Likely intersections
Abstract.
We prove a general likely intersections theorem, a counterpart to the Zilber-Pink conjectures, under the assumption that the Ax-Schanuel property and some mild additional conditions are known to hold for a given category of complex quotient spaces definable in some fixed o-minimal expansion of the ordered field of real numbers.
For an instance of our general result, consider the case of subvarieties of Shimura varieties. Let be a Shimura variety. Let realize as a quotient of , a homogeneous space for the action of a real algebraic group , by the action of , an arithmetic subgroup. Let be a special subvariety of realized as for a homogeneous space for an algebraic subgroup of . Let be an irreducible subvariety of not contained in any proper weakly special subvariety of . Assume that the intersection of with is persistently likely meaning that whenever and are maps of Shimura varieties (meaning regular maps of varieties induced by maps of the corresponding Shimura data) with finite, . Then is dense in for the Euclidean topology.
1991 Mathematics Subject Classification:
03C64, 11G18, 14D07, 14G351. Introduction
If is an complex manifold and and are two sufficiently nice maps from complex analytic spaces to , then we say that an intersection between and is unlikely if 111 We will be working with cases in which is automatically a finite Boolean combination of (at least) real analytic subvarieties of . This happens, for example, when is proper or definable in an o-minimal expansion of the real field. In this case, we could take to mean the o-minimal dimension or the maximal dimension of a submanifold of . and that it is likely otherwise. Conjectures of Zilber-Pink type predict that in many cases of interest there are very few unlikely intersections. For example, the Zilber-Pink conjecture for Shimura varieties takes the following form. We let be a Shimura variety and a Hodge generic subvariety, meaning that there is no proper weakly special subvariety (in the sense of Shimura varieties) with . The conclusion of the conjecture in this case is that the union of all unlikely intersections between and special subvarieties of is not Zariski dense in . In this paper, we address the complementary question of describing the likely intersections.
This is not the first time that the likely intersection problem has been addressed. In Section 4 we discuss several specializations of our main theorem and the relation between our results and those which appear in the literature. Three notable instances are the “all or nothing” theorem of Baldi, Klingler, and Ullmo [8] on the density of the typical Hodge locus, the results of Tayou and Tholozan [29] describing the typical Hodge locus of polarized variations of Hodge structure over a smooth complex quasiprojective variety, and the work of Gao [12, 14] (following on the work of André, Corvaja, and Zannier [1]) on the generic rank of the Betti map, from which a sufficient condition for the density of torsion in subvarieties of abelian schemes is derived.
We formulate and prove our main theorem (Theorem 3.4) in terms of definable analytic maps from complex algebraic varieties to definable complex quotient spaces. See Section 2 for details. In brief, a definable complex quotient space is a complex analytic space which may be presented as a double coset space where is the connected component of the real points of an algebraic group, is a suitable subgroup, and is a discrete subgroup together with a choice of a definable (in a fixed o-minimal expansion of the real numbers) fundamental set . Examples of such definable complex quotient spaces include complex tori, Shimura varieties, Hopf manifolds, and (mixed) period spaces. We then define a special subvariety of a definable complex quotient space to be the image of a map of definable complex quotient space (or possibly taken only from some subcategory of definable complex quotient spaces) which is a definable map induced by an algebraic group homomorphism followed by translation. For example, the special subvarieties of complex tori would be translates by torsion points of subtori and the special subvarieties of period spaces would come from period subdomains. We must also consider a more general class of weakly special varieties, which are the fibers of maps of definable complex quotient spaces and the images of these fibers under other maps of definable complex quotient spaces. We identify a condition we call well-parameterization of weakly special subvarieties whereby all of the weakly special varieties in some given definable complex quotient space come from those appearing in countably many families of weakly special subvarieties. In our applications this condition is easy to verify as we restrict to subcategories of definable complex quotient spaces for which there are only countably many morphisms all told.
Consider and a homogeneous space for a subgroup . We will write for the quotient map. It may happen that is a special subvariety of for many choices of , where “many” might mean that the set of such is dense in . For example, if is the Cartesian square of the upper half plane, and is the diagonal, then all modular plane curves may be expressed as as ranges through acting via pairs of rational linear transformations. In such a situation we might expect that if is a quasiprojective complex algebraic variety and is a definable complex analytic map from the analytification of to , then the set of special intersections is dense in provided that the intersections are likely in the sense that . This is not quite right as the intersection may become unlikely after transformation through a special correspondence. We account for this complication with the notion of persistently likely intersections. Our main Theorem 3.4 asserts that if we know that the Ax-Schanuel theorem holds for our given category of definable complex quotient spaces (which satisfies some natural closure properties and the well-parameterization of weakly special subvarieties condition), then, in fact, when the intersection of with is persistently likely, the set of special intersections is dense in .
When working in the setting of variations of Hodge structures, our result improves the “all or nothing” theorem of [8] first by giving a criterion for the typical Hodge locus to be dense (namely persistent likeliness), and secondly by obtaining a euclidean dense set of likely intersections using a specific subcollection of the family of all special varieties. On the other hand, the general setting of definable quotients spaces does not allow for measure theoretic techniques, so we are unable to recover the equidistribution results appearing in, for example, [8] and [29]. But our result does imply some of the applications of these equidistribution results. For example, in the setting of the moduli space of principally polarized abelian varieties of dimension , given two subvarieties and of of complimentary dimensions and so that , our result implies the part of [29, Theorem 1.22] stating that the set of points in isogenous to a point in is euclidean dense in . See also Section 4 for more applications. An advantage of the present work is that our results apply directly to the likely intersections problem to contexts not covered by [29], such as for mixed Shimura varieties, or more generally, for variations of mixed Hodge structures.
Unsurprisingly, Ax-Schanuel theorems play key roles in the existing proofs of the density of special intersections. What may be surprising is that our argument is not an abstraction of the proofs appearing, for example, in [8], [29] or [12]. Instead, our arguments are inspired by the work of Aslanyan and Kirby [3], especially with the proof of their Theorem 3.1. The reader will recognize a resemblance between our notion of persistent likeliness and the -broad and -free conditions of [3], which themselves extend freeness and rotundity conditions from earlier works on existential closedness as a converse to Schanuel-type statements. While the contents of our arguments differ, the structure of many of the results of Daw and Ren in [11] inspired our approach.
Our own interest in the likely intersections problem was motivated by [24, Conjecture 3.13] in the second author’s work with Pila on effective versions of the Zilber-Pink conjecture.
This paper is organized as follows. In Section 2 we define definable quotient spaces and develop some of their basic theory. These definitions and results owe their form to the formalism of Bakker, Brunebarbe, Klingler and Tsimerman [5, 4] used to study arithmetic quotients and more generally mixed period spaces. In that section we express precisely what the Ax-Schanuel condition means and show that under the well-parameterization of weakly special subvarieties hypothesis, it implies a uniform version of itself. Section 3 is devoted to the statement and proof of our main theorem on dense special intersections. In Section 4 we detail several specializations of the main theorem, including to Hodge loci, intersections with modular varieties, and density of torsion in subvarieties of abelian schemes.
Acknowledgments
During the writing of this paper S.E. was partially supported through the NSF grant RTG DMS-1646385 and T.S. was partially supported by NSF grants DMS-1800492, FRG DMS-1760414, and DMS-22010405. The authors have no competing interests, financial or otherwise. S.E. thanks E. Ullmo and G. Baldi for hosting him at IHÉS, for sharing an early version of [8], and for discussing problems around the density of special intersections. T.S. thanks Z. Gao, M. Orr, and U. Zannier for offering detailed accounts of the state of the art on likely intersection problems for modular curves and torsion in abelian schemes. Both authors thank an anonymous referee of an earlier version of this paper for suggesting several improvements.
2. Complex quotient spaces and -special varieties
We express our theorems on likely intersections in terms of classes of definable complex quotient spaces. Our formalism is similar to what appears in [5], though we explicitly include the fundamental domain giving the definable structure as part of our data.
Throughout we work in an appropriate o-minimal expansion of the real field (usually ), and the word definable is meant with respect to this choice of o-minimal structure.
Definition 2.1.
A definable quotient space is given by the data of
-
•
a definable group ,
-
•
a definable compact subgroup of ,
-
•
a discrete subgroup of , and
-
•
a definable open fundamental set for the action of on (that is, and there is a finite subset so that if and for some , then ) for which the closure of is contained in for some finite subset .
We write both for the quotient space regarded as a definable, real analytic space where the definable structure comes from , and for the data giving this space. We denote the corresponding quotient map by . When the data are understood, we suppress them and write for and for the quotient map.
The class of definable quotient spaces forms a category with the following notion of a morphism.
Definition 2.2.
A morphism is given by a definable map of groups and an element for which
-
•
,
-
•
, and
-
•
there is a finite set with , where is the map induced on the quotient space.
The induced map on the double quotient space is a definable real analytic map which we also denote by .
Remark 2.3.
If is a morphism of definable quotient spaces as in Definition 2.2, then it is definable in the sense that its graph on the specified fundamental domains
is definable. Indeed, it is clear that the induced map is definable and for we have that if and only if .
The category has a terminal object and is closed under fiber products.
Proposition 2.4.
There is a terminal object in .
Proof.
Take to be trivial group and then let , which is a singleton. The one point space is a definable quotient space and for any definable quotient space , the unique set theoretic map is induced by the unique map of groups and . ∎
Proposition 2.5.
The category has pullbacks.
Proof.
More concretely, we need to show that given maps and of definable quotient spaces, there are maps of definable quotient spaces and fitting into the following Cartesian square.
Express and as and as where and are definable group homomorphisms, and . Let , , and . For the fiber products defining and we use the maps (and its restriction to ) and (and its restriction to ). In defining , we use the maps and . Regarding as a subgroup of and then as a subset of , we let . ∎
A useful observation is that morphisms of definable quotient spaces always factor as a surjective map followed by a map induced by an inclusion of subgroups.
Proposition 2.6.
Every map of definable quotient spaces fits into a commutative diagram
where is surjective and is induced by an inclusion of subgroups. If we further assume that the group in the presentation of is compact, then has compact fibers.
Proof.
Take for and let be a definable homomorphism and so that is given by .
Define , a definable group. Set , , and where is the induced map. Since is a definable open map, is open and clearly is a fundamental set of the action of on . Set and let be the map . The map is then given by . The fibers of are contained in a finite union of sets definably homeomorphic to the homogeneous space , which is compact if is. ∎
For the problems we consider in this paper we require that our definable quotient spaces come equipped with a complex structure and for the domain to arise as a subset of an algebraic variety.
Definition 2.7.
A definable complex quotient space is a definable quotient space together with the data of a real algebraic group and an algebraic subgroup of the base change of to for which is the connected component of the identity in the real points of a real algebraic group , , and is an open domain in the complex points of the algebraic variety .
A morphism of definable complex quotient spaces is a morphism of definable quotient spaces for which the definable map of groups is given by a map of algebraic groups for which .
The class of definable complex quotient spaces with this notion of morphism forms a category .
We leave it to the reader to check that the proofs of the basic closure properties for the category , such as the existence of a terminal object and closure under fiber products, go through for the category . In practice, the morphisms in we consider satisfy a stronger conclusion than what Proposition 2.6 gives. That is, in practice, a map in is given by a map of algebraic groups (and an element ) for which is a finite index subgroup of . It then follows from the proof of Proposition 2.6 that factors as where is a surjective morphism and is a morphism with finite fibers.
As we have defined definable complex quotient spaces, such a space may have singularities. When we restrict to the case that the group is compact, then the singularities are at worst locally isomorphic to those coming from a quotient by a finite group. In our applications, we will consider only cases where these quotients may be desingularized by passing to a finite cover by another definable complex quotient space.
For some purposes we may wish to restrict to an even smaller category . We always assume about our given category of definable complex quotient spaces that it satisfies some basic closure properties. Let us specify these with the following convention.
Convention 2.8.
The category is a subcategory of satisfying the following conditions.
-
•
The one point space is a terminal object of .
-
•
The category is closed under fiber products.
-
•
Every -morphism factors as where is a surjective -morphism and is an -morphism with finite fibers.
In Section 3 we will impose an additional restriction on .
Definition 2.9.
If is an -morphism, then the image is called an -special subvariety of .
In Definition 2.9 we refer to as a special subvariety. It is, in fact, always a complex analytic subvariety of . Indeed, if we factor as given by Convention 2.8, then is a finite, and hence, proper, map of complex analytic spaces so that by Remmert’s proper mapping theorem its image is a complex analytic subvariety of . Using this observation, we may modify Definition 2.9 to require the morphism witnessing that is an -special subvariety of to be finite.
Definition 2.10.
An -family of weakly special subvarieties of is given by a pair of -morphisms
for which is a finite map over its image and is surjective. For each , the image is a weakly -special subvariety of .
Remark 2.11.
An -special variety is weakly -special as if expresses as an -special subvariety of with finite, then we can take , , the unique map to . If we take to be the category , then the converse that every weakly -special variety is actually -special holds.
In another extreme, every singleton in is a weakly special variety witnessed by and .
For our results it will be important that all -weakly special varieties come from countably many -families of weakly special subvarieties. We isolate it as an hypothesis and verify it in cases of interest.
Definition 2.12.
We say that the weakly -special subvarieties of are well-parameterized if there are countably many -families of weakly special subvarieties of ,
for so that for every weakly -special subvariety of there is some and so that . More generally, we say that the weakly -special subvarieties are well-parameterized if for every the weakly -special subvarieties of are well-parameterized.
Remark 2.13.
When is itself countable, by which we mean that there are only countably many objects in and the set of -morphisms between any two such objects is itself countable, then the weakly -special subvarieties are well-parameterized.
Remark 2.14.
The well-paramaterization condition may fail in some cases. From the theory of Douady spaces, we know that all complex analytic subvarieties of a given compact definable complex quotient space are parameterized by countably many complex analytic families of analytic spaces. However, when parameterizing weakly special varieties, the Douady universal family need not arise as a weakly special family as defined in Definition 2.10. Moreover, there need not be natural parameterizations of the weakly special varieties in the cases that is non-compact. The Hopf manifold construction is be instructive here. For example, consider and two multiplicatively independent complex numbers of modulus less than one, , is the subgroup of generated by the map , and . The Hopf manifold may be realized as a definable complex quotient space and is a compact complex manifold with a noncompact, connected group of automorphisms. One by one, these automorphisms define special subvarieties of , but they cannot be parameterized by a family of weakly special varieties.
Let us indicate now the key functional transcendence condition which may hold in a category of definable complex quotient spaces.
Definition 2.15.
Fix a category of definable complex quotient spaces. We say that , a definable complex analytic map from the analytification of a complex algebraic variety to a weakly -special variety satisfies the Ax-Schanuel condition relative to if whenever is a positive integer, and is a complex analytic map, where
with , then either
or is contained in a proper weakly -special subvariety of .
Under the hypothesis that the weakly -special subvarieties are well-parameterized, the Ax-Schanuel condition implies a uniform version of itself.
Since our statement of this uniform version, expressed as Proposition 2.20 below, is a bit dense, we take this opportunity to explain it with a few words. Basically, what it says is that if we are given an -family of weakly special varieties and a family of algebraic varieties which might witness the failure of the transcendence degree lower bound in the Ax-Schanuel property, then weakly special variety in the alternative provided by the Ax-Schanuel property may be chosen from one of finitely many preassigned -families of weakly special varieties.
Before proving Proposition 2.20 we require two lemmas. The first describes families of weakly special subvarieties algebraically. The second allows us to recast Ax-Schanuel in differential algebraic terms.
Lemma 2.16.
Let be a definable complex analytic map from a complex algebraic variety to a definable complex quotient space in . Let
be an -family of weakly special subvarieties. Then there are algebraically constructible sets and so that the set of fibers is equal to .
Proof.
By the Riemann Existence Theorem [25, Théorème 5.2], there is an algebraic variety , a regular map of algebraic varieties and an analytic map realizing as the fiber product . The fiber equivalence relation
is analytic and definable, and hence algebraic by the definable Chow theorem [23]. The quotient may be realized within the category of constructible sets as a constructible set. Let us write for the quotient map. We may then take
∎
Definition 2.17.
Let be a definable complex analytic map from the complex algebraic variety to the definable complex quotient space in . We say that a subvariety is relatively weakly -special of relative dimension at most if there is a weakly -special of dimension at most for which .
Note that in Definition 2.17, because we allow for the possibility that is not a finite map, it could happen that the dimension of itself is greater than . On the other hand, the intersection of with may even be empty! Thus, the dimension of could be less than .
It follows from Lemma 2.16 that if the weakly -special subvarieties are well-parameterized, then for any definable complex analytic map from a complex algebraic variety to some definable complex quotient space in that we can recognize the pullbacks under of prespecials in the sense that for each number the collection of relatively weakly -special subvarieties of dimension at most comprise a countable collection of algebraic families of subvarieties of .
Definition 2.18.
Let be a definable complex analytic map from the complex algebraic variety to to a definable complex quotient space in . Given any field over which and a countable collection of families of relatively weakly -special subvarieties of including all such relatively weakly -special subvarieties are defined, by an -relatively weakly -special variety of dimension at most we mean an -variety of the form where is an algebraic family of weakly -special subvarieties of dimension at most and .
Using the Seidenberg embedding theorem we may reformulate the Ax-Schanuel property in differential algebraic terms. To be completely honest, the embedding theorem as stated and proven by Seidenberg in [27, 28] is not quite sufficient in that he starts with a finitely generated differential subfield of a differential field of meromorphic functions on the open domain and then shows that for any finitely generated differential field extension of at the cost of shrinking to some open subdomain we may embed into over the embedding of . For our purposes, we will need to start with a possibly countably generated differential field . The necessary extension of embeddings theorem is a consequence of the the Cauchy-Kovalevskaya theorem [30] and appears as Theorem 1 of [22]. Iterating this construction countably many steps, we see that if is a countable differential subfield of the meromorphic functions on some open domain in and is a countably generated differential field extension of , then embeds into the differential field of germs of meromorphic functions at some point over the embedding of .
We recall the generalized Schwartzian and generalized logarithmic derivative constructions from [26]. Consider a definable complex quotient space and fix an integer . From the action of the algebraic group on the quasiprojective algebraic variety and a positive integer , there is a differentially constructible map from to some algebraic variety so that for any differential field extending (where the derivations commute and vanish on we have that for ,
where
is the common constant field of . In particular, if this common constant field is , then we may express the quotient of by the as the image of . When is a definable, complex analytic map from the analytification of a quasiprojective algebraic variety to , then we may define a differentially analytically constructible function by the rule that for any meromorphic (where is an open domain in ), where is any branch of the inverse of . Theorem 3.12 of [26] shows that is actually differentially constructible. (That theorem is stated in the case where and , but the proof goes through in the more general case.)
Lemma 2.19.
Fix a category of definable complex quotient spaces. Suppose that the weakly -special subvarieties are well-parameterized and that is a definable complex analytic map from the analytification of a complex algebraic variety to an -weakly special subvariety of a definable complex quotient space in . Let be a countable subfield of over which and a complete collection of algebraic families of relatively weakly -special varieties are defined. Let and be two positive integers. Then for any differential field with -commuting derivations for which is a subfield of the constants of and is algebraically closed and any -relatively weakly -special of dimension at most , if satisfies , and , then there is a -relatively weakly -special for which .
Proof.
Consider and as in the statement of the lemma. Let be a countable differential subfield of with an algebraically closed field of constants containing and over which and the point are defined. By the embedding theorem, we may realize as a differential field of germs of meromorphic functions. Let be the -weakly special variety of dimension at most for which . By Ax-Schanuel applied to and regarded as meromorphic functions, there is a proper weakly -special subvariety with the image of contained in . The algebraic variety is then relatively -weakly special. Let be the differential closure of and be the differential closure of the differential field generated over by . In , satisfies the condition that it belongs to a -relatively weakly special variety of relative dimension strictly less than where is the constant field. As is an elementary extension of , the same is true in . Since the constant field of the differential closure is the algebraic of the constant field of the initial field, we see that belongs to a -relatively weakly special variety of relative dimension strictly less than . ∎
A uniform version of the Ax-Schanuel condition follows from Lemma 2.16 using the compactness theorem.
Proposition 2.20.
Let be a definable complex analytic map from the analytification of a complex algebraic variety to a definable complex quotient space satisfying the same hypotheses as in Lemma 2.19.
Given an -family
of weakly -special subvarieties of , a positive integer , and a family of subvarieties of , then there are finitely many -families of weakly special subvarieties
for so that for any pair of parameters and and analytic map with , , , and , there is some and for which .
3. Density of special intersections
In this section we state and prove our general theorem that, when persistently likely, intersections with special varieties are dense.
Throughout this section is a category of definable complex quotient spaces satisfying our usual hypotheses from Convention 2.8 and some further requirements. Let us specify with the following convention the properties we require.
Convention 3.1.
The category of definable complex quotient spaces satisfies the following conditions.
-
•
is closed under fiber products.
-
•
The terminal definable complex quotient (one point) space belongs to as do the unique maps for .
-
•
If is an -morphism then there are -morphisms and so that , is surjective, and has finite fibers.
-
•
For every there is some smooth and a finite surjective -morphism .
-
•
The -weakly special varieties are well-parameterized.
-
•
Every definable analytic map from the analytification of an algebraic variety to a definable complex quotient space in considered in this section satisfies the Ax-Schanuel condition relative to .
With the following definition we specify what is meant by intersections being persistently likely.
Definition 3.2.
Let . We suppose that the inclusion is an -morphism where if and , then . Suppose that and . Let be a definable complex analytic map. We say that and have likely intersection if , where here, the dimension is the o-minimal dimension. We say the intersection is persistently likely if whenever and are -morphisms with finite and surjective, then .
Note that this definition does not actually require to be non-empty for the intersection to be likely.
Remark 3.3.
In Definition 3.2, we have consider only one special subvariety of , but we really intend to consider many. With the notation as in Definition 3.2, for any , the set is locally (away from ) a complex analytic subvariety of of dimension equal to that of . Indeed, in several cases of interest for many choices of (where this may mean that the set of suitable is dense in ) the set is actually a special subvariety of . Persistent likelihood of the intersection of with is equivalent to the persistent likelihood of the intersection of with these .
With this definition in place we may now state our main theorem.
Theorem 3.4.
Let and let be a definable complex analytic map from the irreducible quasi-projective complex algebraic variety to . Suppose that is a special subvariety expressible as where is a homogeneous space in . Suppose moreover that the intersection of with is persistently likely. Let be an open subset of the complex points of . Then the set has nontrivial interior. In particular, if the set is Euclidean dense in , then the set of special intersections,
is dense in for the Euclidean topology.
Proof.
We break the proof of Theorem 3.4 into several claims. The claims at the beginning of the proof are really just reductions permitting us to consider a simpler situation. The main steps of the proof begin with Claim 3.4.9 in which we compute the dimension of the incidence correspondence . We then use this computation to show that and have the same o-minimal dimension, so that has nontrivial interior in .
Claim 3.4.1.
We may assume that is an embedding.
Proof of Claim: The equivalence relation is a definable and complex analytic subset of the quasi-projective algebraic variety . Hence, by the definable Chow theorem, is itself algebraic. Let be a nonempty Zariski open subset of , considered as constructible set. Then induces an embedding for which the image of is dense in . Shrinking , we may assume that and then replacing by , we see that if the theorem holds for and , then it also holds for and .
With Claim 3.4.1 in place, from now on we will regard as a locally closed subvariety of . With the next claim we record the simple observations that it suffices to prove the theorem for any given open subset of in place of and that we take to be definable.
Claim 3.4.2.
If Theorem 3.4 holds for some nonempty open in place of , then it holds as stated. Moreover, we may assume that is definable.
Proof of Claim: The set is a subset of . Hence, if the former set has nonempty interior, so does the latter. For the “moreover” clause apply the main body of the claim to the case that is a nonempty open ball.
From now on we will take to be definable and will continue to refer to the open subset of under consideration as even after taking various steps to shrink it.
Another basic reduction we shall employ is that it suffices to prove the theorem for a finite cover of .
Claim 3.4.3.
Proof of Claim: Filling the Cartesian square
we obtain a complex analytic space . Since is finite, is itself the analytification of an algebraic variety. Let be a component of this algebraic variety and then let .
The map comes from a homomorphism of algebraic groups and some element where . This map induces a map . Let be a component of . If we succeed in showing that contains some nonempty open set , then would be a nonempty open subset of , as required.
Let us record a useful consequence of Claim 3.4.3.
Claim 3.4.4.
We may assume that is smooth.
Proof of Claim: By Convention 3.1, we may find a finite and surjective -morphism with smooth. By Claim 3.4.3, if we know the theorem for , then we may deduce it for .
Another useful consequence of Claim 3.4.3 is that we may assume that is dense in with respect to the weakly special topology.
Claim 3.4.5.
We may assume that there is no proper weakly special variety with .
Proof of Claim: Let us prove Theorem 3.4 by induction on the dimension of . If where is a weakly special variety, then we could find a finite -morphism , a surjective -morphism , and a point so that . By Claim 3.4.3, we may assume that and . That is, . By Convention 3.1, the weakly special variety , being the fiber product of with the the one-point space over , is definably isomorphic to a space in . By induction on dimension, Theorem 3.4 already holds for .
By our hypothesis that the intersection between and is persistently likely, it is, in particular, likely. Since we have reduced to the case that is an embedding by Claim 3.4.1, we may express the likeliness of this intersection by an equation
for some nonnegative integer . Notice that we have expressed this equality with dimensions as complex analytic spaces. Later, when we write “” without qualification we mean the o-minimal dimension, for which we would have
By the uniform Ax-Schanuel condition, which holds in by Convention 3.1 and Proposition 2.20, there is a finite list of families of weakly special varieties
for where is a finite -morphism and is a surjective -morphism and if is a natural number with is complex analytic with , , and for some , then for some and we have .
Claim 3.4.6.
We may assume that is surjective for each .
Proof of Claim: By Claim 3.4.5, we have reduced to the case that is not contained in any proper weakly special subvariety of . For any with , we would thus have that is a proper complex analytic subvariety of . Thus, we may shrink so that for such an we have . We will thus never encounter weakly special varieties of the form with . Thus, we may omit these families of weakly special varieties from our list.
We may adjust our family of weakly special varieties to remember only the maps .
Claim 3.4.7.
We may assume that and is the identity map .
Proof of Claim: Work by induction on . In the inductive case of , apply Claim 3.4.3 to replace by . We then need to replace for by . Conclude by induction.
We shrink once again to ensure that all of the fibers of have the same dimension when restricted to .
Claim 3.4.8.
We may shrink to a smaller nonempty open set so that for all there is some number so that for all we have .
Proof of Claim: For each and each natural number , let
Here refers to the o-minimal dimension at .
The definable set is the finite disjoint union of the definable sets
as ranges through . We may cell decompose subjacent to these definable sets. Let be an open cell in this cell decomposition. Then for some sequence we have . Because is an open cell, for each , we have
Apply Claim 3.4.2 to conclude.
Consider now the following incidence correspondence.
Note that is definable.
Claim 3.4.9.
We have
Proof of Claim: Fix the base point corresponding to in the coset space. For , let with . Let with . We will check that is a homogenous space for with fibers isomorphic to . Indeed, if and , we have , demonstrating that . That is, . On the other hand, if , then we can find some so that . That is, , the stabilizer of in . That is, . We compute that for and , we have only if .
Using the fiber dimension theorem, since all fibers over have the same dimension, , we now compute that
Abusing notation somewhat, for we will also write for the fiber . Note that is definably, complex analytically isomorphic to which is a locally closed complex analytic subset of . It follows that the o-minimal dimension of is always even.
For each , let us define
Claim 3.4.10.
For , we have .
Proof of Claim: We have reduced through Claim 3.4.4 to the case that is smooth. Hence, each component of has complex dimension at least .
The set of the statement of the theorem may be expressed as
By Claim 3.4.10, we actually have
With the next claim we show that (again by shrinking ) we may arrange that .
Claim 3.4.11.
Possibly after shrinking , we have for .
Proof of Claim: Suppose that and . Then the complex analytic set has a component of complex dimension . Let be a complex analytic map with and . By our choice of the witnesses to the Ax-Schanuel property for , for some and we have
By our reduction from Claim 3.4.8 and the fiber dimension theorem, we have
Moreover, by the homogeneity of , we also have
and, of course,
By our hypothesis of persistent intersection, we have
for some .
Written in terms of o-minimal dimension this says
Combining this equalities, we compute that
Since, , this means that the expected (complex) dimension of a component of is at most , but is such a component of complex dimension greater than . That is, is an atypical component of the intersection inside . Applying uniform Ax-Schanuel again, we may extend the family of weakly special varieties
for so that each such atypical component will satisfy for some and .
Repeating the reductions of the earlier claims and this extension of the list of weakly special witnesses to Ax-Schanuel times, we reach a contradiction to the hypothesis that is nonempty for some .
Thus, . So we have
Subtracting from both sides, we conclude that . Hence, by cell decomposition, contains an open subset of . ∎
4. Applications
In this section we illustrate Theorem 3.4 by considering various situations in which it applies.
4.1. Arithmetic quotients
Our formalism is derived from that of Bakker, Klingler, and Tsimerman in [5] for the study of arithmetic quotients. They consider definable complex quotient spaces in which the algebraic group is a semisimple -algebraic group, is arithmetic (so commensurable with for some / any choice of an integral model for ), and is compact. They often require to be neat; we will return to that issue in a moment. The definable fundamental domain is not chosen sufficiently carefully in [5], an issue that was then addressed and fixed in [6]. A similar issue is addressed in [21] in that one needs to take to be constructed from a Siegel set associated to a maximal compact subgroup of containing .
If we drop the neatness requirement on , then an arithmetic quotient need not be smooth, but because every arithmetic group has a neat subgroup of finite index, for any arithmetic quotient we may find a smooth arithmetic quotient and a finite surjective map of arithmetic quotients .
The one point space is clearly a terminal object in the category of arithmetic quotients and the pullback construction of Proposition 2.5 specializes to the category of arithmetic quotients. Since there are only countably many arithmetic quotients all told and at most countably many maps of algebraic groups between algebraic groups defined over the rational numbers, it follows that the weakly special varieties are well-parameterized within the category of arithmetic quotients.
The main theorem of [7] is that period mappings associated to polarized variations of integral Hodge structures satisfy the Ax-Schanuel condition with respect to . It is an interesting open question whether every definable analytic map where is an arithmetic definable complex quotient space necessarily satisfies the Ax-Schanuel condition with respect to .
Our last observation in verifying Convention 3.1 and the hypotheses of Theorem 3.4 for arithmetic quotients is that if is a homogeneous space for which is a special variety, then for every , is also special. Thus the set of for which is special is dense in for the Euclidean topology.
Returning to the case where we know to be a period mapping and to be Hodge generic in , that is, not contained in any proper weakly special subvariety, the union of ranging over all proper special subvarieties is called the Hodge locus. In [19], a dichotomy theorem is proven for a modified form of the Hodge locus which they call the Hodge locus of positive period dimension: either this locus is Zariski dense in or it is itself a proper algebraic subvariety of . In a very recent preprint [17], tight conditions for the density of the Hodge loci are established.
Because the special subvarieties of come from -semisimple algebraic subgroups of and there are only finitely many such subgroups up to -conjugacy, all special subvarieties of come from finitely many families of homogeneous spaces in the sense of Theorem 3.4. That is, we can find finitely many homogeneous spaces so that for any special subvariety there is some and with . Thus, if for some special subvariety the intersection of with is persistently likely, then the Hodge locus is Euclidean dense in . In fact, we may take so that for some and we see that the subset of the Hodge locus of the form is Euclidean dense in . In [8] a theorem of a similar flavor is proven. They show that if the typical Hodge locus is nonempty, then it is analytically dense in . Here the typical Hodge locus is the union of all components of of expected dimension as ranges through the special subvarieties of . The proof in [8] uses some elements in common with ours. Notably, Ax-Schanuel plays a central role in both proofs. To pass from a nonempty typical locus to one which is dense, they argue an analysis of Lie algebras to find enough special varieties. Such a technique is not available to us in general as we must postulate the existence of special varieties of a given shape. On the other hand, such an argument does not immediately lend itself to a study of intersections with a restricted class of special varieties.
Definability of the period mappings associated to admissible, graded polarized, variation of mixed Hodge structures has been established by Bakker, Brunebarbe, Klingler, and Tsimerman in [4] and then Ax-Schanuel for these maps was proven independently by Chiu [10] and Gao and Klingler [15]. Indeed, Chiu has established a stronger Ax-Schanuel theorem with derivatives for such period maps associated to variations of mixed Hodge structures [9]. These results give the necessary ingredients to extend our result on the density of Hodge loci to variations of mixed Hodge structures. We will return to the special case of universal abelian schemes over moduli spaces in Section 4.3 to draw a conclusion from the combination of our Theorem 3.4 and Ax-Schanuel in the context of mixed Shimura varieties.
4.2. Modular varieties
For the sake of illustration, let us consider a very special case of Theorem 3.4. We take . That is, is affine -space (for some positive integer ) regarded as the coarse moduli space of -tuples of elliptic curves. We may see as an arithmetic quotient space, taking , , and the homogeneous space may be identified with where is the upper half plane.
If is a finite sequence taking values in we may define by . For we list the elements of in order as and write for . For a singleton , we just write for .
For a special subvariety of provided that for each the projection map is dominant, then defines a partition of by the rule that and lie in a common element of the partition if and only if . Given a partition of we say that is a special variety of type if . Let us observe that a special variety of type has dimension equal to .
Fix a partition with . Let be the homogeneous subspace of defined by if and only there is some with . Then is the corresponding multi-diagonal subvariety of and is a special variety of type . Indeed, the special varieties of the form as ranges through are exactly the special varieties of type .
For a partition of and subset of , we define
to be the restriction of the partition to .
It is easy to check that for any special variety and subset , the projection is a special variety and . With these combinatorial preliminaries in place, we may state the specialization of Theorem 3.4 to the case of .
Proposition 4.1.
Let be a positive integer and a partition of . If is an irreducible complex algebraic subvariety of affine -space, regarded as the coarse moduli space of -tuples of elliptic curves, and for every we have , then
is dense in for the Euclidean topology.
Proof.
Let us check that the intersection between and is persistently likely. Let
be a pair of surjective maps of arithmetic quotients with finite. The arithmetic quotients and will take the form and with where is an arithmetic group in for and and the corresponding homogeneous spaces are and , respectively. Since each factor is simple, the maps of algebraic groups corresponding and are given by coordinate projections followed by an inner automorphism defined over . That is, the map of groups corresponding to is given by followed by an inner automorphism of defined over for some collection of distinct numbers between and , and likewise for . Let , then permuting coordinates we see that this family of weakly special varieties fits into the commuting square
where is finite. We then have
Since is the commensurator of and is dense in , the concluding “in particular” clause of Theorem 3.4 applies and we find that the intersections of with special varieties of type is dense in in the Euclidean topology. ∎
Instances of Proposition 4.1 appear in the literature. Habegger shows in [16, Theorem 1.2] that if is a curve defined over the algebraic numbers, then there is a constants and so that for every prime number there is an algebraic point with logarithmic height where here is the modular curve parametrizing the isomorphism classes of pairs of elliptic curves for which there is an isogeny of degree . Habegger’s result implies in particular that for and , if is an affine plane curve defined over the algebraic numbers, then the intersection of with the special varieties of type is Zariski dense in . Using equidistribution results, this Zariski density could be upgraded to Euclidean density.
In the discussion after Remark 3.4.5 in [31], Zannier sketches an argument showing that if is a rational affine plane curve, then the intersections of with special curves of type , as in the previous paragraph, are dense in in the Euclidean topology.
4.3. Torsion in families of abelian varieties
If is an abelian scheme of relative dimension over the irreducible quasiprojective complex algebraic variety and is a quasi-section of , by which we mean that restricts to a generically finite map on , then under some mild nondegeneracy conditions, we expect that the set
of points on the base over which meets the torsion subgroup of the fiber is dense in if and only if . Masser-Zannier prove in [20] that when and is the square of the Legendre family of elliptic curves defined in affine coordinates by and where ranges over , and is the curve defined by and , then the set is finite. This theorem sparked much work on torsion in families of abelian varieties culminating in a result announced by Gao and Habegger that, at least for such abelian schemes defined over , if is an algebraic variety, also defined over so that the group generated by is Zariski dense in and is Zariski dense in , then .
In the opposite direction, André, Corvaja, and Zannier study in [1] the problem of density of torsion through an analysis of the rank of the Betti map. In an appendix to that paper written by Gao, it is shown that if is a principally polarized abelian scheme of relative dimension which has no non-trivial endomorphism (on any finite covering), and for which the image of in the moduli space of abelian varieties of dimenion itself has dimension at least and is the image of a section of , then is dense in in the Euclidean topology. The proof of this result made use of the Ax-Schanuel theorem for pure Shimura varieties and was subsequently upgraded. See in particular Gao’s work on the Ax-Schanuel theorem for the universal abelian variety [13] and on the Betti map in [12, 14].
Gao’s main theorem, Theorem 1.1, in [12] may be seen as a geometric elaboration of what Theorem 3.4 means for the density of torsion. Gao considers an abelian scheme of relative dimension over a quasiprojective complex algebraic variety and a closed irreducible subvarierty and then establishes the conditions under which the generic rank of the Betti map restricted to may be smaller than expected. It is noted with [1, Proposition 2.2.1] that density of the torsion in follows from the Betti map, generically, having rank on . Thus, the converse of Gao’s condition gives a criterion for when the torsion is dense.
In more detail, taking finite covers if necessary, one may pass from the problem of density of torsion in as a subvariety of , to the density of torsion in in where is a universal abelian variety over a moduli space of abelian varieties of some fixed polarization type with some fixed level structure and the Cartesian square
expresses as coming from this universal family. To ease notation, we replace by and by . Shrinking the moduli space, possibly taking covers, and moving to an abelian subscheme of , we may arrange that is not contained in any proper weakly special varieties. At this point, Theorem 3.4 says that is dense in in the Euclidean topology if the intersection of with the zero section is persistently likely. Gao’s criterion expresses geometrically what persistent likelihood means here: for any abelian subscheme of , if is the quotient map, then is at least the relative dimension of over .
References
- [1] Y. André, P. Corvaja, and U. Zannier, The Betti map associated to a section of an abelian scheme, Invent. Math., 222 (2020), pp. 161–202.
- [2] V. Aslanyan, Weak modular Zilber-Pink with derivatives, Math. Ann., 383 (2022), pp. 433–474.
- [3] V. Aslanyan and J. Kirby, Blurrings of the -function, Q. J. Math., 73 (2022), pp. 461–475.
- [4] B. Bakker, Y. Brunebarbe, B. Klingler, and J. Tsimerman, Definability of mixed period maps. arXiv:2006.12403, to appear in the Journal of the European Mathematical Society, 2021.
- [5] B. Bakker, B. Klingler, and J. Tsimerman, Tame topology of arithmetic quotients and algebraicity of Hodge loci, J. Amer. Math. Soc., 33 (2020), pp. 917–939.
- [6] , Erratum to “Tame topology of arithmetic quotients and algebraicity of Hodge loci”, J. Amer. Math. Soc., 36 (2023), pp. 1305–1308.
- [7] B. Bakker and J. Tsimerman, The Ax-Schanuel conjecture for variations of Hodge structures, Invent. Math., 217 (2019), pp. 77–94.
- [8] G. Baldi, B. Klingler, and E. Ullmo, On the distribution of the Hodge locus. arXiv:2107.08838, 19 July 2021.
- [9] K. C. T. Chiu, Ax-Schanuel for variations of mixed Hodge structures. arXiv:2101.10968, 2021.
- [10] , Ax-Schanuel with derivatives for mixed period mappings. arXiv:2110.03489, 2021.
- [11] C. Daw and J. Ren, Applications of the hyperbolic Ax-Schanuel conjecture, Compos. Math., 154 (2018), pp. 1843–1888.
- [12] Z. Gao, Generic rank of Betti map and unlikely intersections, Compos. Math., 156 (2020), pp. 2469–2509.
- [13] , Mixed Ax-Schanuel for the universal abelian varieties and some applications, Compos. Math., 156 (2020), pp. 2263–2297.
- [14] , Corrigendum: Generic rank of Betti map and unlikely intersections, Compos. Math., 157 (2021), pp. 2747–2748.
- [15] Z. Gao and B. Klingler, The Ax-Schanuel conjecture for variations of mixed Hodge structures. arXiv:2101.10938, 2021.
- [16] P. Habegger, Weakly bounded height on modular curves, Acta Math. Vietnam., 35 (2010), pp. 43–69.
- [17] N. Khelifa and D. Urbanik, Existence and density of typical Hodge loci. arXiv:2303.16179, 2023.
- [18] J. Kirby, The theory of the exponential differential equations of semiabelian varieties, Selecta Math. (N.S.), 15 (2009), pp. 445–486.
- [19] B. Klingler and A. Otwinowska, On the closure of the hodge locus of positive period dimension, Invent. Math., 225 (2021), pp. 857–883.
- [20] D. Masser and U. Zannier, Torsion anomalous points and families of elliptic curves, Amer. J. Math., 132 (2010), pp. 1677–1691.
- [21] M. Orr and C. Schnell, Correction to “Height bounds and the Siegel property”. https://www.math.stonybrook.edu/~cschnell/pdf/papers/correction.pdf, 2021.
- [22] D. Pavlov, G. Pogudin, and Y. P. Razmyslov, From algebra to analysis: New proofs of theorems by Ritt and Seidenberg, Proceedings of the American Mathematical Society, (2022). https://doi.org/10.1090/proc/16065.
- [23] Y. Peterzil and S. Starchenko, Complex analytic geometry and analytic-geometric categories, J. Reine Angew. Math., 626 (2009), pp. 39–74.
- [24] J. Pila and T. Scanlon, Effective transcendental Zilber-Pink for variations of Hodge structures. arXiv:2105.05845v1, 2021.
- [25] M. Raynaud, Exposé XII: Géométrie algébrique et géométrie analytique, in Séminaire de géométrie algébrique du Bois Marie: Revêtements étale et groupe fondamental (SGA 1), 1961.
- [26] T. Scanlon, Algebraic differential equations from covering maps, Adv. Math., 330 (2018), pp. 1071–1100.
- [27] A. Seidenberg, Abstract differential algebra and the analytic case, Proc. Amer. Math. Soc., 9 (1958), pp. 159–164.
- [28] , Abstract differential algebra and the analytic case. II, Proc. Amer. Math. Soc., 23 (1969), pp. 689–691.
- [29] S. Tayou and N. Tholozan, Equidistribution of Hodge loci II, Compos. Math., 159 (2023), pp. 1–52.
- [30] S. von Kowalevsky, Zur Theorie der partiellen Differentrialgleichung, Journal für die reine und angewandte Mathematik, 80 (1875), pp. 1 – 32.
- [31] U. Zannier, Some problems of unlikely intersections in arithmetic and geometry, vol. 181 of Annals of Mathematics Studies, Princeton University Press, Princeton, NJ, 2012. With appendixes by David Masser.