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

Effective Methods for Diophantine Finiteness

David Urbanik
Abstract

Let KK\subset\mathbb{C} be a number field, and let 𝒪K,N=𝒪K[N1]\mathcal{O}_{K,N}=\mathcal{O}_{K}[N^{-1}] be its ring of NN-integers. Recently, Lawrence and Venkatesh proposed a general strategy for proving the Shafarevich conjecture for the fibres of a smooth projective family f:XSf:X\to S defined over 𝒪K,N\mathcal{O}_{K,N}. To carry out their strategy, one needs to be able to decide whether the algebraic monodromy group 𝐇Z\mathbf{H}_{Z} of any positive-dimensional geometrically irreducible subvariety ZSZ\subset S_{\mathbb{C}} is “large enough”, in the sense that a certain orbit of 𝐇Z\mathbf{H}_{Z} in a variety of Hodge flags has dimension bounded from below by a certain quantity. In this article we give an effective method for deciding this question. Combined with the effective methods of Lawrence-Venkatesh for understanding semisimplifications of global Galois representations using pp-adic Hodge theory, this gives a fully effective strategy for solving Shafarevich-type problems for arbitrary families ff.

1 Introduction

Let f:XSf:X\to S be a smooth projective family defined over 𝒪K,N=𝒪K[N1]\mathcal{O}_{K,N}=\mathcal{O}_{K}[N^{-1}], with KK\subset\mathbb{C} a number field, 𝒪K\mathcal{O}_{K} its ring of integers, and NN any non-zero element of \mathbb{Z}. In a recent paper [LV20], Lawrence and Venkatesh proposed a strategy for bounding the dimension of the Zariski closure S(𝒪K,N)¯Zar\overline{S(\mathcal{O}_{K,N})}^{\textrm{Zar}} of S(𝒪K,N)S(\mathcal{O}_{K,N}). Since for any point sS(𝒪K,N)s\in S(\mathcal{O}_{K,N}) the fibre XsX_{s} has good reduction away from NN, the problem of bounding the dimension of S(𝒪K,N)¯Zar\overline{S(\mathcal{O}_{K,N})}^{\textrm{Zar}} can be interpreted as Shafarevich-type problem for the family ff. In particular, the Shaferevich conjectures for curves, abelian varieties, and K3 surfaces are equivalent to showing that dimS(𝒪K,N)¯Zar=0\dim\overline{S(\mathcal{O}_{K,N})}^{\textrm{Zar}}=0 for particular choices of ff.

The Lawrence-Venkatesh strategy has been implemented to reprove various classical diophantine finiteness results, including the Mordell conjecture [LV20], the finiteness of SS-units [LV20], and Seigel’s theorem [Noo21]. However it is of particular interest to apply the strategy in situations where dimS>1\dim S>1. In this situation the problem becomes much harder, because one needs some understanding of the monodromy of the family ff over essentially arbitrary geometrically irreducible subvarieties ZSZ\subset S_{\mathbb{C}}. To explain what we mean, let 𝕍=Rif\mathbb{V}=R^{i}f_{*}\mathbb{Z} and define for each such ZSZ\subset S_{\mathbb{C}}:

Definition 1.1.

The algebraic monodromy group 𝐇Z\mathbf{H}_{Z} of ZZ is the identity component of Zariski closure of the monodromy representation associated to 𝕍|Znor{\left.\kern-1.2pt\mathbb{V}\vphantom{\big{|}}\right|_{Z^{\textrm{nor}}}}, where ZnorZZ^{\textrm{nor}}\to Z is the normalization.

The variation 𝕍\mathbb{V} determines a flag variety Lˇ\widecheck{{L}}, on which the group 𝐇Z\mathbf{H}_{Z} can be said to act after identifying Lˇ\widecheck{{L}} with the variety of Hodge flags on some fibre 𝕍s\mathbb{V}_{s} for sZ()s\in Z(\mathbb{C}). Let φ:SΓ\D\varphi:S\to\Gamma\backslash D be a period map determined by 𝕍\mathbb{V} with DLˇD\subset\widecheck{{L}} the complex submanifold of polarized Hodge flags. Then the key quantity relevant for the Lawrence-Venkatesh method is

Δ=minZ,s[dim(𝐇ZFs)dimφ(Z)],\Delta=\min_{Z,s}\left[\dim(\mathbf{H}_{Z}\cdot F^{\bullet}_{s})-\dim\varphi(Z)\right], (1)

where the minimum is taken over all positive-dimensional geometrically irreducible subvarieties ZSZ\subset S_{\mathbb{C}}, points sZ()s\in Z(\mathbb{C}), and where FsF^{\bullet}_{s} is the Hodge flag on 𝕍s\mathbb{V}_{s}.111We not here that φ(Z)\varphi(Z) is analytically constructible, for instance by applying the main result of [BBT18], so its dimension makes sense. In particular, the Lawrence-Venkatesh method produces an integer kk, and shows that if Δk\Delta\geq k, then the Shafarevich conjecture holds for ff.

For successful applications of the Lawrence-Venkatesh strategy for the Shafarevich problem in situations when dimS>1\dim S>1 we know of only the paper [LS20] of Lawrence and Sawin, who are able to apply this strategy beyond the first induction step to prove a Shafarevich conjecture for hypersurfaces lying inside a fixed abelian variety AA. Their methods require the auxilliary use of a Tannakian category associated to AA, and it seems unclear what to do without this abelian variety structure present.

Our main result is as follows:

Theorem 1.2.

Consider a smooth projective family f:XSf:X\to S defined over 𝒪K,N\mathcal{O}_{K,N} and with SS smooth and quasi-projective, and given an integer dd, define

Δd=minZ,s[dim(𝐇ZFs)dimφ(Z)],\Delta_{d}=\min_{Z,s}\left[\dim(\mathbf{H}_{Z}\cdot F^{\bullet}_{s})-\dim\varphi(Z)\right], (2)

where the minimum ranges over all geometrically-irreducible subvarieties ZSZ\subset S_{\mathbb{C}} with dimZ>d\dim Z>d and points sZ()s\in Z(\mathbb{C}). Then there exists an effective procedure which outputs an infinite sequence of integers

κ(1)κ(2)κ(r)<Δd\kappa(1)\leq\kappa(2)\leq\cdots\leq\kappa(r)\leq\cdots<\Delta_{d}

such that for some r=r0r=r_{0} we have κ(r0)=Δd1\kappa(r_{0})=\Delta_{d}-1.

If moreover the period map φ\varphi is quasi-finite, one can determine r0r_{0}.

Remark.

Let us make absolutely clear what is meant by the term “effective procedure” in 1.2. We mean that there exists an infinite-time algorithm (for instance, a non-halting Turing machine), which outputs a sequence of integers {κ(r)}r=1\{\kappa(r)\}_{r=1}^{\infty}, with the integer κ(r)\kappa(r) being outputted at some finite time trt_{r} depending on rr. Moreover, one also has at time trt_{r} a proof that κ(r)<Δd\kappa(r)<\Delta_{d}. Therefore, after time trt_{r}, one can stop the algorithm and use the bound κ(r)<Δd\kappa(r)<\Delta_{d} as input to the Lawrence-Venkatesh method. One is also guaranteed that there is some r0r_{0} so that at time tr0t_{r_{0}} the bound κ(r0)<Δd\kappa(r_{0})<\Delta_{d} is best possible, however one does not necessarily have a method to determine r0r_{0} unless φ\varphi is quasi-finite. Finally, the algorithm can be described entirely in terms of algebro-geometric computations involving algebraically constructible sets, and implicit in the proof is a description of how to implement it.

There is no fundamental obstruction which requires us to restrict to quasi-finite φ\varphi for the second part of 1.2. Rather, the second part of 1.2 references some delicate arguments in [Urb21b] which are only enough to handle the quasi-finite case directly, and recalling enough of the machinery of [Urb21b] to carry out the argument for the general case would lead us too far astray from the main ideas. We note that one does not actually need to determine the integer r0r_{0} referenced in 1.2 to apply the machinery of Lawrence and Venkatesh: one wants to be able to compute the best possible lower bound for Δd\Delta_{d}, but one is not required to actually prove that the bound one has is optimal in order to deduce diophantine finiteness results.

1.1 The Approach of Lawrence and Venkatesh

We begin with a preliminary observation. To show that dimS(𝒪K,N)¯Zard\dim\overline{S(\mathcal{O}_{K,N})}^{\textrm{Zar}}\leq d, it suffices to show, for any irreducible subscheme TST\subset S of dimension >d>d and defined over 𝒪K,N\mathcal{O}_{K,N}, that the Zariski closure T(𝒪K,N)¯Zar\overline{T(\mathcal{O}_{K,N})}^{\textrm{Zar}} is a proper algebraic subscheme of TT. We therefore fix such a subscheme TST\subset S with dimT>d\dim T>d, and seek to show that dimT(𝒪K,N)¯Zar<dimT\dim\overline{T(\mathcal{O}_{K,N})}^{\textrm{Zar}}<\dim T.

Fix a prime pp\in\mathbb{Z} not dividing NN, and let tT(𝒪K,N)t\in T(\mathcal{O}_{K,N}) be a point. It is conjectured that for any ii the representation of Gal(K¯/K)\textrm{Gal}(\overline{K}/K) on Héti(Xt,K¯,p)H^{i}_{\textrm{\'{e}t}}(X_{t,\overline{K}},\mathbb{Q}_{p}) is semisimple. If this result were to be known for all such tt, an argument of Faltings shows that for tT(𝒪K,N)t\in T(\mathcal{O}_{K,N}) there are at most finitely many possibilities for the isomorphism class of the representation of Gal(K¯/K)\textrm{Gal}(\overline{K}/K) on Héti(Xt,K¯,p)H^{i}_{\textrm{\'{e}t}}(X_{t,\overline{K}},\mathbb{Q}_{p}). To establish the non Zariski-density of T(𝒪K,N)T(\mathcal{O}_{K,N}) it would then suffice to show that the fibres of the map

tT(𝒪K,N)𝜏{Gal(K¯/K)-rep. on Héti(Xt,K¯,p)}/iso.t\in T(\mathcal{O}_{K,N})\hskip 10.00002pt\xrightarrow{\tau}\hskip 10.00002pt\big{\{}\textrm{Gal}(\overline{K}/K)\textrm{-rep. on }H^{i}_{\textrm{\'{e}t}}(X_{t,\overline{K}},\mathbb{Q}_{p})\big{\}}\hskip 5.0pt\big{/}\hskip 5.0pt\textrm{iso}.

are not Zariski dense. As explained by Lawrence and Venkatesh in [LV20], this is essentially the original argument for the Mordell conjecture due to Faltings.

The problem with applying this strategy in general is twofold. First, the semisimplicity of Héti(Xt,K¯,p)H^{i}_{\textrm{\'{e}t}}(X_{t,\overline{K}},\mathbb{Q}_{p}) is not known, and for most choices of ff appears out of reach. Secondly, without a good geometric interpretation of the étale cohomology Héti(Xt,K¯,p)H^{i}_{\textrm{\'{e}t}}(X_{t,\overline{K}},\mathbb{Q}_{p}) it is difficult to understand τ\tau. The key insight in the paper of Lawrence and Venkatesh is that one may potentially overcome both problems by passing to a pp-adic setting where they are more managable.

Instead of considering the global Galois representation ρt:Gal(K¯/K)Héti(Xt,K¯,p)\rho_{t}:\textrm{Gal}(\overline{K}/K)\to H^{i}_{\textrm{\'{e}t}}(X_{t,\overline{K}},\mathbb{Q}_{p}), we consider its semisimplification ρtss\rho^{\textrm{ss}}_{t}, and restrict ρt\rho_{t} along the map Gal(Kv¯/Kv)Gal(K¯/K)\textrm{Gal}(\overline{K_{v}}/K_{v})\to\textrm{Gal}(\overline{K}/K) induced by a fixed embedding K¯Kv¯\overline{K}\hookrightarrow\overline{K_{v}} to obtain ρt,v\rho_{t,v}, where vv is a fixed place above pp. The functors of pp-adic Hodge theory tell us that the representation ρt,v\rho_{t,v} determines a triple (HdRi(Xt,Kv),ϕt,Ft)(H^{i}_{\textrm{dR}}(X_{t,K_{v}}),\phi_{t},F^{\bullet}_{t}), where FtF^{\bullet}_{t} is the Hodge filtration on de Rham cohomology and ϕt\phi_{t} is the crystalline Frobenius at tt. If we somehow manage to consider the data (HdRi(Xt,Kv),ϕt,Ft)(H^{i}_{\textrm{dR}}(X_{t,K_{v}}),\phi_{t},F^{\bullet}_{t}) up to “semisimplification” (in the sense that we identify such triples when the associated global Galois representations have isomorphic semisimplifications), our problem is then to study the fibres of the map

tT(𝒪K,N)τp{“semisimplifications” of (HdRi(Xt,Kv),ϕt,Ft)}/iso.t\in T(\mathcal{O}_{K,N})\hskip 5.0pt\xrightarrow{\tau_{p}}\hskip 5.0pt\big{\{}\textrm{``semisimplifications'' of }(H^{i}_{\textrm{dR}}(X_{t,K_{v}}),\phi_{t},F^{\bullet}_{t})\big{\}}\hskip 5.0pt\big{/}\hskip 5.0pt\textrm{iso}.

and show they lie in a Zariski closed subscheme of smaller dimension.

Next, we make the elementary observation that to bound the dimension of the Zariski closure of T(𝒪K,N)T(\mathcal{O}_{K,N}), it suffices to cover T(𝒪K,v)T(\mathcal{O}_{K,v}) by finitely many vv-adic disks D1,,DkD_{1},\ldots,D_{k} and bound the dimension of the Zariski closure of DiT(𝒪K,N)D_{i}\cap T(\mathcal{O}_{K,N}) for each ii; here 𝒪K,v\mathcal{O}_{K,v} is the ring of vv-adic integers. It can then be shown that there exists such a cover for which the Hodge bundle =RifΩX/S\mathcal{H}=R^{i}f_{*}\Omega^{\bullet}_{X/S} can be trivialized rigid-analytically over each DiD_{i}, moreover with respect to each such trivialilzation the Frobenius operator ϕt\phi_{t} is independent of tDit\in D_{i}. The problem then reduces to studying a varying filtration FtF^{\bullet}_{t} on a fixed vector space Vp=HdRi(Xt0)V_{p}=H^{i}_{\textrm{dR}}(X_{t_{0}}) for some t0Dit_{0}\in D_{i}. In particular, we obtain a rigid-analytic map

Diψp{Hodge filtrations on Vp}Lˇp,D_{i}\hskip 5.0pt\xrightarrow{\psi_{p}}\hskip 5.0pt\underbrace{\{\hskip 3.00003pt\textrm{Hodge filtrations on }V_{p}\hskip 3.00003pt\}}_{\widecheck{{L}}_{p}},

and those points of Lˇp\widecheck{{L}}_{p} arising from points tT(𝒪K,N)Dit\in T(\mathcal{O}_{K,N})\cap D_{i} lie inside finitely many subvarieties Oi1,,OiO_{i1},\ldots,O_{i\ell} of Lˇp\widecheck{{L}}_{p} corresponding to the finitely many possible isomorphism classes of ρtss\rho^{\textrm{ss}}_{t}.

We are now faced with the problem of understanding the intersections ψp(Di)Oij\psi_{p}(D_{i})\cap O_{ij}, and showing that their inverse images under ψp1\psi_{p}^{-1} lie in an algebraic subscheme of smaller dimension. One part of this problem is to understand the dimensions of the varieties OijO_{ij}, and to show that they are sufficiently small: this step is carried out successfully for the families of hypersurfaces studied both by Lawrence-Venkatesh and Lawrence-Sawin, and appears to be managable in general. The more difficult object to control is ψp(Di)\psi_{p}(D_{i}), for which we need to understand the variation of the filtration FF^{\bullet} over DiD_{i}. This, in turn, is governed by the Gauss-Manin connection :ΩT1\nabla:\mathcal{H}\otimes\Omega^{1}_{T}\to\mathcal{H}, which exists universally over SS after possibly increasing NN; we may adjust pp so that it does not divide NN if necessary. The fact that \nabla exists universally over 𝒪K,N\mathcal{O}_{K,N} means that the same system of differential equations satisfied by ψp\psi_{p} at tT(𝒪K,N)Dit\in T(\mathcal{O}_{K,N})\cap D_{i} is also satisfied by a Hodge-theoretic period map ψ:BLˇ\psi:B\to\widecheck{{L}} on a sufficiently small analytic neighbourhood BT()B\subset T(\mathbb{C}) containing tt, where Lˇ\widecheck{{L}} is a variety of Hodge flags. In particular, one can prove that the Zariski closures of ψp(Di)\psi_{p}(D_{i}) and ψ(B)\psi(B) have the same dimension.

The final step, which is to show that the Zariski closure of T(𝒪K,N)DiT(\mathcal{O}_{K,N})\cap D_{i} in TT has smaller dimension, is completed as follows. The Ax-Schanuel Theorem [BT17] for variations of Hodge structures due to Bakker-Tsimerman shows that if VV is an algebraic subvariety of Lˇ\widecheck{{L}} of dimension at most dimψ(B)¯Zardimψ(B)\dim\overline{\psi(B)}^{\textrm{Zar}}-\dim\psi(B),222The dimension dimψ(B)\dim\psi(B) can once again, at least for open neighbourhoods BB with a sufficiently mild geometry, be made sense of the dimension of a locally constructible analytic set. Alternatively one can replace dimψ(B)\dim\psi(B) with dimφ(Z)\dim\varphi(Z), where φ\varphi is as before and ZTZ\subset T_{\mathbb{C}} is a component containing BB. then the inverse image under ψ\psi of the intersection ψ(B)V\psi(B)\cap V lies in an algebraic subvariety of TT_{\mathbb{C}} of smaller dimension. Choosing an isomorphism Kv¯\mathbb{C}\cong\overline{K_{v}} one can transfer this fact to the same statement for the map ψp\psi_{p} and in particular for V=OijV=O_{ij}. Our problem is finally reduced to giving a lower bound for the difference dimψ(B)¯Zardimψ(B)\dim\overline{\psi(B)}^{\textrm{Zar}}-\dim\psi(B). Our main result now reads:

Theorem 1.3.

Define the quantity

Δd:=minZ,ψ[dimψ(B)¯Zardimψ(B)],\Delta_{d}:=\min_{Z,\psi}\left[\dim\overline{\psi(B)}^{\textrm{Zar}}-\dim\psi(B)\right],

where ZZ ranges over all irreducible complex algebraic subvarieties ZSZ\subset S_{\mathbb{C}} of dimension greater than dd, and where ψ\psi is any complex analytic period map determined by the variation of Hodge structures 𝕍=Rif\mathbb{V}=R^{i}f_{*}\mathbb{Z} and defined on a neighbourhood BZ()B\subset Z(\mathbb{C}). Then there exists an effective procedure which outputs an infinite sequence of lower bounds

κ(1)κ(2)κ(r)<Δd\kappa(1)\leq\kappa(2)\leq\cdots\leq\kappa(r)\leq\cdots<\Delta_{d}

such that for some r=r0r=r_{0} we have κ(r0)=Δd1\kappa(r_{0})=\Delta_{d}-1.

If moreover the period map φ\varphi is quasi-finite, one can determine r0r_{0}.

We note that by [Urb21b, Lem. 4.10(ii)] it is also a consequence of the Bakker-Tsimerman Theorem that when ZZ is geometrically irreducible, we have ψ(B)¯Zar=𝐇Zψ(t)\overline{\psi(B)}^{\textrm{Zar}}=\mathbf{H}_{Z}\cdot\psi(t) for any point tZ()t\in Z(\mathbb{C}), which recovers the statement of 1.2.

1.2 Basic Idea of the Method

We may observe that the computation of the bound described in 1.3 is a purely Hodge-theoretic problem, i.e., it concerns only properties of the integral variation 𝕍=Rif\mathbb{V}=R^{i}f_{*}\mathbb{Z} of Hodge structures on SS_{\mathbb{C}}. Let 𝒬:𝕍𝕍\mathcal{Q}:\mathbb{V}\otimes\mathbb{V}\to\mathbb{Z} be a polarization of 𝕍\mathbb{V}, and let (V,Q)(V,Q) be a fixed polarized lattice isomorphic to one (hence any) fibre of (𝕍,𝒬)(\mathbb{V},\mathcal{Q}); as it causes no harm, we will assume that V=mV=\mathbb{Z}^{m} for some mm, and therefore sometimes write GLm\textrm{GL}_{m} for GL(V)\textrm{GL}(V). Let DD be the complex manifold parametrizing polarized Hodge structures on (V,Q)(V,Q) with the same Hodge numbers as (𝕍,𝒬)(\mathbb{V},\mathcal{Q}). A point hDh\in D we may view as a morphism h:𝕊GL(V)h:\mathbb{S}\to\textrm{GL}(V)_{\mathbb{R}}, where 𝕊\mathbb{S} is the Deligne torus, and the Mumford-Tate group MT(h)\textrm{MT}(h) is the \mathbb{Q}-Zariski closure of h(𝕊)h(\mathbb{S}).

To present our method, we introduce some terminology:

Notation.

We denote by Lˇ\widecheck{{L}} the \mathbb{Q}-algebraic variety of all Hodge flags on the lattice VV, not necessarily polarized. We note that DD is an open submanifold of a closed \mathbb{Q}-algebraic subvariety DˇLˇ\widecheck{{D}}\subset\widecheck{{L}}.

Definition 1.4.

Given two subvarieties W1,W2LˇW_{1},W_{2}\subset\widecheck{{L}}, we say that W1GLW2W_{1}\sim_{\textrm{GL}}W_{2} if there exists gGLm()g\in\textrm{GL}_{m}(\mathbb{C}) such that gW1=W2g\cdot W_{1}=W_{2}. Given a variety WLˇW\subset\widecheck{{L}}, we call the equivalence class 𝒞(W)\mathcal{C}(W) under GL\sim_{\textrm{GL}} a type. The dimension of a type 𝒞(W)\mathcal{C}(W) is the dimension of WW.

Definition 1.5.

We say that a type 𝒞\mathcal{C} is Hodge-theoretic if 𝒞=𝒞(W)\mathcal{C}=\mathcal{C}(W), where W=N()hW=N(\mathbb{C})\cdot h for hDh\in D and NN a \mathbb{Q}-algebraic normal subgroup of MT(h)\textrm{MT}(h).

The first step in our algorithm is:

Step One: Compute a finite list of types 𝒞1,,𝒞\mathcal{C}_{1},\ldots,\mathcal{C}_{\ell} such that every Hodge-theoretic type appears somewhere in the list.

When we say to compute a type 𝒞\mathcal{C}, we mean to compute a representative WLˇW\subset\widecheck{{L}} such that 𝒞=𝒞(W)\mathcal{C}=\mathcal{C}(W). That there are only finitely many Hodge-theoretic types is shown in 2.1 below.

The problem given in Step One is solved in [Urb21b, Prop. 5.4]; we will say little about it here. It is related to the problem of classifying Mumford-Tate groups up to conjugacy by GLm()\textrm{GL}_{m}(\mathbb{C}), for which one can use a constructive version of the proof in [Voi12, Thm. 4.14]. It is also similar to the problem of classifying Mumford-Tate domains as studied in [GGK12, Chap. VII]. We note that the methods of [GGK12, Chap. VII], when they can be carried out effectively, result in an approach for which 𝒞1,,𝒞\mathcal{C}_{1},\ldots,\mathcal{C}_{\ell} will be exactly the set of Hodge-theoretic types.

The second step is more involved, and is the crux of our method. To describe it we need to introduce some terminology.

Definition 1.6.

A local period map is a map ψ:BLˇ\psi:B\to\widecheck{{L}} obtained as a composition ψ=qA\psi=q\circ A, where:

  • (i)

    The set BS()B\subset S(\mathbb{C}) is a connected analytic neighbourhood on which 𝕍\mathbb{V} is constant and FkF^{k}\mathcal{H} is trivial for each kk, where =𝕍𝒪San\mathcal{H}=\mathbb{V}\otimes\mathcal{O}_{S^{\textrm{an}}}.

  • (ii)

    The map A:BGLm()A:B\to\textrm{GL}_{m}(\mathbb{C}) is a varying filtration-compatible period matrix over BB. More precisely, there exists a basis v1,,vmv^{1},\ldots,v^{m} for (B)\mathcal{H}(B), compatible with the filtration in the sense that Fk(B)F^{k}\mathcal{H}(B) is spanned by v1,,vikv^{1},\ldots,v^{i_{k}} for some iki_{k}, and a flat frame b1,,bmb^{1},\ldots,b^{m} for 𝕍(B)\mathbb{V}_{\mathbb{C}}(B), such that A(s)A(s) is the change-of-basis matrix from vs1,,vsmv^{1}_{s},\ldots,v^{m}_{s} to bs1,,bsmb^{1}_{s},\ldots,b^{m}_{s}.

  • (iii)

    The map q:GLmLˇq:\textrm{GL}_{m}\to\widecheck{{L}} sends a matrix MM to the Hodge flag FMF_{M}^{\bullet} defined by the property that FMkF^{k}_{M} is spanned by the first iki_{k} columns.

To summarize the preceding definition: a local period map is exactly a period map on BB except one does not necessarily compute periods with respect to the integral lattice 𝕍(B)𝕍(B)\mathbb{V}(B)\subset\mathbb{V}_{\mathbb{C}}(B) but is instead allowed to consider periods with respect to a more general complex flat frame. There is a natural GLm()\textrm{GL}_{m}(\mathbb{C})-action on the set of germs of local period maps at a point sS()s\in S(\mathbb{C}), where MGLm()M\in\textrm{GL}_{m}(\mathbb{C}) acts on the map ψ=qA\psi=q\circ A to give Mψ=q(MA)M\cdot\psi=q\circ(M\cdot A). This action corresponds exactly to a change of the flat frame b1,,bmb^{1},\ldots,b^{m}, and all germs of local period maps at ss lie in a single GLm()\textrm{GL}_{m}(\mathbb{C})-orbit.

The construction of a local period map ψ:BLˇ\psi:B\to\widecheck{{L}} involves picking a basis b1,,bmb^{1},\ldots,b^{m} of 𝕍(B)\mathbb{V}_{\mathbb{C}}(B), and hence choosing an isomorphism 𝕍(B)m\mathbb{V}_{\mathbb{C}}(B)\simeq\mathbb{C}^{m}. When working with a local period map, we will always assume that such a basis has been choosen, and hence identify subgroups of GL(𝕍(B))\textrm{GL}(\mathbb{V}_{\mathbb{C}}(B)) with subgroups of GLm()\textrm{GL}_{m}(\mathbb{C}). In particular, if ZSZ\subset S_{\mathbb{C}} is a geometrically irreducible subvariety which intersects BB, we have an induced action of 𝐇Z\mathbf{H}_{Z} on Lˇ\widecheck{{L}}.

Lastly, we need:

Definition 1.7.

Given two types 𝒞1\mathcal{C}_{1} and 𝒞2\mathcal{C}_{2}, we say that 𝒞1𝒞2\mathcal{C}_{1}\leq\mathcal{C}_{2} if there exists WiLˇW_{i}\subset\widecheck{{L}} for 1=1,21=1,2 such that 𝒞i=𝒞(Wi)\mathcal{C}_{i}=\mathcal{C}(W_{i}) and W1W2W_{1}\subset W_{2}.

Definition 1.8.

Given a local period map ψ:BLˇ\psi:B\to\widecheck{{L}} and a geometrically irreducible subvariety ZSZ\subset S_{\mathbb{C}} intersecting BB at ss, we call 𝒞(ψ(BZ)¯Zar)=𝒞(𝐇Zψ(s))\mathcal{C}(\overline{\psi(B\cap Z)}^{\textrm{Zar}})=\mathcal{C}(\mathbf{H}_{Z}\cdot\psi(s)) the type of ZZ, and denote it by 𝒞(Z)\mathcal{C}(Z).

For well-definedness, see 2.2 below. From Step One, we have computed a finite list 𝒞1,,𝒞\mathcal{C}_{1},\ldots,\mathcal{C}_{\ell} of types containing all types that can arise from the variation 𝕍\mathbb{V}. Our next task is then:

Step Two: For each type 𝒞i\mathcal{C}_{i} appearing in the list, compute a differential system 𝒯(𝒞i)\mathcal{T}(\mathcal{C}_{i}) on SS characterized by the property that an algebraic subvariety ZSZ\subset S_{\mathbb{C}} is an integral subvariety for 𝒯(𝒞i)\mathcal{T}(\mathcal{C}_{i}) if and only if 𝒞(Z)𝒞i\mathcal{C}(Z)\leq\mathcal{C}_{i}, and determine the dimension of a maximal integral subvariety for this system.

We explain precisely what we mean by “differential system” in section 3; actually our method does something more subtle than Step Two due to the fact that we can only approximate 𝒯(𝒞i)\mathcal{T}(\mathcal{C}_{i}) up to some finite order, but for expository purposes this is the essential point. After this, we will see the problem is reduced to analyzing which of the differential systems 𝒯(𝒞i)\mathcal{T}(\mathcal{C}_{i}) admit algebraic solutions of “exceptional” dimension, which can be carried out using tools from functional transcendence.

1.3 Acknowledgements

The author thanks Brian Lawrence, Akshay Venkatesh, and Will Sawin for comments on a draft of this manuscript.

2 Algebraic Monodromy Orbits up to Conjugacy

In this section we describe an effective method for solving “Step One” as posed in subsection 1.2. We will also prove some preliminary facts about types used in the introduction, and we continue with the notation established there. We will work in the context of a general polarizable integral variation of Hodge structure 𝕍\mathbb{V} on the complex algebraic variety SS, not necessarily coming from a projective family as in the introduction.

2.1 Basic Properties of Types

Lemma 2.1.

For any geometrically irreducible subvariety ZSZ\subset S and any local period map ψ:BLˇ\psi:B\to\widecheck{{L}} with ZBZ\cap B non-empty, we have

ψ(ZB)¯Zar=𝐇Zψ(s),\overline{\psi(Z\cap B)}^{\textrm{Zar}}=\mathbf{H}_{Z}\cdot\psi(s),

for any point sZ()s\in Z(\mathbb{C}).

Proof.

It suffices to show that

ψ(C)¯Zar=𝐇Zψ(s),\overline{\psi(C)}^{\textrm{Zar}}=\mathbf{H}_{Z}\cdot\psi(s),

for each analytic component CZBC\subset Z\cap B separately, with ss a point of CC. By acting on ψ\psi by an element of GLm()\textrm{GL}_{m}(\mathbb{C}), the claim can be reduced to the situation where the periods which determine ψ\psi are computed with respect to a basis for the integral lattice 𝕍(B)\mathbb{V}(B), and then the claim follows from [Urb21b, Lem. 4.10(ii)]. ∎

Lemma 2.2.

The equivalence class under GL\sim_{\textrm{GL}} of ψ(BZ)¯Zar\overline{\psi(B\cap Z)}^{\textrm{Zar}} is independent of ψ\psi; i.e., the type of ZSZ\subset S is well-defined.

Proof.

Let p:ZsmZp:Z^{\textrm{sm}}\to Z be a smooth resolution, and consider the variation p𝕍p^{*}\mathbb{V}. From the fact that germs of local period maps on ZsmZ^{\textrm{sm}} with respect to the variation p𝕍p^{*}\mathbb{V} factor through germs of local period maps on SS, we may reduce to the same problem for ZsmZ^{\textrm{sm}} and the variation p𝕍p^{*}\mathbb{V}, i.e., we may assume Z=SZ=S. By analytically continuing a fixed local period map ψ\psi to the universal covering S~S\widetilde{S}\to S, we learn from the irreducibility of S~\widetilde{S} that at each point sSs\in S, there exists a local period map ψs:BsLˇ\psi_{s}:B_{s}\to\widecheck{{L}} such that ψs(Bs)¯Zar=ψ(B)¯Zar\overline{\psi_{s}(B_{s})}^{\textrm{Zar}}=\overline{\psi(B)}^{\textrm{Zar}}. Since the Zariski closure of ψs(Bs)\psi_{s}(B_{s}) is determined by the germ of ψs\psi_{s} at ss, and because all germs of local period maps at ss lie in a single GLm()\textrm{GL}_{m}(\mathbb{C})-orbit, the result follows. ∎

Lemma 2.3.

There are only finitely many Hodge-theoretic types.

Proof.

We observe that the problem reduces to the following: show they are finitely many GLm()\textrm{GL}_{m}(\mathbb{C})-equivalence classes of pairs (h,N)(h,N), where

  • (i)

    hDh\in D is a polarized Hodge structure; and

  • (ii)

    NN is a \mathbb{Q}-algebraic connected normal subgroup of MT(h)\textrm{MT}(h);

where we regard GLm()\textrm{GL}_{m}(\mathbb{C}) as acting on hh through its action on Lˇ\widecheck{{L}}, and on NN by conjugation. Note that two such equivalent pairs will generate orbits in Lˇ\widecheck{{L}} equivalent under GL\sim_{\textrm{GL}}. Since the groups MT(h)\textrm{MT}(h) are reductive and have finitely many connected normal algebraic factors, this reduces to the same problem for pairs of the form (h,MT(h))(h,\textrm{MT}(h)). We recall that DD is an open submanifold of Dˇ\widecheck{{D}}, the flag variety of flags satisfying the first Hodge-Riemann bilinear relation (the isotropy condition), and that Dˇ\widecheck{{D}} is an algebraic subvariety of Lˇ\widecheck{{L}}. We then use the fact that there are finitely many Mumford-Tate groups up to GLm()\textrm{GL}_{m}(\mathbb{C})-conjugacy (see [Voi12, Thm. 4.14]), and that for a fixed Mumford-Tate group MM the Hodge structures in DD with Mumford-Tate contained in MM lie inside finitely many M()M(\mathbb{C})-orbits in Dˇ\widecheck{{D}}, see [GGK12, VI.B.9]. ∎

2.2 Computing Types up to Conjugacy

In this section we give some references for carrying out Step One as described in the introduction.

Proposition 2.4.

There exists an algorithm to compute subvarieties W1,,WLˇW_{1},\ldots,W_{\ell}\subset\widecheck{{L}} such that the set of Hodge-theoretic types is a (possibly proper) subset of {𝒞(W1),,𝒞(W)}\{\mathcal{C}(W_{1}),\ldots,\mathcal{C}(W_{\ell})\}.

Proof.

This is solved in [Urb21b, Prop. 5.4]. ∎

Let us comment briefly on a different approach to Step One given in [GGK12, Chap. VII]. In [GGK12, Chap. VII], the authors describe a method for classifying both Mumford-Tate groups and Mumford-Tate domains (orbits of points hDh\in D under MT(h)()\textrm{MT}(h)(\mathbb{R}) and MT(h)()\textrm{MT}(h)(\mathbb{C})). Given an appropriate such classification, one can easily solve Step One by computing the decompositions of the groups MT(h)\textrm{MT}(h) that arise into \mathbb{Q}-simple factors. The method of [GGK12, Chap. VII] is to first classify CM Hodge structures hCMDh_{\textrm{CM}}\in D, and then give a criterion for deciding when a Lie subalgebra of 𝔤𝔩(V)\mathfrak{gl}(V) corresponds to a Mumford-Tate group generating a Mumford-Tate domain containing hCMh_{\textrm{CM}}. They carry out this classification procedure successfully when dimV=4\dim V=4, and so for variations with Hodge numbers (2,2)(2,2) and (1,1,1,1)(1,1,1,1).

The method given for classifying CM Hodge structures given in [GGK12, Chap. VII] is to observe that CM Hodge structures up to isogeny are determined by certain data associated to embeddings of CM fields, and hence the first step of the procedure in [GGK12, Chap. VII] is to “classify all CM fields of rank up to [dimV\dim V] by [their] Galois group”. We are not aware of an effective method for carrying out this step.333The paper [Dod84] gives a potential approach by giving a method to classify certain abstract structures associated with Galois groups of CM fields. However one still needs to determine which such structures are actually associated to a concrete CM field. It is also not clear to us precisely the sense in which the term “classify” is being used; i.e., we do not know what form the data of a “classification of CM Hodge structures” takes, and consequently what form the resulting classification of Mumford-Tate domains will have. For this reason, we were unable to apply the methods of [GGK12, Chap. VII] to prove 2.4.

3 Differential Tools and a Jet Criterion

In this section we introduce a collection of effectively computable algebro-geometric correspondences which can be used for studying systems of differential equations on SS induced by the variation 𝕍\mathbb{V}, and then use it to solve the main problem. We have already carried out most of the work in two preceding papers [Urb21a] and [Urb21b], so we will first need to collect some results. In this section we assume that SS is a KK-variety for KK\subset\mathbb{C} a number field, and that 𝕍\mathbb{V} is a polarizable integral variation of Hodge structure on SS_{\mathbb{C}} such that the vector bundle =𝕍𝒪San\mathcal{H}=\mathbb{V}\otimes_{\mathbb{Z}}\mathcal{O}_{S_{\mathbb{C}}^{\textrm{an}}}, the filtration FF^{\bullet}, and the connection :ΩS1\nabla:\mathcal{H}\to\mathcal{H}\otimes\Omega^{1}_{S} all admit KK-algebraic models. Moreover, we assume that we may effectively compute a description of these objects in terms of finitely-presented KK-modules over an affine cover of SS; for a justification of this assumption in the situation where 𝕍\mathbb{V} comes from a smooth projective KK-algebraic family f:XSf:X\to S see [Urb21b, §2].

3.1 The Constructive Period-Jet Correspondence

Our algebro-geometric correspondences will be formulated using the language of jets. Let Ard=K[t1,,td]/t1,,tdr+1A^{d}_{r}=K[t_{1},\ldots,t_{d}]/\langle t_{1},\ldots,t_{d}\rangle^{r+1}, and define 𝔻rd=SpecArd\mathbb{D}^{d}_{r}=\textrm{Spec}\hskip 1.49994ptA^{d}_{r} to be the dd-dimensional disk of order rr; we suppress the field KK in the notation. A jet space associated to a space XX is a space which parametrizes maps 𝔻rdX\mathbb{D}^{d}_{r}\to X. More formally, for XX a finite-type KK-scheme, we have:

Definition 3.1.

We define JrdXJ^{d}_{r}X to be the scheme representing the contravariant functor SchKSet\textrm{Sch}_{K}\to\textrm{Set} given by

THomK(T×K𝔻rd,X),[TT][HomK(T×K𝔻rd,X)HomK(T×K𝔻rd,X)],T\mapsto\textrm{Hom}_{K}(T\times_{K}\mathbb{D}^{d}_{r},X),\hskip 15.00002pt[T\to T^{\prime}]\mapsto[\textrm{Hom}_{K}(T^{\prime}\times_{K}\mathbb{D}^{d}_{r},X)\to\textrm{Hom}_{K}(T\times_{K}\mathbb{D}^{d}_{r},X)],

where the natural map HomK(T×K𝔻rd,X)HomK(T×K𝔻rd,X)\textrm{Hom}_{K}(T^{\prime}\times_{K}\mathbb{D}^{d}_{r},X)\to\textrm{Hom}_{K}(T\times_{K}\mathbb{D}^{d}_{r},X) obtained by pulling back along T×K𝔻rdT×K𝔻rdT\times_{K}\mathbb{D}^{d}_{r}\to T^{\prime}\times_{K}\mathbb{D}^{d}_{r}.

That the functor defining JrdXJ^{d}_{r}X in 3.1 is representable is handled by [Urb21a, §2]. Moreover, JrdJ^{d}_{r} is itself a functor, sending a map g:XYg:X\to Y to the map Jrdg:JrdXJrdYJ^{d}_{r}g:J^{d}_{r}X\to J^{d}_{r}Y that acts on points by post-composition. For XX an analytic space, there is an analogous construction that appears in [Urb21a, §2.3]. If KK\subset\mathbb{C} is a subfield, this construction is compatible with analytification.

The purpose of introducing jets is the following result, proven in [Urb21b], building on [Urb21a].

Theorem 3.2.

For each d,r0d,r\geq 0, a variation of Hodge structure 𝕍\mathbb{V} on SS gives rise to a canonical map

ηrd:JrdSGLm\JrdLˇ,\eta^{d}_{r}:J^{d}_{r}S\to\textrm{GL}_{m}\backslash J^{d}_{r}\widecheck{{L}},

of algebraic stacks characterized by the property that for any local period map ψ:BLˇ\psi:B\to\widecheck{{L}} and any jet jJrdBj\in J^{d}_{r}B we have ψj=ηrd(j)\psi\circ j=\eta^{d}_{r}(j) modulo GLm()\textrm{GL}_{m}(\mathbb{C}).

Moreover, if the data (,F,)(\mathcal{H},F^{\bullet},\nabla) associated to the variation 𝕍\mathbb{V} admits a KK-algebraic model, the map ηrd\eta^{d}_{r} is defined over KK, and there exists an algorithm to compute the GLm\textrm{GL}_{m}-torsor prd:𝒫rdJrdSp^{d}_{r}:\mathcal{P}^{d}_{r}\to J^{d}_{r}S and the GLm\textrm{GL}_{m}-invariant map αrd:𝒫rdJrdLˇ\alpha^{d}_{r}:\mathcal{P}^{d}_{r}\to J^{d}_{r}\widecheck{{L}} which defines ηrd\eta^{d}_{r} from a presentation of the data (,F,)(\mathcal{H},F^{\bullet},\nabla) in terms of finitely-presented KK-modules.

We note that the computability of the torsor 𝒫rd\mathcal{P}^{d}_{r} in 3.2 has in particular the following consequence: if 𝒮(GLm\JrdLˇ)()\mathcal{S}\subset(\textrm{GL}_{m}\backslash J^{d}_{r}\widecheck{{L}})(\mathbb{C}) is a subset which is the image under the quotient of a constructible LL-algebraic set JrdLˇ\mathcal{F}\subset J^{d}_{r}\widecheck{{L}}, where KLK\subset L is a computable extension, then we can compute (ηrd)1(𝒮)(\eta^{d}_{r})^{-1}(\mathcal{S}) by computing prd((αrd)1())p^{d}_{r}((\alpha^{d}_{r})^{-1}(\mathcal{F})). Thus if we define

Definition 3.3.

For a constructible LL-algebraic set JrdLˇ\mathcal{F}\subset J^{d}_{r}\widecheck{{L}}, with KLK\subset L an extension, we write

𝒯rd():=(ηrd)1(GLm).\mathcal{T}^{d}_{r}(\mathcal{F}):=(\eta^{d}_{r})^{-1}(\textrm{GL}_{m}\cdot\mathcal{F}).

Moreover, for a type 𝒞=𝒞(W)\mathcal{C}=\mathcal{C}(W), we will write either 𝒯rd(𝒞)\mathcal{T}^{d}_{r}(\mathcal{C}) or 𝒯rd(W)\mathcal{T}^{d}_{r}(W) for the set 𝒯rd(JrdW)\mathcal{T}^{d}_{r}(J^{d}_{r}W).

then the main consequence of the preceding discussion for our situation is the following, which is immediate from what we have said:

Proposition 3.4.

For each d,r0d,r\geq 0 there exists an algorithm which, given a constructible LL-algebraic set JrdLˇ\mathcal{F}\subset J^{d}_{r}\widecheck{{L}} with KLK\subset L a computable extension, computes 𝒯rd()JrdS\mathcal{T}^{d}_{r}(\mathcal{F})\subset J^{d}_{r}S. ∎

3.2 Jet Conditions and Types

Let us now try to understand how computing the “differential constraints” induced by types 𝒞(W)\mathcal{C}(W) as in 3.4 can help us carry out Step Two of subsection 1.2. Let Γ=Aut(V,Q)()\Gamma=\textrm{Aut}(V,Q)(\mathbb{Z}), and let φ:SΓ\D\varphi:S_{\mathbb{C}}\to\Gamma\backslash D be the canonical period map which sends a point sS()s\in S(\mathbb{C}) to the isomorphism class of the polarized Hodge structure on 𝕍s\mathbb{V}_{s}. By [BBT18], the map φ\varphi factors as ιp\iota\circ p, where p:STp:S_{\mathbb{C}}\to T is a dominant map of algebraic varieties and ι\iota is a closed embedding of analytic spaces; it follows that for each subvariety ZSZ\subset S_{\mathbb{C}} the dimension of the image φ(Z)\varphi(Z) makes sense as the dimension of a constructible algebraic set.

Fix a sequence of compatible embeddings

SpecK=𝔻r0ι0𝔻r1ι1𝔻r2ι2𝔻r3ι3𝔻r4ι4\textrm{Spec}\hskip 1.49994ptK=\mathbb{D}^{0}_{r}\xhookrightarrow{\iota_{0}}\mathbb{D}^{1}_{r}\xhookrightarrow{\iota_{1}}\mathbb{D}^{2}_{r}\xhookrightarrow{\iota_{2}}\mathbb{D}^{3}_{r}\xhookrightarrow{\iota_{3}}\mathbb{D}^{4}_{r}\xhookrightarrow{\iota_{4}}\cdots

of formal disks. By acting on points via pullback, we obtain natural transformations of functors resed:JrdJre\textrm{res}^{d}_{e}:J^{d}_{r}\to J^{e}_{r} which produce maps JrdXJreXJ^{d}_{r}X\to J^{e}_{r}X that take jets j:𝔻rdXj:\mathbb{D}^{d}_{r}\to X to their restrictions jιd1ιej\circ\iota_{d-1}\circ\cdots\circ\iota_{e}. We are now ready to present the key proposition for our method:

Definition 3.5.

For a scheme XX (resp. analytic space XX) denote by Jr,nddXJrdXJ^{d}_{r,nd}X\subset J^{d}_{r}X the subscheme (resp. the analytic subspace) parametrizing those maps j:𝔻rdXj:\mathbb{D}^{d}_{r}\to X which are injective on the level of tangent spaces. We call such jj non-degenerate jets.

Proposition 3.6.

Let 𝒮\mathcal{S} be a set of types containing all the Hodge-theoretic types, and let ee and kk be non-negative integers. Then the following are equivalent:

  • (i)

    there exists a geometrically irreducible subvariety ZSZ\subset S_{\mathbb{C}} with dimZ>d\dim Z>d, dimφ(Z)e\dim\varphi(Z)\geq e, and such that dim𝒞(Z)dimφ(Z)k\dim\mathcal{C}(Z)-\dim\varphi(Z)\leq k;

  • (ii)

    there exists 𝒞𝒮\mathcal{C}\in\mathcal{S} with dim𝒞ek\dim\mathcal{C}-e\leq k, and such that the intersection

    𝒦rd(𝒞,e,k):=𝒯rd+1(𝒞)𝒯rd+1((resed+1)1(Jr,ndeLˇ))Jr,ndd+1S\mathcal{K}^{d}_{r}(\mathcal{C},e,k):=\mathcal{T}^{d+1}_{r}(\mathcal{C})\cap\mathcal{T}^{d+1}_{r}((\textrm{res}^{d+1}_{e})^{-1}(J^{e}_{r,nd}\widecheck{{L}}))\cap J^{d+1}_{r,nd}S

    is non-empty for each r0r\geq 0.

Remark.

In the situation that the variation 𝕍\mathbb{V} admits a local Torelli theorem, one can drop the distinction between dimZ\dim Z and dimφ(Z)\dim\varphi(Z) and consider instead the intersections 𝒯rd+1(𝒞)Jr,ndd+1S\mathcal{T}^{d+1}_{r}(\mathcal{C})\cap J^{d+1}_{r,nd}S in part (ii), ignoring the middle term.

The rest of this section we devote to proving 3.6, identifying SS with SS_{\mathbb{C}} for ease of notation. To begin with, let us check that (i) implies (ii) by applying the definitions. If g:SSg:S^{\prime}\to S is an étale cover and we consider the variation 𝕍=g𝕍\mathbb{V}^{\prime}=g^{*}\mathbb{V}, then the maps ηrd\eta^{d}_{r} and ηrd\eta^{\prime d}_{r} obtained from 3.2 are related by ηrd=ηrd(Jrdg)\eta^{\prime d}_{r}=\eta^{d}_{r}\circ(J^{d}_{r}g). Choosing a finite index subgroup ΓΓ\Gamma^{\prime}\subset\Gamma and passing to such a cover, we can reduce to the case where we have a period map φ:SΓ\D\varphi:S\to\Gamma\backslash D with DΓ\DD\to\Gamma\backslash D a local isomorphism. Applying [BBT18] the map φ:SΓ\D\varphi:S\to\Gamma\backslash D factors as φ=ιp\varphi=\iota\circ p, where p:STp:S\to T is a dominant map of algebraic varieties and ι\iota is an analytic closed embedding. Then via pp, the variety ZZ is dominant over a closed subvariety YTY\subset T of dimension dimφ(Z)e\dim\varphi(Z)\geq e. Shrinking SS (and hence ZZ) we may assume that ZZ is smooth, and that ZZ is surjective onto a dense open subset YYY^{\circ}\subset Y. Shrinking SS even further we may assume that ZYZ\to Y^{\circ} is smooth. The smoothness of ZYZ\to Y^{\circ} implies in particular that the induced jet space maps JrdZJrdYJ^{d}_{r}Z\to J^{d}_{r}Y^{\circ} for all choices of dd and rr are surjective.

We may choose neighbourhoods BS()B\subset S(\mathbb{C}) and UDU\subset D such that π|U:Uπ(U){\left.\kern-1.2pt\pi\vphantom{\big{|}}\right|_{U}}:U\to\pi(U) is an isomorphism, both BZB\cap Z and π(U)Y\pi(U)\cap Y^{\circ} are non-empty, and we have a local lift ψ:BU\psi:B\to U of φ\varphi. Choose a jet σJr,nde(Yπ(U))\sigma\in J^{e}_{r,nd}(Y^{\circ}\cap\pi(U)) and lift it along pp to a jet σ~Jr,nde(ZB)\widetilde{\sigma}\in J^{e}_{r,nd}(Z\cap B) landing at the point sS()s\in S(\mathbb{C}). Using the fact that the germ (Z,s)(Z,s) is smooth of dimension dimZ>d\dim Z>d the jet σ~\widetilde{\sigma} can be extended to a jet jJr,ndd+1(ZB)j\in J^{d+1}_{r,nd}(Z\cap B) such that resed+1(j)=σ~\textrm{res}^{d+1}_{e}(j)=\widetilde{\sigma}, and hence resed+1(φj)=σ\textrm{res}^{d+1}_{e}(\varphi\circ j)=\sigma. From the fact that φ|B=πψ{\left.\kern-1.2pt\varphi\vphantom{\big{|}}\right|_{B}}=\pi\circ\psi and the defining property of the map ηrd\eta^{d}_{r} it follows that jj lies inside 𝒯rd+1((resed+1)1(Jr,ndeLˇ))Jr,ndd+1S\mathcal{T}^{d+1}_{r}((\textrm{res}^{d+1}_{e})^{-1}(J^{e}_{r,nd}\widecheck{{L}}))\cap J^{d+1}_{r,nd}S. We can then take 𝒞=𝒞(Z)\mathcal{C}=\mathcal{C}(Z), and the fact that jj factors through ZZ implies that j𝒯rd+1(𝒞)j\in\mathcal{T}^{d+1}_{r}(\mathcal{C}) as well.

To prove the reverse implication, we review some preliminary facts relating to jets.

Definition 3.7.

We say a sequence {jr}r0\{j_{r}\}_{r\geq 0} with jrJrdXj_{r}\in J^{d}_{r}X is compatible if the projections JrdXJr1dXJ^{d}_{r}X\to J^{d}_{r-1}X map jrj_{r} to jr1j_{r-1}.

Lemma 3.8.

Suppose that 𝒯rJrdX\mathcal{T}_{r}\subset J^{d}_{r}X is a collection of non-empty constructible algebraic sets such that the projections JrdXJr1dXJ^{d}_{r}X\to J^{d}_{r-1}X map 𝒯r\mathcal{T}_{r} into 𝒯r1\mathcal{T}_{r-1}. Then there exists a compatible sequence {jr}r0\{j_{r}\}_{r\geq 0} with jr𝒯rj_{r}\in\mathcal{T}_{r} for all r0r\geq 0.

Proof.

See [Urb21a, Lem. 5.3]. ∎

Definition 3.9.

Given a variety ZZ (algebraic or analytic) and zZz\in Z a point, we denote by (JrdZ)z(J^{d}_{r}Z)_{z} the fibre above zz of the natural projection map JrdZZJ^{d}_{r}Z\to Z.

Lemma 3.10.

If g:(Z,z)(Y,y)g:(Z,z)\to(Y,y) is a map of analytic germs with dim(Z,z)=d\dim(Z,z)=d and (Z,z)(Z,z) smooth, we have an infinite compatible family jr(Jr,nddZ)zj_{r}\in(J^{d}_{r,nd}Z)_{z}, and gjr(JrdX)yg\circ j_{r}\in(J^{d}_{r}X)_{y} for some germ (X,y)(Y,y)(X,y)\subset(Y,y) and all r0r\geq 0, then gg factors through the inclusion (X,x)(Y,y)(X,x)\subset(Y,y).

Proof.

See [Urb21b, Lem. 4.5]. ∎

Lemma 3.11.

Suppose that XX is an algebraic variety (resp. analytic space) and xXx\in X is a point for which the fibre (Jr,nddX)x(J^{d}_{r,nd}X)_{x} above xx is non-empty for all r0r\geq 0. Then the germ (X,x)(X,x) has dimension at least dd.

Proof.

See [Urb21a, Prop. 2.7]. ∎

Proof of 3.6:.

By what we have said, we are reduced to showing that (ii) implies (i). The statement is unchanged by replacing SS with a finite étale covering g:SSg:S^{\prime}\to S and the variation 𝕍\mathbb{V} with g𝕍g^{*}\mathbb{V}; as before this does not affect the hypothesis (ii) since the maps ηrd\eta^{d}_{r} and ηrd\eta^{\prime d}_{r} associated to SS and SS^{\prime} are related by ηrd=ηrd(Jrdg)\eta^{\prime d}_{r}=\eta^{d}_{r}\circ(J^{d}_{r}g). Choosing a finite index subgroup ΓΓ\Gamma^{\prime}\subset\Gamma and choosing gg so the monodromy of g𝕍g^{*}\mathbb{V} lies in Γ\Gamma^{\prime} we may reduce to the case where DΓ\DD\to\Gamma\backslash D is a local isomorphism. Moreover, taking a futher finite étale cover we may apply [CPMS03, Cor. 13.7.6] to reduce to the case where φ\varphi is proper; this requires possibly extending SS^{\prime} to a variety S′′S^{\prime\prime} by adding a closed subvariety at infinity, but as long as we are careful to work only with jets that factor through SS^{\prime} our proof will produce a variety ZZ intersecting SS^{\prime}; in particular, we now assume that φ:SΓ\D\varphi:S\to\Gamma\backslash D is proper but redefine the sets 𝒦rd\mathcal{K}^{d}_{r} to equal

𝒯rd+1(𝒞)𝒯rd+1((res)ed+1)1(Jr,nddLˇ))Jd+1r,ndS,\mathcal{T}^{d+1}_{r}(\mathcal{C})\cap\mathcal{T}^{d+1}_{r}((\textrm{res})^{d+1}_{e})^{-1}(J^{d}_{r,nd}\widecheck{{L}}))\cap J^{d+1}_{r,nd}S^{\circ},

for some open subvariety SSS^{\circ}\subset S.

Applying the main result of [BBT18], the map φ\varphi once again factors as φ=ιp\varphi=\iota\circ p with p:STp:S\to T a dominant (now proper) map of algebraic varieties. We can then consider the Stein factorization S𝑞U𝑟TS\xrightarrow{q}U\xrightarrow{r}T of pp; note that qq is proper with connected fibres, UU is normal, and rr is finite. One can define the type of a subvariety YUY\subset U exactly as in 1.8 with respect to the period map UΓ\DU\to\Gamma\backslash D. From 3.8 above, the assumption (ii) entitles us to a compatible sequence {jr}r0\{j_{r}\}_{r\geq 0} of jets such that jr𝒦rd(𝒞,e,k)j_{r}\in\mathcal{K}^{d}_{r}(\mathcal{C},e,k) for all r0r\geq 0. Let us write 𝒞=𝒞(W)\mathcal{C}=\mathcal{C}(W) for some subvariety WLˇW\subset\widecheck{{L}}.

By construction, the jets σr=resed+1jr\sigma_{r}=\textrm{res}^{d+1}_{e}j_{r} are non-degenerate, and remain so after composing with any local period map ψ:BD\psi:B\to D for which σr\sigma_{r} factors through BB. This in particular implies (since DΓ\DD\to\Gamma\backslash D is a local isomorphism) that the jets φσr\varphi\circ\sigma_{r} are non-degenerate, and hence so are the jets qσrq\circ\sigma_{r}. Let YUY\subset U be the smallest algebraic subvariety such that qjrJrd+1Yq\circ j_{r}\in J^{d+1}_{r}Y for all rr. We observe that there exists a component ZZ of q1(Y)q^{-1}(Y) of dimension at least d+1d+1 that contains the image of the jets {jr}r0\{j_{r}\}_{r\geq 0}: one can see this by picking a neighbourhood of j0j_{0} of the form ×d+1\mathbb{C}^{\ell}\times\mathbb{C}^{d+1} such that jrj_{r} is constant on the first factor, and applying 3.10 above to see that the restriction of qq to {0}×d+1\{0\}\times\mathbb{C}^{d+1} factors through YY. Moreover, we must have q(Z)=Yq(Z)=Y by minimality, and by applying 3.11 to the non-degenerate sequence {qσr}r0\{q\circ\sigma_{r}\}_{r\geq 0} that dimYe\dim Y\geq e. Since rr is finite, this means dimφ(Z)e\dim\varphi(Z)\geq e. From the fact that local period maps on SS factor through local period maps on UU we learn that 𝒞(Z)=𝒞(Y)\mathcal{C}(Z)=\mathcal{C}(Y), so the result will follow if we can show that dim𝒞(Y)dimYk\dim\mathcal{C}(Y)-\dim Y\leq k. For ease of notation let us now write τr=qjr\tau_{r}=q\circ j_{r}.

Fix a local lift ψ:BD\psi:B\to D of the period map UΓ\DU\to\Gamma\backslash D with BU()B\subset U(\mathbb{C}) an analytic ball such that the jets τr\tau_{r} factor through BB. Consider the set 𝒢rGLm()\mathcal{G}_{r}\subset\textrm{GL}_{m}(\mathbb{C}) consisting of those gGLm()g\in\textrm{GL}_{m}(\mathbb{C}) for which ψτrg(Jrd+1W)\psi\circ\tau_{r}\in g\cdot(J^{d+1}_{r}W). Then for each rr the set 𝒢r\mathcal{G}_{r} is algebraically-constructible, and using the fact that jr𝒯rd+1(W)j_{r}\in\mathcal{T}^{d+1}_{r}(W) the set 𝒢r\mathcal{G}_{r} is necessarily non-empty. Let gg_{\infty} be an element of this intersection. Extend ψ\psi to a lift φ~Y:Y~D\widetilde{\varphi}_{Y}:\widetilde{Y}\to D of YΓ\DY\to\Gamma\backslash D to the universal covering. Then φ~Y(Y~)D\widetilde{\varphi}_{Y}(\widetilde{Y})\subset D is a closed analytic set containing the jets ψτr\psi\circ\tau_{r}, and hence the non-degenerate jets resed+1(ψτr)\textrm{res}^{d+1}_{e}(\psi\circ\tau_{r}). Letting Aφ~Y(Y~)(gW)A\subset\widetilde{\varphi}_{Y}(\widetilde{Y})\cap(g_{\infty}\cdot W) be the minimal analytic germ through which ψτr\psi\circ\tau_{r} (and hence resed+1(ψτr)\textrm{res}^{d+1}_{e}(\psi\circ\tau_{r})) factors, it follows from 3.11 that AA has dimension at least ee.

Consider the Zariski closure VYV\subset Y of ψ1(A)\psi^{-1}(A). We claim that V=YV=Y. Because YY was chosen minimal containing the compatible family of jets {τr}r0\{\tau_{r}\}_{r\geq 0}, it suffices to show that each τr\tau_{r} factors through VV. Consider the component of q1(B)q^{-1}(B) containing j0j_{0}; by choosing coordinates we may assume q1(B)×d+1q^{-1}(B)\subset\mathbb{C}^{\ell}\times\mathbb{C}^{d+1} is an open neighbourhood and identify j0j_{0} with the origin. After a further change of coordinates we may assume jrj_{r} is constant on the first factor, and let F=q1(B)({0}×d+1)F=q^{-1}(B)\cap(\{0\}\times\mathbb{C}^{d+1}). Applying 3.10 we find that ψ(q(F))A\psi(q(F))\subset A, and hence q(F)Vq(F)\subset V. Using proper base change the map q1(B)Bq^{-1}(B)\to B is proper, so q(F)q(F) is an analytic subvariety of BB, and by construction the jets {τr}r0\{\tau_{r}\}_{r\geq 0} factor through it, hence through VV.

We are now ready to apply the Bakker-Tsimerman transcendence theorem; the jets are no longer needed. It follows from the structure theorem for period mappings [GGK12, III.A], the closed analytic set φ~Y(Y~)\widetilde{\varphi}_{Y}(\widetilde{Y}) lies inside an orbit Dˇ=𝐇Yψ(τ0)\widecheck{{D}}^{\prime}=\mathbf{H}_{Y}\cdot\psi(\tau_{0}) of the algebraic monodromy of YY. Consider the graph EE of φ~Y\widetilde{\varphi}_{Y} in Y×DˇY\times\widecheck{{D}}^{\prime}. Then as AA has dimension ee and ψ1(A)\psi^{-1}(A) is Zariski dense, there exists a component CC of E(Y×(DˇgW))E\cap(Y\times(\widecheck{{D}}^{\prime}\cap g_{\infty}\cdot W)) of dimension at least ee and projecting to a Zariski dense subset of YY. Applying the main theorem of [BT17] we learn that

codimY×Dˇ(Y×(DˇgW))+codimY×DˇE\displaystyle\textrm{codim}_{Y\times\widecheck{{D}}^{\prime}}(Y\times(\widecheck{{D}}^{\prime}\cap g_{\infty}\cdot W))+\textrm{codim}_{Y\times\widecheck{{D}}^{\prime}}E codimY×DˇC\displaystyle\leq\textrm{codim}_{Y\times\widecheck{{D}}^{\prime}}C
(dimDˇdimW)+dimDˇ\displaystyle(\dim\widecheck{{D}}^{\prime}-\dim W)+\dim\widecheck{{D}}^{\prime} dimY+dimDˇdimC\displaystyle\leq\dim Y+\dim\widecheck{{D}}^{\prime}-\dim C
dimDˇdimY\displaystyle\dim\widecheck{{D}}^{\prime}-\dim Y dimWdimC\displaystyle\leq\dim W-\dim C
dim𝒞(Y)dimY\displaystyle\dim\mathcal{C}(Y)-\dim Y dim𝒞e\displaystyle\leq\dim\mathcal{C}-e
k\displaystyle\leq k

4 Main Results

4.1 Computing Bounds on Δd\Delta_{d}

4.1.1 Computing Lower Bounds

Let us explain the significance of 3.6 in proving 1.2, i.e., giving an effective method to compute bounds for

Δd=mindimZ>d[dim𝒞(Z)dimφ(Z)],\Delta_{d}=\min_{\dim Z>d}[\dim\mathcal{C}(Z)-\dim\varphi(Z)],

where we have used 1.3 and 1.8 to give this equivalent expression for Δd\Delta_{d}. Since Δd\Delta_{d} is a integer bounded by dimD\dim D, giving an effective method to compute it amounts to developing a procedure to decide, for any integer kk, whether we have Δdk\Delta_{d}\leq k. This in turn amounts to deciding, for any integer 0edimφ(S)0\leq e\leq\dim\varphi(S), whether (ii) holds in 3.6.

Let us take 𝒮\mathcal{S} to be the set up types computed by Step One, and let us suppose that in fact Δd>k\Delta_{d}>k. Then by the equivalence in 3.6, we should find that for any 𝒞𝒮\mathcal{C}\in\mathcal{S} with dim𝒞ek\dim\mathcal{C}-e\leq k, there must be some r=r(𝒞,e,k)r=r(\mathcal{C},e,k) such that 𝒦rd(𝒞,e,k)\mathcal{K}^{d}_{r}(\mathcal{C},e,k) is empty. Moreover, verifying that such an rr exists for each such 𝒞\mathcal{C} and ee proves, again by the same equivalence, that Δd>k\Delta_{d}>k. Consequently, we obtain the following result, which is the first half of 1.2:

Proposition 4.1.

By computing the sets 𝒦rd(𝒞,e,k)\mathcal{K}^{d}_{r}(\mathcal{C},e,k) described in 3.6 in parallel, we may compute a non-decreasing sequence of lower bounds

κ(1)κ(2)κ(r)<Δd\kappa(1)\leq\kappa(2)\leq\cdots\leq\kappa(r)\leq\cdots<\Delta_{d}

such that for some r=r0r=r_{0} we have κ(r0)=Δd1\kappa(r_{0})=\Delta_{d}-1.

Proof.

At the rr’th stage we compute all the sets 𝒦rd(𝒞,e,k)\mathcal{K}^{d}_{r}(\mathcal{C},e,k) for all applicable choices of 𝒞\mathcal{C}, ee and kk, and set κ(r)\kappa(r) to be the smallest kk for which all the sets 𝒦rd(𝒞,e,k)\mathcal{K}^{d}_{r}(\mathcal{C},e,k) are empty. From the discussion preceding the Proposition, the result follows. ∎

4.1.2 Computing Upper Bounds

4.1 does not actually give an algorithm for computing Δd\Delta_{d}, since no way is given to decide when r=r0r=r_{0}. For applications to the Lawrence-Venktesh method this doesn’t matter: one wants to be able to compute an optimal lower bound for Δd\Delta_{d}, but one does not actually have to prove that this lower bound actually equals Δd\Delta_{d} in order to apply the diophantine finiteness machinery. Nevertheless, let us explain how one can do this in the case where φ\varphi is quasi-finite; under this assumption, we may drop the distinction between dimZ\dim Z and dimφ(Z)\dim\varphi(Z), and we are instead interested in computing

mindimZ>d[dim𝒞(Z)dimZ].\min_{\dim Z>d}\,\left[\dim\mathcal{C}(Z)-\dim Z\right].

What is needed is the following:

Proposition 4.2.

Suppose that SS is quasi-projective and φ\varphi is quasi-finite. Then there exists a procedure that outputs an infinite sequence of upper bounds

τ(1)τ(2)τ(i)Δd\tau(1)\geq\tau(2)\geq\cdots\geq\tau(i)\geq\cdots\geq\Delta_{d}

such that for some i=i0i=i_{0} we have τ(i)=Δd\tau(i)=\Delta_{d}.

Given both 4.1 and 4.2 we obtain an algorithm for computing Δd\Delta_{d} by running both procedures in parallel and terminating when κ(r)+1=τ(i)\kappa(r)+1=\tau(i).

4.2 Finding Varieties that Exhibit Δd\Delta_{d}

In this section we prove 4.2, assuming throughout that SS is quasi-projective and φ\varphi is quasi-finite. Let us fix a projective compactification SS¯S\subset\overline{S} of SS and consider the Hilbert scheme Hilb(S¯)\textrm{Hilb}(\overline{S}). There exist algorithms, for instance by working with the Plüker coordinates of the appropriate Grassmannian, for computing any finite subset of components of Hilb(S¯)\textrm{Hilb}(\overline{S}). By [Urb21b, Lem. 5.10] we obtain the same fact for the open locus Var(S)Hilb(S¯)\textrm{Var}(S)\subset\textrm{Hilb}(\overline{S}) consisting of just those points [Z¯][\overline{Z}] for which Z=SZ¯Z=S\cap\overline{Z} is a non-empty geometrically irreducible algebraic subvariety of SS. What we will show is that there exists a procedure which outputs an infinite sequence {𝒲i}i=1\{\mathcal{W}_{i}\}_{i=1}^{\infty} of constructible algebraic loci 𝒲iVar(S)\mathcal{W}_{i}\subset\textrm{Var}(S), with the following two properties:

  • (i)

    for each ii, the type 𝒞(Z)\mathcal{C}(Z) and dimension dimZ\dim Z are constant over all [Z]𝒲i[Z]\in\mathcal{W}_{i};

  • (ii)

    there exists some i=i0i=i_{0} such that

    Δd=dim𝒞(Z)Z,\Delta_{d}=\dim\mathcal{C}(Z)-Z,

    for some (hence any) point [Z]𝒲i[Z]\in\mathcal{W}_{i}.

Given such an algorithm the problem of computing the bound τ(i)\tau(i) that appears in 4.2 reduces to choosing a point [Z]𝒲i[Z]\in\mathcal{W}_{i}, computing dim𝒞(Z)dimZ\dim\mathcal{C}(Z)-\dim Z, and setting

τ(i):=min{τ(i1),dim𝒞(Z)dimZ}.\tau(i):=\textrm{min}\{\tau(i-1),\dim\mathcal{C}(Z)-\dim Z\}.

(We note that the problem of computing dim𝒞(Z)\dim\mathcal{C}(Z) from ZZ and the restriction (,F,)|Z{\left.\kern-1.2pt(\mathcal{H},F^{\bullet},\nabla)\vphantom{\big{|}}\right|_{Z}} of the algebraic data on SS is solved for us by [Urb21b, Lem. 5.8] by taking the family gg in the statement of [Urb21b, Lem. 5.8] to be a trivial family; we will say little about this problem here.)

In fact, an algorithm for computing the sets 𝒲i\mathcal{W}_{i} has already been given in a previous paper by the author. We begin by recalling the necessary background. We regard SS as a complex algebraic variety in what follows. Given two (geometrically) irreducible subvarieties Z1,Z2SZ_{1},Z_{2}\subset S with Z1Z2Z_{1}\subset Z_{2}, the algebraic monodromy group 𝐇Z1\mathbf{H}_{Z_{1}} may be naturally regarded as a subgroup of 𝐇Z2\mathbf{H}_{Z_{2}} (after choosing a base point sZ1()s\in Z_{1}(\mathbb{C})). Using this we define:

Definition 4.3.

An irreducible complex subvariety ZSZ\subset S is said to be weakly special if it is maximal among such subvarieties for its algebraic monodromy group.

The key fact is then the following:

Lemma 4.4.

For each integer d>0d>0, there exists a weakly special subvariety ZSZ\subset S such that

Δd=dim𝒞(Z)dimZ.\Delta_{d}=\dim\mathcal{C}(Z)-\dim Z.
Proof.

By [Urb21b, Prop. 4.18], the condition that ZZ be weakly special is equivalent to ZZ being a maximal irreducible complex subvariety of SS of type 𝒞(Z)\mathcal{C}(Z). Thus if we have any YY which is not weakly special, there exists a weakly special ZZ properly containing YY with 𝒞(Y)=𝒞(Z)\mathcal{C}(Y)=\mathcal{C}(Z), hence

dim𝒞(Z)dimZ<dim𝒞(Y)dimY.\dim\mathcal{C}(Z)-\dim Z<\dim\mathcal{C}(Y)-\dim Y.

It follows that the value of Δd\Delta_{d} can only be achieved by a weakly special variety. ∎

Proof of 4.2.

By the degree of a subvariety ZSZ\subset S we will mean the degree of its closure Z¯\overline{Z} inside S¯\overline{S}. For any integer bb, denote by Var(S)bVar(S)\textrm{Var}(S)_{b}\subset\textrm{Var}(S) the finite-type subscheme parametrizing varieties of degree at most bb. Denote by 𝒲Var(S)\mathcal{W}\subset\textrm{Var}(S) the locus of weakly special subvarieties. Then given an integer bb, the algorithm that appears in [Urb21b, Thm. 5.15] computes the intersection 𝒲Var(S)b\mathcal{W}\cap\textrm{Var}(S)_{b} as a constructible algebraic locus.

Let us describe the algorithm appearing in [Urb21b, Thm. 5.15] more precisely. Consider the types 𝒞1,,𝒞\mathcal{C}_{1},\ldots,\mathcal{C}_{\ell} computed by Step One, and define for each such type 𝒞j\mathcal{C}_{j} the locus

𝒲(𝒞j):={[Z]Var(S):𝒞(Z)𝒞j}.\mathcal{W}(\mathcal{C}_{j}):=\{[Z]\in\textrm{Var}(S):\mathcal{C}(Z)\leq\mathcal{C}_{j}\}.

It is shown in [Urb21b, Prop. 4.31] that for each jj the locus 𝒲(𝒞j)\mathcal{W}(\mathcal{C}_{j}) is closed algebraic. We can then consider the sublocus 𝒲(𝒞j)opt𝒲(𝒞j)\mathcal{W}(\mathcal{C}_{j})_{\textrm{opt}}\subset\mathcal{W}(\mathcal{C}_{j}) consisting of just those components C𝒲(𝒞j)C\subset\mathcal{W}(\mathcal{C}_{j}) for which a generic point [Z]C[Z]\in C satisfies 𝒞(Z)=𝒞j\mathcal{C}(Z)=\mathcal{C}_{j}.

In [Urb21b, Prop. 5.14], an algorithm is given for computing 𝒲(𝒞j)optVar(S)b\mathcal{W}(\mathcal{C}_{j})_{\textrm{opt}}\cap\textrm{Var}(S)_{b} for each jj. Using this, one can compute all the finitely many closed algebraic loci C1,,CibC_{1},\ldots,C_{i_{b}} which arise as a component of 𝒲(𝒞j)optVar(S)b\mathcal{W}(\mathcal{C}_{j})_{\textrm{opt}}\cap\textrm{Var}(S)_{b} for some jj. The problem of computing 𝒲Var(S)b\mathcal{W}\cap\textrm{Var}(S)_{b} is reduced to computing constructible algebraic conditions on each component Ci𝒲(𝒞j)optC_{i}\subset\mathcal{W}(\mathcal{C}_{j})_{\textrm{opt}} which define the locus 𝒲iCi\mathcal{W}_{i}\subset C_{i} of points [Z]Ci[Z]\in C_{i} that are weakly special of type 𝒞j\mathcal{C}_{j}. This is taken care of by the proof of [Urb21b, Thm. 5.15]. By construction, the points in 𝒲i\mathcal{W}_{i} all have the same type and the same dimension, so we complete the proof by computing these loci for increasing values of bb. ∎

5 Application to Lawrence-Venkatesh

We now show how the bound of 1.2 can be used to establish diophantine finiteness results. Similar arguments appear in [LV20] and [LS20], but as they are not precisely adapted to our setup, we give our own version. We recall the situation: we have a smooth projective family f:XSf:X\to S over the smooth base SS, with everything defined over 𝒪K,N\mathcal{O}_{K,N}.444Note in particular we are assuming now that SS is smooth over 𝒪K,N\mathcal{O}_{K,N}, which we can achieve by increasing NN if necessary. The relative algebraic de Rham cohomology =RifΩX/S\mathcal{H}=R^{i}f_{*}\Omega^{\bullet}_{X/S} gives a model for the Hodge bundle 𝕍𝒪San\mathbb{V}\otimes\mathcal{O}_{S^{\textrm{an}}}, where 𝕍=Rif\mathbb{V}=R^{i}f_{*}\mathbb{Z}. By a result of Katz and Oda [KO68], the flat connection associated to the local system 𝕍\mathbb{V}_{\mathbb{C}} by the Riemann-Hilbert correspondence admits a model :ΩS1\nabla:\mathcal{H}\to\Omega^{1}_{S}\otimes\mathcal{H} after possibly increasing NN. Likewise, we may also assume the Hodge filtration FF^{\bullet} gives a filtration of \mathcal{H} by vector subbundles.

Fix a prime pp not dividing NN, and a place vv of KK above pp. Then for each integral point sS(𝒪K,N)s\in S(\mathcal{O}_{K,N}), we have a Galois representation ρs:Gal(K¯/K)Aut(Héti(XK¯,s,p))\rho_{s}:\textrm{Gal}(\overline{K}/K)\to\textrm{Aut}(H^{i}_{\textrm{\'{e}t}}(X_{\overline{K},s},\mathbb{Q}_{p})), and an argument of Faltings [LV20, Lem 2.3] shows that the semisimplifications of the representations ρs\rho_{s} belong to a finite set of isomorphism classes. From crystalline cohomology, each sS(𝒪K,N)s\in S(\mathcal{O}_{K,N}), viewed as a point of S(𝒪K,v)S(\mathcal{O}_{K,v}) where 𝒪K,v\mathcal{O}_{K,v} is the vv-adic ring of integers, gives rise to a triple (HdRi(Xs),ϕs,Fs)(H^{i}_{\textrm{dR}}(X_{s}),\phi_{s},F^{\bullet}_{s}) where ϕs\phi_{s} is the crystalline Frobenius. Moreover, using the functor DcrisD_{\textrm{cris}} of pp-adic Hodge theory [Fon94, Exposé III], the triple (HdRi(Xs),ϕs,Fs)(H^{i}_{\textrm{dR}}(X_{s}),\phi_{s},F^{\bullet}_{s}) is determined up to isomorphism by the restriction ρs,v\rho_{s,v} along the map Gal(Kv¯/Kv)Gal(K¯/K)\textrm{Gal}(\overline{K_{v}}/K_{v})\to\textrm{Gal}(\overline{K}/K) determined by a fixed embedding K¯Kv¯\overline{K}\hookrightarrow\overline{K_{v}}. We denote by (s)\mathcal{I}(s) all those triples (V,ϕ,F)(V,\phi,F^{\bullet}) which are of the form Dcris(ρ|p)D_{\textrm{cris}}({\left.\kern-1.2pt\rho\vphantom{\big{|}}\right|_{\mathbb{Q}_{p}}}), where ρ\rho is a global Galois representation whose semisimplification is isomorphic to the semisimplification of ρs\rho_{s}.

Recall that we have fixed the integral lattice V=mV=\mathbb{Z}^{m}, where mm is the dimension of the cohomology of the fibres of ff, and a \mathbb{Q}-algebraic flag variety Lˇ\widecheck{{L}} of Hodge flags on VV. In what follows we write VpV_{p} for VpV\otimes\mathbb{Q}_{p}, and Lˇp\widecheck{{L}}_{p} for Lˇp\widecheck{{L}}_{\mathbb{Q}_{p}}. Then the key idea of the Lawrence-Venkatesh method is the following:

Proposition 5.1.

Suppose that for each sS(𝒪K,N)s\in S(\mathcal{O}_{K,N}), whenever we have an endomorphism ϕs:VpVp\phi_{s}:V_{p}\to V_{p} and a flag FsF^{\bullet}_{s} on VpV_{p} such that (Vp,ϕs,Fs)(V_{p},\phi_{s},F^{\bullet}_{s}) represents (s)\mathcal{I}(s), the Hodge flags FF^{\bullet} on VpV_{p} for which (Vp,ϕs,F)(s)(V_{p},\phi_{s},F^{\bullet})\in\mathcal{I}(s) lie in an algebraic subvariety OsLˇpO_{s}\subset\widecheck{{L}}_{p} satisfying ΔddimOs\Delta_{d}\geq\dim O_{s}. Then dimS(𝒪K,N)¯Zard\dim\overline{S(\mathcal{O}_{K,N})}^{\textrm{Zar}}\leq d.

To prove 5.1 we will need a rigid-analytic version of the Bakker-Tsimerman transcendence theorem, which we will see can be deduced formally from the complex analytic one. To set things up, let us revisit the term local period map, this time in the rigid analytic setting (c.f. 1.6). We will denote by p\mathbb{C}_{p} the completion of the algebraic closure Kv¯\overline{K_{v}}. In what follows we sometimes identify algebraic varieties with their rigid-analytifications when the context is clear.

Definition 5.2.

Let KpK_{p} be a local field containing KvK_{v}, let SKpanS_{K_{p}}^{\textrm{an}} be the rigid-analytification of the base-change SKpS_{K_{p}} of SS, and suppose that BpSKpanB_{p}\subset S_{K_{p}}^{\textrm{an}} is an affinoid subdomain. Then a (rigid-analytic) local period map ψ:BpLˇKpan\psi:B_{p}\to\widecheck{{L}}_{K_{p}}^{\textrm{an}} is a rigid-analytic map obtained as a composition ψ=qKpanAp\psi=q_{K_{p}}^{\textrm{an}}\circ A_{p}, where:

  • (i)

    The rigid analytifications FkKpanF^{k}\mathcal{H}_{K_{p}}^{\textrm{an}} are all trivial on BpB_{p}.

  • (ii)

    The map Ap:BpGLm,KpanA_{p}:B_{p}\to\textrm{GL}_{m,K_{p}}^{\textrm{an}} is a varying filtration-compatible pp-adic period matrix over BpB_{p}. More precisely, there exists a basis v1,,vmv^{1},\ldots,v^{m} for Kpan(Bp)\mathcal{H}_{K_{p}}^{\textrm{an}}(B_{p}), compatible with the filtration in the sense that FkKpan(Bp)F^{k}\mathcal{H}_{K_{p}}^{\textrm{an}}(B_{p}) is spanned by v1,,vikv^{1},\ldots,v^{i_{k}} for some iki_{k}, and a flat (for Kpan\nabla_{K_{p}}^{\textrm{an}}) frame b1,,bmb^{1},\ldots,b^{m} such that ApA_{p} gives a varying change-of-basis matrix from v1,,vmv^{1},\ldots,v^{m} to b1,,bmb^{1},\ldots,b^{m}.

  • (iii)

    The map q:GLmLˇq:\textrm{GL}_{m}\to\widecheck{{L}} is the map that sends a matrix MM to the Hodge flag FMF^{\bullet}_{M} defined by the property that FMkF^{k}_{M} is spanned by the first iki_{k} columns.

To prove 5.1 we will need a version of the Bakker-Tsimerman transcendence result for rigid-analytic local period maps, which we prove by formally transferring the same result for complex analytic local period maps. To avoid certain minor pathologies that can occur in the complex analytic case we will restrict to local period maps ψ:BLˇ\psi:B\to\widecheck{{L}} which are definable in the structure an,exp\mathbb{R}_{\textrm{an},\textrm{exp}}; for background on definability and definable analytic spaces we refer to [VdDM96] and [BBT18]. We note that this is not a serious restriction: given any local period map ψ\psi and any point sBs\in B there exists a definable restriction of ψ\psi to a neighbourhood of ss, a fact which is for instance easily deduced from [Urb21b, Prop. 4.27].

Lemma 5.3.
  • (i)

    Suppose that ψ:BLˇan\psi:B\to\widecheck{{L}}_{\mathbb{C}}^{\textrm{an}} is a definable analytic local period map on SanS_{\mathbb{C}}^{\textrm{an}}. Let VLˇV\subset\widecheck{{L}}_{\mathbb{C}} be an algebraic subvariety satisfying ΔddimV\Delta_{d}\geq\dim V. Then ψ1(V)\psi^{-1}(V) lies in an algebraic subvariety of SS_{\mathbb{C}} of dimension at most dd.

  • (ii)

    Suppose that ψp:BpLˇpan\psi_{p}:B_{p}\to\widecheck{{L}}_{\mathbb{C}_{p}}^{\textrm{an}} is a rigid-analytic local period map on SpanS_{\mathbb{C}_{p}}^{\textrm{an}}. Let VpLˇp,pV_{p}\subset\widecheck{{L}}_{p,\mathbb{C}_{p}} be an algebraic subvariety satisfying ΔddimVp\Delta_{d}\geq\dim V_{p}. Then ψp1(Vp)\psi_{p}^{-1}(V_{p}) lies in an algebraic subvariety of SpS_{\mathbb{C}_{p}} of dimension at most dd.

Proof of 5.3(i):

This is an application of the Bakker-Tsimerman transcendence theorem. Let ZSZ\subset S_{\mathbb{C}} be the Zariski closure of ψ1(V)\psi^{-1}(V). We assume for contradiction that dimZ>d\dim Z>d, and let Z0ZZ_{0}\subset Z be a component of maximal dimension. Let φ:SanΓ\D\varphi:S_{\mathbb{C}}^{\textrm{an}}\to\Gamma\backslash D be the canonical period map with Γ=Aut(V,Q)()\Gamma=\textrm{Aut}(V,Q)(\mathbb{Z}). The statement is invariant under replacing ψ\psi with a GLm()\textrm{GL}_{m}(\mathbb{C})-translate gψg\cdot\psi and VV with gVg\cdot V, so we may assume that ψ\psi is a local lift of φ\varphi. Arguing as in [CPMS03, Cor. 13.7.6] we may assume that φ\varphi is proper, hence the image T=φ(S)T=\varphi(S) is algebraic by [BBT18], and we may consider the Stein factorization S𝑞U𝑟TS\xrightarrow{q}U\xrightarrow{r}T of the map STS\to T.

Let Y=q(Z0)Y=q(Z_{0}), and note that dimY=dimφ(Z0)\dim Y=\dim\varphi(Z_{0}). By assumption we have Δddim𝒞(Z0)dimφ(Z0)=dim𝒞(Y)dimY\Delta_{d}\leq\dim\mathcal{C}(Z_{0})-\dim\varphi(Z_{0})=\dim\mathcal{C}(Y)-\dim Y, where the type of YY is taken with respect to the period map UΓ\DU\to\Gamma\backslash D. Moreover, this continues to hold if we replace YY with a smooth resolution YY^{\prime}. The variation of Hodge structure on SS descends to UU, and hence shrinking BB if necessary we may factor ψ\psi through a definable local lift on UU. By pulling back along the resolution we obtain a definable local lift ψ:BD\psi^{\prime}:B^{\prime}\to D of the period map φ:YΓ\D\varphi^{\prime}:Y^{\prime}\to\Gamma\backslash D such that ψ1(V)\psi^{\prime-1}(V) is Zariski dense in YY^{\prime}. We are reduced to the following situation: we have smooth variety YY^{\prime} with a period map φ:YΓ\D\varphi^{\prime}:Y^{\prime}\to\Gamma\backslash D, a local lift ψ:BD\psi^{\prime}:B^{\prime}\to D such that ψ1(V)\psi^{\prime-1}(V) is Zariski dense, and such that dim𝒞(Y)dimYdimV\dim\mathcal{C}(Y^{\prime})-\dim Y^{\prime}\geq\dim V.

We now contradict the Bakker-Tsimerman theorem. In particular, we may extend the local lift ψ\psi^{\prime} to a lift φ~:Y~D\widetilde{\varphi^{\prime}}:\widetilde{Y^{\prime}}\to D of φ\varphi^{\prime} to the universal cover, and consider the graph WY×DˇW\subset Y^{\prime}\times\widecheck{{D}}^{\prime} of the map φ~\widetilde{\varphi^{\prime}}, where Dˇ\widecheck{{D}}^{\prime} is the orbit 𝐇Yψ(y)\mathbf{H}_{Y^{\prime}}\cdot\psi^{\prime}(y) for some yY()y\in Y^{\prime}(\mathbb{C}). We then have that

codimY×Dˇ(Y×(VDˇ))+codimY×DˇW\displaystyle\textrm{codim}_{Y^{\prime}\times\widecheck{{D}}^{\prime}}(Y^{\prime}\times(V\cap\widecheck{{D}}^{\prime}))+\textrm{codim}_{Y^{\prime}\times\widecheck{{D}}^{\prime}}W dimDˇdimV+dimDˇ\displaystyle\geq\dim\widecheck{{D}}^{\prime}-\dim V+\dim\widecheck{{D}}^{\prime}
=dim𝒞(Y)dimV+dim𝒞(Y)\displaystyle=\dim\mathcal{C}(Y^{\prime})-\dim V+\dim\mathcal{C}(Y^{\prime})
dimY+dim𝒞(Y).\displaystyle\geq\dim Y^{\prime}+\dim\mathcal{C}(Y^{\prime}).

Since ψ\psi^{\prime} is definable, ψ1(V)\psi^{\prime-1}(V) has finitely many components, and hence there exists an analytic component CBC\subset B^{\prime} of ψ1(V)\psi^{\prime-1}(V) such that CC is Zariski dense in YY^{\prime}. Let C~Y×Dˇ\widetilde{C}\subset Y^{\prime}\times\widecheck{{D}}^{\prime} be its graph under ψ\psi^{\prime}. If dimY=0\dim Y^{\prime}=0 there is nothing to show, so we may assume that dim𝒞(Y)>0\dim\mathcal{C}(Y^{\prime})>0. Hence we find that dimY+dim𝒞(Y)>codimY×DˇC~\dim Y^{\prime}+\dim\mathcal{C}(Y^{\prime})>\textrm{codim}_{Y^{\prime}\times\widecheck{{D}}^{\prime}}\widetilde{C}, and by the Bakker-Tsimerman theorem [BT17] the component CC lies in a proper subvariety of YY^{\prime}, giving a contradiction. ∎

To prove 5.3(ii) we first translate 5.3(i) into a claim about rings of formal power series. In particular let ψ:BLˇ\psi:B\to\widecheck{{L}} be a local period map with VLˇV\subset\widecheck{{L}} an algebraic subvariety, and choose a point sBs\in B such that ψ(s)=tV\psi(s)=t\in V. Then ψ\psi induces a map on formal power series rings ψ^:𝒪^Lˇ,t𝒪^S,s\widehat{\psi}^{\sharp}:\widehat{\mathcal{O}}_{\widecheck{{L}}_{\mathbb{C}},t}\to\widehat{\mathcal{O}}_{S_{\mathbb{C}},s}. The claim of 5.3(i) then says that if IV𝒪Lˇ,tI_{V}\subset\mathcal{O}_{\widecheck{{L}}_{\mathbb{C}},t} is the ideal defining VV with extension I^V\widehat{I}_{V} inside 𝒪^Lˇ,t\widehat{\mathcal{O}}_{\widecheck{{L}}_{\mathbb{C}},t}, then the ideal generated by ψ^(I^V)\widehat{\psi}^{\sharp}(\widehat{I}_{V}) contains an ideal I^Z\widehat{I}_{Z} which is the extension of an ideal IZ𝒪S,sI_{Z}\subset\mathcal{O}_{S_{\mathbb{C}},s} defining the germ of a subvariety of dimension at most dd.

Proof of 5.3(ii):

The claim is Zariski-local on SS, so we can in particular assume that the bundles FkF^{k}\mathcal{H} for varying kk are algebraically trivial over SS, that SS is affine, and by smoothness that ΩS1\Omega^{1}_{S} is free. By definition, the map ψ:BpLˇpan\psi:B_{p}\to\widecheck{{L}}_{\mathbb{C}_{p}}^{\textrm{an}} is associated to the following data: a filtration-compatible frame v1,,vmv^{1},\ldots,v^{m}, where v1,,vikv^{1},\ldots,v^{i_{k}} spans Fkpan(Bp)F^{k}\mathcal{H}_{\mathbb{C}_{p}}^{\textrm{an}}(B_{p}), and a flat frame b1,,bmb^{1},\ldots,b^{m} spanning pan(Bp)\mathcal{H}_{\mathbb{C}_{p}}^{\textrm{an}}(B_{p}), where flatness means panbi=0\nabla_{\mathbb{C}_{p}}^{\textrm{an}}b_{i}=0 for all 1im1\leq i\leq m. This data satisfies the property that ψ=qpanAp\psi=q_{\mathbb{C}_{p}}^{\textrm{an}}\circ A_{p}, where ApA_{p} is the change-of-basis matrix from the frame v1,,vmv^{1},\ldots,v^{m} to the frame b1,,bmb^{1},\ldots,b^{m}, and qq is the map GLm,panLˇpan\textrm{GL}_{m,\mathbb{C}_{p}}^{\textrm{an}}\to\widecheck{{L}}_{\mathbb{C}_{p}}^{\textrm{an}} sending a matrix to the Hodge flag it represents. We note that changing the frame v1,,vmv^{1},\ldots,v^{m} to another filtration-compatible frame v1,,vmv^{\prime 1},\ldots,v^{\prime m} does not change the local period map: such a change has the effect of replacing the map ApA_{p} with ApCA_{p}\cdot C, where CC is a varying matrix over BpB_{p} whose right-action on ApA_{p} preserves the span of the first iki_{k} columns for each kk, and hence q(ApC)=qApq\circ(A_{p}\cdot C)=q\circ A_{p}. We threfore lose no generality by assuming the filtration-compatible frame is the restriction to BpB_{p} of an algebraic filtration-compatible frame over SS.

The affinoid neighbourhood BpB_{p} is of the form SpT\textrm{Sp}\,T, where TT is an affinoid p\mathbb{C}_{p}-algebra. The inverse image ψ1(V)\psi^{-1}(V) is then a closed affinoid subdomain of BpB_{p}, i.e., it corresponds to an ideal ITI\subset T such that ψ1(V)\psi^{-1}(V) may be identified with SpT/I\textrm{Sp}\,T/I. If RR is the coordinate ring of SpS_{\mathbb{C}_{p}}, then the map BpSpanSpB_{p}\hookrightarrow S_{\mathbb{C}_{p}}^{\textrm{an}}\to S_{\mathbb{C}_{p}} induces a map ι:RT\iota:R\to T, and the claim to be shown is that there exists an ideal JRJ\subset R defining a subvariety of dimension at most dd such that ι(J)I\iota(J)\subset I. The ring TT is Noetherian, so the ideal II admits a primary decomposition. Taking radicals, we obtain finitely many prime ideals I1,,II_{1},\ldots,I_{\ell} containing II such that the problem reduces, for each 1j1\leq j\leq\ell, to finding JjRJ_{j}\subset R defining subvarieties of dimension at most dd such that ι(Jj)Ij\iota(J_{j})\subset I_{j} for each jj. The analytification map SpanSpS_{\mathbb{C}_{p}}^{\textrm{an}}\to S_{\mathbb{C}_{p}} is bijective onto the set of closed points of SpS_{\mathbb{C}_{p}} and induces isomorphisms on completed local rings [see whatever]. It follows that if we choose a maximal ideal 𝔪T\mathfrak{m}\subset T containing IjI_{j} we obtain a commuting diagram

R{R}T{T}R^ι1(𝔪){\widehat{R}_{\iota^{-1}(\mathfrak{m})}}𝒪^Bp,𝔪,{\widehat{\mathcal{O}}_{B_{p},\mathfrak{m}},}ι\scriptstyle{\iota}\scriptstyle{\sim}

where the bottom arrow is an isomorphism of completed local rings, and the vertical arrows are injections. In particular, if we denote by I^j\widehat{I}_{j} the extension of IjI_{j} in 𝒪^Bp,𝔪\widehat{\mathcal{O}}_{B_{p},\mathfrak{m}}, it suffices to show that ι^(Jj)I^j\widehat{\iota}(J_{j})\subset\widehat{I}_{j}, where ι^\widehat{\iota} is the composition of the left and bottom arrow; here we have used the fact that Ij=I^jTI_{j}=\widehat{I}_{j}\cap T.

Fix an isomorphism τ:p\tau:\mathbb{C}_{p}\xrightarrow{\sim}\mathbb{C}, which we choose to preserve the embeddings KK\subset\mathbb{C} and KpK\subset\mathbb{C}_{p}. Using the model for SS over KK, the isomorphism τ\tau allows us to identify RR with the coordinate ring of SS_{\mathbb{C}}, the ideal ι1(𝔪)\iota^{-1}(\mathfrak{m}) with a complex point sS()s\in S(\mathbb{C}), the ring R^ι1(𝔪)\widehat{R}_{\iota^{-1}(\mathfrak{m})} with the completed local ring 𝒪^S,s\widehat{\mathcal{O}}_{S_{\mathbb{C}},s}. Let tpt_{p} be the image of the point corresponding to 𝔪\mathfrak{m} under ψ\psi, and let tt be the composition tpτ1t_{p}\circ\tau^{-1}. Applying the isomorphism τ\tau at the level of formal power series, the rigid-analytic local period map ψ\psi induces a map

𝒪^Lˇ,t𝜏𝒪^Lˇp,tpψ^𝒪^Bp,𝔪R^ι1(𝔪)𝜏𝒪^S,s,\widehat{\mathcal{O}}_{\widecheck{{L}}_{\mathbb{C}},t}\xrightarrow{\tau}\widehat{\mathcal{O}}_{\widecheck{{L}}_{\mathbb{C}_{p}},t_{p}}\xrightarrow{\widehat{\psi}^{\sharp}}\widehat{\mathcal{O}}_{B_{p},\mathfrak{m}}\xrightarrow{\sim}\widehat{R}_{\iota^{-1}(\mathfrak{m})}\xrightarrow{\tau}\widehat{\mathcal{O}}_{S_{\mathbb{C}},s},

whose composition we denote by η^\widehat{\eta}. In what follows we identify the ideals I^j\widehat{I}_{j} with their images in 𝒪^S,s\widehat{\mathcal{O}}_{S_{\mathbb{C}},s}; by construction they are the extensions along η^\widehat{\eta} of an ideal in 𝒪^Lˇ,t\widehat{\mathcal{O}}_{\widecheck{{L}}_{\mathbb{C}},t} associated to the base-change of VV using τ\tau. By part (i) of this theorem and our reformulation of it in terms of completed local rings, it suffices to show that η^\widehat{\eta} is induced by a complex analytic local period map defined on a neighbourhood of ss.

Recall that we have a decomposition ψ=qpanAp\psi=q_{\mathbb{C}_{p}}^{\textrm{an}}\circ A_{p}, where qq is the rigid-analytification of a \mathbb{Q}-algebraic map, and ApA_{p} gives a varying change-of-basis matrix between a filtration-compatible frame v1,,vmv^{1},\ldots,v^{m} and a rigid-analytic flat frame. Recall also that we have chosen v1,,vmv^{1},\ldots,v^{m} so that it is the rigid-analytification of a KK-algebraic filtration-compatible frame w1,,wmw^{1},\ldots,w^{m} over SS. Using the decomposition ψ=qpanAp\psi=q_{\mathbb{C}_{p}}^{\textrm{an}}\circ A_{p} and the isomorphism τ\tau we may factor η^\widehat{\eta} as q^κ^\widehat{q}\circ\widehat{\kappa}, where κ^:𝒪^S,s𝒪^GLm,,P\widehat{\kappa}:\widehat{\mathcal{O}}_{S_{\mathbb{C}},s}\to\widehat{\mathcal{O}}_{\textrm{GL}_{m,\mathbb{C}},P} is the base-change under τ\tau of the map induced by ApA_{p}. From our definition of local period map in 1.6, it suffices to show that κ^\widehat{\kappa} is induced by a varying change-of-basis matrix A:BGLm,anA:B\to\textrm{GL}_{m,\mathbb{C}}^{\textrm{an}} from w1,,wmw^{1},\ldots,w^{m} to a complex-analytic flat frame.

The result will follow from the fact that AA and ApA_{p} satisfy a common set of KK-algebraic differential equations whose solutions are uniquely determined by the period matrix they assign to a point in BB. To see this, let us write wi=j=1mcijwj\nabla w^{i}=\sum_{j=1}^{m}c_{ij}\otimes w^{j} for KK-algebraic sections cijΩSK1c_{ij}\in\Omega^{1}_{S_{K}}. Suppose then that bk=i=1mfikwib^{k}=\sum_{i=1}^{m}f_{ik}w^{i} is a flat frame on some complex analytic or rigid-analytic neighbourhood. We then have that

bk\displaystyle\nabla b^{k} =(i=1mfikwi)\displaystyle=\nabla\left(\sum_{i=1}^{m}f_{ik}w^{i}\right)
=j=1mdfjkwj+i=1mfik(j=1mcijvj)\displaystyle=\sum_{j=1}^{m}df_{jk}\otimes w^{j}+\sum_{i=1}^{m}f_{ik}\left(\sum_{j=1}^{m}c_{ij}\otimes v^{j}\right)
=j=1m(dfjk+i=1mfikcij)vj,\displaystyle=\sum_{j=1}^{m}\left(df_{jk}+\sum_{i=1}^{m}f_{ik}c_{ij}\right)\otimes v^{j},

from which we see that bkb^{k} giving a flat frame is equivalent to fjkf_{jk} satisfying the system of differential equations dfjk=i=1mfikcijdf_{jk}=-\sum_{i=1}^{m}f_{ik}c_{ij} for all 1j,km1\leq j,k\leq m. If we choose a trivialization dz1,,dzndz_{1},\ldots,dz_{n} of ΩSK1\Omega^{1}_{S_{K}}, we may write the cijc_{ij} in terms of their coefficients cij,c_{ij,\ell} with respect to this trivialization, and the same system of differential equations becomes

fjk=ifikcij,;\partial_{\ell}f_{jk}=-\sum_{i}f_{ik}c_{ij,\ell}; (3)

here the operator \partial_{\ell} is defined using the dual basis to dz1,,dzndz_{1},\ldots,dz_{n}. By differentiating Equation 3 and substituting the lower-order differential equations into the higher-order ones, we obtain, for each sequence {i}i=1e\{\ell_{i}\}_{i=1}^{e} with 1in1\leq\ell_{i}\leq n and e1e\geq 1, a set of KK-algebraic polynomials ξ1,,e;jk\xi_{\ell_{1},\ldots,\ell_{e};jk} in the functions fuvf_{uv} for 1u,vm1\leq u,v\leq m with coefficients in the coordinate ring of SKS_{K} such that

1efjk=ξ1,,e;jk([fuv]).\partial_{\ell_{1}}\cdots\partial_{\ell_{e}}f_{jk}=\xi_{\ell_{1},\ldots,\ell_{e};jk}([f_{uv}]). (4)

Because SKS_{K} is smooth, given a point ss of SKS_{K} the functions z1s1,,znsnz_{1}-s_{1},\ldots,z_{n}-s_{n}, where sis_{i} is the value of ziz_{i} on ss, induce a coordinate system in the local and formal power series rings associated to SKS_{K} at ss. In these coordinates, the map ApA_{p} is given by fp1f^{-1}_{p}, where fp=[fuv]f_{p}=[f_{uv}] is a rigid-analytic matrix-valued solution to the differential equations Equation 3. The formal map κ^\widehat{\kappa} obtained using the isomorphism τ\tau then satisfies the same set of differential equations, and in particular its derivatives of all orders at ss are determined using Equation 4 by the initial condition f1(s)=Pf^{-1}(s)=P. If we then construct an analytic solution to the differential system in Equation 3 in a neighbourhood of ss satisfying f1(s)=Pf^{-1}(s)=P, the resulting analytic map induces the map κ^\widehat{\kappa} on formal power series rings. It follows that η^\widehat{\eta} is induced by a local period map, which completes the proof. ∎

Proof of 5.1:

We denote by 𝒪K,(v)\mathcal{O}_{K,(v)} the ring of integers localized at the prime ideal 𝔭\mathfrak{p} of 𝒪K\mathcal{O}_{K} corresponding to vv. We begin by showing that (base changes to KvK_{v} of) the points of S(𝒪K,(v))S(\mathcal{O}_{K,(v)}) lie inside finitely many distinguished open affinoids BpSKvanB_{p}\subset S_{K_{v}}^{\textrm{an}} admitting local period maps ψp:BpLˇKvan\psi_{p}:B_{p}\to\widecheck{{L}}_{K_{v}}^{\textrm{an}}. This reduces to showing that there are finitely many distinguished open affinoids BpSKvanB_{p}\subset S_{K_{v}}^{\textrm{an}} containing the points in S(𝒪K,(v))S(\mathcal{O}_{K,(v)}) over which Kvan\mathcal{H}_{K_{v}}^{\textrm{an}} admits a rigid-analytic flat frame. We may cover SS by finitely many open subschemes USU\subset S such that ΩU1\Omega^{1}_{U} and the bundles FkF^{k}\mathcal{H} for varying kk are all trivial. Then any point sS(𝒪K,(v))s\in S(\mathcal{O}_{K,(v)}) factors through some element of this cover, so we may reduce to the case where ΩS1\Omega^{1}_{S} and the bundles FkF^{k}\mathcal{H} are all trivial.

Proceeding as in the proof of 5.3, we can choose algebraic functions z1,,znz_{1},\ldots,z_{n} on SS such that dz1,,dzndz_{1},\ldots,dz_{n} trivializes ΩS1\Omega^{1}_{S}. We obtain differential equations as in (4), where the polynomials ξ1,,e;jk\xi_{\ell_{1},\ldots,\ell_{e};jk} are functions in the coordinate ring RR of S×GLmS\times\textrm{GL}_{m}, and so in particular we may view them after base-changing as elements of R𝒪K,vR_{\mathcal{O}_{K,v}}, where 𝒪K,v\mathcal{O}_{K,v} is the ring of vv-adic integers. Choose a point s0¯S(𝒪K,N/𝔭𝒪K,N)\overline{s_{0}}\in S(\mathcal{O}_{K,N}/\mathfrak{p}\mathcal{O}_{K,N}). Then as SS has (by assumption) good reduction modulo 𝔭\mathfrak{p}, we obtain by [DG67, IV. 18.5.17] a lift s0S(𝒪K,v)s_{0}\in S(\mathcal{O}_{K,v}). Choosing an initial condition PGLm(𝒪K,v)P\in\textrm{GL}_{m}(\mathcal{O}_{K,v}), we may use (4) to construct a map ψs0=qKvanf1\psi_{s_{0}}=q_{K_{v}}^{\textrm{an}}\circ f^{-1}, where the partial derivatives of ff at s0s_{0} are given by evaluating the polynomials ξ1,,e;jk\xi_{\ell_{1},\ldots,\ell_{e};jk} at (s0,P)(s_{0},P). As the coefficients of the power series defining ψs0\psi_{s_{0}} lie in 𝒪K,v\mathcal{O}_{K,v}, the map ψs0\psi_{s_{0}} is defined on a residue disk Bp,s0¯B_{p,\overline{s_{0}}} of radius |p|1/[Kv:p]|p|^{1/[K_{v}:\mathbb{Q}_{p}]}, where |||\cdot| is the absolute value on p\mathbb{Q}_{p}. Varying s0s_{0} over the finitely many elements of S(𝒪K,N/𝔭𝒪K,N)S(\mathcal{O}_{K,N}/\mathfrak{p}\mathcal{O}_{K,N}), we obtain the desired cover.

Now we wish to show that dimS(𝒪K,N)Zard\dim S(\mathcal{O}_{K,N})^{\textrm{Zar}}\leq d. Recall that for each sS(𝒪K,N)s\in S(\mathcal{O}_{K,N}), we have a set (s)S(𝒪K,N)\mathcal{I}(s)\subset S(\mathcal{O}_{K,N}) of points whose associated Galois representations have isomorphic semisimplifications. As there are finitely many possibilities for the semisimplification, it suffices to consider the sets 𝒮(s)S(𝒪K,N)\mathcal{S}(s)\subset S(\mathcal{O}_{K,N}) defined by

𝒮(s)={sS(𝒪K,N):(s)=(s) and ss mod 𝔭},\mathcal{S}(s)=\{s^{\prime}\in S(\mathcal{O}_{K,N}):\mathcal{I}(s)=\mathcal{I}(s^{\prime})\textrm{ and }s\equiv s^{\prime}\textrm{ mod }\mathfrak{p}\},

and show that dim𝒮(s)¯Zard\dim\overline{\mathcal{S}(s)}^{\textrm{Zar}}\leq d. In particular, we can consider the Zariski closure of just those elements of 𝒮(s)\mathcal{S}(s) whose associated points in S(𝒪K,v)S(\mathcal{O}_{K,v}) lie inside one of the neighbourhoods Bp,s0¯B_{p,\overline{s_{0}}} constructed above on which we have a local period map ψs0:Bp,s0¯LˇKvan\psi_{s_{0}}:B_{p,\overline{s_{0}}}\to\widecheck{{L}}_{K_{v}}^{\textrm{an}}. The hypothesis of the proposition tells us that the image under ψs0\psi_{s_{0}} of the points in 𝒮(s)\mathcal{S}(s) lie in a subvariety OsLˇpO_{s}\subset\widecheck{{L}}_{\mathbb{C}_{p}} satisfying dimΔddimOs\dim\Delta_{d}\geq\dim O_{s}; here we use the fact that the flat frame on Bp,s0¯B_{p,\overline{s_{0}}} is compatible with the Frobenius endomorphism (c.f. the discussion in [LV20, §3]). Base-changing to p\mathbb{C}_{p} and applying 5.3 above, we find that dimψs01(Os)¯Zard\dim\overline{\psi_{s_{0}}^{-1}(O_{s})}^{\textrm{Zar}}\leq d, hence the result.

References

  • [BBT18] Benjamin Bakker, Yohan Brunebarbe, and Jacob Tsimerman. o-minimal GAGA and a conjecture of Griffiths. arXiv e-prints, page arXiv:1811.12230, November 2018.
  • [BT17] Benjamin Bakker and Jacob Tsimerman. The Ax-Schanuel conjecture for variations of Hodge structures. Inventiones mathematicae, 217:77–94, 2017.
  • [CPMS03] James A Carlson, C. (Chris) Peters, and Stefan Müller-Stach. Period mappings and period domains. Cambridge, U.K. : Cambridge University Press, 2003.
  • [DG67] Jean Dieudonné and Alexander Grothendieck. Éléments de géométrie algébrique. Inst. Hautes Études Sci. Publ. Math., 4, 8, 11, 17, 20, 24, 28, 32, 1961–1967.
  • [Dod84] Bruce Dodson. The structure of Galois groups of CM-fields. Transactions of the American Mathematical Society, 283(1):1–32, 1984.
  • [Fon94] Jean-Marc Fontaine. Le corps des périodes p-adiques. Astérisque, 223:59–111, 1994.
  • [GGK12] Mark Green, Phillip Griffiths, and Matt Kerr. Mumford-Tate Groups and Domains: Their Geometry and Arithmetic (AM-183). Princeton University Press, 2012.
  • [KO68] Nicholas M. Katz and Tadao Oda. On the differentiation of De Rham cohomology classes with respect to parameters. J. Math. Kyoto Univ., 8(2):199–213, 1968.
  • [LS20] Brian Lawrence and Will Sawin. The Shafarevich conjecture for hypersurfaces in abelian varieties, 2020.
  • [LV20] Brian Lawrence and Akshay Venkatesh. Diophantine problems and p-adic period mappings. Inventiones mathematicae, 221:893–999, 2020.
  • [Noo21] Marc Paul Noordman. Siegel’s theorem via the Lawrence-Venkatesh method, 2021.
  • [Urb21a] David Urbanik. On the Transcendence of Period Images. arXiv preprint arXiv:2106.09342, 2021.
  • [Urb21b] David Urbanik. Sets of Special Subvarieties of Bounded Degree. arXiv preprint arXiv:2109.07663, 2021.
  • [VdDM96] Lou Van den Dries and Chris Miller. Geometric categories and o-minimal structures. Duke Mathematical Journal, 84(2):497–540, 1996.
  • [Voi12] Claire Voisin. Hodge loci. In Handbook of Moduli Vol. III, pages 520 – 559. 2012.