebraic number theory: proceedings of an instructional conference. Academic Pr, 1967.
  • [Con] Brian Conrad. Étale cohomology. Available at virtualmath1.stanford.edu/ conrad/Weil2seminar/Notes/etnotes.pdf.
  • [CPMS03] James A Carlson, C. (Chris) Peters, and Stefan Müller-Stach. Period mappings and period domains. Cambridge, U.K. : Cambridge University Press, 2003. Includes bibliographical references and index.
  • [CS14] Francois Charles and Christian Schnell. Notes on Absolute Hodge Classes. In Hodge Theory, pages 469 – 530. Princeton University Press, Princeton, 31 Dec. 2014.
  • [Del71] Pierre Deligne. Théorie de Hodge : II. Publications Mathématiques de l’IHÉS, 40:5–57, 1971.
  • [Del82a] P. Deligne. Hodge Cycles on Abelian Varieties, pages 9–100. Springer Berlin Heidelberg, Berlin, Heidelberg, 1982.
  • [Del82b] Pierre Deligne. Hodge cycles on abelian varieties (notes by J. Milne). In Hodge Cycles, Motives, and Shimura Varieties. Lecture Notes in Mathematics, vol 900. Springer. 1982.
  • [Har75] R. Hartshorne. On the de Rham cohomology of algebraic varieties. Publications Mathématiques de l’IHÉS, 45:6–99, 1975.
  • [Har77] Robin Hartshorne. Algebraic Geometry. Springer New York, New York, NY, 1977.
  • [KO68] Nicholas M. Katz and Tadao Oda. On the differentiation of De Rham cohomology classes with respect to parameters. J. Math. Kyoto Univ., 8(2):199–213, 1968.
  • [Kol19] János Kollár. Fundamental groups and path lifting for algebraic varieties. arXiv e-prints, page arXiv:1906.11816, June 2019.
  • [KOU20] Bruno Klingler, Anna Otwinowska, and David Urbanik. On the fields of definition of Hodge loci, 2020.
  • [Mil80] J. S. Milne. Etale Cohomology (PMS-33). Princeton University Press, 1980.
  • [Sta20] The Stacks project authors. The stacks project. https://stacks.math.columbia.edu, 2020.
  • [Voi07] Claire Voisin. Hodge loci and absolute Hodge classes. Compositio Mathematica, 143(4):945–958, 2007.