Effective Methods for Diophantine Finiteness
Abstract
Let be a number field, and let be its ring of -integers. Recently, Lawrence and Venkatesh proposed a general strategy for proving the Shafarevich conjecture for the fibres of a smooth projective family defined over . To carry out their strategy, one needs to be able to decide whether the algebraic monodromy group of any positive-dimensional geometrically irreducible subvariety is “large enough”, in the sense that a certain orbit of in a variety of Hodge flags has dimension bounded from below by a certain quantity. In this article we give an effective method for deciding this question. Combined with the effective methods of Lawrence-Venkatesh for understanding semisimplifications of global Galois representations using -adic Hodge theory, this gives a fully effective strategy for solving Shafarevich-type problems for arbitrary families .
1 Introduction
Let be a smooth projective family defined over , with a number field, its ring of integers, and any non-zero element of . In a recent paper [LV20], Lawrence and Venkatesh proposed a strategy for bounding the dimension of the Zariski closure of . Since for any point the fibre has good reduction away from , the problem of bounding the dimension of can be interpreted as Shafarevich-type problem for the family . In particular, the Shaferevich conjectures for curves, abelian varieties, and K3 surfaces are equivalent to showing that for particular choices of .
The Lawrence-Venkatesh strategy has been implemented to reprove various classical diophantine finiteness results, including the Mordell conjecture [LV20], the finiteness of -units [LV20], and Seigel’s theorem [Noo21]. However it is of particular interest to apply the strategy in situations where . In this situation the problem becomes much harder, because one needs some understanding of the monodromy of the family over essentially arbitrary geometrically irreducible subvarieties . To explain what we mean, let and define for each such :
Definition 1.1.
The algebraic monodromy group of is the identity component of Zariski closure of the monodromy representation associated to , where is the normalization.
The variation determines a flag variety , on which the group can be said to act after identifying with the variety of Hodge flags on some fibre for . Let be a period map determined by with the complex submanifold of polarized Hodge flags. Then the key quantity relevant for the Lawrence-Venkatesh method is
(1) |
where the minimum is taken over all positive-dimensional geometrically irreducible subvarieties , points , and where is the Hodge flag on .111We not here that is analytically constructible, for instance by applying the main result of [BBT18], so its dimension makes sense. In particular, the Lawrence-Venkatesh method produces an integer , and shows that if , then the Shafarevich conjecture holds for .
For successful applications of the Lawrence-Venkatesh strategy for the Shafarevich problem in situations when we know of only the paper [LS20] of Lawrence and Sawin, who are able to apply this strategy beyond the first induction step to prove a Shafarevich conjecture for hypersurfaces lying inside a fixed abelian variety . Their methods require the auxilliary use of a Tannakian category associated to , and it seems unclear what to do without this abelian variety structure present.
Our main result is as follows:
Theorem 1.2.
Consider a smooth projective family defined over and with smooth and quasi-projective, and given an integer , define
(2) |
where the minimum ranges over all geometrically-irreducible subvarieties with and points . Then there exists an effective procedure which outputs an infinite sequence of integers
such that for some we have .
If moreover the period map is quasi-finite, one can determine .
Remark.
Let us make absolutely clear what is meant by the term “effective procedure” in 1.2. We mean that there exists an infinite-time algorithm (for instance, a non-halting Turing machine), which outputs a sequence of integers , with the integer being outputted at some finite time depending on . Moreover, one also has at time a proof that . Therefore, after time , one can stop the algorithm and use the bound as input to the Lawrence-Venkatesh method. One is also guaranteed that there is some so that at time the bound is best possible, however one does not necessarily have a method to determine unless is quasi-finite. Finally, the algorithm can be described entirely in terms of algebro-geometric computations involving algebraically constructible sets, and implicit in the proof is a description of how to implement it.
There is no fundamental obstruction which requires us to restrict to quasi-finite for the second part of 1.2. Rather, the second part of 1.2 references some delicate arguments in [Urb21b] which are only enough to handle the quasi-finite case directly, and recalling enough of the machinery of [Urb21b] to carry out the argument for the general case would lead us too far astray from the main ideas. We note that one does not actually need to determine the integer referenced in 1.2 to apply the machinery of Lawrence and Venkatesh: one wants to be able to compute the best possible lower bound for , but one is not required to actually prove that the bound one has is optimal in order to deduce diophantine finiteness results.
1.1 The Approach of Lawrence and Venkatesh
We begin with a preliminary observation. To show that , it suffices to show, for any irreducible subscheme of dimension and defined over , that the Zariski closure is a proper algebraic subscheme of . We therefore fix such a subscheme with , and seek to show that .
Fix a prime not dividing , and let be a point. It is conjectured that for any the representation of on is semisimple. If this result were to be known for all such , an argument of Faltings shows that for there are at most finitely many possibilities for the isomorphism class of the representation of on . To establish the non Zariski-density of it would then suffice to show that the fibres of the map
are not Zariski dense. As explained by Lawrence and Venkatesh in [LV20], this is essentially the original argument for the Mordell conjecture due to Faltings.
The problem with applying this strategy in general is twofold. First, the semisimplicity of is not known, and for most choices of appears out of reach. Secondly, without a good geometric interpretation of the étale cohomology it is difficult to understand . The key insight in the paper of Lawrence and Venkatesh is that one may potentially overcome both problems by passing to a -adic setting where they are more managable.
Instead of considering the global Galois representation , we consider its semisimplification , and restrict along the map induced by a fixed embedding to obtain , where is a fixed place above . The functors of -adic Hodge theory tell us that the representation determines a triple , where is the Hodge filtration on de Rham cohomology and is the crystalline Frobenius at . If we somehow manage to consider the data up to “semisimplification” (in the sense that we identify such triples when the associated global Galois representations have isomorphic semisimplifications), our problem is then to study the fibres of the map
and show they lie in a Zariski closed subscheme of smaller dimension.
Next, we make the elementary observation that to bound the dimension of the Zariski closure of , it suffices to cover by finitely many -adic disks and bound the dimension of the Zariski closure of for each ; here is the ring of -adic integers. It can then be shown that there exists such a cover for which the Hodge bundle can be trivialized rigid-analytically over each , moreover with respect to each such trivialilzation the Frobenius operator is independent of . The problem then reduces to studying a varying filtration on a fixed vector space for some . In particular, we obtain a rigid-analytic map
and those points of arising from points lie inside finitely many subvarieties of corresponding to the finitely many possible isomorphism classes of .
We are now faced with the problem of understanding the intersections , and showing that their inverse images under lie in an algebraic subscheme of smaller dimension. One part of this problem is to understand the dimensions of the varieties , and to show that they are sufficiently small: this step is carried out successfully for the families of hypersurfaces studied both by Lawrence-Venkatesh and Lawrence-Sawin, and appears to be managable in general. The more difficult object to control is , for which we need to understand the variation of the filtration over . This, in turn, is governed by the Gauss-Manin connection , which exists universally over after possibly increasing ; we may adjust so that it does not divide if necessary. The fact that exists universally over means that the same system of differential equations satisfied by at is also satisfied by a Hodge-theoretic period map on a sufficiently small analytic neighbourhood containing , where is a variety of Hodge flags. In particular, one can prove that the Zariski closures of and have the same dimension.
The final step, which is to show that the Zariski closure of in has smaller dimension, is completed as follows. The Ax-Schanuel Theorem [BT17] for variations of Hodge structures due to Bakker-Tsimerman shows that if is an algebraic subvariety of of dimension at most ,222The dimension can once again, at least for open neighbourhoods with a sufficiently mild geometry, be made sense of the dimension of a locally constructible analytic set. Alternatively one can replace with , where is as before and is a component containing . then the inverse image under of the intersection lies in an algebraic subvariety of of smaller dimension. Choosing an isomorphism one can transfer this fact to the same statement for the map and in particular for . Our problem is finally reduced to giving a lower bound for the difference . Our main result now reads:
Theorem 1.3.
Define the quantity
where ranges over all irreducible complex algebraic subvarieties of dimension greater than , and where is any complex analytic period map determined by the variation of Hodge structures and defined on a neighbourhood . Then there exists an effective procedure which outputs an infinite sequence of lower bounds
such that for some we have .
If moreover the period map is quasi-finite, one can determine .
1.2 Basic Idea of the Method
We may observe that the computation of the bound described in 1.3 is a purely Hodge-theoretic problem, i.e., it concerns only properties of the integral variation of Hodge structures on . Let be a polarization of , and let be a fixed polarized lattice isomorphic to one (hence any) fibre of ; as it causes no harm, we will assume that for some , and therefore sometimes write for . Let be the complex manifold parametrizing polarized Hodge structures on with the same Hodge numbers as . A point we may view as a morphism , where is the Deligne torus, and the Mumford-Tate group is the -Zariski closure of .
To present our method, we introduce some terminology:
Notation.
We denote by the -algebraic variety of all Hodge flags on the lattice , not necessarily polarized. We note that is an open submanifold of a closed -algebraic subvariety .
Definition 1.4.
Given two subvarieties , we say that if there exists such that . Given a variety , we call the equivalence class under a type. The dimension of a type is the dimension of .
Definition 1.5.
We say that a type is Hodge-theoretic if , where for and a -algebraic normal subgroup of .
The first step in our algorithm is:
Step One: Compute a finite list of types such that every Hodge-theoretic type appears somewhere in the list.
When we say to compute a type , we mean to compute a representative such that . That there are only finitely many Hodge-theoretic types is shown in 2.1 below.
The problem given in Step One is solved in [Urb21b, Prop. 5.4]; we will say little about it here. It is related to the problem of classifying Mumford-Tate groups up to conjugacy by , for which one can use a constructive version of the proof in [Voi12, Thm. 4.14]. It is also similar to the problem of classifying Mumford-Tate domains as studied in [GGK12, Chap. VII]. We note that the methods of [GGK12, Chap. VII], when they can be carried out effectively, result in an approach for which will be exactly the set of Hodge-theoretic types.
The second step is more involved, and is the crux of our method. To describe it we need to introduce some terminology.
Definition 1.6.
A local period map is a map obtained as a composition , where:
-
(i)
The set is a connected analytic neighbourhood on which is constant and is trivial for each , where .
-
(ii)
The map is a varying filtration-compatible period matrix over . More precisely, there exists a basis for , compatible with the filtration in the sense that is spanned by for some , and a flat frame for , such that is the change-of-basis matrix from to .
-
(iii)
The map sends a matrix to the Hodge flag defined by the property that is spanned by the first columns.
To summarize the preceding definition: a local period map is exactly a period map on except one does not necessarily compute periods with respect to the integral lattice but is instead allowed to consider periods with respect to a more general complex flat frame. There is a natural -action on the set of germs of local period maps at a point , where acts on the map to give . This action corresponds exactly to a change of the flat frame , and all germs of local period maps at lie in a single -orbit.
The construction of a local period map involves picking a basis of , and hence choosing an isomorphism . When working with a local period map, we will always assume that such a basis has been choosen, and hence identify subgroups of with subgroups of . In particular, if is a geometrically irreducible subvariety which intersects , we have an induced action of on .
Lastly, we need:
Definition 1.7.
Given two types and , we say that if there exists for such that and .
Definition 1.8.
Given a local period map and a geometrically irreducible subvariety intersecting at , we call the type of , and denote it by .
For well-definedness, see 2.2 below. From Step One, we have computed a finite list of types containing all types that can arise from the variation . Our next task is then:
Step Two: For each type appearing in the list, compute a differential system on characterized by the property that an algebraic subvariety is an integral subvariety for if and only if , and determine the dimension of a maximal integral subvariety for this system.
We explain precisely what we mean by “differential system” in section 3; actually our method does something more subtle than Step Two due to the fact that we can only approximate up to some finite order, but for expository purposes this is the essential point. After this, we will see the problem is reduced to analyzing which of the differential systems admit algebraic solutions of “exceptional” dimension, which can be carried out using tools from functional transcendence.
1.3 Acknowledgements
The author thanks Brian Lawrence, Akshay Venkatesh, and Will Sawin for comments on a draft of this manuscript.
2 Algebraic Monodromy Orbits up to Conjugacy
In this section we describe an effective method for solving “Step One” as posed in subsection 1.2. We will also prove some preliminary facts about types used in the introduction, and we continue with the notation established there. We will work in the context of a general polarizable integral variation of Hodge structure on the complex algebraic variety , not necessarily coming from a projective family as in the introduction.
2.1 Basic Properties of Types
Lemma 2.1.
For any geometrically irreducible subvariety and any local period map with non-empty, we have
for any point .
Proof.
It suffices to show that
for each analytic component separately, with a point of . By acting on by an element of , the claim can be reduced to the situation where the periods which determine are computed with respect to a basis for the integral lattice , and then the claim follows from [Urb21b, Lem. 4.10(ii)]. ∎
Lemma 2.2.
The equivalence class under of is independent of ; i.e., the type of is well-defined.
Proof.
Let be a smooth resolution, and consider the variation . From the fact that germs of local period maps on with respect to the variation factor through germs of local period maps on , we may reduce to the same problem for and the variation , i.e., we may assume . By analytically continuing a fixed local period map to the universal covering , we learn from the irreducibility of that at each point , there exists a local period map such that . Since the Zariski closure of is determined by the germ of at , and because all germs of local period maps at lie in a single -orbit, the result follows. ∎
Lemma 2.3.
There are only finitely many Hodge-theoretic types.
Proof.
We observe that the problem reduces to the following: show they are finitely many -equivalence classes of pairs , where
-
(i)
is a polarized Hodge structure; and
-
(ii)
is a -algebraic connected normal subgroup of ;
where we regard as acting on through its action on , and on by conjugation. Note that two such equivalent pairs will generate orbits in equivalent under . Since the groups are reductive and have finitely many connected normal algebraic factors, this reduces to the same problem for pairs of the form . We recall that is an open submanifold of , the flag variety of flags satisfying the first Hodge-Riemann bilinear relation (the isotropy condition), and that is an algebraic subvariety of . We then use the fact that there are finitely many Mumford-Tate groups up to -conjugacy (see [Voi12, Thm. 4.14]), and that for a fixed Mumford-Tate group the Hodge structures in with Mumford-Tate contained in lie inside finitely many -orbits in , see [GGK12, VI.B.9]. ∎
2.2 Computing Types up to Conjugacy
In this section we give some references for carrying out Step One as described in the introduction.
Proposition 2.4.
There exists an algorithm to compute subvarieties such that the set of Hodge-theoretic types is a (possibly proper) subset of .
Proof.
This is solved in [Urb21b, Prop. 5.4]. ∎
Let us comment briefly on a different approach to Step One given in [GGK12, Chap. VII]. In [GGK12, Chap. VII], the authors describe a method for classifying both Mumford-Tate groups and Mumford-Tate domains (orbits of points under and ). Given an appropriate such classification, one can easily solve Step One by computing the decompositions of the groups that arise into -simple factors. The method of [GGK12, Chap. VII] is to first classify CM Hodge structures , and then give a criterion for deciding when a Lie subalgebra of corresponds to a Mumford-Tate group generating a Mumford-Tate domain containing . They carry out this classification procedure successfully when , and so for variations with Hodge numbers and .
The method given for classifying CM Hodge structures given in [GGK12, Chap. VII] is to observe that CM Hodge structures up to isogeny are determined by certain data associated to embeddings of CM fields, and hence the first step of the procedure in [GGK12, Chap. VII] is to “classify all CM fields of rank up to [] by [their] Galois group”. We are not aware of an effective method for carrying out this step.333The paper [Dod84] gives a potential approach by giving a method to classify certain abstract structures associated with Galois groups of CM fields. However one still needs to determine which such structures are actually associated to a concrete CM field. It is also not clear to us precisely the sense in which the term “classify” is being used; i.e., we do not know what form the data of a “classification of CM Hodge structures” takes, and consequently what form the resulting classification of Mumford-Tate domains will have. For this reason, we were unable to apply the methods of [GGK12, Chap. VII] to prove 2.4.
3 Differential Tools and a Jet Criterion
In this section we introduce a collection of effectively computable algebro-geometric correspondences which can be used for studying systems of differential equations on induced by the variation , and then use it to solve the main problem. We have already carried out most of the work in two preceding papers [Urb21a] and [Urb21b], so we will first need to collect some results. In this section we assume that is a -variety for a number field, and that is a polarizable integral variation of Hodge structure on such that the vector bundle , the filtration , and the connection all admit -algebraic models. Moreover, we assume that we may effectively compute a description of these objects in terms of finitely-presented -modules over an affine cover of ; for a justification of this assumption in the situation where comes from a smooth projective -algebraic family see [Urb21b, §2].
3.1 The Constructive Period-Jet Correspondence
Our algebro-geometric correspondences will be formulated using the language of jets. Let , and define to be the -dimensional disk of order ; we suppress the field in the notation. A jet space associated to a space is a space which parametrizes maps . More formally, for a finite-type -scheme, we have:
Definition 3.1.
We define to be the scheme representing the contravariant functor given by
where the natural map obtained by pulling back along .
That the functor defining in 3.1 is representable is handled by [Urb21a, §2]. Moreover, is itself a functor, sending a map to the map that acts on points by post-composition. For an analytic space, there is an analogous construction that appears in [Urb21a, §2.3]. If is a subfield, this construction is compatible with analytification.
Theorem 3.2.
For each , a variation of Hodge structure on gives rise to a canonical map
of algebraic stacks characterized by the property that for any local period map and any jet we have modulo .
Moreover, if the data associated to the variation admits a -algebraic model, the map is defined over , and there exists an algorithm to compute the -torsor and the -invariant map which defines from a presentation of the data in terms of finitely-presented -modules.
We note that the computability of the torsor in 3.2 has in particular the following consequence: if is a subset which is the image under the quotient of a constructible -algebraic set , where is a computable extension, then we can compute by computing . Thus if we define
Definition 3.3.
For a constructible -algebraic set , with an extension, we write
Moreover, for a type , we will write either or for the set .
then the main consequence of the preceding discussion for our situation is the following, which is immediate from what we have said:
Proposition 3.4.
For each there exists an algorithm which, given a constructible -algebraic set with a computable extension, computes . ∎
3.2 Jet Conditions and Types
Let us now try to understand how computing the “differential constraints” induced by types as in 3.4 can help us carry out Step Two of subsection 1.2. Let , and let be the canonical period map which sends a point to the isomorphism class of the polarized Hodge structure on . By [BBT18], the map factors as , where is a dominant map of algebraic varieties and is a closed embedding of analytic spaces; it follows that for each subvariety the dimension of the image makes sense as the dimension of a constructible algebraic set.
Fix a sequence of compatible embeddings
of formal disks. By acting on points via pullback, we obtain natural transformations of functors which produce maps that take jets to their restrictions . We are now ready to present the key proposition for our method:
Definition 3.5.
For a scheme (resp. analytic space ) denote by the subscheme (resp. the analytic subspace) parametrizing those maps which are injective on the level of tangent spaces. We call such non-degenerate jets.
Proposition 3.6.
Let be a set of types containing all the Hodge-theoretic types, and let and be non-negative integers. Then the following are equivalent:
-
(i)
there exists a geometrically irreducible subvariety with , , and such that ;
-
(ii)
there exists with , and such that the intersection
is non-empty for each .
Remark.
In the situation that the variation admits a local Torelli theorem, one can drop the distinction between and and consider instead the intersections in part (ii), ignoring the middle term.
The rest of this section we devote to proving 3.6, identifying with for ease of notation. To begin with, let us check that (i) implies (ii) by applying the definitions. If is an étale cover and we consider the variation , then the maps and obtained from 3.2 are related by . Choosing a finite index subgroup and passing to such a cover, we can reduce to the case where we have a period map with a local isomorphism. Applying [BBT18] the map factors as , where is a dominant map of algebraic varieties and is an analytic closed embedding. Then via , the variety is dominant over a closed subvariety of dimension . Shrinking (and hence ) we may assume that is smooth, and that is surjective onto a dense open subset . Shrinking even further we may assume that is smooth. The smoothness of implies in particular that the induced jet space maps for all choices of and are surjective.
We may choose neighbourhoods and such that is an isomorphism, both and are non-empty, and we have a local lift of . Choose a jet and lift it along to a jet landing at the point . Using the fact that the germ is smooth of dimension the jet can be extended to a jet such that , and hence . From the fact that and the defining property of the map it follows that lies inside . We can then take , and the fact that factors through implies that as well.
To prove the reverse implication, we review some preliminary facts relating to jets.
Definition 3.7.
We say a sequence with is compatible if the projections map to .
Lemma 3.8.
Suppose that is a collection of non-empty constructible algebraic sets such that the projections map into . Then there exists a compatible sequence with for all .
Proof.
See [Urb21a, Lem. 5.3]. ∎
Definition 3.9.
Given a variety (algebraic or analytic) and a point, we denote by the fibre above of the natural projection map .
Lemma 3.10.
If is a map of analytic germs with and smooth, we have an infinite compatible family , and for some germ and all , then factors through the inclusion .
Proof.
See [Urb21b, Lem. 4.5]. ∎
Lemma 3.11.
Suppose that is an algebraic variety (resp. analytic space) and is a point for which the fibre above is non-empty for all . Then the germ has dimension at least .
Proof.
See [Urb21a, Prop. 2.7]. ∎
Proof of 3.6:.
By what we have said, we are reduced to showing that (ii) implies (i). The statement is unchanged by replacing with a finite étale covering and the variation with ; as before this does not affect the hypothesis (ii) since the maps and associated to and are related by . Choosing a finite index subgroup and choosing so the monodromy of lies in we may reduce to the case where is a local isomorphism. Moreover, taking a futher finite étale cover we may apply [CPMS03, Cor. 13.7.6] to reduce to the case where is proper; this requires possibly extending to a variety by adding a closed subvariety at infinity, but as long as we are careful to work only with jets that factor through our proof will produce a variety intersecting ; in particular, we now assume that is proper but redefine the sets to equal
for some open subvariety .
Applying the main result of [BBT18], the map once again factors as with a dominant (now proper) map of algebraic varieties. We can then consider the Stein factorization of ; note that is proper with connected fibres, is normal, and is finite. One can define the type of a subvariety exactly as in 1.8 with respect to the period map . From 3.8 above, the assumption (ii) entitles us to a compatible sequence of jets such that for all . Let us write for some subvariety .
By construction, the jets are non-degenerate, and remain so after composing with any local period map for which factors through . This in particular implies (since is a local isomorphism) that the jets are non-degenerate, and hence so are the jets . Let be the smallest algebraic subvariety such that for all . We observe that there exists a component of of dimension at least that contains the image of the jets : one can see this by picking a neighbourhood of of the form such that is constant on the first factor, and applying 3.10 above to see that the restriction of to factors through . Moreover, we must have by minimality, and by applying 3.11 to the non-degenerate sequence that . Since is finite, this means . From the fact that local period maps on factor through local period maps on we learn that , so the result will follow if we can show that . For ease of notation let us now write .
Fix a local lift of the period map with an analytic ball such that the jets factor through . Consider the set consisting of those for which . Then for each the set is algebraically-constructible, and using the fact that the set is necessarily non-empty. Let be an element of this intersection. Extend to a lift of to the universal covering. Then is a closed analytic set containing the jets , and hence the non-degenerate jets . Letting be the minimal analytic germ through which (and hence ) factors, it follows from 3.11 that has dimension at least .
Consider the Zariski closure of . We claim that . Because was chosen minimal containing the compatible family of jets , it suffices to show that each factors through . Consider the component of containing ; by choosing coordinates we may assume is an open neighbourhood and identify with the origin. After a further change of coordinates we may assume is constant on the first factor, and let . Applying 3.10 we find that , and hence . Using proper base change the map is proper, so is an analytic subvariety of , and by construction the jets factor through it, hence through .
We are now ready to apply the Bakker-Tsimerman transcendence theorem; the jets are no longer needed. It follows from the structure theorem for period mappings [GGK12, III.A], the closed analytic set lies inside an orbit of the algebraic monodromy of . Consider the graph of in . Then as has dimension and is Zariski dense, there exists a component of of dimension at least and projecting to a Zariski dense subset of . Applying the main theorem of [BT17] we learn that
∎
4 Main Results
4.1 Computing Bounds on
4.1.1 Computing Lower Bounds
Let us explain the significance of 3.6 in proving 1.2, i.e., giving an effective method to compute bounds for
where we have used 1.3 and 1.8 to give this equivalent expression for . Since is a integer bounded by , giving an effective method to compute it amounts to developing a procedure to decide, for any integer , whether we have . This in turn amounts to deciding, for any integer , whether (ii) holds in 3.6.
Let us take to be the set up types computed by Step One, and let us suppose that in fact . Then by the equivalence in 3.6, we should find that for any with , there must be some such that is empty. Moreover, verifying that such an exists for each such and proves, again by the same equivalence, that . Consequently, we obtain the following result, which is the first half of 1.2:
Proposition 4.1.
By computing the sets described in 3.6 in parallel, we may compute a non-decreasing sequence of lower bounds
such that for some we have .
Proof.
At the ’th stage we compute all the sets for all applicable choices of , and , and set to be the smallest for which all the sets are empty. From the discussion preceding the Proposition, the result follows. ∎
4.1.2 Computing Upper Bounds
4.1 does not actually give an algorithm for computing , since no way is given to decide when . For applications to the Lawrence-Venktesh method this doesn’t matter: one wants to be able to compute an optimal lower bound for , but one does not actually have to prove that this lower bound actually equals in order to apply the diophantine finiteness machinery. Nevertheless, let us explain how one can do this in the case where is quasi-finite; under this assumption, we may drop the distinction between and , and we are instead interested in computing
What is needed is the following:
Proposition 4.2.
Suppose that is quasi-projective and is quasi-finite. Then there exists a procedure that outputs an infinite sequence of upper bounds
such that for some we have .
4.2 Finding Varieties that Exhibit
In this section we prove 4.2, assuming throughout that is quasi-projective and is quasi-finite. Let us fix a projective compactification of and consider the Hilbert scheme . There exist algorithms, for instance by working with the Plüker coordinates of the appropriate Grassmannian, for computing any finite subset of components of . By [Urb21b, Lem. 5.10] we obtain the same fact for the open locus consisting of just those points for which is a non-empty geometrically irreducible algebraic subvariety of . What we will show is that there exists a procedure which outputs an infinite sequence of constructible algebraic loci , with the following two properties:
-
(i)
for each , the type and dimension are constant over all ;
-
(ii)
there exists some such that
for some (hence any) point .
Given such an algorithm the problem of computing the bound that appears in 4.2 reduces to choosing a point , computing , and setting
(We note that the problem of computing from and the restriction of the algebraic data on is solved for us by [Urb21b, Lem. 5.8] by taking the family in the statement of [Urb21b, Lem. 5.8] to be a trivial family; we will say little about this problem here.)
In fact, an algorithm for computing the sets has already been given in a previous paper by the author. We begin by recalling the necessary background. We regard as a complex algebraic variety in what follows. Given two (geometrically) irreducible subvarieties with , the algebraic monodromy group may be naturally regarded as a subgroup of (after choosing a base point ). Using this we define:
Definition 4.3.
An irreducible complex subvariety is said to be weakly special if it is maximal among such subvarieties for its algebraic monodromy group.
The key fact is then the following:
Lemma 4.4.
For each integer , there exists a weakly special subvariety such that
Proof.
By [Urb21b, Prop. 4.18], the condition that be weakly special is equivalent to being a maximal irreducible complex subvariety of of type . Thus if we have any which is not weakly special, there exists a weakly special properly containing with , hence
It follows that the value of can only be achieved by a weakly special variety. ∎
Proof of 4.2.
By the degree of a subvariety we will mean the degree of its closure inside . For any integer , denote by the finite-type subscheme parametrizing varieties of degree at most . Denote by the locus of weakly special subvarieties. Then given an integer , the algorithm that appears in [Urb21b, Thm. 5.15] computes the intersection as a constructible algebraic locus.
Let us describe the algorithm appearing in [Urb21b, Thm. 5.15] more precisely. Consider the types computed by Step One, and define for each such type the locus
It is shown in [Urb21b, Prop. 4.31] that for each the locus is closed algebraic. We can then consider the sublocus consisting of just those components for which a generic point satisfies .
In [Urb21b, Prop. 5.14], an algorithm is given for computing for each . Using this, one can compute all the finitely many closed algebraic loci which arise as a component of for some . The problem of computing is reduced to computing constructible algebraic conditions on each component which define the locus of points that are weakly special of type . This is taken care of by the proof of [Urb21b, Thm. 5.15]. By construction, the points in all have the same type and the same dimension, so we complete the proof by computing these loci for increasing values of . ∎
5 Application to Lawrence-Venkatesh
We now show how the bound of 1.2 can be used to establish diophantine finiteness results. Similar arguments appear in [LV20] and [LS20], but as they are not precisely adapted to our setup, we give our own version. We recall the situation: we have a smooth projective family over the smooth base , with everything defined over .444Note in particular we are assuming now that is smooth over , which we can achieve by increasing if necessary. The relative algebraic de Rham cohomology gives a model for the Hodge bundle , where . By a result of Katz and Oda [KO68], the flat connection associated to the local system by the Riemann-Hilbert correspondence admits a model after possibly increasing . Likewise, we may also assume the Hodge filtration gives a filtration of by vector subbundles.
Fix a prime not dividing , and a place of above . Then for each integral point , we have a Galois representation , and an argument of Faltings [LV20, Lem 2.3] shows that the semisimplifications of the representations belong to a finite set of isomorphism classes. From crystalline cohomology, each , viewed as a point of where is the -adic ring of integers, gives rise to a triple where is the crystalline Frobenius. Moreover, using the functor of -adic Hodge theory [Fon94, Exposé III], the triple is determined up to isomorphism by the restriction along the map determined by a fixed embedding . We denote by all those triples which are of the form , where is a global Galois representation whose semisimplification is isomorphic to the semisimplification of .
Recall that we have fixed the integral lattice , where is the dimension of the cohomology of the fibres of , and a -algebraic flag variety of Hodge flags on . In what follows we write for , and for . Then the key idea of the Lawrence-Venkatesh method is the following:
Proposition 5.1.
Suppose that for each , whenever we have an endomorphism and a flag on such that represents , the Hodge flags on for which lie in an algebraic subvariety satisfying . Then .
To prove 5.1 we will need a rigid-analytic version of the Bakker-Tsimerman transcendence theorem, which we will see can be deduced formally from the complex analytic one. To set things up, let us revisit the term local period map, this time in the rigid analytic setting (c.f. 1.6). We will denote by the completion of the algebraic closure . In what follows we sometimes identify algebraic varieties with their rigid-analytifications when the context is clear.
Definition 5.2.
Let be a local field containing , let be the rigid-analytification of the base-change of , and suppose that is an affinoid subdomain. Then a (rigid-analytic) local period map is a rigid-analytic map obtained as a composition , where:
-
(i)
The rigid analytifications are all trivial on .
-
(ii)
The map is a varying filtration-compatible -adic period matrix over . More precisely, there exists a basis for , compatible with the filtration in the sense that is spanned by for some , and a flat (for ) frame such that gives a varying change-of-basis matrix from to .
-
(iii)
The map is the map that sends a matrix to the Hodge flag defined by the property that is spanned by the first columns.
To prove 5.1 we will need a version of the Bakker-Tsimerman transcendence result for rigid-analytic local period maps, which we prove by formally transferring the same result for complex analytic local period maps. To avoid certain minor pathologies that can occur in the complex analytic case we will restrict to local period maps which are definable in the structure ; for background on definability and definable analytic spaces we refer to [VdDM96] and [BBT18]. We note that this is not a serious restriction: given any local period map and any point there exists a definable restriction of to a neighbourhood of , a fact which is for instance easily deduced from [Urb21b, Prop. 4.27].
Lemma 5.3.
-
(i)
Suppose that is a definable analytic local period map on . Let be an algebraic subvariety satisfying . Then lies in an algebraic subvariety of of dimension at most .
-
(ii)
Suppose that is a rigid-analytic local period map on . Let be an algebraic subvariety satisfying . Then lies in an algebraic subvariety of of dimension at most .
Proof of 5.3(i):
This is an application of the Bakker-Tsimerman transcendence theorem. Let be the Zariski closure of . We assume for contradiction that , and let be a component of maximal dimension. Let be the canonical period map with . The statement is invariant under replacing with a -translate and with , so we may assume that is a local lift of . Arguing as in [CPMS03, Cor. 13.7.6] we may assume that is proper, hence the image is algebraic by [BBT18], and we may consider the Stein factorization of the map .
Let , and note that . By assumption we have , where the type of is taken with respect to the period map . Moreover, this continues to hold if we replace with a smooth resolution . The variation of Hodge structure on descends to , and hence shrinking if necessary we may factor through a definable local lift on . By pulling back along the resolution we obtain a definable local lift of the period map such that is Zariski dense in . We are reduced to the following situation: we have smooth variety with a period map , a local lift such that is Zariski dense, and such that .
We now contradict the Bakker-Tsimerman theorem. In particular, we may extend the local lift to a lift of to the universal cover, and consider the graph of the map , where is the orbit for some . We then have that
Since is definable, has finitely many components, and hence there exists an analytic component of such that is Zariski dense in . Let be its graph under . If there is nothing to show, so we may assume that . Hence we find that , and by the Bakker-Tsimerman theorem [BT17] the component lies in a proper subvariety of , giving a contradiction. ∎
To prove 5.3(ii) we first translate 5.3(i) into a claim about rings of formal power series. In particular let be a local period map with an algebraic subvariety, and choose a point such that . Then induces a map on formal power series rings . The claim of 5.3(i) then says that if is the ideal defining with extension inside , then the ideal generated by contains an ideal which is the extension of an ideal defining the germ of a subvariety of dimension at most .
Proof of 5.3(ii):
The claim is Zariski-local on , so we can in particular assume that the bundles for varying are algebraically trivial over , that is affine, and by smoothness that is free. By definition, the map is associated to the following data: a filtration-compatible frame , where spans , and a flat frame spanning , where flatness means for all . This data satisfies the property that , where is the change-of-basis matrix from the frame to the frame , and is the map sending a matrix to the Hodge flag it represents. We note that changing the frame to another filtration-compatible frame does not change the local period map: such a change has the effect of replacing the map with , where is a varying matrix over whose right-action on preserves the span of the first columns for each , and hence . We threfore lose no generality by assuming the filtration-compatible frame is the restriction to of an algebraic filtration-compatible frame over .
The affinoid neighbourhood is of the form , where is an affinoid -algebra. The inverse image is then a closed affinoid subdomain of , i.e., it corresponds to an ideal such that may be identified with . If is the coordinate ring of , then the map induces a map , and the claim to be shown is that there exists an ideal defining a subvariety of dimension at most such that . The ring is Noetherian, so the ideal admits a primary decomposition. Taking radicals, we obtain finitely many prime ideals containing such that the problem reduces, for each , to finding defining subvarieties of dimension at most such that for each . The analytification map is bijective onto the set of closed points of and induces isomorphisms on completed local rings [see whatever]. It follows that if we choose a maximal ideal containing we obtain a commuting diagram
where the bottom arrow is an isomorphism of completed local rings, and the vertical arrows are injections. In particular, if we denote by the extension of in , it suffices to show that , where is the composition of the left and bottom arrow; here we have used the fact that .
Fix an isomorphism , which we choose to preserve the embeddings and . Using the model for over , the isomorphism allows us to identify with the coordinate ring of , the ideal with a complex point , the ring with the completed local ring . Let be the image of the point corresponding to under , and let be the composition . Applying the isomorphism at the level of formal power series, the rigid-analytic local period map induces a map
whose composition we denote by . In what follows we identify the ideals with their images in ; by construction they are the extensions along of an ideal in associated to the base-change of using . By part (i) of this theorem and our reformulation of it in terms of completed local rings, it suffices to show that is induced by a complex analytic local period map defined on a neighbourhood of .
Recall that we have a decomposition , where is the rigid-analytification of a -algebraic map, and gives a varying change-of-basis matrix between a filtration-compatible frame and a rigid-analytic flat frame. Recall also that we have chosen so that it is the rigid-analytification of a -algebraic filtration-compatible frame over . Using the decomposition and the isomorphism we may factor as , where is the base-change under of the map induced by . From our definition of local period map in 1.6, it suffices to show that is induced by a varying change-of-basis matrix from to a complex-analytic flat frame.
The result will follow from the fact that and satisfy a common set of -algebraic differential equations whose solutions are uniquely determined by the period matrix they assign to a point in . To see this, let us write for -algebraic sections . Suppose then that is a flat frame on some complex analytic or rigid-analytic neighbourhood. We then have that
from which we see that giving a flat frame is equivalent to satisfying the system of differential equations for all . If we choose a trivialization of , we may write the in terms of their coefficients with respect to this trivialization, and the same system of differential equations becomes
(3) |
here the operator is defined using the dual basis to . By differentiating Equation 3 and substituting the lower-order differential equations into the higher-order ones, we obtain, for each sequence with and , a set of -algebraic polynomials in the functions for with coefficients in the coordinate ring of such that
(4) |
Because is smooth, given a point of the functions , where is the value of on , induce a coordinate system in the local and formal power series rings associated to at . In these coordinates, the map is given by , where is a rigid-analytic matrix-valued solution to the differential equations Equation 3. The formal map obtained using the isomorphism then satisfies the same set of differential equations, and in particular its derivatives of all orders at are determined using Equation 4 by the initial condition . If we then construct an analytic solution to the differential system in Equation 3 in a neighbourhood of satisfying , the resulting analytic map induces the map on formal power series rings. It follows that is induced by a local period map, which completes the proof. ∎
Proof of 5.1:
We denote by the ring of integers localized at the prime ideal of corresponding to . We begin by showing that (base changes to of) the points of lie inside finitely many distinguished open affinoids admitting local period maps . This reduces to showing that there are finitely many distinguished open affinoids containing the points in over which admits a rigid-analytic flat frame. We may cover by finitely many open subschemes such that and the bundles for varying are all trivial. Then any point factors through some element of this cover, so we may reduce to the case where and the bundles are all trivial.
Proceeding as in the proof of 5.3, we can choose algebraic functions on such that trivializes . We obtain differential equations as in (4), where the polynomials are functions in the coordinate ring of , and so in particular we may view them after base-changing as elements of , where is the ring of -adic integers. Choose a point . Then as has (by assumption) good reduction modulo , we obtain by [DG67, IV. 18.5.17] a lift . Choosing an initial condition , we may use (4) to construct a map , where the partial derivatives of at are given by evaluating the polynomials at . As the coefficients of the power series defining lie in , the map is defined on a residue disk of radius , where is the absolute value on . Varying over the finitely many elements of , we obtain the desired cover.
Now we wish to show that . Recall that for each , we have a set of points whose associated Galois representations have isomorphic semisimplifications. As there are finitely many possibilities for the semisimplification, it suffices to consider the sets defined by
and show that . In particular, we can consider the Zariski closure of just those elements of whose associated points in lie inside one of the neighbourhoods constructed above on which we have a local period map . The hypothesis of the proposition tells us that the image under of the points in lie in a subvariety satisfying ; here we use the fact that the flat frame on is compatible with the Frobenius endomorphism (c.f. the discussion in [LV20, §3]). Base-changing to and applying 5.3 above, we find that , hence the result.
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.
- [BT17] Benjamin Bakker and Jacob Tsimerman. The Ax-Schanuel conjecture for variations of Hodge structures. Inventiones mathematicae, 217:77–94, 2017.
- [CPMS03] James A Carlson, C. (Chris) Peters, and Stefan Müller-Stach. Period mappings and period domains. Cambridge, U.K. : Cambridge University Press, 2003.
- [DG67] Jean Dieudonné and Alexander Grothendieck. Éléments de géométrie algébrique. Inst. Hautes Études Sci. Publ. Math., 4, 8, 11, 17, 20, 24, 28, 32, 1961–1967.
- [Dod84] Bruce Dodson. The structure of Galois groups of CM-fields. Transactions of the American Mathematical Society, 283(1):1–32, 1984.
- [Fon94] Jean-Marc Fontaine. Le corps des périodes p-adiques. Astérisque, 223:59–111, 1994.
- [GGK12] Mark Green, Phillip Griffiths, and Matt Kerr. Mumford-Tate Groups and Domains: Their Geometry and Arithmetic (AM-183). Princeton University Press, 2012.
- [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.
- [LS20] Brian Lawrence and Will Sawin. The Shafarevich conjecture for hypersurfaces in abelian varieties, 2020.
- [LV20] Brian Lawrence and Akshay Venkatesh. Diophantine problems and p-adic period mappings. Inventiones mathematicae, 221:893–999, 2020.
- [Noo21] Marc Paul Noordman. Siegel’s theorem via the Lawrence-Venkatesh method, 2021.
- [Urb21a] David Urbanik. On the Transcendence of Period Images. arXiv preprint arXiv:2106.09342, 2021.
- [Urb21b] David Urbanik. Sets of Special Subvarieties of Bounded Degree. arXiv preprint arXiv:2109.07663, 2021.
- [VdDM96] Lou Van den Dries and Chris Miller. Geometric categories and o-minimal structures. Duke Mathematical Journal, 84(2):497–540, 1996.
- [Voi12] Claire Voisin. Hodge loci. In Handbook of Moduli Vol. III, pages 520 – 559. 2012.