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

The Shafarevich conjecture for hypersurfaces in abelian varieties

Brian Lawrence  and  Will Sawin
Abstract.

Faltings proved that there are finitely many abelian varieties of genus gg over a number field KK, with good reduction outside a finite set of primes SS. Fixing one of these abelian varieties AA, we prove that there are finitely many smooth hypersurfaces in AA, with good reduction outside SS, representing a given ample class in the Néron-Severi group of AA, up to translation, as long as the dimension of AA is at least 44. Our approach builds on the approach of [47] which studies pp-adic variations of Hodge structure to turn finiteness results for pp-adic Galois representations into geometric finiteness statements. A key new ingredient is an approach to proving big monodromy for the variations of Hodge structure arising from the middle cohomology of these hypersurfaces using the Tannakian theory of sheaf convolution on abelian varieties.

1. Introduction

Fix a number field KK with ring of integers 𝒪K\mathcal{O}_{K}, and let SS be a finite set of primes of 𝒪K\mathcal{O}_{K}. Fix an abelian variety AA, defined over KK, with good reduction at all primes outside SS. We say a hypersurface HAH\subseteq A has good reduction at pSp\notin S if the closure of HH in the unique smooth projective model of AA over 𝒪K[1/S]\mathcal{O}_{K}[1/S] is smooth at pp. Our main result is the following.

Theorem 1.1 (Theorem 9.4).

Suppose dimA4\dim A\geq 4. Fix an ample class ϕ\phi in the Néron-Severi group of AA. There are, up to translation, only finitely many smooth hypersurfaces HAH\subseteq A representing ϕ\phi, with good reduction outside SS.

If we fix a Picard class ψ\psi, rather than a Néron-Severi class, this theorem becomes a finiteness result for a Diophantine equation, in principle concrete. The theorem is equivalent to the statement that there are only finitely many HH representing a given Picard class ψ\psi, because only finitely many translates of a given HH will represent ψ\psi. The hypersurfaces in a given Picard class form a projective space, and the singular ones form an irreducible divisor as soon as ψ\psi is very ample, by a classical result (e.g. [65, Theorem 1.18]) which uses the fact that AA is not ruled by projective spaces. Thus, the singular hypersurfaces are the vanishing locus of some discriminant polynomial Δ(x1,,xN)\Delta(x_{1},\dots,x_{N}) in the homogenous coordinates of that projective space. Theorem 9.4 is equivalent to the statement that, for uu any SS-unit in 𝒪K\mathcal{O}_{K}, there are only finitely many solutions of the equation Δ(x1,,xN)=u\Delta(x_{1},...,x_{N})=u with all xi𝒪K[1/S]x_{i}\in\mathcal{O}_{K}[1/S].

For dimA=3\dim A=3 there are additional combinatorial difficulties, leading to a more complicated result. Let a(i)a(i) be the sequence

1,5,20,76,285,1065,1,5,20,76,285,1065,\dots

satisfying

a(1)=1,a(2)=5,a(i+2)=4a(i+1)+1a(i)a(1)=1,a(2)=5,a(i+2)=4a(i+1)+1-a(i)

Let d(i)d(i) be the sequence

d(i)=(a(i)+a(i+1)a(i))d(i)={a(i)+a(i+1)\choose a(i)}

so that

d(1)=6,d(2)=53130,d(3)=216182590635135019896,d(4)=2.5×1079,d(1)=6,d(2)=53130,d(3)=216182590635135019896,d(4)=2.5\ldots\times 10^{79},\cdots
Theorem 1.2 (Theorem 9.5).

Suppose dimA=3\dim A=3. Fix an ample class ϕ\phi in the Néron-Severi group of AA. Assume that the intersection number ϕϕϕ\phi\cdot\phi\cdot\phi is not divisible by d(i)d(i) for any i2i\geq 2. There are only finitely many smooth hypersurfaces HAH\subseteq A representing ϕ\phi, with good reduction outside SS, up to translation.

Since a(i)a(i) increases exponentially, d(i)d(i) increases superexponentially. Because of this rapid rate of increase, and because d(2)d(2) is already large, a very small proportion of possible intersection numbers are not covered by Theorem 9.5.

If dimA=2\dim A=2, then hypersurfaces in AA are curves, and the analogue of Theorems 9.4 and 9.5 follows from the Shafarevich conjecture for curves; see Theorem 9.6.

Our result is analogous to the Shafarevich conjecture for curves, now a theorem of Faltings [21], but (except in dimension 2) it doesn’t seem to follow from Faltings’s work; we’ll say more about the relationship below. Instead, the proof uses a study of variation of Galois representations based on the work of one of the authors (B.L.) and Venkatesh [47], and the sheaf convolution formalism of Krämer and Weissauer [44].

The original Shafarevich conjecture (proved by Faltings) says that there are only finitely many isomorphism classes of curves of fixed genus gg, defined over KK, and having good reduction outside SS. Similar results are now known for various families of varieties: abelian varieties ([21]), K3 surfaces ([2] and [63]), del Pezzo surfaces ([59]), flag varieties ([32]), complete intersections of Hodge level at most 1 ([31]), surfaces fibered smoothly over a curve ([29]), Fano threefolds ([33]), and some general type surfaces ([30]).

As a consequence of a hyperbolicity result of Zuo [71], Javanpeykar and Loughran have suggested that the Shafarevich conjecture should hold in broad generality (see for example [31, Conj. 1.4]); the present result is further evidence in this direction. They show that the Lang–Vojta conjecture implies the Shafarevich conjecture for certain families of complete intersections [31, Thm. 1.5]. One expects the implication to hold for still more general families of varieties: for any family that gives rise to a locally injective period map, Zuo’s theorem shows that the base must be hyperbolic, and the argument of [31] applies. Indeed, in our proof we use a big monodromy statement (Corollary 4.10) that may be seen as a strong form of injectivity of the period map. In fact, we show that this big monodromy statement implies the quasi-finiteness of a certain period map in Proposition 4.11 below.

To understand the relationship between our work and previous work, it is helpful to compare and contrast with two previous finiteness theorems, both due to Faltings, involving abelian varieties. The first is the Shafarevich conjecture for abelian varieties [21], i.e. the result that there are only finitely many isomorphism classes of abelian varieties of dimension nn over KK with good reduction outside SS. The second is the result, in [22], that any closed subvariety of an abelian variety defined over KK that does not contain a positive-dimensional translate of an abelian subvariety contains only finitely many KK-rational points.

Both of these have been very useful for proving further arithmetic finiteness theorems. The result of [21] was applied, using the natural maps from the moduli space of curves, certain moduli spaces of K3 surfaces, and moduli spaces of complete intersections of Hodge level 1 to the moduli space of abelian varieties, to prove most of the Shafarevich-type statements discussed above. Similarly, finiteness results for points on curves over number fields of fixed degree are proven using [22] and the maps from symmetric powers of a curve to the Jacobian variety.

There does not seem to be any logical relation between our work and these two finiteness theorems. There is no reason to believe that there exists a nonconstant map from the moduli space of smooth hypersurfaces HAH\subseteq A to any moduli space of abelian varieties (except when dimA=2\dim A=2). Thus, our result does not seem to follow from [21]. There does exist a map from the moduli space of hypersurfaces to an abelian variety – in fact AA^{\vee} – by sending each hypersurface to its Picard class, but this is surjective so [22] is not helpful. Instead, this map can be used to reduce the finiteness problem to the moduli space of smooth hypersurfaces in a given Picard class, which is an open subset of projective space. Because an open subset of projective space does not have a nonconstant map to any abelian variety, [22] cannot be applied at this point.

Indeed, our main result seems to be synergistic with prior finiteness results in abelian varieties. Faltings proved that there are only finitely many abelian varieties AA of a given dimension with good reduction outside SS. One can check that each of these abelian varieties has only finitely many ample Néron-Severi classes of a given intersection number, up to automorphism. We have proven that each of these ample classes contains only finitely many smooth hypersurfaces with good reduction outside SS, up to translation. Finally Faltings proved that each of these hypersurfaces contains only finitely many KK-rational points, outside of finitely many translates of abelian subvarieties.

The present work uses general machinery introduced by B.L. and Venkatesh in [47] to study period maps and Galois representations applicable to cohomology in arbitrary degree. Significant work is required to apply this machinery in our setting. We develop a version of the sheaf convolution Tannakian category, and use it to prove a uniform big monodromy statement. We extend the methods of [47] to non-connected reductive groups. Finally, we need to do some difficult combinatorial calculations related to Hodge numbers. All of this will be explained in more detail after we recall some general ideas from [47].

The paper [47] introduces a method to bound integral points on a variety XX, assuming one can find a family over XX whose cohomology has big monodromy. Suppose YXY\rightarrow X is a smooth proper family of varieties, extending to a smooth proper SS-integral model 𝒴𝒳\mathcal{Y}\rightarrow\mathcal{X} over 𝒪K,S\mathcal{O}_{K,S}. Then for every integral point x𝒳(𝒪K,S)x\in\mathcal{X}(\mathcal{O}_{K,S}) with associated geometric point x¯X\overline{x}\in X, the étale cohomology of the fiber Yx¯Y_{\overline{x}} gives rise to a global Galois representation

ρx:GalKAut(Heti(Yx¯,p)).\rho_{x}\colon\operatorname{Gal}_{K}\rightarrow\operatorname{Aut}(H^{i}_{et}({Y}_{\overline{x}},\mathbb{Q}_{p})).

A lemma of Faltings shows that there are only finitely many possibilities for ρx\rho_{x}, up to semisimplification. In various settings, it is possible to show that the representation ρx\rho_{x} varies pp-adically in xx, and deduce that the SS-integral points 𝒳(𝒪K,S)\mathcal{X}(\mathcal{O}_{K,S}) are not Zariski dense in XX.

A key input to the methods of [47] is control on the image of the monodromy representation

π1(X,x0)Aut(Hsingi(𝒴x0)).\pi_{1}(X,x_{0})\rightarrow\operatorname{Aut}(H^{i}_{sing}(\mathcal{Y}_{x_{0}})).

(The idea that big monodromy statements might have interesting Diophantine consequences goes back at least to Deligne’s proof of the Weil conjectures [15].) In order to show that a certain period map has big image, we need to know that the Zariski closure of the image of monodromy is “big” in a certain sense. In particular, in the case studied in this paper the image of monodromy is sufficiently big if its Zariski closure includes one of the classical groups SLN,SpN,SONSL_{N},Sp_{N},SO_{N}. Because this is the sufficient condition we use in our argument, one can think of “bigness” in terms of classical groups, but the precise condition in [47] is substantially more flexible, which might prove useful elsewhere. For example, Theorem 8.17 requires that the monodromy group be “strongly cc-balanced” in the sense of Definition 6.6, as well as two numerical conditions that are more easily satisfied when the monodromy group is larger.

A major technical difficulty in this present work is the need to prove a big monodromy statement that applies uniformly to all positive-dimensional subvarieties of the moduli space Hilb\mathrm{Hilb} of hypersurfaces in AA. For the monodromy groups of the universal family over Hilb\mathrm{Hilb} itself, there are multiple geometric and topological arguments that could demonstrate that the monodromy contains a classical group. This would be sufficient to prove Zariski nondensity of the integral points Hilb(𝒪K,S)\mathrm{Hilb}(\mathcal{O}_{K,S}). Then, one hopes to improve from Zariski nondensity of Hilb(𝒪K,S)\mathrm{Hilb}(\mathcal{O}_{K,S}) to finiteness by passing to a subvariety. (This idea was suggested in [47, Sec. 10.2].) Specifically, we may take XX an irreducible component of the Zariski closure of Hilb(𝒪K,S)\mathrm{Hilb}(\mathcal{O}_{K,S}) in XX. If, under the assumption that XX is positive dimensional modulo translation, we can show that 𝒳(𝒪K,S)\mathcal{X}(\mathcal{O}_{K,S}) is not Zariski dense in XX, we obtain a contradiction. We thus deduce that XX is zero-dimensional modulo translation, and thus contains only finitely many distinct hypersurfaces modulo translation, and hence Hilb(𝒪K,S)\mathrm{Hilb}(\mathcal{O}_{K,S}) contains only finitely many hypersurfaces up to translation. However, this requires us to prove large monodromy, not just over Hilb\mathrm{Hilb}, but for every positive-dimensional-modulo-translation subvariety XHilbX\subseteq\mathrm{Hilb}.

For most families of varieties, such as hypersurfaces in projective space, this problem would seem totally insurmountable. Either we know almost nothing about the monodromy groups of arbitrary subvarieties, or, as in the universal family of abelian varieties, we can construct explicit subfamilies with too-small monodromy, e.g. low-dimensional Shimura subvarieties. However, working with a family of subvarieties of a fixed abelian variety AA provides us with a way out. The inverse image of HH under the multiplication-by-n\ell^{n} map AAA\to A has good reduction everywhere HH does, except possibly at \ell, and we can run the argument with its étale cohomology. The n\ell^{n}-torsion points A[n]A[\ell^{n}] act on this inverse image, and thus on its cohomology; this action splits the cohomology into a sum of eigenspaces, each with its own monodromy representation. It suffices for our purposes to show that one of these representations has big monodromy.

This additional freedom allows a new type of argument, based on the Tannakian theory of sheaf convolution developed by Krämer and Weissauer [44]. They defined a group, the “Tannakian monodromy group”, associated to a subvariety in an abelian variety (and in fact to much more general objects). Its definition is subtler than the definition of the usual monodromy group, but it is a better tool to work with because it depends only on a single hypersurface in the family, whose geometry can be controlled, rather than an arbitrary family of hypersurfaces, whose geometry is far murkier. We prove a group-theoretic relationship (Theorem 4.7) between the usual monodromy group of a typical A[n]A[\ell^{n}]-eigenspace in the cohomology of a family of hypersurfaces and the Tannakian monodromy group of a typical member of the family of hypersurfaces. One can think of this as analogous to the relationship between the monodromy groups of the generic horizontal and vertical fibers of a family of varieties over (an open subset of) a product X1×X2X_{1}\times X_{2}. Using purely geometric arguments involving the results of [42] and [43], we show that the Tannakian monodromy group contains a classical group, and then using Theorem 4.7, we show that the usual monodromy group does as well.

We believe the problem of proving big monodromy for the restriction of a local system to an arbitrary subvariety to be very difficult without this Tannakian method, but owing to its arithmetic applications, it would be very interesting to look for new examples where this can be established by a different method. The following vague toy problem illustrates the sort of difficulty that arises. Suppose given a variety XX of dimension at least two and a smooth family YXY\rightarrow X. One wants to obtain a strong lower bound, over all pointed curves (Z,z)X(Z,z)\subseteq X, on the dimension of the Zariski closure of the image of monodromy

π1(Z,z)AutH1(Yz).\pi_{1}(Z,z)\rightarrow\operatorname{Aut}H^{1}(Y_{z}).

This seems difficult in all but a few special cases (such as products of curves, where finiteness already follows immediately from Faltings’s theorem), but see the preprint [66] for a promising approach.

The methods of this paper can likely be applied to many different classes of subvarieties of abelian varieties, beyond hypersurfaces. To make this generalization, the additional inputs needed are a result giving some control on the Tannakian monodromy group associated to the subvariety and the verification of a certain inequality involving the Hodge numbers and this group (see Lemma A.1).

1.1. Outline of the proof

The argument of [47] derives bounds on 𝒳(𝒪K,S)\mathcal{X}(\mathcal{O}_{K,S}) from a family f:YXf\colon Y\rightarrow X, through a study of various cohomology objects Rif()R^{i}f_{*}(-) on XX. The étale local system Rif(p)R^{i}f_{*}(\mathbb{Q}_{p}) gives rise to the global Galois representations to which Faltings’s lemma is applied; a filtered FF-isocrystal coming from crystalline cohomology is used to study the pp-adic variation of these Galois representations; and a variation of Hodge structure allows one to relate a pp-adic period map to topological monodromy. The method allows one to conclude that 𝒳(𝒪K,S)\mathcal{X}(\mathcal{O}_{K,S}) is not Zariski dense in XX.

In the present setting, we will apply these results to Rif(𝖫)R^{i}f_{*}(\mathsf{L}), where 𝖫\mathsf{L} is a nontrivial local system on YY. We now outline the main construction; details are in the proof of Theorem 9.2. For technical reasons, 111Most importantly, Lemma 8.1 holds over \mathbb{Q} but not over an arbitrary number field. Additionally, pp-adic Hodge theory plays well with tensor categories over p\mathbb{Q}_{p}. Since passing to extensions (even unramified extensions) gives rise to semilinear operators, we need to set up the Tannakian formalism over p\mathbb{Q}_{p}. we prefer to work with XX a variety over \mathbb{Q}, so we take XX to be a Zariski-closed subset of the Weil restriction from KK to \mathbb{Q} of the moduli space of smooth hypersurfaces in AA. In fact, we simply take XX to be an irreducible component the Zariski closure of the set of integral points. We take YY the universal family of smooth hypersurfaces over XX; that is, points xX()x\in X(\mathbb{Q}) are in bijection with smooth hypersurfaces YxAY_{x}\subseteq A over KK. To conclude Theorem 9.2, we need to show that YY is a translate of a constant family over XX.

Let n=dimAn=\dim A. We will study cohomology objects Rif(𝖫)R^{i}f_{*}(\mathsf{L}), where 𝖫\mathsf{L} is a local system on YY, given as a direct sum of characters. Any finite-order character χ\chi on π1(A)\pi_{1}(A) defines a local system on YY, by pullback via YAY\hookrightarrow A. If χ\chi is defined over some field LL, the same is true of the corresponding local system 𝖫χ\mathsf{L}_{\chi}. We construct a local system 𝖫\mathsf{L} defined over \mathbb{Q} by descent, as a sum of Galois conjugates of 𝖫χ\mathsf{L}_{\chi}; and we take 𝖵=Rn1f(𝖫)\mathsf{V}=R^{n-1}f_{*}(\mathsf{L}). The construction is given in Lemma 5.29 in §5.5.

We think of 𝖵\mathsf{V} and 𝖫\mathsf{L} as “motives” with various realizations satisfying some compatibilities, as in Deligne [16]. The precise realizations and compatibilities we need are formalized in the notion of a Hodge–Deligne system (Definition 5.2).

The étale realization of 𝖵=Rn1f(𝖫)\mathsf{V}=R^{n-1}f_{*}(\mathsf{L}) gives, for every x𝒳([1/S])x\in\mathcal{X}(\mathbb{Z}[1/S]), a Galois representation

ρx:GalAutHk(Yx,p).\rho_{x}\colon\operatorname{Gal}_{\mathbb{Q}}\rightarrow\operatorname{Aut}H^{k}(Y_{x},\mathbb{Q}_{p}).

By Faltings’s lemma (Lemma 5.49), there are only finitely many possibilities for the semisimplification ρxss\rho_{x}^{ss} of ρx\rho_{x}, as xx varies over 𝒳([1/S])\mathcal{X}(\mathbb{Z}[1/S]).

As in [47], we want to show that the fibers of the map

xρxssx\mapsto\rho_{x}^{ss}

are not Zariski dense. To do this, we consider the map that takes a pp-adic point xX(p)x\in X(\mathbb{Q}_{p}) to a local Galois representation. By pp-adic Hodge theory, the local Galois representation

ρx,p:GalpAutHk(Yx,p)\rho_{x,p}\colon\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow\operatorname{Aut}H^{k}(Y_{x},\mathbb{Q}_{p})

determines the filtered ϕ\phi-module

(Vx,ϕx,Fx)=Hcrisk(Yx).(V_{x},\phi_{x},F_{x})=H^{k}_{cris}(Y_{x}).

We recall from [47, §3] some facts about the variation of (Vx,ϕx,Fx)(V_{x},\phi_{x},F_{x}) with xx. For xx in a fixed mod-pp residue disk Ω\Omega, the pair (Vx,ϕx)(V_{x},\phi_{x}) is constant: these spaces are canonically identified with the crystalline cohomology of the mod-pp reduction of YxY_{x}. The filtration FxF_{x} varies with xx. The assignment xFxx\mapsto F_{x} defines a pp-adic period map

Φp:Ω\Phi_{p}\colon\Omega\rightarrow\mathcal{H}

to a certain flag variety. The pp-adic period map is analogous to the classical complex-analytic period map of Hodge theory, and indeed the two maps are closely related, a fact we will exploit in Section 6 (see for example the proof of Lemma 6.1).

The global semisimplification ρxρxss\rho_{x}\mapsto\rho_{x}^{ss} causes substantial technical difficulties. Before our main argument in Section 8, we give (Section 7) an alternative, simpler proof under the additional assumption that every relevant representation ρx\rho_{x} is semisimple. For this sketch, to illustrate ideas, let us make the same assumption; that is, let us imagine that every global representation ρx\rho_{x} is semisimple. Then there are literally only finitely many possibilities for the isomorphism class ρx\rho_{x}, so (restricting to the local Galois representation and applying the crystalline Dieudonné functor) there are only finitely many possibilities (up to isomorphism) for the filtered ϕ\phi-module (Vx,ϕx,Fx)(V_{x},\phi_{x},F_{x}), as xx ranges over all integral points. In this simplified setting, we need only show that

{xX(p)(Vx,ϕx,Fx)(V0,ϕ0,F0)}\{x\in X(\mathbb{Q}_{p})\mid(V_{x},\phi_{x},F_{x})\cong(V_{0},\phi_{0},F_{0})\}

is contained in a positive-codimension algebraic subset of XX.

Isomorphism classes of triples (Vx,ϕx,Fx)(V_{x},\phi_{x},F_{x}) correspond to orbits of the Frobenius centralizer Z(ϕ)Z(\phi) on \mathcal{H}, so we want to control Φp1(Z)\Phi_{p}^{-1}(Z), where ZZ is an orbit of the Frobenius centralizer. We’ll have the result we want if we can prove precise versions of the following two conditions.

  1. (a)

    The Frobenius centralizer is small.

  2. (b)

    The image of Φp\Phi_{p} is not contained in a small algebraic set.

In fact, since we don’t know that the global Galois representations are semisimple, we need a stronger form of a.

  1. (a’)

    (See Lemma 8.16.) Fix a ϕ\phi-module (V,ϕ)(V,\phi) and a semisimple global Galois representation ρss\rho^{ss}. Consider all global Galois representations ρ\rho whose semisimplification is ρss\rho^{ss}, and such that Dcris(ρ|p)(V,ϕ,F)D_{cris}(\rho|_{\mathbb{Q}_{p}})\cong(V,\phi,F), for some filtration FF on VV.

    The FF that arise in this way all lie in a subvariety ZZ\subseteq\mathcal{H} of low dimension.

Once we have items a and b, we know that X([1/S])X(\mathbb{Z}[1/S]) is contained in Φp1(Z)\Phi_{p}^{-1}(Z). A pp-adic version of the Bakker–Tsimerman transcendence theorem (Theorem 6.4) will imply that X([1/S])X(\mathbb{Z}[1/S]) is not Zariski dense.

Condition a comes from two ingredients. First, the semilinearity of Frobenius gives an upper bound on its centralizer (Lemma 5.33). Second, the possible subrepresentations of a global Galois representation are constrained by purity (Lemma 8.1), which restricts the structure of local Galois representations coming from global ρ\rho having a given semisimplification. It is this latter result that requires us to work over \mathbb{Q} (or at least a number field that has no CM subfield). As mentioned above, we can always pass to this situation by restriction of scalars.

We introduce “H0H^{0}-algebras” (§5.3) to package the cohomological data that arise in this situation. The Hodge–Deligne system 𝖵\mathsf{V} will have the structure of module over a certain algebra object 𝖤\mathsf{E} in the category of Hodge–Deligne systems; this module structure allows us to keep track of the Galois actions on embeddings of the field KK, isomorphism classes of local systems 𝖫\mathsf{L}, and on the coefficient field of the local system, in a uniform and convenient way.

Condition b is a question about the monodromy of the variation of Hodge structure given by 𝖵\mathsf{V}. As mentioned above, we only need a very weak lower bound on the Zariski closure of the monodromy group. We call the relevant condition “strongly cc-balanced” (Definition 6.6; see Corollary 4.10 and Lemma 6.8 for precise statements). It depends on a parameter cc which must be taken sufficiently large. The technical difficulty in Corollary 4.10 is that it applies uniformly to any family of hypersurfaces in an abelian variety, as is required to prove finiteness.

It is now crucial that, in our case, YY is a subvariety of A×XA\times X, with the map ff the restriction of the projection map A×XA\times X to XX. The abelian variety AA has many rank-one local systems 𝖫\mathsf{L}, each of which we can pull back to YY, push forward to XX, and apply this machinery to. These local systems are associated to characters of the fundamental group π1(A)\pi_{1}(A).

To apply the pp-adic Hodge theory argument described above, it suffices to have a local system 𝖫\mathsf{L} on YY such that Rn1f(𝖫)R^{n-1}f_{*}(\mathsf{L}) has big monodromy in our sense. (There are some additional technical conditions that we suppress here to focus on the main difficulty.) In fact we will show big monodromy for almost all rank one local systems 𝖫\mathsf{L}, in a precise sense (Theorem 3.5 and Corollary 4.10). To do this, it is necessary to have a framework in which the vector spaces Rn1f(𝖫)xR^{n-1}f_{*}(\mathsf{L})_{x} for different local systems 𝖫\mathsf{L} can be studied all at once. This is accomplished by the Tannakian theory of sheaf convolution [44].

The fundamental objects of the Tannakian theory of sheaf convolution are perverse sheaves. The fundamental perverse sheaf for us is the constant sheaf on YxY_{x}, pushed forward to AA, and placed in degree 1n1-n. The vector space Rn1f(𝖫)xR^{n-1}f_{*}(\mathsf{L})_{x} can be recovered from this by applying the functor KH0(A,K𝖫)K\mapsto H^{0}(A,K\otimes\mathsf{L}). The theory of [44] views (a slight modification of) the category of perverse sheaves on AA as the category of representations of a certain group; the functors KH0(A,K𝖫)K\mapsto H^{0}(A,K\otimes\mathsf{L}) are almost all isomorphic to the functor taking a representation to the underlying vector space. The image of this group on the representation associated to a perverse sheaf KK is the convolution monodromy group.

We show that, if the convolution monodromy group of the constant sheaf on YxY_{x} contains a classical group for some xXx\in X, and if the family YY over XX is not equal to a translate of the constant family, then for almost all 𝖫\mathsf{L}, the monodromy groups of Rn1f(𝖫)xR^{n-1}f_{*}(\mathsf{L})_{x} contain a classical group (Theorem 4.7, Corollary 4.10). To check the condition that the convolution monodromy group of the smooth hypersurface YxY_{x} contains a classical group, we use recent results of Krämer [42, 43], to reduce to a small number of cases—essentially, the simple algebraic groups acting by their minuscule representations—and then some intricate but elementary combinatorics involving Hodge numbers to eliminate the non-classical cases.

We now conclude the argument by taking Hilb\mathrm{Hilb} to be the moduli space of hypersurfaces in a given Néron-Severi class in AA and XX to be an irreducible component of the Zariski closure of Hilb(𝒪K,S)\mathrm{Hilb}(\mathcal{O}_{K,S}). Assuming that the universal family over XX is not equal to a translate of the constant family, we find a sheaf Rn1f(𝖫)xR^{n-1}f_{*}(\mathsf{L})_{x} with big monodromy and, using pp-adic Hodge theory, show the integral points of XX are not Zariski dense. This contradicts the definition of XX as an irreducible component, so we conclude the universal family over each component is equal to a translate of the constant family. It follows that all the fibers YxY_{x} contained in each irreducible component are identical up to translation, so because there are finitely many irreducible components, there are finitely many hypersurfaces up to translation – our desired conclusion.

1.2. Sheaf convolution and uniform big monodromy

Given an abelian variety AA over an algebraically closed field, Krämer and Weissauer [44] construct a Tannakian category as a quotient of the category of perverse sheaves on AA. A perverse sheaf NN on AA is said to be negligible if its Euler characteristic is zero; the negligible sheaves form a thick subcategory, and the sheaf convolution category is defined as the quotient of the category of all perverse sheaves by the negligible sheaves. The convolution of two perverse sheaves has negligible perverse homology in nonzero degrees; in other words, it is “perverse up to negligible sheaves,” and convolution defines a tensor structure on this quotient category.

One original motivation for this construction was the Schottky problem [45]. Given a principally polarized abelian variety AA (say of dimension gg) with theta divisor Θ\Theta, one wants to know whether AA is isomorphic to a Jacobian, say JacC\operatorname{Jac}C. In this case, Θ\Theta would be the (g1)(g-1)-st convolution power of CC. Informally, the role of the Tannakian formalism here is to determine whether Θ\Theta is “a (g1)(g-1)-st convolution power of something.”

An alternate motivation for the sheaf convolution theory comes from work of Katz. This time, one works with an abelian variety AA over a finite field 𝔽q\mathbb{F}_{q}. A perverse sheaf KK on AA has a trace function fKf_{K} on A(𝔽q)A(\mathbb{F}_{q}). Associated to a character χ\chi of A(𝔽q)A(\mathbb{F}_{q}) is the character sum xA(𝔽q)fK(x)χ(x)\sum_{x\in A(\mathbb{F}_{q})}f_{K}(x)\chi(x). Katz showed (in unpublished work analogous to [37]) that the distribution of xA(𝔽q)fK(x)χ(x)\sum_{x\in A(\mathbb{F}_{q})}f_{K}(x)\chi(x), viewed as a random variable for uniformly random χ\chi, converges to a distribution determined by the convolution monodromy group, in the limit as qq goes to \infty. More precisely, the distribution is like the trace of a random element in the maximal compact subgroup of the convolution monodromy group. To gain some intuition for this, note that given representations V1V_{1}, V2V_{2}, we have tr(g,V1V2)=tr(g,V1)tr(g,V2)\operatorname{tr}(g,V_{1}\otimes V_{2})=\operatorname{tr}(g,V_{1})\operatorname{tr}(g,V_{2}); that is, taking the tensor product of representations has the effect of multiplying the traces. For the character sums xA(𝔽q)fK(x)χ(x)\sum_{x\in A(\mathbb{F}_{q})}f_{K}(x)\chi(x), convolution has the same effect:

xA(𝔽q)(fK1fK2)(x)χ(x)=(xA(𝔽q)fK1(x)χ(x))(xA(𝔽q)fK2(x)χ(x)).\sum_{x\in A(\mathbb{F}_{q})}(f_{K_{1}}*f_{K_{2}})(x)\chi(x)=\left(\sum_{x\in A(\mathbb{F}_{q})}f_{K_{1}}(x)\chi(x)\right)\left(\sum_{x\in A(\mathbb{F}_{q})}f_{K_{2}}(x)\chi(x)\right).

In other words, convolution of these functions fKf_{K} has a similar effect on this sum as tensor product of the representations VV has on the trace. It stands to reason that a framework where perverse sheaves correspond to representations, and convolution of sheaves correspond to tensor product of representations, would have relevance to the distribution of the trace. In particular, this should be plausible if one is familiar with Deligne’s equidistribution theorem [15, Theorem 3.5.3], whose proof is similar to the argument Katz uses to establish the relationship between the distribution and the convolution monodromy group [37, Corollary 7.4].

For non-algebraically closed fields, such as finite fields, we can construct a Tannakian category in almost the same way as Krämer and Weissauer did—again defining negligible sheaves as those with zero Euler characteristic. The key facts (for example, that the convolution of two perverse sheaves has negligible perverse cohomology in nonzero degrees) hold over the base field once checked over its algebraic closure.

To relate these two categories, it is convenient to restrict attention to geometrically semisimple perverse sheaves on AkA_{k}, and to perverse sheaves on Ak¯A_{\overline{k}} which are summands of the pullback from AkA_{k} to Ak¯A_{\overline{k}} of geometrically semisimple perverse sheaves. Having done this, we obtain (in Lemma 2.8) an exact sequence of pro-algebraic groups

(1) 1Gk¯GkGalk11\to G_{\overline{k}}\to G_{k}\to\operatorname{Gal}_{k}\to 1

where GkG_{k} is the Tannakian group of a suitable category of perverse sheaves on AkA_{k}, Gk¯G_{\overline{k}} is the Tannakian group of a suitable category of perverse sheaves on Ak¯A_{\overline{k}}, and Galk\operatorname{Gal}_{k} is the Tannakian group of the category of \ell-adic Gal(k¯/k)\operatorname{Gal}(\overline{k}/k)-representations – in other words, the Zariski closure of Gal(k¯/k)\operatorname{Gal}(\overline{k}/k) in the product of the general linear groups of all its finite-dimensional \ell-adic representations. We think of this as a close analogue of the usual exact sequence

1π1geom(X)π1arith(X)Gal(k¯/k)11\to\pi_{1}^{geom}(X)\to\pi_{1}^{arith}(X)\to\operatorname{Gal}(\overline{k}/k)\to 1

for a variety XX over a field kk.

Just like this usual exact sequence, (1) often has splittings. In our case, splittings arise from certain local systems 𝖫\mathsf{L} on AA defined over kk, as the cohomology of a perverse sheaf twisted by a local system is a Galois representation, on which Galk\operatorname{Gal}_{k} acts, and we can check that this action factors through the Tannakian group GkG_{k}, giving the splitting.

Fix now a subvariety XX of the moduli space of smooth hypersurfaces in an abelian variety AA. Let kk be the field of functions on the generic point of XX. Let HH be the universal hypersurface in AA, defined over kk. Let KK be the constant sheaf on HH, pushed forward to AA, placed in degree 1n1-n; this is our perverse sheaf of interest. Associated to kk is a representation of GkG_{k}. The action of Gk¯G_{\overline{k}} on this representation is a purely geometric object. By geometric methods, we will show in Theorem 3.5 that the image of Gk¯G_{\overline{k}} acting on this representation contains SLN,SONSL_{N},SO_{N}, or SpNSp_{N} as a normal subgroup. So the image of GkG_{k} on the representation associated to kk contains the same classical group as a normal subgroup. Because the action of Galk\operatorname{Gal}_{k} in this setting matches the action of the fundamental group, it will suffice for our big monodromy theorem to show that the action of Galk\operatorname{Gal}_{k} also contains (as a normal subgroup) the same classical group.

To do this, we construct in Lemma 4.5 a battery of tests, each consisting of a representation of the normalizer of the classical group, such that any subgroup of the normalizer contains the classical group if and only if it has no invariants on any of these representations. Associated to each of these representations is a perverse sheaf on AkA_{k}. We prove Lemma 4.4 showing that the action of Galk\operatorname{Gal}_{k} on the cohomology of a perverse sheaf, defined using a generic local system 𝖫\mathsf{L}, has invariants if and only if the perverse sheaf has a very special form. Using Lemma 4.1, we check that the relevant perverse sheaves do not have this special form unless the family of hypersurfaces over XX is constant, up to translation by a section of AA, completing the proof of Theorem 4.7.

Next we describe how we check in Theorem 3.5 that the image of the Gk¯G_{\overline{k}}-action on the representation associated to a smooth hypersurface in AA contains a classical group acting by the standard representation as a normal subgroup. This proceeds in two steps. The first step shows (in Lemmas 3.8 and 3.9) that the commutator of the identity component of this image group is a simple algebraic group acting by a minuscule representation. (Recall that a minuscule representation is one where the eigenvalues of the maximal torus action are conjugate under the Weyl group.) The second step eliminates (in Lemmas 3.12 and 3.14 and Proposition 3.15) all such pairs of a group and a representation except the standard representations of the classical groups. The first step is a conceptual proof using sophisticated machinery from [42, 43], while the second uses no additional machinery (except a bit of Hodge theory) but involves an intricate combinatorial argument.

For the first step, we apply results of Krämer that study the characteristic cycle of a perverse sheaf. This is a fundamental invariant of any perverse sheaf on a smooth variety, defined as an algebraic cycle on the cotangent bundle of that variety. (For abelian varieties, the cotangent bundle is a trivial vector bundle.) By examining how the characteristic cycle of a perverse sheaf changes when it is convolved with another perverse sheaf, Krämer was able to relate the convolution monodromy group to the characteristic cycle. In particular, he gave criteria for the commutator subgroup of the identity component to be a simple group, and for the representation of it to be minuscule. The fact that our hypersurface is smooth makes its characteristic cycle relatively simple—it is simply the conormal bundle to the hypersurface. This makes Krämer’s minisculeness criterion straightforward to check, but to check simplicity we must make a small modification to Krämer’s argument. The reason for this is that Krämer, motivated by the theta divisor and the Schottky problem, assumed that a hypersurface in AA was invariant under the inversion map, while we do not wish to assume this.

For the second step, the exceptional groups and spin groups are not too hard to eliminate, as they only occur for representations of very specific dimensions. The Tannakian dimension in our setting is the topological Euler characteristic of the hypersurface, which we have an explicit formula for. Comparing these, we can see in Lemma 3.12 that the problematic cases only occur for curves in an abelian surface, which are excluded by the assumption dim(A)3\dim(A)\geq 3. The only remaining case, except for the good classical cases, is the case of a special linear group acting by a wedge power representation. For this representation, the Euler characteristic formula is not sufficient, but we are eventually able to rule this case out using a more sophisticated numerical invariant, the Hodge numbers. If the convolution monodromy group acts on the representation associated to HH by the kk-th wedge power of an mm-dimensional representation, we might expect that the Hodge structure on the cohomology of HH, or the cohomology of HH twisted by a rank one local system, is the kk-th wedge power of an mm-dimensional Hodge structure. This would place some restrictions on the Hodge numbers. We don’t prove this, but instead prove in Lemma 3.14 a pp-adic Hodge-theoretic analogue, using the Galk\operatorname{Gal}_{k}-action discussed earlier. On the other hand, we can calculate the Hodge numbers of the cohomology of HH twisted by a rank one local system using the Hirzebruch-Riemann-Roch formula. Working this out gives a complicated set of combinatorial relations between the Hodge numbers of the original mm-dimensional Hodge structure. By a lengthy combinatorial argument in Appendix B, we find all solutions of these relations, noting in particular that they occur only for abelian varieties of dimension less than four. This completes the proof.

1.3. Outline of the paper

The argument proceeds in three parts.

First, we use the sheaf convolution formalism to prove a big monodromy result for families of hypersurfaces. In Section 2 we introduce the sheaf convolution category, a Tannakian category of perverse sheaves on an abelian variety. In Section 3 we investigate the convolution monodromy group of a hypersurface; we show in many cases that this group must be as big as possible. In Section 4 we relate the convolution monodromy group to the geometric monodromy group, which gives the big monodromy statement we need.

Sections 58 explain how to deduce non-density of integral points, following the strategy in [47]. Section 5 contains some technical preliminaries. We introduce the notion of Hodge–Deligne system, which is closely related to Deligne’s “system of realizations” of a motive, although we include only the realizations that are relevant for our argument. We discuss “H0H^{0}-algebras”, roughly, algebra objects in the category of Artin motives with rational coefficients, which we need to express the semilinearity of Frobenius. We also recall some facts from the theory of not-necessarily-connected reductive groups. Section 6 relates the big monodromy statement from Section 4 to the pp-adic period map, via the theorem of Bakker and Tsimerman ([3]). In Section 7, we deduce the non-density of integral points, under the simplifying assumption that all the global representations that arise are semisimple. In Section 8, we prove the theorem in full generality. The argument used to handle the global semisimplification involves combinatorics on reductive groups, analogous to [47, §11]. We conclude with Theorem 8.17, which is analogous to Lemma 4.2, Prop. 5.3, and Thm. 10.1 in [47].

Finally, we wrap up the proof of our main theorem in Section 9.

Appendices A, B, and C contain some purely combinatorial calculations involving Eulerian numbers. Appendix A verifies the two numerical conditions in the hypotheses of Theorem 8.17. Appendix B is devoted to the proof of Prop. 3.15, which is used to show that the representation of the Tannakian group associated to a smooth hypersurface is not the wedge power of a smaller-dimensional representation—the last remaining case where the Tannakian group could be too small, and Appendix C contains inequalities that are used in Appendix B.

1.4. Acknowledgements

We would like to thank Johan de Jong, Matthew Emerton, Sergey Gorchinskiy, Ariyan Javanpeykar, Caleb Ji, Shizhang Li, Benjamin Martin, Bjorn Poonen, Akshay Venkatesh, Thomas Krämer, and Marco Maculan for interesting discussions related to this project. We would like to thank the three anonymous referees for numerous helpful comments.

This work was conducted while Will Sawin served as a Clay Research Fellow, and, later, was supported by NSF grant DMS-2101491. Brian Lawrence would like to acknowledge support from the National Science Foundation. We met to work on this project at the Oberwolfach Research Institute for Mathematics, Columbia University, and the University of Chicago; we would like to thank these institutions for their hospitality.

2. Sheaf convolution over a field

A Tannakian category over a field FF of characteristic 0 is a rigid symmetric monoidal FF-linear abelian category with a faithful exact tensor functor to the category of vector spaces over FF. The point of these conditions is that Tannakian categories are necessarily equivalent to the category of representations of some pro-algebraic group (the group of automorphisms of the functor), together with the forgetful functor to the category of vector spaces. Thus, associated to each object is some representation of this pro-algebraic group. For such a representation VV, we refer to the image of the Tannakian group inside GL(V)GL(V) as the Tannakian monodromy group.

Krämer and Weissauer [44] constructed a Tannakian category as a quotient of the category of perverse sheaves on an abelian variety over an algebraically closed field (initially of characteristic zero, but Weissauer [70] later extended it to characteristic pp), where the tensor operation is sheaf convolution. We will use the Tannakian monodromy groups from their theory, which we call the convolution monodromy groups, to control usual monodromy groups.

In this section, we check that these convolution monodromy groups behave similarly to the usual monodromy groups with respect to the distinction between the geometric and arithmetic monodromy groups. In the setting of the étale fundamental group, we can define both geometric and arithmetic monodromy groups, with the geometric a normal subgroup of the arithmetic. We will check that the same works here. The Tannakian group of the category defined by Krämer and Weissauer will function as the geometric group, and we will define a Tannakian category of perverse sheaves over a non-algebraically closed field whose Tannakian monodromy group will function as the arithmetic group. We will verify that the geometric group is a normal subgroup of the arithmetic group.

Our construction of the Tannakian category over a non-algebraically closed field will follow a version of the strategy of Krämer and Weissauer, and thus will also serve as a very brief review of their construction.

Let AA be an abelian variety over a field kk of characteristic zero. Fix a prime pp. Let Dcb(A,¯p)D^{b}_{c}(A,\overline{\mathbb{Q}}_{p}) be the derived category of bounded complexes of pp-adic sheaves on AA with constructible cohomology. Define a sheaf convolution functor :Dcb(A,¯p)×Dcb(A,¯p)Dcb(A,¯p)*\colon D^{b}_{c}(A,\overline{\mathbb{Q}}_{p})\times D^{b}_{c}(A,\overline{\mathbb{Q}}_{p})\to D^{b}_{c}(A,\overline{\mathbb{Q}}_{p}) that sends complexes K1,K2K_{1},K_{2} to

K1K2=Ra(K1K2)K_{1}*K_{2}=Ra_{*}(K_{1}\boxtimes K_{2})

for a:A×AAa\colon A\times A\to A the group law.

Lemma 2.1.

(Dcb(A,¯p),)(D^{b}_{c}(A,\overline{\mathbb{Q}}_{p}),*) is a rigid symmetric monoidal category, where the unit object is the skyscraper sheaf at 0, and the dual of a complex KK is

K=[1]DKK^{\vee}=[-1]^{*}DK

where DD is Verdier duality and [1]:AA[-1]\colon A\to A is the inversion map.

Proof.

These were proved in [67, §2.1] (the symmetric monoidality and unit) and [69, Proposition] (the rigidity and description of the dual). These results are stated in the case where kk is an algebraically closed field, but they proceed without modification in the general case. ∎

Let 𝒫\mathcal{P} be the category of perverse sheaves on AA with ¯p\overline{\mathbb{Q}}_{p}-coefficients. Let 𝒩\mathcal{N} be the subcategory of perverse sheaves with Euler characteristic zero. We similarly write 𝒫k¯\mathcal{P}_{\overline{k}} and 𝒩k¯\mathcal{N}_{\overline{k}} for the category of perverse sheaves on Ak¯A_{\overline{k}} and its subcategory of objects with Euler characteristic zero, respectively. Let Db(𝒩)D^{b}(\mathcal{N}) be the category of complexes in Dcb(A,¯p)D^{b}_{c}(A,\overline{\mathbb{Q}}_{p}) whose perverse homology objects lie in 𝒩\mathcal{N}.

The Tannakian category will be constructed by combining this rigid symmetric monoidal structure with the abelian structure on the category of perverse sheaves. This requires modifying the category of perverse sheaves slightly because it is not quite stable under convolution. Instead one verifies that it is stable under convolution “up to” 𝒩\mathcal{N}, i.e. that the convolution of two perverse sheaves has all perverse homology objects in nonzero degrees lying in 𝒩\mathcal{N}. This lets us give 𝒫/𝒩\mathcal{P}/\mathcal{N} the structure of a rigid symmetric monoidal ¯p\overline{\mathbb{Q}}_{p}-linear abelian category.

Lemma 2.2.
  1. (1)

    Perverse sheaves on AA have nonnegative Euler characteristics.

  2. (2)

    𝒩\mathcal{N} is a thick subcategory of 𝒫\mathcal{P} (i.e. it is stable under taking subobjects, quotients, and extensions).

  3. (3)

    Db(𝒩)D^{b}(\mathcal{N}) is a thick subcategory of Dcb(A,¯p)D^{b}_{c}(A,\overline{\mathbb{Q}}_{p}) (i.e. for any distinguished triangle with two objects in Db(𝒩)D^{b}(\mathcal{N}), the third one is in Db(𝒩)D^{b}(\mathcal{N}) as well).

  4. (4)

    For K1,K2Dcb(A,¯p)K_{1},K_{2}\in D^{b}_{c}(A,\overline{\mathbb{Q}}_{p}), if K1K_{1} or K2K_{2} lies in Db(𝒩)D^{b}(\mathcal{N}), then K1K2K_{1}*K_{2} lies in Db(𝒩)D^{b}(\mathcal{N}).

  5. (5)

    For K1,K2𝒫K_{1},K_{2}\in\mathcal{P}, ip(K1K2)𝒩{}^{p}\mathcal{H}^{i}(K_{1}*K_{2})\in\mathcal{N} if i0i\neq 0.

  6. (6)

    Convolution descends to a functor

    Dcb(A,¯p)/Db(𝒩)×Dcb(A,¯p)/Db(𝒩)Dcb(A,¯p)/Db(𝒩).D^{b}_{c}(A,\overline{\mathbb{Q}}_{p})/D^{b}(\mathcal{N})\times D^{b}_{c}(A,\overline{\mathbb{Q}}_{p})/D^{b}(\mathcal{N})\to D^{b}_{c}(A,\overline{\mathbb{Q}}_{p})/D^{b}(\mathcal{N}).
  7. (7)

    The essential image of 𝒫\mathcal{P} in Dcb(A,¯p)/Db(𝒩)D^{b}_{c}(A,\overline{\mathbb{Q}}_{p})/D^{b}(\mathcal{N}) is equivalent to 𝒫/N\mathcal{P}/N.

  8. (8)

    The essential image of 𝒫\mathcal{P} in Dcb(A,¯p)/Db(𝒩)D^{b}_{c}(A,\overline{\mathbb{Q}}_{p})/D^{b}(\mathcal{N}) is stable under convolution.

  9. (9)

    (𝒫/𝒩,)(\mathcal{P}/\mathcal{N},*) is a rigid symmetric monoidal ¯p\overline{\mathbb{Q}}_{p}-linear abelian category.

Proof.

It suffices to check the first five statements after passing to Ak¯A_{\overline{k}}, where they were checked in [24, Corollary 1.4], [44, Prop  10.1 and preceding paragraph], and [44, Lemma 13.1]. The remainder follow from the first five by the arguments in, e.g., [41, p. 90, Theorem 5.1, Theorem 5.2].

We will work with lisse rank-one sheaves on an abelian variety. It will be convenient to parametrize them by representations of the fundamental group.

Definition 2.3.

Let AA be an abelian variety over a field kk. Fix a continuous character χ:π1et(Ak¯)¯p×\chi\colon\pi_{1}^{et}(A_{\overline{k}})\to\overline{\mathbb{Q}}_{p}^{\times}. We define the character sheaf χ\mathcal{L}_{\chi} to be the unique rank-one sheaf on Ak¯A_{\overline{k}} whose monodromy representation is χ\chi.

We also have a canonical way to descend these sheaves to AkA_{k}:

Definition 2.4.

Let AA be an abelian variety over a field kk. Let χ\chi be a character of π1et(Ak¯)\pi_{1}^{et}(A_{\overline{k}}) that is Gal(k¯|k)\operatorname{Gal}(\overline{k}|k)-invariant. We define the character sheaf χ\mathcal{L}_{\chi} to be the unique lisse rank-one sheaf on AkA_{k} whose associated representation of the fundamental group restricts to χ\chi on π1et(Ak¯)\pi_{1}^{et}(A_{\overline{k}}) and whose stalk at the identity has trivial Galois action.

In other words, we take the splitting of the exact sequence 1π1et(Ak¯)π1et(Ak)Gal(k¯|k)11\to\pi_{1}^{et}(A_{\overline{k}})\to\pi_{1}^{et}(A_{k})\to\operatorname{Gal}(\overline{k}|k)\to 1 induced by the identity at 11, and use it to extend χ\chi from π1et(Ak¯)\pi_{1}^{et}(A_{\overline{k}}) to π1et(Ak)\pi_{1}^{et}(A_{k}).

For χ\chi a character of π1et(Ak¯)\pi_{1}^{et}(A_{\overline{k}}), let 𝒫χ\mathcal{P}^{\chi} be the subcategory of 𝒫\mathcal{P} consisting of perverse sheaves KK with Hi(Ak¯,Qχ)=0H^{i}(A_{\overline{k}},Q\otimes\mathcal{L}_{\chi})=0 for all i0i\neq 0 and QQ a subquotient of Kk¯K_{\overline{k}}, and let 𝒩χ=𝒫χ𝒩\mathcal{N}^{\chi}=\mathcal{P}^{\chi}\cap\mathcal{N}.

Lemma 2.5.
  1. (1)

    The essential image of 𝒫χ\mathcal{P}^{\chi} in 𝒫/𝒩\mathcal{P}/\mathcal{N} is equivalent to 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi}.

  2. (2)

    𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi} contains the unit and is stable under convolution and duality.

  3. (3)

    KH0(Ak¯,Kχ)K\mapsto H^{0}(A_{\overline{k}},K\otimes\mathcal{L}_{\chi}) is an exact tensor functor from 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi} to ¯p\overline{\mathbb{Q}}_{p}-vector spaces.

  4. (4)

    The category 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi}, convolution, and the functor H0(Ak¯,Kχ)H^{0}(A_{\overline{k}},K\otimes\mathcal{L}_{\chi}) are a rigid symmetric monoidal ¯p\overline{\mathbb{Q}}_{p}-linear abelian category with a faithful exact tensor functor to ¯p\overline{\mathbb{Q}}_{p}-vector spaces.

Proof.

(1) follows from [44, Lemma 12.3] and the fact that 𝒫χ\mathcal{P}^{\chi}, by construction, is a thick subcategory.

The claims in (2) may be checked after passing to an algebraically closed field. To check that it contains the unit, we must check that the skyscraper sheaf at zero has cohomology only in degree zero, which is obvious. To check that it is closed under duality, it suffices to observe that

Hi(Ak¯,[1]DQχ)\displaystyle H^{i}(A_{\overline{k}},[-1]^{*}DQ\otimes\mathcal{L}_{\chi}) =\displaystyle= Hi(Ak¯,DQ[1]χ)=Hi(Ak¯,DQχ1)\displaystyle H^{i}(A_{\overline{k}},DQ\otimes[-1]_{*}\mathcal{L}_{\chi})=H^{i}(A_{\overline{k}},DQ\otimes\mathcal{L}_{\chi}^{-1})
=\displaystyle= Hi(Ak¯,D(Qχ))=Hi(Ak¯,Qχ)\displaystyle H^{i}(A_{\overline{k}},D(Q\otimes\mathcal{L}_{\chi}))=H^{-i}(A_{\overline{k}},Q\otimes\mathcal{L}_{\chi})^{\vee}

so if one vanishes for all i0i\neq 0 the other does. That it is closed under convolution is checked in [44, Theorem 13.2].

The claims in (3) may be checked after passing to an algebraically closed field, where they are proved in [44, Theorem 13.2]. Specifically, [44, Proposition 4.1] reduces this to the case where χ\chi is trivial. In this case, exactness follows from the long exact sequence of cohomology, which reduces to a short exact sequence because higher and lower cohomology groups vanish, and tensorness follows from the Künneth formula, which gives

H0(Ak¯,a(K1K2))=H0(Ak¯×Ak¯,K1K2)=H0(Ak¯,K1)H0(Ak¯,K2)H^{0}(A_{\overline{k}},a_{*}(K_{1}\boxtimes K_{2}))=H^{0}(A_{\overline{k}}\times A_{\overline{k}},K_{1}\boxtimes K_{2})=H^{0}(A_{\overline{k}},K_{1})\otimes H^{0}(A_{\overline{k}},K_{2})

again using the vanishing of higher and lower cohomology.

(4) The category 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi} is rigid symmetric monoidal by part (2) and Lemma 2.2(4). It is ¯p\overline{\mathbb{Q}}_{p}-linear abelian because it is the quotient of a ¯p\overline{\mathbb{Q}}_{p}-linear abelian category by a thick subcategory. The functor is an exact tensor functor by part (3), and is faithful since exact ¯p\overline{\mathbb{Q}}_{p}-linear tensor functors between rigid abelian ¯p\overline{\mathbb{Q}}_{p}-linear tensor categories are automatically faithful if the endomorphisms of the unit are ¯p\overline{\mathbb{Q}}_{p}. ∎

By [44, Theorem 1.1], for any K𝒫K\in\mathcal{P}, there exists χ\chi such that K𝒫χK\in\mathcal{P}^{\chi}. We use the fiber functor KH0(Ak¯,Kχ)K\mapsto H^{0}(A_{\overline{k}},K\otimes\mathcal{L}_{\chi}) on 𝒫χ\mathcal{P}^{\chi} to define the “Tannakian group” of KK. This group is independent of the choice of χ\chi, since any two fiber functors on the same Tannakian category over an algebraically closed field are equivalent [17, Theorem 3.2], and give rise to equivalent Tannnakian groups.

For AA an abelian variety over a field kk with algebraic closure k¯\overline{k}, we say that a perverse sheaf KK on AA is geometrically semisimple if its pullback to Ak¯A_{\overline{k}} is a sum of irreducible perverse sheaves.

Lemma 2.6.

Let K1K_{1} and K2K_{2} be geometrically semisimple perverse sheaves on AA. Then Hom𝒫/𝒩(K1,K2)\operatorname{Hom}_{\mathcal{P}/\mathcal{N}}(K_{1},K_{2}) is the quotient of the space of homorphisms K1K2K_{1}\to K_{2} by the subspace of homomorphisms factoring through an element of 𝒩\mathcal{N}.

Proof.

Without loss of generality, we may assume that K1K_{1} and K2K_{2} are indecomposable.

We first check that the set of isomorphism classes of irreducible components of the pullback K1,k¯K_{1,\overline{k}} of K1K_{1} to k¯\overline{k} forms a single Gal(k¯|k)\operatorname{Gal}(\overline{k}|k)-orbit. Suppose not; then we can fix a Gal(k¯|k)\operatorname{Gal}(\overline{k}|k)-orbit and consider an endomorphism of K1,k¯K_{1,\overline{k}} defined as the idempotent projector onto the sum of all irreducible components in that orbit. This endomorphism is, by construction, stable under Gal(k¯|k)\operatorname{Gal}(\overline{k}|k). Because K1K_{1} is perverse, Rom(K1,K1)R{\mathcal{H}}om(K_{1},K_{1}) is concentrated in degrees 0\geq 0 and thus

Hom(K1,K1)=H0(A,Rom(K1,K1))=H0(Ak¯,Rom(K1,K1))Gal(k¯|k)=Hom(K1,k¯,K1,k¯)Gal(k¯|k)\operatorname{Hom}(K_{1},K_{1})=H^{0}(A,R{\mathcal{H}}om(K_{1},K_{1}))=H^{0}(A_{\overline{k}},R{\mathcal{H}}om(K_{1},K_{1}))^{\operatorname{Gal}(\overline{k}|k)}=\operatorname{Hom}(K_{1,\overline{k}},K_{1,\overline{k}})^{\operatorname{Gal}(\overline{k}|k)}

and hence this endomorphism arises from a nontrivial idempotent endomorphism of K1K_{1}, contradicting the irreducibility of K1K_{1}.

It follows that either all irreducible components of K1,k¯K_{1,\overline{k}} are in 𝒩k¯\mathcal{N}_{\overline{k}} or none of them are. The same is true for K2,k¯K_{2,\overline{k}} by the same argument.

If all irreducible components of K1,k¯K_{1,\overline{k}} or K2,k¯K_{2,\overline{k}} are in 𝒩k¯\mathcal{N}_{\overline{k}}, then K1K_{1} or K2K_{2} is in 𝒩\mathcal{N}, so maps in the quotient category are zero and all maps factor through elements of 𝒩\mathcal{N}, and the statement holds.

Thus, we may assume that no irreducible components of K1K_{1} and K2K_{2} are in 𝒩\mathcal{N}. By definition, Hom𝒫/𝒩(K1,K2)\operatorname{Hom}_{\mathcal{P}/\mathcal{N}}(K_{1},K_{2}) is the limit of Hom(K1,K2)\operatorname{Hom}(K_{1}^{\prime},K_{2}^{\prime}) where K1K_{1}^{\prime} is a subobject of K1K_{1} whose quotient lies in 𝒩\mathcal{N} and K2K_{2}^{\prime} is a quotient of K2K_{2} by an object in 𝒩\mathcal{N}. By assumption, we must have K1=K1K_{1}^{\prime}=K_{1} and K2=K2K_{2}^{\prime}=K_{2}, so Hom𝒫/𝒩(K1,K2)=Hom(K1,K2)\operatorname{Hom}_{\mathcal{P}/\mathcal{N}}(K_{1},K_{2})=\operatorname{Hom}(K_{1},K_{2}). Again because no irreducible components of K1K_{1} and K2K_{2} lie in 𝒩\mathcal{N}, no nonzero map from K1K_{1} to K2K_{2} factors through an object in 𝒩\mathcal{N}, so the statement holds in this case as well. ∎

Lemma 2.7.
  1. (1)

    The full subcategory of 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi} consisting of geometrically semisimple perverse sheaves is a Tannakian subcategory of 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi}.

  2. (2)

    The full subcategory of 𝒫k¯χ/𝒩k¯χ\mathcal{P}_{\overline{k}}^{\chi}/\mathcal{N}_{\overline{k}}^{\chi} consisting of summands of the pullbacks to Ak¯A_{\overline{k}} of geometrically semisimple elements of 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi} on AkA_{k} is a Tannakian subcategory of 𝒫k¯χ/𝒩k¯χ\mathcal{P}^{\chi}_{\overline{k}}/\mathcal{N}^{\chi}_{\overline{k}}.

Proof.

To prove part (1), we must check that this subcategory contains the unit, and is closed under kernels, cokernels, direct sums, convolution, duals. The unit, direct sum, and dual steps are straightforward. For kernels and cokernels, by Lemma 2.6 it suffices to check that kernels and cokernels of morphisms between geometrically semisimple sheaves are geometrically semisimple, which is clear. For convolution, this follows from Kashiwara’s conjecture, proven in [18] and [27].

To prove part (2), the argument for the units, direct sum, convolution, and dual steps is straightforward, again using Kashiwara’s conjecture. For the kernel and cokernel, the key is that summands of the pullback of geometrically semisimple perverse sheaves on AkA_{k} remain semisimple. Semisimplicity allows us to apply Lemma 2.6 to reduce to kernels and cokernels of honest morphisms and then shows that those kernels and cokernels are themselves summands. ∎

Fix an abelian variety AA over kk and a character χ\chi of π1et(Ak¯)\pi_{1}^{et}(A_{\overline{k}}).

Let GkG_{k} be the Tannakian fundamental group of the full subcategory of 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi} consisting of geometrically semisimple perverse sheaves.

Let Gk¯G_{\overline{k}} be the Tannakian fundamental group of the full subcategory of 𝒫k¯χ/𝒩k¯χ\mathcal{P}_{\overline{k}}^{\chi}/\mathcal{N}_{\overline{k}}^{\chi} consisting of summands of the pullbacks to Ak¯A_{\overline{k}} of geometrically semisimple perverse sheaves on AkA_{k}.

Let Galk\operatorname{Gal}_{k} be the Tannakian group of the category of pp-adic Galois representations over kk.

Lemma 2.8.

The group Gk¯G_{\overline{k}} is a normal subgroup of GkG_{k}, with quotient Galk\operatorname{Gal}_{k}.

Proof.

There is a functor from the Tannakian category of Galois representations over kk to the geometrically semisimple objects of 𝒫χ/𝒩χ\mathcal{P}^{\chi}/\mathcal{N}^{\chi} that sends a Galois representation to the corresponding skyscraper sheaf at the identity.

There is a functor from geometrically semisimple perverse sheaves on AkA_{k} to summands of pullbacks of geometrically semisimple perverse sheaves to Ak¯A_{\overline{k}}, given by pullback to Ak¯A_{\overline{k}}.

Because these functors are both exact tensor functors, they define homomorphisms Gk¯GkGalkG_{\overline{k}}\to G_{k}\to\operatorname{Gal}_{k}. We wish to show that this is an exact sequence of groups, i.e. that Gk¯G_{\overline{k}} is a normal subgroup of GkG_{k} whose quotient is Galk\operatorname{Gal}_{k}. To do this, we check the criteria of [19, Theorem A.1] (which incorporate earlier results of [17, Proposition 2.21]).

First, to check that GkGalkG_{k}\to\operatorname{Gal}_{k} is surjective, it suffices to check that the functor from Galois representations to skyscraper sheaves at the origin is full, and that a subquotient of a skyscraper sheaf at the origin is a skyscraper sheaf at the origin. These are both easy to check.

Second, to check that Gk¯GkG_{\overline{k}}\to G_{k} is a closed immersion, we must check that every representation of Gk¯G_{\overline{k}} is a subquotient of a pullback to Gk¯G_{\overline{k}} of a representation of GkG_{k}. This is automatic, as the Tannakian category of representations of Gk¯G_{\overline{k}} is defined to consist of perverse sheaves that are summands of pullbacks of perverse, geometrically semisimple sheaves on AkA_{k} that lie in PχP^{\chi}, which by definition are representations of GkG_{k}, and because all summands are subquotients.

Third we must check that a perverse sheaf on AkA_{k} is a skyscraper sheaf at the origin if and only if is trivial when pulled back to Ak¯A_{\overline{k}}. This is obvious.

Fourth, we must check that given a geometrically semisimple perverse sheaf on AkA_{k}, its maximal trivial subobject over Ak¯A_{\overline{k}} (i.e. the maximal sub-perverse sheaf that is a skyscraper sheaf at the origin) is a subobject over AkA_{k}. By duality, it is equivalent to check this with quotient objects, where the maximal trivial quotient is simply the stalk at zero of the zeroth homology and hence is certainly defined over kk.

The fifth condition is simply the second condition with “subquotient” replaced with “subobject”. This follows again because summands are subobjects.∎

It is likely possible to prove the analogous theorem, without the “geometrically semisimple” conditions in the definitions of the key Tannakian categories, by a similar but more complicated argument. However, this additional level of generality is not needed for our paper, and so we did not pursue this.

Using the fact that Gk¯G_{\overline{k}} is a normal subgroup of GkG_{k}, we will show that the Galois action on our fiber functor normalizes the Tannakian monodromy.

Lemma 2.9.

Let AA be an abelian variety over a field kk. Let χ\chi be a character of π1et(Ak¯)\pi_{1}^{et}(A_{\overline{k}}) that is Gal(k¯|k)\operatorname{Gal}(\overline{k}|k)-invariant. Let χ\mathcal{L}_{\chi} be the associated character sheaf.

Let KK be a geometrically semisimple perverse sheaf on AA such that Hi(Ak¯,Kχ)H^{i}(A_{\overline{k}},K\otimes\mathcal{L}_{\chi}) vanishes for i0i\neq 0. Then the action of Gal(k¯|k)\operatorname{Gal}(\overline{k}|k) on H0(Ak¯,Kχ)H^{0}(A_{\overline{k}},K\otimes\mathcal{L}_{\chi}) normalizes the commutator subgroup of the identity component of the geometric convolution monodromy group of KK.

Proof.

We first prove that the action of Gal(k¯|k)\operatorname{Gal}(\overline{k}|k) normalizes the geometric convolution monodromy group of KK. For this, note that Gal(k¯|k)\operatorname{Gal}(\overline{k}|k) acts by automorphisms of the fiber functor Hci(Ak¯,Kχ)H^{i}_{c}(A_{\overline{k}},K\otimes\mathcal{L}_{\chi}) of the arithmetic Tannakian category, giving a homomorphism Gal(k¯|k)Gk\operatorname{Gal}(\overline{k}|k)\rightarrow G_{k}. Since the geometric convolution monodromy group is a normal subgroup of the arithmetic Tannakian group, it follows that Gal(k¯|k)\operatorname{Gal}(\overline{k}|k) normalizes the geometric convolution monodromy group.

It follows that Gal(k¯|k)\operatorname{Gal}(\overline{k}|k) also normalizes the commutator subgroup of the identity component, since the commutator subgroup of the identity component is a characteristic subgroup. ∎

3. The convolution monodromy group of a hypersurface

In this section, we fix an abelian variety AA of dimension nn and a smooth hypersurface HH in AA, which we will take except where noted to be defined over the complex numbers. Let i:HAi\colon H\to A be the inclusion. Let GHG_{H} be the convolution monodromy group of the perverse sheaf i[n1]i_{*}\mathbb{Q}[n-1]. The main goal of this section is to compute GHG_{H}.

To begin, we compute various Euler characteristics of HH - its arithmetic Euler characteristic, its topological Euler characteristic, and the Euler characteristics of the wedge powers of its cotangent sheaf. Using these, we will calculate the dimension, and Hodge numbers, of the cohomology of HH with coefficients in a rank-one lisse sheaf. These Hodge numbers will be used to compute the convolution monodromy group, and also used in later sections.

Lemma 3.1.

Let LL be a line bundle on AA. We have

χ(A,L)=c1()n/n!\chi(A,{L})=c_{1}(\mathcal{L})^{n}/{n!}
Proof.

By Hirzebruch-Riemann-Roch, the Euler characteristic of the coherent sheaf LL is the integral of its Chern character against the Todd class. By definition, the Chern character of LL is ec1[L]=k=0nc1(L)k/k!e^{c_{1}[L]}=\sum_{k=0}^{n}c_{1}(L)^{k}/k!. Because the tangent bundle of AA is trivial, its Todd class is 11. Integrating is equivalent to taking the degree nn term, which is c1(L)n/n!c_{1}({L})^{n}/{n!}. ∎

Lemma 3.2.

The arithmetic Euler characteristic χ(H,𝒪H)\chi(H,\mathcal{O}_{H}) of HH is (1)n1[H]n/n!(-1)^{n-1}[H]^{n}/n!.

Proof.

Using the exact sequence 0𝒪A(H)𝒪A𝒪H00\to\mathcal{O}_{A}(-H)\to\mathcal{O}_{A}\to\mathcal{O}_{H}\to 0, we observe that

χ(H,𝒪H)=χ(A,𝒪A)χ(A,𝒪A(H))=0([H])n/n!=(1)n1[H]n/n!\chi(H,\mathcal{O}_{H})=\chi(A,\mathcal{O}_{A})-\chi(A,\mathcal{O}_{A}(-H))=0-(-[H])^{n}/n!=(-1)^{n-1}[H]^{n}/n!

by Lemma 3.1. ∎

Lemma 3.3.

The topological Euler characteristic of HH is (1)n1[H]n(-1)^{n-1}[H]^{n}.

Proof.

The topological Euler characteristic of HH is the top Chern class of the tangent bundle of HH. Using the exact sequence 0𝒪([H])ΩA1ΩH100\to\mathcal{O}(-[H])\to\Omega^{1}_{A}\to\Omega^{1}_{H}\to 0, and the fact that all Chern classes of ΩA1\Omega^{1}_{A} vanish, we see that the top Chern class of ΩH1\Omega^{1}_{H} is (1)n1[H]n(-1)^{n-1}[H]^{n}. ∎

Motivated by Lemma 3.2, we define the degree dd of HH to be [H]nn!\frac{[H]^{n}}{n!}, which is always a positive integer for HH an ample hypersurface.

The Hodge numbers of HH can be computed in terms of Eulerian numbers. For a general reference on Eulerian numbers in combinatorics, see [52, Chap. 1].

Lemma 3.4.

We have

(2) χ(H,ΩHi)=(1)n1idA(n,i)\chi(H,\Omega^{i}_{H})=(-1)^{n-1-i}dA(n,i)

where A(n,i)A(n,i) is the Eulerian number.

Proof.

By [68, p. 272, Some recursion formulas], we have

χ(H,ΩHi)=(1)nj=0i(nij)(1)j(jn(j+1)n)χ()\chi(H,\Omega^{i}_{H})=(-1)^{n}\sum_{j=0}^{i}{n\choose i-j}(-1)^{j}(j^{n}-(j+1)^{n})\chi(\mathcal{L})
=(1)nj=0i+1(1)j((nij)+(nij+1))jnχ()=(1)nj=0i+1(1)j(n+1i+1j)jnχ()=(-1)^{n}\sum_{j=0}^{i+1}(-1)^{j}\left({n\choose i-j}+{n\choose i-j+1}\right)j^{n}\chi(\mathcal{L})=(-1)^{n}\sum_{j=0}^{i+1}(-1)^{j}{n+1\choose i+1-j}j^{n}\chi(\mathcal{L})

We now use the combinatorial identity ([52, Cor. 1.3])

(1)nj=0i+1(1)j(n+1i+1j)jn=(1)n1iA(n,i)(-1)^{n}\sum_{j=0}^{i+1}(-1)^{j}{n+1\choose i+1-j}j^{n}=(-1)^{n-1-i}A(n,i)

and Lemma 3.1

χ()=d\chi(\mathcal{L})=d

to derive (2). ∎

Let N=(n!)d=[H]nN=(n!)d=[H]^{n}.

Recall from the introduction that a(i)a(i) is the sequence

1,5,20,76,285,1065,1,5,20,76,285,1065,\dots

satisfying

a(1)=1,a(2)=5,a(i+2)=4a(i+1)+1a(i)a(1)=1,a(2)=5,a(i+2)=4a(i+1)+1-a(i)

(OEIS A061278).

Theorem 3.5.

Assume that HH is not equal to the translate of HH by any nontrivial point of AA and n>2n>2. Assume also that neither of the following holds:

  1. (1)

    n=2n=2 and d=28d=28.

  2. (2)

    n=3n=3 and d=(a(i)+a(i+1)a(i))/6d={a(i)+a(i+1)\choose a(i)}/6 for some i2i\geq 2.

Then GHG_{H} contains as a normal subgroup either SLN,SpNSL_{N},Sp_{N}, or SONSO_{N}. If HH is not equal to any translate of [1]H[-1]^{*}H then the SLNSL_{N} case holds. If HH is equal to such a translate, then if nn is even SpNSp_{N} holds and if nn is odd SONSO_{N} holds.

Remark 3.6.

In the exceptional cases, there are only a few other possibilities for GHG_{H}. Suppose GHG_{H} does not contain SLNSL_{N}, SpNSp_{N}, or SONSO_{N}. Then in case (1), GHG_{H} contains E7E_{7} acting by its 5656-dimensional representation. In (2), GHG_{H} contains as a normal subgroup SLa(i)+a(i+1)SL_{a(i)+a(i+1)} acting by the representation a(i)\wedge^{a(i)}.

The proof occupies the remainder of this section.

We will call the representation of GHG_{H} associated to the object i[n1]i_{*}\mathbb{Q}[n-1] the distinguished representation.

Lemma 3.7.

The dimension of the distinguished representation of GHG_{H} is NN.

Proof.

By construction, the dimension of the representation associated to any object in the Tannakian category is the Euler characteristic of the corresponding perverse sheaf, which is (1)n1(-1)^{n-1} times the topological Euler characteristic of HH. This now follows from Lemma 3.3.∎

Lemma 3.8.

Assume that HH is not equal to the translate of HH by any nontrivial point of AA. Then the distinguished representation of GHG_{H} is an irreducible minuscule representation of the Lie algebra of HH.

Proof.

Because HH is smooth, the characteristic cycle of i[n]i_{*}\mathbb{Q}[n] is simply the conormal bundle of HH with multiplicity 11, hence has a single irreducible component, with multiplicity one. The representation is then irreducible minuscule by [42, Corollary 1.0]

Lemma 3.9.

Assume HH is not translation-invariant by any nontrivial point of AA and n>2n>2. Then the identity component of GHG_{H} is simple modulo its center.

Proof.

We follow the argument of [43, Theorem 6.2.1], with minor modifications. That theorem is not directly applicable because it assumes that HH is symmetric (i.e. stable under inversion), which is not necessarily true here. Thus we restate the proof, which we can also simplify somewhat because our assumption that HH is smooth is stronger than the analogous assumption in [43]. First, we review some notation and terminology from [43].

The proof relies on the notion of the characteristic cycle of a perverse sheaf. Classically, the characteristic cycle of a perverse sheaf on a variety XX of dimension nn is an effective Lagrangian cycle in the cotangent bundle TXT^{*}X of XX. In other words, it is a nonnegative-integer-weighted sum of irreducible nn-dimensional subvarieties of TXT^{*}X whose tangent space at a generic point is isotropic for the natural symplectic form on TXT^{*}X. Any such subvariety is automatically the conormal bundle to an irreducible subvariety of XX, of arbitrary dimension, i.e. its support.

For AA an abelian variety, because TAT^{*}A is a trivial bundle, we can express it as a product A×H0(A,ΩA1)A\times H^{0}(A,\Omega^{1}_{A}). The projection onto the second factor is called the Gauss map. Krämer considers an irreducible component negligible if its image under the Gauss map is not dense, and a cycle clean if none of its components are negligible [43, Definition 1.2.2]. He defines cc(K)cc(K) for a perverse sheaf KK to be the usual characteristic cycle but ignoring any negligible components, making it automatically clean [43, Definition 2.1.1].

The degree of a cycle is the degree of the Gauss map restricted to that cycle. It is manifestly a sum over components of the degree of the Gauss map on that component, which vanishes if and only if the component is negligible.

A clean cycle is determined by its restriction to any open set UH0(A,ΩA1)U\subseteq H^{0}(A,\Omega^{1}_{A}). In particular, given two such cycles Λ1,Λ2\Lambda_{1},\Lambda_{2}, because

dimΛ1=dimΛ2=n=dimH0(A,ΩA1),\dim\Lambda_{1}=\dim\Lambda_{2}=n=\dim H^{0}(A,\Omega^{1}_{A}),

one can find an open set over which both Λ1\Lambda_{1} and Λ2\Lambda_{2} are finite. The fiber product Λ1×UΛ2\Lambda_{1}\times_{U}\Lambda_{2} then maps to UU by the obvious projection and to AA by composing the two projections Λ1A,Λ2A\Lambda_{1}\to A,\Lambda_{2}\to A with the multiplication map A×AAA\times A\to A. Hence Λ1×UΛ2\Lambda_{1}\times_{U}\Lambda_{2} maps to A×UA\times U. Its image has a unique clean extension to A×H0(A,ΩA1)A\times H^{0}(A,\Omega^{1}_{A}). Krämer defines this Λ1Λ2\Lambda_{1}\circ\Lambda_{2} to be this extension [43, Example 1.3.2].

A key property of this convolution product is that deg(Λ1Λ2)=deg(Λ1)deg(Λ2)\deg(\Lambda_{1}\circ\Lambda_{2})=\deg(\Lambda_{1})\deg(\Lambda_{2}); as a consequence, if Λ1,Λ2\Lambda_{1},\Lambda_{2} are clean and nonzero, then Λ1Λ2\Lambda_{1}\circ\Lambda_{2} is nonzero as well.

We are now ready to begin the argument.

Assume for contradiction that the identity component of GHG_{H} is not simple modulo its center. Then its Lie algebra is not simple modulo its center. By [43, Proposition 6.1.1], it follows from this that there exist mm\in\mathbb{N} and effective clean cycles Λ1,Λ2\Lambda_{1},\Lambda_{2} on AA, with deg(Λi)>1\deg(\Lambda_{i})>1, such that

[m]cc(i[n])=Λ1Λ2[m]_{*}cc(i_{*}\mathbb{Q}[n])=\Lambda_{1}\circ\Lambda_{2}

where [m][m] is the multiplication-by-nn map. Because HH is smooth, the characteristic cycle of i[n]i_{*}\mathbb{Q}[n] is simply the conormal bundle ΛH\Lambda_{H} of HH, which is irreducible. Its degree is dn!d\cdot n!, because the degree of the Gauss map of the conormal bundle to HH is the sum of the multiplicities of vanishing of a general 11-form on HH, which is the Euler characteristic of HH, which is dn!d\cdot n!. In particular, this degree is nonzero, so cc(i[n])=ΛHcc(i_{*}\mathbb{Q}[n])=\Lambda_{H}.

Because HH is not translation-invariant, the map from HH to its image under [m][m] is generically one-to-one, and so [m]ΛH[m]_{*}\Lambda_{H} is an irreducible cycle with multiplicity one in the cotangent bundle of AA. This implies Λ1\Lambda_{1} and Λ2\Lambda_{2} are irreducible: If not, say if Λ1=Λ1a+Λ1b\Lambda_{1}=\Lambda_{1}^{a}+\Lambda_{1}^{b}, we would have

ΛH=Λ1Λ2=Λ1aΛ2+Λ1bΛ2\Lambda_{H}=\Lambda_{1}\circ\Lambda_{2}=\Lambda_{1}^{a}\circ\Lambda_{2}+\Lambda_{1}^{b}\circ\Lambda_{2}

with both Λ1aΛ2\Lambda_{1}^{a}\circ\Lambda_{2} and Λ1bΛ2\Lambda_{1}^{b}\circ\Lambda_{2} nonzero, contradicting irreducibility.

It follows that Λ1\Lambda_{1} and Λ2\Lambda_{2} must be the conormal bundles ΛZ1\Lambda_{Z_{1}} and ΛZ2\Lambda_{Z_{2}} of varieties Z1Z_{1} and Z2Z_{2}. Because deg(Λi)>1\deg(\Lambda_{i})>1, neither Z1Z_{1} nor Z2Z_{2} can be a point, as the conormal bundle to a point is simply an affine space, and its Gauss map is an isomorphism, and thus has degree 11.

Let YY be the image of HH under [m][m]. By [43, Lemma 5.2.2] there is a dominant rational map from YY to Z1Z_{1} (say), and thus a dominant rational map from HH to Z1Z_{1}. Because HH is smooth and Z1Z_{1} is a subvariety of an abelian variety, this dominant rational map automatically extends to a surjective morphism [12, Theorem 4.4.1]. Moreover, by the Lefschetz hyperplane theorem (since n>2n>2), AA is the Albanese of HH, so the surjective morphism f1:HZ1Af_{1}\colon H\to Z_{1}\subseteq A extends to a homomorphism g1:AAg_{1}\colon A\to A, giving a commutative diagram

H{H}Z1{Z_{1}}A{A}A{A}f1\scriptstyle{f_{1}}i\scriptstyle{i}g1\scriptstyle{g_{1}}

Let B1B_{1} be the image of g1g_{1}. Because f1f_{1} is surjective, Z1B1Z_{1}\subseteq B_{1}.

If Z1=B1Z_{1}=B_{1} then Z1Z_{1} is an abelian variety, and the conormal bundle to any nontrivial abelian subvariety of AA has Gauss map of degree 0, contradicting deg(Λi)>1\deg(\Lambda_{i})>1. Otherwise, by commutativity of the diagram, Hg1Z1H\subseteq g_{1}^{*}Z_{1}. Because HH is a hypersurface, it is a maximal proper subvariety of AA, so H=g1Z1H=g_{1}^{*}Z_{1}. This contradicts ampleness of HH unless g1g_{1} is finite, and contradicts HH not being translation-invariant unless gig_{i} is an isomorphism. This means Z1Z_{1} and HH are isomorphic as subvarieties of an abelian variety. Thus deg(ΛZ1)=deg(TZ1Z1)=deg(THH)\deg(\Lambda_{Z_{1}})=\deg(T^{*}_{Z_{1}}Z_{1})=\deg(T^{*}_{H}H) and so because deg(ΛZ1)deg(ΛZ2)=deg(THH)\deg(\Lambda_{Z_{1}})\deg(\Lambda_{Z_{2}})=\deg(T^{*}_{H}H), we have deg(ΛZ2=1)\deg(\Lambda_{Z_{2}}=1), contradicting deg(Λi)>1\deg(\Lambda_{i})>1.

Because we have a contradiction in every case, we have shown that GG is simple modulo its center.∎

Lemma 3.10.

Assume that HH is not equal to the translate of HH by any nontrivial point of AA. Then the commutator subgroup of the identity component of GHG_{H}, together with its distinguished representation, is one of the following:

  1. (1)

    SLN,SpNSL_{N},Sp_{N} or SONSO_{N}, with its standard representation distinguished.

  2. (2)

    SOmSO_{m} with one of its spin representations distinguished, or E6E_{6} or E7E_{7} with one of its lowest-dimensional nontrivial representations distingusihed.

  3. (3)

    SLmSL_{m} with the representation k\wedge^{k} distinguished for some 2km/22\leq k\leq m/2.

Proof.

It follows from the Lemma 3.9 that the commutator subgroup of the identity component of GHG_{H} is a simple Lie group. Furthermore, from Lemma 3.8 its distinguished representation must be irreducible and minuscule. But the above is an exhaustive list of minuscule representations of simple Lie groups (see e.g. [42, p. 7]).∎

Lemma 3.11.

Assume that GHG_{H} contains as a normal subgroup one of SLNSL_{N}, SpNSp_{N}, or SONSO_{N}. Then it contains SLNSL_{N} only if HH is not equal to any translate of [1]H[-1]^{*}H, it contains SpNSp_{N} only if nn is even, and it contains SONSO_{N} only if nn is odd.

Proof.

Note first that the distinguished representation of any subgroup of GLNGL_{N} which contains SpNSp_{N} or SONSO_{N} as a normal subgroup is equal to the tensor product of its dual representation with a one-dimensional representation, since it is contained in the normalizer GSpNGSp_{N} or GONGO_{N} respectively. Conversely, if N>2N>2 then the distinguished representation of any subgroup of GLNGL_{N} which contains SLNSL_{N} as a normal subgroup is not equal to the tensor product of its dual representation with any one-dimensional representation.

Translating into the language of the Tannakian category, we see that under this assumption, GHG_{H} contains SLNSL_{N} as a normal subgroup if, and only if, the perverse sheaf i[n1]i_{*}\mathbb{Q}[n-1] is not isomorphic, up to negligible factors, to the convolution of its dual [1]Di[n1]=[1]i[n1][-1]^{*}Di_{*}\mathbb{Q}[n-1]=[-1]^{*}i_{*}\mathbb{Q}[n-1] with any perverse sheaf corresponding to a one-dimensional representation. Now perverse sheaves corresponding to a one-dimensional representation are always skyscraper sheaves [44, Proposition 10.1], and convolution with a skyscraper sheaf is equivalent to translation, so it is equivalent to say that i[n1]i_{*}\mathbb{Q}[n-1] is not isomorphic, up to negligible factors, to any translate of [1]i[n1][-1]^{*}i_{*}\mathbb{Q}[n-1]. Because i[n1]i_{*}\mathbb{Q}[n-1] and [1]i[n1][-1]^{*}i_{*}\mathbb{Q}[n-1] are both irreducible perverse sheaves, there can be no negligible factors, and so this happens if and only if they are isomorphic (up to translation). Because i[n1]i_{*}\mathbb{Q}[n-1] and [1]i[n1][-1]^{*}i_{*}\mathbb{Q}[n-1] are each constant sheaves on their support, they are isomorphic (up to translation) if and only if their supports are equal (up to translation), which happens exactly when HH is equal to a translate of [1]H[-1]^{*}H. This handles the SLNSL_{N} case.

The argument to distinguish SpNSp_{N} and SONSO_{N} is identical to [45, Lemma 2.1], which is stated only in the case where HH is a theta divisor, but the assumption is never used in the proof, except that they write g!g! instead of NN since, for a theta-divisor, N=g!N=g!. ∎

To prove the main theorem, it remains to give a complete list of n,dn,d for which GHG_{H} can contain as a normal subgroup one of the groups in Lemma 3.10, cases (2) or (3).

Case (2) is relatively easy as for these groups the dimensions have a special form.

Lemma 3.12.

If n>2n>2, we cannot have the commutator subgroup of the identity component of GHG_{H} be SOmSO_{m} with a spin representation distinguished, or E6E_{6} or E7E_{7} with one of their lowest-dimensional nontrivial irreducible representations distinguished.

Proof.

For n>2n>2, (n!)d(n!)d is always a multiple of 3!=63!=6. However, the stated representations cannot have dimension a multiple of 66, contradicting Lemma 3.7. Indeed, the spin representations have dimension a power of 22, while the lowest-dimensional representations of E6E_{6} and E7E_{7} have dimension 2727 and 5656 respectively, and none of these are multiples of 66. ∎

The remainder of the section is devoted to restricting the case of wedge powers. We will obtain further numerical obstructions by introducing a Hodge torus into our convolution monodromy group. The action of the Hodge torus is obtained using the Galois action on Hn1(Xκ¯,χ)H^{n-1}(X_{\overline{\kappa}},\mathcal{L}_{\chi}) for a field κ\kappa and pp-adic Hodge theory, and thus relies on Lemma 2.9 and our earlier construction of a Tannakian category of perverse sheaves over a non-algebraically closed field, as well as a calculation of Hodge-Tate weights.

An alternate approach should be possible, using classical Hodge theory and a Tannakian category of mixed Hodge modules on AA_{\mathbb{C}}, as suggested by [44, Example 5.2], but we did not take this approach as we found the Galois action useful elsewhere in the argument.

We first review some pp-adic Hodge theory. For VV a representation of the Galois group of a pp-adic field KK with coefficients in KK, p\mathbb{C}_{p} the completion of the algebraic closure of KK, and p(q)\mathbb{C}_{p}(q) the Tate twist by qq, the Hodge-Tate weights of VV are the integers qq such that (VKp)Gal(K¯/K)0(V\otimes_{K}\mathbb{C}_{p})^{\operatorname{Gal}(\overline{K}/K)}\neq 0, and the multiplicity of the Hodge-Tate weight qq is the dimension of (Vp¯p)Gal(K¯/K)(V\otimes_{\overline{\mathbb{Q}_{p}}}\mathbb{C}_{p})^{\operatorname{Gal}(\overline{K}/K)} over KK.

We say VV is a Hodge-Tate representation if the sum of the Hodge-Tate weights is dimV\dim V. The map DD sending VV to q(VKp)Gal(K¯/K)\bigoplus_{q\in\mathbb{Z}}(V\otimes_{K}\mathbb{C}_{p})^{\operatorname{Gal}(\overline{K}/K)} gives a functor from the category of Hodge-Tate representations to graded vector spaces. The category of Hodge-Tate representations is stable under direct sums, tensor products, duals, subobjects, and quotients, and DD is a faithful exact tensor functor. It follows that the category of Hodge-Tate representations is Tannakian, and DD corresponds to a map from the Tannakian group 𝔾m\mathbb{G}_{m} of the category of graded vector spaces to the Tannakian group of the category of Hodge-Tate representations.

Lemma 3.13.

Let HH be a smooth hypersurface, defined over a pp-adic field kk, in an abelian variety AA of dimension nn. Let χ:π1(Ak¯)p¯×\chi\colon\pi_{1}(A_{\overline{k}})\to\overline{\mathbb{Q}_{p}}^{\times} be a finite-order character such that Hi(Hk¯,χ)=0H^{i}(H_{\overline{k}},\mathcal{L}_{\chi})=0 for in1i\neq n-1. Let kk^{\prime} be a finite extension of kk containing the coefficient field and field of definition of χ\chi.

Then the Hodge-Tate weights of the Gal(k¯|k)\operatorname{Gal}(\overline{k}^{\prime}|k^{\prime}) action on Hn1(Hk¯,χ)H^{n-1}(H_{\overline{k}^{\prime}},\mathcal{L}_{\chi}) are 0,,n10,\dots,n-1, where the multiplicity of the weight qq is dA(n,q)dA(n,q).

Proof.

As a finite-order character of π1(A)\pi_{1}(A), χ\chi factors through A[m]A[m] for some mm. Let HH^{\prime} be the inverse image of HH under the multiplication-by-mm map of AA. Then we can express H(Hk¯,χ)H^{*}(H_{\overline{k}},\mathcal{L}_{\chi}) as the part of the étale cohomology H(Hk¯,p)H^{*}(H^{\prime}_{\overline{k}},\mathbb{Q}_{p}) where A[m]A[m] acts by the character χ\chi. Applying pp-adic Hodge theory, we see that the dimension of the Hodge-Tate weight qq subspace of Hp+q(Hk¯,χ)H^{p+q}(H_{\overline{k}},\mathcal{L}_{\chi}) is equal to the dimension of the subspace of Hp(Hk,ΩHq)H^{p}(H^{\prime}_{k^{\prime}},\Omega^{q}_{H^{\prime}}) on which A[m]A[m] acts by the character χ\chi. By descent, this is the dimension of Hp(Hk,ΩHqL)H^{p}(H_{k^{\prime}},\Omega^{q}_{H}\otimes L) for a torsion line bundle LL on HH. Because Hp+q(Hk¯,χ)H^{p+q}(H_{\overline{k}^{\prime}},\mathcal{L}_{\chi}) vanishes for pn1qp\neq n-1-q, the dimension of Hp(H,ΩHqL)H^{p}(H,\Omega^{q}_{H}\otimes L) for p=n1qp=n-1-q is equal to (1)n1q(-1)^{n-1-q} times the Euler characteristic χ(Hk,ΩHqL)=χ(H,ΩHq)\chi(H_{k^{\prime}},\Omega^{q}_{H}\otimes L)=\chi(H,\Omega^{q}_{H}) because, as LL is torsion, its Chern class vanishes. By Lemma 3.4, this Euler characteristic is (1)n1qdA(n,q)(-1)^{n-1-q}dA(n,q), so the dimension and Hodge-Tate multiplicity are both dA(n,q)dA(n,q) . ∎

Lemma 3.14.

Let HH be a hypersurface on on abelian variety AA of dimension nn and let d=[H]n/n!d=[H]^{n}/n!. Suppose that the commutator subgroup of the identity component of GHG_{H} is SLmSL_{m}, with distinguished representation k\wedge^{k}. Then there exists a function mHm_{H} from the integers to the natural numbers and an integer ss such that imH(i)=m\sum_{i}m_{H}(i)=m and such that, for all qq\in\mathbb{Z},

(3) mS:0mS(i)mH(i)imS(i)=kiimS(i)=s+qi(mH(i)mS(i))=dA(n,q).\sum_{\begin{subarray}{c}m_{S}\colon\mathbb{Z}\to\mathbb{Z}\\ 0\leq m_{S}(i)\leq m_{H}(i)\\ \sum_{i}m_{S}(i)=k\\ \sum_{i}im_{S}(i)=s+q\end{subarray}}\prod_{i}{m_{H}(i)\choose m_{S}(i)}=dA(n,q).

Here we use the convention that A(n,q)=0A(n,q)=0 unless 0q<n0\leq q<n.

Proof.

Fix a finite-order character χ\chi of the geometric fundamental group of AA such that the cohomology of HH with coefficients in χ\mathcal{L}_{\chi} is concentrated in degree 0. Let κ\kappa be a finitely generated field over which A,HA,H, and χ\chi may all be defined.

By our assumption on the convolution monodromy group GHG_{H} and Lemma 2.9(2), it follows that the Gal(κ¯/κ)\operatorname{Gal}(\overline{\kappa}/\kappa) action on Hn1(Xk¯,χ)H^{n-1}(X_{\overline{k}^{\prime}},\mathcal{L}_{\chi}) lies in the normalizer of SLmSL_{m} inside the general linear group of the representation k\wedge^{k}. It suffices to check that this condition implies the existence of mHm_{H} and ss. For this we may spread out HH to a family over a variety defined over a number field and specialize to a closed point. Since the property that the image of Galois is contained in the normalizer is preserved under specialization, we may assume κ\kappa is a number field.

This normalizer of SLmSL_{m} inside the general linear group is GLm/μkGL_{m}/\mu_{k} if km/2k\neq m/2 and GLm/μk(/2)GL_{m}/\mu_{k}\ltimes(\mathbb{Z}/2) if k=m/2k=m/2.

The category of Galois representations generated by Hn1(Xκ¯,χ)H^{n-1}(X_{\overline{\kappa}},\mathcal{L}_{\chi}) under tensor products, direct sums, subobjects, quotients, and duals is a Tannakian category isomorphic to the category of representations of the Zariski closure of the image of Gal(κ¯|κ)\operatorname{Gal}(\overline{\kappa}|\kappa) acting on Hn1(Xκ¯,χ)H^{n-1}(X_{\overline{\kappa}},\mathcal{L}_{\chi}) (a quotient of the group Galκ\operatorname{Gal}_{\kappa} defined in the previous section). Because Hn1(Xκ¯,χ)H^{n-1}(X_{\overline{\kappa}},\mathcal{L}_{\chi}) is Hodge-Tate, this category is a Tannakian subcategory of the category of Hodge-Tate representations, so the group is a quotient of the Tannakian group of the category of Hodge-Tate representations. Thus, the map from 𝔾m\mathbb{G}_{m} constructed via the associated graded vector space functor gives a map from 𝔾m\mathbb{G}_{m} to the Zariski closure of the image of Gal(κ¯|κ)\operatorname{Gal}(\overline{\kappa}|\kappa), whose weights are the Hodge-Tate weights of Hn1(Xκ¯,χ)H^{n-1}(X_{\overline{\kappa}},\mathcal{L}_{\chi}).

Thus we have a homomorphism from 𝔾m\mathbb{G}_{m} to GLm/μkGL_{m}/\mu_{k} (or its semidirect product with /2\mathbb{Z}/2). Because 𝔾m\mathbb{G}_{m} is connected, this defines a homomorphism 𝔾mGLm/μk\mathbb{G}_{m}\to GL_{m}/\mu_{k}. Its weight multiplicities on the representation k\wedge^{k} must equal the Hodge-Tate multiplicities of Hn1(Xκ¯,χ)H^{n-1}(X_{\overline{\kappa}},\mathcal{L}_{\chi}).

Homomorphisms 𝔾mGLm\mathbb{G}_{m}\to GL_{m} are parameterized by their weights, an mm-tuple of integers with w1wmw_{1}\leq\dots\leq w_{m}. Any homomorphism α:𝔾mGLm/μk\alpha\colon\mathbb{G}_{m}\to GL_{m}/\mu_{k} has a unique lift α¯:𝔾mGLm\overline{\alpha}\colon\mathbb{G}_{m}\to GL_{m} forming a commutative diagram

𝔾m{\mathbb{G}_{m}}GLm{GL_{m}}𝔾m{\mathbb{G}_{m}}GLm/μk{GL_{m}/\mu_{k}}xxk\scriptstyle{x\mapsto x^{k}}α¯\scriptstyle{\overline{\alpha}}α\scriptstyle{\alpha}

If we let w¯1w¯m\overline{w}_{1}\leq\dots\leq\overline{w}_{m} be the weights of α¯\overline{\alpha}, then w¯iw¯jmodk\overline{w}_{i}\equiv\overline{w}_{j}\mod k since their difference is a weight of the adjoint representation, which factors through GLm/μkGL_{m}/\mu_{k} and thus factors through xxkx\mapsto x^{k}.

Thus w¯1k,,w¯mk\frac{\overline{w}_{1}}{k},\dots,\frac{\overline{w}_{m}}{k} are rational numbers with the same fractional part, equal to the “weights” of the standard representation of GLmGL_{m}. We let wi=w¯ikw_{i}=\lfloor\frac{\overline{w}_{i}}{k}\rfloor and let s=kwiw¯is=kw_{i}-\overline{w}_{i}, which is independent of ii, so that w¯ik=wisk\frac{\overline{w}_{i}}{k}=w_{i}-\frac{s}{k} .

Then the weights of the representation k\wedge^{k} are exactly (iSwi)s(\sum_{i\in S}w_{i})-s for all S{1,m}S\subseteq\{1,\dots m\} with |S|=k|S|=k.

It follows that the multiplicity of the weight qq inside k\wedge^{k} of the standard representation of GLmGL_{m} is the number of subsets S{1,,m}S\subseteq\{1,\dots,m\} with |S|=k|S|=k and iSwi=s+q\sum_{i\in S}w_{i}=s+q.

To calculate this, let mH(j)m_{H}(j) be the number of ii such that wi=jw_{i}=j. For S{1,,m}S\subseteq\{1,\dots,m\}, let mS(j)m_{S}(j) be the number of iSi\in S such that wi=jw_{i}=j. Then |S|=jmS(j)|S|=\sum_{j}m_{S}(j) and iSwi=jjmS(j)\sum_{i\in S}w_{i}=\sum_{j}jm_{S}(j). Furthermore the number of sets SS attaining a given function mSm_{S} is i(mH(i)mS(i))\prod_{i}{m_{H}(i)\choose m_{S}(i)}.

The stated identities then follow from Lemma 3.13.

To complete the argument, we have the following purely combinatorial proposition, which is proven in Appendix B.

Proposition 3.15 (Appendix B).

Suppose that there exists a natural number kk, function mHm_{H} from the integers to the natural numbers and an integer ss such that 1<k<1+imH(i)1<k<-1+\sum_{i}m_{H}(i) and Equation (3) is satisfied for all qq\in\mathbb{Z}. Then we have one of the cases

  1. (1)

    m=4m=4 and k=2k=2,

  2. (2)

    n=2n=2 and d=(2k1k)d={2k-1\choose k} for some k>2k>2, or

  3. (3)

    n=3n=3 and d=(a(i)+a(i+1)a(i))/6d={a(i)+a(i+1)\choose a(i)}/6 for some i2i\geq 2. (Here a(i)a(i) is defined by

    a(1)=1,a(2)=5,a(i+2)=4a(i+1)+1a(i).)a(1)=1,a(2)=5,a(i+2)=4a(i+1)+1-a(i).)

We are now ready to prove the main theorem of this section:

Proof of Theorem 3.5.

The commutator subgroup of the identity component of GHG_{H} must be one of the groups listed in the three cases of Lemma 3.10. In case (1), we know immediately that GHG_{H} contains as a normal subgroup one of SLNSL_{N}, SpNSp_{N} or SONSO_{N}, and we conclude by Lemma 3.11.

We rule out case (2) by Lemma 3.12 and case (3) by Proposition 3.15. Note that Proposition 3.15 does not rule out the case m=4,k=2m=4,k=2, but 2SL4\wedge^{2}SL_{4} is SO6SO_{6} so GHG_{H} contains SO6SO_{6} as a normal subgroup in this case. ∎

4. Big monodromy from big convolution monodromy

Let XX be a variety, AA an abelian variety, and YX×AY\subseteq X\times A a family over XX of smooth hypersurfaces in AA, with XX, AA, YY all defined over \mathbb{C}. Let n=dimAn=\dim A, so that n1n-1 is the relative dimension of YY over XX. Let η\eta be the generic point of XX and η¯\overline{\eta} a geometric generic point.

Let i:Yη¯Aη¯i\colon Y_{\overline{\eta}}\to A_{\overline{\eta}} be the inclusion, and let KK be the perverse sheaf K=i¯p[n1]K=i_{*}\overline{\mathbb{Q}}_{p}[n-1] on Aη¯A_{\overline{\eta}}. Let GG be the convolution monodromy group of KK. Let GG^{*} be the commutator subgroup of the identity component of GG. We continue to call the distinguished representation of GG the representation arising from the object KK, and let its restriction to GG^{*} be the distinguished representation of GG^{*}.

Lemma 4.1.

Assume that GG^{*} is a simple algebraic group with irreducible distinguished representation, and that YY is not equal to a constant family of hypersurfaces translated by a section of AA.

Let KK^{\prime} be an irreducible perverse sheaf on Aη¯A_{\overline{\eta}} in the Tannakian category generated by KK. Assume that KK^{\prime} is a pullback from AA_{\mathbb{C}} to Aη¯A_{\overline{\eta}} of a perverse sheaf on AA_{\mathbb{C}}. Then GG^{*} acts trivially on the irreducible representation of GG corresponding to KK^{\prime}.

Proof.

Let η×η¯\overline{\eta\times\eta} be a geometric generic point of X×XX\times X. Let pr1,pr2:Aη×η¯Aη¯pr_{1},pr_{2}\colon A_{\overline{\eta\times\eta}}\to A_{\overline{\eta}} be the maps induced by the two projections η×η¯η¯\overline{\eta\times\eta}\to\overline{\eta}. The convolution monodromy groups of pr1Kpr_{1}^{*}K and pr2Kpr_{2}^{*}K are both isomorphic to GG, so the convolution monodromy group of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K is isomorphic to a subgroup of G×GG\times G whose projection onto each factor is surjective. By Goursat’s lemma, there are normal subgroups H1,H2H_{1},H_{2} in GG and an isomorphism a:(G/H1)(G/H2)a\colon(G/H_{1})\to(G/H_{2}) such that the convolution monodromy group of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K is isomorphic to

{(g1,g2)G×G|a(g1)=g2modH2}.\{(g_{1},g_{2})\in G\times G|a(g_{1})=g_{2}\mod H_{2}\}.

Note that GG has a unique factor in its Jordan-Hölder decomposition which is a nonabelian connected simple group. Hence this factor appears either in H1H_{1} and H2H_{2} or in G/H1G/H_{1} and G/H2G/H_{2}.

In the first case, we must have GH1G^{*}\subseteq H_{1} and GH2G^{*}\subseteq H_{2}. This is because, if the nonabelian connected simple factor appears in HiH_{i}, then it must appear in HiGH_{i}\cap G^{*} as Hi/(HiG)G/GH_{i}/(H_{i}\cap G^{*})\subseteq G/G^{*} which, modulo scalars, is contained in the outer automorphism group of GG^{*} and thus is virtually abelian and cannot contain a nonabelian connected simple factor. Furthermore HiGH_{i}\cap G^{*} is a normal subgroup of GG^{*}, and since it cannot be a finite group, it must be GG^{*}.

Now, using the fact that GH1G^{*}\subseteq H_{1} and GH2G^{*}\subseteq H_{2}, we will show that GG^{*} acts trivially on the irreducible representation corresponding to KK^{\prime}. To do this, observe that pr1Kpr_{1}^{*}K^{\prime} and pr2Kpr_{2}^{*}K^{\prime} are isomorphic because KK^{\prime} is a pullback from AA_{\mathbb{C}}. These correspond to two representations of the convolution monodromy group of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K that factor through the projection onto the first and second factors respectively. Because GG^{*} lies in H1H_{1} and H2H_{2}, the convolution monodromy group of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K contains two copies of GG^{*}. The first copy of GG^{*} acts trivially on pr2Kpr_{2}^{*}K^{\prime}, so it must act trivially on pr1Kpr_{1}^{*}K^{\prime}, so GG^{*} acts trivially on KK^{\prime}, as desired.

In the second case, H1H_{1} and H2H_{2} must both be contained in the scalars. To see this, because the scalars are the centralizers of GG^{*}, it suffices to show that the image of HiH_{i} in the automorphisms of GG^{*} vanishes. Equivalently, we must show that the image of HiH_{i} in the automorphisms of the Lie algebra of GG^{*} vanishes. This automorphism group is an extension of the finite outer automorphism group of GG^{*} by GG^{*} mod its center. Because the image of HiH_{i} in the automorphism group is normalized by GG^{*}, it either contains GG^{*} or is finite, and it cannot contain GG^{*}, so it is finite. Because it is finite and normalized by GG^{*}, it commutes with GG^{*}. Because the Lie algebra of GG^{*} is an irreducible representation of GG^{*}, this forces the image of HiH_{i} to act as scalars. But there are no nontrivial scalar automorphisms of a nonabelian Lie algebra, as they would never preserve any equation [x,y]=z[x,y]=z, and so the image of HiH_{i} is trivial, as desired.

Now, using the fact that H1H_{1} and H2H_{2} are both contained in the scalars, we will derive a contradiction. Thus the convolution monodromy group pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K is contained in the set {(g1,g2)G2|a(g1)=g2}\{(g_{1},g_{2})\in G^{2}|a(g_{1})=g_{2}\} for some automorphism aa of GG mod scalars. Let us first check that the automorphism aa is inner. To do this, let η¯m\overline{\eta}_{m} be the geometric generic point of XmX^{m}. Let pr1,,prmpr_{1},\dots,pr_{m} be the projections to η\eta. For each ii we have a homomorphism ρi\rho_{i} from the convolution monodromy group of i=1mpriK\bigoplus_{i=1}^{m}pr_{i}^{*}K to the convolution monodromy group of priKpr_{i}^{*}K modulo scalars, and let ρ1\rho^{1} and ρ2\rho^{2} be the similarly-defined homomorphisms from the convolution monodromy group of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K to the convolution monodromy groups of pr1Kpr_{1}^{*}K and pr2Kpr_{2}^{*}K modulo scalars. Then because for any pair i,ji,j there is a projection XmX2X^{m}\to X^{2} onto the iith and jjth copies of XX, and priKprjKpr_{i}^{*}K\oplus pr_{j}^{*}K is isomorphic to the pullback of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K along this projection, the convolution monodromy group of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K is isomorphic to to the convolution monodromy group of priKprjKpr_{i}^{*}K\oplus pr_{j}^{*}K. Hence for any pair ii and jj, there exists an automorphism σij\sigma_{ij} of GG modulo scalars that sends ρi\rho_{i} to ρj\rho_{j}. Since ρi\rho_{i} and ρj\rho_{j} are surjective, this automorphism is unique, and so σik=σjkσij\sigma_{ik}=\sigma_{jk}\sigma_{ij}. Hence if mm is greater than the order of the outer automorphism group of GG mod scalars, there are ii and jj such that σij\sigma_{ij} is an inner automorphism, say conjugation by gGg\in G. Because priKprjKpr_{i}^{*}K\oplus pr_{j}^{*}K is isomorphic to the pullback of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K, conjugation by gg sends ρ1\rho^{1} to ρ2\rho^{2}. Because the convolution monodromy group is well-defined only up to inner automorphisms in the first place, we may assume ρ1=ρ2\rho^{1}=\rho^{2}. It follows that the map from the convolution monodromy group of pr1Kpr2Kpr_{1}^{*}K\oplus pr_{2}^{*}K to G×GG\times G has image consisting of pairs (g1,g2)(g_{1},g_{2}) where g2=λg1g_{2}=\lambda g_{1} for a scalar λ\lambda.

Now the representation associated to pr1K[1]Dpr2Kpr_{1}^{*}K*[-1]^{*}Dpr_{2}^{*}K is the standard representation of the first GG tensored with the dual of the standard representation of the second GG. Because the convolution monodromy group consists of pairs (g1,g2)(g_{1},g_{2}) where g2=λg1g_{2}=\lambda g_{1} for a scalar λ\lambda, this tensor product contains a one-dimensional subrepresentation χ\chi. (Viewing this tensor product as the space of endomorphisms of the standard representation, the one-dimensional subrepresentation consists of scalar endomorphisms). Because χ\chi admits a nontrivial homomorphism to std1std2\operatorname{std}_{1}\otimes\operatorname{std}_{2}^{\vee}, we have a natural map χstd2std1\chi\otimes\operatorname{std}_{2}\to\operatorname{std}_{1}, which must be an isomorphism because both sides are irreducible. Any one-dimensional representation of the convolution monodromy group must be a skyscraper sheaf δx\delta_{x} [44, Proposition 10.1], so we have an isomorphism K2δx=K1K_{2}*\delta_{x}=K_{1} for some xA(η×η¯)x\in A(\overline{\eta\times\eta}). Considering the support, we see that the translation of Yη2Y_{\eta_{2}} by xx is Yη1Y_{\eta_{1}}. Spreading out this identity and then specializing η2\eta_{2} to a sufficiently general point, we see that YY is generically the translation of a constant variety by a section xx of AA. We can extend this section to some open set, and then YY over some open set will be the translation of a constant variety by a section, and then because YY is a smooth proper family this will be true globally, contradicting the assumption. ∎

Lemma 4.2.

Let A1,A2A_{1},A_{2} be abelian varieties. For j{1,2}j\in\{1,2\}, let zj:AjA1×A2z_{j}\colon A_{j}\to A_{1}\times A_{2} be the inclusion map and let KjK_{j} be a perverse sheaf on AjA_{j} with sheaf convolution group GjG_{j}. Then the sheaf convolution group of z1A1z2A2z_{1*}A_{1}\oplus z_{2*}A_{2} is G1×G2G_{1}\times G_{2}.

Proof.

Since z1A1z_{1*}A_{1} and z2A2z_{2*}A_{2} have Tannakian groups G1G_{1} and G2G_{2}, the Tannakian group of their sum is a subgroup of G1×G2G_{1}\times G_{2} whose projection onto both factors is surjective, so by Goursat’s lemma there exists a group HH and quotient maps q1:G1Hq_{1}\colon G_{1}\to H and q2:G2Hq_{2}\colon G_{2}\to H such the sheaf convolution group of z1A1z2A2z_{1*}A_{1}\oplus z_{2*}A_{2} is {(a1,a2)G1×G2q1(a1)=q2(a2)}\{(a_{1},a_{2})\in G_{1}\times G_{2}\mid q_{1}(a_{1})=q_{2}(a_{2})\}. For L1,L2L_{1},L_{2} perverse sheaves corresponding to a faithful representation of HH, this implies z1L1z2L2z_{1*}L_{1}\cong z_{2*}L_{2}, absurd unless L1L_{1} and L2L_{2} are both skyscraper sheaves at the identity, implying HH is trivial and so the Tannakian group is a product, as desired.∎

Corollary 4.3.

Assume that GG^{*} is a simple algebraic group with irreducible distinguished representation, and that YY is not equal to a constant family of hypersurfaces translated by a section of AA.

Let cc be a positive integer, and let i1,,ici_{1},\dots,i_{c} be the inclusions of AA into AcA^{c} that send AA to one of the cc coordinate axes.

Then the convolution monodromy group of j=1cijK\bigoplus_{j=1}^{c}i_{j*}K is GcG^{c} and thus contains (G)c(G^{*})^{c} as a normal subgroup. The representation associated to j=1cijK\bigoplus_{j=1}^{c}i_{j*}K is isomorphic to the sums of the distinguished representations of the cc factors of GG. Finally, this normal subgroup acts trivially on any representation of the convolution monodromy group corresponding to a perverse sheaf that is pulled back from AcA^{c}_{\mathbb{C}} to Aη¯cA^{c}_{\overline{\eta}}.

Proof.

The convolution monodromy group is GcG^{c} by induction on Lemma 4.2.

If any irreducible representation corresponding to a perverse sheaf which is a pullback of a perverse sheaf from AcA^{c}_{\mathbb{C}} to Aη¯cA^{c}_{\overline{\eta}} is nontrivial on (G)c(G^{*})^{c}, then it is nontrivial on the iith copy of GG^{*} for some ii. Tensor product with χ\mathcal{L}_{\chi} for generic χ\chi and then pushforward to AA gives a faithful exact tensor functor, whose associated map on Tannakian groups is the iith inclusion of GG into GcG^{c}, so the image under this map is nontrivial, which contradicts Lemma 4.1 and the fact that the pushforward to Aη¯A_{\overline{\eta}} of a pullback from AcA^{c}_{\mathbb{C}} is itself a pullback from AA_{\mathbb{C}} by proper base change.∎

We write Π(A)\Pi(A) for the set of continuous characters π1et(A)¯p×\pi_{1}^{et}(A)\to\overline{\mathbb{Q}}_{p}^{\times}. We call Π(A)\Pi(A) the dual torus of π1et(A)\pi_{1}^{et}(A).

We call the set of characters of π1(A)\pi_{1}(A) trivial on the fundamental group of a nontrivial abelian subvariety of AA a proper subtorus of the dual torus of π1et(A)\pi_{1}^{et}(A).

The generic vanishing theorem of Krämer and Weissauer [44, Lemma 11.2] states that for KK a perverse sheaf on AA, Hi(A,Kχ)=0H^{i}(A_{\mathbb{C}},K\otimes\mathcal{L}_{\chi})=0 for i0i\neq 0 for χ\chi outside a finite set of torsion translates proper subtori of Π(A)\Pi(A). (In fact it states this for characters of the topological fundamental group, but since every character of the et́ale fundamental group can be restricted to the topological one, that statement is stronger.) Building on this theorem, we prove a lemma that combines that statement with some useful information about H0H^{0}:

Lemma 4.4.

Let KDcb(A(η),¯p)K\in D^{b}_{c}(A_{\mathbb{C}(\eta)},\overline{\mathbb{Q}}_{p}) be a perverse sheaf of geometric origin. If no irreducible component of KK is a pullback from AA_{\mathbb{C}}, then for all characters χ:π1et(A(η)¯)¯p×\chi\colon\pi_{1}^{et}(A_{\overline{\mathbb{C}(\eta)}})\to\overline{\mathbb{Q}}_{p}^{\times} outside a finite set of torsion translates of proper subtori of Π(A)\Pi(A), we have Hi(A(η)¯,Kχ)=0H^{i}(A_{\overline{\mathbb{C}(\eta)}},K\otimes\mathcal{L}_{\chi})=0 for i0i\neq 0 and (H0(A(η)¯,Kχ))Gal((η)¯|(η))=0\left(H^{0}(A_{\overline{\mathbb{C}(\eta)}},K\otimes\mathcal{L}_{\chi})\right)^{\operatorname{Gal}(\overline{\mathbb{C}(\eta)}|\mathbb{C}(\eta))}=0.

Proof.

The first claim follows from [44, Lemma 11.2], which is stated for varieties over \mathbb{C} and singular cohomology, but we may embed (η)\mathbb{C}(\eta) into \mathbb{C} and then base change from the étale to the analytic site.

The second claim follows from the same theorem, but indirectly. By restricting to an open subset of XX, we may assume XX is smooth. Let mm be the dimension of XX. We may spread KK out (using the fact that it is of geometric origin) to a sheaf KK^{\prime} over A×XA\times X such that K[m]K^{\prime}[m] is perverse. Let π:A×XA\pi\colon A\times X\to A and ρ:A×XX\rho\colon A\times X\to X be the projections.

We will prove the second claim by contradiction. We will first assume that (H0(A(η)¯,Kχ))Gal((η)¯|(η))0\left(H^{0}(A_{\overline{\mathbb{C}(\eta)}},K\otimes\mathcal{L}_{\chi})\right)^{\operatorname{Gal}(\overline{\mathbb{C}(\eta)}|\mathbb{C}(\eta))}\neq 0 for a particular χ\chi such that Hi(A(η)¯,Kχ)=0H^{i}(A_{\overline{\mathbb{C}(\eta)}},K\otimes\mathcal{L}_{\chi})=0 for i0i\neq 0, and derive some conclusions from this. We will then define a finite set of torsion translates of proper subtori of Π(A)\Pi(A), assume that this nonvanishing holds for some χ\chi outside their union, and derive a contradiction from that.

Let us first see how to interpret the nonvanishing of monodromy invariants in terms of the perverse sheaf K[m]K^{\prime}[m]. This will essentially be the usual observation that sheaves with monodromy invariants have global sections, and thus have nonzero H0H^{0}. Additional care must be taken because Rρ(Kχ)R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}) is a complex of perverse sheaves, but the decomposition theorem will give us exactly what is needed.

The stalk of Rρ(Kχ)R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}) at the generic point is the complex H(A(η)¯,Kχ)H^{*}(A_{\overline{\mathbb{C}(\eta)}},K\otimes\mathcal{L}_{\chi}). By the assumption that this cohomology group vanishes in degree 0\neq 0, the stalk of Rρ(Kχ)R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}) at the generic point is supported in degree 0. There is some open subset of YY over which Rρ(Kχ)R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}) remains a lisse sheaf in degree 0, and the Galois action matches the monodromy action on that open subset. By the decomposition theorem, Rρ(Kχ)R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}) is a sum of shifts of irreducible perverse sheaves. In particular, this monodromy action is semisimple.

Now we assume that the Galois invariants are nonzero. It follows that the monodromy invariants are nonzero, and thus, by semisimplicity, there is a rank-one monodromy-invariant summand. Equivalently, there is a summand of Rρ(Kχ)R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}), restricted to this open set, that is isomorphic to the constant sheaf ¯p\overline{\mathbb{Q}}_{p}. Because Rρ(Kχ)R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}) is a sum of shifts of irreducible perverse sheaves, this irreducible summand ¯p\overline{\mathbb{Q}}_{p} on an open set must extend to a shift of an irreducible perverse sheaf on the whole space. Because XX is smooth, the unique irreducible extension of the constant sheaf ¯p\overline{\mathbb{Q}}_{p} from an open subset to all of XX is ¯p\overline{\mathbb{Q}}_{p}, which is a perverse sheaf shifted by mm. (In general, it would be the IC sheaf of XX.)

Because H0(X,¯p)0H^{0}(X,\overline{\mathbb{Q}}_{p})\neq 0, and ¯p\overline{\mathbb{Q}}_{p} is a summand of Rρ(Kχ)R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}), it follows that H0(X,Rρ(Kχ))0H^{0}(X,R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}))\neq 0.

Now that we have interpreted the existence of nontrivial monodromy invariants cohomologically, we can re-express the cohomology group in terms of shaves on AA, which will enable us to understand its dependence on χ\chi using the generic vanishing theorem. It follows from the Leray spectral sequence and the projection formula that

0H0(X,Rρ(Kχ))=H0(A×X,Kχ)=H0(A,RπKχ).0\neq H^{0}(X,R\rho_{*}(K^{\prime}\otimes\mathcal{L}_{\chi}))=H^{0}(A\times X,K^{\prime}\otimes\mathcal{L}_{\chi})=H^{0}(A,R\pi_{*}K^{\prime}\otimes\mathcal{L}_{\chi}).

Now we choose our finite union of torsion translates of subtori. We apply the generic vanishing theorem to every perverse sheaf in sight. For all jj\in\mathbb{Z} with jp(RπK)0{}^{p}\mathcal{H}^{j}(R\pi_{*}K^{\prime})\neq 0, we take from [44, Lemma 11.2] a finite set of torsion translates of subtori such that Hi(A,jp(RπK)χ)=0H^{i}(A,{}^{p}\mathcal{H}^{j}(R\pi_{*}K^{\prime})\otimes\mathcal{L}_{\chi})=0 for all χ\chi not in this set and all i0i\neq 0. Because there are only finitely many jj where jp(RπK)0{}^{p}\mathcal{H}^{j}(R\pi_{*}K^{\prime})\neq 0, the union of all of these is again a finite set.

Assume that (H0(A(η)¯,Kχ))Gal((η)¯|(η))0\left(H^{0}(A_{\overline{\mathbb{C}(\eta)}},K\otimes\mathcal{L}_{\chi})\right)^{\operatorname{Gal}(\overline{\mathbb{C}(\eta)}|\mathbb{C}(\eta))}\neq 0 for a particular χ\chi not in this set. The vanishing of Hi(A,jp(RπK)χ)H^{i}(A,{}^{p}\mathcal{H}^{j}(R\pi_{*}K^{\prime})\otimes\mathcal{L}_{\chi}) for all jj and all i0i\neq 0 forces the spectral sequence

Hi(A,pj(RπK)χ)Hi+j(A,RπKχ)H^{i}(A,^{p}\mathcal{H}^{j}(R\pi_{*}K^{\prime})\otimes\mathcal{L}_{\chi})\mapsto H^{i+j}(A,R\pi_{*}K^{\prime}\otimes\mathcal{L}_{\chi})

to degenerate, giving

0H0(A,RπKχ)=H0(A,p0(RπK)χ).0\neq H^{0}(A,R\pi_{*}K^{\prime}\otimes\mathcal{L}_{\chi})=H^{0}(A,^{p}\mathcal{H}^{0}(R\pi_{*}K^{\prime})\mathcal{L}_{\chi}).

In particular, we can conclude that

0p(RπK)0.{}^{p}\mathcal{H}^{0}(R\pi_{*}K^{\prime})\neq 0.

We will now derive a contradiction from this simpler statement, which notably is independent of χ\chi.

Note first that because K[m]K^{\prime}[m] is perverse, and π\pi has fibers of dimension at most mm, we have ip(RπK)=0{}^{p}\mathcal{H}^{i}(R\pi_{*}K^{\prime})=0 for i<0i<0 by [8, 4.2.4]. Hence there is a natural map

0p(RπK)RπK{}^{p}\mathcal{H}^{0}(R\pi_{*}K^{\prime})\to R\pi_{*}K^{\prime}

arising from the perverse tt-structure. Because 0p(RπK){}^{p}\mathcal{H}^{0}(R\pi_{*}K^{\prime}) is nonzero, this map must be nonzero. Applying adjunction, we obtain a nonzero map π0p(RπK)K\pi^{*}{}^{p}\mathcal{H}^{0}(R\pi_{*}K^{\prime})\to K^{\prime}, and and shifting by mm, a nonzero map

π0p(RπK)[m]K[m].\pi^{*}{}^{p}\mathcal{H}^{0}(R\pi_{*}K^{\prime})[m]\to K^{\prime}[m].

Because this is a nonzero map between perverse sheaves, some irreducible component of the source is equal to some irreducible component of the target. This cannot happen because by assumption no irreducible component of KK is a pullback from AA_{\mathbb{C}}, giving a contradiction.∎

Lemma 4.5.

Let GG^{*} be a simple algebraic group and cc a natural number. Fix an irreducible representation of GG^{*}. Let N(G)N(G^{*}) be the normalizer of GG^{*} inside the group of automorphisms of this representation. Then there is a finite list of irreducible representations of N(G)cN(G^{*})^{c} such that a reductive subgroup BB of N(G)cN(G^{*})^{c} contains (G)c(G^{*})^{c} if and only if BB has no invariants on any of these representations.

Proof.

Let mm be such that any subgroup of N(G)N(G^{*}) which is finite modulo scalars contains an abelian subgroup of index mm. Such mm exists by Jordan-Schur. Let VV be an irreducible representation of N(G)N(G^{*}) of dimension >m>m which remains irreducible on restriction to GG^{*}. Let πi\pi_{i} be the ii-th projection (N(G)c)N(G)(N(G^{*})^{c})\to N(G^{*}).

We take our list to be, for each ii from 11 to cc, all irreducible N(G)N(G^{*})-subrepresentations of (VV)/(V\otimes V^{\vee})/\mathbb{C} and (adGadG)/(\operatorname{ad}G^{*}\otimes\operatorname{ad}G^{*\vee})/\mathbb{C} composed with πi\pi_{i}, together with, for 1i<jc1\leq i<j\leq c, the representation adGπiadGπj\operatorname{ad}G^{*}\circ\pi_{i}\otimes\operatorname{ad}G^{{*\vee}}\circ\pi_{j}. It is straightforward to check that (G)c(G^{*})^{c} has no invariants on these representations, so if BB contains (G)c(G^{*})^{c} then BB has no invariants. We must check the converse.

If BB has no invariants, then πi(B)\pi_{i}(B) acts irreducibly on VV and ad(G)\operatorname{ad}(G^{*}). The Lie algebra of πi(B)\pi_{i}(B) is an invariant subspace of ad(N(G))ad(G)\operatorname{ad}(N(G^{*}))\cong\operatorname{ad}(G^{*})\oplus\mathbb{C}, thus either contains ad(G)\operatorname{ad}(G^{*}), in which case πi(B)\pi_{i}(B) contains GG^{*}, or is contained in \mathbb{C}, in which case πi(B)\pi_{i}(B) is finite modulo scalars, hence has an abelian subgroup of index m\leq m, thus cannot act irreducibly on VV, a contradiction. (Compare [39, Theorem 2.2.2], due originally to Larsen.)

Since πi(B)\pi_{i}(B) contains πi\pi_{i} for all ii, we have πi([B0,B0])=G\pi_{i}([B^{0},B^{0}])=G^{*}. By [40, Theorem on p. 1152], if [B0,B0](G)c[B^{0},B^{0}]\neq(G^{*})^{c}, then there exists i,ji,j with 1i<jc1\leq i<j\leq c and an isomorphism GGG^{*}\to G^{*} that sends πi\pi_{i} to πj\pi_{j}. This isomorphism is unique and thus BB-invariant, so adGπi\operatorname{ad}G^{*}\circ\pi_{i} and adGπj\operatorname{ad}G^{{*}}\circ\pi_{j} are isomorphic as representations of BB. Thus adGπiadGπj\operatorname{ad}G^{*}\circ\pi_{i}\otimes\operatorname{ad}G^{{*\vee}}\circ\pi_{j} has invariants, contradicting our assumption and showing (G)c=[B0,B0]B(G^{*})^{c}=[B^{0},B^{0}]\subseteq B. ∎

Remark 4.6.

Lists of representations satisfying the condition of Lemma 4.5 with smaller dimension follow from Larsen’s conjecture [46, Theorem 1.4] in the case that GG^{*} is a classical group, but these depend on the classification of finite simple groups. For some applications of results of this type, optimizing the dimension of the representations is relevant, but not here.

Let f:YXf\colon Y\to X and g:YAg\colon Y\to A be the projection maps associated to YX×AY\subseteq X\times A a family of smooth projective hypersurfaces parameterized by XX.

Recall that the geometric monodromy group of a constructible pp-adic sheaf on an irreducible scheme is the Zariski closure of the image of the natural map from the geometric étale fundamental group of the largest open set on which the sheaf is lisse to the general linear group of the stalk at the generic point.

The following theorem is the analogue of Pink’s specialization theorem [38, Theorem 8.18.2], which shows, given any sheaf on the total space of a family of schemes, for “most” schemes in the family (i.e. for the fibers over a dense open subset), the monodromy of the restricted sheaf is equal to a generic monodromy group. Our analogue shows that for “most” characters χ\chi, the monodromy of i=1cRn1f(gχi)\bigoplus_{i=1}^{c}R^{n-1}f_{*}(g^{*}\mathcal{L}_{\chi_{i}}) is (roughly) equal to a generic convolution monodromy group. This will connect our earlier investigations of the convolution monodromy group to our later arguments, which require control on monodromy groups.

Theorem 4.7.

Assume that Yη¯Y_{\overline{\eta}} is not translation-invariant by any nonzero element of AA, that GG^{*} is a simple algebraic group with irreducible distinguished representation, and that YY is not equal to a constant family of hypersurfaces translated by a section of AA. Fix cc\in\mathbb{N}.

Then for χ1,,χc\chi_{1},\dots,\chi_{c} characters of π1et(A)\pi_{1}^{et}(A), with (χ1,,χc)(\chi_{1},\dots,\chi_{c}) avoiding some finite set of torsion translates of proper subtori of the dual torus Π(A)c\Pi(A)^{c} to π1et(Ac)\pi_{1}^{et}(A^{c}), the following conditions are satisfied:

  • Rkf(gχi)=0R^{k}f_{*}(g^{*}\mathcal{L}_{\chi_{i}})=0 for kn1k\neq n-1, and

  • the geometric monodromy group of i=1cRn1f(gχi)\bigoplus_{i=1}^{c}R^{n-1}f_{*}(g^{*}\mathcal{L}_{\chi_{i}}) contains (G)c(G^{*})^{c} as a normal subgroup, where the representation associated to i=1cRn1f(gχi)\bigoplus_{i=1}^{c}R^{n-1}f_{*}(g^{*}\mathcal{L}_{\chi_{i}}), restricted to (G)c(G^{*})^{c}, is isomorphic to a sum of cc copies of the distinguished representation of GG^{*}.

Proof.

First, we use the generic vanishing theorem to find a finite set of torsion translates of proper subtori of Π(A)\Pi(A) such that, for χ\chi avoiding them, Rkf(gχ)=0R^{k}f_{*}(g^{*}\mathcal{L}_{\chi})=0 for kn1k\neq n-1. We will then take the inverse images of these subtori under the duals of the cc projections Π(Ac)Π(A)\Pi(A^{c})\to\Pi(A) to be in our finite set of torsion translates of subtori of Π(A)c\Pi(A)^{c}.

Next, to calculate the monodromy, we will use the fact that the monodromy group is equal to the Zariski closure of the image of Gal((η)¯|(η))\operatorname{Gal}(\overline{\mathbb{C}(\eta)}|\mathbb{C}(\eta)) acting on the stalk at the geometric generic point.

Let us first check that the geometric monodromy group of i=1cRn1f(gχi)\bigoplus_{i=1}^{c}R^{n-1}f_{*}(g^{*}\mathcal{L}_{\chi_{i}}) is contained in N(G)cN(G^{*})^{c}. It suffices to show for each ii that the geometric monodromy group of Rn1f(gχi)R^{n-1}f_{*}(g^{*}\mathcal{L}_{\chi_{i}}) is contained in N(G)N(G^{*}). This follows from Lemma 2.9(2).

Now by Lemma 4.5 we can find an explicit list of representations of N(G)cN(G^{*})^{c} such that any reductive subgroup of N(G)cN(G^{*})^{c} contains (G)c(G^{*})^{c} if and only if its action on all these representations has no invariants. By Deligne’s theorem, the monodromy group of i=1cRn1f(gχi)\bigoplus_{i=1}^{c}R^{n-1}f_{*}(g^{*}\mathcal{L}_{\chi_{i}}) is reductive, and so we can apply this lemma.

By Lemma 2.8, each representation from the list of Lemma 4.5 corresponds to a perverse sheaf on AηcA^{c}_{\eta} in the Tannakian subcategory generated by j=1ciji¯p[n1]\bigoplus_{j=1}^{c}i_{j*}i_{*}\overline{\mathbb{Q}}_{p}[n-1] inside the arithmetic Tannakian category constructed in Lemma 2.7 of perverse sheaves on AηcA^{c}_{\eta} modulo negligible sheaves. (We have to check that j=1ciji¯p[n1]\bigoplus_{j=1}^{c}i_{j*}i_{*}\overline{\mathbb{Q}}_{p}[n-1] is geometrically semisimple, but this is clear as the constant sheaf on any closed subvariety is semisimple. It follows that the action of GkG_{k} on the representation associated to this complex factors through the normalizer of the action of GkG_{k^{\prime}}, and thus factors through N(G)cN(G^{*})^{c}, and so all representations of N(G)cN(G^{*})^{c} correspond to geometrically semisimple perverse sheaves.)

Because (G)c(G^{*})^{c} acts nontrivially on all representations from Lemma 4.5, by Lemma 4.3 none of these perverse sheaves is a pullback from AcA^{c}_{\mathbb{C}}, so by Lemma 4.4, outside some finite set of torsion translates of proper subtori of Π(A)c\Pi(A)^{c}, the Galois group has no invariants for these representations. Thus, outside some finite set of torsion translates, the Galois group contains (G)c(G^{*})^{c}. ∎

We now specialize to the case that AA is defined over a number field KK. Fix an algebraic closure ¯\overline{\mathbb{Q}} of \mathbb{Q}. Let ΠK/(A)\Pi^{K/\mathbb{Q}}(A) be the set of pairs of an embedding ι\iota of AA into ¯\overline{\mathbb{Q}} and a character χ\chi of π1(Aι)\pi_{1}(A_{\iota}) where Aι=A×K,ι¯A_{\iota}=A\times_{K,\iota}\overline{\mathbb{Q}}. Then ΠK/(A)\Pi^{K/\mathbb{Q}}(A) naturally admits an action of Gal×Galcyc/\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}}/\mathbb{Q}}, where Gal\operatorname{Gal}_{\mathbb{Q}} acts on ¯\overline{\mathbb{Q}} and Galcyc/\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}}/\mathbb{Q}} acts on the values of the character χ\chi (which necessarily lie in cyc\mathbb{Q}^{\textrm{cyc}}. This action extends the more straightforward action of GalK×Galcyc/\operatorname{Gal}_{K}\times\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}}/\mathbb{Q}} on Π(A)\Pi(A).

Fix inside the group Gal×Galcyc/\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}}/\mathbb{Q}} an element Frobp\operatorname{Frob}_{p} of elements projecting to a lift of Frobenius inside the decomposition group at pp in both Gal×Galcyc/\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}}/\mathbb{Q}}.

For applications later in the paper, we will be interested in proving our large monodromy statement uniformly in an orbit of Gal×Galcyc/\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}}/\mathbb{Q}} and this statement will involve a monodromy group of a direct sum of sheaves produced by cc characters in the orbit of Frobp\operatorname{Frob}_{p}. We therefore prove a series of lemmas that enable us to obtain a monodromy statement of this type.

Lemma 4.8.

Suppose AA is defined over a number field KK, and let pp be a prime at which AA has good reduction. Let mm be a natural number such that Frobpm\operatorname{Frob}_{p}^{m} fixes the Galois closure of KK. For any positive integer cc and set SS of torsion translates of proper subtori of the dual torus Π(A)c\Pi(A)^{c} to π1(Ac)\pi_{1}(A^{c}), there is a finite set SS^{\prime} of torsion translates of proper subtori of Π(A)\Pi(A) such that for any χ\chi outside the union of SS^{\prime}, there exist e1,,ece_{1},\dots,e_{c}\in\mathbb{Z} such that the tuple (Frobpme1(χ),,Frobpmec(χ))(\operatorname{Frob}_{p}^{me_{1}}(\chi),\dots,\operatorname{Frob}_{p}^{me_{c}}(\chi)) does not lie in the union of SS.

Proof.

We can freely replace mm by any multiple. By choosing a suitable multiple, we may assume Frobp\operatorname{Frob}_{p} fixes all elements of SS.

Let σ=Frobpm\sigma=\operatorname{Frob}_{p}^{m}. We will prove this lemma as a consequence of the fact that the action of σ\sigma on Π(A)\Pi(A) is by invertible linear transformations, with no roots of unity as eigenvalues. (The action of FrobpmGalcyc/\operatorname{Frob}_{p}^{m}\in\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}}/\mathbb{Q}} is by multiplication by pmp^{m}, while the action of FrobpmGalKGal\operatorname{Frob}_{p}^{m}\in\operatorname{Gal}_{K}\subseteq\operatorname{Gal}_{\mathbb{Q}} is invertible, with eigenvalues the inverses of Weil numbers of absolute value pm/2p^{m/2}. Hence their product has eigenalues Weil numbers of absolute value qm/2q^{m/2}, which are in particular not roots of unity.)

It suffices to show for each torsion translate of a torus T+ξT+\xi in SS, for all χ\chi outside some finite set of torsion transltaes of proper subtori of Π(A)\Pi(A), the number of tuples e1,,ec{0,,|S|}e_{1},\dots,e_{c}\in\{0,\dots,|S|\} such that the tuple (σe1(χ),,σec(χ))(\sigma^{e_{1}}(\chi),\dots,\sigma^{e_{c}}(\chi)) does not lie in T+ξT+\xi is at most (|S|+1)c1(|S|+1)^{c-1}. Indeed, we can then take SS^{\prime} to be the union of all these finite sets, and then the number of tuples e1,,ec{0,,|S|}e_{1},\dots,e_{c}\in\{0,\dots,|S|\} such that (σe1(χ),,σec(χ))(\sigma^{e_{1}}(\chi),\dots,\sigma^{e_{c}}(\chi)) does not lie in the union of SS will be at least

(|S|+1)c|S|(|S|+1)c1>0.(|S|+1)^{c}-|S|(|S|+1)^{c-1}>0.

For each rr from 11 to cc, let TrT^{r} be the inverse image of TT under the rr’th inclusion map Π(A)Π(Ac)\Pi(A)\to\Pi(A^{c}). Each TrT^{r} is a subtorus of Π(A)\Pi(A) and, since TT is proper and the images of the inclusion maps Π(A)Π(Ac)\Pi(A)\to\Pi(A^{c}) generate π(Ac)\pi(A^{c}) at least one of the TrT^{r} must be proper. Fix this rr. We will first check that

{χσe(χ)σe(χ)1Tr for some 0e<e|S|}\{\chi\mid\sigma^{e}(\chi)\sigma^{e^{\prime}}(\chi)^{-1}\in T^{r}\textrm{ for some }0\leq e<e^{\prime}\leq|S|\}

is a finite union of torsion translates of TrT^{r} and then check that for χ\chi not in this set, the number of tuples e1,,ec{0,,|S|}e_{1},\dots,e_{c}\in\{0,\dots,|S|\} such that the tuple (σe1(χ),,σec(χ))(\sigma^{e_{1}}(\chi),\dots,\sigma^{e_{c}}(\chi)) lies in T+ξT+\xi is at most (|S|+1)c1(|S|+1)^{c-1}

For the first claim, TrT^{r} consists of characters restricting to the trivial character on some abelian subvariety BB. If σe(χ)σe(χ)1Tr\sigma^{e}(\chi)\sigma^{e^{\prime}}(\chi)^{-1}\in T^{r} then σe(χ)\sigma^{e}(\chi) and σe(χ)\sigma^{e^{\prime}}(\chi) agree on restriction to BB. Since the eigenvalues of σee\sigma^{e^{\prime}-e} are algebraic numbers but not roots of unity, the fixed points of σee\sigma^{e^{\prime}-e} on BB are torsion and finite in number. The same is true for their inverse image under σe\sigma^{e}, i.e. the elements of Π(B)\Pi(B) whose images under σe\sigma^{e} and σe\sigma^{e^{\prime}} agree. Characters of π1(A)\pi_{1}(A) whose restriction to π1(B)\pi_{1}(B) take a fixed torsion value form a torsion translate of TT, completing the proof of the first claim.

For the second claim, the number of choices of eie_{i} for all iri\neq r is (|S|+1)c1(|S|+1)^{c-1} and these choices, together with the condition that (σe1(χ),,σec(χ))(\sigma^{e_{1}}(\chi),\dots,\sigma^{e_{c}}(\chi)) lies in T+ξT+\xi, determine ere_{r}, since if both (σe1(χ),,σer(χ),σec(χ))(\sigma^{e_{1}}(\chi),\dots,\sigma^{e_{r}}(\chi),\dots\sigma^{e_{c}}(\chi)) and (σe1(χ),,σer(χ),σec(χ))(\sigma^{e_{1}}(\chi),\dots,\sigma^{e_{r}^{\prime}}(\chi),\dots\sigma^{e_{c}}(\chi)) lie in T+ξT+\xi by dividing, without loss of generality with er<ere_{r}<e_{r}^{\prime}, it would follow that (1,,1,σer(χ)σer(χ)1,1,,1)(1,\dots,1,\sigma^{e_{r}}(\chi)\sigma^{e_{r}^{\prime}}(\chi)^{-1},1,\dots,1) lies in TT and hence σer(χ)σer(χ)1\sigma^{e_{r}}(\chi)\sigma^{e_{r}^{\prime}}(\chi)^{-1} lies in TrT^{r}, contradicting the assumption on χ\chi. ∎

Lemma 4.9.

Suppose AA is defined over a number field KK. For each tuple indexed by embeddings ι:K¯\iota\colon K\to\overline{\mathbb{Q}} of finite subsets SιS_{\iota} of torsion translates of proper subtori of Π(Aι)\Pi(A_{\iota}), there exists an embedding ι\iota of KK into ¯\overline{\mathbb{Q}} and a torsion character χ\chi of π1(Aι)\pi_{1}(A_{\iota}) such that for no Gal×Galcyc/\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}/\mathbb{Q}}}-conjugate (ι,χ)(\iota^{\prime},\chi^{\prime}) of (ι,χ)(\iota,\chi) does χ\chi^{\prime} lie in the union of SιS_{\iota^{\prime}}. Furthermore, we can arrange that χ\chi is of order a power of \ell, for any given prime \ell.

Proof.

For χ\chi to have no Galois conjugate in any element of any SιS_{\iota^{\prime}}, it suffices that χ\chi not lie in any Galois conjugate of these translates of SιS_{\iota^{\prime}}.

By definition, each proper subtorus of Π(A)\Pi(A) corresponds to some abelian subvariety, which must be defined over a number field. Since every torsion point is defined over a number field, every torsion translate of a proper subtorus is defined over a number field. Hence they have finitely many conjugates under Gal\operatorname{Gal}_{\mathbb{Q}}. The action of Galcyc/\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}/\mathbb{Q}}} fixes each subtorus and gives each torsion point finitely many conjuates, so the total number of conjugates under both actions is finite. Thus the union of all these conjugates is a finite union of torsion translates of proper subtori.

Each proper subtorus can contain at most 2k(n1)\ell^{2k(n-1)} k\ell^{k}-torsion characters, and so any translate of a proper subtorus can contain at most 2k(n1)\ell^{2k(n-1)} k\ell^{k}-torsion characters, while there are 2kn\ell^{2kn} 2k\ell^{2k}-torsion characters in total, so as soon as 2k\ell^{2k} is greater than this finite number of tori, there will be an k\ell^{k}-torsion character not in any of them. ∎

For the remainder of this section, we’ll suppose AA is an abelian variety over a number field KK, XX is a smooth scheme over \mathbb{Q}, and

YX×A=XK×KAY\subseteq X\times_{\mathbb{Q}}A=X_{K}\times_{K}A

is a family of hypersurfaces in AA, smooth, proper and flat over XKX_{K}. For every embedding ι:K\iota\colon K\rightarrow\mathbb{C}, we can form schemes

Aι=ASpecK,ιSpecA_{\iota}=A\otimes_{\operatorname{Spec}K,\iota}\operatorname{Spec}\mathbb{C}

and

Yι=YSpecK,ιSpec;Y_{\iota}=Y\otimes_{\operatorname{Spec}K,\iota}\operatorname{Spec}\mathbb{C};

these are both schemes over \mathbb{C}, and YιY_{\iota} has a projection to

X=XSpecSpec.X_{\mathbb{C}}=X\otimes_{\operatorname{Spec}\mathbb{Q}}\operatorname{Spec}\mathbb{C}.

Let f:YιXf\colon Y_{\iota}\rightarrow X_{\mathbb{C}} and g:YιAιg\colon Y_{\iota}\rightarrow A_{\iota} be the projections; for every torsion character χ\chi of π1et(Aι)\pi_{1}^{et}(A_{\iota}), let (ι,χ)\mathcal{L}_{(\iota,\chi)} be the corresponding character sheaf on AιA_{\iota}.

The next result will be what we use from this section later in the paper. This gives a big monodromy statement which holds simultaneously for multiple characters – as many as desired – in a Galois orbit. The big monodromy will be an input into the Lawrence-Venkatesh method, and being able to work with multiple characters in a Galois orbit helps to control the variation of a global Galois representation. (Specifically, in order to apply Lemma 7.2 or Lemma 8.16, it is advantageous to be able to take cc as large as desired.)

Corollary 4.10.

Assume that Yη¯Y_{\overline{\eta}} is not translation-invariant by any nonzero element of AA, that GG^{*} is a simple algebraic group with irreducible distinguished representation, and that YY is not equal to a constant family of hypersurfaces translated by a section of AA.

Let mm be a natural number such that Frobpm\operatorname{Frob}_{p}^{m} fixes the Galois closure of KK.

Then for any prime \ell, positive integer cc, and prime pp where AA has good reduction, there exist an embedding ι:K\iota\colon K\rightarrow\mathbb{C} and a torsion character χ\chi of π1et(Aι)\pi_{1}^{et}(A_{\iota}), of order a power of \ell, such that for every conjugate (ι,χ)(\iota^{\prime},\chi^{\prime}) of (ι,χ)(\iota,\chi) by an element of Gal()×Galcyc/\operatorname{Gal}(\mathbb{Q})\times\operatorname{Gal}_{\mathbb{Q}^{\textrm{cyc}/\mathbb{Q}}}:

  • for kn1k\neq n-1, we have Rkfι(gι(ι,χ))=0R^{k}{f_{\iota^{\prime}}}_{*}(g_{\iota^{\prime}}^{*}\mathcal{L}_{(\iota^{\prime},\chi^{\prime})})=0, and

  • there exist e1,,ece_{1},\dots,e_{c}\in\mathbb{Z} such that the monodromy group of i=1cRn1fι(gιFrobpmei(ι,χ))\bigoplus_{i=1}^{c}R^{n-1}{f_{\iota^{\prime}}}_{*}(g_{\iota^{\prime}}^{*}\mathcal{L}_{\operatorname{Frob}_{p}^{me_{i}}(\iota^{\prime},\chi^{\prime})}) contains (G)c(G^{*})^{c} as a normal subgroup, where the representation associated to i=1cRn1f(gFrobpmei(ι,χ))\bigoplus_{i=1}^{c}R^{n-1}f_{*}(g^{*}\mathcal{L}_{\operatorname{Frob}_{p}^{me_{i}}(\iota^{\prime},\chi^{\prime})}), restricted to (G)c(G^{*})^{c}, is isomorphic to a sum of cc copies of the distinguished representation of GG^{*}.

We will eventually apply this result with the parameter cc taken sufficiently large to ensure the inequalities stated in Theorem 8.17 are satisfied. For this reason, the parameter cc will depend on the varieties AA, XX, and YY involved, and then Corollary 4.10 will give us a character χ\chi depending on cc. However, until we choose the parameter cc in this way, all our results will be valid for any positive integer cc.

Proof.

This follows from the previous 3 results, applying Theorem 4.7 to each of the finitely many pairs (Yι,Aι)(Y_{\iota},A_{\iota}). ∎

Using Theorem 4.7, we can also prove a result on the period maps of certain variations of Hodge structures associated to families of hypersurfaces in an abelian variety. This is not used anywhere in this paper. Instead, it provides a different perspective on our main result, showing that it is compatible with the “Shafarevich” philosophy that varieties with a quasi-finite period map should have finitely many 𝒪K[1/S]\mathcal{O}_{K}[1/S]-points for any number field KK and set SS of prime ideals.

Proposition 4.11.

Let AA be an abelian variety of dimension n2n\geq 2 over \mathbb{C}. Let ϕ\phi be an ample class in the Picard group of AA. For a positive integer mm, let [m]:AA[m]\colon A\to A be the multiplication-by-mm map.

There exists a positive integer mm such that the natural period map from the moduli space of smooth hypersurfaces HH in AA of class ϕ\phi to a period domain, which sends HH to the Hodge structure on Hn1([m]1H,)H^{n-1}([m]^{-1}H,\mathbb{Q}), is quasi-finite.

Proof.

Let \mathcal{M} be this moduli space. Suppose that, for some mm (to be chosen below), the period map is not quasi-finite. Then its fiber over some point must contain a positive dimensional analytic subvariety WmW_{m}.

Consider the variation of Hodge structures over ×\mathcal{M}\times\mathcal{M} whose fiber over a pair of hypersurfaces H1,H2H_{1},H_{2} is Hn1([m]1H1,)Hn1([m]1H2,)H^{n-1}([m]^{-1}H_{1},\mathbb{Q})\otimes H^{n-1}([m]^{-1}H_{2},\mathbb{Q})^{\vee}. Over the diagonal in ×\mathcal{M}\times\mathcal{M}, this variation of Hodge structures has a Hodge class representing the identity isomorphism between the two Hodge structures. Let Zm×Z_{m}\subset\mathcal{M}\times\mathcal{M} be the projection from the universal cover of the locus where this cohomology class is Hodge.

By [13, Corollary 1.3], ZmZ_{m} is a Zariski closed subset. If m1m_{1} divides m2m_{2} then we have Zm2Zm1Z_{m_{2}}\subseteq Z_{m_{1}}. Because ×\mathcal{M}\times\mathcal{M} is Noetherian, it follows that there exists mm such that Zm=ZmZ_{m^{\prime}}=Z_{m} whenever mm^{\prime} is a multiple of mm. Fix such an mm.

ZmZ_{m} certainly contains the square of the positive-dimensional analytic subvariety WmW_{m} discussed earlier. Fix xWmx\in W_{m}. It follows that the fiber of ZmZ_{m} over xx (for the second projection Zm×Z_{m}\subseteq\mathcal{M}\times\mathcal{M}\rightarrow\mathcal{M}) has positive dimension. Let XX be the smooth locus of some irreducible component of the fiber of ZmZ_{m} over xx. Let f:YXf\colon Y\to X be the universal family of hypersurfaces H2H_{2} over XX and g:YAg\colon Y\to A the projection map. By assumption, for all multiples mm^{\prime} of mm, for all yXy\in X, we have an isomorphism of Hodge structures

Hn1([m]1H1,x,)=Hn1([m]1H1,y,)Hn1([m]1H2,y,).H^{n-1}([m^{\prime}]^{-1}H_{1,x},\mathbb{Q})=H^{n-1}([m^{\prime}]^{-1}H_{1,y},\mathbb{Q})\cong H^{n-1}([m]^{-1}H_{2,y},\mathbb{Q}).

Hence the variation of Hodge structures Hn1([m]1H2,y,)H^{n-1}([m^{\prime}]^{-1}H_{2,y},\mathbb{Q}) is constant, and thus has finite monodromy. This is the sum, over characters χ\chi of π1(A)\pi_{1}(A) of order dividing mm^{\prime}, of Rn1f(gχ)R^{n-1}f^{*}(g^{*}\mathcal{L}_{\chi}), and so all these individual summands have finite monodromy.

The family YXY\to X is a family of smooth hypersurfaces. Because we have fixed an ample class in the Picard group, and there are only finitely many translates of a given hypersurface in a given Picard class, and because XX is a positive-dimensional subvariety of the moduli space \mathcal{M}, it is not the constant family up to translation. From this fact, and the finiteness of the mondromy of Rn1f(gχ)R^{n-1}f^{*}(g^{*}\mathcal{L}_{\chi}) for all torsion characters χ\chi, we will derive a contradiction.

Before proceeding, we consider the case where Yη¯Y_{\overline{\eta}} is translation-invariant by a nonzero element of AA, for η¯\overline{\eta} the generic point of XX. It follows that the whole family is invariant under the same element. In this case, we consider the subgroup of all such elements and quotient AA by it. The family YY is then a pullback from a family YY^{\prime} of hypersurfaces in this quotient AA^{\prime} of AA, and the pushforward from YY^{\prime} of gχg^{{}^{\prime}*}\mathcal{L}_{\chi^{\prime}} is a summand of the pushforward from YY of gχg^{*}\mathcal{L}_{\chi}, where χ\chi is the composition of χ\chi^{\prime} with the map AAA\to A^{\prime}, so our finite monodromy assumption remains true for YY^{\prime}. Hence we may assume that Yη¯Y_{\overline{\eta}} is not translation-invariant by a nonzero element of AA.

Let GG be the convolution monodromy group of Yη¯Y_{\overline{\eta}} and let GG^{*} be the commutator subgroup of the identity component of GG. By Lemmas 3.8 and 3.9, GG^{*} is a simple algebraic group acting by an irreducible representation. We have thus verified all the assumptions of Theorem 4.7. It follows that for all χ\chi outside some finite set of proper subtori Π(A)\Pi(A), which necessarily includes at least one torsion character, the geometric monodromy group of Rn1f(gχ)R^{n-1}f_{*}(g^{*}\mathcal{L}_{\chi}) contains GG^{*}, contradicting our assumption that it is finite, as desired. ∎

5. Hodge–Deligne systems

The goal of the next few sections is to prove Theorem 8.17, which is analogous to Lemma 4.2, Prop. 5.3, and Thm. 10.1 in [47]. Roughly, the theorem says that, if a smooth variety XX over \mathbb{Q} admits a Hodge–Deligne system that has big monodromy and satisfies two numerical conditions, then the integral points of XX are not Zariski dense. We follow the same strategy as [47], but we’ll need to work in greater generality. First, [47] works only with the primitive cohomology of a family of varieties, but we’ll need to work with the cohomology with coefficients in a local system. Second, we’ll need to work with Galois representations valued in a disconnected reductive group. Finally, we are unable to precisely identify the Zariski closure of the image of monodromy; we only know that it is a cc-balanced subgroup of 𝐆dR\mathbf{G}_{dR} (Definition 6.6).

We’ll begin by defining the notion of “Hodge–Deligne system”, which will figure in our statement of Theorem 8.17. Let XX be a variety over a number field KK (which will eventually be taken to be \mathbb{Q}). A smooth, projective family of varieties over XX gives rise to various cohomology objects on XX. The argument of [47] relies on the interplay among several of these objects: a complex period map, a pp-adic period map, and a family of pp-adic global Galois representations on XX. Deligne has called the collection of these cohomology objects a “system of realizations” for a motive [16]; our notion of “Hodge–Deligne system” will be closely related to Deligne’s systems of realizations.

5.1. Summary of constructions and notation

In this section we define terms like “Hodge-Deligne system” and “H0H^{0}-algebra” in a level of generality which is natural but is greater than what we need for the rest of the argument. Here, we briefly review the specific setup we will use in the proof so that one can have a concrete case in mind when reading the definitions. Thus, the meaning of this summary should not be completely clear before the definitions are read.

We begin (see Section 5.5) with AA an abelian variety of dimension nn over a number field KK. Let XX be an arbitrary smooth variety over \mathbb{Q}, and XKX_{K} its base change to KK. Let

YX×A=XK×KAY\subseteq X\times_{\mathbb{Q}}A=X_{K}\times_{K}A

be a subscheme, smooth, proper and flat over XKX_{K}. (In our application, we will take XX to be the Weil restriction, from KK to \mathbb{Q}, of a subvariety of the moduli space of hypersurfaces on AA, and YY the universal hypersurface over XX. See the proof of Theorem 9.2.)

We can choose finite sets SS of places of KK, and SS^{\prime} of places of \mathbb{Q}, and spread everything out to a family

f:𝒴𝒳×[1/S]𝒜,f\colon\mathcal{Y}\subseteq\mathcal{X}\times_{\mathbb{Z}[1/S^{\prime}]}\mathcal{A},

in such a way that 𝒪K,S\mathcal{O}_{K,S} is finite étale over [1/S]\mathbb{Z}[1/S^{\prime}], 𝒜\mathcal{A} is a smooth abelian scheme over 𝒪K,S\mathcal{O}_{K,S}, 𝒳\mathcal{X} is smooth over [1/S]\mathbb{Z}[1/S^{\prime}], and 𝒴𝒳𝒪K,S\mathcal{Y}\rightarrow\mathcal{X}_{\mathcal{O}_{K},S} is smooth, proper, and flat.

Fix a prime pp. Let LL be a field, containing KK, Galois over \mathbb{Q}, and over which A[r]A[r] splits.

Fix some embedding ι0:KL\iota_{0}\colon K\rightarrow L. Fix a natural number cc. (The choice of cc will be made in the proof of Theorem 9.2, depending only on the numerics of certain Hodge numbers; everything we do until then is independent of the choice of cc.) Corollary 4.10 gives a torsion character χ0\chi_{0} of π1et(Aι0)\pi_{1}^{et}(A_{\iota_{0}}), of some order rr, satisfying a big monodromy condition.

In Lemma 5.29, we construct an H0H^{0}-algebra 𝖤I\mathsf{E}_{I} on 𝒪K,S\mathcal{O}_{K,S} and an 𝖤I\mathsf{E}_{I}-module 𝖵I\mathsf{V}_{I} on 𝒳\mathcal{X}, where II is the full (Gal×Galcyc/)(\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{cyc}/\mathbb{Q}})-orbit containing (ι0,χ0)(\iota_{0},\chi_{0}). The construction is roughly as follows. Any character χ\chi of π1et(Aι)\pi_{1}^{et}(A_{\iota}), defined over LL, defines a local system 𝖫χ\mathsf{L}_{\chi} on 𝒜L\mathcal{A}_{L}. By definition, 𝒴\mathcal{Y} is a subvariety of 𝒳𝒜\mathcal{X}\rightarrow\mathcal{A}; let

g:𝒴𝒜g\colon\mathcal{Y}\rightarrow\mathcal{A}

be the second projection. By Galois descent,

(ι,χ)IRkfιgι𝖫χ\bigoplus_{(\iota,\chi)\in I}R^{k}{f_{\iota}}_{*}g_{\iota}^{*}\mathsf{L}_{\chi}

descends to a Hodge–Deligne system 𝖵I\mathsf{V}_{I} on 𝒳\mathcal{X}, which is a module for the algebra

(ι,χ)I\bigoplus_{(\iota,\chi)\in I}\mathbb{Q}

which descends to an H0H^{0}-algebra 𝖤I\mathsf{E}_{I} on 𝒳\mathcal{X}. We’ll fix II, and suppress the subscript II from 𝖤I\mathsf{E}_{I} and 𝖵I\mathsf{V}_{I}.

In Section 5.7 we elaborate the structure of 𝖤\mathsf{E} and 𝖵\mathsf{V}. Let E0=pE_{0}=\mathbb{Q}_{p}, and let EE be the p\mathbb{Q}_{p}-algebra underlying either 𝖤et\mathsf{E}_{et} or 𝖤dR\mathsf{E}_{dR}. In Section 5.6, we define 𝐇\mathbf{H} to be one of the groups GLNGL_{N}, GSpNGSp_{N}, or GONGO_{N}, viewed as an algebraic group over EE. We take 𝐆0\mathbf{G}^{0} to be the Weil restriction 𝐆0=ResE0E𝐇\mathbf{G}^{0}={\operatorname{Res}^{E}_{E_{0}}}\mathbf{H}. This group 𝐆0\mathbf{G}^{0} has an action on a free EE-module VV (coming from the standard representation of GLNGL_{N}, GSpNGSp_{N}, or GONGO_{N}), and we take 𝐆\mathbf{G} to be the normalizer of 𝐆0\mathbf{G}^{0} in the group of E0E_{0}-linear automorphisms of VV. Whether the de Rham or étale version is meant is devoted by subscripts, as in 𝐆et\mathbf{G}_{et} and 𝐆dR\mathbf{G}_{dR}.

5.2. Hodge–Deligne systems

Definition 5.1.

Let kk be an integer, and qq a prime power. A rational qq-Weil number of weight kk is an algebraic number 222It is important that we allow Weil numbers that are not algebraic integers, since we want Hodge–Deligne systems to form a Tannakian category. In particular, we want the dual of a Hodge–Deligne system to again be a Hodge–Deligne system., all of whose conjugates have complex absolute value qk/2q^{k/2}.

An integral qq-Weil number is a rational qq-Weil number that is an algebraic integer.

When \ell is a prime of 𝒪K\mathcal{O}_{K}, we write qq_{\ell} for the cardinality of the residue field at \ell.

Definition 5.2.

Suppose given a number field KK with a chosen embedding KK\rightarrow\mathbb{C}. Let XX be a smooth variety over KK. Let SS be a finite set of primes of KK, and let 𝒳\mathcal{X} be a smooth model of XX over 𝒪K[1S]\mathcal{O}_{K}\left[\frac{1}{S}\right]. Let pp be a prime of \mathbb{Q} not lying below any place of SS, such that KK is unramified over pp; let vv be a place of KK lying over pp.

A Hodge–Deligne system 333The name is meant to evoke variations of Hodge structure and Deligne’s systems of realizations. on 𝒳\mathcal{X} at vv consists of the following structures:

  • A singular local system 𝖵Sing\mathsf{V}_{Sing} of \mathbb{Q}-vector spaces on XX_{\mathbb{C}}.

  • An étale local system 𝖵et\mathsf{V}_{et} of p\mathbb{Q}_{p}-vector spaces on 𝒳et×SpecSpec[1/p]\mathcal{X}_{\text{et}}\times_{\operatorname{Spec}\mathbb{Z}}\operatorname{Spec}\mathbb{Z}[1/p].

  • A vector bundle 𝖵dR\mathsf{V}_{dR} on XX, an integrable connection \nabla on 𝖵dR\mathsf{V}_{dR}, and a descending filtration Fili𝖵dR\operatorname{Fil}^{i}\mathsf{V}_{dR} of 𝖵dR\mathsf{V}_{dR} by subbundles

    𝖵dR=FilM𝖵dRFilM+1𝖵dRFilM𝖵dR=0\mathsf{V}_{dR}=\operatorname{Fil}^{-M}\mathsf{V}_{dR}\supseteq\operatorname{Fil}^{-M+1}\mathsf{V}_{dR}\supseteq\cdots\supseteq\operatorname{Fil}^{M}\mathsf{V}_{dR}=0

    (not necessarily \nabla-stable), each of which is locally a direct summand of 𝖵dR\mathsf{V}_{dR}.

  • A filtered FF-isocrystal 𝖵cris\mathsf{V}_{cris} on XKvX_{K_{v}} (see, for example, [64, end of §3.1]),

with the following isomorphisms:

  1. (1)

    An isomorphism on X,anX_{\mathbb{C},an} between 𝖵Singp\mathsf{V}_{Sing}\otimes_{\mathbb{Q}}\mathbb{Q}_{p} and the pullback of 𝖵et\mathsf{V}_{et} to X,anX_{\mathbb{C},an}. 444We do not use the étale-singular comparison; we could have left it out.

  2. (2)

    An isomorphism on X,anX_{\mathbb{C},an} between 𝖵SingmissingOX,an\mathsf{V}_{Sing}\otimes_{\mathbb{Q}}\mathcal{\mathcal{missing}}O_{X_{\mathbb{C},an}} and 𝖵dR𝒪X𝒪X,an\mathsf{V}_{dR}\otimes_{\mathcal{O}_{X}}\mathcal{O}_{X_{\mathbb{C},an}}.

  3. (3)

    An isomorphism on an open neighborhood of the rigid analytic generic fiber of XKvX_{K_{v}} between the underlying vector bundle to 𝖵cris\mathsf{V}_{cris} and the pullback of 𝖵dR\mathsf{V}_{dR}.

  4. (4)

    An isomorphism on XKv,proetX_{K_{v},proet} between the 𝒪𝔹cris\mathcal{O}\mathbb{B}_{cris}-modules 𝖵cris𝒪XKv𝒪𝔹cris\mathsf{V}_{cris}\otimes_{\mathcal{O}_{X_{K_{v}}}}\mathcal{O}\mathbb{B}_{cris} and 𝖵etp𝒪𝔹cris\mathsf{V}_{et}\otimes_{\mathbb{Q}_{p}}\mathcal{O}\mathbb{B}_{cris}.

and an increasing filtration WiW_{i} of all four objects, compatible with all the isomorphisms, such that all this data satisfies the axioms:

  • Fili𝖵dR\operatorname{Fil}^{i}\mathsf{V}_{dR} and \nabla satisfy Griffiths transversality.

  • The connection \nabla is induced under the isomorphism (2) by the trivial connection on 𝒪X\mathcal{O}_{X}.

  • For each point of XX_{\mathbb{C}}, the ii-th associated graded under WiW_{i} of the stalk of (𝖵Sing,𝖵DRK,Fili,(2))(\mathsf{V}_{Sing},\mathsf{V}_{DR}\otimes_{K}\mathbb{C},\operatorname{Fil}^{i},(2)) at that point is a pure Hodge structure of weight ii.

  • The ii-th associated graded under WiW_{i} of 𝖵et\mathsf{V}_{et} is pure of weight ii, i.e. for each closed point xx of 𝒳×SpecSpec[1/p]\mathcal{X}\times_{\operatorname{Spec}\mathbb{Z}}\operatorname{Spec}\mathbb{Z}[1/p] with residue field κx\kappa_{x}, the eigenvalues of Frobκx\operatorname{Frob}_{\kappa_{x}} on the ii-th associated graded of 𝖵et,x\mathsf{V}_{et,x} are |κx||\kappa_{x}|-Weil numbers of weight ii.

  • The ii-th associated graded of 𝖵cris\mathsf{V}_{cris} under WiW_{i} is pure of weight ii, i.e. for each closed point xx of 𝒳\mathcal{X} lying over pp with residue field κx\kappa_{x}, the eigenvalues of Frobenius on the ii-th associated graded of 𝖵cris,x\mathsf{V}_{cris,x} are |κx||\kappa_{x}|-Weil numbers of weight ii.

  • The connection \nabla has regular singularities in a smooth simple normal crossings compactification of XKX_{K}.

  • The isomorphism (3) is compatible with the connection.

  • The isomorphism (4) is compatible with connection, filtration, and Frobenius.

We note that the isomorphism (2) and the first three axioms make up the definition of a variation of Hodge structure; we will denote by 𝖵H\mathsf{V}_{H} the variation of Hodge structure given by 𝖵Sing\mathsf{V}_{Sing}, 𝖵dR\mathsf{V}_{dR}, Fili𝖵dR\operatorname{Fil}^{i}\mathsf{V}_{dR} and \nabla. The isomorphism (4) and the last axiom make up the definition of a crystalline local system [64, §1]. (Faltings calls these objects “dual-crystalline sheaves” [20, Theorem 2.6], at least in the situation where the Hodge–Tate weights are bounded between 0 and p2p-2.)

We say a Hodge-Deligne system is pure of weight ww if Ww1W_{w-1} vanishes and WwW_{w} is the whole system.

The rank of a Hodge–Deligne system 𝖵\mathsf{V} is the rank of the local system 𝖵sing\mathsf{V}_{sing} of \mathbb{Q}-vector spaces. By the various isomorphisms, this is equal to the ranks of 𝖵et\mathsf{V}_{et}, 𝖵dR\mathsf{V}_{dR}, and 𝖵cris\mathsf{V}_{cris}.

We will also need to work with polarized and integral variations of Hodge structure, and for that we need the following slight modifications of the notion of Hodge–Deligne system.

Definition 5.3.

Let KK, XX, SS, 𝒳\mathcal{X}, vv be as above. An integral Hodge–Deligne system on 𝒳\mathcal{X} consists of a Hodge-Deligne system on 𝒳\mathcal{X} together with an integral structure on 𝖵sing\mathsf{V}_{sing} (i.e. a singular local system 𝖵int\mathsf{V}_{int} of free \mathbb{Z}-modules on XX_{\mathbb{C}} together with an isomorphism 𝖵int𝖵sing\mathsf{V}_{int}\otimes_{\mathbb{Z}}\mathbb{Q}\cong\mathsf{V}_{sing}.)

Definition 5.4.

Let KK, XX, SS, 𝒳\mathcal{X}, vv be as above. A polarized Hodge-Deligne system on 𝒳\mathcal{X} consists of a Hodge-Deligne system on 𝒳\mathcal{X}, pure of some weight, together with a polarization of the variation of Hodge structures (𝖵sing,𝖵dR,(2))(\mathsf{V}_{sing},\mathsf{V}_{dR},(2)) (i.e. a morphism of local systems 𝖵sing𝖵sing\mathsf{V}_{sing}\otimes\mathsf{V}_{sing}\to\mathbb{Q} which restricted to the stalk at any point of X()X(\mathbb{C}) defines a polarization of the pure Hodge structure at that point.)

Definition 5.5.

We say that a Hodge–Deligne system 𝖵\mathsf{V} has integral Frobenius eigenvalues if the Weil numbers appearing as eigenvalues of Frobenius on 𝖵et,x\mathsf{V}_{et,x} and 𝖵cris,x\mathsf{V}_{cris,x} are integral, for all closed points x𝒳×SpecSpec[1/p]x\in\mathcal{X}\times_{\operatorname{Spec}\mathbb{Z}}\operatorname{Spec}\mathbb{Z}[1/p] and all closed points xx of 𝒳\mathcal{X} lying over pp, respectively.

Definition 5.6.

The differential Galois group of a Hodge–Deligne system 𝖵\mathsf{V} is the differential Galois group of the underlying vector bundle with connection 𝖵dR\mathsf{V}_{dR}. (For the definition of differential Galois group, see [54, §§1.2 – 1.4]. A vector bundle with connection gives rise to a linear differential equation; by the differential Galois group of the vector bundle with connection, we mean the differential Galois group of a Picard–Vessiot ring of the corresponding differential equation.)

The differential Galois group is the Zariski closure of the monodromy group of the variation of Hodge structure 𝖵H\mathsf{V}_{H}; this follows from the Riemann–Hilbert correspondence, and the fact that the period map has regular singularities along a smooth normal crossings compactification.

Remark 5.7.

Let kvk_{v} be the residue field of KK at vv. The “Frobenius automorphism” of KvK_{v} is the element of GalKv/p\operatorname{Gal}_{K_{v}/\mathbb{Q}_{p}} that acts as the pp-th power map on kvk_{v}.

A filtered FF-isocrystal 𝖵cris\mathsf{V}_{cris} gives, for every x¯𝒳(kv)\overline{x}\in\mathcal{X}(k_{v}), a pair

(Vx¯,ϕx¯),(V_{\overline{x}},\phi_{\overline{x}}),

where Vx¯V_{\overline{x}} is a KvK_{v}-vector space, and ϕx¯\phi_{\overline{x}} is an endomorphism of Vx¯V_{\overline{x}}, semilinear over Frobenius. Furthermore, for every x𝒳(𝒪Kv)x\in\mathcal{X}(\mathcal{O}_{K_{v}}) belonging to the residue class of x¯\overline{x}, the object 𝖵cris\mathsf{V}_{cris} defines a filtration on Vx¯V_{\overline{x}}, with an isomorphism to the filtered vector space 𝖵dRKv\mathsf{V}_{dR}\otimes K_{v}. We’ll call the resulting data

𝖵cris,x=(Vcris,x,ϕcris,x,Fcris,x).\mathsf{V}_{cris,x}=(V_{cris,x},\phi_{cris,x},F_{cris,x}).
Example 5.8.

(The trivial Hodge–Deligne system.)

Let KK and EE be number fields, and let SS be a finite set of places of KK. Take 𝒳=Spec𝒪K,S\mathcal{X}=\operatorname{Spec}\mathcal{O}_{K,S}, and define the trivial Hodge–Deligne system 𝖮E\mathsf{O}_{E} on Spec𝒪K,S\operatorname{Spec}\mathcal{O}_{K,S} by:

  • 𝖮E,sing=E\mathsf{O}_{E,sing}=E.

  • 𝖮E,et=Ep\mathsf{O}_{E,et}=E\otimes_{\mathbb{Q}}\mathbb{Q}_{p}, with the trivial Galois action.

  • 𝖮E,dR=EK\mathsf{O}_{E,dR}=E\otimes_{\mathbb{Q}}K, with trivial connection and filtration (i.e. Fil0𝖮E,dR=𝖮E,dR\operatorname{Fil}^{0}\mathsf{O}_{E,dR}=\mathsf{O}_{E,dR}, and Fil1𝖮E,dR=0\operatorname{Fil}^{1}\mathsf{O}_{E,dR}=0).

  • 𝖮E,cris\mathsf{O}_{E,cris} is determined by 𝖵dR\mathsf{V}_{dR} and the requirement that Frobenius act on EpKvE\otimes_{\mathbb{Q}_{p}}K_{v} through the trivial action on EE and the Frobenius automorphism of KvK_{v}.

  • The weight filtration is such that 𝖮E\mathsf{O}_{E} is concentrated in weight zero.

When XX is an arbitrary smooth KK-variety, we define the system 𝖮E\mathsf{O}_{E} on XX by pullback from SpecK\operatorname{Spec}K.

Remark 5.9.

In general, to give 𝖮E,cris\mathsf{O}_{E,cris} the structure of filtered FF-isocrystal on XKvX_{K_{v}}, it is enough to give a “Frobenius” isomorphism between the vector bundle with connection (𝖮E,dR,)(\mathsf{O}_{E,dR},\nabla) and its pullback under a lift of Frobenius to XKvX_{K_{v}}.

In Example 5.8, with 𝖮E,dR\mathsf{O}_{E,dR} constant and =0\nabla=0, the Frobenius is simply given as an automorphism of EpKvE\otimes_{\mathbb{Q}_{p}}K_{v}.

In general, Hodge–Deligne systems will come from families of varieties by taking cohomology.

Example 5.10.

(Pushforward of Hodge–Deligne systems.)

Let XX be a smooth variety over a number field KK, and let 𝒳\mathcal{X} be a smooth model for XX over 𝒪K,S\mathcal{O}_{K,S}, for some finite set SS of places of KK. Let π:𝒴𝒳\pi\colon\mathcal{Y}\rightarrow\mathcal{X} be a smooth, projective family of relative dimension nn; let YY be the base change of 𝒴\mathcal{Y} to KK. Let 𝖵\mathsf{V} be a Hodge–Deligne system on 𝒴\mathcal{Y}, and choose some kk with 0kn0\leq k\leq n. Let pp be a prime of \mathbb{Q} not lying below any place of SS, such that KK is unramified over pp, and let vv be a place of KK lying over pp. We define a Hodge–Deligne system 𝖶=𝖱kπ(𝖵)\mathsf{W}=\mathsf{R}^{k}\pi_{*}(\mathsf{V}) on 𝒳\mathcal{X} at vv, as follows.

  • Take 𝖶sing=Rkπ𝖵sing\mathsf{W}_{sing}=R^{k}\pi_{*}\mathsf{V}_{sing}, with the pushforward taken in the analytic topology on XX and YY. (This is again a local system, by Ehresmann’s theorem.)

  • Take 𝖶et=Rkπ𝖵et\mathsf{W}_{et}=R^{k}\pi_{*}\mathsf{V}_{et}, with the pushforward taken in the étale topology. (See [25] for an introduction to the étale topology.)

  • Take 𝖶dR\mathsf{W}_{dR} to be the relative de Rham cohomology of 𝖵dR\mathsf{V}_{dR} over XX, i.e. the pushforward of 𝖵dR\mathsf{V}_{dR} as a DD-module, with its Hodge filtration Fili\operatorname{Fil}^{i}. This is a filtered vector bundle by Hodge theory.

  • Take 𝖶cris\mathsf{W}_{cris} to be the vv-adic crystalline cohomology of YY. This is a filtered FF-isocrystal by pp-adic Hodge theory ([60, Thm. 8.8]). (See [9] or [11] for the construction of crystalline cohomology, and [51, §1] for its structure as filtered FF-isocrystal.)

  • The isomorphisms (1), (2), (3), (4) follow from Artin’s comparison theorem, Hodge theory, the de Rham-crystalline comparison, and relative pp-adic Hodge theory ([60, Thm. 8.8], [64, Thm. 5.5]) respectively.

  • The filtration WiW_{i} is induced from the filtration of 𝖵\mathsf{V} after shifting by kk. In particular, if 𝖵\mathsf{V} is pure of weight ww, then 𝖶\mathsf{W} is pure of weight w+kw+k.

Definition 5.11.

If 𝖵\mathsf{V} and 𝖶\mathsf{W} are two Hodge–Deligne systems on 𝒳\mathcal{X}, a morphism from 𝖵\mathsf{V} to 𝖶\mathsf{W} consists of:

  • A map of analytic local systems 𝖵sing𝖶sing\mathsf{V}_{sing}\rightarrow\mathsf{W}_{sing},

  • A map of étale local systems 𝖵et𝖶et\mathsf{V}_{et}\rightarrow\mathsf{W}_{et},

  • A map of vector bundles 𝖵dR𝖶dR\mathsf{V}_{dR}\rightarrow\mathsf{W}_{dR}, flat with respect to the connections on 𝖵dR\mathsf{V}_{dR} and 𝖶dR\mathsf{W}_{dR}, and respecting the filtrations Fil𝖵dR\operatorname{Fil}^{*}\mathsf{V}_{dR} and Fil𝖶dR\operatorname{Fil}^{*}\mathsf{W}_{dR}, and

  • A map of filtered F-isocrystals 𝖵cris𝖶cris\mathsf{V}_{cris}\rightarrow\mathsf{W}_{cris},

compatible with all the comparison isomorphisms (1), (2), (3), (4).

Lemma 5.12.

The Hodge–Deligne systems on 𝒳\mathcal{X} at vv form a Tannakian category with fiber functor given by 𝖵sing,x\mathsf{V}_{sing,x} for some x𝒳()x\in\mathcal{X}(\mathbb{C}).

In this Tannakian category, the tensor product of two systems will be defined by separately tensoring the individual objects 𝖵sing,𝖵et,𝖵dR,\mathsf{V}_{sing},\mathsf{V}_{et},\mathsf{V}_{dR}, and 𝖵cris\mathsf{V}_{cris}, and similarly for the dual of a system.

Proof.

Most of the argument is standard. Only two points require special attention.

The first is existence of a cokernel for fdRf_{dR}. In general the cokernel of a morphism of vector bundles need not be a vector bundle. But if f:𝖵𝖶f\colon\mathsf{V}\rightarrow\mathsf{W} is a morphism of Hodge–Deligne systems, the cokernel of fdR:𝖵dR𝖶dRf_{dR}\colon\mathsf{V}_{dR}\rightarrow\mathsf{W}_{dR} must be a vector bundle because fdRf_{dR} is a flat map of vector bundles with connection.

The second is the equality of images and coimages. A priori, the image of fdRf_{dR} has two possibly different filtrations, the image filtration and the coimage filtration. But these filtrations agree because variations of Hodge structure form an abelian category.

The remaining verifications are tedious but routine. ∎

If f:𝒳𝒳f\colon\mathcal{X}^{\prime}\rightarrow\mathcal{X} is a morphism, then for any system 𝖵\mathsf{V} on 𝒳\mathcal{X}, we can define the pullback f𝖵f^{*}\mathsf{V} on 𝒳\mathcal{X}^{\prime} by pulling back the four components separately. When the map ff is clear, we’ll sometimes write 𝖵|𝒳\mathsf{V}|_{\mathcal{X}^{\prime}} instead of f𝖵f^{*}\mathsf{V}.

Definition 5.13.

Let 𝒳\mathcal{X} be a smooth 𝒪K,S\mathcal{O}_{K,S}-scheme. A constant Hodge–Deligne system on 𝒳\mathcal{X} is a system of the form f𝖵f^{*}\mathsf{V}, where 𝖵\mathsf{V} is a Hodge–Deligne system on Spec𝒪K,S\operatorname{Spec}\mathcal{O}_{K,S}.

Constant Hodge-Deligne systems will be much more general than most notions of motives. For instance nothing prevents us from combining the étale and crystalline cohomology of one variety with the Hodge structure of a different variety, as long as they have the same Hodge numbers. Despite this, the notion of Hodge-Deligne system is strong enough for the arguments that we will make.

5.3. H0H^{0}-algebras

Throughout this section, KK will denote a number field, and SS a finite set of places of KK.

In order to make the arguments of [47] work, we need bounds on the centralizer of Frobenius.

The paper [47] works with Hodge–Deligne systems of the form 𝖱kπ(𝖮)\mathsf{R}^{k}\pi_{*}(\mathsf{O}_{\mathbb{Q}}), for π:𝒴𝒳\pi\colon\mathcal{Y}\rightarrow\mathcal{X} a family with not necessarily geometrically connected fibers. In this context, the zeroth cohomology H0(𝒴x)H^{0}(\mathcal{Y}_{x}) of a fiber (over any x𝒳(K)x\in\mathcal{X}(K)) has nontrivial Galois structure. The action of H0(𝒴x)H^{0}(\mathcal{Y}_{x}) on Hk(𝒴x)H^{k}(\mathcal{Y}_{x}) gives rise to the bounds we need on the Frobenius centralizer, by means of the semilinearity of Frobenius. Specifically, Hcrisk(𝒴x)H^{k}_{cris}(\mathcal{Y}_{x}) has a natural structure of Hcris0(𝒴x)H^{0}_{cris}(\mathcal{Y}_{x})-module, and the Frobenius endomorphism of Hcrisk(𝒴x)H^{k}_{cris}(\mathcal{Y}_{x}) is semilinear over the Frobenius endomorphism of Hcris0(𝒴x)H^{0}_{cris}(\mathcal{Y}_{x}).

We need an analogous statement the Frobenius centralizer in our situation. Let 𝒜\mathcal{A} be a smooth proper model of AA over 𝒪K,S\mathcal{O}_{K,S}. Suppose 𝒪K,S\mathcal{O}_{K,S} is finite étale over some [1/S]\mathbb{Z}[1/S^{\prime}], and 𝒳\mathcal{X} is smooth over [1/S]\mathbb{Z}[1/S^{\prime}]. Suppose 𝒴𝒳×𝒜\mathcal{Y}\subseteq\mathcal{X}\times_{\mathbb{Q}}\mathcal{A} is a smooth, proper, and flat over 𝒳\mathcal{X}; in the application, each fiber 𝒴x\mathcal{Y}_{x} will be a hypersurface in AA. An order-rr character χ\chi of π1(A)\pi_{1}(A), defined over some field LL, gives rise to a Hodge–Deligne system 𝖫χ\mathsf{L}_{\chi} on the base change of 𝒴x\mathcal{Y}_{x} to 𝒪L,S\mathcal{O}_{L,S} (Lemma 5.28); considering conjugates of χ\chi, we can descend from LL to \mathbb{Q} (Lemma 5.29). Taking cohomology, we will study a Hodge–Deligne system 𝖵\mathsf{V} on 𝒳\mathcal{X}, whose fiber over a point is (a descent to \mathbb{Q} of) Hk(𝒴x,𝖫χ)H^{k}(\mathcal{Y}_{x},\mathsf{L}_{\chi}). This will be an algebra over H0(𝒴x,𝖫χ)H^{0}(\mathcal{Y}_{x},\mathsf{L}_{\chi}); we will study this structure in detail.

The object H0(𝒴x,𝖫χ)H^{0}(\mathcal{Y}_{x},\mathsf{L}_{\chi}) is the cohomology of a motive with coefficients in [μr]\mathbb{Q}[\mu_{r}], defined over KK and having an algebra structure coming from the group scheme A[r]A[r]. The cohomology Hk(𝒴x,𝖫χ)H^{k}(\mathcal{Y}_{x},\mathsf{L}_{\chi}) is naturally a module over the stalk of 𝖫χ\mathsf{L}_{\chi} at the identity thanks to the compatibility of 𝖫χ\mathsf{L}_{\chi} with convolution ; the purpose of this section and the next is to clarify the structure of the stalk of 𝖫χ\mathsf{L}_{\chi} and modules over it.

To this end, we will define a general notion of “H0H^{0}-algebras.” Loosely speaking, an H0H^{0}-algebra is a weight-zero algebra object in the category of Hodge–Deligne systems. We will soon see that motives over extensions of KK (Example 5.21), motives with coefficients in a number field EE (Examples 5.15 and 5.23), and motives with an action of a finite abelian group (Examples 5.20 and 5.23) are all modules over various H0H^{0}-algebras. Our construction (Lemma 5.29) of Hodge–Deligne systems coming from families of hypersurfaces on abelian varieties will combine these ideas.

Definition 5.14.

A commutative H0H^{0}-algebra is a Hodge–Deligne system 𝖤\mathsf{E} on a smooth 𝒪K,S\mathcal{O}_{K,S}-scheme 𝒳\mathcal{X}, equipped with morphisms

e:𝖮𝖤e\colon\mathsf{O}_{\mathbb{Q}}\rightarrow\mathsf{E}

and

m:𝖤𝖤𝖤,m\colon\mathsf{E}\otimes\mathsf{E}\rightarrow\mathsf{E},

satisfying the following properties.

  • 𝖤\mathsf{E} is pure of weight 0.

  • The filtration on 𝖤dR\mathsf{E}_{dR} is trivial: Fil0𝖤dR=𝖤dR\operatorname{Fil}^{0}\mathsf{E}_{dR}=\mathsf{E}_{dR} and Fil1𝖤dR=0\operatorname{Fil}^{1}\mathsf{E}_{dR}=0.

  • The morphisms ee and mm make 𝖤\mathsf{E} into a commutative algebra object.

When we say “H0H^{0}-algebra”, we will mean “commutative H0H^{0}-algebra”.

Example 5.15.

(Trivial Hodge–Deligne system 𝖮E\mathsf{O}_{E}.)

Let KK and EE be number fields, and let SS be a finite set of places of KK. The trivial Hodge–Deligne system 𝖮E\mathsf{O}_{E} of Example 5.8 has an H0H^{0}-algebra structure coming functorially from the algebra structure on EE.

Definition 5.16.

If 𝒳=Spec𝒪K,S\mathcal{X}=\operatorname{Spec}\mathcal{O}_{K,S}, we say that 𝖤\mathsf{E} is étale if 𝖤sing\mathsf{E}_{sing} is an étale \mathbb{Q}-algebra.

Example 5.17.

(Étale H0H^{0}-algebras over a field.)

Let 𝒳=Spec𝒪K,S\mathcal{X}=\operatorname{Spec}\mathcal{O}_{K,S}, with KK a number field and SS a finite set of places of KK. In this setting we can give a concrete description of étale H0H^{0}-algebras 𝖤\mathsf{E} over Spec𝒪K,S\operatorname{Spec}\mathcal{O}_{K,S}.

The singular realization E=𝖤singE=\mathsf{E}_{sing} has the structure of \mathbb{Q}-algebra, which we assume is étale; 𝖤dR\mathsf{E}_{dR} is determined by

𝖤dR=EK\mathsf{E}_{dR}=E\otimes_{\mathbb{Q}}K

with trivial filtration.

The étale realization 𝖤et\mathsf{E}_{et} is the p\mathbb{Q}_{p}-algebra

Ep,E\otimes_{\mathbb{Q}}\mathbb{Q}_{p},

equipped with a continuous action of the Galois group GalK\operatorname{Gal}_{K}. By assumption, EE is an étale \mathbb{Q}-algebra, so Aut(Ep)\operatorname{Aut}(E\otimes_{\mathbb{Q}}\mathbb{Q}_{p}) is a finite group. The action of GalK\operatorname{Gal}_{K} descends to the maximal quotient GalK,S\operatorname{Gal}_{K,S^{\prime}} of GalK\operatorname{Gal}_{K} unramified outside the union of SS and the set of places of KK lying over pp.

Finally we turn to 𝖤cris\mathsf{E}_{cris}. The structure of KvK_{v}-algebra is given by an isomorphism

𝖤crisEpKv.\mathsf{E}_{cris}\cong E\otimes_{\mathbb{Q}_{p}}K_{v}.

The filtration is trivial, and the Frobenius (which we will notate σ\sigma) is the endomorphism σ1σ2\sigma_{1}\otimes\sigma_{2} of EpKvE\otimes_{\mathbb{Q}_{p}}K_{v}, where σ1\sigma_{1} gives the action of FrobvGK\operatorname{Frob}_{v}\in G_{K} on EE, and σ2\sigma_{2} is the endomorphism of KvK_{v} that acts as the pp-th power map on residue fields. The Frobenius FrobvGK\operatorname{Frob}_{v}\in G_{K} is only well-defined up to conjugacy, but that is enough to determine 𝖤cris\mathsf{E}_{cris} up to isomorphism.

Example 5.18.

(H0H^{0} of a family.)

If π:𝒴𝒳\pi\colon\mathcal{Y}\rightarrow\mathcal{X} is a proper map, then the degree-zero cohomology of 𝒴\mathcal{Y}, equipped with the cup product, gives an H0H^{0}-algebra on 𝒳\mathcal{X}.

The underlying Hodge-Deligne system is constructed as in Example 5.10 as 𝖱0π(𝖮)\mathsf{R}^{0}\pi_{*}(\mathsf{O}_{\mathbb{Q}}). The Hodge filtration is trivial because the Hodge filtration on HdR0H^{0}_{dR} of any smooth scheme is trivial. The map 𝖮𝖤\mathsf{O}_{\mathbb{Q}}\to\mathsf{E} is the unit of the adjunction between π\pi_{*} and π\pi^{*}, and the map 𝖤𝖤𝖤\mathsf{E}\otimes\mathsf{E}\to\mathsf{E} is given by cup product.

Example 5.19.

(Group algebra of a finite abelian group.)

For a finite abelian group GG, define the H0H^{0}-algebra 𝖮[G]\mathsf{O}[G] as follows.

  • 𝖮[G]sing=[G]¯\mathsf{O}[G]_{sing}=\underline{\mathbb{Q}[G]}.

  • 𝖮[G]et=p[G]¯\mathsf{O}[G]_{et}=\underline{\mathbb{Q}_{p}[G]}, with trivial Galois action.

  • 𝖮[G]dR=K[G]\mathsf{O}[G]_{dR}=K[G] is the trivial vector bundle, with trivial connection.

  • The filtration on 𝖮[G]dR\mathsf{O}[G]_{dR} is 𝖮[G]dR=Fil0Fil1=0\mathsf{O}[G]_{dR}=\operatorname{Fil}^{0}\supseteq\operatorname{Fil}^{1}=0.

  • The filtered FF-isocrystal is the constant vector bundle 𝒪[G]\mathcal{O}[G] with trivial connection. Its fiber at any point is the group algebra Kv[G]K_{v}[G], with Frobenius action coming from the Frobenius on KvK_{v}.

Note that, for a group GG, it is natural to view the group algebra [G]\mathbb{Q}[G] as a space of measures on GG, and thus dual to the space of functions on GG. Then the multiplication in the group algebra corresponds to convolution of measures. This suggests the right way to generalize the group algebra to group schemes, as the dual to their ring of functions. (Of course, the trace map makes their ring of functions self-dual.)

Example 5.20.

(Group algebra of a finite commutative group scheme over 𝒪K,S\mathcal{O}_{K,S}.)

Let GG be a finite étale commutative group scheme over some 𝒪K,S\mathcal{O}_{K,S}. The group operation G×GGG\times G\to G defines a Hopf algebra comultiplication Γ(G,𝒪G)Γ(G,𝒪G)Γ(G,𝒪G)\Gamma(G,\mathcal{O}_{G})\to\Gamma(G,\mathcal{O}_{G})\otimes\Gamma(G,\mathcal{O}_{G}). The dual map (Γ(G,𝒪G))(Γ(G,𝒪G))(Γ(G,𝒪G))\left(\Gamma(G,\mathcal{O}_{G})\right)^{\vee}\otimes\left(\Gamma(G,\mathcal{O}_{G})\right)^{\vee}\to\left(\Gamma(G,\mathcal{O}_{G})\right)^{\vee} gives (Γ(G,𝒪G))\left(\Gamma(G,\mathcal{O}_{G})\right)^{\vee} the structure of an 𝒪K,S\mathcal{O}_{K,S}-algebra.

Denoting by π:GSpec𝒪K,S\pi\colon G\rightarrow\operatorname{Spec}\mathcal{O}_{K,S} the structure map, we define 𝖤=(π𝖮|G)\mathsf{E}=(\pi_{*}\mathsf{O}_{\mathbb{Q}}|_{G})^{\vee}. A concrete description is as follows.

  • 𝖤sing=[G()]\mathsf{E}_{sing}=\mathbb{Q}[G(\mathbb{C})] with the usual algebra structure.

  • 𝖤et=p[G(K¯)]\mathsf{E}_{et}=\mathbb{Q}_{p}[G(\overline{K})] with its natural Galois action and structure of p\mathbb{Q}_{p}-algebra.

  • 𝖤dR=(Γ(GK,𝒪G))\mathsf{E}_{dR}=\left(\Gamma(G_{K},\mathcal{O}_{G})\right)^{\vee} with is natural algebra structure.

  • 𝖤cris=(Γ(GKv,𝒪GKv))\mathsf{E}_{cris}=\left(\Gamma(G_{K_{v}},\mathcal{O}_{G_{K_{v}}})\right)^{\vee} with its natural algebra structure and Frobenius coming from the Galois action on GG.

After passing to an extension of KK over which GG splits, an element aGa\in G gives an element of each of 𝖤sing\mathsf{E}_{sing}, 𝖤et\mathsf{E}_{et}, 𝖤dR\mathsf{E}_{dR}, and 𝖤cris\mathsf{E}_{cris}, and each of these four realizations is generated (as a vector space over the appropriate field) by GG.

Example 5.21.

Let L/KL/K be an extension of fields, and let SS and SS^{\prime} be finite sets of places of KK and LL, respectively, such that 𝒪L,S\mathcal{O}_{L,S^{\prime}} is finite étale over 𝒪K,S\mathcal{O}_{K,S}. Let π\pi be the map of schemes π:Spec𝒪L,SSpec𝒪K,S\pi\colon\operatorname{Spec}\mathcal{O}_{L,S^{\prime}}\rightarrow\operatorname{Spec}\mathcal{O}_{K,S}. If 𝖤\mathsf{E} is an H0H^{0}-algebra on 𝒪L,S\mathcal{O}_{L,S^{\prime}}, then π𝖤\pi_{*}\mathsf{E} is an H0H^{0}-algebra on 𝒪K,S\mathcal{O}_{K,S}.

5.4. Modules over an H0H^{0}-algebra

Definition 5.22.

Let 𝖤\mathsf{E} be an H0H^{0}-algebra. An 𝖤\mathsf{E}-module is an 𝖤\mathsf{E}-module object in the category of Hodge–Deligne systems; in other words, it is a Hodge–Deligne system 𝖵\mathsf{V} with a morphism m𝖵:𝖤𝖵𝖵m_{\mathsf{V}}\colon\mathsf{E}\otimes\mathsf{V}\rightarrow\mathsf{V}, such that the composition

𝖵𝖮𝖵e1𝖤𝖵m𝖵𝖵\mathsf{V}\cong\mathsf{O}_{\mathbb{Q}}\otimes\mathsf{V}\xrightarrow{e\otimes 1}\mathsf{E}\otimes\mathsf{V}\xrightarrow{m_{\mathsf{V}}}\mathsf{V}

is the identity map, and the diagram

(4) 𝖤𝖤𝖵\textstyle{\mathsf{E}\otimes\mathsf{E}\otimes\mathsf{V}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1m𝖵\scriptstyle{1\otimes m_{\mathsf{V}}}m1\scriptstyle{m\otimes 1}𝖤𝖵\textstyle{\mathsf{E}\otimes\mathsf{V}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m\scriptstyle{m}𝖤𝖵\textstyle{\mathsf{E}\otimes\mathsf{V}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m𝖵\scriptstyle{m_{\mathsf{V}}}𝖵\textstyle{\mathsf{V}}

commutes.

Remark 5.23.

An 𝖮E\mathsf{O}_{E}-module (Example 5.15) is a motive with coefficients in EE. An 𝖮[G]\mathsf{O}[G]-module (Example 5.20) is a motive with an action of the group scheme GG.

If 𝖵\mathsf{V} and 𝖶\mathsf{W} are 𝖤\mathsf{E}-modules, then we define the tensor product

𝖵𝖤𝖶\mathsf{V}\otimes_{\mathsf{E}}\mathsf{W}

as the coequalizer of two maps

𝖤𝖵𝖶𝖵𝖶,\mathsf{E}\otimes\mathsf{V}\otimes\mathsf{W}\rightrightarrows\mathsf{V}\otimes\mathsf{W},

the first of which is induced from 𝖤𝖵𝖵\mathsf{E}\otimes\mathsf{V}\rightarrow\mathsf{V}, and the second from 𝖤𝖶𝖶\mathsf{E}\otimes\mathsf{W}\rightarrow\mathsf{W}. (See [14, §2.3].) Then 𝖵𝖤𝖶\mathsf{V}\otimes_{\mathsf{E}}\mathsf{W} also has the structure of 𝖤\mathsf{E}-module.

Definition 5.24.

Let EE be a finite étale algebra over a field E0E_{0}. We say an EE-module VV is equidimensional if it is free of finite rank.

Equivalently, writing EE as a product of fields EiE_{i}, we say that VV is equidimensional if

dimEi(VEEi)\dim_{E_{i}}(V\otimes_{E}E_{i})

is independent of ii.

In this case, we call that dimension the rank of VV.

Definition 5.25.

Suppose 𝖤\mathsf{E} is a constant étale H0H^{0}-algebra, and 𝖵\mathsf{V} is an 𝖤\mathsf{E}-module. We say 𝖵\mathsf{V} is equidimensional if the stalks of 𝖵sing\mathsf{V}_{sing} are equidimensional modules over the stalks of 𝖤sing\mathsf{E}_{sing}. In this case, the ranks of the stalks as 𝖤sing\mathsf{E}_{sing}-modules are constant; we call that rank the rank of 𝖵\mathsf{V}.

Lemma 5.26.

If 𝖵\mathsf{V} is an equidimensional 𝖤\mathsf{E}-module of rank NN on XX, then the following statements hold.

  • The stalks of 𝖵et\mathsf{V}_{et} are equidimensional modules of rank NN over the stalks of 𝖤et\mathsf{E}_{et}.

  • The stalks of 𝖵dR\mathsf{V}_{dR} are equidimensional modules of rank NN over the stalks of 𝖤dR\mathsf{E}_{dR}.

  • The stalks of 𝖵cris\mathsf{V}_{cris} are equidimensional modules of rank NN over the stalks of 𝖤cris\mathsf{E}_{cris}.

Proof.

Follows from the “comparison isomorphisms” in the definition of Hodge–Deligne system. ∎

Definition 5.27.

Suppose 𝖵\mathsf{V} is an equidimensional 𝖤\mathsf{E}-module of rank NN. Then we say that 𝖵\mathsf{V} is an 𝖤\mathsf{E}-module with GLNGL_{N}-structure, or has GLNGL_{N}-structure over 𝖤\mathsf{E}.

Suppose additionally that there exists a Hodge-Deligne system 𝖫\mathsf{L} of rank 1, with a nondegenerate bilinear pairing

𝖵𝖤𝖵𝖫.\mathsf{V}\otimes_{\mathsf{E}}\mathsf{V}\rightarrow\mathsf{L}.

If the pairing is alternating, we say that 𝖵\mathsf{V} has GSpNGSp_{N}-structure over 𝖤\mathsf{E}; if symmetric, we say that 𝖵\mathsf{V} has GONGO_{N}-structure over 𝖤\mathsf{E}.

In Lemma 5.30, we will see that if YX×AY\subseteq X\times_{\mathbb{Q}}A is equal to a translate of [1]Y[-1]^{*}Y, then the cup-product pairing gives either GSpNGSp_{N}-structure or GONGO_{N}-structure on its middle cohomology; here we take 𝖫\mathsf{L} to be the top-degree cohomology of YY.

Below (Definition 5.38) we will define a notion of “object with GG-structure” in a Tannakian category. The Tannakian notion (Definition 5.38) only applies to objects of an EE-linear tensor category, with EE a field; and 𝖤\mathsf{E} is not a field. We will use the words “over 𝖤\mathsf{E}” to emphasize this distinction. The two notions are compatible in that, when 𝖤Sing\mathsf{E}_{Sing} is a field, viewed as a constant local system, a GSpNGSp_{N}-structure on 𝖵\mathsf{V} over 𝖤\mathsf{E} gives a GSpnGSp_{n}-structure on 𝖵Sing\mathsf{V}_{Sing} over 𝖤Sing\mathsf{E}_{Sing}, and similar statements are true for GONGO_{N} and for the other realizations. In Section 5.7 we’ll give a detailed description of these objects in Tannakian terms.

5.5. Local systems on an abelian variety: construction of a Hodge–Deligne system

Let AA be an abelian variety of dimension nn over a number field KK, XX an arbitrary smooth variety over \mathbb{Q}, and XKX_{K} its base change to KK. Let SS be a finite set of primes of 𝒪K\mathcal{O}_{K}, and SS^{\prime} a finite set of primes of \mathbb{Z}, such that 𝒪K,S\mathcal{O}_{K,S} is finite étale over [1/S]\mathbb{Z}[1/S^{\prime}]. Let 𝒜\mathcal{A} be a smooth proper model for AA over 𝒪K,S\mathcal{O}_{K,S}, and 𝒳\mathcal{X} a smooth model for XX over [1/S]\mathbb{Z}[1/S^{\prime}]. Let

𝒴𝒳×[1/S]𝒜=𝒳𝒪K,S×𝒪K,S𝒜\mathcal{Y}\subseteq\mathcal{X}\times_{\mathbb{Z}[1/S^{\prime}]}\mathcal{A}=\mathcal{X}_{\mathcal{O}_{K,S}}\times_{\mathcal{O}_{K,S}}\mathcal{A}

be a subscheme, smooth, proper and flat over 𝒳𝒪K,S\mathcal{X}_{\mathcal{O}_{K},S}. Let f:𝒴𝒳f\colon\mathcal{Y}\rightarrow\mathcal{X} and g:𝒴𝒜g\colon\mathcal{Y}\rightarrow\mathcal{A} be the projections. Fix a prime pp over which KK is unramified and a positive integer rr prime to pp. Let LL be a field, containing KK, Galois over \mathbb{Q}, and over which A[r]A[r] splits; we suppose that SS has been chosen so that 𝒪L,S\mathcal{O}_{L,S} is unramified over 𝒪K,S\mathcal{O}_{K,S}. Let χ\chi be an order-rr character of π1et(A)\pi_{1}^{et}(A); let χ\mathcal{L}_{\chi} be the corresponding character sheaf on 𝒜\mathcal{A}. (It is a p[μr]\mathbb{Q}_{p}[\mu_{r}]-local system on the étale site of 𝒜𝒪L,S[1/p]\mathcal{A}_{\mathcal{O}_{L,S}[1/p]}.) Let

k=n1=dimA1=dimY.k=n-1=\dim A-1=\dim Y.

We want to create a Hodge–Deligne system on 𝒳\mathcal{X} whose base change to 𝒪L,S\mathcal{O}_{L,S} has RkfgχR^{k}f_{*}g^{*}\mathcal{L}_{\chi} as a direct summand.

The tensor product 𝒪K,S[1/S]𝒪L,S\mathcal{O}_{K,S}\otimes_{\mathbb{Z}[1/S^{\prime}]}\mathcal{O}_{L,S} splits as a direct sum

𝒪K,S[1/S]𝒪L,Sι𝒪L,S(ι),\mathcal{O}_{K,S}\otimes_{\mathbb{Z}[1/S^{\prime}]}\mathcal{O}_{L,S}\cong\bigoplus_{\iota}\mathcal{O}_{L,S}^{(\iota)},

indexed by the [K:][K\colon\mathbb{Q}] embeddings of KK into LL. Here each 𝒪L,S(ι)\mathcal{O}_{L,S}^{(\iota)} is an isomorphic copy of 𝒪L,S\mathcal{O}_{L,S}; the superscript (ι)(\iota) is merely an index. We have the corresponding splitting

𝒜×Spec[1/S]Spec𝒪L,Sι𝒜ι,\mathcal{A}\times_{\operatorname{Spec}\mathbb{Z}[1/S^{\prime}]}\operatorname{Spec}\mathcal{O}_{L,S}\cong\coprod_{\iota}\mathcal{A}_{\iota},

where for each ι\iota, we define

𝒜ι=𝒜×Spec𝒪K,S;ιSpec𝒪L,S.\mathcal{A}_{\iota}=\mathcal{A}\times_{\operatorname{Spec}\mathcal{O}_{K,S};\iota}\operatorname{Spec}\mathcal{O}_{L,S}.

Similarly, define

𝒴ι=𝒴×Spec𝒪K,S;ιSpec𝒪L,S,\mathcal{Y}_{\iota}=\mathcal{Y}\times_{\operatorname{Spec}\mathcal{O}_{K,S};\iota}\operatorname{Spec}\mathcal{O}_{L,S},

the base change of 𝒴\mathcal{Y} along ι\iota. Then we have the Cartesian diagram

ι𝒴ι\textstyle{\coprod_{\iota}\mathcal{Y}_{\iota}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒴\textstyle{\mathcal{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι𝒳𝒪L,S×𝒪L,S𝒜ι\textstyle{\coprod_{\iota}\mathcal{X}_{\mathcal{O}_{L,S}}\times_{\mathcal{O}_{L,S}}\mathcal{A}_{\iota}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳×𝒜\textstyle{\mathcal{X}\times\mathcal{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳𝒪L,S\textstyle{\mathcal{X}_{\mathcal{O}_{L,S}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳.\textstyle{\mathcal{X}.}

Let fι:𝒴ι𝒳𝒪L,Sf_{\iota}\colon\mathcal{Y}_{\iota}\rightarrow\mathcal{X}_{\mathcal{O}_{L,S}} and gι:𝒴ι𝒜ιg_{\iota}\colon\mathcal{Y}_{\iota}\rightarrow\mathcal{A}_{\iota} be the projections.

Let rr be prime to pp, and let ΠK/(A)[r]\Pi^{K/\mathbb{Q}}(A)[r] be the set of all pairs (ι,χ)(\iota,\chi), where ι:KL\iota\colon K\rightarrow L is a \mathbb{Q}-linear embedding, and χ\chi is a character of π1(Aι)\pi_{1}(A_{\iota}) of order dividing rr. For fixed ι\iota, the set of characters χ\chi is naturally identified with the set of homomorphisms

Aι[r]μr;A_{\iota}[r]\rightarrow\mu_{r};

thus, ΠK/(A)[r]\Pi^{K/\mathbb{Q}}(A)[r] has a natural action of

Gal×Gal[μr]/,\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}[\mu_{r}]/\mathbb{Q}},

where Gal\operatorname{Gal}_{\mathbb{Q}} acts on the pairs (ι,χ)(\iota,\chi) via its action on LL, and Gal[μr]/\operatorname{Gal}_{\mathbb{Q}[\mu_{r}]/\mathbb{Q}} acts on murmu_{r}.

Lemma 5.28.

Let χ\chi be character of π1et(A)\pi_{1}^{et}(A) of some finite order rr. Then there exists an 𝖮[μr]\mathsf{O}_{\mathbb{Q}[\mu_{r}]}-module 𝖫χ\mathsf{L}_{\chi} on 𝒜|𝒪L,S\mathcal{A}|_{\mathcal{O}_{L,S}} such that (𝖫χ)et(\mathsf{L}_{\chi})_{et} is the character sheaf associated with the character χ\chi, and (𝖫χ)sing(\mathsf{L}_{\chi})_{sing} is the analytic [μr]\mathbb{Q}[\mu_{r}]-local system associated with χ\chi.

Proof.

Descent. ∎

Lemma 5.29.

(Construction of 𝖤I\mathsf{E}_{I} and 𝖵I\mathsf{V}_{I}.)

Let II be an orbit of Gal×Gal[μr]/\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}[\mu_{r}]/\mathbb{Q}} on ΠK/(A)[r]\Pi^{K/\mathbb{Q}}(A)[r]. There exist an H0H^{0}-algebra 𝖤I\mathsf{E}_{I} on 𝒪K,S\mathcal{O}_{K,S} and an 𝖤I\mathsf{E}_{I}-module 𝖵I\mathsf{V}_{I} on 𝒳\mathcal{X} with the following properties.

  • After base change to 𝒪L,S\mathcal{O}_{L,S} and extension of coefficients to [μr]\mathbb{Q}[\mu_{r}], we have the direct sum decomposition

    𝖤I|𝒪L,S𝖮𝖮[μr](ι,χ)I𝖮[μr](ι,χ),\mathsf{E}_{I}|_{\mathcal{O}_{L,S}}\otimes_{\mathsf{O}_{\mathbb{Q}}}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}\cong\bigoplus_{(\iota,\chi)\in I}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}^{(\iota,\chi)},

    where each 𝖮[μr](ι,χ)\mathsf{O}_{\mathbb{Q}[\mu_{r}]}^{(\iota,\chi)} is a copy of 𝖮[μr]\mathsf{O}_{\mathbb{Q}[\mu_{r}]}.

  • The Galois representation 𝖤I,et\mathsf{E}_{I,et} is compatible with the isomorphism

    𝖤I,et[μr](ι,χ)Ip[μr],\mathsf{E}_{I,et}\otimes_{\mathbb{Q}}\mathbb{Q}[\mu_{r}]\cong\bigoplus_{(\iota,\chi)\in I}\mathbb{Q}_{p}\otimes_{\mathbb{Q}}\mathbb{Q}[\mu_{r}],

    where GalK\operatorname{Gal}_{K} acts on the right-hand side by permutation of the characters χ\chi.

    Similarly, the Frobenius endomorphism of 𝖤I,cris\mathsf{E}_{I,cris} is compatible with the isomorphism

    𝖤I,cris[μr](ι,χ)Ip[μr],\mathsf{E}_{I,cris}\otimes_{\mathbb{Q}}\mathbb{Q}[\mu_{r}]\cong\bigoplus_{(\iota,\chi)\in I}\mathbb{Q}_{p}\otimes_{\mathbb{Q}}\mathbb{Q}[\mu_{r}],

    where Frobenius acts on the right-hand side by permuting the pairs (ι,χ)(\iota,\chi), via the Gal\operatorname{Gal}_{\mathbb{Q}} action, with trivial action on [μr]\mathbb{Q}[\mu_{r}].

  • After base change to 𝒪L,S\mathcal{O}_{L,S} and extension of coefficients to [μr]\mathbb{Q}[\mu_{r}], the module 𝖵I\mathsf{V}_{I} decomposes as the direct sum

    𝖵I|𝒪L,S𝖮𝖮[μr]=(ι,χ)IRkfιgι𝖫χ.\mathsf{V}_{I}|_{\mathcal{O}_{L,S}}\otimes_{\mathsf{O}_{\mathbb{Q}}}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}=\bigoplus_{(\iota,\chi)\in I}R^{k}{f_{\iota}}_{*}g_{\iota}^{*}\mathsf{L}_{\chi}.

    Furthermore, this decomposition is compatible with the decomposition of

    𝖤I|L𝖮𝖮[μr]\mathsf{E}_{I}|_{L}\otimes_{\mathsf{O}_{\mathbb{Q}}}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}

    into fields.

  • 𝖵I\mathsf{V}_{I} can be made into a polarized, integral Hodge–Deligne system.

Proof.

We have the map

idX×[r]:𝒳×[1/S]𝒜𝒳×[1/S]𝒜id_{X}\times[r]\colon\mathcal{X}\times_{\mathbb{Z}[1/S^{\prime}]}\mathcal{A}\rightarrow\mathcal{X}\times_{\mathbb{Z}[1/S^{\prime}]}\mathcal{A}

where [r]:𝒜𝒜[r]\colon\mathcal{A}\rightarrow\mathcal{A} is multiplication by rr; let

𝒴r=(id𝒳×[r])1(𝒴).\mathcal{Y}_{r}=(id_{\mathcal{X}}\times[r])^{-1}(\mathcal{Y}).
𝒴r\textstyle{\mathcal{Y}_{r}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h}𝒴\textstyle{\mathcal{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}-\crvi-\crvi-\crvif\scriptstyle{f}𝒳×𝒜\textstyle{\mathcal{X}\times\mathcal{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id𝒳×[r]\scriptstyle{id_{\mathcal{X}}\times[r]}𝒳×𝒜\textstyle{\mathcal{X}\times\mathcal{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒜\textstyle{\mathcal{A}}𝒳\textstyle{\mathcal{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}𝒳\textstyle{\mathcal{X}}

We have an isomorphism of Hodge–Deligne systems

Rkh𝖮Rkfg[r]𝖮.R^{k}h_{*}\mathsf{O}_{\mathbb{Q}}\cong R^{k}f_{*}g^{*}[r]_{*}\mathsf{O}_{\mathbb{Q}}.

Furthermore, on 𝒜×Spec[1/S]Spec𝒪L,S\mathcal{A}\times_{\operatorname{Spec}\mathbb{Z}[1/S^{\prime}]}\operatorname{Spec}\mathcal{O}_{L,S}, we have the direct sum decomposition of Hodge–Deligne systems

[r]𝖮|𝒜×Spec[1/S]Spec𝒪L,S𝖮𝖮[μr](ι,χ)ΠK/(A)[r]𝖫ι,χ,[r]_{*}\mathsf{O}_{\mathbb{Q}}|_{\mathcal{A}\times_{\operatorname{Spec}\mathbb{Z}[1/S^{\prime}]}\operatorname{Spec}\mathcal{O}_{L,S}}\otimes_{\mathsf{O}_{\mathbb{Q}}}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}\cong\bigoplus_{(\iota,\chi)\in\Pi^{K/\mathbb{Q}}(A)[r]}\mathsf{L}_{\iota,\chi},

where 𝖫ι,χ\mathsf{L}_{\iota,\chi} is the system 𝖫χ\mathsf{L}_{\chi} on 𝒜ι\mathcal{A}_{\iota}, and the trivial system on all other components of 𝒜×Spec[1/S]Spec𝒪L,S\mathcal{A}\times_{\operatorname{Spec}\mathbb{Z}[1/S^{\prime}]}\operatorname{Spec}\mathcal{O}_{L,S}. This gives a decomposition on 𝒳𝒪L,S\mathcal{X}_{\mathcal{O}_{L,S}}

(Rkh𝖮)|𝒪L,S𝖮[μr]=(ι,χ)Rkfg𝖫ι,χ=(ι,χ)Rkfιgι𝖫χ.(R^{k}h_{*}\mathsf{O}_{\mathbb{Q}})|_{\mathcal{O}_{L,S}}\otimes\mathsf{O}_{\mathbb{Q}}[\mu_{r}]=\bigoplus_{(\iota,\chi)}R^{k}f_{*}g^{*}\mathsf{L}_{\iota,\chi}=\bigoplus_{(\iota,\chi)}R^{k}{f_{\iota}}_{*}g_{\iota}^{*}\mathsf{L}_{\chi}.

Let

𝖵=Rkh𝖮Rkfg[r]𝖮.\mathsf{V}=R^{k}h_{*}\mathsf{O}_{\mathbb{Q}}\cong R^{k}f_{*}g^{*}[r]_{*}\mathsf{O}_{\mathbb{Q}}.

Let 𝖤\mathsf{E} be the pullback to 𝒳\mathcal{X} of the Hodge–Deligne system on [1/S]\mathbb{Z}[1/S^{\prime}] coming from

𝖤=(π𝖮|A[r]),\mathsf{E}=(\pi_{*}\mathsf{O}_{\mathbb{Q}}|_{A[r]})^{\vee},

where π:SpecKSpec\pi\colon\operatorname{Spec}K\rightarrow\operatorname{Spec}\mathbb{Q} is the projection. This 𝖤\mathsf{E} has a structure of H0H^{0}-algebra, coming from the structure of group scheme on A[r]A[r] (see Examples 5.20 and 5.21). Furthermore, the group action 𝒜[r]×𝒴r𝒴r\mathcal{A}[r]\times\mathcal{Y}_{r}\rightarrow\mathcal{Y}_{r} makes 𝖵\mathsf{V} into a module over 𝖤\mathsf{E}.

After base change from [1/S]\mathbb{Z}[1/S^{\prime}] to 𝒪L,S\mathcal{O}_{L,S} and extension of coefficients, we have a decomposition of H0H^{0}-algebras on 𝒳\mathcal{X}

𝖤|𝒪L,S𝖮[μr](ι,χ)ΠK/(A)[r]𝖮[μr](ι,χ).\mathsf{E}|_{\mathcal{O}_{L,S}}\otimes_{\mathbb{Q}}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}\cong\bigoplus_{(\iota,\chi)\in\Pi^{K/\mathbb{Q}}(A)[r]}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}^{(\iota,\chi)}.

The set ΠK/(A)[r]\Pi^{K/\mathbb{Q}}(A)[r] has commuting actions of Gal\operatorname{Gal}_{\mathbb{Q}} and Galcyc/\operatorname{Gal}_{\mathbb{Q}^{cyc}/\mathbb{Q}}, the former coming from the base field \mathbb{Q}, and the latter from the coefficient field cyc\mathbb{Q}^{cyc}. For each orbit II of Gal×Galcyc/\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{cyc}/\mathbb{Q}}, the direct sum (ι,χ)I𝖮[μr](ι,χ)\bigoplus_{(\iota,\chi)\in I}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}^{(\iota,\chi)} descends to an H0H^{0}-algebra over \mathbb{Q}, with \mathbb{Q}-coefficients. More precisely, choose any (ι0,χ0)I(\iota_{0},\chi_{0})\in I, and let 𝖤I\mathsf{E}_{I} be the pushforward of 𝖮[μr]\mathsf{O}_{\mathbb{Q}[\mu_{r}]} from KK to \mathbb{Q}. Then there is an isomorphism

𝖤I|𝒪L,S𝖮[μr](ι,χ)I𝖮[μr](ι,χ).\mathsf{E}_{I}|_{\mathcal{O}_{L,S}}\otimes\mathsf{O}_{\mathbb{Q}[\mu_{r}]}\cong\bigoplus_{(\iota,\chi)\in I}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}^{(\iota,\chi)}.

Over \mathbb{Q} and with \mathbb{Q}-coefficients, 𝖤\mathsf{E} splits as the direct sum of the algebras 𝖤I\mathsf{E}_{I}.

The 𝖤\mathsf{E}-module 𝖵\mathsf{V} also splits as the direct sum of the objects 𝖵I=𝖵𝖤𝖤I\mathsf{V}_{I}=\mathsf{V}\otimes_{\mathsf{E}}\mathsf{E}_{I}, and after base change and extension of coefficients we have

𝖵I|𝒪L,S𝖮𝖮[μr]=(ι,χ)IRkfιgι𝖫χ.\mathsf{V}_{I}|_{\mathcal{O}_{L,S}}\otimes_{\mathsf{O}_{\mathbb{Q}}}\mathsf{O}_{\mathbb{Q}[\mu_{r}]}=\bigoplus_{(\iota,\chi)\in I}R^{k}{f_{\iota}}_{*}g_{\iota}^{*}\mathsf{L}_{\chi}.

Finally we have to explain the polarization and integral structure on 𝖵I,sing\mathsf{V}_{I,sing}. On 𝖵sing\mathsf{V}_{sing} we have a standard polarization and integral structure: the polarization comes from Poincaré duality, and the integral structure is simply the integral structure on the singular cohomology of 𝒴r\mathcal{Y}_{r}. These structures induce a polarization and integral structure on 𝖵I,sing\mathsf{V}_{I,sing}, by restricting the polarization and intersecting the integral lattice with 𝖵I,sing\mathsf{V}_{I,sing}. ∎

Lemma 5.30.

Fix notation as in Lemma 5.29. Then 𝖵I\mathsf{V}_{I} is an 𝖤I\mathsf{E}_{I}-module with GLNGL_{N}-structure, in the sense of Definition 5.27.

Furthermore, if YY is equal to a translate of [1]Y[-1]^{*}Y, then 𝖵I\mathsf{V}_{I} has GSpNGSp_{N}-structure over 𝖤I\mathsf{E}_{I} if nn is even and GONGO_{N}-structure over 𝖤I\mathsf{E}_{I} if nn is odd.

Note that [1]Y[-1]^{*}Y is dual to YY in the Tannakian formalism; see 3.11.

Proof.

To prove that 𝖵I\mathsf{V}_{I} has GLNGL_{N}-structure we only need to check that 𝖵I\mathsf{V}_{I} is equidimensional; this is a consequence of the transitive Galois action on the index set II.

Suppose YY is equal to a translate of [1]Y[-1]^{*}Y (so also 𝒴\mathcal{Y} is equal to a translate of [1]𝒴[-1]^{*}\mathcal{Y}). This equality gives an involution

ι:𝖵𝖵.\iota\colon\mathsf{V}\rightarrow\mathsf{V}.

The cup product pairing and the trace map compose to give a map

,:𝖵𝖵=Rkh𝖮Rkh𝖮R2kh𝖮𝖮(k).\langle-,-\rangle\colon\mathsf{V}\otimes\mathsf{V}=R^{k}h_{*}\mathsf{O}_{\mathbb{Q}}\otimes R^{k}h_{*}\mathsf{O}_{\mathbb{Q}}\rightarrow R^{2k}h_{*}\mathsf{O}_{\mathbb{Q}}\rightarrow\mathsf{O}_{\mathbb{Q}}(-k).

This pairing does not factor through 𝖵𝖤𝖵\mathsf{V}\otimes_{\mathsf{E}}\mathsf{V}, which would be equivalent to the identity

(5) ev,w=v,ew\langle ev,w\rangle=\langle v,ew\rangle

of maps 𝖵×𝖤×𝖵𝖮(k)\mathsf{V}\times\mathsf{E}\times\mathsf{V}\to\mathsf{O}_{\mathbb{Q}}(-k) (where v,e,v,e, and ww are local sections of the sheaves underlying 𝖵,𝖤,\mathsf{V},\mathsf{E}, and 𝖵\mathsf{V}.) Instead, for aA[r]a\in A[r], the pairing satisfies

v,w=av,aw.\langle v,w\rangle=\langle av,aw\rangle.

However, when YY is equal to a translate of [1]Y[-1]^{*}Y, and ι\iota is the involution of 𝖵\mathsf{V} described above, we can form the pairing

(vw)v,ιw(v\otimes w)\mapsto\langle v,\iota w\rangle

as the composition

𝖵𝖵1ι𝖵𝖵,R2kfg𝖮|Y.\mathsf{V}\otimes\mathsf{V}\stackrel{{\scriptstyle 1\otimes\iota}}{{\rightarrow}}\mathsf{V}\otimes\mathsf{V}\stackrel{{\scriptstyle\langle-,-\rangle}}{{\rightarrow}}R^{2k}f_{*}g^{*}\mathsf{O}_{\mathbb{Q}}|_{Y}.

We claim that this pairing does satisfy (5), so that it descends to a pairing

𝖵𝖤𝖵R2kfg𝖮|Y.\mathsf{V}\otimes_{\mathsf{E}}\mathsf{V}\rightarrow R^{2k}f_{*}g^{*}\mathsf{O}_{\mathbb{Q}}|_{Y}.

It is enough to check (5) after extending both the base field and the coefficient field, so we can assume 𝖤\mathsf{E} is the group algebra of the finite abelian group A[r]A[r], and that 𝖤\mathsf{E} splits as a direct sum over characters of that group. Now (5) follows from

av,ιw=v,a1ιw=v,ι(aw),\langle av,\iota w\rangle=\langle v,a^{-1}\iota w\rangle=\langle v,\iota(aw)\rangle,

where aA[r]a\in A[r].

We have

w,ιv=ιw,ι2v=ιw,v=(1)kv,ιw\langle w,\iota v\rangle=\langle\iota w,\iota^{2}v\rangle=\langle\iota w,v\rangle=(-1)^{k}\langle v,\iota w\rangle

so this pairing is symplectic if nn is even (and thus kk is odd) and symmetric if nn is odd (and thus kk is even). ∎

For a fixed (Gal×Galcyc/)(\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{cyc}/\mathbb{Q}})-orbit II on ΠK/(A)[r]\Pi^{K/\mathbb{Q}}(A)[r], define the reductive group 𝐇\mathbf{H} as follows. Let NN be the rank of 𝖵I\mathsf{V}_{I}. If YY is equal to a translate of [1]Y[-1]^{*}Y and nn is even, let 𝐇=GSpN\mathbf{H}=GSp_{N}. If YY is equal to a translate of [1]Y[-1]^{*}Y and nn is odd, let 𝐇=GON\mathbf{H}=GO_{N}. Otherwise, let 𝐇=GLN\mathbf{H}=GL_{N}. Then by Lemma 5.30, 𝖵I\mathsf{V}_{I} has 𝐇\mathbf{H}-structure over 𝖤I\mathsf{E}_{I}. In the following sections we will analyze this structure in some detail.

5.6. An algebraic group

Let E0E_{0} be a field – we will mostly be interested in the case E0=pE_{0}=\mathbb{Q}_{p} – and let EE be a finite étale E0E_{0}-algebra. In this subsection we will define some groups 𝐆0\mathbf{G}^{0} and 𝐆\mathbf{G}. Specializing EE to 𝖤et\mathsf{E}_{et}, the group 𝐆\mathbf{G} will contain the monodromy group of 𝖵et\mathsf{V}_{et}, and specializing EE to 𝖤dR\mathsf{E}_{dR}, the group 𝐆0\mathbf{G}^{0} will contain the differential Galois group of 𝖵dR\mathsf{V}_{dR}.

Let 𝐇\mathbf{H} be a reductive group over E0E_{0}, and choose a representation VsimpV_{simp} of 𝐇\mathbf{H}. We will assume that 𝐇\mathbf{H} is one of GLNGL_{N}, GSpNGSp_{N}, or GONGO_{N}, and VsimpV_{simp} is the standard representation. A subtlety is that the definition of GONGO_{N} depends on the choice of a symmetric bilinear form on VsimpV_{simp}. Here, and at all future points, when we assume that 𝐇\mathbf{H} is GONGO_{N}, we mean that there exists a nondegenerate symmetric E0E_{0}-linear form on VsimpV_{simp} for which 𝐇\mathbf{H} is the group of similitudes. Since this is cumbersome, and the differences between different quadratic forms are of little importance to our argument, we leave the quadratic form implicit and say simply 𝐇=GON\mathbf{H}=GO_{N}.

Let 𝐆0\mathbf{G}^{0} be the Weil restriction 𝐆0=ResE0E𝐇E\mathbf{G}^{0}={\operatorname{Res}^{E}_{E_{0}}}\mathbf{H}_{E}, and let V=EE0VsimpV=E\otimes_{E_{0}}V_{simp}. By restriction of scalars, we will view VV as an E0E_{0}-vector space with actions of EE and 𝐆0\mathbf{G}^{0}. Let 𝐆\mathbf{G} be the normalizer of 𝐆0\mathbf{G}^{0} in the group of E0E_{0}-linear automorphisms of VV.

Definition 5.31.

Let σ\sigma be a E0E_{0}-linear automorphism of EE. We say that an E0E_{0}-linear endomorphism ϕ:VV\phi\colon V\rightarrow V is σ\sigma-semilinear (or semilinear over σ\sigma) if

ϕ(λv)=σ(λ)ϕ(v)\phi(\lambda v)=\sigma(\lambda)\phi(v)

for all λE\lambda\in E and vVv\in V.

Lemma 5.32.

If 𝐇=GLN\mathbf{H}=GL_{N}, then 𝐆\mathbf{G} is the algebraic group whose E0E_{0}-points correspond to endomorphisms ϕ\phi of VV, semilinear over some E0E_{0}-linear automorphism σ\sigma of EE.

If 𝐇\mathbf{H} is GSpNGSp_{N} or GONGO_{N}, then 𝐆0\mathbf{G}^{0} preserves, up to scaling, an (alternating or symmetric) EE-valued pairing ,\langle-,-\rangle on VV. In this setting 𝐆\mathbf{G} is the group of endomorphisms ϕ\phi of VV, semilinear over some E0E_{0}-linear automorphism σ\sigma of EE, and satisfying

(6) ϕ(v1),ϕ(v2)=Cσ(v1,v2)\langle\phi(v_{1}),\phi(v_{2})\rangle=C\sigma(\langle v_{1},v_{2}\rangle)

for some CEC\in E, for all v1,v2Vv_{1},v_{2}\in V.

In particular, we have an exact sequence of groups

1𝐆0𝐆AutE0E1.1\rightarrow\mathbf{G}^{0}\rightarrow\mathbf{G}\rightarrow\operatorname{Aut}_{E_{0}}E\rightarrow 1.
Proof.

Any element of the normalizer of 𝐆0\mathbf{G}^{0} normalizes the center of 𝐆0\mathbf{G}^{0}, which is EE, acting by scalar multiplication. The E0E_{0}-linear automorphisms of VV that normalize the action of EE are exactly the semilinear automorphisms.

If 𝐇\mathbf{H} is GSpNGSp_{N} or GONGO_{N}, then for ϕ𝐆\phi\in\mathbf{G} which is EE-semilinear over σ\sigma, the bilinear form ϕ(v1),ϕ(v2)\langle\phi(v_{1}),\phi(v_{2})\rangle is EE-semilinear over σ\sigma in both variables and is preserved up to scaling by 𝐆0\mathbf{G}^{0}. Thus σ1(ϕ(v1),ϕ(v2))\sigma^{-1}\left(\langle\phi(v_{1}),\phi(v_{2})\rangle\right) is EE-bilinear and preserved up to scaling by 𝐆0\mathbf{G}^{0}. Hence it is a scalar multiple of v1,v2\langle v_{1},v_{2}\rangle, which is the unique EE-linear 𝐆0\mathbf{G}^{0}-equivariant form. This gives (6).

Conversely, any EE-semilinear automorphism which, if 𝐇\mathbf{H} is GSpNGSp_{N} or GONGO_{N}, satisfies (6), manifestly normalizes 𝐆0\mathbf{G}^{0}. ∎

We need a generalization of [47, Lemma 2.1].

Lemma 5.33.

Suppose σAutE0E\sigma\in\operatorname{Aut}_{E_{0}}E is such that Eσ=E0E^{\sigma}=E_{0}, and suppose ϕ𝐆\phi\in\mathbf{G} is semilinear over σ\sigma. Then the centralizer Z𝐆(ϕ)Z_{\mathbf{G}}(\phi) satisfies

dimZ𝐆(ϕ)dim𝐇.\dim Z_{\mathbf{G}}(\phi)\leq\dim\mathbf{H}.
Proof.

The proof goes through exactly as in [47]. By passing to the algebraic closure of E0E_{0} we may assume that E=E0dE=E_{0}^{d}, so 𝐆0=𝐇d\mathbf{G}^{0}=\mathbf{H}^{d}. The hypothesis implies that σ\sigma acts transitively (i.e. cyclically) on the dd factors E0E_{0} of E0dE_{0}^{d}. We may assume the factors are ordered so that σ\sigma takes the ii-th factor to the (i+1)(i+1)-st factor (modulo dd). Then ϕ𝐆\phi\in\mathbf{G} has the off-diagonal matrix form

ϕ=(0000ϕ1ϕ200000ϕ300000ϕ400000ϕd0).\phi=\begin{pmatrix}0&0&0&\dots&0&\phi_{1}\\ \phi_{2}&0&0&\cdots&0&0\\ 0&\phi_{3}&0&\cdots&0&0\\ 0&0&\phi_{4}&\cdots&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&\cdots&\phi_{d}&0\\ \end{pmatrix}.

Now if g=(g1,,gd)Z𝐆0(ϕ)g=(g_{1},\ldots,g_{d})\in Z_{\mathbf{G}^{0}}(\phi), then

giϕi=ϕigi1g_{i}\phi_{i}=\phi_{i}g_{i-1}

for all ii (indices modulo dd). Thus, g1g_{1} determines gig_{i} for all ii. In other words, the projection Z𝐆0(ϕ)𝐇Z_{\mathbf{G}^{0}}(\phi)\rightarrow\mathbf{H} onto any single factor 𝐇\mathbf{H} of 𝐇d\mathbf{H}^{d} is injective. The result follows because Z𝐆0(ϕ)Z_{\mathbf{G}^{0}}(\phi) has finite index in Z𝐆(ϕ)Z_{\mathbf{G}}(\phi). ∎

5.7. Structure of 𝖤\mathsf{E}-modules

Let KK be a number field, SS a finite set of places of KK, and 𝒳\mathcal{X} a smooth 𝒪K,S\mathcal{O}_{K,S}-scheme. Let 𝖤\mathsf{E} be an étale H0H^{0}-algebra on Spec𝒪K,S\operatorname{Spec}\mathcal{O}_{K,S}, and 𝖵\mathsf{V} an 𝖤\mathsf{E}-module on 𝒳\mathcal{X} with 𝐇\mathbf{H}-structure, where 𝐇\mathbf{H} is GLNGL_{N}, GSpNGSp_{N}, or GONGO_{N}. We described the structure of 𝖤\mathsf{E} in Example 5.17. Now we are ready to describe 𝖵\mathsf{V}.

The local system 𝖤et\mathsf{E}_{et} is given by a Galois representation on SpecK\operatorname{Spec}K, unramified outside SS, which we’ll also call 𝖤et\mathsf{E}_{et}. By definition this is a finite étale p\mathbb{Q}_{p}-algebra with an action of GalK\operatorname{Gal}_{K}.

Let E0=pE_{0}=\mathbb{Q}_{p}, and let 𝐆et\mathbf{G}_{et} and 𝐆et0\mathbf{G}^{0}_{et} be the groups 𝐆\mathbf{G} and 𝐆0\mathbf{G}^{0} of Section 5.6 taking E=𝖤etE=\mathsf{E}_{et}. At any xX(K)x\in X(K), the fiber 𝖵et,x\mathsf{V}_{et,x} of the étale local system is a representation of the Galois group GalK\operatorname{Gal}_{K}. This VV has the structure of EE-algebra, and if 𝐇\mathbf{H} is GSpGSp or GOGO then there is a pairing

𝖵et,x𝖤𝖵et,xL,\mathsf{V}_{et,x}\otimes_{\mathsf{E}}\mathsf{V}_{et,x}\rightarrow L,

where LL is a one-dimensional p\mathbb{Q}_{p}-vector space with an action of GalK\operatorname{Gal}_{K}. The action of GalK\operatorname{Gal}_{K} respects the pairing, and acts on 𝖵et,x\mathsf{V}_{et,x} semilinearly over its action on 𝖤\mathsf{E}. It follows (Lemma 5.32) that the representation of GalK\operatorname{Gal}_{K} on VV is given by a homomorphism

GalK𝐆et,\operatorname{Gal}_{K}\rightarrow\mathbf{G}_{et},

and the quotient

GalK𝐆et/𝐆et0AutE0𝖤et\operatorname{Gal}_{K}\rightarrow\mathbf{G}_{et}/\mathbf{G}^{0}_{et}\rightarrow\operatorname{Aut}_{E_{0}}\mathsf{E}_{et}

is exactly the representation of GalK\operatorname{Gal}_{K} on EE given by the structure of 𝖤et\mathsf{E}_{et}.

Now we turn to de Rham cohomology and its cousins. Similarly, 𝖤dR\mathsf{E}_{dR} is a finite p\mathbb{Q}_{p}-algebra which is isomorphic to 𝖤et\mathsf{E}_{et} over BdRB_{dR} (and hence also over p¯)\overline{\mathbb{Q}_{p}}) and thus is also étale. Let 𝐆dR\mathbf{G}_{dR} and 𝐆dR0\mathbf{G}^{0}_{dR} be the groups 𝐆\mathbf{G} and 𝐆0\mathbf{G}^{0} of Section 5.6 taking E=𝖤dRE=\mathsf{E}_{dR}.

For a point xX(K)x\in X(K), Lemma 5.32 implies that 𝐆dR0\mathbf{G}^{0}_{dR} is isomorphic to the group of automorphisms of 𝖵dR,x\mathsf{V}_{dR,x} respecting the 𝖤dR\mathsf{E}_{dR}-action and (where applicable) the bilinear pairing.

The isomorphism between 𝖤dR\mathsf{E}_{dR} and 𝖤et\mathsf{E}_{et} after base change to ¯p\overline{\mathbb{Q}}_{p} implies that 𝐆et0\mathbf{G}^{0}_{et} and 𝐆dR0\mathbf{G}^{0}_{dR} become isomorphic after base change to ¯p\overline{\mathbb{Q}}_{p}, and the same is true for 𝐆et\mathbf{G}_{et} and 𝐆dR\mathbf{G}_{dR}.

The differential Galois group 𝐆mon\mathbf{G}_{mon} satisfies 𝐆mon𝐆dR0\mathbf{G}_{mon}\subseteq\mathbf{G}^{0}_{dR}. The weaker inclusion 𝐆mon𝐆dR\mathbf{G}_{mon}\subseteq\mathbf{G}_{dR} is immediate from the Tannakian formalism: the local system 𝖵H\mathsf{V}_{H} is naturally an algebra over 𝖤H\mathsf{E}_{H}, and if 𝐇\mathbf{H} is GSpNGSp_{N} or GONGO_{N}, then 𝖵H\mathsf{V}_{H} admits a bilinear pairing

𝖵H𝖤H𝖵H(R2kfg𝖮|Y)H\mathsf{V}_{H}\otimes_{\mathsf{E}_{H}}\mathsf{V}_{H}\rightarrow(R^{2k}f_{*}g^{*}\mathsf{O}_{\mathbb{Q}}|_{Y})_{H}

(see Lemma 5.30). Since the monodromy action on 𝖤H\mathsf{E}_{H} is trivial, 𝐆mon\mathbf{G}_{mon} is in fact contained in the finite-index subgroup 𝐆dR0\mathbf{G}^{0}_{dR}.

Lastly, we describe the filtered ϕ\phi-module 𝖵cris,x\mathsf{V}_{cris,x}. For this, we assume that Kv=pK_{v}=\mathbb{Q}_{p}. Recall the structure of 𝖤cris\mathsf{E}_{cris} from Example 5.17: it is the p\mathbb{Q}_{p}-algebra EE, equipped with a Frobenius endomorphism σ\sigma. Then 𝖵cris,x\mathsf{V}_{cris,x} is a vector space over EE, and it is naturally a filtered ϕ\phi-module with 𝐆dR\mathbf{G}_{dR}-structure. Its Frobenius endomorphism ϕ\phi is semilinear over σAutE\sigma\in\operatorname{Aut}E.

5.8. Disconnected reductive groups

We need to apply pp-adic Hodge theory in a Tannakian setting: we’ll work with Galois representations valued in the disconnected algebraic group 𝐆et\mathbf{G}_{et}, and the corresponding filtered ϕ\phi-modules. To this end, we need some general results about groups with reductive identity component.

We will use a notion of parabolic subgroup due to Richardson [56] (see also [48] and the survey [6]). Let GG be an algebraic group over a field of characteristic zero whose identity component G0G^{0} is reductive.

Definition 5.34.

Let f:𝔾mGf\colon\mathbb{G}_{m}\rightarrow G be a morphism of schemes. We say that limt0f(t)\lim_{t\rightarrow 0}f(t) exists if ff extends to a morphism f~:𝔸1G\tilde{f}\colon\mathbb{A}^{1}\rightarrow G; in this case we write

limt0f(t)=f~(0).\lim_{t\rightarrow 0}f(t)=\tilde{f}(0).

Let μ:𝔾mG\mu\colon\mathbb{G}_{m}\rightarrow G be a cocharacter. Define the subgroups PμP_{\mu}, LμL_{\mu}, and UμU_{\mu} of GG as follows.

  • Pμ={gG|limt0μ(t)gμ(t)1 exists}P_{\mu}=\{g\in G|\lim_{t\rightarrow 0}\mu(t)g\mu(t)^{-1}\mbox{ exists}\}

  • Uμ={gG|limt0μ(t)gμ(t)1=1}U_{\mu}=\{g\in G|\lim_{t\rightarrow 0}\mu(t)g\mu(t)^{-1}=1\}

  • Lμ={limt0μ(t)gμ(t)1|gPμ}L_{\mu}=\{\lim_{t\rightarrow 0}\mu(t)g\mu(t)^{-1}|g\in P_{\mu}\}

We say that a subgroup PGP\subseteq G is parabolic if it is of the form PμP_{\mu} for some μ\mu; in this case we say LμL_{\mu} is a Levi subgroup associated to PP, for any μ\mu such that Pμ=PP_{\mu}=P.

In this setting, UμU_{\mu} is unipotent, and PμP_{\mu} is the semidirect product of LμL_{\mu} by UμU_{\mu}; in particular, there is a natural projection PμLμP_{\mu}\rightarrow L_{\mu}, given by

glimt0μ(t)gμ(t)1.g\mapsto\lim_{t\rightarrow 0}\mu(t)g\mu(t)^{-1}.

Furthermore, LμL_{\mu} is the centralizer of μ\mu in GG [48, Prop. 5.2].

Example 5.35.

The purpose of this example is to show that a parabolic subgroup PGP\subseteq G is not uniquely determined by the parabolic subgroup P0=PG0P^{0}=P\cap G^{0} of G0G^{0}.

Let GG be the normalizer in GL2NGL_{2N} of GLN×GLNGL_{N}\times GL_{N}. Then G0=GLN×GLNG^{0}=GL_{N}\times GL_{N}. Let cocharacters μ1,μ2:𝔾mGLN×GLN\mu_{1},\mu_{2}\colon\mathbb{G}_{m}\rightarrow GL_{N}\times GL_{N} be given by

μ1(t)=(t3,t2)\mu_{1}(t)=(t^{3},t^{2})
μ2(t)=(t2,t2).\mu_{2}(t)=(t^{2},t^{2}).

Then Pμ1=G0P_{\mu_{1}}=G^{0} but Pμ2=GP_{\mu_{2}}=G.

Now we turn to the notion of semisimplification of subgroups of GG. See [62] for a discussion of this notion in the connected setting; it is applied in [47, §2.3]. For the general theory of complete reducibility for disconnected reductive groups, see [5], [7], [4], and [6]. We warn the reader that the theory is developed there over an arbitrary field, and many complications arise in positive characteristic that are irrelevant to us here.

Definition 5.36.

We say that an algebraic subgroup HGH\subseteq G is GG-completely reducible if, for every parabolic subgroup PP containing HH, there is some Levi subgroup LL associated to PP that also contains HH.

If HGH\subseteq G is an arbitrary algebraic subgroup, we define its semisimplification with respect to GG as follows. Let PP be a parabolic subgroup of GG, minimal containing HH. Choose a Levi subgroup LL of PP, and let HssH^{ss} be the image of HH under the projection PLP\rightarrow L.

We say that a representation valued in GG is semisimple if the Zariski closure of its image is GG-completely reducible.

Lemma 5.37.

Let HGH\subseteq G be an algebraic subgroup. Then HH is GG-completely reducible if and only if the identity component of HH is reductive. If we choose an embedding of GG into GLnGL_{n}, then HH is GG-completely reducible if and only if it is GLnGL_{n}-completely reducible.

For a general algebraic subgroup HGH\subseteq G, the GG-semisimplification HssH^{ss} is well-defined up to GG-conjugacy, and it is GG-completely reducible.

Proof.

This is Cor. 3.5 and Thm. 4.5 of [6]. ∎

5.8.1. Some Tannakian formalism

We recall the notion of object with GG-structure. See [49, Def. 1.3]; a general reference for the Tannakian formalism is [57].

Definition 5.38.

Let GG be an algebraic group over a field EE of characteristic zero, and let CC be an EE-linear rigid abelian tensor category with fiber functor. An object in CC with GG-structure is an exact faithful tensor functor ω:Rep¯GC\omega\colon\underline{\operatorname{Rep}}_{G}\rightarrow C, where Rep¯G\underline{\operatorname{Rep}}_{G} is the category of finite-dimensional EE-linear representations of GG, together with an isomorphism between the composition of ω\omega with the fiber functor of CC and the forgetful functor from Rep¯G\underline{\operatorname{Rep}}_{G} to the category Vec¯E\underline{\operatorname{Vec}}_{E} of vector spaces over EE.

Two objects in CC with GG-structure are GG-conjugate if they correspond to isomorphic functors ω\omega. Note that GG-conjugate objects with GG-structure differ exactly by an automorphism of the forgetful functor Rep¯GVec¯E\underline{\operatorname{Rep}}_{G}\to\underline{\operatorname{Vec}}_{E}, i.e. by an element of GG.

If G1G2G_{1}\rightarrow G_{2} is a morphism of algebraic groups, then an object with G1G_{1}-structure gives rise to an object with G2G_{2}-structure by functoriality.

Remark 5.39.

If VV is a faithful representation of GG, then an object ω\omega with GG-structure is determined by ω(V)\omega(V). In practice, we will find it useful to specify ω\omega by describing ω(V)\omega(V).

5.8.2. Filtrations

(Filtrations on GG.)

We want to define a notion of “filtered vector space with GG-structure” or “filtration on GG.” Definition 5.38 does not apply, because the category of filtered vector spaces is not abelian.555In Section 5.8.4 below, we will apply Definition 5.38 to the category of weakly admissible filtered ϕ\phi-modules, which is abelian. Instead, we will use the formalism of filtrations from [57, §IV.2]. Throughout this section, GG will be an algebraic group over a field EE of characteristic zero.

Definition 5.40.

A filtration on GG (or GG-filtration) is an exact tensor filtration of the forgetful functor ω\omega from Rep¯G\underline{\operatorname{Rep}}_{G} to Vec¯E\underline{\operatorname{Vec}}_{E}, in the sense of [57, IV.2.1.1].

In other words, a GG-filtration is a sequence of exact subfunctors FnωF^{n}\omega of ω\omega, such that:

  1. (1)

    For each object EE of Rep¯G\underline{\operatorname{Rep}}_{G}, the objects Fnω(E)F^{n}\omega(E) form a decreasing filtration of ω(E)\omega(E).

  2. (2)

    The associated graded functor grF(ω)\operatorname{gr}_{F}(\omega), which assigns to EE the associated graded of the filtration Fnω(E)F^{n}\omega(E), is exact.

  3. (3)

    For every nn\in\mathbb{Z}, and all objects E,EE,E^{\prime} of Rep¯G\underline{\operatorname{Rep}}_{G}, we have

    Fnω(EE)=a+b=nFaω(E)Fbω(E).F^{n}\omega(E\otimes E^{\prime})=\sum_{a+b=n}F^{a}\omega(E)\otimes F^{b}\omega(E^{\prime}).

Loosely speaking, a GG-filtration is a choice of descending filtration on every finite-dimensional representation of GG, compatible with tensor products, duals, and passage to subquotients.

If G1G2G_{1}\rightarrow G_{2} is a morphism of algebraic groups, then a G1G_{1}-filtration gives rise to a G2G_{2}-filtration by functoriality.

Since the base field EE is of characteristic zero, every GG-filtration is “scindable” [57, Théorème IV.2.4] and hence also “admissible” [57, IV.2.2.1]. In particular, every filtration comes from a gradation (in general not unique) on GG.

Let us explain this more carefully. A representation of 𝔾m\mathbb{G}_{m} on a finite-dimensional vector space VV gives a decomposition V=VjV=\bigoplus V_{j}, where VjV_{j} is the tjt^{j}-eigenspace of 𝔾m\mathbb{G}_{m} on VV. A cocharacter μ:𝔾mG\mu\colon\mathbb{G}_{m}\rightarrow G defines a filtration on GG by defining, for all representations VV of GG,

FiliV=jiVj.\operatorname{Fil}^{i}V=\bigoplus_{j\geq i}V_{j}.

That every filtration on GG is “scindable” means that every filtration comes from some cocharacter μ\mu (in general not unique).

Given a filtration on GG, we can define two distinguished subgroups PP and UU of GG [57, §IV.2.1.3]. The subgroup PGP\subseteq G is the group of elements that stabilize the filtration on every representation VV of GG, and UPU\subseteq P is the group of elements that stabilize the filtration and furthermore act as the identity on the associated graded. When GG is a group whose identity component is reductive, and the filtration is defined by a cocharacter μ\mu, the group PP coincides with the Richardson parabolic PμP_{\mu} (Definition 5.34), and UU is its unipotent radical UμU_{\mu}.

Again supposing GG is a group whose identity component is reductive, define an equivalence relation on the set of cocharacters of GG as follows: we say that μ1μ2\mu_{1}\sim\mu_{2} if Pμ1=Pμ2P_{\mu_{1}}=P_{\mu_{2}} and μ1\mu_{1} is conjugate to μ2\mu_{2} in this parabolic. Then μ1\mu_{1} and μ2\mu_{2} give rise to the same filtration on GG if and only if μ1μ2\mu_{1}\sim\mu_{2} [57, §IV.2.2.4].

If GG acts faithfully on VV, then a GG-filtration is determined by the corresponding filtration on VV, as in Remark 5.39; we will use this without comment.

We mention in passing that, even if GG is reductive, the parabolic PP does not quite determine the filtration. If GG is reductive, then PP determines a filtration on every representation of GG only up to reindexing.

5.8.3. GG-filtrations vs. G0G_{0}-filtrations

Let GG be an algebraic group over a field of characteristic zero whose identity component G0G^{0} is reductive.

A G0G_{0}-filtration F0F^{0} gives rise to a GG-filtration FF, by functoriality; in terms of cocharacters, the correspondence is given by composition

𝔾mG0G.\mathbb{G}_{m}\rightarrow G_{0}\rightarrow G.
Definition 5.41.

We say that a G0G_{0}-filtration F0F^{0} is associated to a GG-filtration FF if FF is induced from F0F^{0} by functoriality with respect to the inclusion G0GG_{0}\hookrightarrow G.

Lemma 5.42.

Being associated defines a bijection between G0G_{0}-filtrations and GG-filtrations.

Proof.

We need to show that every filtration on GG comes from a unique filtration on G0G_{0}.

Existence is easy. A filtration on GG is defined by a cocharacter 𝔾mG\mathbb{G}_{m}\rightarrow G; since 𝔾m\mathbb{G}_{m} is connected, every such cocharacter factors through G0G_{0}.

Conversely, suppose

μ1,μ2:𝔾mG\mu_{1},\mu_{2}\colon\mathbb{G}_{m}\rightarrow G

are two cocharacters defining the same filtration on GG. We need to show that μ1\mu_{1} and μ2\mu_{2} define the same filtration on G0G_{0}.

We know there is some gPμ1g\in P_{\mu_{1}} such that

gμ1g1=μ2.g\mu_{1}g^{-1}=\mu_{2}.

We need to show that we can in fact take

gPμ10=Pμ1G0.g\in P^{0}_{\mu_{1}}=P_{\mu_{1}}\cap G_{0}.

Let gLg_{L} be the image of gg under the projection Pμ1Lμ1P_{\mu_{1}}\twoheadrightarrow L_{\mu_{1}}. Then gLg_{L} centralizes μ1\mu_{1} by [48, Prop. 5.2(a)], and from the limit formula

gL=limt0μ1(t)gμ1(t)1g_{L}=\lim_{t\rightarrow 0}\mu_{1}(t)g\mu_{1}(t)^{-1}

we see that ggL1G0gg_{L}^{-1}\in G_{0}. Thus ggL1G0gg_{L}^{-1}\in G_{0} conjugates μ1\mu_{1} to μ2\mu_{2}. ∎

5.8.4. Weakly admissible filtered ϕ\phi-modules with GG-structure

Let GG be an algebraic group over p\mathbb{Q}_{p}. (It is important that we work over p\mathbb{Q}_{p}, and not an extension, because the category of filtered ϕ\phi-modules is only p\mathbb{Q}_{p}-linear.) A weakly admissible filtered ϕ\phi-module with GG-structure gives rise to a GG-filtration and an element (the Frobenius endomorphism) in GG. In the spirit of Remark 5.39, we can describe such an object as a triple (V,ϕ,F)(V,\phi,F), where VV is a vector space on which GG acts faithfully, ϕG\phi\in G is an automorphism of VV, and FF is a GG-filtration on VV, such that (V,ϕ,F)(V,\phi,F) is weakly admissible. Somewhat imprecisely, we will call such objects “filtered ϕ\phi-modules with GG-structure.”

5.8.5. pp-adic Hodge theory

The next result is a Tannakian form of the crystalline comparison theorem, for representations valued in an arbitrary algebraic group GG. In order to avoid problems with semilinearity in the target category, we restrict to representations of Galp\operatorname{Gal}_{\mathbb{Q}_{p}}.

Lemma 5.43.

Let GGLNG\subseteq GL_{N} be an algebraic group over p\mathbb{Q}_{p}. We say that a local Galois representation

GalpG\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow G

is crystalline if the representation

GalpGLN\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow GL_{N}

is crystalline in the usual sense. This property is independent of the choice of embedding GGLNG\hookrightarrow GL_{N}.

  1. (1)

    The functor D¯cris\underline{D}_{\mathrm{cris}} of pp-adic Hodge theory [23, Expose III] extends to a functor from crystalline representations GalpG\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow G to pairs of an inner form GG^{\prime} of GG and an admissible filtered ϕ\phi-modules over p\mathbb{Q}_{p} with GG^{\prime}-structure. (Here morphisms of such pairs are given by isomorphisms of inner forms of GG together to isomorphisms of admissible filtered ϕ\phi-modules compatible with the isomorphism of inner forms.)

  2. (2)

    A homomorphism of groups G1G2G_{1}\rightarrow G_{2} and a crystalline Galois representation GalpG1\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow G_{1} induces a morphism of inner forms G1G2G_{1}^{\prime}\to G_{2}^{\prime} compatible with the filtered ϕ\phi-modules with G1G_{1}^{\prime}-structure and G2G_{2}^{\prime}-structure associated to GalpG1\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow G_{1} and GalpG1G2\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow G_{1}\rightarrow G_{2} respectively. The map G1G2G_{1}^{\prime}\to G_{2}^{\prime} is the twist of G1G2G_{1}\to G_{2} by some G1G_{1}-torsor. In particular, if G1G2G_{1}\to G_{2} has any property preserved by such twists then G1G2G_{1}^{\prime}\to G_{2}^{\prime} does as well.

  3. (3)

    For G=𝐆etG=\mathbf{G}_{et} and a Galois representation ρ:Galp𝐆et\rho\colon\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow\mathbf{G}_{et} whose composition with the natural map 𝐆etAutp𝖤et\mathbf{G}_{et}\rightarrow\operatorname{Aut}_{\mathbb{Q}_{p}}\mathsf{E}_{et} is the usual action of Galois on 𝖤et\mathsf{E}_{et}, the relevant inner twist of 𝐆et\mathbf{G}_{et} is 𝐆dR\mathbf{G}_{dR}.

Proof.

The general claims follow from the Tannakian formalism and the fact that D¯cris\underline{D}_{\mathrm{cris}} is a fiber functor on the Tannakian category of crystalline Galois representations [23, Exposé III, Proposition 1.5.2].

Indeed, a Galois representation ρ:GalpG\rho\colon\operatorname{Gal}_{\mathbb{Q}_{p}}\rightarrow G gives rise to an exact tensor functor from Rep¯G\underline{\operatorname{Rep}}_{G} to the category of Galois representations. The crystallinity assumption implies that all representations in the image are crystalline. The composition of this functor with D¯cris\underline{D}_{\mathrm{cris}} then gives an exact tensor functor from representations of GG to admissible filtered ϕ\phi-modules, and the underlying vector space is a fiber functor. However, this fiber functor need not be isomorphic to the forgetful functor of Rep¯G\underline{\operatorname{Rep}}_{G}, so we do not immediately obtain a filtered ϕ\phi-module with GG-structure. Instead, for GG^{\prime} the group of automorphisms of this fiber functor, [17, Theorem 2.11(b)] implies that the Tannakian category of representations of GG is equivalent to the Tannakian category of representations of GG^{\prime}. Hence this functor defines an admissible filtered ϕ\phi-module with GG^{\prime} structure. Furthermore, [17, Theorem 3.2(b)] implies that the category of fiber functors is equivalent to the category of GG-torsors. Since the automorphism group of any GG-torsor is an inner form of GG, it follows that GG^{\prime} is an inner form of GG.

Functoriality is the statement that if two Galois representations are GG-conjugate, then the associated groups GG^{\prime} are isomorphic and the associated filtered ϕ\phi-modules with GG^{\prime}-structure are GG^{\prime}-conjugate. This is because the two exact tensor functors from Rep¯G\underline{\operatorname{Rep}}_{G} to the category of Galois representations are isomorphic, and hence the functors from Rep¯G\underline{\operatorname{Rep}}_{G} to the category of admissible filtered ϕ\phi-modules are isomorphic. This completes tthe proof of (1).

For (2), we now have a composition of functors, first from Rep¯G2\underline{\operatorname{Rep}}_{G_{2}} to Rep¯G1\underline{\operatorname{Rep}}_{G_{1}}, then to the category of Galois representations, then to the category of filtered ϕ\phi-modules, then to the category of vector spaces. The fiber functor of G2G_{2} we use to construct G2G_{2}^{\prime} is given by the composition of all these functors, and the fiber functor of G1G_{1} we use to construct G1G_{1}^{\prime} is given by the composition of all these functors but one. Since these are compatible with the functor Rep¯G2Rep¯G1\underline{\operatorname{Rep}}_{G_{2}}\to\underline{\operatorname{Rep}}_{G_{1}}, we obtain a map of Tannakian categories Rep¯G2Rep¯G1\underline{\operatorname{Rep}}_{G_{2}^{\prime}}\to\underline{\operatorname{Rep}}_{G_{1}^{\prime}}, hence a map of groups G1G2G_{1}^{\prime}\to G_{2}. The fiber functor of Rep¯G1\underline{\operatorname{Rep}}_{G_{1}} corresponds to a G1G_{1}-torsor, the induced fiber functor of Rep¯G2\underline{\operatorname{Rep}}_{G_{2}} comes from the pushforward of that torsor from G1G_{1} to G2G_{2}, from which it follows that G1G2G_{1}^{\prime}\to G_{2}^{\prime} arises from twisting G1G2G_{1}\to G_{2} by that G1G_{1}-torsor.

For (3), observe that the representation ρ\rho corresponds to a functor from the Tannakian category of representations of 𝐆et\mathbf{G}_{et} to Galois representations. This Tannakian category includes the standard representation of 𝐆et\mathbf{G}_{et} on 𝖤etpVsimp\mathsf{E}_{et}\otimes_{\mathbb{Q}_{p}}V_{simp}, which may admit a bilinear form, as well as the representation 𝖤et\mathsf{E}_{et}. Call these objects 𝖵rep\mathsf{V}_{rep} and 𝖤rep\mathsf{E}_{rep}. The assumptions imply that this functor sends 𝖤rep\mathsf{E}_{rep} to the usual Galois representation 𝖤et\mathsf{E}_{et}. This Tannakian category admits a second fiber functor fdRf_{dR} which takes each Galois representation to its associated filtered ϕ\phi-module and takes the underlying vector space. The fiber functor fdRf_{dR} takes 𝖤rep\mathsf{E}_{rep} to the filtered ϕ\phi-module associated to 𝖤et\mathsf{E}_{et}, which is 𝖤dR\mathsf{E}_{dR}.

The relevant inner twist is the automorphism group of fdRf_{dR}, which is clearly contained in the group of automorphisms of fdR(𝖵rep)f_{dR}(\mathsf{V}_{rep}) that are semiilinear over some automorphism of 𝖤rep\mathsf{E}_{rep} and respect the bilinear form if it appears. The group of automorphisms satisfying these semilinearity and bilinear form conditions is 𝐆dR\mathbf{G}_{dR}. After passing to an algebraically closed field, the two fiber functors are isomorphic, and so the automorphism group of fdRf_{dR} becomes equal to 𝐆et\mathbf{G}_{et} and thus equal to 𝐆dR\mathbf{G}_{dR}. Thus the automorphism group of fdRf_{dR} must be all of 𝐆dR\mathbf{G}_{dR} over the base field p\mathbb{Q}_{p} as well. ∎

5.9. Adjoint Hodge numbers

Definition 5.44.

(Adjoint Hodge numbers; see beginning of §10 of [47].)

Let GG be a reductive group, and suppose FF is a filtration on GG.

In this setting, we define the adjoint Hodge numbers as follows: The filtration FF on GG gives, by definition, a filtration on every representation of GG. We apply this to the adjoint representation of GG on LieG\operatorname{Lie}G, and call the adjoint Hodge numbers hah^{a} the dimensions of the associated graded of the resulting filtration666We use the notation hah^{a}, rather than the more common hph^{p}, to avoid conflict with the prime pp used throughout the proof.:

ha=dimFaV/Fa1V.h^{a}=\dim F^{a}V/F^{a-1}V.
Definition 5.45.

For any real x[0,aha]x\in[0,\sum_{a}h^{a}], we define the “sum of the topmost xx Hodge numbers” T(x)T(x) to be the continuous, piecewise-linear function T:[0,aha]T\colon[0,\sum_{a}h^{a}]\rightarrow\mathbb{R} satisfying T(0)=0T(0)=0 and

T(x)=kT^{\prime}(x)=k

for a=k+1ha<x<a=kha\sum_{a=k+1}^{\infty}h^{a}<x<\sum_{a=k}^{\infty}h^{a}. (The sums are finite because only finitely many hah^{a} are nonzero.)

When we want to emphasize the dependence on the group GG, we will write TG(x)T_{G}(x) and hGah_{G}^{a} instead of T(x)T(x) and hah^{a}.

Definition 5.46.

(Uniform Hodge numbers.)

Recall notation from Section 5.6 and 5.7. In particular, 𝐇\mathbf{H} is a reductive group over p\mathbb{Q}_{p}, and 𝐆dR0=Resp𝖤dR𝐇\mathbf{G}^{0}_{dR}={\operatorname{Res}^{\mathsf{E}_{dR}}_{\mathbb{Q}_{p}}}\mathbf{H}, is a form of 𝐇d\mathbf{H}^{d}; the groups 𝐆dR0\mathbf{G}^{0}_{dR} and 𝐆dR\mathbf{G}_{dR} act on the vector space 𝖵dR,x\mathsf{V}_{dR,x}. Let FF be a 𝐆dR0\mathbf{G}^{0}_{dR}-filtration on 𝖵dR,x\mathsf{V}_{dR,x}.

After base change to p¯\overline{\mathbb{Q}_{p}}, we obtain

Ep¯ι(p¯)ι,E\otimes\overline{\mathbb{Q}_{p}}\cong\bigoplus_{\iota}(\overline{\mathbb{Q}_{p}})_{\iota},

the direct sum taken over all embeddings ι:Ep¯\iota\colon E\rightarrow\overline{\mathbb{Q}_{p}}. (The subscript on (p¯)ι(\overline{\mathbb{Q}_{p}})_{\iota} is purely for notational convenience: each (p¯)ι(\overline{\mathbb{Q}_{p}})_{\iota} is a copy of p¯\overline{\mathbb{Q}_{p}}.) Similarly, the group 𝐆dR,p¯0\mathbf{G}^{0}_{dR,\overline{\mathbb{Q}_{p}}} splits as a direct sum of copies of 𝐇p¯\mathbf{H}_{\overline{\mathbb{Q}_{p}}}, and (V,F)p¯(V,F)\otimes\overline{\mathbb{Q}_{p}} splits as a direct sum of filtered p¯\overline{\mathbb{Q}_{p}}-vector spaces (Vι,Fι)(V_{\iota},F_{\iota}), indexed by ι\iota.

For each ι\iota, let hιah^{a}_{\iota} be the adjoint Hodge numbers of the 𝐇\mathbf{H}-filtration FιF_{\iota}. We say that FF is uniform if the numbers hιah^{a}_{\iota} are independent of ι\iota. In this situation, we write the Hodge numbers and associated TT-function as

ha=h𝐇a(Vι) and T(x)=T𝐇(x).h^{a}=h^{a}_{\mathbf{H}}(V_{\iota})\mbox{ and }T(x)=T_{\mathbf{H}}(x).
Example 5.47.

Let E0=pE_{0}=\mathbb{Q}_{p} and E=E02E=E_{0}^{2}. Any EE-module VV can be written as a direct sum V=V1V2V=V_{1}\oplus V_{2}, where V1V_{1} is a vector space over the first factor E0E_{0}, and V2V_{2} a vector space over the second. A filtration FF on VV is determined by a filtration FiF_{i} on ViV_{i}, for i=1,2i=1,2. Then FF is uniform if and only if the adjoint Hodge numbers of 𝐇\mathbf{H} on (V1,F1)(V_{1},F_{1}) are the same as those of 𝐇\mathbf{H} on (V2,F2)(V_{2},F_{2}).

Lemma 5.48.

In the setting of Definition 5.46, suppose FF is a 𝐆dR0\mathbf{G}^{0}_{dR}-filtration on VV with uniform Hodge numbers, and let c=[E:p]c=[E\colon\mathbb{Q}_{p}]. We have

h𝐆dR0a=ch𝐇ah^{a}_{\mathbf{G}^{0}_{dR}}=ch^{a}_{\mathbf{H}}

and

T𝐆dR0(cx)=cT𝐇(x).T_{\mathbf{G}^{0}_{dR}}(cx)=cT_{\mathbf{H}}(x).
Proof.

The Hodge numbers can be computed after base change to p¯\overline{\mathbb{Q}_{p}}. ∎

5.10. Galois representations

The next result is a form of the Faltings finiteness lemma. Compare also [47, Lemma 2.4], which applies when GG is a connected reductive group.

Lemma 5.49.

Let E0=pE_{0}=\mathbb{Q}_{p}, and let EE, NN, 𝐇\mathbf{H}, 𝐆et\mathbf{G}_{et}, 𝐆et0\mathbf{G}^{0}_{et} be as in Section 5.6. The group 𝐆et\mathbf{G}_{et} has a standard embedding into GLN[E:p](p)GL_{N[E:\mathbb{Q}_{p}]}(\mathbb{Q}_{p}), coming from its action by endomorphisms on a free EE-module of rank NN.

Fix a number field KK, a finite set SS of primes of 𝒪K\mathcal{O}_{K}, and an integer ww. There are, up to 𝐆et\mathbf{G}_{et}-conjugacy, only finitely many Galois representations

ρ:GalK𝐆et\rho\colon\operatorname{Gal}_{K}\rightarrow\mathbf{G}_{et}

satisfying the following conditions.

  • The representation ρ\rho is semisimple (in the sense of Definition 5.36).

  • The representation ρ\rho is unramified outside SS.

  • The representation ρ\rho is pure of weight ww with integral Weil numbers: for every prime S\ell\not\in S, the characteristic polynomial of Frob\operatorname{Frob}_{\ell} has all roots algebraic integers of complex absolute value qw/2q_{\ell}^{w/2}, where qq_{\ell} is the order of the residue field at \ell.

Proof.

If ρ\rho is semisimple in the sense of Section 5.8, then it is semisimple in the usual sense (as a representation into GLN[E:p](p)GL_{N[E:\mathbb{Q}_{p}]}(\mathbb{Q}_{p})); see Definition 5.36 and Lemma 5.37.

Let GalLGalK\operatorname{Gal}_{L}\subseteq\operatorname{Gal}_{K} be the kernel of

GalK𝐆et𝐆et/𝐆et0.\operatorname{Gal}_{K}\rightarrow\mathbf{G}_{et}\rightarrow\mathbf{G}_{et}/\mathbf{G}^{0}_{et}.

The representation ρ\rho restricts to a 𝐆et0\mathbf{G}^{0}_{et}-valued representation

ρ|L:GalL𝐆et0,\rho|_{L}\colon\operatorname{Gal}_{L}\rightarrow\mathbf{G}^{0}_{et},

which is also semisimple (in the usual sense). Now 𝐆et0\mathbf{G}^{0}_{et} is a connected reductive group, so we can use [47, Lemma 2.6] to conclude that there are only finitely many possibilities for ρ|L\rho|_{L}.

For each fixed choice of ρ|L\rho|_{L} an extension to a map ρ\rho compatible with the map GalK𝐆et/𝐆et0\operatorname{Gal}_{K}\rightarrow\mathbf{G}_{et}/\mathbf{G}^{0}_{et} is determined by its value at a system of coset representatives for GalK/GalL\operatorname{Gal}_{K}/\operatorname{Gal}_{L}, so the set of extensions forms an algebraic variety. To show this set is finite, it suffices to show that conjugation by Z𝐆et0(Im(ρ|L)))Z_{\mathbf{G}^{0}_{et}}(\operatorname{Im}(\rho|_{L}))) acts transitively on each component of the variety, or equivalently that the tangent space of each point in the variety modulo the tangent space of Z𝐆et0(Im(ρ|L)))Z_{\mathbf{G}^{0}_{et}}(\operatorname{Im}(\rho|_{L}))) vanishes. This tangent space may be calculated to be H1H^{1} of GalL/K\operatorname{Gal}_{L/K} with coefficients in the Lie algebra of Z𝐆et0(Im(ρ|L))Z_{\mathbf{G}^{0}_{et}}(\operatorname{Im}(\rho|_{L})), which vanishes as it is the cohomology of a finite group with coefficients in a vector space of characteristic zero. ∎

Lemma 5.50.

Let GG be an algebraic group over a field EE whose identity component is reductive.

Let HGH\subseteq G be an algebraic subgroup.

Then the set of Levi subgroups of GG containing HH and defined over EE forms finitely many orbits under conjugation by the EE-points of the centralizer ZG(H)Z_{G}(H).

Proof.

A Levi subgroup LL of GG is the centralizer of a cocharacter μ:𝔾mG\mu\colon\mathbb{G}_{m}\rightarrow G; the subgroup LL contains HH if and only if μ\mu takes values in ZG(H)Z_{G}(H). Since all maximal EE-split tori in ZG(H)Z_{G}(H) are conjugate, we can assume that μ\mu takes values in a fixed torus TT.

So we need to know that cocharacters μ\mu of GG, taking values in a given torus TT, define only finitely many different Levi subgroups LL of GG. It is well-known that there are only finitely many possibilities for L0=LG0L^{0}=L\cap G^{0}. But now we have

L0LNG(L0),L^{0}\subseteq L\subseteq N_{G}(L^{0}),

and since NG(L0)N_{G}(L^{0}) contains L0L^{0} with finite index, there are only finitely many possibilities for LL. ∎

Remark 5.51.

Even if G0G^{0} is a torus, and thus has only a single Levi subgroup, there can be multiple Levi subgroups of GG if the component group of GG acts nontrivially on the cocharacters of G0G^{0}, because a component will be in the Levi if and only if it fixes the cocharacter μ\mu. However, there will only be finitely many.

Lemma 5.52.

(only finitely many Levis to give the semisimplification)

Let

ρ0:GalK𝐆et\rho_{0}\colon\operatorname{Gal}_{K}\rightarrow\mathbf{G}_{et}

be a semisimple Galois representation. Then there exists a finite collection of Levi subgroups L𝐆etL\subseteq\mathbf{G}_{et} with the following property: for any Galois representation

ρ:GalK𝐆et\rho\colon\operatorname{Gal}_{K}\rightarrow\mathbf{G}_{et}

whose semisimplification is 𝐆et\mathbf{G}_{et}-conjugate to ρ0\rho_{0}, there exist g𝐆etg\in\mathbf{G}_{et}, an LL from the finite collection, and a parabolic subgroup PP containing LL, such that gρg1g\rho g^{-1} takes values in PP, and the composite map

𝐆etgρg1PL\mathbf{G}_{et}\stackrel{{\scriptstyle g\rho g^{-1}}}{{\rightarrow}}P\rightarrow L

is ρ0\rho_{0}.

Proof.

This is an immediate consequence of Lemma 5.50. ∎

6. Period maps and monodromy

6.1. Compatible period maps and Bakker–Tsimerman

The Bakker–Tsimerman theorem [3] is a strong result on the transcendence of complex period mappings. It implies a pp-adic analogue, which we recall here. (See also [47, §9].)

The pp-adic Bakker-Tsimerman theorem is an unlikely intersection statement for the pp-adic period map attached to a Hodge–Deligne system. (For a detailed discussion of this period map, see [47, §3.3-3.4].) Suppose 𝖵\mathsf{V} is a Hodge–Deligne system on 𝒳\mathcal{X}, and let ΩX(Kv)\Omega\subseteq X(K_{v}) be a vv-adic residue disk. For all xΩx\in\Omega,

𝖵cris,x\mathsf{V}_{cris,x}

is a filtered ϕ\phi-module (Vx,ϕx,Filx)(V_{x},\phi_{x},\operatorname{Fil}_{x}). The structure of FF-isocrystal means that, for all xΩx\in\Omega, the vector spaces VxV_{x} are canonically identified, in such a way that ϕx\phi_{x} is constant on Ω\Omega. Using this identification, the filtration Filx\operatorname{Fil}_{x} varies pp-adically in xx. A priori, this defines a map

Φp:ΩGLN/P\Phi_{p}\colon\Omega\rightarrow\mathcal{H}\cong GL_{N}/P

into some flag variety, where NN is the rank of 𝖵\mathsf{V}. We will show that the period map actually takes values in a smaller flag variety 𝐆mon/Pmon\mathbf{G}_{mon}/P_{mon}, where 𝐆mon\mathbf{G}_{mon} is the differential Galois group of 𝖵dR\mathsf{V}_{dR} in the sense of [36, §IV], i.e. the Tannakian group of 𝖵dR\mathsf{V}_{dR} in the Tannakian category of vector bundles with flat connection.

Lemma 6.1.

If 𝐆mon\mathbf{G}_{mon} is the differential Galois group of 𝖵dR\mathsf{V}_{dR}, then in fact the image of Φp\Phi_{p} is contained in a single orbit of 𝐆mon\mathbf{G}_{mon}.

Proof.

(This argument was suggested by Sergey Gorchinskiy.)

Fix a basepoint x0X(K)x_{0}\in X(K^{\prime}), for some field KK^{\prime} with fixed embeddings into both KvK_{v} and \mathbb{C}. Let \mathcal{H} be the flag variety classifying filtrations with the same dimensional data as the filtration on 𝖵dR,x0\mathsf{V}_{dR,x_{0}}; we have an isomorphism GLN/P\mathcal{H}\cong GL_{N}/P, for PP some parabolic subgroup of GLNGL_{N}.

Consider the complex period map Φ\Phi_{\mathbb{C}} from the universal cover Xan~\widetilde{X^{an}} of XanX^{an} to \mathcal{H}. Its image lands in a single orbit of 𝐆mon\mathbf{G}_{mon}. (See [28, III.A, item (ii) on p. 73].)

To transfer the result from Φ\Phi_{\mathbb{C}} to Φp\Phi_{p}, we use Picard–Vessiot theory. (For an introduction to Picard–Vessiot theory, see [54].)

To the vector bundle with connection underlying 𝖵\mathsf{V}, Picard–Vessiot theory attaches an (algebraic) 𝐆mon\mathbf{G}_{mon}-torsor PVXKPV\rightarrow X_{K^{\prime}}, whose fiber over any LL-point xx classifies vector space isomorphisms

𝖵dR,x𝖵dR,x0\mathsf{V}_{dR,x}\cong\mathsf{V}_{dR,x_{0}}

respecting the 𝐆mon\mathbf{G}_{mon}-structure on both sides (but not, in general, the filtrations). Furthermore, we obtain a 𝐆mon\mathbf{G}_{mon}-equivariant map

ΦPV:PVGLN/P\Phi_{PV}\colon PV\rightarrow GL_{N}/P

where a point of PVPV over xXx\in X gives an isomorphism 𝖵dR,x𝖵dR,x0\mathsf{V}_{dR,x}\cong\mathsf{V}_{dR,x_{0}}, and we use that isomorphism to identify the filtration on 𝖵dR,x\mathsf{V}_{dR,x} with a point of the flag variety \mathcal{H}.

In the complex setting, fix a lift x0~\widetilde{x_{0}} of x0x_{0} to Xan~\widetilde{X^{an}}. For every x~Xan~\widetilde{x}\in\widetilde{X^{an}} lying over some point xXanx\in X^{an}, integrating the connection from x0~\widetilde{x_{0}} to x~\widetilde{x} gives an identification 𝖵dR,x𝖵dR,x0\mathsf{V}_{dR,x}\cong\mathsf{V}_{dR,x_{0}}, which gives by definition a point of PVanPV^{an} lying over xx. Thus, we obtain a complex-analytic map

ι:Xan~PVan\iota_{\mathbb{C}}\colon\widetilde{X^{an}}\rightarrow PV^{an}

lifting the projection Xan~Xan\widetilde{X^{an}}\rightarrow X^{an}.

Similarly, in the pp-adic setting, integration gives a rigid-analytic section

ιp:ΩPVrig\iota_{p}\colon\Omega\rightarrow PV^{rig}

to the torsor PVrigιpPV^{rig}\rightarrow\iota_{p}.

By definition, the complex and pp-adic period maps are given by ΦPVι\Phi_{PV}\circ\iota_{\mathbb{C}} and ΦPVιp\Phi_{PV}\circ\iota_{p}, respectively.

The image of Φ=ΦPVι\Phi_{\mathbb{C}}=\Phi_{PV}\circ\iota_{\mathbb{C}} is contained in a single 𝐆mon\mathbf{G}_{mon}-orbit on \mathcal{H}, and the image of ι\iota_{\mathbb{C}} intersects every 𝐆mon\mathbf{G}_{mon}-orbit on PVanPV^{an}. Thus, by 𝐆mon\mathbf{G}_{mon}-equivariance, the image of ΦPV\Phi_{PV} is itself contained in a single 𝐆mon\mathbf{G}_{mon}-orbit on \mathcal{H}, so the same is true of the image of Φp\Phi_{p}. ∎

Lemma 6.2.

The construction above defines a pp-adic period map

Φp:Ω𝐆mon/Pmon,\Phi_{p}\colon\Omega\to\mathbf{G}_{mon}/P_{mon},

where PmonP_{mon} is a parabolic subgroup of 𝐆mon\mathbf{G}_{mon}.

Proof.

By Lemma 6.1, the pp-adic period map takes values in 𝐆mon/(P𝐆mon)\mathbf{G}_{mon}/(P\cap\mathbf{G}_{mon}), where PGLNP\subseteq GL_{N} is the parabolic subgroup determined by the Hodge cocharacter μ:𝔾mGLN\mu\colon\mathbb{G}_{m}\rightarrow GL_{N}. What remains is to show that P𝐆monP\cap\mathbf{G}_{mon} is parabolic in 𝐆mon\mathbf{G}_{mon}.

We know that μ\mu lies in the generic Mumford-Tate group, and hence normalizes 𝐆mon\mathbf{G}_{mon} [1, §5 Thm. 1]. Since the outer automorphism group of 𝐆mon\mathbf{G}_{mon} is finite and 𝔾m\mathbb{G}_{m} is connected, the cocharacter μ\mu acts on 𝐆mon\mathbf{G}_{mon} by inner automorphisms. Thus the adjoint action of μ\mu by conjugation on 𝐆mon\mathbf{G}_{mon} gives a homomorphism

𝔾mInn𝐆mon\mathbb{G}_{m}\rightarrow\operatorname{Inn}\mathbf{G}_{mon}

to the group of inner automorphisms of 𝐆mon\mathbf{G}_{mon}. Raising to a power if necessary, we can lift μ\mu to a cocharacter

ν:𝔾m𝐆mon.\nu\colon\mathbb{G}_{m}\rightarrow\mathbf{G}_{mon}.

Then the cocharater ν\nu defines the parabolic subgroup Pmon=P𝐆monP_{mon}=P\cap\mathbf{G}_{mon}. ∎

From now on, the parabolic subgroup Pmon𝐆monP_{mon}\subseteq\mathbf{G}_{mon} will simply be called PP. We have defined a pp-adic period map, valued in 𝐆mon/P\mathbf{G}_{mon}/P. To study this map, we’ll need to recall a corollary of the complex Bakker–Tsimerman theorem [3].

Lemma 6.3.

(Complex Bakker–Tsimerman theorem). In the above setting, suppose that Z(𝐆mon/P)Z\subseteq(\mathbf{G}_{mon}/P)_{\mathbb{C}} is an algebraic subvariety, and

codim(𝐆mon/P)Zdim(X).\operatorname{codim}_{(\mathbf{G}_{mon}/P)}Z\geq\dim(X).

Then any irreducible component of Φ1(Z)\Phi_{\mathbb{C}}^{-1}(Z) is contained inside the preimage, in Xan~\widetilde{X^{an}}, of the complex points of a proper subvariety of XX_{\mathbb{C}}.

Proof.

(This is a mild generalization of [47, Corollary 9.2]. The proof is the same; we reproduce it here for the reader’s convenience.)

We will apply [3, Theorem 1.1]. Let Dˇ=(𝐆mon/P)\check{D}=(\mathbf{G}_{mon}/P)_{\mathbb{C}} and V=X×ZXan×DˇV=X\times Z\subseteq X^{an}\times\check{D}. Let WW be the image of Xan~\widetilde{X^{an}} under the analytic map Xan~Xan×Dˇ\widetilde{X^{an}}\mapsto X^{an}\times\check{D} . Let QQ be an irreducible component of Φ1(Z)\Phi_{\mathbb{C}}^{-1}(Z); its image under Xan~Xan×Dˇ\widetilde{X^{an}}\mapsto X^{an}\times\check{D} is contained in some irreducible component of WVW\cap V; call this component UU.

Now we apply [3, Theorem 1.1]. Note that

codimXan×DˇV=codim(𝐆mon/P)ZdimX\operatorname{codim}_{X^{an}\times\check{D}}V=\operatorname{codim}_{(\mathbf{G}_{mon}/P)}Z\geq\dim X

by hypothesis, while

codimXan×DˇW=dimDˇ.\operatorname{codim}_{X^{an}\times\check{D}}W=\dim\check{D}.

On the other hand, we may as well assume dimU>0\dim U>0; otherwise QQ is a point and there is nothing to prove. So we have the strict inequality

codimXan×DˇU<dimX+dimDˇcodimXan×DˇV+codimXan×DˇW,\operatorname{codim}_{X^{an}\times\check{D}}U<\dim X+\dim\check{D}\leq\operatorname{codim}_{X^{an}\times\check{D}}V+\operatorname{codim}_{X^{an}\times\check{D}}W,

and we conclude by [3, Theorem 1.1]. ∎

Theorem 6.4.

(pp-adic Bakker–Tsimerman theorem). Let 𝖵\mathsf{V} be a polarized, integral Hodge–Deligne system on a scheme 𝒳\mathcal{X} smooth over 𝒪K,S\mathcal{O}_{K,S}, and let 𝐆mon\mathbf{G}_{mon} be the differential Galois group of 𝖵\mathsf{V}. Let XX be the base change of 𝒳\mathcal{X} to KK.

Choose a pp-adic residue disk Ω𝒳(𝒪Kv)\Omega\subseteq\mathcal{X}(\mathcal{O}_{K_{v}}), and a basepoint x0Ωx_{0}\in\Omega. By Lemma 6.2, these give rise to a period map

Φp:Ω𝐆mon/P,\Phi_{p}\colon\Omega\rightarrow\mathbf{G}_{mon}/P,

where PP is a parabolic subgroup of 𝐆mon\mathbf{G}_{mon}.

Suppose Z𝐆mon/PZ\subseteq\mathbf{G}_{mon}/P is a closed subscheme such that

codim𝐆mon/PZdimX.\operatorname{codim}_{\mathbf{G}_{mon}/P}Z\geq\dim X.

Then Φp1(Z)ΩX(𝒪Kv)\Phi_{p}^{-1}(Z)\subseteq\Omega\subseteq X(\mathcal{O}_{K_{v}}) is not Zariski-dense in XX.

Theorem 6.4 is stated over a general number field KK for generality, but we will apply it with K=K=\mathbb{Q}.

Proof.

This is a mild generalization of [47, Lemma 9.3]. Specifically, [47, Lemma 9.3] imposes the following additional hypotheses:

  • The Hodge–Deligne system 𝖵\mathsf{V} arises as the cohomology (primitive in middle dimension) of a smooth, proper family of varieties.

  • The Zariski closure 𝐆mon\mathbf{G}_{mon} of the image of monodromy coincides with the full orthogonal or symplectic group.

  • The base field is K=K=\mathbb{Q}.

The proof from loc. cit. goes through in our more general setting essentially without change; here we present the same argument, couched in slightly different language.

First of all, note that for any fixed n>0n>0, ΩX(𝒪Kv)\Omega\subseteq X(\mathcal{O}_{K_{v}}) is covered by finitely many mod-pnp^{n} residue disks; we choose nn such that the image of each of these residue disks under Φp\Phi_{p} is contained in an affine subset UU of 𝐆mon/P\mathbf{G}_{mon}/P. There is no harm in replacing Ω\Omega with one such mod-pnp^{n} residue disk, and we will do so.

Now suppose Z𝐆mon/PZ\subseteq\mathbf{G}_{mon}/P is defined by equations Fi=0F_{i}=0, with each FiF_{i} a regular function on UU. Let Gi=FiΦpG_{i}=F_{i}\circ\Phi_{p}; this is an element of 𝒪(Ωrig)\mathcal{O}(\Omega^{rig}), i.e. an element of the Tate algebra RR of the affinoid disk Ωrig\Omega^{rig}.

We want to show that the common vanishing locus Φp1ZΩrig\Phi_{p}^{-1}Z\cap\Omega^{rig} of the functions GiG_{i} on Ωrig\Omega^{rig} is contained in the zero-set of some regular function on XX. Now Φp1ZΩrig\Phi_{p}^{-1}Z\cap\Omega^{rig} is the union of finitely many irreducible components, so it suffices to show that any one such irreducible component (call it WiW_{i}) is contained in the zero-set of some regular function on XX. As we are only interested in KvK_{v}-points, we may as well assume that Wi(Kv)W_{i}(K_{v}) is nonempty.

Choose a basepoint x0Wi(Kv)x_{0}\in W_{i}(K_{v}) and an isomorphism σ:p¯\sigma\colon\overline{\mathbb{Q}_{p}}\cong\mathbb{C}, and repeat the construction of Lemma 6.1 with respect to the basepoint x0x_{0} (taking K=KvK^{\prime}=K_{v}). In particular, we have the algebraic Picard–Vessiot torsor PVXKvPV\rightarrow X_{K_{v}} under the group 𝐆mon\mathbf{G}_{mon} and the “Picard–Vessiot period map”

ΦPV:PVGLN/P.\Phi_{PV}\colon PV\rightarrow GL_{N}/P.

We also have maps

ι:Xan~PVan\iota_{\mathbb{C}}\colon\widetilde{X^{an}}\rightarrow PV^{an}

and

ιp:ΩPVrig,\iota_{p}\colon\Omega\rightarrow PV^{rig},

such that the complex and pp-adic period maps are given by

Φ=ΦPVι\Phi_{\mathbb{C}}=\Phi_{PV}\circ\iota_{\mathbb{C}}

and

Φp=ΦPVιp,\Phi_{p}=\Phi_{PV}\circ\iota_{p},

respectively.

Recall that both Φ\Phi_{\mathbb{C}} and Φp\Phi_{p} are defined by integrating the connection on 𝖵dR\mathsf{V}_{dR}. In particular, with respect to some fixed coordinates on PVPV and local coordinates on XX, both Φ\Phi_{\mathbb{C}} and Φp\Phi_{p} are given by the same power series (which then has positive radius of convergence both for the complex and the pp-adic metric).

Let us rephrase this, more precisely, in geometric terms. Let X^\hat{X} be the formal completion of XKvX_{K_{v}} at the point x0x_{0}; it is noncanonically isomorphic to (the formal spectrum of) a power series ring in dimX\dim X variables over KvK_{v}. The formal completions of XrigX^{rig} and XanX^{an} at x0x_{0} are related to X^\hat{X} in the obvious way. If we may abuse notation by making a diagram out of objects of three different categories, we have the following inclusions:

Xrig\textstyle{X^{rig}}X^\textstyle{\hat{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Xan.\textstyle{X^{an}.}

Over the formal scheme X^\hat{X}, the Picard–Vessiot torsor PVPV admits a canonical section ιform\iota_{form}, given (as in the proof of Lemma 6.1) by integrating the connection; the sections ιp\iota_{p} and ι\iota_{\mathbb{C}} over XrigX^{rig} and XanX^{an}, respectively, both restrict to ιform\iota_{form} when pulled back to X^\hat{X}. Let Φform=ΦPVιform\Phi_{form}=\Phi_{PV}\circ\iota_{form}; this is the restriction to X^\hat{X} of both the complex and pp-adic period maps. This is the sense in which Φp\Phi_{p} and Φ\Phi_{\mathbb{C}} have the same power series.

Returning to the main argument, we will apply Lemma 6.3 over XanX^{an}, then “pull back to X^\hat{X} and push forward to XrigX^{rig}.”

Pulling ZZ back via σ\sigma gives a variety Zσ(𝐆mon/Pmon)Z^{\sigma}\subseteq(\mathbf{G}_{mon}/P_{mon})_{\mathbb{C}}; its inverse image in Xan~\widetilde{X^{an}} is cut out by analytic functions GiσG_{i}^{\sigma}. (The functions have the same power series as the functions GiG_{i}, regarded now as complex power series via the homomorphism σ\sigma.)

Now we apply the complex Bakker–Tsimerman theorem (Lemma 6.3) to the variety ZσZ^{\sigma}. As a result, we obtain some regular function HH on the scheme XX_{\mathbb{C}} that vanishes (pointwise) on Φ1(Z)\Phi_{\mathbb{C}}^{-1}(Z). By the locally analytic Nullstellensatz [35, §3.4], some power HmH^{m} lies inside the ideal generated by the functions GiσG_{i}^{\sigma} inside the ring of germs of analytic functions at x0x_{0}.

Pulling back to X^\hat{X}, it is clear that HmH^{m} also vanishes on the formal subscheme Φform1(Z)\Phi_{form}^{-1}(Z) of X^\hat{X}.

By means of our chosen isomorphism σ:p¯\sigma\colon\overline{\mathbb{Q}_{p}}\rightarrow\mathbb{C}, we obtain a regular function (Hm)σ1(H^{m})^{\sigma^{-1}} on Xp¯X_{\overline{\mathbb{Q}_{p}}} that again vanishes on Φform1(Z)\Phi_{form}^{-1}(Z). This (Hm)σ1(H^{m})^{\sigma^{-1}} must in fact be defined over some finite extension of KvK_{v}; taking a norm, we may assume that it is in fact a regular function on XKvX_{K_{v}}. For simplicity, let us call this function HH.

At this point we have a regular function HH on XKvX_{K_{v}}, and we know that it vanishes on Φform1(Z)X^\Phi_{form}^{-1}(Z)\subseteq\hat{X}. To conclude, we need to show that HH vanishes on the rigid-analytic neighborhood WiW_{i} of x0x_{0}. To this end, recall that RR was the Tate algebra of the affinoid disk Ωrig\Omega^{rig}; let IRI\subseteq R be the ideal defining WiW_{i}. “Clearing denominators” if necessary in the regular function HH, we may assume HH lies in the Tate algebra RR. Let 𝔪\mathfrak{m} denote the ideal of R/IR/I defining the point x0x_{0}. We know that H𝔪i(R/I)H\in\mathfrak{m}^{i}(R/I) for every integer i0i\geq 0; by Krull’s Intersection Theorem, we conclude that HH vanishes in R/IR/I, that is, that HIH\in I. Thus WiW_{i} is contained in the vanishing locus of the algebraic function HH. ∎

6.2. Complex monodromy

Our next goal is Lemma 6.8, which shows that for any positive integer cc, we can construct our 𝖵\mathsf{V} in such a way that its differential Galois group is strongly cc-balanced (Definition 6.6 below). This is a “big monodromy” statement, analogous to [47, Lemma 4.3] and [47, Theorem 8.1].

Lemma 6.5.

Let HH be one of the algebraic groups SLNSL_{N}, SpNSp_{N}, or SONSO_{N}. Let GG be a subgroup of HdH^{d} such that each of the dd coordinate projections πi:GH\pi_{i}\colon G\rightarrow H (for 1id1\leq i\leq d) is surjective.

Define a relation \sim on the index set {1,,d}\{1,\ldots,d\} by declaring that iji\sim j if and only if the projection

(πi,πj):GH2(\pi_{i},\pi_{j})\colon G\rightarrow H^{2}

is not surjective.

  1. (1)

    The relation \sim is an equivalence relation.

  2. (2)

    If i1,,ici_{1},\ldots,i_{c} are a complete set of representatives for \sim, then the map

    (πi1,,πic):GHc(\pi_{i_{1}},\ldots,\pi_{i_{c}})\colon G\rightarrow H^{c}

    is surjective with finite kernel.

Proof.

This is an algebraic version of Goursat’s lemma. See [47, Lemma 2.12] and [55, Lemma 5.2.1].

The group HH has a finite center; call it ZZ. For any two indices ii and jj (possibly equal), there are two possibilities for the image of the projection

G(πi,πj)H2(H/Z)2.G\stackrel{{\scriptstyle(\pi_{i},\pi_{j})}}{{\rightarrow}}H^{2}\rightarrow(H/Z)^{2}.

Either the map is surjective, or its image is the graph of an automorphism of H/ZH/Z. In the former case, (πi,πj)(\pi_{i},\pi_{j}) must surject (onto H2H^{2}) as well.

An easy calculation shows that \sim is an equivalence relation. If i1,,ici_{1},\ldots,i_{c} is a complete system of representatives for \sim, then repeated application of Goursat’s lemma shows that

(πi1,,πic):GHc(\pi_{i_{1}},\ldots,\pi_{i_{c}})\colon G\rightarrow H^{c}

is surjective with finite kernel. ∎

Definition 6.6.

Let HH be one of the algebraic groups SLNSL_{N}, SpNSp_{N}, or ONO_{N}, and GG a subgroup of HdH^{d}. For 1id1\leq i\leq d, let πi:GH\pi_{i}\colon G\rightarrow H be the coordinate projection, and suppose that each πi\pi_{i} is surjective. The index classes of GG are the equivalence classes of the relation \sim of Lemma 6.5.

Let cc be a positive integer777not necessarily the same as the cc of Lemma 6.5. We say that GG is cc-balanced (as a subgroup of HdH^{d}) if its index classes are all of equal size, and there are at least cc of them.

Suppose now we are given a permutation σ\sigma of the index set {1,,d}\{1,\ldots,d\}. (In the sequel, σ\sigma will come from Frobenius.) We say GG is strongly cc-balanced (with respect to σ\sigma) if it is cc-balanced, each orbit of σ\sigma on {1,,d}\{1,\ldots,d\} contains elements of at least cc of the index classes of GG, and σ\sigma preserves the partition of {1,,d}\{1,\ldots,d\} into index classes.

Finally, let E0,E,𝐇,𝐆dR,𝐆dR0E_{0},E,\mathbf{H},\mathbf{G}_{dR},\mathbf{G}^{0}_{dR} be as in Section 5.6, and let 𝐆dR1\mathbf{G}^{1}_{dR} be:

  • in the case 𝐇=GLN\mathbf{H}=GL_{N}, the kernel of the determinant map 𝐆dR0𝔾m,E\mathbf{G}^{0}_{dR}\rightarrow\mathbb{G}_{m,E}

  • in the case 𝐇=GSpN\mathbf{H}=GSp_{N}, the kernel of the similitude character 𝐆dR0𝔾m,E\mathbf{G}^{0}_{dR}\rightarrow\mathbb{G}_{m,E}

  • in the case 𝐇=GON\mathbf{H}=GO_{N}, the intersection of the kernels of the determinant map 𝐆dR0𝔾m,E\mathbf{G}^{0}_{dR}\rightarrow\mathbb{G}_{m,E} and the similitude character 𝐆dR0𝔾m,E\mathbf{G}^{0}_{dR}\rightarrow\mathbb{G}_{m,E} .

Then 𝐆dR1\mathbf{G}^{1}_{dR} is a form of SLNdSL_{N}^{d}, SpNdSp_{N}^{d}, or ONdO_{N}^{d}. We say that an algebraic subgroup G𝐆dR0G\subseteq\mathbf{G}^{0}_{dR} is cc-balanced (resp. strongly cc-balanced with respect to σ\sigma) if (G𝐆dR1)E¯(G\cap\mathbf{G}^{1}_{dR})_{\overline{E}} is cc-balanced (resp. strongly cc-balanced with respect to σ\sigma) as a subgroup of SLndSL_{n}^{d}, SpNdSp_{N}^{d} or ONdO_{N}^{d}.

Note that the condition of being strongly cc-balanced gets stronger as cc grows. We will later choose cc to be sufficiently large to satisfy some inequalities given in Theorem 8.17. We do that in the proof of Theorem 9.2; until then, all our statements will be proven for an arbitrary natural number cc.

Example 6.7.

Let ΔH2\Delta\subseteq H^{2} be the diagonal, and let G=Δ×Δ×Δ×ΔH8G=\Delta\times\Delta\times\Delta\times\Delta\subseteq H^{8}.

The index classes of GG are {1,2},{3,4},{5,6},{7,8}\{1,2\},\{3,4\},\{5,6\},\{7,8\}; thus GG is cc-balanced for c=1,2,3,4c=1,2,3,4.

Let σ=(13)(24)(57)(68)\sigma=(13)(24)(57)(68). Then GG is strongly cc-balanced with respect to σ\sigma only for c=1,2c=1,2.

(We thank the anonymous referee for this example.)

Lemma 6.8.

Recall notation from Section 5.1. Specifically, let KK, pp, AA, XX, YY, 𝒜\mathcal{A}, 𝒳\mathcal{X}, 𝒴\mathcal{Y}, LL be as in Section 5.5, and let vv be a place of KK over pp. Fix some embedding ι0:KL\iota_{0}\colon K\rightarrow L. Fix a natural number cc, and pick χ0\chi_{0} as in Corollary 4.10 (depending on cc), with G=SLN,SpNG^{*}=SL_{N},Sp_{N}, or SONSO_{N}, and let E0,E,𝐇,𝐆dR,𝐆dR0E_{0},E,\mathbf{H},\mathbf{G}_{dR},\mathbf{G}^{0}_{dR} be as in Sections 5.6 and 5.7. Let II be the full (Gal×Galcyc/)(\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{cyc}/\mathbb{Q}})-orbit containing (ι0,χ0)(\iota_{0},\chi_{0}), and let 𝖵=𝖵I\mathsf{V}=\mathsf{V}_{I} be the corresponding Hodge–Deligne system.

The Frobenius at vv, acts on the set II through the diagonal action as an element of (Gal×Galcyc/)(\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{cyc}/\mathbb{Q}}); call this permutation σ\sigma.

Then the differential Galois group 𝐆mon\mathbf{G}_{mon} of 𝖵\mathsf{V} (base-changed to p\mathbb{Q}_{p}) is a strongly cc-balanced subgroup of 𝐆dR0\mathbf{G}^{0}_{dR}, with respect to σ\sigma.

Proof.

This is little more than a restatement of Corollary 4.10.

After base change to \mathbb{C}, we have that 𝐇\mathbf{H}_{\mathbb{C}} is either GSpGSp, GOGO, or GLGL; and 𝐆dR,0\mathbf{G}^{0}_{dR,\mathbb{C}} splits as a product of copies of 𝐇\mathbf{H}_{\mathbb{C}}, indexed by pairs (ι,χ)I(\iota,\chi)\in I. We’ll write each of these direct factors of 𝐆dR,0\mathbf{G}^{0}_{dR,\mathbb{C}} as 𝐇(ι,χ)\mathbf{H}_{(\iota,\chi)}.

The differential Galois group, after base change to \mathbb{C}, is the Zariski closure of the monodromy of the variation of Hodge structure. The variation of Hodge structure 𝖵H\mathsf{V}_{H} splits as the direct sum of 𝖵H,(ι,χ)=Rkfιgιχ\mathsf{V}_{H,(\iota,\chi)}=R^{k}{f_{\iota}}_{*}g_{\iota}^{*}\mathcal{L}_{\chi}. Clearly monodromy acts trivially on the set of pairs (ι,χ)(\iota,\chi), and when 𝐇\mathbf{H} is GSpGSp or GOGO, there is a bilinear pairing on RkfιgιχR^{k}{f_{\iota}}_{*}g_{\iota}^{*}\mathcal{L}_{\chi}. Thus, 𝐆mon𝐆dR0\mathbf{G}_{mon}\subseteq\mathbf{G}^{0}_{dR}.

Now Corollary 4.10 implies that the geometric monodromy group of 𝖵H,(ι0,χ0)\mathsf{V}_{H,(\iota_{0},\chi_{0})} is all of 𝐇()\mathbf{H}(\mathbb{C}). By symmetry under the action of (Gal×Galcyc/)(\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{cyc}/\mathbb{Q}}), the same is true for all (ι,χ)I(\iota,\chi)\in I, so 𝐆mon\mathbf{G}_{mon} surjects onto each factor 𝐇\mathbf{H} of 𝐆dR0\mathbf{G}^{0}_{dR}. Thus, can talk about the relation \sim and the index classes of Definition 6.6.

The action of (Gal×Galcyc/)(\operatorname{Gal}_{\mathbb{Q}}\times\operatorname{Gal}_{\mathbb{Q}^{cyc}/\mathbb{Q}}) respects the relation \sim; in particular the index classes are all of the same size, and Frobenius respects the relation \sim. Corollary 4.10 shows that every σ\sigma-orbit in II contains elements of at least cc index classes. Specifically, note that we chose pp such that this Galois action is unramified at pp, and thus the local Galois group is generated by Frobenius. So for any (ι,χ)I(\iota^{\prime},\chi^{\prime})\in I, the conclusion of Corollary 4.10 gives cc elements (ι1,χ1),,(ιc,χc)(\iota_{1},\chi_{1}),\ldots,(\iota_{c},\chi_{c}) of the σ\sigma-orbit of (ι,χ)(\iota^{\prime},\chi^{\prime}) in II, such that the projection

𝐆mon𝐆dR0i=1c𝐇(ιi,χi)\mathbf{G}_{mon}\hookrightarrow\mathbf{G}^{0}_{dR}\twoheadrightarrow\prod_{i=1}^{c}\mathbf{H}_{(\iota_{i},\chi_{i})}

contains (G)c(G^{*})^{c}, where GG^{*} is either SLNSL_{N}, SpNSp_{N}, or ONO_{N}. It follows that (𝐆mon𝐆dR1)(\mathbf{G}_{mon}\cap\mathbf{G}^{1}_{dR})_{\mathbb{C}} is strongly cc-balanced with respect to σ\sigma. ∎

7. Hodge–Deligne systems and integral points, assuming global semisimplicity

In this section we prove Theorem 7.3, a variant of Theorem 8.17 that assumes the semisimplicity of certain global Galois representations. This material is not logically needed for the main argument; we include it to illustrate the main ideas of Section 8, without the complications coming from semisimplification.

Lemma 7.1.

(Compare Lemma 8.10.)

Let pp be a prime. A semisimple representation

ρ0:G𝐆et\rho_{0}\colon G_{\mathbb{Q}}\rightarrow\mathbf{G}_{et}

of the global Galois group GG_{\mathbb{Q}}, crystalline at pp, such that the composition G𝐆etAutp𝖤etG_{\mathbb{Q}}\rightarrow\mathbf{G}_{et}\rightarrow\operatorname{Aut}_{\mathbb{Q}_{p}}\mathsf{E}_{et} agrees with the usual action of Galois on 𝖤et\mathsf{E}_{et}, gives rise by pp-adic Hodge theory to an admissible filtered ϕ\phi-module (V0,ϕ0,F0)(V_{0},\phi_{0},F_{0}) with 𝐆dR\mathbf{G}_{dR}-structure. Suppose another crystalline global representation ρ:G𝐆\rho\colon G_{\mathbb{Q}}\rightarrow\mathbf{G} is isomorphic to ρ0\rho_{0}, and call the corresponding filtered ϕ\phi-module (V,ϕ,F)(V,\phi,F). Then there is an isomorphism of filtered ϕ\phi-modules

(V0,ϕ0,F0)(V,ϕ,F).(V_{0},\phi_{0},F_{0})\cong(V,\phi,F).

In particular, if (V0,ϕ0)=(V,ϕ)(V_{0},\phi_{0})=(V,\phi), then there exists an automorphism f:VVf\colon V\rightarrow V such that

  • f𝐆dRf\in\mathbf{G}_{dR},

  • ff commutes with ϕ\phi, and

  • fF0=FfF_{0}=F.

Proof.

The fact that representations 𝐆et\mathbf{G}_{et} structure are sent to fitlered ϕ\phi-modules with 𝐆dR\mathbf{G}_{dR}-structure is from Lemma 5.43.

The existence of isomorphisms is because functors of pp-adic Hodge theory take isomorphic objects to isomorphic objects. ∎

Lemma 7.2.

(Compare Lemma 8.16.)

Assume we are in the setting of Section 5.6. Fix an admissible filtered ϕ\phi-module with 𝐆dR\mathbf{G}_{dR}-structure (V0,ϕ0,F0)(V_{0},\phi_{0},F_{0}), and another ϕ\phi-module (V,ϕ)(V,\phi) with 𝐆dR\mathbf{G}_{dR}-structure; suppose both ϕ\phi and ϕ0\phi_{0} are semilinear over some σAutE0E\sigma\in\operatorname{Aut}_{E_{0}}E.

Let 𝐆mon\mathbf{G}_{mon} be a subgroup of 𝐆dR0\mathbf{G}^{0}_{dR}, strongly cc-balanced with respect to σ\sigma for some positive integer cc.

Suppose F0F_{0} is uniform in the sense of Definition 5.46, and let ha=hsimpah^{a}=h^{a}_{simp} be the adjoint Hodge numbers on 𝐇\mathbf{H}. Suppose ee is a positive integer satisfying the following numerical condition.

  • (Numerical condition.)

    a>0ha1c(e+dim𝐇).\sum_{a>0}h^{a}\geq\frac{1}{c}(e+\dim\mathbf{H}).

Let =𝐆mon/(Q0𝐆mon)\mathcal{H}=\mathbf{G}_{mon}/(Q^{0}\cap\mathbf{G}_{mon}) be the flag variety parametrizing filtrations on 𝐆dR0\mathbf{G}^{0}_{dR} that are conjugate to F0F_{0} under the conjugation of 𝐆mon\mathbf{G}_{mon}. Then the filtrations FF such that (V,ϕ,F)(V,\phi,F) is isomorphic to (V0,ϕ0,F0)(V_{0},\phi_{0},F_{0}) are of codimension at least ee in \mathcal{H}.

Proof.

By Lemma 7.1, the filtrations FF satisfying the condition described form at most one orbit under the action of Z(ϕ)Z(\phi) on \mathcal{H}. We will show that any orbit of Z(ϕ)Z(\phi) on \mathcal{H} has codimension at least ee. This is a question about the dimension of a variety over E0E_{0}; by passing to an extension, we may assume that 𝐆dR0=𝐇d\mathbf{G}^{0}_{dR}=\mathbf{H}^{d} is split, and σ\sigma acts by permuting the factors. Call the dd factors 𝐇1,,𝐇d\mathbf{H}_{1},\ldots,\mathbf{H}_{d}.

The Frobenius element σAutE0E\sigma\in\operatorname{Aut}_{E_{0}}E gives a permutation of the index set {1,,d}\{1,\ldots,d\}, which we also call σ\sigma. Semilinearity over σ\sigma means that the map ϕ\phi permutes the dd factors according to the permutation σ\sigma.

Let JJ be an orbit of σ\sigma on the index set {1,,d}\{1,\ldots,d\}. By Definition 6.6, JJ must contain elements of at least cc distinct index classes. Let IJI\subseteq J be a system of representatives for the index classes appearing in JJ. This II must have at least cc elements, and in fact there is no harm in increasing cc so that #I=c\#I=c exactly.

Now everything in sight splits as a direct product. Let 𝐇I=iI𝐇i\mathbf{H}^{I}=\prod_{i\in I}\mathbf{H}_{i}; define 𝐇J\mathbf{H}^{J} similarly. Let E0JE_{0}^{J} be the direct summand of EE corresponding to the σ\sigma-orbit JJ.

Since the elements of II belong to distinct index classes, the projection

𝐆mon(𝐇/Z(𝐇))I\mathbf{G}_{mon}\rightarrow(\mathbf{H}/Z(\mathbf{H}))^{I}

has image a union of connected components of the target, so it is smooth with equidimensional fibers, and the same is true of

=𝐆mon/(Q0𝐆mon)𝐇I/(Q0,I𝐇I)=:I.\mathcal{H}=\mathbf{G}_{mon}/(Q^{0}\cap\mathbf{G}_{mon})\rightarrow\mathbf{H}^{I}/(Q^{0,I}\cap\mathbf{H}^{I})=\colon\mathcal{H}^{I}.

Let Z𝐇I(ϕ)Z_{\mathbf{H}^{I}}(\phi) be projection to 𝐇I\mathbf{H}^{I} of ZG(ϕ)Z_{G}(\phi); this is the set of elements of 𝐇I\mathbf{H}^{I} that commute with ϕ\phi, when 𝐇I\mathbf{H}^{I} is viewed as a direct summand of 𝐆dR0\mathbf{G}^{0}_{dR}. Define Z𝐇J(ϕ)Z_{\mathbf{H}^{J}}(\phi) similarly.

To finish the proof, it is enough to show that any orbit of Z𝐇I(ϕ)Z_{\mathbf{H}^{I}}(\phi) on I\mathcal{H}^{I} has codimension at least ee. By Lemma 5.33, applied with E0JE_{0}^{J} in place of EE, we have

dimZ𝐇J(ϕ)dim𝐇.\dim Z_{\mathbf{H}^{J}}(\phi)\leq\dim\mathbf{H}.

Since Z𝐇I(ϕ)Z_{\mathbf{H}^{I}}(\phi) is the projection of Z𝐇J(ϕ)Z_{\mathbf{H}^{J}}(\phi) to 𝐇I\mathbf{H}^{I}, we deduce

dimZ𝐇I(ϕ)dim𝐇.\dim Z_{\mathbf{H}^{I}}(\phi)\leq\dim\mathbf{H}.

On the other hand, for any reductive group GG and filtration FF on GG, corresponding to a parabolic subgroup QQ, the sum a>0ha\sum_{a>0}h^{a} of the adjoint Hodge numbers (Definition 5.44) is precisely dimG/Q\dim G/Q. Apply this with G=𝐇IG=\mathbf{H}^{I}; since the Hodge numbers F0F_{0} are uniform, we can compute the adjoint Hodge numbers on I\mathcal{H}^{I} by Lemma 5.48; we find

dimI=ca>0ha.\dim\mathcal{H}^{I}=c\sum_{a>0}h^{a}.

The result follows. ∎

Theorem 7.3.

(Compare Theorem 8.17.)

Let XX be a variety over \mathbb{Q}, let SS be a finite set of primes of \mathbb{Z}, and let 𝒳\mathcal{X} be a smooth model of XX over [1/S]\mathbb{Z}[1/S].

Let 𝖤\mathsf{E} be a constant H0H^{0}-algebra on 𝒳\mathcal{X}, and let 𝐇\mathbf{H} be one of GLNGL_{N}, GSpNGSp_{N}, or GONGO_{N}. Let 𝖵\mathsf{V} be a polarized, integral, 𝖤\mathsf{E}-module with 𝐇\mathbf{H}-structure on 𝒳\mathcal{X}, in the sense of Definition 5.27, having integral Frobenius eigenvalues (Def. 5.5). Suppose the Hodge numbers of 𝖵\mathsf{V} are uniform in the sense of Definition 5.46, and let ha=hsimpah^{a}=h^{a}_{simp} be the adjoint Hodge numbers on 𝐇\mathbf{H}. Let 𝐆dR0\mathbf{G}^{0}_{dR}, 𝐆dR\mathbf{G}_{dR} be as in Section 5.7.

Suppose there is a positive integer cc such that 𝖵\mathsf{V} satisfies the following conditions.

  • (Big monodromy.) If 𝐆mon𝐆dR0\mathbf{G}_{mon}\subseteq\mathbf{G}^{0}_{dR} is the differential Galois group of 𝖵\mathsf{V}, then 𝐆mon𝐆dR0\mathbf{G}_{mon}\subseteq\mathbf{G}^{0}_{dR} is strongly cc-balanced with respect to Frobenius. (The Frobenius is determined from the structure of 𝖤\mathsf{E}; see Section 5.7.)

  • (Numerical condition.)

    a>0ha1c(dimX+dim𝐇).\sum_{a>0}h^{a}\geq\frac{1}{c}(\dim X+\dim\mathbf{H}).

Let 𝒳([1/S])ss\mathcal{X}(\mathbb{Z}[1/S])^{ss} be the subset of 𝒳([1/S])\mathcal{X}(\mathbb{Z}[1/S]) consisting of those xx for which the global Galois representation 𝖵et,x\mathsf{V}_{et,x} is semisimple. Then the image of 𝒳([1/S])ss\mathcal{X}(\mathbb{Z}[1/S])^{ss} is not Zariski dense in XX.

Proof.

For every x𝒳([1/S])x\in\mathcal{X}(\mathbb{Z}[1/S]), consider the semisimple global Galois representation ρx=𝖵et,x\rho_{x}=\mathsf{V}_{et,x}. By Lemma 5.49, there are only finitely many possible isomorphism classes for the semisimple representation ρx\rho_{x}. So it is enough to show, for any fixed ρ0\rho_{0}, that the set

𝒳([1/S],ρ0):={xX([1/S])|ρxρ0}\mathcal{X}(\mathbb{Z}[1/S],\rho_{0}):=\{x\in X(\mathbb{Z}[1/S])|\rho_{x}\cong\rho_{0}\}

is not Zariski dense in XX. By Lemma 7.1, it is enough, for each residue disk ΩX(p)\Omega\subseteq X(\mathbb{Z}_{p}), to show that the set

X(Ω,(V0,ϕ0,F0))={xΩ|(V,ϕ,F)(V0,ϕ0,F0)}X(\Omega,(V_{0},\phi_{0},F_{0}))=\{x\in\Omega|(V,\phi,F)\cong(V_{0},\phi_{0},F_{0})\}

is not Zariski dense in XX.

We are now in the setting of Theorem 6.4. For xΩx\in\Omega, the filtered ϕ\phi-module 𝖵cris,x\mathsf{V}_{cris,x} is of the form (V,ϕ,Fx)(V,\phi,F_{x}), where (V,ϕ)(V,\phi) is independent of xx. (This is a general property of FF-isocrystals; see remarks in the proof of Theorem 8.17.) The variation of FxF_{x} with xx is classified by a pp-adic period map

Φp:Ω𝐆dR0/Q,\Phi_{p}\colon\Omega\rightarrow\mathbf{G}^{0}_{dR}/Q,

where 𝐆dR0/Q\mathbf{G}^{0}_{dR}/Q is the flag variety classifying 𝐆dR0\mathbf{G}^{0}_{dR}-filtrations on VV.

In fact, the pp-adic period map lands in a single 𝐆mon\mathbf{G}_{mon}-orbit on 𝐆dR0/Q\mathbf{G}^{0}_{dR}/Q; we can write the orbit as 𝐆mon/(Q𝐆mon)\mathbf{G}_{mon}/(Q\cap\mathbf{G}_{mon}). By Lemma 7.2 with e=dimXe=\dim X, there is a Zariski-closed set Z𝐆mon/(Q𝐆mon)Z\subseteq\mathbf{G}_{mon}/(Q\cap\mathbf{G}_{mon}) of codimension at least dimX\dim X such that, if (V,ϕ,Fx)(V0,ϕ0,F0)(V,\phi,F_{x})\cong(V_{0},\phi_{0},F_{0}), then xΦp1(Z)x\in\Phi_{p}^{-1}(Z). We conclude by Theorem 6.4. ∎

8. Hodge–Deligne systems and integral points

The goal of this section is to prove Theorem 8.17, which gives Zariski non-density of integral points on XX. Recall the setup from Section 5: we have a smooth variety XX over \mathbb{Q}888 By Weil restriction, we will assume that XX is defined over \mathbb{Q}, not a general number field; see the proof of Theorem 9.2. and a Hodge–Deligne system 𝖵\mathsf{V} on an integral model 𝒳\mathcal{X} of XX, satisfying certain conditions. Lemma 5.49 tells us that, as xx ranges over the integral points of XX, there are only finitely many possibilities, up to semisimplification, for the global Galois representation 𝖵x,et\mathsf{V}_{x,et}. We will use this to bound the integral points of XX.

A significant technical obstacle (both in [47] and here) is that Lemma 5.49 only applies to semisimple global Galois representations; there may be many different Galois representations arising as fibers of 𝖵et\mathsf{V}_{et}, all of which have the same semisimplification. If we assume the Grothendieck–Serre conjecture—that all Galois representations that arise are semisimple—this difficulty does not arise. Under this assumption, the argument is very simple; see Section 7.

To apply Lemma 5.49 without assuming semisimplicity, we need to recognize when two global Galois representations might have the same semisimplification. We only have access to the local representations, which are generally far from semisimple. The key idea is that any filtration on the global representation, when restricted to the local representation, must be of a special form.

Passing from local Galois representations via pp-adic Hodge theory to filtered ϕ\phi-modules, we have the following situation. We’re given a ϕ\phi-module (V,ϕ)(V,\phi) with varying filtration F=FF=F^{\bullet}; the variation of FF is described by a “monodromy group” 𝐆mon\mathbf{G}_{mon} (the differential Galois group attached to our Hodge–Deligne system). We will consider the associated graded of (V,ϕ,F)(V,\phi,F) with respect to a ϕ\phi-stable filtration 𝔣\mathfrak{f} – eventually we will take 𝔣\mathfrak{f} to be a semisimplification filtration on the global Galois representation. By Lemma 8.7 below, if 𝔣\mathfrak{f} comes from a filtration on the global representation, then the associated graded of FF with respect to 𝔣\mathfrak{f} is a balanced filtration (Definition 8.4). So we want to know, for how many choices of FF does there exist a ϕ\phi-stable 𝔣\mathfrak{f}, such that the associated graded of (V,ϕ,F)(V,\phi,F) with respect to 𝔣\mathfrak{f} lies in a given balanced isomorphism class? We will bound the dimension of such FF in the flag variety. This material is very similar to the combinatorial arguments in [47, §10-11].

To start with, we’ll recall some results from [47] to limit the reducibility of global representations. The following result is the reason we need to work with \mathbb{Q}-varieties XX, instead of varieties over an arbitrary number field. The natural generalization to representations of GalK\operatorname{Gal}_{K}, with KK a number field, is false – for a counterexample, take KK a CM field, and VV a one-dimensional representation coming from a CM elliptic curve.

Lemma 8.1.

Let pp be a prime, and let VV be a representation of Gal\operatorname{Gal}_{\mathbb{Q}} on a p\mathbb{Q}_{p}-vector space which is crystalline at pp, and such that at all primes \ell outside of a finite set SS, the characteristic polynomial of Frobenius has algebraic coefficients and all roots rational \ell-Weil numbers of weight ww.

Let VdR=(VpBcris)GalpV_{dR}=(V\otimes_{\mathbb{Q}_{p}}B_{\mathrm{cris}})^{\operatorname{Gal}_{\mathbb{Q}_{p}}} be the filtered p\mathbb{Q}_{p}-vector space that is associated to ρ|p\rho|_{\mathbb{Q}_{p}} by the pp-adic Hodge theory functor D¯cris\underline{D}_{\mathrm{cris}} of [23, Expose III].

Then the average weight of the Hodge filtration on VdRV_{dR} equals w/2w/2.

Proof.

This is [47, Lemma 2.9] applied to \mathbb{Q}, which has no CM subfield; the condition that pp be a “friendly place” is automatically satisfied over \mathbb{Q}. ∎

Our first goal is to rephrase Lemma 8.1 in terms of filtrations on reductive groups; the resulting statement is Lemma 8.7 below.

We work with filtrations and semisimplifications relative to a group GG whose identity component is reductive. When GG is disconnected, recall the notions of “parabolic subgroup,” “filtration” and so forth from Section 5.8.

We use the following notation (consistent with [47]): QQ is the parabolic subgroup of GG corresponding to the Hodge filtration FF, while PP corresponds to a semisimplification filtration 𝔣\mathfrak{f}. The group MM is a Levi subgroup associated to PP, corresponding to the associated graded of 𝔣\mathfrak{f}.

Fix GG, PP, and MM. In the (connected) reductive case [47, Lemma 11.2] defines a map from filtrations FF on GG to filtrations FMF_{M} on MM; we need to extend this result to non-connected groups GG as well. Recall (Section 5.8.2) that for any GG-filtration FF, there is a cocharacter μ:𝔾mG\mu\colon\mathbb{G}_{m}\rightarrow G defining FF. The substance of [47, Lemma 11.2] is that μ\mu can be chosen to take values in PP. Projecting from PP to the Levi subgroup MM gives a filtration on MM, which is independent of the choice of μ\mu.

Lemma 8.2.

Suppose GG is an algebraic group, whose identity component is reductive, over a field of characteristic zero. Let PP be a parabolic subgroup of GG, and MM a Levi subgroup associated to PP.

Fix a filtration FF on GG. Then there exists a cocharacter μ:𝔾mP\mu\colon\mathbb{G}_{m}\rightarrow P (with image in PP) defining FF. Furthermore, if FMF_{M} is the filtration on MM defined by the composite map

𝔾mPM,\mathbb{G}_{m}\rightarrow P\rightarrow M,

then FMF_{M} is independent of the choice of μ\mu.

Proof.

For GG connected reductive this is [47, Lemma 11.2].

In the general case, by [47, Lemma 11.2] applied to the identity component G0G^{0}, we know that μ\mu can be chosen with image in PG0P\cap G^{0}, and the corresponding filtration on MG0M\cap G^{0} is independent of the choice of μ\mu. But the filtration defined by μ\mu on MG0M\cap G^{0} determines the filtration defined by μ\mu on MM; see Section 5.8.3. ∎

Definition 8.3.

Given GG, PP, MM, FF as in Lemma 8.2, we call the filtration FMF_{M} on GG the associated graded filtration, and write it as FM=GrMFF_{M}=\operatorname{Gr}_{M}F. (It is well-defined by Lemma 8.2.)

We need a generalization of the notion of “balanced filtration” from [47, §11.1, 11.4]. Given a group SS whose identity component is reductive, we define

𝔞S=X(ZS0)=(X(S0)),\mathfrak{a}_{S}=X_{*}(Z_{S^{0}})\otimes\mathbb{Q}=(X^{*}(S^{0})\otimes\mathbb{Q})^{\vee},

where ZS0Z_{S^{0}} is the center of the identity component S0S^{0} of SS. A cocharacter μ\mu of SS defines a class w(μ)=wS(μ)𝔞Sw(\mu)=w_{S}(\mu)\in\mathfrak{a}_{S} by

(χμ)(t)=twS(μ),χ(\chi\circ\mu)(t)=t^{\langle w_{S}(\mu),\chi\rangle}

for all χX(S0)\chi\in X^{*}(S_{0}). In other words: χμ\chi\circ\mu is an automorphism of 𝔾m\mathbb{G}_{m}, so it is of the form ttαt\mapsto t^{\alpha}; we choose wS(μ)w_{S}(\mu) so that wS(μ),χ=α\langle w_{S}(\mu),\chi\rangle=\alpha, for all χ\chi. We call w(μ)=wS(μ)w(\mu)=w_{S}(\mu) the weight of μ\mu.

Let GG, PP, MM be as above. Then the inclusion Z(G0)Z(M0)Z(G^{0})\rightarrow Z(M^{0}) gives a map ιGM:𝔞G𝔞M\iota_{GM}\colon\mathfrak{a}_{G}\rightarrow\mathfrak{a}_{M}. Furthermore, the parabolic PP defines a preferred element of (𝔞M)(\mathfrak{a}_{M})^{\vee}, the modular character γP\gamma_{P}, defined999Since the sign convention is important, we provide an example. If PGL2P\subseteq GL_{2} is the group of upper triangular matrices, then – identifying MM with the group of diagonal matrices – the character γP\gamma_{P} is given by γP((a00b))=a1b.\gamma_{P}\left(\begin{pmatrix}a&0\\ 0&b\end{pmatrix}\right)=a^{-1}b. as the inverse of the determinant of the adjoint representation of MM on the Lie algebra of PP.

Definition 8.4.

Suppose given GG an algebraic group whose identity component is reductive, PP a parabolic subgroup, and MM a Levi subgroup associated to PP. Let FMF_{M} be a filtration on MM, given by a cocharacter μ:𝔾mM\mu\colon\mathbb{G}_{m}\rightarrow M.

  • We say that FMF_{M} is balanced with respect to PP if wM(μ)=ιGM(wG(μG))w_{M}(\mu)=\iota_{GM}(w_{G}(\mu_{G})), where μG\mu_{G} is the cocharacter 𝔾mμMG\mathbb{G}_{m}\stackrel{{\scriptstyle\mu}}{{\rightarrow}}M\hookrightarrow G.

  • We say that FMF_{M} is weakly balanced if γP(μ)=0\gamma_{P}(\mu)=0 for γP\gamma_{P} the modular character of PP.

  • We say that FMF_{M} is semibalanced if γP(μ)0\gamma_{P}(\mu)\leq 0, for γP\gamma_{P} the modular character of PP.

We say a GG-filtration FF is balanced (resp. weakly balanced, semibalanced) with respect to PP (or MM, or 𝔣\mathfrak{f}) if the associated graded GrMF\operatorname{Gr}_{M}F is so.

We remark that a GG-filtration FF is balanced with respect to PP if and only if the associated G0G^{0}-filtration is balanced with respect to P0P^{0}.

Remark 8.5.

Balanced implies weakly balanced because

γP(ιGM(wG(μ)))=0;\gamma_{P}(\iota_{GM}(w_{G}(\mu)))=0;

this identity boils down to the fact that the center of GG acts trivially through the adjoint representation on the Lie algebra of PP.

Furthermore, weakly balanced implies semibalanced.

Example 8.6.

Let G=GL6G=GL_{6}, acting on the space VV with standard basis vectors e1,,e6e_{1},\ldots,e_{6}. Let 𝔣\mathfrak{f} be the filtration with

𝔣0\displaystyle\mathfrak{f}_{0} =\displaystyle= V\displaystyle V
𝔣1\displaystyle\mathfrak{f}_{1} =\displaystyle= span(e1,e2,e3,e4)\displaystyle\operatorname{span}(e_{1},e_{2},e_{3},e_{4})
𝔣2\displaystyle\mathfrak{f}_{2} =\displaystyle= span(e1,e2)\displaystyle\operatorname{span}(e_{1},e_{2})
𝔣3\displaystyle\mathfrak{f}_{3} =\displaystyle= 0,\displaystyle 0,

let PP be the subgroup of GG that stabilizes 𝔣\mathfrak{f}, and let MM be the Levi subgroup that fixes the subspaces span(e1,e2)\operatorname{span}(e_{1},e_{2}), span(e3,e4)\operatorname{span}(e_{3},e_{4}), and span(e5,e6)\operatorname{span}(e_{5},e_{6}).

An element of X(ZG0)X_{*}(Z_{G^{0}}) is given by an action of 𝔾m\mathbb{G}_{m} on VV by some integral power tnt^{n} (where tt is the coordinate on 𝔾m\mathbb{G}_{m}); tensoring with \mathbb{Q}, we can write elements of 𝔞G\mathfrak{a}_{G} formally as scalar matrices tat^{a}, with aa\in\mathbb{Q}. Similarly, an element of 𝔞M\mathfrak{a}_{M} can be written as (ta0,ta1,ta2)(t^{a_{0}},t^{a_{1}},t^{a_{2}}), where 𝔾m\mathbb{G}_{m} is understood to act on 𝔣k/𝔣k+1\mathfrak{f}_{k}/\mathfrak{f}_{k+1} as takt^{a_{k}}.

The modular character γP\gamma_{P} takes (ta0,ta1,ta2)(t^{a_{0}},t^{a_{1}},t^{a_{2}}) to t4(a0a2)t^{4(a_{0}-a_{2})}.

We will consider filtrations FF such that F0=VF_{0}=V, dimF1=3\dim F_{1}=3, and F2=0F_{2}=0. For such filtration FF, with corresponding cocharacter μ\mu, we have wG(μ)=1/2w_{G}(\mu)=1/2.

Now let

ak=dim(F1𝔣k)dim(F1𝔣k+1),a_{k}=\dim(F_{1}\cap\mathfrak{f}_{k})-\dim(F_{1}\cap\mathfrak{f}_{k+1}),

for k=0,1,2k=0,1,2. We have a0+a1+a2=3a_{0}+a_{1}+a_{2}=3, and

wM(μ)=(ta0/2,ta1/2,ta2/2).w_{M}(\mu)=(t^{a_{0}/2},t^{a_{1}/2},t^{a_{2}/2}).

Thus μ\mu is balanced if and only if (a0,a1,a2)=(1,1,1)(a_{0},a_{1},a_{2})=(1,1,1). It is weakly balanced if and only if a0=a2a_{0}=a_{2}; it is semibalanced if and only if a0a2a_{0}\leq a_{2}.

Note that for FF generic in the Grassmannian, we have (a0,a1,a2)=(2,1,0)(a_{0},a_{1},a_{2})=(2,1,0).

In general, the condition that FF be semibalanced is a strong condition on FF, only satisfied for FF in a high-codimension subset of the flag variety. Lemma 8.12 will give a precise bound on the codimension, in the context of interest to us.

The notion of a semibalanced (rather than balanced) filtration will be important in the proof of Lemma 8.16. Our method requires us to work with a period domain on which the monodromy group acts transitively. We cannot guarantee that the monodromy group is all of 𝐆0\mathbf{G}^{0}; we know only that it is a cc-balanced subgroup; thus (after changing coefficients to arrange that 𝐆0=𝐇d\mathbf{G}^{0}=\mathbf{H}^{d}) we will work with a period domain of the form

𝐇I/(Q0I𝐇I),\mathbf{H}^{I}/(Q_{0}^{I}\cap\mathbf{H}^{I}),

for some index set I{1,,d}I\subseteq\{1,\ldots,d\} of cardinality cdc\leq d.

Lemma 8.7 tells us that the filtered ϕ\phi-modules coming from global Galois representations are balanced; this amounts to a condition on the Hodge numbers averaged over the dd different indices. In the proof of Lemma 8.12, we will pass to a subset II of cardinality cc, on which the filtration is semibalanced.

Lemma 8.7.

(Filtered ϕ\phi-modules coming from global representations are balanced.)

Let GG be an algebraic group over p\mathbb{Q}_{p} whose identity component is reductive, and fix an embedding GGLNG\hookrightarrow GL_{N}. Let

ρ:GalG(p)GLN(p)\rho\colon\operatorname{Gal}_{\mathbb{Q}}\rightarrow G(\mathbb{Q}_{p})\subseteq GL_{N}(\mathbb{Q}_{p})

be a representation satisfying the hypotheses of Lemma 8.1. Suppose ρ\rho has image contained in some parabolic subgroup P(p)G(p)P(\mathbb{Q}_{p})\subseteq G(\mathbb{Q}_{p}), and let MM be a Levi subgroup associated to PP.

Let (V,ϕ,F)(V,\phi,F) be the filtered ϕ\phi-module over p\mathbb{Q}_{p} that is associated to the local representation ρ|p\rho|_{\mathbb{Q}_{p}}. Then FF is a GG-filtration on VV, and the associated graded FM=GrM(F)F_{M}=\operatorname{Gr}_{M}(F) is a balanced filtration on MM.

Proof.

(Compare [47, Prop. 10.6(b), §11.4, §11.6].)

That FF is a GG-filtration on VV is a consequence of the Tannakian formalism (see Lemma 5.43).

To see that FMF_{M} is balanced, let μ:𝔾mP\mu\colon\mathbb{G}_{m}\rightarrow P be a cocharacter defining FF. It is enough to show that every character χ:P𝔾m\chi\colon P\rightarrow\mathbb{G}_{m} annihilating Z(G)Z(G) also kills μ\mu.

For every such character χ\chi, we will show that χρ\chi\circ\rho is pure of weight zero: the Frobenius eigenvalues at unramified primes are rational Weil numbers of weight zero.

Indeed, let ρ=ρχ\rho^{\prime}=\rho\oplus\chi as a representation of PP. Let Frobss\operatorname{Frob}_{\ell}^{ss} be the semisimplification of a Frobenius element at \ell acting on ρ\rho^{\prime}. Then Frobss\operatorname{Frob}_{\ell}^{ss} lies in PP. Fix an eigenbasis of Frobss\operatorname{Frob}_{\ell}^{ss} – with the last eigenvector lying in χ\chi – and let TT be the subgroup of PP consisting of elements which have each element of this eigenbasis as eigenvectors. This is an algebraic subgroup of the torus with coordinates λ1,,λN+1\lambda_{1},\dots,\lambda_{N+1}, hence is defined by relations iλiei=1\prod_{i}\lambda_{i}^{e_{i}}=1 for eie_{i}\in\mathbb{Z}; we have chosen indices so that λ1,,λN\lambda_{1},\ldots,\lambda_{N} are the eigenvalues on ρ\rho, and λN+1\lambda_{N+1} is the eigenvalue on χ\chi.

Since Z(G)Z(G) acts trivially on χ\chi, every element of TT whose eigenvalues λ1,,λN\lambda_{1},\dots,\lambda_{N} are all equal acts by scalars on ρ\rho, hence lies in Z(G)Z(G), and thus acts trivially on χ\chi. It follows by restricting all relations to the subtorus where λ1==λN\lambda_{1}=\dots=\lambda_{N} that there must be some relation with eN+10e_{N+1}\neq 0 and i=1Nei=0\sum_{i=1}^{N}e_{i}=0. Applying this relation to the eigenvalues of Frobss\operatorname{Frob}_{\ell}^{ss}, we see that

|χ(Frobss)|eN+1=|χ(Frobss)eN|=|i=1Nλi(Frobss)ei|=i=1N|λi(Frobss)|ei|\chi(\operatorname{Frob}_{\ell}^{ss})|^{e_{N+1}}=|\chi(\operatorname{Frob}_{\ell}^{ss})^{e_{N}}|=|\prod_{i=1}^{N}\lambda_{i}(\operatorname{Frob}_{\ell}^{ss})^{-e_{i}}|=\prod_{i=1}^{N}|\lambda_{i}(\operatorname{Frob}_{\ell}^{ss})|^{-e_{i}}
=i=1N(pw/2)ei=p(w/2)i=1Nei=p0=1=\prod_{i=1}^{N}(p^{w/2})^{-e_{i}}=p^{-(w/2)\sum_{i=1}^{N}e_{i}}=p^{0}=1

and so χ\chi is pure of weight 0.

By Lemma 8.1, the weight of the corresponding filtered ϕ\phi-module is also zero, which is what we needed to prove. ∎

Definition 8.8.

Let E0E_{0}, EE, 𝐆0\mathbf{G}^{0} and 𝐆\mathbf{G} be as in Section 5.6, and fix σAutE0E\sigma\in\operatorname{Aut}_{E_{0}}E. A 𝐆\mathbf{G}-bifiltered ϕ\phi-module is a quadruple (V,ϕ,F,𝔣)(V,\phi,F,\mathfrak{f}), with VV as in Section 5.6, FF and 𝔣\mathfrak{f} two 𝐆\mathbf{G}-filtrations on VV, and ϕ𝐆\phi\in\mathbf{G} a σ\sigma-semilinear endomorphism of VV respecting 𝔣\mathfrak{f}. In this setting, let PP and QQ denote the parabolic subgroups of 𝐆\mathbf{G} corresponding to 𝔣\mathfrak{f} and FF, respectively; to say that ϕ\phi respects 𝔣\mathfrak{f} means that ϕP\phi\in P.

A graded 𝐆\mathbf{G}-bifiltered ϕ\phi-module 101010In the notation (V,ϕ,𝔣,FM)(V,\phi,\mathfrak{f},F_{M}), the filtration FMF_{M} comes after the semisimplification filtration 𝔣\mathfrak{f} to indicate that 𝔣\mathfrak{f} is logically prior to FMF_{M}. is a quadruple (V,ϕ,𝔣,FM)(V,\phi,\mathfrak{f},F_{M}), where VV is as in Notation 5.6, 𝔣\mathfrak{f} is a 𝐆\mathbf{G}-filtration on VV with associated Levi subgroup MM, and FMF_{M} is a filtration on MM.

We say that two graded 𝐆\mathbf{G}-bifiltered ϕ\phi-modules (V1,ϕ1,𝔣1,FM,1)(V_{1},\phi_{1},\mathfrak{f}_{1},F_{M,1}) and (V2,ϕ2,𝔣2,FM,2)(V_{2},\phi_{2},\mathfrak{f}_{2},F_{M,2}) are equivalent if they agree up to 𝐆\mathbf{G}-conjugacy. More precisely, let P1P_{1} and P2P_{2} be the parabolic subgroups attached to 𝔣1\mathfrak{f}_{1} and 𝔣2\mathfrak{f}_{2}, and let M1M_{1} and M2M_{2} be Levi subgroups associated to P1P_{1} and P2P_{2}, respectively. Then (V1,ϕ1,𝔣1,FM,1)(V_{1},\phi_{1},\mathfrak{f}_{1},F_{M,1}) and (V2,ϕ2,𝔣2,FM,2)(V_{2},\phi_{2},\mathfrak{f}_{2},F_{M,2}) are equivalent if there exists g𝐆g\in\mathbf{G} satisfying the following conditions.

  • gP1g1=P2gP_{1}g^{-1}=P_{2}.

  • The filtrations gFM,1g1gF_{M,1}g^{-1} and FM,2F_{M,2} on M2M_{2} agree.

  • The two elements gϕ1g1g\phi_{1}g^{-1} and ϕ2\phi_{2} of P2P_{2} project to the same element of M2M_{2}.

There is an obvious functor

(V,ϕ,F,𝔣)(V,ϕ,𝔣,GrMF)(V,\phi,F,\mathfrak{f})\mapsto(V,\phi,\mathfrak{f},\operatorname{Gr}_{M}F)

from 𝐆\mathbf{G}-bifiltered ϕ\phi-modules to graded 𝐆\mathbf{G}-bifiltered ϕ\phi-modules (with FGrMFF\mapsto\operatorname{Gr}_{M}F given by Definition 8.3). We say that two 𝐆\mathbf{G}-bifiltered ϕ\phi-modules are semisimply equivalent if the corresponding graded 𝐆\mathbf{G}-bifiltered ϕ\phi-modules are equivalent.

We say that a 𝐆\mathbf{G}-filtered ϕ\phi-module (V,ϕ,F)(V,\phi,F) is of the semisimplicity type (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}) if there exists a 𝐆\mathbf{G}-filtration 𝔣\mathfrak{f} on VV such that (V,ϕ,F,𝔣)(V,\phi,F,\mathfrak{f}) and (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}) are semisimply equivalent.

Remark 8.9.

To illustrate ideas, consider the case where E=E0E=E_{0} and 𝐆=𝐆0=GLn\mathbf{G}=\mathbf{G}^{0}=GL_{n}.

A 𝐆\mathbf{G}-bifiltered ϕ\phi-module comes (by pp-adic Hodge theory) from a filtered Galois representation (i.e. a Galois representation on a vector space VV, and a Galois-stable filtration 𝔣iV\mathfrak{f}_{i}V on VV).

A graded 𝐆\mathbf{G}-bifiltered ϕ\phi-module comes from the associated graded to a filtered Galois representation (i.e. the Galois representation on i(𝔣iV/𝔣i+1V)\bigoplus_{i}(\mathfrak{f}_{i}V/\mathfrak{f}_{i+1}V)).

Two graded 𝐆\mathbf{G}-bifiltered ϕ\phi-modules are equivalent if the corresponding Galois representations are isomorphic; two 𝐆\mathbf{G}-filtered ϕ\phi-modules are of the same semisimplicity type if the corresponding representations i(𝔣iV/𝔣i+1V)\bigoplus_{i}(\mathfrak{f}_{i}V/\mathfrak{f}_{i+1}V) are isomorphic.

Lemma 8.10.

(Compare Lemma 7.1.)

Let pp be a prime. A representation  111111The restriction to \mathbb{Q}, instead of an arbitrary number field KK, is for two reasons. First, in the general setting, a filtered ϕ\phi-module would be semilinear over KvK_{v}, and we have not defined filtered ϕ\phi-modules with GG-structure in the semilinear setting; this restriction is inessential. Second, we will need to apply Lemma 8.1, and for that we need KK to have no CM subfield.

ρ0:G𝐆et\rho_{0}\colon G_{\mathbb{Q}}\rightarrow\mathbf{G}_{et}

of the global Galois group GG_{\mathbb{Q}}, crystalline at pp, such that the composition G𝐆etAutp𝖤etG_{\mathbb{Q}}\rightarrow\mathbf{G}_{et}\rightarrow\operatorname{Aut}_{\mathbb{Q}_{p}}\mathsf{E}_{et} agrees with the usual action of Galois on 𝖤et\mathsf{E}_{et}, gives rise by pp-adic Hodge theory to an admissible filtered ϕ\phi-module (V0,ϕ0,F0)(V_{0},\phi_{0},F_{0}) with 𝐆dR\mathbf{G}_{dR}-structure. Suppose ρ0\rho_{0} is semisimple. Suppose another crystalline global representation ρ:G𝐆et\rho\colon G_{\mathbb{Q}}\rightarrow\mathbf{G}_{et} with the same composition with 𝐆etAutp𝖤et\mathbf{G}_{et}\rightarrow\operatorname{Aut}_{\mathbb{Q}_{p}}\mathsf{E}_{et} has semisimplification ρ0\rho_{0}, and call the corresponding filtered ϕ\phi-module (V,ϕ,F)(V,\phi,F). Then there exist 𝐆dR\mathbf{G}_{dR}-filtrations 𝔣0\mathfrak{f}_{0} on V0V_{0} and 𝔣\mathfrak{f} on VV such that (V,ϕ,F,𝔣)(V,\phi,F,\mathfrak{f}) and (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}) are semisimply equivalent.

Furthermore, we can take 𝔣0\mathfrak{f}_{0} one of a list of finitely many candidates, depending only on ρ0\rho_{0}, and the filtration F0F_{0} is balanced with respect to 𝔣0\mathfrak{f}_{0}.

Proof.

By restriction to GpG_{\mathbb{Q}_{p}} and Lemma 5.43, a crystalline representation G𝐆etG_{\mathbb{Q}}\rightarrow\mathbf{G}_{et} gives rise to a filtered ϕ\phi-module with 𝐆dR\mathbf{G}_{dR}-structure (V,ϕ,F)(V,\phi,F), with ϕ\phi semilinear over σ\sigma.

To say that ρ0\rho_{0} is the semisimplification of ρ\rho means (see Definition 5.36 and Lemma 5.37) that there exist a parabolic subgroup Pet𝐆etP_{et}\subseteq\mathbf{G}_{et} with associated Levi LetL_{et}, and g𝐆etg\in\mathbf{G}_{et} such that ρ0\rho_{0} takes values in LetL_{et}, and g1ρgg^{-1}\rho g takes values in PetP_{et}, and the composition of g1ρgg^{-1}\rho g with the quotient map PetLetP_{et}\rightarrow L_{et} is exactly ρ0\rho_{0}.

Lemma 5.52 shows that, given ρ0\rho_{0}, we can take LetL_{et} to be one of finitely many possible subgroups. For each such LetL_{et}, there are finitely many parabolic subgroups PetP_{et} with LetPetL_{et}\subseteq P_{et} as a Levi subgroup.

Since the kernel of PetLetP_{et}\to L_{et} is a unipotent group, the natural map H1(p,Pet)H1(p,Let)H^{1}(\mathbb{Q}_{p},P_{et})\to H^{1}(\mathbb{Q}_{p},L_{et}) is a bijection, so an inner form of PetP_{et} is determined by the corresponding inner form of LetL_{et}.

Now fix ρ0\rho_{0}, PetP_{et}, and LetL_{et}. Let PdRP_{dR} and LdRL_{dR} be the inner twists of PetP_{et} and LetL_{et} arising by Lemma 5.43 from ρ0\rho_{0}. Note that PdRP_{dR} is a parabolic of 𝐆dR\mathbf{G}_{dR} and LdRL_{dR} is the associated Levi by Lemma 5.43(2) and the fact that the property of being an inclusion of a parabolic or an inclusion of a Levi is preserved by inner twists.

Lemma 5.43 produces inner twists of PetP_{et} and LetL_{et} associated to g1ρgg^{-1}\rho g. But since the projection of g1ρgg^{-1}\rho g under PetLetP_{et}\to L_{et} agrees with ρ0\rho_{0}, the inner twists of LetL_{et} is again LdRL_{dR}, and hence the inner twist of PetP_{et} is again PdRP_{dR}.

By Lemma 5.43, we find that ρ0\rho_{0} (resp. g1ρgg^{-1}\rho g) must give rise to filtered ϕ\phi-modules with PdRP_{dR}-structure, such that the corresponding filtered ϕ\phi-modules with LdRL_{dR}-structure (obtained by functoriality via the quotient map PLP\rightarrow L) are isomorphic.

But a filtered ϕ\phi-module with PdRP_{dR}-structure is precisely the same as a filtered ϕ\phi-module with 𝐆dR\mathbf{G}_{dR}-structure, equipped with a filtration 𝔣0\mathfrak{f}_{0} whose associated parabolic is PdRP_{dR}. The filtered ϕ\phi-modules with 𝐆dR\mathbf{G}_{dR}-structure arising from g1ρgg^{-1}\rho g by this process is 𝐆dR\mathbf{G}_{dR}-conjugate to the filtered ϕ\phi-module with 𝐆dR\mathbf{G}_{dR}-structure arising from g1ρgg^{-1}\rho g directly by Lemma 5.43(2) and hence 𝐆dR\mathbf{G}_{dR}-conjugate to the filtered ϕ\phi-module with 𝐆dR\mathbf{G}_{dR}-structure arising from ρ\rho by Lemma 5.43(1) since they correspond to isomorphic Galois representations into 𝐆et\mathbf{G}_{et}.

So take 𝔣0\mathfrak{f}_{0} a filtration on V0V_{0} attached to PdRP_{dR}, and let 𝔣\mathfrak{f} be its image under this 𝐆dR\mathbf{G}_{dR}-conjugation. Then (V,ϕ,F,𝔣)(V,\phi,F,\mathfrak{f}) and (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}) are semisimply equivalent.

Finally, F0F_{0} is balanced with respect to 𝔣0\mathfrak{f}_{0} by Lemma 8.7. ∎

Remark 8.11.

Recall (Definition 5.41 and Lemma 5.42) that there is an equivalence between 𝐆dR\mathbf{G}_{dR}-filtrations and 𝐆dR0\mathbf{G}^{0}_{dR}-filtrations; we will use this without comment, and work with 𝐆dR0\mathbf{G}^{0}_{dR}-filtrations on VV.

Lemma 8.12.

Take notation as in Section 5.6, let cc be a positive integer, and suppose E=E0cE=E_{0}^{c}.

Let PP be a parabolic subgroup of 𝐆0=𝐇c\mathbf{G}^{0}=\mathbf{H}^{c}, corresponding to a 𝐆0\mathbf{G}^{0}-filtration 𝔣\mathfrak{f}, and let MM be a Levi subgroup associated to PP.

Let Q0Q_{0} be another parabolic subgroup of 𝐆0\mathbf{G}^{0}, corresponding to some filtration F0F_{0}, so that 𝐆0/Q0\mathbf{G}^{0}/Q_{0} parametrizes filtrations FF that are 𝐆0\mathbf{G}^{0}-conjugate to F0F_{0}. Suppose F0F_{0} is uniform in the sense of Definition 5.46, and let ha=h𝐇ah^{a}=h^{a}_{\mathbf{H}} be the adjoint Hodge numbers on 𝐇\mathbf{H}. Let tt be the dimension of a maximal torus in 𝐇\mathbf{H}. Suppose ee is a positive integer such that:

  • (First numerical condition.)

    a>0haec\sum_{a>0}h^{a}\geq\frac{e}{c}

    and

  • (Second numerical condition.)

    a>0aha>T(ec)+T(12(h0t)+ec).\sum_{a>0}ah^{a}>T\left(\frac{e}{c}\right)+T\left(\frac{1}{2}(h^{0}-t)+\frac{e}{c}\right).

Then for any semibalanced filtration F0,MF_{0,M} on MM, the set of filtrations FF on 𝐆0\mathbf{G}^{0} that are 𝐆0\mathbf{G}^{0}-conjugate to F0F_{0} and satisfy GrMF=F0,M\operatorname{Gr}_{M}F=F_{0,M} is of codimension at least ee in 𝐆0/Q0\mathbf{G}^{0}/Q_{0}.

Proof.

This is essentially [47, Prop. 11.3], applied to 𝐆0=𝐇c\mathbf{G}^{0}=\mathbf{H}^{c}. Note that the Hodge numbers hah^{a} and the function TT used in [47] are h𝐇ch_{\mathbf{H}^{c}} and T𝐇cT_{\mathbf{H}^{c}}. By Lemma 5.48, they are related to h𝐇h_{\mathbf{H}} and T𝐇T_{\mathbf{H}} by

h𝐇ca=ch𝐇ah_{\mathbf{H}^{c}}^{a}=ch_{\mathbf{H}}^{a}

and

T𝐇c(cx)=cT𝐇(x).T_{\mathbf{H}^{c}}(cx)=cT_{\mathbf{H}}(x).

Similarly, the dimension of a maximal torus in 𝐇c\mathbf{H}^{c} is ctct.

Two small modifications need to be made to the proof in [47]. First, we have replaced 12h0\frac{1}{2}h^{0} with 12(h0t)\frac{1}{2}(h^{0}-t). To get this stronger bound, replace the final inequality of [47, Equation 11.15] with

dim(Q/B)+e12(a0t)+e.\dim(Q/B)+e\leq\frac{1}{2}(a_{0}-t)+e.

(Here Q=Q0Q=Q_{0}, a0a_{0} the dimension of an associated Levi subgroup, and BB a Borel. There is no new idea here; this bound is stronger only because the bound in [47] was not sharp.)

Second, our hypothesis is weaker: in the above-referenced proposition, F0,MF_{0,M} is assumed to be balanced, while here it is only assumed to be semibalanced. This is not a problem, since the inequalities work in our favor. Recall from [47, proof of Proposition 11.3] that TT is a maximal torus contained in PP, Σ\Sigma is the set of roots of TT on 𝐆0\mathbf{G}^{0}, ΣPΣ\Sigma_{P}\subseteq\Sigma is the set of roots of TT on PP, and μ\mu is a cocharacter defining the parabolic subgroup Q0Q_{0}. In our context, [47, Equation 11.14] is replaced with the inequality

γΣΣPwμ,γ0,\sum_{\gamma\in\Sigma-\Sigma_{P}}\langle w\mu,\gamma\rangle\leq 0,

and [47, Equation 11.16] becomes

βXμ,w1βXμ,w1β.\sum_{\beta\in X}\langle\mu,w^{-1}\beta\rangle\leq\sum_{X^{\prime}}-\langle\mu,w^{-1}\beta\rangle.

The rest of the proof goes through as in [47]. ∎

Our next goal (Lemma 8.15) is a slight generalization of [47, Equation 11.18]: the result in [47] only holds with ϕ\phi contained in a connected reductive group (e.g. 𝐆dR0\mathbf{G}^{0}_{dR}), but here ϕ\phi is semilinear, so it is contained in 𝐆dR\mathbf{G}_{dR}, but not in 𝐆dR0\mathbf{G}^{0}_{dR}.

Lemma 8.13.

Let UU be a unipotent algebraic group over a field of characteristic zero, and ψ\psi an automorphism of UU, such that ψr=idU\psi^{r}=\operatorname{id}_{U}. Suppose uUu\in U is such that

uψ(u)ψ2(u)ψr1(u)=1.u\psi(u)\psi^{2}(u)\cdots\psi^{r-1}(u)=1.

Then there exists vUv\in U such that

u=v1ψ(v).u=v^{-1}\psi(v).
Proof.

By induction on dimU\dim U; reduce to the case where U=𝔾akU=\mathbb{G}_{a}^{k} is abelian. ∎

Lemma 8.14.

Let GG be an algebraic group whose identity component G0G^{0} is reductive. Let PP a parabolic subgroup of G0G^{0} and MM an associated Levi subgroup. Suppose ϕG\phi\in G is semisimple and normalizes PP.

Then ϕ\phi is PP-conjugate to an element that normalizes MM.

Proof.

Suppose ϕrG0\phi^{r}\in G^{0}. Since ϕr\phi^{r} normalizes PP, ϕrP\phi^{r}\in P. Since ϕr\phi^{r} is semisimple, we may conjugate by an element of PP to assure that ϕrM\phi^{r}\in M.

Since ϕ\phi normalizes PP, and all Levi subgroups in PP are PP-conjugate, we can write ϕMϕ1=u1Mu\phi M\phi^{-1}=u^{-1}Mu for some uPu\in P. Multiplying uu by an element of MM from the left, we may assume that uu is unipotent. Since (uϕ)M(uϕ)1=M(u\phi)M(u\phi)^{-1}=M, it will suffice to show that uϕu\phi is PP-conjugate to ϕ\phi.

Now (uϕ)r(u\phi)^{r} is an element of UϕrU\phi^{r} that normalizes MM, so

(uϕ)r=ϕr.(u\phi)^{r}=\phi^{r}.

We can apply Lemma 8.13, with

ψ(x)=ϕxϕ1,\psi(x)=\phi x\phi^{-1},

to conclude that

u=v1ψ(v)=v1ϕvϕ1u=v^{-1}\psi(v)=v^{-1}\phi v\phi^{-1}

for some vPv\in P. Then v1ϕv=uϕv^{-1}\phi v=u\phi normalizes MM, and we are done. ∎

Lemma 8.15.

(Compare [47, Equation 11.18].)

Assume we are in the setting of Section 5.6. Fix a 𝐆\mathbf{G}-bifiltered ϕ\phi-module (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}), and another ϕ\phi-module (V,ϕ)(V,\phi) with 𝐆\mathbf{G}-structure; suppose both ϕ\phi and ϕ0\phi_{0} are semilinear over some σAutE0E\sigma\in\operatorname{Aut}_{E_{0}}E.

Consider all pairs (𝔣,FM)(\mathfrak{f},F_{M}), where 𝔣\mathfrak{f} is a filtration on 𝐆\mathbf{G}, MM is an associated Levi subgroup, and (V,ϕ,𝔣,FM)(V,\phi,\mathfrak{f},F_{M}) is equivalent to (V0,ϕ0,𝔣0,(F0)M0)(V_{0},\phi_{0},\mathfrak{f}_{0},(F_{0})_{M_{0}}), in the sense of Definition 8.8. The set of such pairs has dimension at most dimZ𝐆dR0(ϕss)\dim Z_{\mathbf{G}^{0}_{dR}}(\phi^{ss}).

(Recall that any element ϕ\phi of an algebraic group has a unique Jordan decomposition ϕ=ϕssϕunip\phi=\phi^{ss}\phi^{unip}, where ϕss\phi^{ss} is semisimple, ϕunip\phi^{unip} is unipotent, and ϕss\phi^{ss} and ϕunip\phi^{unip} commute.)

Proof.

This is a question about the dimension of a variety, so we can pass to a finite extension of KK. In particular, we may assume E=KdE=K^{d} and σ\sigma acts by permutation on the factors of E=KdE=K^{d}. We index the factors K1K_{1} through KdK_{d}, and we regard σ\sigma as a permutation of {1,,d}\{1,\ldots,d\}. Write Vi=VEKiV_{i}=V\otimes_{E}K_{i} for the ii-th factor of VV, and write 𝐇i\mathbf{H}_{i} for the corresponding factor of 𝐆0=𝐇d\mathbf{G}^{0}=\mathbf{H}^{d} (it satisfies 𝐇i𝐇\mathbf{H}_{i}\cong\mathbf{H}). The semilinearity condition means that ϕ\phi decomposes as a sum of maps

ϕi:ViVσi.\phi_{i}\colon V_{i}\rightarrow V_{\sigma i}.

It follows from functoriality of the Jordan decomposition that, if (V,ϕ,𝔣,FM)(V,\phi,\mathfrak{f},F_{M}) is equivalent to (V0,ϕ0,𝔣0,(F0)M0)(V_{0},\phi_{0},\mathfrak{f}_{0},(F_{0})_{M_{0}}), then (V,ϕss,𝔣,FM)(V,\phi^{ss},\mathfrak{f},F_{M}) is equivalent to (V0,ϕ0ss,𝔣0,(F0)M0)(V_{0},\phi_{0}^{ss},\mathfrak{f}_{0},(F_{0})_{M_{0}}). Thus, as in [47, §11.6, third paragraph], we can assume ϕ\phi is semisimple. Otherwise, replacing ϕ\phi with ϕss\phi^{ss} will only increase the dimension of the set of (𝔣,FM)(\mathfrak{f},F_{M}) for which (V,ϕ,𝔣,FM)(V,\phi,\mathfrak{f},F_{M}) is equivalent to (V0,ϕ0,𝔣0,(F0)M0)(V_{0},\phi_{0},\mathfrak{f}_{0},(F_{0})_{M_{0}}).

A filtration 𝔣\mathfrak{f} on 𝐆0\mathbf{G}^{0} (which is the same as a filtration on 𝐆\mathbf{G}) decomposes as the product of filtrations 𝔣i\mathfrak{f}_{i} on 𝐇i\mathbf{H}_{i}. If 𝔣\mathfrak{f} is ϕ\phi-stable, then 𝔣i\mathfrak{f}_{i} determines 𝔣σi\mathfrak{f}_{\sigma i}. So to specify 𝔣\mathfrak{f} it is enough to specify 𝔣i\mathfrak{f}_{i}, for a single ii in each σ\sigma-orbit.

Since everything in sight splits as a direct product over σ\sigma-orbits, we may as well restrict attention to a single σ\sigma-orbit; call it {i1,,ir}\{i_{1},\ldots,i_{r}\}. A ϕ\phi-stable filtration 𝔣\mathfrak{f} on 𝐆0\mathbf{G}^{0} is uniquely determined by a ϕr\phi^{r}-stable filtration 𝔣1\mathfrak{f}_{1} on 𝐇1\mathbf{H}_{1}. As in [47, §11.6, fourth and fifth paragraphs], the dimension of the set of such filtrations is exactly

dimZ𝐆0(ϕ)dimZP(ϕ),\dim Z_{\mathbf{G}^{0}}(\phi)-\dim Z_{P}(\phi),

where Z𝐆0(ϕ)=Z𝐆(ϕ)𝐆0Z_{\mathbf{G}^{0}}(\phi)=Z_{\mathbf{G}}(\phi)\cap\mathbf{G}^{0}, and ZP(ϕ)Z_{P}(\phi) is defined similarly.

To conclude, we need to show that, given 𝔣\mathfrak{f} (and an associated Levi subgroup MM), the set of filtrations FMF_{M} such that (V,ϕ,𝔣,FM)(V,\phi,\mathfrak{f},F_{M}) is equivalent to (V0,ϕ0,𝔣0,(F0)M0)(V_{0},\phi_{0},\mathfrak{f}_{0},(F_{0})_{M_{0}}) has dimension at most dimZP(ϕ)\dim Z_{P}(\phi). The proof of this is the same as in [47, end of §11.6], using Lemma 8.14 in place of [47, Equation 2.1]. ∎

Lemma 8.16.

(Compare Lemma 7.2.)

Assume we are in the setting of Section 5.6 and Section 5.7. Fix a 𝐆dR0\mathbf{G}^{0}_{dR}-bifiltered ϕ\phi-module (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}), and another ϕ\phi-module (V,ϕ)(V,\phi) with 𝐆dR\mathbf{G}_{dR}-structure; suppose both ϕ\phi and ϕ0\phi_{0} are semilinear over some σAutE0E\sigma\in\operatorname{Aut}_{E_{0}}E, and F0F_{0} is balanced with respect to 𝔣0\mathfrak{f}_{0}.

Let 𝐆mon\mathbf{G}_{mon} be a subgroup of 𝐆dR0\mathbf{G}^{0}_{dR}, strongly cc-balanced with respect to σ\sigma for some positive integer cc.

Suppose F0F_{0} is uniform in the sense of Definition 5.46, and let ha=hsimpah^{a}=h^{a}_{simp} be the adjoint Hodge numbers on 𝐇\mathbf{H}. Let tt be the dimension of a maximal torus in 𝐇\mathbf{H}. Suppose ee is a positive integer satisfying the following numerical conditions.

  • (First numerical condition.)

    a>0ha1c(e+dim𝐇)\sum_{a>0}h^{a}\geq\frac{1}{c}(e+\dim\mathbf{H})

    and

  • (Second numerical condition.)

    a>0aha>T(1c(e+dim𝐇))+T(12(h0t)+1c(e+dim𝐇)).\sum_{a>0}ah_{a}>T\left(\frac{1}{c}(e+\dim\mathbf{H})\right)+T\left(\frac{1}{2}(h^{0}-t)+\frac{1}{c}(e+\dim\mathbf{H})\right).

Let =𝐆mon/(Q0𝐆mon)\mathcal{H}=\mathbf{G}_{mon}/(Q_{0}\cap\mathbf{G}_{mon}) be the flag variety parametrizing filtrations on 𝐆dR0\mathbf{G}^{0}_{dR} that are conjugate to F0F_{0} under the conjugation of 𝐆mon\mathbf{G}_{mon}. Then the filtrations FF such that (V,ϕ,F)(V,\phi,F) is of semisimplicity type (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}) are of codimension at least ee in \mathcal{H}.

Proof.

Let PP be the parabolic of 𝐆dR0\mathbf{G}^{0}_{dR}, and MM an associated Levi subgroup, associated to a hypothetical semisimplification filtration 𝔣\mathfrak{f}. There are only finitely many possibilities for the parabolic group PP, up to 𝐆dR0\mathbf{G}^{0}_{dR}-conjugacy, so we may as well as fix a single PP.

The dimension in question can be calculated after base change to an extension of KK, so we can assume that 𝐆dR0=𝐇d\mathbf{G}^{0}_{dR}=\mathbf{H}^{d}. As in the proof of Lemma 8.15, we’ll call the dd factors 𝐇1,,𝐇d\mathbf{H}_{1},\ldots,\mathbf{H}_{d}. The whole setup factors over these dd factors, so we can write PP as the direct sum of parabolics Pi𝐇iP_{i}\subseteq\mathbf{H}_{i}, and so forth. Again, σAutE0E\sigma\in\operatorname{Aut}_{E_{0}}E gives a permutation of the index set {1,,d}\{1,\ldots,d\}, which we also call σ\sigma. Semilinearity over σ\sigma means that the map ϕ\phi permutes the dd factors according to the permutation σ\sigma.

The strategy is as follows. We want to apply a result like Lemma 8.12, but we don’t have full monodromy group 𝐆dR0\mathbf{G}^{0}_{dR}. Instead, we know the group 𝐆mon\mathbf{G}_{mon} is cc-balanced; we’ll project onto cc of the dd factors, so that 𝐆mon\mathbf{G}_{mon} projects onto the full group (𝐇/Z(𝐇))c(\mathbf{H}/Z(\mathbf{H}))^{c}. We need the projection of F0F_{0} to the cc factors to be semibalanced, and we can apply Lemmas 8.12 and 8.15 to the group 𝐇c\mathbf{H}^{c} to finish.

For any subset II of the index set {1,,d}\{1,\ldots,d\}, let 𝐇I=iI𝐇i\mathbf{H}^{I}=\prod_{i\in I}\mathbf{H}_{i}; define PIP^{I}, MIM^{I}, and F0IF_{0}^{I} similarly. Any filtrations FF and 𝔣\mathfrak{f} on 𝐆dR0\mathbf{G}^{0}_{dR} can be written as products of factors FiF_{i} and 𝔣i\mathfrak{f}_{i}, so we can define FIF^{I} and 𝔣I\mathfrak{f}^{I}.

We claim that, for any 𝔣\mathfrak{f} and any FF that is balanced with respect to 𝔣\mathfrak{f}, we can find I{1,,d}I\subseteq\{1,\ldots,d\} satisfying the following properties.

  • II consists of exactly cc elements, from cc distinct index classes for 𝐆mon\mathbf{G}_{mon}.

  • The elements of II belong to a single orbit of σ\sigma on {1,,d}\{1,\ldots,d\}.

  • FIF^{I} is semibalanced with respect to 𝔣I\mathfrak{f}^{I}.

To see this, let μ\mu be a cocharacter defining FF as in Definition 8.4. We can write μ\mu as a product of μi\mu_{i} over i{1,,d}i\in\{1,\ldots,d\}. Similarly, the character γP\gamma_{P} splits over the factors, and we have:

γP(μ)=iγPi(μi).\gamma_{P}(\mu)=\sum_{i}\gamma_{P_{i}}(\mu_{i}).

Hence there exists an orbit JJ of σ\sigma on the dd factors such that FJF^{J} is semibalanced. Since 𝐆mon\mathbf{G}_{mon} is strongly cc-balanced, we can find a subset

IJ{1,,d}I\subset J\subset\{1,\ldots,d\}

of the index set such that #I=c\#I=c, the elements of II belong to cc distinct index classes, and FIF^{I} is semibalanced. Since the elements of II belong to distinct index classes, the projection

𝐆mon(𝐇/Z(𝐇))I\mathbf{G}_{mon}\rightarrow(\mathbf{H}/Z(\mathbf{H}))^{I}

has image a union of connected components of the target, so it is smooth with equidimensional fibers, and the same is true of

𝐆mon/(Q0𝐆mon)𝐇I/(Q0I𝐇I).\mathbf{G}_{mon}/(Q_{0}\cap\mathbf{G}_{mon})\rightarrow\mathbf{H}^{I}/(Q_{0}^{I}\cap\mathbf{H}^{I}).

We want to estimate the codimension in =𝐆mon/(Q0𝐆mon)\mathcal{H}=\mathbf{G}_{mon}/(Q_{0}\cap\mathbf{G}_{mon}) of the set of FF such that (V,ϕ,F,𝔣)(V,\phi,F,\mathfrak{f}) is semisimply equivalent to (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}), for some choice of 𝔣\mathfrak{f}. Consider the projection

(V,ϕ,F,𝔣)(𝔣,FM).(V,\phi,F,\mathfrak{f})\rightarrow(\mathfrak{f},F_{M}).

By Lemma 8.15 applied to 𝐇J\mathbf{H}^{J}, the set of pairs (𝔣J,FMJ)(\mathfrak{f}^{J},F_{M}^{J}) such that (V0J,ϕ0J,(F0)M0J,𝔣0J)(V_{0}^{J},\phi_{0}^{J},(F_{0})_{M_{0}}^{J},\mathfrak{f}_{0}^{J}) is equivalent to (VJ,ϕJ,FMJ,𝔣J)(V^{J},\phi^{J},F_{M}^{J},\mathfrak{f}^{J}) has dimension bounded by dimZGJ((ϕJ)ss)\dim Z_{G^{J}}((\phi^{J})^{ss}); this dimension is at most dim𝐇\dim\mathbf{H} by Lemma 5.33.

Now fix (𝔣J,FMJ)(\mathfrak{f}^{J},F_{M}^{J}). By Lemma 8.12 applied to 𝐇I\mathbf{H}^{I}, the set of FIF^{I} such that

GrMFI=FMI\operatorname{Gr}_{M}F^{I}=F_{M}^{I}

has codimension at least e+dim𝐇e+\dim\mathbf{H} among all 𝐇I\mathbf{H}^{I}-filtrations FIF^{I}. The map FJFIF^{J}\mapsto F^{I} from 𝐇J\mathbf{H}^{J}-filtrations in a given 𝐆mon\mathbf{G}_{mon}-conjugacy class to 𝐇I\mathbf{H}^{I}-filtrations is smooth with equidimensional fibers, so the set of FJF^{J} such that

GrM(FJ)=FMJ\operatorname{Gr}_{M}(F^{J})=F_{M}^{J}

again has codimension at least e+dim𝐇e+\dim\mathbf{H}.

It follows that the set of FF satisfying the desired condition has codimension at least ee. ∎

Theorem 8.17.

(Basic theorem giving non-density of integral points. Compare Theorem 7.3.)

Let XX be a variety over \mathbb{Q}, let SS be a finite set of primes of \mathbb{Z}, and let 𝒳\mathcal{X} be a smooth model of XX over [1/S]\mathbb{Z}[1/S].

Let 𝖤\mathsf{E} be a constant H0H^{0}-algebra on 𝒳\mathcal{X}, and let 𝐇\mathbf{H} be one of GLNGL_{N}, GSpNGSp_{N}, or GONGO_{N}. Let 𝖵\mathsf{V} be a polarized, integral 𝖤\mathsf{E}-module with 𝐇\mathbf{H}-structure, in the sense of Definition 5.27, having integral Frobenius eigenvalues (Def. 5.5). Suppose the Hodge numbers of 𝖵\mathsf{V} are uniform in the sense of Definition 5.46, and let ha=hsimpah^{a}=h^{a}_{simp} be the adjoint Hodge numbers on 𝐇\mathbf{H}. Let tt be the dimension of a maximal torus in 𝐇\mathbf{H}. Let pp be as in Definition 5.2. Let 𝐆dR0\mathbf{G}^{0}_{dR} and 𝐆dR\mathbf{G}_{dR} be as in Section 5.7.

Suppose there is a positive integer cc such that 𝖵\mathsf{V} satisfies the following conditions.

  • (Big monodromy.) If 𝐆mon𝐆dR0\mathbf{G}_{mon}\subseteq\mathbf{G}^{0}_{dR} is the differential Galois group of 𝖵\mathsf{V}, then 𝐆mon𝐆dR0\mathbf{G}_{mon}\subseteq\mathbf{G}^{0}_{dR} is strongly cc-balanced with respect to Frobenius. (The Frobenius is determined from the structure of 𝖤\mathsf{E}; see Section 5.7.)

  • (First numerical condition.)

    a>0ha1c(dimX+dim𝐇)\sum_{a>0}h^{a}\geq\frac{1}{c}(\dim X+\dim\mathbf{H})

    and

  • (Second numerical condition.)

    a>0aha>TG(1c(dimX+dim𝐇))+TG(12(h0t)+1c(dimX+dim𝐇)).\sum_{a>0}ah_{a}>T_{G}\left(\frac{1}{c}(\dim X+\dim\mathbf{H})\right)+T_{G}\left(\frac{1}{2}(h^{0}-t)+\frac{1}{c}(\dim X+\dim\mathbf{H})\right).

Then the image of 𝒳([1/S])\mathcal{X}(\mathbb{Z}[1/S]) is not Zariski dense in XX.

Proof.

This follows from Lemmas 8.10 and 8.16 in the same way that Theorem 7.3 follows from Lemmas 7.1 and 7.2.

For every x𝒳([1/S])x\in\mathcal{X}(\mathbb{Z}[1/S]), the fiber ρx=𝖵et,x\rho_{x}=\mathsf{V}_{et,x} of the pp-adic étale local system is a global Galois representation valued in 𝐆et\mathbf{G}_{et}, having good reduction outside SS and all Frobenius eigenvalues Weil numbers, by hypothesis. By Faltings’s finiteness lemma (in the form of Lemma 5.49), there are only finitely many possible isomorphism classes for the semisimplified representation ρxss\rho^{ss}_{x}. So it is enough to show, for any fixed ρ0\rho_{0}, that the set

𝒳([1/S],ρ0):={xX(𝒪K,S)|ρxssρ0}\mathcal{X}(\mathbb{Z}[1/S],\rho_{0}):=\{x\in X(\mathcal{O}_{K,S})|\rho^{ss}_{x}\cong\rho_{0}\}

is not Zariski dense in XX.

Lemma 8.10 gives a finite list of semisimplicity types such that, for every x𝒳([1/S],ρ0)x\in\mathcal{X}(\mathbb{Z}[1/S],\rho_{0}), the filtered ϕ\phi-module 𝖵cris,x\mathsf{V}_{cris,x} belongs to one of them. Let Ω𝒳([1/S],ρ0)\Omega\subseteq\mathcal{X}(\mathbb{Z}[1/S],\rho_{0}) be a mod-pp residue disk, and fix a semisimplicity type (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}). By Lemma 8.10, F0F_{0} is balanced with respect to 𝔣0\mathfrak{f}_{0}. It is enough to show that the set

X(Ω,(V0,ϕ0,F0,𝔣0))={xΩ|𝖵cris,x is of semisimplicity type (V0,ϕ0,F0,𝔣0)}X(\Omega,(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}))=\{x\in\Omega|\text{$\mathsf{V}_{cris,x}$ is of semisimplicity type $(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0})$}\}

is not Zariski dense in XX.

We are now in the setting of Theorem 6.4. For xΩx\in\Omega, the filtered ϕ\phi-module 𝖵cris,x\mathsf{V}_{cris,x} is of the form (V,ϕ,Fx)(V,\phi,F_{x}), where (V,ϕ)(V,\phi) is independent of xx. (This is a property of FF-isocrystals in general; it reflects the fact that if (V,ϕ,F)(V,\phi,F) is the crystalline cohomology of a scheme over p\mathbb{Z}_{p}, then (V,ϕ)(V,\phi) can be recovered from its reduction modulo pp. See [47, Section 3.3] for further discussion.) The variation of FxF_{x} with xx is classified by a pp-adic period map

Φp:Ω𝐆dR0/Q,\Phi_{p}\colon\Omega\rightarrow\mathbf{G}^{0}_{dR}/Q,

where 𝐆dR0/Q\mathbf{G}^{0}_{dR}/Q is the flag variety classifying 𝐆dR0\mathbf{G}^{0}_{dR}-filtrations on VV.

In fact, the pp-adic period map lands in a single 𝐆mon\mathbf{G}_{mon}-orbit on 𝐆dR0/Q\mathbf{G}^{0}_{dR}/Q; we can write the orbit as 𝐆mon/(Q𝐆mon)\mathbf{G}_{mon}/(Q\cap\mathbf{G}_{mon}). By Lemma 8.16 with e=dimXe=\dim X, there is a Zariski-closed set Z𝐆mon/(Q𝐆mon)Z\subseteq\mathbf{G}_{mon}/(Q\cap\mathbf{G}_{mon}) of codimension at least dimX\dim X such that, if (V,ϕ,Fx)(V,\phi,F_{x}) is of semisimplicity type (V0,ϕ0,F0,𝔣0)(V_{0},\phi_{0},F_{0},\mathfrak{f}_{0}), then xΦp1(Z)x\in\Phi_{p}^{-1}(Z). We conclude by Theorem 6.4. ∎

9. Proof of main theorem

Definition 9.1.

Let AA be an abelian variety over a number field KK, and HAH\subseteq A a hypersurface. We say that HH is primitive if it is not invariant under translation by any xA(¯)x\in A(\overline{\mathbb{Q}}).

Recall the sequence a(i)a(i) defined in Theorem 3.5.

Theorem 9.2.

Let AA be an abelian variety of dimension nn over a number field KK. Let SS be a finite set of primes of 𝒪K\mathcal{O}_{K} including all the places of bad reduction for AA.

Let ϕ\phi be an ample class in the Néron-Severi group of AA, and let d=ϕn/n!d=\phi^{n}/n!.

Suppose that either n4n\geq 4 or n=3n=3 and dd is not (a(i)+a(i+1)a(i+1)/6)\binom{a(i)+a(i+1)}{a(i+1)/6} for any i2i\geq 2.

Then, up to translation there are only finitely many smooth primitive hypersurfaces HAH\subseteq A representing ϕ\phi, defined over KK with good reduction outside SS.

Proof.

Choose a smooth proper model 𝒜\mathcal{A} of AA over 𝒪K,S\mathcal{O}_{K,S}. Working over 𝒪K,S\mathcal{O}_{K,S}, let Hilb\mathrm{Hilb} be the Hilbert scheme of smooth hypersurfaces of class ϕ\phi, and let HunivHilb×𝒜H_{\mathrm{univ}}\subseteq\mathrm{Hilb}\times\mathcal{A} be the universal family over Hilb\mathrm{Hilb}. Let

Hilb(𝒪K,S)primHilb(𝒪K,S)\mathrm{Hilb}(\mathcal{O}_{K,S})^{prim}\subseteq\mathrm{Hilb}(\mathcal{O}_{K,S})

be the set of primitive hypersurfaces in AA (Definition 9.1). (Note that, by definition, a hypersurface defined over KK with good reduction outside SS extends to an 𝒪K,S\mathcal{O}_{K,S}-point of Hilb\mathrm{Hilb}. However, we are interested in the set of hypersurfaces which are primitive over KK, not the set of 𝒪K,S\mathcal{O}_{K,S}-points of the scheme of primitive hypersurfaces, which may be smaller.) The group A(K)=𝒜(𝒪K,S)A(K)=\mathcal{A}(\mathcal{O}_{K,S}) acts on Hilb\mathrm{Hilb} by translation; we need to show that Hilb(𝒪K,S)prim\mathrm{Hilb}(\mathcal{O}_{K,S})^{prim} is contained in the union of finitely many orbits of A(K)A(K).

Theorem 8.17 only works over \mathbb{Q}; we’ll pass from KK to \mathbb{Q} by Weil restriction. (See [12, Theorem 7.6.4] for the notion of Weil restriction relative to a finite étale extension of rings.) Enlarging SS if necessary, we can arrange that 𝒪K,S\mathcal{O}_{K,S} is finite étale over [1/S]\mathbb{Z}[1/S^{\prime}], where SS^{\prime} is a finite set of primes of \mathbb{Z}. Over the Weil restriction

Res[1/S]𝒪K,SHilb{\operatorname{Res}^{\mathcal{O}_{K,S}}_{\mathbb{Z}[1/S^{\prime}]}}\mathrm{Hilb}

we have the family

Huniv,=Huniv×Hilb(Res[1/S]𝒪K,SHilb×[1/S]𝒪K,S),H_{\mathrm{univ},\mathbb{Q}}=H_{\mathrm{univ}}\times_{\mathrm{Hilb}}({\operatorname{Res}^{\mathcal{O}_{K,S}}_{\mathbb{Z}[1/S^{\prime}]}}\mathrm{Hilb}\times_{\mathbb{Z}[1/S^{\prime}]}\mathcal{O}_{K,S}),

and the fibers of this family are the same as the fibers of the original family HunivH_{\mathrm{univ}}. More precisely, the sets Hilb(𝒪K,S)\mathrm{Hilb}(\mathcal{O}_{K,S}) and

Res[1/S]𝒪K,SHilb([1/S]){\operatorname{Res}^{\mathcal{O}_{K,S}}_{\mathbb{Z}[1/S^{\prime}]}}\mathrm{Hilb}(\mathbb{Z}[1/S^{\prime}])

are in canonical bijection. Let

Res[1/S]𝒪K,SHilb([1/S])primRes[1/S]𝒪K,SHilb([1/S]){\operatorname{Res}^{\mathcal{O}_{K,S}}_{\mathbb{Z}[1/S^{\prime}]}}\mathrm{Hilb}(\mathbb{Z}[1/S^{\prime}])^{prim}\subseteq{\operatorname{Res}^{\mathcal{O}_{K,S}}_{\mathbb{Z}[1/S^{\prime}]}}\mathrm{Hilb}(\mathbb{Z}[1/S^{\prime}])

be the subset corresponding to Hilb(𝒪K,S)prim\mathrm{Hilb}(\mathcal{O}_{K,S})^{prim} under this bijection. For any xHilb(𝒪K,S)x\in\mathrm{Hilb}(\mathcal{O}_{K,S}), call Resx\operatorname{Res}x the corresponding point of Res[1/S]𝒪K,SHilb([1/S]){\operatorname{Res}^{\mathcal{O}_{K,S}}_{\mathbb{Z}[1/S^{\prime}]}}\mathrm{Hilb}(\mathbb{Z}[1/S^{\prime}]). Then we have a canonical isomorphism of schemes

(Huniv)x(Huniv,)Resx;(H_{\mathrm{univ}})_{x}\cong(H_{\mathrm{univ},\mathbb{Q}})_{\operatorname{Res}x};

here (Huniv)x(H_{\mathrm{univ}})_{x} is an 𝒪K,S\mathcal{O}_{K,S}-scheme, (Huniv,)Resx(H_{\mathrm{univ},\mathbb{Q}})_{\operatorname{Res}x} is a [1/S]\mathbb{Z}[1/S^{\prime}]-scheme, and the structure maps are related by the fact that the diagram

(7) (Huniv)x\textstyle{(H_{\mathrm{univ}})_{x}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}(Huniv,)Resx\textstyle{(H_{\mathrm{univ},\mathbb{Q}})_{\operatorname{Res}x}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec𝒪K,S\textstyle{\operatorname{Spec}\mathcal{O}_{K,S}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec[1/S]\textstyle{\operatorname{Spec}\mathbb{Z}[1/S^{\prime}]}

commutes.

Let XsingX_{sing} be an irreducible component of the Zariski closure of (ResKHilb)([1/S])prim({\operatorname{Res}^{K}_{\mathbb{Q}}}\mathrm{Hilb})(\mathbb{Z}[1/S])^{prim} in ResKHilb{\operatorname{Res}^{K}_{\mathbb{Q}}}\mathrm{Hilb}, and choose a resolution of singularities XXsingX\rightarrow X_{sing}. Let Y=Huniv,×ResKHilbXY=H_{\mathrm{univ},\mathbb{Q}}\times_{{\operatorname{Res}^{K}_{\mathbb{Q}}}\mathrm{Hilb}}X be the pullback of our family of hypersurfaces to XX. Note that YY is a hypersuface in X×A=XK×KAX\times_{\mathbb{Q}}A=X_{K}\times_{K}A.

We will apply Theorem 8.17 to show that YXK×KAY\subset X_{K}\times_{K}A is the translate of a constant hypersurface H0AH_{0}\in A by a section sA(XK)s\in A(X_{K}). In this case, all the points of (ResKHilb)([1/S])prim({\operatorname{Res}^{K}_{\mathbb{Q}}}\mathrm{Hilb})(\mathbb{Z}[1/S])^{prim} lying in XX would correspond to points in the orbit of H0H_{0}.

The map XXsingX\rightarrow X_{sing} is proper, so it spreads out to a proper map

𝒳Res[1/S]𝒪K,SHilb,\mathcal{X}\rightarrow{\operatorname{Res}^{\mathcal{O}_{K,S}}_{\mathbb{Z}[1/S^{\prime}]}}\mathrm{Hilb},

possibly after further enlarging the finite sets SS and SS^{\prime}. Possibly after further enlargement of SS and SS^{\prime}, the family YY spreads out to a smooth proper family 𝒴𝒳\mathcal{Y}\rightarrow\mathcal{X}, which is a hypersurface in 𝒜×[1/S]𝒳\mathcal{A}\times_{\mathbb{Z}[1/S^{\prime}]}\mathcal{X}. Now, by properness, every SS-integral point of Hilb\mathrm{Hilb} lifts to an SS-integral point of 𝒳\mathcal{X}, so by construction the SS-integral points are Zariski dense in 𝒳\mathcal{X}.

Fix some pSp\notin S^{\prime}. Assume that YXK×KAY\subseteq X_{K}\times_{K}A is not equal to the translate of a constant hypersurface H0AH_{0}\in A by a section sA(XK)s\in A(X_{K}). Let η\eta be the generic point of XKX_{K}. The fibers of YY over points in a dense subset of XX are primitive hypersurfaces, so Yη¯Y_{\overline{\eta}} is not translation-invariant by any nonzero element of AA. Let GG be the Tannakian group of the constant sheaf on Yη¯Y_{\overline{\eta}}, and let GG^{*} be the commutator subgroup of the identity component of GG. By Theorem 3.5, because of our assumptions on nn and dd, we have G=SLN,SpN,G^{*}=SL_{N},Sp_{N}, or SONSO_{N}, acting by its standard representation. Furthermore the SpNSp_{N} case occurs exactly when Yη¯Y_{\overline{\eta}} is equal to a translate of [1]Yη[-1]^{*}Y_{\eta} and nn is even and the SONSO_{N} case occurs exactly when Yη¯Y_{\overline{\eta}} is equal to a translate of [1]Yη[-1]^{*}Y_{\eta} and nn is odd. In particular, GG^{*} is a simple algebraic group acting by an irreducible representation.

We can now apply Corollary 4.10 for a positive integer cc to be chosen shortly. This gives us, for any fixed cc, the existence of an embedding ι:K\iota\colon K\rightarrow\mathbb{C} and a torsion character χ\chi of π1(Aι)\pi_{1}(A_{\iota}), depending on cc, satisfying the big monodromy condition that is needed for Lemma 6.8. (Note a small subtlety here. The Tannakian monodromy groups SOSO, SpSp, or SLSL are calculated over an algebraically closed field ¯p\overline{\mathbb{Q}}_{p}, but to apply them we need them over a smaller field. Thus we use the straightforward fact that any form of Sp,SL,Sp,SL, or SOSO, together with its standard representation, over any field is the split form, or in the SOSO case, the special orthogonal group associated to some nondegenerate quadratic form.) In Section 5.5 and Lemma 5.29 we have constructed a Hodge–Deligne system 𝖵I\mathsf{V}_{I} on 𝒳\mathcal{X} attached to the orbit II containing (ι,χ)(\iota,\chi) and the family 𝒴𝒳𝒪K,S×𝒪K,S𝒜\mathcal{Y}\subseteq\mathcal{X}_{\mathcal{O}_{K,S}}\times_{\mathcal{O}_{K,S}}\mathcal{A}. By Lemma 6.8, the differential Galois group of 𝖵I\mathsf{V}_{I} is a strongly cc-balanced subgroup of 𝖦0\mathsf{G}^{0}. Corollary 4.10 gives vanishing of cohomology outside degree n1n-1, so the Hodge numbers of 𝖵I\mathsf{V}_{I} are given by Lemma 3.13.

We apply Theorem 8.17 to 𝖵=𝖵I\mathsf{V}=\mathsf{V}_{I}. That the eigenvalues of Frobenius on 𝖵\mathsf{V} are integral Weil numbers is a consequence of the Weil conjectures, since 𝖵\mathsf{V} comes from geometry; the polarization and integral structure are given in Lemma 5.29. Lemma 5.30 gives an 𝐇\mathbf{H}-structure over 𝖤I\mathsf{E}_{I}, with 𝐇\mathbf{H} chosen as in Lemma 5.30. (Note that this matches the group GG^{*} by our earlier calculation, because Yη¯Y_{\overline{\eta}} is equal to a translate of [1]Yη[-1]^{*}Y_{\eta} if and only if YY is equal to a translate of [1]Y[-1]^{*}Y.) Lemma 6.8 gives that 𝖦mon\mathsf{G}_{mon} is strongly cc-balanced with respect to Frobenius. The Hodge numbers of 𝖵\mathsf{V} are uniform because we have explicitly computed them, independently of ι\iota, in Lemma 3.13.

The numerical conditions in the hypothesis of Theorem 8.17 will hold for big enough cc. To verify this, since every term except cc is independent of cc (because the Hodge numbers calculated in Lemma 3.13 are independent of χ\chi and thus of cc), it suffices to have

a>0ha>0\sum_{a>0}h^{a}>0
a>0aha>TG(12(h0t)).\sum_{a>0}ah^{a}>T_{G}\left(\frac{1}{2}(h^{0}-t)\right).

By Lemma A.1, we have

a>0ha>12(h0t).\sum_{a>0}h^{a}>\frac{1}{2}(h^{0}-t).

This implies the first condition because th0t\leq h^{0}. It implies the second condition because, by the definition of TGT_{G}, TG(x)T_{G}(x) is strictly increasing for xa>0hax\leq\sum_{a>0}h^{a} and TG(a>0ha)=a>0ahaT_{G}(\sum_{a>0}h^{a})=\sum_{a>0}ah^{a}.

Hence the hypotheses of Theorem 8.17 are satisfied, showing that the integral points are not Zariski dense, contradicting our construction. We thus conclude that our assumption for contradiction is false, i.e. that YXK×KAY\subseteq X_{K}\times_{K}A is equal to the translate of a constant hypersurface H0AH_{0}\in A by a section sA(XK)s\in A(X_{K}). It follows that every point of (ResKHilb)([1/S])prim({\operatorname{Res}^{K}_{\mathbb{Q}}}\mathrm{Hilb})(\mathbb{Z}[1/S])^{prim} that is contained in XsingX^{sing} corresponds to a hypersurface that is a translate of H0H_{0} (by the value of the section ss on the inverse image of that point). Since there are finitely many irreducible components of the Zariski closure, every point of (ResKHilb)([1/S])prim({\operatorname{Res}^{K}_{\mathbb{Q}}}\mathrm{Hilb})(\mathbb{Z}[1/S])^{prim} is contained in at least one irreducible component, and our argument applies to each irreducible component, we see that there exist finitely many hypersurfaces such that every primitive hypersurface with good reduction away from SS is isomorphic to a translate of one of them. In other words, there are only finitely many primitive hypersurfaces, up to translation. ∎

Corollary 9.3.

Let KK, AA, SS, ϕ\phi be as in Theorem 9.2.

Suppose that either n4n\geq 4 or n=3n=3 and dd is not a multiple of (a(i)+a(i+1)a(i+1))/6\binom{a(i)+a(i+1)}{a(i+1)}/6 for any i2i\geq 2

There are only finitely many smooth hypersurfaces HAH\subseteq A representing ϕ\phi, with good reduction outside SS, up to translation.

Proof.

Any hypersurface HAH\subseteq A is of the form π1H\pi^{-1}H^{\prime}, where π:AA\pi\colon A\rightarrow A^{\prime} is an isogeny defined over KK, and HAH^{\prime}\subseteq A^{\prime} is a primitive hypersurface defined over KK. In this case ϕ\phi is the pullback of an ample class ϕ\phi^{\prime} along π\pi, and we have d=ϕn/n!=(ϕn/n!)degπd=\phi^{n}/n!=(\phi^{\prime n}/n!)\cdot\deg\pi. Hence degπ\deg\pi is bounded, so there are only finitely many possibilities for π\pi. For each one there is a unique ϕ\phi^{\prime} with ϕ=πϕ\phi=\pi^{*}\phi^{\prime}. Furthermore, d(ϕ)(a(i)+a(i+1)a(i+1))/6d(\phi^{\prime})\neq\binom{a(i)+a(i+1)}{a(i+1)}/6 for any i2i\geq 2. We conclude by applying Theorem 9.2 to each (A,ϕ)(A^{\prime},\phi^{\prime}). ∎

Theorem 9.4.

Suppose dimA4\dim A\geq 4. Fix an ample class ϕ\phi in the Néron-Severi group of AA. There are only finitely many smooth hypersurfaces HAH\subseteq A representing ϕ\phi, with good reduction outside SS, up to translation.

Proof.

This is one case of Corollary 9.3 . ∎

Theorem 9.5.

Suppose dimA=3\dim A=3. Fix an ample class ϕ\phi in the Néron-Severi group of AA. Assume that the intersection number ϕϕϕ\phi\cdot\phi\cdot\phi is not divisible by d(i)d(i) for any i2i\geq 2. There are only finitely many smooth hypersurfaces HAH\subseteq A representing ϕ\phi, with good reduction outside SS, up to translation.

Proof.

This is one case of Corollary 9.3, once we cancel the factor of 3!=63!=6 from the denominators. ∎

Theorem 9.6.

Suppose dimA=2\dim A=2. Fix an ample class ϕ\phi in the Néron-Severi group of AA. There are only finitely many smooth curves CAC\subseteq A representing ϕ\phi, with good reduction outside SS, up to translation.

Proof.

This is a consequence of the Shafarevich conjecture for curves. By the Shafarevich conjecture, there are only finitely many possibilities for the isomorphism class of CC. Consider a fixed CC; enlarging KK if necessary, we may assume that C(K)C(K) is nonempty, and choose a basepoint x0C(K)x_{0}\in C(K). It is enough to show that there are only finitely many maps CAC\rightarrow A taking x0x_{0} to the origin of AA, for which the image of CC represents the class ϕ\phi.

By the Albanese property, pointed maps (C,x0)(A,0)(C,x_{0})\rightarrow(A,0) are in bijection with maps of abelian varieties JacCA\operatorname{Jac}C\rightarrow A; the set of such maps forms a finitely generated free abelian group. Fix an ample class ψ\psi on AA; define the degree of any map f:JacCAf\colon\operatorname{Jac}C\rightarrow A by

degf=f(C)ψ.\deg f=f(C)\cdot\psi.

This intersection number is a positive definite quadratic form on Hom(JacC,A)\operatorname{Hom}(\operatorname{Jac}C,A), so there are only finitely many maps of given degree, and any map representing ϕ\phi has degree ϕψ\phi\cdot\psi. ∎

Appendix A Verifying the numerical conditions

The goal of this section is to verify the two numerical conditions of Theorem 8.17 for cc sufficiently large. This is very similar to the estimates in [47, Section 10.2].

Our Hodge-Deligne systems arise from pushforwards of character sheaves of an abelian variety along families of hypersurfaces in that abelian variety. Thus, the Hodge structure on the stalk at a point arises from the cohomology of a hypersurface, twisted by a rank-one character sheaf.

For AA an abelian variety of dimension nn, YY a degree dd hypersurface in AA, χ\chi a finite-order character of π1(A)\pi_{1}(A) such that Hi(Y,χ)H^{i}(Y,\mathcal{L}_{\chi}) vanishes for in1i\neq n-1, the kkth Hodge number of the Hodge filtration on Hn1(Y,χ)H^{n-1}(Y,\mathcal{L}_{\chi}) is dA(n,k)dA(n,k), where A(n,k)A(n,k) is the Eulerian number, by Lemma 3.4. (The degree of a hypersurface in an abelian variety is the degree of the corresponding polarization.)

Recall from Section 5.5 that the various realizations of Hn1(Y,χ)H^{n-1}(Y,\mathcal{L}_{\chi}) form a Hodge–Deligne system with 𝐇\mathbf{H}-structure, for 𝐇\mathbf{H} one of GL\operatorname{GL}, GSp\operatorname{GSp}, and GO\operatorname{GO}. This system has uniform Hodge numbers (Definition 5.46), as computed in Lemma 3.4, so it makes sense to talk about the Hodge filtration as a filtration on 𝐇\mathbf{H}.

Lemma A.1.

Let YY be a smooth hypersurface in an nn-dimensional abelian variety AA, let χ\chi be a finite-order character of π1(A)\pi_{1}(A) such that Hi(Y,χ)H^{i}(Y,\mathcal{L}_{\chi}) vanishes for in1i\neq n-1, and let VV be the Hodge structure on Hn1(Y,χ)H^{n-1}(Y,\mathcal{L}_{\chi}), regarded as a filtered vector space with 𝐇\mathbf{H}-structure, for 𝐇\mathbf{H} one of GL\operatorname{GL}, GSp\operatorname{GSp}, GO\operatorname{GO}. Let hkh^{k} be the adjoint Hodge numbers (Definition 5.44), and let tt be the dimension of a maximal torus in 𝐇\mathbf{H}.

If n2n\geq 2, we have 12(h0t)<k>0hk\frac{1}{2}(h^{0}-t)<\sum_{k>0}h^{k}.

The relevance of this inequality is that TG(k>0hk)=k>0khkT_{G}(\sum_{k>0}h^{k})=\sum_{k>0}kh^{k} by the definition of TGT_{G} (Definition 5.45), so by the monotonicity of TGT_{G} this implies

TG(12(h0t))<TG(k>0hk)=k>0khk,T_{G}(\frac{1}{2}(h^{0}-t))<T_{G}(\sum_{k>0}h^{k})=\sum_{k>0}kh^{k},

which for cc sufficiently large is equivalent to the second numerical condition of Theorem 8.17

Proof.

(See also [47, Lemma 10.3].)

Since the adjoint Hodge numbers satisfy hk=hkh^{k}=h^{-k}, we have

k>0hk=12((khk)h0)=12(dim𝐇h0).\sum_{k>0}h^{k}=\frac{1}{2}\left((\sum_{k\in\mathbb{Z}}h^{k})-h^{0}\right)=\frac{1}{2}\left(\dim\mathbf{H}-h^{0}\right).

Thus it is enough to show that

(8) 2h0<dim𝐇+t.2h^{0}<\dim\mathbf{H}+t.

In each case, we will calculate h0h^{0} and dim𝐇\dim\mathbf{H} in terms of Eulerian numbers A(n,k)A(n,k), and then prove (8) by using the following inequality of Eulerian numbers, to be established later:

(9) 2kd2A(n,k)2d2(kA(n,k))22\sum_{k}d^{2}A(n,k)^{2}\leq d^{2}\left(\sum_{k}A(n,k)\right)^{2}

We adopt the convention that A(n,k)=0A(n,k)=0 when k{0,,n1}k\not\in\{0,\ldots,n-1\}.

  • If 𝐇=GL\mathbf{H}=\operatorname{GL} then

    h0\displaystyle h^{0} =\displaystyle= kd2A(n,k)2\displaystyle\sum_{k}d^{2}A(n,k)^{2}
    dim𝐇\displaystyle\dim\mathbf{H} =\displaystyle= d2(kA(n,k))2\displaystyle d^{2}\left(\sum_{k}A(n,k)\right)^{2}
    t\displaystyle t =\displaystyle= dkA(n,k).\displaystyle d\sum_{k}A(n,k).

    so

    2h0=2kd2A(n,k)2d2(kA(n,k))2<d2(kA(n,k))2+dkA(n,k)=dim𝐇+t.2h^{0}=2\sum_{k}d^{2}A(n,k)^{2}\leq d^{2}\left(\sum_{k}A(n,k)\right)^{2}<d^{2}\left(\sum_{k}A(n,k)\right)^{2}+d\sum_{k}A(n,k)=\dim\mathbf{H}+t.
  • If 𝐇=GSp\mathbf{H}=\operatorname{GSp} then

    h0\displaystyle h^{0} =\displaystyle= 12[(kd2A(n,k)2)+dA(n,n12)]+1\displaystyle\frac{1}{2}\left[\left(\sum_{k}d^{2}A(n,k)^{2}\right)+dA(n,\frac{n-1}{2})\right]+1
    dim𝐇\displaystyle\dim\mathbf{H} =\displaystyle= 12d(kA(n,k))[d(kA(n,k))+1]+1\displaystyle\frac{1}{2}d\left(\sum_{k}A(n,k)\right)\left[d\left(\sum_{k}A(n,k)\right)+1\right]+1
    t\displaystyle t =\displaystyle= 12dkA(n,k)+1.\displaystyle\frac{1}{2}d\sum_{k}A(n,k)+1.

    so

    2h0=(kd2A(n,k)2)+dA(n,n12)+2d22(kA(n,k))2+dA(n,n12)+22h^{0}=\left(\sum_{k}d^{2}A(n,k)^{2}\right)+dA(n,\frac{n-1}{2})+2\leq\frac{d^{2}}{2}\left(\sum_{k}A(n,k)\right)^{2}+dA(n,\frac{n-1}{2})+2
    <d22(kA(n,k))2+d(kA(n,k))+2=dim𝐇+t<\frac{d^{2}}{2}\left(\sum_{k}A(n,k)\right)^{2}+d\left(\sum_{k}A(n,k)\right)+2=\dim\mathbf{H}+t
  • If 𝐇=GO\mathbf{H}=\operatorname{GO} then

    h0\displaystyle h^{0} =\displaystyle= 12[(kd2A(n,k)2)dA(n,n12)]+1\displaystyle\frac{1}{2}\left[\left(\sum_{k}d^{2}A(n,k)^{2}\right)-dA(n,\frac{n-1}{2})\right]+1
    dim𝐇\displaystyle\dim\mathbf{H} =\displaystyle= 12d(kA(n,k))[d(kA(n,k))1]+1\displaystyle\frac{1}{2}d\left(\sum_{k}A(n,k)\right)\left[d\left(\sum_{k}A(n,k)\right)-1\right]+1
    t\displaystyle t =\displaystyle= 12dkA(n,k)+1,\displaystyle\frac{1}{2}d\sum_{k}A(n,k)+1,

    (The formula for tt in the GO\operatorname{GO} case holds only for even-dimensional orthogonal groups, but dkA(n,k)=dn!d\sum_{k}A(n,k)=d\cdot n! is even.) Thus

    2h0=(kd2A(n,k)2)dA(n,n12)+2<(kd2A(n,k)2)+2d22(kA(n,k))2+2=dim𝐇+t.2h^{0}=\left(\sum_{k}d^{2}A(n,k)^{2}\right)-dA(n,\frac{n-1}{2})+2<\left(\sum_{k}d^{2}A(n,k)^{2}\right)+2\leq\frac{d^{2}}{2}\left(\sum_{k}A(n,k)\right)^{2}+2=\dim\mathbf{H}+t.

We now prove (9). It is known that, as nn grows large, the Eulerian numbers approximate a normal distribution with variance n/12\sqrt{n}/12. This purely qualitative result implies that (9) holds for sufficiently large nn.

To get precise bounds, we’ll use log concavity, together with a calculation of the second moment. The key idea is that a sequence of numbers that (i) is log-concave, and (ii) has large second moment, cannot be too concentrated at the middle term. Let

ai=kA(n,k)A(n,ki)(kA(n,k))2;a_{i}=\frac{\sum_{k}A(n,k)A(n,k-i)}{(\sum_{k}A(n,k))^{2}};

this is normalized so that ai=1\sum a_{i}=1.

Now we’ll prove that a01/2a_{0}\leq 1/2 for all n2n\geq 2; this will be Lemma A.6. This will be a consequence of log-concavity and a formula for the second moment.

Lemma A.2.

The sequence (ai)(a_{i}) is log-concave and satisfies ai=aia_{-i}=a_{i}.

Proof.

This is proved in the first paragraph of the proof of [47, Lemma 10.3]. Symmetry is elementary; log-concavity follows from the classical fact that the Eulerian numbers are log-concave (see, for example, [52, Problems 4.6 and 4.8]) and the fact that log-concavity is preserved under convolution [34, Thms. 1.4, 3.3]. ∎

Lemma A.3.

The second moment of (ai)(a_{i}) is

ii2ai=n+16.\sum_{i}i^{2}a_{i}=\frac{n+1}{6}.
Proof.

This is [47, Eqn. 10.10]. ∎

Lemma A.4.

Suppose a0>1/2a_{0}>1/2. Then for all k1k\geq 1 we have

i=kai<i=k123i.\sum_{i=k}^{\infty}a_{i}<\sum_{i=k}^{\infty}\frac{1}{2\cdot 3^{i}}.
Proof.

Because a0>1/2a_{0}>1/2, by symmetry, we have

i=1ai=1a02<14=i=1123i.\sum_{i=1}^{\infty}a_{i}=\frac{1-a_{0}}{2}<\frac{1}{4}=\sum_{i=1}^{\infty}\frac{1}{2\cdot 3^{i}}.

so if

i=kaii=k123i\sum_{i=k}^{\infty}a_{i}\geq\sum_{i=k}^{\infty}\frac{1}{2\cdot 3^{i}}

we must have aj<123ja_{j}<\frac{1}{2\cdot 3^{j}} for some 1j<k1\leq j<k.

But then by log-concavity and the fact that a0>12a_{0}>\frac{1}{2}, we have ai<123ja_{i}<\frac{1}{2\cdot 3^{j}} for all i>ji>j, and thus in particular for all iki\geq k, so

i=kai<i=k123i.\sum_{i=k}^{\infty}a_{i}<\sum_{i=k}^{\infty}\frac{1}{2\cdot 3^{i}}.

Lemma A.5.

If a0>1/2a_{0}>1/2 then kk2ak<32\sum_{k}k^{2}a_{k}<\frac{3}{2}.

Proof.

We have

kk2ak=2k=1k2ak=2k=1(2k1)i=kai>2k=1(2k1)i=k123i=2k=1k2123i=32.\sum_{k}k^{2}a_{k}=2\sum_{k=1}^{\infty}k^{2}a_{k}=2\sum_{k=1}^{\infty}(2k-1)\sum_{i=k}^{\infty}a_{i}>2\sum_{k=1}^{\infty}(2k-1)\sum_{i=k}^{\infty}\frac{1}{2\cdot 3^{i}}=2\sum_{k=1}^{\infty}k^{2}\frac{1}{2\cdot 3^{i}}=\frac{3}{2}.

Here we use symmetry, Lemma A.4, and the identity

k>0k2λk=λ(1+λ)(1λ)3.\sum_{k>0}k^{2}\lambda^{k}=\frac{\lambda(1+\lambda)}{(1-\lambda)^{3}}.

Lemma A.6.

If n2n\geq 2 then a01/2a_{0}\leq 1/2.

Proof.

If n9n\geq 9, this is immediate from Lemmas A.3 and A.5.

For 2n82\leq n\leq 8, we can prove this by computation. For even nn this follows immediately from the symmetry A(n,k)=A(n,n1k)A(n,k)=A(n,n-1-k), so it suffices to check n=3,5,7n=3,5,7, which can be done by hand. ∎

This proves Lemma A.1. ∎

Appendix B Combinatorics involving binomial coefficients and Eulerian numbers

This section is devoted to proving Proposition 3.15. Thus, throughout this section, we preserve the notation and assumptions of Proposition 3.15, which we review here.

We use A(n,q)A(n,q) for the Eulerian numbers. We adopt the convention that A(n,q)=0A(n,q)=0 unless 0q<n0\leq q<n; similarly, we take (nq){n\choose q} to vanish whenever nn is positive but q<0q<0 or qnq\geq n.

Recall from the introduction that a(i)a(i) is the sequence satisfying

a(1)=1,a(2)=5,a(i+2)=4a(i+1)+1a(i).a(1)=1,a(2)=5,a(i+2)=4a(i+1)+1-a(i).
Proposition B.1 (Proposition 3.15).

Let n2n\geq 2 and d1d\geq 1 be integers. Suppose that there exists a natural number kk, function mHm_{H} from the integers to the natural numbers and an integer ss. Write m=imH(i)m=\sum_{i}m_{H}(i), and suppose that 1<k<m11<k<m-1. Suppose the equation

(10) mS:0mS(i)mH(i)imS(i)=kiimS(i)=s+qi(mH(i)mS(i))=dA(n,q).\sum_{\begin{subarray}{c}m_{S}:\mathbb{Z}\to\mathbb{Z}\\ 0\leq m_{S}(i)\leq m_{H}(i)\\ \sum_{i}m_{S}(i)=k\\ \sum_{i}im_{S}(i)=s+q\end{subarray}}\prod_{i}{m_{H}(i)\choose m_{S}(i)}=dA(n,q).

is satisfied for all qq\in\mathbb{Z}. Then we have one of the cases

  1. (1)

    m=4m=4 and k=2k=2

  2. (2)

    n=2n=2 and d=(2k1k)d={2k-1\choose k} for some k>2k>2

  3. (3)

    n=3n=3 and d=(a(i)+a(i+1)a(i))/6d={a(i)+a(i+1)\choose a(i)}/6 for some i2i\geq 2.

A tuple (n,d,m,k,mH,s)(n,d,m,k,m_{H},s) satisfying the conditions of the proposition will be called a solution.

We first handle the cases n=2,3,4n=2,3,4 directly, then give a general argument that handles cases n5n\geq 5. Thus Proposition B.1 will follow immediately once we have proven Lemmas B.4 (the n=2n=2 case), B.8 (the n=3n=3 case), B.9 (the n=4n=4 case), B.18, B.21, B.25, B.27, and B.31 (which collectively handle the n5n\geq 5 case).

We can assume without loss of generality that km/2k\leq m/2.

Let mmaxm_{max} and mminm_{min} be the functions mS:m_{S}:\mathbb{Z}\to\mathbb{Z} satisfying 0mS(i)mH(i),imS(i)=k0\leq m_{S}(i)\leq m_{H}(i),\sum_{i}m_{S}(i)=k, and maximizing, respectively, minimizing iimS(i)\sum_{i}im_{S}(i).

Lemma B.2.

There is a unique ww such that mmin(i)=mH(i)m_{min}(i)=m_{H}(i) for all i<wi<w, mmin(i)=0m_{min}(i)=0 for all i>wi>w, and mmin(i)>0m_{min}(i)>0 for i=wi=w.

Similarly, there is a unique ww^{\prime} such that mmax(i)=mH(i)m_{max}(i)=m_{H}(i) for all i>wi>w^{\prime}, mmax=0m_{max}=0 for all i<wi<w^{\prime} and mmax>0m_{max}>0 for i=wi=w^{\prime}.

Furthermore www\leq w^{\prime}.

Proof.

We take ww to be the largest ii such that mmin(i)>0m_{min}(i)>0. The last two conditions are then obvious and the fact that mmin(i)=mH(i)m_{min}(i)=m_{H}(i) for i<wi<w follows by minimality – if it were not so, we could increase mmin(i)m_{min}(i) by 11, reduce mmin(w)m_{min}(w) by 11, and thereby reduce iimS(i)\sum_{i}im_{S}(i) by wiw-i.

We take ww^{\prime} to be the least ii such that mmax(i)>0m_{max}(i)>0, and make a symmetrical argument.

Finally, for contradiction, assume w>ww>w^{\prime}. Then

imH(i)\displaystyle\sum_{i}m_{H}(i) =\displaystyle= m2k=immin(i)+immax(i)\displaystyle m\geq 2k=\sum_{i}m_{min}(i)+\sum_{i}m_{max}(i)
>\displaystyle> i<wmmin(i)+i>wmmax(i)=i<wmH(i)+i>wmH(i)imH(i),\displaystyle\sum_{i<w}m_{min}(i)+\sum_{i>w^{\prime}}m_{max}(i)=\sum_{i<w}m_{H}(i)+\sum_{i>w^{\prime}}m_{H}(i)\geq\sum_{i}m_{H}(i),

a contradiction. ∎

Lemma B.3.

We have

(11) i<w(wi)mH(i)+i>w(iw)mH(i)+k(ww)=n1\sum_{i<w}(w-i)m_{H}(i)+\sum_{i>w^{\prime}}(i-w^{\prime})m_{H}(i)+k(w^{\prime}-w)=n-1

Note that all the terms on the left side of this equation are nonnegative.

Proof.

Because the function qdE(n,q)q\mapsto dE(n,q) is supported on qq ranging from 0 to n1n-1, we must have

n1=(s+n1)s=iimmax(i)iimmin(i)=i(iw)mmax(i)i(iw)mmin(i)+k(ww)n-1=(s+n-1)-s=\sum_{i}im_{max}(i)-\sum_{i}im_{min}(i)=\sum_{i}(i-w^{\prime})m_{max}(i)-\sum_{i}(i-w)m_{min}(i)+k(w^{\prime}-w)
=i<w(wi)mH(i)+i>w(iw)mH(i)+k(ww)=\sum_{i<w}(w-i)m_{H}(i)+\sum_{i>w^{\prime}}(i-w^{\prime})m_{H}(i)+k(w^{\prime}-w)

B.1. The case n4n\leq 4

Lemma B.4.

Suppose n=2n=2. Then we must have d=(2k1k)d={2k-1\choose k}. Furthermore if k=2k=2 then m=4m=4.

Proof.

In (11), because the summands on the left side are nonnegative, one must be 11 and the others must vanish. The one that is 11 can only be the summand associated to i=w1i=w-1 or i=w+1i=w^{\prime}+1 as the other summands are integer multiples of something at least 22. By symmetry, we may assume the 11 comes from i=w+1i=w+1. Because the last summand vanishes, we must have w=ww=w^{\prime}.

This gives mH(i)=0m_{H}(i)=0 unless i=wi=w or w+1w+1 and mH(w+1)=1.m_{H}(w+1)=1.

Then the only possible solutions to 0mS(i)mH(i)0\leq m_{S}(i)\leq m_{H}(i) and imS(i)=k\sum_{i}m_{S}(i)=k are mS(w)=k,mS(w+1)=0m_{S}(w)=k,m_{S}(w+1)=0 and mS(w)=k1,mS(w+1)=1m_{S}(w)=k-1,m_{S}(w+1)=1. These have imS(i)\sum_{i}m_{S}(i) equal to kwkw and kw+1kw+1 respectively, so we must have s=kws=kw. This implies

(mS(w)k)=(mS(w)k)(10)=dE(2,0)=d{m_{S}(w)\choose k}={m_{S}(w)\choose k}{1\choose 0}=dE(2,0)=d
(mS(w)k1)=(mS(w)k1)(11)=dE(2,1)=d{m_{S}(w)\choose k-1}={m_{S}(w)\choose k-1}{1\choose 1}=dE(2,1)=d

and thus

(mS(w)k1)=(mS(w)k){m_{S}(w)\choose k-1}={m_{S}(w)\choose k}

which implies mS(w)=2k1m_{S}(w)=2k-1 and thus d=(2k1k).d={2k-1\choose k}.

In the k=2k=2 case, we have m=mS(w)+mS(w+1)=2k=4m=m_{S}(w)+m_{S}(w+1)=2k=4, as desired.

Lemma B.5.

Suppose n=3n=3. Then we must have mH(i)=0m_{H}(i)=0 for iw1,w,w+1i\neq w-1,w,w+1 and mH(w1)=mH(w+1)=1m_{H}(w-1)=m_{H}(w+1)=1 unless m=4m=4 and k=2k=2.

Proof.

In (11), either two terms are 11 and the rest zero or one term is 22 and the rest are zero. In the first case, since the only terms that can be 11 are i=w1i=w-1 and i=w+1i=w^{\prime}+1, implying in particular that w=ww=w^{\prime} we have the stated conclusion. So it suffices to eliminate the case that one term is 22. The only possibilities are i=w2,w1,w+1,w+2i=w-2,w-1,w^{\prime}+1,w^{\prime}+2, and the last term if k=2k=2. By symmetry, we are reduced to eliminating i=w+1i=w^{\prime}+1, i=w+2i=w^{\prime}+2, and the final term. In the first two cases we have w=ww^{\prime}=w and in the last case we have k=2,ww=1k=2,w^{\prime}-w=1.

If mH(i)=0m_{H}(i)=0 for all ii except w,w+1w,w+1, and mH(w+1)=2m_{H}(w+1)=2, then there are three possibilities for mSm_{S}: (mS(w),mS(w+1))(m_{S}(w),m_{S}(w+1)) must equal (k,0),(k1,1),(k,0),(k-1,1), or (k2,2)(k-2,2). Using mH(w)=m2m_{H}(w)=m-2, this gives

(m2k)=(m2k)(20)=dE(3,0)=d{m-2\choose k}={m-2\choose k}{2\choose 0}=dE(3,0)=d
2(m2k1)=(m2k1)(21)=dE(3,1)=4d2{m-2\choose k-1}={m-2\choose k-1}{2\choose 1}=dE(3,1)=4d
(m2k2)=(m2k2)(22)=dE(3,2)=d{m-2\choose k-2}={m-2\choose k-2}{2\choose 2}=dE(3,2)=d

and this implies

12=(m2k)(m2k1)=mk1k\frac{1}{2}=\frac{{m-2\choose k}}{{m-2\choose k-1}}=\frac{m-k-1}{k}
12=(m2k2)(m2k1)=k1mk\frac{1}{2}=\frac{{m-2\choose k-2}}{{m-2\choose k-1}}=\frac{k-1}{m-k}

so mk=2k2m-k=2k-2 and k=2m2k2k=2m-2k-2 giving m=4,k=2m=4,k=2.

If mH(i)=0m_{H}(i)=0 for all ii except w,w+2w,w+2, then iimS(i)kwmod2\sum_{i}im_{S}(i)\equiv kw\mod 2 whenever 0mS(i)mH(i)0\leq m_{S}(i)\leq m_{H}(i) and imS(i)=k\sum_{i}m_{S}(i)=k. Hence the left side of (3) is nonzero only when s+qkwmod2s+q\equiv kw\mod 2. This contradicts the fact that E(2,q)E(2,q) is nonzero for qq of both parities.

If mH(i)=0m_{H}(i)=0 for all ii except w,w+1w,w+1 and k=2k=2, then there are three possibilities for mSm_{S}: (mS(w),mS(w+1))(m_{S}(w),m_{S}(w+1)) must be (2,0),(1,1),(2,0),(1,1), or (0,2)(0,2). This gives

mH(w)(mH(w)1)2=(mH(w)2)(mH(w+1)0)=dA(3,0)=d\frac{m_{H}(w)(m_{H}(w)-1)}{2}={m_{H}(w)\choose 2}{m_{H}(w+1)\choose 0}=dA(3,0)=d
mH(w)mH(w+1)=(mH(w)1)(mH(w+1)1)=dA(3,1)=4dm_{H}(w)m_{H}(w+1)={m_{H}(w)\choose 1}{m_{H}(w+1)\choose 1}=dA(3,1)=4d
mH(w+1)(mH(w+1)1)2=(mH(w)0)(mH(w+1)2)=dA(3,2)=d\frac{m_{H}(w+1)(m_{H}(w+1)-1)}{2}={m_{H}(w)\choose 0}{m_{H}(w+1)\choose 2}=dA(3,2)=d

Because d>0d>0 this implies mH(w)=mH(w+1)m_{H}(w)=m_{H}(w+1) and thus 2mH(w)(mH(w)1))=mH(w)22m_{H}(w)(m_{H}(w)-1))=m_{H}(w)^{2} which implies mH(w)=mH(w+1)=2m_{H}(w)=m_{H}(w+1)=2 and thus m=4m=4, k=2k=2. ∎

Lemma B.6.

Suppose n=3n=3. Then unless m=4m=4 and k=2k=2, we have

(mk)24(mk)k+k2=m.(m-k)^{2}-4(m-k)k+k^{2}=m.
Proof.

By Lemma B.5, we have mH(i)=0m_{H}(i)=0 for iw1,w,w+1i\neq w-1,w,w+1 and mH(w1)=mH(w+1)=1m_{H}(w-1)=m_{H}(w+1)=1. This means there are four possibilities for mSm_{S}: (mS(w1),mS(w),mS(w+1))(m_{S}(w-1),m_{S}(w),m_{S}(w+1)) must equal (0,k,0)(0,k,0), (0,k1,1)(0,k-1,1), (1,k1,0)(1,k-1,0), or (1,k2,1)(1,k-2,1). Using mH(w)=m2m_{H}(w)=m-2, and (10)=(11)=1{1\choose 0}={1\choose 1}=1, we get

(m2k1)=dE(3,0)=d{m-2\choose k-1}=dE(3,0)=d
(m2k)+(m2k2)=dE(3,1)=4d{m-2\choose k}+{m-2\choose k-2}=dE(3,1)=4d
(m2k1)=dE(3,2)=d{m-2\choose k-1}=dE(3,2)=d

so in other words we have

(m2k)+(m2k2)=4(m2k1){m-2\choose k}+{m-2\choose k-2}=4{m-2\choose k-1}

which dividing by (m2)!(m-2)! and multiplying by k!(mk)!k!(m-k)! is

(mk)(mk1)+k(k1)=4k(mk)(m-k)(m-k-1)+k(k-1)=4k(m-k)

which is exactly the stated Diophantine equation. ∎

Lemma B.7.

The positive integer solutions to a24ab+b2=a+ba^{2}-4ab+b^{2}=a+b with aba\leq b have the form (a,b)=(a(i),a(i+1))(a,b)=(a(i),a(i+1)) for some i1i\geq 1.

Proof.

Let aa and bb be positive integers with a24ab+b2a^{2}-4ab+b^{2} and bab\geq a. We will show that there exists i1i\geq 1 such that a=a(i)a=a(i) and b=a(i+1)b=a(i+1). By induction, it suffices to prove that either a=1,b=5a=1,b=5 or that there exists an integer bb^{\prime} with 0<b<a0<b^{\prime}<a, b=4a+1bb=4a+1-b^{\prime}, and (b,a)(b^{\prime},a) solving the same equation, as if b=a(i),a=a(i+1)b^{\prime}=a(i^{\prime}),a=a(i^{\prime}+1) then b=a(i+2)b=a(i^{\prime}+2).

To do this, let b=4a+1bb^{\prime}=4a+1-b. Then because we can rewrite the equation as

b(4a+1b)=a2a,b(4a+1-b)=a^{2}-a,

as bb is a solution then bb^{\prime} is a solution as well. Furthermore, if a>1a>1 then a2a>0a^{2}-a>0 so b=a2ab>0b^{\prime}=\frac{a^{2}-a}{b}>0, and we must have b<ab^{\prime}<a because if bab^{\prime}\geq a we have

a2a=b(4a+1b)aa>a2a,a^{2}-a=b(4a+1-b)\geq a\cdot a>a^{2}-a,

so if a>1a>1 we always have such a bb^{\prime}. On the other hand, if a=1a=1 then b(5b)=0b(5-b)=0 so, because b>0b>0, we have b=5b=5, the base case. ∎

Lemma B.8.

Suppose n=3n=3. Then either m=4m=4 and k=2k=2 or m=a(i)+a(i+1)m=a(i)+a(i+1) and k=a(i)k=a(i) for some i2i\geq 2.

Proof.

This follows from Lemma B.6 and B.7 once we observe that because km/2k\leq m/2, we have kmkk\leq m-k, and because k2k\geq 2, the case k=a(1),m=a(2)k=a(1),m=a(2) cannot occur. ∎

Lemma B.9.

There are no solutions for n=4n=4.

Proof.

We first consider the contribution to (11) from k(ww)k(w^{\prime}-w). Because k2k\geq 2 and this contribution is at most (n1)=3(n-1)=3, we can only have w=ww^{\prime}=w or w=w+1w^{\prime}=w+1, k=2k=2 or 33. Let us eliminate the w=w+1w^{\prime}=w+1 cases first.

In the k=3k=3 case we have mH(i)=0m_{H}(i)=0 unless i=wi=w or w+1w+1. Thus we have four possibilities for mSm_{S} – we must have (mS(w),mS(w+1))=(3,0),(2,1),(1,2),(m_{S}(w),m_{S}(w+1))=(3,0),(2,1),(1,2), or (0,3)(0,3). This gives

(mH(w)3)(mH(w+1)0)=dE(4,0)=d{m_{H}(w)\choose 3}{m_{H}(w+1)\choose 0}=dE(4,0)=d
(mH(w)2)(mH(w+1)1)=dE(4,1)=11d{m_{H}(w)\choose 2}{m_{H}(w+1)\choose 1}=dE(4,1)=11d
(mH(w)1)(mH(w+1)2)=dE(4,2)=11d{m_{H}(w)\choose 1}{m_{H}(w+1)\choose 2}=dE(4,2)=11d
(mH(w)0)(mH(w+1)3)=dE(4,3)=d{m_{H}(w)\choose 0}{m_{H}(w+1)\choose 3}=dE(4,3)=d

Combining the first and last equations with d>0d>0, we see that mH(w)=mH(w+1)m_{H}(w)=m_{H}(w+1). Dividing the second equation by the first, we get

3mH(w)mH(w)2=11\frac{3m_{H}(w)}{m_{H}(w)-2}=11

which implies mH(w)=114m_{H}(w)=\frac{11}{4}, a contradiction.

In the k=2k=2 case, one more term must be 11, which is i=w1i=w-1 or i=w+1i=w^{\prime}+1. Without loss of generality, it is i=w1i=w-1. Then we have mH(i)=0m_{H}(i)=0 unless i=w1,w,w+1i=w-1,w,w+1 and mH(w1)=1m_{H}(w-1)=1. Then the possible values of (mS(w1),mS(w),mS(w+1))(m_{S}(w-1),m_{S}(w),m_{S}(w+1)) are (1,1,0)(1,1,0), (1,0,1)(1,0,1), (0,2,0)(0,2,0), (0,1,1)(0,1,1), and (0,0,2)(0,0,2). This gives (ignoring factors of the form (n0)n\choose 0 or (11)1\choose 1)

(mH(w)1)=dE(4,0)=d{m_{H}(w)\choose 1}=dE(4,0)=d
(mH(w+1)1)+(mH(w)2)=dE(4,1)=11d{m_{H}(w+1)\choose 1}+{m_{H}(w)\choose 2}=dE(4,1)=11d
(mH(w)1)(mH(w+1)1)=dE(4,2)=11d{m_{H}(w)\choose 1}{m_{H}(w+1)\choose 1}=dE(4,2)=11d
(mH(w+1)2)=d{m_{H}(w+1)\choose 2}=d

Dividing the third equation by the first, we get mH(w+1)=11m_{H}(w+1)=11. Thus by the fourth equation d=(112)=55d={11\choose 2}=55, and by the first equation mH(w)=55m_{H}(w)=55. Then the second equation gives 11+(552)=115511+{55\choose 2}=11\cdot 55, which is false, so there are no solutions.

We now handle the case w=ww=w^{\prime}. In this case, because the total sum of (11) is 33, and only the two terms i=w+1,i=w1i=w+1,i=w-1 can contribute a 11, we must have one term contributing 33 or one 22 and one 11. There are four terms that might contribute 33: i=w+3i=w+3, i=w+1i=w+1, i=w1i=w-1, and i=w3i=w-3, and four that might contribute 22: i=w+2i=w+2, i=w+1i=w+1, i=w1i=w-1, and i=w2i=w-2. This gives a total of 1010 possibilities for mHm_{H}, or 55 up to symmetry: we may assume (mH(w1),mH(w),mH(w+1),mH(w+2),mH(w+3))=(0,m1,0,0,1),(0,m3,3,0,0),(0,m2,1,1,0),(1,m2,0,1,0),(m_{H}(w-1),m_{H}(w),m_{H}(w+1),m_{H}(w+2),m_{H}(w+3))=(0,m-1,0,0,1),(0,m-3,3,0,0),(0,m-2,1,1,0),(1,m-2,0,1,0), or (1,m2,2,0,0)(1,m-2,2,0,0).

The case (0,m1,0,0,1)(0,m-1,0,0,1) is easy to eliminate as it implies that the left side of (3) is nonvanishing only in a single residue class mod 33, but we know that the right side does not have that possibility.

The case (0,m3,3,0,0)(0,m-3,3,0,0) gives us four possibilities for mSm_{S}, implying

(m3k)(30)=dE(4,0)=d{m-3\choose k}{3\choose 0}=dE(4,0)=d
(m3k1)(31)=dE(4,0)=11d{m-3\choose k-1}{3\choose 1}=dE(4,0)=11d
(m3k2)(32)=dE(4,0)=11d{m-3\choose k-2}{3\choose 2}=dE(4,0)=11d
(m3k3)(33)=dE(4,0)=d{m-3\choose k-3}{3\choose 3}=dE(4,0)=d

The first and fourth equations gives (m3k)=(m3k3){m-3\choose k}={m-3\choose k-3}, which implies that m3k=k3m-3-k=k-3 or m=2km=2k. Dividing the third equation by the fourth, we then obtain

11=(2k3k2)(2k3k3)(32)(33)=kk2311=\frac{{2k-3\choose k-2}}{{2k-3\choose k-3}}\frac{{3\choose 2}}{{3\choose 3}}=\frac{k}{k-2}\cdot 3

whose unique solution is k=114k=\frac{11}{4}, a contradiction.

The case (0,m2,1,1,0)(0,m-2,1,1,0) gives us four possibilities for mSm_{S}, implying

(m2k)(10)(10)=dE(4,0)=d{m-2\choose k}{1\choose 0}{1\choose 0}=dE(4,0)=d
(m2k1)(11)(10)=dE(4,1)=11d{m-2\choose k-1}{1\choose 1}{1\choose 0}=dE(4,1)=11d
(m2k1)(10)(11)=dE(4,2)=11d{m-2\choose k-1}{1\choose 0}{1\choose 1}=dE(4,2)=11d
(m2k2)(11)(11)=dE(4,3)=d{m-2\choose k-2}{1\choose 1}{1\choose 1}=dE(4,3)=d

By the first and fourth equations we have (m2k)=(m2k2){m-2\choose k}={m-2\choose k-2} which implies m2k=k2m-2-k=k-2 or m=2km=2k. Dividing the second equation by the first we get 11=kk111=\frac{k}{k-1} so k=1110k=\frac{11}{10}, a contradiction.

The case (1,m2,0,1,0)(1,m-2,0,1,0) gives us four possibilities for mSm_{S}, implying

(11)(m2k1)(10)=dE(4,0)=d{1\choose 1}{m-2\choose k-1}{1\choose 0}=dE(4,0)=d
(10)(m2k)(10)=dE(4,1)=11d{1\choose 0}{m-2\choose k}{1\choose 0}=dE(4,1)=11d
(11)(m2k2)(11)=dE(4,2)=11d{1\choose 1}{m-2\choose k-2}{1\choose 1}=dE(4,2)=11d
(10)(m2k1)(11)=dE(4,3)=d{1\choose 0}{m-2\choose k-1}{1\choose 1}=dE(4,3)=d

By the second and third equations, we have (m2k)=(m2k2){m-2\choose k}={m-2\choose k-2} which again implies m=2km=2k. But then (2k2k1)>(2k2k2){2k-2\choose k-1}>{2k-2\choose k-2} which means d>11dd>11d, a contradiction.

The case (1,m2,2,0,0)(1,m-2,2,0,0) has six possibilities for mSm_{S}, implying

(11)(m2k1)(20)=dE(4,0)=d{1\choose 1}{m-2\choose k-1}{2\choose 0}=dE(4,0)=d
(10)(m2k)(20)+(11)(m2k2)(21)=dE(4,1)=11d{1\choose 0}{m-2\choose k}{2\choose 0}+{1\choose 1}{m-2\choose k-2}{2\choose 1}=dE(4,1)=11d
(11)(m2k3)(22)+(10)(m2k1)(21)=dE(4,2)=11d{1\choose 1}{m-2\choose k-3}{2\choose 2}+{1\choose 0}{m-2\choose k-1}{2\choose 1}=dE(4,2)=11d
(10)(m2k2)(22)=dE(4,3)=d{1\choose 0}{m-2\choose k-2}{2\choose 2}=dE(4,3)=d

By the first and fourth equations, we have (m2k1)=(m2k2){m-2\choose k-1}={m-2\choose k-2} which implies m=2k1m=2k-1. But then since (2k2k3)(2k2k2){2k-2\choose k-3}\leq{2k-2\choose k-2} we have by the third and fourth equations

11d=(2k2k3)+2(2k2k1)(2k2k2)+2(2k2k2)=3d11d={2k-2\choose k-3}+2{2k-2\choose k-1}\leq{2k-2\choose k-2}+2{2k-2\choose k-2}=3d

a contradiction. ∎

B.2. The case n5n\geq 5: general setup

We’ll write m0=mminm_{0}=m_{min} for the rest of this section. The key equality for q=0q=0 gives

(12) d=(mH(w)m0(w)).d={m_{H}(w)\choose m_{0}(w)}.

For any mS:m_{S}:\mathbb{Z}\to\mathbb{Z} as in Equation 10 we’ll write

N(mS)=N(mH,mS)=i(mH(i)mS(i)).N(m_{S})=N(m_{H},m_{S})=\prod_{i}{m_{H}(i)\choose m_{S}(i)}.

Now the key equality (10) becomes

(13) mS:0mS(i)mH(i)imS(i)=kiimS(i)=s+qN(mS)N(m0)=A(n,q).\sum_{\begin{subarray}{c}m_{S}:\mathbb{Z}\to\mathbb{Z}\\ 0\leq m_{S}(i)\leq m_{H}(i)\\ \sum_{i}m_{S}(i)=k\\ \sum_{i}im_{S}(i)=s+q\end{subarray}}\frac{N(m_{S})}{N(m_{0})}=A(n,q).

We’re going to get a contradiction from combinatorial considerations involving the terms associated to small qq in Equation (3).

By abuse of notation, we’ll let [i][i] denote the function m{i}m_{\{i\}} taking the value 11 on ii and zero elsewhere; so any function \mathbb{Z}\rightarrow\mathbb{Z} is a linear combination of the elementary functions [i][i].

There are at most two functions m1m_{1} that contribute to the q=1q=1 case in Equation (13). These are

m1a=m0+[w][w1]m_{1}^{a}=m_{0}+[w]-[w-1]

and

m1b=m0+[w+1][w].m_{1}^{b}=m_{0}+[w+1]-[w].

We compute

N(m1a)N(m0)=mH(w1)(mH(w)m0(w))(m0(w)+1)\frac{N(m_{1}^{a})}{N(m_{0})}=\frac{m_{H}(w-1)(m_{H}(w)-m_{0}(w))}{(m_{0}(w)+1)}
N(m1b)N(m0)=mH(w+1)m0(w)(mH(w)m0(w)+1);\frac{N(m_{1}^{b})}{N(m_{0})}=\frac{m_{H}(w+1)m_{0}(w)}{(m_{H}(w)-m_{0}(w)+1)};

the q=1q=1 case of Equation (13) gives

N(m1a)N(m0)+N(m1b)N(m0)=A(n,1).\frac{N(m_{1}^{a})}{N(m_{0})}+\frac{N(m_{1}^{b})}{N(m_{0})}=A(n,1).

Similarly, there are at most five nonzero terms in the q=2q=2 case Equation (13):

m2a\displaystyle m_{2}^{a} =\displaystyle= m0+[w][w2]\displaystyle m_{0}+[w]-[w-2]
m2b\displaystyle m_{2}^{b} =\displaystyle= m0+2[w]2[w1]\displaystyle m_{0}+2[w]-2[w-1]
m2c\displaystyle m_{2}^{c} =\displaystyle= m0+[w+1][w1]\displaystyle m_{0}+[w+1]-[w-1]
m2d\displaystyle m_{2}^{d} =\displaystyle= m0+2[w+1]2[w]\displaystyle m_{0}+2[w+1]-2[w]
m2e\displaystyle m_{2}^{e} =\displaystyle= m0+[w+2][w].\displaystyle m_{0}+[w+2]-[w].

We have the following equalities.

N(m2a)N(m0)\displaystyle\frac{N(m_{2}^{a})}{N(m_{0})} =\displaystyle= mH(w2)(mH(w)m0(w))(m0(w)+1)\displaystyle\frac{m_{H}(w-2)(m_{H}(w)-m_{0}(w))}{(m_{0}(w)+1)}
N(m2b)N(m0)\displaystyle\frac{N(m_{2}^{b})}{N(m_{0})} =\displaystyle= mH(w1)(mH(w1)1)(mH(w)m0(w))(mH(w)m0(w)1)2(m0(w)+1)(m0(w)+2)\displaystyle\frac{m_{H}(w-1)(m_{H}(w-1)-1)(m_{H}(w)-m_{0}(w))(m_{H}(w)-m_{0}(w)-1)}{2(m_{0}(w)+1)(m_{0}(w)+2)}
N(m2c)N(m0)\displaystyle\frac{N(m_{2}^{c})}{N(m_{0})} =\displaystyle= mH(w1)mH(w+1)\displaystyle m_{H}(w-1)m_{H}(w+1)
N(m2d)N(m0)\displaystyle\frac{N(m_{2}^{d})}{N(m_{0})} =\displaystyle= (mH(w+1))(mH(w+1)1)(m0(w))(m0(w)1)2(mH(w)m0(w)+1)(mH(w)m0(w)+2)\displaystyle\frac{(m_{H}(w+1))(m_{H}(w+1)-1)(m_{0}(w))(m_{0}(w)-1)}{2(m_{H}(w)-m_{0}(w)+1)(m_{H}(w)-m_{0}(w)+2)}
N(m2e)N(m0)\displaystyle\frac{N(m_{2}^{e})}{N(m_{0})} =\displaystyle= m0(w)mH(w+2)mH(w)m0(w)+1.\displaystyle\frac{m_{0}(w)m_{H}(w+2)}{m_{H}(w)-m_{0}(w)+1}.

Equation (13) gives

(14) =a,b,c,d,eN(m2)N(m0)=A(n,2).\sum_{*=a,b,c,d,e}\frac{N(m_{2}^{*})}{N(m_{0})}=A(n,2).

We conclude with a lemma that will be useful at several points in the argument.

Lemma B.10.

We have

N(m1a)N(m0)+N(m1b)N(m0)=A(n,1)\frac{N(m_{1}^{a})}{N(m_{0})}+\frac{N(m_{1}^{b})}{N(m_{0})}=A(n,1)

and

(N(m1a)N(m0))(N(m1b)N(m0))<mH(w1)mH(w+1).\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)\left(\frac{N(m_{1}^{b})}{N(m_{0})}\right)<m_{H}(w-1)m_{H}(w+1).

In particular, if

mH(w1)mH(w+1)<14A(n,1)2,m_{H}(w-1)m_{H}(w+1)<\frac{1}{4}A(n,1)^{2},

let α<β\alpha<\beta be the real roots of

X2A(n,1)X+mH(w1)mH(w+1).X^{2}-A(n,1)X+m_{H}(w-1)m_{H}(w+1).

Then we have

min(N(m1a)N(m0),N(m1b)N(m0))α\min\left(\frac{N(m_{1}^{a})}{N(m_{0})},\frac{N(m_{1}^{b})}{N(m_{0})}\right)\leq\alpha

and

max(N(m1a)N(m0),N(m1b)N(m0))β\max\left(\frac{N(m_{1}^{a})}{N(m_{0})},\frac{N(m_{1}^{b})}{N(m_{0})}\right)\geq\beta
Proof.

The first equality is the q=1q=1 case of Equation (13). The second follows from the explicit formulas for the two quotients N(m1)N(m0)\frac{N(m_{1}^{*})}{N(m_{0})}.

The bounds in terms of α\alpha and β\beta follow from the first two inequalities. ∎

B.3. The case n5n\geq 5, with N(m1a)N(m_{1}^{a}) big

In the following sections we will use without proof a number of inequalities involving Eulerian numbers. The proofs of such inequalities are routine; we discuss the technique in Section C.

Recall notation from the beginning of Section B, and the beginning of Section B.2.

Lemma B.11.

If n5n\geq 5 and

(15) N(m1a)N(m0)A(n,1)2\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}

then mH(w1)=1m_{H}(w-1)=1.

Proof.

We’ll estimate N(m2b)N(m_{2}^{b}). The key point is that, combining (15) and (14), we have

(16) N(m2b)N(m0)N(m1a)24A(n,2)A(n,1)2.\frac{N(m_{2}^{b})N(m_{0})}{N(m_{1}^{a})^{2}}\leq 4\frac{A(n,2)}{A(n,1)^{2}}.

We will focus on proving a lower bound for the left side, assuming mH(w1)>1m_{H}(w-1)>1, which which will give a contradiction for sufficiently large nn.

By hypothesis, we have

A(n,1)2mH(w1)(mH(w)m0(w))(m0(w)+1).\frac{A(n,1)}{2}\leq\frac{m_{H}(w-1)(m_{H}(w)-m_{0}(w))}{(m_{0}(w)+1)}.

By Lemma B.3 we have mH(w1)n1m_{H}(w-1)\leq n-1. Also, we know m0(w)1m_{0}(w)\geq 1, so

mH(w)m0(w)A(n,1)n1.m_{H}(w)-m_{0}(w)\geq\frac{A(n,1)}{n-1}.

If n6n\geq 6, then A(n,1)n19\frac{A(n,1)}{n-1}\geq 9, which gives the bound

(17) mH(w)m0(w)189(mH(w)m0(w))m_{H}(w)-m_{0}(w)-1\geq\frac{8}{9}(m_{H}(w)-m_{0}(w))

Assuming mH(w1)2m_{H}(w-1)\geq 2, we find

(18) N(m2b)N(m0)N(m1a)2=mH(w1)(mH(w1)1)(mH(w)m0(w))(mH(w)m0(w)1)2(m0(w)+1)(m0(w)+2)(m0(w)+1)2mH2(w1)(mH(w)m0(w))2=12m0(w)+1m0(w)+2mH(w1)1mH(w1)mH(w)m0(w)1mH(w)m0(w)122312mH(w)m0(w)1mH(w)m0(w)=16mH(w)m0(w)1mH(w)m0(w)\displaystyle\begin{split}&\frac{N(m_{2}^{b})N(m_{0})}{N(m_{1}^{a})^{2}}\\ =&\frac{m_{H}(w-1)(m_{H}(w-1)-1)(m_{H}(w)-m_{0}(w))(m_{H}(w)-m_{0}(w)-1)}{2(m_{0}(w)+1)(m_{0}(w)+2)}\cdot\frac{(m_{0}(w)+1)^{2}}{m_{H}^{2}(w-1)(m_{H}(w)-m_{0}(w))^{2}}\\ =&\frac{1}{2}\cdot\frac{m_{0}(w)+1}{m_{0}(w)+2}\cdot\frac{m_{H}(w-1)-1}{m_{H}(w-1)}\cdot\frac{m_{H}(w)-m_{0}(w)-1}{m_{H}(w)-m_{0}(w)}\\ \geq&\frac{1}{2}\cdot\frac{2}{3}\cdot\frac{1}{2}\cdot\frac{m_{H}(w)-m_{0}(w)-1}{m_{H}(w)-m_{0}(w)}=\frac{1}{6}\cdot\frac{m_{H}(w)-m_{0}(w)-1}{m_{H}(w)-m_{0}(w)}\\ \end{split}

By combining this with (17), we conclude that

(19) N(m2b)N(m0)N(m1a)24/27\frac{N(m_{2}^{b})N(m_{0})}{N(m_{1}^{a})^{2}}\geq 4/27

which combines with (16) to give

A(n,2)A(n,1)21/27\frac{A(n,2)}{A(n,1)^{2}}\geq 1/27

which is impossible for n11n\geq 11. (See the discussion in Appendix C, and bound-A3a in the Python code.)

For smaller nn, we do a more precise version of the above analysis. For 5n105\leq n\leq 10, we will improve on the bound (15) by using Lemma B.10. For n=5n=5, we will also need to replace the bound (17) by a slightly weaker one without the assumption n6n\geq 6. With these modifications, the argument will work for nn from 55 to 1010.

To apply Lemma B.10, we must give an upper bound for mH(w1)mH(w+1)m_{H}(w-1)m_{H}(w+1), showing that the two roots α,β\alpha,\beta are far apart. Consider

m4a=m0+2[w+1]2[w1].m_{4}^{a}=m_{0}+2[w+1]-2[w-1].

We have

A(n,4)N(m4a)N(m0)=[mH(w+1)(mH(w+1)1)][mH(w1)(mH(w1)1)]4,A(n,4)\geq\frac{N(m_{4}^{a})}{N(m_{0})}=\frac{\left[m_{H}(w+1)(m_{H}(w+1)-1)\right]\left[m_{H}(w-1)(m_{H}(w-1)-1)\right]}{4},

so (still assuming mH(w1)2m_{H}(w-1)\geq 2) we conclude that

2A(n,4)mH(w+1)(mH(w+1)1).2A(n,4)\geq m_{H}(w+1)(m_{H}(w+1)-1).

For 5n105\leq n\leq 10 we have

mH(w+1)2A(n,4)+1148A(n,1)2n1m_{H}(w+1)\leq\sqrt{2A(n,4)}+1\leq\frac{1}{48}\frac{A(n,1)^{2}}{n-1}

(see bound-A3b in the Python code). Thus we have

mH(w1)mH(w+1)(n1)148A(n,1)2n1=148A(n,1)2,m_{H}(w-1)m_{H}(w+1)\leq(n-1)\cdot\frac{1}{48}\frac{A(n,1)^{2}}{n-1}=\frac{1}{48}A(n,1)^{2},

so by Lemma B.10,

(20) N(m1a)N(m0)4546A(n,1).\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{45}{46}A(n,1).

When 6n106\leq n\leq 10 we have from above (19) that

N(m2b)N(m0)N(m1a)24/27,\frac{N(m_{2}^{b})N(m_{0})}{N(m_{1}^{a})^{2}}\geq 4/27,

so

A(n,2)A(n,1)2427(4546)2,\frac{A(n,2)}{A(n,1)^{2}}\geq\frac{4}{27}\cdot\left(\frac{45}{46}\right)^{2},

which does not hold for any 6n106\leq n\leq 10.

Finally when n=5n=5 we use the estimate

mH(w)m0(w)A(n,1)n1=132m_{H}(w)-m_{0}(w)\geq\frac{A(n,1)}{n-1}=\frac{13}{2}

to deduce

mH(w)m0(w)11113(mH(w)m0(w)).m_{H}(w)-m_{0}(w)-1\geq\frac{11}{13}(m_{H}(w)-m_{0}(w)).

Now (LABEL:eqn-line415) implies

N(m2b)N(m0)N(m1a)2161113;\frac{N(m_{2}^{b})N(m_{0})}{N(m_{1}^{a})^{2}}\geq\frac{1}{6}\cdot\frac{11}{13};

combining this with (20), we get

A(n,2)A(n,1)2161113(4546)2,\frac{A(n,2)}{A(n,1)^{2}}\geq\frac{1}{6}\cdot\frac{11}{13}\cdot\left(\frac{45}{46}\right)^{2},

which is not true for n=5n=5.

Thus we arrive at a contradiction in every case. ∎

Lemma B.12.

If

N(m1a)N(m0)A(n,1)2\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}

then mH(w2)2m_{H}(w-2)\leq 2.

Proof.

Assume mH(w2)3m_{H}(w-2)\geq 3, and consider

m7a=m03[w2][w1]+4[w].m_{7}^{a}=m_{0}-3[w-2]-[w-1]+4[w].

We will show that N(m7a)N(m_{7}^{a}) is too big.

First, note that we must have n8n\geq 8 by Lemma B.3.

By hypothesis, we have N(m1a)N(m0)A(n,1)2\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}. From Lemma B.11, we know mH(w1)=1m_{H}(w-1)=1, so

N(m1a)N(m0)=(mH(w)m0(w))(m0(w)+1)\frac{N(m_{1}^{a})}{N(m_{0})}=\frac{(m_{H}(w)-m_{0}(w))}{(m_{0}(w)+1)}

and thus

(mH(w)m0(w))A(n,1).(m_{H}(w)-m_{0}(w))\geq A(n,1).

Since n8n\geq 8, this gives

(mH(w)m0(w))247.(m_{H}(w)-m_{0}(w))\geq 247.

Now

N(m7a)N(m0)=Ci=14(mH(w)m0(w)i+1)(m0(w)+i),\frac{N(m_{7}^{a})}{N(m_{0})}=C\prod_{i=1}^{4}\frac{(m_{H}(w)-m_{0}(w)-i+1)}{(m_{0}(w)+i)},

where

C=mH(w2)(mH(w2)1)(mH(w2)2)(mH(w1))61.C=\frac{m_{H}(w-2)(m_{H}(w-2)-1)(m_{H}(w-2)-2)(m_{H}(w-1))}{6}\geq 1.

Using the bounds (mH(w)m0(w)3)>.98(mH(w)m0(w))(m_{H}(w)-m_{0}(w)-3)>.98(m_{H}(w)-m_{0}(w)) and

(m0(w)+1)4(m0(w)+1)(m0(w)+2)(m0(w)+3)(m0(w)+4)215,\frac{(m_{0}(w)+1)^{4}}{(m_{0}(w)+1)(m_{0}(w)+2)(m_{0}(w)+3)(m_{0}(w)+4)}\geq\frac{2}{15},

we deduce that

N(m7a)N(m0).92A(n,1)4120A(n,1)4140.\frac{N(m_{7}^{a})}{N(m_{0})}\geq\frac{.92A(n,1)^{4}}{120}\geq\frac{A(n,1)^{4}}{140}.

The inequality

A(n,7)<1140A(n,1)4A(n,7)<\frac{1}{140}A(n,1)^{4}

(see Appendix C, and bound-A3c in the Python code) gives a contradiction. ∎

Lemma B.13.

If n5n\geq 5 and

N(m1a)N(m0)A(n,1)2\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}

then

mH(w+1)(mH(w+1)1)4811A(n,3)N(m1a)N(m0)4811A(n,3)A(n,1).m_{H}(w+1)(m_{H}(w+1)-1)\leq\frac{48}{11}A(n,3)\frac{N(m_{1}^{a})}{N(m_{0})}\leq\frac{48}{11}A(n,3)A(n,1).

Furthermore, if m0(w)2m_{0}(w)\geq 2 then we have the stronger bound

mH(w+1)(mH(w+1)1)3611A(n,3)N(m1a)N(m0)3611A(n,3)A(n,1).m_{H}(w+1)(m_{H}(w+1)-1)\leq\frac{36}{11}A(n,3)\frac{N(m_{1}^{a})}{N(m_{0})}\leq\frac{36}{11}A(n,3)A(n,1).
Proof.

Let

m3a=m0+2[w+1][w][w1].m_{3}^{a}=m_{0}+2[w+1]-[w]-[w-1].

The result will follow from

N(m3a)N(m0)A(n,3).\frac{N(m_{3}^{a})}{N(m_{0})}\leq A(n,3).

We have

A(n,3)N(m3a)N(m0)=mH(w+1)(mH(w+1)1)mH(w1)m0(w)2(mH(w)m0(w)+1).A(n,3)\geq\frac{N(m_{3}^{a})}{N(m_{0})}=\frac{m_{H}(w+1)(m_{H}(w+1)-1)m_{H}(w-1)m_{0}(w)}{2(m_{H}(w)-m_{0}(w)+1)}.

On the other hand,

mH(w1)(mH(w)m0(w))(m0(w)+1)=N(m1a)N(m0)A(n,1).\frac{m_{H}(w-1)(m_{H}(w)-m_{0}(w))}{(m_{0}(w)+1)}=\frac{N(m_{1}^{a})}{N(m_{0})}\leq A(n,1).

We know mH(w1)=1m_{H}(w-1)=1 by Lemma B.11, so mH(w)m0(w)A(n,1)11m_{H}(w)-m_{0}(w)\geq A(n,1)\geq 11. Thus we can estimate

m0(w)2(mH(w)m0(w)+1)(mH(w)m0(w))(m0(w)+1)141112=1148.\frac{m_{0}(w)}{2(m_{H}(w)-m_{0}(w)+1)}\cdot\frac{(m_{H}(w)-m_{0}(w))}{(m_{0}(w)+1)}\geq\frac{1}{4}\cdot\frac{11}{12}=\frac{11}{48}.

The first bound follows.

To get the second bound, note that if m0(w)2m_{0}(w)\geq 2 then

m0(w)m0(w)+123,\frac{m_{0}(w)}{m_{0}(w)+1}\geq\frac{2}{3},

so

m0(w)2(mH(w)m0(w)+1)(mH(w)m0(w))(m0(w)+1)131112=1136.\frac{m_{0}(w)}{2(m_{H}(w)-m_{0}(w)+1)}\cdot\frac{(m_{H}(w)-m_{0}(w))}{(m_{0}(w)+1)}\geq\frac{1}{3}\cdot\frac{11}{12}=\frac{11}{36}.

Lemma B.14.

If n5n\geq 5 and

N(m1a)N(m0)A(n,1)2\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}

then

mH(w+2)(mH(w+2)1)4811A(n,5)N(m1a)N(m0)4811A(n,5)A(n,1).m_{H}(w+2)(m_{H}(w+2)-1)\leq\frac{48}{11}A(n,5)\frac{N(m_{1}^{a})}{N(m_{0})}\leq\frac{48}{11}A(n,5)A(n,1).

Furthermore, if m0(w)2m_{0}(w)\geq 2 then we have the stronger bound

mH(w+2)(mH(w+2)1)3611A(n,5)N(m1a)N(m0)3611A(n,5)A(n,1).m_{H}(w+2)(m_{H}(w+2)-1)\leq\frac{36}{11}A(n,5)\frac{N(m_{1}^{a})}{N(m_{0})}\leq\frac{36}{11}A(n,5)A(n,1).
Proof.

Let

m5a=m0+2[w+2][w][w1].m_{5}^{a}=m_{0}+2[w+2]-[w]-[w-1].

The result follows from

N(m5a)N(m0)A(n,5);\frac{N(m_{5}^{a})}{N(m_{0})}\leq A(n,5);

the proof is exactly analogous to the proof of Lemma B.13. ∎

Lemma B.15.

If

N(m1a)N(m0)A(n,1)2\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}

then n10n\leq 10.

Proof.

Assume n5n\geq 5. We’ll estimate each N(m2)N(m0)\frac{N(m_{2}^{*})}{N(m_{0})}, and show that for large nn their sum is too small.

We have the following bounds. By Lemma B.12, we have mH(w2)2m_{H}(w-2)\leq 2, so

N(m2a)N(m0)2N(m1a)N(m0)2A(n,1).\frac{N(m_{2}^{a})}{N(m_{0})}\leq 2\frac{N(m_{1}^{a})}{N(m_{0})}\leq 2A(n,1).

Lemma B.11 tells us that mH(w1)=1m_{H}(w-1)=1 (if n5n\geq 5), so

N(m2b)N(m0)=0.\frac{N(m_{2}^{b})}{N(m_{0})}=0.

By Lemma B.13, we know that

mH(w+1)(mH(w+1)1)4811A(n,3)N(m1a)N(m0)4811A(n,3)A(n,1),m_{H}(w+1)(m_{H}(w+1)-1)\leq\frac{48}{11}A(n,3)\frac{N(m_{1}^{a})}{N(m_{0})}\leq\frac{48}{11}A(n,3)A(n,1),

so

N(m2c)N(m0)=mH(w1)mH(w+1)=mH(w+1)4811A(n,3)A(n,1)+1\frac{N(m_{2}^{c})}{N(m_{0})}=m_{H}(w-1)m_{H}(w+1)=m_{H}(w+1)\leq\sqrt{\frac{48}{11}A(n,3)A(n,1)}+1

and

N(m2d)N(m0)\displaystyle\frac{N(m_{2}^{d})}{N(m_{0})} =\displaystyle= (mH(w+1))(mH(w+1)1)(m0(w))(m0(w)1)2(mH(w)m0(w)+1)(mH(w)m0(w)+2)\displaystyle\frac{(m_{H}(w+1))(m_{H}(w+1)-1)(m_{0}(w))(m_{0}(w)-1)}{2(m_{H}(w)-m_{0}(w)+1)(m_{H}(w)-m_{0}(w)+2)}
\displaystyle\leq 12(4811A(n,3)N(m1a)N(m0))(N(m1a)N(m0))2\displaystyle\frac{1}{2}\left(\frac{48}{11}A(n,3)\frac{N(m_{1}^{a})}{N(m_{0})}\right)\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)^{-2}
\displaystyle\leq 4811A(n,3)A(n,1).\displaystyle\frac{48}{11}\frac{A(n,3)}{A(n,1)}.

Finally, from Lemma B.14, we find

mH(w+2)4811A(n,5)N(m1a)N(m0)+1m_{H}(w+2)\leq\sqrt{\frac{48}{11}A(n,5)\frac{N(m_{1}^{a})}{N(m_{0})}}+1

so

N(m2e)N(m0)\displaystyle\frac{N(m_{2}^{e})}{N(m_{0})} =\displaystyle= m0(w)mH(w+2)mH(w)m0(w)+1\displaystyle\frac{m_{0}(w)m_{H}(w+2)}{m_{H}(w)-m_{0}(w)+1}
\displaystyle\leq (N(m1a)N(m0))1(4811A(n,5)N(m1a)N(m0)+1)\displaystyle\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)^{-1}\left(\sqrt{\frac{48}{11}A(n,5)\frac{N(m_{1}^{a})}{N(m_{0})}}+1\right)
\displaystyle\leq 9611A(n,5)A(n,1)+1.\displaystyle\sqrt{\frac{96}{11}\frac{A(n,5)}{A(n,1)}}+1.

These five bounds combine (see Appendix C, and bound-A3d in the Python code) to give

N(m2a)N(m0)+N(m2b)N(m0)+N(m2c)N(m0)+N(m2d)N(m0)+N(m2e)N(m0)<A(n,2)\frac{N(m_{2}^{a})}{N(m_{0})}+\frac{N(m_{2}^{b})}{N(m_{0})}+\frac{N(m_{2}^{c})}{N(m_{0})}+\frac{N(m_{2}^{d})}{N(m_{0})}+\frac{N(m_{2}^{e})}{N(m_{0})}<A(n,2)

for n11n\geq 11, a contradiction. ∎

Lemma B.16.

If 5n195\leq n\leq 19 and

N(m1a)N(m0)A(n,1)2\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}

then m0(w1)=1m_{0}(w-1)=1 and m0(ws)=0m_{0}(w-s)=0 for s2s\geq 2.

Proof.

We have already seen in Lemma B.11 that m0(w1)=1m_{0}(w-1)=1. Suppose for a contradiction that m0(ws)>0m_{0}(w-s)>0 for some s2s\geq 2. Let

ms+3=m0+2[w+1][w1][ws].m_{s+3}=m_{0}+2[w+1]-[w-1]-[w-s].

Then

n!A(n,s+3)N(ms+3)N(m0)mH(w+1)(mH(w+1)1)2,n!\geq A(n,s+3)\geq\frac{N(m_{s+3})}{N(m_{0})}\geq\frac{m_{H}(w+1)(m_{H}(w+1)-1)}{2},

so

mH(w+1)2n!+1.m_{H}(w+1)\leq\sqrt{2n!}+1.

We will use this stronger bound to redo the estimates in Lemma B.15.

We have

N(m2c)N(m0)=mH(w+1)2n!+1\frac{N(m_{2}^{c})}{N(m_{0})}=m_{H}(w+1)\leq\sqrt{2n!}+1

and

N(m2d)N(m0)\displaystyle\frac{N(m_{2}^{d})}{N(m_{0})} =\displaystyle= (mH(w+1))(mH(w+1)1)(m0(w))(m0(w)1)2(mH(w)m0(w)+1)(mH(w)m0(w)+2)\displaystyle\frac{(m_{H}(w+1))(m_{H}(w+1)-1)(m_{0}(w))(m_{0}(w)-1)}{2(m_{H}(w)-m_{0}(w)+1)(m_{H}(w)-m_{0}(w)+2)}
\displaystyle\leq 12mH(w+1)2(N(m1a)N(m0))2\displaystyle\frac{1}{2}m_{H}(w+1)^{2}\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)^{-2}
\displaystyle\leq 12(2n!+1)2(A(n,1)2)2.\displaystyle\frac{1}{2}\left(\sqrt{2n!}+1\right)^{2}\left(\frac{A(n,1)}{2}\right)^{-2}.

We use the same bounds on

N(m2a)N(m0),N(m2b)N(m0),N(m2e)N(m0)\frac{N(m_{2}^{a})}{N(m_{0})},\frac{N(m_{2}^{b})}{N(m_{0})},\frac{N(m_{2}^{e})}{N(m_{0})}

as in Lemma B.15; we conclude (see bound-A3e in the Python code) that

N(m2a)N(m0)+N(m2b)N(m0)+N(m2c)N(m0)+N(m2d)N(m0)+N(m2e)N(m0)<A(n,2)\frac{N(m_{2}^{a})}{N(m_{0})}+\frac{N(m_{2}^{b})}{N(m_{0})}+\frac{N(m_{2}^{c})}{N(m_{0})}+\frac{N(m_{2}^{d})}{N(m_{0})}+\frac{N(m_{2}^{e})}{N(m_{0})}<A(n,2)

for 6n196\leq n\leq 19.

Finally, suppose n=5n=5. Then since s+34s+3\geq 4, we have the much stronger bound A(n,s+3)1A(n,s+3)\leq 1, which implies mH(w+1)2m_{H}(w+1)\leq 2. We deduce as above

N(m2c)N(m0)=mH(w+1)2\frac{N(m_{2}^{c})}{N(m_{0})}=m_{H}(w+1)\leq 2
N(m2d)N(m0)(N(m1a)N(m0))2<1\frac{N(m_{2}^{d})}{N(m_{0})}\leq\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)^{-2}<1

and arrive at a contradiction as before. (See bound-A3f in the Python code.) ∎

Lemma B.17.

If n5n\geq 5 and

N(m1a)N(m0)A(n,1)2\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}

then m0(w)=1m_{0}(w)=1.

Proof.

Suppose m0(w)2m_{0}(w)\geq 2. Again, we’ll redo the estimates in Lemma B.15.

Since n5n\geq 5 and (by Lemma B.15) n10n\leq 10, Lemma B.16 implies that

N(m2a)N(m0)=N(m2b)N(m0)=0.\frac{N(m_{2}^{a})}{N(m_{0})}=\frac{N(m_{2}^{b})}{N(m_{0})}=0.

Lemmas B.13 and B.14 give stronger bounds in case m0(w)2m_{0}(w)\geq 2. As in Lemma B.15, we deduce in this case that:

N(m2c)N(m0)3611A(n,3)A(n,1)+1\frac{N(m_{2}^{c})}{N(m_{0})}\leq\sqrt{\frac{36}{11}A(n,3)A(n,1)}+1
N(m2d)N(m0)3611A(n,3)A(n,1)\frac{N(m_{2}^{d})}{N(m_{0})}\leq\frac{36}{11}\frac{A(n,3)}{A(n,1)}
N(m2e)N(m0)7211A(n,5)A(n,1)+1.\frac{N(m_{2}^{e})}{N(m_{0})}\leq\sqrt{\frac{72}{11}\frac{A(n,5)}{A(n,1)}}+1.

Again, we conclude for nn in the given range that

N(m2a)N(m0)+N(m2b)N(m0)+N(m2c)N(m0)+N(m2d)N(m0)+N(m2e)N(m0)<A(n,2),\frac{N(m_{2}^{a})}{N(m_{0})}+\frac{N(m_{2}^{b})}{N(m_{0})}+\frac{N(m_{2}^{c})}{N(m_{0})}+\frac{N(m_{2}^{d})}{N(m_{0})}+\frac{N(m_{2}^{e})}{N(m_{0})}<A(n,2),

a contradiction. (See Appendix C, and bound-A3g in the Python code.) ∎

Lemma B.18.

We cannot have n5n\geq 5 and

N(m1a)N(m0)A(n,1)2.\frac{N(m_{1}^{a})}{N(m_{0})}\geq\frac{A(n,1)}{2}.
Proof.

Again, we’ll redo the estimates in Lemma B.15, in light of everything we now know. By Lemma B.15, we may assume that n11n\leq 11.

Now Lemma B.16 implies that

N(m2a)N(m0)=N(m2b)N(m0)=0,\frac{N(m_{2}^{a})}{N(m_{0})}=\frac{N(m_{2}^{b})}{N(m_{0})}=0,

while Lemma B.17 gives us

N(m2d)N(m0)=0.\frac{N(m_{2}^{d})}{N(m_{0})}=0.

Yet again (see proof of Lemma B.15) we have the bounds

N(m2c)N(m0)4811A(n,3)A(n,1)+1\frac{N(m_{2}^{c})}{N(m_{0})}\leq\sqrt{\frac{48}{11}A(n,3)A(n,1)}+1

and

N(m2e)N(m0)9611A(n,5)A(n,1)+1.\frac{N(m_{2}^{e})}{N(m_{0})}\leq\sqrt{\frac{96}{11}\frac{A(n,5)}{A(n,1)}}+1.

Yet again, we conclude that

N(m2a)N(m0)+N(m2b)N(m0)+N(m2c)N(m0)+N(m2d)N(m0)+N(m2e)N(m0)<A(n,2),\frac{N(m_{2}^{a})}{N(m_{0})}+\frac{N(m_{2}^{b})}{N(m_{0})}+\frac{N(m_{2}^{c})}{N(m_{0})}+\frac{N(m_{2}^{d})}{N(m_{0})}+\frac{N(m_{2}^{e})}{N(m_{0})}<A(n,2),

a contradiction. (See Appendix C, and bound-A3h in the Python code.) ∎

B.4. The case n5n\geq 5, with knk\geq n

Recall notation from the beginning of Section B, and the beginning of Section B.2.

We’ll treat the case where knk\geq n next. By Lemma B.3, if knk\geq n, then w=ww=w^{\prime}.

Lemma B.19.

If knk\geq n and n5n\geq 5 then one of the two ratios

N(m1a)N(m0),N(m1b)N(m0)\frac{N(m_{1}^{a})}{N(m_{0})},\frac{N(m_{1}^{b})}{N(m_{0})}

is less than 1, and the other is greater than A(n,1)1A(n,1)-1.

Proof.

From Lemma B.3, we have

mH(w1)+mH(w+1)n1m_{H}(w-1)+m_{H}(w+1)\leq n-1

so

mH(w1)mH(w+1)(n1)22.m_{H}(w-1)m_{H}(w+1)\leq\frac{(n-1)^{2}}{2}.

Then apply Lemma B.10, and the inequality

(n1)22<A(n,1)1\frac{(n-1)^{2}}{2}<A(n,1)-1

(see Appendix C). ∎

Lemma B.20.

If n5n\geq 5 and

N(m1a)N(m0)>A(n,1)1\frac{N(m_{1}^{a})}{N(m_{0})}>A(n,1)-1

then mH(w1)1m_{H}(w-1)\leq 1.

If n5n\geq 5 and

N(m1b)N(m0)>A(n,1)1\frac{N(m_{1}^{b})}{N(m_{0})}>A(n,1)-1

then mH(w+1)1m_{H}(w+1)\leq 1.

Proof.

The first case follows from Lemma B.11; we’ll prove the second. (As an alternative to Lemma B.11, the first case could be proven by an argument analagous to the argument below.)

So suppose N(m1b)N(m0)>A(n,1)1\frac{N(m_{1}^{b})}{N(m_{0})}>A(n,1)-1 and mH(w+1)2m_{H}(w+1)\geq 2, and consider

N(m2d)N(m1b)=(mH(w+1)1)(m0(w)1)2(mH(w)m0(w)+2).\frac{N(m_{2}^{d})}{N(m_{1}^{b})}=\frac{(m_{H}(w+1)-1)(m_{0}(w)-1)}{2(m_{H}(w)-m_{0}(w)+2)}.
N(m2d)N(m0)N(m1b)2=12(mH(w+1)1)mH(w+1)(mH(w)m0(w)+1)(mH(w)m0(w)+2)(m0(w)1)m0(w).\frac{N(m_{2}^{d})N(m_{0})}{N(m_{1}^{b})^{2}}=\frac{1}{2}\cdot\frac{(m_{H}(w+1)-1)}{m_{H}(w+1)}\cdot\frac{(m_{H}(w)-m_{0}(w)+1)}{(m_{H}(w)-m_{0}(w)+2)}\cdot\frac{(m_{0}(w)-1)}{m_{0}(w)}.

From

(m0(w))(mH(w)m0(w)+1)=1mH(w+1)N(m1b)N(m0)A(n,1)1n1>6\frac{(m_{0}(w))}{(m_{H}(w)-m_{0}(w)+1)}=\frac{1}{m_{H}(w+1)}\cdot\frac{N(m_{1}^{b})}{N(m_{0})}\geq\frac{A(n,1)-1}{n-1}>6

we deduce that m0(w)>6m_{0}(w)>6, so by integrality m0(w)7m_{0}(w)\geq 7, so

m0(w)167m0(w).m_{0}(w)-1\geq\frac{6}{7}m_{0}(w).

We also know that (mH(w)m0(w)+2)2(mH(w)m0(w)+1)(m_{H}(w)-m_{0}(w)+2)\leq 2(m_{H}(w)-m_{0}(w)+1) and (mH(w+1)1)12mH(w+1)(m_{H}(w+1)-1)\geq\frac{1}{2}m_{H}(w+1), so

A(n,2)(A(n,1)1)2N(m2d)N(m0)N(m1b)212121267=328.\frac{A(n,2)}{(A(n,1)-1)^{2}}\geq\frac{N(m_{2}^{d})N(m_{0})}{N(m_{1}^{b})^{2}}\geq\frac{1}{2}\cdot\frac{1}{2}\cdot\frac{1}{2}\cdot\frac{6}{7}=\frac{3}{28}.

For n5n\geq 5, this contradicts

A(n,2)<328(A(n,1)1)2.A(n,2)<\frac{3}{28}(A(n,1)-1)^{2}.

(See Appendix C and bound-A4a in the Python code.) ∎

Lemma B.21.

For n5n\geq 5 we cannot have knk\geq n.

Proof.

Assume knk\geq n.

We’ll consider the two cases given in Lemma B.19. The first case is ruled out by Lemma B.18; we’ll prove the second. (Alternatively, the first case could be proven by an argument analagous to the argument below.)

So, suppose

N(m1b)N(m0)>A(n,1)1.\frac{N(m_{1}^{b})}{N(m_{0})}>A(n,1)-1.

From

m0(w)mH(w)m0(w)+1>A(n,1)1\frac{m_{0}(w)}{m_{H}(w)-m_{0}(w)+1}>A(n,1)-1

and the bounds

mH(w2),mH(w1)n1<A(n,1)1m_{H}(w-2),m_{H}(w-1)\leq n-1<A(n,1)-1

we deduce that

N(m2a)N(m0)<mH(w2)(m0(w)mH(w)m0(w)+1)1<1\frac{N(m_{2}^{a})}{N(m_{0})}<m_{H}(w-2)\left(\frac{m_{0}(w)}{m_{H}(w)-m_{0}(w)+1}\right)^{-1}<1

and

N(m2b)N(m0)<12mH(w1)2(m0(w)mH(w)m0(w)+1)2<1.\frac{N(m_{2}^{b})}{N(m_{0})}<\frac{1}{2}m_{H}(w-1)^{2}\left(\frac{m_{0}(w)}{m_{H}(w)-m_{0}(w)+1}\right)^{-2}<1.

By Lemma B.20, we have mH(w+1)1m_{H}(w+1)\leq 1, so

N(m2c)N(m0)=mH(w1)mH(w+1)n1.\frac{N(m_{2}^{c})}{N(m_{0})}=m_{H}(w-1)m_{H}(w+1)\leq n-1.

Also, mH(w+1)1m_{H}(w+1)\leq 1 implies that

N(m2d)N(m0)=0.\frac{N(m_{2}^{d})}{N(m_{0})}=0.

Finally, Lemma B.3 gives mH(w+2)n12m_{H}(w+2)\leq\frac{n-1}{2}, so

N(m2e)N(m0)(n1)2A(n,1).\frac{N(m_{2}^{e})}{N(m_{0})}\leq\frac{(n-1)}{2}A(n,1).

Adding these bounds, we see that

A(n,2)==a,b,c,d,eN(m2)N(m0)n+1+(n1)2A(n,1).A(n,2)=\sum_{*=a,b,c,d,e}\frac{N(m_{2}^{*})}{N(m_{0})}\leq n+1+\frac{(n-1)}{2}A(n,1).

But this contradicts the inequality

n+1+(n1)2A(n,1)<A(n,2),n+1+\frac{(n-1)}{2}A(n,1)<A(n,2),

which holds for n5n\geq 5 (see Appendix C and bound-A4b in the Python code). ∎

B.5. The case n5n\geq 5, with N(m1b)N(m_{1}^{b}) big and k<nk<n.

Recall notation from the beginning of Section B, and the beginning of Section B.2.

Lemma B.22.

Suppose kn1k\leq n-1 and

N(m1b)N(m0)A(n,1)/2.\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2.

Then

mH(w+1)A(n,1)2(n1).m_{H}(w+1)\geq\frac{A(n,1)}{2(n-1)}.

In particular, if n5n\geq 5 then

mH(w+1)4.m_{H}(w+1)\geq 4.
Proof.

We have

mH(w+1)N(m1b)N(m0)1m0(w)A(n,1)2m0(w)A(n,1)2(n1).m_{H}(w+1)\geq\frac{N(m_{1}^{b})}{N(m_{0})}\cdot\frac{1}{m_{0}(w)}\geq\frac{A(n,1)}{2m_{0}(w)}\geq\frac{A(n,1)}{2(n-1)}.

The “in particular” follows from the fact that mH(w+1)m_{H}(w+1) is an integer. ∎

Lemma B.23.

Suppose n5n\geq 5, kn1k\leq n-1, and

N(m1b)N(m0)A(n,1)/2.\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2.

Then we cannot have simultaneously N(m1a)N(m0)>0\frac{N(m_{1}^{a})}{N(m_{0})}>0 and m0(w)2m_{0}(w)\geq 2.

Proof.

Suppose N(m1a)N(m0)>0\frac{N(m_{1}^{a})}{N(m_{0})}>0 and m0(w)2m_{0}(w)\geq 2; the first condition implies mH(w1)1m_{H}(w-1)\geq 1 and mH(w)m0(w)1m_{H}(w)-m_{0}(w)\geq 1. Consider

m4b=m0[w1]2[w]+3[w+1].m_{4}^{b}=m_{0}-[w-1]-2[w]+3[w+1].

We have

N(m4b)N(m1b)=mH(w1)(mH(w+1)1)(mH(w+1)2)(m0(w)1)6(mH(w)m0(w)+2).\frac{N(m_{4}^{b})}{N(m_{1}^{b})}=\frac{m_{H}(w-1)(m_{H}(w+1)-1)(m_{H}(w+1)-2)(m_{0}(w)-1)}{6(m_{H}(w)-m_{0}(w)+2)}.

By Lemma B.10 we have

(N(m1a)N(m0))(N(m1b)N(m0))<mH(w1)mH(w+1).\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)\left(\frac{N(m_{1}^{b})}{N(m_{0})}\right)<m_{H}(w-1)m_{H}(w+1).

Thus we obtain

N(m4b)N(m1b)>16(N(m1a)N(m0))(N(m1b)N(m0))2(mH(w+1)1)(mH(w+1)2)mH(w+1)2mH(w)m0(w)+1mH(w)m0(w)+2m0(w)1m0(w).\frac{N(m_{4}^{b})}{N(m_{1}^{b})}>\frac{1}{6}\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)\left(\frac{N(m_{1}^{b})}{N(m_{0})}\right)^{2}\frac{(m_{H}(w+1)-1)(m_{H}(w+1)-2)}{m_{H}(w+1)^{2}}\frac{m_{H}(w)-m_{0}(w)+1}{m_{H}(w)-m_{0}(w)+2}\frac{m_{0}(w)-1}{m_{0}(w)}.

Using mH(w+1)4m_{H}(w+1)\geq 4, mH(w)m0(w)1m_{H}(w)-m_{0}(w)\geq 1, and m0(w)2m_{0}(w)\geq 2, the three fractional factors on the right can be bounded below by 38\frac{3}{8}, 23\frac{2}{3}, and 12\frac{1}{2}, respectively.

On the other hand, as soon as N(m1a)N(m_{1}^{a}) is nonzero, we have

N(m1a)N(m0)=mH(w1)(mH(w)m0(w))m0(w)+11n.\frac{N(m_{1}^{a})}{N(m_{0})}=\frac{m_{H}(w-1)(m_{H}(w)-m_{0}(w))}{m_{0}(w)+1}\geq\frac{1}{n}.

We conclude that

N(m4b)N(m1b)>16(N(m1a)N(m0))(N(m1b)N(m0))2382312=148(N(m1a)N(m0))(A(n,1)N(m1a)N(m0))2\frac{N(m_{4}^{b})}{N(m_{1}^{b})}>\frac{1}{6}\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)\left(\frac{N(m_{1}^{b})}{N(m_{0})}\right)^{2}\frac{3}{8}\frac{2}{3}\frac{1}{2}=\frac{1}{48}\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)\left(A(n,1)-\frac{N(m_{1}^{a})}{N(m_{0})}\right)^{2}

so

N(m4b)N(m0)>148(N(m1a)N(m0))(A(n,1)N(m1a)N(m0))3\frac{N(m_{4}^{b})}{N(m_{0})}>\frac{1}{48}\left(\frac{N(m_{1}^{a})}{N(m_{0})}\right)\left(A(n,1)-\frac{N(m_{1}^{a})}{N(m_{0})}\right)^{3}

As a function of N(m1a)N(m0)\frac{N(m_{1}^{a})}{N(m_{0})}, this right-hand side is minimized when N(m1a)N(m0)=1n\frac{N(m_{1}^{a})}{N(m_{0})}=\frac{1}{n}, so we have

A(n,4)N(m4b)N(m1b)>148n(A(n,1)1n)3.A(n,4)\geq\frac{N(m_{4}^{b})}{N(m_{1}^{b})}>\frac{1}{48n}\left(A(n,1)-\frac{1}{n}\right)^{3}.

This contradicts the inequality

A(n,4)<178n(A(n,1)1)3,A(n,4)<\frac{1}{78n}(A(n,1)-1)^{3},

which is valid for all nn. (See Appendix C and bound-A5a in the Python code.) ∎

Lemma B.24.

If n5n\geq 5, N(m1a)N(m0)=0\frac{N(m_{1}^{a})}{N(m_{0})}=0, kn1k\leq n-1, and

N(m1b)N(m0)A(n,1)/2,\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2,

then m0(w)=1m_{0}(w)=1 and N(m2d)=0N(m_{2}^{d})=0.

Proof.

Since N(m1a)=0N(m_{1}^{a})=0, we have

N(m1b)N(m0)=A(n,1).\frac{N(m_{1}^{b})}{N(m_{0})}=A(n,1).

Assuming m0(w)2m_{0}(w)\geq 2, let’s look at m2dm_{2}^{d}. We have

N(m2d)N(m0)=(mH(w+1))(mH(w+1)1)(m0(w))(m0(w)1)2(mH(w)m0(w)+1)(mH(w)m0(w)+2).\frac{N(m_{2}^{d})}{N(m_{0})}=\frac{(m_{H}(w+1))(m_{H}(w+1)-1)(m_{0}(w))(m_{0}(w)-1)}{2(m_{H}(w)-m_{0}(w)+1)(m_{H}(w)-m_{0}(w)+2)}.

We want to compare this with the inequality

N(m2d)N(m0)A(n,2)<110A(n,1)2,\frac{N(m_{2}^{d})}{N(m_{0})}\leq A(n,2)<\frac{1}{10}A(n,1)^{2},

which is valid for all nn. (See Appendix C and bound-A5b in the Python code.)

We have certainly

N(m2d)N(m0)18(mH(w+1)1)mH(w+1)(N(m1b)N(m0))2=18(mH(w+1)1)mH(w+1)A(n,1)2.\frac{N(m_{2}^{d})}{N(m_{0})}\geq\frac{1}{8}\cdot\frac{(m_{H}(w+1)-1)}{m_{H}(w+1)}\left(\frac{N(m_{1}^{b})}{N(m_{0})}\right)^{2}=\frac{1}{8}\cdot\frac{(m_{H}(w+1)-1)}{m_{H}(w+1)}A(n,1)^{2}.

So we will be done if mH(w+1)5m_{H}(w+1)\geq 5.

But as in the proof of Lemma B.22, we find (using now N(m1b)N(m0)=A(n,1)\frac{N(m_{1}^{b})}{N(m_{0})}=A(n,1) in place of a weaker bound) that

mH(w+1)A(n,1)(n1),m_{H}(w+1)\geq\frac{A(n,1)}{(n-1)},

and so in particular mH(w+1)>5m_{H}(w+1)>5 whenever n5n\geq 5. ∎

Lemma B.25.

If n5n\geq 5, kn1k\leq n-1, and

N(m1b)N(m0)A(n,1)/2,\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2,

then m0(w)=1m_{0}(w)=1.

In particular, there must be some r>0r>0 with m0(wr)>0m_{0}(w-r)>0.

Proof.

That m0(w)=1m_{0}(w)=1 follows from Lemmas B.23 and B.24; the second claim follows because k>1k>1. ∎

Let wrw-r be the highest Hodge weight below ww. In other words, take r>0r>0 minimal such that mH(wr)>0m_{H}(w-r)>0. (Such rr exists by Lemma B.25.) For example, if N(m1a)0N(m_{1}^{a})\neq 0 then we must have mH(w1)0m_{H}(w-1)\neq 0, so r=1r=1.

B.5.1. Case: m1bm_{1}^{b} big and r=1r=1.

Recall the definitions of m2m_{2}^{*} from above (=a,b,c,d,e*=a,b,c,d,e).

Lemma B.26.

If kn1k\leq n-1,

N(m1b)N(m0)A(n,1)/2,\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2,

and m0(w)=1m_{0}(w)=1, then m0(w1)1m_{0}(w-1)\leq 1.

Proof.

Suppose for a contradiction m0(w1)2m_{0}(w-1)\geq 2, and consider

m5b=m0+3[w+1][w]2[w1].m_{5}^{b}=m_{0}+3[w+1]-[w]-2[w-1].

We have

N(m5b)N(m1b)\displaystyle\frac{N(m_{5}^{b})}{N(m_{1}^{b})} =\displaystyle= (mH(w1)(mH(w1)1)2)((mH(w+1)1)(mH(w+1)2)6)\displaystyle\left(\frac{m_{H}(w-1)(m_{H}(w-1)-1)}{2}\right)\left(\frac{(m_{H}(w+1)-1)(m_{H}(w+1)-2)}{6}\right)
\displaystyle\geq (mH(w+1)1)(mH(w+1)2)6.\displaystyle\frac{(m_{H}(w+1)-1)(m_{H}(w+1)-2)}{6}.

From

N(m1b)N(m0)=mH(w+1)m0(w)mH(w)m0(w)+1\frac{N(m_{1}^{b})}{N(m_{0})}=\frac{m_{H}(w+1)m_{0}(w)}{m_{H}(w)-m_{0}(w)+1}

and m0(w)=1m_{0}(w)=1, we find that

mH(w+1)A(n,1)/2.m_{H}(w+1)\geq A(n,1)/2.

In particular, since n5n\geq 5, we have mH(w+1)13m_{H}(w+1)\geq 13, so certainly

(mH(w+1)1)(mH(w+1)2)12mH(w+1)2.(m_{H}(w+1)-1)(m_{H}(w+1)-2)\geq\frac{1}{2}m_{H}(w+1)^{2}.

We conclude that

N(m5b)N(m0)=(N(m5b)N(m1b))(N(m1b)N(m0))112mH(w+1)2A(n,1)2A(n,1)396.\frac{N(m_{5}^{b})}{N(m_{0})}=\left(\frac{N(m_{5}^{b})}{N(m_{1}^{b})}\right)\left(\frac{N(m_{1}^{b})}{N(m_{0})}\right)\geq\frac{1}{12}m_{H}(w+1)^{2}\frac{A(n,1)}{2}\geq\frac{A(n,1)^{3}}{96}.

This contradicts the inequality

A(n,1)3>341A(n,5),A(n,1)^{3}>341A(n,5),

which is valid for all nn. (See Appendix C and bound-A5c in the Python code.) ∎

Lemma B.27.

We cannot have n5n\geq 5, kn1k\leq n-1,

N(m1b)N(m0)A(n,1)/2,\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2,

and r=1r=1.

Proof.

Assume the stated conditions hold. We will bound the five ratios N(m2)N(m0)\frac{N(m_{2}^{*})}{N(m_{0})}, and show that their sum is too small. Lemmas B.25 and B.26 tell us that m0(w1)=m0(w)=1m_{0}(w-1)=m_{0}(w)=1. Hence, N(m2b)=N(m2d)=0N(m_{2}^{b})=N(m_{2}^{d})=0.

We have

N(m2a)N(m0)=m0(w2)(mH(w)1)2\frac{N(m_{2}^{a})}{N(m_{0})}=\frac{m_{0}(w-2)(m_{H}(w)-1)}{2}
N(m2c)N(m0)=mH(w+1)\frac{N(m_{2}^{c})}{N(m_{0})}=m_{H}(w+1)
N(m2e)N(m0)=mH(w+2)mH(w).\frac{N(m_{2}^{e})}{N(m_{0})}=\frac{m_{H}(w+2)}{m_{H}(w)}.

Since

A(n,1)2N(m1a)N(m0)=mH(w)12\frac{A(n,1)}{2}\geq\frac{N(m_{1}^{a})}{N(m_{0})}=\frac{m_{H}(w)-1}{2}

and m0(w2)n1m_{0}(w-2)\leq n-1, we have

N(m2a)N(m0)(n1)A(n,1)2.\frac{N(m_{2}^{a})}{N(m_{0})}\leq\frac{(n-1)A(n,1)}{2}.

Let

m3b=m0+2[w+1][w][w1].m_{3}^{b}=m_{0}+2[w+1]-[w]-[w-1].

Dividing the bound

A(n,3)N(m3b)N(m0)=(mH(w+1))(mH(w+1)1)2(mH(w)m0(w)+1).A(n,3)\geq\frac{N(m_{3}^{b})}{N(m_{0})}=\frac{(m_{H}(w+1))(m_{H}(w+1)-1)}{2(m_{H}(w)-m_{0}(w)+1)}.

by

N(m1b)N(m0)=mH(w+1)m0(w)(mH(w)m0(w)+1)A(n,1)2,\frac{N(m_{1}^{b})}{N(m_{0})}=\frac{m_{H}(w+1)m_{0}(w)}{(m_{H}(w)-m_{0}(w)+1)}\geq\frac{A(n,1)}{2},

we obtain

(mH(w+1)1)4A(n,3)A(n,1),(m_{H}(w+1)-1)\leq\frac{4A(n,3)}{A(n,1)},

so

N(m2c)N(m0)4A(n,3)A(n,1)+1.\frac{N(m_{2}^{c})}{N(m_{0})}\leq\frac{4A(n,3)}{A(n,1)}+1.

Finally, let

m5c=m0+2[w+2][w][w1].m_{5}^{c}=m_{0}+2[w+2]-[w]-[w-1].

We have

A(n,5)N(m5c)N(m0)=mH(w+2)(mH(w+2)1)2mH(w)12(N(m2e)N(m0)1)2,A(n,5)\geq\frac{N(m_{5}^{c})}{N(m_{0})}=\frac{m_{H}(w+2)(m_{H}(w+2)-1)}{2m_{H}(w)}\geq\frac{1}{2}\left(\frac{N(m_{2}^{e})}{N(m_{0})}-1\right)^{2},

so

N(m2e)N(m0)2A(n,5)+1.\frac{N(m_{2}^{e})}{N(m_{0})}\leq\sqrt{2A(n,5)}+1.

We conclude that

A(n,2)==a,b,cN(m2)N(m0)(nA(n,1)2)+(4A(n,3)A(n,1)+1)+(2A(n,5)+1),A(n,2)=\sum_{*=a,b,c}\frac{N(m_{2}^{*})}{N(m_{0})}\leq\left(\frac{nA(n,1)}{2}\right)+\left(\frac{4A(n,3)}{A(n,1)}+1\right)+\left(\sqrt{2A(n,5)}+1\right),

which contradicts the inequality

A(n,2)>((n1)A(n,1)2)+(4A(n,3)A(n,1)+1)+(2A(n,5)+1),A(n,2)>\left(\frac{(n-1)A(n,1)}{2}\right)+\left(\frac{4A(n,3)}{A(n,1)}+1\right)+\left(\sqrt{2A(n,5)}+1\right),

valid for all n5n\geq 5. (See Appendix C and bound-A5d in the Python code.) ∎

B.5.2. Case: N(m1b)N(m_{1}^{b}) big and r2r\geq 2.

Suppose n5n\geq 5 and

N(m1b)N(m0)A(n,1)/2.\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2.

From Lemma B.25, we have m0(w)=1m_{0}(w)=1, and because r2r\geq 2, we have m0(w1)=0m_{0}(w-1)=0.

Lemma B.28.

Suppose r2r\geq 2, m0(w)=1m_{0}(w)=1, and

N(m1b)N(m0)A(n,1)/2.\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2.

We have

m0(w)=1,m_{0}(w)=1,
mH(w+i)=dA(n,i)m_{H}(w+i)=dA(n,i)

for 0ir10\leq i\leq r-1, and

mH(w+r)=dA(n,r)d(d1)2mH(wr).m_{H}(w+r)=dA(n,r)-\frac{d(d-1)}{2}m_{H}(w-r).
Proof.

We have already seen in Lemma B.25 that m0(w)=1m_{0}(w)=1.

For wqw+r1w\leq q\leq w+r-1, there is only one nonzero term in Equation (3); for q=w+rq=w+r, there are only two. ∎

Lemma B.29.

Suppose n5n\geq 5, r2r\geq 2, m0(w)=1m_{0}(w)=1, and

N(m1b)N(m0)A(n,1)/2.\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2.

Then

A(n,3r2)A(n,r1)(dA(n,r1)1)2.A(n,3r-2)\geq\frac{A(n,r-1)(dA(n,r-1)-1)}{2}.
Proof.

Let

m3r2=m0+2[w+r1][w][wr].m_{3r-2}=m_{0}+2[w+r-1]-[w]-[w-r].

We have

A(n,3r2)N(m3r2)d=(dA(n,r1))(dA(n,r1)1)mH(wr)2d(mH(w)m0(w)+1)A(n,r1)(dA(n,r1)1)2.A(n,3r-2)\geq\frac{N(m_{3r-2})}{d}=\frac{(dA(n,r-1))(dA(n,r-1)-1)m_{H}(w-r)}{2d(m_{H}(w)-m_{0}(w)+1)}\geq\frac{A(n,r-1)(dA(n,r-1)-1)}{2}.

Lemma B.30.

Suppose n5n\geq 5, r2r\geq 2, m0(w)=1m_{0}(w)=1, and

N(m1b)N(m0)A(n,1)/2.\frac{N(m_{1}^{b})}{N(m_{0})}\geq A(n,1)/2.

Then r=2r=2 and

d<3A(n,4)A(n,1)2.d<\frac{3A(n,4)}{A(n,1)^{2}}.
Proof.

By Lemma B.29, we have

A(n,3r2)A(n,r1)(dA(n,r1)1)2A(n,r1)(A(n,r1)1)2.A(n,3r-2)\geq\frac{A(n,r-1)(dA(n,r-1)-1)}{2}\geq\frac{A(n,r-1)(A(n,r-1)-1)}{2}.

If r3r\geq 3, this is not possible by Lemma C.4.

If r=2r=2 the result follows from Lemma B.29 and

dA(n,r1)1>23dA(n,r1).dA(n,r-1)-1>\frac{2}{3}dA(n,r-1).

Lemma B.31.

We cannot have n5n\geq 5, kn1k\leq n-1, r2r\geq 2, and m0(w)=1m_{0}(w)=1.

Proof.

Assume we had a solution satisfying the given conditions. By Lemma B.30 we know that r=2r=2 and

d<3A(n,4)A(n,1)2.d<\frac{3A(n,4)}{A(n,1)^{2}}.

This implies

d<A(n,2)2n,d<\frac{A(n,2)}{2n},

by the inequality

6nA(n,4)<A(n,1)2A(n,2),6nA(n,4)<A(n,1)^{2}A(n,2),

which holds for all n3n\geq 3. (See Appendix C and bound-A5e in the Python code.) Thus, since mH(w2)<nm_{H}(w-2)<n, we have

mH(w+2)=dA(n,2)d(d1)2mH(w2)34dA(n,2).m_{H}(w+2)=dA(n,2)-\frac{d(d-1)}{2}m_{H}(w-2)\geq\frac{3}{4}dA(n,2).

Finally, taking

m6a=m0+2[w+2][w][w2],m_{6}^{a}=m_{0}+2[w+2]-[w]-[w-2],

we find that

A(n,6)\displaystyle A(n,6) \displaystyle\geq N(m6a)N(m0)\displaystyle\frac{N(m_{6}^{a})}{N(m_{0})}
\displaystyle\geq mH(w+2)(mH(w+2)1)m0(w)mH(w2)2(mH(w)m0(w)+1)\displaystyle\frac{m_{H}(w+2)(m_{H}(w+2)-1)m_{0}(w)m_{H}(w-2)}{2(m_{H}(w)-m_{0}(w)+1)}
=\displaystyle= mH(w+2)(mH(w+2)1)mH(w2)2d\displaystyle\frac{m_{H}(w+2)(m_{H}(w+2)-1)m_{H}(w-2)}{2d}
\displaystyle\geq 1234A(n,2)(34dA(n,2)1)\displaystyle\frac{1}{2}\cdot\frac{3}{4}A(n,2)(\frac{3}{4}dA(n,2)-1)
\displaystyle\geq 932(A(n,2)1)2.\displaystyle\frac{9}{32}(A(n,2)-1)^{2}.

This contradicts the inequality

(A(n,2)1)2>100A(n,6),(A(n,2)-1)^{2}>100A(n,6),

which holds for all nn. (See Appendix C and bound-A5f in the Python code.) ∎

Appendix C Collected inequalities involving Eulerian numbers

The argument in Section B used several dozen inequalities involving Eulerian numbers. We will not give detailed proofs of them; aside from the inequality in Lemma C.4, each of the inequalities used can be proven using Lemma C.1 or Lemma C.3 for large nn, and then verifying by hand the finite number of remaining cases. The Python code used to verify these remaining cases has been posted as an ancillary file alongside the arXiv submission. One proof is presented as Lemma C.5 to illustrate the method.

The reader is encouraged to verify the plausibility of such inequalities for large nn, using the asymptotic approximation A(n,q)(q+1)nA(n,q)\sim(q+1)^{n}, which is valid for fixed qq and large nn.

Recall our convention that A(n,q)=0A(n,q)=0 if qnq\geq n.

Lemma C.1.

For all n1n\geq 1 and q0q\geq 0, we have

(q+1)n(n+1)qnA(n,q)(q+1)n.(q+1)^{n}-(n+1)q^{n}\leq A(n,q)\leq(q+1)^{n}.
Proof.

Recall ([52, §1.3]) that the Eulerian number A(n,q)A(n,q) counts the number of permutations of {1,2,,n}\{1,2,\ldots,n\} with exactly qq ascents.

If we label the integers 11 to nn with labels 11 through q+1q+1, then we get a permutation with at most qq ascents by giving all the numbers of label 11 in decreasing order, then all the numbers of label 22 in decreasing order, and so on. Every permutation with at most qq ascents arises in this way; this proves the right-hand inequality.

If a permutation constructed this way has fewer than qq ascents, then there must exist adjacent labels ii and i+1i+1 where all the numbers with label ii are less than all the numbers with label i+1i+1. If that happens, we can record a number jj from 0 to nn which is the number of elements in the sequence with label at most ii, then subtract one from the labeling of everything with label greater than ii. There are (n+1)qn(n+1)q^{n} possibilities of this new data, and we can recover the original labeling by adding 11 to the label of everything that comes after the first jj elements in the sequence. So the number of labelings giving permutations with fewer than qq ascents is at most (n+1)qn(n+1)q^{n}. This proves the left-hand inequality. ∎

Lemma C.2.

For all n1n\geq 1 and q0q\geq 0, we have

A(n,q)n!.A(n,q)\leq n!.
Proof.

Follows from

qA(n,q)=n!.\sum_{q}A(n,q)=n!.

The two bounds given patch well enough for our modest needs: if qn/lognq\sim n/\log n, then the bound in Lemma C.1 is close to n!n!, at least in a power sense. Surely more precise asymptotics are known, but these weak bounds are enough for the proof of Lemma C.4.

Lemma C.3.

If n2n\geq 2 and

q<nlog(n+1)+11q<\frac{n}{log(n+1)+1}-1

then

(11/e)(q+1)nA(n,q)(q+1)n.(1-1/e)(q+1)^{n}\leq A(n,q)\leq(q+1)^{n}.
Proof.

We have

1/(q+1)<log(q+1q)<1/q1/(q+1)<\log\left(\frac{q+1}{q}\right)<1/q
en/(q+1)<(q+1q)n<en/q.e^{n/(q+1)}<\left(\frac{q+1}{q}\right)^{n}<e^{n/q}.

If n/(q+1)>log(n+1)+1n/{(q+1)}>\log(n+1)+1 then

(n+1)qn<(q+1)n/e.(n+1)q^{n}<(q+1)^{n}/e.

Now use Lemma C.1. ∎

Lemma C.4.

For nn arbitrary and 3rn23\leq r\leq n-2 we have

A(n,r1)(A(n,r1)1)>2A(n,3r2).A(n,r-1)(A(n,r-1)-1)>2A(n,3r-2).
Proof.

We’ll assume n21n\geq 21; for smaller nn there are only finitely many cases, which can be checked by hand. (See bound-B in the Python code.)

We can also assume rn/2r\leq n/2; otherwise, the right-hand side is zero. We’ll split into two cases: either r<nlog(n+1)+1r<\frac{n}{log(n+1)+1} or r>nr>\sqrt{n}. Our hypothesis on nn guarantees that there is some r0r_{0} with nr0<nlog(n+1)+1\sqrt{n}\leq r_{0}<\frac{n}{log(n+1)+1}; a fortiori, one of the two cases always holds.

If r<nlog(n+1)+1r<\frac{n}{log(n+1)+1}, then Lemma C.3 gives

A(n,r1)(11/e)rn,A(n,r-1)\geq(1-1/e)r^{n},

so

A(n,r1)(A(n,r1)1)(11/e)2r2n(11/e)rn.A(n,r-1)(A(n,r-1)-1)\geq(1-1/e)^{2}r^{2n}-(1-1/e)r^{n}.

On the other hand, by Lemma C.1, we can bound the right-hand side:

2A(n,3r2)2(3r1)n.2A(n,3r-2)\leq 2(3r-1)^{n}.

The reader may verify that

(11/e)2r2n>2(3r1)n+(11/e)rn(1-1/e)^{2}r^{2n}>2(3r-1)^{n}+(1-1/e)r^{n}

whenever r3r\geq 3 and n14n\geq 14; this proves the inequality we want.

Recall that we chose r0r_{0} such that nr0<nlog(n+1)+1\sqrt{n}\leq r_{0}<\frac{n}{log(n+1)+1}. Lemma C.3 gives

A(n,r01)(A(n,r01)1)>310r02n310nn>2n!.A(n,r_{0}-1)(A(n,r_{0}-1)-1)>\frac{3}{10}r_{0}^{2n}\geq\frac{3}{10}n^{n}>2*n!.

Now for any rr with r0rn/2r_{0}\leq r\leq n/2, we have

A(n,r)A(n,r0),A(n,r)\geq A(n,r_{0}),

so

A(n,r1)(A(n,r1)1)>2n!2A(n,3r2)A(n,r-1)(A(n,r-1)-1)>2*n!\geq 2A(n,3r-2)

by Lemma C.2. ∎

We’ll conclude with Lemma C.5, whose proof is given merely to illustrate a routine technique. A number of inequalities were used without proof in Appendix B. They can all be proven by asymptotic estimates using Lemma C.3 for large nn, followed case-by-case verification for small nn. The following bound was used in the proof of Lemma B.11; we give a full proof here to illustrate the general method.

Lemma C.5.

For n11n\geq 11, we have

A(n,2)A(n,1)2<1/27.\frac{A(n,2)}{A(n,1)^{2}}<1/27.
Proof.

Lemma C.3 gives us

A(n,2)3nA(n,2)\leq 3^{n}

and

A(n,1)(11/e)2n,A(n,1)\geq(1-1/e)2^{n},

provided n6n\geq 6. Hence we will be done as soon as we can show that

1(11/e)2(34)n<1/27,\frac{1}{(1-1/e)^{2}}\cdot\left(\frac{3}{4}\right)^{n}<1/27,

which happens for n15n\geq 15. The cases 11n1411\leq n\leq 14 must be checked separately. ∎

References

  • [1] Yves André. Mumford-Tate groups of mixed Hodge structures and the theorem of the fixed part. Comp. Math.  82 (1992), 1–24.
  • [2] Yves André. On the Shafarevich and Tate conjectures for hyperkähler varieties. Math. Ann. 305, 205–248 (1996).
  • [3] Benjamin Bakker and Jacob Tsimerman. The Ax-Schanuel conjecture for variations of Hodge structures. Invent. Math. 217 (2019), 77–94.
  • [4] Michael Bate, Sebastian Herpel, Benjamin Martin, and Gerhard Röhrle. Cocharacter-closure and the rational Hilbert-Mumford Theorem. Math. Zeit. 287 (2017), 39–72.
  • [5] Michael Bate, Benjamin Martin, and Gerhard Röhrle. A geometric approach to complete reducibility. Invent. Math. 161 (2005), 177–218.
  • [6] Michael Bate, Benjamin Martin, and Gerhard Röhrle. Semisimplification for subgroups of reductive algebraic groups. Forum of Mathematics, Sigma, 8, E43 (2020).
  • [7] Michael Bate, Benjamin Martin, Gerhard Röhrle, and Rudolf Tange. Closed orbits and uniform SS-instability in invariant theory.
  • [8] A. A. Beilinson, J. Bernstein, P. Deligne, and O. Gabber. Fascieaux pervers, Astérisque, Société Mathématique de France, Paris. 100 (1982).
  • [9] Pierre Berthelot. Cohomologie cristalline des schémas de caractéristique p>0p>0. Springer, Lecture Notes in Mathematics, 1974.
  • [10] Pierre Berthelot. Cohomologie rigide et cohomologie rigide à supports propres, Première partie. Preprint, 1996.
  • [11] Pierre Berthelot and Arthur Ogus. Notes on Crystalline Cohomology. Princeton University Press, 1978.
  • [12] Siegfried Bosch, Werner Lütkebohmert, Michel Raynaud. Néron Models. Springer, 1990.
  • [13] Eduardo Cattani, Pierre Deligne, and Aroldo Kaplan, On the locus of Hodge classes. Journal of the AMS 8 483–506 (1995).
  • [14] Pierre Deligne, Catégories tensorielles. Mosc. Math. J., 2 (2002), 227–248. Online at https://www.math.ias.edu/files/deligne/Tensorielles.pdf.
  • [15] Pierre Deligne, La conjecture de Weil: II, IHÉS Mathematical Publications, 52, 137–252 (1980).
  • [16] Pierre Deligne, Le groupe fondamental de la droite projective moins trois points. 1989.
  • [17] P. Deligne and J. S. Milne, Tannakian categories: Hodge Cycles, Motives, and Shimura Varieties. Lecture Notes in Mathematics, vol 900. Springer, Berlin, Heidelberg. https://www.jmilne.org/math/xnotes/tc2018.pdf.
  • [18] Vladimir Drinfeld, On a conjecture of Kashiwara, Mathematics Research Letters, 8, 713–728, 2001.
  • [19] Hélène Esnault, Phûng Hô Hai, and Xiaotao Sun, On Nori’s fundamental group scheme, In: Mikhail Kapranov, Yuri Ivanovich Manin, Pieter Moree, Sergiy Kolyada, Leonid Potyagailo, Geometry and Dynamics of Groups and Spaces. Progress in Mathematics, vol 265. Birkhäuser Basel
  • [20] Gerd Faltings. Crystalline cohomology and pp-adic Galois representations. In Algebraic analysis, geometry, and number theory: proceedings of the JAMI inaugural conference, edited by Jun-Ichi Igusa, Johns Hopkins University Press, 1989.
  • [21] Gerd Faltings, Endlichkeitssätze für abelsche Varietäten über Zahlkörpern. Invent. math. 73 (1983), 349–366.
  • [22] Gerd Faltings, Diophantine approximation on abelian varieties, Annals of Mathematics, 133 (1991), 549–576.
  • [23] Jean-Marc Fontaine. Périodes pp-adiques. In Astérisque, volume 223. Société Mathématique de France, 1994.
  • [24] J. Franecki and M. Kapranov, The Gauss map and a noncompact Riemann-Roch formula for constructible sheaves on semiabelian varieties, Duke Mathematical Journal 104 (2000), 171-180.
  • [25] Eberhard Freitag and Reinhardt Kiehl. Étale cohomology and the Weil conjectures. Springer, 1988.
  • [26] Ofer Gabber and François Loeser, Faisceaux perver \ell-adique sur un tore, Duke Mathematical Journal 83 (1996), 501–606.
  • [27] D. Gaitsgory, On de Jong’s conjecture, Israel Journal of Mathematics 157, 155–191 (2007).
  • [28] Mark Green, Phillip A. Griffiths, and Matt Kerr. Mumford–Tate Groups and Domains. Annals of Mathematics Studies, 2012.
  • [29] Ariyan Javanpeykar. Néron models and the arithmetic Shafarevich conjecture for certain canonically polarized surfaces. Bull. London Math. Soc. 47(1):55–64 (2015).
  • [30] A. Javanpeykar and D. Loughran. Arithmetic hyperbolicity and a stacky Chevalley–Weil theorem. Journal of the LMS 101 (2021), 846-849.
  • [31] A. Javanpeykar and D. Loughran. Complete intersections: moduli, Torelli, and good reduction. Math. Ann. 368 (2017), no. 3–4, 1191–1225.
  • [32] A. Javanpeykar and D. Loughran. Good reduction of algebraic groups and flag varieties. Arch. Math. (Basel), 104(2):133–143 (2015).
  • [33] Ariyan Javanpeykar and Daniel Loughran. Good reduction of Fano threefolds and sextic surfaces. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 18 (2018), no. 2, 509–535.
  • [34] Oliver Johnson and Christina Goldschmidt. Preservation of log-concavity on summation. ESAIM Probab. Stat., 10:206–215, 2006.
  • [35] Theo de Jong and Gerhard Pfister. Local analytic geometry. Advanced Lectures in Mathematics. Viehweg and Teubner, 2000.
  • [36] Nicholas M. Katz. A conjecture in the arithmetic theory of differential equations. Bull. Soc. math. France 110 (1982), 203-239.
  • [37] Nicholas M. Katz. Convolution and Equidistribution: Sato-Tate theorems for finite field Mellin transforms. Annals of Mathematics Studies 180, 2012.
  • [38] Nicholas M. Katz. Exponential Sums and Differential Equations. Annals of Mathematics Studies 124, 1991.
  • [39] Nicholas M. Katz. Moments, Monodromy and Perversity: a Diophantine perspective. Annals of Mathematics Studies 159, 2006.
  • [40] E. R. Kolchin, Algebraic groups and algebraic dependence, American Journal of Mathematics 90 (1968), 1151-1164.
  • [41] Thomas Krämer. Perverse sheaves on semiabelian varieties, Rendiconti del Seminario Matematico della Università di Padova 132 (2014), 83–102.
  • [42] Thomas Krämer., Characteristic cycles and the microlocal geometry of the Gauss map, I. https://arxiv.org/pdf/1604.02389v3.pdf, 2016.
  • [43] Thomas Krämer., Characteristic cycles and the microlocal geometry of the Gauss map, II. Journal für die reine und angewandte Mathematik (Crelles Journal, 2021 (2021), 53–92
  • [44] Thomas Krämer. and Rainer Weissauer, Vanishing theorems for constructible sheaves on abelian varieties, J. Algebraic Geom. 24 (2015), 531–568
  • [45] T. Krämer. and R. Weissauer, On the Tannaka group attached to the theta divisor of a generic principally polarized abelian variety, Mathematische Zeitschrift 281 (2015), 723-745.
  • [46] Robert M. Guralnick and Pham Huu Tiep, Decompositions of small tensor powers and Larsen’s conjecture, Representation Theory 9 (2005), 138–208
  • [47] Brian Lawrence and Akshay Venkatesh. Diophantine problems and pp-adic period mappings. Invent. math. 221, (2020) 893–999.
  • [48] Benjamin M. S. Martin. Reductive subgroups of reductive groups in nonzero characteristic. Journal of Algebra 262 (2003) 265–286.
  • [49] James S. Milne, Points on Shimura varieties over finite fields: the conjecture of Langlands and Rapoport. ArXiv preprint 0707.3173, v. 4.
  • [50] M. S. Narasimhan and M. V. Nori. Polarizations of an abelian variety. Proceeings of the Indian Academic Society (Math Society) 90 (1981), 125–128.
  • [51] A. Ogus. FF-crystals and Griffiths transversality. In Proceedings of the international symposium on algebraic geometry, Kyoto, 1977 (a Taniguchi symposium), Kinokuniya Book-Store Co., Ltd., 1978.
  • [52] T. Kyle Petersen. Eulerian numbers. Birkhäuser, 2015.
  • [53] Mihnea Popa and Christian Schnell, Generic vanishing theory via Mixed Hodge modules. For. Math. Sigma, 1 (2013).
  • [54] Marius van der Put and Michael F. Singer. Galois Theory of Linear Differential Equations. Springer, 2003.
  • [55] Kenneth A. Ribet. Galois action on division points of abelian varieties with real multiplications. Amer. J. Math., 98(3): 751–804 (1976).
  • [56] R. W. Richardson. Conjugacy classes of nn-tuples in Lie algebras and algebraic groups. Duke (1988) Vol. 57, No. 1.
  • [57] Neantro Saavedra Rivano, Catégories Tannakiennes. Bulletin de la Société Mathématique de France 100 (1972), 417–430.
  • [58] Wilfried Schmid. Variation of Hodge structure: the singularities of the period mapping. Inv. math. 22, 211–319 (1973).
  • [59] A. J. Scholl. A finiteness theorem for del Pezzo surfaces over algebraic number fields J. London Math. Soc. (1985).
  • [60] Peter Scholze. pp-Adic Hodge theory for rigid analytic varieties. For. Math. Pi, 1 (2013).
  • [61] Jean-Pierre Serre, Cohomologie Galoisienne. Springer, Lecture Notes in Mathematics, 1994.
  • [62] Jean-Pierre Serre, Complète Réducibilité. Séminaire Bourbaki, no. 932, 195–217 (2004).
  • [63] Yiwei She, The unpolarized Shafarevich conjecture for K3 surfaces. ArXiv preprint 1705.09038 (2017).
  • [64] Fucheng Tan and Jilong Tong. Crystalline comparison isomorphisms in pp-adic Hodge theory: the absolutely unramified case. Alg. Num. Th. 13, 1509–1581 (2019).
  • [65] Evgueni V. Tevelev. Projective Duality and Homogeneous Spaces. Encyclopedia of Mathematical Sciences 133, Springer-Verlag (2005).
  • [66] David Urbanik. Effective methods for Diophantine finiteness. ArXiv preprint 2110.14289 (2021).
  • [67] Rainer Weissauer. Brill-Noether sheaves. https://arxiv.org/pdf/math/0610923 (2006).
  • [68] Rainer Weissauer. Tannakian Categories attached to abelian varieties, in Modular Forms on Schiermonnikoog, edited by B. Edixhofen, G. van der Geer and B. Moonen, Cambridge University Press p.267–284, (2008).
  • [69] R. Weissauer. A remark on rigidity of BN-sheaves. https://arxiv.org/pdf/1111.6095 (2011).
  • [70] Rainer Weissauer. Vanishing theorems for constructible sheaves on abelian varieties over finite fields. Math. Ann. 365 (2016), 559–578.
  • [71] Kang Zuo. On the negativity of kernels of Kodaira–Spencer maps on Hodge bundles and applications. Asian J. Math., 4(1): 279–301 (2000).