Absolute Hodge and -adic Monodromy
Abstract
Let be a motivic variation of Hodge structure on a -variety , let be the associated -algebraic Hodge bundle, and let be an automorphism. The absolute Hodge conjecture predicts that given a Hodge vector above which lies inside , the conjugate vector is Hodge and lies inside . We study this problem in the situation where we have an algebraic subvariety containing whose algebraic monodromy group fixes . Using relationships between and coming from the theories of complex and -adic local systems, we establish a criterion that implies the absolute Hodge conjecture for subject to a group-theoretic condition on . We then use our criterion to establish new cases of the absolute Hodge conjecture.
1 Introduction
Let be a smooth projective111We follow the convention in [Har77], so this means that there exists an embedding over . morphism of -varieties, with smooth and quasi-projective, and a subfield. In this setting, the local system underlies a polarizable variation of Hodge structure, which has the property that the vector bundle admits a -algebraic model , where is the complex of relative differentials. Moreover, the underlying integral local system admits a comparison with the -adic local system on the étale site of .
Notation.
For , we denote by and the natural comparison isomorphisms.
The fact that and come from a morphism of -varieties means that for each automorphism and point , there are isomorphisms and induced by conjugation. Combining these isomorphisms with the isomorphisms and , we may consider several possible ways in which the rational structure of may be preserved under conjugation:
Definition 1.1.
Let be a point, be a vector, a subspace, and . Then we say:
-
(i)
That has rational conjugates if (resp. for all ) are rational vectors of . (They lie in the image of under the comparison isomorphisms and .)
-
(ii)
That has a canonical rational conjugate if there exists such that and for all . (The vector works for all comparison isomorphisms at once.)
-
(iii)
That has rational conjugates if each element of does.
-
(iv)
That has a canonical rational conjugate if each element of does.
Remark.
It is clear that if and have (canonical) rational conjugates then so do and for any , so we lose nothing by considering subspaces with (canonical) rational conjugates.
In the case , the property that a Hodge vector has rational conjugates is equivalent to the absolute Hodge conjecture for formulated in [CS14].222Although the formulation in [Del82a], notably, is equivalent to the condition that has a canonical rational conjugate; see also the discussion in section 2.3. The property that 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 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 (or a subspace of the various tensor powers of ) has rational conjugates by studying the monodromy of complex subvarieties . To formulate the statement we introduce some notation.
Notation.
Given an -module , we let , with the analogous definition for -local systems .
Notation.
Given an -module , we let , with the analogous definition for -local systems .
Notation.
Let be a vector space over , and let be a vector subspace. We denote by the -algebraic subgroup defined by the property that it fixes each element of .
Notation.
Let be an algebraic subvariety, and be a local system. We denote by the algebraic monodromy of , which is the identity component of the Zariski closure of the monodromy representation on , where is the normalization of .
Remark.
Given a subvariety , we may regard as an algebraic subgroup of for any point . We will use the notation for each of these subgroups, even though the point 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 is an algebraic group, and is a subspace, we will say that fixes as a -subspace if for each we have .333We wish to stress that this condition is different from fixing pointwise, or fixing as a subspace.
Theorem 1.3.
Let be a motivic variation of Hodge structure on , and let . Suppose that is a subspace with canonical rational conjugates, that is a complex subvariety containing satisfying , and that is a subspace containing . Suppose that the normalizer of inside fixes as a -subspace. Then has rational conjugates.
A simple application of 1.3 is then:
Corollary 1.4.
Suppose is a motivic variation of Hodge structure on , let and suppose that contains . Then if the only tensors fixed by are of the form for some , then the subspace has rational conjugates.
Proof.
We show in section 2.2 that any motivic variation admits a polarization such that the class of inside has a canonical rational conjugate. Take to be the span of . From the fact that preserves we learn it acts on by . Taking to be the span of and , we learn that 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 be the -algebraic variety consisting of all isomorphisms between and . Our proof of 1.3 proceeds through the study of the -subvariety
defined by the property that if and only if and sends each to its canonical rational conjugate . We observe that is a torsor under the normalizer of inside . 1.3 follows immediately from the following two facts about , established in section 3:
Proposition 1.5.
-
(i)
The map
induced by is a -point of .
-
(ii)
For each , the map
induced by is a -point of .
Proof of 1.3 (assuming 1.5):.
We let for ease of notation. That is naturally a torsor under implies that the subspace is independent of the choice of : any two choices and are related by with , and fixes by assumption. As is a -variety, it has a -point , hence is defined over . If we can show that is in fact a -subspace of , then the proof will be complete as 1.5 then implies both that and for each .
By 1.5(ii) the variety has -points for every , hence it follows that under some choice of isomorphism , every element of is defined over . The proof is completed by the following:
Claim 1.6.
Let be a -point in affine space, and suppose that for each there exists an isomorphism under which is identified with a -point. Then is defined over .
Proof.
As , it suffices to consider the case . Let be the minimal polynomial of . The statement is unchanged by replacing with for , so we may assume that is integral and that is monic with integral coefficients. The hypothesis shows (using integrality) that has a root over for every , hence a root modulo for every . As is irreducible, this implies that is linear by [CF67, p. 362, Ex. 6.2]. ∎
∎
Remark.
The proof of 1.3 does not invoke the properties of the filtration 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 , , , , and be as in 1.3, and suppose that is a subspace of Hodge vectors. Then the -absolute Hodge conjecture444See section 2.3 for a definition; the usual absolute Hodge conjecture is the case . holds for each element of .
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 be the family of degree hypersurfaces in stable under the involution , let be the geometric variation on middle cohomology of degree , and let be the subvariation obtained by taking anti-invariants under . Then if is the quotient , the rational Hodge classes for which the absolute Hodge conjecture holds are topologically dense in the real part of .
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 -adic input in our argument.
As a second application, we let be a motivic variation on , let be the algebraic locus of points such that as integral polarized Hodge structures, and we let be an irreducible component containing the diagonal . The variety is a component of the Hodge locus for the variation on (see 5.2). We call an isomorphism with generic if its class is Hodge at a generic point of . We have the following result:
Theorem 1.9.
If is a motivic variation of Hodge structure on the smooth -variety such that (monodromy is Zariski dense), then rational tensors fixed by have rational conjugates, and hence generic isomorphisms between and are -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 -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 essentially conjugates the representation of on the fibre to the representation of on the fibre .
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 obtained from a smooth, projective family of -varieties, with smooth. Let us briefly state some properties of such variations:
-
(i)
The Hodge bundle admits a canonical model as a -algebraic vector bundle , where is the complex of relative algebraic differentials.
-
(ii)
The filtration on admits a canonical -algebraic model, agreeing with the filtration in the analytic setting. This is obtained from a filtration on and an appropriate spectral sequence.
-
(iii)
The connection whose flat sections are given by is -algebraic. This is due to Katz and Oda [KO68].
-
(iv)
The -local system admits a canonical integral subsystem , such that admits a comparison isomorphism with , where is an -adic local system defined on the étale site of . This is the relative version of the comparison between Betti and étale cohomology.
The following additional properties relate to the polarization on , and will be established in the next section:
-
(v)
The variation admits a polarization which is -algebraic, in the sense that we have a morphism of -algebraic vector bundles whose analytification is identified with after applying the isomorphism .
-
(vi)
There exists a bilinear form , compatible with the polarization in (v).
-
(vii)
Let , and denote by the induced automorphism of the étale site of . Then there exists a canonical isomorphism , compatible with the polarization 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 -conjugate of the class is rational for (and given by ) in both the de Rham and -adic setting, where . ∎
2.2 Construction of -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 , and . We will define the primitive cohomology subsystems and , and vector subbundle , as well as polarizations and 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 is a point, our definition of and essentially appears (among other places) in [Del71]. Our goal is to generalize this to the relative setting, where we regard the scheme as an -scheme via the map . The first task is to define the relative Chern classes and of a line bundle on .
Lemma 2.2.
Let be a line bundle on . Then there exists a (necessarily unique) global section of whose restriction to the fibre at is the Chern class .555In the sense that the equality holds after identifying with the algebraic de Rham cohomology of the fibre using proper base change.
Proof.
By [Sta20, Section 0FLE] we obtain a class , where is the cohomology of the complex . Choose an injective resolution . We may compute as the cohomology sheaf of in degree two. From the equality we therefore obtain a morphism , and we may define as the image of under .
Let be a point, and be the restriction of a global section to the fibre at . Let be the canonical base change morphism. As Chern classes are functorial, the map induced by the inclusion sends to . It therefore suffices to check that .
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 with the pullback , and find an injective resolution . We then obtain a commuting diagram
,
where the map is unique up to homotopy. The map (resp. the map ) is constructed from by applying (resp. applying ) and taking the induced map on cohomology. The required equality then follows from the fact that . ∎
Lemma 2.3.
Let be an algebraic closure of inside , and let be a line bundle on . Then there exists a (necessarily unique) global section of whose restriction to the fibre at is the Chern class .666In the same sense as in 2.2.
Proof.
We argue analogously to 2.2. For ease of notation, we assume and so replace by and by . By [Mil80, VI, §10] we obtain a class . Choose an injective resolution .777Here we really mean a compatible system of resolutions , where we regard -adic sheaves as systems of -sheaves in the usual way. Proceeding as before, we may compute from the degree two cohomology of , and as the degree two cohomology sheaf of . From the equality we therefore obtain a map , and define as the image of under .
Letting be a (geometric) point, one defines and analogously to 2.2, and similarly checks that . ∎
Definition 2.4.
Let be a smooth projective morphism of -varieties, with smooth and fibres of dimension . Let , and . Let be a very ample bundle over , and let , and be the analytification of . We define:
-
(i)
the operators
-
(ii)
and the subbundle and subsystems
Now fix isomorphisms
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
Proposition 2.5.
The polarizations and 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 and . For property (vii) this follows from the compatibility of the cup product with conjugation by , as well as the fact that the section is defined over . ∎
2.3 The Absolute Hodge Conjecture
Let be a smooth complex projective variety, and a non-negative integer. We define by the intersection , where is the Hodge decomposition. According to [CS14], the absolute Hodge conjecture says the following:
Conjecture 2.6 (Absolute Hodge).
Let , and let and be the comparison isomorphisms. Then
-
(i)
if is the image of under then is the image of some under ;
-
(ii)
if is the image of under then is the image of some under .
Remark.
Deligne in [Del82a] requires the class 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 replaced by the -absolute Hodge conjecture.
Definition 2.8.
By the (-)absolute Hodge conjecture for , we mean the statement of 2.6 for some fixed vector .
Proposition 2.9.
In the situation of 1.3, if each element of is Hodge, then the -absolute Hodge conjecture holds for each element of .
Proof.
What needs to be checked is that if is a rational conjugate to , then is automatically Hodge. We recall that a rational vector inside is Hodge if and only if it lies inside , where is the Hodge filtration. As is a filtration by -algebraic bundles on , this shows the result in the de Rham case. In the -adic case, we use the fact that is independent of the point of , hence taking we see that the result holds -adically as well. ∎
3 Conjugation Isomorphisms and Monodromy
In this section we establish 1.5. The variety is defined by the conditions conditions and for each , where is a point of and the first equality is in the sense of -subschemes. That and satisfy the second condition is just the assumption that each has a canonical rational conjugate , so 1.5 reduces to the following statement:
Proposition 3.1.
-
(i)
The map
induced by is satisfies the property that .
-
(ii)
For each , the map
induced by satisfies the property that induced by is satisfies the property that .
In both sections that follow, we denote the normalization of an algebraic variety by . If is a subvariety of and is a point lying in , we will denote by a lift of to .
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 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
where the notation 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 from the fact that the connection is defined over . If and are the obvious fibre functors, then the Tannakian groups associated to and have natural faithful representations on and , as follows from the fact that an automorphism of (resp. an automorphism of ) is determined by its induced automorphism of (resp. ).
Under these representations, the Tannakian groups associated to and agree with the complex algebraic monodromy groups of and . It follows from the fact that is an equivalence that the induced map (the notation is in agreement with our previous definition of ) gives an isomorphism of representations, and hence conjugates the (complex) algebraic monodromy groups.
3.2 The -adic Case
We will reduce the problem to a statement about -adic local systems and the étale fundamental group, which can then be solved entirely algebraically.
Definition 3.2.
Let be an -adic local system on the complex algebraic variety . We define to be the identity component of the Zariski closure of inside .
Lemma 3.3.
Let be a -local system on , let be an -adic local system on , and suppose that we have an isomorphism . Then the induced isomorphism on fibres identifies the groups and .
Proof.
Using the analytic (resp. -adic) equivalence between local systems and monodromy representations, and the canonical identification of with the profinite completion , the statement amounts to the following claim: given the commuting diagram
the Zariski closures and of the images and inside coincide. We clearly have from the inclusion , so it suffices to show the reverse inclusion.
It follows from the fact that a group is dense in its profinite completion that if is a morphism to a finite group , then the induced morphism has the same image. As a consequence, the image of inside consists of compatible sequences where each is a reduction modulo of an element in . Letting be a polynomial function vanishing on with coefficients in , it now suffices to show that vanishes on such compatible sequences . Scaling if necessary, we may assume that has coefficients in . But then modulo holds for all , so the result follows. ∎
Applying 3.3, the proof of 3.1(ii) is reduced to showing that the conjugation isomorphism sends the image of inside to the image of inside . Applying the isomorphism , it suffices to show that there exists an isomorphism sending to and a commuting diagram of the following form:
(1) |
where the horizontal arrows are the natural representations and the vertical arrow on the right comes from the natural isomorphism of étale stalks.
Conjugation Isomorphisms:
To describe the map , we begin with a more general construction. We let be a complex variety, and . Then defines a categorical equivalence . Denote by the fibre functor at . Given , we may define an automorphism of as follows: for each cover of , choose an isomorphic cover in the essential image of ; this is the conjugate of a cover of . Then define . One checks that extends uniquely to a well-defined automorphism of , and that the map defined by is a group homomorphism. We define the map to be the case with and , and the map to be the case and .
Completing the Proof:
Let and be the natural maps. It now suffices check that and that diagram (1) commutes. In the first case, one immediately checks that if acts on fibres of above by base changing to a cover of and acting via , then the same is true for with respect to and . In the second case, it is immediate from the explicit description of the isomorphism , that together with the map , the map gives a map of representations (i.e., acting by then applying is the same as applying and then acting by ). But 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 established in the introduction.
Definition 4.1.
Let be a point, a collection of subvarieties of containing , an automorphism, and a collection of tensors with canonical rational conjugates. Then we define
by the property that if and only if for all , and sends each to its canonical rational conjugate .
We observe that the normalizer in the statement of 1.3 is equal to , so we may view as its natural generalization. The following generalization of 1.3 is then immediate from 3.1:
Theorem 4.2.
Let be a motivic variation of Hodge structure on , and let . Suppose that is a subspace with canonical rational conjugates, that for is a collection of complex subvarieties containing , and that is a subspace containing such that either
-
(i)
fixes as a -subspace; or
-
(ii)
fixes pointwise.
Then
-
(i)
each element of has rational conjugates; or
-
(ii)
each element of 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 and for all are points of . It follows as in 1.3 that the points of are defined over and (under an appropriate isomorphism ) over for every . By 1.6 we conclude in that is a -subspace, where . In Case (ii) we additionally know that is independent of the choice of , making the conjugates canonical. ∎
Finally let us give a similar, but simpler argument which establishes cases of a kind of “-absolute Hodge” conjecture. We note that a “-absolute Hodge” conjecture may in fact be enough for certain applications of absolute Hodge to the algebraicity of periods. The following requires no -adic input.
Theorem 4.3.
Let be a variation of Hodge structure on , suppose that , and are defined over , and choose . Let be a -subspace with canonical rational conjugates, let for be a collection of complex subvarieties containing , and suppose that is a subspace containing such that the orbit of under inside the appropriate Grassmanian is finite. Then has -rational conjugates.
Proof.
The proof follows immediately, as 3.1 ensures that (and , if an appropriate comparison to an -adic local system exists) lies inside and the assumptions ensure that is defined over for any point of . ∎
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 be a motivic variation on the smooth -variety , and be the -span of the polarization ; by 2.1 the subspace has canonical rational conjugates. Suppose that there exists a subvariety containing such that the fixed locus of inside is the line spanned by the Hodge class . Then if is a complex point of the normalizer of inside , we have that for some , and from the fact that we learn that . 1.7 therefore applies, and the tensor 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 (the case); the variety is taken to be a special subvariety (irreducible component of the Hodge locus); and the condition that the fixed locus of consist of exactly the line spanned by takes the form of the condition that the restricted variation has as its constant subvariation. Voisin also only obtains the weaker conclusion that the conjugate is a -scalar multiple of a Hodge class; the essential difference, it seems to us, is the presence of -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 be a polarizable variation of Hodge structure of weight . Denote by the -linear map induced by the connection, let be a Hodge class with Hodge locus , and let be its projection to . Suppose that
-
(i)
the map is surjective;
-
(ii)
for the restriction of to the the tangent space is injective;
-
(iii)
and the restriction of to the tangent space has kernel equal to the span of .
Then the fixed locus in of is exactly the line spanned by .
Proof of 1.8:.
We observe that the variation is motivic: taking anti-invariants defines an appropriate subbundle and subsystem , 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 whose projections lie in a certain topologically dense subset of the underlying real subbundle of . 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 and be algebraic varieties, and let and be variations of Hodge structure on and , respectively. We define to be the variation on , where is the projection.
Suppose that is an integral variation of Hodge structure on , and let be the algebraic subvariety whose points are pairs such that is isomorphic to as an integral polarized Hodge structure. These isomorphisms are Hodge tensors of the variation . When a Torelli theorem is avaliable, the locus is simply equal to the diagonal , and its algebraic monodromy is easily determined to be equal to the image of under the diagonal action. In the general case, let be an irreducible component of containing the diagonal . We determine 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 be a proper, dominant morphism of irreducible complex algebraic varieties, and let be a local system on . Then for any point , the isomorphism induces an isomorphism between and .
Proof.
That is proper means that it has a Stein factorization , where is proper with connected fibres and is finite. We are thus reduced to showing the theorem in two cases:
-
(i)
when additionally has connected fibres;
-
(ii)
and when is additionally finite.
Let us first assume that both and are normal. If is a normal variety, then the fundamental group of any open subvariety surjects onto the fundamental group of , so we may replace both and by Zariski open sets, restricting appropriately. In case (i) this lets us assume that 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 is étale, from which the result follows as the monodromy of will be finite index in the monodromy of , 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 on agrees with the algebraic monodromy of the restriction of to the normal locus of . The result then follows by restricting . ∎
Lemma 5.4.
Let be a polarizable integral variation of Hodge structure on the smooth complex algebraic variety , and let and be as above. Then .
Proof.
The statement is unchanged under replacing with a finite étale cover, so we may assume that has unipotent monodromy at infinity. Let be the period map on ; here is the full period domain of polarized integral Hodge structures on a fixed lattice , and where . Then may be interpreted as the moduli space for polarized integral Hodge structures of the same type as the fibres of , and the map sends to the isomorphism class of . Arguing as in [CPMS03, Corollary 13.7.6] we may complete the variety by adding finitely many points so that is proper. Applying the main theorem of [BBT18], we find that there exists a factorization where is a dominant proper map of algebraic varieties and is a closed embedding.
The variety admits a variation such that may be identified with , and is the period map of . The variety is then the inverse image of the diagonal under the product morphism . If and are the respective diagonals, the maps and 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 has canonical rational conjugates, then so does every rational tensor fixed by . 1.9 now follows from the more general:
Theorem 5.5.
If is a motivic variation of Hodge structure on the smooth -variety such that rational tensors fixed by have canonical rational conjugates and the points of the center are defined over , then tensors fixed by have rational conjugates, and generic isomorphisms between and are -absolutely Hodge.
Proof.
Let be the subspace of -tensors fixed by , and let be its dual. As the subspace (resp. the subspace ) is fixed under algebraic monodromy, parallel transport gives a path-independent translate of (resp. ) to any fibre of (resp. ). It follows that we may unambiguously refer to the subspace inside for any pair . The fact that has canonical rational conjugates implies the same fact for . We then have:
Claim 5.6.
The groups and are equal as subgroups of .
Proof.
Let be the subspace of tensors fixed by . It suffices to show that . Let us first argue that the space is spanned by pure tensors; i.e., it has a basis of the form where and . Indeed, suppose that fixes , where and form linearly independent sets. Let be a point. Then from the fact that fixes we learn that fixes each , and analogously we learn that fixes each . It follows that fixes each term in the sum.
We are reduced to arguing that any element of the form with and fixed by must satisfy and . From the fact that acts on and separately we learn immediately that (resp. ) acts on (resp. ) through a character. As (resp. ) is a connected, semisimple group (isomorphic to ), we learn that (resp. ) fixes (resp. ). ∎
We may now complete the proof by computing the normalizer of inside . It follows from 5.4 that is the diagonal subgroup of . The normalizer of is the group generated by and the product of centers . It follows that preserves the subspace of all rational tensors fixed by , 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.