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

Degrees of Hodge Loci

David Urbanik
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 3\mathbb{P}^{3}, 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 \mathbb{C} unless otherwise stated. The typical exception will be algebraic groups, which will almost always be \mathbb{Q}-algebraic. Moreover all Mumford-Tate groups (defined below), will always be understood to be special Mumford-Tate groups. For two real sequences ana_{n} and bnb_{n}, we write anbna_{n}\sim b_{n} if an/bnn1a_{n}/b_{n}\xrightarrow{n\to\infty}1 and anbna_{n}\asymp b_{n} if an/bna_{n}/b_{n} is bounded away from both 0 and ±\pm\infty as nn\to\infty. Finally we write anbna_{n}\lesssim b_{n} if for any ε>0\varepsilon>0 there exists NN such that bn(1ε)anb_{n}\geq(1-\varepsilon)a_{n} for all nNn\geq N.

1 Introduction

Given a smooth projective family of varieties f:XSf:X\to S, the cohomology groups 𝕍s:=Hi(Xs,)/tor.\mathbb{V}_{s}:=H^{i}(X_{s},\mathbb{Z})/\textrm{tor.} in degree ii carry a natural Hodge structure. Via the Hodge conjecture, rational vectors inside 𝕍s,\mathbb{V}_{s,\mathbb{Q}} and its tensor powers are predicted to characterize, up to cohomological equivalence, the algebraic cycles associated to XsX_{s} and its self-products. A question of much current and classical interest is to understand for which ss the fibre XsX_{s} carries “more than the expected number” of algebraic cycles, i.e., more than at a very general point sS()s\in S(\mathbb{C}).

The loci in SS where the fibres XsX_{s} 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 SS. 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 SS 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 Q:𝕍𝕍Q:\mathbb{V}\otimes\mathbb{V}\to\mathbb{Z}, where 𝕍=Rif/tor.\mathbb{V}=R^{i}f_{*}\mathbb{Z}/\textrm{tor}. is the natural local system interpolating the fibres 𝕍s\mathbb{V}_{s}. Then one can consider the reduced Hodge locus ZnSZ_{n}\subset S defined by primitive integral vectors vv with Q(v,v)=nQ(v,v)=n for some integer nn. The locus ZnZ_{n} is algebraic, so after choosing a quasi-projective embedding of SS it makes sense to ask for its degree; more generally, it makes sense to ask

Question: What are the asymptotics of degZn\deg Z_{n} as nn\to\infty?

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 UdNU_{d}\subset\mathbb{P}^{N} be the parameter space of smooth degree dd hypersurfaces in 3\mathbb{P}^{3} for some integer d5d\geq 5, where UdU_{d} is the complement in N\mathbb{P}^{N} of the discriminant locus Δ\Delta. Write f:XS:=Udf:X\to S:=U_{d} for the universal family. Using the natural map i:X3i:X\to\mathbb{P}^{3} we obtain a family :=i𝒪(1)\mathcal{L}:=i^{*}\mathcal{O}(1) of line bundles in the fibres of ff. We define the primitive subsystem

𝕍:=ker[R2f()c1()R4f]\mathbb{V}:=\ker\left[R^{2}f_{*}\mathbb{Z}\xrightarrow{(-)\cup c^{1}(\mathcal{L})}R^{4}f_{*}\mathbb{Z}\right]

where c1()c^{1}(\mathcal{L}) is the global section of R2fR^{2}f_{*}\mathbb{Z} coming from the relative Chern class of \mathcal{L}. The local system 𝕍\mathbb{V} is torsion-free, and cup product induces a non-degenerate symmetric pairing Q:𝕍𝕍Q:\mathbb{V}\otimes\mathbb{V}\to\mathbb{Z} which makes, as a consequence of Poincaré duality, each fibre (𝕍s,Qs)(\mathbb{V}_{s},Q_{s}) into a unimodular lattice. We write (r,s)(r,s) for the signature of this lattice, which is related to the Hodge numbers (h2,0,h1,1,h0,2)(h^{2,0},h^{1,1},h^{0,2}) of 𝕍\mathbb{V} by r=2h2,0=2h0,2r=2h^{2,0}=2h^{0,2} and s=h1,1s=h^{1,1}.

The Noether-Lefschetz locus of 𝕍\mathbb{V} is constructed as follows. For each point sS()s\in S(\mathbb{C}) and each integral vector v𝕍sv\in\mathbb{V}_{s}, there is a locus Z(v)SZ(v)\subset S consisting of all sS()s^{\prime}\in S(\mathbb{C}) where some flat translate of vv in 𝕍s\mathbb{V}_{s^{\prime}} is Hodge. By [CDK94] the locus Z(v)Z(v) is algebraic, and we will choose to equip it with its underlying reduced structure. Moreover [CDK94] also shows that for \ell\in\mathbb{Z} the locus Q(v,v)=Z(v)\bigcup_{Q(v,v)=\ell}Z(v) is algebraic.

Definition 1.1.

We say a vector vv in a \mathbb{Z}-lattice LL is primitive if vv has non-zero image in L𝔽pL_{\mathbb{F}_{p}} for each pp.

Working with the union over all vv such that Q(v,v)=Q(v,v)=\ell results in overcounting: if λ\lambda\in\mathbb{Z} then we have Z(v)=Z(λv)Z(v)=Z(\lambda v). To remedy this, we define a modified scalar-invariant quantity better suited for counting Hodge loci which agrees with Q(v,v)Q(v,v) whenever vv is primitive.

Definition 1.2.

For vv a non-zero vector in some fibre of 𝕍\mathbb{V}_{\mathbb{Q}}, we define a(v)×a(v)\in\mathbb{Q}^{\times} to be the unique positive rational scalar for which a(v)va(v)v is a primitive integral vector in the corresponding fibre of 𝕍\mathbb{V}. We moreover define

Q¯:𝕍{0},Q¯(v)=a(v)2Q(v,v)=Q(a(v)v,a(v)v)\overline{Q}:\mathbb{V}_{\mathbb{Q}}\setminus\{0\}\to\mathbb{Z},\hskip 20.00003pt\overline{Q}(v)=a(v)^{2}Q(v,v)=Q(a(v)v,a(v)v)

which is a map of flat bundles over SS. We write u(v)u(v) and ν(v)\nu(v) for the unique integers such that Q¯(v)=u(v)ν(v)2\overline{Q}(v)=u(v)\nu(v)^{2} with ν(v)\nu(v) positive and u(v)u(v) square-free.

It is clear that Q¯(λv)=Q¯(v)\overline{Q}(\lambda v)=\overline{Q}(v) for each λ×\lambda\in\mathbb{Q}^{\times}, and that u(v)=Q¯(v)=Q(v,v)u(v)=\overline{Q}(v)=Q(v,v) modulo scaling by (×)2(\mathbb{Q}^{\times})^{2}. We may then define

NLν,u\displaystyle\operatorname{NL}_{\nu,u} :={sS(): there exists a Hodge vector v𝕍s with u(v)=u mod (×)2,ν(v)|ν}\displaystyle:=\left\{s\in S(\mathbb{C}):\begin{array}[]{c}\textrm{ there exists a Hodge vector }v\in\mathbb{V}_{s}\\ \textrm{ with }u(v)=u\textrm{ mod }(\mathbb{Q}^{\times})^{2},\,\,\nu(v)|\nu\end{array}\right\}
={sS(): there exists a Hodge vector v𝕍s with Q(v,v)=u mod (×)2,ν(v)|ν}.\displaystyle=\left\{s\in S(\mathbb{C}):\begin{array}[]{c}\textrm{ there exists a Hodge vector }v\in\mathbb{V}_{s}\\ \textrm{ with }Q(v,v)=u\textrm{ mod }(\mathbb{Q}^{\times})^{2},\,\,\nu(v)|\nu\end{array}\right\}.

For a closed subvariety of UdU_{d} we define its degree to be the degree of its closure in N\mathbb{P}^{N}. We set S=UdS=U_{d} and m=r+sm=r+s.

Theorem 1.3 (Upper bound).

For each fixed square-free u>0u\in\mathbb{Z}_{>0} we have

deg(NLν,u)cuνm+2h0,22\deg(\operatorname{NL}_{\nu,u})\lesssim c_{u}\nu^{m+2h^{0,2}-2}

as ν\nu\to\infty, where cu>0c_{u}>0 is a real constant depending on uu.

Remark 1.4.

The quantity mm is b21b_{2}-1, where b2b_{2} is the second Betti number of a fibre of ff. It is known that b2=d34d2+6d2b_{2}=d^{3}-4d^{2}+6d-2, and by [Ara12, Ch 17.3] one has that h0,2=(d1)(d2)(d3)/6h^{0,2}=(d-1)(d-2)(d-3)/6. Combining these estimates one easily expresses m+2h0,22m+2h^{0,2}-2 as a polynomial in dd.

Definition 1.5.

The tensorial Hodge locus of 𝕍\mathbb{V} is the collection of all Hodge loci (i.e., Noether-Lefschetz loci) associated to the variations 𝕍k\mathbb{V}^{\otimes k} for all k0k\geq 0.

We say a subvariety of SS 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 TS=UdT\subset S=U_{d}, with period dimension h2,0+1\geq h^{2,0}+1;

  • -

    a relatively compact open neighbourhood BTB\subset T; and

  • -

    a sufficiently general hyperplane section LBL\cap B of BB with codimSL=dimTh2,0\operatorname{codim}_{S}L=\dim T-h^{2,0}.

Then for each fixed square-free u>0u\in\mathbb{Z}_{>0}, and for ν\nu with ν>02\nu\in\mathbb{Z}_{>0}\setminus 2\mathbb{Z}, one has

cuνm2\displaystyle c_{u}\nu^{m-2} |NLν,uLB|\displaystyle\lesssim|\operatorname{NL}_{\nu,u}\cap L\cap B|

as ν\nu\to\infty, where cu>0c_{u}\in\mathbb{R}_{>0} is a constant depending only on u,Bu,B and LL.

In the special case where S=TS=T, one has

Theorem 1.7 (Lower Bound).

For each fixed square-free u>0u\in\mathbb{Z}_{>0}, and for ν\nu with ν>02\nu\in\mathbb{Z}_{>0}\setminus 2\mathbb{Z}, one has

cuνm2\displaystyle c_{u}\nu^{m-2} degNLν,u\displaystyle\lesssim\deg\operatorname{NL}_{\nu,u}

as ν\nu\to\infty, where cu>0c_{u}\in\mathbb{R}_{>0} 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 ν\nu. We give the precise calculation in §6.2.1. With the optimal formula appearing there, one can replace the symbol \lesssim with \asymp as long as BB does not intersect a certain finite union of tensorial Hodge loci of TT.

It is a consequence of the theory developed in [BKU24b] (c.f. [BKU24a, Thm. 6]) that each component of NLν,uT\operatorname{NL}_{\nu,u}\cap T has codimension exactly h2,0h^{2,0} in TT outside of a closed algebraic locus in TT, so away from this locus the “sufficiently general” hypothesis in 1.6 ensures that the cardinality |NLν,uLB||\operatorname{NL}_{\nu,u}\cap L\cap B| is finite. The left-hand side of the inequality in 1.6 diverges as a function of ν\nu, so one obtains infinitely many Hodge loci defined by vectors vv with u(v)=uu(v)=u in any neighbourhood of TT. In fact, away from the atypical Hodge locus, one can obtain exact asymptotics in terms of ν\nu for the size of the Hodge locus in the prescribed region.

The period dimension hypothesis of h2,0+1\geq h^{2,0}+1 on the other hand implies that each component of NLν,uT\operatorname{NL}_{\nu,u}\cap T has positive period dimension; this will be used to verify that a general component of NLν,uT\operatorname{NL}_{\nu,u}\cap T carries a unique global 11-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 f:XS:=gf:X\to S:=\mathcal{M}_{g} of genus gg 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 g¯\overline{\mathcal{M}_{g}} of (the coarse space of) g\mathcal{M}_{g} and define the degree of an algebraic locus in g\mathcal{M}_{g} to be the degree of its closure in g¯\overline{\mathcal{M}_{g}}.

Definition 1.9.

For a complex abelian variety AA, we say that AA has a kk-factor of degree ν\nu if there is an isogeny φ:AA×A′′\varphi:A\to A^{\prime}\times A^{\prime\prime} such that degφ=ν\deg\varphi=\nu and dimA=k\dim A^{\prime}=k.

For a curve CC, we write J(C)J(C) for the associated Jaocbian variety. As an analogue of the Noether-Lefschetz loci considered above, we consider

SPk,ν:={sS():J(Xs) has an k-factor of degree dividing ν}.\displaystyle\textrm{SP}_{k,\nu}:=\{s\in S(\mathbb{C}):J(X_{s})\textrm{ has an }k\textrm{-factor of degree dividing }\nu\}.
Theorem 1.10 (Upper Bound).

For each integer kk such that 1kg/21\leq k\leq g/2 we have

deg(SPk,ν)ckν8k(gk)\deg(\textrm{SP}_{k,\nu})\lesssim c_{k}\nu^{8k(g-k)}

as ν\nu\to\infty, where ck>0c_{k}>0 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 𝕍=R1f\mathbb{V}=R^{1}f_{*}\mathbb{Z}. We consider the pair (Spg,g)(\operatorname{Sp}_{g},\mathbb{H}_{g}) consisting of the symplectic group of a 2g2g-dimensional integral symplectic space (V,Q)(V,Q), and the Siegel upper half space, respectively. This pair can be viewed as the “generic Hodge datum” of the variation of Hodge structure 𝕍\mathbb{V} (we review this notion below). The complex manifold g\mathbb{H}_{g} has natural closed submanifolds k×gk\mathbb{H}_{k}\times\mathbb{H}_{g-k} whose images in 𝒜g()=Spg()\g\mathcal{A}_{g}(\mathbb{C})=\operatorname{Sp}_{g}(\mathbb{Z})\backslash\mathbb{H}_{g} consist exactly of those closed algebraic loci parameterizing non-simple principally polarized abelian varieties. Each such submanifold corresponds to a decomposition VV1V2V_{\mathbb{Q}}\simeq V_{1}\oplus V_{2}, compatible with the symplectic form, and the associated symplectic idempotents e1e_{1} and e2e_{2} belong to only finitely many Spg()\operatorname{Sp}_{g}(\mathbb{R}) orbits as the submanifolds k×gkg\mathbb{H}_{k}\times\mathbb{H}_{g-k}\hookrightarrow\mathbb{H}_{g} vary. We let OkO_{k} be the Spg\operatorname{Sp}_{g}-orbit of such an idempotent eke_{k} whose image has dimension 2k2k. We view OkO_{k} as a \mathbb{Q}-algebraic subvariety of the affine space associated to the free \mathbb{Z}-module End(V)\textrm{End}(V).

Theorem 1.11 (Lower Bound).

Let 1kg11\leq k\leq g-1 and g2g\geq 2. Fix the data of:

  • -

    a locally closed irreducible Hodge generic subvariety TST\subset S of dimension at least k(gk)+1k(g-k)+1;

  • -

    a relatively compact open neighbourhood BTB\subset T; and

  • -

    a sufficiently general hyperplane section LBL\cap B of BB with codimSL=dimTk(gk)\operatorname{codim}_{S}L=\dim T-k(g-k).

Then with O=OkO=O_{k} one has

ckpμp,O(O(p)1νEnd(V)p)|SPk,ν2gLB|c_{k}\prod_{p}\mu_{p,O}\left(O(\mathbb{Q}_{p})\cap\frac{1}{\nu}\textrm{End}(V)_{\mathbb{Z}_{p}}\right)\lesssim|\textrm{SP}_{k,\nu^{2g}}\cap L\cap B| (1)

as ν\nu\to\infty, where

  • -

    ckc_{k} is a positive real constant depending only on k,Bk,B and LL; and

  • -

    μp,O\mu_{p,O} is the unique Spg(p)\operatorname{Sp}_{g}(\mathbb{Q}_{p})-invariant measure on O(p)O(\mathbb{Q}_{p}) compatible with the Tamagawa measures on Spg\operatorname{Sp}_{g} and the stabilizer in Spg\operatorname{Sp}_{g} of eke_{k}.

Remark 1.12.

In the proof of 1.11, it will be more natural to work with a certain locus SPk,ν\textrm{SP}^{\prime}_{k,\nu} defined Hodge-theoretically. We have SPk,νSPk,ν2g\textrm{SP}^{\prime}_{k,\nu}\subset\textrm{SP}_{k,\nu^{2g}}. Moreover if BB does not intersect a certain finite union of tensorial Hodge loci of TT, then the statement of 1.11 holds with SPk,ν2g\textrm{SP}_{k,\nu^{2g}} replaced by SPk,ν\textrm{SP}^{\prime}_{k,\nu} and the symbol \lesssim replaced by \asymp.

Remark 1.13.

One can show that the left-hand side of (1) diverges as ν\nu\to\infty.

Let us take some time to unpack the statement. The group Spg\operatorname{Sp}_{g} is semisimple, simply-connected, and unimodular. By [Wei59, pg.26] one also knows that the locally compact topological group Spg(𝔸)\operatorname{Sp}_{g}(\mathbb{A}) is unimodular, where we write 𝔸\mathbb{A} for the adele ring. The semisimplicity and simply-connected properties imply that Spg()\operatorname{Sp}_{g}(\mathbb{Q}) is a lattice in 𝐇S(𝔸)\mathbf{H}_{S}(\mathbb{A}) [Bor61, Thm. 1] [BHC61], and this lets us put a unique Haar measure μ\mu, the Tamagawa measure, on Spg(𝔸)\operatorname{Sp}_{g}(\mathbb{A}) such that Spg(𝔸)/Spg()\operatorname{Sp}_{g}(\mathbb{A})/\operatorname{Sp}_{g}(\mathbb{Q}) has unit volume. The measure μ\mu splits as a product μ×pμp\mu_{\infty}\times\prod_{p}\mu_{p} of measures defined at each place of \mathbb{Q}.

The stabilizer in Spg\operatorname{Sp}_{g} of a Hodge-theoretic idempotent eke_{k} is abstractly isomorphic to a product H=Spk×SpgkH=\textrm{Sp}_{k}\times\textrm{Sp}_{g-k} of symplectic groups, and the same reasoning applies, mutatis mutandis, to HH. Moreover by [GO11, A.1.2, A.5.1] the space (Spg/H)(𝔸)(\operatorname{Sp}_{g}/H)(\mathbb{A}) consists of finitely many Spg(𝔸)\operatorname{Sp}_{g}(\mathbb{A}) orbits; let us fix such an orbit (Spg/H)(𝔸)(\operatorname{Sp}_{g}/H)(\mathbb{A})^{\circ}. Using [Gar18, Thm. 5.2.1, 5.2.2] there is a unique invariant measure on this quotient determined by the measures μ\mu on Spg(𝔸)\operatorname{Sp}_{g}(\mathbb{A}) and H(𝔸)H(\mathbb{A}) which again splits as a product of measures associated to each place of \mathbb{Q}. Using the natural identification Spg/HO=Ok\operatorname{Sp}_{g}/H\simeq O=O_{k} this induces in a natural way a measure on a component O(𝔸)O(𝔸)O(\mathbb{A})^{\circ}\subset O(\mathbb{A}) and on O(p)O(\mathbb{Q}_{p}) for each pp (note here that O(p)O(\mathbb{Q}_{p}) consists of a single Spg(p)\operatorname{Sp}_{g}(\mathbb{Q}_{p}) 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 Spg(𝔸)\operatorname{Sp}_{g}(\mathbb{A}) can also be described using a Spg\operatorname{Sp}_{g}-invariant \mathbb{Q}-algebraic form ω\omega on Spg\operatorname{Sp}_{g} of top degree, in such a way so that the measures at each place of \mathbb{Q} are obtained by integrating against ω\omega. Such a form is called a gauge form. It turns out that the measures associated to HH and the quotient Spg/H\operatorname{Sp}_{g}/H 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 pp-adic manifolds O(p)O(\mathbb{Q}_{p}).

In the Noether-Lefschetz locus case, essentially the same reasoning applies. In that case, the relevant group is SO(r,s)\operatorname{SO}(r,s) instead of Spg\operatorname{Sp}_{g}. In particular, up to the scaling action of ×\mathbb{Q}^{\times}, the orbits of the group SO(r,s)()\operatorname{SO}(r,s)(\mathbb{R}) on fibres of 𝕍\mathbb{V} are actually classified by the invariant uu, with two rational vectors v,vv,v^{\prime} lying in the same ×SO(r,s)()\mathbb{Q}^{\times}\operatorname{SO}(r,s)(\mathbb{R})-orbit if and only if u(v)=u(v)u(v)=u(v^{\prime}). (This follows from the fact that u(v)=Q(v,v)=Q¯(v)u(v)=Q(v,v)=\overline{Q}(v) modulo (×)2(\mathbb{Q}^{\times})^{2}, and then by using that u(v)u(v) is the unique integral square-free representative of the coset u(v)(×)2u(v)(\mathbb{Q}^{\times})^{2}.) 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 SO(r,s)\operatorname{SO}(r,s). The difference being that, in this case, we have taken the time to actually compute the relevant pp-adic volume integrals, and the factor at pp gives a contribution lower bounded by p(m2)ep(ν)p^{(m-2)e_{p}(\nu)}, where ep(ν)e_{p}(\nu) is the exponent of pp in the prime factorization of ν\nu.

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 𝕍\mathbb{V} with polarization QQ on a smooth quasi-projective complex variety SS. This will require some additional setup. We will simplify the situation by studying the asymptotics only of Hodge loci associated to vectors of 𝕍\mathbb{V}; the case of loci associated to general Hodge tensors reduces to this by replacing 𝕍\mathbb{V} with some tensor construction. We may also assume the weight of 𝕍\mathbb{V} is zero, Tate-twisting if necessary. In particular we have a Hodge filtration FF^{\bullet} on :=𝕍𝒪San\mathcal{H}:=\mathbb{V}\otimes_{\mathbb{Z}}\mathcal{O}_{S^{\textrm{an}}} and the space of Hodge vectors above ss is the subspace Fs0𝕍,sF^{0}_{s}\cap\mathbb{V}_{\mathbb{Q},s}. Given a vector vv in some fibre of 𝕍\mathbb{V}_{\mathbb{Q}}, we write Z(v)Z(v) for the corresponding (reduced) Hodge locus. We define Q¯\overline{Q}, uu and ν\nu as in 1.2.

For each point sS()s\in S(\mathbb{C}) we write hsh_{s} for the Hodge structure on 𝕍s\mathbb{V}_{s}, which we view as a map hs:𝕌Aut(𝕍s,Qs)h_{s}:\mathbb{U}\to\textrm{Aut}(\mathbb{V}_{s},Q_{s})_{\mathbb{R}}. Here 𝕌\mathbb{U} is the “circle group”, a real-algebraic subgroup of the Deligne torus 𝕊=Res/𝔾m,\mathbb{S}=\textrm{Res}_{\mathbb{C}/\mathbb{R}}\mathbb{G}_{m,\mathbb{C}} constructed as in [GGK12, I.A]. The Mumford-Tate group 𝐆s\mathbf{G}_{s} at ss is the \mathbb{Q}-Zariski closure of hs(𝕌)h_{s}(\mathbb{U}), and the Mumford-Tate domain DsD_{s} at ss is the 𝐆s()\mathbf{G}_{s}(\mathbb{R})-conjugacy class of the morphism hs:𝕌𝐆s,h_{s}:\mathbb{U}\to\mathbf{G}_{s,\mathbb{R}}. We call the pair (𝐆s,Ds)(\mathbf{G}_{s},D_{s}) the Hodge datum at ss. The generic Hodge datum (𝐆S,DS)(\mathbf{G}_{S},D_{S}) is the is the abstract Hodge datum which is isomorphic to (𝐆s,Ds)(\mathbf{G}_{s},D_{s}) at a point sS()s\in S(\mathbb{C}) outside the Hodge locus of all the tensor powers 𝕍k\mathbb{V}^{\otimes k} for k1k\geq 1. The flat structure of the local system locally identifies the realizations (𝐆S,s,DS,s)(\mathbf{G}_{S,s},D_{S,s}) of the generic Hodge datum and allows one to view the pair (𝐆S,DS)(\mathbf{G}_{S},D_{S}) as an abstract object with a realization in each fibre of 𝕍\mathbb{V}.

Definition 1.14.

For a local system 𝕃\mathbb{L} on a smooth complex algebraic variety SS, the algebraic monodromy group 𝐇𝕃,S,s\mathbf{H}_{\mathbb{L},S,s} is the identity component of the Zariski closure of the map

π1(S,s)GL(𝕃s)\pi_{1}(S,s)\to\operatorname{GL}(\mathbb{L}_{s})

where sS()s\in S(\mathbb{C}) is a point. Usually either some base point ss, the local system 𝕃\mathbb{L}, or both, are understood, in which case we write 𝐇𝕃,S,𝐇S,s\mathbf{H}_{\mathbb{L},S},\mathbf{H}_{S,s} or 𝐇S\mathbf{H}_{S}, as appropriate.

Write 𝐇S\mathbf{H}_{S} for the algebraic monodromy group of 𝕍\mathbb{V}. By [And92, §5], 𝐇S\mathbf{H}_{S} is a normal subgroup of the derived subgroup 𝐃S\mathbf{D}_{S} of 𝐆S\mathbf{G}_{S}. We note that, by [And92, Lem. 2], the group 𝐆S\mathbf{G}_{S} is always at least reductive, which implies that 𝐇S\mathbf{H}_{S} is always semisimple.

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 SS; 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 SS. 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 SS with a finite cover g:SSg:S^{\prime}\to S and 𝕍\mathbb{V} with 𝕍=g𝕍\mathbb{V}^{\prime}=g^{*}\mathbb{V}, we will freely make such changes.

Associated Period Maps:

We may replace SS with a finite étale covering such that 𝕍\mathbb{V} induces a period map φ:SΓ\DS\varphi:S\to\Gamma\backslash D_{S}, where Γ𝐆S()\Gamma\subset\mathbf{G}_{S}(\mathbb{Q}) 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 φ=φ1××φn\varphi=\varphi_{1}\times\cdots\times\varphi_{n}, such that

  • (i)

    the factorization of φ\varphi is induced by a factorization Γ\DS=Γ1\D1××Γn\Dn\Gamma\backslash D_{S}=\Gamma_{1}\backslash D_{1}\times\cdots\times\Gamma_{n}\backslash D_{n} with Γ=Γ1××Γn\Gamma=\Gamma_{1}\times\cdots\times\Gamma_{n}; and

  • (ii)

    the factorizations in (i) are induced by the almost-direct product factorization 𝐃S=𝐇1𝐇n\mathbf{D}_{S}=\mathbf{H}_{1}\cdots\mathbf{H}_{n}, where we note that the center of 𝐆S\mathbf{G}_{S} acts trivially on DS=D1××DnD_{S}=D_{1}\times\cdots\times D_{n}.

Given a subset I{1,,n}I\subset\{1,\ldots,n\}, we write pI:Γ\DSΓI\DIp_{I}:\Gamma\backslash D_{S}\to\Gamma_{I}\backslash D_{I} where ΓI=iIΓi\Gamma_{I}=\prod_{i\in I}\Gamma_{i} and DI=iIDiD_{I}=\prod_{i\in I}D_{i}. The map φI\varphi_{I} is defined similarly.

Definition 1.15.

For a subvariety TST\subset S, its period dimension is the dimension of φ(T)\varphi(T).

Likeliness:

Given the generic Hodge datum (𝐆S,DS)(\mathbf{G}_{S},D_{S}), a Hodge sub-datum is a pair (𝐌,DM)(\mathbf{M},D_{M}) consisting of a subgroup 𝐌𝐆S\mathbf{M}\subset\mathbf{G}_{S} which is the Mumford-Tate group of some Hodge structure hDSh\in D_{S} and the associated orbit DM=𝐌()hD_{M}=\mathbf{M}(\mathbb{R})\cdot h under the conjugation action. The Hodge datum of a point hDSh\in D_{S} is the pair (𝐌h,𝐌h()h)(\mathbf{M}_{h},\mathbf{M}_{h}(\mathbb{R})\cdot h), where 𝐌h\mathbf{M}_{h} is the Mumford-Tate group of hh.

Definition 1.16.

We say the Hodge sub-datum (𝐌,DM)(𝐆S,DS)(\mathbf{M},D_{M})\subset(\mathbf{G}_{S},D_{S}) is 𝕍\mathbb{V}-likely if for every non-empty set I{1,,n}I\subset\{1,\ldots,n\} the inequality dimπI(DM)+dimφI(San)dimDI0\dim\pi_{I}(D_{M})+\dim\varphi_{I}(S^{\textrm{an}})-\dim D_{I}\geq 0 holds, and we say it is strongly 𝕍\mathbb{V}-likely if these inequalities are strict.

Subdata of Definition:

A component ZSZ\subset S of the (tensorial) Hodge locus of SS can be defined by many different Hodge subdata. This means the following:

Definition 1.17.

We say a Hodge subdatum (𝐌,DM)(𝐆S,DS)(\mathbf{M},D_{M})\subset(\mathbf{G}_{S},D_{S}) defines a closed algebraic locus ZSZ\subset S if ZZ is an irreducible component of φ1(π(DM))\varphi^{-1}(\pi(D_{M})).

Note that 𝐌\mathbf{M} need not be the Mumford-Tate group of ZZ in (1.17). A Hodge locus ZZ is always defined by a Hodge subdatum (𝐆Z,DZ)(𝐆S,DS)(\mathbf{G}_{Z},D_{Z})\subset(\mathbf{G}_{S},D_{S}) which arises from the pair (𝐆s,Ds)(\mathbf{G}_{s},D_{s}) at a very general point sZ()s\in Z(\mathbb{C}), but it is often useful to view ZZ as defined by a potentially larger subdatum.

1.2.3 Equivalence Classes of Hodge Subdata

The group 𝐆S()\mathbf{G}_{S}(\mathbb{R}) acts on Hodge subdata (𝐌,DM)(𝐆S,DS)(\mathbf{M},D_{M})\subset(\mathbf{G}_{S},D_{S}) by conjugation, with (𝐌,DM)(\mathbf{M},D_{M}) being equivalent to (𝐌,DM)(\mathbf{M}^{\prime},D_{M^{\prime}}) if there is g𝐆S()g\in\mathbf{G}_{S}(\mathbb{R}) such that 𝐌=g𝐌g1\mathbf{M}^{\prime}=g\mathbf{M}g^{-1} and DM=gDMD_{M^{\prime}}=gD_{M}. 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 (𝐌,DM),(𝐌,DM)(𝐆S,DS)(\mathbf{M},D_{M}),(\mathbf{M}^{\prime},D_{M^{\prime}})\subset(\mathbf{G}_{S},D_{S}) are \mathbb{Q}-coset equivalent if there exists g𝐆S()g\in\mathbf{G}_{S}(\mathbb{R}) such that (g𝐌g1,gDM)=(𝐌,DM)(g\mathbf{M}^{\prime}g^{-1},gD_{M^{\prime}})=(\mathbf{M},D_{M}), and the coset g𝐌g\mathbf{M}^{\prime} defines a rational point in (𝐆S/𝐌)()(\mathbf{G}_{S}/\mathbf{M}^{\prime})(\mathbb{Q}).

Lemma 1.19.

\mathbb{Q}-coset equivalence is an equivalence relation on Hodge subdata.

Proof.

We observe that the notion of being \mathbb{Q}-coset equivalent is symmetric: if g𝐌g\mathbf{M}^{\prime} defines a rational point in 𝐆S/𝐌\mathbf{G}_{S}/\mathbf{M}^{\prime}, then g𝐌=𝐌gg\mathbf{M}^{\prime}=\mathbf{M}g is a \mathbb{Q}-algebraic subvariety of 𝐆S\mathbf{G}_{S}. Taking inverses, this implies that g1𝐌g^{-1}\mathbf{M} is a \mathbb{Q}-algebraic subvariety of 𝐆S\mathbf{G}_{S}, and hence corresponds to a point of (𝐆S/𝐌)()(\mathbf{G}_{S}/\mathbf{M})(\mathbb{Q}).

For transitivity, we consider three tuples (𝐌,DM),(𝐌,DM)(\mathbf{M},D_{M}),(\mathbf{M}^{\prime},D_{M^{\prime}}) and (𝐌′′,DM′′)(\mathbf{M}^{\prime\prime},D_{M^{\prime\prime}}) related by \mathbb{Q}-cosets g𝐌g\mathbf{M} and g𝐌g^{\prime}\mathbf{M}^{\prime}. Then for any automorphism σAut(/)\sigma\in\textrm{Aut}(\mathbb{C}/\mathbb{Q}) we have

(gg𝐌)σ\displaystyle(g^{\prime}g\mathbf{M})^{\sigma} =gσg𝐌\displaystyle=g^{\prime\sigma}g\mathbf{M}
=gσ𝐌g\displaystyle=g^{\prime\sigma}\mathbf{M}^{\prime}g
=g𝐌g=gg𝐌.\displaystyle=g^{\prime}\mathbf{M}^{\prime}g=g^{\prime}g\mathbf{M}.

A given real conjugacy class of Hodge subdata, when further partitioned by \mathbb{Q}-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 uu, 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 (𝐌,DM)(𝐆S,DS)(\mathbf{M},D_{M})\subset(\mathbf{G}_{S},D_{S}) be a Hodge subdatum, let g𝐆S()g\in\mathbf{G}_{S}(\mathbb{R}), and suppose that g𝐌(𝐆S/𝐌)()g\mathbf{M}\in(\mathbf{G}_{S}/\mathbf{M})(\mathbb{Q}). Then (g𝐌g1,gDM)(g\mathbf{M}g^{-1},gD_{M}) is a Hodge subdatum of (𝐆S,DS)(\mathbf{G}_{S},D_{S}).

Proof.

See 4.7. ∎

In many situations in practice, this means that counting \mathbb{Q}-coset equivalent Hodge subdata is the same as counting rational points in 𝐆S/𝐌\mathbf{G}_{S}/\mathbf{M}.

1.2.4 From Hodge Loci to Point Counting

We are now ready to describe our main technical results.

Situation 1.21.

Let 𝕍\mathbb{V} be a polarized integral variation of Hodge structure on SS satisfying 𝐇S=𝐃S\mathbf{H}_{S}=\mathbf{D}_{S}. Suppose that (𝐌,DM)(𝐆S,DS)(\mathbf{M},D_{M})\subset(\mathbf{G}_{S},D_{S}) is a Hodge subdatum, and that under the natural Hodge representation ρ:𝐆SGL(𝕍s)\rho:\mathbf{G}_{S}\to\operatorname{GL}(\mathbb{V}_{s})

  • -

    the group 𝐌\mathbf{M} is identified with the stabilizer of some v𝕍sv\in\mathbb{V}_{s};

  • -

    one has (𝐆s,Ds)(𝐌,DM)(\mathbf{G}_{s},D_{s})\subset(\mathbf{M},D_{M}), where (𝐆s,Ds)(\mathbf{G}_{s},D_{s}) is the Hodge datum at ss.

Let O=ρ(𝐆S)vO=\rho(\mathbf{G}_{S})\cdot v be the orbit of vv and write O()O()O(\mathbb{R})^{\circ}\subset O(\mathbb{R}) for the component containing vv, and set O()=O()O()O(\mathbb{Q})^{\circ}=O(\mathbb{Q})\cap O(\mathbb{R})^{\circ}. Fix a projective compactification SS¯S\subset\overline{S} and define the degree of ZSZ\subset S to be the degree of its closure.

Note that the orbit OO is preserved by the action of ΓS\Gamma_{S}, and so has, for each sS()s\in S(\mathbb{C}), a well-defined realization OsO_{s} as a subvariety of the affine space associated to 𝕍s\mathbb{V}_{s}. Likewise it makes sense to consider O()1ν𝕍O(\mathbb{Q})\cap\frac{1}{\nu}\mathbb{V} in each fibre of 𝕍\mathbb{V}: those points vO()v^{\prime}\in O(\mathbb{Q}) with the property that νv\nu v^{\prime} is integral. (Here we are using the notation “𝕍\mathbb{V}” in the same spirit as with 𝐆S\mathbf{G}_{S} and DSD_{S} above: it is an abstract object with a realization above each point sS()s\in S(\mathbb{C}).) Given a vector vO()v^{\prime}\in O(\mathbb{Q}) we obtain, using the fixed isomorphism 𝐆S/𝐌O\mathbf{G}_{S}/\mathbf{M}\simeq O, a Hodge subdatum (𝐌,DM)(\mathbf{M}^{\prime},D_{M^{\prime}}) defined as in 1.20. We then set Z(v)=φ1(π(DM))Z(v^{\prime})=\varphi^{-1}(\pi(D_{M^{\prime}})) 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 𝔊1,,𝔊\mathfrak{G}_{1},\ldots,\mathfrak{G}_{\ell} for 𝐆S\mathbf{G}_{S} associated to Hodge structures h1,,hDSh_{1},\ldots,h_{\ell}\in D_{S}, such that the Hodge datum of each hih_{i} is contained in a Hodge datum \mathbb{Q}-coset equivalent to (𝐌,DM)(\mathbf{M},D_{M}); and

  • -

    Hodge vectors v1,,vO()v_{1},\ldots,v_{\ell}\in O(\mathbb{Q})^{\circ}, with viv_{i} a Hodge vector for hih_{i};

such that we have

deg[v𝒪()ν(v)|νZ(v)]c+i=1|𝔊vi1ν𝕍|,\deg\left[\bigcup_{\begin{subarray}{c}v^{\prime}\in\mathcal{O}(\mathbb{Q})^{\circ}\\ \nu(v^{\prime})|\nu\end{subarray}}Z(v^{\prime})\right]\lesssim c^{+}\sum_{i=1}^{\ell}\left|\mathfrak{G}\cdot v_{i}\cap\frac{1}{\nu}\mathbb{V}\right|, (2)

as ν\nu\to\infty, where c+>0c^{+}>0 is a real constant depending on S,𝕍S,\mathbb{V} and QQ.

Notation.

For each fixed non-negative integer dd and analytic variety ZZ, we write ZdZ_{d} for the union of dd-dimensional components of ZZ.

Theorem 1.23 (Lower Bound).

Work in the situation of 1.21, assume that (𝐌,DM)(\mathbf{M},D_{M}) is strongly 𝕍\mathbb{V}-likely, and that 𝐌\mathbf{M} has finite index in its normalizer in 𝐆S\mathbf{G}_{S}. Let BS()B\subset S(\mathbb{C}) be any open neighbourhood. Then there exists a sufficiently general hyperplane section LSL\subset S which intersects BB and has codimension dimS(dimDSdimDM)\dim S-(\dim D_{S}-\dim D_{M}), and an open subset UO()U\subset O(\mathbb{R})^{\circ} such that

c|U1ν𝕍|v𝒪()ν(v)|ν|(Z(v)LB)0|c^{-}\left|U\cap\frac{1}{\nu}\mathbb{V}\right|\lesssim\sum_{\begin{subarray}{c}v^{\prime}\in\mathcal{O}(\mathbb{Q})^{\circ}\\ \nu(v^{\prime})|\nu\end{subarray}}|(Z(v^{\prime})\cap L\cap B)_{0}| (3)

as ν\nu\to\infty, where c>0c^{-}>0 is a real constant depending on S,𝕍S,\mathbb{V} and QQ.

The basic summary of the the two theorems is that Hodge loci defined by vectors vv^{\prime} with ν=ν(v)\nu=\nu(v^{\prime}) are roughly in correspondence with rational vectors in certain subsets of O()O(\mathbb{R})^{\circ} with denominators dividing ν\nu. These subsets are at worst as large as a Siegel set orbit, and in the strongly 𝕍\mathbb{V}-likely case contain open neighbourhoods corresponding to any chosen open subset of S()S(\mathbb{C}). The Siegel sets 𝔊i\mathfrak{G}_{i} referred to in 1.22 are “Hodge-theoretic” in the sense that their associated maximal compact subgroups KiK_{i} are naturally constructed from the Hodge metric at a point hiDSh_{i}\in D_{S}; we give a precise description in §2.1. Moreover we act with these sets on Hodge vectors viv_{i} for the associated hih_{i}.

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 O()O(\mathbb{R})^{\circ}, but nevertheless satisfy the property that, for any positive integer ν\nu, they contain only finitely many rational points with denominators of size at most ν\nu. 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 𝔊\mathfrak{G} associated to a Hodge structure hh on a polarized integral lattice LL, a \mathbb{Q}-algebraic subvariety XX of the affine space associated to LL, and orbit 𝔊v\mathfrak{G}\cdot v of a Hodge vector vv for hh, there exists a constant c>0c>0 such that

#((𝔊v)X()L1ν)cνmin{m+s2,m+r2, 2dimX},\#\left((\mathfrak{G}\cdot v)\cap X(\mathbb{Q})\cap L\frac{1}{\nu}\right)\lesssim c\cdot\nu^{\min\{m+s-2,\,m+r-2,\,2\dim X\}},

where (r,s)(r,s) 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 vv 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 𝕍\mathbb{V} and orbit OO of 𝐆S\mathbf{G}_{S}, there exists a constant cc depending on (O,𝕍,Q)(O,\mathbb{V},Q) such that

degNLO,νcνmin{m+r2,m+s2, 2dimO}\deg\operatorname{NL}_{O,\nu}\lesssim c\cdot\nu^{\min\{m+r-2,\,m+s-2,\,2\dim O\}}

where (r,s)(r,s) is the signature of (𝕍,Q)(\mathbb{V},Q).

In the above result NLO,ν\operatorname{NL}_{O,\nu} is the Noether-Lefschetz locus associated to OO and ν\nu, defined as

NLO,ν={sS(): there exists a Hodge vector v𝕍s with vO(),ν(v)|ν}.\operatorname{NL}_{O,\nu}=\left\{s\in S(\mathbb{C}):\begin{array}[]{c}\textrm{ there exists a Hodge vector }v\in\mathbb{V}_{s}\\ \textrm{ with }v\in O(\mathbb{Q}),\,\,\nu(v)|\nu\end{array}\right\}.

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 𝕍\mathbb{V} there exists a constant cc depending only on (𝕍,Q)(\mathbb{V},Q) such that

degNLqcqm2+min{r2+s2,2rs}2,\deg\operatorname{NL}_{q}\lesssim c\cdot q^{m^{2}+\textrm{min}\{r^{2}+s^{2},2rs\}-2},

where (r,s)(r,s) is the signature of (𝕍,Q)(\mathbb{V},Q).

In the above we define NLq\operatorname{NL}_{q} by

NLq:={sS(): there exists a Hodge vector v𝕍s with Q¯(v)=q}.\operatorname{NL}_{q}:=\left\{s\in S(\mathbb{C}):\textrm{ there exists a Hodge vector }v\in\mathbb{V}_{s}\textrm{ with }\overline{Q}(v)=q\right\}.

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 𝐆S\mathbf{G}_{S}, 𝐇S\mathbf{H}_{S}, and 𝐌\mathbf{M}, 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 π:DSΓ\DS\pi:D_{S}\to\Gamma\backslash D_{S} the natural projection, and fix a Hodge subdatum (𝐌,DM)(𝐆S,DS)(\mathbf{M},D_{M})\subset(\mathbf{G}_{S},D_{S}). Suppose one wanted to merely count loci in Γ\DS\Gamma\backslash D_{S} which are “generalized Hecke translates” of the image of some DMDD_{M}\subset D: i.e., images under π\pi of DMD_{M^{\prime}}, where (𝐌,DM)(\mathbf{M}^{\prime},D_{M^{\prime}}) is a Hodge subdatum of (𝐆S,DS)(\mathbf{G}_{S},D_{S}) which is \mathbb{Q}-coset equivalent to (𝐌,DM)(\mathbf{M},D_{M}). Then as we have discussed, one can view each such generalized Hecke translate as corresponding to a rational point of 𝐆S/𝐌\mathbf{G}_{S}/\mathbf{M}. One can then try to pick a fundamental set for the action of Γ\Gamma on 𝐆S()\mathbf{G}_{S}(\mathbb{R}), and count generalized Hecke translates by counting rational points in Γ\G()/𝐌()\Gamma\backslash G(\mathbb{R})/\mathbf{M}(\mathbb{R}). (The space Γ\G()/𝐌()\Gamma\backslash G(\mathbb{R})/\mathbf{M}(\mathbb{R}) 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 SΓ\DSS\to\Gamma\backslash D_{S}, and to make the strategy work one faces at least two potential problems.

  • (1)

    The loci π(DM)\pi(D_{M^{\prime}}) might not intersect φ(S)\varphi(S). More generally, if one looks at the locus of g𝐌()𝐆S()/𝐌()g\mathbf{M}(\mathbb{R})\in\mathbf{G}_{S}(\mathbb{R})/\mathbf{M}(\mathbb{R}) for which π(gDM)\pi(gD_{M}) intersects φ(S)\varphi(S) 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 π(gDM)\pi(gD_{M}), in particular infinitely many, intersect φ(S)\varphi(S) in the same locus. In other words, one has many different rational points of 𝐆S/𝐌\mathbf{G}_{S}/\mathbf{M} corresponding to the same Hodge locus, leading to overcounting.

The first problem (1) is solved using by the “𝕍\mathbb{V}-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 𝐌\mathbf{M} in 𝐆S\mathbf{G}_{S}) away from a strict Zariski closed locus in SS. This being unavailable, we need to assume the “strongly 𝕍\mathbb{V}-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 SS with an appropriately chosen hyperplane section so that the Hodge loci of interest are zero dimensional. One then carefully partitions SS into analytic neighbourhoods BS()B\subset S(\mathbb{C}) such that a Hodge locus in BB of the form Bφ1(gDM)B\cap\varphi^{-1}(g\cdot D_{M}) always consists of a single point. That this can be done for BB belonging to a finite definable analytic cover S()=i=1nBiS(\mathbb{C})=\bigcup_{i=1}^{n}B_{i} is a consequence of the definability in an,exp\mathbb{R}_{\textrm{an,exp}} of Hodge-theoretic period maps proven in [BKT20]. Then by counting Hodge loci in each BiB_{i} and summing over all ii, 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 BiB_{i}. Following the theory in [BKT20], we can reduce to the case where each BiB_{i} 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 \mathbb{Q}. One can essentially always apply such techniques to count rational points inside compact subsets of O()O(\mathbb{R}), 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 𝐆S()\mathbf{G}_{S}(\mathbb{R})-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 ν\nu lie inside a compact subset of volume polynomially bounded by ν\nu. 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 (S,𝕍)(S,\mathbb{V}), while [Tay20] and [TT23] both essentially estimate the degree of a corresponding “locus of Hodge classes” of (S,𝕍)(S,\mathbb{V}). Here we are borrowing the terminology of [CDK94] and Voisin [Voi10] to describe the analytic loci in the vector bundle an=𝕍𝒪San\mathcal{H}^{\textrm{an}}=\mathbb{V}\otimes_{\mathbb{Z}}\mathcal{O}_{S^{\textrm{an}}} consisting of points (s,v)(s,v) with vv a Hodge class for the Hodge structure on 𝕍s\mathbb{V}_{s}. The vector bundle an\mathcal{H}^{\textrm{an}} has a canonical algebraic model \mathcal{H} 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 SS 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 S\mathcal{H}\to S 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 f:XSf:X\to S of smooth projective degree d5d\geq 5 hypersurfaces in 3\mathbb{P}^{3} with SS a curve, the Noether-Lefschetz locus of SS is a finite subset of S()S(\mathbb{C}). Formulated for the locus of Hodge classes, or when counting intersections with Noether-Lefschetz components in UdU_{d} with multiplicity, the statement is false: if s0Ss_{0}\in S is a point mapping to the Fermat surface in UdU_{d} (recall §1.1.1), the locus of Hodge classes above s0s_{0} consists of all points in 𝕍,s0\mathbb{V}_{\mathbb{Z},s_{0}} belonging to an infinite lattice of large rank (c.f. [AMVL19, Thm. 1]), and counting with multiplicity each 11-dimensional subspace is counted separately. And situations where the Hodge locus of SS is infinite and the two types of loci should have different asymptotics also occur: if one takes SS to be a Shimura curve in 𝒜g\mathcal{A}_{g} with g2g\geq 2, every Hodge locus point s0Ss_{0}\in S arises in infinitely many different ways as the intersection of SS with a special locus in 𝒜g\mathcal{A}_{g}, 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 SS. (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 S()S(\mathbb{C}) of uncontrolled size, to deduce a global point count on all of S()S(\mathbb{C}) one must, at least in the case of non-compact S()S(\mathbb{C}), 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 (𝐆S/𝐌)()(\mathbf{G}_{S}/\mathbf{M})(\mathbb{R}) whereas [TT23] instead integrates a differential form on SS 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 (G,D)=(𝐆S,DS)(G,D)=(\mathbf{G}_{S},D_{S}) be the generic Hodge datum associated to 𝕍\mathbb{V}, and φ:SΓ\D\varphi:S\to\Gamma\backslash D the period map induced by 𝕍\mathbb{V} with ΓG()\Gamma\subset G(\mathbb{Q}) a neat arithmetic lattice containing the image of monodromy. By embedding DD as an open subvariety of its compact dual Dˇ\widecheck{{D}} the set DD inherits a natural alg\mathbb{R}_{\textrm{alg}}-definable structure.

To define a Siegel set of DD we require some setup. Each such set will be defined as a certain orbit 𝔒D\mathfrak{O}\subset D associated to a maximal compact subgroup KG()K\subset G(\mathbb{R}) and a minimal parabolic \mathbb{Q}-subgroup PGP\subset G. The maximal compact group will be defined using a Hodge structure oDo\in D by the following lemma:

Lemma 2.1.

Suppose that oDo\in D is a point, and let MoG()M_{o}\subset G(\mathbb{R}) be its stabilizer. Then there is a unique maximal compact subgroup KoG()K_{o}\subset G(\mathbb{R}) containing MoM_{o}.

Proof.

The Hodge structure o:𝕌Go:\mathbb{U}\to G_{\mathbb{R}} induces a grading 𝔤=i𝔤i,i\mathfrak{g}_{\mathbb{C}}=\bigoplus_{i}\mathfrak{g}^{i,-i} through the adjoint action, and we define KoK_{o} to be the exponential of the real Lie subalgebra 𝔨o\mathfrak{k}_{o} underlying 𝔨o,=i even𝔤i,i\mathfrak{k}_{o,\mathbb{C}}=\bigoplus_{i\textrm{ even}}\mathfrak{g}^{i,-i}. As the Lie algebra 𝔪o\mathfrak{m}_{o} of MoM_{o} may be identified with the 𝔤0,0\mathfrak{g}^{0,0}_{\mathbb{R}} summand one clearly has MoKoM_{o}\subset K_{o}. A polarization on the Hodge structure oo induces a polarization on each simple summand 𝔲\mathfrak{u} of 𝔤\mathfrak{g} which is ad𝔲\textrm{ad}\,\mathfrak{u}-invariant when restricted to that summand, see [GGK12, III.A.5, pg.75]. Necessarily such a form is proportional to the Killing form BB on each summand, which implies that the restriction of the Killing form to 𝔨o\mathfrak{k}_{o} is definite of a common sign, and hence negative definite because the Killing form is negative definite on 𝔪o\mathfrak{m}_{o}. 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 𝔨o\mathfrak{k}_{o} is a maximal compact subgroup containing 𝔪o\mathfrak{m}_{o}. That it is unique, which amounts to the statement that any other subalgebra of 𝔤\mathfrak{g} containing 𝔪o\mathfrak{m}_{o} on which BB is negative definite lies in 𝔨o\mathfrak{k}_{o}, follows from the Hodge-Riemann bilinear relations for the polarizing form on 𝔤\mathfrak{g}. ∎

Lemma 2.2.

For any minimal parabolic \mathbb{Q}-subgroup PP of a reductive \mathbb{Q}-group GG and maximal compact subgroup KG()K\subset G(\mathbb{R}), there exists a unique real torus SP,KPS_{P,K}\subset P_{\mathbb{R}} satisfying:

  • (i)

    the torus SP,KS_{P,K} is P()P(\mathbb{R})-conjugate to a maximal \mathbb{Q}-split torus of PP; and

  • (ii)

    SP,KS_{P,K} is stabilized by the Cartan involution associated to KK.

Proof.

This is [Orr18, Lem 2.1]. ∎

Definition 2.3.

Suppose that PGP\subset G is a minimal parabolic \mathbb{Q}-subgroup, and let S=SP,oS=S_{P,o} be as in 2.2. For any real number t>0t>0 write

At={αS()+:χ(α)t for all χΔ},A_{t}=\{\alpha\in S(\mathbb{R})^{+}:\chi(\alpha)\geq t\textrm{ for all }\chi\in\Delta\},

where Δ\Delta is the set of simple roots of GG with respect to SS, using the ordering induced by PP. Then we define a Siegel set 𝔖G()\mathfrak{S}\subset G(\mathbb{R}) (associated to (P,o,t)(P,o,t)) to be a set ΩAtKo\Omega A_{t}K_{o} where ΩP()\Omega\subset P(\mathbb{R}) is compact. We likewise define a Siegel set 𝔒D\mathfrak{O}\subset D (associated to (P,o,t)(P,o,t)) to be an orbit ΩAtKoo\Omega A_{t}K_{o}\cdot o.

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 𝔉\mathfrak{F} is any definable fundamental set for Γ\Gamma, there is a unique induced definable structure on Γ\D\Gamma\backslash D for which the map 𝔉Γ\D\mathfrak{F}\to\Gamma\backslash D 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] + ε\varepsilon).

There exists a definable fundamental set 𝔉D\mathfrak{F}\subset D for Γ\Gamma, obtained as a finite union of Siegel sets 𝔒i\mathfrak{O}_{i}, such that the map φ:SΓ\D\varphi:S\to\Gamma\backslash D is an,exp\mathbb{R}_{\textrm{an,exp}}-definable for the induced definable structure on Γ\D\Gamma\backslash D. In particular, there exists a finite cover S=i=1nBiS=\bigcup_{i=1}^{n}B_{i} by definable simply-connected opens and an,exp\mathbb{R}_{\textrm{an,exp}}-definable local lifts ψi:Bi𝔒i𝔉\psi_{i}:B_{i}\to\mathfrak{O}_{i}\subset\mathfrak{F} of φ\varphi.

Fix a set of points 𝒮S()\mathcal{S}\subset S(\mathbb{C}). Then we may choose the above data such that for each BiB_{i} intersecting the topological closure 𝒮top\mathcal{S}^{\textrm{top}} of 𝒮\mathcal{S} in S()S(\mathbb{C}), the Siegel set 𝔒i=ΩiAt,iKoioi\mathfrak{O}_{i}=\Omega_{i}A_{t,i}K_{o_{i}}\cdot o_{i} is associated to a point oiψi(Bi𝒮)o_{i}\in\psi_{i}(B_{i}\cap\mathcal{S}).

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 𝒮\mathcal{S} at the end.

Applying Hironaka’s Theorem we may reduce to the case where SS is the complement of a simple normal crossing divisor EE in a smooth projective variety S¯\overline{S}. By passing to a finite étale cover we may assume monodromy at infinity is unipotent.

It is clear we can construct such a ψ\psi in a small enough neighbourhood of any sSs\in S, so it suffices to show this on a small enough punctured neighbourhood B=Δa×(Δ)bB=\Delta^{a}\times(\Delta^{\circ})^{b} of the simple normal crossing divisor EE; 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 a=0a=0 by allowing factors with trivial monodromy. We let exp:b(Δ)b\exp:\mathbb{H}^{b}\to(\Delta^{\circ})^{b} be the standard universal covering and let \mathfrak{H}\subset\mathbb{H} be the standard Siegel fundamental set as in [BKT20, §4.2]. Then we may construct a map ψ~:bD\widetilde{\psi}:\mathfrak{H}^{b}\to D by lifting φ\varphi, and define ψ\psi by inverting exp|b{\left.\kern-1.2pt\exp\vphantom{\big{|}}\right|_{\mathfrak{H}^{b}}} after choosing a branch cut and and composing with ψ~\widetilde{\psi}. By expanding the map exp|b{\left.\kern-1.2pt\exp\vphantom{\big{|}}\right|_{\mathfrak{H}^{b}}} in terms of its real-analytic components one sees it is an,exp\mathbb{R}_{\textrm{an,exp}}-definable, so to show that ψ\psi is an,exp\mathbb{R}_{\textrm{an,exp}}-definable it suffices to check the definability of ψ~\widetilde{\psi}.

To do this we apply the Nilpotent orbit theorem [Sch73], which tells us that

ψ~=exp(i=1bziNi)ψlim(exp(z))\widetilde{\psi}=\exp\left(\sum_{i=1}^{b}z_{i}N_{i}\right)\cdot\psi_{\textrm{lim}}(\exp(z))

where z=(z1,,zb)z=(z_{1},\ldots,z_{b}) are the natural coordinates for b\mathfrak{H}^{b}, the NiN_{i} are nilpotent operators, and ψlim\psi_{\textrm{lim}} is an analytic function on Δb\Delta^{b}, hence definable. Then as the left factor is just a polynomial in zz, the result follows.

Following [BKT20, §4.5], it now suffices to check that the image of ψ~\widetilde{\psi} lies in a finite union of Siegel sets. The monodromy representation associated to 𝕍\mathbb{V} gives a faithful representation ρ:GSL(V)\rho:G\to\operatorname{SL}(V) where VV is an integral lattice, and hence induces a map ι:DDV\iota:D\to D_{V}, where DVD_{V} is the symmetric space of positive-definite symmetric forms on VV_{\mathbb{R}}; here the the map ι\iota is given by sending the polarized Hodge structure oo to the positive-definite symmetric form Qo:(u,v)Q(Cou,v)Q_{o}:(u,v)\mapsto Q_{\mathbb{C}}(C_{o}u,v), where CoC_{o} is the Weil operator associated to oo. From [BGST23, Prop 3.4] one learns that the inverse image under ι\iota of a Siegel set of DVD_{V} associated to K~o\widetilde{K}_{o} is contained in finitely many Siegel sets of DD for KoK_{o}. Thus we may reduce to the same claim for the image of ιψ~\iota\circ\widetilde{\psi}.

Using the explicit description of Siegel sets for DVD_{V} in terms of the Gram-Schmidt process, the condition that the set 𝔅=im(ιψ~)DV\mathfrak{B}=\textrm{im}(\iota\circ\widetilde{\psi})\subset D_{V} lie inside a Siegel for DVD_{V} is equivalent to the following claim: there exists a basis ={e1,,em}\mathcal{E}=\{e_{1},\ldots,e_{m}\} for VV and a constant κ\kappa such that the following inequalities hold for all b𝔅b\in\mathfrak{B}:

  • (i)

    |b(ei,ej)|<κb(ei,ei)|b(e_{i},e_{j})|<\kappa\,b(e_{i},e_{i}) for all i,ji,j;

  • (ii)

    b(ei,ei)<κb(ej,ej)b(e_{i},e_{i})<\kappa\,b(e_{j},e_{j}) for i<ji<j; and

  • (iii)

    ib(ei,ei)<κdetb\prod_{i}b(e_{i},e_{i})<\kappa\det b.

The corresponding Siegel subset of DVD_{V} may be taken to be associated to the maximal compact group K~o\widetilde{K}_{o} associated to a chosen point oDSo\in D_{S} (c.f. [BGST23, Thm. 3.3] and its proof). In particular we can take oψ(B)o\in\psi(B).

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 e1,,eme_{1},\ldots,e_{m} so that condition (iii) holds for some κ\kappa as a consequence of [BKT20, Thm. 4.8], and (ii) can be assumed for the same κ\kappa by reordering the basis and partitioning the image as necessary. Schmid’s 11-dimensional result shows all of these conditions after restricting to any curve τ\tau in b\mathfrak{H}^{b} for a constant κτ\kappa_{\tau} depending only on τ\tau, and so in particular shows (1) on such curves. One can then make the constant κτ\kappa_{\tau} appearing in (1) independent of τ\tau using the fact that the coordinates of Q(Cψ~(z)(),())Q_{\mathbb{C}}(C_{\widetilde{\psi}(z)}(-),(-)) are “roughly polynomial” (see [BKT20, Lem. 4.5]) in the coordinates zz 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 𝒮\mathcal{S}. The above argument has showed how, around any point sS¯()s\in\overline{S}(\mathbb{C}), and given any ball B¯S¯()\overline{B}\subset\overline{S}(\mathbb{C}) centred at ss with B¯=Δa+b\overline{B}=\Delta^{a+b}, we can construct a period map ψ\psi on B=B¯S()B=\overline{B}\cap S(\mathbb{C}) landing inside a Siegel set 𝔊=𝔊(s)\mathfrak{G}=\mathfrak{G}(s) depending on ss. (Up to shrinking BB and taking a branch cut.) This in particular applies to points s𝒮¯tops\in\overline{\mathcal{S}}^{\textrm{top}}, where we write 𝒮¯top\overline{\mathcal{S}}^{\textrm{top}} for the topological closure of 𝒮\mathcal{S} in S¯()\overline{S}(\mathbb{C}). Now as discussed above, the maximal compact of the resulting Siegel set can be chosen to be associated to any point oψ(B)o\in\psi(B). Since 𝒮\mathcal{S} is dense in 𝒮¯top\overline{\mathcal{S}}^{\textrm{top}}, we can in particular choose oψ(B𝒮)o\in\psi(B\cap\mathcal{S}).

The set 𝒮¯top\overline{\mathcal{S}}^{\textrm{top}} is a closed subset of a compact set, so is in particular compact. Thus, after applying this argument at every point s𝒮¯tops\in\overline{\mathcal{S}}^{\textrm{top}} and obtaining an appropriate cover {B¯s}s𝒮¯top\{\overline{B}_{s}\}_{s\in\overline{\mathcal{S}}^{\textrm{top}}} of 𝒮¯top\overline{\mathcal{S}}^{\textrm{top}}, we can find a finite subcover. This finite subcover may then be extended to a finite cover of all of S¯\overline{S} which induces the cover of SS 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 SO(r,s)\operatorname{SO}(r,s).

3.1 Standard Siegel Sets

We set m=r+sm=r+s, and G=SO(r,s)G=\operatorname{SO}(r,s). We will view GG as the stabilizer of the bilinear form Q=IrIsQ=-I_{r}\oplus I_{s}. Replacing QQ with Q-Q and reordering the standard coordiantes x1,,xmx_{1},\ldots,x_{m} if necessary, we may assume that rsr\geq s; none of our results on Siegel set orbits will depend on this convention.

We will assume that 𝔖\mathfrak{S} is a Siegel set constructed with respect to a certain special choice of maximal compact subgroup KK and parabolic subgroup PP; 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 K=S(O(r)×O(s))K=S(O(r)\times O(s)). To construct PP we introduce a second set of coordinates, w1,,wmw_{1},\ldots,w_{m}, which are related to the coordinates x1,,xmx_{1},\ldots,x_{m} by

wi\displaystyle w_{i} =xixr+i\displaystyle=x_{i}-x_{r+i} 1is\displaystyle 1\leq i\leq s (4)
wi\displaystyle w_{i} =xi\displaystyle=x_{i} s+1ir\displaystyle s+1\leq i\leq r (5)
wi\displaystyle w_{i} =xi+xir\displaystyle=x_{i}+x_{i-r} r+1im.\displaystyle r+1\leq i\leq m. (6)

The change of basis matrices from x¯\overline{x}-coordinates to w¯\overline{w}-coordinates are given by

S=(Is0Is0Irs0Is0Is),S1=12(Is0Is02Irs0Is0Is).S=\begin{pmatrix}I_{s}&0&-I_{s}\\ 0&I_{r-s}&0\\ I_{s}&0&I_{s}\end{pmatrix},\hskip 20.00003ptS^{-1}=\frac{1}{2}\begin{pmatrix}I_{s}&0&I_{s}\\ 0&2I_{r-s}&0\\ -I_{s}&0&I_{s}\end{pmatrix}.

And in the new coordinate frame, the bilinear form QQ is given by

QS\displaystyle Q_{S} :=14(Is0Is02Irs0Is0Is)(Is000Irs000Is)(Is0Is02Irs0Is0Is)\displaystyle:=\frac{1}{4}\begin{pmatrix}I_{s}&0&-I_{s}\\ 0&2I_{r-s}&0\\ I_{s}&0&I_{s}\end{pmatrix}\begin{pmatrix}-I_{s}&0&0\\ 0&-I_{r-s}&0\\ 0&0&I_{s}\end{pmatrix}\begin{pmatrix}I_{s}&0&I_{s}\\ 0&2I_{r-s}&0\\ -I_{s}&0&I_{s}\end{pmatrix}
=14(Is0Is02Irs0Is0Is)(Is0Is02Irs0Is0Is)\displaystyle=\frac{1}{4}\begin{pmatrix}I_{s}&0&-I_{s}\\ 0&2I_{r-s}&0\\ I_{s}&0&I_{s}\end{pmatrix}\begin{pmatrix}-I_{s}&0&-I_{s}\\ 0&-2I_{r-s}&0\\ -I_{s}&0&I_{s}\end{pmatrix}
=14(002Is04Irs02Is00).\displaystyle=\frac{1}{4}\begin{pmatrix}0&0&-2I_{s}\\ 0&-4I_{r-s}&0\\ -2I_{s}&0&0\end{pmatrix}.

We compute the Lie algebra in the w¯\overline{w}-coordinate system, which is given by matrices WW satisfying WtQS+QSW=0W^{t}Q_{S}+Q_{S}W=0. For such WW, one calculates that

0\displaystyle 0 =(W11tW21tW31tW12tW22tW32tW13tW23tW33t)(00Is02Irs0Is00)+(00Is02Irs0Is00)(W11W12W13W21W22W23W31W32W33)\displaystyle=\begin{pmatrix}W^{t}_{11}&W^{t}_{21}&W^{t}_{31}\\ W^{t}_{12}&W^{t}_{22}&W^{t}_{32}\\ W^{t}_{13}&W^{t}_{23}&W^{t}_{33}\end{pmatrix}\begin{pmatrix}0&0&I_{s}\\ 0&2I_{r-s}&0\\ I_{s}&0&0\end{pmatrix}+\begin{pmatrix}0&0&I_{s}\\ 0&2I_{r-s}&0\\ I_{s}&0&0\end{pmatrix}\begin{pmatrix}W_{11}&W_{12}&W_{13}\\ W_{21}&W_{22}&W_{23}\\ W_{31}&W_{32}&W_{33}\end{pmatrix}
=(W31t+W312W21t+W32W11t+W33W32t+2W212(W22t+W22)W12t+2W23W33t+W112W23t+W12W13t+W13).\displaystyle=\begin{pmatrix}W^{t}_{31}+W_{31}&2W^{t}_{21}+W_{32}&W^{t}_{11}+W_{33}\\ W^{t}_{32}+2W_{21}&2(W^{t}_{22}+W_{22})&W^{t}_{12}+2W_{23}\\ W^{t}_{33}+W_{11}&2W^{t}_{23}+W_{12}&W^{t}_{13}+W_{13}\end{pmatrix}.

So in the w¯\overline{w}-coordinate system one obtains the description

𝔰𝔬(r,s)w¯={(W112W23tW13W21W22W23W312W21tW11t):W22=W22t,W31=W31t,W13=W13t}.\displaystyle\mathfrak{so}(r,s)_{\overline{w}}=\left\{\begin{pmatrix}W_{11}&-2W^{t}_{23}&W_{13}\\ W_{21}&W_{22}&W_{23}\\ W_{31}&-2W^{t}_{21}&-W^{t}_{11}\end{pmatrix}:W_{22}=-W^{t}_{22},\hskip 5.0ptW_{31}=-W^{t}_{31},\hskip 5.0ptW_{13}=-W^{t}_{13}\right\}.

Write UsGLsU_{s}\subset\operatorname{GL}_{s} the group of s×ss\times s upper triangular matrices, and 𝔲s\mathfrak{u}_{s} for its Lie algebra. We define a Lie algebra 𝔭\mathfrak{p} by the equations

𝔭\displaystyle\mathfrak{p} ={W𝔰𝔬(r,s)w¯:W21=0,W31=0,W11𝔲s()}.\displaystyle=\{W\in\mathfrak{so}(r,s)_{\overline{w}}:\hskip 5.0ptW_{21}=0,\hskip 5.0ptW_{31}=0,\hskip 5.0ptW_{11}\in\mathfrak{u}_{s}(\mathbb{R})\}.

In w¯\overline{w}-coordinates, the corresponding matrices have the form

={(Y2W23tW130W22W2300Yt):YUs(),W13=W13t}.\displaystyle=\left\{\begin{pmatrix}Y&-2W^{t}_{23}&W_{13}\\ 0&W_{22}&W_{23}\\ 0&0&-Y^{t}\end{pmatrix}:Y\in U_{s}(\mathbb{R}),\hskip 5.0ptW_{13}=-W^{t}_{13}\right\}.

We write 𝔫𝔭\mathfrak{n}\subset\mathfrak{p} for the nilpotent Lie algebra defined by setting W22=0W_{22}=0 and mandating that the diagonal entries of YY are all equal to 0, and set U=exp(𝔫)U=\textrm{exp}(\mathfrak{n}).

Lemma 3.1.

The set 𝔭\mathfrak{p} is the Lie algebra of a minimal parabolic subalgebra of 𝔤\mathfrak{g}.

Proof.

[He13, §2.3] gives a description of a minimal parabolic subgroup of GG. One immediately checks that our algebra 𝔭\mathfrak{p} and the group appearing in [He13, §2.3] have the same dimension, so it suffices to check that 𝔭\mathfrak{p} is parabolic. By [BDPP11, Def. 1.1] it suffices to check that the orthogonal complement 𝔭\mathfrak{p}^{\perp} with respect to the Killing form is nilpotent, which is easily checked. ∎

Write PP for the associated Parabolic subgroup of GG. By exponentiating, we see that

P{(ABC0DE00At,1):AUs(),B,C,EGLm(),DO(rs)()}.P\subset\left\{\begin{pmatrix}A&B&C\\ 0&D&E\\ 0&0&A^{t,-1}\end{pmatrix}:A\in U_{s}(\mathbb{R}),B,C,E\in\operatorname{GL}_{m}(\mathbb{R}),D\in O(r-s)(\mathbb{R})\right\}. (7)

We now describe the split torus SPS\subset P determined by 2.2. The Cartan involution associated to KK in x¯\overline{x}-coordinates is given by θ(X)=Xt\theta(X)=-X^{t}. In the w¯\overline{w}-coordinates this becomes

θ(W)\displaystyle\theta(W) =Sθ(S1WS)S1\displaystyle=S\theta(S^{-1}WS)S^{-1}
=SStWtS1,tS1.\displaystyle=-SS^{t}W^{t}S^{-1,t}S^{-1}.

The matrices SStSS^{t} and S1,tS1=(SSt)1S^{-1,t}S^{-1}=(SS^{t})^{-1} are both diagonal, so one sees that the diagonal matrices of PP are preserved under the involution. We then take the Lie algebra 𝔞\mathfrak{a} of SS to be the intersection in the w¯\overline{w}-coordinates of 𝔭\mathfrak{p} and the diagonal subgelbra of 𝔤𝔩m\mathfrak{gl}_{m}. It is then clear that SS is split over \mathbb{Q} of the correct dimension (compare with the torus in [He13, §2.3]), and stable under θ\theta.

The rank of the group SO(r,s)\operatorname{SO}(r,s) is ss (recall we have assumed rsr\geq s), so we should have ss simple roots χ1,,χs\chi_{1},\ldots,\chi_{s} matching 2.3. Since our torus SS lies in the diagonal subgroup DmD_{m} of GLm\operatorname{GL}_{m}, we may obtain these roots from restrictions of roots of DD acting on 𝔤𝔩m\mathfrak{gl}_{m}, in particular, we have

At:={[aij]Dm()+SO(r,s)():aiita(i+1)(i+1) for 1is1,ast}.\displaystyle A_{t}:=\{[a_{ij}]\in D_{m}(\mathbb{R})^{+}\cap\operatorname{SO}(r,s)(\mathbb{R}):a_{ii}\geq ta_{(i+1)(i+1)}\hskip 5.0pt\textrm{ for }\hskip 5.0pt1\leq i\leq s-1,\hskip 10.00002pta_{s}\geq t\}.
Lemma 3.2.

Let KG()K^{\prime}\subset G(\mathbb{R}) be a maximal compact subgroup and 𝔖\mathfrak{S}^{\prime} a Siegel set for G()G(\mathbb{R}) relative to KK^{\prime} and a minimal paraoblic PP^{\prime}. Then there exists γG()\gamma\in G(\mathbb{Q}), τU()\tau\in U(\mathbb{R}) and βS()\beta\in S(\mathbb{R}), and a Siegel set 𝔖\mathfrak{S} for (P,S,K)(P,S,K) such that γ𝔖γ1σ𝔖\gamma\mathfrak{S}^{\prime}\gamma^{-1}\sigma\subset\mathfrak{S} where σ=τβ\sigma=\tau\beta. We have γPγ1=P\gamma P^{\prime}\gamma^{-1}=P and σKσ1=γ1Kγ\sigma K\sigma^{-1}=\gamma^{-1}K^{\prime}\gamma.

Proof.

Our argument is a mild variant of [Orr18, Lem. 3.8]. Let PP^{\prime} be the parabolic group associated to 𝔖=ΩAtK\mathfrak{S}^{\prime}=\Omega^{\prime}A^{\prime}_{t}K^{\prime}. Then as both PP and PP^{\prime} are minimal parabolic \mathbb{Q}-subgroups, there exists γG()\gamma\in G(\mathbb{Q}) such that γPγ1=P\gamma P^{\prime}\gamma^{-1}=P. Since KK and γKγ1\gamma K^{\prime}\gamma^{-1} are both maximal compact subgroups, there exists σG()\sigma\in G(\mathbb{R}) such that σKσ1=γ1Kγ\sigma K\sigma^{-1}=\gamma^{-1}K^{\prime}\gamma.

We now consider the Iwasawa decomposition of G()G(\mathbb{R}) with respect to the Cartan involution θ\theta. We recall that this is induced by a decomposition 𝔤=𝔫𝔞𝔨\mathfrak{g}=\mathfrak{n}\oplus\mathfrak{a}\oplus\mathfrak{k}, where 𝔮=𝔫𝔞\mathfrak{q}=\mathfrak{n}\oplus\mathfrak{a} and 𝔨\mathfrak{k} are both eigenspaces for θ\theta, and 𝔞\mathfrak{a} is a choice of maximal abelian subalgebra of 𝔮\mathfrak{q}. Since the Lie algebra of SS is stabilized by θ\theta, we may choose 𝔞\mathfrak{a} to be the Lie algebra of SS. It is then directly checked that 𝔫\mathfrak{n} is a sum of root spaces for SS. It follows that G()=U()S()+KG(\mathbb{R})=U(\mathbb{R})S(\mathbb{R})^{+}K, and that we may choose σ=τβ\sigma=\tau\beta with βS+()\beta\in S^{+}(\mathbb{R}) and τU()\tau\in U(\mathbb{R}). It follows that (γσ)1P(γσ)=P(\gamma\sigma)^{-1}P^{\prime}(\gamma\sigma)=P and hence by 2.2 that (γσ)1S(γσ)=S(\gamma\sigma)^{-1}S^{\prime}(\gamma\sigma)=S, so we may write At=σ1γ1AtγσA_{t}=\sigma^{-1}\gamma^{-1}A^{\prime}_{t}\gamma\sigma. Finally, as in the proof of [Orr18, Lem. 3.8]

γ1𝔖γσ\displaystyle\gamma^{-1}\mathfrak{S}^{\prime}\gamma\sigma =(γ1Ωγ)σσ1γ1Atγσσ1γ1Kγσ\displaystyle=(\gamma^{-1}\Omega^{\prime}\gamma)\sigma\sigma^{-1}\gamma^{-1}A^{\prime}_{t}\gamma\sigma\sigma^{-1}\gamma^{-1}K^{\prime}\gamma\sigma
=(γ1Ωγ)τβAtK.\displaystyle=(\gamma^{-1}\Omega^{\prime}\gamma)\tau\beta A_{t}K.

so the result follows choosing Ω(γ1Ωγ)τ\Omega\supset(\gamma^{-1}\Omega^{\prime}\gamma)\tau and AsβAtA_{s}\supset\beta A_{t} 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 h:𝕌SO(r,s)()h:\mathbb{U}\to\operatorname{SO}(r,s)(\mathbb{R}), let KhK_{h} be the associated maximal compact, and let (Ph,Sh,Kh)(P_{h},S_{h},K_{h}) be an associated Siegel triple. As in §2.1, the group KhK_{h} is the intersection with SO(r,s)()\operatorname{SO}(r,s)(\mathbb{R}) of O(Qh)O(Q_{h}), where Qh(u,w)=Q(Chu,w)Q_{h}(u,w)=Q(C_{h}u,w) is the Hodge-metric at hh.

Notation.

Given a vector vmv\in\mathbb{R}^{m}, we write either viv_{i} or [v]i[v]_{i} for its ii’th entry. We let =(v){r+1,,m}\ell=\ell(v)\in\{r+1,\ldots,m\} be the smallest index ii in the specified range for which vi0v_{i}\neq 0. If we write [v][v]_{\ell} we understand [v](v)[v]_{\ell(v)}. If no such index exists, then we say =(v)\ell=\ell(v) is undefined.

Proposition 3.3.

Fix a Siegel set 𝔊\mathfrak{G} for (Ph,Sh,Kh)(P_{h},S_{h},K_{h}) and a real Hodge vector vv for hh. Then there exists a basis ={f1,,fm}\mathcal{F}=\{f_{1},\ldots,f_{m}\} for m\mathbb{Q}^{m} with the following properties:

  • (1)

    in the basis \mathcal{F}, the form QQ is equal to IrIs-I_{r}\oplus I_{s} and the special orthogonal group of QQ is equal to SO(r,s)\operatorname{SO}(r,s);

  • (2)

    writing x¯\overline{x} for the coordinates with respect to \mathcal{F} and defining w¯\overline{w} as in §3.1, there exists real constants c,κ>0c,\kappa>0, independent of w𝔊vw\in\mathfrak{G}\cdot v, such that

    |w|1+κ\displaystyle|w_{\ell}|^{-1}+\kappa >c|wi|,\displaystyle>c|w_{i}|, 1is,\displaystyle 1\leq i\leq s, (8)
    κ\displaystyle\kappa >|wi|,\displaystyle>|w_{i}|, s+1im.\displaystyle s+1\leq i\leq m. (9)

    if =(w)\ell=\ell(w) is defined, and w<κ\|w\|<\kappa otherwise.

Proof.

Define γ\gamma and σ\sigma as in 3.2 so that γ𝔊γ1σ𝔖\gamma\mathfrak{G}\gamma^{-1}\sigma\subset\mathfrak{S}, where 𝔖\mathfrak{S} is one of the standard Siegel sets constructed above. We define \mathcal{F} as γ\gamma\cdot\mathcal{E}, where ={e1,,em}\mathcal{E}=\{e_{1},\ldots,e_{m}\} is the standard basis. Then we have

(γ𝔊γ1)γv𝔖σ1γv.(\gamma\mathfrak{G}\gamma^{-1})\gamma v\subset\mathfrak{S}\sigma^{-1}\gamma v.

The left-hand side is nothing more than the orbit 𝔊v\mathfrak{G}\cdot v except in the coordinates determined by \mathcal{F}, so it suffices to prove the inequalities for the points inside the right-hand side. The vector σ1γv\sigma^{-1}\gamma v is a real Hodge vector for the Hodge structure σ1γh\sigma^{-1}\gamma\cdot h. Moreover because σKσ1=γ1Khγ\sigma K\sigma^{-1}=\gamma^{-1}K_{h}\gamma, the maximal compact associated to 𝔖\mathfrak{S} agrees with the stabilizer of the Hodge metric for σ1γh\sigma^{-1}\gamma\cdot h. It thus suffices to prove the statement of the proposition when 𝔊=𝔖\mathfrak{G}=\mathfrak{S} is a standard Siegel set which is also a Hodge-theoretic Siegel set for hh.

We thus reduce to considering the orbit ΩAtKv\Omega A_{t}K\cdot v, with KK as in the previous section. Let m=Hp,q\mathbb{C}^{m}=\bigoplus H^{p,q} be the Hodge decomposition associated to hh, where we assume the Hodge structure has weight w=p+qw=p+q where w=2kw=2k is even. Define Ve=p evenHp,wpV_{e}=\bigoplus_{p\textrm{ even}}H^{p,w-p} and Vo=p oddHp,wpV_{o}=\bigoplus_{p\textrm{ odd}}H^{p,w-p}, which we may view as subspaces of m\mathbb{R}^{m}. The Weil operator ChC_{h} acts as i2pw=(1)pki^{2p-w}=(-1)^{p-k} on Hp,qH^{p,q}, so the forms QQ and the Hodge metric QhQ_{h} agree on one of the two summands {Ve,Vo}\{V_{e},V_{o}\} and agree up to a sign on the other.

The maximal compact group associated to hh preserves both VeV_{e} and VoV_{o}. Since we have reduced to the setting where the maximal compact associated to hh is K=S(O(r)×O(s))K=S(O(r)\times O(s)), it follows that {Ve,Vo}={span{e1,,er},span{er+1,,em}}\{V_{e},V_{o}\}=\{\textrm{span}\{e_{1},\ldots,e_{r}\},\textrm{span}\{e_{r+1},\ldots,e_{m}\}\}. Then because vv is Hodge, hence vVev\in V_{e}, we observe that all the vectors in KvK\cdot v lie inside VeV_{e} as well. In particular, if kvKvkv\in K\cdot v, then [kv]i=±[kv]r+i[kv]_{i}=\pm[kv]_{r+i} for 1is1\leq i\leq s in w¯\overline{w}-coordinates, with the sign depending on which of the two summands in {span{e1,,er},span{er+1,,em}}\{\textrm{span}\{e_{1},\ldots,e_{r}\},\textrm{span}\{e_{r+1},\ldots,e_{m}\}\} the space VeV_{e} is identified with. We will assume the sign is positive, with the other case handled analogously.

We now consider a vector wΩAtKvw\in\Omega A_{t}K\cdot v which we may write as w=ωakvw=\omega ak\cdot v with kKk\in K, aAta\in A_{t}, and ωΩ\omega\in\Omega. After possibly adjusting aa and ω\omega and decreasing tt, we may assume that all diagonal entries of ω\omega are equal to 11 except for possibly those that lie in the central block. We write aia_{i} for the ii’th diagonal entry of aa in the w¯\overline{w}-coordinate system. We note that

[akv]i=ai[kv]i=ai[kv]r+i=ai2[akv]r+i[ak\cdot v]_{i}=a_{i}[k\cdot v]_{i}=a_{i}[k\cdot v]_{r+i}=a^{2}_{i}[ak\cdot v]_{r+i}

for 1is1\leq i\leq s. We may assume 0=(kv)\ell_{0}=\ell(k\cdot v) is defined: otherwise [akv]=[kv][ak\cdot v]=[k\cdot v] and the entries of [ωakv]=[ωkv][\omega ak\cdot v]=[\omega k\cdot v] are universally bounded as both ω\omega and kk range over a compact set.

We then have

[akv]01=[kv]0r2[akv]0r.\displaystyle[ak\cdot v]^{-1}_{\ell_{0}}=[k\cdot v]^{-2}_{\ell_{0}-r}[ak\cdot v]_{\ell_{0}-r}.

Choose jj such that 0jr+i\ell_{0}\leq j\leq r+i and [kv]jr0[k\cdot v]_{j-r}\neq 0. From the inequalities defining AtA_{t} we then get

[akv]01tj0[kv]0r2[kv]jr1[akv]jr,\displaystyle[ak\cdot v]^{-1}_{\ell_{0}}\geq t^{j-\ell_{0}}[k\cdot v]^{-2}_{\ell_{0}-r}[k\cdot v]^{-1}_{j-r}[ak\cdot v]_{j-r},

We now take cc such that 0<c[kv]0r1[kv]jr1tj00<c\leq[k\cdot v]^{-1}_{\ell_{0}-r}[k\cdot v]^{-1}_{j-r}t^{j-\ell_{0}}, which we can do independently of ww and ii since [kv]0r1[kv]jr1[k\cdot v]^{-1}_{\ell_{0}-r}[k\cdot v]^{-1}_{j-r} is bounded from below independently of ww. Taking j=r+ij=r+i in the above gives

[akv]01c[akv]i,\displaystyle[ak\cdot v]^{-1}_{\ell_{0}}\geq c[ak\cdot v]_{i},

for all 1is1\leq i\leq s. One computes immediately from the description of PP in (7) that (ωakv)=(kv)=0\ell(\omega ak\cdot v)=\ell(k\cdot v)=\ell_{0} and that [ωakv]0=[akv]0[\omega ak\cdot v]_{\ell_{0}}=[ak\cdot v]_{\ell_{0}}; in particular (w)=(ωakv)\ell(w)=\ell(\omega ak\cdot v) is defined. It thus follows that

[w]1c[akv]i,[w]^{-1}_{\ell}\geq c[ak\cdot v]_{i}, (10)

for all 1is1\leq i\leq s.

Now we also have asta_{s}\geq t, and hence aitsi+1a_{i}\geq t^{s-i+1} for all 1is1\leq i\leq s. This implies that ar+i=ai1tsi+1a_{r+i}=a^{-1}_{i}\leq t^{s-i+1} for all 1is1\leq i\leq s, in particular, there is an absolute bound κ\kappa such that

κ>[akv]i\kappa>[ak\cdot v]_{i}

for all ir+1i\geq r+1. We can even assume this is true for all is+1i\geq s+1 using that AtA_{t} acts trivially on the coordinates indexed by s+1,,rs+1,\ldots,r.

We conclude that for all 1im1\leq i\leq m we have

[w]1+κ\displaystyle[w]^{-1}_{\ell}+\kappa c[akv]i,\displaystyle\geq c[ak\cdot v]_{i}, 1is\displaystyle 1\leq i\leq s
κ\displaystyle\kappa >[akv]i,\displaystyle>[ak\cdot v]_{i}, s+1im.\displaystyle s+1\leq i\leq m.

Then since ω\omega lies in a compact subset ΩP()\Omega\subset P(\mathbb{R}), the inequalities continue to hold with [ωakv]i[\omega ak\cdot v]_{i} replacing [akv]i[ak\cdot v]_{i} after possibly adjusting cc and κ\kappa. ∎

3.3 Point Counting Upper Bounds

We continue with the notation and setup of the preceding section. We view m\mathbb{R}^{m} as the set of real points of the standard mm-dimensional affine space 𝔸m\mathbb{A}^{m}.

Proposition 3.4.

Fix a Siegel set 𝔊\mathfrak{G} for (Ph,Sh,Kh)(P_{h},S_{h},K_{h}) and a real Hodge vector vv for hh. Then there exists a real constant ρ\rho such that

|(𝔊v)(1νm)|ρνm+s2\left|(\mathfrak{G}\cdot v)\cap\left(\frac{1}{\nu}\mathbb{Z}^{m}\right)\right|\lesssim\rho\cdot\nu^{m+s-2}

as ν\nu\to\infty.

Proof.

The statement is unchanged (up to adjusting ρ\rho) after a \mathbb{Q}-linear change of coordinates, so we may work in the coordinate system w¯\overline{w} of 3.3. In the region of 𝔊v\mathfrak{G}\cdot v where (w)\ell(w) is undefined we then know that w\|w\| 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 =(w)\ell=\ell(w). Rearranging these inequalities one obtains

κ\displaystyle\kappa^{\prime} >c|wi||w|,\displaystyle>c|w_{i}||w_{\ell}|, 1is,\displaystyle 1\leq i\leq s, (11)
κ\displaystyle\kappa >|wi|,\displaystyle>|w_{i}|, s+1im,\displaystyle s+1\leq i\leq m, (12)

where we take κ>1+κ2>1+κ|w|\kappa^{\prime}>1+\kappa^{2}>1+\kappa|w_{\ell}| using that κ>|w|\kappa>|w_{\ell}| from (12).

For each index r+1mr+1\leq\ell\leq m we define a map τ:𝔸m𝔸m1\tau_{\ell}:\mathbb{A}^{m}\to\mathbb{A}^{m-1} via the coordinate functions

(w1w,,wr1w,wr+1w,,wsw,ws+1,,wm).(w_{1}w_{\ell},\ldots,w_{\ell-r-1}w_{\ell},w_{\ell-r+1}w_{\ell},\ldots,w_{s}w_{\ell},w_{s+1},\ldots,w_{m}). (13)

Letting X𝔸mX_{\ell}\subset\mathbb{A}^{m} be the locus where w0w_{\ell}\neq 0, we observe that the induced map

X(SO(r,s)v)τ𝔸m1X_{\ell}\cap(\operatorname{SO}(r,s)\cdot v)\xrightarrow{\tau_{\ell}}\mathbb{A}^{m-1}

is injective. Indeed the locus SO(r,s)v\operatorname{SO}(r,s)\cdot v is defined by an equation

i=1swiwr+i+j=1rswj2=c\sum_{i=1}^{s}w_{i}w_{r+i}+\sum_{j=1}^{r-s}w_{j}^{2}=c (14)

for some constant cc. Given the coordinates (13) and the condition that w0w_{\ell}\neq 0, one recovers by dividing by ww_{\ell} all the coordinates wiw_{i} except wrw_{\ell-r}. But then one can determine each term appearing in (14) except for wrww_{\ell-r}w_{\ell}, so one can again divide through by ww_{\ell} and recover wrw_{\ell-r}.

The problem reduces to bounding rational points in the image of τ\tau_{\ell}. From the inequalities (11) and (12) we see that τ(𝔊vX())\tau_{\ell}(\mathfrak{G}\cdot v\cap X_{\ell}(\mathbb{R})) lies inside some compact set 𝒦\mathcal{K}, and the rational points in 1νm\frac{1}{\nu}\mathbb{Z}^{m} have image inside 1ν2s11νms\frac{1}{\nu^{2}}\mathbb{Z}^{s-1}\oplus\frac{1}{\nu}\mathbb{Z}^{m-s}. The number of such points is then at most ρν2(s1)νms=ρνm+s2\rho\cdot\nu^{2(s-1)}\nu^{m-s}=\rho\cdot\nu^{m+s-2} choosing ρ\rho large enough. ∎

Proposition 3.5.

Fix a Siegel set 𝔊\mathfrak{G} for (Ph,Sh,Kh)(P_{h},S_{h},K_{h}) and a real Hodge vector vv for hh. Suppose that Y𝔸mY\subset\mathbb{A}^{m} is an irreducible \mathbb{Q}-algebraic subvariety. Then there exists a real constant cc, dependent on YY, such that

|(𝔊v)Y()(1νm)|cν2dimY\left|(\mathfrak{G}\cdot v)\cap Y(\mathbb{Q})\cap\left(\frac{1}{\nu}\mathbb{Z}^{m}\right)\right|\lesssim c\cdot\nu^{2\dim Y}

as ν\nu\to\infty.

Proof.

Once again we work in the coordinate system w¯\overline{w} of 3.3. We write Vi𝔸mV_{i}\subset\mathbb{A}^{m} for the locus where wiw_{i} vanishes, and let Xi=𝔸mViX_{i}=\mathbb{A}^{m}\setminus V_{i} be the complement. Let E[m]={1,,m}E\subset[m]=\{1,\ldots,m\} be the subset of indices ii such that YViY\subset V_{i}. Then by arguing inductively on YVjY\cap V_{j} for j[m]Ej\in[m]\setminus E, we may reduce to considering just those rational points in Y:=Yj[m]EXjY^{\prime}:=Y\cap\bigcap_{j\in[m]\setminus E}X_{j}.

We may assume ([m]E){r+1,,m}([m]\setminus E)\cap\{r+1,\ldots,m\} is non-empty. Otherwise (w)\ell(w) is undefined for each wY()w\in Y(\mathbb{R}), and the desired result follows by projecting (𝔊v)Y()(\mathfrak{G}\cdot v)\cap Y(\mathbb{Q}) to a coordinate hyperplane and using that w\|w\| is absolutely bounded. Then we let \ell be the smallest entry of ([m]E){r+1,,m}([m]\setminus E)\cap\{r+1,\ldots,m\}.

We consider a projection τ:𝔸m𝔸dimY\tau^{\prime}_{\ell}:\mathbb{A}^{m}\to\mathbb{A}^{\dim Y} defined by composing the map τ\tau_{\ell} defined in the proof of 3.4 with a further projection 𝔸m1𝔸dimY\mathbb{A}^{m-1}\to\mathbb{A}^{\dim Y} onto a coordinate hyperplane and a translation 𝔸dimY𝔸dimY\mathbb{A}^{\dim Y}\to\mathbb{A}^{\dim Y} by an integral vector in dimY\mathbb{Z}^{\dim Y}. We may choose these projections and translations such that Y𝔸dimYY\to\mathbb{A}^{\dim Y} is quasi-finite away from a closed \mathbb{Q}-algebraic sublocus of YY which we handle inductively. For the remaining points we are reduced by taking the image under τ\tau^{\prime}_{\ell} to counting points in 1ν2dimY\frac{1}{\nu^{2}}\mathbb{Z}^{\dim Y} which lie in a compact region, so the result follows. ∎

Proof of (1.24):.

This follows from 3.4 and 3.5 after recalling that we have adopted the convention that rsr\geq s throughout §3, and replacing QQ by Q-Q does not change the group preserving the form. ∎

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 n\mathbb{R}^{n} into definable subsets, called cells, defined inductively as follows (c.f. [vdD98, Ch. 3, 2.3]):

  • (i)

    when n=1n=1, the cells are open intervals and singleton sets;

  • (ii)

    given a cell Dn1D\subset\mathbb{R}^{n-1}, a cell of n\mathbb{R}^{n} is either:

    • -

      the graph of a continuous definable function f:Df:D\to\mathbb{R} viewed in the natural way as a subset of n\mathbb{R}^{n}; or

    • -

      the region in n\mathbb{R}^{n} defined by

      {(x,y)D×:f(x)<y<g(x)}\{(x,y)\in D\times\mathbb{R}:f(x)<y<g(x)\}

      where f,g:Df,g:D\to\mathbb{R} are continuous definable functions satisfying f<gf<g.

Then one has [vdD98, Ch. 3, 2.11]:

Theorem 4.1.

For any definable subset FnF\subset\mathbb{R}^{n} there exists a cell decomposition of n\mathbb{R}^{n} such that FF is a finite union of cells.

We will find the following lemma useful in later arguments.

Lemma 4.2.

Let AA and BB be definable sets, and FA×BF\subset A\times B a definable subset such that the intersections F{a}×BF\cap\{a\}\times B are all finite. Then there exists a definable partition F=F1FkF=F_{1}\sqcup\cdots\sqcup F_{k} such that the intersections Fi{a}×BF_{i}\cap\{a\}\times B have cardinality at most 11 for each aa and ii.

Proof.

We may reduce to the case where A=nA=\mathbb{R}^{n} and B=mB=\mathbb{R}^{m}, so FF is a definable subset of n+m\mathbb{R}^{n+m}. We then consider a cell decomposition of (F,n+m)(F,\mathbb{R}^{n+m}). We let DFD\subset F be a cell, and prove that D{a}×mD\cap\{a\}\times\mathbb{R}^{m} has cardinality at most 11 for each bb. From the definition, the cell DD is obtained as a sequence D1,D2,,Dn,Dn+1,,Dn+mD_{1},D_{2},\ldots,D_{n},D_{n+1},\ldots,D_{n+m} where DiD_{i} is a cell in i\mathbb{R}^{i} and Di+1D_{i+1} is obtained from DiD_{i} in one of the two ways described in (ii) above.

For each 1km1\leq k\leq m we have a natural projection

πk:Dn+k{a}×kDn+(k1){a}×k1.\pi_{k}:D_{n+k}\cap\{a\}\times\mathbb{R}^{k}\to D_{n+(k-1)}\cap\{a\}\times\mathbb{R}^{k-1}. (15)

For k=1k=1 this is just a map Dn+1{a}×{a}D_{n+1}\cap\{a\}\times\mathbb{R}\to\{a\}. We observe that all such maps are necessarily surjective, which is immediate from the two possibilities for the construction of Dn+kD_{n+k} from Dn+(k1)D_{n+(k-1)} in (ii). Thus the hypothesis that Dn+m{a}×mD_{n+m}\cap\{a\}\times\mathbb{R}^{m} is finite in fact implies that Dn+k{a}×k{a}D_{n+k}\cap\{a\}\times\mathbb{R}^{k}\to\{a\} is finite for all 0km0\leq k\leq m, since each map Dn+m{a}×mDn+k{a}×kD_{n+m}\cap\{a\}\times\mathbb{R}^{m}\to D_{n+k}\cap\{a\}\times\mathbb{R}^{k} is surjective as it is a composition of surjective maps.

It then suffices to prove inductively on 0km0\leq k\leq m that Dn+k{a}×kD_{n+k}\cap\{a\}\times\mathbb{R}^{k} is a singleton, which at each stage amounts to proving that the (necessarily unique, by induction) fibre of πk\pi_{k} is a singleton. Notice the finiteness of Dn+m{a}×kD_{n+m}\cap\{a\}\times\mathbb{R}^{k} means that Dn+kD_{n+k} cannot be obtained from Dn+(k1)D_{n+(k-1)} by a construction of the second type in (ii) above, since then Dn+k{a}×kD_{n+k}\cap\{a\}\times\mathbb{R}^{k} would contain the infinite set

{(x0,y):f(x0)<y<g(x0)}\{(x_{0},y):f(x_{0})<y<g(x_{0})\}

for the (necessarily unique, by induction) point x0Dn+(k1){a}×k1x_{0}\in D_{n+(k-1)}\cap\{a\}\times\mathbb{R}^{k-1}.

Hence Dn+kD_{n+k} is obtained from Dn+(k1)D_{n+(k-1)} by a construction of the form

Dn+k={(x,y)Dn+(k1)×:y=f(x)}.D_{n+k}=\{(x,y)\in D_{n+(k-1)}\times\mathbb{R}:y=f(x)\}.

But then the induction hypothesis that Dn+(k1){a}×k1={x0}D_{n+(k-1)}\cap\{a\}\times\mathbb{R}^{k-1}=\{x_{0}\} is a singleton implies that Dn+k{a}×k={(x0,f(x0))}D_{n+k}\cap\{a\}\times\mathbb{R}^{k}=\{(x_{0},f(x_{0}))\}, so the result follows. ∎

Corollary 4.3.

Let f:BAf:B\to A be a definable function with finite fibres. Then there exists a definable partition B=B1BkB=B_{1}\sqcup\cdots\sqcup B_{k} such that BiAB_{i}\to A is injective.

Proof.

Let FB×AF\subset B\times A be the graph of ff. Then after swapping the order of the coordinates, we obtain from 4.2 a definable partition F=F1FkF=F_{1}\sqcup\cdots\sqcup F_{k} such that FiB×{a}F_{i}\cap B\times\{a\} has cardinality at most 11 for each ii and aa. Define BiB_{i} to be the projection of FiF_{i} and observe that FiF_{i} is just the graph of BiAB_{i}\to A. ∎

4.2 Stripes

4.2.1 Parametrization of Stripes

We work in the setup of §2.1. We additionally fix a fibre V=𝕍s0V=\mathbb{V}_{s_{0}}, the choice of which is unimportant, and identify (G,D)=(𝐆S,DS)(G,D)=(\mathbf{G}_{S},D_{S}) with their realizations at s0s_{0}; in particular, GG is a subgroup of GL(V)\operatorname{GL}(V) which lies in the group SAut(V,Q)\textrm{SAut}(V,Q) of linear mappings which preserve the polarizing form QQ on VV and have determinant one. Fix a Hodge-theoretic Siegel set 𝔒=ΩAtKooD\mathfrak{O}=\Omega A_{t}K_{o}\cdot o\subset D associated to (P,o,t)(P,o,t), and set 𝔊=ΩAtKo\mathfrak{G}=\Omega A_{t}K_{o}. We will denote the natural orbit map G()DG(\mathbb{R})\to D given by ggog\mapsto g\cdot o by qq. In what follows all definable sets are regarded as definable relative to the structure an,exp\mathbb{R}_{\textrm{an,exp}}. We let hDh\in D be a point with Mumford-Tate group MhGM_{h}\subset G, and fix another point oDM=Mh()ho\in D_{M}=M_{h}(\mathbb{R})\cdot h. Then oo has Mumford-Tate group contained in M:=MhM:=M_{h}. We write Dˇ\widecheck{{D}} for the compact dual of DD, which is a projective algebraic variety containing DD as an open submanifold. It can be identified with the orbit of G()G(\mathbb{C}) on FoF^{\bullet}_{o} in the space of all Hodge flags on VV, where FoF^{\bullet}_{o} is the Hodge flag corresponding to oo.

Notation.

Suppose XX is a \mathbb{Q}-algebraic variety and 𝔖X()\mathfrak{S}\subset X(\mathbb{R}) is a subset. We write 𝔖()\mathfrak{S}(\mathbb{Q}) for the intersection 𝔖X()\mathfrak{S}\cap X(\mathbb{Q}).

Notation.

Given a group GG with subgroup HH, we write 𝒩G(H)\mathcal{N}_{G}(H) (resp. 𝒞G(H)\mathcal{C}_{G}(H)) for the normalizer (resp. centralizer) in GG of HH. If both GG and HH are algebraic groups, we interpret 𝒩G(H)\mathcal{N}_{G}(H) (resp. 𝒞G(H)\mathcal{C}_{G}(H)) as an algebraic group.

Notation.

For an algebraic group GG, we write GG^{\circ} for its identity component.

Remark 4.4.

If HH is a connected reductive subgroup of a connected reductive group GG, then H𝒞G(H)=𝒩G(H)H\cdot\mathcal{C}_{G}(H)^{\circ}=\mathcal{N}_{G}(H)^{\circ} (c.f. [hh]). This is often useful for understanding 𝒩𝐆S(𝐌)\mathcal{N}_{\mathbf{G}_{S}}(\mathbf{M}) when 𝐌\mathbf{M} is a Mumford-Tate subgroup of 𝐆S\mathbf{G}_{S}.

We set DM=M()oD_{M}=M(\mathbb{R})\cdot o to be the M()M(\mathbb{R})-orbit of oo in DD, and DˇM=M()FoDˇ\widecheck{{D}}_{M}=M(\mathbb{C})\cdot F^{\bullet}_{o}\subset\widecheck{{D}} to be the corresponding orbit in Dˇ\widecheck{{D}}, where FoF^{\bullet}_{o} is the Hodge flag corresponding to oo. We write NLˇMDˇ\widecheck{{\operatorname{NL}}}_{M}\subset\widecheck{{D}} for the \mathbb{Q}-algebraic locus considering all Hodge flags whose set of Hodge tensors contains those tensors fixed by MM (c.f. [GGK12, Ch. 2]).

Proposition 4.5.

Let (DM)\mathcal{F}(D_{M}) be the Zariski closure in GG of

{gG():gDMDM},\{g\in G(\mathbb{R}):gD_{M}\subset D_{M}\},

and (DˇM)\mathcal{F}(\widecheck{{D}}_{M}) the algebraic group defined by

(DˇM)():={gG():gDˇMDˇM}.\mathcal{F}(\widecheck{{D}}_{M})(\mathbb{C}):=\{g\in G(\mathbb{C}):g\widecheck{{D}}_{M}\subset\widecheck{{D}}_{M}\}.

Then if II is the identity component of 𝒩G(M)\mathcal{N}_{G}(M), we have

I(DM)(DˇM).I\subset\mathcal{F}(D_{M})\subset\mathcal{F}(\widecheck{{D}}_{M}). (16)

Moreover, the identity components of all three groups agree.

Proof.

We first observe that gDMDMgD_{M}\subset D_{M} implies gDˇMDˇMg\widecheck{{D}}_{M}\subset\widecheck{{D}}_{M} because DMD_{M} is open in the irreducible variety DˇM\widecheck{{D}}_{M}. This implies the latter inclusion (DM)(DˇM)\mathcal{F}(D_{M})\subset\mathcal{F}(\widecheck{{D}}_{M}). For the former we use [GGK12, VI.A.3] to obtain that 𝒩G(M)DˇMNLˇM\mathcal{N}_{G}(M)\widecheck{{D}}_{M}\subset\widecheck{{\operatorname{NL}}}_{M}, and note that by [GGK12, VI.B.1] DˇM\widecheck{{D}}_{M} is a (geometrically) connected component of NLˇM\widecheck{{\operatorname{NL}}}_{M}. As a consequence of [GGK12, VI.B.1], the (geometrically) irreducible components of NLˇM\widecheck{{\operatorname{NL}}}_{M} agree with the (geometrically) connected components. This implies that geometrically connected group II preserves DˇM\widecheck{{D}}_{M}. Then if gI()g\in I(\mathbb{R}) one necessarily has that gDMDDˇM=DMgD_{M}\subset D\cap\widecheck{{D}}_{M}=D_{M} (using [GGK12, VI.B.11]), hence I()(DM)()I(\mathbb{R})\subset\mathcal{F}(D_{M})(\mathbb{R}). But I()I(\mathbb{R}) is Zariski dense in the connected \mathbb{Q}-group II 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 (DM)():=(DM)()(DM)()\mathcal{F}(D_{M})(\mathbb{Q})^{\circ}:=\mathcal{F}(D_{M})(\mathbb{Q})\cap\mathcal{F}(D_{M})^{\circ}(\mathbb{R}) is Zariski dense in (DM)\mathcal{F}(D_{M})^{\circ}. Now consider an element g(DM)()g\in\mathcal{F}(D_{M})(\mathbb{Q})^{\circ}. We have gDMDDˇM=DMgD_{M}\subset D\cap\widecheck{{D}}_{M}=D_{M}, and so gg sends a generic point of DMD_{M} with Mumford-Tate group MM to another such point. Applying [GGK12, VI.A.3] this tells us that gNLM()g\in\operatorname{NL}_{M}(\mathbb{R}). Then necessarily gI()g\in I(\mathbb{R}) since (DM)\mathcal{F}(D_{M})^{\circ} is connected. The analogous argument works for (DˇM)\mathcal{F}(\widecheck{{D}}_{M}). ∎

Definition 4.6.

A \mathbb{Q}-subgroup H𝒩G(M)H\subset\mathcal{N}_{G}(M) is said to generate DMD_{M} if H()o=DMH(\mathbb{R})\cdot o=D_{M}.

We set V=a,b0Va(V)bV^{\otimes}=\bigoplus_{a,b\geq 0}V^{\otimes a}\otimes(V^{*})^{\otimes b}. Each flag FDˇF^{\bullet}\in\widecheck{{D}} induces a filtration on VV^{\otimes} which we denote with the same notation.

Lemma 4.7.

Suppose that H𝒩G(M)H\subset\mathcal{N}_{G}(M) generates DMD_{M} and consider a point gH(G/H)()gH\in(G/H)(\mathbb{Q}) such that Dg=gDˇMDD_{g}=g\widecheck{{D}}_{M}\cap D is non-empty. Then there exists a Hodge structure hgDgh_{g}\in D_{g} with Mumford-Tate group MggMg1M_{g}\subset gMg^{-1} such that Dg=DMg:=Mg()hgD_{g}=D_{M_{g}}:=M_{g}(\mathbb{R})\cdot h_{g}. If gHgH admits a representative gG()g\in G(\mathbb{R}), one has Dg=gDMD_{g}=gD_{M} and Mg=gMg1M_{g}=gMg^{-1}.

Proof.

Note that because H𝒩G(M)H\subset\mathcal{N}_{G}(M), the group gMg1gMg^{-1} is independent of the representative gg chosen. Thus gMg1gMg^{-1} is \mathbb{Q}-algebraic because gHgH is. Let gFgF^{\bullet} be a Hodge flag corresponding to a point hgDgh_{g}\in D_{g}, where FF^{\bullet} is a Hodge flag corresponding to a point of DˇM\widecheck{{D}}_{M}. Write Fix(M)V\textrm{Fix}(M)\subset V^{\otimes}_{\mathbb{C}} for the vector subspace of tensors fixed by MM. Choosing F0F^{0} sufficiently generic, this is exactly the \mathbb{C}-span of F0VF^{0}\cap V^{\otimes}_{\mathbb{Q}}. One has gFix(M)=Fix(gMg1)g\textrm{Fix}(M)=\textrm{Fix}(gMg^{-1}). As the group gMg1gMg^{-1} is defined over \mathbb{Q}, the complex vector space gFix(M)g\textrm{Fix}(M) has a natural underlying \mathbb{Q}-structure [gFix(M)][g\textrm{Fix}(M)]_{\mathbb{Q}}, and by construction [gFix(M)]gF0V[g\textrm{Fix}(M)]_{\mathbb{Q}}\subset gF^{0}\cap V^{\otimes}_{\mathbb{Q}}. Then the Mumford-Tate group MgM_{g} of hgh_{g} fixes all tensors in gF0VgF^{0}\cap V^{\otimes}_{\mathbb{Q}}, hence lies inside gMg1gMg^{-1}. Taking hgh_{g} sufficiently generic and using that gDˇMg\widecheck{{D}}_{M} is irreducible, the points of DgD_{g} have Mumford-Tate group contained in a common \mathbb{Q}-algebraic Mumford-Tate group MggMg1M_{g}\subset gMg^{-1}. Let DMg=Mg()hgD_{M_{g}}=M_{g}(\mathbb{R})\cdot h_{g}. We then have

DgDNLˇMg=DMg,D_{g}\subset D\cap\widecheck{{\operatorname{NL}}}_{M_{g}}=D_{M_{g}},

where we have applied [GGK12, VI.B.11].

Now gNLˇMg\widecheck{{\operatorname{NL}}}_{M} is exactly the locus of Hodge flags whose set of Hodge tensors contains [gFix(M)][g\textrm{Fix}(M)]_{\mathbb{Q}}. It thus follows that NLˇMggNLˇM\widecheck{{\operatorname{NL}}}_{M_{g}}\subset g\widecheck{{\operatorname{NL}}}_{M}. For both the varieties NLˇMg\widecheck{{\operatorname{NL}}}_{M_{g}} and gNLˇMg\widecheck{{\operatorname{NL}}}_{M} their geometrically irreducible components agree with their geometrically connected components (this is a consequence of [GGK12, VI.B.1]), so the inclusion NLˇMggNLˇM\widecheck{{\operatorname{NL}}}_{M_{g}}\subset g\widecheck{{\operatorname{NL}}}_{M} also implies an inclusion DˇMggDˇM\widecheck{{D}}_{M_{g}}\subset g\widecheck{{D}}_{M} of components passing through hgh_{g}. Then

DMg=DDˇMgDgDˇM=Dg.D_{M_{g}}=D\cap\widecheck{{D}}_{M_{g}}\subset D\cap g\widecheck{{D}}_{M}=D_{g}.

It remains to show that Mg=gMg1M_{g}=gMg^{-1} and that Dg=gDMD_{g}=gD_{M} under the assumption gG()g\in G(\mathbb{R}). For the second claim, observe that

g1Dg=g1(gDˇMD)=DˇMD=DM,g^{-1}D_{g}=g^{-1}(g\widecheck{{D}}_{M}\cap D)=\widecheck{{D}}_{M}\cap D=D_{M},

where we again apply [GGK12, VI.B.11]. Hence gDM=DggD_{M}=D_{g}. Finally g1Mggg^{-1}M_{g}g is the generic Mumford-Tate group of a point in g1Dg=DMg^{-1}D_{g}=D_{M}, hence g1Mgg=Mg^{-1}M_{g}g=M. ∎

Definition 4.8.

A Mumford-Tate domain DgD_{g} obtained as in 4.7 is called an HH-translate of DMD_{M}.

We thus obtain a map

η:(G/H)(){H-translates of DM}\eta:(G/H)(\mathbb{Q})\to\{H\textrm{-translates of }D_{M}\}

which sends gHgH to DMgD_{M_{g}}. We write (G/H)()(G/H)()(G/H)(\mathbb{R})^{\circ}\subset(G/H)(\mathbb{R}) for the connected component containing HH. We then write η\eta^{\circ} for the restriction of η\eta to (G/H)()=(G/H)()(G/H)()(G/H)(\mathbb{Q})^{\circ}=(G/H)(\mathbb{Q})\cap(G/H)(\mathbb{R})^{\circ}.

Lemma 4.9.

Each fibre of η\eta lies inside a fibre of (G/H)()(G/(DˇM))()(G/H)(\mathbb{C})\to(G/\mathcal{F}(\widecheck{{D}}_{M}))(\mathbb{C}). Each fibre of η\eta^{\circ} lies inside a fibre of G()/H()G()/(DM)()G(\mathbb{R})/H(\mathbb{R})\to G(\mathbb{R})/\mathcal{F}(D_{M})(\mathbb{R}).

Proof.

If η(gH)=η(gH)\eta(gH)=\eta(g^{\prime}H) then gDˇM=gDˇMg\widecheck{{D}}_{M}=g^{\prime}\widecheck{{D}}_{M} which implies that g1g((DˇM))()g^{\prime-1}g\in(\mathcal{F}(\widecheck{{D}}_{M}))(\mathbb{C}). Similarly if η(gH)=η(gH)\eta^{\circ}(gH)=\eta^{\circ}(g^{\prime}H) then gDM=gDMgD_{M}=g^{\prime}D_{M} and hence g1g(DM)()g^{\prime-1}g\in\mathcal{F}(D_{M})(\mathbb{R}). ∎

Definition 4.10.

Suppose that 𝔗𝔒\mathfrak{T}\subset\mathfrak{O} is a subset containing oo, and that HGH\subset G is a \mathbb{Q}-algebraic subgroup which generates DMD_{M}. By a stripe of HH in 𝔗\mathfrak{T} we mean a non-empty intersection η(gH)𝔗\eta(gH)\cap\mathfrak{T} for some gH(G/H)()gH\in(G/H)(\mathbb{Q}). A stripe is said to be principal if we can choose gH(G/H)()(G/H)()gH\in(G/H)(\mathbb{Q})\cap(G/H)(\mathbb{R})^{\circ}.

Definition 4.11.

Given a subset 𝔗𝔒\mathfrak{T}\subset\mathfrak{O} containing oo and a \mathbb{Q}-subgroup HGH\subset G generating DMD_{M}, we say that a stripe

g:=η(gH)𝔗=DMg𝔗\mathfrak{H}_{g}:=\eta(gH)\cap\mathfrak{T}=D_{M_{g}}\cap\mathfrak{T}

of HH in 𝔗\mathfrak{T} is generic if there is a point of g\mathfrak{H}_{g} with Mumford-Tate group MgM_{g}. We say it is uniformly generic if every connected component of g\mathfrak{H}_{g} has such a point.

Lemma 4.12.

In the context of 4.11, the genericity of the stripe g\mathfrak{H}_{g} does not depend on gg; if g=g\mathfrak{H}_{g}=\mathfrak{H}_{g^{\prime}} and there exists a point tgt\in\mathfrak{H}_{g} with Mumford-Tate group MgM_{g}, then in fact Mg=MgM_{g}=M_{g^{\prime}} and DMg=DMgD_{M_{g}}=D_{M_{g^{\prime}}}.

Proof.

Let tgDMgt\in\mathfrak{H}_{g}\subset D_{M_{g}} be a point with Mumford-Tate group equal to MgM_{g}. Then tgDMgt\in\mathfrak{H}_{g^{\prime}}\subset D_{M_{g^{\prime}}} so MgMgM_{g}\subset M_{g^{\prime}}. On the other hand dimDMg=dimDM=dimDMg\dim D_{M_{g}}=\dim D_{M}=\dim D_{M_{g^{\prime}}}, so Mg()tM_{g}(\mathbb{R})\cdot t is open in DMgD_{M_{g^{\prime}}}. Since any open subset of DMgD_{M_{g^{\prime}}} contains a Hodge-generic point it follows that Mg=MgM_{g}=M_{g^{\prime}}. Likewise, DMg=DMgD_{M_{g}}=D_{M_{g^{\prime}}}. ∎

We now obtain a diagram

{generic stripes of H in 𝔗}{\{\textrm{generic stripes of }H\textrm{ in }\mathfrak{T}\}}(G/H)(){(G/H)(\mathbb{Q})}{H-translates of DM},{\{H\textrm{-translates of }D_{M}\},}β=β𝔗\scriptstyle{\beta=\beta_{\mathfrak{T}}}η\scriptstyle{\eta} (17)

where the map β\beta sends a stripe g\mathfrak{H}_{g} to the (necessarily unique by 4.12) HH-translate of DMD_{M} which induces it. We wish to understand the inverse image η1(im(β))\eta^{-1}(\textrm{im}(\beta)) by constructing, using 𝔗\mathfrak{T}, a subset of (G/H)()(G/H)(\mathbb{R}) which contains it. For this we define

:=im[q1(𝔗)(G/H)()](G/H)()(G/H)().\displaystyle\mathfrak{Z}:=\textrm{im}[q^{-1}(\mathfrak{T})\to(G/H)(\mathbb{R})]\cap(G/H)(\mathbb{R})^{\circ}\subset(G/H)(\mathbb{R}).
Lemma 4.13.

The restriction of η\eta to ()\mathfrak{Z}(\mathbb{Q}) induces a surjection onto the principal stripes of HH in 𝔗\mathfrak{T}.

Proof.

Suppose that gH(G/H)()(G/H)()gH\in(G/H)(\mathbb{Q})\cap(G/H)(\mathbb{R})^{\circ} is such that η(gH)𝔗\eta(gH)\cap\mathfrak{T} is non-empty. Since gH(G/H)()G()HgH\in(G/H)(\mathbb{R})^{\circ}\subset G(\mathbb{R})\cdot H, we may choose gG()g\in G(\mathbb{R}). From 4.7 we then know that gDM=Dg=η(gH)gD_{M}=D_{g}=\eta(gH), so gDM𝔗gD_{M}\cap\mathfrak{T} is non-empty. Since DM=H()oD_{M}=H(\mathbb{R})\cdot o, we may therefore choose hH()h\in H(\mathbb{R}) such that gho=t𝔗gh\cdot o=t\in\mathfrak{T}. Then ghq1(𝔗)gh\in q^{-1}(\mathfrak{T}) and gHgH\in\mathfrak{Z}. ∎

4.3 Point Counting in Coset Spaces

In this section we explain how to count rational points inside (G/H)()(G/H)(\mathbb{R}), at least for appropriately chosen GG and HH. 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 HGH\subset G is an inclusion of \mathbb{Q}-algebraic scss groups and write G/HG/H for the quotient \mathbb{Q}-variety. Given Haar measures μH\mu_{H} and μG\mu_{G} on H(𝔸)H(\mathbb{A}) and G(𝔸)G(\mathbb{A}), respectively, we say that a measure μG/H:=μZ\mu_{G/H}:=\mu_{Z} on Z:=G(𝔸)/H(𝔸)Z:=G(\mathbb{A})/H(\mathbb{A}) is compatible with μH\mu_{H} and μG\mu_{G} if for any compactly supported function ff on G(𝔸)G(\mathbb{A}) we have

G(𝔸)f(t)μG(t)=ZH(𝔸)f(gu)𝑑μH(u)𝑑μZ(gH(𝔸)).\int_{G(\mathbb{A})}f(t)\,\mu_{G}(t)=\int_{Z}\int_{H(\mathbb{A})}f(gu)\,d\mu_{H}(u)\,d\mu_{Z}(gH(\mathbb{A})).

In our case, because both GG and HH 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 μH\mu_{H} (resp. μG\mu_{G}) induces a probability measure on the quotient H(𝔸)/H()H(\mathbb{A})/H(\mathbb{Q}) (resp. G(𝔸)/G()G(\mathbb{A})/G(\mathbb{Q})); 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 (G,H,Z=G(𝔸)/H(𝔸))(G,H,Z=G(\mathbb{A})/H(\mathbb{A})) with GG and HH scss that the measures μG,μH\mu_{G},\mu_{H} and μG/H=μZ\mu_{G/H}=\mu_{Z} have been chosen in this way.

In a situation where we have a representation ρ:GGL(V)\rho:G\to\operatorname{GL}(V) on a \mathbb{Q}-vector space VV such that HH is the stabilizer in GG of some vVv\in V, we will also denote by μG/H\mu_{G/H} the induced measure on the adelic points of the orbit variety GvG/HG\cdot v\simeq G/H. Using the inclusion p𝔸\mathbb{Q}_{p}\hookrightarrow\mathbb{A} we also obtain induced measures μG/H,p\mu_{G/H,p} on (G/H)(p)(G/H)(\mathbb{Q}_{p}) for each prime pp, and likewise a measure μG/H,\mu_{G/H,\infty} 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 Cc()C_{c}(-) refers to compactly supported continuous functions on the topological space inside the brackets. In what follows we set μ=μZ\mu=\mu_{Z}.

Proposition 4.15 (Prop. 5.3 in [GO11]).

Suppose that both GG and HH are scss \mathbb{Q}-groups with HGH\subset G a maximal connected \mathbb{Q}-subgroup. Set Z=G(𝔸)/H(𝔸)Z=G(\mathbb{A})/H(\mathbb{A}). Then for any well-rounded sequence {BTZ}\{B_{T}\subset Z\} of compact subsets whose volume diverges as TT\to\infty, we have,

#(G/H)()BT=#(G()[1]BT)μ(BT),\#(G/H)(\mathbb{Q})^{\circ}\cap B_{T}=\#(G(\mathbb{Q})[1]\cap B_{T})\sim\mu(B_{T}),

where [1]Z[1]\in Z is the class of the identity.

The statement of the proposition uses the following definition:

Definition 4.16.

A family BTZB_{T}\subset Z of compact subsets is called well-rounded if there exists c>0c>0 such that for every small ε>0\varepsilon>0, there exists a neighbourhood UεU_{\varepsilon} of 11 in G(𝔸)G(\mathbb{A}) such that for all sufficiently large TT,

(1cε)μ(UεBT)μ(BT)(1+cε)μ(uUεuBT).(1-c\cdot\varepsilon)\mu(U_{\varepsilon}B_{T})\leq\mu(B_{T})\leq(1+c\cdot\varepsilon)\mu(\cap_{u\in U_{\varepsilon}}uB_{T}).

We note that our definition of “well-rounded” is the special case of the notion “WW-well-rounded” appearing in [GO11, Def. 5.1] with W=1W=1.

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 GG where we use G(𝔸)G(\mathbb{A}), LL where we use H(𝔸)H(\mathbb{A}), and we take Γ=G()\Gamma=G(\mathbb{Q}). Here we have used the normalization μG(G(𝔸)/G())=1\mu_{G}(G(\mathbb{A})/G(\mathbb{Q}))=1 as well as μH(H(𝔸)/H())=1\mu_{H}(H(\mathbb{A})/H(\mathbb{Q}))=1. 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 GW=G(𝔸)G_{W}=G(\mathbb{A}) for us.

For the first equality we note that, because HH is simply connected, there is exactly one G()G(\mathbb{Q}) orbit in each G()G(\mathbb{R}) orbit inside G/HG/H (see the proof of [GO11, Cor. 1.9]). Thus it makes sense to identify (G/H)()(G/H)(\mathbb{Q})^{\circ} with G()H()G()/H()G(\mathbb{Q})\cdot H(\mathbb{R})\subset G(\mathbb{R})/H(\mathbb{R}), 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 ρ:GGLm\rho:G\to\operatorname{GL}_{m} be a representation of GG satisfying the condition that HGH\subset G is identified with the stabilizer of some vector vmv\in\mathbb{Q}^{m}. Fix a compact measurable subset ΩG()v\Omega\subset G(\mathbb{R})\cdot v with boundary measure zero and positive volume. We then consider, for each positive integer \ell, the sets

B:={(xp)G(𝔸)v:xΩ,xpppep for p a finite prime},=ppep.B_{\ell}:=\{(x_{p})\in G(\mathbb{A})\cdot v:x_{\infty}\in\Omega,\ \|x_{p}\|_{p}\leq p^{e_{p}}\textrm{ for }p\textrm{ a finite prime}\},\hskip 15.00002pt\ell=\prod_{p}p^{e_{p}}. (18)

We observe that BB_{\ell} is well-rounded with c=1c=1. Indeed, we can consider the subgroup pG(p)G(𝔸)\prod_{p}G(\mathbb{Z}_{p})\subset G(\mathbb{A}), which we may observe preserves BB_{\ell}. Then taking a neighbourhood of 11 of the form Uε=Kε×pG(p)U_{\varepsilon}=K_{\varepsilon}\times\prod_{p}G(\mathbb{Z}_{p}) for KεK_{\varepsilon} a sufficiently small compact neighbourhood of 11\in\mathbb{R} depending on ε\varepsilon we easily see that BB_{\ell} 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 μ(B)\mu(B_{\ell})\to\infty. We conclude from 4.15 that

Corollary 4.17.

In the above setup,

#(G/H)()Bμ(B).\#(G/H)(\mathbb{Q})^{\circ}\cap B_{\ell}\sim\mu(B_{\ell}). (19)

We remark that in the argument appearing in [GO11, Pf. of Cor. 1.9] the representation ρ:GGLm\rho:G\to\operatorname{GL}_{m} 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

B:={(xp)G(𝔸)v:xΩ,xpppep for p a finite prime},=ppepB^{\circ}_{\ell}:=\{(x_{p})\in G(\mathbb{A})\cdot v:x_{\infty}\in\Omega^{\circ},\ \|x_{p}\|_{p}\leq p^{e_{p}}\textrm{ for }p\textrm{ a finite prime}\},\hskip 15.00002pt\ell=\prod_{p}p^{e_{p}} (20)

where ΩΩ\Omega^{\circ}\subset\Omega denotes the interior of Ω\Omega, we also have

Corollary 4.18.

In the above setup,

#(G/H)()Bμ(B)=μ(B).\#(G/H)(\mathbb{Q})^{\circ}\cap B^{\circ}_{\ell}\sim\mu(B^{\circ}_{\ell})=\mu(B_{\ell}). (21)
Proof.

Using the structure of the product measure, we have μ(B)=μ(Ω)μf(B)\mu(B_{\ell})=\mu(\Omega)\cdot\mu_{f}(B^{\prime}_{\ell}), where BG(𝔸f)vB^{\prime}_{\ell}\subset G(\mathbb{A}_{f})\cdot v is defined as BB_{\ell} is except with the condition xΩx_{\infty}\in\Omega^{\circ} at the real place. We may approximate Ω\Omega^{\circ} from below as a union of compact subsets Ωj\Omega_{j}, and by applying 4.15 to a sequence of sets BjB_{\ell j} constructed with the Ωj\Omega_{j} we learn that #(G/H)()Ω\#(G/H)(\mathbb{Q})^{\circ}\cap\Omega^{\circ} is asymptotically greater than μ(Ωj)μf(B)\mu(\Omega_{j})\cdot\mu_{f}(B^{\prime}_{\ell}) for each jj. Since μ(Ωj)μ(Ω)\mu(\Omega_{j})\to\mu(\Omega) we obtain the result. ∎

5 Proofs of the General Theorems

We define a subset 𝒮()\mathcal{S}(\mathbb{C}) of points in the tensorial Hodge locus which will be of interest.

𝒮={sS():(𝐆s,Ds)(𝐌,D) where (𝐌,D) is a  Hodge subdatum -coset equivalent to (𝐌,DM)}.\mathcal{S}=\left\{s\in S(\mathbb{C}):\begin{matrix}(\mathbf{G}_{s},D_{s})\subset(\mathbf{M}^{\prime},D^{\prime})\textrm{ where }(\mathbf{M}^{\prime},D^{\prime})\textrm{ is a }\\ \textrm{ Hodge subdatum }\mathbb{Q}\textrm{-coset equivalent to }(\mathbf{M},D_{M})\end{matrix}\right\}.
Definition 5.1.

We say a hyperplane section LL of SS of codimension dd is sufficiently general if all of the following conditions hold:

  • (i)

    LL is smooth and irreducible;

  • (ii)

    LL intersects every Hodge locus component of dimension dd in a reduced set of points of cardinality equal to its degree;

  • (iii)

    LL does not intersect any (possibly tensorial) Hodge locus component of dimension smaller than dd;

  • (iv)

    LL is not contained in any component of the tensorial Hodge locus, and the algebraic monodromy group of 𝕍\mathbb{V} agrees with that of 𝕍|SL{\left.\kern-1.2pt\mathbb{V}\vphantom{\big{|}}\right|_{S\cap L}}; and

  • (v)

    the restriction of the period map φ\varphi to SLS\cap L is quasi-finite.

Note that sufficiently general hyperplanes sections exist assuming ddimSdimφ(S)d\geq\dim S-\dim\varphi(S). To explain why one can achieve (v), we note that the map φ\varphi factors as φ:STΓ\DS\varphi:S\to T\hookrightarrow\Gamma\backslash D_{S} with STS\to T algebraic (see [BBT23]), and all the components of the (tensorial) Hodge locus are pulled back from algebraic subvarieties of TT. In particular to achieve (v) it is enough to require that LL does not intersect any fibre of STS\to T in a positive dimensional locus, which is true away from a closed locus in the parameter space of hyperplane sections assuming ddimSdimφ(S)d\geq\dim S-\dim\varphi(S).

5.1 Proof of 1.22

We will fix a dd, and prove the same statement but with Z(v)Z(v^{\prime}) replaced by Z(v)dZ(v^{\prime})_{d}. This suffices, since one can always obtain (2) by summing over such inequalities for all possible dd and combining the constants.

Reduction to Local Point Counting: Now letting LSL\subset S be a sufficiently general hyperplane section of codimension dd, we can replace SS with SLS\cap L, 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 LZ(v)L\cap Z(v^{\prime}) are equal to the Mumford-Tate group of the corresponding component of Z(v)Z(v^{\prime}). After all these changes one can take d=0d=0 and reduce to estimating the degrees of the union of components Z(v)0Z(v^{\prime})_{0}.

We now use 2.4 to take an open cover S=i=1nBiS=\bigcup_{i=1}^{n}B_{i} such that each BiB_{i} admits a definable local period map ψi:BiDS\psi_{i}:B_{i}\to D_{S} landing inside a Siegel set 𝔒iDS\mathfrak{O}_{i}\subset D_{S}, with 𝔒i=ΩiAt,iKoioi\mathfrak{O}_{i}=\Omega_{i}A_{t,i}K_{o_{i}}\cdot o_{i}, with 𝒮\mathcal{S} the set defined above matching the one in 2.4. We set 𝔊i=ΩiAt,iKoi\mathfrak{G}_{i}=\Omega_{i}A_{t,i}K_{o_{i}}. We therefore have, for each BiB_{i} which intersects 𝒮\mathcal{S}, a point siBis_{i}\in B_{i} such that the Hodge datum of oi:=ψi(si)o_{i}:=\psi_{i}(s_{i}) is contained in a Hodge datum \mathbb{Q}-coset equivalent to (𝐌,DM)(\mathbf{M},D_{M}). We may moreover use 4.3 and the fact that φ\varphi is quasi-finite to ensure that each map ψi\psi_{i} is actually injective (we no longer require that oio_{i} is the image of some point siBis_{i}\in B_{i}, just that the Hodge datum of oio_{i} be contained in a Hodge datum \mathbb{Q}-coset equivalent to (𝐌,DM)(\mathbf{M},D_{M})).

By shrinking the BiB_{i} to closed subsets CiBiC_{i}\subset B_{i}, we may assume that S=i=1nCiS=\bigcup_{i=1}^{n}C_{i} but that the sets {C1,,Cn}\{C_{1},\ldots,C_{n}\} have disjoint interiors, which we denote by CiC^{\circ}_{i}. Moreover, we can choose the CiC_{i} so that the boundary CiCiC_{i}\setminus C^{\circ}_{i} avoids the countably many points vZ(v)0\bigcup_{v^{\prime}}Z(v^{\prime})_{0} for each ii. We then have that

[vO()ν(v)|νdegZ(v)0]=i=1n|Ci(vO()ν(v)|νZ(v)0)|.\left[\sum_{\begin{subarray}{c}v^{\prime}\in O(\mathbb{Q})^{\circ}\\ \nu(v^{\prime})|\nu\end{subarray}}\deg Z(v^{\prime})_{0}\right]=\sum_{i=1}^{n}\left|C^{\circ}_{i}\cap\left(\bigcup_{\begin{subarray}{c}v^{\prime}\in O(\mathbb{Q})^{\circ}\\ \nu(v^{\prime})|\nu\end{subarray}}Z(v^{\prime})_{0}\right)\right|. (22)

It will therefore suffice to estimate the asymptotic size as ν\nu\to\infty of each of the summands on the right.

Relating to Counting Stripes: We now let qi:𝐆S()DSq_{i}:\mathbf{G}_{S}(\mathbb{R})\to D_{S} be the natural orbit map ggoig\mapsto g\cdot o_{i}. By our choice above, the Hodge structure oio_{i} has Hodge datum contained in a datum \mathbb{Q}-coset equivalent to (𝐌,DM)(\mathbf{M},D_{M}) we denote by (𝐌i,DMi)(\mathbf{M}_{i},D_{M_{i}}). Because (𝐌i,DMi)(\mathbf{M}_{i},D_{M_{i}}) is \mathbb{Q}-coset equivalent to (𝐌,DM)(\mathbf{M},D_{M}), there is g𝐆S()g\in\mathbf{G}_{S}(\mathbb{R}) relating the two Hodge subdatums such that g𝐌g\mathbf{M} is defined over \mathbb{Q}. Using the isomorphism 𝐆S/𝐌O\mathbf{G}_{S}/\mathbf{M}\cong O, this means that vi:=(g𝐌)vv_{i}:=(g\mathbf{M})\cdot v is a \mathbb{Q}-vector and 𝐌i\mathbf{M}_{i} is exactly the stabilizer in 𝐆S\mathbf{G}_{S} of viv_{i}. Note this means that viv_{i} is Hodge for oio_{i}.

Let ri:𝐆S()(𝐆S/𝐌i)()r_{i}:\mathbf{G}_{S}(\mathbb{R})\to(\mathbf{G}_{S}/\mathbf{M}_{i})(\mathbb{R}) be the projection. We set 𝔗i=ψi(Ci)\mathfrak{T}_{i}=\psi_{i}(C^{\circ}_{i}) and i=ri(qi1(𝔗i))\mathfrak{Z}_{i}=r_{i}(q^{-1}_{i}(\mathfrak{T}_{i})). Using the injectivity of ψi\psi_{i} we may then compute the summands in (22) by

|Ci(vO()ν(v)|νZ(v)0)|=|g(𝐆S/𝐌i)()ν(gvi)|ν[𝔗igDMi]0|.\left|C^{\circ}_{i}\cap\left(\bigcup_{\begin{subarray}{c}v^{\prime}\in O(\mathbb{Q})^{\circ}\\ \nu(v^{\prime})|\nu\end{subarray}}Z(v^{\prime})_{0}\right)\right|=\left|\bigcup_{\begin{subarray}{c}g\in(\mathbf{G}_{S}/\mathbf{M}_{i})(\mathbb{Q})^{\circ}\\ \nu(g\cdot v_{i})|\nu\end{subarray}}[\mathfrak{T}_{i}\cap gD_{M_{i}}]_{0}\right|. (23)

By construction, one also has

i={g(𝐆S/𝐌i)():(𝔗igDMi)}.\mathfrak{Z}_{i}=\{g\in(\mathbf{G}_{S}/\mathbf{M}_{i})(\mathbb{R})^{\circ}:(\mathfrak{T}_{i}\cap gD_{M_{i}})\neq\varnothing\}. (24)

Because 𝔗i\mathfrak{T}_{i} is definable and the sets gDMigD_{M_{i}} lie in a common definable family, the number of isolated points in 𝔗igDMi\mathfrak{T}_{i}\cap gD_{M_{i}} is uniformly bounded by some constant κ\kappa independent of gg. This implies that

|g(𝐆S/𝐌i)()ν(gvi)|ν[𝔗igDMi]0|\displaystyle\left|\bigcup_{\begin{subarray}{c}g\in(\mathbf{G}_{S}/\mathbf{M}_{i})(\mathbb{Q})^{\circ}\\ \nu(g\cdot v_{i})|\nu\end{subarray}}[\mathfrak{T}_{i}\cap gD_{M_{i}}]_{0}\right| κ|{gi(𝐆S/𝐌i)():ν(gv)|ν}|\displaystyle\leq\kappa\cdot|\{g\in\mathfrak{Z}_{i}\cap(\mathbf{G}_{S}/\mathbf{M}_{i})(\mathbb{Q})^{\circ}:\nu(g\cdot v)|\nu\}|
κ|{v(𝔊ivi)():ν(v)|ν}|,\displaystyle\leq\kappa\cdot|\{v^{\prime}\in(\mathfrak{G}_{i}\cdot v_{i})(\mathbb{Q}):\nu(v^{\prime})|\nu\}|,

where we use that qi1(𝔗i)𝔊iq_{i}^{-1}(\mathfrak{T}_{i})\subset\mathfrak{G}_{i} as well as the isomorphism 𝐆S/𝐌i𝐆Svi\mathbf{G}_{S}/\mathbf{M}_{i}\cong\mathbf{G}_{S}\cdot v_{i}. This is what we wanted to show.

5.2 Proof of 1.23

The statement is clearly local, and it suffices to take BB relatively compact admitting a local period map ψ:BDS\psi:B\to D_{S} which lifts φ\varphi. The strongly 𝕍\mathbb{V}-likely hypothesis implies that 𝒮\mathcal{S} is dense in S()S(\mathbb{C}) by [KU23, Thm. 3.5] (in fact even just 𝕍\mathbb{V}-likely suffices here). Thus BB contains a point s𝒮s\in\mathcal{S}. We set o=ψ(s)o=\psi(s). After relabelling we may assume that the Hodge datum at oo is contained in (𝐌,DM)(\mathbf{M},D_{M}). We set q:𝐆S()DSq:\mathbf{G}_{S}(\mathbb{R})\to D_{S} to be the orbit map ggog\mapsto g\cdot o.

We set =ψ(B)\mathcal{I}=\psi(B), and follow the proof given in [KU23, §4.4]. The proof given there explains that we have an open neighbourhood in 𝐆S()×\mathbf{G}_{S}(\mathbb{R})\times\mathcal{I} such that for each (g,x)(g,x) in this neighbourhood the intersection gDMgD_{M}\cap\mathcal{I} has dimension d=dim+dimDMdimDSd=\dim\mathcal{I}+\dim D_{M}-\dim D_{S} at xx. Let us be more precise about what this neighbourhood looks like. One can start by defining 𝒱𝐆S()\mathcal{V}\subset\mathbf{G}_{S}(\mathbb{R}) as the locus of g𝐆S()g\in\mathbf{G}_{S}(\mathbb{R}) for which gDMgD_{M}\cap\mathcal{I} is non-empty. Then the argument in [KU23, §4.3] shows that the condition that dimx(gDM)=d\dim_{x}(gD_{M}\cap\mathcal{I})=d is open on (g,x)𝒱×(g,x)\in\mathcal{V}\times\mathcal{I}. The proof of [KU23, §4.4] then shows that, after removing a closed locus 𝒞\mathcal{C} of smaller dimension from 𝒱×\mathcal{V}\times\mathcal{I}, the intersection germs (gDM,x)(gD_{M}\cap\mathcal{I},x) all have pure dimension dd and do not lie inside any translates of period subdomains of DSD_{S} of smaller dimension. In particular, whenever (g,x)(g,x) is outside of 𝒞\mathcal{C} and gDMgD_{M} is a Mumford-Tate domain, the intersection gDMgD_{M}\cap\mathcal{I} is a generic stripe of H=𝐌H=\mathbf{M} in \mathcal{I}, where we use the language of §4.2.1. Using the definition of the product topology we can find a product 𝒱×(𝒱×)𝒞\mathcal{V}^{\prime}\times\mathcal{I}^{\prime}\subset(\mathcal{V}\times\mathcal{I})\setminus\mathcal{C}. Replacing BB with B=ψ1()B^{\prime}=\psi^{-1}(\mathcal{I}^{\prime}) we may then assume that 𝒞=\mathcal{C}=\varnothing.

Now choose LL sufficiently general. This means in particular that, for each of the intersections LZ(v)L\cap Z(v^{\prime}), each point of LZ(v)L\cap Z(v^{\prime}) has the same Mumford-Tate group as the component of Z(v)Z(v^{\prime}) in which it lies. In particular, each intersection gDMψ(BL)gD_{M}\cap\psi(B\cap L) with (g𝐌g1,gDM)(g\mathbf{M}g^{-1},gD_{M}) a Hodge datum \mathbb{Q}-coset equivalent to (𝐌,DM)(\mathbf{M},D_{M}) and g𝒱g\in\mathcal{V} is a generic stripe of DMD_{M} in ψ(BL)\psi(B\cap L). Moreover the condition that gDMgD_{M} intersect ψ(BL)\psi(B\cap L) is still open on 𝒱×ψ(BL)\mathcal{V}\times\psi(B\cap L) (the analytic varieties ψ(BL)\psi(B\cap L) and gDMgD_{M} have complementary dimension inside DSD_{S}), and so after possibly shrinking 𝒱\mathcal{V} we can ensure that for each g𝒱g\in\mathcal{V} the intersection gDMψ(BL)gD_{M}\cap\psi(B\cap L) is a generic stripe of DMD_{M} in 𝔗:=ψ(BL)\mathfrak{T}:=\psi(B\cap L).

Now let UU be the image of 𝒱\mathcal{V} in O()𝐆S()/𝐌()O(\mathbb{R})^{\circ}\cong\mathbf{G}_{S}(\mathbb{R})/\mathbf{M}(\mathbb{R}), where for the isomorphism we take the natural orbit map. Then the set UU is open inside the set \mathfrak{Z} of §4.2.1. We then consider the map

U()η{generic stripes of H in 𝔗}U(\mathbb{Q})\xrightarrow{\eta^{\circ}}\{\textrm{generic stripes of }H\textrm{ in }\mathfrak{T}\}

induced by the restriction of the map η\eta^{\circ} of §4.2.1. Combining 4.9, 4.5 and the assumption that 𝐌\mathbf{M} has finite index in its normalizer, we learn that η\eta^{\circ} is quasi-finite with fibres of uniformly bounded size. Identifying 𝐆S/𝐌\mathbf{G}_{S}/\mathbf{M} with OO using the natural representation, we get that

|U1ν𝕍|gU()ν(gv)|ν|[ψ(BL)gDM]0|.\left|U\cap\frac{1}{\nu}\mathbb{V}\right|\asymp\sum_{\begin{subarray}{c}g\in U(\mathbb{Q})\\ \nu(g\cdot v)|\nu\end{subarray}}\left|[\psi(B\cap L)\cap gD_{M}]_{0}\right|. (25)

Using the genericity of LL, the sum on the right is then a lower bound for the sum in the statement. Since the symbol \asymp implies the same expression with the symbol \lesssim up to a positive real constant, the result follows.

5.3 Proof of 1.24 and 1.25

1.24 was shown in §3.3. Then 1.25 follows directly by combining 1.24 with 1.22.

5.4 Proof of 1.26

Given a Hodge structure hh on a lattice VV, a vector vVv\in V is Hodge for hh if and only if vvv\otimes v is Hodge for the Hodge structure hhh\otimes h on VVV\otimes V. (Indeed, the Mumford-Tate group MhM_{h} of hh fixes vv if and only if it fixes vvv\otimes v acting diagonally, and MhhM_{h\otimes h} is identified with the image of MhM_{h} under the diagonal action since it is the \mathbb{Q}-Zariski closure of the diagonal image of 𝕌\mathbb{U}.) If VV comes with a polarization form QQ, then Q(vv,vv)=Q(v,v)2Q(v\otimes v,v\otimes v)=Q(v,v)^{2} for the induced polarization on VVV\otimes V. Likewise, Q¯(vv)=Q¯(v)2\overline{Q}(v\otimes v)=\overline{Q}(v)^{2}. Thus to bound NLq\operatorname{NL}_{q}, we may reduce to bounding the Noether-Lefschetz locus NL1,q2\operatorname{NL}^{\otimes 2}_{1,q} associated to 𝕍𝕍\mathbb{V}\otimes\mathbb{V} (since Q¯(vv)=ν(vv)2=q2\overline{Q}(v\otimes v)=\nu(v\otimes v)^{2}=q^{2}, as there is no square-free part). It then follows from 1.25 that

degNLqdegNL1,q2cqm2+min{r2+s2,2rs}2\deg\operatorname{NL}_{q}\lesssim\deg\operatorname{NL}^{\otimes 2}_{1,q}\lesssim c\cdot q^{m^{2}+\textrm{min}\{r^{2}+s^{2},2rs\}-2}

where we use that the signature of the lattices appearing in the fibres of 𝕍𝕍\mathbb{V}\otimes\mathbb{V} is (2rs,r2+s2)(2rs,r^{2}+s^{2}).

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 TT in the former two statements to be equal to the SS 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 T=ST=S. We start by observing that, in the proof of 1.23, LL was required to be sufficiently general intersecting some BBB^{\prime}\subset B constructed in the proof. This BB^{\prime} could be chosen to be any open subset of BB such that gDMψ(B)gD_{M}\cap\psi(B^{\prime}) always has the “expected” dimension dd if non-empty.

We observe that BB^{\prime} has this property outside of a finite union of tensorial Hodge loci of 𝕍\mathbb{V}. This follows from the “Ax-Schanuel in families” statement in [BU24, §3.3], where the bundle PP is constructed from 𝕍\mathbb{V} as in [BU24, §4.6]. Indeed, the varieties gDˇMg\widecheck{{D}}_{M} belong to finitely many families of algebraic subvarieties of Dˇ\widecheck{{D}}, hence induce finitely many algebraic families f:𝒵𝒴f:\mathcal{Z}\to\mathcal{Y} of subvarieties of PP using the procedure outlined in [BU24, §4.6]. Then the images in TT of intersections of fibres of ff with leaves of PP, which in particular include all the germs of the form ψ1(gDM)\psi^{-1}(gD_{M}), are contained in the fibres of finitely many families {hi:CiAi}i=1m\{h_{i}:C_{i}\to A_{i}\}_{i=1}^{m} of weakly special subvarieties of TT as a consequence of [BU24, Thm 3.12]. We may assume that the fibres of the hih_{i} all have the same algebraic monodromy groups using [BU24, Lem. 7.6] (c.f. [BU24, Lem. 7.5]). The fibres of the hih_{i} which contain the ψ1(gDM)\psi^{-1}(gD_{M}) do not have zero period dimension (since dimφ(ψ1(gDM))>0\dim\varphi(\psi^{-1}(gD_{M}))>0 as dimDdimDM<dimφ(T)\dim D-\dim D_{M}<\dim\varphi(T) in both cases). It then follows from the \mathbb{Q}-simplicity of the algebraic monodromy group 𝐇T\mathbf{H}_{T} in our two examples that each fibre of each hih_{i} lies in a weakly non-factor weakly special subvariety of TT 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 TT. Thus the images of the CiC_{i} in TT are contained in a union of strict special subvarieties (i.e., tensorial Hodge locus components) of TT. There are countably many such varieties and the images of the CiC_{i} are algebraic, so we can take this union to be finite.

Now to prove 1.6 and 1.11, it suffices to replace TT with the complement in TT of this collection of tensorial Hodge loci. Note that by construction, any sufficiently general hyperplane section of the original TT also intersects this complement. After doing this, the locus 𝒞\mathcal{C} appearing in the proof of 1.23 is automatically empty, and one can take UU to be the image of q1(ψ(B))q^{-1}(\psi(B)) in O()O(\mathbb{R})^{\circ}. 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 𝕍\mathbb{V}-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 (a,b)=(r,s)(a,b)=(r,s). We may identify 𝐆S=SO(a,b)\mathbf{G}_{S}=\operatorname{SO}(a,b) and the natural inclusion 𝐌𝐆S\mathbf{M}\hookrightarrow\mathbf{G}_{S} with 𝐌=SO(a,b1)SO(a,b)\mathbf{M}=\operatorname{SO}(a,b-1)\hookrightarrow\operatorname{SO}(a,b).

We let DvDSD_{v}\subset D_{S} be the natural Mumford-Tate subdomain consisting of Hodge structures with Mumford-Tate group SO(a,b1)\operatorname{SO}(a,b-1) (the subscript vv denotes a primitive \mathbb{Q}-vector which SO(a,b1)\operatorname{SO}(a,b-1) 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 (𝐌,DM)(SO(a,b),DS)(\mathbf{M},D_{M})\subset(\operatorname{SO}(a,b),D_{S}) which is conjugation-equivalent to the Hodge subdatum (SO(a,b1),Dv)(\operatorname{SO}(a,b-1),D_{v}) is \mathbb{Q}-coset equivalent to the same subdatum if and only if u(v)=u(v)u(v)=u(v^{\prime}), where vv^{\prime} is a non-zero element in the \mathbb{Q}-vector space stabilized by 𝐌\mathbf{M}.

Proof.

By scaling vv we can assume that Q(v,v)=u(v)Q(v,v)=u(v). We also let gSO(a,b)()g\in\operatorname{SO}(a,b)(\mathbb{R}) be a point such that (g1𝐌g,g1DM)=(SO(a,b1),Dv)(g^{-1}\mathbf{M}g,g^{-1}D_{M})=(\operatorname{SO}(a,b-1),D_{v}).

Suppose first that (𝐌,DM)(\mathbf{M},D_{M}) is \mathbb{Q}-coset equivalent to (SO(a,b1),Dv)(\operatorname{SO}(a,b-1),D_{v}). Then we may choose gg such that v=gvv^{\prime}=g\cdot v is defined over \mathbb{Q}. Choose λ×\lambda\in\mathbb{Q}^{\times} such that λ2Q(v,v)=Q(λv,λv)=u(v)\lambda^{2}Q(v^{\prime},v^{\prime})=Q(\lambda v^{\prime},\lambda v^{\prime})=u(v^{\prime}). Then since Q(v,v)=Q(gv,gv)=Q(v,v)Q(v,v)=Q(gv,gv)=Q(v^{\prime},v^{\prime}), necessarily λ2Q(v,v)=u(v)\lambda^{2}Q(v,v)=u(v^{\prime}), and thus λ2u(v)=u(v)\lambda^{2}u(v)=u(v^{\prime}). Since u(v)u(v) and u(v)u(v^{\prime}) are both square free, it follows that u(v)=u(v)u(v)=u(v^{\prime}).

Conversely, suppose that u(v)=u(v)u(v)=u(v^{\prime}). Because the subspace stabilized by SO(a,b1)\operatorname{SO}(a,b-1) is 11-dimensional, we have gv=λvg\cdot v=\lambda\cdot v^{\prime} for some λ×\lambda\in\mathbb{R}^{\times}, and we may assume vv^{\prime} is chosen such that Q(v,v)=u(v)Q(v^{\prime},v^{\prime})=u(v^{\prime}). We then have Q(v,v)=Q(v,v)Q(v,v)=Q(v^{\prime},v^{\prime}). Then

Q(v,v)=Q(gv,gv)=λ2Q(v,v)=λ2Q(v,v)Q(v,v)=Q(gv,gv)=\lambda^{2}Q(v^{\prime},v^{\prime})=\lambda^{2}Q(v,v)

so it follows that λ=±1\lambda=\pm 1 and gvg\cdot v is a \mathbb{Q}-vector. It follows that the coset gSO(a,b1)g\operatorname{SO}(a,b-1) is defined over \mathbb{Q}, or that the two Hodge data are \mathbb{Q}-coset equivalent. ∎

We are therefore reduced to estimating the asymptotic number of rational points in a compact subset of (SO(a,b)/SO(a,b1))()(\operatorname{SO}(a,b)/\operatorname{SO}(a,b-1))(\mathbb{R}). We set G=SO(a,b)G=\operatorname{SO}(a,b) and H=SO(a,b1)H=\operatorname{SO}(a,b-1) to match the notation of §4.2.1. We start by giving an alternative description of the quotient SO(a,b)/SO(a,b1)\operatorname{SO}(a,b)/\operatorname{SO}(a,b-1). First, consider the universal cover π:G~SO(a,b)\pi:\widetilde{G}\to\operatorname{SO}(a,b), which is an isogeny of \mathbb{Q}-algebraic groups. If we consider the scalar extension π\pi_{\mathbb{C}} we have that SO(a,b)SO(a+b)\operatorname{SO}(a,b)_{\mathbb{C}}\cong\operatorname{SO}(a+b)_{\mathbb{C}}. The fundamental group of SO(a+b)\operatorname{SO}(a+b)_{\mathbb{C}} is /2\mathbb{Z}/2\mathbb{Z} and its universal cover is the Spin group Spin(a+b)\textrm{Spin}(a+b); it follows that G~\widetilde{G} is a \mathbb{Q}-form of Spin(a+b)\textrm{Spin}(a+b).

Now consider H~=π1(SO(a,b1))\widetilde{H}=\pi^{-1}(\operatorname{SO}(a,b-1)). We claim that H~\widetilde{H} is (geometrically) connected. Once again it suffices to scalar extend to \mathbb{C}, after which we can reduce to the same problem for SO(a+b1)SO(a+b)\operatorname{SO}(a+b-1)\subset\operatorname{SO}(a+b) and the universal covering by Spin(a+b)\textrm{Spin}(a+b). 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 SO(a+b1)\operatorname{SO}(a+b-1) in Spin(a+b)\textrm{Spin}(a+b) is identified with a copy of Spin(a+b1)\textrm{Spin}(a+b-1).

We thus have an identification SO(a,b)/SO(a,b1)G~/H~\operatorname{SO}(a,b)/\operatorname{SO}(a,b-1)\cong\widetilde{G}/\widetilde{H}. The groups G~\widetilde{G} and H~\widetilde{H} are both semisimple and simply-connected, so we may estimate the number of rational points a definable (relatively) compact subset of (SO(a,b)/SO(a,b1))()(\operatorname{SO}(a,b)/\operatorname{SO}(a,b-1))(\mathbb{R}) using the results of §4.3. In particular, letting Ω\Omega be such a subset and applying our discussion in §4.3.2 to the pair (G~,H~)(\widetilde{G},\widetilde{H}) and the natural representation G~GLm\widetilde{G}\to\operatorname{GL}_{m}, we obtain from 4.18 that

(Gv)()Ω1νmμ(Bν)=μ(Ω)pμp(Bν,p)(G\cdot v)(\mathbb{Q})^{\circ}\cap\Omega\cap\frac{1}{\nu}\mathbb{Z}^{m}\simeq\mu(B_{\nu})=\mu(\Omega)\prod_{p}\mu_{p}(B_{\nu,p}) (26)

where Bν,p={xpG(p)v:xpppep}B_{\nu,p}=\{x_{p}\in G(\mathbb{Q}_{p})\cdot v:\|x_{p}\|_{p}\leq p^{e_{p}}\}. It therefore suffices to compute the volumes μp(Bν,p)\mu_{p}(B_{\nu,p}).

6.2.1 Volume at pp

We recall that the variety O=GvO=G\cdot v is a closed subvariety of 𝔸m\mathbb{A}^{m} defined by the relation Q(w,w)=Q(v,v)=:uQ(w,w)=Q(v,v)=:u for w=(w1,,wm)w=(w_{1},\ldots,w_{m}) coordinates of 𝔸m\mathbb{A}^{m}. To compute using the measure μp\mu_{p} on O(p)O(\mathbb{Q}_{p}) we may use Gauge forms. The Gauge forms Ω\Omega (resp ω)\omega) for the measures on SO(a,b)\operatorname{SO}(a,b) (resp. OO) are given on [Sat92, pg. 149]. The expression for ω\omega in particular, which is what interests us, is

ω=dx1dxmd(x¯tQx¯).\displaystyle\omega=\frac{dx_{1}\wedge\cdots\wedge dx_{m}}{d(\overline{x}^{t}Q\overline{x})}.

(Here we interpret the division of differential forms α/β\alpha/\beta as representing a differential form γ\gamma, if it exists, for which α=γβ\alpha=\gamma\wedge\beta. The wedge product is computed in the algebra of differentials on 𝔸m\mathbb{A}^{m}.)

We write L=mL=\mathbb{Z}^{m} and let B:LB:L\to\mathbb{Z} be the quadratic form induced by QQ. The lattice (L,B)(L,B) is unimodular and indefinite. Unimodular and indefinite lattices are fully classified by the signature of the form QQ, and whether the lattice is “even” or “odd” (see [MH73, Ch. II, Thm. 5.3]). (The lattice (L,B)(L,B) is even if B(v,v)2B(v,v)\in 2\mathbb{Z} for all v,vLv,v^{\prime}\in L, and odd otherwise.) If the lattice is odd, then QQ can be diagonalized [MH73, Ch. II, Thm. 4.3]. If we tensor with [21]\mathbb{Z}[2^{-1}] the distinction disappears, and (L[21],B[21])(L_{\mathbb{Z}[2^{-1}]},B_{\mathbb{Z}[2^{-1}]}) becomes isomorphic over [21]\mathbb{Z}[2^{-1}] to an odd unimodular indefinite lattice. In particular (L,Q)[21](L,Q)_{\mathbb{Z}[2^{-1}]} is diagonalizable. As we are only interested in computing the factors in (26) away from the prime 22, we may therefore work over [21]\mathbb{Z}[2^{-1}] and assume that QQ is diagonal.

After diagonalizing QQ we may write B(x¯)=x¯tQx¯=iaixi2B(\overline{x})=\overline{x}^{t}Q\overline{x}=\sum_{i}a_{i}x_{i}^{2}. Then

d(iaixi2)dx2dxmx1\displaystyle d\left(\sum_{i}a_{i}x_{i}^{2}\right)\wedge\frac{dx_{2}\wedge\cdots\wedge dx_{m}}{x_{1}} =2a1x1dx1dx2dxmx1\displaystyle=2a_{1}\,x_{1}dx_{1}\wedge\frac{dx_{2}\wedge\cdots\wedge dx_{m}}{x_{1}}
=2a1dx1dxm,\displaystyle=2a_{1}\,dx_{1}\wedge\cdots\wedge dx_{m},

so we in fact have

ω=12a1dx2dxmx1.\displaystyle\omega=\frac{1}{2a_{1}}\frac{dx_{2}\wedge\cdots\wedge dx_{m}}{x_{1}}.

To integrate this form over the pp-adics with p2p\neq 2 we apply the main result of [Yam83], which says that

Theorem 6.2.

Let fip[z1,,zn]f_{i}\in\mathbb{Z}_{p}[z_{1},\ldots,z_{n}] be polynomials for 1iN1\leq i\leq N and write f¯:𝔸n𝔸N\overline{f}:\mathbb{A}^{n}\to\mathbb{A}^{N} for the associated map. Let t=(t1,,tN)pNt=(t_{1},\ldots,t_{N})\in\mathbb{Z}^{N}_{p} be a point such that the fibre f¯1(t)\overline{f}^{-1}(t) has at least one p\mathbb{Q}_{p}-point at which the map f¯1(t)𝔸N\overline{f}^{-1}(t)\to\mathbb{A}^{N} is submersive. Then one has

f¯1(t)pn|Θt|=lime#{apn/pe:fi(a)=ti mod pe for 1iN}p(nN)e=:αp(t,f¯),\int_{\overline{f}^{-1}(t)\cap\mathbb{Z}^{n}_{p}}|\Theta_{t}|=\lim_{e\to\infty}\frac{\#\{a\in\mathbb{Z}^{n}_{p}/p^{e}:f_{i}(a)=t_{i}\textrm{ mod }p^{e}\textrm{ for }1\leq i\leq N\}}{p^{(n-N)e}}=:\alpha_{p}(t,\overline{f}), (27)

where Θt\Theta_{t} is defined by

Θt=sgn(λ1,,λn)|det[fizΛ]|1dzΛc.\Theta_{t}=\textrm{sgn}(\lambda_{1},\ldots,\lambda_{n})\left|\det\left[\frac{\partial f_{i}}{\partial z_{\Lambda}}\right]\right|^{-1}dz_{\Lambda^{c}}. (28)

Here Λ=(λ1,,ΛN)\Lambda=(\lambda_{1},\ldots,\Lambda_{N}) is an ordered NN-tuple of distinct increasing integers, and Λc=(λN+1,,λn)\Lambda^{c}=(\lambda_{N+1},\ldots,\lambda_{n}) is the complementary tuple (c.f. [Yam83, §2]).

We will apply this theorem with N=1N=1, Λ=(1)\Lambda=(1), t=p2kut=p^{2k}u, n=mn=m, and f1=Bf_{1}=B for some fixed integer k0k\geq 0. We consider the sets A=B1(u)(1pkpm)A=B^{-1}(u)\cap\left(\frac{1}{p^{k}}\mathbb{Z}^{m}_{p}\right) and Ak=B1(p2ku)pmA_{k}=B^{-1}(p^{2k}u)\cap\mathbb{Z}^{m}_{p}. Let zi=pkxiz_{i}=p^{k}x_{i} for each ii. We have that

A|ω|\displaystyle\int_{A}|\omega| =A|dx2dxm2a1x1|\displaystyle=\int_{A}\left|\frac{dx_{2}\wedge\cdots\wedge dx_{m}}{2a_{1}x_{1}}\right|
=Ak|d(pkz2)d(pkzm)pk2a1z1|\displaystyle=\int_{A_{k}}\left|\frac{d(p^{-k}z_{2})\wedge\cdots\wedge d(p^{-k}z_{m})}{p^{-k}2a_{1}z_{1}}\right|
=p(m2)kAk|dz2dzm2a1z1|\displaystyle=p^{(m-2)k}\int_{A_{k}}\left|\frac{dz_{2}\wedge\cdots\wedge dz_{m}}{2a_{1}z_{1}}\right|
=p(m2)kAk|Θt|\displaystyle=p^{(m-2)k}\int_{A_{k}}\left|\Theta_{t}\right|
=p(m2)kαp(p2ku,B)\displaystyle=p^{(m-2)k}\,\alpha_{p}(p^{2k}u,B)

We now use the main theorem of [SH00] to compute αp(p2ku,B)\alpha_{p}(p^{2k}u,B).

Proposition 6.3.

We have

αp(p2ku,B)\displaystyle\alpha_{p}(p^{2k}u,B) =1+j=1j even2k+1pw(j)+j=1j odd2k+1pw(j)((1)mdetQp)()\displaystyle=1+\sum_{\begin{subarray}{c}j=1\\ j\textrm{ even}\end{subarray}}^{2k+1}p^{w(j)}+\sum_{\begin{subarray}{c}j=1\\ j\textrm{ odd}\end{subarray}}^{2k+1}p^{w(j)}\left(\frac{(-1)^{m}\det Q}{p}\right)(\star)
()\displaystyle(\star) =k=1m{02kj0,m odd(1p1)(1p)m/22kj0,m even(1p)(m+1)/2(up)2kj=1,m oddp1/2(1p)m/22kj=1,m even,\displaystyle=\prod_{k=1}^{m}\begin{cases}0&2k-j\geq 0,m\textrm{ odd}\\ (1-p^{-1})\cdot\left(\frac{-1}{p}\right)^{m/2}&2k-j\geq 0,m\textrm{ even}\\ \left(\frac{-1}{p}\right)^{(m+1)/2}\left(\frac{u}{p}\right)&2k-j=-1,m\textrm{ odd}\\ -p^{-1/2}\cdot\left(\frac{-1}{p}\right)^{m/2}&2k-j=-1,m\textrm{ even},\end{cases}
w(j)\displaystyle w(j) :=mj/2+j+min{2kj,0}/2.\displaystyle:=-mj/2+j+\textrm{min}\{2k-j,0\}/2.
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 T=p2kuT=p^{2k}u and S=QS=Q. We write

αp(T,S)=?F1(?)k=0r+1F2(?,k){ν}kF3(?,k,{ν}k)\alpha_{p}(T,S)=\sum_{?}F_{1}(?)\sum_{k=0}^{r+1}F_{2}(?,k)\sum^{\prime}_{\{\nu\}_{k}}F_{3}(?,k,\{\nu\}_{k})

in accordance with the expression for αp(T,S)\alpha_{p}(T,S) appearing in [SH00]. We have n=1n=1, so 𝔖n=1\mathfrak{S}_{n}=1 is the trivial group; c1(σ)=1c_{1}(\sigma)=1 always and c2(σ)=0c_{2}(\sigma)=0 always; the set I={1}I=\{1\} and admits a unique 𝔖n\mathfrak{S}_{n} stable partition of length r+1=1r+1=1. This means that the functions tt and τ\tau are always zero. From all this it follows that F1=1/2F_{1}=1/2, the first summation sign can be removed, and we are left to estimate an expression of the form

αp(T,S)=12k=01F2(k){ν}kF3(k,{ν}k).\alpha_{p}(T,S)=\frac{1}{2}\sum_{k=0}^{1}F_{2}(k)\sum^{\prime}_{\{\nu\}_{k}}F_{3}(k,\{\nu\}_{k}).

Now n,n()n_{\ell},n^{(\ell)} and n()n(\ell) are equal 11 when 0\ell\leq 0, and when 1\ell\geq 1 we have n=n()=n()=0n_{\ell}=n^{(\ell)}=n(\ell)=0. Moreover c1(0)c^{(0)}_{1} is always just c1c_{1}, so equal to 11, and c1(1)=0c^{(1)}_{1}=0. It follows that F2(0)=2F_{2}(0)=2 and F2(1)=1F_{2}(1)=1. When k=0k=0 the third summation is equal to 11 by convention, so we obtain

αp(T,S)=1+12{ν}1F3(1,{ν}1).\alpha_{p}(T,S)=1+\frac{1}{2}\sum^{\prime}_{\{\nu\}_{1}}F_{3}(1,\{\nu\}_{1}).

Now {ν}1\{\nu\}_{1} is just a single integer ν\nu satisfying b0(id,T)ν1-b_{0}(\textrm{id},T)\leq\nu\leq-1, where b0(id,T)=β1+1b_{0}(\textrm{id},T)=\beta_{1}+1. We thus get

{ν}1F3(1,{ν}1)\displaystyle\sum^{\prime}_{\{\nu\}_{1}}F_{3}(1,\{\nu\}_{1}) =ν=1β1+1pνΞ0,ν(id;T,S).\displaystyle=\sum_{-\nu=1}^{\beta_{1}+1}p^{-\nu}\Xi_{0,\nu}(\textrm{id};T,S).

Now with ν=λ<0\nu=\lambda<0 one has ρ0,λ=mλ2+12min{β1+λ,0}\rho_{0,\lambda}=\frac{m\lambda}{2}+\frac{1}{2}\textrm{min}\{\beta_{1}+\lambda,0\}, and Ξ0,λ(id,T,S)=pρ0,λξ1,λ(T,S)\Xi_{0,\lambda}(\textrm{id},T,S)=p^{\rho_{0,\lambda}}\xi_{1,\lambda}(T,S). In our setting Bi(λ)=B_{i}(\lambda)=\varnothing, and A(λ)={k:1km,λ0 mod 2}A(\lambda)=\{k:1\leq k\leq m,\lambda\neq 0\textrm{ mod }2\}, i.e., either A(λ)={1,,m}A(\lambda)=\{1,\ldots,m\} or A(λ)=A(\lambda)=\varnothing, depending on the parity of λ\lambda.

Thus ξ1,λ(T,S)\xi_{1,\lambda}(T,S) is equal to 22 when λ\lambda is even, and for λ\lambda odd is given by the expression

ξ1,λ=2((1)mdetQp)k=1m{0β1+λ0,m odd(1p1)(1p)m/2β1+λ0,m even(1p)(m+1)/2(up)β1+λ=1,m oddp1/2(1p)m/2β1+λ=1,m even.\displaystyle\xi_{1,\lambda}=2\left(\frac{(-1)^{m}\det Q}{p}\right)\prod_{k=1}^{m}\begin{cases}0&\beta_{1}+\lambda\geq 0,m\textrm{ odd}\\ (1-p^{-1})\cdot\left(\frac{-1}{p}\right)^{m/2}&\beta_{1}+\lambda\geq 0,m\textrm{ even}\\ \left(\frac{-1}{p}\right)^{(m+1)/2}\left(\frac{u}{p}\right)&\beta_{1}+\lambda=-1,m\textrm{ odd}\\ -p^{-1/2}\cdot\left(\frac{-1}{p}\right)^{m/2}&\beta_{1}+\lambda=-1,m\textrm{ even}.\end{cases}

Putting everything together we obtain the result in the statement. ∎

Corollary 6.4.

We have αp(p2ku,B)12p(1m/2)1p1m/2\alpha_{p}(p^{2k}u,B)\geq\frac{1-2p^{(1-m/2)}}{1-p^{1-m/2}}.

Proof.

From 6.3 we get that

αp(p2ku,B)\displaystyle\alpha_{p}(p^{2k}u,B) 1j=12k+1p(1m/2)j\displaystyle\geq 1-\sum_{j=1}^{2k+1}p^{(1-m/2)j}
1(11p(1m/2)1)\displaystyle\geq 1-\left(\frac{1}{1-p^{(1-m/2)}}-1\right)
=12p(1m/2)1p1m/2.\displaystyle=\frac{1-2p^{(1-m/2)}}{1-p^{1-m/2}}.

As a consequence of the corollary and the calculation preceding 6.3, one has μp(Bν,p)(12p1m/21p1m/2)p(m2)k\mu_{p}(B_{\nu,p})\geq\left(\frac{1-2p^{1-m/2}}{1-p^{1-m/2}}\right)p^{(m-2)k}. Taking the product over all pp the product p(12p1m/21p1m/2)\prod_{p}\left(\frac{1-2p^{1-m/2}}{1-p^{1-m/2}}\right) is just a positive constant, so one obtains pμp(Bν,p)cνm2\prod_{p}\mu_{p}(B_{\nu,p})\geq c\cdot\nu^{m-2} for some constant c>0c>0, hence the result.

6.3 Splittings of Jacobians

In this case we have (𝐆S=Spg,g)(\mathbf{G}_{S}=\operatorname{Sp}_{g},\mathbb{H}_{g}), with g\mathbb{H}_{g} the Siegel upper half space, and (𝐌=Spk×Spgk,k×gk)(\mathbf{M}=\operatorname{Sp}_{k}\times\operatorname{Sp}_{g-k},\mathbb{H}_{k}\times\mathbb{H}_{g-k}).

Lemma 6.5.

There are only finitely many \mathbb{Q}-coset equivalence classes of Hodge subdata of (Spg,g)(\operatorname{Sp}_{g},\mathbb{H}_{g}) which are isomorphic to (Spk×Spgk,k×gk)(\operatorname{Sp}_{k}\times\operatorname{Sp}_{g-k},\mathbb{H}_{k}\times\mathbb{H}_{g-k}).

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 \mathbb{Q}-coset equivalence. Given two such subdatum (𝐌,DM)(\mathbf{M},D_{M}) and (𝐌,DM)(\mathbf{M}^{\prime},D_{M^{\prime}}) of the Hodge datum (Spg,g)(\operatorname{Sp}_{g},\mathbb{H}_{g}), one knows that the subset of End(2g)\textrm{End}(\mathbb{Q}^{2g}) stabilized by the conjugation action of 𝐌\mathbf{M} (resp. 𝐌\mathbf{M}^{\prime}) is the 22-dimensional subspace spanned by {e,1e}\{e,1-e\} (resp. {e,1e}\{e^{\prime},1-e^{\prime}\}) where e:mme:\mathbb{Q}^{m}\to\mathbb{Q}^{m} (resp. e:mme^{\prime}:\mathbb{Q}^{m}\to\mathbb{Q}^{m}) is a non-trivial idempotent. A real element gSpg()g\in\operatorname{Sp}_{g}(\mathbb{R}) that conjugates the two Hodge subdata then induces an equality of algebras

gspan{e,1e}g1=span{e,1e}.g\,\textrm{span}_{\mathbb{R}}\{e,1-e\}g^{-1}=\textrm{span}_{\mathbb{R}}\{e^{\prime},1-e^{\prime}\}.

In particular geg1geg^{-1} is a non-trivial idempotent in the algebra span{e,1e}\textrm{span}_{\mathbb{R}}\{e^{\prime},1-e^{\prime}\}, necessarily equal to either ee^{\prime} or 1e1-e^{\prime}. Since 𝐌\mathbf{M} (resp. 𝐌\mathbf{M}^{\prime}) is exactly the stabilizer in Spg\operatorname{Sp}_{g} of ee (resp. ee^{\prime}), it follows after applying the isomorphism Spg/𝐌Spge\operatorname{Sp}_{g}/\mathbf{M}\simeq\operatorname{Sp}_{g}\cdot e that the coset g𝐌g\mathbf{M} is defined over \mathbb{Q}. ∎

To deduce our results, we compare the degrees of isogenies associated to kk-factors of an abelian variety AA with denominators of Hodge-theoretic idempotents.

Definition 6.6.

For a weight one integral Hodge structure VV, we say that VV has a kk-factor of denominator ν\nu if there exists a rank 2k2k-Hodge-theoretic idempotent e:VVe:V_{\mathbb{Q}}\to V_{\mathbb{Q}} such that νeEnd(V)\nu e\in\textrm{End}(V) is primitive.

Lemma 6.7.

Suppose that AA is a complex abelian variety, and that VV is the associated weight one integral Hodge structure. Then

  • -

    if AA has a kk-factor of degree dividing ν\nu, then VV has a kk-factor of denominator dividing ν\nu; and

  • -

    if VV has a kk-factor of denominator dividing ν\nu, then AA has a kk-factor of degree dividing ν2g\nu^{2g}.

Proof.

Suppose first that AA has a kk-factor of degree dd dividing ν\nu. Let VV^{\prime} be an integral weight one Hodge structure with α:VV\alpha:V\to V^{\prime} an isogeny of degree dd such that V=V1V2V^{\prime}=V_{1}\oplus V_{2} is a decomposition of weight 11 integral Hodge structures, and V1V_{1} has rank 2k2k. Let e:VV1e^{\prime}:V^{\prime}\to V_{1} be the associated Hodge-theoretic idempotent. Then e=α1eαe=\alpha^{-1}e^{\prime}\alpha is a Hodge-theoretic idempotent VVV_{\mathbb{Q}}\to V_{\mathbb{Q}} of rank 2k2k and (detα)e(\det\alpha)e is integral, so the denominator of ee divides (detα)=d(\det\alpha)=d, and d|νd|\nu by assumption.

For the second statement we start with a rank 2k2k Hodge-theoretic idempotent e:VVe:V_{\mathbb{Q}}\to V_{\mathbb{Q}} of VV such that νeEnd(V)\nu e\in\textrm{End}(V) and construct an isogenous Hodge structure VV^{\prime}. As a lattice, we define V=span{e(V),(1e)(V)}V^{\prime}=\textrm{span}_{\mathbb{Z}}\{e(V),(1-e)(V)\}. It is clear that VV^{\prime} is preserved by ee, and that the Hodge decomposition on VV_{\mathbb{C}} induces a Hodge-decomposition on VV^{\prime}_{\mathbb{C}}. Moreover one has VVV\subset V^{\prime}, since v=ev+(1e)vv=ev+(1-e)v for any vVv\in V. Then the inclusion VVV\subset V^{\prime} is an isogeny α\alpha. Its degree is the index [V:V][V^{\prime}:V]. We then have

ν2g=[1νV:V]=[1νV:V][V:V],\nu^{2g}=\left[\frac{1}{\nu}V:V\right]=\left[\frac{1}{\nu}V:V^{\prime}\right][V^{\prime}:V],

hence (degα)|ν2g(\deg\alpha)|\nu^{2g}. ∎

We define

SPk,ν:={sS():𝕍s has a k-factor of denominator dividing ν}.\displaystyle\textrm{SP}^{\prime}_{k,\nu}:=\{s\in S(\mathbb{C}):\mathbb{V}_{s}\textrm{ has a }k\textrm{-factor of denominator dividing }\nu\}.

Note that our lemma shows that SPk,νSPk,ν\textrm{SP}_{k,\nu}\subset\textrm{SP}^{\prime}_{k,\nu} and SPk,νSPk,ν2g\textrm{SP}^{\prime}_{k,\nu}\subset\textrm{SP}_{k,\nu^{2g}}.

Our result 1.10 now follows immediately from 1.25 and 6.7, where we note that components of the split Jacobian locus become components of the Noether-Lefschetz locus upon replacing 𝕍\mathbb{V} with End(𝕍)\textrm{End}(\mathbb{V}). In particular, one has

2dimOk\displaystyle 2\dim O_{k} =2(dimSpgdim(Spk×Spgk))\displaystyle=2(\dim\operatorname{Sp}_{g}-\dim(\operatorname{Sp}_{k}\times\operatorname{Sp}_{g-k}))
=2(g(2g+1)(k(2k+1)+(gk)(2(gk)+1)))\displaystyle=2(g(2g+1)-(k(2k+1)+(g-k)(2(g-k)+1)))
=8k(gk).\displaystyle=8k(g-k).

For the lower bound, we may first prove it for SPk,ν\textrm{SP}^{\prime}_{k,\nu}, and then use that SPk,ν2g\textrm{SP}_{k,\nu^{2g}}. For this we use that both 𝐆S\mathbf{G}_{S} and 𝐌\mathbf{M} are semisimple and simply-connected, so we can reason exactly as before to deduce the result from 4.18 and 1.23, our discussion in §6.1.

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 RR-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 SS-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 pp-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.