Compact subvarieties of the moduli space of complex abelian varieties
Abstract.
We determine the maximal dimension of compact subvarieties of , the moduli space of complex principally polarized abelian varieties of dimension , and the maximal dimension of a compact subvariety through a very general point of . This also allows us to draw some conclusions for compact subvarieties of the moduli space of complex curves of compact type.
1. Introduction
Given an irreducible quasi-projective variety , it is natural to ask how far is from being affine or projective. Perhaps the simplest way to measure this is the maximal dimension of a compact subvariety of , which we denote , though perhaps the maximal dimension of a compact subvariety of passing through a very general point of , which we denote , is more natural. We observe that . Clearly, if is affine, and if is projective.
In this paper we determine and for the moduli space of complex principally polarized abelian varieties (ppav) of dimension , and we draw some consequences for the moduli of complex curves of compact type, and for the locus of indecomposable abelian varieties.
Throughout the paper we work over .
1.1. Compact subvarieties through a very general point of
If is a projective compactification of , such that the codimension of the boundary , then by embedding and recursively choosing very general hypersurfaces such that for any , it follows that is a compact subvariety, which moreover can be chosen to go through any given finite collection of points of . Thus the existence of such a compactification implies . We note, however, that in general there is no implication going the other way, as is easily seen by considering for various .
The Satake compactification of has boundary of codimension , which by the above implies . Our first result is that this is sharp.
Theorem A.
For a very general point, the maximal dimension of a compact subvariety of containing is equal to ; that is,
What we prove is in fact a more precise statement, showing that the maximal dimension of a compact subvariety containing a Hodge-generic point (which, as we recall below, is equivalent to saying that does not belong to any of the countably many proper Shimura subvarieties of , see also the end of Section 2.2) is equal to . This will be the special easier case of 3.4, which establishes this statement for a Hodge-generic point in any non-compact Shimura subvariety .
1.2. Compact subvarieties of
Turning to the question of maximal dimension of all compact subvarieties, the best known upper bound for is the 20 year old result of Keel and Sadun [KS03], who proved for any . This was conjectured by Oort, and underscored the difference between and , as in finite characteristic there does exist a complete subvariety of of codimension — the locus of ppav whose subscheme of -torsion points is supported at zero.
We determine precisely the maximal dimension of compact subvarieties of , for all genera.
Theorem B.
The maximal dimension of a compact subvariety of is
In Table 1 we give our results and the bound of Keel-Sadun, for comparison, in some small genera, and also for to underscore the difference of the growth rates.
What we actually prove is much stronger: we show that any compact subvariety of of dimension at least is contained in the product of a -dimensional compact subvariety of , for some , and a compact Shimura subvariety of (see 2.9 and 3.4). We then determine the maximal dimension of compact Shimura subvarieties of . In fact our proof and analyses of compact Shimura subvarieties of allows us to describe all maximal-dimensional compact subvarieties of — they are either Hodge-generic, or compact Shimura subvarieties of a specific type. See 5.10 for a precise statement, and 5.11 for very explicit examples of maximal-dimensional compact Shimura subvarieties of .
1.3. Compact subvarieties of
Recall that the Torelli map sending a smooth complex curve to its Jacobian is an injection of the coarse moduli spaces (and is 2:1 onto its image as a map of stacks, but this does not matter for discussing compact subvarieties). Its image is contained in the moduli space of indecomposable abelian varieties, and determining and would also be very interesting, see Section 6.1.
Remark 1.1.
The classes of the loci of products inside have recently been studied in [CMOP24] and [COP24]. It would be interesting to investigate how our approach, described in Section 1.4, relates to this study (see also Section 6.1).
The Torelli morphism extends to a proper morphism from the moduli space of curves of compact type. As a corollary of their bound (and the theorem of Diaz [Dia84]), Keel and Sadun [KS03] deduce the bound for any . Our results also have implications for .
Corollary C.
The following equality and upper bound hold:
for any , | ||||
for any . |
The point of the difference between these numbers is that is injective at a generic point of , but not along , as it sends a stable curve to the product of Jacobians of its irreducible components, forgetting the location of the nodes.
To obtain the first equality, for any we construct a compact subvariety of of dimension , starting from a compact curve in .
The second inequality implies the bound on the maximal dimension of compact subvarieties of that intersect , and follows from B (in fact B also improves Keel and Sadun’s bound to for ).
It is tempting to conjecture that in fact for all , which would then imply for all . Notice that Krichever [Kri12] claimed exactly such a bound, but unfortunately that proof was never completed.
1.4. Idea of the proofs of A and B
The inspiration for the proof of A is the following. For contradiction, assume that is a compact subvariety of dimension at least . Note that for a fixed elliptic curve , the codimension of the locus is equal to . If we could ensure that the intersection of with were non-empty and transverse, then such property would hold for a Zariski open subset of , i.e. for all but finitely many . By taking the limit as approaches the boundary point , this contradicts the compactness of , thus proving that .
The recent advances in weakly special subvarieties show that this method essentially works for Hodge-generic subvarieties, up to relaxing the transversality condition to “the intersection having a component of the expected dimension”. An essential tool we will use for this is a simplified version of [KU23, Theorem 1.6(i)], whose proof relies on the Ax-Schanuel conjecture, proven for by Bakker and the fourth author [BT19], and proven for an arbitrary Shimura variety by Mok, Pila and the fourth author [MPT19].
To deduce B, we first show that for any Hodge-generic compact subvariety of a non-compact Shimura variety , the dimension of is bounded above as . This argument, applied to , itself gives a stronger version of A, for all Hodge-generic compact subvarieties . We then show that, if a compact is not Hodge-generic, then it must either be contained in a compact Shimura variety, or must be contained (up to isogeny) in a product of Shimura varieties , with non-compact, and , with . Thus a maximal-dimensional non-Hodge-generic is either or , where is a -dimensional subvariety of . We thus complete the proof of B by examining the maximal possible dimension of compact Shimura subvarieties of , using the lists of symplectic representations provided by Milne [Mil11] and Lan [Lan17].
1.5. Structure of the paper
In Section 2 we recall basic facts on algebraic groups, symplectic representations and Shimura varieties. We also prove a structure theorem for Shimura subvarieties that are products of two Shimura varieties, one of which is non-compact (see 2.9).
In Section 3 we recall the statement of Ax-Schanuel for Shimura varieties, and we use it to prove 3.4, which is a stronger and more general version of A. This permits us to give a first estimate for the maximal dimension of a compact subvariety of (see 3.7).
In Section 4 we investigate compact Shimura subvarieties of that are induced by certain symplectic representations that we call “decoupled”, and determine which of them have dimension (see 4.14).
In Section 5 we investigate compact Shimura subvarieties of that are induced by symplectic representations that are not decoupled. We show that such subvarieties have dimension either smaller than , or smaller than some Shimura variety originating from a decoupled representation. We thus conclude the proof of B.
1.6. Acknowledgments
The first author is grateful to Università Roma “La Sapienza” for hospitality in June and October 2023, when part of this work was done. We are grateful to Sorin Popescu who endowed the Stony Brook lectures in Algebraic Geometry, which allowed the first and fourth author to meet and think about these topics. We are very grateful to Salim Tayou and Nicolas Tholozan for enlightening discussions and pointers to the literature on these topics.
2. Connected Shimura subvarieties of
In this section we recall some useful notions on algebraic groups, Shimura data, and Shimura varieties and symplectic representations. We refer to [Mil17, Mil11, Lan17, PR94] for all information relevant to the current paper.
2.1. Algebraic groups
Let be an algebraic group over a number field ; is called simple (resp. almost-simple) if every proper closed normal subgroup of is trivial (resp. finite).
The group is called geometrically simple (resp. geometrically almost-simple) if is simple (resp. almost-simple), and it is semisimple if is, namely if all connected normal closed subgroups of are trivial.
If is almost-simple, then is finite and is simple;
hence, cannot be isogenous to for any positive-dimensional
algebraic groups and .
An isogeny is a surjective homomorphism of algebraic groups with finite kernel; the kernel is then necessarily contained in . An algebraic group is simply connected if every isogeny is an isomorphism.
If is semisimple and are all the almost-simple closed normal subgroups of , then the product map is an isogeny (see [Mil17, Theorem 22.121]). The Galois group fixes , and thus it acts on by permuting its almost-simple factors and by fixing . As a consequence, the stabilizer of the factors of has a number field as fixed field, and so splits as a product of geometrically almost-simple factors.
If is a geometrically almost-simple group over a totally real number field and is the -group obtained from by restriction of the scalars, then is an isogeny, where ranges over all embeddings and .
A split torus is a product of a number of copies of the multiplicative group . An algebraic group is called anisotropic if it contains no positive-dimensional split tori. A semisimple group over a number field is anisotropic if and only if contains no nontrivial unipotent elements.
2.2. Connected Shimura varieties
In this paper we will only deal with connected Shimura varieties. So, when we mention a Shimura variety we mean that it is connected.
We will denote by the connected component of that contains the identity.
Definition 2.1 ([Mil05, 4.4]).
A connected Shimura datum is a pair where is a connected semisimple algebraic group over a number field, and is a -conjugacy class of homomorphisms of real algebraic groups
such that
-
(i)
there is a direct sum decomposition of the Lie algebra of , such that acts trivially on , as on and as on , via the representation
-
(ii)
The conjugation by is a Cartan involution, namely is a compact real form of (and so an inner form of )
-
(iii)
has no nontrivial -simple factor such that is compact.
A morphism of connected Shimura data is a homomorphism of algebraic groups such that the map induced by induces a morphism between and .
Remark 2.2.
In the above definition of connected Shimura datum, the relevant information seem to be carried by and . The relevance of will be clear in Section 2.3 below, where we will see that a morphism from a connected Shimura variety to will be associated to a homomorphism that does not necessarily descend to .
Example 2.3.
Let , where is the identity matrix, and denotes the Siegel space of complex symmetric matrices with positive-definite imaginary part. Then decomposes as the orthogonal sum , where is the Lie algebra of the stabilizer of , and . The infinitesimal action of on at induces the isomorphism . After complexifying, we have , where , , . Note that the homomorphism defined as
fixes . Moreover acts trivially on , as on and as on .
The group splits (up to isogeny) into a product of geometrically almost-simple factors: denote by the product of all its compact factors, and by the product of all its non-compact factors. Now, is a connected Hermitian symmetric domain, and acts on transitively, isometrically, and holomorphically. The group acts on via by fixing the point and by complex multiplication on . The compact subgroup is the connected component of the identity of the stabilizer of .
Note that with , where is a compact inner form of , i.e. has an inner automorphism which is a Cartan involution. We claim that is geometrically almost-simple (see the proof of [Mil11, Theorem 3.13]). Indeed, if were not geometrically almost-simple, then would be isomorphic to the real group obtained from a complex group by the restriction of scalars. But then such has no compact inner form by the following:
Lemma 2.4.
Let be a complex group of positive dimension, and let be the real group obtained from by the restriction of scalars. Then there is no element of the complexification such that is a Cartan involution.
Proof.
Note that the homomorphism , defined as , induces an isomorphism of the complexification . Moreover, the image of is fixed under the involution of .
For contradiction, suppose that is a Cartan involution of , for some . This would mean and thus that can be written as . It is immediate to check that then if and only if . Thus is an isomorphism. This is a contradiction, since is non-compact, while by definition of the Shimura datum must be compact. ∎
Recall that a subgroup is called arithmetic if there exists a simply connected rational algebraic group and a surjective homomorphism with compact kernel such that is commensurable to . An arithmetic subgroup is automatically a lattice. Moreover, denoting , the product is then a finite index subgroup of , and so the map is a finite étale cover.
As every arithmetic subgroup of contains a torsion-free normal subgroup of finite index, the connected Shimura variety is a quotient of by the finite group (see [Mil11, Theorem 3.6]). Moreover, is a smooth quasi-projective variety over (see [Mil11, Theorems 4.2-4.3]) and so can be seen as an irreducible quasi-projective variety too, but also as an irreducible, smooth, Deligne-Mumford stack.
We say that a connected Shimura variety associated to the Shimura datum is indecomposable if is almost-simple. As for the compactness, we recall the following criterion.
Lemma 2.5 ([PR94, page 210, Theorem 4.12]).
The Shimura varieties and are compact if and only if is anisotropic.
If is a Shimura variety associated
to the Shimura datum for ,
a morphism of Shimura varieties
is a map induced by a homomorphism of algebraic groups
such that and .
A point of the Shimura variety corresponding to the homomorphism is called Hodge-generic if the smallest algebraic -subgroup of such that contains the image of is the whole . A subvariety is called Hodge-generic (within ) if its general point is. This is equivalent to requiring to contain a point that is Hodge-generic in .
A special subvariety of is a subvariety induced by a Shimura sub-datum of . A weakly special subvariety of is either a point, or a special subvariety, or a subvariety of a special subvariety of , where is a Hodge-generic point in , up to isogeny.
2.3. Symplectic representations
Let be a connected semisimple algebraic group over and let be a rational symplectic representation of (i.e. a homomorphism , for a vector space over with a symplectic form ). A symplectic subspace of is a vector subspace on which restricts to a (non-degenerate) symplectic form.
Definition 2.6.
Given a representation of , an invariant linear summand of is a -invariant linear subspace; a linear summand is irreducible if it contains no nontrivial -invariant linear subspaces. Given a symplectic representation of , a sub-representation of a -invariant symplectic subspace; such sub-representation is -irreducible if it has no nontrivial sub-representations.
Lemma 2.7.
Any symplectic representation is a direct sum of -irreducible sub-representations.
Proof.
We proceed by induction on . The case is trivial.
Suppose and let be a nontrivial irreducible linear summand. The pairing induces a map which is either zero or an isomorphism, since is a linear summand of .
If is an isomorphism, then is a symplectic representation and hence -irreducible. Moreover, is a direct sum of -irreducibles by induction, and we are done.
We may thus assume that for any irreducible linear summand of , the restriction of to it is zero. Then the pairing against with gives a map . Since is non-degenerate, the map is nonzero and so surjective (by the irreducibility of ). Let be an irreducible linear summand not contained inside . Then the induced map is an isomorphism. Since is isotropic for , it follows that and so is a symplectic representation. Moreover, it must be -irreducible, as its only proper non-zero invariant subspaces are themselves irreducible, and hence by our assumption the restriction of to them is zero. It follows that is a sub-representation of and so it is a direct sum of -irreducibles by induction, and again we are done. ∎
We will often invoke the following useful result.
Lemma 2.8 ([Mil11, Corollary 10.7]).
Suppose that and are non-compact almost-simple real algebraic groups, and let be an irreducible complex symplectic representation of . Then either or acts trivially on .
2.4. Products of Shimura subvarieties in
Using the results recalled above, we now study Shimura subvarieties of that are a product of a non-compact Shimura variety and another Shimura variety, showing that then each factor embeds into a suitable , and that, up to isogeny, the embedding of the product is the product of the embeddings.
Theorem 2.9.
Let be a non-compact indecomposable Shimura variety, and let be any other Shimura variety. Then any embedding of Shimura varieties factors, up to isogeny, as , , followed by the natural embedding , for some .
Proof.
Denote the groups associated to (which thus have no rational compact factor), and denote the rational symplectic representation of that induces the map . Then decomposes as a direct sum of irreducible -invariant linear subspaces; in particular, by suitably regrouping the summands, we can write , where acts nontrivially on each irreducible linear summand of and trivially on , and acts nontrivially on each irreducible linear summand of and trivially on . It is easy to see that , , and are sub-representations.
We claim that ; then the conclusion of the theorem follows, as and are rational representations.
In order to prove the claim, for contradiction let be a nontrivial irreducible linear summand . We write , where is a nontrivial irreducible linear representation of for .
Up to isogeny, decomposes as , where (resp. ) is a product of almost-simple compact (resp. non-compact) real groups. Since have positive dimension (otherwise the theorem is trivial), both and are nontrivial. On the other hand, , where is a geometrically almost-simple algebraic group defined over a totally real number field . Hence is an isogeny.
Now, there exists such that acts nontrivially on : set . Since is indecomposable and non-compact, it follows that is almost-simple and isotropic: hence all are non-compact, and in particular is. By 2.8 it follows that all direct factors of must act trivially on . Let such that acts nontrivially on : set . Since has no compact -factors, there exists such that is non-compact. Hence, contains a factor , which is acted upon nontrivially by the non-compact factor of and by the non-compact factor of . This contradicts 2.8. ∎
3. Ax-Schanuel for Shimura varieties, and Hodge-generic compact subvarieties
In this section we recall the statement of the (weak) Ax-Schanuel conjecture for Shimura varieties, proven in [MPT19], and a consequence of it, which is essentially [KU23, Theorem 1.6(i)]. We will then use it to prove a more precise version of A.
3.1. A consequence of Ax-Schanuel for Shimura varieties
Let be a connected Shimura datum and let be a lattice in . Denote by the Shimura variety and by the natural projection, and let be the compact Hermitian domain dual to , which is a projective variety.
Theorem 3.1 (Weak Ax-Schanuel [MPT19]).
Let be an algebraic subvariety of (namely, obtained by intersecting with an algebraic subvariety of ) and let be an algebraic subvariety of . If has an analytic irreducible component of dimension larger than expected, then is contained inside a proper weakly special subvariety of .
This will be usually applied when is the image inside of the map induced by a morphism of Shimura data.
As a consequence of 3.1, one obtains 3.2 below, which is essentially a simplified version of [KU23, Theorem 1.6(i)], whose proof we include for completeness.
Recall that if is an irreducible subvariety and is its (connected) smooth locus, then the algebraic monodromy of at a point is the Zariski closure of the image inside of the induced homomorphism .
Moreover, recall that for a subvariety , denoting , for any the image is a subvariety of called the -translate of .
Theorem 3.2.
Let be a subvariety whose generic point has -simple algebraic monodromy , and let be the smallest Shimura subvariety containing . If is a morphism of Shimura varieties such that
-
(a)
, and
-
(b)
,
then there exists a -translate of inside whose intersection with has a component of the expected dimension.
Proof.
Denote by the universal cover of , by the preimage of inside , and by the preimage of inside .
Fix a Hodge-generic point of , let be a preimage of , and let be such that .
Since there are no proper positive-dimensional weakly special subvarieties of containing , and since by (a), it follows from 3.1 that (which is non-empty) has an analytic irreducible component of the expected dimension containing .
Thus, for any sufficiently close to , there will also exist an analytic irreducible component of that still has the expected (non-negative) dimension. In particular, since is dense in , such property holds for some rational sufficiently close to . ∎
The following application will be useful for us.
Corollary 3.3.
With the same hypotheses as in 3.2, assume moreover that and
-
(a)
-
(b)
for all .
Then
-
(i)
there exists a -translate of inside (for some ) such that the intersection of the -translate of its slice with has a component of the expected dimension, for every in a (non-empty) Zariski-open subset of ;
-
(ii)
if is non-compact, then is non-compact.
Proof.
(i) For every denote the preimage of inside . Fix . By 3.2 there exists such that the intersection has a component of the expected dimension (and, in particular, is non-empty). Thus, the same holds for every in a sufficiently small neighbourhood of . By analyticity of this condition, this then also holds for all outside a countable union of proper analytic subvarieties.
(ii) Let be as in (i) and let be a sequence of points such that , while the corresponding sequence of images diverges. Then contains a diverging sequence, which shows that is not compact. ∎
3.2. Hodge-generic compact subvarieties of
Given an irreducible subvariety , by definition is Hodge-generic within the smallest Shimura subvariety of containing . In this section we prove the following.
Theorem 3.4.
Let be an indecomposable non-compact Shimura subvariety, and let be a compact subvariety that is Hodge-generic within . Then .
The first main result of the present paper is an immediate consequence.
Note that we of course get a stronger version of A, for compact subvarieties through any Hodge-generic point of , not just through an abstract very general point of . Note also that the proof of 3.4 in this special case does not require either technical 3.5 or 3.6, and implements the idea discussed in the introduction, making use of 3.3 to see that the intersections are indeed of expected dimension.
The idea of the proof of 3.4, for arbitrary non-compact , is to first produce a morphism of Shimura varieties with and such that intersects for almost every (where denotes the moduli of ppav with a full level structure). By 3.3(ii), the compactness of then forces .
The construction of uses the following preliminary lemma, in which we construct a modular curve mapping to .
Lemma 3.5.
Let be a non-compact connected Shimura variety corresponding to a connected semisimple group over . Then there is a homomorphism of -groups and a point such that induces a holomorphic map of modular curves , for some large enough .
Proof.
Let , for a Shimura datum and a lattice in . Then the adjoint action of on determines a canonical variation of Hodge structure on (see [Mil11, §5]). Since is non-compact, the group contains a unipotent element, and the variation associated to degenerates. The conclusion now follows from [Sch73, Cor 5.19]. (Technically Schmid works only for periods domains corresponding to the full orthogonal/symplectic group, but the argument goes through unchanged in our context). ∎
We can now construct the desired morphism .
Lemma 3.6.
Let be a non-compact Shimura variety. Then there is a morphism of Shimura varieties such that for each the intersection is non-empty.
Proof.
Let be the homomorphism over that induces . Let be the homomorphism, over , provided by 3.5, and denote .
For each , denote the stabilizer of , which is a maximal compact subgroup isomorphic to . The homomorphism induces a holomorphic map , which sends to the point of corresponding to .
Denoting the standard two-dimensional representation of , recall that all irreducible representations of are of the form for some integer . The group acts on via with weights , , or . This implies that all irreducible linear summands of are isomorphic to either or , and so decomposes as an orthogonal direct sum of rational representations, where , and is a trivial representation. Note that , because the map from the modular curve to provided by 3.5 is not constant. The polarization on induces a quadratic form on , which can be diagonalized over . We can thus decompose with , so that moreover the polarization form is orthogonal with respect to this direct sum decomposition.
Let denote the ppav corresponding to . Then can be identified to , and its Hodge filtration is determined by the subspace .
Since acts on preserving the Hodge filtration, we can decompose into irreducible representations of the subgroup . This gives , with , and .
The universal cover identifies the abelian variety with , where is a discrete subgroup of of rank . Denoting , and , since is a direct sum decomposition over , it follows that has finite index in . Hence, the abelian variety is isogenous to , where , and . More precisely, , where is a finite subgroup of . In particular, , where are elliptic curves. Note that is an orthogonal decomposition with respect to the pullback of the polarization of . Since every polarized abelian variety is isogenous to a ppav (via an isogeny respecting the polarization), there exist isogenies and such that the pullback polarization on and on can be written as and for some integer , where and are principal polarizations on and on . Moreover, this can be done globally in the family over , with a constant isogeny, so that and are the fibers over of two families of abelian varieties and over endowed with principal polarizations and .
Thus, we can write , with being a family of torsion subgroups of the family of abelian varieties over , and then the polarization descends to a principal polarization on .
Denote the order of the group . The action of on extends to a flat action of on the universal family over the level cover , and taking this quotient by gives the morphism , so that for any we have , and in particular . ∎
We are now ready to prove the theorem, implementing the original idea.
Proof of 3.4.
Assume, for contradiction, that is a compact Hodge-generic irreducible subvariety of , of dimension at least . Let be the morphism constructed in 3.6, so that for any the intersection is non-empty. Since and since is non-compact, by 3.3(ii) it follows that must also be non-compact, contradicting our hypotheses. ∎
Proposition 3.7.
Let denote the maximal dimension of a compact Shimura subvariety of . Then the maximal dimension of a compact subvariety of is
Proof.
Let be a compact subvariety and let be the smallest Shimura subvariety containing . Since , it is enough to deal with the case , in which case from 3.4 it follows that .
Decompose with all indecomposable non-compact Shimura varieties, and with a compact Shimura variety. Then by applying 2.9 repeatedly, it follows that the embedding must factor, up to isogeny, via
for suitable and . In particular (the isogeny image of) embeds into the product . The projection of to each non-compact factor is of dimension at most , by 3.4, and thus the projection of to has dimension at most . Thus .
On the other hand, the maximal dimension of a compact Hodge-generic subvariety of is , and can be realized by some complete intersection . Hence, the compact subvariety of has dimension . The conclusion follows, as . ∎
4. Compact Shimura subvarieties of from decoupled representations
In this section we begin investigating the maximal dimension of compact Shimura subvarieties , by going through the lists of the real, simply connected almost-simple groups and their irreducible linear representations that occur as constituent of a complex symplectic representation.
4.1. Setting
Let be a connected semisimple group corresponding to a connected Shimura variety , and let be a symplectic representation over inducing a map of to .
We say that is the pair associated to (or just to ) if and . We recall that there exist (Hodge-generic) -dimensional compact subvarieties of for every . Hence we call a pair negligible if and eligible if .
Moreover, we introduce a partial ordering on by declaring
If , we will say that is dominated by .
Let be an irreducible linear summand of . Then is acted upon nontrivially by almost-simple factors of , at most one of which is non-compact by 2.8. Since the case will be the most relevant, we introduce the following.
Definition 4.1.
A complex symplectic representation of is decoupled if, for every irreducible linear summand, there exists an almost-simple factor of such that the action of on such summand agrees with its restriction to this almost-simple factor. A rational symplectic representation of is decoupled if its complexification is. ∎
In this section we study compact Shimura subvarieties associated to decoupled representations of . In view of 2.5, we will thus assume that
-
•
is anisotropic.
4.2. -irreducible decoupled -representations
Later in this section we will classify dominating pairs of the following type.
Definition 4.2.
Let be a decoupled symplectic representation of over . The pair corresponding to is d(ecouplably)-achieved if is anisotropic and is -irreducible. ∎
Geometrically, the -irreducibility of is equivalent to the fact that is not contained, up to isogeny, in the decomposable locus of .
A relevant property that follows from the -irreducibility of is the following.
Lemma 4.3.
Assume that is -irreducible. Then all the irreducible linear summands of are either Galois conjugate to each other or to each other’s duals. In particular, they all have the same dimension, and are all of the same type (orthogonal, or symplectic, or not self-dual).
Proof.
Let be an irreducible linear summand, and let denote the direct sum of all subspaces of that are isomorphic to a Galois conjugate of or . Then is Galois stable, and thus is defined over . Moreover, decomposes as with not having any of the irreducible linear summands that appear in , and thus is symplectic. Since is assumed -irreducible, we must have . This proves the claim. ∎
Note that, if the representation is -irreducible and decoupled, then the group must be almost -simple.
4.3. Almost -simple groups and decoupled symplectic representations
Let be an almost -simple group. It follows that for some totally real field , and that is an isogeny, where for . So all are -forms of the same complex group .
Let be a symplectic representation of over , which is decoupled but not necessarily -irreducible. Then decomposes as , where each is a complex linear representation of . If is non-compact, then every -irreducible sub-representation of corresponds to a “symplectic node”, see [Mil11, §10].
Denote by a -invariant irreducible linear subspace of .
Remark 4.4.
The -representation is symplectic. Hence, if is not a symplectic subspace, then has a -invariant subspace isomorphic to .
In Table 2 we list all possible types of complex groups , real groups , domains associated to the non-compact , and -irreducible linear summands that can occur in the symplectic representation (see, for example, [Mil11, §10] and [Lan17, §3.7]).
Type of | Type of | Hermitian | Domain | Repr. | Non-self-dual/ |
complex | non-compact | symmetric | /Symplectic/ | ||
group | real group | space | /Orthogonal | ||
Symp | |||||
, | |||||
, | NSD | ||||
, | |||||
Symp | |||||
, | |||||
, | Orth |
For Table 2, we also recall that
can be identified with the group of automorphisms of the -dimensional vector space over the quaternion skew algebra with center that preserve
the standard skew-Hermitian pairing on .
Moreover, in the case we have .
In the dimension estimates contained in Section 4.4, we will often make use of the following result.
Lemma 4.5.
Let be a rational anisotropic group, where is a totally real number field, and is one of the entries in Table 2. If has no compact factors, then each is isomorphic to either of the following
-
(1)
for some indefinite Hermitian form on ,
-
(2)
for some symmetric bilinear form on of signature .
Proof.
The cases we are interested in of of classical type are summarized in [PR94, page 92]. In the case, the group must be of type , while otherwise the group is isomorphic to either:
-
(a)
for , or
-
(b)
for ,
where is either (and is symmetric or alternating), or an imaginary extension of (and is a Hermitian form), or a quaternion skew field with center (and is a Hermitian or a skew-Hermitian form).
For any the group contains nontrivial unipotent elements. If and is alternating, then , in which case it also contains unipotent elements.
In the remaining cases with symmetric/Hermitian/skew-Hermitian, it is known that contains nontrivial unipotent elements if and only if has nontrivial isotropic vectors, which happens if and only if nontrivially represents zero.
If is symmetric and , then [PR94, page 342, Claim 6.1] implies that . In this case, is abelian (for ), or of type (for or of type (for ). The abelian case can clearly be discarded, while the other two cases do contain nontrivial unipotent elements, and thus can also be ruled out.
If is Hermitian, then [PR94, page 343, Claim 6.2] implies that for , and for . Hence is either of type or isomorphic to . Since , such case can be ruled out.
If is skew-Hermitian and , then [PR94, page 343, Claim 6.3] implies that . The case can be ruled out since is of type , and the case corresponds to of type .
Hence, can only be of type or , and for to be non-compact, the Hermitian form must be indefinite. ∎
4.4. Key estimates
The following result contains the key dimension estimates. We state it in a version that will be useful for Section 5 too.
Lemma 4.6.
Let be a -almost-simple, anisotropic algebraic group and let be a decoupled complex symplectic representation of whose irreducible linear summands are all Galois conjugate of each other or Galois conjugate to each other’s dual. If either of the following is satisfied
-
(a)
is not of type;
-
(b)
is the complexification of a rational representation ;
-
(c)
each consists of at least irreducible linear summands;
then the pair is either negligible or dominated by one of the -achievable pairs given in Table 3 below.
Type of | Hermitian | Dominating |
complex group | symmetric | d-achievable |
group | space | pairs |
Negligible | ||
, | Negligible | |
, | , | |
, | Negligible | |
Negligible | ||
, | Negligible | |
, |
In Section 4.5 below we will use the following consequence of the above result.
Corollary 4.7.
Let be a -almost-simple, anisotropic algebraic group, and let be an -irreducible decoupled -representation of . Then the associated pair is either negligible or dominated by one of the d-achievable pairs given in Table 3.
Proof.
It is an immediate consequence of 4.6, since satisfies hypothesis (b) of that lemma. ∎
Proof of 4.6.
Note that all the almost-simple factors of are Galois conjugate to each other. Since all the irreducible linear summands of are Galois conjugate to each other or to each other’s dual, then there exists an integer independent of such that every is the direct sum of irreducible linear summands.
We go through the following case-by-case analysis, starting with the special cases where by 4.5 anisotropic forms with no compact factors may appear.
4.4.1. Special case
In this case the groups are isomorphic either to or to , and by 4.5 there may be anisotropic forms which have no compact factors.
Let be of type . Since the action of on each irreducible linear summand of has weight , or , each must be isomorphic to the standard representation of .
Now, is isomorphic either to or to , and each could be isomorphic either to or to .
Note that , since such group is isotropic. Hence . Then one factor of is isomorphic to .
Assume that is defined over . Then acts nontrivially on a -dimensional -vector space. Since does not have nontrivial 2-dimensional representations over , it follows that . Thus we have and . But in fact all such pairs with are negligible, which gives the first row of Table 3.
Note that the only such pair satisfying is .
Indeed, if we endow with an indefinite anisotropic norm and we let
act on by left-multiplication, the couple determines
a compact Shimura curve in .
4.4.2. Special case
In this case the groups are isomorphic either to or to , and by 4.5 there may be anisotropic forms which have no compact factors.
By Table 2 the representation is not symplectic (it is in fact orthogonal) and -dimensional. Hence , and so . Thus the d-achievable pairs are dominated by , which are negligible.
4.4.3. Case for
The factors are of type . Since we have already treated the case with all non-compact above, by 4.5 we can assume that at least one factor must be compact (namely, it must be ).
Case (i): all factors are of type or . By the classification, has dimension for some .
If , or if and is even, then is not symplectic and so by 4.4. It follows that . The corresponding pair is then dominated by , which is negligible.
If and is odd, then , and the corresponding pair is dominated by
Note that , so this is also negligible.
Case (ii): at least one factor is of type with . By Table 2 the irreducible linear summands of are all the standard representation or its dual: in particular, has dimension and is not self-dual. Thus by 4.4 and so . It follows that , and so .
Likewise, there are at most non-compact factors, with associated symmetric spaces each of dimension .
Thus .
Hence, the d-achievable pairs are dominated by
.
Existence. Let be a totally imaginary quadratic extension, which comes with a natural involution, and let , where is a Hermitian form on . By [BH78, Theorem B], the form can be chosen to have signature at all but one place of , and definite signature at the remaining place . At the compact place , a homomorphism can be defined using the action of on , where is an extension of , as in [Mil11, Theorem 10.14].
4.4.4. Case for
The factors are of type or . By 4.5, at least one factor must be compact, namely of type .
If , then is not self-dual, and so by 4.4. It follows that must have dimension at least , which gives a d-achievable pair dominated by . Since , this is negligible.
If , then is self-dual and so . This gives a d-achievable pair dominated by , which is again negligible since .
4.4.5. Case for
The factors are of type or . By 4.5, at least one factor is compact, namely of type .
Now, each is isomorphic to the fundamental representation, which is defined over . If were isomorphic to (i.e. ), then would be the complexification of a real representation, and the compact factor would inject into . This contradiction shows that (and, in fact, one could show that must be even).
Assume now that .
Thus and so the corresponding d-achievable pair
is dominated by .
Existence. Let be a quaternion algebra over and let for some Hermitian form on . By [BH78, Theorem B], the form can be chosen to be definite at exactly one place of , and (non-degenerate) indefinite at all the other places. Then the pair is d-achieved by the group .
4.4.6. Case for
The factors are of type or , and there is at least one compact factor (of type ) by 4.5.
Suppose . Then the representation is not self-dual, and so by 4.4. It follows that , and so , whereas the domain has dimension at most . Hence the d-achievable pairs satisfy , while , so they are negligible.
Suppose now . Since , we get d-achievable pairs dominated by , which is again negligible as .
4.4.7. Case for
The groups are of type or , and there is at least one compact factor (of type ) by 4.5.
Moreover, is not symplectic, and so by 4.4.
Hence , and all such d-achievable pairs are dominated by
.
Existence.
Let be a quaternion algebra over and let
for some skew-Hermitian form on .
By [BH78, Theorem B] the form can be chosen
to have definite signature at exactly one place of , and to be (non-degenerate) indefinite at
all other places.
The pair is d-achieved by the group .
This completes the proof of 4.6. ∎
Remark 4.8.
As Table 3 shows, both cases and are dominated by the case .
Remark 4.9.
The only d-achievable pair with occurs for , see Section 4.4.1.
Remark 4.10.
We illustrated the above construction in the and cases because they provide remarkable examples of high-dimensional (though not maximal-dimensional) compact subvarieties of .
It would be interesting to determine all maximal irreducible compact subvarieties (i.e. those not properly contained in another irreducible compact subvariety) of , in addition to our determination of all maximal-dimensional compact subvarieties, see 5.10 below.
4.5. Consequences
For every define
Lemma 4.11.
For all we have
Moreover, the above inequality is strict unless and is even.
Proof.
Suppose first that . Then and the inequality is strict unless and is even. So we can assume .
Suppose that . Then .
Suppose now .
It is easy to check that, for every even smaller than , we have . This implies that it is enough to verify the statement for or .
For we have and so . For we have and so . ∎
As a consequence of the analysis done in Section 4.4, we obtain the following bound for the d-achievable pairs.
Proposition 4.12.
For a d-achievable pair we have .
Proof.
To obtain the best possible d-achievable pairs , we invoke 4.7.
We first observe that for and for all cases listed in Table 3 are negligible, i.e. then . Moreover, we note that
Now assume .
We observe that by 4.8 the three possible eligible cases can be easily reduced to a single one. Hence we have to consider only the case with , which gives the d-achievable pairs
Case even.
In this case is even, and . Hence the pair is dominated by
the case , namely by .
Case odd. Then , and we note that is strictly decreasing as a function of .
If is even, then must be even and so the dominating pairs are obtained for : in this case .
If is odd with , then and . Observe that and that for . Hence, for odd. ∎
Remark 4.13.
We observe that the pair as defined in 4.12 is d-achieved if and only if is even and .
In view of 4.12 and 3.7, a compact subvariety of of maximal dimension can be either a Hodge-generic compact subvariety (for example, a complete intersection), or a compact Shimura subvariety, or a product of the two types. For all cases in which the Shimura subvariety is a product of Shimura subvarieties induced by an irreducible decoupled representations, we obtain the following result.
Proposition 4.14.
Let be a compact subvariety of of maximal possible dimension, which is either Hodge-generic, or a Shimura subvariety induced by an irreducible decoupled representation, or a product of these two types. Then .
Proof.
Recall that, by 3.4, the largest possible dimension of a compact Hodge-generic subvariety of is . As in 4.12, we separately analyze a few cases.
If and , all d-achievable pairs are negligible and so a maximal-dimensional compact subvariety can be obtained by taking Hodge-generic.
For or we have , and so the maximal dimension of is achieved using one indecomposable Shimura variety by 4.11.
Thus, for even, the bound is achieved by a compact Shimura variety obtained by an irreducible decoupled representation (see 4.13).
For odd, the bound is achieved by , where is a compact Shimura subvariety of of the type considered above in the case of even . ∎
Remark 4.15.
There are two interesting cases in which the optimal bound in 4.14 is achieved in two different ways.
-
(a)
For a compact curve in can be constructed as complete intersection, or as a Shimura variety, as in Section 4.4.1.
-
(b)
For a compact subvariety of of largest dimension can be constructed again as a complete intersection (or can be a more general Hodge-generic subvariety), or as the product of a point in and a largest (16-dimensional) compact Shimura subvariety of , see Section 4.4.3(ii).
In all the other cases, the construction described in the proof of 4.14 is essentially unique (see 4.9).
Corollary 4.16.
Let be an anisotropic -algebraic group and let be a decoupled representation of such that the number of irreducible linear summands nontrivially acted on by the same component of is at least . Then the associated pair satisfies .
Moreover, if , then must be almost-simple.
Proof.
Decompose (up to isogeny) as a product of -(almost) simple, anisotropic groups and as , where is acted on nontrivially by the factor only. Hence is a decoupled representation of .
Now, up to isogenies decomposes as a product of real almost-simple factors as and we can write with acted on nontrivially by only. By construction, consists of at least irreducible linear summands. Hence each satisfies the hypotheses of 4.6 and condition (c).
Denote by the dimension of the domain associated to , by the dimension of , by the dimension of . Then by 4.6 the pairs are either negligible or dominated by a d-achievable pair as in Table 3. It follows from 4.12 that Since and , by 4.11 we conclude that .
If , then must be isogenous to the product of at most two almost-simple factors by 4.11. Suppose, for the sake of contradiction, that is isogenous to , with and anisotropic and almost-simple. Up to reordering the factors, 4.11 implies that , which gives a contradiction since does not contain positive-dimensional compact subvarieties. ∎
5. Compact Shimura subvarieties from non-decoupled representations
In this section we study the dimension of compact Shimura subvarieties of arising from non-decoupled representations. We show that such Shimura subvarieties never have higher dimension than the already constructed Shimura subvarieties, and so 4.14 indeed gives the dimension of the largest compact subvariety of .
5.1. Setting
Denote by a connected semisimple algebraic group over , and let be the decomposition (up to isogeny) of into almost simple factors. Denote by an -irreducible, symplectic -representation of .
Then decomposes as a direct sum of -invariant linear subspaces , with all conjugate to each other or to each other’s dual under the Galois group . Moreover, each can be written as , where is an irreducible linear representation of .
Remark 5.1.
Since all irreducible linear summands of are Galois conjugate to each other, or to each other’s duals, either all are self-dual or none of them is. Also, in the self-dual case, either all are symplectic or all are orthogonal.
For a fixed consider the unordered string of numbers , which is independent of since is -irreducible, and let be the substring consisting of numbers strictly greater than one. Denote by and respectively the product and the sum of the elements of .
Then for any we have .
5.2. Two decoupling tricks
Now we introduce two complex decoupled symplectic representations of obtained from . We begin with the following simple observation.
Remark 5.2.
Let be a representation and let be the symplectic form on defined by
for all and . Then acts symplectically on via .
First decoupling trick.
For every , define as
if is a nontrivial representation of ,
and as if is trivial. Put on
a symplectic structure as in 5.2.
Then let
and , which is thus a complex symplectic representation of .
Second decoupling trick. Assume that the ’s are not symplectic. Then, for every , define as if is a nontrivial representation of , and as otherwise. Then let and .
Since the irreducible linear summands of are not symplectic,
for every there must be another irreducible linear summand of isomorphic to : let be one such summand.
In particular, for every
there must be another summand of dual to : let be one such summand.
As finding such dual summands is bijective, we can define a symplectic structure on
by pairing each to as in 5.2, and by declaring orthogonal to all linear summands different from .
Here we stress that the complex symplectic representations and
-
•
need not be defined over , and
-
•
need not be -irreducible over ,
but they have the following properties.
Lemma 5.3.
The complex symplectic representations and of are decoupled, and the following hold.
-
(i)
The number of irreducible linear summands of nontrivially acted on by a factor of is at least . Moreover for each .
-
(ii)
Assume that some is not symplectic. Then the number of irreducible linear summands of nontrivially acted on by a factor of is at least . Moreover and the equality holds only if .
Proof.
The representations and are decoupled by construction.
(i) follows directly from the construction of the summands . Indeed, if the factor nontrivially acts on , then so does on .
(ii) again follows directly from the definition of , observing that . ∎
5.3. Shimura subvarieties from decouplable representations
In view of 5.3, we need to treat representations in different ways depending on the values of and .
Definition 5.4.
We call a finite collection of integers greater than one decouplable if . A representation is decouplable if its associated is. An irreducible representation is decouplable if all of its linear summands are.
Dimension estimates for decouplable representations can be easily reduced to the decoupled case, and the following lemma shows that decouplable representations are never associated to a compact Shimura subvariety of of maximal dimension.
Lemma 5.5.
Let be a decouplable representation of corresponding to a map to . Then the dimension of the domain associated to satisfies .
Proof.
We wish to use the first decoupling trick. By 5.3(i) we have a decoupled complex representation such that , exactly because is decouplable.
Now, might not be defined over . Nevertheless, we can conclude since and satisfy the hypotheses of 4.16. ∎
Remark 5.6.
Note that if is a symplectic representation of corresponding to a map to , and satisfies , then the same argument as in 5.5 yields .
5.4. Shimura subvarieties from non-decouplable representations
Now we have to deal with non-decouplable representations. To that end we begin with the following numerical lemma.
Lemma 5.7.
A finite collection of integers larger than is non-decouplable if and only if it belongs to the following list:
-
(i)
;
-
(ii)
;
-
(iii)
with ;
-
(iv)
;
-
(v)
with ;
-
(vi)
.
Proof.
First, if , then is decouplable. Indeed, for we have , and increasing each number in by one results in increasing by and by at least .
Let now and note that, for , we have and so is non-decouplable. Moreover, increasing each number in by one results in increasing by and by at least . So all the other cases are decouplable.
Next, if , then is non-decouplable since . Now, increasing a number in by one results in increasing by and by at least ; so, in order to find non-decouplable sets of three elements, we can only do it at most twice. This shows that and are the only non-decouplable sets of 3 elements, since is decouplable by inspection.
Now, suppose that , with . Then is non-decouplable if and only if , which is equivalent to . So either , or , or
Finally, if , then it is certainly non-decouplable. ∎
Now we show that non-decouplable representations are never associated to compact Shimura subvarieties of of maximal dimension.
Lemma 5.8.
If is non-decouplable, then the pair associated to the -representation satisfies .
In the proof, we will often use the following observation.
Remark 5.9.
Let (resp. ) be a real algebraic group and let (resp. ) be a complex irreducible linear representation of (resp. of ). Then is a self-dual -representation if and only if is a self-dual -representation and is a self-dual -representation. Moreover, is symplectic if and only if is symplectic and is orthogonal, or if is orthogonal and is symplectic.
Proof of 5.8.
We proceed by separately analyzing each non-decouplable case.
Cases , and .
This can only occur if all factors are of type and the dimension of the corresponding symmetric space is for each non-compact . Hence, the dimension of the domain is exactly the number of non-compact factors in . On the other hand, each linear summand is nontrivially acted on by at most one non-compact factor of by 2.8. It follows that there must be at least such linear summands, each one of dimension . Hence, and so . Thus this case is negligible.
We will show now by contradiction that the case does not occur.
Indeed, if , we must have and . This implies that is almost-simple and so of type , where is a totally real quadratic extension. So is a product of the unit quaternions and .
However, seen as an -representation, must have no trivial summands and so it must be isomorphic
to the direct sum of 2 copies of standard representation .
The centralizer inside of such -representation is isomorphic to , that
acts on via its natural action on .
Hence the action of on factors through .
Since has no 2-dimensional real representations,
it follows that acts trivially.
This is clearly a contradiction.
Case .
Here at least two of the factors correspond to type with
dimension of the corresponding symmetric space , and at least one to type , with dimension of the corresponding symmetric space .
Let be the number of non-compact factors of of type
and the number of non-compact factors of type .
The dimension of the domain is then .
On the other hand, each linear summand is nontrivially acted on
by at most one non-compact factor of by 2.8.
It follows that there must be at least such linear summands, each one of dimension . Hence, and so .
Since , this case is negligible.
Case .
We proceed as in the case above.
At least two factors must be of type ;
and at least one factor must be of type or (with corresponding symmetric spaces of dimension or )
or (with corresponding symmetric spaces of dimension ).
Let (resp. , ) be the number of non-compact factors of of type
(resp. or , ), so that
the domain has dimension .
Since each linear summand is nontrivially acted on by at most one non-compact factor of ,
the representation must consist of at least
linear summands , each of dimension . It follows that , and so . This case is thus negligible.
Case with .
Suppose that and .
Then is not self-dual, and so is not symplectic
by 5.9.
Hence we can use the second decoupling trick, and we consider the representation
constructed in Section 5.2.
Since , the pairs , arising from and satisfy .
Though the representation might not be defined over , the
conclusion still follows from
5.3(ii) and 4.16.
Case .
The factors of must be either of type or .
Consider a linear summand ,
with .
By Table 2 either is not self-dual,
or is not self-dual, or both and are symplectic.
In all cases, no linear summand is not symplectic.
Hence we can use the second decoupling trick and consider
the representation produced in Section 5.2.
Since and , we have
. Hence the pairs and arising from and satisfy .
Again, the representation might not be defined over , but the
conclusion follows from 5.3(ii) and 4.16.
Case with .
In this case is isogenous to , where is almost -simple of type and is almost -simple of type different from .
Since has no nontrivial 2-dimensional representation, must be of type and must be the standard representation, which is thus symplectic. It also follows that must have at least one compact factor.
We separately analyze two cases, depending on whether
is symplectic.
Case not symplectic.
We can use the second decoupling trick and consider produced in Section 5.2.
Since , the pairs , arising from and satisfy . As before, the representation might not be defined over , but the
conclusion follows from 5.3(ii) and 4.16.
Case symplectic.
By 5.9, the factor is orthogonal. We proceed case by case, analyzing the possibilities for from Table 2.
We denote by the number of factors of and by the number of factors in , so that .
Subcase of type with .
In this case with even. Moreover must have at least one compact factor by 4.5.
According to Table 2, we have and , since .
Since every is nontrivially acted on by at most two factors of , the representation must consist of at least irreducible linear summands, each of dimension . It follows that
It can be easily checked that for all ,
and so this case is negligible.
Subcase of type with .
Since every irreducible linear summand of is nontrivially acted on by at most one factor of and one factor of , there are at least irreducible linear summands, each of dimension .
Hence, .
Since , we have , and so
this case is negligible.
Subcase of type with .
According to Table 2,
we have .
As in the previous case, must consist of at least linear summands, of dimension each, and so
.
Again , and so this case is negligible.
Subcase of type with .
Since each is nontrivially acted on by at most two factors of , the representation consists of at least irreducible linear summands, each of dimension according to Table 2. It follows that .
Since must have at least one compact factor by 4.5, we have .
The pair is dominated by , which is achieved in the case, according to Table 3. ∎
5.5. Compact Shimura subvarieties of maximal dimension
Now we can prove our second main result.
Proof of B.
4.15 and the proof of B in fact yield more information than B, allowing us to describe all maximal-dimensional compact subvarieties of , not just determining their dimensions. We collect such information in the following statement.
Theorem 5.10.
All maximal-dimensional compact subvarieties of , in each genus, are described as follows:
-
(i)
for , the maximal-dimensional compact subvarieties of are either Hodge-generic curves (for example, general complete intersections), or Shimura curves (see 4.4.1);
-
(ii)
for , all maximal-dimensional compact subvarieties of must be Hodge-generic: for example, general complete intersections work;
-
(iii)
for even, all maximal-dimensional compact subvarieties of are compact Shimura subvarieties of the type constructed in Section 4.4.3(ii) with a totally real quadratic extension (see also 4.12 and 4.13);
-
(iv)
for odd, all maximal-dimensional compact subvarieties of are products of a point in and a Shimura subvariety of of maximal dimension of the type discussed in (iii);
-
(v)
for , the maximal-dimensional compact subvarieties of are either Hodge-generic (for example, general complete intersections), or the product of a point in with a compact 16-dimensional Shimura subvariety of of the type discussed in (iii).
To be completely explicit, we now give a simplest example of the compact Shimura varieties mentioned in cases (iii-iv-v) of the theorem above.
Example 5.11.
Let and .
-
(i)
Assume . Let be the Hermitian form on defined by Take , so that is isogenous to . Then the domain maps to and determines a compact Shimura subvariety of , as described in Section 4.4.3(ii).
-
(ii)
Assume and let be the Hermitian form on defined by . Take , so that is isogenous to . Then the domain maps to and determines a compact Shimura subvariety of , as described in Section 4.4.3(ii).
-
(iii)
Assume odd. Choose a point and denote by the compact subvariety of .
Then for any this is a maximal-dimensional compact subvariety of .
6. The indecomposable locus and the locus of Jacobians
In this section we discuss some consequences of our results, and some open problems concerning the locus of indecomposable ppav and the locus of Jacobians inside , in particular proving C.
6.1. The locus of indecomposable ppav
As mentioned in Section 1.3, the moduli space of indecomposable ppav of dimension is also very natural to consider, see [KS03, §1] for a further discussion of the motivation.
Thinking of the Satake compactification as a compactification of , we see that the boundary has one irreducible component that has maximal dimension, which has codimension . Thus , while A of course implies the following.
Corollary 6.1.
For the following holds:
It would be very interesting to know which value it in fact is. As we will now see, for , and it is natural to wonder if this is the case for all .
Indeed, the cases of are classical: is affine, while does not contain a compact surface, for example by [Dia84]. In general, recall that denotes the ’th Chern class of the Hodge rank vector bundle on , and that the tautological ring is the subring generated by the classes .
For , from the results of [HT12, HT18] it follows that the classes of all algebraic subvarieties of lie in , and in fact the class of is a nonzero multiple of the class . Thus the main result of [KS03] shows that any compact subvariety of must satisfy , since . This implies .
However, the situation in higher genus remains mysterious. By [vdG99], for any the homology class of is a multiple of the top Hodge class , which in fact vanishes on . As detailed in [CMOP24], it is possible to define a projection map . Then in [COP24] the authors show that, in general, the projection of the class of to the tautological ring of is a nonzero multiple of , but also that for the class of does not lie in . This makes considerations similar to the case above impossible in higher genus.
Note also that as the maximal-dimensional compact Shimura subvarieties of constructed in Section 4.4.3 are not contained in , our results do not suffice to determine for any , and it would be interesting to find explicit high-dimensional compact Shimura subvarieties of .
6.2. The locus of Jacobians, and the moduli space of curves
We now prove the results on compact subvarieties of and .
Proof of C.
As already mentioned in the introduction, the upper bound for simply follows from the inclusion and the case of B, giving for that range of .
To show that for , we recall the construction of a compact subvariety contained in the boundary , already pointed out by the first author for [Kri12].
Indeed, since and contain a compact curve and surface, respectively, the statement is true for and . For , we proceed by induction, assuming the result for all . Indeed, by B, for in this range we have , and so an irreducible compact subvariety that satisfies (and thus maps generically 1-to-1 to its image in under ) has dimension strictly smaller than . Thus it is enough to deal with irreducible compact . Such an must then be contained in some irreducible component of the boundary . But since the forgetful map has compact curve fibers, the maximal dimension of a compact subvariety of is equal to plus the maximal dimension of a compact subvariety of , it then follows from the inductive assumption that
which proves the upper bound for . We now construct an explicit compact subvariety of this dimension. For even we consider inside the boundary stratum
the product of two compact surfaces in , and compact threefolds in , that are preimages of a compact curve in , giving altogether a product variety of dimension . For odd, we do the same except taking the last factor to be a compact threefold in that is the preimage of a compact surface in .
Thus for any we construct a compact subvariety of of dimension . If for a given this dimension is greater than , then it follows that for this we have . This happens to be the case if and only if , as , while . ∎
The compact subvariety costructed above yields for any , and considering its image under gives , which is obtained using a variety contained in the boundary, and we do not know any construction of a higher-dimensional compact that intersects both and .
Remark 6.2.
We see no reason to expect the bound for to be sharp. It would be very interesting to improve it, eg. by bounding the dimensions of the intersections of compact Shimura varieties with , extending the spirit of the Coleman-Oort conjecture [Oor97, Section 5].
For easier future reference, and to summarize the current state of the art, in Table 4 we give the results of C for small genera, and summarize all the prior knowledge on compact subvarieties of and .
Here we recall that Keel and Sadun proved for any , while the lower bounds on and simply follow from considering the Satake compactification , which is the closure of in , so that and .
The lower bounds for are obtained by covering constructions starting either from a compact curve in or from a one-dimensional compact family of pairs of distinct points on a fixed curve of genus . The best known results are due to Zaal, following ideas by González-Díez and Harvey in [GDH91, Section 4], and are described in detail in Zaal’s thesis, where in particular it is shown in [Zaa05, Thm. 2.3] that a compact -fold exists in for any . Finally, the best known upper bound for is the famous 40 year old theorem of Diaz [Dia84] (which by now has multiple proofs): for any .
References
- [BH78] A. Borel and G. Harder, Existence of discrete cocompact subgroups of reductive groups over local fields., Journal für die reine und angewandte Mathematik 298 (1978), 53–64.
- [BT19] B. Bakker and J. Tsimerman, The Ax-Schanuel conjecture for variations of Hodge structures, Invent. Math. (2019), no. 217, 77–94.
- [CMOP24] S. Canning, S. Molcho, D. Oprea, and R Pandharipande, Tautological projection for cycles on the moduli space of abelian varieties, 2024.
- [COP24] S. Canning, D. Oprea, and R Pandharipande, Cycles on moduli of abelian varieties, 2024, in preparation.
- [Dia84] S. Diaz, A bound on the dimensions of complete subvarieties of , Duke Math. J. 51 (1984), no. 2, 405–408. MR MR747872 (85j:14042)
- [GDH91] G González Díez and W. Harvey, On complete curves in moduli space. I, Math. Proc. Cambridge Philos. Soc. 110 (1991), no. 3, 461–466.
- [HT12] K. Hulek and O. Tommasi, Cohomology of the second Voronoi compactification of , Doc. Math. 17 (2012), 195–244. MR 2946823
- [HT18] by same author, The topology of and its compactifications, Geometry of moduli, Abel Symp., vol. 14, Springer, Cham, 2018, With an appendix by Olivier Taïbi, pp. 135–193. MR 3968046
- [Kri12] I. Krichever, Real normalized differentials and compact cycles in the moduli space of curves, 2012, arXiv:1204.2192.
- [KS03] S. Keel and L. Sadun, Oort’s conjecture for , J. Amer. Math. Soc. 16 (2003), no. 4, 887–900 (electronic). MR MR1992828 (2004d:14064)
- [KU23] N. Khelifa and D. Urbanik, Existence and density of typical Hodge loci, 2023, arXiv:2303.16179.
- [Lan17] K.-W. Lan, An example-based introduction to Shimura varieties, Preprint (2017), available at https://www-users.cse.umn.edu/~kwlan/articles/intro-sh-ex.pdf.
- [Mil05] James S Milne, Introduction to shimura varieties, Harmonic analysis, the trace formula, and Shimura varieties 4 (2005), 265–378.
- [Mil11] J. S. Milne, Shimura varieties and moduli, arXiv:1105.0887 (2011).
- [Mil17] by same author, Algebraic groups, Cambridge Studies in Advanced Mathematics, vol. 170, Cambridge University Press, Cambridge, 2017, The theory of group schemes of finite type over a field. MR 3729270
- [MPT19] N. Mok, J. Pila, and J. Tsimerman, Ax-Schanuel for Shimura varieties, Annals of Mathematics 189 (2019), no. 3, 945–978.
- [Oor97] Frans Oort, Canonical liftings and dense sets of cm-points, Arithmetic geometry (Cortona, 1994) 37 (1997), 228–234.
- [PR94] V. Platonov and A. Rapinchuk, Algebraic groups and number theory, Pure and Applied Mathematics, vol. 139, Academic Press, Inc., Boston, MA, 1994, Translated from the 1991 Russian original by Rachel Rowen. MR 1278263
- [Sch73] W. Schmid, Variation of Hodge structure: the singularities of the period mapping, Inventiones mathematicae 22 (1973), no. 3-4, 211–319.
- [vdG99] G. van der Geer, Cycles on the moduli space of abelian varieties, Aspects Math., E33, pp. 65–89, Vieweg, Braunschweig, 1999. MR MR1722539 (2001b:14009)
- [Zaa05] C. Zaal, Complete subvarieties of moduli spaces of algebraic curves, PhD Dissertation, University of Amsterdam, 2005. Available at https://pure.uva.nl/ws/files/3915897/36235_Thesis.pdf.