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

Compact subvarieties of the moduli space of complex abelian varieties

Samuel Grushevsky Department of Mathematics and Simons Center for Geometry and Physics, Stony Brook University, Stony Brook, NY 11794-3651 [email protected] Gabriele Mondello Dipartimento di Matematica, Piazzale Aldo Moro, 2, I-00185 Roma, Italy [email protected] Riccardo Salvati Manni Dipartimento di Matematica, Piazzale Aldo Moro, 2, I-00185 Roma, Italy [email protected]  and  Jacob Tsimerman Dept. of Mathematics, University of Toronto, Toronto, Canada [email protected]
Abstract.

We determine the maximal dimension of compact subvarieties of 𝒜g{\mathcal{A}}_{g}, the moduli space of complex principally polarized abelian varieties of dimension gg, and the maximal dimension of a compact subvariety through a very general point of 𝒜g{\mathcal{A}}_{g}. This also allows us to draw some conclusions for compact subvarieties of the moduli space of complex curves of compact type.

Research of the first author is supported in part by NSF grant DMS-21-01631. The second author is partially supported by INdAM GNSAGA research group, the second and third authors are partially supported by PRIN 2022 research project “Moduli spaces and special varieties”.

1. Introduction

Given an irreducible quasi-projective variety 𝒱\mathcal{V}, it is natural to ask how far is 𝒱\mathcal{V} from being affine or projective. Perhaps the simplest way to measure this is the maximal dimension of a compact subvariety of 𝒱\mathcal{V}, which we denote mdimc(𝒱)\operatorname{mdim_{c}}(\mathcal{V}), though perhaps the maximal dimension of a compact subvariety of 𝒱\mathcal{V} passing through a very general point of XX, which we denote mdimc,gen(𝒱)\operatorname{mdim_{c,gen}}(\mathcal{V}), is more natural. We observe that 0mdimc,gen(𝒱)mdimc(𝒱)dim𝒱0\leq\operatorname{mdim_{c,gen}}(\mathcal{V})\leq\operatorname{mdim_{c}}(\mathcal{V})\leq\dim\mathcal{V}. Clearly, mdimc,gen(𝒱)=mdimc(𝒱)=0\operatorname{mdim_{c,gen}}(\mathcal{V})=\operatorname{mdim_{c}}(\mathcal{V})=0 if 𝒱\mathcal{V} is affine, and mdimc,gen(𝒱)=mdimc(𝒱)=dim𝒱\operatorname{mdim_{c,gen}}(\mathcal{V})=\operatorname{mdim_{c}}(\mathcal{V})=\dim\mathcal{V} if 𝒱\mathcal{V} is projective.

In this paper we determine mdimc,gen(𝒜g)\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g}) and mdimc(𝒜g)\operatorname{mdim_{c}}({\mathcal{A}}_{g}) for the moduli space 𝒜g{\mathcal{A}}_{g} of complex principally polarized abelian varieties (ppav) of dimension gg, 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 {\mathbb{C}}.

1.1. Compact subvarieties through a very general point of 𝒜g{\mathcal{A}}_{g}

If 𝒱¯𝒱\overline{\mathcal{V}}\supsetneq\mathcal{V} is a projective compactification of 𝒱\mathcal{V}, such that the codimension of the boundary codim𝒱¯(𝒱¯𝒱)=d\operatorname{codim}_{\overline{\mathcal{V}}}(\overline{\mathcal{V}}\setminus\mathcal{V})=d, then by embedding 𝒱¯N\overline{\mathcal{V}}\hookrightarrow{\mathbb{P}}^{N} and recursively choosing very general hypersurfaces H1,,Hnd+1H_{1},\dots,H_{n-d+1} such that dimH1Hi(𝒱¯𝒱)=(nd)i\dim H_{1}\cap\dots\cap H_{i}\cap(\overline{\mathcal{V}}\setminus\mathcal{V})=(n-d)-i for any ii, it follows that H1Hnd+1𝒱¯𝒱H_{1}\cap\dots\cap H_{n-d+1}\cap\overline{\mathcal{V}}\subset\mathcal{V} is a compact subvariety, which moreover can be chosen to go through any given finite collection of points of 𝒱\mathcal{V}. Thus the existence of such a compactification 𝒱¯\overline{\mathcal{V}} implies mdimc,gen(𝒱)d1\operatorname{mdim_{c,gen}}(\mathcal{V})\geq d-1. We note, however, that in general there is no implication going the other way, as is easily seen by considering a×ba+b{\mathbb{P}}^{a}\times{\mathbb{C}}^{b}\subset{\mathbb{P}}^{a+b} for various a,ba,b.

The Satake compactification 𝒜g=𝒜g𝒜g1𝒜0{\mathcal{A}}_{g}^{*}={\mathcal{A}}_{g}\sqcup{\mathcal{A}}_{g-1}\sqcup\dots\sqcup{\mathcal{A}}_{0} of 𝒜g{\mathcal{A}}_{g} has boundary of codimension gg, which by the above implies mdimc,gen(𝒜g)g1\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g})\geq g-1. Our first result is that this is sharp.

Theorem A.

For p𝒜gp\in{\mathcal{A}}_{g} a very general point, the maximal dimension of a compact subvariety of 𝒜g{\mathcal{A}}_{g} containing pp is equal to g1g-1; that is,

mdimc,gen(𝒜g)=g1.\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g})=g-1\,.

What we prove is in fact a more precise statement, showing that the maximal dimension of a compact subvariety X𝒜gX\subsetneq{\mathcal{A}}_{g} containing a Hodge-generic point p𝒜gp\in{\mathcal{A}}_{g} (which, as we recall below, is equivalent to saying that pp does not belong to any of the countably many proper Shimura subvarieties of 𝒜g{\mathcal{A}}_{g}, see also the end of Section 2.2) is equal to g1g-1. This will be the special easier case of 3.4, which establishes this statement for a Hodge-generic point p𝒮p\in{\mathcal{S}} in any non-compact Shimura subvariety 𝒮𝒜g{\mathcal{S}}\subseteq{\mathcal{A}}_{g}.

1.2. Compact subvarieties of 𝒜g{\mathcal{A}}_{g}

Turning to the question of maximal dimension of all compact subvarieties, the best known upper bound for 𝒜g{\mathcal{A}}_{g} is the 20 year old result of Keel and Sadun [KS03], who proved mdimc(𝒜g)g(g1)21\operatorname{mdim_{c}}({\mathcal{A}}_{g})\leq\tfrac{g(g-1)}{2}-1 for any g3g\geq 3. This was conjectured by Oort, and underscored the difference between 𝒜g{\mathcal{A}}_{g}\otimes{\mathbb{C}} and 𝒜g𝔽p{\mathcal{A}}_{g}\otimes{\mathbb{F}}_{p}, as in finite characteristic there does exist a complete subvariety of 𝒜g𝔽p{\mathcal{A}}_{g}\otimes{\mathbb{F}}_{p} of codimension gg — the locus of ppav whose subscheme of pp-torsion points is supported at zero.

We determine precisely the maximal dimension of compact subvarieties of 𝒜g{\mathcal{A}}_{g}, for all genera.

Theorem B.

The maximal dimension of a compact subvariety of 𝒜g{\mathcal{A}}_{g} is

mdimc(𝒜g)={g1if g<16g216for even g16(g1)216for odd g17.\operatorname{mdim_{c}}({\mathcal{A}}_{g})=\begin{cases}\ \ g-1\ &\mbox{if }g<16\\[4.0pt] \ \ \left\lfloor\tfrac{g^{2}}{16}\right\rfloor&\mbox{for even }g\geq 16\\[10.0pt] \left\lfloor\tfrac{(g-1)^{2}}{16}\right\rfloor&\mbox{for odd }g\geq 17\,.\end{cases}

In Table 1 we give our results and the bound of Keel-Sadun, for comparison, in some small genera, and also for g=100g=100 to underscore the difference of the growth rates.

g345615161718100Theorem A:mdimc,gen(𝒜g)=23451415161799Theorem B:mdimc(𝒜g)=234514161620625Keel-Sadun:mdimc(𝒜g)259141041191351522449\begin{array}[]{|lr||rrrr||rrrr||r|}\hline\cr&g&3&4&5&6&15&16&17&18&100\\ \hline\cr\text{Theorem A:}&\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g})=&2&3&4&5&14&15&16&17&99\\ \text{Theorem B:}&\operatorname{mdim_{c}}({\mathcal{A}}_{g})=&2&3&4&5&14&16&16&20&625\\ \text{\it Keel-Sadun:}&\operatorname{mdim_{c}}({\mathcal{A}}_{g})\leq&2&5&9&14&104&119&135&152&2449\\ \hline\cr\end{array}

Table 1. Maximal dimensions of compact subvarieties of 𝒜g{\mathcal{A}}_{g}

What we actually prove is much stronger: we show that any compact subvariety of 𝒜g{\mathcal{A}}_{g} of dimension at least gg is contained in the product of a (k1)(k-1)-dimensional compact subvariety of 𝒜k{\mathcal{A}}_{k}, for some 0k<g0\leq k<g, and a compact Shimura subvariety of 𝒜gk{\mathcal{A}}_{g-k} (see 2.9 and 3.4). We then determine the maximal dimension of compact Shimura subvarieties of 𝒜g{\mathcal{A}}_{g}. In fact our proof and analyses of compact Shimura subvarieties of 𝒜g{\mathcal{A}}_{g} allows us to describe all maximal-dimensional compact subvarieties of 𝒜g{\mathcal{A}}_{g} — 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 𝒜g{\mathcal{A}}_{g}.

1.3. Compact subvarieties of gct{\mathcal{M}}_{g}^{\rm ct}

Recall that the Torelli map sending a smooth complex curve to its Jacobian is an injection J:g𝒜gJ:{\mathcal{M}}_{g}\hookrightarrow{\mathcal{A}}_{g} 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 𝒜gind:=𝒜g(𝒜1×𝒜g1𝒜2×𝒜g2){\mathcal{A}}_{g}^{\rm ind}:={\mathcal{A}}_{g}\setminus({\mathcal{A}}_{1}\times{\mathcal{A}}_{g-1}\cup{\mathcal{A}}_{2}\times{\mathcal{A}}_{g-2}\cup\dots) of indecomposable abelian varieties, and determining mdimc(𝒜gind)\operatorname{mdim_{c}}({\mathcal{A}}_{g}^{\rm ind}) and mdimc,gen(𝒜gind)\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g}^{\rm ind}) would also be very interesting, see Section 6.1.

Remark 1.1.

The classes of the loci of products 𝒜g×𝒜gg{\mathcal{A}}_{g^{\prime}}\times{\mathcal{A}}_{g-g^{\prime}} inside 𝒜g{\mathcal{A}}_{g} 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 J:gct𝒜gJ:{\mathcal{M}}_{g}^{\rm ct}\to{\mathcal{A}}_{g} from the moduli space of curves of compact type. As a corollary of their bound mdimc(𝒜3)2\operatorname{mdim_{c}}({\mathcal{A}}_{3})\leq 2 (and the theorem of Diaz [Dia84]), Keel and Sadun [KS03] deduce the bound mdimc(gct)2g4\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct})\leq 2g-4 for any g3g\geq 3. Our results also have implications for gct{\mathcal{M}}_{g}^{\rm ct}.

Corollary C.

The following equality and upper bound hold:

mdimc(gct)\displaystyle\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct}) =3g/22\displaystyle=\lfloor 3g/2\rfloor-2 for any 2g232\leq g\leq 23 ,
mdimc(J(gct))\displaystyle\operatorname{mdim_{c}}(J({\mathcal{M}}_{g}^{\rm ct})) g1\displaystyle\leq g-1 for any 2g152\leq g\leq 15 .

The point of the difference between these numbers is that JJ is injective at a generic point of gct{\mathcal{M}}_{g}^{\rm ct}, but not along gct\partial{\mathcal{M}}_{g}^{\rm ct}, 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 g3g\geq 3 we construct a compact subvariety of gctg{\mathcal{M}}_{g}^{\rm ct}\setminus{\mathcal{M}}_{g} of dimension 3g/22\lfloor 3g/2\rfloor-2, starting from a compact curve in 2ct{\mathcal{M}}_{2}^{\rm ct}.

The second inequality implies the bound on the maximal dimension of compact subvarieties of gct{\mathcal{M}}_{g}^{\rm ct} that intersect g{\mathcal{M}}_{g}, and follows from B (in fact B also improves Keel and Sadun’s bound mdimc(gct)2g4\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct})\leq 2g-4 to mdimc(gct)(g/2)2/4\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct})\leq\lfloor(\lfloor g/2\rfloor)^{2}/4\rfloor for g28g\leq 28).

It is tempting to conjecture that in fact mdimc(J(gct))g1\operatorname{mdim_{c}}(J({\mathcal{M}}_{g}^{\rm ct}))\leq g-1 for all g2g\geq 2, which would then imply mdimc(gct)=3g/22\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct})=\lfloor 3g/2\rfloor-2 for all g2g\geq 2. 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 X𝒜gX\subsetneq{\mathcal{A}}_{g} is a compact subvariety of dimension at least gg. Note that for a fixed elliptic curve EE, the codimension of the locus [E]×𝒜g1𝒜g[E]\times{\mathcal{A}}_{g-1}\subsetneq{\mathcal{A}}_{g} is equal to gg. If we could ensure that the intersection of [E]×𝒜g1[E]\times{\mathcal{A}}_{g-1} with XX were non-empty and transverse, then such property would hold for a Zariski open subset of [E]𝒜1[E]\in{\mathcal{A}}_{1}, i.e. for all but finitely many [E]𝒜1[E]\in{\mathcal{A}}_{1}. By taking the limit as [E][E] approaches the boundary point [i]𝒜1[i\infty]\in\partial{\mathcal{A}}_{1}^{*}, this contradicts the compactness of XX, thus proving that dimXg1\dim X\leq g-1.

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 𝒜g{\mathcal{A}}_{g} 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 XX^{\prime} of a non-compact Shimura variety 𝒮𝒜g{\mathcal{S}}\subseteq{\mathcal{A}}_{g^{\prime}}, the dimension of XX^{\prime} is bounded above as dimXg1\dim X^{\prime}\leq g^{\prime}-1. This argument, applied to 𝒮=𝒜g{\mathcal{S}}={\mathcal{A}}_{g}, itself gives a stronger version of A, for all Hodge-generic compact subvarieties X𝒜gX\subsetneq{\mathcal{A}}_{g}. We then show that, if a compact X𝒜gX\subsetneq{\mathcal{A}}_{g} 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 𝒮×𝒮′′{\mathcal{S}}^{\prime}\times{\mathcal{S}}^{\prime\prime}, with 𝒮𝒜g{\mathcal{S}}^{\prime}\subseteq{\mathcal{A}}_{g^{\prime}} non-compact, and 𝒮′′𝒜g′′{\mathcal{S}}^{\prime\prime}\subseteq{\mathcal{A}}_{g^{\prime\prime}}, with g+g′′=gg^{\prime}+g^{\prime\prime}=g. Thus a maximal-dimensional non-Hodge-generic XX is either 𝒮′′{\mathcal{S}}^{\prime\prime} or X×𝒮′′X^{\prime}\times{\mathcal{S}}^{\prime\prime}, where XX^{\prime} is a (g1)(g^{\prime}-1)-dimensional subvariety of 𝒮𝒜g{\mathcal{S}}^{\prime}\subseteq{\mathcal{A}}_{g^{\prime}}. We thus complete the proof of B by examining the maximal possible dimension of compact Shimura subvarieties of 𝒜g{\mathcal{A}}_{g}, 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 𝒮𝒜g{\mathcal{S}}\subset{\mathcal{A}}_{g} 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 𝒜g{\mathcal{A}}_{g} (see 3.7).

In Section 4 we investigate compact Shimura subvarieties of 𝒜g{\mathcal{A}}_{g} that are induced by certain symplectic representations that we call “decoupled”, and determine which of them have dimension dgd\geq g (see 4.14).

In Section 5 we investigate compact Shimura subvarieties of 𝒜g{\mathcal{A}}_{g} that are induced by symplectic representations that are not decoupled. We show that such subvarieties have dimension either smaller than gg, or smaller than some Shimura variety originating from a decoupled representation. We thus conclude the proof of B.

In Section 6 we discuss related problems for the moduli space of indecomposable abelian varieties, and we prove C for compact subvarieties of the moduli space gct{\mathcal{M}}_{g}^{\rm ct} of curves of compact type.

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 𝒜g{\mathcal{A}}_{g}

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 GG be an algebraic group over a number field 𝔽{\mathbb{F}}; GG is called simple (resp. almost-simple) if every proper closed normal subgroup of GG is trivial (resp. finite).

The group GG is called geometrically simple (resp. geometrically almost-simple) if GG_{\mathbb{C}} is simple (resp. almost-simple), and it is semisimple if GG_{\mathbb{C}} is, namely if all connected normal closed subgroups of GG_{\mathbb{C}} are trivial.

If GG is almost-simple, then Z(G)Z(G) is finite and Gad:=G/Z(G)G^{\mathrm{ad}}:=G/Z(G) is simple; hence, GG cannot be isogenous to G1×G2G_{1}\times G_{2} for any positive-dimensional algebraic groups G1G_{1} and G2G_{2}.

An isogeny is a surjective homomorphism of algebraic groups GGG^{\prime}\rightarrow G with finite kernel; the kernel is then necessarily contained in Z(G)Z(G^{\prime}). An algebraic group GG is simply connected if every isogeny GGG^{\prime}\rightarrow G is an isomorphism.

If GG is semisimple and G1,,GrG_{1},\dots,G_{r} are all the almost-simple closed normal subgroups of GG, then the product map G1××GrGG_{1}\times\dots\times G_{r}\rightarrow G is an isogeny (see [Mil17, Theorem 22.121]). The Galois group Gal(/𝔽)\operatorname{Gal}({\mathbb{C}}/{\mathbb{F}}) fixes 𝔽{\mathbb{F}}, and thus it acts on GG_{\mathbb{C}} by permuting its almost-simple factors and by fixing G𝔽G_{\mathbb{F}}. As a consequence, the stabilizer of the factors of GG_{\mathbb{C}} has a number field 𝔽𝔽{\mathbb{F}}^{\prime}\supseteq{\mathbb{F}} as fixed field, and so G𝔽G_{{\mathbb{F}}^{\prime}} splits as a product of geometrically almost-simple factors.

If H𝔽H_{\mathbb{F}} is a geometrically almost-simple group over a totally real number field 𝔽{\mathbb{F}} and G=H𝔽/G=H_{{\mathbb{F}}/{\mathbb{Q}}} is the {\mathbb{Q}}-group obtained from H𝔽H_{\mathbb{F}} by restriction of the scalars, then ProdσHσ,G\operatorname{Prod}_{\sigma}H_{\sigma,{\mathbb{R}}}\rightarrow G_{\mathbb{R}} is an isogeny, where σ\sigma ranges over all embeddings 𝔽{\mathbb{F}}\hookrightarrow{\mathbb{R}} and Hσ,:=σ(H𝔽)σ(𝔽)H_{\sigma,{\mathbb{R}}}:=\sigma(H_{{\mathbb{F}}})\otimes_{\sigma({\mathbb{F}})}{\mathbb{R}}.

A split torus TT is a product of a number of copies of the multiplicative group 𝔾m\mathbb{G}_{m}. An algebraic group is called anisotropic if it contains no positive-dimensional split tori. A semisimple group GG over a number field 𝔽{\mathbb{F}} is anisotropic if and only if G𝔽G_{\mathbb{F}} 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.

In this section we recall definitions and basic properties, see also [Mil05] and [Mil11].

We will denote by (Gad)+(G^{\mathrm{ad}}_{{\mathbb{R}}})^{+} the connected component of GadG^{\mathrm{ad}}_{\mathbb{R}} that contains the identity.

Definition 2.1 ([Mil05, 4.4]).

A connected Shimura datum is a pair (G,𝒟+)(G,\mathcal{D}^{+}) where GG is a connected semisimple algebraic group over a number field, and 𝒟+\mathcal{D}^{+} is a (Gad)+(G^{\mathrm{ad}}_{\mathbb{R}})^{+}-conjugacy class of homomorphisms of real algebraic groups

u:S1(Gad)+u:S^{1}\longrightarrow(G^{\mathrm{ad}}_{\mathbb{R}})^{+}

such that

  • (i)

    there is a direct sum decomposition 𝔤=𝔩𝔭+𝔭\mathfrak{g}_{{\mathbb{C}}}=\mathfrak{l}\oplus\mathfrak{p}^{+}\oplus\mathfrak{p}^{-} of the Lie algebra of GG_{\mathbb{C}}, such that Ad(u(z))\operatorname{Ad}(u(z)) acts trivially on 𝔩\mathfrak{l}, as zz on 𝔭+\mathfrak{p}^{+} and as z¯\bar{z} on 𝔭\mathfrak{p}^{-}, via the representation

    S1u(Gad)+AdEnd(𝔤)S^{1}\stackrel{{\scriptstyle u}}{{\longrightarrow}}(G^{\mathrm{ad}}_{\mathbb{R}})^{+}\stackrel{{\scriptstyle\operatorname{Ad}}}{{\longrightarrow}}\operatorname{End}(\mathfrak{g}_{\mathbb{C}})
  • (ii)

    The conjugation by θ:=u(1)\theta:=u(-1) is a Cartan involution, namely Gθ:={gGad|θg¯θ1=g}G^{\theta}:=\{g\in G^{\mathrm{ad}}_{\mathbb{C}}\,|\,\theta\bar{g}\theta^{-1}=g\} is a compact real form of GadG^{\mathrm{ad}}_{\mathbb{C}} (and so an inner form of GadG^{\mathrm{ad}}_{\mathbb{R}})

  • (iii)

    GadG^{\mathrm{ad}} has no nontrivial {\mathbb{Q}}-simple factor HH such that HH_{\mathbb{R}} is compact.

A morphism of connected Shimura data (G1,𝒟1+)(G2,𝒟2+)(G_{1},\mathcal{D}^{+}_{1})\rightarrow(G_{2},\mathcal{D}^{+}_{2}) is a homomorphism f:G1G2f:G_{1}\rightarrow G_{2} of algebraic groups such that the map f¯:(G1,ad)+(G2,ad)+\bar{f}:(G^{\mathrm{ad}}_{1,{\mathbb{R}}})^{+}\rightarrow(G^{\mathrm{ad}}_{2,{\mathbb{R}}})^{+} induced by ff induces a morphism between 𝒟1+\mathcal{D}^{+}_{1} and 𝒟2+\mathcal{D}^{+}_{2}.

Remark 2.2.

In the above definition of connected Shimura datum, the relevant information seem to be carried by GadG^{\mathrm{ad}} and uu. The relevance of GG will be clear in Section 2.3 below, where we will see that a morphism from a connected Shimura variety to 𝒜g{\mathcal{A}}_{g} will be associated to a homomorphism GSp2g,G\rightarrow\operatorname{Sp}_{2g,{\mathbb{Q}}} that does not necessarily descend to GadG^{\mathrm{ad}}.

Example 2.3.

Let τ0:=iIgg\tau_{0}:=i\cdot I_{g}\in\mathcal{H}_{g}, where IgI_{g} is the g×gg\times g identity matrix, and g\mathcal{H}_{g} denotes the Siegel space of complex symmetric matrices with positive-definite imaginary part. Then 𝔰𝔭2g()\mathfrak{sp}_{2g}({\mathbb{R}}) decomposes as the orthogonal sum 𝔰𝔱\mathfrak{s}\oplus\mathfrak{t}, where 𝔰\mathfrak{s} is the Lie algebra of the stabilizer of τ0\tau_{0}, and 𝔱=𝔰\mathfrak{t}=\mathfrak{s}^{\perp}. The infinitesimal action of PSp2g()\operatorname{PSp}_{2g}({\mathbb{R}}) on g\mathcal{H}_{g} at τ0\tau_{0} induces the isomorphism Tτ0g𝔱T_{\tau_{0}}\mathcal{H}_{g}\cong\mathfrak{t}. After complexifying, we have psp2g()=𝔰𝔱1,0𝔱0,1\mathrm{psp}_{2g}({\mathbb{C}})=\mathfrak{s}_{{\mathbb{C}}}\oplus\mathfrak{t}^{1,0}\oplus\mathfrak{t}^{0,1}, where 𝔰=𝔰\mathfrak{s}_{{\mathbb{C}}}=\mathfrak{s}\otimes_{\mathbb{R}}{\mathbb{C}}, 𝔱1,0Tτ01,0g\mathfrak{t}^{1,0}\cong T^{1,0}_{\tau_{0}}\mathcal{H}_{g}, 𝔱0,1Tτ00,1g\mathfrak{t}^{0,1}\cong T^{0,1}_{\tau_{0}}\mathcal{H}_{g}. Note that the homomorphism u:S1PSp2g()u:S^{1}\rightarrow\operatorname{PSp}_{2g}({\mathbb{R}}) defined as

u(eiϑ):=[cos(ϑ/2)Igsin(ϑ/2)Igsin(ϑ/2)Igcos(ϑ/2)Ig]u(e^{i\vartheta}):=\left[\begin{array}[]{cc}\cos(\vartheta/2)I_{g}&-\sin(\vartheta/2)I_{g}\\ \sin(\vartheta/2)I_{g}&\cos(\vartheta/2)I_{g}\end{array}\right]

fixes τ0\tau_{0}. Moreover u(z)u(z) acts trivially on 𝔰\mathfrak{s}_{{\mathbb{C}}}, as zz on 𝔭1,0\mathfrak{p}^{1,0} and as z¯\bar{z} on 𝔭0,1\mathfrak{p}^{0,1}.

The group GG_{\mathbb{R}} splits (up to isogeny) into a product of geometrically almost-simple factors: denote by GcG_{c} the product of all its compact factors, and by Gnc=G1××GkG_{nc}=G_{1}\times\dots\times G_{k} the product of all its non-compact factors. Now, 𝒟+\mathcal{D}^{+} is a connected Hermitian symmetric domain, and (Gad)+(G^{\mathrm{ad}}_{\mathbb{R}})^{+} acts on 𝒟+\mathcal{D}^{+} transitively, isometrically, and holomorphically. The group S1S^{1} acts on 𝒟+\mathcal{D}^{+} via Adu\operatorname{Ad}\circ u by fixing the point uu and by complex multiplication on Tu𝒟+T_{u}\mathcal{D}^{+}. The compact subgroup Gθ(Gad)+G^{\theta}\cap(G^{\mathrm{ad}}_{\mathbb{R}})^{+} is the connected component of the identity of the stabilizer of u𝒟+u\in\mathcal{D}^{+}.

Note that 𝒟+=𝒟1+××𝒟k+\mathcal{D}^{+}=\mathcal{D}_{1}^{+}\times\dots\times\mathcal{D}^{+}_{k} with 𝒟i+=Giad/Ki\mathcal{D}^{+}_{i}=G^{\mathrm{ad}}_{i}/K_{i}, where Ki=GθGiadK_{i}=G^{\theta}\cap G^{\mathrm{ad}}_{i} is a compact inner form of GiadG^{\mathrm{ad}}_{i}, i.e. GiadG^{\mathrm{ad}}_{i} has an inner automorphism which is a Cartan involution. We claim that GiG_{i} is geometrically almost-simple (see the proof of [Mil11, Theorem 3.13]). Indeed, if GiG_{i} were not geometrically almost-simple, then Gi,G_{i,{\mathbb{C}}} would be isomorphic to the real group H/H_{{\mathbb{C}}/{\mathbb{R}}} obtained from a complex group HH_{\mathbb{C}} by the restriction of scalars. But then such GiG_{i} has no compact inner form by the following:

Lemma 2.4.

Let HH_{\mathbb{C}} be a complex group of positive dimension, and let H/H_{{\mathbb{C}}/{\mathbb{R}}} be the real group obtained from HH_{\mathbb{C}} by the restriction of scalars. Then there is no element γ\gamma of the complexification (H/)(H_{{\mathbb{C}}/{\mathbb{R}}})_{\mathbb{C}} such that Adγ\operatorname{Ad}_{\gamma} is a Cartan involution.

Proof.

Note that the homomorphism ι:H/H×H\iota:H_{{\mathbb{C}}/{\mathbb{R}}}\hookrightarrow H_{\mathbb{C}}\times H_{\mathbb{C}}, defined as ι(h):=(h,h¯)\iota(h):=(h,\bar{h}), induces an isomorphism of the complexification (H/)H×H(H_{{\mathbb{C}}/{\mathbb{R}}})_{\mathbb{C}}\cong H_{\mathbb{C}}\times H_{\mathbb{C}}. Moreover, the image of ι\iota is fixed under the involution c(g,g):=(g¯,g¯)c(g,g^{\prime}):=(\bar{g^{\prime}},\bar{g}) of H×HH_{\mathbb{C}}\times H_{\mathbb{C}}.

For contradiction, suppose that θ=Adγ\theta=\operatorname{Ad}_{\gamma} is a Cartan involution of (H/)(H_{{\mathbb{C}}/{\mathbb{R}}})_{\mathbb{C}}, for some γH/\gamma\in H_{{\mathbb{C}}/{\mathbb{R}}}. This would mean γ2Z(H/)\gamma^{2}\in Z(H_{{\mathbb{C}}/{\mathbb{R}}}) and thus that HθH^{\theta} can be written as Hθ={(h,h)H×H|Adγ(c(h,h))=(h,h)}H^{\theta}=\{(h,h^{\prime})\in H\times H\,|\,\operatorname{Ad}_{\gamma}(c(h,h^{\prime}))=(h,h^{\prime})\}. It is immediate to check that then (h,h)Hθ(h,h^{\prime})\in H^{\theta} if and only if h=Adγ¯h¯h^{\prime}=\operatorname{Ad}_{\bar{\gamma}}\bar{h}. Thus H/h(h,Adγ¯h¯)HθH_{{\mathbb{C}}/{\mathbb{R}}}\ni h\mapsto(h,\operatorname{Ad}_{\bar{\gamma}}\bar{h})\in H^{\theta} is an isomorphism. This is a contradiction, since H/H_{{\mathbb{C}}/{\mathbb{R}}} is non-compact, while by definition of the Shimura datum HθH^{\theta} must be compact. ∎

Recall that a subgroup ΓGnc\Gamma\subset G_{nc} is called arithmetic if there exists a simply connected rational algebraic group GG^{\prime} and a surjective homomorphism φ:GGnc\varphi:G^{\prime}_{\mathbb{R}}\rightarrow G_{nc} with compact kernel such that Γ\Gamma is commensurable to φ(G)\varphi(G^{\prime}_{\mathbb{Z}}). An arithmetic subgroup is automatically a lattice. Moreover, denoting Γi:=ΓGi(F)\Gamma_{i}:=\Gamma\cap G_{i}(F), the product Γ1××Γk\Gamma_{1}\times\dots\times\Gamma_{k} is then a finite index subgroup of Γ\Gamma, and so the map i(Γi\𝒟i+)Γ\𝒟+\prod_{i}(\Gamma_{i}\backslash\mathcal{D}^{+}_{i})\rightarrow\Gamma\backslash\mathcal{D}^{+} is a finite étale cover.

As every arithmetic subgroup Γ\Gamma of GncG_{nc} contains a torsion-free normal subgroup Γ\Gamma^{\prime} of finite index, the connected Shimura variety 𝒮=Γ\𝒟+{\mathcal{S}}=\Gamma\backslash\mathcal{D}^{+} is a quotient of 𝒮=Γ\𝒟+{\mathcal{S}}^{\prime}=\Gamma^{\prime}\backslash\mathcal{D}^{+} by the finite group Γ/Γ\Gamma/\Gamma^{\prime} (see [Mil11, Theorem 3.6]). Moreover, 𝒮{\mathcal{S}}^{\prime} is a smooth quasi-projective variety over {\mathbb{C}} (see [Mil11, Theorems 4.2-4.3]) and so 𝒮{\mathcal{S}} 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 (G,𝒟+)(G,\mathcal{D}^{+}) is indecomposable if GG is almost-simple. As for the compactness, we recall the following criterion.

Lemma 2.5 ([PR94, page 210, Theorem 4.12]).

The Shimura varieties 𝒮{\mathcal{S}} and 𝒮{\mathcal{S}}^{\prime} are compact if and only if G𝔽G_{\mathbb{F}} is anisotropic.

If 𝒮j=Γj\𝒟j+{\mathcal{S}}_{j}=\Gamma_{j}\backslash\mathcal{D}_{j}^{+} is a Shimura variety associated to the Shimura datum (Gj,𝒟j+)(G_{j},\mathcal{D}^{+}_{j}) for j=1,2j=1,2, a morphism of Shimura varieties 𝒮1𝒮2{\mathcal{S}}_{1}\rightarrow{\mathcal{S}}_{2} is a map induced by a homomorphism ρ:G1G2\rho:G_{1}\rightarrow G_{2} of algebraic groups such that ρ(𝒟1+)𝒟2+\rho(\mathcal{D}_{1}^{+})\subseteq\mathcal{D}_{2}^{+} and ρ(Γ1)Γ2\rho(\Gamma_{1})\subseteq\Gamma_{2}.

A point of the Shimura variety 𝒮{\mathcal{S}} corresponding to the homomorphism u:S1(Gad)+u:S^{1}\rightarrow(G^{\mathrm{ad}}_{\mathbb{R}})^{+} is called Hodge-generic if the smallest algebraic {\mathbb{Q}}-subgroup GG^{\prime} of GadG^{\mathrm{ad}} such that GG^{\prime}_{\mathbb{R}} contains the image of uu is the whole GadG^{\mathrm{ad}}. A subvariety X𝒮X\subseteq{\mathcal{S}} is called Hodge-generic (within 𝒮{\mathcal{S}}) if its general point is. This is equivalent to requiring XX to contain a point that is Hodge-generic in 𝒮{\mathcal{S}}.

A special subvariety of 𝒮{\mathcal{S}} is a subvariety induced by a Shimura sub-datum of (G,𝒟+)(G,\mathcal{D}^{+}). A weakly special subvariety of 𝒮{\mathcal{S}} is either a point, or a special subvariety, or a subvariety 𝒮H×{p}{\mathcal{S}}_{H}\times\{p\} of a special subvariety 𝒮H×𝒮K{\mathcal{S}}_{H}\times{\mathcal{S}}_{K} of 𝒮{\mathcal{S}}, where pp is a Hodge-generic point in 𝒮K{\mathcal{S}}_{K}, up to isogeny.

2.3. Symplectic representations

Let GG be a connected semisimple algebraic group over {\mathbb{Q}} and let (V,ω)(V,\omega) be a rational symplectic representation of GG (i.e. a homomorphism GSp(V,ω)G\to\operatorname{Sp}(V,\omega), for VV a vector space over {\mathbb{Q}} with a symplectic form ω\omega). A symplectic subspace of VV is a vector subspace on which ω\omega restricts to a (non-degenerate) symplectic form.

Definition 2.6.

Given a representation VV of GG, an invariant linear summand of VV is a GG-invariant linear subspace; a linear summand is irreducible if it contains no nontrivial GG-invariant linear subspaces. Given a symplectic representation VV of GG, a sub-representation of VV a GG-invariant symplectic subspace; such sub-representation is Sp\operatorname{Sp}-irreducible if it has no nontrivial sub-representations.

Lemma 2.7.

Any symplectic representation VV is a direct sum of Sp\operatorname{Sp}-irreducible sub-representations.

Proof.

We proceed by induction on dimV\dim V. The case dimV=0\dim V=0 is trivial.

Suppose dimV>0\dim V>0 and let WVW\subset V be a nontrivial irreducible linear summand. The pairing ωW\omega\mid_{W} induces a map ϕ:WW\phi:W\rightarrow W^{\vee} which is either zero or an isomorphism, since ker(ϕ)\mathrm{ker}(\phi) is a linear summand of WW.

If ϕ\phi is an isomorphism, then WW is a symplectic representation and hence Sp\operatorname{Sp}-irreducible. Moreover, WW^{\perp} is a direct sum of Sp\operatorname{Sp}-irreducibles by induction, and we are done.

We may thus assume that for any irreducible linear summand of VV, the restriction of ω\omega to it is zero. Then the pairing against WW with ω\omega gives a map ψ:VW\psi:V\rightarrow W^{\vee}. Since ω\omega is non-degenerate, the map ψ\psi is nonzero and so surjective (by the irreducibility of WW^{\vee}). Let WVW^{\prime}\subset V be an irreducible linear summand not contained inside ker(ψ)\ker(\psi). Then the induced map ψ:WW\psi^{\prime}:W^{\prime}\rightarrow W^{\vee} is an isomorphism. Since WW is isotropic for ω\omega, it follows that WW={0}W\cap W^{\prime}=\{0\} and so (WW,ω)(W\oplus W^{\prime},\omega) is a symplectic representation. Moreover, it must be Sp\operatorname{Sp}-irreducible, as its only proper non-zero invariant subspaces are themselves irreducible, and hence by our assumption the restriction of ω\omega to them is zero. It follows that (WW)(W\oplus W^{\prime})^{\perp} is a sub-representation of VV and so it is a direct sum of Sp\operatorname{Sp}-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 G1,G_{1,{\mathbb{R}}} and G2,G_{2,{\mathbb{R}}} are non-compact almost-simple real algebraic groups, and let VV_{\mathbb{C}} be an irreducible complex symplectic representation of G1,×G2,G_{1,{\mathbb{R}}}\times G_{2,{\mathbb{R}}}. Then either G1,G_{1,{\mathbb{R}}} or G2,G_{2,{\mathbb{R}}} acts trivially on VV_{\mathbb{C}}.

2.4. Products of Shimura subvarieties in 𝒜g{\mathcal{A}}_{g}

Using the results recalled above, we now study Shimura subvarieties of 𝒜g{\mathcal{A}}_{g} that are a product of a non-compact Shimura variety and another Shimura variety, showing that then each factor embeds into a suitable 𝒜k{\mathcal{A}}_{k}, and that, up to isogeny, the embedding of the product is the product of the embeddings.

Theorem 2.9.

Let 𝒮1{\mathcal{S}}_{1} be a non-compact indecomposable Shimura variety, and let 𝒮2{\mathcal{S}}_{2} be any other Shimura variety. Then any embedding 𝒮1×𝒮2𝒜g{\mathcal{S}}_{1}\times{\mathcal{S}}_{2}\hookrightarrow{\mathcal{A}}_{g} of Shimura varieties factors, up to isogeny, as 𝒮1𝒜g1{\mathcal{S}}_{1}\hookrightarrow{\mathcal{A}}_{g_{1}}, 𝒮2𝒜g2{\mathcal{S}}_{2}\hookrightarrow{\mathcal{A}}_{g_{2}}, followed by the natural embedding 𝒜g1×𝒜g2𝒜g{\mathcal{A}}_{g_{1}}\times{\mathcal{A}}_{g_{2}}\hookrightarrow{\mathcal{A}}_{g}, for some g1+g2=gg_{1}+g_{2}=g.

Proof.

Denote G1,G2G_{1},G_{2} the groups associated to 𝒮1,𝒮2{\mathcal{S}}_{1},{\mathcal{S}}_{2} (which thus have no rational compact factor), and denote VV the rational symplectic representation of G1×G2G_{1}\times G_{2} that induces the map 𝒮1×𝒮2𝒜g{\mathcal{S}}_{1}\times{\mathcal{S}}_{2}\hookrightarrow{\mathcal{A}}_{g}. Then VV decomposes as a direct sum of irreducible G1×G2G_{1}\times G_{2}-invariant linear subspaces; in particular, by suitably regrouping the summands, we can write V=V0V1V2V1,2V=V_{0}\oplus V_{1}\oplus V_{2}\oplus V_{1,2}, where G1G_{1} acts nontrivially on each irreducible linear summand of V1V1,2V_{1}\oplus V_{1,2} and trivially on V0V2V_{0}\oplus V_{2}, and G2G_{2} acts nontrivially on each irreducible linear summand of V2V1,2V_{2}\oplus V_{1,2} and trivially on V0V1V_{0}\oplus V_{1}. It is easy to see that V0V_{0}, V1V_{1}, V2V_{2} and V1,2V_{1,2} are sub-representations.

We claim that V1,2={0}V_{1,2}=\{0\}; then the conclusion of the theorem follows, as G1Sp(V1)G_{1}\rightarrow\operatorname{Sp}(V_{1}) and G2Sp(V0V2)G_{2}\rightarrow\operatorname{Sp}(V_{0}\oplus V_{2}) are rational representations.

In order to prove the claim, for contradiction let UU be a nontrivial irreducible linear summand V1,2,V_{1,2,{\mathbb{C}}}. We write U=U1U2U=U_{1}\otimes U_{2}, where UjU_{j} is a nontrivial irreducible linear representation of GjG_{j} for j=1,2j=1,2.

Up to isogeny, Gj,G_{j,{\mathbb{R}}} decomposes as Gj,c×Gj,ncG_{j,c}\times G_{j,nc}, where Gj,cG_{j,c} (resp. Gj,ncG_{j,nc}) is a product of almost-simple compact (resp. non-compact) real groups. Since 𝒮1,𝒮2{\mathcal{S}}_{1},{\mathcal{S}}_{2} have positive dimension (otherwise the theorem is trivial), both G1,ncG_{1,nc} and G2,ncG_{2,nc} are nontrivial. On the other hand, Gj=Res𝔽j/HjG_{j}=\operatorname{Res}_{{\mathbb{F}}_{j}/{\mathbb{Q}}}H_{j}, where HjH_{j} is a geometrically almost-simple algebraic group defined over a totally real number field 𝔽j{\mathbb{F}}_{j}. Hence σ:𝔽jHj,σ,Gj,\prod_{\sigma:{\mathbb{F}}_{j}\hookrightarrow{\mathbb{R}}}H_{j,\sigma,{\mathbb{R}}}\rightarrow G_{j,{\mathbb{R}}} is an isogeny.

Now, there exists σHom(𝔽1,)\sigma\in\operatorname{Hom}({\mathbb{F}}_{1},{\mathbb{R}}) such that Hj,σH_{j,\sigma} acts nontrivially on U1U_{1}: set H1o:=Hj,σH^{o}_{1}:=H_{j,\sigma}. Since 𝒮1{\mathcal{S}}_{1} is indecomposable and non-compact, it follows that G1G_{1} is almost-simple and isotropic: hence all G1,σG_{1,\sigma} are non-compact, and in particular H1oH_{1}^{o} is. By 2.8 it follows that all direct factors of G2,ncG_{2,nc} must act trivially on U2U_{2}. Let σHom(𝔽2,)\sigma^{\prime}\in\operatorname{Hom}({\mathbb{F}}_{2},{\mathbb{R}}) such that H2,σH_{2,\sigma^{\prime}} acts nontrivially on U2U_{2}: set H2o:=H2,σH^{o}_{2}:=H_{2,\sigma^{\prime}}. Since G2G_{2} has no compact {\mathbb{Q}}-factors, there exists ςGal(/)\varsigma\in\mathrm{Gal}({\mathbb{C}}/{\mathbb{Q}}) such that ς(H2o)\varsigma(H^{o}_{2})_{\mathbb{R}} is non-compact. Hence, V1,2,V_{1,2,{\mathbb{C}}} contains a factor ς(U)=ς(U1)ς(U2)\varsigma(U)=\varsigma(U_{1})\otimes\varsigma(U_{2}), which is acted upon nontrivially by the non-compact factor ς(G1o)\varsigma(G_{1}^{o})_{\mathbb{R}} of G1,G_{1,{\mathbb{R}}} and by the non-compact factor ς(G2o)\varsigma(G_{2}^{o})_{\mathbb{R}} of G2,G_{2,{\mathbb{R}}}. 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 (G,𝒟+)(G,\mathcal{D}^{+}) be a connected Shimura datum and let Γ\Gamma be a lattice in GG_{\mathbb{Z}}. Denote by 𝒮{\mathcal{S}} the Shimura variety Γ\𝒟+\Gamma\backslash\mathcal{D}^{+} and by π:𝒟+𝒮\pi:\mathcal{D}^{+}\rightarrow{\mathcal{S}} the natural projection, and let 𝒟ˇ+\check{\mathcal{D}}^{+} be the compact Hermitian domain dual to 𝒟+\mathcal{D}^{+}, which is a projective variety.

Theorem 3.1 (Weak Ax-Schanuel [MPT19]).

Let WW be an algebraic subvariety of 𝒟+\mathcal{D}^{+} (namely, obtained by intersecting 𝒟+\mathcal{D}^{+} with an algebraic subvariety of 𝒟ˇ+\check{\mathcal{D}}^{+}) and let XX be an algebraic subvariety of 𝒮{\mathcal{S}}. If π1(X)W\pi^{-1}(X)\cap W has an analytic irreducible component UU of dimension larger than expected, then π(U)\pi(U) is contained inside a proper weakly special subvariety of 𝒮{\mathcal{S}}.

This will be usually applied when WW is the image inside 𝒟+\mathcal{D}^{+} 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 X𝒮X\subseteq{\mathcal{S}} is an irreducible subvariety and XsmX_{\mathrm{sm}} is its (connected) smooth locus, then the algebraic monodromy of XX at a point xXsmx\in X_{\mathrm{sm}} is the Zariski closure of the image inside GG_{\mathbb{Q}} of the induced homomorphism π1(Xsm,x)Γ\pi_{1}(X_{\mathrm{sm}},x)\rightarrow\Gamma.

Moreover, recall that for a subvariety Y𝒮Y\subseteq{\mathcal{S}}, denoting Y~:=π1(Y)𝒟+\widetilde{Y}:=\pi^{-1}(Y)\subseteq\mathcal{D}^{+}, for any γG\gamma\in G_{\mathbb{Q}} the image π(γY~)\pi(\gamma\cdot\widetilde{Y}) is a subvariety of 𝒮{\mathcal{S}} called the γ\gamma-translate of YY.

Theorem 3.2.

Let X𝒜gX\subseteq{\mathcal{A}}_{g} be a subvariety whose generic point has {\mathbb{Q}}-simple algebraic monodromy GG, and let 𝒮𝒜g{\mathcal{S}}\subseteq{\mathcal{A}}_{g} be the smallest Shimura subvariety containing XX. If Φ:𝒮𝒜g\Phi:{\mathcal{S}}^{\prime}\rightarrow{\mathcal{A}}_{g} is a morphism of Shimura varieties such that

  • (a)

    dimX+dimΦ(𝒮)dim𝒜g\dim X+\dim\Phi({\mathcal{S}}^{\prime})\geq\dim{\mathcal{A}}_{g}, and

  • (b)

    Φ(𝒮)𝒮\Phi({\mathcal{S}}^{\prime})\cap{\mathcal{S}}\neq\emptyset,

then there exists a GG_{\mathbb{Q}}-translate of Φ(𝒮)𝒮\Phi({\mathcal{S}}^{\prime})\cap{\mathcal{S}} inside 𝒮{\mathcal{S}} whose intersection with XX has a component of the expected dimension.

Proof.

Denote by 𝒮~\widetilde{{\mathcal{S}}} the universal cover of 𝒮{\mathcal{S}}, by X~\widetilde{X} the preimage of XX inside 𝒮~\widetilde{{\mathcal{S}}}, and by 𝒮~′′\widetilde{{\mathcal{S}}}^{\prime\prime} the preimage of Φ(𝒮)𝒮\Phi({\mathcal{S}}^{\prime})\cap{\mathcal{S}} inside 𝒮~\widetilde{{\mathcal{S}}}.

Fix x0Xx_{0}\in X a Hodge-generic point of XX, let x~0X~\tilde{x}_{0}\in\widetilde{X} be a preimage of x0x_{0}, and let γ0G\gamma_{0}\in G_{\mathbb{R}} be such that x~0γ0𝒮~′′\tilde{x}_{0}\in\gamma_{0}\cdot\widetilde{{\mathcal{S}}}^{\prime\prime}.

Since there are no proper positive-dimensional weakly special subvarieties of 𝒮{\mathcal{S}} containing x0x_{0}, and since dimX~+dim𝒮~′′dim𝒮~\dim\widetilde{X}+\dim\widetilde{{\mathcal{S}}}^{\prime\prime}\geq\dim\widetilde{{\mathcal{S}}} by (a), it follows from 3.1 that X~(γ0𝒮~′′)\widetilde{X}\cap(\gamma_{0}\cdot\widetilde{{\mathcal{S}}}^{\prime\prime}) (which is non-empty) has an analytic irreducible component of the expected dimension containing x~0\tilde{x}_{0}.

Thus, for any γG\gamma\in G_{\mathbb{R}} sufficiently close to γ0\gamma_{0}, there will also exist an analytic irreducible component of X~(γ𝒮~′′)\widetilde{X}\cap(\gamma\cdot\widetilde{{\mathcal{S}}}^{\prime\prime}) that still has the expected (non-negative) dimension. In particular, since GG_{\mathbb{Q}} is dense in GG_{\mathbb{R}}, such property holds for some rational γG\gamma\in G_{\mathbb{Q}} sufficiently close to γ0\gamma_{0}. ∎

The following application will be useful for us.

Corollary 3.3.

With the same hypotheses as in 3.2, assume moreover that 𝒮=𝒮1×𝒮2{\mathcal{S}}^{\prime}={\mathcal{S}}^{\prime}_{1}\times{\mathcal{S}}^{\prime}_{2} and

  • (a)

    dimX+dim𝒮2dim𝒜g\dim X+\dim{\mathcal{S}}^{\prime}_{2}\geq\dim{\mathcal{A}}_{g}

  • (b)

    Φ({y}×𝒮2)𝒮\Phi(\{y\}\times{\mathcal{S}}^{\prime}_{2})\cap{\mathcal{S}}\neq\emptyset for all y𝒮1y\in{\mathcal{S}}^{\prime}_{1}.

Then

  • (i)

    there exists a γ\gamma-translate of Φ(𝒮)𝒮\Phi({\mathcal{S}}^{\prime})\cap{\mathcal{S}} inside 𝒮{\mathcal{S}} (for some γG\gamma\in G_{\mathbb{Q}}) such that the intersection of the γ\gamma-translate of its slice Φ({y}×𝒮2)𝒮\Phi(\{y\}\times{\mathcal{S}}^{\prime}_{2})\cap{\mathcal{S}} with XX has a component of the expected dimension, for every yy in a (non-empty) Zariski-open subset of 𝒮1{\mathcal{S}}^{\prime}_{1};

  • (ii)

    if 𝒮1{\mathcal{S}}^{\prime}_{1} is non-compact, then XX is non-compact.

Proof.

(i) For every y~𝒮~1\tilde{y}\in\widetilde{{\mathcal{S}}}^{\prime}_{1} denote 𝒮~y~′′\widetilde{{\mathcal{S}}}^{\prime\prime}_{\tilde{y}} the preimage of Φ({y}×𝒮2)𝒮\Phi(\{y\}\times{\mathcal{S}}^{\prime}_{2})\cap{\mathcal{S}} inside 𝒮~\widetilde{{\mathcal{S}}}. Fix y~0𝒮~1\tilde{y}_{0}\in\widetilde{{\mathcal{S}}}^{\prime}_{1}. By 3.2 there exists γG\gamma\in G_{\mathbb{Q}} such that the intersection Xπ(γ𝒮~y~0′′)X\cap\pi(\gamma\cdot\widetilde{{\mathcal{S}}}^{\prime\prime}_{\tilde{y}_{0}}) has a component of the expected dimension (and, in particular, is non-empty). Thus, the same holds for every y~𝒮~1\tilde{y}\in\widetilde{{\mathcal{S}}}^{\prime}_{1} in a sufficiently small neighbourhood of y~0\tilde{y}_{0}. By analyticity of this condition, this then also holds for all y~𝒮~1\tilde{y}\in\widetilde{{\mathcal{S}}}^{\prime}_{1} outside a countable union of proper analytic subvarieties.

(ii) Let γ\gamma be as in (i) and let (y~n)𝒮~1(\tilde{y}_{n})\subset\widetilde{{\mathcal{S}}}^{\prime}_{1} be a sequence of points such that Xπ(γ𝒮~yn~′′)X\cap\pi(\gamma\cdot\widetilde{{\mathcal{S}}}^{\prime\prime}_{\tilde{y_{n}}})\neq\emptyset, while the corresponding sequence of images (yn)𝒮1(y_{n})\subset{\mathcal{S}}^{\prime}_{1} diverges. Then XX contains a diverging sequence, which shows that XX is not compact. ∎

3.2. Hodge-generic compact subvarieties of 𝒜g{\mathcal{A}}_{g}

Given an irreducible subvariety X𝒜gX\subset{\mathcal{A}}_{g}, by definition XX is Hodge-generic within the smallest Shimura subvariety SS of 𝒜g{\mathcal{A}}_{g} containing XX. In this section we prove the following.

Theorem 3.4.

Let 𝒮𝒜g{\mathcal{S}}\subseteq{\mathcal{A}}_{g} be an indecomposable non-compact Shimura subvariety, and let X𝒮X\subset{\mathcal{S}} be a compact subvariety that is Hodge-generic within 𝒮{\mathcal{S}}. Then dimXg1\dim X\leq g-1.

The first main result of the present paper is an immediate consequence.

Proof of A.

This is simply the 𝒮=𝒜g{\mathcal{S}}={\mathcal{A}}_{g} case of 3.4. ∎

Note that we of course get a stronger version of A, for compact subvarieties through any Hodge-generic point of 𝒜g{\mathcal{A}}_{g}, not just through an abstract very general point of 𝒜g{\mathcal{A}}_{g}. Note also that the proof of 3.4 in this special case 𝒮=𝒜g{\mathcal{S}}={\mathcal{A}}_{g} 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 𝒮{\mathcal{S}}, is to first produce a morphism of Shimura varieties Φ:𝒮1×𝒮2𝒜g\Phi:{\mathcal{S}}^{\prime}_{1}\times{\mathcal{S}}^{\prime}_{2}\rightarrow{\mathcal{A}}_{g} with 𝒮1=𝒜1(N){\mathcal{S}}^{\prime}_{1}={\mathcal{A}}_{1}(N) and 𝒮2=𝒜g1(N){\mathcal{S}}^{\prime}_{2}={\mathcal{A}}_{g-1}(N) such that Φ({y}×𝒮2)\Phi(\{y\}\times{\mathcal{S}}^{\prime}_{2}) intersects 𝒮{\mathcal{S}} for almost every y𝒮1y\in{\mathcal{S}}^{\prime}_{1} (where 𝒜g(N){\mathcal{A}}_{g}(N) denotes the moduli of ppav with a full level NN structure). By 3.3(ii), the compactness of XX then forces dimXg1\dim X\leq g-1.

The construction of Φ\Phi uses the following preliminary lemma, in which we construct a modular curve mapping to 𝒮{\mathcal{S}}.

Lemma 3.5.

Let 𝒮{\mathcal{S}} be a non-compact connected Shimura variety corresponding to a connected semisimple group GG over {\mathbb{Q}}. Then there is a homomorphism of {\mathbb{Q}}-groups ϕ:SL2G\phi:\operatorname{SL}_{2}\rightarrow G and a point p𝒮p\in{\mathcal{S}} such that gϕ(g)pg\mapsto\phi(g)\cdot p induces a holomorphic map of modular curves X(N)𝒮X(N^{\prime})\rightarrow{\mathcal{S}}, for some large enough NN^{\prime}.

Proof.

Let 𝒮=Γ\𝒟+{\mathcal{S}}=\Gamma\backslash\mathcal{D}^{+}, for a Shimura datum (G,𝒟+)(G,\mathcal{D}^{+}) and a lattice Γ\Gamma in (Gad)+(G^{\mathrm{ad}}_{\mathbb{R}})^{+}. Then the adjoint action of GG on 𝔤\mathfrak{g} determines a canonical variation of Hodge structure Γ\(𝒟+×𝔤)\Gamma\backslash(\mathcal{D}^{+}\times\mathfrak{g}_{\mathbb{C}}) on 𝒮{\mathcal{S}} (see [Mil11, §5]). Since 𝒮{\mathcal{S}} is non-compact, the group GG_{\mathbb{Q}} contains a unipotent element, and the variation associated to 𝒮{\mathcal{S}} 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 Φ\Phi.

Lemma 3.6.

Let 𝒮𝒜g{\mathcal{S}}\subset{\mathcal{A}}_{g} be a non-compact Shimura variety. Then there is a morphism of Shimura varieties Φ:𝒜1(N)×𝒜g1(N)𝒜g\Phi:{\mathcal{A}}_{1}(N)\times{\mathcal{A}}_{g-1}(N)\rightarrow{\mathcal{A}}_{g} such that for each [E]𝒜1(N)[E]\in{\mathcal{A}}_{1}(N) the intersection 𝒮Φ([E]×𝒜g1(N)){\mathcal{S}}\cap\Phi\left([E]\times{\mathcal{A}}_{g-1}(N)\right) is non-empty.

Proof.

Let ι:GSp2g\iota:G\rightarrow\operatorname{Sp}_{2g} be the homomorphism over {\mathbb{Q}} that induces 𝒮𝒜g{\mathcal{S}}\hookrightarrow{\mathcal{A}}_{g}. Let ϕ:SL2G\phi:\operatorname{SL}_{2}\rightarrow G be the homomorphism, over {\mathbb{Q}}, provided by 3.5, and denote ψ:=ιϕ:SL2Sp2g=Sp(V)\psi:=\iota\circ\phi:\operatorname{SL}_{2}\rightarrow\operatorname{Sp}_{2g}=\operatorname{Sp}(V).

For each p1p\in\mathcal{H}_{1}, denote KpSL2()K_{p}\subset\operatorname{SL}_{2}({\mathbb{R}}) the stabilizer of pp, which is a maximal compact subgroup isomorphic to S1S^{1}. The homomorphism ψ\psi induces a holomorphic map 1S~g\mathcal{H}_{1}\rightarrow\widetilde{S}\rightarrow\mathcal{H}_{g}, which sends p1p\in\mathcal{H}_{1} to the point of g\mathcal{H}_{g} corresponding to ψ|Kp:KpS1Sp2g()\psi|_{K_{p}}:K_{p}\cong S^{1}\rightarrow\operatorname{Sp}_{2g}({\mathbb{R}}).

Denoting V1V_{1} the standard two-dimensional representation of SL2\operatorname{SL}_{2}, recall that all irreducible representations of SL2\operatorname{SL}_{2} are of the form V:=SymV1V_{\ell}:=\operatorname{Sym}^{\ell}V_{1} for some integer 0\ell\geq 0. The group KpK_{p} acts on VV_{\mathbb{C}} via ψ\psi with weights 0, 11, or 1-1. This implies that all irreducible linear summands of VV_{\mathbb{C}} are isomorphic to either V0,V_{0,{\mathbb{C}}} or V1,V_{1,{\mathbb{C}}}, and so VV decomposes as an orthogonal direct sum V=VV′′V=V^{\prime}\oplus V^{\prime\prime} of rational representations, where VV1mV^{\prime}\cong V_{1}\otimes{\mathbb{Q}}^{m}, and V′′V^{\prime\prime} is a trivial representation. Note that m1m\geq 1, because the map from the modular curve to 𝒜g{\mathcal{A}}_{g} provided by 3.5 is not constant. The polarization on VV^{\prime} induces a quadratic form on m{\mathbb{Q}}^{m}, which can be diagonalized over {\mathbb{Q}}. We can thus decompose V=j=1mVjV^{\prime}=\bigoplus_{j=1}^{m}V^{\prime}_{j} with VjV1V^{\prime}_{j}\cong V_{1}, so that moreover the polarization form is orthogonal with respect to this direct sum decomposition.

Let ApA_{p} denote the ppav corresponding to ψ(p)\psi(p). Then VV_{\mathbb{C}} can be identified to H1(Ap;)H^{1}(A_{p};{\mathbb{C}}), and its Hodge filtration is determined by the subspace Fp1V=H1,0(Ap)F^{1}_{p}V_{\mathbb{C}}=H^{1,0}(A_{p}).

Since ψ(Kp)\psi(K_{p}) acts on VV preserving the Hodge filtration, we can decompose Fp1VF_{p}^{1}V_{\mathbb{C}} into irreducible representations of the subgroup ψ(Kp)ψ(SL2())\psi(K_{p})\subset\psi(\operatorname{SL}_{2}({\mathbb{R}})). This gives Fp1V=F1VFp1V′′F_{p}^{1}V_{\mathbb{C}}=F^{1}V^{\prime}_{\mathbb{C}}\oplus F_{p}^{1}V^{\prime\prime}_{\mathbb{C}}, with Fp1V=j=1mFp1Vj,F_{p}^{1}V^{\prime}_{\mathbb{C}}=\bigoplus_{j=1}^{m}F_{p}^{1}V^{\prime}_{j,{\mathbb{C}}}, Fp1Vj,=Vj,Fp1VF^{1}_{p}V^{\prime}_{j,{\mathbb{C}}}=V^{\prime}_{j,{\mathbb{C}}}\cap F_{p}^{1}V_{\mathbb{C}} and Fp1V′′=V′′Fp1VF_{p}^{1}V^{\prime\prime}_{\mathbb{C}}=V^{\prime\prime}_{\mathbb{C}}\cap F_{p}^{1}V_{\mathbb{C}}.

The universal cover V/Fp1VApV_{\mathbb{C}}/F_{p}^{1}V_{\mathbb{C}}\rightarrow A_{p} identifies the abelian variety ApA_{p} with Λ\V/Fp1V\Lambda\backslash V_{\mathbb{C}}/F_{p}^{1}V_{\mathbb{C}}, where Λ\Lambda is a discrete subgroup of VV of rank 2g2g. Denoting Λj:=ΛVj\Lambda^{\prime}_{j}:=\Lambda\cap V^{\prime}_{j}, Λ:=j=1mΛj\Lambda^{\prime}:=\bigoplus_{j=1}^{m}\Lambda^{\prime}_{j} and Λ′′:=ΛV′′\Lambda^{\prime\prime}:=\Lambda\cap V^{\prime\prime}, since V=VV′′V=V^{\prime}\oplus V^{\prime\prime} is a direct sum decomposition over {\mathbb{Q}}, it follows that ΛΛ′′\Lambda^{\prime}\oplus\Lambda^{\prime\prime} has finite index in Λ\Lambda. Hence, the abelian variety ApA_{p} is isogenous to ApAp′′A_{p}^{\prime}\oplus A_{p}^{\prime\prime}, where Ap=Λ\V/F1VA_{p}^{\prime}=\Lambda^{\prime}\backslash V^{\prime}_{\mathbb{C}}/F^{1}V^{\prime}_{\mathbb{C}}, and Ap′′=Λ′′\V′′/F1V′′A_{p}^{\prime\prime}=\Lambda^{\prime\prime}\backslash V^{\prime\prime}_{\mathbb{C}}/F^{1}V^{\prime\prime}_{\mathbb{C}}. More precisely, Ap=(ApAp′′)/PA_{p}=(A_{p}^{\prime}\oplus A_{p}^{\prime\prime})/P, where P=Λ/(ΛΛ′′)P=\Lambda/(\Lambda^{\prime}\oplus\Lambda^{\prime\prime}) is a finite subgroup of ApAp′′A_{p}^{\prime}\oplus A_{p}^{\prime\prime}. In particular, Apj=1mEp,jA_{p}^{\prime}\cong\bigoplus_{j=1}^{m}E_{p,j}, where Ep,jΛj\Vj,/Fp1Vj,E_{p,j}\cong\Lambda^{\prime}_{j}\backslash V^{\prime}_{j,{\mathbb{C}}}/F_{p}^{1}V^{\prime}_{j,{\mathbb{C}}} are elliptic curves. Note that j=1mEp,jAp′′\bigoplus_{j=1}^{m}E_{p,j}\oplus A_{p}^{\prime\prime} is an orthogonal decomposition with respect to the pullback of the polarization of ApA_{p}. Since every polarized abelian variety is isogenous to a ppav (via an isogeny respecting the polarization), there exist isogenies EpEp,1E_{p}\rightarrow E_{p,1} and Bpj=2mEp,jAp′′B_{p}\rightarrow\bigoplus_{j=2}^{m}E_{p,j}\oplus A_{p}^{\prime\prime} such that the pullback polarization on EpE_{p} and on BpB_{p} can be written as kωEpk\cdot\omega_{E_{p}} and kωBpk\cdot\omega_{B_{p}} for some integer kk, where ωEp\omega_{E_{p}} and ωBp\omega_{B_{p}} are principal polarizations on Ep𝒜1E_{p}\in{\mathcal{A}}_{1} and on Bp𝒜g1B_{p}\in{\mathcal{A}}_{g-1}. Moreover, this can be done globally in the family over 1\mathcal{H}_{1}, with a constant isogeny, so that EpE_{p} and BpB_{p} are the fibers over pp of two families of abelian varieties EE and BB over 1\mathcal{H}_{1} endowed with principal polarizations ωE\omega_{E} and ωB\omega_{B}.

Thus, we can write A(EB)/P1A\cong(E\oplus B)/P_{1}, with P1P_{1} being a family of torsion subgroups of the family of abelian varieties EBE\oplus B over 1\mathcal{H}_{1}, and then the polarization k(ωEωB)k\cdot(\omega_{E}\oplus\omega_{B}) descends to a principal polarization on AA.

Denote NN the order of the group P1P_{1}. The action of P1P_{1} on EBE\oplus B extends to a flat action of P1P_{1} on the universal family over the level cover 𝒜1(N)×𝒜g1(N){\mathcal{A}}_{1}(N)\times{\mathcal{A}}_{g-1}(N), and taking this quotient by P1P_{1} gives the morphism Φ:𝒜1(N)×𝒜g1(N)𝒜g\Phi:{\mathcal{A}}_{1}(N)\times{\mathcal{A}}_{g-1}(N)\to{\mathcal{A}}_{g}, so that for any p1p\in\mathcal{H}_{1} we have Ap𝒮Φ([Ep]×𝒜g1(N))A_{p}\in{\mathcal{S}}\cap\Phi([E_{p}]\times{\mathcal{A}}_{g-1}(N)), and in particular 𝒮Φ([Ep]×𝒜g1(N)){\mathcal{S}}\cap\Phi([E_{p}]\times{\mathcal{A}}_{g-1}(N))\neq\emptyset. ∎

We are now ready to prove the theorem, implementing the original idea.

Proof of 3.4.

Assume, for contradiction, that X𝒮X\subsetneq{\mathcal{S}} is a compact Hodge-generic irreducible subvariety of 𝒮{\mathcal{S}}, of dimension at least gg. Let Φ:𝒜1(N)×𝒜g1(N)𝒜g\Phi:{\mathcal{A}}_{1}(N)\times{\mathcal{A}}_{g-1}(N)\rightarrow{\mathcal{A}}_{g} be the morphism constructed in 3.6, so that for any [E]𝒜1(N)[E]\in{\mathcal{A}}_{1}(N) the intersection 𝒮Φ([E]×𝒜g1(N)){\mathcal{S}}\cap\Phi\left([E]\times{\mathcal{A}}_{g-1}(N)\right) is non-empty. Since dimX+dim𝒜g1(N)dim𝒜g\dim X+\dim{\mathcal{A}}_{g-1}(N)\geq\dim{\mathcal{A}}_{g} and since 𝒜1(N){\mathcal{A}}_{1}(N) is non-compact, by 3.3(ii) it follows that XX must also be non-compact, contradicting our hypotheses. ∎

The combination of 2.9 and 3.4 leads to the following exact expression.

Proposition 3.7.

Let F(g)F(g) denote the maximal dimension of a compact Shimura subvariety of 𝒜g{\mathcal{A}}_{g}. Then the maximal dimension of a compact subvariety of 𝒜g{\mathcal{A}}_{g} is

mdimc(𝒜g)=max(F(g),max0g<g(gg1+F(g))).\operatorname{mdim_{c}}({\mathcal{A}}_{g})=\max\left(F(g),\max_{0\leq g^{\prime}<g}\left(g-g^{\prime}-1+F(g^{\prime})\right)\right).
Proof.

Let X𝒜gX\subsetneq{\mathcal{A}}_{g} be a compact subvariety and let 𝒮𝒜g{\mathcal{S}}\subseteq{\mathcal{A}}_{g} be the smallest Shimura subvariety containing XX. Since mdimc(𝒜g)mdimc,gen(𝒜g)=g1\operatorname{mdim_{c}}({\mathcal{A}}_{g})\geq\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g})=g-1, it is enough to deal with the case dimXg\dim X\geq g, in which case from 3.4 it follows that 𝒮𝒜g{\mathcal{S}}\subsetneq{\mathcal{A}}_{g}.

Decompose 𝒮=𝒮1××𝒮n×𝒮{\mathcal{S}}={\mathcal{S}}_{1}\times\dots\times{\mathcal{S}}_{n}\times{\mathcal{S}}^{\prime} with all 𝒮i{\mathcal{S}}_{i} indecomposable non-compact Shimura varieties, and with 𝒮{\mathcal{S}}^{\prime} a compact Shimura variety. Then by applying 2.9 repeatedly, it follows that the embedding 𝒮𝒜g{\mathcal{S}}\hookrightarrow{\mathcal{A}}_{g} must factor, up to isogeny, via

𝒮1×𝒮n×𝒮𝒜g1××𝒜gn×𝒜g𝒜g{\mathcal{S}}_{1}\times\cdots{\mathcal{S}}_{n}\times{\mathcal{S}}^{\prime}\rightarrow{\mathcal{A}}_{g_{1}}\times\dots\times{\mathcal{A}}_{g_{n}}\times{\mathcal{A}}_{g^{\prime}}\hookrightarrow{\mathcal{A}}_{g}

for suitable 𝒮i𝒜gi{\mathcal{S}}_{i}\rightarrow{\mathcal{A}}_{g_{i}} and 𝒮𝒜g{\mathcal{S}}^{\prime}\rightarrow{\mathcal{A}}_{g^{\prime}}. In particular (the isogeny image of) XX embeds into the product 𝒜g1××𝒜gn×𝒮{\mathcal{A}}_{g_{1}}\times\dots\times{\mathcal{A}}_{g_{n}}\times{\mathcal{S}}^{\prime}. The projection of XX to each non-compact factor 𝒮i𝒜gi{\mathcal{S}}_{i}\subset{\mathcal{A}}_{g_{i}} is of dimension at most gi1g_{i}-1, by 3.4, and thus the projection of XX to 𝒜g1××𝒜gn{\mathcal{A}}_{g_{1}}\times\dots\times{\mathcal{A}}_{g_{n}} has dimension at most i=1n(gi1)=(i=1ngi)n\sum_{i=1}^{n}(g_{i}-1)=(\sum_{i=1}^{n}g_{i})-n. Thus dimX(i=1ngi)n+dim𝒮\dim X\leq(\sum_{i=1}^{n}g_{i})-n+\dim{\mathcal{S}}^{\prime}.

On the other hand, the maximal dimension of a compact Hodge-generic subvariety of 𝒜gg{\mathcal{A}}_{g-g^{\prime}} is gg1g-g^{\prime}-1, and can be realized by some complete intersection YY. Hence, the compact subvariety Y×𝒮Y\times{\mathcal{S}}^{\prime} of 𝒜gg×𝒜g𝒜g{\mathcal{A}}_{g-g^{\prime}}\times{\mathcal{A}}_{g^{\prime}}\subset{\mathcal{A}}_{g} has dimension gg1+dim𝒮dimXg-g^{\prime}-1+\dim{\mathcal{S}}^{\prime}\geq\dim X. The conclusion follows, as dim𝒮F(g)\dim{\mathcal{S}}^{\prime}\leq F(g^{\prime}). ∎

4. Compact Shimura subvarieties of 𝒜g{\mathcal{A}}_{g} from decoupled representations

In this section we begin investigating the maximal dimension of compact Shimura subvarieties 𝒮𝒜g{\mathcal{S}}\subsetneq{\mathcal{A}}_{g}, 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 G/G/{\mathbb{Q}} be a connected semisimple group corresponding to a connected Shimura variety 𝒮{\mathcal{S}}, and let VV be a symplectic representation over {\mathbb{Q}} inducing a map of 𝒮{\mathcal{S}} to 𝒜g{\mathcal{A}}_{g}.

We say that (d,g)(d,g) is the pair associated to (G,V)(G,V) (or just to 𝒮{\mathcal{S}}) if d=dim𝒮d=\dim{\mathcal{S}} and 2g=dimV2g=\dim V. We recall that there exist (Hodge-generic) dd-dimensional compact subvarieties of 𝒜g{\mathcal{A}}_{g} for every d<gd<g. Hence we call a pair (d,g)(d,g) negligible if d<gd<g and eligible if dgd\geq g.

Moreover, we introduce a partial ordering on ×{\mathbb{N}}\times{\mathbb{N}} by declaring

(d,g)(d,g)ddandgg,(d,g)\preceq(d^{\prime},g^{\prime})\iff d\leq d^{\prime}\ {\rm and}\ g\geq g^{\prime},

If (d,g)(d,g)(d,g)\preceq(d^{\prime},g), we will say that (d,g)(d,g) is dominated by (d,g)(d^{\prime},g).

Let VαV_{\alpha} be an irreducible linear summand of VV_{\mathbb{C}}. Then VαV_{\alpha} is acted upon nontrivially by lαl_{\alpha} almost-simple factors of GG_{\mathbb{R}}, at most one of which is non-compact by 2.8. Since the case lα=1l_{\alpha}=1 will be the most relevant, we introduce the following.

Definition 4.1.

A complex symplectic representation of GG is decoupled if, for every irreducible linear summand, there exists an almost-simple factor of GG_{\mathbb{R}} such that the action of GG_{\mathbb{R}} on such summand agrees with its restriction to this almost-simple factor. A rational symplectic representation of GG is decoupled if its complexification is. ∎

In this section we study compact Shimura subvarieties 𝒮𝒜g{\mathcal{S}}\rightarrow{\mathcal{A}}_{g} associated to decoupled representations VV of GG. In view of 2.5, we will thus assume that

  • GG is anisotropic.

4.2. Sp\operatorname{Sp}-irreducible decoupled {\mathbb{Q}}-representations

Later in this section we will classify dominating pairs of the following type.

Definition 4.2.

Let VV be a decoupled symplectic representation of GG over {\mathbb{Q}}. The pair (d,g)(d,g) corresponding to (G,V)(G,V) is d(ecouplably)-achieved if GG is anisotropic and VV is Sp\operatorname{Sp}-irreducible. ∎

Geometrically, the Sp\operatorname{Sp}-irreducibility of VV is equivalent to the fact that 𝒮{\mathcal{S}} is not contained, up to isogeny, in the decomposable locus of 𝒜g{\mathcal{A}}_{g}.

A relevant property that follows from the Sp\operatorname{Sp}-irreducibility of VV is the following.

Lemma 4.3.

Assume that VV is Sp\operatorname{Sp}-irreducible. Then all the irreducible linear summands of VV_{\mathbb{C}} 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 W0VW_{0}\subset V_{\mathbb{C}} be an irreducible linear summand, and let WVW\subset V denote the direct sum of all subspaces of VV_{\mathbb{C}} that are isomorphic to a Galois conjugate of W0W_{0} or W0W_{0}^{\vee}. Then WW is Galois stable, and thus is defined over {\mathbb{Q}}. Moreover, VV decomposes as WWW\oplus W^{\prime} with WW^{\prime} not having any of the irreducible linear summands that appear in WW, and thus WW is symplectic. Since VV is assumed Sp\operatorname{Sp}-irreducible, we must have W=VW=V. This proves the claim. ∎

Note that, if the representation VV is Sp\operatorname{Sp}-irreducible and decoupled, then the group GG must be almost {\mathbb{Q}}-simple.

4.3. Almost {\mathbb{Q}}-simple groups and decoupled symplectic representations

Let GG be an almost {\mathbb{Q}}-simple group. It follows that G=H𝔽/G=H_{{\mathbb{F}}/{\mathbb{Q}}} for some totally real field 𝔽{\mathbb{F}}, and that i=1kGiG\prod_{i=1}^{k}G_{i}\rightarrow G_{\mathbb{R}} is an isogeny, where Gi=HσiG_{i}=H_{\sigma_{i}} for {σ1,,σk}=Hom(𝔽,)\{\sigma_{1},\dots,\sigma_{k}\}=\operatorname{Hom}({\mathbb{F}},{\mathbb{R}}). So all GiG_{i} are 𝔽{\mathbb{F}}-forms of the same complex group HH_{{\mathbb{C}}}.

Let VV be a symplectic representation of GG over {\mathbb{Q}}, which is decoupled but not necessarily Sp\operatorname{Sp}-irreducible. Then VV_{\mathbb{C}} decomposes as V=i=1kViV_{\mathbb{C}}=\oplus_{i=1}^{k}V_{i}, where each ViV_{i} is a complex linear representation of GiG_{i}. If GiG_{i} is non-compact, then every Sp\operatorname{Sp}-irreducible sub-representation of ViV_{i} corresponds to a “symplectic node”, see [Mil11, §10].

Denote by UiU_{i} a GiG_{i}-invariant irreducible linear subspace of ViV_{i}.

Remark 4.4.

The GiG_{i}-representation ViV_{i} is symplectic. Hence, if UiU_{i} is not a symplectic subspace, then ViV_{i} has a GiG_{i}-invariant subspace isomorphic to UiUiU_{i}\oplus U_{i}^{\vee}.

In Table 2 we list all possible types of complex groups HH_{{\mathbb{C}}}, real groups GiG_{i}, domains 𝒟i\mathcal{D}_{i} associated to the non-compact GiG_{i}, and GiG_{i}-irreducible linear summands UiU_{i} that can occur in the symplectic representation VV_{\mathbb{C}} (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 HH_{\mathbb{C}} real group GiG_{i} space dim𝒟i\dim\mathcal{D}_{i} dimUi\dim U_{i} /Orthogonal
SL2()\operatorname{SL}_{2}({\mathbb{C}}) SL2()\operatorname{SL}_{2}({\mathbb{R}}) 𝐀1{\bf{A}}_{1} 11 22 Symp
SLn+1()\operatorname{SL}_{n+1}({\mathbb{C}}), n2n\geq 2 SUn,1()\operatorname{SU}_{n,1}({\mathbb{R}}) 𝐀n(1){\bf{A}}_{n}(1) nn (n+1h)1hn\begin{array}[]{c}\\ \displaystyle\binom{n+1}{h}\\ \\ 1\leq h\leq n\end{array} {hn+12NSDh=n+12 evenOrth h=n+12 oddSymp\begin{cases}h\neq\frac{n+1}{2}&\textrm{NSD}\\ h=\frac{n+1}{2}\textrm{ even}&\textrm{Orth }\\ h=\frac{n+1}{2}\textrm{ odd}&\textrm{Symp }\end{cases}
SLn+1()\operatorname{SL}_{n+1}({\mathbb{C}}), n2n\geq 2 SUh,n+1h()2hn+12\begin{array}[]{c}\\ \operatorname{SU}_{h,n+1-h}({\mathbb{R}})\\ \\ 2\leq h\leq\frac{n+1}{2}\end{array} 𝐀n(h){\bf{A}}_{n}(h) h(n+1h)h(n+1-h) n+1n+1 NSD
SO2n+1()\operatorname{SO}_{2n+1}({\mathbb{C}}), n3n\geq 3 SO2n1,2()\operatorname{SO}_{2n-1,2}({\mathbb{R}}) 𝐁n(1){\bf{B}}_{n}(1) 2n12n-1 2n2^{n} {n0,3(4)Orthn1,2(4)Symp\begin{cases}n\equiv 0,3(4)&\textrm{Orth}\\ n\equiv 1,2(4)&\textrm{Symp}\end{cases}
Sp2n()\operatorname{Sp}_{2n}({\mathbb{C}}) Sp2n()\operatorname{Sp}_{2n}({\mathbb{R}}) 𝐂n{\bf{C}}_{n} n(n+1)2\frac{n(n+1)}{2} 2n2n Symp
SO2n()\operatorname{SO}_{2n}({\mathbb{C}}), n4n\geq 4 SO2n2,2()\operatorname{SO}_{2n-2,2}({\mathbb{R}}) 𝐃n(1){\bf{D}}_{n}(1) 2n22n-2 2n12^{n-1} {n2(4)Sympn0(4)Orthn1,3(4)NSD\begin{cases}n\equiv 2(4)&\textrm{Symp}\\ n\equiv 0(4)&\textrm{Orth}\\ n\equiv 1,3(4)&\textrm{NSD}\end{cases}
SO2n()\operatorname{SO}_{2n}({\mathbb{C}}), n5n\geq 5 SO2n()\operatorname{SO}^{*}_{2n}({\mathbb{R}}) 𝐃n(n1)or𝐃n(n)\begin{array}[]{c}{\displaystyle{\bf{D}}_{n}(n-1)}\\ {\displaystyle\text{or}\ {\bf{D}}_{n}(n)}\end{array} n2n2\frac{n^{2}-n}{2} 2n2n Orth

Table 2. All irreducible Hermitian domains that map nontrivially to some Siegel upper half-space.

For Table 2, we also recall that

SO2n():=SO2n()SUn,n()\operatorname{SO}^{*}_{2n}({\mathbb{R}}):=\operatorname{SO}_{2n}({\mathbb{C}})\cap\operatorname{SU}_{n,n}({\mathbb{R}})

can be identified with the group of automorphisms of the nn-dimensional vector space n{\mathbb{H}}^{n}_{\mathbb{R}} over the quaternion skew algebra {\mathbb{H}}_{\mathbb{R}} with center {\mathbb{R}} that preserve the standard skew-Hermitian pairing on {\mathbb{H}}_{\mathbb{R}}. Moreover, in the 𝐃4{\bf{D}}_{4} case we have SO8()SO6,2()\operatorname{SO}^{*}_{8}({\mathbb{R}})\cong\operatorname{SO}_{6,2}({\mathbb{R}}).

In the dimension estimates contained in Section 4.4, we will often make use of the following result.

Lemma 4.5.

Let G=H𝔽/G=H_{{\mathbb{F}}/{\mathbb{Q}}} be a rational anisotropic group, where 𝔽{\mathbb{F}} is a totally real number field, and HH_{{\mathbb{C}}} is one of the entries in Table 2. If GG_{\mathbb{R}} has no compact factors, then each HH is isomorphic to either of the following

  • (1)

    H𝔽1={γ𝔽|f(γ)=1}\mathrm{H}^{1}_{{\mathbb{F}}}=\{\gamma\in{\mathbb{H}}_{\mathbb{F}}\,|\,f(\gamma)=1\} for some indefinite Hermitian form ff on 𝔽{\mathbb{H}}_{\mathbb{F}},

  • (2)

    SO6,2(𝔽,f)\operatorname{SO}_{6,2}({\mathbb{F}},f) for some symmetric bilinear form ff on 𝔽8{\mathbb{F}}^{8} of signature (6,2)(6,2).

Proof.

The cases we are interested in of HH of classical type are summarized in [PR94, page 92]. In the 𝐃4{\bf{D}}_{4} case, the group HH must be of type SO6,2\operatorname{SO}_{6,2}, while otherwise the group HH is isomorphic to either:

  • (a)

    SLq(𝔻)\operatorname{SL}_{q}({\mathbb{D}}) for q2q\geq 2, or

  • (b)

    SUq(𝔻,f)\operatorname{SU}_{q}({\mathbb{D}},f) for q1q\geq 1,

where 𝔻{\mathbb{D}} is either 𝔽{\mathbb{F}} (and ff is symmetric or alternating), or an imaginary extension 𝔼{\mathbb{E}} of 𝔽{\mathbb{F}} (and ff is a Hermitian form), or a quaternion skew field {\mathbb{H}} with center 𝔽{\mathbb{F}} (and ff is a Hermitian or a skew-Hermitian form).

For any q2q\geq 2 the group SLq(𝔻)\operatorname{SL}_{q}({\mathbb{D}}) contains nontrivial unipotent elements. If 𝔻=𝔽{\mathbb{D}}={\mathbb{F}} and ff is alternating, then SUq(𝔻,f)=Sp2q(𝔽)\operatorname{SU}_{q}({\mathbb{D}},f)=\operatorname{Sp}_{2q}({\mathbb{F}}), in which case it also contains unipotent elements.

In the remaining cases with ff symmetric/Hermitian/skew-Hermitian, it is known that SUq(𝔻,f)\operatorname{SU}_{q}({\mathbb{D}},f) contains nontrivial unipotent elements if and only if 𝔻q{\mathbb{D}}^{q} has nontrivial isotropic vectors, which happens if and only if ff nontrivially represents zero.

If ff is symmetric and 𝔻=𝔽{\mathbb{D}}={\mathbb{F}}, then [PR94, page 342, Claim 6.1] implies that q4q\leq 4. In this case, SOq(𝔽,f)\operatorname{SO}_{q}({\mathbb{F}},f) is abelian (for q=2q=2), or of SL2\operatorname{SL}_{2} type (for q=3)q=3) or of SL2×SL2\operatorname{SL}_{2}\times\operatorname{SL}_{2} type (for q=4q=4). 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 ff is Hermitian, then [PR94, page 343, Claim 6.2] implies that q2q\leq 2 for 𝔻=𝔼{\mathbb{D}}={\mathbb{E}}, and q1q\leq 1 for 𝔻={\mathbb{D}}={\mathbb{H}}. Hence SUq(𝔻,f)\operatorname{SU}_{q}({\mathbb{D}},f) is either of type SU1,1\operatorname{SU}_{1,1} or isomorphic to H𝔽1\mathrm{H}^{1}_{\mathbb{F}}. Since SU1,1SL2\operatorname{SU}_{1,1}\cong\operatorname{SL}_{2}, such case can be ruled out.

If ff is skew-Hermitian and 𝔻={\mathbb{D}}={\mathbb{H}}, then [PR94, page 343, Claim 6.3] implies that q3q\leq 3. The case q=2q=2 can be ruled out since SU4(,f)\operatorname{SU}_{4}({\mathbb{H}},f) is of type SU2,2\operatorname{SU}_{2,2}, and the case q=3q=3 corresponds to SU6(,f)\operatorname{SU}_{6}({\mathbb{H}},f) of type SO6,2\operatorname{SO}_{6,2}.

Hence, HH can only be of type H1\mathrm{H}^{1} or SO6,2\operatorname{SO}_{6,2}, and for H1\mathrm{H}^{1} to be non-compact, the Hermitian form ff 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 GG be a {\mathbb{Q}}-almost-simple, anisotropic algebraic group and let VV_{\mathbb{C}} be a decoupled complex symplectic representation of GG 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)

    GG is not of 𝐀1{\bf{A}}_{1} type;

  • (b)

    VV_{\mathbb{C}} is the complexification of a rational representation VV;

  • (c)

    each ViV_{i} consists of at least 22 irreducible linear summands;

then the pair (d,g)(d,g) is either negligible or dominated by one of the dd-achievable pairs given in Table 3 below.

Type of Hermitian Dominating
complex group symmetric d-achievable
group HH_{\mathbb{C}} space pairs (d,g)(d,g)
SL2()\operatorname{SL}_{2}({\mathbb{C}}) 𝐀1(1){\bf{A}}_{1}(1) Negligible
SLn+1()\operatorname{SL}_{n+1}({\mathbb{C}}), n2n\geq 2 𝐀n(1)[SUn,1]{\bf{A}}_{n}(1)\ [\operatorname{SU}_{n,1}] Negligible
SLn+1()\operatorname{SL}_{n+1}({\mathbb{C}}), n2n\geq 2 𝐀n(h)[SUh,n+1h]{\bf{A}}_{n}(h)\ [\operatorname{SU}_{h,n+1-h}], h=n+12h=\lfloor\frac{n+1}{2}\rfloor ((k1)n+12n+12,(n+1)k)\large((k-1)\lceil\frac{n+1}{2}\rceil\cdot\lfloor\frac{n+1}{2}\rfloor,(n+1)k\large)
SO2n+1()\operatorname{SO}_{2n+1}({\mathbb{C}}), n3n\geq 3 𝐁n(1)[SO2n1,2]{\bf{B}}_{n}(1)\ [\operatorname{SO}_{2n-1,2}] Negligible
Sp2n()\operatorname{Sp}_{2n}({\mathbb{C}}) 𝐂n[Sp2n]{\bf{C}}_{n}\ [\operatorname{Sp}_{2n}] ((k1)n(n+1)2,2nk)((k-1)\frac{n(n+1)}{2},2nk)
SO8()\operatorname{SO}_{8}({\mathbb{C}}) 𝐃4[SO6,2]{\bf{D}}_{4}\ [\operatorname{SO}_{6,2}] Negligible
SO2n()\operatorname{SO}_{2n}({\mathbb{C}}), n5n\geq 5 𝐃n(1)[SO2n2,2]{\bf{D}}_{n}(1)\ [\operatorname{SO}_{2n-2,2}] Negligible
SO2n()\operatorname{SO}_{2n}({\mathbb{C}}), n5n\geq 5 𝐃n(n1)or𝐃n(n)[SO2n]\begin{array}[]{c}{\displaystyle{\bf{D}}_{n}(n-1)}\\ {\displaystyle\text{or}\ {\bf{D}}_{n}(n)}\end{array}\ [\operatorname{SO}^{*}_{2n}] ((k1)n(n1)2,2nk)((k-1)\frac{n(n-1)}{2},2nk)

Table 3. List of d-achievable pairs

In Section 4.5 below we will use the following consequence of the above result.

Corollary 4.7.

Let GG be a {\mathbb{Q}}-almost-simple, anisotropic algebraic group, and let VV be an Sp\operatorname{Sp}-irreducible decoupled {\mathbb{Q}}-representation of GG. Then the associated pair (d,g)(d,g) 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 VV_{\mathbb{C}} satisfies hypothesis (b) of that lemma. ∎

Proof of 4.6.

Note that all the almost-simple factors of GG_{\mathbb{R}} are Galois conjugate to each other. Since all the irreducible linear summands of VV_{\mathbb{C}} are Galois conjugate to each other or to each other’s dual, then there exists an integer m1m\geq 1 independent of ii such that every ViV_{i} is the direct sum of mm 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 SL2\operatorname{SL}_{2}

In this case the groups GiG_{i} are isomorphic either to H1SU2()\mathrm{H}^{1}_{\mathbb{R}}\cong\operatorname{SU}_{2}({\mathbb{R}}) or to SL2()\operatorname{SL}_{2}({\mathbb{R}}), and by 4.5 there may be anisotropic forms which have no compact factors.

Let GG be of type 𝐀1{\bf{A}}_{1}. Since the action of S1S^{1} on each irreducible linear summand of VV_{\mathbb{C}} has weight 0, 11 or 1-1, each UiU_{i} must be isomorphic to the standard representation of GiG_{i}.

Now, GG is isomorphic either to Res𝔽/H𝔽1\operatorname{Res}_{{\mathbb{F}}/{\mathbb{Q}}}\mathrm{H}^{1}_{\mathbb{F}} or to Res𝔽/SL2(𝔽)\operatorname{Res}_{{\mathbb{F}}/{\mathbb{Q}}}\operatorname{SL}_{2}({\mathbb{F}}), and each GiG_{i} could be isomorphic either to H1\mathrm{H}^{1}_{\mathbb{R}} or to SL2()\operatorname{SL}_{2}({\mathbb{R}}).

Note that GRes𝔽/SL2(𝔽)G\neq\operatorname{Res}_{{\mathbb{F}}/{\mathbb{Q}}}\operatorname{SL}_{2}({\mathbb{F}}), since such group is isotropic. Hence G=Res𝔽/H𝔽1G=\operatorname{Res}_{{\mathbb{F}}/{\mathbb{Q}}}\mathrm{H}^{1}_{\mathbb{F}}. Then one factor of G𝔽G_{{\mathbb{F}}} is isomorphic to H𝔽1\mathrm{H}^{1}_{\mathbb{F}}.

Assume that VV_{\mathbb{C}} is defined over {\mathbb{Q}}. Then G𝔽H𝔽1G_{{\mathbb{F}}}\cong\mathrm{H}^{1}_{{\mathbb{F}}} acts nontrivially on a 2m2m-dimensional 𝔽{\mathbb{F}}-vector space. Since H𝔽1\mathrm{H}^{1}_{\mathbb{F}} does not have nontrivial 2-dimensional representations over 𝔽{\mathbb{F}}, it follows that m2m\geq 2. Thus we have dkd\leq k and g=mk2kg=mk\geq 2k. But in fact all such pairs (d,g)=(k,mk)(d,g)=(k,mk) with m2m\geq 2 are negligible, which gives the first row of Table 3.

Note that the only such pair (d,g)(d,g) satisfying d=g1d=g-1 is (1,2)(1,2). Indeed, if we endow {\mathbb{H}}_{\mathbb{Q}} with an indefinite anisotropic norm and we let G:=H1G:=\mathrm{H}^{1}_{\mathbb{Q}} act on V:=V:={\mathbb{H}}_{\mathbb{Q}} by left-multiplication, the couple (G,V)(G,V) determines a compact Shimura curve in 𝒜2{\mathcal{A}}_{2}.

4.4.2. Special case SO8\operatorname{SO}_{8}

In this case the groups GiG_{i} are isomorphic either to SO8()\operatorname{SO}_{8}({\mathbb{R}}) or to SO6,2()\operatorname{SO}_{6,2}({\mathbb{R}}), and by 4.5 there may be anisotropic forms which have no compact factors.

By Table 2 the representation UiU_{i} is not symplectic (it is in fact orthogonal) and 88-dimensional. Hence m2m\geq 2, and so dimVi16\dim V_{i}\geq 16. Thus the d-achievable pairs are dominated by (6k,8k)(6k,8k), which are negligible.

4.4.3. Case SLn+1\operatorname{SL}_{n+1} for n2n\geq 2

The factors GiG_{i} are of type SUh,n+1h()\operatorname{SU}_{h,n+1-h}({\mathbb{R}}). Since we have already treated the case n=3n=3 with all GiG_{i} non-compact above, by 4.5 we can assume that at least one factor must be compact (namely, it must be SUn()\operatorname{SU}_{n}({\mathbb{R}})).

Case (i): all factors GiG_{i} are of type SUn+1()\operatorname{SU}_{n+1}({\mathbb{R}}) or SUn,1()\operatorname{SU}_{n,1}({\mathbb{R}}). By the classification, UiU_{i} has dimension (n+1h)\binom{n+1}{h} for some hh.

If hn+12h\neq\frac{n+1}{2}, or if h=n+12h=\frac{n+1}{2} and is even, then UiU_{i} is not symplectic and so m2m\geq 2 by 4.4. It follows that dimVi2dimUi=2(n+1h)\dim V_{i}\geq 2\dim U_{i}=2\binom{n+1}{h}. The corresponding pair (d,g)(d,g) is then dominated by (kn,k(n+1h))(kn,\,k\binom{n+1}{h}), which is negligible.

If h=n+12h=\frac{n+1}{2} and is odd, then n5n\geq 5, and the corresponding pair (d,g)(d,g) is dominated by

((k1)n,k2(n+1n+12)).((k-1)n,\,\frac{k}{2}\binom{n+1}{\frac{n+1}{2}}).

Note that 12(n+1n+12)>n\frac{1}{2}\binom{n+1}{\frac{n+1}{2}}>n, so this is also negligible.

Case (ii): at least one factor GiG_{i} is of type SUh,n+1h()\operatorname{SU}_{h,n+1-h}({\mathbb{R}}) with 2hn+122\leq h\leq\frac{n+1}{2}. By Table 2 the irreducible linear summands UiU_{i} of ViV_{i} are all the standard representation or its dual: in particular, UiU_{i} has dimension n+1n+1 and is not self-dual. Thus m2m\geq 2 by 4.4 and so dimVi2(n+1)\dim V_{i}\geq 2(n+1). It follows that dimV2(n+1)k\dim V\geq 2(n+1)k, and so g(n+1)kg\geq(n+1)k.

Likewise, there are at most k1k-1 non-compact factors, with associated symmetric spaces each of dimension h(n+1h)n+12n+12h(n+1-h)\leq\lceil\frac{n+1}{2}\rceil\cdot\lfloor\frac{n+1}{2}\rfloor. Thus dim𝒟+(k1)n+12n+12\dim\mathcal{D}^{+}\leq(k-1)\lceil\frac{n+1}{2}\rceil\cdot\lfloor\frac{n+1}{2}\rfloor. Hence, the d-achievable pairs are dominated by ((k1)n+12n+12,(n+1)k)\large((k-1)\lceil\frac{n+1}{2}\rceil\cdot\lfloor\frac{n+1}{2}\rfloor,(n+1)k\large).

Existence. Let 𝔼/𝔽{\mathbb{E}}/{\mathbb{F}} be a totally imaginary quadratic extension, which comes with a natural involution, and let H:=SUn+1(𝔼,f)H:=\operatorname{SU}_{n+1}({\mathbb{E}},f), where ff is a Hermitian form on 𝔼n+1{\mathbb{E}}^{n+1}. By [BH78, Theorem B], the form ff can be chosen to have signature (n+12,n+12)(\lceil\frac{n+1}{2}\rceil,\lfloor\frac{n+1}{2}\rfloor) at all but one place of 𝔽{\mathbb{F}}, and definite signature at the remaining place σ\sigma. At the compact place σ\sigma, a homomorphism uu can be defined using the action of S1S^{1} on σ~(𝔼)σ(𝔽)\tilde{\sigma}({\mathbb{E}})\otimes_{\sigma({\mathbb{F}})}{\mathbb{R}}\cong{\mathbb{C}}, where σ~:𝔼\tilde{\sigma}:{\mathbb{E}}\hookrightarrow{\mathbb{C}} is an extension of σ\sigma, as in [Mil11, Theorem 10.14].

4.4.4. Case SO2n+1\operatorname{SO}_{2n+1} for n3n\geq 3

The factors GiG_{i} are of type SO2n1,2()\operatorname{SO}_{2n-1,2}({\mathbb{R}}) or SO2n+1()\operatorname{SO}_{2n+1}({\mathbb{R}}). By 4.5, at least one factor must be compact, namely of type SO2n+1()\operatorname{SO}_{2n+1}({\mathbb{R}}).

If n0,3(4)n\equiv 0,3(4), then UiU_{i} is not self-dual, and so m2m\geq 2 by 4.4. It follows that ViV_{i} must have dimension at least 22n2\cdot 2^{n}, which gives a d-achievable pair dominated by ((2n1)(k1),2nk)((2n-1)(k-1),2^{n}k). Since 2n1<2n2n-1<2^{n}, this is negligible.

If n1,2(4)n\equiv 1,2(4), then UiU_{i} is self-dual and so dimVi2n\dim V_{i}\geq 2^{n}. This gives a d-achievable pair dominated by ((2n1)(k1),2n1k)((2n-1)(k-1),2^{n-1}k), which is again negligible since n5n\geq 5.

4.4.5. Case Sp2n\operatorname{Sp}_{2n} for n2n\geq 2

The factors GiG_{i} are of type Sp2n()\operatorname{Sp}_{2n}({\mathbb{R}}) or Sp(n)=Sp2n()SU2n\operatorname{Sp}^{*}(n)=\operatorname{Sp}_{2n}({\mathbb{C}})\cap\operatorname{SU}_{2n}. By 4.5, at least one factor GjG_{j} is compact, namely of type Sp(n)\operatorname{Sp}^{*}(n).

Now, each UjU_{j} is isomorphic to the fundamental representation, which is defined over {\mathbb{R}}. If VjV_{j} were isomorphic to UjU_{j} (i.e. m=1m=1), then VjV_{j} would be the complexification of a real representation, and the compact factor GjG_{j} would inject into Sp2n()\operatorname{Sp}_{2n}({\mathbb{R}}). This contradiction shows that m2m\geq 2 (and, in fact, one could show that mm must be even).

Assume now that m2m\geq 2. Thus dimV2k2n\dim V\geq 2k\cdot 2n and so the corresponding d-achievable pair is dominated by ((k1)n(n+1)2,2nk)((k-1)\frac{n(n+1)}{2},2nk).

Existence. Let 𝔽{\mathbb{H}}_{\mathbb{F}} be a quaternion algebra over 𝔽{\mathbb{F}} and let H𝔽:=SUn(𝔽,f)H_{\mathbb{F}}:=\operatorname{SU}_{n}({\mathbb{H}}_{\mathbb{F}},f) for some Hermitian form ff on 𝔽n{\mathbb{H}}^{n}_{\mathbb{F}}. By [BH78, Theorem B], the form ff can be chosen to be definite at exactly one place of 𝔽{\mathbb{F}}, and (non-degenerate) indefinite at all the other places. Then the pair ((k1)n(n+1)2,2nk)((k-1)\frac{n(n+1)}{2},2nk) is d-achieved by the group H𝔽H_{\mathbb{F}}.

4.4.6. Case SO2n2,2\operatorname{SO}_{2n-2,2} for n4n\geq 4

The factors GiG_{i} are of type SO2n2,2()\operatorname{SO}_{2n-2,2}({\mathbb{R}}) or SO2n()\operatorname{SO}_{2n}({\mathbb{R}}), and there is at least one compact factor (of type SO2n()\operatorname{SO}_{2n}({\mathbb{R}})) by 4.5.

Suppose n=4n=4. Then the representation UiU_{i} is not self-dual, and so m2m\geq 2 by 4.4. It follows that dimVi=8m16\dim V_{i}=8m\geq 16, and so dimV=8mk16k\dim V=8mk\geq 16k, whereas the domain has dimension at most 6(k1)6(k-1). Hence the d-achievable pairs (d,g)(d,g) satisfy d6(k1)d\leq 6(k-1), while g16kg\geq 16k, so they are negligible.

Suppose now n5n\geq 5. Since dimVi2n1\dim V_{i}\geq 2^{n-1}, we get d-achievable pairs dominated by ((2n2)(k1),2n2k)((2n-2)(k-1),2^{n-2}k), which is again negligible as 2n22n22^{n-2}\geq 2n-2.

4.4.7. Case SO2n\operatorname{SO}^{*}_{2n} for n5n\geq 5

The groups GiG_{i} are of type SO2n()\operatorname{SO}^{*}_{2n}({\mathbb{R}}) or SO2n()\operatorname{SO}_{2n}({\mathbb{R}}), and there is at least one compact factor (of type SO2n()\operatorname{SO}_{2n}({\mathbb{R}})) by 4.5. Moreover, UiU_{i} is not symplectic, and so m2m\geq 2 by 4.4. Hence dimVi4n\dim V_{i}\geq 4n, and all such d-achievable pairs are dominated by ((k1)n(n1)2,2nk)((k-1)\frac{n(n-1)}{2},2nk).

Existence. Let 𝔽{\mathbb{H}}_{\mathbb{F}} be a quaternion algebra over 𝔽{\mathbb{F}} and let H𝔽:=SUn(𝔽,f)H_{\mathbb{F}}:=\operatorname{SU}_{n}({\mathbb{H}}_{\mathbb{F}},f) for some skew-Hermitian form ff on 𝔽n{\mathbb{H}}^{n}_{\mathbb{F}}. By [BH78, Theorem B] the form ff can be chosen to have definite signature at exactly one place of 𝔽{\mathbb{F}}, and to be (non-degenerate) indefinite at all other places. The pair ((k1)n(n1)2,2nk)((k-1)\frac{n(n-1)}{2},2nk) is d-achieved by the group H𝔽H_{\mathbb{F}}.

This completes the proof of 4.6. ∎

Remark 4.8.

As Table 3 shows, both cases 𝐂n{\bf{C}}_{n} and 𝐃n{\bf{D}}_{n} are dominated by the case 𝐀2n1(n){\bf{A}}_{2n-1}(n).

Remark 4.9.

The only d-achievable pair (g1,g)(g-1,g) with g17g\leq 17 occurs for g=2g=2, see Section 4.4.1.

Remark 4.10.

We illustrated the above construction in the 𝐂n{\bf{C}}_{n} and 𝐃n{\bf{D}}_{n} cases because they provide remarkable examples of high-dimensional (though not maximal-dimensional) compact subvarieties of 𝒜g{\mathcal{A}}_{g}.

It would be interesting to determine all maximal irreducible compact subvarieties (i.e. those not properly contained in another irreducible compact subvariety) of 𝒜g{\mathcal{A}}_{g}, in addition to our determination of all maximal-dimensional compact subvarieties, see 5.10 below.

4.5. Consequences

For every g1g\geq 1 define

dmax(g):={g1if g<16;g216for even g16;(g1)216for odd g17.d_{\mathrm{max}}(g):=\begin{cases}\ \ g-1&\mbox{if }g<16\,;\\ \ \ \left\lfloor\tfrac{g^{2}}{16}\right\rfloor&\mbox{for even }g\geq 16\,;\\ \left\lfloor\tfrac{(g-1)^{2}}{16}\right\rfloor&\mbox{for odd }g\geq 17\,.\end{cases}
Lemma 4.11.

For all 1g1g21\leq g_{1}\leq g_{2} we have

dmax(g1+g2)dmax(g1)+dmax(g2).d_{\mathrm{max}}(g_{1}+g_{2})\geq d_{\mathrm{max}}(g_{1})+d_{\mathrm{max}}(g_{2}).

Moreover, the above inequality is strict unless g1=1g_{1}=1 and g216g_{2}\geq 16 is even.

Proof.

Suppose first that g1=1g_{1}=1. Then dmax(1+g2)dmax(g2)d_{\mathrm{max}}(1+g_{2})\geq d_{\mathrm{max}}(g_{2}) and the inequality is strict unless g216g_{2}\geq 16 and is even. So we can assume g22g_{2}\geq 2.

Suppose that g1+g215g_{1}+g_{2}\leq 15. Then dmax(g1+g2)=g1+g21(g11)+(g21)=dmax(g1)+dmax(g2)d_{\mathrm{max}}(g_{1}+g_{2})=g_{1}+g_{2}-1\leq(g_{1}-1)+(g_{2}-1)=d_{\mathrm{max}}(g_{1})+d_{\mathrm{max}}(g_{2}).

Suppose now g1+g216g_{1}+g_{2}\geq 16.

It is easy to check that, for every even gg^{\prime} smaller than g1g_{1}, we have dmax(g1)dmax(g1g)dmax(g2)dmax(g2g)d_{\mathrm{max}}(g_{1})-d_{\mathrm{max}}(g_{1}-g^{\prime})\leq d_{\mathrm{max}}(g_{2})-d_{\mathrm{max}}(g_{2}-g^{\prime}). This implies that it is enough to verify the statement for g1=2g_{1}=2 or g1=3g_{1}=3.

For g1=2g_{1}=2 we have dmax(2+g2)2d_{\mathrm{max}}(2+g_{2})\geq 2 and so dmax(2+g2)>dmax(2)+dmax(g2)d_{\mathrm{max}}(2+g_{2})>d_{\mathrm{max}}(2)+d_{\mathrm{max}}(g_{2}). For g1=3g_{1}=3 we have dmax(3+g2)3d_{\mathrm{max}}(3+g_{2})\geq 3 and so dmax(3+g2)>dmax(3)+dmax(g2)d_{\mathrm{max}}(3+g_{2})>d_{\mathrm{max}}(3)+d_{\mathrm{max}}(g_{2}). ∎

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 (d,g)(d,g) we have ddmax(g)d\leq d_{\mathrm{max}}(g).

Proof.

To obtain the best possible d-achievable pairs (d,g)(d,g), we invoke 4.7.

We first observe that for g15g\leq 15 and for g=17g=17 all cases listed in Table 3 are negligible, i.e. then dg1d\leq g-1. Moreover, we note that

{g216>g1g16for g even(g1)216g1g17for g odd.\begin{cases}\left\lfloor\tfrac{g^{2}}{16}\right\rfloor>g-1\quad\iff g\geq 16&\text{for $g$ even}\\ \left\lfloor\tfrac{(g-1)^{2}}{16}\right\rfloor\geq g-1\quad\iff g\geq 17&\text{for $g$ odd.}\end{cases}

Now assume g16g\geq 16.

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 SLn\operatorname{SL}_{n} case with n3n\geq 3, which gives the d-achievable pairs

(d,g)=((k1)n2n2,kn)(d,g)=\left((k-1)\left\lceil\frac{n}{2}\right\rceil\cdot\left\lfloor\frac{n}{2}\right\rfloor,kn\right)

Case nn even. In this case g=kng=kn is even, and d=g2(k1)4k2d=\frac{g^{2}(k-1)}{4k^{2}}. Hence the pair (d,g)(d,g) is dominated by the case k=2k=2, namely by (g216,g)\left(\frac{g^{2}}{16},g\right).

Case nn odd. Then d=k14(g2k21)d=\frac{k-1}{4}\left(\frac{g^{2}}{k^{2}}-1\right), and we note that dd is strictly decreasing as a function of kk.

If gg is even, then kk must be even and so the dominating pairs are obtained for k=2k=2: in this case d=g216116=g216d=\frac{g^{2}}{16}-\frac{1}{16}=\lfloor\frac{g^{2}}{16}\rfloor.

If g=2+1g=2\ell+1 is odd with 8\ell\geq 8, then k3k\geq 3 and d12(g291)=(2+1)21812d\leq\frac{1}{2}\left(\frac{g^{2}}{9}-1\right)=\frac{(2\ell+1)^{2}}{18}-\frac{1}{2}. Observe that 2414(g1)216\frac{\ell^{2}}{4}-\frac{1}{4}\leq\lfloor\frac{(g-1)^{2}}{16}\rfloor and that (2+1)21812<2414\frac{(2\ell+1)^{2}}{18}-\frac{1}{2}<\frac{\ell^{2}}{4}-\frac{1}{4} for 8\ell\geq 8. Hence, d(g1)216d\leq\lfloor\frac{(g-1)^{2}}{16}\rfloor for g17g\geq 17 odd. ∎

Remark 4.13.

We observe that the pair (d,dmax(g))(d,d_{\mathrm{max}}(g)) as defined in 4.12 is d-achieved if and only if gg is even and g16g\geq 16.

In view of 4.12 and 3.7, a compact subvariety of 𝒜g{\mathcal{A}}_{g} 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 XX be a compact subvariety of 𝒜g{\mathcal{A}}_{g} 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 dimX=dmax(g)\dim X=d_{\mathrm{max}}(g).

Proof.

Recall that, by 3.4, the largest possible dimension of a compact Hodge-generic subvariety of 𝒜g{\mathcal{A}}_{g^{\prime}} is g1g^{\prime}-1. As in 4.12, we separately analyze a few cases.

If g15g\leq 15 and g=17g=17, all d-achievable pairs are negligible and so a maximal-dimensional compact subvariety can be obtained by taking X𝒜gX\subset{\mathcal{A}}_{g} Hodge-generic.

For g=16g=16 or g18g\geq 18 we have dmax(g)>g1d_{\mathrm{max}}(g)>g-1, and so the maximal dimension of XX is achieved using one indecomposable Shimura variety by 4.11.

Thus, for g16g\geq 16 even, the bound is achieved by X=𝒮X={\mathcal{S}} a compact Shimura variety 𝒮{\mathcal{S}} obtained by an irreducible decoupled representation (see 4.13).

For g17g\geq 17 odd, the bound is achieved by X={p}×𝒮𝒜1×𝒜g1𝒜gX=\{p\}\times{\mathcal{S}}\subset{\mathcal{A}}_{1}\times{\mathcal{A}}_{g-1}\subset{\mathcal{A}}_{g}, where 𝒮{\mathcal{S}} is a compact Shimura subvariety of 𝒜g1{\mathcal{A}}_{g-1} of the type considered above in the case of even g16g\geq 16. ∎

Remark 4.15.

There are two interesting cases in which the optimal bound in 4.14 is achieved in two different ways.

  • (a)

    For g=2g=2 a compact curve in 𝒜2{\mathcal{A}}_{2} can be constructed as complete intersection, or as a Shimura variety, as in Section 4.4.1.

  • (b)

    For g=17g=17 a compact subvariety of 𝒜17{\mathcal{A}}_{17} 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 𝒜1{\mathcal{A}}_{1} and a largest (16-dimensional) compact Shimura subvariety of 𝒜16{\mathcal{A}}_{16}, 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).

Finally, we state another consequence of 4.6 which will be used in Section 5.

Corollary 4.16.

Let GG be an anisotropic {\mathbb{Q}}-algebraic group and let VV_{\mathbb{C}} be a decoupled representation of GG such that the number of irreducible linear summands nontrivially acted on by the same component of GG_{\mathbb{R}} is at least 22. Then the associated pair (d,g)(d,g) satisfies ddmax(g)d\leq d_{\mathrm{max}}(g).

Moreover, if d=dmax(g)d=d_{\mathrm{max}}(g), then GG must be almost-simple.

Proof.

Decompose GG (up to isogeny) as a product jG(j)\prod_{j}G_{(j)} of {\mathbb{Q}}-(almost) simple, anisotropic groups and VV as jV(j)\bigoplus_{j}V_{(j)}, where V(j)V_{(j)} is acted on nontrivially by the factor G(j)G_{(j)} only. Hence V(j)V_{(j)} is a decoupled representation of G(j)G_{(j)}.

Now, up to isogenies G(j),G_{(j),{\mathbb{R}}} decomposes as a product of real almost-simple factors as iG(j),i\prod_{i}G_{(j),i} and we can write V(j)=iV(j),iV_{(j)}=\bigoplus_{i}V_{(j),i} with V(j),iV_{(j),i} acted on nontrivially by G(j),iG_{(j),i} only. By construction, V(j),iV_{(j),i} consists of at least 22 irreducible linear summands. Hence each (G(j),V(j))(G_{(j)},V_{(j)}) satisfies the hypotheses of 4.6 and condition (c).

Denote by djd_{j} the dimension of the domain associated to G(j)G_{(j)}, by 2gj2g_{j} the dimension of V(j)V^{\prime}_{(j)}, by 2g2g the dimension of VV_{\mathbb{C}}. Then by 4.6 the pairs (dj,gj)(d_{j},g_{j}) are either negligible or dominated by a d-achievable pair as in Table 3. It follows from 4.12 that djdmax(gj)d_{j}\leq d_{\mathrm{max}}(g_{j}) Since d=jdjd=\sum_{j}d_{j} and g=jgjg=\sum_{j}g_{j}, by 4.11 we conclude that d=jdjjdmax(gj)dmax(g)d=\sum_{j}d_{j}\leq\sum_{j}d_{\mathrm{max}}(g_{j})\leq d_{\mathrm{max}}(g).

If d=dmax(g)d=d_{\mathrm{max}}(g), then GG must be isogenous to the product of at most two almost-simple factors by 4.11. Suppose, for the sake of contradiction, that GG is isogenous to G1×G2G_{1}\times G_{2}, with G1G_{1} and G2G_{2} anisotropic and almost-simple. Up to reordering the factors, 4.11 implies that g1=1g_{1}=1, which gives a contradiction since 𝒜1{\mathcal{A}}_{1} 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 𝒜g{\mathcal{A}}_{g} 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 𝒜g{\mathcal{A}}_{g}.

5.1. Setting

Denote by GG a connected semisimple algebraic group over {\mathbb{Q}}, and let i=1kGi\prod_{i=1}^{k}G_{i} be the decomposition (up to isogeny) of GG_{\mathbb{R}} into almost simple factors. Denote by VV an Sp\operatorname{Sp}-irreducible, symplectic {\mathbb{Q}}-representation of GG.

Then V=αVαV_{\mathbb{C}}=\bigoplus_{\alpha}V_{\alpha} decomposes as a direct sum of GG_{\mathbb{R}}-invariant linear subspaces VαV_{\alpha}, with all VαV_{\alpha} conjugate to each other or to each other’s dual under the Galois group Gal(/)\operatorname{Gal}({\mathbb{C}}/{\mathbb{Q}}). Moreover, each VαV_{\alpha} can be written as Vα=Uα,1Uα,kV_{\alpha}=U_{\alpha,1}\otimes\dots\otimes U_{\alpha,k}, where Uα,iU_{\alpha,i} is an irreducible linear representation of GiG_{i}.

Remark 5.1.

Since all irreducible linear summands of VV_{\mathbb{C}} are Galois conjugate to each other, or to each other’s duals, either all VαV_{\alpha} are self-dual or none of them is. Also, in the self-dual case, either all VαV_{\alpha} are symplectic or all VαV_{\alpha} are orthogonal.

For a fixed α\alpha consider the unordered string of numbers {dimUα,1,,dimUα,k}\{\dim U_{\alpha,1},\dots,\dim U_{\alpha,k}\}, which is independent of α\alpha since VV is Sp\operatorname{Sp}-irreducible, and let NN be the substring consisting of numbers strictly greater than one. Denote by Prod(N)\operatorname{Prod}(N) and Sum(N)\operatorname{Sum}(N) respectively the product and the sum of the elements of NN.

Then for any α\alpha we have dimVα=Prod(N)\dim V_{\alpha}=\operatorname{Prod}(N).

5.2. Two decoupling tricks

Now we introduce two complex decoupled symplectic representations of GG obtained from VV_{\mathbb{C}}. We begin with the following simple observation.

Remark 5.2.

Let ρ:GiGL(Uα,i)\rho:G_{i}\rightarrow\operatorname{GL}(U_{\alpha,i}) be a representation and let ω\omega be the symplectic form on Uα,iUα,iU_{\alpha,i}\oplus U^{\vee}_{\alpha,i} defined by

ω(v,v)=ω(η,η)=0,ω(v,η)=ω(η,v)=η(v)\omega(v,v^{\prime})=\omega(\eta,\eta^{\prime})=0,\qquad\omega(v,\eta)=-\omega(\eta,v)=\eta(v)

for all v,vUα,iv,v^{\prime}\in U_{\alpha,i} and η,ηUα,i\eta,\eta^{\prime}\in U^{\vee}_{\alpha,i}. Then GiG_{i} acts symplectically on Uα,iUα,iU_{\alpha,i}\oplus U^{\vee}_{\alpha,i} via (ρ,ρt)(\rho,\rho^{t}).

First decoupling trick. For every α\alpha, define Uα,iU^{\prime}_{\alpha,i} as Uα,iUα,iU_{\alpha,i}\oplus U^{\vee}_{\alpha,i} if Uα,iU_{\alpha,i} is a nontrivial representation of GiG_{i}, and as {0}\{0\} if Uα,iU_{\alpha,i} is trivial. Put on Uα,iU^{\prime}_{\alpha,i} a symplectic structure as in 5.2. Then let Vα:=i=1kUα,iV^{\prime}_{\alpha}:=\bigoplus_{i=1}^{k}U^{\prime}_{\alpha,i} and V:=αVαV^{\prime}:=\bigoplus_{\alpha}V^{\prime}_{\alpha}, which is thus a complex symplectic representation of GG.

Second decoupling trick. Assume that the VαV_{\alpha}’s are not symplectic. Then, for every α\alpha, define Uα,i′′U^{\prime\prime}_{\alpha,i} as Uα,iU_{\alpha,i} if Uα,iU_{\alpha,i} is a nontrivial representation of GiG_{i}, and as {0}\{0\} otherwise. Then let Vα′′:=i=1kUα,i′′V^{\prime\prime}_{\alpha}:=\bigoplus_{i=1}^{k}U^{\prime\prime}_{\alpha,i} and V′′:=αVα′′V^{\prime\prime}:=\bigoplus_{\alpha}V^{\prime\prime}_{\alpha}.

Since the irreducible linear summands of VV_{\mathbb{C}} are not symplectic, for every VαV_{\alpha} there must be another irreducible linear summand of VV_{\mathbb{C}} isomorphic to VαV^{\vee}_{\alpha}: let VαV_{\alpha^{\vee}} be one such summand. In particular, for every Uα,i′′U^{\prime\prime}_{\alpha,i} there must be another summand of Vα′′V^{\prime\prime}_{\alpha^{\vee}} dual to Uα,i′′U^{\prime\prime}_{\alpha,i}: let Uα,i′′U^{\prime\prime}_{\alpha^{\vee},i^{\vee}} be one such summand. As finding such dual summands is bijective, we can define a symplectic structure on V′′V^{\prime\prime} by pairing each Uα,i′′U^{\prime\prime}_{\alpha,i} to Uα,i′′U^{\prime\prime}_{\alpha^{\vee},i^{\vee}} as in 5.2, and by declaring Uα,i′′U^{\prime\prime}_{\alpha,i} orthogonal to all linear summands different from Uα,i′′U^{\prime\prime}_{\alpha^{\vee},i^{\vee}}.

Here we stress that the complex symplectic representations VV^{\prime} and V′′V^{\prime\prime}

  • need not be defined over {\mathbb{Q}}, and

  • need not be Sp\operatorname{Sp}-irreducible over {\mathbb{C}},

but they have the following properties.

Lemma 5.3.

The complex symplectic representations VV^{\prime} and V′′V^{\prime\prime} of GG are decoupled, and the following hold.

  • (i)

    The number of irreducible linear summands of VV^{\prime} nontrivially acted on by a factor of GG_{\mathbb{R}} is at least 22. Moreover dimVα2Sum(N)\dim V^{\prime}_{\alpha}\leq 2\cdot\operatorname{Sum}(N) for each α\alpha.

  • (ii)

    Assume that some VαV_{\alpha} is not symplectic. Then the number of irreducible linear summands of V′′V^{\prime\prime} nontrivially acted on by a factor of GG_{\mathbb{R}} is at least 22. Moreover dimV′′dimV\dim V^{\prime\prime}\leq\dim V and the equality holds only if N={2,2}N=\{2,2\}.

Proof.

The representations VV^{\prime} and V′′V^{\prime\prime} are decoupled by construction.

(i) follows directly from the construction of the summands VαV^{\prime}_{\alpha}. Indeed, if the factor GiG_{i} nontrivially acts on Uα,iU_{\alpha,i}, then so does on Uα,iU^{\vee}_{\alpha,i}.

(ii) again follows directly from the definition of V′′V^{\prime\prime}, observing that dimVα′′Sum(N)\dim V^{\prime\prime}_{\alpha}\leq\operatorname{Sum}(N). ∎

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 Prod(N)\operatorname{Prod}(N) and Sum(N)\operatorname{Sum}(N).

Definition 5.4.

We call a finite collection NN of integers greater than one decouplable if Prod(N)>2Sum(N)\operatorname{Prod}(N)>2\cdot\operatorname{Sum}(N). A representation VαV_{\alpha} is decouplable if its associated NN is. An irreducible representation VV is decouplable if all of its linear summands VαV_{\alpha} 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 𝒜g{\mathcal{A}}_{g} of maximal dimension.

Lemma 5.5.

Let VV be a decouplable representation of GG corresponding to a map to 𝒜g{\mathcal{A}}_{g}. Then the dimension dd of the domain associated to GG satisfies d<dmax(g)d<d_{\mathrm{max}}(g).

Proof.

We wish to use the first decoupling trick. By 5.3(i) we have a decoupled complex representation VV^{\prime} such that dimV<dimV\dim V^{\prime}<\dim V, exactly because VV is decouplable.

Now, VV^{\prime} might not be defined over {\mathbb{Q}}. Nevertheless, we can conclude since GG and VV^{\prime} satisfy the hypotheses of 4.16. ∎

Remark 5.6.

Note that if VV is a symplectic representation of GG corresponding to a map to 𝒜g{\mathcal{A}}_{g}, and VV satisfies Prod(N)=2Sum(N)\operatorname{Prod}(N)=2\cdot\operatorname{Sum}(N), then the same argument as in 5.5 yields ddmax(g)d\leq d_{\mathrm{max}}(g).

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 NN of integers larger than 11 is non-decouplable if and only if it belongs to the following list:

  1. (i)

    {b}\{b\};

  2. (ii)

    {2,b}\{2,b\};

  3. (iii)

    {3,b}\{3,b\} with 3b63\leq b\leq 6;

  4. (iv)

    {4,4}\{4,4\};

  5. (v)

    {2,2,b}\{2,2,b\} with 2b42\leq b\leq 4;

  6. (vi)

    {2,2,2,2}\{2,2,2,2\}.

Proof.

First, if |N|5|N|\geq 5, then NN is decouplable. Indeed, for N={2,,2}N=\{2,\dots,2\} we have Prod(N)=2|N|>4|N|=2Sum(N)\operatorname{Prod}(N)=2^{|N|}>4|N|=2\cdot\operatorname{Sum}(N), and increasing each number in NN by one results in increasing 2Sum2\cdot\operatorname{Sum} by 22 and Prod\operatorname{Prod} by at least 2|N|12^{|N|-1}.

Let now |N|=4|N|=4 and note that, for N={2,2,2,2}N=\{2,2,2,2\}, we have Prod(N)=16=2Sum(N)\operatorname{Prod}(N)=16=2\cdot\operatorname{Sum}(N) and so {2,2,2,2}\{2,2,2,2\} is non-decouplable. Moreover, increasing each number in NN by one results in increasing 2Sum2\cdot\operatorname{Sum} by 22 and Prod\operatorname{Prod} by at least 88. So all the other cases are decouplable.

Next, if |N|=3|N|=3, then N={2,2,2}N=\{2,2,2\} is non-decouplable since Prod(N)=8<12=2Sum(N)\operatorname{Prod}(N)=8<12=2\cdot\operatorname{Sum}(N). Now, increasing a number in NN by one results in increasing 2Sum2\cdot\operatorname{Sum} by 22 and Prod\operatorname{Prod} by at least 44; so, in order to find non-decouplable sets of three elements, we can only do it at most twice. This shows that {2,2,2},{2,2,3}\{2,2,2\},\{2,2,3\} and {2,2,4}\{2,2,4\} are the only non-decouplable sets of 3 elements, since {2,3,3}\{2,3,3\} is decouplable by inspection.

Now, suppose that N={a,b}N=\{a,b\}, with aba\leq b. Then NN is non-decouplable if and only if ab2a+2bab\leq 2a+2b, which is equivalent to (a2)(b2)4(a-2)(b-2)\leq 4. So either 2=ab2=a\leq b, or 3=ab<63=a\leq b<6, or {a,b}={4,4}\{a,b\}=\{4,4\}

Finally, if |N|=1|N|=1, then it is certainly non-decouplable. ∎

Now we show that non-decouplable representations are never associated to compact Shimura subvarieties of 𝒜g{\mathcal{A}}_{g} of maximal dimension.

Lemma 5.8.

If NN is non-decouplable, then the pair (d,g)(d,g) associated to the GG-representation VV satisfies d<dmax(g)d<d_{\mathrm{max}}(g).

In the proof, we will often use the following observation.

Remark 5.9.

Let GiG_{i} (resp. GjG_{j}) be a real algebraic group and let UiU_{i} (resp. UjU_{j}) be a complex irreducible linear representation of GiG_{i} (resp. of GjG_{j}). Then UiUjU_{i}\otimes U_{j} is a self-dual Gi×GjG_{i}\times G_{j}-representation if and only if UiU_{i} is a self-dual GiG_{i}-representation and UjU_{j} is a self-dual GjG_{j}-representation. Moreover, UiUjU_{i}\otimes U_{j} is symplectic if and only if UiU_{i} is symplectic and UjU_{j} is orthogonal, or if UiU_{i} is orthogonal and UjU_{j} is symplectic.

Proof of 5.8.

We proceed by separately analyzing each non-decouplable case.

Cases N={2,2,2,2}N=\{2,2,2,2\}, N={2,2,2}N=\{2,2,2\} and N={2,2}N=\{2,2\}.

This can only occur if all factors GiG_{i} are of type 𝐀1(1){\bf{A}}_{1}(1) and the dimension of the corresponding symmetric space is 11 for each non-compact GiG_{i}. Hence, the dimension dd of the domain 𝒟+\mathcal{D}^{+} is exactly the number of non-compact factors in GG_{\mathbb{R}}. On the other hand, each linear summand VαV_{\alpha} is nontrivially acted on by at most one non-compact factor of GG_{\mathbb{R}} by 2.8. It follows that there must be at least dd such linear summands, each one of dimension 2|N|2^{|N|}. Hence, dimV2|N|d\dim V\geq 2^{|N|}d and so g2|N|1dg\geq 2^{|N|-1}d. Thus this case is negligible.

We will show now by contradiction that the case d=g1d=g-1 does not occur. Indeed, if d=g1d=g-1, we must have N={2,2}N=\{2,2\} and d=1d=1. This implies that GG is almost-simple and so of type H𝔽/1\mathrm{H}^{1}_{{\mathbb{F}}/{\mathbb{Q}}}, where 𝔽/{\mathbb{F}}/{\mathbb{Q}} is a totally real quadratic extension. So GG_{\mathbb{R}} is a product of the unit quaternions H1\mathrm{H}^{1}_{\mathbb{R}} and SL2,\operatorname{SL}_{2,{\mathbb{R}}}. However, seen as an SL2,\operatorname{SL}_{2,{\mathbb{R}}}-representation, VV_{\mathbb{R}} must have no trivial summands and so it must be isomorphic to the direct sum of 2 copies of standard representation WW_{\mathbb{R}}. The centralizer ZZ inside Sp(V)\operatorname{Sp}(V_{\mathbb{R}}) of such SL2,\operatorname{SL}_{2,{\mathbb{R}}}-representation is isomorphic to SL2,\operatorname{SL}_{2,{\mathbb{R}}}, that acts on VW2V_{\mathbb{R}}\cong W_{\mathbb{R}}\otimes_{\mathbb{R}}{\mathbb{R}}^{2} via its natural action on 2{\mathbb{R}}^{2}. Hence the action of H1\mathrm{H}^{1}_{\mathbb{R}} on VV_{\mathbb{R}} factors through ZZ. Since H1\mathrm{H}^{1}_{\mathbb{R}} has no 2-dimensional real representations, it follows that H1\mathrm{H}^{1}_{\mathbb{R}} acts trivially. This is clearly a contradiction.

Case N={2,2,3}N=\{2,2,3\}.

Here at least two of the factors GiG_{i} correspond to type 𝐀1(1){\bf{A}}_{1}(1) with dimension of the corresponding symmetric space 11, and at least one GjG_{j} to type 𝐀2(1){\bf{A}}_{2}(1), with dimension of the corresponding symmetric space 22. Let 12\ell_{1}\geq 2 be the number of non-compact factors of GG_{\mathbb{R}} of type 𝐀1(1){\bf{A}}_{1}(1) and 21\ell_{2}\geq 1 the number of non-compact factors of type 𝐀2(1){\bf{A}}_{2}(1). The dimension of the domain is then d=1+22d=\ell_{1}+2\ell_{2}. On the other hand, each linear summand VαV_{\alpha} is nontrivially acted on by at most one non-compact factor of GG_{\mathbb{R}} by 2.8. It follows that there must be at least 1+2\ell_{1}+\ell_{2} such linear summands, each one of dimension 12=22312=2\cdot 2\cdot 3. Hence, dimV12(1+2)\dim V\geq 12(\ell_{1}+\ell_{2}) and so g6(1+2)g\geq 6(\ell_{1}+\ell_{2}). Since g3dg\geq 3d, this case is negligible.

Case N={2,2,4}N=\{2,2,4\}.

We proceed as in the case N={2,2,3}N=\{2,2,3\} above. At least two factors GiG_{i} must be of type 𝐀1(1){\bf{A}}_{1}(1); and at least one factor GjG_{j} must be of type 𝐀3(1){\bf{A}}_{3}(1) or 𝐀3(2){\bf{A}}_{3}(2) (with corresponding symmetric spaces of dimension 33 or 44) or 𝐂2{\bf{C}}_{2} (with corresponding symmetric spaces of dimension 33). Let 12\ell_{1}\geq 2 (resp. 2\ell_{2}, 3\ell_{3}) be the number of non-compact factors of GG_{\mathbb{R}} of type 𝐀1(1){\bf{A}}_{1}(1) (resp. 𝐀3(1){\bf{A}}_{3}(1) or 𝐂2{\bf{C}}_{2}, 𝐀3(2){\bf{A}}_{3}(2)), so that the domain has dimension d=1+32+43d=\ell_{1}+3\ell_{2}+4\ell_{3}. Since each linear summand VαV_{\alpha} is nontrivially acted on by at most one non-compact factor of GG_{\mathbb{R}}, the representation VV_{\mathbb{C}} must consist of at least 1+2+3\ell_{1}+\ell_{2}+\ell_{3} linear summands VαV_{\alpha}, each of dimension 1616. It follows that dimV16(1+2+3)>4d\dim V\geq 16(\ell_{1}+\ell_{2}+\ell_{3})>4d, and so g>2dg>2d. This case is thus negligible.

Case N={3,b}N=\{3,b\} with 3b63\leq b\leq 6.

Suppose that dimUα,i=3\dim U_{\alpha,i}=3 and dimUα,j=b\dim U_{\alpha,j}=b. Then Uα,iU_{\alpha,i} is not self-dual, and so VαV_{\alpha} is not symplectic by 5.9. Hence we can use the second decoupling trick, and we consider the representation V′′V^{\prime\prime} constructed in Section 5.2. Since dimVα′′=3+b<3b=dimVα\dim V^{\prime\prime}_{\alpha}=3+b<3b=\dim V_{\alpha}, the pairs (d,g)(d,g), (d,g′′)(d,g^{\prime\prime}) arising from VV and V′′V^{\prime\prime} satisfy g′′2g3g^{\prime\prime}\leq\frac{2g}{3}. Though the representation V′′V^{\prime\prime} might not be defined over {\mathbb{Q}}, the conclusion still follows from 5.3(ii) and 4.16.

Case N={4,4}N=\{4,4\}.

The factors GiG_{i} of GG_{\mathbb{R}} must be either of type 𝐀3(2){\bf{A}}_{3}(2) or 𝐂2{\bf{C}}_{2}.

Consider a linear summand VαUα,i1Uα,i2V_{\alpha}\cong U_{\alpha,i_{1}}\otimes U_{\alpha,i_{2}}, with dimUα,i1=dimUα,i2=4\dim U_{\alpha,i_{1}}=\dim U_{\alpha,i_{2}}=4. By Table 2 either Uα,i1U_{\alpha,i_{1}} is not self-dual, or Uα,i2U_{\alpha,i_{2}} is not self-dual, or both Uα,i1U_{\alpha,i_{1}} and Uα,i2U_{\alpha,i_{2}} are symplectic. In all cases, no linear summand VαV_{\alpha} is not symplectic. Hence we can use the second decoupling trick and consider the representation V′′V^{\prime\prime} produced in Section 5.2. Since dimVα′′=8\dim V^{\prime\prime}_{\alpha}=8 and dimVα=16\dim V_{\alpha}=16, we have dimV′′=12dimV\dim V^{\prime\prime}=\frac{1}{2}\dim V. Hence the pairs (d,g)(d,g) and (d,g′′)(d,g^{\prime\prime}) arising from VV and VV^{\prime} satisfy g′′=g/2g^{\prime\prime}=g/2. Again, the representation V′′V^{\prime\prime} might not be defined over {\mathbb{Q}}, but the conclusion follows from 5.3(ii) and 4.16.

Case N={2,b}N=\{2,b\} with b3b\geq 3.

In this case GG is isogenous to G1×G2G_{1}\times G_{2}, where G1G_{1} is almost {\mathbb{Q}}-simple of type 𝐀1(1){\bf{A}}_{1}(1) and G2G_{2} is almost {\mathbb{Q}}-simple of type different from 𝐀1(1){\bf{A}}_{1}(1).

Since H𝔽1\mathrm{H}^{1}_{{\mathbb{F}}} has no nontrivial 2-dimensional representation, G1G_{1} must be of type Res𝔽/SL2(𝔽)\operatorname{Res}_{{\mathbb{F}}/{\mathbb{Q}}}\operatorname{SL}_{2}({\mathbb{F}}) and Uα,1U_{\alpha,1} must be the standard representation, which is thus symplectic. It also follows that (G1)(G_{1})_{\mathbb{R}} must have at least one compact factor.

We separately analyze two cases, depending on whether VαV_{\alpha} is symplectic.

Case VαV_{\alpha} not symplectic.

We can use the second decoupling trick and consider V′′V^{\prime\prime} produced in Section 5.2. Since dimV′′<dimV\dim V^{\prime\prime}<\dim V, the pairs (d,g)(d,g), (d,g′′)(d,g^{\prime\prime}) arising from VV and V′′V^{\prime\prime} satisfy g′′<gg^{\prime\prime}<g. As before, the representation V′′V^{\prime\prime} might not be defined over {\mathbb{Q}}, but the conclusion follows from 5.3(ii) and 4.16.

Case VαV_{\alpha} symplectic.

By 5.9, the factor Uα,2U_{\alpha,2} is orthogonal. We proceed case by case, analyzing the possibilities for G2G_{2} from Table 2.

We denote by k1k_{1} the number of factors of (G1)(G_{1})_{\mathbb{R}} and by k2k_{2} the number of factors in (G2)(G_{2})_{\mathbb{R}}, so that k=k1+k2k=k_{1}+k_{2}.

Subcase G2G_{2} of type 𝐀n(1){\bf{A}}_{n}(1) with n3(4)n\equiv 3(4).

In this case n=2h1n=2h-1 with h2h\geq 2 even. Moreover (G2)(G_{2})_{\mathbb{R}} must have at least one compact factor by 4.5.

According to Table 2, we have b=dimUα,2=(2hh)b=\dim U_{\alpha,2}=\binom{2h}{h} and d(k11)+(k21)n1+(k3)(2h1)d\leq(k_{1}-1)+(k_{2}-1)n\leq 1+(k-3)(2h-1), since k12k_{1}\geq 2.

Since every Vα=Uα,1Uα,2V_{\alpha}=U_{\alpha,1}\otimes U_{\alpha,2} is nontrivially acted on by at most two factors of GG_{\mathbb{R}}, the representation VV_{\mathbb{C}} must consist of at least k2\frac{k}{2} irreducible linear summands, each of dimension 2(2hh)2\cdot\binom{2h}{h}. It follows that

2g=dimVk22(2hh)=k(2hh).2g=\dim V\geq\frac{k}{2}\cdot 2\cdot\binom{2h}{h}=k\binom{2h}{h}.

It can be easily checked that d1+(k3)(2h1)<k2(2hh)1g1d\leq 1+(k-3)(2h-1)<\frac{k}{2}\binom{2h}{h}-1\leq g-1 for all kk, and so this case is negligible.

Subcase G2G_{2} of type 𝐁n(1){\bf{B}}_{n}(1) with n0,3(4)n\equiv 0,3(4).

As before, (G2)(G_{2})_{\mathbb{R}} must have at least one compact factor by 4.5. From Table 2 we have d(k11)+(k21)(2n+1)<(2n+2)max(k1,k2)1d\leq(k_{1}-1)+(k_{2}-1)(2n+1)<(2n+2)\max(k_{1},k_{2})-1.

Since every irreducible linear summand of VV_{\mathbb{C}} is nontrivially acted on by at most one factor of (G1)(G_{1})_{\mathbb{R}} and one factor of (G2)(G_{2})_{\mathbb{R}}, there are at least max(k1,k2)\max(k_{1},k_{2}) irreducible linear summands, each of dimension 22n2\cdot 2^{n}. Hence, dimVmax(k1,k2)2n+1\dim V\geq\max(k_{1},k_{2})\cdot 2^{n+1}. Since 2n2n+22^{n}\geq 2n+2, we have d<(2n+2)max(k1,k2)12nmax(k1,k2)1g1d<(2n+2)\max(k_{1},k_{2})-1\leq 2^{n}\max(k_{1},k_{2})-1\leq g-1, and so this case is negligible.

Subcase G2G_{2} of type 𝐃n(1){\bf{D}}_{n}(1) with n0(4)n\equiv 0(4).

According to Table 2, we have d(k11)+k2(2n2)<(2n1)max(k1,k2)d\leq(k_{1}-1)+k_{2}(2n-2)<(2n-1)\max(k_{1},k_{2}). As in the previous case, VV_{\mathbb{C}} must consist of at least max(k1,k2)\max(k_{1},k_{2}) linear summands, of dimension 22n12\cdot 2^{n-1} each, and so dimV2nmax(k1,k2)\dim V\geq 2^{n}\cdot\max(k_{1},k_{2}). Again d<(2n1)max(k1,k2)2n1max(k1,k2)1g1d<(2n-1)\max(k_{1},k_{2})\leq 2^{n-1}\max(k_{1},k_{2})-1\leq g-1, and so this case is negligible.

Subcase G2G_{2} of type 𝐃n(n1 or n){\bf{D}}_{n}(n-1\textrm{ or }n) with n5n\geq 5.

Since each VαV_{\alpha} is nontrivially acted on by at most two factors of GG_{\mathbb{R}}, the representation VV_{\mathbb{C}} consists of at least k2\frac{k}{2} irreducible linear summands, each of dimension 22n2\cdot 2n according to Table 2. It follows that dimV2kn\dim V\geq 2kn.

Since (G2)(G_{2})_{\mathbb{R}} must have at least one compact factor by 4.5, we have d(k11)+(k21)n2n21+(k3)n2n2d\leq(k_{1}-1)+(k_{2}-1)\frac{n^{2}-n}{2}\leq 1+(k-3)\frac{n^{2}-n}{2}.

The pair (1+(k3)n2n2,2kn)(1+(k-3)\frac{n^{2}-n}{2},2kn) is dominated by ((k1)n2+n2,2kn)((k-1)\frac{n^{2}+n}{2},2kn), which is achieved in the 𝐂n{\bf{C}}_{n} case, according to Table 3. ∎

5.5. Compact Shimura subvarieties of maximal dimension

Now we can prove our second main result.

Proof of B.

By 5.5, 5.8 and 5.6 a compact subvariety of 𝒜g{\mathcal{A}}_{g} of dimension dmax(g)d_{\mathrm{max}}(g) must be either a Hodge-generic subvariety, or a compact Shimura subvariety associated to a decoupled representation, or a product of the two types.

The conclusion is then a consequence of 4.14 and of 4.15. ∎

4.15 and the proof of B in fact yield more information than B, allowing us to describe all maximal-dimensional compact subvarieties of 𝒜g{\mathcal{A}}_{g}, not just determining their dimensions. We collect such information in the following statement.

Theorem 5.10.

All maximal-dimensional compact subvarieties of 𝒜g{\mathcal{A}}_{g}, in each genus, are described as follows:

  • (i)

    for g=2g=2, the maximal-dimensional compact subvarieties of 𝒜2{\mathcal{A}}_{2} are either Hodge-generic curves (for example, general complete intersections), or Shimura curves (see 4.4.1);

  • (ii)

    for 3g153\leq g\leq 15, all maximal-dimensional compact subvarieties of 𝒜g{\mathcal{A}}_{g} must be Hodge-generic: for example, general complete intersections work;

  • (iii)

    for g16g\geq 16 even, all maximal-dimensional compact subvarieties of 𝒜g{\mathcal{A}}_{g} are compact Shimura subvarieties of the type constructed in Section 4.4.3(ii) with 𝔽/{\mathbb{F}}/{\mathbb{Q}} a totally real quadratic extension (see also 4.12 and 4.13);

  • (iv)

    for g19g\geq 19 odd, all maximal-dimensional compact subvarieties of 𝒜g{\mathcal{A}}_{g} are products of a point in 𝒜1{\mathcal{A}}_{1} and a Shimura subvariety of 𝒜g1{\mathcal{A}}_{g-1} of maximal dimension of the type discussed in (iii);

  • (v)

    for g=17g=17, the maximal-dimensional compact subvarieties of 𝒜17{\mathcal{A}}_{17} are either Hodge-generic (for example, general complete intersections), or the product of a point in 𝒜1{\mathcal{A}}_{1} with a compact 16-dimensional Shimura subvariety of 𝒜16{\mathcal{A}}_{16} 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 𝔽=(2){\mathbb{F}}={\mathbb{Q}}(\sqrt{2}) and 𝔼=(2,i){\mathbb{E}}={\mathbb{Q}}(\sqrt{2},i).

  • (i)

    Assume g=4g=4\ell. Let ff be the Hermitian form on 𝔼2{\mathbb{E}}^{2\ell} defined by f(z):=j=1|zj|2+h=12|z+h|2f(z):=\sum_{j=1}^{\ell}|z_{j}|^{2}+\sum_{h=1}^{\ell}\sqrt{2}|z_{\ell+h}|^{2} Take G=SU(𝔼2,f)G=\operatorname{SU}({\mathbb{E}}^{2\ell},f), so that GG_{\mathbb{R}} is isogenous to SU,×SU2\operatorname{SU}_{\ell,\ell}\times\operatorname{SU}_{2\ell}. Then the domain 𝒟+SU,/S(U×U)\mathcal{D}^{+}\cong\operatorname{SU}_{\ell,\ell}/\mathrm{S}(\mathrm{U}_{\ell}\times\mathrm{U}_{\ell}) maps to g\mathcal{H}_{g} and determines a compact Shimura subvariety of 𝒜g{\mathcal{A}}_{g}, as described in Section 4.4.3(ii).

  • (ii)

    Assume g=4+2g=4\ell+2 and let ff be the Hermitian form on 𝔼2+1{\mathbb{E}}^{2\ell+1} defined by f(z):=j=1+1|zj|2+h=12|z+1+h|2f(z):=\sum_{j=1}^{\ell+1}|z_{j}|^{2}+\sum_{h=1}^{\ell}\sqrt{2}|z_{\ell+1+h}|^{2}. Take G=SU(𝔼2+1,f)G=\operatorname{SU}({\mathbb{E}}^{2\ell+1},f), so that GG_{\mathbb{R}} is isogenous to SU+1,×SU2+1\operatorname{SU}_{\ell+1,\ell}\times\operatorname{SU}_{2\ell+1}. Then the domain 𝒟+SU+1,/S(U+1×U)\mathcal{D}^{+}\cong\operatorname{SU}_{\ell+1,\ell}/\mathrm{S}(\mathrm{U}_{\ell+1}\times\mathrm{U}_{\ell}) maps to g\mathcal{H}_{g} and determines a compact Shimura subvariety of 𝒜g{\mathcal{A}}_{g}, as described in Section 4.4.3(ii).

  • (iii)

    Assume gg odd. Choose a point p𝒜1p\in{\mathcal{A}}_{1} and denote by 𝒳g{\mathcal{X}}_{g} the compact subvariety {p}×𝒳g1\{p\}\times{\mathcal{X}}_{g-1} of 𝒜1×𝒜g1𝒜g{\mathcal{A}}_{1}\times{\mathcal{A}}_{g-1}\subset{\mathcal{A}}_{g}.

Then for any g16g\geq 16 this 𝒳g{\mathcal{X}}_{g} is a maximal-dimensional compact subvariety of 𝒜g{\mathcal{A}}_{g}.

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 𝒜gind{\mathcal{A}}_{g}^{\rm ind} and the locus of Jacobians inside 𝒜g{\mathcal{A}}_{g}, in particular proving C.

6.1. The locus 𝒜gind{\mathcal{A}}_{g}^{\rm ind} of indecomposable ppav

As mentioned in Section 1.3, the moduli space 𝒜gind{\mathcal{A}}_{g}^{\rm ind} of indecomposable ppav of dimension gg 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 𝒜gind𝒜g{\mathcal{A}}_{g}^{\rm ind}\subsetneq{\mathcal{A}}_{g}^{*}, we see that the boundary 𝒜g𝒜gind{\mathcal{A}}_{g}^{*}\setminus{\mathcal{A}}_{g}^{\rm ind} has one irreducible component 𝒜1×𝒜g1{\mathcal{A}}_{1}^{*}\times{\mathcal{A}}_{g-1}^{*} that has maximal dimension, which has codimension g1g-1. Thus g2mdimc,gen(𝒜gind)g-2\leq\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g}^{\rm ind}), while A of course implies the following.

Corollary 6.1.

For g2g\geq 2 the following holds:

g2mdimc,gen(𝒜gind)g1.g-2\leq\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g}^{\rm ind})\leq g-1.

It would be very interesting to know which value it in fact is. As we will now see, mdimc(𝒜gind)=mdimc,gen(𝒜gind)=g2\operatorname{mdim_{c}}({\mathcal{A}}_{g}^{\rm ind})=\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g}^{\rm ind})=g-2 for g=2,3,4g=2,3,4, and it is natural to wonder if this is the case for all gg.

Indeed, the cases of g=2,3g=2,3 are classical: 𝒜2ind2{\mathcal{A}}_{2}^{\rm ind}\cong{\mathcal{M}}_{2} is affine, while 𝒜3ind3{\mathcal{A}}_{3}^{\rm ind}\cong{\mathcal{M}}_{3} does not contain a compact surface, for example by [Dia84]. In general, recall that λi\lambda_{i} denotes the ii’th Chern class of the Hodge rank gg vector bundle on 𝒜g{\mathcal{A}}_{g}, and that the tautological ring R(𝒜g)CH(𝒜g)R^{*}({\mathcal{A}}_{g})\subset CH^{*}_{\mathbb{Q}}({\mathcal{A}}_{g}) is the subring generated by the classes λi\lambda_{i}.

For g=4g=4, from the results of [HT12, HT18] it follows that the classes of all algebraic subvarieties of 𝒜4{\mathcal{A}}_{4} lie in R(𝒜4)R^{*}({\mathcal{A}}_{4}), and in fact the class of 𝒜1×𝒜3{\mathcal{A}}_{1}\times{\mathcal{A}}_{3} is a nonzero multiple of the class λ3\lambda_{3}. Thus the main result of [KS03] shows that any compact subvariety XX of 𝒜4𝒜1×𝒜3{\mathcal{A}}_{4}\setminus{\mathcal{A}}_{1}\times{\mathcal{A}}_{3} must satisfy dimX32/21=2\dim X\leq 3\cdot 2/2-1=2, since λ3|X=0\lambda_{3}|_{X}=0. This implies mdimc(𝒜gind)=2\operatorname{mdim_{c}}({\mathcal{A}}_{g}^{\rm ind})=2.

However, the situation in higher genus remains mysterious. By [vdG99], for any gg the homology class of [E]×𝒜g1[E]\times{\mathcal{A}}_{g-1} is a multiple of the top Hodge class λg\lambda_{g}, which in fact vanishes on 𝒜g{\mathcal{A}}_{g}. As detailed in [CMOP24], it is possible to define a projection map CH(𝒜g)R(𝒜g)CH^{*}_{\mathbb{Q}}({\mathcal{A}}_{g})\to R^{*}({\mathcal{A}}_{g}). Then in [COP24] the authors show that, in general, the projection of the class of 𝒜1×𝒜g1{\mathcal{A}}_{1}\times{\mathcal{A}}_{g-1} to the tautological ring of 𝒜g{\mathcal{A}}_{g} is a nonzero multiple of λg1\lambda_{g-1}, but also that for g=6g=6 the class of 𝒜1×𝒜5{\mathcal{A}}_{1}\times{\mathcal{A}}_{5} does not lie in R(𝒜6)R^{*}({\mathcal{A}}_{6}). This makes considerations similar to the g=4g=4 case above impossible in higher genus.

Note also that as the maximal-dimensional compact Shimura subvarieties of 𝒜g{\mathcal{A}}_{g} constructed in Section 4.4.3 are not contained in 𝒜gind{\mathcal{A}}_{g}^{\rm ind}, our results do not suffice to determine mdimc(𝒜gind)\operatorname{mdim_{c}}({\mathcal{A}}_{g}^{\rm ind}) for any g5g\geq 5, and it would be interesting to find explicit high-dimensional compact Shimura subvarieties of 𝒜gind{\mathcal{A}}_{g}^{\rm ind}.

6.2. The locus of Jacobians, and the moduli space of curves

We now prove the results on compact subvarieties of gct{\mathcal{M}}_{g}^{\rm ct} and J(gct)J({\mathcal{M}}_{g}^{\rm ct}).

Proof of C.

As already mentioned in the introduction, the upper bound mdimc(J(gct))g1\operatorname{mdim_{c}}(J({\mathcal{M}}_{g}^{\rm ct}))\leq g-1 for g15g\leq 15 simply follows from the inclusion J(gct)𝒜gJ({\mathcal{M}}_{g}^{\rm ct})\subset{\mathcal{A}}_{g} and the g15g\leq 15 case of B, giving mdimc(𝒜g)=mdimc,gen(𝒜g)=g1\operatorname{mdim_{c}}({\mathcal{A}}_{g})=\operatorname{mdim_{c,gen}}({\mathcal{A}}_{g})=g-1 for that range of gg.

To show that mdimc,gen(gct)=3g22\operatorname{mdim_{c,gen}}({\mathcal{M}}_{g}^{\rm ct})=\lfloor\tfrac{3g}{2}\rfloor-2 for 2g232\leq g\leq 23, we recall the construction of a compact subvariety contained in the boundary gctg{\mathcal{M}}_{g}^{\rm ct}\setminus{\mathcal{M}}_{g}, already pointed out by the first author for [Kri12].

Indeed, since 𝒜2=J(2ct){\mathcal{A}}_{2}=J({\mathcal{M}}_{2}^{\rm ct}) and 𝒜3=J(3ct){\mathcal{A}}_{3}=J({\mathcal{M}}_{3}^{\rm ct}) contain a compact curve and surface, respectively, the statement is true for g=2g=2 and g=3g=3. For 4g234\leq g\leq 23, we proceed by induction, assuming the result for all g<gg^{\prime}<g. Indeed, by B, for gg in this range we have mdimc(𝒜g)<3g22\operatorname{mdim_{c}}({\mathcal{A}}_{g})<\lfloor\tfrac{3g}{2}\rfloor-2, and so an irreducible compact subvariety XgctX\subset{\mathcal{M}}_{g}^{\rm ct} that satisfies XgX\cap{\mathcal{M}}_{g}\neq\emptyset (and thus maps generically 1-to-1 to its image in 𝒜g{\mathcal{A}}_{g} under JJ) has dimension strictly smaller than 3g22\lfloor\tfrac{3g}{2}\rfloor-2. Thus it is enough to deal with irreducible compact XgctX\subset\partial{\mathcal{M}}_{g}^{\rm ct}. Such an XX must then be contained in some irreducible component g,1ct×gg,1ct{\mathcal{M}}_{g^{\prime},1}^{\rm ct}\times{\mathcal{M}}_{g-g^{\prime},1}^{\rm ct} of the boundary gct=gctg\partial{\mathcal{M}}_{g}^{\rm ct}={\mathcal{M}}_{g}^{\rm ct}\setminus{\mathcal{M}}_{g}. But since the forgetful map k,1ct{\mathcal{M}}_{k,1}^{\rm ct} has compact curve fibers, the maximal dimension of a compact subvariety of k,1ct{\mathcal{M}}_{k,1}^{\rm ct} is equal to 11 plus the maximal dimension of a compact subvariety of kct{\mathcal{M}}_{k}^{\rm ct}, it then follows from the inductive assumption that

dimX(3g22)+1+(3(gg)22)+13g22,\dim X\leq\left(\lfloor\tfrac{3g^{\prime}}{2}\rfloor-2\right)+1+\left(\lfloor\tfrac{3(g-g^{\prime})}{2}\rfloor-2\right)+1\leq\lfloor\tfrac{3g}{2}\rfloor-2\,,

which proves the upper bound mdimc,gen(gct)3g22\operatorname{mdim_{c,gen}}({\mathcal{M}}_{g}^{\rm ct})\leq\lfloor\tfrac{3g}{2}\rfloor-2 for g23g\leq 23. We now construct an explicit compact subvariety of this dimension. For even g=2kg=2k we consider inside the boundary stratum

2,1ct×2,2ct××2,2ct×2,1ctgct{\mathcal{M}}_{2,1}^{\rm ct}\times{\mathcal{M}}_{2,2}^{\rm ct}\times\dots\times{\mathcal{M}}_{2,2}^{\rm ct}\times{\mathcal{M}}_{2,1}^{\rm ct}\subset{\mathcal{M}}_{g}^{\rm ct}

the product of two compact surfaces in 2,1ct{\mathcal{M}}_{2,1}^{\rm ct}, and k2k-2 compact threefolds in 2,2ct{\mathcal{M}}_{2,2}^{\rm ct}, that are preimages of a compact curve in 2ct{\mathcal{M}}_{2}^{\rm ct}, giving altogether a product variety of dimension 22+3(k2)=3g222\cdot 2+3\cdot(k-2)=\tfrac{3g}{2}-2. For g=2k+1g=2k+1 odd, we do the same except taking the last factor to be a compact threefold in 3,1ct{\mathcal{M}}_{3,1}^{\rm ct} that is the preimage of a compact surface in 3ct{\mathcal{M}}_{3}^{\rm ct}.

Thus for any g3g\geq 3 we construct a compact subvariety of gctg{\mathcal{M}}_{g}^{\rm ct}\setminus{\mathcal{M}}_{g} of dimension 3g22\lfloor\tfrac{3g}{2}\rfloor-2. If for a given gg this dimension is greater than mdimc(𝒜g)\operatorname{mdim_{c}}({\mathcal{A}}_{g}), then it follows that for this gg we have mdimc(gct)=3g/22\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct})=\lfloor 3g/2\rfloor-2. This happens to be the case if and only if g23g\leq 23, as 32422=34<242/16=36\lfloor\tfrac{3\cdot 24}{2}\rfloor-2=34<24^{2}/16=36, while 32322=32>32222=31>22216=30\lfloor\tfrac{3\cdot 23}{2}\rfloor-2=32>\lfloor\tfrac{3\cdot 22}{2}\rfloor-2=31>\lfloor\tfrac{22^{2}}{16}\rfloor=30. ∎

The compact subvariety costructed above yields mdimc(gct)3g22\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct})\geq\lfloor\tfrac{3g}{2}\rfloor-2 for any g3g\geq 3, and considering its image under JJ gives mdimc(J(gct))g2\operatorname{mdim_{c}}(J({\mathcal{M}}_{g}^{\rm ct}))\geq\lfloor\tfrac{g}{2}\rfloor, which is obtained using a variety contained in the boundary, and we do not know any construction of a higher-dimensional compact XJ(gct)X\subset J({\mathcal{M}}_{g}^{\rm ct}) that intersects both J(g)J({\mathcal{M}}_{g}) and J(gct)J(\partial{\mathcal{M}}_{g}^{\rm ct}).

Remark 6.2.

We see no reason to expect the bound mdimc(J(gct))g1\operatorname{mdim_{c}}(J({\mathcal{M}}_{g}^{\rm ct}))\leq g-1 for g15g\leq 15 to be sharp. It would be very interesting to improve it, eg. by bounding the dimensions of the intersections of compact Shimura varieties with J(gct)J({\mathcal{M}}_{g}^{\rm ct}), 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 g{\mathcal{M}}_{g} and gct{\mathcal{M}}_{g}^{\rm ct}.

g3456151617182324100mdimc,gen(gct)22222222222mdimc(gct)=2457202223253234148mdimc(J(gct))2345141616203036198mdimc,gen(g)11111111111covers: mdimc(g)11112333335Diaz: mdimc(g)123413141516212298\begin{array}[]{|r||rrrr||rrrr||rr||r|}\hline\cr g&3&4&5&6&15&16&17&18&23&24&100\\ \hline\cr\hbox{$\operatorname{mdim_{c,gen}}({\mathcal{M}}_{g}^{\rm ct})\geq$}&2&2&2&2&2&2&2&2&2&2&2\\ \hbox{$\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct})=$}&2&4&5&7&20&22&23&25&32&\geq 34&\geq 148\\ \hbox{$\operatorname{mdim_{c}}(J({\mathcal{M}}_{g}^{\rm ct}))\leq$}&2&3&4&5&14&16&16&20&30&36&198\\ \hline\cr\hbox{$\operatorname{mdim_{c,gen}}({\mathcal{M}}_{g})\geq$}&1&1&1&1&1&1&1&1&1&1&1\\ \hbox{\it covers: $\operatorname{mdim_{c}}({\mathcal{M}}_{g})\geq$}&1&1&1&1&2&3&3&3&3&3&5\\ \hbox{\it Diaz: $\operatorname{mdim_{c}}({\mathcal{M}}_{g})\leq$}&1&2&3&4&13&14&15&16&21&22&98\\ \hline\cr\end{array}

Table 4. Known results for maximal dimensions of compact subvarieties of gct{\mathcal{M}}_{g}^{\rm ct} and g{\mathcal{M}}_{g}

Here we recall that Keel and Sadun proved mdimc(gct)2g4\operatorname{mdim_{c}}({\mathcal{M}}_{g}^{\rm ct})\leq 2g-4 for any g3g\geq 3, while the lower bounds on mdimc,gen(g)\operatorname{mdim_{c,gen}}({\mathcal{M}}_{g}) and mdimc,gen(gct)\operatorname{mdim_{c,gen}}({\mathcal{M}}_{g}^{\rm ct}) simply follow from considering the Satake compactification g{\mathcal{M}}_{g}^{*}, which is the closure of J(g)J({\mathcal{M}}_{g}) in 𝒜g{\mathcal{A}}_{g}^{*}, so that codim(gJ(g))=2\operatorname{codim}({\mathcal{M}}_{g}^{*}\setminus J({\mathcal{M}}_{g}))=2 and codim(gJ(gct))=3\operatorname{codim}({\mathcal{M}}_{g}^{*}\setminus J({\mathcal{M}}_{g}^{\rm ct}))=3.

The lower bounds for mdimc(g)\operatorname{mdim_{c}}({\mathcal{M}}_{g}) are obtained by covering constructions starting either from a compact curve in 3{\mathcal{M}}_{3} or from a one-dimensional compact family of pairs of distinct points on a fixed curve of genus 22. 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 dd-fold exists in g{\mathcal{M}}_{g} for any g2d+1g\geq 2^{d+1}. Finally, the best known upper bound for mdimc(g)\operatorname{mdim_{c}}({\mathcal{M}}_{g}) is the famous 40 year old theorem of Diaz [Dia84] (which by now has multiple proofs): mdimc(g)g2\operatorname{mdim_{c}}({\mathcal{M}}_{g})\leq g-2 for any g3g\geq 3.

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 g{\mathcal{M}}_{g}, 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 𝒜4{\mathcal{A}}_{4}, Doc. Math. 17 (2012), 195–244. MR 2946823
  • [HT18] by same author, The topology of 𝒜g{\mathcal{A}}_{g} 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 𝒜g{\mathcal{A}}_{g}\otimes{\mathbb{C}}, 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.