The Shafarevich conjecture for hypersurfaces in abelian varieties
Abstract.
Faltings proved that there are finitely many abelian varieties of genus over a number field , with good reduction outside a finite set of primes . Fixing one of these abelian varieties , we prove that there are finitely many smooth hypersurfaces in , with good reduction outside , representing a given ample class in the Néron-Severi group of , up to translation, as long as the dimension of is at least . Our approach builds on the approach of [47] which studies -adic variations of Hodge structure to turn finiteness results for -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 with ring of integers , and let be a finite set of primes of . Fix an abelian variety , defined over , with good reduction at all primes outside . We say a hypersurface has good reduction at if the closure of in the unique smooth projective model of over is smooth at . Our main result is the following.
Theorem 1.1 (Theorem 9.4).
Suppose . Fix an ample class in the Néron-Severi group of . There are, up to translation, only finitely many smooth hypersurfaces representing , with good reduction outside .
If we fix a Picard class , 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 representing a given Picard class , because only finitely many translates of a given will represent . The hypersurfaces in a given Picard class form a projective space, and the singular ones form an irreducible divisor as soon as is very ample, by a classical result (e.g. [65, Theorem 1.18]) which uses the fact that is not ruled by projective spaces. Thus, the singular hypersurfaces are the vanishing locus of some discriminant polynomial in the homogenous coordinates of that projective space. Theorem 9.4 is equivalent to the statement that, for any -unit in , there are only finitely many solutions of the equation with all .
For there are additional combinatorial difficulties, leading to a more complicated result. Let be the sequence
satisfying
Let be the sequence
so that
Theorem 1.2 (Theorem 9.5).
Suppose . Fix an ample class in the Néron-Severi group of . Assume that the intersection number is not divisible by for any . There are only finitely many smooth hypersurfaces representing , with good reduction outside , up to translation.
Since increases exponentially, increases superexponentially. Because of this rapid rate of increase, and because is already large, a very small proportion of possible intersection numbers are not covered by Theorem 9.5.
If , then hypersurfaces in 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 , defined over , and having good reduction outside . 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 over with good reduction outside . The second is the result, in [22], that any closed subvariety of an abelian variety defined over that does not contain a positive-dimensional translate of an abelian subvariety contains only finitely many -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 to any moduli space of abelian varieties (except when ). 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 – 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 of a given dimension with good reduction outside . 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 , up to translation. Finally Faltings proved that each of these hypersurfaces contains only finitely many -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 , assuming one can find a family over whose cohomology has big monodromy. Suppose is a smooth proper family of varieties, extending to a smooth proper -integral model over . Then for every integral point with associated geometric point , the étale cohomology of the fiber gives rise to a global Galois representation
A lemma of Faltings shows that there are only finitely many possibilities for , up to semisimplification. In various settings, it is possible to show that the representation varies -adically in , and deduce that the -integral points are not Zariski dense in .
A key input to the methods of [47] is control on the image of the monodromy representation
(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 . 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 -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 of hypersurfaces in . For the monodromy groups of the universal family over 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 . Then, one hopes to improve from Zariski nondensity of to finiteness by passing to a subvariety. (This idea was suggested in [47, Sec. 10.2].) Specifically, we may take an irreducible component of the Zariski closure of in . If, under the assumption that is positive dimensional modulo translation, we can show that is not Zariski dense in , we obtain a contradiction. We thus deduce that is zero-dimensional modulo translation, and thus contains only finitely many distinct hypersurfaces modulo translation, and hence contains only finitely many hypersurfaces up to translation. However, this requires us to prove large monodromy, not just over , but for every positive-dimensional-modulo-translation subvariety .
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 provides us with a way out. The inverse image of under the multiplication-by- map has good reduction everywhere does, except possibly at , and we can run the argument with its étale cohomology. The -torsion points 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 -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 . 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 of dimension at least two and a smooth family . One wants to obtain a strong lower bound, over all pointed curves , on the dimension of the Zariski closure of the image of monodromy
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 from a family , through a study of various cohomology objects on . The étale local system gives rise to the global Galois representations to which Faltings’s lemma is applied; a filtered -isocrystal coming from crystalline cohomology is used to study the -adic variation of these Galois representations; and a variation of Hodge structure allows one to relate a -adic period map to topological monodromy. The method allows one to conclude that is not Zariski dense in .
In the present setting, we will apply these results to , where is a nontrivial local system on . 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 but not over an arbitrary number field. Additionally, -adic Hodge theory plays well with tensor categories over . Since passing to extensions (even unramified extensions) gives rise to semilinear operators, we need to set up the Tannakian formalism over . we prefer to work with a variety over , so we take to be a Zariski-closed subset of the Weil restriction from to of the moduli space of smooth hypersurfaces in . In fact, we simply take to be an irreducible component the Zariski closure of the set of integral points. We take the universal family of smooth hypersurfaces over ; that is, points are in bijection with smooth hypersurfaces over . To conclude Theorem 9.2, we need to show that is a translate of a constant family over .
Let . We will study cohomology objects , where is a local system on , given as a direct sum of characters. Any finite-order character on defines a local system on , by pullback via . If is defined over some field , the same is true of the corresponding local system . We construct a local system defined over by descent, as a sum of Galois conjugates of ; and we take . The construction is given in Lemma 5.29 in §5.5.
We think of and 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 gives, for every , a Galois representation
By Faltings’s lemma (Lemma 5.49), there are only finitely many possibilities for the semisimplification of , as varies over .
As in [47], we want to show that the fibers of the map
are not Zariski dense. To do this, we consider the map that takes a -adic point to a local Galois representation. By -adic Hodge theory, the local Galois representation
determines the filtered -module
We recall from [47, §3] some facts about the variation of with . For in a fixed mod- residue disk , the pair is constant: these spaces are canonically identified with the crystalline cohomology of the mod- reduction of . The filtration varies with . The assignment defines a -adic period map
to a certain flag variety. The -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 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 is semisimple. For this sketch, to illustrate ideas, let us make the same assumption; that is, let us imagine that every global representation is semisimple. Then there are literally only finitely many possibilities for the isomorphism class , 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 -module , as ranges over all integral points. In this simplified setting, we need only show that
is contained in a positive-codimension algebraic subset of .
Isomorphism classes of triples correspond to orbits of the Frobenius centralizer on , so we want to control , where 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.
-
(a)
The Frobenius centralizer is small.
-
(b)
The image of 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.
-
(a’)
(See Lemma 8.16.) Fix a -module and a semisimple global Galois representation . Consider all global Galois representations whose semisimplification is , and such that , for some filtration on .
The that arise in this way all lie in a subvariety of low dimension.
Once we have items a and b, we know that is contained in . A -adic version of the Bakker–Tsimerman transcendence theorem (Theorem 6.4) will imply that 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 having a given semisimplification. It is this latter result that requires us to work over (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 “-algebras” (§5.3) to package the cohomological data that arise in this situation. The Hodge–Deligne system will have the structure of module over a certain algebra object in the category of Hodge–Deligne systems; this module structure allows us to keep track of the Galois actions on embeddings of the field , isomorphism classes of local systems , 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 . 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 -balanced” (Definition 6.6; see Corollary 4.10 and Lemma 6.8 for precise statements). It depends on a parameter 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, is a subvariety of , with the map the restriction of the projection map to . The abelian variety has many rank-one local systems , each of which we can pull back to , push forward to , and apply this machinery to. These local systems are associated to characters of the fundamental group .
To apply the -adic Hodge theory argument described above, it suffices to have a local system on such that 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 , 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 for different local systems 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 , pushed forward to , and placed in degree . The vector space can be recovered from this by applying the functor . The theory of [44] views (a slight modification of) the category of perverse sheaves on as the category of representations of a certain group; the functors 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 is the convolution monodromy group.
We show that, if the convolution monodromy group of the constant sheaf on contains a classical group for some , and if the family over is not equal to a translate of the constant family, then for almost all , the monodromy groups of contain a classical group (Theorem 4.7, Corollary 4.10). To check the condition that the convolution monodromy group of the smooth hypersurface 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 to be the moduli space of hypersurfaces in a given Néron-Severi class in and to be an irreducible component of the Zariski closure of . Assuming that the universal family over is not equal to a translate of the constant family, we find a sheaf with big monodromy and, using -adic Hodge theory, show the integral points of are not Zariski dense. This contradicts the definition of 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 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 over an algebraically closed field, Krämer and Weissauer [44] construct a Tannakian category as a quotient of the category of perverse sheaves on . A perverse sheaf on 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 (say of dimension ) with theta divisor , one wants to know whether is isomorphic to a Jacobian, say . In this case, would be the -st convolution power of . Informally, the role of the Tannakian formalism here is to determine whether is “a -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 over a finite field . A perverse sheaf on has a trace function on . Associated to a character of is the character sum . Katz showed (in unpublished work analogous to [37]) that the distribution of , viewed as a random variable for uniformly random , converges to a distribution determined by the convolution monodromy group, in the limit as goes to . 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 , , we have ; that is, taking the tensor product of representations has the effect of multiplying the traces. For the character sums , convolution has the same effect:
In other words, convolution of these functions has a similar effect on this sum as tensor product of the representations 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 , and to perverse sheaves on which are summands of the pullback from to of geometrically semisimple perverse sheaves. Having done this, we obtain (in Lemma 2.8) an exact sequence of pro-algebraic groups
(1) |
where is the Tannakian group of a suitable category of perverse sheaves on , is the Tannakian group of a suitable category of perverse sheaves on , and is the Tannakian group of the category of -adic -representations – in other words, the Zariski closure of in the product of the general linear groups of all its finite-dimensional -adic representations. We think of this as a close analogue of the usual exact sequence
for a variety over a field .
Just like this usual exact sequence, (1) often has splittings. In our case, splittings arise from certain local systems on defined over , as the cohomology of a perverse sheaf twisted by a local system is a Galois representation, on which acts, and we can check that this action factors through the Tannakian group , giving the splitting.
Fix now a subvariety of the moduli space of smooth hypersurfaces in an abelian variety . Let be the field of functions on the generic point of . Let be the universal hypersurface in , defined over . Let be the constant sheaf on , pushed forward to , placed in degree ; this is our perverse sheaf of interest. Associated to is a representation of . The action of on this representation is a purely geometric object. By geometric methods, we will show in Theorem 3.5 that the image of acting on this representation contains , or as a normal subgroup. So the image of on the representation associated to contains the same classical group as a normal subgroup. Because the action of in this setting matches the action of the fundamental group, it will suffice for our big monodromy theorem to show that the action of 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 . We prove Lemma 4.4 showing that the action of on the cohomology of a perverse sheaf, defined using a generic local system , 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 is constant, up to translation by a section of , completing the proof of Theorem 4.7.
Next we describe how we check in Theorem 3.5 that the image of the -action on the representation associated to a smooth hypersurface in 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 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 . 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 by the -th wedge power of an -dimensional representation, we might expect that the Hodge structure on the cohomology of , or the cohomology of twisted by a rank one local system, is the -th wedge power of an -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 -adic Hodge-theoretic analogue, using the -action discussed earlier. On the other hand, we can calculate the Hodge numbers of the cohomology of 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 -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 5–8 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 “-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 -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 of characteristic is a rigid symmetric monoidal -linear abelian category with a faithful exact tensor functor to the category of vector spaces over . 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 , we refer to the image of the Tannakian group inside 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 ), 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 be an abelian variety over a field of characteristic zero. Fix a prime . Let be the derived category of bounded complexes of -adic sheaves on with constructible cohomology. Define a sheaf convolution functor that sends complexes to
for the group law.
Lemma 2.1.
is a rigid symmetric monoidal category, where the unit object is the skyscraper sheaf at , and the dual of a complex is
where is Verdier duality and is the inversion map.
Proof.
Let be the category of perverse sheaves on with -coefficients. Let be the subcategory of perverse sheaves with Euler characteristic zero. We similarly write and for the category of perverse sheaves on and its subcategory of objects with Euler characteristic zero, respectively. Let be the category of complexes in whose perverse homology objects lie in .
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” , i.e. that the convolution of two perverse sheaves has all perverse homology objects in nonzero degrees lying in . This lets us give the structure of a rigid symmetric monoidal -linear abelian category.
Lemma 2.2.
-
(1)
Perverse sheaves on have nonnegative Euler characteristics.
-
(2)
is a thick subcategory of (i.e. it is stable under taking subobjects, quotients, and extensions).
-
(3)
is a thick subcategory of (i.e. for any distinguished triangle with two objects in , the third one is in as well).
-
(4)
For , if or lies in , then lies in .
-
(5)
For , if .
-
(6)
Convolution descends to a functor
-
(7)
The essential image of in is equivalent to .
-
(8)
The essential image of in is stable under convolution.
-
(9)
is a rigid symmetric monoidal -linear abelian category.
Proof.
It suffices to check the first five statements after passing to , 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 be an abelian variety over a field . Fix a continuous character . We define the character sheaf to be the unique rank-one sheaf on whose monodromy representation is .
We also have a canonical way to descend these sheaves to :
Definition 2.4.
Let be an abelian variety over a field . Let be a character of that is -invariant. We define the character sheaf to be the unique lisse rank-one sheaf on whose associated representation of the fundamental group restricts to on and whose stalk at the identity has trivial Galois action.
In other words, we take the splitting of the exact sequence induced by the identity at , and use it to extend from to .
For a character of , let be the subcategory of consisting of perverse sheaves with for all and a subquotient of , and let .
Lemma 2.5.
-
(1)
The essential image of in is equivalent to .
-
(2)
contains the unit and is stable under convolution and duality.
-
(3)
is an exact tensor functor from to -vector spaces.
-
(4)
The category , convolution, and the functor are a rigid symmetric monoidal -linear abelian category with a faithful exact tensor functor to -vector spaces.
Proof.
(1) follows from [44, Lemma 12.3] and the fact that , 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
so if one vanishes for all 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 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
again using the vanishing of higher and lower cohomology.
(4) The category is rigid symmetric monoidal by part (2) and Lemma 2.2(4). It is -linear abelian because it is the quotient of a -linear abelian category by a thick subcategory. The functor is an exact tensor functor by part (3), and is faithful since exact -linear tensor functors between rigid abelian -linear tensor categories are automatically faithful if the endomorphisms of the unit are . ∎
By [44, Theorem 1.1], for any , there exists such that . We use the fiber functor on to define the “Tannakian group” of . This group is independent of the choice of , 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 an abelian variety over a field with algebraic closure , we say that a perverse sheaf on is geometrically semisimple if its pullback to is a sum of irreducible perverse sheaves.
Lemma 2.6.
Let and be geometrically semisimple perverse sheaves on . Then is the quotient of the space of homorphisms by the subspace of homomorphisms factoring through an element of .
Proof.
Without loss of generality, we may assume that and are indecomposable.
We first check that the set of isomorphism classes of irreducible components of the pullback of to forms a single -orbit. Suppose not; then we can fix a -orbit and consider an endomorphism of defined as the idempotent projector onto the sum of all irreducible components in that orbit. This endomorphism is, by construction, stable under . Because is perverse, is concentrated in degrees and thus
and hence this endomorphism arises from a nontrivial idempotent endomorphism of , contradicting the irreducibility of .
It follows that either all irreducible components of are in or none of them are. The same is true for by the same argument.
If all irreducible components of or are in , then or is in , so maps in the quotient category are zero and all maps factor through elements of , and the statement holds.
Thus, we may assume that no irreducible components of and are in . By definition, is the limit of where is a subobject of whose quotient lies in and is a quotient of by an object in . By assumption, we must have and , so . Again because no irreducible components of and lie in , no nonzero map from to factors through an object in , so the statement holds in this case as well. ∎
Lemma 2.7.
-
(1)
The full subcategory of consisting of geometrically semisimple perverse sheaves is a Tannakian subcategory of .
-
(2)
The full subcategory of consisting of summands of the pullbacks to of geometrically semisimple elements of on is a Tannakian subcategory of .
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 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 over and a character of .
Let be the Tannakian fundamental group of the full subcategory of consisting of geometrically semisimple perverse sheaves.
Let be the Tannakian fundamental group of the full subcategory of consisting of summands of the pullbacks to of geometrically semisimple perverse sheaves on .
Let be the Tannakian group of the category of -adic Galois representations over .
Lemma 2.8.
The group is a normal subgroup of , with quotient .
Proof.
There is a functor from the Tannakian category of Galois representations over to the geometrically semisimple objects of that sends a Galois representation to the corresponding skyscraper sheaf at the identity.
There is a functor from geometrically semisimple perverse sheaves on to summands of pullbacks of geometrically semisimple perverse sheaves to , given by pullback to .
Because these functors are both exact tensor functors, they define homomorphisms . We wish to show that this is an exact sequence of groups, i.e. that is a normal subgroup of whose quotient is . 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 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 is a closed immersion, we must check that every representation of is a subquotient of a pullback to of a representation of . This is automatic, as the Tannakian category of representations of is defined to consist of perverse sheaves that are summands of pullbacks of perverse, geometrically semisimple sheaves on that lie in , which by definition are representations of , and because all summands are subquotients.
Third we must check that a perverse sheaf on is a skyscraper sheaf at the origin if and only if is trivial when pulled back to . This is obvious.
Fourth, we must check that given a geometrically semisimple perverse sheaf on , its maximal trivial subobject over (i.e. the maximal sub-perverse sheaf that is a skyscraper sheaf at the origin) is a subobject over . 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 .
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 is a normal subgroup of , we will show that the Galois action on our fiber functor normalizes the Tannakian monodromy.
Lemma 2.9.
Let be an abelian variety over a field . Let be a character of that is -invariant. Let be the associated character sheaf.
Let be a geometrically semisimple perverse sheaf on such that vanishes for . Then the action of on normalizes the commutator subgroup of the identity component of the geometric convolution monodromy group of .
Proof.
We first prove that the action of normalizes the geometric convolution monodromy group of . For this, note that acts by automorphisms of the fiber functor of the arithmetic Tannakian category, giving a homomorphism . Since the geometric convolution monodromy group is a normal subgroup of the arithmetic Tannakian group, it follows that normalizes the geometric convolution monodromy group.
It follows that 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 of dimension and a smooth hypersurface in , which we will take except where noted to be defined over the complex numbers. Let be the inclusion. Let be the convolution monodromy group of the perverse sheaf . The main goal of this section is to compute .
To begin, we compute various Euler characteristics of - 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 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 be a line bundle on . We have
Proof.
By Hirzebruch-Riemann-Roch, the Euler characteristic of the coherent sheaf is the integral of its Chern character against the Todd class. By definition, the Chern character of is . Because the tangent bundle of is trivial, its Todd class is . Integrating is equivalent to taking the degree term, which is . ∎
Lemma 3.2.
The arithmetic Euler characteristic of is .
Proof.
Lemma 3.3.
The topological Euler characteristic of is .
Proof.
The topological Euler characteristic of is the top Chern class of the tangent bundle of . Using the exact sequence , and the fact that all Chern classes of vanish, we see that the top Chern class of is . ∎
Motivated by Lemma 3.2, we define the degree of to be , which is always a positive integer for an ample hypersurface.
The Hodge numbers of 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) |
where is the Eulerian number.
Proof.
Let .
Recall from the introduction that is the sequence
satisfying
(OEIS A061278).
Theorem 3.5.
Assume that is not equal to the translate of by any nontrivial point of and . Assume also that neither of the following holds:
-
(1)
and .
-
(2)
and for some .
Then contains as a normal subgroup either , or . If is not equal to any translate of then the case holds. If is equal to such a translate, then if is even holds and if is odd holds.
Remark 3.6.
In the exceptional cases, there are only a few other possibilities for . Suppose does not contain , , or . Then in case (1), contains acting by its -dimensional representation. In (2), contains as a normal subgroup acting by the representation .
The proof occupies the remainder of this section.
We will call the representation of associated to the object the distinguished representation.
Lemma 3.7.
The dimension of the distinguished representation of is .
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 times the topological Euler characteristic of . This now follows from Lemma 3.3.∎
Lemma 3.8.
Assume that is not equal to the translate of by any nontrivial point of . Then the distinguished representation of is an irreducible minuscule representation of the Lie algebra of .
Proof.
Because is smooth, the characteristic cycle of is simply the conormal bundle of with multiplicity , hence has a single irreducible component, with multiplicity one. The representation is then irreducible minuscule by [42, Corollary 1.0] ∎
Lemma 3.9.
Assume is not translation-invariant by any nontrivial point of and . Then the identity component of 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 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 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 of dimension is an effective Lagrangian cycle in the cotangent bundle of . In other words, it is a nonnegative-integer-weighted sum of irreducible -dimensional subvarieties of whose tangent space at a generic point is isotropic for the natural symplectic form on . Any such subvariety is automatically the conormal bundle to an irreducible subvariety of , of arbitrary dimension, i.e. its support.
For an abelian variety, because is a trivial bundle, we can express it as a product . 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 for a perverse sheaf 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 . In particular, given two such cycles , because
one can find an open set over which both and are finite. The fiber product then maps to by the obvious projection and to by composing the two projections with the multiplication map . Hence maps to . Its image has a unique clean extension to . Krämer defines this to be this extension [43, Example 1.3.2].
A key property of this convolution product is that ; as a consequence, if are clean and nonzero, then is nonzero as well.
We are now ready to begin the argument.
Assume for contradiction that the identity component of 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 and effective clean cycles on , with , such that
where is the multiplication-by- map. Because is smooth, the characteristic cycle of is simply the conormal bundle of , which is irreducible. Its degree is , because the degree of the Gauss map of the conormal bundle to is the sum of the multiplicities of vanishing of a general -form on , which is the Euler characteristic of , which is . In particular, this degree is nonzero, so .
Because is not translation-invariant, the map from to its image under is generically one-to-one, and so is an irreducible cycle with multiplicity one in the cotangent bundle of . This implies and are irreducible: If not, say if , we would have
with both and nonzero, contradicting irreducibility.
It follows that and must be the conormal bundles and of varieties and . Because , neither nor 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 .
Let be the image of under . By [43, Lemma 5.2.2] there is a dominant rational map from to (say), and thus a dominant rational map from to . Because is smooth and 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 ), is the Albanese of , so the surjective morphism extends to a homomorphism , giving a commutative diagram
Let be the image of . Because is surjective, .
If then is an abelian variety, and the conormal bundle to any nontrivial abelian subvariety of has Gauss map of degree , contradicting . Otherwise, by commutativity of the diagram, . Because is a hypersurface, it is a maximal proper subvariety of , so . This contradicts ampleness of unless is finite, and contradicts not being translation-invariant unless is an isomorphism. This means and are isomorphic as subvarieties of an abelian variety. Thus and so because , we have , contradicting .
Because we have a contradiction in every case, we have shown that is simple modulo its center.∎
Lemma 3.10.
Assume that is not equal to the translate of by any nontrivial point of . Then the commutator subgroup of the identity component of , together with its distinguished representation, is one of the following:
-
(1)
or , with its standard representation distinguished.
-
(2)
with one of its spin representations distinguished, or or with one of its lowest-dimensional nontrivial representations distingusihed.
-
(3)
with the representation distinguished for some .
Proof.
It follows from the Lemma 3.9 that the commutator subgroup of the identity component of 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 contains as a normal subgroup one of , , or . Then it contains only if is not equal to any translate of , it contains only if is even, and it contains only if is odd.
Proof.
Note first that the distinguished representation of any subgroup of which contains or 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 or respectively. Conversely, if then the distinguished representation of any subgroup of which contains 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, contains as a normal subgroup if, and only if, the perverse sheaf is not isomorphic, up to negligible factors, to the convolution of its dual 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 is not isomorphic, up to negligible factors, to any translate of . Because and 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 and 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 is equal to a translate of . This handles the case.
The argument to distinguish and is identical to [45, Lemma 2.1], which is stated only in the case where is a theta divisor, but the assumption is never used in the proof, except that they write instead of since, for a theta-divisor, . ∎
To prove the main theorem, it remains to give a complete list of for which 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 , we cannot have the commutator subgroup of the identity component of be with a spin representation distinguished, or or with one of their lowest-dimensional nontrivial irreducible representations distinguished.
Proof.
For , is always a multiple of . However, the stated representations cannot have dimension a multiple of , contradicting Lemma 3.7. Indeed, the spin representations have dimension a power of , while the lowest-dimensional representations of and have dimension and respectively, and none of these are multiples of . ∎
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 for a field and -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 , 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 -adic Hodge theory. For a representation of the Galois group of a -adic field with coefficients in , the completion of the algebraic closure of , and the Tate twist by , the Hodge-Tate weights of are the integers such that , and the multiplicity of the Hodge-Tate weight is the dimension of over .
We say is a Hodge-Tate representation if the sum of the Hodge-Tate weights is . The map sending to 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 is a faithful exact tensor functor. It follows that the category of Hodge-Tate representations is Tannakian, and corresponds to a map from the Tannakian group of the category of graded vector spaces to the Tannakian group of the category of Hodge-Tate representations.
Lemma 3.13.
Let be a smooth hypersurface, defined over a -adic field , in an abelian variety of dimension . Let be a finite-order character such that for . Let be a finite extension of containing the coefficient field and field of definition of .
Then the Hodge-Tate weights of the action on are , where the multiplicity of the weight is .
Proof.
As a finite-order character of , factors through for some . Let be the inverse image of under the multiplication-by- map of . Then we can express as the part of the étale cohomology where acts by the character . Applying -adic Hodge theory, we see that the dimension of the Hodge-Tate weight subspace of is equal to the dimension of the subspace of on which acts by the character . By descent, this is the dimension of for a torsion line bundle on . Because vanishes for , the dimension of for is equal to times the Euler characteristic because, as is torsion, its Chern class vanishes. By Lemma 3.4, this Euler characteristic is , so the dimension and Hodge-Tate multiplicity are both . ∎
Lemma 3.14.
Let be a hypersurface on on abelian variety of dimension and let . Suppose that the commutator subgroup of the identity component of is , with distinguished representation . Then there exists a function from the integers to the natural numbers and an integer such that and such that, for all ,
(3) |
Here we use the convention that unless .
Proof.
Fix a finite-order character of the geometric fundamental group of such that the cohomology of with coefficients in is concentrated in degree . Let be a finitely generated field over which , and may all be defined.
By our assumption on the convolution monodromy group and Lemma 2.9(2), it follows that the action on lies in the normalizer of inside the general linear group of the representation . It suffices to check that this condition implies the existence of and . For this we may spread out 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 is a number field.
This normalizer of inside the general linear group is if and if .
The category of Galois representations generated by 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 acting on (a quotient of the group defined in the previous section). Because 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 constructed via the associated graded vector space functor gives a map from to the Zariski closure of the image of , whose weights are the Hodge-Tate weights of .
Thus we have a homomorphism from to (or its semidirect product with ). Because is connected, this defines a homomorphism . Its weight multiplicities on the representation must equal the Hodge-Tate multiplicities of .
Homomorphisms are parameterized by their weights, an -tuple of integers with . Any homomorphism has a unique lift forming a commutative diagram
If we let be the weights of , then since their difference is a weight of the adjoint representation, which factors through and thus factors through .
Thus are rational numbers with the same fractional part, equal to the “weights” of the standard representation of . We let and let , which is independent of , so that .
Then the weights of the representation are exactly for all with .
It follows that the multiplicity of the weight inside of the standard representation of is the number of subsets with and .
To calculate this, let be the number of such that . For , let be the number of such that . Then and . Furthermore the number of sets attaining a given function is .
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 , function from the integers to the natural numbers and an integer such that and Equation (3) is satisfied for all . Then we have one of the cases
-
(1)
and ,
-
(2)
and for some , or
-
(3)
and for some . (Here is defined by
We are now ready to prove the main theorem of this section:
4. Big monodromy from big convolution monodromy
Let be a variety, an abelian variety, and a family over of smooth hypersurfaces in , with , , all defined over . Let , so that is the relative dimension of over . Let be the generic point of and a geometric generic point.
Let be the inclusion, and let be the perverse sheaf on . Let be the convolution monodromy group of . Let be the commutator subgroup of the identity component of . We continue to call the distinguished representation of the representation arising from the object , and let its restriction to be the distinguished representation of .
Lemma 4.1.
Assume that is a simple algebraic group with irreducible distinguished representation, and that is not equal to a constant family of hypersurfaces translated by a section of .
Let be an irreducible perverse sheaf on in the Tannakian category generated by . Assume that is a pullback from to of a perverse sheaf on . Then acts trivially on the irreducible representation of corresponding to .
Proof.
Let be a geometric generic point of . Let be the maps induced by the two projections . The convolution monodromy groups of and are both isomorphic to , so the convolution monodromy group of is isomorphic to a subgroup of whose projection onto each factor is surjective. By Goursat’s lemma, there are normal subgroups in and an isomorphism such that the convolution monodromy group of is isomorphic to
Note that has a unique factor in its Jordan-Hölder decomposition which is a nonabelian connected simple group. Hence this factor appears either in and or in and .
In the first case, we must have and . This is because, if the nonabelian connected simple factor appears in , then it must appear in as which, modulo scalars, is contained in the outer automorphism group of and thus is virtually abelian and cannot contain a nonabelian connected simple factor. Furthermore is a normal subgroup of , and since it cannot be a finite group, it must be .
Now, using the fact that and , we will show that acts trivially on the irreducible representation corresponding to . To do this, observe that and are isomorphic because is a pullback from . These correspond to two representations of the convolution monodromy group of that factor through the projection onto the first and second factors respectively. Because lies in and , the convolution monodromy group of contains two copies of . The first copy of acts trivially on , so it must act trivially on , so acts trivially on , as desired.
In the second case, and must both be contained in the scalars. To see this, because the scalars are the centralizers of , it suffices to show that the image of in the automorphisms of vanishes. Equivalently, we must show that the image of in the automorphisms of the Lie algebra of vanishes. This automorphism group is an extension of the finite outer automorphism group of by mod its center. Because the image of in the automorphism group is normalized by , it either contains or is finite, and it cannot contain , so it is finite. Because it is finite and normalized by , it commutes with . Because the Lie algebra of is an irreducible representation of , this forces the image of to act as scalars. But there are no nontrivial scalar automorphisms of a nonabelian Lie algebra, as they would never preserve any equation , and so the image of is trivial, as desired.
Now, using the fact that and are both contained in the scalars, we will derive a contradiction. Thus the convolution monodromy group is contained in the set for some automorphism of mod scalars. Let us first check that the automorphism is inner. To do this, let be the geometric generic point of . Let be the projections to . For each we have a homomorphism from the convolution monodromy group of to the convolution monodromy group of modulo scalars, and let and be the similarly-defined homomorphisms from the convolution monodromy group of to the convolution monodromy groups of and modulo scalars. Then because for any pair there is a projection onto the th and th copies of , and is isomorphic to the pullback of along this projection, the convolution monodromy group of is isomorphic to to the convolution monodromy group of . Hence for any pair and , there exists an automorphism of modulo scalars that sends to . Since and are surjective, this automorphism is unique, and so . Hence if is greater than the order of the outer automorphism group of mod scalars, there are and such that is an inner automorphism, say conjugation by . Because is isomorphic to the pullback of , conjugation by sends to . Because the convolution monodromy group is well-defined only up to inner automorphisms in the first place, we may assume . It follows that the map from the convolution monodromy group of to has image consisting of pairs where for a scalar .
Now the representation associated to is the standard representation of the first tensored with the dual of the standard representation of the second . Because the convolution monodromy group consists of pairs where for a scalar , this tensor product contains a one-dimensional subrepresentation . (Viewing this tensor product as the space of endomorphisms of the standard representation, the one-dimensional subrepresentation consists of scalar endomorphisms). Because admits a nontrivial homomorphism to , we have a natural map , which must be an isomorphism because both sides are irreducible. Any one-dimensional representation of the convolution monodromy group must be a skyscraper sheaf [44, Proposition 10.1], so we have an isomorphism for some . Considering the support, we see that the translation of by is . Spreading out this identity and then specializing to a sufficiently general point, we see that is generically the translation of a constant variety by a section of . We can extend this section to some open set, and then over some open set will be the translation of a constant variety by a section, and then because is a smooth proper family this will be true globally, contradicting the assumption. ∎
Lemma 4.2.
Let be abelian varieties. For , let be the inclusion map and let be a perverse sheaf on with sheaf convolution group . Then the sheaf convolution group of is .
Proof.
Since and have Tannakian groups and , the Tannakian group of their sum is a subgroup of whose projection onto both factors is surjective, so by Goursat’s lemma there exists a group and quotient maps and such the sheaf convolution group of is . For perverse sheaves corresponding to a faithful representation of , this implies , absurd unless and are both skyscraper sheaves at the identity, implying is trivial and so the Tannakian group is a product, as desired.∎
Corollary 4.3.
Assume that is a simple algebraic group with irreducible distinguished representation, and that is not equal to a constant family of hypersurfaces translated by a section of .
Let be a positive integer, and let be the inclusions of into that send to one of the coordinate axes.
Then the convolution monodromy group of is and thus contains as a normal subgroup. The representation associated to is isomorphic to the sums of the distinguished representations of the factors of . Finally, this normal subgroup acts trivially on any representation of the convolution monodromy group corresponding to a perverse sheaf that is pulled back from to .
Proof.
The convolution monodromy group is by induction on Lemma 4.2.
If any irreducible representation corresponding to a perverse sheaf which is a pullback of a perverse sheaf from to is nontrivial on , then it is nontrivial on the th copy of for some . Tensor product with for generic and then pushforward to gives a faithful exact tensor functor, whose associated map on Tannakian groups is the th inclusion of into , so the image under this map is nontrivial, which contradicts Lemma 4.1 and the fact that the pushforward to of a pullback from is itself a pullback from by proper base change.∎
We write for the set of continuous characters . We call the dual torus of .
We call the set of characters of trivial on the fundamental group of a nontrivial abelian subvariety of a proper subtorus of the dual torus of .
The generic vanishing theorem of Krämer and Weissauer [44, Lemma 11.2] states that for a perverse sheaf on , for for outside a finite set of torsion translates proper subtori of . (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 :
Lemma 4.4.
Let be a perverse sheaf of geometric origin. If no irreducible component of is a pullback from , then for all characters outside a finite set of torsion translates of proper subtori of , we have for and .
Proof.
The first claim follows from [44, Lemma 11.2], which is stated for varieties over and singular cohomology, but we may embed into 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 , we may assume is smooth. Let be the dimension of . We may spread out (using the fact that it is of geometric origin) to a sheaf over such that is perverse. Let and be the projections.
We will prove the second claim by contradiction. We will first assume that for a particular such that for , and derive some conclusions from this. We will then define a finite set of torsion translates of proper subtori of , assume that this nonvanishing holds for some 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 . This will essentially be the usual observation that sheaves with monodromy invariants have global sections, and thus have nonzero . Additional care must be taken because is a complex of perverse sheaves, but the decomposition theorem will give us exactly what is needed.
The stalk of at the generic point is the complex . By the assumption that this cohomology group vanishes in degree , the stalk of at the generic point is supported in degree . There is some open subset of over which remains a lisse sheaf in degree , and the Galois action matches the monodromy action on that open subset. By the decomposition theorem, 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 , restricted to this open set, that is isomorphic to the constant sheaf . Because is a sum of shifts of irreducible perverse sheaves, this irreducible summand on an open set must extend to a shift of an irreducible perverse sheaf on the whole space. Because is smooth, the unique irreducible extension of the constant sheaf from an open subset to all of is , which is a perverse sheaf shifted by . (In general, it would be the IC sheaf of .)
Because , and is a summand of , it follows that .
Now that we have interpreted the existence of nontrivial monodromy invariants cohomologically, we can re-express the cohomology group in terms of shaves on , which will enable us to understand its dependence on using the generic vanishing theorem. It follows from the Leray spectral sequence and the projection formula that
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 with , we take from [44, Lemma 11.2] a finite set of torsion translates of subtori such that for all not in this set and all . Because there are only finitely many where , the union of all of these is again a finite set.
Assume that for a particular not in this set. The vanishing of for all and all forces the spectral sequence
to degenerate, giving
In particular, we can conclude that
We will now derive a contradiction from this simpler statement, which notably is independent of .
Note first that because is perverse, and has fibers of dimension at most , we have for by [8, 4.2.4]. Hence there is a natural map
arising from the perverse -structure. Because is nonzero, this map must be nonzero. Applying adjunction, we obtain a nonzero map , and and shifting by , a nonzero map
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 is a pullback from , giving a contradiction.∎
Lemma 4.5.
Let be a simple algebraic group and a natural number. Fix an irreducible representation of . Let be the normalizer of inside the group of automorphisms of this representation. Then there is a finite list of irreducible representations of such that a reductive subgroup of contains if and only if has no invariants on any of these representations.
Proof.
Let be such that any subgroup of which is finite modulo scalars contains an abelian subgroup of index . Such exists by Jordan-Schur. Let be an irreducible representation of of dimension which remains irreducible on restriction to . Let be the -th projection .
We take our list to be, for each from to , all irreducible -subrepresentations of and composed with , together with, for , the representation . It is straightforward to check that has no invariants on these representations, so if contains then has no invariants. We must check the converse.
If has no invariants, then acts irreducibly on and . The Lie algebra of is an invariant subspace of , thus either contains , in which case contains , or is contained in , in which case is finite modulo scalars, hence has an abelian subgroup of index , thus cannot act irreducibly on , a contradiction. (Compare [39, Theorem 2.2.2], due originally to Larsen.)
Since contains for all , we have . By [40, Theorem on p. 1152], if , then there exists with and an isomorphism that sends to . This isomorphism is unique and thus -invariant, so and are isomorphic as representations of . Thus has invariants, contradicting our assumption and showing . ∎
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 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 and be the projection maps associated to a family of smooth projective hypersurfaces parameterized by .
Recall that the geometric monodromy group of a constructible -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 , the monodromy of 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 is not translation-invariant by any nonzero element of , that is a simple algebraic group with irreducible distinguished representation, and that is not equal to a constant family of hypersurfaces translated by a section of . Fix .
Then for characters of , with avoiding some finite set of torsion translates of proper subtori of the dual torus to , the following conditions are satisfied:
-
•
for , and
-
•
the geometric monodromy group of contains as a normal subgroup, where the representation associated to , restricted to , is isomorphic to a sum of copies of the distinguished representation of .
Proof.
First, we use the generic vanishing theorem to find a finite set of torsion translates of proper subtori of such that, for avoiding them, for . We will then take the inverse images of these subtori under the duals of the projections to be in our finite set of torsion translates of subtori of .
Next, to calculate the monodromy, we will use the fact that the monodromy group is equal to the Zariski closure of the image of acting on the stalk at the geometric generic point.
Let us first check that the geometric monodromy group of is contained in . It suffices to show for each that the geometric monodromy group of is contained in . This follows from Lemma 2.9(2).
Now by Lemma 4.5 we can find an explicit list of representations of such that any reductive subgroup of contains if and only if its action on all these representations has no invariants. By Deligne’s theorem, the monodromy group of 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 in the Tannakian subcategory generated by inside the arithmetic Tannakian category constructed in Lemma 2.7 of perverse sheaves on modulo negligible sheaves. (We have to check that is geometrically semisimple, but this is clear as the constant sheaf on any closed subvariety is semisimple. It follows that the action of on the representation associated to this complex factors through the normalizer of the action of , and thus factors through , and so all representations of correspond to geometrically semisimple perverse sheaves.)
Because acts nontrivially on all representations from Lemma 4.5, by Lemma 4.3 none of these perverse sheaves is a pullback from , so by Lemma 4.4, outside some finite set of torsion translates of proper subtori of , the Galois group has no invariants for these representations. Thus, outside some finite set of torsion translates, the Galois group contains . ∎
We now specialize to the case that is defined over a number field . Fix an algebraic closure of . Let be the set of pairs of an embedding of into and a character of where . Then naturally admits an action of , where acts on and acts on the values of the character (which necessarily lie in . This action extends the more straightforward action of on .
Fix inside the group an element of elements projecting to a lift of Frobenius inside the decomposition group at in both .
For applications later in the paper, we will be interested in proving our large monodromy statement uniformly in an orbit of and this statement will involve a monodromy group of a direct sum of sheaves produced by characters in the orbit of . We therefore prove a series of lemmas that enable us to obtain a monodromy statement of this type.
Lemma 4.8.
Suppose is defined over a number field , and let be a prime at which has good reduction. Let be a natural number such that fixes the Galois closure of . For any positive integer and set of torsion translates of proper subtori of the dual torus to , there is a finite set of torsion translates of proper subtori of such that for any outside the union of , there exist such that the tuple does not lie in the union of .
Proof.
We can freely replace by any multiple. By choosing a suitable multiple, we may assume fixes all elements of .
Let . We will prove this lemma as a consequence of the fact that the action of on is by invertible linear transformations, with no roots of unity as eigenvalues. (The action of is by multiplication by , while the action of is invertible, with eigenvalues the inverses of Weil numbers of absolute value . Hence their product has eigenalues Weil numbers of absolute value , which are in particular not roots of unity.)
It suffices to show for each torsion translate of a torus in , for all outside some finite set of torsion transltaes of proper subtori of , the number of tuples such that the tuple does not lie in is at most . Indeed, we can then take to be the union of all these finite sets, and then the number of tuples such that does not lie in the union of will be at least
For each from to , let be the inverse image of under the ’th inclusion map . Each is a subtorus of and, since is proper and the images of the inclusion maps generate at least one of the must be proper. Fix this . We will first check that
is a finite union of torsion translates of and then check that for not in this set, the number of tuples such that the tuple lies in is at most
For the first claim, consists of characters restricting to the trivial character on some abelian subvariety . If then and agree on restriction to . Since the eigenvalues of are algebraic numbers but not roots of unity, the fixed points of on are torsion and finite in number. The same is true for their inverse image under , i.e. the elements of whose images under and agree. Characters of whose restriction to take a fixed torsion value form a torsion translate of , completing the proof of the first claim.
For the second claim, the number of choices of for all is and these choices, together with the condition that lies in , determine , since if both and lie in by dividing, without loss of generality with , it would follow that lies in and hence lies in , contradicting the assumption on . ∎
Lemma 4.9.
Suppose is defined over a number field . For each tuple indexed by embeddings of finite subsets of torsion translates of proper subtori of , there exists an embedding of into and a torsion character of such that for no -conjugate of does lie in the union of . Furthermore, we can arrange that is of order a power of , for any given prime .
Proof.
For to have no Galois conjugate in any element of any , it suffices that not lie in any Galois conjugate of these translates of .
By definition, each proper subtorus of 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 . The action of 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 -torsion characters, and so any translate of a proper subtorus can contain at most -torsion characters, while there are -torsion characters in total, so as soon as is greater than this finite number of tori, there will be an -torsion character not in any of them. ∎
For the remainder of this section, we’ll suppose is an abelian variety over a number field , is a smooth scheme over , and
is a family of hypersurfaces in , smooth, proper and flat over . For every embedding , we can form schemes
and
these are both schemes over , and has a projection to
Let and be the projections; for every torsion character of , let be the corresponding character sheaf on .
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 as large as desired.)
Corollary 4.10.
Assume that is not translation-invariant by any nonzero element of , that is a simple algebraic group with irreducible distinguished representation, and that is not equal to a constant family of hypersurfaces translated by a section of .
Let be a natural number such that fixes the Galois closure of .
Then for any prime , positive integer , and prime where has good reduction, there exist an embedding and a torsion character of , of order a power of , such that for every conjugate of by an element of :
-
•
for , we have , and
-
•
there exist such that the monodromy group of contains as a normal subgroup, where the representation associated to , restricted to , is isomorphic to a sum of copies of the distinguished representation of .
We will eventually apply this result with the parameter taken sufficiently large to ensure the inequalities stated in Theorem 8.17 are satisfied. For this reason, the parameter will depend on the varieties , , and involved, and then Corollary 4.10 will give us a character depending on . However, until we choose the parameter in this way, all our results will be valid for any positive integer .
Proof.
This follows from the previous 3 results, applying Theorem 4.7 to each of the finitely many pairs . ∎
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 -points for any number field and set of prime ideals.
Proposition 4.11.
Let be an abelian variety of dimension over . Let be an ample class in the Picard group of . For a positive integer , let be the multiplication-by- map.
There exists a positive integer such that the natural period map from the moduli space of smooth hypersurfaces in of class to a period domain, which sends to the Hodge structure on , is quasi-finite.
Proof.
Let be this moduli space. Suppose that, for some (to be chosen below), the period map is not quasi-finite. Then its fiber over some point must contain a positive dimensional analytic subvariety .
Consider the variation of Hodge structures over whose fiber over a pair of hypersurfaces is . Over the diagonal in , this variation of Hodge structures has a Hodge class representing the identity isomorphism between the two Hodge structures. Let be the projection from the universal cover of the locus where this cohomology class is Hodge.
By [13, Corollary 1.3], is a Zariski closed subset. If divides then we have . Because is Noetherian, it follows that there exists such that whenever is a multiple of . Fix such an .
certainly contains the square of the positive-dimensional analytic subvariety discussed earlier. Fix . It follows that the fiber of over (for the second projection ) has positive dimension. Let be the smooth locus of some irreducible component of the fiber of over . Let be the universal family of hypersurfaces over and the projection map. By assumption, for all multiples of , for all , we have an isomorphism of Hodge structures
Hence the variation of Hodge structures is constant, and thus has finite monodromy. This is the sum, over characters of of order dividing , of , and so all these individual summands have finite monodromy.
The family 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 is a positive-dimensional subvariety of the moduli space , it is not the constant family up to translation. From this fact, and the finiteness of the mondromy of for all torsion characters , we will derive a contradiction.
Before proceeding, we consider the case where is translation-invariant by a nonzero element of , for the generic point of . 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 by it. The family is then a pullback from a family of hypersurfaces in this quotient of , and the pushforward from of is a summand of the pushforward from of , where is the composition of with the map , so our finite monodromy assumption remains true for . Hence we may assume that is not translation-invariant by a nonzero element of .
Let be the convolution monodromy group of and let be the commutator subgroup of the identity component of . By Lemmas 3.8 and 3.9, 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 outside some finite set of proper subtori , which necessarily includes at least one torsion character, the geometric monodromy group of contains , 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 over admits a Hodge–Deligne system that has big monodromy and satisfies two numerical conditions, then the integral points of 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 -balanced subgroup of (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 be a variety over a number field (which will eventually be taken to be ). A smooth, projective family of varieties over gives rise to various cohomology objects on . The argument of [47] relies on the interplay among several of these objects: a complex period map, a -adic period map, and a family of -adic global Galois representations on . 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 “-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 an abelian variety of dimension over a number field . Let be an arbitrary smooth variety over , and its base change to . Let
be a subscheme, smooth, proper and flat over . (In our application, we will take to be the Weil restriction, from to , of a subvariety of the moduli space of hypersurfaces on , and the universal hypersurface over . See the proof of Theorem 9.2.)
We can choose finite sets of places of , and of places of , and spread everything out to a family
in such a way that is finite étale over , is a smooth abelian scheme over , is smooth over , and is smooth, proper, and flat.
Fix a prime . Let be a field, containing , Galois over , and over which splits.
Fix some embedding . Fix a natural number . (The choice of 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 .) Corollary 4.10 gives a torsion character of , of some order , satisfying a big monodromy condition.
In Lemma 5.29, we construct an -algebra on and an -module on , where is the full -orbit containing . The construction is roughly as follows. Any character of , defined over , defines a local system on . By definition, is a subvariety of ; let
be the second projection. By Galois descent,
descends to a Hodge–Deligne system on , which is a module for the algebra
which descends to an -algebra on . We’ll fix , and suppress the subscript from and .
In Section 5.7 we elaborate the structure of and . Let , and let be the -algebra underlying either or . In Section 5.6, we define to be one of the groups , , or , viewed as an algebraic group over . We take to be the Weil restriction . This group has an action on a free -module (coming from the standard representation of , , or ), and we take to be the normalizer of in the group of -linear automorphisms of . Whether the de Rham or étale version is meant is devoted by subscripts, as in and .
5.2. Hodge–Deligne systems
Definition 5.1.
Let be an integer, and a prime power. A rational -Weil number of weight 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 .
An integral -Weil number is a rational -Weil number that is an algebraic integer.
When is a prime of , we write for the cardinality of the residue field at .
Definition 5.2.
Suppose given a number field with a chosen embedding . Let be a smooth variety over . Let be a finite set of primes of , and let be a smooth model of over . Let be a prime of not lying below any place of , such that is unramified over ; let be a place of lying over .
A Hodge–Deligne system 333The name is meant to evoke variations of Hodge structure and Deligne’s systems of realizations. on at consists of the following structures:
-
•
A singular local system of -vector spaces on .
-
•
An étale local system of -vector spaces on .
-
•
A vector bundle on , an integrable connection on , and a descending filtration of by subbundles
(not necessarily -stable), each of which is locally a direct summand of .
-
•
A filtered -isocrystal on (see, for example, [64, end of §3.1]),
with the following isomorphisms:
-
(1)
An isomorphism on between and the pullback of to . 444We do not use the étale-singular comparison; we could have left it out.
-
(2)
An isomorphism on between and .
-
(3)
An isomorphism on an open neighborhood of the rigid analytic generic fiber of between the underlying vector bundle to and the pullback of .
-
(4)
An isomorphism on between the -modules and .
and an increasing filtration of all four objects, compatible with all the isomorphisms, such that all this data satisfies the axioms:
-
•
and satisfy Griffiths transversality.
-
•
The connection is induced under the isomorphism (2) by the trivial connection on .
-
•
For each point of , the -th associated graded under of the stalk of at that point is a pure Hodge structure of weight .
-
•
The -th associated graded under of is pure of weight , i.e. for each closed point of with residue field , the eigenvalues of on the -th associated graded of are -Weil numbers of weight .
-
•
The -th associated graded of under is pure of weight , i.e. for each closed point of lying over with residue field , the eigenvalues of Frobenius on the -th associated graded of are -Weil numbers of weight .
-
•
The connection has regular singularities in a smooth simple normal crossings compactification of .
-
•
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 the variation of Hodge structure given by , , and . 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 and .)
We say a Hodge-Deligne system is pure of weight if vanishes and is the whole system.
The rank of a Hodge–Deligne system is the rank of the local system of -vector spaces. By the various isomorphisms, this is equal to the ranks of , , and .
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 , , , , be as above. An integral Hodge–Deligne system on consists of a Hodge-Deligne system on together with an integral structure on (i.e. a singular local system of free -modules on together with an isomorphism .)
Definition 5.4.
Let , , , , be as above. A polarized Hodge-Deligne system on consists of a Hodge-Deligne system on , pure of some weight, together with a polarization of the variation of Hodge structures (i.e. a morphism of local systems which restricted to the stalk at any point of defines a polarization of the pure Hodge structure at that point.)
Definition 5.5.
We say that a Hodge–Deligne system has integral Frobenius eigenvalues if the Weil numbers appearing as eigenvalues of Frobenius on and are integral, for all closed points and all closed points of lying over , respectively.
Definition 5.6.
The differential Galois group of a Hodge–Deligne system is the differential Galois group of the underlying vector bundle with connection . (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 ; 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 be the residue field of at . The “Frobenius automorphism” of is the element of that acts as the -th power map on .
A filtered -isocrystal gives, for every , a pair
where is a -vector space, and is an endomorphism of , semilinear over Frobenius. Furthermore, for every belonging to the residue class of , the object defines a filtration on , with an isomorphism to the filtered vector space . We’ll call the resulting data
Example 5.8.
(The trivial Hodge–Deligne system.)
Let and be number fields, and let be a finite set of places of . Take , and define the trivial Hodge–Deligne system on by:
-
•
.
-
•
, with the trivial Galois action.
-
•
, with trivial connection and filtration (i.e. , and ).
-
•
is determined by and the requirement that Frobenius act on through the trivial action on and the Frobenius automorphism of .
-
•
The weight filtration is such that is concentrated in weight zero.
When is an arbitrary smooth -variety, we define the system on by pullback from .
Remark 5.9.
In general, to give the structure of filtered -isocrystal on , it is enough to give a “Frobenius” isomorphism between the vector bundle with connection and its pullback under a lift of Frobenius to .
In Example 5.8, with constant and , the Frobenius is simply given as an automorphism of .
In general, Hodge–Deligne systems will come from families of varieties by taking cohomology.
Example 5.10.
(Pushforward of Hodge–Deligne systems.)
Let be a smooth variety over a number field , and let be a smooth model for over , for some finite set of places of . Let be a smooth, projective family of relative dimension ; let be the base change of to . Let be a Hodge–Deligne system on , and choose some with . Let be a prime of not lying below any place of , such that is unramified over , and let be a place of lying over . We define a Hodge–Deligne system on at , as follows.
-
•
Take , with the pushforward taken in the analytic topology on and . (This is again a local system, by Ehresmann’s theorem.)
-
•
Take , with the pushforward taken in the étale topology. (See [25] for an introduction to the étale topology.)
-
•
Take to be the relative de Rham cohomology of over , i.e. the pushforward of as a -module, with its Hodge filtration . This is a filtered vector bundle by Hodge theory.
- •
- •
-
•
The filtration is induced from the filtration of after shifting by . In particular, if is pure of weight , then is pure of weight .
Definition 5.11.
If and are two Hodge–Deligne systems on , a morphism from to consists of:
-
•
A map of analytic local systems ,
-
•
A map of étale local systems ,
-
•
A map of vector bundles , flat with respect to the connections on and , and respecting the filtrations and , and
-
•
A map of filtered F-isocrystals ,
compatible with all the comparison isomorphisms (1), (2), (3), (4).
Lemma 5.12.
The Hodge–Deligne systems on at form a Tannakian category with fiber functor given by for some .
In this Tannakian category, the tensor product of two systems will be defined by separately tensoring the individual objects and , 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 . In general the cokernel of a morphism of vector bundles need not be a vector bundle. But if is a morphism of Hodge–Deligne systems, the cokernel of must be a vector bundle because is a flat map of vector bundles with connection.
The second is the equality of images and coimages. A priori, the image of 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 is a morphism, then for any system on , we can define the pullback on by pulling back the four components separately. When the map is clear, we’ll sometimes write instead of .
Definition 5.13.
Let be a smooth -scheme. A constant Hodge–Deligne system on is a system of the form , where is a Hodge–Deligne system on .
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. -algebras
Throughout this section, will denote a number field, and a finite set of places of .
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 , for a family with not necessarily geometrically connected fibers. In this context, the zeroth cohomology of a fiber (over any ) has nontrivial Galois structure. The action of on gives rise to the bounds we need on the Frobenius centralizer, by means of the semilinearity of Frobenius. Specifically, has a natural structure of -module, and the Frobenius endomorphism of is semilinear over the Frobenius endomorphism of .
We need an analogous statement the Frobenius centralizer in our situation. Let be a smooth proper model of over . Suppose is finite étale over some , and is smooth over . Suppose is a smooth, proper, and flat over ; in the application, each fiber will be a hypersurface in . An order- character of , defined over some field , gives rise to a Hodge–Deligne system on the base change of to (Lemma 5.28); considering conjugates of , we can descend from to (Lemma 5.29). Taking cohomology, we will study a Hodge–Deligne system on , whose fiber over a point is (a descent to of) . This will be an algebra over ; we will study this structure in detail.
The object is the cohomology of a motive with coefficients in , defined over and having an algebra structure coming from the group scheme . The cohomology is naturally a module over the stalk of at the identity thanks to the compatibility of with convolution ; the purpose of this section and the next is to clarify the structure of the stalk of and modules over it.
To this end, we will define a general notion of “-algebras.” Loosely speaking, an -algebra is a weight-zero algebra object in the category of Hodge–Deligne systems. We will soon see that motives over extensions of (Example 5.21), motives with coefficients in a number field (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 -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 -algebra is a Hodge–Deligne system on a smooth -scheme , equipped with morphisms
and
satisfying the following properties.
-
•
is pure of weight .
-
•
The filtration on is trivial: and .
-
•
The morphisms and make into a commutative algebra object.
When we say “-algebra”, we will mean “commutative -algebra”.
Example 5.15.
(Trivial Hodge–Deligne system .)
Let and be number fields, and let be a finite set of places of . The trivial Hodge–Deligne system of Example 5.8 has an -algebra structure coming functorially from the algebra structure on .
Definition 5.16.
If , we say that is étale if is an étale -algebra.
Example 5.17.
(Étale -algebras over a field.)
Let , with a number field and a finite set of places of . In this setting we can give a concrete description of étale -algebras over .
The singular realization has the structure of -algebra, which we assume is étale; is determined by
with trivial filtration.
The étale realization is the -algebra
equipped with a continuous action of the Galois group . By assumption, is an étale -algebra, so is a finite group. The action of descends to the maximal quotient of unramified outside the union of and the set of places of lying over .
Finally we turn to . The structure of -algebra is given by an isomorphism
The filtration is trivial, and the Frobenius (which we will notate ) is the endomorphism of , where gives the action of on , and is the endomorphism of that acts as the -th power map on residue fields. The Frobenius is only well-defined up to conjugacy, but that is enough to determine up to isomorphism.
Example 5.18.
( of a family.)
If is a proper map, then the degree-zero cohomology of , equipped with the cup product, gives an -algebra on .
The underlying Hodge-Deligne system is constructed as in Example 5.10 as . The Hodge filtration is trivial because the Hodge filtration on of any smooth scheme is trivial. The map is the unit of the adjunction between and , and the map is given by cup product.
Example 5.19.
(Group algebra of a finite abelian group.)
For a finite abelian group , define the -algebra as follows.
-
•
.
-
•
, with trivial Galois action.
-
•
is the trivial vector bundle, with trivial connection.
-
•
The filtration on is .
-
•
The filtered -isocrystal is the constant vector bundle with trivial connection. Its fiber at any point is the group algebra , with Frobenius action coming from the Frobenius on .
Note that, for a group , it is natural to view the group algebra as a space of measures on , and thus dual to the space of functions on . 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 .)
Let be a finite étale commutative group scheme over some . The group operation defines a Hopf algebra comultiplication . The dual map gives the structure of an -algebra.
Denoting by the structure map, we define . A concrete description is as follows.
-
•
with the usual algebra structure.
-
•
with its natural Galois action and structure of -algebra.
-
•
with is natural algebra structure.
-
•
with its natural algebra structure and Frobenius coming from the Galois action on .
After passing to an extension of over which splits, an element gives an element of each of , , , and , and each of these four realizations is generated (as a vector space over the appropriate field) by .
Example 5.21.
Let be an extension of fields, and let and be finite sets of places of and , respectively, such that is finite étale over . Let be the map of schemes . If is an -algebra on , then is an -algebra on .
5.4. Modules over an -algebra
Definition 5.22.
Let be an -algebra. An -module is an -module object in the category of Hodge–Deligne systems; in other words, it is a Hodge–Deligne system with a morphism , such that the composition
is the identity map, and the diagram
(4) |
commutes.
Remark 5.23.
If and are -modules, then we define the tensor product
as the coequalizer of two maps
the first of which is induced from , and the second from . (See [14, §2.3].) Then also has the structure of -module.
Definition 5.24.
Let be a finite étale algebra over a field . We say an -module is equidimensional if it is free of finite rank.
Equivalently, writing as a product of fields , we say that is equidimensional if
is independent of .
In this case, we call that dimension the rank of .
Definition 5.25.
Suppose is a constant étale -algebra, and is an -module. We say is equidimensional if the stalks of are equidimensional modules over the stalks of . In this case, the ranks of the stalks as -modules are constant; we call that rank the rank of .
Lemma 5.26.
If is an equidimensional -module of rank on , then the following statements hold.
-
•
The stalks of are equidimensional modules of rank over the stalks of .
-
•
The stalks of are equidimensional modules of rank over the stalks of .
-
•
The stalks of are equidimensional modules of rank over the stalks of .
Proof.
Follows from the “comparison isomorphisms” in the definition of Hodge–Deligne system. ∎
Definition 5.27.
Suppose is an equidimensional -module of rank . Then we say that is an -module with -structure, or has -structure over .
Suppose additionally that there exists a Hodge-Deligne system of rank 1, with a nondegenerate bilinear pairing
If the pairing is alternating, we say that has -structure over ; if symmetric, we say that has -structure over .
In Lemma 5.30, we will see that if is equal to a translate of , then the cup-product pairing gives either -structure or -structure on its middle cohomology; here we take to be the top-degree cohomology of .
Below (Definition 5.38) we will define a notion of “object with -structure” in a Tannakian category. The Tannakian notion (Definition 5.38) only applies to objects of an -linear tensor category, with a field; and is not a field. We will use the words “over ” to emphasize this distinction. The two notions are compatible in that, when is a field, viewed as a constant local system, a -structure on over gives a -structure on over , and similar statements are true for 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 be an abelian variety of dimension over a number field , an arbitrary smooth variety over , and its base change to . Let be a finite set of primes of , and a finite set of primes of , such that is finite étale over . Let be a smooth proper model for over , and a smooth model for over . Let
be a subscheme, smooth, proper and flat over . Let and be the projections. Fix a prime over which is unramified and a positive integer prime to . Let be a field, containing , Galois over , and over which splits; we suppose that has been chosen so that is unramified over . Let be an order- character of ; let be the corresponding character sheaf on . (It is a -local system on the étale site of .) Let
We want to create a Hodge–Deligne system on whose base change to has as a direct summand.
The tensor product splits as a direct sum
indexed by the embeddings of into . Here each is an isomorphic copy of ; the superscript is merely an index. We have the corresponding splitting
where for each , we define
Similarly, define
the base change of along . Then we have the Cartesian diagram
Let and be the projections.
Let be prime to , and let be the set of all pairs , where is a -linear embedding, and is a character of of order dividing . For fixed , the set of characters is naturally identified with the set of homomorphisms
thus, has a natural action of
where acts on the pairs via its action on , and acts on .
Lemma 5.28.
Let be character of of some finite order . Then there exists an -module on such that is the character sheaf associated with the character , and is the analytic -local system associated with .
Proof.
Descent. ∎
Lemma 5.29.
(Construction of and .)
Let be an orbit of on . There exist an -algebra on and an -module on with the following properties.
-
•
After base change to and extension of coefficients to , we have the direct sum decomposition
where each is a copy of .
-
•
The Galois representation is compatible with the isomorphism
where acts on the right-hand side by permutation of the characters .
Similarly, the Frobenius endomorphism of is compatible with the isomorphism
where Frobenius acts on the right-hand side by permuting the pairs , via the action, with trivial action on .
-
•
After base change to and extension of coefficients to , the module decomposes as the direct sum
Furthermore, this decomposition is compatible with the decomposition of
into fields.
-
•
can be made into a polarized, integral Hodge–Deligne system.
Proof.
We have the map
where is multiplication by ; let
We have an isomorphism of Hodge–Deligne systems
Furthermore, on , we have the direct sum decomposition of Hodge–Deligne systems
where is the system on , and the trivial system on all other components of . This gives a decomposition on
Let
Let be the pullback to of the Hodge–Deligne system on coming from
where is the projection. This has a structure of -algebra, coming from the structure of group scheme on (see Examples 5.20 and 5.21). Furthermore, the group action makes into a module over .
After base change from to and extension of coefficients, we have a decomposition of -algebras on
The set has commuting actions of and , the former coming from the base field , and the latter from the coefficient field . For each orbit of , the direct sum descends to an -algebra over , with -coefficients. More precisely, choose any , and let be the pushforward of from to . Then there is an isomorphism
Over and with -coefficients, splits as the direct sum of the algebras .
The -module also splits as the direct sum of the objects , and after base change and extension of coefficients we have
Finally we have to explain the polarization and integral structure on . On 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 . These structures induce a polarization and integral structure on , by restricting the polarization and intersecting the integral lattice with . ∎
Lemma 5.30.
Furthermore, if is equal to a translate of , then has -structure over if is even and -structure over if is odd.
Note that is dual to in the Tannakian formalism; see 3.11.
Proof.
To prove that has -structure we only need to check that is equidimensional; this is a consequence of the transitive Galois action on the index set .
Suppose is equal to a translate of (so also is equal to a translate of ). This equality gives an involution
The cup product pairing and the trace map compose to give a map
This pairing does not factor through , which would be equivalent to the identity
(5) |
of maps (where and are local sections of the sheaves underlying and .) Instead, for , the pairing satisfies
However, when is equal to a translate of , and is the involution of described above, we can form the pairing
as the composition
We claim that this pairing does satisfy (5), so that it descends to a pairing
It is enough to check (5) after extending both the base field and the coefficient field, so we can assume is the group algebra of the finite abelian group , and that splits as a direct sum over characters of that group. Now (5) follows from
where .
We have
so this pairing is symplectic if is even (and thus is odd) and symmetric if is odd (and thus is even). ∎
For a fixed -orbit on , define the reductive group as follows. Let be the rank of . If is equal to a translate of and is even, let . If is equal to a translate of and is odd, let . Otherwise, let . Then by Lemma 5.30, has -structure over . In the following sections we will analyze this structure in some detail.
5.6. An algebraic group
Let be a field – we will mostly be interested in the case – and let be a finite étale -algebra. In this subsection we will define some groups and . Specializing to , the group will contain the monodromy group of , and specializing to , the group will contain the differential Galois group of .
Let be a reductive group over , and choose a representation of . We will assume that is one of , , or , and is the standard representation. A subtlety is that the definition of depends on the choice of a symmetric bilinear form on . Here, and at all future points, when we assume that is , we mean that there exists a nondegenerate symmetric -linear form on for which 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 .
Let be the Weil restriction , and let . By restriction of scalars, we will view as an -vector space with actions of and . Let be the normalizer of in the group of -linear automorphisms of .
Definition 5.31.
Let be a -linear automorphism of . We say that an -linear endomorphism is -semilinear (or semilinear over ) if
for all and .
Lemma 5.32.
If , then is the algebraic group whose -points correspond to endomorphisms of , semilinear over some -linear automorphism of .
If is or , then preserves, up to scaling, an (alternating or symmetric) -valued pairing on . In this setting is the group of endomorphisms of , semilinear over some -linear automorphism of , and satisfying
(6) |
for some , for all .
In particular, we have an exact sequence of groups
Proof.
Any element of the normalizer of normalizes the center of , which is , acting by scalar multiplication. The -linear automorphisms of that normalize the action of are exactly the semilinear automorphisms.
If is or , then for which is -semilinear over , the bilinear form is -semilinear over in both variables and is preserved up to scaling by . Thus is -bilinear and preserved up to scaling by . Hence it is a scalar multiple of , which is the unique -linear -equivariant form. This gives (6).
Conversely, any -semilinear automorphism which, if is or , satisfies (6), manifestly normalizes . ∎
We need a generalization of [47, Lemma 2.1].
Lemma 5.33.
Suppose is such that , and suppose is semilinear over . Then the centralizer satisfies
Proof.
The proof goes through exactly as in [47]. By passing to the algebraic closure of we may assume that , so . The hypothesis implies that acts transitively (i.e. cyclically) on the factors of . We may assume the factors are ordered so that takes the -th factor to the -st factor (modulo ). Then has the off-diagonal matrix form
Now if , then
for all (indices modulo ). Thus, determines for all . In other words, the projection onto any single factor of is injective. The result follows because has finite index in . ∎
5.7. Structure of -modules
Let be a number field, a finite set of places of , and a smooth -scheme. Let be an étale -algebra on , and an -module on with -structure, where is , , or . We described the structure of in Example 5.17. Now we are ready to describe .
The local system is given by a Galois representation on , unramified outside , which we’ll also call . By definition this is a finite étale -algebra with an action of .
Let , and let and be the groups and of Section 5.6 taking . At any , the fiber of the étale local system is a representation of the Galois group . This has the structure of -algebra, and if is or then there is a pairing
where is a one-dimensional -vector space with an action of . The action of respects the pairing, and acts on semilinearly over its action on . It follows (Lemma 5.32) that the representation of on is given by a homomorphism
and the quotient
is exactly the representation of on given by the structure of .
Now we turn to de Rham cohomology and its cousins. Similarly, is a finite -algebra which is isomorphic to over (and hence also over and thus is also étale. Let and be the groups and of Section 5.6 taking .
For a point , Lemma 5.32 implies that is isomorphic to the group of automorphisms of respecting the -action and (where applicable) the bilinear pairing.
The isomorphism between and after base change to implies that and become isomorphic after base change to , and the same is true for and .
The differential Galois group satisfies . The weaker inclusion is immediate from the Tannakian formalism: the local system is naturally an algebra over , and if is or , then admits a bilinear pairing
(see Lemma 5.30). Since the monodromy action on is trivial, is in fact contained in the finite-index subgroup .
Lastly, we describe the filtered -module . For this, we assume that . Recall the structure of from Example 5.17: it is the -algebra , equipped with a Frobenius endomorphism . Then is a vector space over , and it is naturally a filtered -module with -structure. Its Frobenius endomorphism is semilinear over .
5.8. Disconnected reductive groups
We need to apply -adic Hodge theory in a Tannakian setting: we’ll work with Galois representations valued in the disconnected algebraic group , and the corresponding filtered -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 be an algebraic group over a field of characteristic zero whose identity component is reductive.
Definition 5.34.
Let be a morphism of schemes. We say that exists if extends to a morphism ; in this case we write
Let be a cocharacter. Define the subgroups , , and of as follows.
-
•
-
•
-
•
We say that a subgroup is parabolic if it is of the form for some ; in this case we say is a Levi subgroup associated to , for any such that .
In this setting, is unipotent, and is the semidirect product of by ; in particular, there is a natural projection , given by
Furthermore, is the centralizer of in [48, Prop. 5.2].
Example 5.35.
The purpose of this example is to show that a parabolic subgroup is not uniquely determined by the parabolic subgroup of .
Let be the normalizer in of . Then . Let cocharacters be given by
Then but .
Now we turn to the notion of semisimplification of subgroups of . 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 is -completely reducible if, for every parabolic subgroup containing , there is some Levi subgroup associated to that also contains .
If is an arbitrary algebraic subgroup, we define its semisimplification with respect to as follows. Let be a parabolic subgroup of , minimal containing . Choose a Levi subgroup of , and let be the image of under the projection .
We say that a representation valued in is semisimple if the Zariski closure of its image is -completely reducible.
Lemma 5.37.
Let be an algebraic subgroup. Then is -completely reducible if and only if the identity component of is reductive. If we choose an embedding of into , then is -completely reducible if and only if it is -completely reducible.
For a general algebraic subgroup , the -semisimplification is well-defined up to -conjugacy, and it is -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 -structure. See [49, Def. 1.3]; a general reference for the Tannakian formalism is [57].
Definition 5.38.
Let be an algebraic group over a field of characteristic zero, and let be an -linear rigid abelian tensor category with fiber functor. An object in with -structure is an exact faithful tensor functor , where is the category of finite-dimensional -linear representations of , together with an isomorphism between the composition of with the fiber functor of and the forgetful functor from to the category of vector spaces over .
Two objects in with -structure are -conjugate if they correspond to isomorphic functors . Note that -conjugate objects with -structure differ exactly by an automorphism of the forgetful functor , i.e. by an element of .
If is a morphism of algebraic groups, then an object with -structure gives rise to an object with -structure by functoriality.
Remark 5.39.
If is a faithful representation of , then an object with -structure is determined by . In practice, we will find it useful to specify by describing .
5.8.2. Filtrations
(Filtrations on .)
We want to define a notion of “filtered vector space with -structure” or “filtration on .” 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 -modules, which is abelian. Instead, we will use the formalism of filtrations from [57, §IV.2]. Throughout this section, will be an algebraic group over a field of characteristic zero.
Definition 5.40.
A filtration on (or -filtration) is an exact tensor filtration of the forgetful functor from to , in the sense of [57, IV.2.1.1].
In other words, a -filtration is a sequence of exact subfunctors of , such that:
-
(1)
For each object of , the objects form a decreasing filtration of .
-
(2)
The associated graded functor , which assigns to the associated graded of the filtration , is exact.
-
(3)
For every , and all objects of , we have
Loosely speaking, a -filtration is a choice of descending filtration on every finite-dimensional representation of , compatible with tensor products, duals, and passage to subquotients.
If is a morphism of algebraic groups, then a -filtration gives rise to a -filtration by functoriality.
Since the base field is of characteristic zero, every -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 .
Let us explain this more carefully. A representation of on a finite-dimensional vector space gives a decomposition , where is the -eigenspace of on . A cocharacter defines a filtration on by defining, for all representations of ,
That every filtration on is “scindable” means that every filtration comes from some cocharacter (in general not unique).
Given a filtration on , we can define two distinguished subgroups and of [57, §IV.2.1.3]. The subgroup is the group of elements that stabilize the filtration on every representation of , and is the group of elements that stabilize the filtration and furthermore act as the identity on the associated graded. When is a group whose identity component is reductive, and the filtration is defined by a cocharacter , the group coincides with the Richardson parabolic (Definition 5.34), and is its unipotent radical .
Again supposing is a group whose identity component is reductive, define an equivalence relation on the set of cocharacters of as follows: we say that if and is conjugate to in this parabolic. Then and give rise to the same filtration on if and only if [57, §IV.2.2.4].
If acts faithfully on , then a -filtration is determined by the corresponding filtration on , as in Remark 5.39; we will use this without comment.
We mention in passing that, even if is reductive, the parabolic does not quite determine the filtration. If is reductive, then determines a filtration on every representation of only up to reindexing.
5.8.3. -filtrations vs. -filtrations
Let be an algebraic group over a field of characteristic zero whose identity component is reductive.
A -filtration gives rise to a -filtration , by functoriality; in terms of cocharacters, the correspondence is given by composition
Definition 5.41.
We say that a -filtration is associated to a -filtration if is induced from by functoriality with respect to the inclusion .
Lemma 5.42.
Being associated defines a bijection between -filtrations and -filtrations.
Proof.
We need to show that every filtration on comes from a unique filtration on .
Existence is easy. A filtration on is defined by a cocharacter ; since is connected, every such cocharacter factors through .
Conversely, suppose
are two cocharacters defining the same filtration on . We need to show that and define the same filtration on .
We know there is some such that
We need to show that we can in fact take
Let be the image of under the projection . Then centralizes by [48, Prop. 5.2(a)], and from the limit formula
we see that . Thus conjugates to . ∎
5.8.4. Weakly admissible filtered -modules with -structure
Let be an algebraic group over . (It is important that we work over , and not an extension, because the category of filtered -modules is only -linear.) A weakly admissible filtered -module with -structure gives rise to a -filtration and an element (the Frobenius endomorphism) in . In the spirit of Remark 5.39, we can describe such an object as a triple , where is a vector space on which acts faithfully, is an automorphism of , and is a -filtration on , such that is weakly admissible. Somewhat imprecisely, we will call such objects “filtered -modules with -structure.”
5.8.5. -adic Hodge theory
The next result is a Tannakian form of the crystalline comparison theorem, for representations valued in an arbitrary algebraic group . In order to avoid problems with semilinearity in the target category, we restrict to representations of .
Lemma 5.43.
Let be an algebraic group over . We say that a local Galois representation
is crystalline if the representation
is crystalline in the usual sense. This property is independent of the choice of embedding .
-
(1)
The functor of -adic Hodge theory [23, Expose III] extends to a functor from crystalline representations to pairs of an inner form of and an admissible filtered -modules over with -structure. (Here morphisms of such pairs are given by isomorphisms of inner forms of together to isomorphisms of admissible filtered -modules compatible with the isomorphism of inner forms.)
-
(2)
A homomorphism of groups and a crystalline Galois representation induces a morphism of inner forms compatible with the filtered -modules with -structure and -structure associated to and respectively. The map is the twist of by some -torsor. In particular, if has any property preserved by such twists then does as well.
-
(3)
For and a Galois representation whose composition with the natural map is the usual action of Galois on , the relevant inner twist of is .
Proof.
The general claims follow from the Tannakian formalism and the fact that is a fiber functor on the Tannakian category of crystalline Galois representations [23, Exposé III, Proposition 1.5.2].
Indeed, a Galois representation gives rise to an exact tensor functor from to the category of Galois representations. The crystallinity assumption implies that all representations in the image are crystalline. The composition of this functor with then gives an exact tensor functor from representations of to admissible filtered -modules, and the underlying vector space is a fiber functor. However, this fiber functor need not be isomorphic to the forgetful functor of , so we do not immediately obtain a filtered -module with -structure. Instead, for the group of automorphisms of this fiber functor, [17, Theorem 2.11(b)] implies that the Tannakian category of representations of is equivalent to the Tannakian category of representations of . Hence this functor defines an admissible filtered -module with structure. Furthermore, [17, Theorem 3.2(b)] implies that the category of fiber functors is equivalent to the category of -torsors. Since the automorphism group of any -torsor is an inner form of , it follows that is an inner form of .
Functoriality is the statement that if two Galois representations are -conjugate, then the associated groups are isomorphic and the associated filtered -modules with -structure are -conjugate. This is because the two exact tensor functors from to the category of Galois representations are isomorphic, and hence the functors from to the category of admissible filtered -modules are isomorphic. This completes tthe proof of (1).
For (2), we now have a composition of functors, first from to , then to the category of Galois representations, then to the category of filtered -modules, then to the category of vector spaces. The fiber functor of we use to construct is given by the composition of all these functors, and the fiber functor of we use to construct is given by the composition of all these functors but one. Since these are compatible with the functor , we obtain a map of Tannakian categories , hence a map of groups . The fiber functor of corresponds to a -torsor, the induced fiber functor of comes from the pushforward of that torsor from to , from which it follows that arises from twisting by that -torsor.
For (3), observe that the representation corresponds to a functor from the Tannakian category of representations of to Galois representations. This Tannakian category includes the standard representation of on , which may admit a bilinear form, as well as the representation . Call these objects and . The assumptions imply that this functor sends to the usual Galois representation . This Tannakian category admits a second fiber functor which takes each Galois representation to its associated filtered -module and takes the underlying vector space. The fiber functor takes to the filtered -module associated to , which is .
The relevant inner twist is the automorphism group of , which is clearly contained in the group of automorphisms of that are semiilinear over some automorphism of and respect the bilinear form if it appears. The group of automorphisms satisfying these semilinearity and bilinear form conditions is . After passing to an algebraically closed field, the two fiber functors are isomorphic, and so the automorphism group of becomes equal to and thus equal to . Thus the automorphism group of must be all of over the base field as well. ∎
5.9. Adjoint Hodge numbers
Definition 5.44.
(Adjoint Hodge numbers; see beginning of §10 of [47].)
Let be a reductive group, and suppose is a filtration on .
In this setting, we define the adjoint Hodge numbers as follows: The filtration on gives, by definition, a filtration on every representation of . We apply this to the adjoint representation of on , and call the adjoint Hodge numbers the dimensions of the associated graded of the resulting filtration666We use the notation , rather than the more common , to avoid conflict with the prime used throughout the proof.:
Definition 5.45.
For any real , we define the “sum of the topmost Hodge numbers” to be the continuous, piecewise-linear function satisfying and
for . (The sums are finite because only finitely many are nonzero.)
When we want to emphasize the dependence on the group , we will write and instead of and .
Definition 5.46.
(Uniform Hodge numbers.)
Recall notation from Section 5.6 and 5.7. In particular, is a reductive group over , and , is a form of ; the groups and act on the vector space . Let be a -filtration on .
After base change to , we obtain
the direct sum taken over all embeddings . (The subscript on is purely for notational convenience: each is a copy of .) Similarly, the group splits as a direct sum of copies of , and splits as a direct sum of filtered -vector spaces , indexed by .
For each , let be the adjoint Hodge numbers of the -filtration . We say that is uniform if the numbers are independent of . In this situation, we write the Hodge numbers and associated -function as
Example 5.47.
Let and . Any -module can be written as a direct sum , where is a vector space over the first factor , and a vector space over the second. A filtration on is determined by a filtration on , for . Then is uniform if and only if the adjoint Hodge numbers of on are the same as those of on .
Lemma 5.48.
In the setting of Definition 5.46, suppose is a -filtration on with uniform Hodge numbers, and let . We have
and
Proof.
The Hodge numbers can be computed after base change to . ∎
5.10. Galois representations
The next result is a form of the Faltings finiteness lemma. Compare also [47, Lemma 2.4], which applies when is a connected reductive group.
Lemma 5.49.
Let , and let , , , , be as in Section 5.6. The group has a standard embedding into , coming from its action by endomorphisms on a free -module of rank .
Fix a number field , a finite set of primes of , and an integer . There are, up to -conjugacy, only finitely many Galois representations
satisfying the following conditions.
-
•
The representation is semisimple (in the sense of Definition 5.36).
-
•
The representation is unramified outside .
-
•
The representation is pure of weight with integral Weil numbers: for every prime , the characteristic polynomial of has all roots algebraic integers of complex absolute value , where is the order of the residue field at .
Proof.
If is semisimple in the sense of Section 5.8, then it is semisimple in the usual sense (as a representation into ); see Definition 5.36 and Lemma 5.37.
Let be the kernel of
The representation restricts to a -valued representation
which is also semisimple (in the usual sense). Now is a connected reductive group, so we can use [47, Lemma 2.6] to conclude that there are only finitely many possibilities for .
For each fixed choice of an extension to a map compatible with the map is determined by its value at a system of coset representatives for , so the set of extensions forms an algebraic variety. To show this set is finite, it suffices to show that conjugation by 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 vanishes. This tangent space may be calculated to be of with coefficients in the Lie algebra of , which vanishes as it is the cohomology of a finite group with coefficients in a vector space of characteristic zero. ∎
Lemma 5.50.
Let be an algebraic group over a field whose identity component is reductive.
Let be an algebraic subgroup.
Then the set of Levi subgroups of containing and defined over forms finitely many orbits under conjugation by the -points of the centralizer .
Proof.
A Levi subgroup of is the centralizer of a cocharacter ; the subgroup contains if and only if takes values in . Since all maximal -split tori in are conjugate, we can assume that takes values in a fixed torus .
So we need to know that cocharacters of , taking values in a given torus , define only finitely many different Levi subgroups of . It is well-known that there are only finitely many possibilities for . But now we have
and since contains with finite index, there are only finitely many possibilities for . ∎
Remark 5.51.
Even if is a torus, and thus has only a single Levi subgroup, there can be multiple Levi subgroups of if the component group of acts nontrivially on the cocharacters of , because a component will be in the Levi if and only if it fixes the cocharacter . However, there will only be finitely many.
Lemma 5.52.
(only finitely many Levis to give the semisimplification)
Let
be a semisimple Galois representation. Then there exists a finite collection of Levi subgroups with the following property: for any Galois representation
whose semisimplification is -conjugate to , there exist , an from the finite collection, and a parabolic subgroup containing , such that takes values in , and the composite map
is .
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 -adic analogue, which we recall here. (See also [47, §9].)
The -adic Bakker-Tsimerman theorem is an unlikely intersection statement for the -adic period map attached to a Hodge–Deligne system. (For a detailed discussion of this period map, see [47, §3.3-3.4].) Suppose is a Hodge–Deligne system on , and let be a -adic residue disk. For all ,
is a filtered -module . The structure of -isocrystal means that, for all , the vector spaces are canonically identified, in such a way that is constant on . Using this identification, the filtration varies -adically in . A priori, this defines a map
into some flag variety, where is the rank of . We will show that the period map actually takes values in a smaller flag variety , where is the differential Galois group of in the sense of [36, §IV], i.e. the Tannakian group of in the Tannakian category of vector bundles with flat connection.
Lemma 6.1.
If is the differential Galois group of , then in fact the image of is contained in a single orbit of .
Proof.
(This argument was suggested by Sergey Gorchinskiy.)
Fix a basepoint , for some field with fixed embeddings into both and . Let be the flag variety classifying filtrations with the same dimensional data as the filtration on ; we have an isomorphism , for some parabolic subgroup of .
Consider the complex period map from the universal cover of to . Its image lands in a single orbit of . (See [28, III.A, item (ii) on p. 73].)
To transfer the result from to , we use Picard–Vessiot theory. (For an introduction to Picard–Vessiot theory, see [54].)
To the vector bundle with connection underlying , Picard–Vessiot theory attaches an (algebraic) -torsor , whose fiber over any -point classifies vector space isomorphisms
respecting the -structure on both sides (but not, in general, the filtrations). Furthermore, we obtain a -equivariant map
where a point of over gives an isomorphism , and we use that isomorphism to identify the filtration on with a point of the flag variety .
In the complex setting, fix a lift of to . For every lying over some point , integrating the connection from to gives an identification , which gives by definition a point of lying over . Thus, we obtain a complex-analytic map
lifting the projection .
Similarly, in the -adic setting, integration gives a rigid-analytic section
to the torsor .
By definition, the complex and -adic period maps are given by and , respectively.
The image of is contained in a single -orbit on , and the image of intersects every -orbit on . Thus, by -equivariance, the image of is itself contained in a single -orbit on , so the same is true of the image of . ∎
Lemma 6.2.
The construction above defines a -adic period map
where is a parabolic subgroup of .
Proof.
By Lemma 6.1, the -adic period map takes values in , where is the parabolic subgroup determined by the Hodge cocharacter . What remains is to show that is parabolic in .
We know that lies in the generic Mumford-Tate group, and hence normalizes [1, §5 Thm. 1]. Since the outer automorphism group of is finite and is connected, the cocharacter acts on by inner automorphisms. Thus the adjoint action of by conjugation on gives a homomorphism
to the group of inner automorphisms of . Raising to a power if necessary, we can lift to a cocharacter
Then the cocharater defines the parabolic subgroup . ∎
From now on, the parabolic subgroup will simply be called . We have defined a -adic period map, valued in . 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 is an algebraic subvariety, and
Then any irreducible component of is contained inside the preimage, in , of the complex points of a proper subvariety of .
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 and . Let be the image of under the analytic map . Let be an irreducible component of ; its image under is contained in some irreducible component of ; call this component .
On the other hand, we may as well assume ; otherwise is a point and there is nothing to prove. So we have the strict inequality
and we conclude by [3, Theorem 1.1]. ∎
Theorem 6.4.
(-adic Bakker–Tsimerman theorem). Let be a polarized, integral Hodge–Deligne system on a scheme smooth over , and let be the differential Galois group of . Let be the base change of to .
Choose a -adic residue disk , and a basepoint . By Lemma 6.2, these give rise to a period map
where is a parabolic subgroup of .
Suppose is a closed subscheme such that
Then is not Zariski-dense in .
Theorem 6.4 is stated over a general number field for generality, but we will apply it with .
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 arises as the cohomology (primitive in middle dimension) of a smooth, proper family of varieties.
-
•
The Zariski closure of the image of monodromy coincides with the full orthogonal or symplectic group.
-
•
The base field is .
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 , is covered by finitely many mod- residue disks; we choose such that the image of each of these residue disks under is contained in an affine subset of . There is no harm in replacing with one such mod- residue disk, and we will do so.
Now suppose is defined by equations , with each a regular function on . Let ; this is an element of , i.e. an element of the Tate algebra of the affinoid disk .
We want to show that the common vanishing locus of the functions on is contained in the zero-set of some regular function on . Now is the union of finitely many irreducible components, so it suffices to show that any one such irreducible component (call it ) is contained in the zero-set of some regular function on . As we are only interested in -points, we may as well assume that is nonempty.
Choose a basepoint and an isomorphism , and repeat the construction of Lemma 6.1 with respect to the basepoint (taking ). In particular, we have the algebraic Picard–Vessiot torsor under the group and the “Picard–Vessiot period map”
We also have maps
and
such that the complex and -adic period maps are given by
and
respectively.
Recall that both and are defined by integrating the connection on . In particular, with respect to some fixed coordinates on and local coordinates on , both and are given by the same power series (which then has positive radius of convergence both for the complex and the -adic metric).
Let us rephrase this, more precisely, in geometric terms. Let be the formal completion of at the point ; it is noncanonically isomorphic to (the formal spectrum of) a power series ring in variables over . The formal completions of and at are related to 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:
Over the formal scheme , the Picard–Vessiot torsor admits a canonical section , given (as in the proof of Lemma 6.1) by integrating the connection; the sections and over and , respectively, both restrict to when pulled back to . Let ; this is the restriction to of both the complex and -adic period maps. This is the sense in which and have the same power series.
Returning to the main argument, we will apply Lemma 6.3 over , then “pull back to and push forward to .”
Pulling back via gives a variety ; its inverse image in is cut out by analytic functions . (The functions have the same power series as the functions , regarded now as complex power series via the homomorphism .)
Now we apply the complex Bakker–Tsimerman theorem (Lemma 6.3) to the variety . As a result, we obtain some regular function on the scheme that vanishes (pointwise) on . By the locally analytic Nullstellensatz [35, §3.4], some power lies inside the ideal generated by the functions inside the ring of germs of analytic functions at .
Pulling back to , it is clear that also vanishes on the formal subscheme of .
By means of our chosen isomorphism , we obtain a regular function on that again vanishes on . This must in fact be defined over some finite extension of ; taking a norm, we may assume that it is in fact a regular function on . For simplicity, let us call this function .
At this point we have a regular function on , and we know that it vanishes on . To conclude, we need to show that vanishes on the rigid-analytic neighborhood of . To this end, recall that was the Tate algebra of the affinoid disk ; let be the ideal defining . “Clearing denominators” if necessary in the regular function , we may assume lies in the Tate algebra . Let denote the ideal of defining the point . We know that for every integer ; by Krull’s Intersection Theorem, we conclude that vanishes in , that is, that . Thus is contained in the vanishing locus of the algebraic function . ∎
6.2. Complex monodromy
Our next goal is Lemma 6.8, which shows that for any positive integer , we can construct our in such a way that its differential Galois group is strongly -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 be one of the algebraic groups , , or . Let be a subgroup of such that each of the coordinate projections (for ) is surjective.
Define a relation on the index set by declaring that if and only if the projection
is not surjective.
-
(1)
The relation is an equivalence relation.
-
(2)
If are a complete set of representatives for , then the map
is surjective with finite kernel.
Proof.
The group has a finite center; call it . For any two indices and (possibly equal), there are two possibilities for the image of the projection
Either the map is surjective, or its image is the graph of an automorphism of . In the former case, must surject (onto ) as well.
An easy calculation shows that is an equivalence relation. If is a complete system of representatives for , then repeated application of Goursat’s lemma shows that
is surjective with finite kernel. ∎
Definition 6.6.
Let be one of the algebraic groups , , or , and a subgroup of . For , let be the coordinate projection, and suppose that each is surjective. The index classes of are the equivalence classes of the relation of Lemma 6.5.
Let be a positive integer777not necessarily the same as the of Lemma 6.5. We say that is -balanced (as a subgroup of ) if its index classes are all of equal size, and there are at least of them.
Suppose now we are given a permutation of the index set . (In the sequel, will come from Frobenius.) We say is strongly -balanced (with respect to ) if it is -balanced, each orbit of on contains elements of at least of the index classes of , and preserves the partition of into index classes.
Finally, let be as in Section 5.6, and let be:
-
•
in the case , the kernel of the determinant map
-
•
in the case , the kernel of the similitude character
-
•
in the case , the intersection of the kernels of the determinant map and the similitude character .
Then is a form of , , or . We say that an algebraic subgroup is -balanced (resp. strongly -balanced with respect to ) if is -balanced (resp. strongly -balanced with respect to ) as a subgroup of , or .
Note that the condition of being strongly -balanced gets stronger as grows. We will later choose 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 .
Example 6.7.
Let be the diagonal, and let .
The index classes of are ; thus is -balanced for .
Let . Then is strongly -balanced with respect to only for .
(We thank the anonymous referee for this example.)
Lemma 6.8.
Recall notation from Section 5.1. Specifically, let , , , , , , , , be as in Section 5.5, and let be a place of over . Fix some embedding . Fix a natural number , and pick as in Corollary 4.10 (depending on ), with , or , and let be as in Sections 5.6 and 5.7. Let be the full -orbit containing , and let be the corresponding Hodge–Deligne system.
The Frobenius at , acts on the set through the diagonal action as an element of ; call this permutation .
Then the differential Galois group of (base-changed to ) is a strongly -balanced subgroup of , with respect to .
Proof.
This is little more than a restatement of Corollary 4.10.
After base change to , we have that is either , , or ; and splits as a product of copies of , indexed by pairs . We’ll write each of these direct factors of as .
The differential Galois group, after base change to , is the Zariski closure of the monodromy of the variation of Hodge structure. The variation of Hodge structure splits as the direct sum of . Clearly monodromy acts trivially on the set of pairs , and when is or , there is a bilinear pairing on . Thus, .
Now Corollary 4.10 implies that the geometric monodromy group of is all of . By symmetry under the action of , the same is true for all , so surjects onto each factor of . Thus, can talk about the relation and the index classes of Definition 6.6.
The action of respects the relation ; in particular the index classes are all of the same size, and Frobenius respects the relation . Corollary 4.10 shows that every -orbit in contains elements of at least index classes. Specifically, note that we chose such that this Galois action is unramified at , and thus the local Galois group is generated by Frobenius. So for any , the conclusion of Corollary 4.10 gives elements of the -orbit of in , such that the projection
contains , where is either , , or . It follows that is strongly -balanced with respect to . ∎
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 be a prime. A semisimple representation
of the global Galois group , crystalline at , such that the composition agrees with the usual action of Galois on , gives rise by -adic Hodge theory to an admissible filtered -module with -structure. Suppose another crystalline global representation is isomorphic to , and call the corresponding filtered -module . Then there is an isomorphism of filtered -modules
In particular, if , then there exists an automorphism such that
-
•
,
-
•
commutes with , and
-
•
.
Proof.
The fact that representations structure are sent to fitlered -modules with -structure is from Lemma 5.43.
The existence of isomorphisms is because functors of -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 -module with -structure , and another -module with -structure; suppose both and are semilinear over some .
Let be a subgroup of , strongly -balanced with respect to for some positive integer .
Suppose is uniform in the sense of Definition 5.46, and let be the adjoint Hodge numbers on . Suppose is a positive integer satisfying the following numerical condition.
-
•
(Numerical condition.)
Let be the flag variety parametrizing filtrations on that are conjugate to under the conjugation of . Then the filtrations such that is isomorphic to are of codimension at least in .
Proof.
By Lemma 7.1, the filtrations satisfying the condition described form at most one orbit under the action of on . We will show that any orbit of on has codimension at least . This is a question about the dimension of a variety over ; by passing to an extension, we may assume that is split, and acts by permuting the factors. Call the factors .
The Frobenius element gives a permutation of the index set , which we also call . Semilinearity over means that the map permutes the factors according to the permutation .
Let be an orbit of on the index set . By Definition 6.6, must contain elements of at least distinct index classes. Let be a system of representatives for the index classes appearing in . This must have at least elements, and in fact there is no harm in increasing so that exactly.
Now everything in sight splits as a direct product. Let ; define similarly. Let be the direct summand of corresponding to the -orbit .
Since the elements of belong to distinct index classes, the projection
has image a union of connected components of the target, so it is smooth with equidimensional fibers, and the same is true of
Let be projection to of ; this is the set of elements of that commute with , when is viewed as a direct summand of . Define similarly.
To finish the proof, it is enough to show that any orbit of on has codimension at least . By Lemma 5.33, applied with in place of , we have
Since is the projection of to , we deduce
On the other hand, for any reductive group and filtration on , corresponding to a parabolic subgroup , the sum of the adjoint Hodge numbers (Definition 5.44) is precisely . Apply this with ; since the Hodge numbers are uniform, we can compute the adjoint Hodge numbers on by Lemma 5.48; we find
The result follows. ∎
Theorem 7.3.
(Compare Theorem 8.17.)
Let be a variety over , let be a finite set of primes of , and let be a smooth model of over .
Let be a constant -algebra on , and let be one of , , or . Let be a polarized, integral, -module with -structure on , in the sense of Definition 5.27, having integral Frobenius eigenvalues (Def. 5.5). Suppose the Hodge numbers of are uniform in the sense of Definition 5.46, and let be the adjoint Hodge numbers on . Let , be as in Section 5.7.
Suppose there is a positive integer such that satisfies the following conditions.
-
•
(Big monodromy.) If is the differential Galois group of , then is strongly -balanced with respect to Frobenius. (The Frobenius is determined from the structure of ; see Section 5.7.)
-
•
(Numerical condition.)
Let be the subset of consisting of those for which the global Galois representation is semisimple. Then the image of is not Zariski dense in .
Proof.
For every , consider the semisimple global Galois representation . By Lemma 5.49, there are only finitely many possible isomorphism classes for the semisimple representation . So it is enough to show, for any fixed , that the set
is not Zariski dense in . By Lemma 7.1, it is enough, for each residue disk , to show that the set
is not Zariski dense in .
We are now in the setting of Theorem 6.4. For , the filtered -module is of the form , where is independent of . (This is a general property of -isocrystals; see remarks in the proof of Theorem 8.17.) The variation of with is classified by a -adic period map
where is the flag variety classifying -filtrations on .
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 . Recall the setup from Section 5: we have a smooth variety over 888 By Weil restriction, we will assume that is defined over , not a general number field; see the proof of Theorem 9.2. and a Hodge–Deligne system on an integral model of , satisfying certain conditions. Lemma 5.49 tells us that, as ranges over the integral points of , there are only finitely many possibilities, up to semisimplification, for the global Galois representation . We will use this to bound the integral points of .
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 , 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 -adic Hodge theory to filtered -modules, we have the following situation. We’re given a -module with varying filtration ; the variation of is described by a “monodromy group” (the differential Galois group attached to our Hodge–Deligne system). We will consider the associated graded of with respect to a -stable filtration – eventually we will take to be a semisimplification filtration on the global Galois representation. By Lemma 8.7 below, if comes from a filtration on the global representation, then the associated graded of with respect to is a balanced filtration (Definition 8.4). So we want to know, for how many choices of does there exist a -stable , such that the associated graded of with respect to lies in a given balanced isomorphism class? We will bound the dimension of such 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 -varieties , instead of varieties over an arbitrary number field. The natural generalization to representations of , with a number field, is false – for a counterexample, take a CM field, and a one-dimensional representation coming from a CM elliptic curve.
Lemma 8.1.
Let be a prime, and let be a representation of on a -vector space which is crystalline at , and such that at all primes outside of a finite set , the characteristic polynomial of Frobenius has algebraic coefficients and all roots rational -Weil numbers of weight .
Let be the filtered -vector space that is associated to by the -adic Hodge theory functor of [23, Expose III].
Then the average weight of the Hodge filtration on equals .
Proof.
This is [47, Lemma 2.9] applied to , which has no CM subfield; the condition that be a “friendly place” is automatically satisfied over . ∎
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 whose identity component is reductive. When is disconnected, recall the notions of “parabolic subgroup,” “filtration” and so forth from Section 5.8.
We use the following notation (consistent with [47]): is the parabolic subgroup of corresponding to the Hodge filtration , while corresponds to a semisimplification filtration . The group is a Levi subgroup associated to , corresponding to the associated graded of .
Fix , , and . In the (connected) reductive case [47, Lemma 11.2] defines a map from filtrations on to filtrations on ; we need to extend this result to non-connected groups as well. Recall (Section 5.8.2) that for any -filtration , there is a cocharacter defining . The substance of [47, Lemma 11.2] is that can be chosen to take values in . Projecting from to the Levi subgroup gives a filtration on , which is independent of the choice of .
Lemma 8.2.
Suppose is an algebraic group, whose identity component is reductive, over a field of characteristic zero. Let be a parabolic subgroup of , and a Levi subgroup associated to .
Fix a filtration on . Then there exists a cocharacter (with image in ) defining . Furthermore, if is the filtration on defined by the composite map
then is independent of the choice of .
Proof.
For connected reductive this is [47, Lemma 11.2].
Definition 8.3.
We need a generalization of the notion of “balanced filtration” from [47, §11.1, 11.4]. Given a group whose identity component is reductive, we define
where is the center of the identity component of . A cocharacter of defines a class by
for all . In other words: is an automorphism of , so it is of the form ; we choose so that , for all . We call the weight of .
Let , , be as above. Then the inclusion gives a map . Furthermore, the parabolic defines a preferred element of , the modular character , defined999Since the sign convention is important, we provide an example. If is the group of upper triangular matrices, then – identifying with the group of diagonal matrices – the character is given by as the inverse of the determinant of the adjoint representation of on the Lie algebra of .
Definition 8.4.
Suppose given an algebraic group whose identity component is reductive, a parabolic subgroup, and a Levi subgroup associated to . Let be a filtration on , given by a cocharacter .
-
•
We say that is balanced with respect to if , where is the cocharacter .
-
•
We say that is weakly balanced if for the modular character of .
-
•
We say that is semibalanced if , for the modular character of .
We say a -filtration is balanced (resp. weakly balanced, semibalanced) with respect to (or , or ) if the associated graded is so.
We remark that a -filtration is balanced with respect to if and only if the associated -filtration is balanced with respect to .
Remark 8.5.
Balanced implies weakly balanced because
this identity boils down to the fact that the center of acts trivially through the adjoint representation on the Lie algebra of .
Furthermore, weakly balanced implies semibalanced.
Example 8.6.
Let , acting on the space with standard basis vectors . Let be the filtration with
let be the subgroup of that stabilizes , and let be the Levi subgroup that fixes the subspaces , , and .
An element of is given by an action of on by some integral power (where is the coordinate on ); tensoring with , we can write elements of formally as scalar matrices , with . Similarly, an element of can be written as , where is understood to act on as .
The modular character takes to .
We will consider filtrations such that , , and . For such filtration , with corresponding cocharacter , we have .
Now let
for . We have , and
Thus is balanced if and only if . It is weakly balanced if and only if ; it is semibalanced if and only if .
Note that for generic in the Grassmannian, we have .
In general, the condition that be semibalanced is a strong condition on , only satisfied for 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 ; we know only that it is a -balanced subgroup; thus (after changing coefficients to arrange that ) we will work with a period domain of the form
for some index set of cardinality .
Lemma 8.7 tells us that the filtered -modules coming from global Galois representations are balanced; this amounts to a condition on the Hodge numbers averaged over the different indices. In the proof of Lemma 8.12, we will pass to a subset of cardinality , on which the filtration is semibalanced.
Lemma 8.7.
(Filtered -modules coming from global representations are balanced.)
Let be an algebraic group over whose identity component is reductive, and fix an embedding . Let
be a representation satisfying the hypotheses of Lemma 8.1. Suppose has image contained in some parabolic subgroup , and let be a Levi subgroup associated to .
Let be the filtered -module over that is associated to the local representation . Then is a -filtration on , and the associated graded is a balanced filtration on .
Proof.
(Compare [47, Prop. 10.6(b), §11.4, §11.6].)
That is a -filtration on is a consequence of the Tannakian formalism (see Lemma 5.43).
To see that is balanced, let be a cocharacter defining . It is enough to show that every character annihilating also kills .
For every such character , we will show that is pure of weight zero: the Frobenius eigenvalues at unramified primes are rational Weil numbers of weight zero.
Indeed, let as a representation of . Let be the semisimplification of a Frobenius element at acting on . Then lies in . Fix an eigenbasis of – with the last eigenvector lying in – and let be the subgroup of consisting of elements which have each element of this eigenbasis as eigenvectors. This is an algebraic subgroup of the torus with coordinates , hence is defined by relations for ; we have chosen indices so that are the eigenvalues on , and is the eigenvalue on .
Since acts trivially on , every element of whose eigenvalues are all equal acts by scalars on , hence lies in , and thus acts trivially on . It follows by restricting all relations to the subtorus where that there must be some relation with and . Applying this relation to the eigenvalues of , we see that
and so is pure of weight .
By Lemma 8.1, the weight of the corresponding filtered -module is also zero, which is what we needed to prove. ∎
Definition 8.8.
Let , , and be as in Section 5.6, and fix . A -bifiltered -module is a quadruple , with as in Section 5.6, and two -filtrations on , and a -semilinear endomorphism of respecting . In this setting, let and denote the parabolic subgroups of corresponding to and , respectively; to say that respects means that .
A graded -bifiltered -module 101010In the notation , the filtration comes after the semisimplification filtration to indicate that is logically prior to . is a quadruple , where is as in Notation 5.6, is a -filtration on with associated Levi subgroup , and is a filtration on .
We say that two graded -bifiltered -modules and are equivalent if they agree up to -conjugacy. More precisely, let and be the parabolic subgroups attached to and , and let and be Levi subgroups associated to and , respectively. Then and are equivalent if there exists satisfying the following conditions.
-
•
.
-
•
The filtrations and on agree.
-
•
The two elements and of project to the same element of .
There is an obvious functor
from -bifiltered -modules to graded -bifiltered -modules (with given by Definition 8.3). We say that two -bifiltered -modules are semisimply equivalent if the corresponding graded -bifiltered -modules are equivalent.
We say that a -filtered -module is of the semisimplicity type if there exists a -filtration on such that and are semisimply equivalent.
Remark 8.9.
To illustrate ideas, consider the case where and .
A -bifiltered -module comes (by -adic Hodge theory) from a filtered Galois representation (i.e. a Galois representation on a vector space , and a Galois-stable filtration on ).
A graded -bifiltered -module comes from the associated graded to a filtered Galois representation (i.e. the Galois representation on ).
Two graded -bifiltered -modules are equivalent if the corresponding Galois representations are isomorphic; two -filtered -modules are of the same semisimplicity type if the corresponding representations are isomorphic.
Lemma 8.10.
(Compare Lemma 7.1.)
Let be a prime. A representation 111111The restriction to , instead of an arbitrary number field , is for two reasons. First, in the general setting, a filtered -module would be semilinear over , and we have not defined filtered -modules with -structure in the semilinear setting; this restriction is inessential. Second, we will need to apply Lemma 8.1, and for that we need to have no CM subfield.
of the global Galois group , crystalline at , such that the composition agrees with the usual action of Galois on , gives rise by -adic Hodge theory to an admissible filtered -module with -structure. Suppose is semisimple. Suppose another crystalline global representation with the same composition with has semisimplification , and call the corresponding filtered -module . Then there exist -filtrations on and on such that and are semisimply equivalent.
Furthermore, we can take one of a list of finitely many candidates, depending only on , and the filtration is balanced with respect to .
Proof.
By restriction to and Lemma 5.43, a crystalline representation gives rise to a filtered -module with -structure , with semilinear over .
To say that is the semisimplification of means (see Definition 5.36 and Lemma 5.37) that there exist a parabolic subgroup with associated Levi , and such that takes values in , and takes values in , and the composition of with the quotient map is exactly .
Lemma 5.52 shows that, given , we can take to be one of finitely many possible subgroups. For each such , there are finitely many parabolic subgroups with as a Levi subgroup.
Since the kernel of is a unipotent group, the natural map is a bijection, so an inner form of is determined by the corresponding inner form of .
Now fix , , and . Let and be the inner twists of and arising by Lemma 5.43 from . Note that is a parabolic of and 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 and associated to . But since the projection of under agrees with , the inner twists of is again , and hence the inner twist of is again .
By Lemma 5.43, we find that (resp. ) must give rise to filtered -modules with -structure, such that the corresponding filtered -modules with -structure (obtained by functoriality via the quotient map ) are isomorphic.
But a filtered -module with -structure is precisely the same as a filtered -module with -structure, equipped with a filtration whose associated parabolic is . The filtered -modules with -structure arising from by this process is -conjugate to the filtered -module with -structure arising from directly by Lemma 5.43(2) and hence -conjugate to the filtered -module with -structure arising from by Lemma 5.43(1) since they correspond to isomorphic Galois representations into .
So take a filtration on attached to , and let be its image under this -conjugation. Then and are semisimply equivalent.
Finally, is balanced with respect to by Lemma 8.7. ∎
Remark 8.11.
Lemma 8.12.
Take notation as in Section 5.6, let be a positive integer, and suppose .
Let be a parabolic subgroup of , corresponding to a -filtration , and let be a Levi subgroup associated to .
Let be another parabolic subgroup of , corresponding to some filtration , so that parametrizes filtrations that are -conjugate to . Suppose is uniform in the sense of Definition 5.46, and let be the adjoint Hodge numbers on . Let be the dimension of a maximal torus in . Suppose is a positive integer such that:
-
•
(First numerical condition.)
and
-
•
(Second numerical condition.)
Then for any semibalanced filtration on , the set of filtrations on that are -conjugate to and satisfy is of codimension at least in .
Proof.
This is essentially [47, Prop. 11.3], applied to . Note that the Hodge numbers and the function used in [47] are and . By Lemma 5.48, they are related to and by
and
Similarly, the dimension of a maximal torus in is .
Two small modifications need to be made to the proof in [47]. First, we have replaced with . To get this stronger bound, replace the final inequality of [47, Equation 11.15] with
(Here , the dimension of an associated Levi subgroup, and 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, 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 is a maximal torus contained in , is the set of roots of on , is the set of roots of on , and is a cocharacter defining the parabolic subgroup . In our context, [47, Equation 11.14] is replaced with the inequality
and [47, Equation 11.16] becomes
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 contained in a connected reductive group (e.g. ), but here is semilinear, so it is contained in , but not in .
Lemma 8.13.
Let be a unipotent algebraic group over a field of characteristic zero, and an automorphism of , such that . Suppose is such that
Then there exists such that
Proof.
By induction on ; reduce to the case where is abelian. ∎
Lemma 8.14.
Let be an algebraic group whose identity component is reductive. Let a parabolic subgroup of and an associated Levi subgroup. Suppose is semisimple and normalizes .
Then is -conjugate to an element that normalizes .
Proof.
Suppose . Since normalizes , . Since is semisimple, we may conjugate by an element of to assure that .
Since normalizes , and all Levi subgroups in are -conjugate, we can write for some . Multiplying by an element of from the left, we may assume that is unipotent. Since , it will suffice to show that is -conjugate to .
Now is an element of that normalizes , so
We can apply Lemma 8.13, with
to conclude that
for some . Then normalizes , and we are done. ∎
Lemma 8.15.
(Compare [47, Equation 11.18].)
Assume we are in the setting of Section 5.6. Fix a -bifiltered -module , and another -module with -structure; suppose both and are semilinear over some .
Consider all pairs , where is a filtration on , is an associated Levi subgroup, and is equivalent to , in the sense of Definition 8.8. The set of such pairs has dimension at most .
(Recall that any element of an algebraic group has a unique Jordan decomposition , where is semisimple, is unipotent, and and commute.)
Proof.
This is a question about the dimension of a variety, so we can pass to a finite extension of . In particular, we may assume and acts by permutation on the factors of . We index the factors through , and we regard as a permutation of . Write for the -th factor of , and write for the corresponding factor of (it satisfies ). The semilinearity condition means that decomposes as a sum of maps
It follows from functoriality of the Jordan decomposition that, if is equivalent to , then is equivalent to . Thus, as in [47, §11.6, third paragraph], we can assume is semisimple. Otherwise, replacing with will only increase the dimension of the set of for which is equivalent to .
A filtration on (which is the same as a filtration on ) decomposes as the product of filtrations on . If is -stable, then determines . So to specify it is enough to specify , for a single in each -orbit.
Since everything in sight splits as a direct product over -orbits, we may as well restrict attention to a single -orbit; call it . A -stable filtration on is uniquely determined by a -stable filtration on . As in [47, §11.6, fourth and fifth paragraphs], the dimension of the set of such filtrations is exactly
where , and is defined similarly.
Lemma 8.16.
(Compare Lemma 7.2.)
Assume we are in the setting of Section 5.6 and Section 5.7. Fix a -bifiltered -module , and another -module with -structure; suppose both and are semilinear over some , and is balanced with respect to .
Let be a subgroup of , strongly -balanced with respect to for some positive integer .
Suppose is uniform in the sense of Definition 5.46, and let be the adjoint Hodge numbers on . Let be the dimension of a maximal torus in . Suppose is a positive integer satisfying the following numerical conditions.
-
•
(First numerical condition.)
and
-
•
(Second numerical condition.)
Let be the flag variety parametrizing filtrations on that are conjugate to under the conjugation of . Then the filtrations such that is of semisimplicity type are of codimension at least in .
Proof.
Let be the parabolic of , and an associated Levi subgroup, associated to a hypothetical semisimplification filtration . There are only finitely many possibilities for the parabolic group , up to -conjugacy, so we may as well as fix a single .
The dimension in question can be calculated after base change to an extension of , so we can assume that . As in the proof of Lemma 8.15, we’ll call the factors . The whole setup factors over these factors, so we can write as the direct sum of parabolics , and so forth. Again, gives a permutation of the index set , which we also call . Semilinearity over means that the map permutes the factors according to the permutation .
The strategy is as follows. We want to apply a result like Lemma 8.12, but we don’t have full monodromy group . Instead, we know the group is -balanced; we’ll project onto of the factors, so that projects onto the full group . We need the projection of to the factors to be semibalanced, and we can apply Lemmas 8.12 and 8.15 to the group to finish.
For any subset of the index set , let ; define , , and similarly. Any filtrations and on can be written as products of factors and , so we can define and .
We claim that, for any and any that is balanced with respect to , we can find satisfying the following properties.
-
•
consists of exactly elements, from distinct index classes for .
-
•
The elements of belong to a single orbit of on .
-
•
is semibalanced with respect to .
To see this, let be a cocharacter defining as in Definition 8.4. We can write as a product of over . Similarly, the character splits over the factors, and we have:
Hence there exists an orbit of on the factors such that is semibalanced. Since is strongly -balanced, we can find a subset
of the index set such that , the elements of belong to distinct index classes, and is semibalanced. Since the elements of belong to distinct index classes, the projection
has image a union of connected components of the target, so it is smooth with equidimensional fibers, and the same is true of
We want to estimate the codimension in of the set of such that is semisimply equivalent to , for some choice of . Consider the projection
By Lemma 8.15 applied to , the set of pairs such that is equivalent to has dimension bounded by ; this dimension is at most by Lemma 5.33.
Now fix . By Lemma 8.12 applied to , the set of such that
has codimension at least among all -filtrations . The map from -filtrations in a given -conjugacy class to -filtrations is smooth with equidimensional fibers, so the set of such that
again has codimension at least .
It follows that the set of satisfying the desired condition has codimension at least . ∎
Theorem 8.17.
(Basic theorem giving non-density of integral points. Compare Theorem 7.3.)
Let be a variety over , let be a finite set of primes of , and let be a smooth model of over .
Let be a constant -algebra on , and let be one of , , or . Let be a polarized, integral -module with -structure, in the sense of Definition 5.27, having integral Frobenius eigenvalues (Def. 5.5). Suppose the Hodge numbers of are uniform in the sense of Definition 5.46, and let be the adjoint Hodge numbers on . Let be the dimension of a maximal torus in . Let be as in Definition 5.2. Let and be as in Section 5.7.
Suppose there is a positive integer such that satisfies the following conditions.
-
•
(Big monodromy.) If is the differential Galois group of , then is strongly -balanced with respect to Frobenius. (The Frobenius is determined from the structure of ; see Section 5.7.)
-
•
(First numerical condition.)
and
-
•
(Second numerical condition.)
Then the image of is not Zariski dense in .
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 , the fiber of the -adic étale local system is a global Galois representation valued in , having good reduction outside 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 . So it is enough to show, for any fixed , that the set
is not Zariski dense in .
Lemma 8.10 gives a finite list of semisimplicity types such that, for every , the filtered -module belongs to one of them. Let be a mod- residue disk, and fix a semisimplicity type . By Lemma 8.10, is balanced with respect to . It is enough to show that the set
is not Zariski dense in .
We are now in the setting of Theorem 6.4. For , the filtered -module is of the form , where is independent of . (This is a property of -isocrystals in general; it reflects the fact that if is the crystalline cohomology of a scheme over , then can be recovered from its reduction modulo . See [47, Section 3.3] for further discussion.) The variation of with is classified by a -adic period map
where is the flag variety classifying -filtrations on .
9. Proof of main theorem
Definition 9.1.
Let be an abelian variety over a number field , and a hypersurface. We say that is primitive if it is not invariant under translation by any .
Recall the sequence defined in Theorem 3.5.
Theorem 9.2.
Let be an abelian variety of dimension over a number field . Let be a finite set of primes of including all the places of bad reduction for .
Let be an ample class in the Néron-Severi group of , and let .
Suppose that either or and is not for any .
Then, up to translation there are only finitely many smooth primitive hypersurfaces representing , defined over with good reduction outside .
Proof.
Choose a smooth proper model of over . Working over , let be the Hilbert scheme of smooth hypersurfaces of class , and let be the universal family over . Let
be the set of primitive hypersurfaces in (Definition 9.1). (Note that, by definition, a hypersurface defined over with good reduction outside extends to an -point of . However, we are interested in the set of hypersurfaces which are primitive over , not the set of -points of the scheme of primitive hypersurfaces, which may be smaller.) The group acts on by translation; we need to show that is contained in the union of finitely many orbits of .
Theorem 8.17 only works over ; we’ll pass from to by Weil restriction. (See [12, Theorem 7.6.4] for the notion of Weil restriction relative to a finite étale extension of rings.) Enlarging if necessary, we can arrange that is finite étale over , where is a finite set of primes of . Over the Weil restriction
we have the family
and the fibers of this family are the same as the fibers of the original family . More precisely, the sets and
are in canonical bijection. Let
be the subset corresponding to under this bijection. For any , call the corresponding point of . Then we have a canonical isomorphism of schemes
here is an -scheme, is a -scheme, and the structure maps are related by the fact that the diagram
(7) |
commutes.
Let be an irreducible component of the Zariski closure of in , and choose a resolution of singularities . Let be the pullback of our family of hypersurfaces to . Note that is a hypersuface in .
We will apply Theorem 8.17 to show that is the translate of a constant hypersurface by a section . In this case, all the points of lying in would correspond to points in the orbit of .
The map is proper, so it spreads out to a proper map
possibly after further enlarging the finite sets and . Possibly after further enlargement of and , the family spreads out to a smooth proper family , which is a hypersurface in . Now, by properness, every -integral point of lifts to an -integral point of , so by construction the -integral points are Zariski dense in .
Fix some . Assume that is not equal to the translate of a constant hypersurface by a section . Let be the generic point of . The fibers of over points in a dense subset of are primitive hypersurfaces, so is not translation-invariant by any nonzero element of . Let be the Tannakian group of the constant sheaf on , and let be the commutator subgroup of the identity component of . By Theorem 3.5, because of our assumptions on and , we have or , acting by its standard representation. Furthermore the case occurs exactly when is equal to a translate of and is even and the case occurs exactly when is equal to a translate of and is odd. In particular, is a simple algebraic group acting by an irreducible representation.
We can now apply Corollary 4.10 for a positive integer to be chosen shortly. This gives us, for any fixed , the existence of an embedding and a torsion character of , depending on , satisfying the big monodromy condition that is needed for Lemma 6.8. (Note a small subtlety here. The Tannakian monodromy groups , , or are calculated over an algebraically closed field , but to apply them we need them over a smaller field. Thus we use the straightforward fact that any form of or , together with its standard representation, over any field is the split form, or in the 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 on attached to the orbit containing and the family . By Lemma 6.8, the differential Galois group of is a strongly -balanced subgroup of . Corollary 4.10 gives vanishing of cohomology outside degree , so the Hodge numbers of are given by Lemma 3.13.
We apply Theorem 8.17 to . That the eigenvalues of Frobenius on are integral Weil numbers is a consequence of the Weil conjectures, since comes from geometry; the polarization and integral structure are given in Lemma 5.29. Lemma 5.30 gives an -structure over , with chosen as in Lemma 5.30. (Note that this matches the group by our earlier calculation, because is equal to a translate of if and only if is equal to a translate of .) Lemma 6.8 gives that is strongly -balanced with respect to Frobenius. The Hodge numbers of are uniform because we have explicitly computed them, independently of , in Lemma 3.13.
The numerical conditions in the hypothesis of Theorem 8.17 will hold for big enough . To verify this, since every term except is independent of (because the Hodge numbers calculated in Lemma 3.13 are independent of and thus of ), it suffices to have
By Lemma A.1, we have
This implies the first condition because . It implies the second condition because, by the definition of , is strictly increasing for and .
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 is equal to the translate of a constant hypersurface by a section . It follows that every point of that is contained in corresponds to a hypersurface that is a translate of (by the value of the section on the inverse image of that point). Since there are finitely many irreducible components of the Zariski closure, every point of 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 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 , , , be as in Theorem 9.2.
Suppose that either or and is not a multiple of for any
There are only finitely many smooth hypersurfaces representing , with good reduction outside , up to translation.
Proof.
Any hypersurface is of the form , where is an isogeny defined over , and is a primitive hypersurface defined over . In this case is the pullback of an ample class along , and we have . Hence is bounded, so there are only finitely many possibilities for . For each one there is a unique with . Furthermore, for any . We conclude by applying Theorem 9.2 to each . ∎
Theorem 9.4.
Suppose . Fix an ample class in the Néron-Severi group of . There are only finitely many smooth hypersurfaces representing , with good reduction outside , up to translation.
Proof.
This is one case of Corollary 9.3 . ∎
Theorem 9.5.
Suppose . Fix an ample class in the Néron-Severi group of . Assume that the intersection number is not divisible by for any . There are only finitely many smooth hypersurfaces representing , with good reduction outside , up to translation.
Proof.
This is one case of Corollary 9.3, once we cancel the factor of from the denominators. ∎
Theorem 9.6.
Suppose . Fix an ample class in the Néron-Severi group of . There are only finitely many smooth curves representing , with good reduction outside , 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 . Consider a fixed ; enlarging if necessary, we may assume that is nonempty, and choose a basepoint . It is enough to show that there are only finitely many maps taking to the origin of , for which the image of represents the class .
By the Albanese property, pointed maps are in bijection with maps of abelian varieties ; the set of such maps forms a finitely generated free abelian group. Fix an ample class on ; define the degree of any map by
This intersection number is a positive definite quadratic form on , so there are only finitely many maps of given degree, and any map representing has degree . ∎
Appendix A Verifying the numerical conditions
The goal of this section is to verify the two numerical conditions of Theorem 8.17 for 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 an abelian variety of dimension , a degree hypersurface in , a finite-order character of such that vanishes for , the th Hodge number of the Hodge filtration on is , where 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 form a Hodge–Deligne system with -structure, for one of , , and . 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 .
Lemma A.1.
Let be a smooth hypersurface in an -dimensional abelian variety , let be a finite-order character of such that vanishes for , and let be the Hodge structure on , regarded as a filtered vector space with -structure, for one of , , . Let be the adjoint Hodge numbers (Definition 5.44), and let be the dimension of a maximal torus in .
If , we have .
The relevance of this inequality is that by the definition of (Definition 5.45), so by the monotonicity of this implies
which for 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 , we have
Thus it is enough to show that
(8) |
In each case, we will calculate and in terms of Eulerian numbers , and then prove (8) by using the following inequality of Eulerian numbers, to be established later:
(9) |
We adopt the convention that when .
-
•
If then
so
-
•
If then
so
-
•
If then
(The formula for in the case holds only for even-dimensional orthogonal groups, but is even.) Thus
We now prove (9). It is known that, as grows large, the Eulerian numbers approximate a normal distribution with variance . This purely qualitative result implies that (9) holds for sufficiently large .
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
this is normalized so that .
Now we’ll prove that for all ; 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 is log-concave and satisfies .
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 is
Proof.
This is [47, Eqn. 10.10]. ∎
Lemma A.4.
Suppose . Then for all we have
Proof.
Because , by symmetry, we have
so if
we must have for some .
But then by log-concavity and the fact that , we have for all , and thus in particular for all , so
∎
Lemma A.5.
If then .
Proof.
Lemma A.6.
If then .
Proof.
For , we can prove this by computation. For even this follows immediately from the symmetry , so it suffices to check , 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 for the Eulerian numbers. We adopt the convention that unless ; similarly, we take to vanish whenever is positive but or .
Recall from the introduction that is the sequence satisfying
Proposition B.1 (Proposition 3.15).
Let and be integers. Suppose that there exists a natural number , function from the integers to the natural numbers and an integer . Write , and suppose that . Suppose the equation
(10) |
is satisfied for all . Then we have one of the cases
-
(1)
and
-
(2)
and for some
-
(3)
and for some .
A tuple satisfying the conditions of the proposition will be called a solution.
We first handle the cases directly, then give a general argument that handles cases . Thus Proposition B.1 will follow immediately once we have proven Lemmas B.4 (the case), B.8 (the case), B.9 (the case), B.18, B.21, B.25, B.27, and B.31 (which collectively handle the case).
We can assume without loss of generality that .
Let and be the functions satisfying , and maximizing, respectively, minimizing .
Lemma B.2.
There is a unique such that for all , for all , and for .
Similarly, there is a unique such that for all , for all and for .
Furthermore .
Proof.
We take to be the largest such that . The last two conditions are then obvious and the fact that for follows by minimality – if it were not so, we could increase by , reduce by , and thereby reduce by .
We take to be the least such that , and make a symmetrical argument.
Finally, for contradiction, assume . Then
a contradiction. ∎
Lemma B.3.
We have
(11) |
Note that all the terms on the left side of this equation are nonnegative.
Proof.
Because the function is supported on ranging from to , we must have
∎
B.1. The case
Lemma B.4.
Suppose . Then we must have . Furthermore if then .
Proof.
In (11), because the summands on the left side are nonnegative, one must be and the others must vanish. The one that is can only be the summand associated to or as the other summands are integer multiples of something at least . By symmetry, we may assume the comes from . Because the last summand vanishes, we must have .
This gives unless or and
Then the only possible solutions to and are and . These have equal to and respectively, so we must have . This implies
and thus
which implies and thus
In the case, we have , as desired.
∎
Lemma B.5.
Suppose . Then we must have for and unless and .
Proof.
In (11), either two terms are and the rest zero or one term is and the rest are zero. In the first case, since the only terms that can be are and , implying in particular that we have the stated conclusion. So it suffices to eliminate the case that one term is . The only possibilities are , and the last term if . By symmetry, we are reduced to eliminating , , and the final term. In the first two cases we have and in the last case we have .
If for all except , and , then there are three possibilities for : must equal or . Using , this gives
and this implies
so and giving .
If for all except , then whenever and . Hence the left side of (3) is nonzero only when . This contradicts the fact that is nonzero for of both parities.
If for all except and , then there are three possibilities for : must be or . This gives
Because this implies and thus which implies and thus , . ∎
Lemma B.6.
Suppose . Then unless and , we have
Proof.
By Lemma B.5, we have for and . This means there are four possibilities for : must equal , , , or . Using , and , we get
so in other words we have
which dividing by and multiplying by is
which is exactly the stated Diophantine equation. ∎
Lemma B.7.
The positive integer solutions to with have the form for some .
Proof.
Let and be positive integers with and . We will show that there exists such that and . By induction, it suffices to prove that either or that there exists an integer with , , and solving the same equation, as if then .
To do this, let . Then because we can rewrite the equation as
as is a solution then is a solution as well. Furthermore, if then so , and we must have because if we have
so if we always have such a . On the other hand, if then so, because , we have , the base case. ∎
Lemma B.8.
Suppose . Then either and or and for some .
Proof.
Lemma B.9.
There are no solutions for .
Proof.
We first consider the contribution to (11) from . Because and this contribution is at most , we can only have or , or . Let us eliminate the cases first.
In the case we have unless or . Thus we have four possibilities for – we must have or . This gives
Combining the first and last equations with , we see that . Dividing the second equation by the first, we get
which implies , a contradiction.
In the case, one more term must be , which is or . Without loss of generality, it is . Then we have unless and . Then the possible values of are , , , , and . This gives (ignoring factors of the form or )
Dividing the third equation by the first, we get . Thus by the fourth equation , and by the first equation . Then the second equation gives , which is false, so there are no solutions.
We now handle the case . In this case, because the total sum of (11) is , and only the two terms can contribute a , we must have one term contributing or one and one . There are four terms that might contribute : , , , and , and four that might contribute : , , , and . This gives a total of possibilities for , or up to symmetry: we may assume or .
The case is easy to eliminate as it implies that the left side of (3) is nonvanishing only in a single residue class mod , but we know that the right side does not have that possibility.
The case gives us four possibilities for , implying
The first and fourth equations gives , which implies that or . Dividing the third equation by the fourth, we then obtain
whose unique solution is , a contradiction.
The case gives us four possibilities for , implying
By the first and fourth equations we have which implies or . Dividing the second equation by the first we get so , a contradiction.
The case gives us four possibilities for , implying
By the second and third equations, we have which again implies . But then which means , a contradiction.
The case has six possibilities for , implying
By the first and fourth equations, we have which implies . But then since we have by the third and fourth equations
a contradiction. ∎
B.2. The case : general setup
We’ll write for the rest of this section. The key equality for gives
(12) |
For any as in Equation 10 we’ll write
Now the key equality (10) becomes
(13) |
We’re going to get a contradiction from combinatorial considerations involving the terms associated to small in Equation (3).
By abuse of notation, we’ll let denote the function taking the value on and zero elsewhere; so any function is a linear combination of the elementary functions .
There are at most two functions that contribute to the case in Equation (13). These are
and
We compute
the case of Equation (13) gives
Similarly, there are at most five nonzero terms in the case Equation (13):
We have the following equalities.
Equation (13) gives
(14) |
We conclude with a lemma that will be useful at several points in the argument.
Lemma B.10.
We have
and
In particular, if
let be the real roots of
Then we have
and
Proof.
The first equality is the case of Equation (13). The second follows from the explicit formulas for the two quotients .
The bounds in terms of and follow from the first two inequalities. ∎
B.3. The case , with 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.
Lemma B.11.
If and
(15) |
then .
Proof.
We will focus on proving a lower bound for the left side, assuming , which which will give a contradiction for sufficiently large .
If , then , which gives the bound
(17) |
Assuming , we find
(18) |
By combining this with (17), we conclude that
(19) |
which combines with (16) to give
which is impossible for . (See the discussion in Appendix C, and bound-A3a in the Python code.)
For smaller , we do a more precise version of the above analysis. For , we will improve on the bound (15) by using Lemma B.10. For , we will also need to replace the bound (17) by a slightly weaker one without the assumption . With these modifications, the argument will work for from to .
To apply Lemma B.10, we must give an upper bound for , showing that the two roots are far apart. Consider
We have
so (still assuming ) we conclude that
For we have
(see bound-A3b in the Python code). Thus we have
so by Lemma B.10,
(20) |
When we have from above (19) that
so
which does not hold for any .
Finally when we use the estimate
to deduce
Now (LABEL:eqn-line415) implies
combining this with (20), we get
which is not true for .
Thus we arrive at a contradiction in every case. ∎
Lemma B.12.
If
then .
Proof.
Assume , and consider
We will show that is too big.
First, note that we must have by Lemma B.3.
Now
where
Using the bounds and
we deduce that
The inequality
(see Appendix C, and bound-A3c in the Python code) gives a contradiction. ∎
Lemma B.13.
If and
then
Furthermore, if then we have the stronger bound
Proof.
Let
The result will follow from
We have
On the other hand,
We know by Lemma B.11, so . Thus we can estimate
The first bound follows.
To get the second bound, note that if then
so
∎
Lemma B.14.
If and
then
Furthermore, if then we have the stronger bound
Proof.
Lemma B.15.
If
then .
Proof.
Assume . We’ll estimate each , and show that for large their sum is too small.
We have the following bounds. By Lemma B.12, we have , so
Lemma B.11 tells us that (if ), so
By Lemma B.13, we know that
so
and
Finally, from Lemma B.14, we find
so
These five bounds combine (see Appendix C, and bound-A3d in the Python code) to give
for , a contradiction. ∎
Lemma B.16.
If and
then and for .
Proof.
We will use this stronger bound to redo the estimates in Lemma B.15.
We have
and
We use the same bounds on
as in Lemma B.15; we conclude (see bound-A3e in the Python code) that
for .
Finally, suppose . Then since , we have the much stronger bound , which implies . We deduce as above
and arrive at a contradiction as before. (See bound-A3f in the Python code.) ∎
Lemma B.17.
If and
then .
Proof.
Suppose . Again, we’ll redo the estimates in Lemma B.15.
Lemma B.18.
We cannot have and
B.4. The case , with
We’ll treat the case where next. By Lemma B.3, if , then .
Lemma B.19.
If and then one of the two ratios
is less than 1, and the other is greater than .
Lemma B.20.
If and
then .
If and
then .
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 and , and consider
From
we deduce that , so by integrality , so
We also know that and , so
For , this contradicts
(See Appendix C and bound-A4a in the Python code.) ∎
Lemma B.21.
For we cannot have .
B.5. The case , with big and .
Lemma B.22.
Suppose and
Then
In particular, if then
Proof.
We have
The “in particular” follows from the fact that is an integer. ∎
Lemma B.23.
Suppose , , and
Then we cannot have simultaneously and .
Proof.
Suppose and ; the first condition implies and . Consider
We have
By Lemma B.10 we have
Thus we obtain
Using , , and , the three fractional factors on the right can be bounded below by , , and , respectively.
On the other hand, as soon as is nonzero, we have
We conclude that
so
As a function of , this right-hand side is minimized when , so we have
This contradicts the inequality
which is valid for all . (See Appendix C and bound-A5a in the Python code.) ∎
Lemma B.24.
If , , , and
then and .
Proof.
Since , we have
Assuming , let’s look at . We have
We want to compare this with the inequality
which is valid for all . (See Appendix C and bound-A5b in the Python code.)
We have certainly
So we will be done if .
But as in the proof of Lemma B.22, we find (using now in place of a weaker bound) that
and so in particular whenever . ∎
Lemma B.25.
If , , and
then .
In particular, there must be some with .
Let be the highest Hodge weight below . In other words, take minimal such that . (Such exists by Lemma B.25.) For example, if then we must have , so .
B.5.1. Case: big and .
Recall the definitions of from above ().
Lemma B.26.
If ,
and , then .
Proof.
Suppose for a contradiction , and consider
We have
From
and , we find that
In particular, since , we have , so certainly
We conclude that
This contradicts the inequality
which is valid for all . (See Appendix C and bound-A5c in the Python code.) ∎
Lemma B.27.
We cannot have , ,
and .
Proof.
Assume the stated conditions hold. We will bound the five ratios , and show that their sum is too small. Lemmas B.25 and B.26 tell us that . Hence, .
We have
Since
and , we have
Let
Dividing the bound
by
we obtain
so
Finally, let
We have
so
We conclude that
which contradicts the inequality
valid for all . (See Appendix C and bound-A5d in the Python code.) ∎
B.5.2. Case: big and .
Lemma B.28.
Suppose , , and
We have
for , and
Proof.
We have already seen in Lemma B.25 that .
For , there is only one nonzero term in Equation (3); for , there are only two. ∎
Lemma B.29.
Suppose , , , and
Then
Proof.
Let
We have
∎
Lemma B.30.
Suppose , , , and
Then and
Proof.
Lemma B.31.
We cannot have , , , and .
Proof.
Assume we had a solution satisfying the given conditions. By Lemma B.30 we know that and
This implies
by the inequality
which holds for all . (See Appendix C and bound-A5e in the Python code.) Thus, since , we have
Finally, taking
we find that
This contradicts the inequality
which holds for all . (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 , 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 , using the asymptotic approximation , which is valid for fixed and large .
Recall our convention that if .
Lemma C.1.
For all and , we have
Proof.
Recall ([52, §1.3]) that the Eulerian number counts the number of permutations of with exactly ascents.
If we label the integers to with labels through , then we get a permutation with at most ascents by giving all the numbers of label in decreasing order, then all the numbers of label in decreasing order, and so on. Every permutation with at most ascents arises in this way; this proves the right-hand inequality.
If a permutation constructed this way has fewer than ascents, then there must exist adjacent labels and where all the numbers with label are less than all the numbers with label . If that happens, we can record a number from to which is the number of elements in the sequence with label at most , then subtract one from the labeling of everything with label greater than . There are possibilities of this new data, and we can recover the original labeling by adding to the label of everything that comes after the first elements in the sequence. So the number of labelings giving permutations with fewer than ascents is at most . This proves the left-hand inequality. ∎
Lemma C.2.
For all and , we have
Proof.
Follows from
∎
The two bounds given patch well enough for our modest needs: if , then the bound in Lemma C.1 is close to , 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 and
then
Proof.
Lemma C.4.
For arbitrary and we have
Proof.
We’ll assume ; for smaller there are only finitely many cases, which can be checked by hand. (See bound-B in the Python code.)
We can also assume ; otherwise, the right-hand side is zero. We’ll split into two cases: either or . Our hypothesis on guarantees that there is some with ; a fortiori, one of the two cases always holds.
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 , followed case-by-case verification for small . 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 , we have
Proof.
Lemma C.3 gives us
and
provided . Hence we will be done as soon as we can show that
which happens for . The cases 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 -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 . 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 -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 -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 -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 -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. -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 -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. -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 -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).