Degrees of Hodge Loci
Abstract
We prove asymptotic estimates for the growth in the degree of the Hodge locus in terms of arithmetic properties of the integral vectors that define it. Our methods are general and apply to most variations of Hodge structures for which the Hodge locus is dense. As applications we give asymptotic formulas controlling the degrees of Noether-Lefschetz loci associated to smooth projective hypersurfaces in , and the degrees of subvarieties of the Torelli locus parameterizing Jacobians split up to isogeny.
0.1 Preamble
We adopt the convention that all algebraic varieties and schemes are defined over unless otherwise stated. The typical exception will be algebraic groups, which will almost always be -algebraic. Moreover all Mumford-Tate groups (defined below), will always be understood to be special Mumford-Tate groups. For two real sequences and , we write if and if is bounded away from both and as . Finally we write if for any there exists such that for all .
1 Introduction
Given a smooth projective family of varieties , the cohomology groups in degree carry a natural Hodge structure. Via the Hodge conjecture, rational vectors inside and its tensor powers are predicted to characterize, up to cohomological equivalence, the algebraic cycles associated to and its self-products. A question of much current and classical interest is to understand for which the fibre carries “more than the expected number” of algebraic cycles, i.e., more than at a very general point .
The loci in where the fibres acquire such cycles are examples of Hodge loci (conjecturally, all Hodge loci arise from families of algebraic cycles). Recent work [BKU24b] [KU23] has provided a conjectural framework — and in many cases largely settled — questions regarding their existence and density. More precisely, following the work in [BKU24b], one can give precise conjectures detailing when exactly such loci should exist, what sorts of vectors and tensors should define them, and whether they should be analytically or Zariski dense in . Moreover the work of [ES22] and [KU23] allows one to verify such existence and density properties in most cases that arise in practice.
However even knowing that such loci are dense in is in some sense only the beginning of understanding how many such loci there are. A more refined problem is to quantify how such loci grow as one increases the number or type of allowable vectors or tensors which define them. A natural way of counting these loci, dating back to at least [CDK94], is to use a polarizing form , where is the natural local system interpolating the fibres . Then one can consider the reduced Hodge locus defined by primitive integral vectors with for some integer . The locus is algebraic, so after choosing a quasi-projective embedding of it makes sense to ask for its degree; more generally, it makes sense to ask
Question: What are the asymptotics of as ?
In this paper we give some general techniques for answering this question.
1.1 Applications
To motivate the more technical results that follow, we give some concrete applications of our methods.
1.1.1 Noether-Lefschetz Loci
Let be the parameter space of smooth degree hypersurfaces in for some integer , where is the complement in of the discriminant locus . Write for the universal family. Using the natural map we obtain a family of line bundles in the fibres of . We define the primitive subsystem
where is the global section of coming from the relative Chern class of . The local system is torsion-free, and cup product induces a non-degenerate symmetric pairing which makes, as a consequence of Poincaré duality, each fibre into a unimodular lattice. We write for the signature of this lattice, which is related to the Hodge numbers of by and .
The Noether-Lefschetz locus of is constructed as follows. For each point and each integral vector , there is a locus consisting of all where some flat translate of in is Hodge. By [CDK94] the locus is algebraic, and we will choose to equip it with its underlying reduced structure. Moreover [CDK94] also shows that for the locus is algebraic.
Definition 1.1.
We say a vector in a -lattice is primitive if has non-zero image in for each .
Working with the union over all such that results in overcounting: if then we have . To remedy this, we define a modified scalar-invariant quantity better suited for counting Hodge loci which agrees with whenever is primitive.
Definition 1.2.
For a non-zero vector in some fibre of , we define to be the unique positive rational scalar for which is a primitive integral vector in the corresponding fibre of . We moreover define
which is a map of flat bundles over . We write and for the unique integers such that with positive and square-free.
It is clear that for each , and that modulo scaling by . We may then define
For a closed subvariety of we define its degree to be the degree of its closure in . We set and .
Theorem 1.3 (Upper bound).
For each fixed square-free we have
as , where is a real constant depending on .
Remark 1.4.
The quantity is , where is the second Betti number of a fibre of . It is known that , and by [Ara12, Ch 17.3] one has that . Combining these estimates one easily expresses as a polynomial in .
Definition 1.5.
The tensorial Hodge locus of is the collection of all Hodge loci (i.e., Noether-Lefschetz loci) associated to the variations for all .
We say a subvariety of is Hodge-generic if it is not contained in the tensorial Hodge locus. In what follows the term “sufficiently general hyperplane section” is understood as in 5.1. For the use of the term “period dimension” in the following statement, see 1.15.
Theorem 1.6 (Lower Bound).
Fix the data of:
-
-
a locally closed irreducible Hodge-generic subvariety , with period dimension ;
-
-
a relatively compact open neighbourhood ; and
-
-
a sufficiently general hyperplane section of with .
Then for each fixed square-free , and for with , one has
as , where is a constant depending only on and .
In the special case where , one has
Theorem 1.7 (Lower Bound).
For each fixed square-free , and for with , one has
as , where is a real constant.
Remark 1.8.
The lower bound given in 1.6 can be made optimal at the cost of a more complicated algebraic expression in the prime factorization of . We give the precise calculation in §6.2.1. With the optimal formula appearing there, one can replace the symbol with as long as does not intersect a certain finite union of tensorial Hodge loci of .
It is a consequence of the theory developed in [BKU24b] (c.f. [BKU24a, Thm. 6]) that each component of has codimension exactly in outside of a closed algebraic locus in , so away from this locus the “sufficiently general” hypothesis in 1.6 ensures that the cardinality is finite. The left-hand side of the inequality in 1.6 diverges as a function of , so one obtains infinitely many Hodge loci defined by vectors with in any neighbourhood of . In fact, away from the atypical Hodge locus, one can obtain exact asymptotics in terms of for the size of the Hodge locus in the prescribed region.
The period dimension hypothesis of on the other hand implies that each component of has positive period dimension; this will be used to verify that a general component of carries a unique global -dimensional subspace of Hodge vectors. For the importance of carrying out such a verification see our discussion in §1.3.1 and §1.4.
1.1.2 Split Jacobians
In this case we start with the universal family of genus curves (strictly speaking, we make sense of this using the language of Deligne-Mumford stacks, but the results would be no different were we to rigidify the moduli problem by adding marked points). We fix a projective compactification of (the coarse space of) and define the degree of an algebraic locus in to be the degree of its closure in .
Definition 1.9.
For a complex abelian variety , we say that has a -factor of degree if there is an isogeny such that and .
For a curve , we write for the associated Jaocbian variety. As an analogue of the Noether-Lefschetz loci considered above, we consider
Theorem 1.10 (Upper Bound).
For each integer such that we have
as , where is a positive real constant.
We define the atypical Hodge locus and the notion of a “sufficiently general” hyperplane as in the previous section except with . We consider the pair consisting of the symplectic group of a -dimensional integral symplectic space , and the Siegel upper half space, respectively. This pair can be viewed as the “generic Hodge datum” of the variation of Hodge structure (we review this notion below). The complex manifold has natural closed submanifolds whose images in consist exactly of those closed algebraic loci parameterizing non-simple principally polarized abelian varieties. Each such submanifold corresponds to a decomposition , compatible with the symplectic form, and the associated symplectic idempotents and belong to only finitely many orbits as the submanifolds vary. We let be the -orbit of such an idempotent whose image has dimension . We view as a -algebraic subvariety of the affine space associated to the free -module .
Theorem 1.11 (Lower Bound).
Let and . Fix the data of:
-
-
a locally closed irreducible Hodge generic subvariety of dimension at least ;
-
-
a relatively compact open neighbourhood ; and
-
-
a sufficiently general hyperplane section of with .
Then with one has
(1) |
as , where
-
-
is a positive real constant depending only on and ; and
-
-
is the unique -invariant measure on compatible with the Tamagawa measures on and the stabilizer in of .
Remark 1.12.
Remark 1.13.
One can show that the left-hand side of (1) diverges as .
Let us take some time to unpack the statement. The group is semisimple, simply-connected, and unimodular. By [Wei59, pg.26] one also knows that the locally compact topological group is unimodular, where we write for the adele ring. The semisimplicity and simply-connected properties imply that is a lattice in [Bor61, Thm. 1] [BHC61], and this lets us put a unique Haar measure , the Tamagawa measure, on such that has unit volume. The measure splits as a product of measures defined at each place of .
The stabilizer in of a Hodge-theoretic idempotent is abstractly isomorphic to a product of symplectic groups, and the same reasoning applies, mutatis mutandis, to . Moreover by [GO11, A.1.2, A.5.1] the space consists of finitely many orbits; let us fix such an orbit . Using [Gar18, Thm. 5.2.1, 5.2.2] there is a unique invariant measure on this quotient determined by the measures on and which again splits as a product of measures associated to each place of . Using the natural identification this induces in a natural way a measure on a component and on for each (note here that consists of a single orbit by [GO11, A.5.3]), and it is these measures we use in the statement of 1.11.
1.1.3 Comparing The Examples
To understand the relationship between our examples, let us explain how one would actually go about computing the asymptotic expression in 1.11. The Haar measure on the group can also be described using a -invariant -algebraic form on of top degree, in such a way so that the measures at each place of are obtained by integrating against . Such a form is called a gauge form. It turns out that the measures associated to and the quotient are also associated to gauge forms, and so the quantities appearing in the product in 1.11 can be computed assuming one is able to compute the relevant volume integrals over the -adic manifolds .
In the Noether-Lefschetz locus case, essentially the same reasoning applies. In that case, the relevant group is instead of . In particular, up to the scaling action of , the orbits of the group on fibres of are actually classified by the invariant , with two rational vectors lying in the same -orbit if and only if . (This follows from the fact that modulo , and then by using that is the unique integral square-free representative of the coset .) The product lower bound in 1.6 is in fact derived from an analogous product of adelic volumes coming from a natural adelic measure on a corresponding orbit of . The difference being that, in this case, we have taken the time to actually compute the relevant -adic volume integrals, and the factor at gives a contribution lower bounded by , where is the exponent of in the prime factorization of .
We expect similar computations can be done to give an explicit form to the left-hand side of 1.11, but this task is beyond the scope of this paper.
1.2 The General Case
1.2.1 Variational Setup
To facilitate the above applications, we now describe some general theorems applying to an arbitrary integral variation of Hodge structure with polarization on a smooth quasi-projective complex variety . This will require some additional setup. We will simplify the situation by studying the asymptotics only of Hodge loci associated to vectors of ; the case of loci associated to general Hodge tensors reduces to this by replacing with some tensor construction. We may also assume the weight of is zero, Tate-twisting if necessary. In particular we have a Hodge filtration on and the space of Hodge vectors above is the subspace . Given a vector in some fibre of , we write for the corresponding (reduced) Hodge locus. We define , and as in 1.2.
For each point we write for the Hodge structure on , which we view as a map . Here is the “circle group”, a real-algebraic subgroup of the Deligne torus constructed as in [GGK12, I.A]. The Mumford-Tate group at is the -Zariski closure of , and the Mumford-Tate domain at is the -conjugacy class of the morphism . We call the pair the Hodge datum at . The generic Hodge datum is the is the abstract Hodge datum which is isomorphic to at a point outside the Hodge locus of all the tensor powers for . The flat structure of the local system locally identifies the realizations of the generic Hodge datum and allows one to view the pair as an abstract object with a realization in each fibre of .
Definition 1.14.
For a local system on a smooth complex algebraic variety , the algebraic monodromy group is the identity component of the Zariski closure of the map
where is a point. Usually either some base point , the local system , or both, are understood, in which case we write or , as appropriate.
1.2.2 Period Maps
An important part of formulating a generalization of the examples we have discussed is deciding when a dense collection of Hodge loci should exist in ; this is not automatic, since for variations of higher level (in the sense of [BKU24b, §4.6]) one expects by [BKU24b, Conj. 3.5] that the Hodge locus should lie in a strict algebraic subvariety of . Recently a general criterion for proving the existence of such loci was given first in [ES22] in an abstract setting, and sharpened in the Hodge-theoretic setting in [KU23]. One of our main technical results will in some sense be a further strengthening of [KU23], so we start by adopting some of the setup. Because the statement of our main results can be deduced after replacing with a finite cover and with , we will freely make such changes.
Associated Period Maps:
We may replace with a finite étale covering such that induces a period map , where is a fixed neat arithmetic lattice containing the image of monodromy. Moreover by applying [GGK12, III.A.1] we may assume that we have a factorization , such that
-
(i)
the factorization of is induced by a factorization with ; and
-
(ii)
the factorizations in (i) are induced by the almost-direct product factorization , where we note that the center of acts trivially on .
Given a subset , we write where and . The map is defined similarly.
Definition 1.15.
For a subvariety , its period dimension is the dimension of .
Likeliness:
Given the generic Hodge datum , a Hodge sub-datum is a pair consisting of a subgroup which is the Mumford-Tate group of some Hodge structure and the associated orbit under the conjugation action. The Hodge datum of a point is the pair , where is the Mumford-Tate group of .
Definition 1.16.
We say the Hodge sub-datum is -likely if for every non-empty set the inequality holds, and we say it is strongly -likely if these inequalities are strict.
Subdata of Definition:
A component of the (tensorial) Hodge locus of can be defined by many different Hodge subdata. This means the following:
Definition 1.17.
We say a Hodge subdatum defines a closed algebraic locus if is an irreducible component of .
Note that need not be the Mumford-Tate group of in (1.17). A Hodge locus is always defined by a Hodge subdatum which arises from the pair at a very general point , but it is often useful to view as defined by a potentially larger subdatum.
1.2.3 Equivalence Classes of Hodge Subdata
The group acts on Hodge subdata by conjugation, with being equivalent to if there is such that and . We say Hodge data that are equivalent in this way are conjugation-equivalent. There are finitely many conjugation equivalence classes of Hodge subdata, a fact which is explained in [KU23, §4.4].
Definition 1.18.
We say that two Hodge subdata are -coset equivalent if there exists such that , and the coset defines a rational point in .
Lemma 1.19.
-coset equivalence is an equivalence relation on Hodge subdata.
Proof.
We observe that the notion of being -coset equivalent is symmetric: if defines a rational point in , then is a -algebraic subvariety of . Taking inverses, this implies that is a -algebraic subvariety of , and hence corresponds to a point of .
For transitivity, we consider three tuples and related by -cosets and . Then for any automorphism we have
∎
A given real conjugacy class of Hodge subdata, when further partitioned by -coset equivalence, can either result in countably-infinitely many subequivalence classes or finitely many. The situation with Noether-Lefschetz loci in §1.1.1 is an example of the first case, where the subequivalence classes are indexed by the invariant , and the second case is exhibited by the split Jacobian loci discussed in §1.1.2.
A useful converse result is the following:
Proposition 1.20.
Let be a Hodge subdatum, let , and suppose that . Then is a Hodge subdatum of .
Proof.
See 4.7. ∎
In many situations in practice, this means that counting -coset equivalent Hodge subdata is the same as counting rational points in .
1.2.4 From Hodge Loci to Point Counting
We are now ready to describe our main technical results.
Situation 1.21.
Let be a polarized integral variation of Hodge structure on satisfying . Suppose that is a Hodge subdatum, and that under the natural Hodge representation
-
-
the group is identified with the stabilizer of some ;
-
-
one has , where is the Hodge datum at .
Let be the orbit of and write for the component containing , and set . Fix a projective compactification and define the degree of to be the degree of its closure.
Note that the orbit is preserved by the action of , and so has, for each , a well-defined realization as a subvariety of the affine space associated to . Likewise it makes sense to consider in each fibre of : those points with the property that is integral. (Here we are using the notation “” in the same spirit as with and above: it is an abstract object with a realization above each point .) Given a vector we obtain, using the fixed isomorphism , a Hodge subdatum defined as in 1.20. We then set to be the corresponding Hodge locus.
Theorem 1.22 (Upper Bound).
Work in the situation of 1.21. Then there exists:
-
-
finitely many Hodge-theoretic Siegel sets for associated to Hodge structures , such that the Hodge datum of each is contained in a Hodge datum -coset equivalent to ; and
-
-
Hodge vectors , with a Hodge vector for ;
such that we have
(2) |
as , where is a real constant depending on and .
Notation.
For each fixed non-negative integer and analytic variety , we write for the union of -dimensional components of .
Theorem 1.23 (Lower Bound).
Work in the situation of 1.21, assume that is strongly -likely, and that has finite index in its normalizer in . Let be any open neighbourhood. Then there exists a sufficiently general hyperplane section which intersects and has codimension , and an open subset such that
(3) |
as , where is a real constant depending on and .
The basic summary of the the two theorems is that Hodge loci defined by vectors with are roughly in correspondence with rational vectors in certain subsets of with denominators dividing . These subsets are at worst as large as a Siegel set orbit, and in the strongly -likely case contain open neighbourhoods corresponding to any chosen open subset of . The Siegel sets referred to in 1.22 are “Hodge-theoretic” in the sense that their associated maximal compact subgroups are naturally constructed from the Hodge metric at a point ; we give a precise description in §2.1. Moreover we act with these sets on Hodge vectors for the associated .
It turns out that such Siegel-set orbits constructed from Hodge-theoretic data have highly constrained geometry: they are in general non-compact, and have infinite volume for the natural invariant measure on , but nevertheless satisfy the property that, for any positive integer , they contain only finitely many rational points with denominators of size at most . This means in particular that counting points in such orbits will make sense. A basic result is then the following:
Proposition 1.24.
For any Hodge-theoretic Siegel set associated to a Hodge structure on a polarized integral lattice , a -algebraic subvariety of the affine space associated to , and orbit of a Hodge vector for , there exists a constant such that
where is the signature of the underlying polarized lattice.
We note that the above result uses both that the Siegel set is Hodge-theoretic and that is a Hodge vector in a crucial way, and we expect that in general the intersections on the left are infinite without these assumptions. As a corollary we obtain:
Corollary 1.25.
For any polarized integral variation of Hodge structure and orbit of , there exists a constant depending on such that
where is the signature of .
In the above result is the Noether-Lefschetz locus associated to and , defined as
Finally, we also have an overall bound on the Noether-Lefschetz locus for an arbitrary polarized VHS:
Corollary 1.26 (General Upper Bound).
For any polarized integral variation of Hodge structure there exists a constant depending only on such that
where is the signature of .
In the above we define by
The bound in 1.26 is deduced from 1.25 via a trick and we expect it can be substantially improved.
In explicit situations all of the above abstract results can be sharpened, and this is what we have done in the examples in the previous section. What happens then is that one knows explicitly the groups , , and , the structure of the Siegel sets becomes more explicit, and correspondingly the point counting becomes easier. We discuss this further below.
1.3 A Sketch of the Ideas behind the Arguments
1.3.1 Reduction to Point Counting
Let us denote by the natural projection, and fix a Hodge subdatum . Suppose one wanted to merely count loci in which are “generalized Hecke translates” of the image of some : i.e., images under of , where is a Hodge subdatum of which is -coset equivalent to . Then as we have discussed, one can view each such generalized Hecke translate as corresponding to a rational point of . One can then try to pick a fundamental set for the action of on , and count generalized Hecke translates by counting rational points in . (The space does not usually have a reasonable topology, so one has to work directly with fundamental sets, but we ignore this for the moment.)
However in general one merely has a map , and to make the strategy work one faces at least two potential problems.
-
(1)
The loci might not intersect . More generally, if one looks at the locus of for which intersects this locus might have empty interior, making it difficult to produce rational points inside it.
-
(2)
It could be that many different varieties of the form , in particular infinitely many, intersect in the same locus. In other words, one has many different rational points of corresponding to the same Hodge locus, leading to overcounting.
The first problem (1) is solved using by the “-likely hypothesis”, which can be used to guarantee intersections, following [KU23]. Then it would in fact follow from the Hodge-theoretic Zilber-Pink conjecture that (2) does not occur (more precisely, is controlled by a finite action of the normalizer of in ) away from a strict Zariski closed locus in . This being unavailable, we need to assume the “strongly -likely” hypothesis, which, following the arguments in [KU23], suffices to demonstrate the predictions of Hodge-theoretic Zilber-Pink in this case.
To go from merely “counting Hodge loci” to a precise degree estimate one replaces with an appropriately chosen hyperplane section so that the Hodge loci of interest are zero dimensional. One then carefully partitions into analytic neighbourhoods such that a Hodge locus in of the form always consists of a single point. That this can be done for belonging to a finite definable analytic cover is a consequence of the definability in of Hodge-theoretic period maps proven in [BKT20]. Then by counting Hodge loci in each and summing over all , one recovers an estimate for the number of points in the Hodge locus, which, having reduced to the zero dimensional case, is the degree estimate we wanted.
The task thus reduces to counting Hodge loci in each . Following the theory in [BKT20], we can reduce to the case where each maps into a Hodge-theoretic Siegel set. After this, one establishes a correspondence between the Hodge locus points of interest and vectors in a Siegel set orbit, and counts points in the manner discussed above.
1.3.2 Counting Points
For the actual point counting itself, there are two tools. The first is the work of Gorodnik-Oh [GO11], which gives general techniques based on adelic equidistribution results for counting points in subsets of homogeneous spaces for algebraic groups. These estimates are often sharp, and describe the asymptotic point counts using Euler-product-like expressions over all places of . One can essentially always apply such techniques to count rational points inside compact subsets of , which leads to the lower bounds in 1.6 and 1.11. However it is unclear how to apply such techniques to count rational points in a general Siegel set orbit, especially given that usually such orbits are non-compact and have infinite volume in the natural -invariant measure.
The upper bounds, therefore, are computed via a different method. We study carefully the geometry of Siegel set orbits, and reduce the problem to an analysis of orbits of certain “standard Siegel sets”. Then we compute the orbits in explicit coordinates, and prove that one can bound the size of rational points in such orbits by the denominators of specially chosen coordinate entries. This in particular implies that rational points in Siegel set orbits whose denominators divide lie inside a compact subset of volume polynomially bounded by . After this, the result follows from elementary bounds on the number of lattice points in a compact region.
1.4 Comparison with Other Work
Our results are similar in spirit to recent work studying the equidistribution of Hodge loci, in particular the papers [Tay20] and [TT23]. There are a few differences and similarities.
The first is that we estimate the degree of Hodge loci (or Noether-Lefschetz loci) for a polarized variation of Hodge structure , while [Tay20] and [TT23] both essentially estimate the degree of a corresponding “locus of Hodge classes” of . Here we are borrowing the terminology of [CDK94] and Voisin [Voi10] to describe the analytic loci in the vector bundle consisting of points with a Hodge class for the Hodge structure on . The vector bundle has a canonical algebraic model for which the components of the locus of Hodge classes are algebraic [CDK94]. That [TT23] count a (subset of the) Hodge class locus rather than the Noether-Lefschetz locus in can be seen in the statements of the asymptotic theorems [TT23, Thm. 1.6, Thm. 1.7, Thm. 1.21, Thm. 1.22]. The situation in [Tay20, Thm. 1.1] is similar, where points in the Hodge locus are “counted with multiplicity”, which in practice amounts to counting points by the number of Hodge classes which define them (but regarding Hodge classes differing by a scalar as equivalent).
Because a Noether-Lefschetz locus is always a projection under of a locus of Hodge classes, asymptotic lower bounds for degrees of Noether-Lefschetz loci imply asymptotic lower bounds for the degrees of Hodge class loci, and asymptotic upper bounds for the degrees of Hodge class loci imply asymptotic upper bounds for the degrees of Noether-Lefschetz loci. But there is in general a gap between the two, and understanding the size of this gap and when it occurs is the subject of non-trivial questions in unlikely intersection theory. For instance, the Zilber-Pink conjecture of [BKU24b] predicts that if one considers a Hodge-generic family of smooth projective degree hypersurfaces in with a curve, the Noether-Lefschetz locus of is a finite subset of . Formulated for the locus of Hodge classes, or when counting intersections with Noether-Lefschetz components in with multiplicity, the statement is false: if is a point mapping to the Fermat surface in (recall §1.1.1), the locus of Hodge classes above consists of all points in belonging to an infinite lattice of large rank (c.f. [AMVL19, Thm. 1]), and counting with multiplicity each -dimensional subspace is counted separately. And situations where the Hodge locus of is infinite and the two types of loci should have different asymptotics also occur: if one takes to be a Shimura curve in with , every Hodge locus point arises in infinitely many different ways as the intersection of with a special locus in , and from this one can show that counting the Hodge class locus (or counting with the multiplicity of the intersections) gives an overcount of the Hodge locus of . (This phenomenon is also what we identified as “potential problem (2)” in §1.3.1.)
The second difference is that our upper bounds are global in nature. Although one could in principle use the methods of [TT23] to give upper-bound asymptotics for Hodge loci inside small neighbourhoods of of uncontrolled size, to deduce a global point count on all of one must, at least in the case of non-compact , sum over infinitely many such neighbourhoods. Without uniformly controlling the rates of convergence on all neighbourhoods at once, it is then difficult to obtain a global estimate. Our methods allow one to reduce the problem to counting rational points associated to finitely many Siegel sets, which crucially uses the limit theory of variations of Hodge structures. This finiteness ensures one can safely combine the asymptotic estimates in each region. An exception to this is [Tay20, Thm. 1.1], which also achieves a global upper bound for variations of K3 type.
On the other hand, the underlying principle in both our approaches is the same. The idea is to use Hodge theory to replace degree estimates with counting problems for rational vectors. Over a compact set, explicit estimates may then be deduced from the work of Eskin, Gorodnik, Oh and others on such counting problems. We also both rely on measure theory, though we do our computations in coset spaces whereas [TT23] instead integrates a differential form on obtained by applying a “push-pull” procedure to a form on such a coset space. Ultimately such volume computations should amount to the same thing and we believe that one could use the results of [TT23] to deduce the analogues of the lower bound statements 1.6 and 1.11 for an associated Hodge class locus (indeed, for 1.6 this is basically contained in [TT23, Thm. 1.6, Thm. 1.7], and for split Jacobians there is [TT23, Thm 1.14]). But for producing a lower bound for the Noether-Lefschetz locus itself, we do not know how to do so without additionally using arguments like those in [KU23] to solve the relevant unlikely intersection problems, and this is the approach to Noether-Lefschetz lower bounds we take in this paper.
1.5 Acknowledgements
The author thanks both Salim Tayou and Nicolas Tholozan for several discussions about their work, and for explaining to him their perspective on Hodge locus asymptotics. He thanks Salim Tayou in particular for discussions at the MSRI (now SLMath) and at the third JNT Biennial Conference in Cetraro, and for pointing him to his paper [Tay20].
2 Recollections on Period Maps
2.1 Siegel Sets
Let be the generic Hodge datum associated to , and the period map induced by with a neat arithmetic lattice containing the image of monodromy. By embedding as an open subvariety of its compact dual the set inherits a natural -definable structure.
To define a Siegel set of we require some setup. Each such set will be defined as a certain orbit associated to a maximal compact subgroup and a minimal parabolic -subgroup . The maximal compact group will be defined using a Hodge structure by the following lemma:
Lemma 2.1.
Suppose that is a point, and let be its stabilizer. Then there is a unique maximal compact subgroup containing .
Proof.
The Hodge structure induces a grading through the adjoint action, and we define to be the exponential of the real Lie subalgebra underlying . As the Lie algebra of may be identified with the summand one clearly has . A polarization on the Hodge structure induces a polarization on each simple summand of which is -invariant when restricted to that summand, see [GGK12, III.A.5, pg.75]. Necessarily such a form is proportional to the Killing form on each summand, which implies that the restriction of the Killing form to is definite of a common sign, and hence negative definite because the Killing form is negative definite on . This implies the restriction of the Killing form must be definite with a common (opposite) sign on the odd part of the Lie algebra. It then follows from [Mos61, §4] that is a maximal compact subgroup containing . That it is unique, which amounts to the statement that any other subalgebra of containing on which is negative definite lies in , follows from the Hodge-Riemann bilinear relations for the polarizing form on . ∎
Lemma 2.2.
For any minimal parabolic -subgroup of a reductive -group and maximal compact subgroup , there exists a unique real torus satisfying:
-
(i)
the torus is -conjugate to a maximal -split torus of ; and
-
(ii)
is stabilized by the Cartan involution associated to .
Proof.
This is [Orr18, Lem 2.1]. ∎
Definition 2.3.
Suppose that is a minimal parabolic -subgroup, and let be as in 2.2. For any real number write
where is the set of simple roots of with respect to , using the ordering induced by . Then we define a Siegel set (associated to ) to be a set where is compact. We likewise define a Siegel set (associated to ) to be an orbit .
2.2 Definable Period Maps
We now recall one the main results of [BKT20] (in the form corrected by [BKT23]), which says roughly that local period maps associated to polarized integral variations of Hodge structures land inside Siegel sets. We also explain a very mild strengthening of this result which we will use in our arguments.
We note that if is any definable fundamental set for , there is a unique induced definable structure on for which the map is definable; see [BBKT24, Prop. 2.3]. The following is a minor variant of [BKT20, Thm. 1.5], whose argument we summarize for completeness.
Proposition 2.4 ([BKT20] + ).
There exists a definable fundamental set for , obtained as a finite union of Siegel sets , such that the map is -definable for the induced definable structure on . In particular, there exists a finite cover by definable simply-connected opens and -definable local lifts of .
Fix a set of points . Then we may choose the above data such that for each intersecting the topological closure of in , the Siegel set is associated to a point .
Proof.
We start by summarizing the proof in [BKT20] of the first paragraph in the statement of 2.4. We then explain how to modify the argument to involve at the end.
Applying Hironaka’s Theorem we may reduce to the case where is the complement of a simple normal crossing divisor in a smooth projective variety . By passing to a finite étale cover we may assume monodromy at infinity is unipotent.
It is clear we can construct such a in a small enough neighbourhood of any , so it suffices to show this on a small enough punctured neighbourhood of the simple normal crossing divisor ; this being done, the finiteness of the cover will follow from compactness. To do this we follow [BKT20, §4]. Without loss of generality we take by allowing factors with trivial monodromy. We let be the standard universal covering and let be the standard Siegel fundamental set as in [BKT20, §4.2]. Then we may construct a map by lifting , and define by inverting after choosing a branch cut and and composing with . By expanding the map in terms of its real-analytic components one sees it is -definable, so to show that is -definable it suffices to check the definability of .
To do this we apply the Nilpotent orbit theorem [Sch73], which tells us that
where are the natural coordinates for , the are nilpotent operators, and is an analytic function on , hence definable. Then as the left factor is just a polynomial in , the result follows.
Following [BKT20, §4.5], it now suffices to check that the image of lies in a finite union of Siegel sets. The monodromy representation associated to gives a faithful representation where is an integral lattice, and hence induces a map , where is the symmetric space of positive-definite symmetric forms on ; here the the map is given by sending the polarized Hodge structure to the positive-definite symmetric form , where is the Weil operator associated to . From [BGST23, Prop 3.4] one learns that the inverse image under of a Siegel set of associated to is contained in finitely many Siegel sets of for . Thus we may reduce to the same claim for the image of .
Using the explicit description of Siegel sets for in terms of the Gram-Schmidt process, the condition that the set lie inside a Siegel for is equivalent to the following claim: there exists a basis for and a constant such that the following inequalities hold for all :
-
(i)
for all ;
-
(ii)
for ; and
-
(iii)
.
The corresponding Siegel subset of may be taken to be associated to the maximal compact group associated to a chosen point (c.f. [BGST23, Thm. 3.3] and its proof). In particular we can take .
We now check conditions (i), (ii) and (iii) by summarizing the argument in [BKT20, §4.5], which we refer to for details. Using the limiting weight filtration one can choose a basis so that condition (iii) holds for some as a consequence of [BKT20, Thm. 4.8], and (ii) can be assumed for the same by reordering the basis and partitioning the image as necessary. Schmid’s -dimensional result shows all of these conditions after restricting to any curve in for a constant depending only on , and so in particular shows (1) on such curves. One can then make the constant appearing in (1) independent of using the fact that the coordinates of are “roughly polynomial” (see [BKT20, Lem. 4.5]) in the coordinates as a consequence of the results [CKS86] and [Kas85] on the asymptotics of Hodge forms. This completes the proof of the first paragraph of 2.4.
Now we explain how to involve the set . The above argument has showed how, around any point , and given any ball centred at with , we can construct a period map on landing inside a Siegel set depending on . (Up to shrinking and taking a branch cut.) This in particular applies to points , where we write for the topological closure of in . Now as discussed above, the maximal compact of the resulting Siegel set can be chosen to be associated to any point . Since is dense in , we can in particular choose .
The set is a closed subset of a compact set, so is in particular compact. Thus, after applying this argument at every point and obtaining an appropriate cover of , we can find a finite subcover. This finite subcover may then be extended to a finite cover of all of which induces the cover of given in the statement. ∎
3 Siegel sets and their Orbits
In this section we study carefully the structure of, and point counting in, Siegel set orbits for the group .
3.1 Standard Siegel Sets
We set , and . We will view as the stabilizer of the bilinear form . Replacing with and reordering the standard coordiantes if necessary, we may assume that ; none of our results on Siegel set orbits will depend on this convention.
We will assume that is a Siegel set constructed with respect to a certain special choice of maximal compact subgroup and parabolic subgroup ; we justify this in 3.2 below by showing that one can always reduce to this setting by a change of coordinates. We fix the maximal compact subgroup . To construct we introduce a second set of coordinates, , which are related to the coordinates by
(4) | |||||
(5) | |||||
(6) |
The change of basis matrices from -coordinates to -coordinates are given by
And in the new coordinate frame, the bilinear form is given by
We compute the Lie algebra in the -coordinate system, which is given by matrices satisfying . For such , one calculates that
So in the -coordinate system one obtains the description
Write the group of upper triangular matrices, and for its Lie algebra. We define a Lie algebra by the equations
In -coordinates, the corresponding matrices have the form
We write for the nilpotent Lie algebra defined by setting and mandating that the diagonal entries of are all equal to , and set .
Lemma 3.1.
The set is the Lie algebra of a minimal parabolic subalgebra of .
Proof.
[He13, §2.3] gives a description of a minimal parabolic subgroup of . One immediately checks that our algebra and the group appearing in [He13, §2.3] have the same dimension, so it suffices to check that is parabolic. By [BDPP11, Def. 1.1] it suffices to check that the orthogonal complement with respect to the Killing form is nilpotent, which is easily checked. ∎
Write for the associated Parabolic subgroup of . By exponentiating, we see that
(7) |
We now describe the split torus determined by 2.2. The Cartan involution associated to in -coordinates is given by . In the -coordinates this becomes
The matrices and are both diagonal, so one sees that the diagonal matrices of are preserved under the involution. We then take the Lie algebra of to be the intersection in the -coordinates of and the diagonal subgelbra of . It is then clear that is split over of the correct dimension (compare with the torus in [He13, §2.3]), and stable under .
The rank of the group is (recall we have assumed ), so we should have simple roots matching 2.3. Since our torus lies in the diagonal subgroup of , we may obtain these roots from restrictions of roots of acting on , in particular, we have
Lemma 3.2.
Let be a maximal compact subgroup and a Siegel set for relative to and a minimal paraoblic . Then there exists , and , and a Siegel set for such that where . We have and .
Proof.
Our argument is a mild variant of [Orr18, Lem. 3.8]. Let be the parabolic group associated to . Then as both and are minimal parabolic -subgroups, there exists such that . Since and are both maximal compact subgroups, there exists such that .
We now consider the Iwasawa decomposition of with respect to the Cartan involution . We recall that this is induced by a decomposition , where and are both eigenspaces for , and is a choice of maximal abelian subalgebra of . Since the Lie algebra of is stabilized by , we may choose to be the Lie algebra of . It is then directly checked that is a sum of root spaces for . It follows that , and that we may choose with and . It follows that and hence by 2.2 that , so we may write . Finally, as in the proof of [Orr18, Lem. 3.8]
so the result follows choosing and sufficiently large. ∎
3.2 Orbits of Hodge Vectors
To compare the above Siegel sets to the ones given in §2.1, we fix a Hodge structure , let be the associated maximal compact, and let be an associated Siegel triple. As in §2.1, the group is the intersection with of , where is the Hodge-metric at .
Notation.
Given a vector , we write either or for its ’th entry. We let be the smallest index in the specified range for which . If we write we understand . If no such index exists, then we say is undefined.
Proposition 3.3.
Fix a Siegel set for and a real Hodge vector for . Then there exists a basis for with the following properties:
-
(1)
in the basis , the form is equal to and the special orthogonal group of is equal to ;
-
(2)
writing for the coordinates with respect to and defining as in §3.1, there exists real constants , independent of , such that
(8) (9) if is defined, and otherwise.
Proof.
Define and as in 3.2 so that , where is one of the standard Siegel sets constructed above. We define as , where is the standard basis. Then we have
The left-hand side is nothing more than the orbit except in the coordinates determined by , so it suffices to prove the inequalities for the points inside the right-hand side. The vector is a real Hodge vector for the Hodge structure . Moreover because , the maximal compact associated to agrees with the stabilizer of the Hodge metric for . It thus suffices to prove the statement of the proposition when is a standard Siegel set which is also a Hodge-theoretic Siegel set for .
We thus reduce to considering the orbit , with as in the previous section. Let be the Hodge decomposition associated to , where we assume the Hodge structure has weight where is even. Define and , which we may view as subspaces of . The Weil operator acts as on , so the forms and the Hodge metric agree on one of the two summands and agree up to a sign on the other.
The maximal compact group associated to preserves both and . Since we have reduced to the setting where the maximal compact associated to is , it follows that . Then because is Hodge, hence , we observe that all the vectors in lie inside as well. In particular, if , then for in -coordinates, with the sign depending on which of the two summands in the space is identified with. We will assume the sign is positive, with the other case handled analogously.
We now consider a vector which we may write as with , , and . After possibly adjusting and and decreasing , we may assume that all diagonal entries of are equal to except for possibly those that lie in the central block. We write for the ’th diagonal entry of in the -coordinate system. We note that
for . We may assume is defined: otherwise and the entries of are universally bounded as both and range over a compact set.
We then have
Choose such that and . From the inequalities defining we then get
We now take such that , which we can do independently of and since is bounded from below independently of . Taking in the above gives
for all . One computes immediately from the description of in (7) that and that ; in particular is defined. It thus follows that
(10) |
for all .
Now we also have , and hence for all . This implies that for all , in particular, there is an absolute bound such that
for all . We can even assume this is true for all using that acts trivially on the coordinates indexed by .
We conclude that for all we have
Then since lies in a compact subset , the inequalities continue to hold with replacing after possibly adjusting and . ∎
3.3 Point Counting Upper Bounds
We continue with the notation and setup of the preceding section. We view as the set of real points of the standard -dimensional affine space .
Proposition 3.4.
Fix a Siegel set for and a real Hodge vector for . Then there exists a real constant such that
as .
Proof.
The statement is unchanged (up to adjusting ) after a -linear change of coordinates, so we may work in the coordinate system of 3.3. In the region of where is undefined we then know that is absolutely bounded, and on this region the result follows by projecting to a coordinate hyperplane. We may thus assume the inequalities (8) and (9) hold for some fixed . Rearranging these inequalities one obtains
(11) | |||||
(12) |
where we take using that from (12).
For each index we define a map via the coordinate functions
(13) |
Letting be the locus where , we observe that the induced map
is injective. Indeed the locus is defined by an equation
Proposition 3.5.
Fix a Siegel set for and a real Hodge vector for . Suppose that is an irreducible -algebraic subvariety. Then there exists a real constant , dependent on , such that
as .
Proof.
Once again we work in the coordinate system of 3.3. We write for the locus where vanishes, and let be the complement. Let be the subset of indices such that . Then by arguing inductively on for , we may reduce to considering just those rational points in .
We may assume is non-empty. Otherwise is undefined for each , and the desired result follows by projecting to a coordinate hyperplane and using that is absolutely bounded. Then we let be the smallest entry of .
We consider a projection defined by composing the map defined in the proof of 3.4 with a further projection onto a coordinate hyperplane and a translation by an integral vector in . We may choose these projections and translations such that is quasi-finite away from a closed -algebraic sublocus of which we handle inductively. For the remaining points we are reduced by taking the image under to counting points in which lie in a compact region, so the result follows. ∎
4 Assorted Tools
4.1 A Definability Lemma
We start with a basic lemma regarding definable sets, which follows from definable cell decomposition. A definable cell decomposition is a partition of into definable subsets, called cells, defined inductively as follows (c.f. [vdD98, Ch. 3, 2.3]):
-
(i)
when , the cells are open intervals and singleton sets;
-
(ii)
given a cell , a cell of is either:
-
-
the graph of a continuous definable function viewed in the natural way as a subset of ; or
-
-
the region in defined by
where are continuous definable functions satisfying .
-
-
Then one has [vdD98, Ch. 3, 2.11]:
Theorem 4.1.
For any definable subset there exists a cell decomposition of such that is a finite union of cells.
We will find the following lemma useful in later arguments.
Lemma 4.2.
Let and be definable sets, and a definable subset such that the intersections are all finite. Then there exists a definable partition such that the intersections have cardinality at most for each and .
Proof.
We may reduce to the case where and , so is a definable subset of . We then consider a cell decomposition of . We let be a cell, and prove that has cardinality at most for each . From the definition, the cell is obtained as a sequence where is a cell in and is obtained from in one of the two ways described in (ii) above.
For each we have a natural projection
(15) |
For this is just a map . We observe that all such maps are necessarily surjective, which is immediate from the two possibilities for the construction of from in (ii). Thus the hypothesis that is finite in fact implies that is finite for all , since each map is surjective as it is a composition of surjective maps.
It then suffices to prove inductively on that is a singleton, which at each stage amounts to proving that the (necessarily unique, by induction) fibre of is a singleton. Notice the finiteness of means that cannot be obtained from by a construction of the second type in (ii) above, since then would contain the infinite set
for the (necessarily unique, by induction) point .
Hence is obtained from by a construction of the form
But then the induction hypothesis that is a singleton implies that , so the result follows. ∎
Corollary 4.3.
Let be a definable function with finite fibres. Then there exists a definable partition such that is injective.
Proof.
Let be the graph of . Then after swapping the order of the coordinates, we obtain from 4.2 a definable partition such that has cardinality at most for each and . Define to be the projection of and observe that is just the graph of . ∎
4.2 Stripes
4.2.1 Parametrization of Stripes
We work in the setup of §2.1. We additionally fix a fibre , the choice of which is unimportant, and identify with their realizations at ; in particular, is a subgroup of which lies in the group of linear mappings which preserve the polarizing form on and have determinant one. Fix a Hodge-theoretic Siegel set associated to , and set . We will denote the natural orbit map given by by . In what follows all definable sets are regarded as definable relative to the structure . We let be a point with Mumford-Tate group , and fix another point . Then has Mumford-Tate group contained in . We write for the compact dual of , which is a projective algebraic variety containing as an open submanifold. It can be identified with the orbit of on in the space of all Hodge flags on , where is the Hodge flag corresponding to .
Notation.
Suppose is a -algebraic variety and is a subset. We write for the intersection .
Notation.
Given a group with subgroup , we write (resp. ) for the normalizer (resp. centralizer) in of . If both and are algebraic groups, we interpret (resp. ) as an algebraic group.
Notation.
For an algebraic group , we write for its identity component.
Remark 4.4.
If is a connected reductive subgroup of a connected reductive group , then (c.f. [hh]). This is often useful for understanding when is a Mumford-Tate subgroup of .
We set to be the -orbit of in , and to be the corresponding orbit in , where is the Hodge flag corresponding to . We write for the -algebraic locus considering all Hodge flags whose set of Hodge tensors contains those tensors fixed by (c.f. [GGK12, Ch. 2]).
Proposition 4.5.
Let be the Zariski closure in of
and the algebraic group defined by
Then if is the identity component of , we have
(16) |
Moreover, the identity components of all three groups agree.
Proof.
We first observe that implies because is open in the irreducible variety . This implies the latter inclusion . For the former we use [GGK12, VI.A.3] to obtain that , and note that by [GGK12, VI.B.1] is a (geometrically) connected component of . As a consequence of [GGK12, VI.B.1], the (geometrically) irreducible components of agree with the (geometrically) connected components. This implies that geometrically connected group preserves . Then if one necessarily has that (using [GGK12, VI.B.11]), hence . But is Zariski dense in the connected -group since connected algebraic groups over a field of characteristic zero are unirational [Bor91, Theorem 18.2].
Now for the claim about identity components. We start by using unirationality to observe that is Zariski dense in . Now consider an element . We have , and so sends a generic point of with Mumford-Tate group to another such point. Applying [GGK12, VI.A.3] this tells us that . Then necessarily since is connected. The analogous argument works for . ∎
Definition 4.6.
A -subgroup is said to generate if .
We set . Each flag induces a filtration on which we denote with the same notation.
Lemma 4.7.
Suppose that generates and consider a point such that is non-empty. Then there exists a Hodge structure with Mumford-Tate group such that . If admits a representative , one has and .
Proof.
Note that because , the group is independent of the representative chosen. Thus is -algebraic because is. Let be a Hodge flag corresponding to a point , where is a Hodge flag corresponding to a point of . Write for the vector subspace of tensors fixed by . Choosing sufficiently generic, this is exactly the -span of . One has . As the group is defined over , the complex vector space has a natural underlying -structure , and by construction . Then the Mumford-Tate group of fixes all tensors in , hence lies inside . Taking sufficiently generic and using that is irreducible, the points of have Mumford-Tate group contained in a common -algebraic Mumford-Tate group . Let . We then have
where we have applied [GGK12, VI.B.11].
Now is exactly the locus of Hodge flags whose set of Hodge tensors contains . It thus follows that . For both the varieties and their geometrically irreducible components agree with their geometrically connected components (this is a consequence of [GGK12, VI.B.1]), so the inclusion also implies an inclusion of components passing through . Then
It remains to show that and that under the assumption . For the second claim, observe that
where we again apply [GGK12, VI.B.11]. Hence . Finally is the generic Mumford-Tate group of a point in , hence . ∎
Definition 4.8.
A Mumford-Tate domain obtained as in 4.7 is called an -translate of .
We thus obtain a map
which sends to . We write for the connected component containing . We then write for the restriction of to .
Lemma 4.9.
Each fibre of lies inside a fibre of . Each fibre of lies inside a fibre of .
Proof.
If then which implies that . Similarly if then and hence . ∎
Definition 4.10.
Suppose that is a subset containing , and that is a -algebraic subgroup which generates . By a stripe of in we mean a non-empty intersection for some . A stripe is said to be principal if we can choose .
Definition 4.11.
Given a subset containing and a -subgroup generating , we say that a stripe
of in is generic if there is a point of with Mumford-Tate group . We say it is uniformly generic if every connected component of has such a point.
Lemma 4.12.
In the context of 4.11, the genericity of the stripe does not depend on ; if and there exists a point with Mumford-Tate group , then in fact and .
Proof.
Let be a point with Mumford-Tate group equal to . Then so . On the other hand , so is open in . Since any open subset of contains a Hodge-generic point it follows that . Likewise, . ∎
We now obtain a diagram
(17) |
where the map sends a stripe to the (necessarily unique by 4.12) -translate of which induces it. We wish to understand the inverse image by constructing, using , a subset of which contains it. For this we define
Lemma 4.13.
The restriction of to induces a surjection onto the principal stripes of in .
Proof.
Suppose that is such that is non-empty. Since , we may choose . From 4.7 we then know that , so is non-empty. Since , we may therefore choose such that . Then and . ∎
4.3 Point Counting in Coset Spaces
In this section we explain how to count rational points inside , at least for appropriately chosen and . For this we will use some results of [GO11].
4.3.1 Measures
Definition 4.14.
We say an algebraic group is scss if it is simply-connected and semisimple.
Suppose that is an inclusion of -algebraic scss groups and write for the quotient -variety. Given Haar measures and on and , respectively, we say that a measure on is compatible with and if for any compactly supported function on we have
In our case, because both and are unimodular, such a measure exists by [Gar18, Thm. 5.2.1, 5.2.2] and is uniquely determined up to a scalar. This scalar can be fixed by requiring that (resp. ) induces a probability measure on the quotient (resp. ); note that these quotients have finite volume as a consequence of [Bor61, Thm. 1] [BHC61]. We will always adopt the convention whenever considering a triple with and scss that the measures and have been chosen in this way.
In a situation where we have a representation on a -vector space such that is the stabilizer in of some , we will also denote by the induced measure on the adelic points of the orbit variety . Using the inclusion we also obtain induced measures on for each prime , and likewise a measure at the infinite place.
4.3.2 Orbit Asymptotics
We now recall a theorem proven in [GO11]. We note that [GO11] uses right-coset spaces instead of left ones, the latter being our convention. The notation refers to compactly supported continuous functions on the topological space inside the brackets. In what follows we set .
Proposition 4.15 (Prop. 5.3 in [GO11]).
Suppose that both and are scss -groups with a maximal connected -subgroup. Set . Then for any well-rounded sequence of compact subsets whose volume diverges as , we have,
where is the class of the identity.
The statement of the proposition uses the following definition:
Definition 4.16.
A family of compact subsets is called well-rounded if there exists such that for every small , there exists a neighbourhood of in such that for all sufficiently large ,
We note that our definition of “well-rounded” is the special case of the notion “-well-rounded” appearing in [GO11, Def. 5.1] with .
Proof of (4.15):.
We start by focusing on the asymptotic claim. This is a special case of [GO11, Prop. 5.3]. Note that the authors of [GO11] use where we use , where we use , and we take . Here we have used the normalization as well as . The equidistribution hypothesis of [GO11] (i.e., the convergence of the expression involving integrals) is a consequence of [GO11, Cor. 4.14], where we note that for us.
For the first equality we note that, because is simply connected, there is exactly one orbit in each orbit inside (see the proof of [GO11, Cor. 1.9]). Thus it makes sense to identify with , and the result follows. ∎
To apply 4.15, it suffices to construct a sequence of well-rounded sets. To do this we proceed as follows, following [GO11, Pf. of Cor. 1.9]. We let be a representation of satisfying the condition that is identified with the stabilizer of some vector . Fix a compact measurable subset with boundary measure zero and positive volume. We then consider, for each positive integer , the sets
(18) |
We observe that is well-rounded with . Indeed, we can consider the subgroup , which we may observe preserves . Then taking a neighbourhood of of the form for a sufficiently small compact neighbourhood of depending on we easily see that is well-rounded. From the calculation in [BO12, Cor. 7.7] (c.f. the proof of Cor. 1.9 in [GO11]), one also sees that . We conclude from 4.15 that
Corollary 4.17.
In the above setup,
(19) |
We remark that in the argument appearing in [GO11, Pf. of Cor. 1.9] the representation is assumed to be faithful, but as we have just seen for the purpose of establishing 4.17 this is not need. Supposing now that we define
(20) |
where denotes the interior of , we also have
Corollary 4.18.
In the above setup,
(21) |
Proof.
Using the structure of the product measure, we have , where is defined as is except with the condition at the real place. We may approximate from below as a union of compact subsets , and by applying 4.15 to a sequence of sets constructed with the we learn that is asymptotically greater than for each . Since we obtain the result. ∎
5 Proofs of the General Theorems
We define a subset of points in the tensorial Hodge locus which will be of interest.
Definition 5.1.
We say a hyperplane section of of codimension is sufficiently general if all of the following conditions hold:
-
(i)
is smooth and irreducible;
-
(ii)
intersects every Hodge locus component of dimension in a reduced set of points of cardinality equal to its degree;
-
(iii)
does not intersect any (possibly tensorial) Hodge locus component of dimension smaller than ;
-
(iv)
is not contained in any component of the tensorial Hodge locus, and the algebraic monodromy group of agrees with that of ; and
-
(v)
the restriction of the period map to is quasi-finite.
Note that sufficiently general hyperplanes sections exist assuming . To explain why one can achieve (v), we note that the map factors as with algebraic (see [BBT23]), and all the components of the (tensorial) Hodge locus are pulled back from algebraic subvarieties of . In particular to achieve (v) it is enough to require that does not intersect any fibre of in a positive dimensional locus, which is true away from a closed locus in the parameter space of hyperplane sections assuming .
5.1 Proof of 1.22
We will fix a , and prove the same statement but with replaced by . This suffices, since one can always obtain (2) by summing over such inequalities for all possible and combining the constants.
Reduction to Local Point Counting: Now letting be a sufficiently general hyperplane section of codimension , we can replace with , and reduce to proving the same result for the restricted variation on the new space. Note that condition (iii) in particular implies that the Mumford-Tate groups of the Hodge structures above the points of are equal to the Mumford-Tate group of the corresponding component of . After all these changes one can take and reduce to estimating the degrees of the union of components .
We now use 2.4 to take an open cover such that each admits a definable local period map landing inside a Siegel set , with , with the set defined above matching the one in 2.4. We set . We therefore have, for each which intersects , a point such that the Hodge datum of is contained in a Hodge datum -coset equivalent to . We may moreover use 4.3 and the fact that is quasi-finite to ensure that each map is actually injective (we no longer require that is the image of some point , just that the Hodge datum of be contained in a Hodge datum -coset equivalent to ).
By shrinking the to closed subsets , we may assume that but that the sets have disjoint interiors, which we denote by . Moreover, we can choose the so that the boundary avoids the countably many points for each . We then have that
(22) |
It will therefore suffice to estimate the asymptotic size as of each of the summands on the right.
Relating to Counting Stripes: We now let be the natural orbit map . By our choice above, the Hodge structure has Hodge datum contained in a datum -coset equivalent to we denote by . Because is -coset equivalent to , there is relating the two Hodge subdatums such that is defined over . Using the isomorphism , this means that is a -vector and is exactly the stabilizer in of . Note this means that is Hodge for .
Let be the projection. We set and . Using the injectivity of we may then compute the summands in (22) by
(23) |
By construction, one also has
(24) |
Because is definable and the sets lie in a common definable family, the number of isolated points in is uniformly bounded by some constant independent of . This implies that
where we use that as well as the isomorphism . This is what we wanted to show.
5.2 Proof of 1.23
The statement is clearly local, and it suffices to take relatively compact admitting a local period map which lifts . The strongly -likely hypothesis implies that is dense in by [KU23, Thm. 3.5] (in fact even just -likely suffices here). Thus contains a point . We set . After relabelling we may assume that the Hodge datum at is contained in . We set to be the orbit map .
We set , and follow the proof given in [KU23, §4.4]. The proof given there explains that we have an open neighbourhood in such that for each in this neighbourhood the intersection has dimension at . Let us be more precise about what this neighbourhood looks like. One can start by defining as the locus of for which is non-empty. Then the argument in [KU23, §4.3] shows that the condition that is open on . The proof of [KU23, §4.4] then shows that, after removing a closed locus of smaller dimension from , the intersection germs all have pure dimension and do not lie inside any translates of period subdomains of of smaller dimension. In particular, whenever is outside of and is a Mumford-Tate domain, the intersection is a generic stripe of in , where we use the language of §4.2.1. Using the definition of the product topology we can find a product . Replacing with we may then assume that .
Now choose sufficiently general. This means in particular that, for each of the intersections , each point of has the same Mumford-Tate group as the component of in which it lies. In particular, each intersection with a Hodge datum -coset equivalent to and is a generic stripe of in . Moreover the condition that intersect is still open on (the analytic varieties and have complementary dimension inside ), and so after possibly shrinking we can ensure that for each the intersection is a generic stripe of in .
Now let be the image of in , where for the isomorphism we take the natural orbit map. Then the set is open inside the set of §4.2.1. We then consider the map
induced by the restriction of the map of §4.2.1. Combining 4.9, 4.5 and the assumption that has finite index in its normalizer, we learn that is quasi-finite with fibres of uniformly bounded size. Identifying with using the natural representation, we get that
(25) |
Using the genericity of , the sum on the right is then a lower bound for the sum in the statement. Since the symbol implies the same expression with the symbol up to a positive real constant, the result follows.
5.3 Proof of 1.24 and 1.25
5.4 Proof of 1.26
Given a Hodge structure on a lattice , a vector is Hodge for if and only if is Hodge for the Hodge structure on . (Indeed, the Mumford-Tate group of fixes if and only if it fixes acting diagonally, and is identified with the image of under the diagonal action since it is the -Zariski closure of the diagonal image of .) If comes with a polarization form , then for the induced polarization on . Likewise, . Thus to bound , we may reduce to bounding the Noether-Lefschetz locus associated to (since , as there is no square-free part). It then follows from 1.25 that
where we use that the signature of the lattices appearing in the fibres of is .
6 Applications
6.1 Specialization of the Lower Bounds
We would like to deduce 1.6 and 1.11 from 1.23 by taking the in the former two statements to be equal to the in the latter. However the statements of 1.6 and 1.11 apply to any sufficiently general hyperplane section in the sense of 5.1, whereas 1.23 just works for some choice of such a section. Let us explain why in the case of our examples, this is not an issue, by following the proof of 1.23 with . We start by observing that, in the proof of 1.23, was required to be sufficiently general intersecting some constructed in the proof. This could be chosen to be any open subset of such that always has the “expected” dimension if non-empty.
We observe that has this property outside of a finite union of tensorial Hodge loci of . This follows from the “Ax-Schanuel in families” statement in [BU24, §3.3], where the bundle is constructed from as in [BU24, §4.6]. Indeed, the varieties belong to finitely many families of algebraic subvarieties of , hence induce finitely many algebraic families of subvarieties of using the procedure outlined in [BU24, §4.6]. Then the images in of intersections of fibres of with leaves of , which in particular include all the germs of the form , are contained in the fibres of finitely many families of weakly special subvarieties of as a consequence of [BU24, Thm 3.12]. We may assume that the fibres of the all have the same algebraic monodromy groups using [BU24, Lem. 7.6] (c.f. [BU24, Lem. 7.5]). The fibres of the which contain the do not have zero period dimension (since as in both cases). It then follows from the -simplicity of the algebraic monodromy group in our two examples that each fibre of each lies in a weakly non-factor weakly special subvariety of in the sense of [KOU20, Def. 1.13]. Then, as a consequence of [KOU20, Lem. 2.5], each such factor lies in a strict special subvariety of . Thus the images of the in are contained in a union of strict special subvarieties (i.e., tensorial Hodge locus components) of . There are countably many such varieties and the images of the are algebraic, so we can take this union to be finite.
Now to prove 1.6 and 1.11, it suffices to replace with the complement in of this collection of tensorial Hodge loci. Note that by construction, any sufficiently general hyperplane section of the original also intersects this complement. After doing this, the locus appearing in the proof of 1.23 is automatically empty, and one can take to be the image of in . The statements of 1.6 and 1.11 are thus reduced to an explicit estimate of the left-hand side of (25), which we turn to in the remaining sections.
Finally, let us note that the “strongly -likely” hypothesis in 1.23 is checked in [KU23, §1.2.2] for the Noether-Lefschetz locus case, and is immediate from a simple dimension count in the split Jacobian case. The hypotheses on the normalizers is an easy consequence of 4.4 and the fact that the centralizers of the subgroups in question are finite.
6.2 The Noether-Lefschetz Locus
In this section we set . We may identify and the natural inclusion with .
We let be the natural Mumford-Tate subdomain consisting of Hodge structures with Mumford-Tate group (the subscript denotes a primitive -vector which stabilizes). The upper bound appearing in 1.3 is then a formal application of 1.25 together with the following lemma.
Lemma 6.1.
A Hodge subdatum which is conjugation-equivalent to the Hodge subdatum is -coset equivalent to the same subdatum if and only if , where is a non-zero element in the -vector space stabilized by .
Proof.
By scaling we can assume that . We also let be a point such that .
Suppose first that is -coset equivalent to . Then we may choose such that is defined over . Choose such that . Then since , necessarily , and thus . Since and are both square free, it follows that .
Conversely, suppose that . Because the subspace stabilized by is -dimensional, we have for some , and we may assume is chosen such that . We then have . Then
so it follows that and is a -vector. It follows that the coset is defined over , or that the two Hodge data are -coset equivalent. ∎
We are therefore reduced to estimating the asymptotic number of rational points in a compact subset of . We set and to match the notation of §4.2.1. We start by giving an alternative description of the quotient . First, consider the universal cover , which is an isogeny of -algebraic groups. If we consider the scalar extension we have that . The fundamental group of is and its universal cover is the Spin group ; it follows that is a -form of .
Now consider . We claim that is (geometrically) connected. Once again it suffices to scalar extend to , after which we can reduce to the same problem for and the universal covering by . Then the result follows from the fact that the construction of the Spin group is functorial in maps of quadratic spaces, which shows that the inverse image of in is identified with a copy of .
We thus have an identification . The groups and are both semisimple and simply-connected, so we may estimate the number of rational points a definable (relatively) compact subset of using the results of §4.3. In particular, letting be such a subset and applying our discussion in §4.3.2 to the pair and the natural representation , we obtain from 4.18 that
(26) |
where . It therefore suffices to compute the volumes .
6.2.1 Volume at
We recall that the variety is a closed subvariety of defined by the relation for coordinates of . To compute using the measure on we may use Gauge forms. The Gauge forms (resp for the measures on (resp. ) are given on [Sat92, pg. 149]. The expression for in particular, which is what interests us, is
(Here we interpret the division of differential forms as representing a differential form , if it exists, for which . The wedge product is computed in the algebra of differentials on .)
We write and let be the quadratic form induced by . The lattice is unimodular and indefinite. Unimodular and indefinite lattices are fully classified by the signature of the form , and whether the lattice is “even” or “odd” (see [MH73, Ch. II, Thm. 5.3]). (The lattice is even if for all , and odd otherwise.) If the lattice is odd, then can be diagonalized [MH73, Ch. II, Thm. 4.3]. If we tensor with the distinction disappears, and becomes isomorphic over to an odd unimodular indefinite lattice. In particular is diagonalizable. As we are only interested in computing the factors in (26) away from the prime , we may therefore work over and assume that is diagonal.
After diagonalizing we may write . Then
so we in fact have
To integrate this form over the -adics with we apply the main result of [Yam83], which says that
Theorem 6.2.
Let be polynomials for and write for the associated map. Let be a point such that the fibre has at least one -point at which the map is submersive. Then one has
(27) |
where is defined by
(28) |
Here is an ordered -tuple of distinct increasing integers, and is the complementary tuple (c.f. [Yam83, §2]).
We will apply this theorem with , , , , and for some fixed integer . We consider the sets and . Let for each . We have that
We now use the main theorem of [SH00] to compute .
Proposition 6.3.
We have
Proof.
This is just a specialization of the main theorem of [SH00]; since the expression appearing there is complicated, we give some details, following the notation appearing there. We thus set and . We write
in accordance with the expression for appearing in [SH00]. We have , so is the trivial group; always and always; the set and admits a unique stable partition of length . This means that the functions and are always zero. From all this it follows that , the first summation sign can be removed, and we are left to estimate an expression of the form
Now and are equal when , and when we have . Moreover is always just , so equal to , and . It follows that and . When the third summation is equal to by convention, so we obtain
Now is just a single integer satisfying , where . We thus get
Now with one has , and . In our setting , and , i.e., either or , depending on the parity of .
Thus is equal to when is even, and for odd is given by the expression
Putting everything together we obtain the result in the statement. ∎
Corollary 6.4.
We have .
Proof.
As a consequence of the corollary and the calculation preceding 6.3, one has . Taking the product over all the product is just a positive constant, so one obtains for some constant , hence the result.
6.3 Splittings of Jacobians
In this case we have , with the Siegel upper half space, and .
Lemma 6.5.
There are only finitely many -coset equivalence classes of Hodge subdata of which are isomorphic to .
Proof.
Because there are finitely many equivalence classes of Hodge subdata under real-conjugation equivalence, it suffices to fix such an equivalence class and show that it contains finitely many sub-equivalence classes under -coset equivalence. Given two such subdatum and of the Hodge datum , one knows that the subset of stabilized by the conjugation action of (resp. ) is the -dimensional subspace spanned by (resp. ) where (resp. ) is a non-trivial idempotent. A real element that conjugates the two Hodge subdata then induces an equality of algebras
In particular is a non-trivial idempotent in the algebra , necessarily equal to either or . Since (resp. ) is exactly the stabilizer in of (resp. ), it follows after applying the isomorphism that the coset is defined over . ∎
To deduce our results, we compare the degrees of isogenies associated to -factors of an abelian variety with denominators of Hodge-theoretic idempotents.
Definition 6.6.
For a weight one integral Hodge structure , we say that has a -factor of denominator if there exists a rank -Hodge-theoretic idempotent such that is primitive.
Lemma 6.7.
Suppose that is a complex abelian variety, and that is the associated weight one integral Hodge structure. Then
-
-
if has a -factor of degree dividing , then has a -factor of denominator dividing ; and
-
-
if has a -factor of denominator dividing , then has a -factor of degree dividing .
Proof.
Suppose first that has a -factor of degree dividing . Let be an integral weight one Hodge structure with an isogeny of degree such that is a decomposition of weight integral Hodge structures, and has rank . Let be the associated Hodge-theoretic idempotent. Then is a Hodge-theoretic idempotent of rank and is integral, so the denominator of divides , and by assumption.
For the second statement we start with a rank Hodge-theoretic idempotent of such that and construct an isogenous Hodge structure . As a lattice, we define . It is clear that is preserved by , and that the Hodge decomposition on induces a Hodge-decomposition on . Moreover one has , since for any . Then the inclusion is an isogeny . Its degree is the index . We then have
hence . ∎
We define
Note that our lemma shows that and .
References
- [AMVL19] Enzo Aljovin, Hossein Movasati, and Roberto Villaflor Loyola. Integral Hodge conjecture for Fermat varieties. J. Symb. Comput., 95:177–184, 2019.
- [And92] Yves André. Mumford-Tate groups of mixed Hodge structures and the theorem of the fixed part. Compositio Mathematica, 82(1):1–24, 1992.
- [Ara12] Donu Arapura. Algebraic geometry over the complex numbers. Universitext. Berlin: Springer, 2012.
- [BBKT24] Benjamin Bakker, Yohan Brunebarbe, Bruno Klingler, and Jacob Tsimerman. Definability of mixed period maps. J. Eur. Math. Soc. (JEMS), 26(6):2191–2209, 2024.
- [BBT23] Benjamin Bakker, Yohan Brunebarbe, and Jacob Tsimerman. o-minimal GAGA and a conjecture of Griffiths. Invent. Math., 232(1):163–228, 2023.
- [BDPP11] Francis E. Burstall, Neil M. Donaldson, Franz Pedit, and Ulrich Pinkall. Isothermic submanifolds of symmetric -spaces. J. Reine Angew. Math., 660:191–243, 2011.
- [BGST23] Benjamin Bakker, Thomas W. Grimm, Christian Schnell, and Jacob Tsimerman. Finiteness for self-dual classes in integral variations of Hodge structure. Épijournal de Géom. Algébr., EPIGA, Spec. Vol.:25, 2023. Id/No 1.
- [BHC61] Armand Borel and Harish-Chandra. Arithmetic subgroups of algebraic groups. Bull. Am. Math. Soc., 67:579–583, 1961.
- [BKT20] Bakker Benjamin, Bruno Klingler, and Jacob Tsimerman. Tame topology of arithmetic quotients and algebraicity of Hodge loci. American Mathematical Society, 33:917–939, 2020.
- [BKT23] B. Bakker, B. Klingler, and J. Tsimerman. Erratum to: “Tame topology of arithmetic quotients and algebraicity of Hodge loci”. J. Am. Math. Soc., 36(4):1305–1308, 2023.
- [BKU24a] Gregorio Baldi, Bruno Klingler, and Emmanuel Ullmo. Non-density of the exceptional components of the noether–lefschetz locus. International Mathematics Research Notices, 2024(21):13642–13650, 10 2024.
- [BKU24b] Gregorio Baldi, Bruno Klingler, and Emmanuel Ullmo. On the distribution of the Hodge locus. Invent. Math., 235(2):441–487, 2024.
- [BO12] Yves Benoist and Hee Oh. Effective equidistribution of -integral points on symmetric varieties. Ann. Inst. Fourier, 62(5):1889–1942, 2012.
- [Bor61] Armand Borel. Some properties of adele groups attached to algebraic groups. Bull. Am. Math. Soc., 67:583–585, 1961.
- [Bor91] A. Borel. Linear algebraic groups., volume 126 of Grad. Texts Math. New York etc.: Springer-Verlag, 2nd enlarged ed. edition, 1991.
- [BU24] Gregorio Baldi and David Urbanik. Effective atypical intersections and applications to orbit closures, 2024.
- [CDK94] Eduardo Cattani, Pierre Deligne, and Aroldo Kaplan. On the locus of Hodge Classes. Journal of the American Mathematical Society, 8, 03 1994.
- [CKS86] Eduardo Cattani, Aroldo Kaplan, and Wilfried Schmid. Degeneration of Hodge structures. Ann. Math. (2), 123:457–535, 1986.
- [ES22] Sebastian Eterović and Thomas Scanlon. Likely intersections. Preprint, arXiv:2211.10592 [math.AG] (2022), 2022.
- [Gar18] Paul Garrett. Modern analysis of automorphic forms by example. Volume 1, volume 173 of Camb. Stud. Adv. Math. Cambridge: Cambridge University Press, 2018.
- [GGK12] Mark Green, Phillip Griffiths, and Matt Kerr. Mumford-Tate Groups and Domains: Their Geometry and Arithmetic (AM-183). Princeton University Press, 2012.
- [GO11] Alex Gorodnik and Hee Oh. Rational points on homogeneous varieties and equidistribution of adelic periods (with appendix by Mikhail Borovoi). Geom. Funct. Anal., 21(2):319–392, 2011.
- [He13] Hongyu He. Certain unitary langlands-vogan parameters for special orthogonal groups i, 2013.
- [hh] Jim Humphreys (https://mathoverflow.net/users/4231/jim humphreys). The normalizer of a reductive subgroup. MathOverflow. URL:https://mathoverflow.net/q/27349 (version: 2016-06-11).
- [Kas85] Masaki Kashiwara. The asymptotic behavior of a variation of polarized Hodge structure. Publ. Res. Inst. Math. Sci., 21:853–875, 1985.
- [KOU20] Bruno Klingler, Anna Otwinowska, and David Urbanik. On the fields of definition of Hodge loci. To appear in Ann. Sci. de l’École Nor. Sup. arXiv:2010.03359, October 2020.
- [KU23] Nazim Khelifa and David Urbanik. Existence and density of typical Hodge loci. Preprint, arXiv:2303.16179 [math.AG] (2023), 2023.
- [MH73] John W. Milnor and Dale H. Husemoller. Symmetric bilinear forms, volume 73 of Ergeb. Math. Grenzgeb. Springer-Verlag, Berlin, 1973.
- [Mos61] G. D. Mostow. On maximal subgroups of real Lie groups. Ann. Math. (2), 74:503–517, 1961.
- [Orr18] Martin Orr. Height bounds and the Siegel property. Algebra Number Theory, 12(2):455–478, 2018.
- [Sat92] Fumihiro Sato. Siegel’s main theorem of homogeneous spaces. Comment. Math. Univ. St. Pauli, 41(2):141–167, 1992.
- [Sch73] Wilfried Schmid. Variation of Hodge Structure: The Singularities of the Period Mapping. Inventiones mathematicae, 22:211–320, 1973.
- [SH00] Fumihiro Sato and Yumiko Hironaka. Local densities of representations of quadratic forms over -adic integers (the non-dyadic case). J. Number Theory, 83(1):106–136, 2000.
- [Tay20] Salim Tayou. On the equidistribution of some Hodge loci. J. Reine Angew. Math., 762:167–194, 2020.
- [TT23] Salim Tayou and Nicolas Tholozan. Equidistribution of Hodge loci. II. Compos. Math., 159(1):1–52, 2023.
- [vdD98] Lou van den Dries. Tame topology and o-minimal structures, volume 248 of Lond. Math. Soc. Lect. Note Ser. Cambridge: Cambridge University Press, 1998.
- [Voi10] Claire Voisin. Hodge loci. Handbook of moduli, 3:507–546, 2010.
- [Wei59] André Weil. Adèles and algebraic groups. Sém. Bourbaki 11(1958/59), Exp. No. 186, 9 p. (1959)., 1959.
- [Yam83] Tadashi Yamazaki. Integrals defining singular series. Mem. Fac. Sci., Kyushu Univ., Ser. A, 37:113–128, 1983.