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

Arithmetic representations of mapping class groups

Eduard Looijenga Mathematisch Instituut, Universiteit Utrecht (Nederland) and Mathematics Department, University of Chicago (USA) e.j.n.looijenga@uu.nl
Abstract.

Let SS be a closed oriented surface and GG a finite group of orientation preserving automorphisms of SS whose orbit space has genus at least 22. There is a natural group homomorphism from the GG-centralizer in Diff+(S)\operatorname{Diff}^{+}(S) to the GG-centralizer in Sp(H1(S))\operatorname{Sp}(H_{1}(S)). We give a sufficient condition for its image to be a subgroup of finite index.

Key words and phrases:
Mapping class group, Arithmetic group
2020 Mathematics Subject Classification:
Primary: 57K20, 57M12; Secondary: 11E39
Part of the research for this paper was done when the author was supported by the Chinese National Science Foundation and by the Jump Trading Mathlab Research Fund

1. Introduction and statement of the main result

Let SS be a closed connected oriented surface of genus 2\geq 2. The group Diff+(S)\operatorname{Diff}^{+}(S) of orientation preserving diffeomorphisms of SS acts on H1(S)\operatorname{H}_{1}(S) via its connected component group π0(Diff+(S))\pi_{0}(\operatorname{Diff}^{+}(S)), known as the mapping class group of SS, and it is a classical fact that the image of this representation is the full symplectic group Sp(H1(S))\operatorname{Sp}(\operatorname{H}_{1}(S)) of integral linear transformations which preserve the intersection form on H1(S)\operatorname{H}_{1}(S). This paper concerns an equivariant version, where it is assumed that we are given a finite subgroup GDiff+(S)G\subset\operatorname{Diff}^{+}(S). The centralizer Diff+(S)G\operatorname{Diff}^{+}(S)^{G} of GG in Diff+(S)\operatorname{Diff}^{+}(S) lands under the above symplectic representation in Sp(H1(S))G\operatorname{Sp}(\operatorname{H}_{1}(S))^{G} and the question we address here is how much smaller the image is. Besides its intrinsic interest, the answer has consequences for understanding the mapping class group of the GG-orbit space of SS. We shall regard the latter as an orbifold surface and denote it by SGS_{G}; the regular orbits then define an open subset SGSGS_{G}^{\circ}\subset S_{G} with finite complement. This punctured surface SGS_{G}^{\circ} has negative Euler characteristic. The image of Diff+(S)G\operatorname{Diff}^{+}(S)^{G} in the mapping class group of the punctured surface SGS_{G}^{\circ} is of finite index and thus makes Sp(H1(S))\operatorname{Sp}(\operatorname{H}_{1}(S)) a ‘virtual representation’ of that mapping class group.

The work of Putman-Wieland [6] relates our question to the Ivanov conjecture as follows. Let us say that the GG-action on SS has the Putman-Wieland property if Diff+(S)G\operatorname{Diff}^{+}(S)^{G} has no finite orbit in H1(S){0}\operatorname{H}_{1}(S)\smallsetminus\{0\}. These authors prove that if that property holds for a given genus hh of SGS_{G} (no matter what SS and GG are), then every finite index subgroup of a mapping class group of a connected oriented surface of finite type of genus >h>h has zero first Betti number. The first part of our main result is about that property.

Theorem 1.1.

Let SSGS\to S_{G} be a GG-cover as above.

  1. (i)

    If this cover is trivial over a compact genus one subsurface of SGS_{G}^{\circ} with connected boundary, then the action of Diff+(S)G\operatorname{Diff}^{+}(S)^{G} on H1(S)\operatorname{H}_{1}(S) has no nonzero finite orbits.

  2. (ii)

    If this cover is trivial over a compact genus two subsurface of SGS_{G}^{\circ} with connected boundary, then the image of Diff+(S)G\operatorname{Diff}^{+}(S)^{G} in Sp(H1(S))G\operatorname{Sp}(\operatorname{H}_{1}(S))^{G} is of finite index.

We will also obtain an arithmeticity property in the setting of (i). See the discussion and the end of this introduction as well as Remark 5.6.

Remark 1.2.

Note that in either case the cover over the complement of such a subsurface of SGS_{G} must be connected (for the subsurface has a connected boundary and SS is connected). Since this complement has a boundary component over which the covering is trivial, we may contract that component (and each of the components lying over it) to obtain a GG-covering SSGS^{\prime}\to S^{\prime}_{G}, where the genus of SGS^{\prime}_{G} is now 1 resp.  2 less than that of SGS_{G}. As this covering represents all the topological input, we may paraphrase our main theorem as saying that the Putman-Wieland property resp. the arithmeticity property holds after a ‘stabilization’ by taking a connected sum of the base orbifold with a closed surface of genus 1 resp. 2.

We will prove this theorem under the apparently weaker assumption that there exists a closed one-dimensional submanifold nonempty ASGA\subset S_{G}^{\circ}, so a disjoint union of say k1k\geq 1 embedded circles (with k=1k=1 in case (i) and k=2k=2 in case (ii)) such that SSGS\to S_{G} is trivial over AA and connected over SGAS_{G}\smallsetminus A. This looks as if this is a more general result, because it is easy to find in the respective cases such an AA inside the postulated subsurface with the property that its complement is connected. But we will see that this generalization is only apparent.

There is also a useful geometric interpretation for this last formulation: given such an AA, then we can obtain the GG-covering SSGS^{\prime}\to S^{\prime}_{G} as above as follows: regard SGAS_{G}\smallsetminus A as a punctured surface (so with two punctures for each component of AA) and let SGSGAS^{\prime}_{G}\supset S_{G}\smallsetminus A be the closed orbifold obtained by filling in these punctures as non-orbifold points. Our assumptions say that SGS^{\prime}_{G} is a closed connected surface (the genus drop is the number of connected components of AA) and that the given GG-covering SSGS\to S_{G} arises from a GG-covering SSGS^{\prime}\to S^{\prime}_{G} with for each component of AA an identification of the fibers of this covering over the two associated points (as principal GG-sets). If we give SGS_{G} a complex structure and thus turn it into a smooth complex-projective curve with an orbifold structure, then an algebraic geometer might be tempted to regard this orbifold curve as being in its moduli space near the Deligne-Mumford stratum where the orbifold acquires kk nodes, but for which the GG-covering stays irreducible and does not ramify over the nodes. The covering SSGS^{\prime}\to S^{\prime}_{G} then appears as the normalization of such a degeneration. No algebraic geometry is used in the proof, though, for the topological part of this paper uses methods that directly generalize those of [3].

Let us compare the above theorem with the work of Grünewald-Larsen-Lubotzky-Malestein [1], whose main motivation was to construct, via the virtual isomorphism mentioned above, new arithmetic quotients of the mapping class group of SGS_{G}. They assume that GG acts freely so that SG=SGS_{G}=S^{\circ}_{G} and impose another, more technical condition, which in our set-up translates into that we are in the context of (i) and demand that the covering SSGS^{\prime}\to S^{\prime}_{G} is of ‘handlebody type’, in the sense that it extends to a handlebody that has SGS^{\prime}_{G} as boundary. They prove that the image of Diff+(S)G\operatorname{Diff}^{+}(S)^{G} in each simple factor of Sp(H1(S,))G\operatorname{Sp}(\operatorname{H}^{1}(S,{\mathbb{Q}}))^{G} is arithmetic. Our approach differs from theirs in several aspects, but mostly in our direct and relatively simple way of constructing GG-equivariant mapping classes.

When speaking of arithmetic subgroups of Sp(H1(S,))G\operatorname{Sp}(\operatorname{H}^{1}(S,{\mathbb{Q}}))^{G}, it is of course tacitly understood that the latter can be regarded as the group of rational points of an algebraic group defined over {\mathbb{Q}}. Let us make this explicit.

Denote by X(G)X({\mathbb{Q}}G) the set of irreducible characters of G{\mathbb{Q}}G and choose for every χX(G)\chi\in X({\mathbb{Q}}G) a representing irreducible (left) G{\mathbb{Q}}G-module VχV_{\chi}. We also fix on every VχV_{\chi} a GG-invariant inner product sχ:Vχ×Vχs_{\chi}:V_{\chi}\times V_{\chi}\to{\mathbb{Q}} (which can be obtained as the GG-average of an arbitrary inner product). This exhibits VχV_{\chi} as a self-dual G{\mathbb{Q}}G-module. Then EndG(Vχ)\operatorname{End}_{{\mathbb{Q}}G}(V_{\chi}) is a skew field which is of finite {\mathbb{Q}}-dimension. We denote its opposite by DχD_{\chi} (meaning that the underling {\mathbb{Q}}-vector space is EndG(Vχ)\operatorname{End}_{{\mathbb{Q}}G}(V_{\chi}), but that composition is taken in opposite order) so that VχV_{\chi} is now a right DχD_{\chi}-module. Adopting as a convention that DχD_{\chi} used as superscript (resp. subscript) indicates that we are dealing with right (resp. left) DχD_{\chi}-module endomorphisms, then the natural map

GχX(G)EndDχ(Vχ),\textstyle{\mathbb{Q}}G\cong\prod_{\chi\in X({\mathbb{Q}}G)}\operatorname{End}^{D_{\chi}}(V_{\chi}),

is an isomorphism of {\mathbb{Q}}-algebras. This is in fact the Wedderburn decomposition of G{\mathbb{Q}}G, as each factor is a minimal 22-sided ideal.

The group algebra G{\mathbb{Q}}G comes with an anti-involution rrr\mapsto r^{\dagger} which takes each basis element ege_{g} (gGg\in G) to the basis element eg1e_{g^{-1}}. This identifies G{\mathbb{Q}}G with its opposite. Since VχV_{\chi} is self-dual, the involution leaves in the above decomposition each factor EndDχ(Vχ)\operatorname{End}^{D_{\chi}}(V_{\chi}) invariant and induces one in the skew-field DχD_{\chi}: the involution on EndDχ(Vχ)\operatorname{End}^{D_{\chi}}(V_{\chi}) is given by taking the sχs_{\chi}-adjoint: sχ(σv,v)=sχ(v,σv)s_{\chi}(\sigma v,v^{\prime})=s_{\chi}(v,\sigma^{\dagger}v^{\prime}). Since we have DχD_{\chi} acting on VχV_{\chi} on the right, this means that sχ(vλ,v)=sχ(v,vλ)s_{\chi}(v\lambda,v^{\prime})=s_{\chi}(v,v^{\prime}\lambda^{\dagger}). (We note in passing that any other GG-invariant inner product sχs^{\prime}_{\chi} on VχV_{\chi} is of the form sχ(vλ,v)s_{\chi}(v\lambda,v^{\prime}) for some nonzero λ\lambda with λ=λ\lambda^{\dagger}=\lambda; the associated anti-involution of DχD_{\chi} is then a conjugate of \dagger.) The center of DχD_{\chi}, which we denote by LχL_{\chi}, is a number field and the fixed point set of \dagger in LχL_{\chi} is a subfield KχLχK_{\chi}\subset L_{\chi} with [Lχ:Kχ]2[L_{\chi}:K_{\chi}]\leq 2.

For a finitely generated G{\mathbb{Q}}G-module HH, we denote by H[χ]H[\chi] the associated χ\chi-isogeny space HomG(Vχ,H)\operatorname{Hom}_{{\mathbb{Q}}G}(V_{\chi},H). The right DχD_{\chi}-module structure on VχV_{\chi} determines a left DχD_{\chi}-module structure on H[χ]H[\chi] and the isotypical decomposition of HH is the assertion that the natural map

χX(G)VχDχH[χ]H,vDχuVχDχH[χ]u(v)\oplus_{\chi\in X({\mathbb{Q}}G)}V_{\chi}\otimes_{D_{\chi}}H[\chi]\to H,\quad v\otimes_{D_{\chi}}u\in V_{\chi}\otimes_{D_{\chi}}H[\chi]\mapsto u(v)

is an isomorphism of G{\mathbb{Q}}G-modules. So the χ\chi-isotypical subspace of HH, ie, the image HχH_{\chi} of VχDχH[χ]V_{\chi}\otimes_{D_{\chi}}H[\chi] in VV, has the structure of a KχK_{\chi}-vector space.

Assume now that HH comes equipped with a nondegenerate GG-invariant symplectic form (a,b)H×Hab(a,b)\in H\times H\mapsto a\cdot b\in{\mathbb{Q}}. Then the isotypical decomposition of HH is symplectic, so that we also have a decomposition Sp(H)G=χX(G)Sp(Hχ)G\operatorname{Sp}(H)^{G}=\prod_{\chi\in X({\mathbb{Q}}G)}\operatorname{Sp}(H_{\chi})^{G}. Every factor Sp(Hχ)G\operatorname{Sp}(H_{\chi})^{G} can be understood with the help of the skew-hermitian form

(f,f)H[χ]×H[χ]f,fχDχ,(f,f^{\prime})\in H[\chi]\times H[\chi]\mapsto\langle f,f^{\prime}\rangle_{\chi}\in D_{\chi},

which is characterized by the property that for all v,vVχv,v^{\prime}\in V_{\chi},

f(v)f(v)=sχ(vf,fχ,v)f(v)\cdot f^{\prime}(v^{\prime})=s_{\chi}(v\langle f,f^{\prime}\rangle_{\chi},v^{\prime})

(skew-hermitian means that the form is DχD_{\chi}-linear in the first variable and f,fχ=f,fχ\langle f^{\prime},f\rangle_{\chi}=-\langle f,f^{\prime}\rangle_{\chi}^{\dagger}). Indeed, for fixed ff and ff^{\prime}, the map (v,v)Vχ×Vχf(v)f(v)(v,v^{\prime})\in V_{\chi}\times V\chi\mapsto f(v)\cdot f(v^{\prime})\in{\mathbb{Q}} is a bilinear form on VχV_{\chi}. Since sχ:Vχ×Vχs_{\chi}:V_{\chi}\times V_{\chi}\to{\mathbb{Q}} is nondegenerate, there exists a unique {\mathbb{Q}}-linear endomorphism σ\sigma of VχV_{\chi} such that f(v)f(v)=sχ(σ(v),v)f(v)\cdot f(v^{\prime})=s_{\chi}(\sigma(v),v^{\prime}). The GG-invariance of both bilinear forms implies that σ\sigma is GG-equivariant, ie, is an element of EndG(Vχ)\operatorname{End}_{{\mathbb{Q}}G}(V_{\chi}). We prefer to regard it as an element of its opposite DχD_{\chi}, so that f(v)f(v)=sχ(vσ,v)f(v)\cdot f(v^{\prime})=s_{\chi}(v\sigma,v^{\prime}). This σ\sigma is evidently {\mathbb{Q}}-linear in both ff and ff^{\prime} and that is why we denote it f,fχDχ\langle f,f^{\prime}\rangle_{\chi}\in D_{\chi}. It is then a little exercise to check that ,χ\langle\;,\;\rangle_{\chi} is skew-hermitian. This form is nondegenerate in an obvious sense. The group of automorphisms H[χ]H[\chi] that preserve this form is a generalized unitary group and therefore denoted U(H[χ])\operatorname{U}(H[\chi]).

Any element of Sp(Hχ)G\operatorname{Sp}(H_{\chi})^{G} acts via the isomorphism HχVχDχH[χ]H_{\chi}\cong V_{\chi}\otimes_{D_{\chi}}H[\chi] as an element of the form 1Vχu1_{V_{\chi}}\otimes u with uU(H[χ])u\in\operatorname{U}(H[\chi]) and this identifies Sp(Hχ)G\operatorname{Sp}(H_{\chi})^{G} with U(H[χ])\operatorname{U}(H[\chi]). The group U(H[χ])\operatorname{U}(H[\chi]) is the group of KχK_{\chi}-points of a reductive algebraic group defined over KχK_{\chi}, whereas Sp(Hχ)G\operatorname{Sp}(H_{\chi})^{G} is the group of {\mathbb{Q}}-points of an algebraic group defined over {\mathbb{Q}}. Indeed, the latter is obtained from the former by the restriction of scalars Kχ|K_{\chi}|{\mathbb{Q}}.

Theorem 1.1 then amounts to the assertion that the image of Diff+(S)G\operatorname{Diff}^{+}(S)^{G} in the product of unitary groups χX(G)U(H1(S;)[χ])\prod_{\chi\in X({\mathbb{Q}}G)}\operatorname{U}(\operatorname{H}_{1}(S;{\mathbb{Q}})[\chi]) is arithmetic. We use this decomposition to prove the theorem, since we first prove arithmeticity for a single factor. This leads to a somewhat stronger result, for we show that in setting of (i) of our main theorem (so when SSGS\to S_{G} is trivial over a genus 1 subsurface) the image of Diff+(S)G\operatorname{Diff}^{+}(S)^{G} in U(H1(S;)[χ])\operatorname{U}(\operatorname{H}_{1}(S;{\mathbb{Q}})[\chi]) is almost always arithmetic (see Remark 5.6).

Here is a brief description of the structure of the paper. Of the four sections, only the last one is topological, but in order to put the constructions given there to work, we need a considerable amount of algebra and that explains the nature of the preceding sections.

Section 2 collects useful (and essentially known) algebraic proprieties of constructs that we encounter in the symplectic representation theory of a finite group over {\mathbb{Q}}. So there is little or no claim of originality here, although it was (for us) a bit of an effort to extract this material from the literature. In Section 3 we introduce and study what we might regard as the basic symplectic module associated to an irreducible G{\mathbb{Q}}G-module, where GG is a finite group. The main result is Proposition 3.1 which states an arithmeticity property and also lists the (few) cases for which this arithmetic group has real rank 1\leq 1. This prepares us for stating and proving the arithmeticity criterion Theorem 4.2 in Section 4, which furnishes the main algebraic input for Section 5. As mentioned, this last section is essentially topological: we there construct sufficiently many GG-equivariant mapping classes to ensure that we can apply the said theorem to obtain our main Theorem 1.1.

At various stages of this work I benefited from correspondence with colleagues on this material. These include many enlightening exchanges with Marco Boggi, who in 2019 provided me with a number of helpful comments on earlier drafts of the present paper. Tyakal Venkataramana explained to me in 2018 some of the implications of [9] and [7]. Justin Malestein helped me to better understand parts of [1]. I thank all of them. I am also very grateful to a number of referees for their sometimes extensive comments on an earlier version. The paper greatly improved as a result.

2. Brief review of special unitary groups

The Albert classification

In this subsection DD is a skew field of finite dimension over {\mathbb{Q}} endowed with an anti-involution \dagger. We assume that the involution \dagger is positive in the sense that λDtrD/(λλ)\lambda\in D\mapsto\operatorname{tr}_{D/{\mathbb{Q}}}(\lambda\lambda^{\dagger}) is a positive definite form. We remark that this is so for the cases that matter here, for the given anti-involution on G{\mathbb{Q}}G is evidently positive: for rGr\in{\mathbb{Q}}G, the {\mathbb{Q}}-trace of rrrr^{\dagger} is |G||G| times the coefficient of e1e_{1} in rrrr^{\dagger} and hence is positive definite. The same is then true for its Wedderburn factors EndDχ(Vχ)\operatorname{End}^{D_{\chi}}(V_{\chi}) and their associated skew fields DχD_{\chi}.

We denote the center of DD by LL (so that is a number field) and by D+D_{+} resp. KK the \dagger-invariant part of DD resp. LL. Albert’s classification of such pairs (D,)(D,\dagger) (see for example [5], Ch. IV, Thm. 2) then tells us that KK is totally real, so that D=σσD{\mathbb{R}}\otimes_{\mathbb{Q}}D=\prod_{\sigma}{\mathbb{R}}\otimes_{\sigma}D, where σ\sigma runs over the distinct field embeddings σ:K\sigma:K\hookrightarrow{\mathbb{R}}, and that there are essentially four cases:

  1. (I)

    D=L=KD=L=K so that σD={\mathbb{R}}\otimes_{\sigma}D={\mathbb{R}} for each σ\sigma,

  2. (II)

    L=KL=K and for each σ\sigma there exists an isomorphism σDEnd(2){\mathbb{R}}\otimes_{\sigma}D\cong\operatorname{End}_{\mathbb{R}}({\mathbb{R}}^{2}) which sends \dagger to taking the transpose (so [D:L]=4[D:L]=4),

  3. (III)

    L=KL=K and for each σ\sigma there exists an isomorphism σD𝕂{\mathbb{R}}\otimes_{\sigma}D\cong{\mathbb{K}}, where 𝕂{\mathbb{K}} denotes the Hamilton quaternions, which sends \dagger to quaternion conjugation (so [D:L]=4[D:L]=4),

  4. (IV)

    LL is a purely imaginary extension of KK (in other words, LL is a CM field) and for each σ\sigma there exists an isomorphism σDEnd(d){\mathbb{R}}\otimes_{\sigma}D\cong\operatorname{End}_{\mathbb{C}}({\mathbb{C}}^{d}), which takes σL{\mathbb{R}}\otimes_{\sigma}L to {\mathbb{C}} (so [D:L]=d2[D:L]=d^{2}) and sends \dagger to taking the conjugate transpose.

Let MM be a left DD-module of finite rank. We write MM^{\dagger} for MM endowed with the structure of a right DD-module via the rule aλ:=λaa\lambda:=\lambda^{\dagger}a (aMa\in M, λD\lambda\in D). So if MM^{\prime} is another left DD-module, then MDMM^{\dagger}\otimes_{D}M^{\prime} is defined. It is a KK-vector space with the property that aDλa=(λa)Daa\otimes_{D}\lambda a^{\prime}=(\lambda^{\dagger}a)\otimes_{D}a^{\prime} for all λD\lambda\in D, aMa\in M and aMa^{\prime}\in M^{\prime}. In particular, we have in MDMM^{\dagger}\otimes_{D}M a KK-linear involution defined by (aDb)=bDa(a\otimes_{D}b)^{\prime}=b\otimes_{D}a. We denote its fixed point set u(M)MDMu(M)\subset M^{\dagger}\otimes_{D}M. As a {\mathbb{Q}}-subspace of MDMM^{\dagger}\otimes_{D}M it is spanned by the symmetric tensors aDaa\otimes_{D}a.

Isotropic transvections and Eichler transformations

Suppose given a skew-hermitian form (a,b)M×Ma,bD(a,b)\in M\times M\mapsto\langle a,b\rangle\in D on MM. We denote its radical by MoM_{o} so that the form descends to a nondegenerate one on M¯:=M/Mo\overline{M}:=M/M_{o}. We define the associated unitary group U(M)\operatorname{U}(M) as the group of DD-linear automorphisms of MM that preserve the form and act as the identity on MoM_{o}. It is the group of KK-points of an algebraic group defined over KK. If the form is nondegenerate (Mo=0M_{o}=0), then U(M)\operatorname{U}(M) is what is called in [2] (§5.2B) a classical unitary group.

If cMc\in M is isotropic (meaning that c,c=0\langle c,c\rangle=0), then we have the associated isotropic transvection Tc=T(cDc)U(M)T_{c}=T(c\otimes_{D}c)\in\operatorname{U}(M) defined by xMx+x,ccx\in M\mapsto x+\langle x,c\rangle c (so cDcc\otimes_{D}c is here understood as an element of MDMM^{\dagger}\otimes_{D}M). It ‘generates’ an abelian unipotent subgroup of U(M)\operatorname{U}(M) defined by

λD+T(cDλc)U(M),T(cDλc)(x)=x+x,λcc.\lambda\in D_{+}\mapsto T(c\otimes_{D}\lambda c)\in\operatorname{U}(M)\;,\;T(c\otimes_{D}\lambda c)(x)=x+\langle x,\lambda c\rangle c.

Isotropic transvections are particular cases of Eichler transformations. These are defined as follows. Let cMc\in M be isotropic, aMa\in M perpendicular to cc and λD\lambda\in D such that λλ=a,a\lambda-\lambda^{\dagger}=\langle a,a\rangle (equivalently, λ12a,aD+\lambda-\frac{1}{2}\langle a,a\rangle\in D_{+}). Then the associated Eichler transformation is

E(c,a,λ):xMx+x,ac+x,ca+x,cλcM.E(c,a,\lambda):x\in M\mapsto x+\langle x,a\rangle c+\langle x,c\rangle a+\langle x,c\rangle\lambda c\in M.

It is a DD-linear transformation which preserves the form. When λ=12a,a\lambda=\frac{1}{2}\langle a,a\rangle, we shall write E(c,a)E(c,a) instead. Since T(cDc)=E(12c,c)T(c\otimes_{D}c)=E(\frac{1}{2}c,c), isotropic transvections are Eichler transformations as asserted.

One checks that each Eichler transformation lies in U(M)\operatorname{U}(M) as defined above. In fact, tKE(tc,a,λ)=E(c,ta,t2λ)t\in K\mapsto E(tc,a,\lambda)=E(c,ta,t^{2}\lambda) is a closed one-parameter subgroup of U(M)\operatorname{U}(M) whose infinitesimal generator is represented by aDc+cDau(M)a\otimes_{D}c+c\otimes_{D}a\in u(M) (or rather by its image in u(M)/u(Mo)u(M)/u(M_{o}), for if both aa and cc lie in MoM_{o}, then we get the identity). By a general property of algebraic groups ([8], Cor. 2.2.7) such subgroups then generate a closed algebraic KK-subgroup of U(M)\operatorname{U}(M). Following [2] we denote that group by EU(M)\operatorname{EU}(M).

We note the commutator identity

(1) [E(c,a1,λ1),E(c,a2,λ2)]=T(cλc), with λ=a1,a2+a1,a2.[E(c,a_{1},\lambda_{1}),E(c,a_{2},\lambda_{2})]=T(c\otimes\lambda c),\text{ with }\lambda=\langle a_{1},a_{2}\rangle+\langle a_{1},a_{2}\rangle^{\dagger}.

It follows that if we fix cc, but let aa and λ\lambda vary (subject to the conditions above, so with aca\in c^{\perp}), then the E(c,a,λ)E(c,a,\lambda) generate a unipotent group that appears as an extension of the vector group c/(Mo+Dc)c^{\perp}/(M_{o}+Dc) by the abelian subgroup of U(M)\operatorname{U}(M) defined by the T(cDλc)T(c\otimes_{D}\lambda c).

The group EU(M)\operatorname{EU}(M) is already generated by the isotropic transvections: When \dagger is nontrivial this follows from (6.3.1) of [2]. The remaining case is the one we labelled (I): this is when D=L=KD=L=K and U(M)\operatorname{U}(M) is a symplectic group over KK, but then there is no issue for then every aMa\in M is isotropic, and then E(c,a)=Ta+cTa1Tc1E(c,a)=T_{a+c}T_{a}^{-1}T_{c}^{-1}.

Unipotent radical and Levi quotient

If the form is nondegenerate (ie, Mo={0}M_{o}=\{0\}), then EU(M)\operatorname{EU}(M) is a KK-form of a classical semisimple algebraic group and hence has finite center. To be precise, it is a group of symplectic type in the cases (I) and (II), of orthogonal type in case (III) and of special linear type in case (IV).

If MoM_{o} is possibly nonzero, then per convention the elements of U(M)\operatorname{U}(M) act trivially on MoM_{o}. The natural map U(M)U(M¯)\operatorname{U}(M)\to\operatorname{U}(\overline{M}) is evidently onto. Its kernel consists of the transformations that act trivially on both MoM_{o} and M/MoM/M_{o} and is therefore the unipotent radical Ru(U(M))\operatorname{R_{u}}(\operatorname{U}(M)) of U(M)\operatorname{U}(M) (recall that U(M¯)\operatorname{U}(\overline{M}) is reductive): we have an exact sequence

1Ru(U(M))U(M)U(M¯)11\to\operatorname{R_{u}}(\operatorname{U}(M))\to\operatorname{U}(M)\to\operatorname{U}(\overline{M})\to 1

The elements of Ru(U(M))\operatorname{R_{u}}(\operatorname{U}(M)) are the Eichler transformations E(c,a)E(c,a) with cMoc\in M_{o} and aMa\in M arbitrary. In this case E(c,a)E(c,a) only depends on the image of cac\otimes_{\mathbb{Z}}a in MoDM¯)M_{o}^{\dagger}\otimes_{D}\overline{M}), so that the resulting map

M0DM¯Ru(U(M))M_{0}^{\dagger}\otimes_{D}\overline{M}\to\operatorname{R_{u}}(\operatorname{U}(M))

is an isomorphism. So Ru(U(M))\operatorname{R_{u}}(\operatorname{U}(M)) is a vector group over KK (ie, a KK-vector space regarded as the group of KK-points of a KK-algebraic group of additive type). Since EU(M¯)\operatorname{EU}(\overline{M}) is a normal semisimple subgroup of U(M)\operatorname{U}(M), it has the same unipotent radical: Ru(EU(M))=Ru(U(M))\operatorname{R_{u}}(\operatorname{EU}(M))=\operatorname{R_{u}}(\operatorname{U}(M)).

The relation between EU(M)\operatorname{EU}(M) and U(M)\operatorname{U}(M)

Since in what follows the notion of real rank of an algebraic group shows up, let us begin with reviewing this concept briefly.

Let 𝒢\mathscr{G} be a reductive algebraic group. Suppose first that 𝒢\mathscr{G} is defined over {\mathbb{R}}. Then the real rank rk(𝒢)\operatorname{rk}_{\mathbb{R}}(\mathscr{G}) of 𝒢\mathscr{G} is by definition the dimension of a Cartan subgroup of 𝒢\mathscr{G} defined over {\mathbb{R}}. For example, if 𝒢\mathscr{G} is the orthogonal group of a nondegenerate quadratic form over {\mathbb{R}}, then its real rank is the Witt index of this form: the dimension of a maximal isotropic subspace defined over {\mathbb{R}}. If 𝒢\mathscr{G} is defined over a number field kk, then we restrict scalars à la Weil so that Resk|𝒢\operatorname{Res}_{k|{\mathbb{Q}}}\mathscr{G} is a group defined over {\mathbb{Q}}. We then regard Resk|𝒢\operatorname{Res}_{k|{\mathbb{Q}}}\mathscr{G} (by base change) as a group over {\mathbb{R}} and define the real rank of 𝒢\mathscr{G} to be the real rank of the latter. Concretely, if σ1,,σr\sigma_{1},\dots,\sigma_{r} are the real embeddings of kk in {\mathbb{R}} and τ1,τ¯1,,τs,τ¯s\tau_{1},\overline{\tau}_{1},\dots,\tau_{s},\overline{\tau}_{s} are the remaining distinct (complex) embeddings (they come in complex conjugate pairs), then the definition comes down to

rk(𝒢)=i=1rrk(𝒢σi)+i=1srk(𝒢τi)\textstyle\operatorname{rk}_{\mathbb{R}}(\mathscr{G})=\sum_{i=1}^{r}\operatorname{rk}_{\mathbb{R}}(\mathscr{G}_{\sigma_{i}})+\sum_{i=1}^{s}\operatorname{rk}_{\mathbb{C}}(\mathscr{G}_{\tau_{i}})

(this is also the sum over all the archimedean valuations of kk, taking as general term the real rank of the corresponding completion of 𝒢(k)\mathscr{G}(k)). The Dirichlet unit theorem often gives lower bounds for the rank. For example, if the skew field DD is as in the Albert classification, the group of units D×D^{\times} is a reductive group defined over KK and its group of real points and its real rank are then as follows: putting e:=[K:]e:=[K:{\mathbb{Q}}], then

  1. (I)

    ResK|D×()\operatorname{Res}_{K|{\mathbb{Q}}}D^{\times}({\mathbb{R}}) is open in (×)e({\mathbb{R}}^{\times})^{e}; the real rank of D×D^{\times} is ee,

  2. (II)

    ResK|D×()\operatorname{Res}_{K|{\mathbb{Q}}}D^{\times}({\mathbb{R}}) is open in GL2()e\operatorname{GL}_{2}({\mathbb{R}})^{e}; the real rank of D×D^{\times} is 2e2e,

  3. (III)

    ResK|D×()\operatorname{Res}_{K|{\mathbb{Q}}}D^{\times}({\mathbb{R}}) is open in (𝕂×)e({\mathbb{K}}^{\times})^{e}; the real rank of D×D^{\times} is ee,

  4. (IV)

    ResK|D×()\operatorname{Res}_{K|{\mathbb{Q}}}D^{\times}({\mathbb{R}}) is open in GLd()e\operatorname{GL}_{d}({\mathbb{C}})^{e}; the real rank of D×D^{\times} is dede.

So D×D^{\times} has real rank 2\geq 2, unless DD equals {\mathbb{Q}} (I) or is a definite quaternion algebra with center {\mathbb{Q}} (III) or is an imaginary quadratic extension of {\mathbb{Q}}.

It is clear that EU(M)\operatorname{EU}(M) is a normal subgroup of U(M)\operatorname{U}(M). We already noticed that it is closed in U(M)\operatorname{U}(M) and hence the quotient U(M)/EU(M)\operatorname{U}(M)/\operatorname{EU}(M) is also an algebraic group. Note however that if M¯\overline{M} has no nonzero isotropic vectors, then EU(M¯)\operatorname{EU}(\overline{M}) is trivial. We mention for future reference a consequence of a theorem of G.E. Wall [10]:

Lemma 2.1.

Suppose that MM is nondegenerate and contains a nonzero isotropic vector. Then U(M)/EU(M)\operatorname{U}(M)/\operatorname{EU}(M) is anisotropic (all its real forms are compact). So EU(M)\operatorname{EU}(M) and U(M)\operatorname{U}(M) have the same real rank and any arithmetic subgroup of U(M)\operatorname{U}(M) will have finite image in U(M)/EU(M)\operatorname{U}(M)/\operatorname{EU}(M).

Proof.

Thm. 1 of [10] identifies U(M)/EU(M)\operatorname{U}(M)/\operatorname{EU}(M) as a quotient of D×D^{\times} by a normal subgroup which contains D×D+D^{\times}\cap D_{+}. It is easy to check that in all four cases (I)-(IV) in the Albert classification such a quotient must be anisotropic. ∎

Wall’s result is more specific and tells us that U(M)/EU(M)\operatorname{U}(M)/\operatorname{EU}(M) is often an anisotropic torus. But this need not be so when dimDM=2\operatorname{dim}_{D}M=2.

Lemma 2.2.

If MM is nondegenerate isotropic, then the KK-algebraic group EU(M)\operatorname{EU}(M) is almost simple (by which we mean that EU(M)\operatorname{EU}(M) is perfect and every proper normal subgroup is contained in its center) unless D=KD=K and MK4M\cong K^{4} is endowed with a nondegenerate symmetric form which admits an isotropic plane defined over KK.

Proof.

This follows from [2], Thm. 6.3.16 combined with Thm.  6.3.15. ∎

The excepted case is genuine, for in that case MM1KM2M\cong M_{1}\otimes_{K}M_{2} as modules endowed with KK-forms, where MiM_{i} is a 2-dimensional KK-vector space endowed with a nondegenerate symplectic form. The resulting map SL(M1)×SL(M2)GL(M)\operatorname{SL}(M_{1})\times\operatorname{SL}(M_{2})\to\operatorname{GL}(M) has image EU(M)O(M)\operatorname{EU}(M)\cong\operatorname{O}(M) and its kernel has (1,1)(-1,-1) as its unique nonidentity element.

Remark 2.3.

The reduced norm is the homomorphism N:U(M)L×N:\operatorname{U}(M)\to L^{\times} characterized by the following property: if TU(M)T\in\operatorname{U}(M), then for some (or equivalently, every) real embedding σ:K\sigma:K\hookrightarrow{\mathbb{R}}, the DD-linear TT induces a linear transformation of the σL{\mathbb{R}}\otimes_{\sigma}L-vector space KM{\mathbb{R}}\otimes_{K}M whose determinant is 1σN(T)1\otimes_{\sigma}N(T). The kernel of NN, usually denoted SU(M)\operatorname{SU}(M), contains EU(M)\operatorname{EU}(M) and is often equal to it. But in our context this group does not show up in a natural manner.

3. The hyperbolic module attached to a finite group

A hermitian extension

Our discussion of symplectic G{\mathbb{Q}}G-modules also applies to orthogonal G{\mathbb{Q}}G-modules. One such module is G{\mathbb{Q}}G itself (regarded as a left module): it comes indeed with GG-invariant pairing G×G{\mathbb{Q}}G\times{\mathbb{Q}}G\to{\mathbb{Q}}, the trace form, which assigns to the pair (r1,r2)(r_{1},r_{2}) the trace of r1r2r_{1}r_{2}^{\dagger} considered as an endomorphism of G{\mathbb{Q}}G as a {\mathbb{Q}}-vector space (this is simply |G||G| times the coefficient of e1e_{1}). This pairing is symmetric and nondegenerate.

The Wedderburn decomposition GχEndDχ(Vχ){\mathbb{Q}}G\cong\prod_{\chi}\operatorname{End}^{D_{\chi}}(V_{\chi}) is also the isotypical decomposition, for the KχK_{\chi}-linear map

VχDχHomDχ(Vχ,Dχ)EndDχ(Vχ),v1Dχf:vVχv1f(v)V_{\chi}\otimes_{D_{\chi}}\operatorname{Hom}^{D_{\chi}}(V_{\chi},D_{\chi})\to\operatorname{End}^{D_{\chi}}(V_{\chi}),\quad v_{1}\otimes_{D_{\chi}}f:v\in V_{\chi}\mapsto v_{1}f(v)

which assigns to v1Dχfv_{1}\otimes_{D_{\chi}}f the endomorphism vVχv1f(v)v\in V_{\chi}\mapsto v_{1}f(v) is well-defined and is an isomorphism of left G{\mathbb{Q}}G-modules (and also as right G{\mathbb{Q}}G-modules). This also shows that G[χ]HomDχ(Vχ,Dχ){\mathbb{Q}}G[\chi]\cong\operatorname{Hom}^{D_{\chi}}(V_{\chi},D_{\chi}) as a right DχD_{\chi}-module.

We claim that HomDχ(Vχ,Dχ)Vχ\operatorname{Hom}^{D_{\chi}}(V_{\chi},D_{\chi})\cong V_{\chi}^{\dagger} as left DχD_{\chi}-modules. This is based on a hermitian extension of sχs_{\chi} to VχV_{\chi}^{\dagger}: if we follow the same recipe as in the introduction for a symplectic representation, then we find that there is a hermitian form hχ:Vχ×VχDχh_{\chi}:V_{\chi}^{\dagger}\times V_{\chi}^{\dagger}\to D_{\chi} characterized by the property that for all v,vVχv,v^{\prime}\in V_{\chi} and f,fVχf,f^{\prime}\in V_{\chi}^{\dagger},

sχ(v,f)sχ(v,f)=sχ(vhχ(f,f),v).s_{\chi}(v,f^{\prime})s_{\chi}(v^{\prime},f)=s_{\chi}(vh_{\chi}(f,f^{\prime}),v^{\prime}).

This formula implies that hχh_{\chi} is GG-invariant (we let GG act on the right of VχV_{\chi}^{\dagger}). By taking v=f=fv=f=f^{\prime}, we also see that hχ(f,f)=sχ(f,f)h_{\chi}(f,f)=s_{\chi}(f,f), so that hχh_{\chi} is a hermitian extension of sχs_{\chi}. For every vVχv\in V_{\chi}^{\dagger}, the expression hχ(v,)h_{\chi}(v,\;-) yields an element of HomDχ(Vχ,Dχ)\operatorname{Hom}^{D_{\chi}}(V_{\chi},D_{\chi}) and defines the stated isomorphism.

Isotropic transvections

Here and in the rest of this paper, we write RR for the integral group ring G{\mathbb{Z}}G. Let MM be a finitely generated (left) RR-module, free over {\mathbb{Z}} and let (a,b)M×Mab(a,b)\in M\times M\mapsto a\cdot b\in{\mathbb{Z}} be a nondegenerate (but not necessarily unimodular) GG-equivariant symplectic form. We extend this in the standard manner to a form

(a,b)M×Ma,b:=gG(g1ab)eg=gG(agb)egR.\textstyle(a,b)\in M\times M\mapsto\langle a,b\rangle:=\sum_{g\in G}(g^{-1}a\cdot b)e_{g}=\sum_{g\in G}(a\cdot gb)e_{g}\in R.

This form is skew-hermitian: it is RR-linear in the first variable and a,b=b,a\langle a,b\rangle=-\langle b,a\rangle^{\dagger}. A {\mathbb{Z}}-linear automorphism of MM is GG-equivariant and preserves the symplectic form if and only if it is an RR-module automorphism which preserves this skew-hermitian form. We denote the group of such automorphisms by U(M)\operatorname{U}(M).

Let R+R_{+} stand for the fixed point set of \dagger in RR; this is an additive subgroup of RR. If aMa\in M is RR-isotropic in the sense that a,a=0\langle a,a\rangle=0, then for every rR+r\in R_{+} the isotropic transvection

(2) Ta(r):xMx+x,araMT_{a}(r):x\in M\mapsto x+\langle x,a\rangle ra\in M

lies in U(M)U(M) and rR+Ta(r)r\in R_{+}\mapsto T_{a}(r) is a homomorphism from (the additively written) R+R_{+} to (the multiplicatively written) U(M)U(M). Since Ta(r)T_{a}(r) only depends on araMRMa\otimes ra\in M^{\dagger}\otimes_{R}M, we also denote this transformation by T(ara)T(a\otimes ra).

The basic hyperbolic module

Let AA be a (not necessarily commutative) unital ring with unit endowed with an anti-involution \dagger. The basic hyperbolic AA-module 2(A)\mathcal{H}^{2}(A) is the free left AA-module of rank 22 (whose generators we denote 𝐞\mathbf{e} and 𝐟\mathbf{f}) endowed with the skew-hermitian form defined by 𝐞,𝐞=𝐟,𝐟=0\langle\mathbf{e},\mathbf{e}\rangle=\langle\mathbf{f},\mathbf{f}\rangle=0 and 𝐞,𝐟=1\langle\mathbf{e},\mathbf{f}\rangle=1. It can be regarded as the AA-form of the standard symplectic module 2{\mathbb{Z}}^{2}. In vector notation:

(3) (aa′′),(bb′′)=ab′′a′′b.\Big{\langle}\binom{a^{\prime}}{a^{\prime\prime}},\binom{b^{\prime}}{b^{\prime\prime}}\Big{\rangle}=a^{\prime}b^{\prime\prime}{}^{\dagger}-a^{\prime\prime}b^{\prime}{}^{\dagger}.

It is unimodular in the sense that a2(A),aHomA(2(A),A)a\in\mathcal{H}^{2}(A)\mapsto\langle-,a\rangle\in\operatorname{Hom}_{A}(\mathcal{H}^{2}(A),A) is an antilinear isomorphism. We will write U2(A)\operatorname{U}_{2}(A) resp.  EU2(A)\operatorname{EU}_{2}(A) for U(2(A))\operatorname{U}(\mathcal{H}^{2}(A)) resp. EU(2(A))\operatorname{EU}(\mathcal{H}^{2}(A)). The latter contains SL2()\operatorname{SL}_{2}({\mathbb{Z}}) in an obvious manner. Let A+AA_{+}\subset A be the set of \dagger-invariant elements. One verifies that T𝐞:x2(A)x+x,𝐞𝐞T_{\mathbf{e}}:x\in\mathcal{H}^{2}(A)\mapsto x+\langle x,\mathbf{e}\rangle\mathbf{e} and the similarly defined T𝐟T_{\mathbf{f}} have the matrix form

T𝐞(a)=(1ρa01) resp. T𝐟(a)=(11ρa1),T_{\mathbf{e}}(a)=(\begin{smallmatrix}1&-\rho_{a}\\ 0&1\\ \end{smallmatrix})\text{ resp.\ }T_{\mathbf{f}}(a)=(\begin{smallmatrix}1&1\\ \rho_{a}&1\\ \end{smallmatrix}),

where ρa\rho_{a} stands for right multiplication with aa.

The group Γ(G)\Gamma(G)

We take A=R(=G)A=R(={\mathbb{Z}}G). The elements of the form r+rr+r^{\dagger} with rRr\in R, make up a subgroup R++R_{++} of R+R_{+} such that R+/R++R_{+}/R_{++} is a finite dimensional 𝔽2{\mathbb{F}}_{2}-vector space. Let Γ+(G)\Gamma_{+}(G) resp.  Γ(G)\Gamma_{-}(G) denote the subgroup of EU2(R)\operatorname{EU}_{2}(R) generated by T𝐞(R++)T_{\mathbf{e}}(R_{++}) resp. T𝐟(R++)T_{\mathbf{f}}(R_{++}) and let Γo(G)EU2(R)\Gamma_{o}(G)\subset\operatorname{EU}_{2}(R) stand for the subgroup generated by Γ+(G)\Gamma_{+}(G) and SL2()\operatorname{SL}_{2}({\mathbb{Z}}). Since Γ(G)\Gamma_{-}(G) is a SL2()\operatorname{SL}_{2}({\mathbb{Z}})-conjugate of Γ+(G)\Gamma_{+}(G), the group Γo(G)\Gamma_{o}(G) contains T𝐟(R++)T_{\mathbf{f}}(R_{++}). The right (inverse) action of GG on 2(R)\mathcal{H}^{2}(R) defines an embedding of GG in U2(R)\operatorname{U}_{2}(R). One checks that

ρgT𝐞(r)ρg=T𝐞(grg1)\rho_{g}T_{\mathbf{e}}(r)\rho_{g^{\dagger}}=T_{\mathbf{e}}(grg^{-1})

and so GG normalizes Γo(G)\Gamma_{o}(G). We put Γ(G):=Γo(G).G\Gamma(G):=\Gamma_{o}(G).G.

Arithmetic nature of Γ(G)\Gamma(G)

The notion of a basic hyperbolic module generalizes in a straightforward manner to 2(Vχ)\mathcal{H}^{2}(V_{\chi}^{\dagger}), the skew-hermitian form being given by

(4) (vv′′),(ww′′)=hχ(v,w′′)hχ(v′′,w).\Big{\langle}\binom{v^{\prime}}{v^{\prime\prime}},\binom{w^{\prime}}{w^{\prime\prime}}\Big{\rangle}=h_{\chi}(v^{\prime},w^{\prime\prime})-h_{\chi}(v^{\prime\prime},w^{\prime}).

So we have defined U2(Vχ)\operatorname{U}_{2}(V_{\chi}^{\dagger}); it is the group of KχK_{\chi}-points of a reductive algebraic group defined over KχK_{\chi}. If we write an element of U2(Vχ)\operatorname{U}_{2}(V_{\chi}^{\dagger}) in block form (ABCD)(\begin{smallmatrix}A&B\\ C&D\\ \end{smallmatrix}), with A,B,C,DA,B,C,D in EndDχ(Vχ)\operatorname{End}_{D_{\chi}}(V_{\chi}^{\dagger}), then the subgroup defined by C=0C=0 is a parabolic subgroup. Its unipotent radical is given by requiring that in addition AA and DD are the identity. The corresponding subgroup is then the vector group (1B01)(\begin{smallmatrix}1&B\\ 0&1\\ \end{smallmatrix}) for which BB is hermitian relative to hχh_{\chi}. In other words, hχh_{\chi} identifies BB with an element of u(Vχ)u(V_{\chi}^{\dagger}) (that is, a symmetric element of VχDχVχV_{\chi}\otimes_{D_{\chi}}V_{\chi}^{\dagger}) and hence defines an isotropic transvection. In particular, this is also the unipotent radical of the corresponding subgroup of EU2(Vχ)\operatorname{EU}_{2}(V_{\chi}^{\dagger}). An opposite parabolic subgroup is defined by B=0B=0 and has a similar description of its unipotent radical. Let us denote these unipotent radicals +(U2(Vχ)){\mathscr{R}}_{+}(\operatorname{U}_{2}(V_{\chi}^{\dagger})) resp.  (U2(Vχ)){\mathscr{R}}_{-}(\operatorname{U}_{2}(V_{\chi}^{\dagger})).

We run into this when we consider the isotypical decomposition of 2(G)\mathcal{H}^{2}({\mathbb{Q}}G). The isogeny space 2(G)[χ]=HomG(Vχ,2(G))\mathcal{H}^{2}({\mathbb{Q}}G)[\chi]=\operatorname{Hom}_{{\mathbb{Q}}G}(V_{\chi},\mathcal{H}^{2}({\mathbb{Q}}G)) is then a left DχD_{\chi}-module that is naturally identified with 2(Vχ)\mathcal{H}^{2}(V_{\chi}^{\dagger}). This gives rise to a decomposition

U2(G)=χX(G)U2(Vχ).\operatorname{U}_{2}({\mathbb{Q}}G)=\prod_{\chi\in X({\mathbb{Q}}G)}\operatorname{U}_{2}(V_{\chi}^{\dagger}).

The image of Γ±(G)\Gamma_{\pm}(G) in U2(Vχ)\operatorname{U}_{2}(V_{\chi}^{\dagger}) clearly lands in ±(U2(Vχ)){\mathscr{R}}_{\pm}(\operatorname{U}_{2}(V_{\chi}^{\dagger})). Since EndDχ(Vχ)=EndDχ(Vχ)\operatorname{End}_{D_{\chi}}(V_{\chi}^{\dagger})=\operatorname{End}^{D_{\chi}}(V_{\chi}) is a Wedderburn factor of G{\mathbb{Q}}G, it follows that the image of RR in EndDχ(Vχ)\operatorname{End}_{D_{\chi}}(V_{\chi}^{\dagger}) is an order (a lattice that is also a unital algebra). This is compatible with the anti-involutions and hence the image of R++R_{++} in EndDχ(Vχ)\operatorname{End}_{D_{\chi}}(V_{\chi}^{\dagger}) is a lattice in the subspace of hermitian matrices. So the image of Γ±(G)\Gamma_{\pm}(G) in U2(Vχ)\operatorname{U}_{2}(V_{\chi}^{\dagger}) is a lattice in ±(U2(Vχ)){\mathscr{R}}_{\pm}(\operatorname{U}_{2}(V_{\chi}^{\dagger})).

It is clear that Γo(G)\Gamma_{o}(G) maps to EU2(Vχ)\operatorname{EU}_{2}(V_{\chi}^{\dagger}).

Proposition 3.1.

The image of Γ(G)\Gamma(G) in U2(Vχ)\operatorname{U}_{2}(V_{\chi}^{\dagger}) is an arithmetic subgroup. The real rank of U2(Vχ)\operatorname{U}_{2}(V_{\chi}^{\dagger}) is 2\geq 2 unless dimDχVχ=1\operatorname{dim}_{D_{\chi}}V_{\chi}=1. In that last case, where we can assume that Vχ=DχV^{\dagger}_{\chi}=D_{\chi} with GG acting on the right and mapping to its group of units, one of the following holds:

  1. (i)

    Dχ=D_{\chi}={\mathbb{Q}} and GG maps to μ2\mu_{2},

  2. (iia)

    DχD_{\chi} is the Gaussian field (1){\mathbb{Q}}(\sqrt{-1}) and GG maps onto μ4\mu_{4}, or

  3. (iib)

    DχD_{\chi} is the Eisenstein field (3){\mathbb{Q}}(\sqrt{-3}) and GG maps onto μ3\mu_{3} or μ6\mu_{6}, or

  4. (iiia)

    Dχ+𝐢+𝐣+𝐤D_{\chi}\cong{\mathbb{Q}}+{\mathbb{Q}}\mathbf{i}+{\mathbb{Q}}\mathbf{j}+{\mathbb{Q}}\mathbf{k} and GG maps onto its group of units (a binary tetrahedral group of order 2424) or onto the quaternion group of order 88, or

  5. (iiib)

    Dχ+3𝐢+𝐣+3𝐤D_{\chi}\cong{\mathbb{Q}}+{\mathbb{Q}}\sqrt{3}\mathbf{i}+{\mathbb{Q}}\mathbf{j}+{\mathbb{Q}}\sqrt{3}\mathbf{k} and GG maps onto the binary dihedral group of order 1212.

In all cases, Γ(G)\Gamma(G) acts {\mathbb{Q}}-irreducibly in 2(Vχ)\mathcal{H}^{2}(V_{\chi}^{\dagger}).

Remark 3.2.

It is well-known that the quaternion group appearing in 3.1-(iiia) is realized as the Galois group of a torus ramified at four points (the covering surface has genus 3). This example is like a Swiss army knife for illustrating (and refuting) statements in complex dynamics, which is why that community refers to it as the eierlegende Wollmilchsau. We do not know whether its appearance here is just a coincidence.

For the proof we need:

Theorem 3.3 (Raghunathan [7], Venkataramana [9]).

Let 𝒢{\mathscr{G}} be an almost simple, simply connected {\mathbb{Q}}-algebraic group of real rank 2\geq 2. Let {\mathscr{R}}_{-} and +{\mathscr{R}}_{+} be {\mathbb{Q}}-subgroups that contain the unipotent radicals of opposite {\mathbb{Q}}-parabolic subgroups of 𝒢{\mathscr{G}}. Then for any pair of lattices Γ++()\Gamma_{+}\subset{\mathscr{R}}_{+}({\mathbb{Q}}), Γ()\Gamma_{-}\subset{\mathscr{R}}_{-}({\mathbb{Q}}), the subgroup of 𝒢(){\mathscr{G}}({\mathbb{Q}}) generated by their union Γ+Γ\Gamma_{+}\cup\Gamma_{-} is a congruence subgroup of 𝒢(){\mathscr{G}}({\mathbb{Q}}).

Proof.

We first prove the arithmeticity property of Γo(G)\Gamma_{o}(G) in EU2(Vχ)\operatorname{EU}_{2}(V_{\chi}^{\dagger}). Let us first observe that the group EU2(Vχ)\operatorname{EU}_{2}(V_{\chi}^{\dagger}) is almost-simple by Lemma 2.2. Indeed, this can only fail if Dχ=KχD_{\chi}=K_{\chi} and the form is symmetric (with dimKχVχ=2\operatorname{dim}_{K_{\chi}}V_{\chi}^{\dagger}=2) and this is clearly not the case.

If the real rank of EU2(Vχ)\operatorname{EU}_{2}(V_{\chi}^{\dagger}) is 2\geq 2, then the theorem of Raghunathan-Venkataramana applies and we conclude that Γo(χ)\Gamma_{o}(\chi) is an arithmetic subgroup of EU2(Vχ)\operatorname{EU}_{2}(V_{\chi}^{\dagger}). It is then also easy to see that Γ(G)\Gamma(G) acts {\mathbb{Q}}-irreducibly in 2(Vχ)\mathcal{H}^{2}(V_{\chi}^{\dagger}).

Since the real rank of EU2(Vχ)\operatorname{EU}_{2}(V_{\chi}^{\dagger}) is dimDχVχ\geq\operatorname{dim}_{D_{\chi}}V_{\chi}^{\dagger}, it remains to treat the case when dimDχVχ=1\operatorname{dim}_{D_{\chi}}V_{\chi}^{\dagger}=1. In other words, we can assume that Vχ=DχV_{\chi}^{\dagger}=D_{\chi}. The real rank of EU2(Dχ)\operatorname{EU}_{2}(D_{\chi}) is then still 2\geq 2 most of the time. As we saw above, the exceptions are the cases for which Kχ=K_{\chi}={\mathbb{Q}} and DχD_{\chi} is either {\mathbb{Q}}, a definite quaternion algebra over {\mathbb{Q}} with center {\mathbb{Q}} or an imaginary quadratic extension of {\mathbb{Q}}. Since DχD_{\chi} is also an irreducible G{\mathbb{Q}}G-module, we have a homomorphism ρ:GDχ×\rho:G\to D_{\chi}^{\times} whose image contains a {\mathbb{Q}}-basis of DχD_{\chi}. In particular, DD is generated over {\mathbb{Q}} by its units. So if DχD_{\chi} is an imaginary quadratic extension of {\mathbb{Q}}, then DχD_{\chi} is either {\mathbb{Q}}, the Gaussian field or the Eisenstein field. In the definite quaternion case, Dχ×()D_{\chi}^{\times}({\mathbb{R}}) is the group of unit quaternions and hence ρ(G)\rho(G) is one of the subgroups classified by Klein: this group must be binary tetrahedral, binary octahedral, binary icosahedral of binary dihedral group (of order 4n4n). In these cases ρ(G)\rho({\mathbb{Q}}G)\cap{\mathbb{R}} equals resp.  {\mathbb{Q}}, (2){\mathbb{Q}}(\sqrt{2}), (5){\mathbb{Q}}(\sqrt{5}), (cos(π/n)){\mathbb{Q}}(\cos(\pi/n)). Since we want this intersection to be {\mathbb{Q}}, only the two groups listed have that property.

Note that in each of these exceptional cases, Dχ,+=Kχ=D_{\chi,+}=K_{\chi}={\mathbb{Q}}. The isotropic subspaces in 2(Dχ)\mathcal{H}^{2}(D_{\chi}) are defined over {\mathbb{Q}} and hence EU2(Dχ)SL2()\operatorname{EU}_{2}(D_{\chi})\cong\operatorname{SL}_{2}({\mathbb{Q}}). The group Γo(χ)\Gamma_{o}(\chi) is then a copy of SL2()\operatorname{SL}_{2}({\mathbb{Z}}). So Γo(G)\Gamma_{o}(G) is arithmetic in EU2(Vχ)\operatorname{EU}_{2}(V_{\chi}^{\dagger}). In view of Lemma 2.1 this also implies the arithmeticity of Γ(G)\Gamma(G) in U2(Vχ)\operatorname{U}_{2}(V_{\chi}^{\dagger}). The actions of GG (on the right) and SL2()\operatorname{SL}_{2}({\mathbb{Z}}) (on the left) on 2(Vχ)\mathcal{H}^{2}(V^{\dagger}_{\chi}) commute and make 2(Vχ)\mathcal{H}^{2}(V^{\dagger}_{\chi}) an exterior tensor product of irreducible {\mathbb{Q}}-representations: it is the right G{\mathbb{Q}}G-module VχV^{\dagger}_{\chi} tensored with the tautological representation of SL2()\operatorname{SL}_{2}({\mathbb{Z}}) on 2{\mathbb{Q}}^{2} (which is absolutely irreducible). Hence 2(Vχ)\mathcal{H}^{2}(V^{\dagger}_{\chi}) is irreducible as a representation of SL2()×G\operatorname{SL}_{2}({\mathbb{Z}})\times G. This implies that 2(Vχ)\mathcal{H}^{2}(V^{\dagger}_{\chi}) is irreducible as a Γ(G)\Gamma(G)-module. ∎

4. An arithmeticity criterion

In this section we fix a rational character χX(G)\chi\in X({\mathbb{Q}}G). We therefore suppress the subscript χ\chi and write DD for DχD_{\chi} and VV for VχV_{\chi}.

Proposition 3.1 tells us that Γ(G)U2(V)\Gamma(G)\subset\operatorname{U}_{2}(V^{\dagger}) is an arithmetic subgroup which acts {\mathbb{Q}}-irreducibly on 2(V)\mathcal{H}^{2}(V^{\dagger}) and that with a few exceptions, the group U2(V)\operatorname{U}_{2}(V^{\dagger}) is of real rank 2\geq 2.

Eichler transformations revisited

Let MM be (left) DD-module of finite rank endowed with a nondegenerate skew-hermitian form ,:M×MD\langle\;,\,\rangle:M\times M\to D. Given a DD-submodule NMN\subset M, we denote by UM(N)\operatorname{U}_{M}(N) the subgroup of the group of transformations that act trivially on NN^{\perp}. This group preserves NN and acts trivially on its radical No=NNN_{o}=N\cap N^{\perp}. Hence ‘restriction to NN’ defines homomorphism UM(N)U(N)\operatorname{U}_{M}(N)\to\operatorname{U}(N). This homomorphism is easily shown to be onto. Its kernel consists of the unitary transformations that act trivially on N+NN+N^{\perp} and one checks that this is the image of u(No)u(N_{o}) under TT. We saw that the homomorphism U(N)U(N¯)\operatorname{U}(N)\to\operatorname{U}(\overline{N}) is also onto and we identified its kernel with the vector group NoN¯N_{o}^{\dagger}\otimes\overline{N}. So the Levi quotient of UM(N)\operatorname{U}_{M}(N) is U(N¯)\operatorname{U}(\overline{N}) and its unipotent radical Ru(UM(N))\operatorname{R_{u}}(\operatorname{U}_{M}(N)) is an extension of vector groups:

1u(No)𝑇Ru(UM(N))NoDN¯0.1\to u(N_{o})\xrightarrow{T}\operatorname{R_{u}}(\operatorname{U}_{M}(N))\to N_{o}^{\dagger}\otimes_{D}\overline{N}\to 0.

As is clear from the Equation (1), this extension is usually nontrivial. In case NoN_{o} is spanned by a single element cc, then we can write this sequence as

(5) 0cD+c𝑇Ru(UM(N))cN¯0.0\to c\otimes D_{+}c\xrightarrow{T}\operatorname{R_{u}}(\operatorname{U}_{M}(N))\to c\otimes\overline{N}\to 0.

Any element of Ru(UM(N))\operatorname{R_{u}}(\operatorname{U}_{M}(N)) is an Eichler transformation E(c,a,λ)E(c,a,\lambda) whose image in cN¯c\otimes\overline{N} is ca¯c\otimes\overline{a} (where a¯N¯\overline{a}\in\overline{N} is the image of aa). We will often use the following lemma.

Lemma 4.1.

Let ΓUM(N)\Gamma\subset\operatorname{U}_{M}(N) be a discrete subgroup whose image in U(N¯)\operatorname{U}(\overline{N}) is arithmetic and which acts {\mathbb{Q}}-irreducibly in N¯\overline{N}. If ΓRu(UM(N))\Gamma\cap\operatorname{R_{u}}(\operatorname{U}_{M}(N)) contains an Eichler transformation E(c,a,λ)E(c,a,\lambda) with aND+ca\in N\smallsetminus D_{+}c, then ΓUM(N)\Gamma\cap\operatorname{U}_{M}(N) is arithmetic in UM(N)\operatorname{U}_{M}(N).

Proof.

We are given that in the exact sequence of algebraic groups

1Ru(UM(N))UM(N)U(N¯)11\to\operatorname{R_{u}}(\operatorname{U}_{M}(N))\to\operatorname{U}_{M}(N)\to\operatorname{U}(\overline{N})\to 1

the image Γ¯\overline{\Gamma} of Γ\Gamma in U(N¯)\operatorname{U}(\overline{N}) is arithmetic. Hence for Γ\Gamma to be arithmetic, it suffices that ΓRu(EUM(N))\Gamma\cap\operatorname{R_{u}}(\operatorname{EU}_{M}(N)) be a lattice. For this we turn to the exact sequence (5). The Eichler transformation E(c,a,λ)E(c,a,\lambda) has image ca¯c\otimes\overline{a} in cN¯c\otimes\overline{N} and this image is nonzero by assumption. The image of the Γ\Gamma-conjugacy class of E(c,a,λ)E(c,a,\lambda) in cN¯c\otimes\overline{N} is equal to cΓ¯a¯c\otimes\overline{\Gamma}\overline{a}. Since our assumptions also imply that Γ¯\overline{\Gamma} acts {\mathbb{Q}}-irreducibly in N¯\overline{N}, it follows that the image of ΓRu(UM(N))\Gamma\cap\operatorname{R_{u}}(\operatorname{U}_{M}(N)) in cN¯c\otimes\overline{N} is a lattice in cN¯c\otimes\overline{N}.

Next observe that if E(c,a1,λ1)E(c,a_{1},\lambda_{1}) and E(c,a2,λ2)E(c,a_{2},\lambda_{2}) lie in ΓRu(UM(N))\Gamma\cap\operatorname{R_{u}}(\operatorname{U}_{M}(N)), then so does their commutator, which by the identity (1) is T(cλc)T(c\otimes\lambda c) with λ=a1,a2+a1,a2\lambda=\langle a_{1},a_{2}\rangle+\langle a_{1},a_{2}\rangle^{\dagger}. Since the a1,a2\langle a_{1},a_{2}\rangle generate a lattice in DD, it follows that the λ\lambda generate a lattice in D+D_{+}. In other words, the preimage of ΓRu(UM(N))\Gamma\cap\operatorname{R_{u}}(\operatorname{U}_{M}(N)) in cD+cc\otimes D_{+}c is also a lattice. Hence ΓRu(UM(N))\Gamma\cap\operatorname{R_{u}}(\operatorname{U}_{M}(N)) is a lattice. ∎

Hyperbolic submodules

If j:2(V)Mj:\mathcal{H}^{2}(V^{\dagger})\hookrightarrow M is an embedding of hermitian DD-modules, then MM is the orthogonal direct sum of the image of jj and its perp (for 2(V)\mathcal{H}^{2}(V^{\dagger}) is nondegenerate), so that jj gives rise to an injective homomorphism of groups j:U2(V)U(M)j_{*}:\operatorname{U}_{2}(V^{\dagger})\hookrightarrow\operatorname{U}(M). Let us refer to such an embedding as a VV^{\dagger}-hyperbolic summand in MM.

The following criterion for arithmeticity will be central to our argument.

Theorem 4.2.

Let MM be a nondegenerate skew-hermitian DD-module of finite rank and

a:VM,{b:VM}ba:V^{\dagger}\hookrightarrow M,\quad\{b:V^{\dagger}\hookrightarrow M\}_{b\in{\mathscr{B}}}

a collection of DD-linear embeddings (with {\mathscr{B}} finite, nonempty) whose images span MM over DD and are such that for each bb\in{\mathscr{B}} the pair (a,b)(a,b) defines a hyperbolic summand of MM. In case dimDV=1\operatorname{dim}_{D}V^{\dagger}=1 and dimDM>2\operatorname{dim}_{D}M>2, assume in addition that there exist b1,b2b_{1},b_{2}\in{\mathscr{B}} for which b1(V)b_{1}(V^{\dagger}) and b2(V)b_{2}(V^{\dagger}) are perpendicular.

Then the subgroup Γ\Gamma of U(M)\operatorname{U}(M) generated by {(a,b)Γ(G)}b\{(a,b)_{*}\Gamma(G)\}_{b\in{\mathscr{B}}} is an arithmetic subgroup of U(M)\operatorname{U}(M) which acts {\mathbb{Q}}-irreducibly in MM.

The proof will be by induction on dimDM\operatorname{dim}_{D}M. As may be inferred from the statement of the theorem, the case when dimDV=1\operatorname{dim}_{D}V^{\dagger}=1 is a bit more delicate. Indeed, the first induction step then requires special care and so we do that case first. Once we have dealt with it, we indicate how to modify the arguments in order to obtain a proof of the unrestricted version of Theorem 4.2.

Let us say that a DD-subspace NMN\subset M is Γ\Gamma-arithmetic if ΓUM(N)\Gamma\cap\operatorname{U}_{M}(N) is arithmetic in UM(N)\operatorname{U}_{M}(N) and acts {\mathbb{Q}}-irreducibly in N¯\overline{N} (the last property is a consequence of the first if the real rank of UM(N)\operatorname{U}_{M}(N) is 2\geq 2).

The case dimDV=1\operatorname{dim}_{D}V^{\dagger}=1

We then identify VV^{\dagger} with DD and aa and each bb\in{\mathscr{B}} with the image of 1D1\in D under these embeddings so that a,b=1\langle a,b\rangle=1 for all bb\in{\mathscr{B}}. Note that {a}M\{a\}\cup{\mathscr{B}}\subset M consists of isotropic elements and generates MM over DD. We write Γ(a,b)\Gamma(a,b) for the image of Γ(G)\Gamma(G) under (a,b)(a,b), so that Γ\Gamma is generated by {Γ(a,b)}b\{\Gamma(a,b)\}_{b\in{\mathscr{B}}}. As any bb\in{\mathscr{B}} lies Γ(a,b)a\Gamma(a,b)a, it follows that {a}Γa\{a\}\cup{\mathscr{B}}\subset\Gamma a.

By Proposition 3.1, Da+DbDa+Db is Γ\Gamma-arithmetic for every bb\in{\mathscr{B}}. We therefore assume that MM is not of the form Da+DbDa+Db. So there exist b1,b2b_{1},b_{2}\in{\mathscr{B}} with b2Da+Db1b_{2}\notin Da+Db_{1} such that b1,b2=0\langle b_{1},b_{2}\rangle=0.

Lemma 4.3.

Put N:=Da+Db1N:=Da+Db_{1}. Then N:=N+Db2N^{\prime}:=N+Db_{2} is Γ\Gamma-arithmetic.

Proof.

We verify that the assumptions of the Lemma 4.1 are satisfied by ΓUM(N)\Gamma\cap\operatorname{U}_{M}(N^{\prime}). It is clear that the radical of NN^{\prime} is spanned by c:=b2b1c:=b_{2}-b_{1} so that N¯\overline{N}^{\prime} is the isomorphic image of NN. We know that NN is Γ\Gamma-arithmetic and so Γ\Gamma has arithmetic image in U(N¯)\operatorname{U}(\overline{N}^{\prime}). Since Tb1T_{b_{1}} and Tb2T_{b_{2}} lie in Γ\Gamma, so does Tb2Tb11T_{b_{2}}T_{b_{1}}^{-1}. We check that

Tb2Tb11(x)=xx,b1+x,b1+c(b1+c)=E(c,b1,1)(x).T_{b_{2}}T_{b_{1}}^{-1}(x)=x-\langle x,b_{1}\rangle+\langle x,b_{1}+c\rangle(b_{1}+c)=E(c,b_{1},1)(x).

So the image of ΓRu(UM(N))\Gamma\cap\operatorname{R_{u}}(\operatorname{U}_{M}(N^{\prime})) in cN¯c\otimes\overline{N}^{\prime} contains cb1c\otimes b_{1}. Now apply Lemma 4.1. ∎

From this point onward the argument will be inductive. The union of Lemmas 4.4 and 4.5 will establish the theorem in case dimDV=1\operatorname{dim}_{D}V^{\dagger}=1.

Lemma 4.4.

Let NMN\subsetneq M be a DD-subspace which contains a,b1,b2a,b_{1},b_{2} and whose the radical is of DD-dimension one. If NN is Γ\Gamma-arithmetic, then there exists a bb\in{\mathscr{B}} such that N:=N+DbN^{\prime}:=N+Db is nondegenerate and Γ\Gamma-arithmetic, and the real rank of U(N)\operatorname{U}(N^{\prime}) is 2\geq 2.

Proof.

Let cNc\in N span the radical of NN. Since MM is nondegenerate and DD-spanned by {a}\{a\}\cup{\mathscr{B}}, there must exist a bb\in{\mathscr{B}} such that cc is not in the radical of N:=N+DbN^{\prime}:=N+Db. Then it is easily seen that NN^{\prime} is nondegenerate so that UM(N)U(N)\operatorname{U}_{M}(N^{\prime})\cong\operatorname{U}(N^{\prime}). In case N=Da+Db1+Db2N=Da+Db_{1}+Db_{2} (where we can take c=b2b1c=b_{2}-b_{1}) one checks that NN is the perpendicular sum of two copies of 2(D)\mathcal{H}^{2}(D). Otherwise, NN^{\prime} contains such a sum. This implies that U(N)\operatorname{U}(N^{\prime}) has real rank 2\geq 2.

The U(N)\operatorname{U}(N^{\prime})-stabilizer U(N)c\operatorname{U}(N^{\prime})_{c} of cc is equal to UM(N)\operatorname{U}_{M}(N) and hence contains ΓU(N)c\Gamma\cap\operatorname{U}(N^{\prime})_{c} as an arithmetic subgroup. Observe that c:=Tb(c)=c+c,bbc^{\prime}:=T_{b}(c)=c+\langle c,b\rangle b is another isotropic element with c,c=c,bb,c0\langle c^{\prime},c\rangle=\langle c,b\rangle\langle b,c\rangle\not=0 and so Dc+Dc2(D)Dc+Dc^{\prime}\cong\mathcal{H}^{2}(D). The two U(N)\operatorname{U}(N^{\prime})-stabilizers of DcDc and DcDc^{\prime} are opposite parabolic subgroups of U(N)\operatorname{U}(N^{\prime}) whose unipotent radicals are contained in U(N)c\operatorname{U}(N^{\prime})_{c} resp.  U(N)c\operatorname{U}(N^{\prime})_{c^{\prime}}. Since U(N)c\operatorname{U}(N^{\prime})_{c^{\prime}} is a Γ\Gamma-conjugate of U(N)c\operatorname{U}(N^{\prime})_{c}, it follows that ΓU(N)c\Gamma\cap\operatorname{U}(N^{\prime})_{c^{\prime}} is an arithmetic subgroup of U(N)c\operatorname{U}(N^{\prime})_{c^{\prime}}. We have thus satisfied the hypotheses of Theorem 3.3 and we conclude that ΓU(N)\Gamma\cap\operatorname{U}(N^{\prime}) is arithmetic in U(N)\operatorname{U}(N^{\prime}). The fact that ΓU(N)\Gamma\cap\operatorname{U}(N^{\prime}) acts {\mathbb{Q}}-irreducibly in NN^{\prime} follows from the fact that ΓU(N)\Gamma\cap\operatorname{U}(N) has this property in NN, for the ΓU(N)\Gamma\cap\operatorname{U}(N^{\prime})-translates of NN span NN^{\prime} over {\mathbb{Q}}, but do not decompose NN^{\prime}. ∎

Lemma 4.5.

Let NMN\subsetneq M be a proper nondegenerate DD-subspace of dimension 4\geq 4 and contain a,b1,b2a,b_{1},b_{2}. If NN is Γ\Gamma-arithmetic, then so is N:=N+DbN^{\prime}:=N+Db^{\prime} for every bNb^{\prime}\in{\mathscr{B}}\smallsetminus N.

Proof in case NN^{\prime} is degenerate.

We verify that the assumptions of the Lemma 4.1 are satisfied by ΓUM(N)\Gamma\cap\operatorname{U}_{M}(N^{\prime}). The radical of NN^{\prime} is necessary spanned by an element of the form c:=bbc:=b^{\prime}-b where bNb\in N is characterized by the property that x,b=x,b\langle x,b\rangle=\langle x,b^{\prime}\rangle for all xNx\in N. So NN maps then isomorphically onto N/Dc=N¯N^{\prime}/Dc=\overline{N}^{\prime}. In particular, the natural map U(N)U(N¯)\operatorname{U}(N)\to\operatorname{U}(\overline{N}^{\prime}) is an isomorphism and hence ΓUM(N)\Gamma\cap\operatorname{U}_{M}(N^{\prime}) maps onto an arithmetic subgroup of U(N¯)\operatorname{U}(\overline{N}^{\prime}).

Let nn be a positive integer such that TbnΓT_{b}^{n}\in\Gamma. Then

TbnTbn(x)=xnx,bb+nx,bb=x+nx,bc+nx,cb+nx,cc=E(c,nb,n)(x)T_{b^{\prime}}^{n}T^{-n}_{b}(x)=x-n\langle x,b\rangle b+n\langle x,b^{\prime}\rangle b^{\prime}=x+n\langle x,b\rangle c+n\langle x,c\rangle b+n\langle x,c\rangle c=E(c,nb,n)(x)

So E(c,nb,n)ΓRu(UM(N))E(c,nb,n)\in\Gamma\cap\operatorname{R_{u}}(\operatorname{U}_{M}(N^{\prime})) and this element has image cnbc\otimes nb in cN¯c\otimes\overline{N}^{\prime}. It then follows from Lemma 4.1 that NN^{\prime} is Γ\Gamma-arithmetic. ∎

Proof in case NN^{\prime} is nondegenerate.

Then N1:=NbN^{\prime}_{1}:=N^{\prime}\cap b^{\prime}{}^{\perp} is degenerate with radical spanned by bb^{\prime}. We first prove that N1N^{\prime}_{1} is Γ\Gamma-arithmetic by verifying the assumptions of the Lemma 4.1. The subspace N1:=NbN_{1}:=N\cap b^{\prime}{}^{\perp} supplements bb^{\prime} in N1N^{\prime}_{1}. It is therefore nondegenerate and maps isomorphically onto N¯1=N1/Db\overline{N}^{\prime}_{1}=N^{\prime}_{1}/Db^{\prime}. This enables us to regard U(N1)\operatorname{U}(N_{1}) as a subgroup of UM(N1)\operatorname{U}_{M}(N^{\prime}_{1}) that acts trivially on both bb^{\prime} and its orthogonal projection in NN.

Since ΓUM(N)\Gamma\cap\operatorname{U}_{M}(N) is arithmetic in UM(N)\operatorname{U}_{M}(N), its subgroup ΓU(N1)\Gamma\cap\operatorname{U}(N_{1}) is arithmetic in U(N1)\operatorname{U}(N_{1}) and has arithmetic image in U(N¯1)\operatorname{U}(\overline{N}^{\prime}_{1}). We show that E(b,c)ΓE(b^{\prime},c^{\prime})\in\Gamma for some cN1Dbc^{\prime}\in N^{\prime}_{1}\smallsetminus Db^{\prime}. Then Lemma 4.1 will imply that N1N^{\prime}_{1} is Γ\Gamma-arithmetic.

For this we recall that c:=b2b1c:=b_{2}-b_{1} is perpendicular to b1b_{1} and has nonzero image in Db1/Db1Db_{1}^{\perp}/Db_{1}. Let γΓ(a,b)Γ(a,b1)Γ(N)\gamma\in\Gamma(a,b^{\prime})\Gamma(a,b_{1})\subset\Gamma(N^{\prime}) take b1b_{1} to bb^{\prime}. Since cc is perpendicular to b1b_{1}, c:=γ(c)Nc^{\prime}:=\gamma(c)\in N^{\prime} is perpendicular to γ(b1)=b\gamma(b_{1})=b^{\prime} (so lies in N1N^{\prime}_{1}) and has nonzero image [c][c^{\prime}] in N¯1N1\overline{N}^{\prime}_{1}\cong N_{1}. Since E(b,c)=E(c,b)1E(b^{\prime},c^{\prime})=E(c^{\prime},b^{\prime})^{-1} is a Γ\Gamma-conjugate of E(c,b1)1E(c,b_{1})^{-1}, it lies in Γ\Gamma.

Now is U(N)b=U(Nb)\operatorname{U}(N^{\prime})_{b^{\prime}}=\operatorname{U}(N^{\prime}\cap b^{\prime}{}^{\perp}) and so ΓU(N)b\Gamma\cap\operatorname{U}(N^{\prime})_{b^{\prime}} is arithmetic in U(N)b\operatorname{U}(N^{\prime})_{b^{\prime}}. As aΓba\in\Gamma b^{\prime}, the same is true for ΓU(N)a\Gamma\cap\operatorname{U}(N^{\prime})_{a}. Since aa and bb^{\prime} span a copy of 2(D)\mathcal{H}^{2}(D), their U(N)\operatorname{U}(N^{\prime})-stabilizers contain the unipotent radicals of opposite parabolic subgroups of U(N)\operatorname{U}(N^{\prime}). The real rank of U(N)\operatorname{U}(N^{\prime}) is 2\geq 2, so that Theorem 3.3 applies and tells us that NN^{\prime} is Γ\Gamma-arithmetic. ∎

The case when dimDV>1\operatorname{dim}_{D}V^{\dagger}>1

The same scheme for the proof works when dimDV>1\operatorname{dim}_{D}V^{\dagger}>1. The difference is that we deal with larger hyperbolic packets, to wit, the images of hyperbolic embeddings (a,b):2(V)M(a,b):\mathcal{H}^{2}(V^{\dagger})\hookrightarrow M. The essential difference is that we start off in a better position, since we begin with a VV^{\dagger}-hyperbolic embedding j:2(V)Mj:\mathcal{H}^{2}(V^{\dagger})\hookrightarrow M and we already know that its image NN is Γ\Gamma-arithmetic and that UM(N)U(N)\operatorname{U}_{M}(N)\cong\operatorname{U}(N) has real rank 2\geq 2.

5. Finding liftable mapping classes

In this section the GG-covering SSGS\to S_{G} is as in the introduction and ASGA\subset S_{G}^{\circ} is a nonempty closed 11-submanifold such that the covering is trivial over AA and is connected over SAS\smallsetminus A. We also choose a connected component α\alpha of AA, so that A:=AαA^{\prime}:=A\smallsetminus\alpha might be empty. We orient α\alpha and regard it as the oriented image of an embedding of the circle in SGS_{G}^{\circ}. We will see that this gives rise to enough copies of Γ(G)\Gamma(G) in the representation of the GG-equivariant mapping classes as to satisfy the hypotheses of our arithmeticity criterion Theorem 4.2.

We denote by SG(α)S_{G}(\alpha) the singular surface obtained from SGS_{G} by contracting α\alpha to a point (that we will denote by \infty). Its topological normalization is a closed connected surface, denoted S^G(α)\widehat{S}_{G}(\alpha), whose genus is one less than that of SGS_{G}. The surface S^G(α)\widehat{S}_{G}(\alpha) comes with two points over \infty and the orientation of α\alpha enables us to tell them apart: we let pp_{-} be ‘to the left’ of α\alpha and p+p_{+} is ‘to the right’ of α\alpha. If we regard AA^{\prime} also as a submanifold of S^G(α)\widehat{S}_{G}^{\circ}(\alpha), then the surjection S^G(α)SG(α)\widehat{S}_{G}^{\circ}(\alpha)\to S^{\circ}_{G}(\alpha) defines a map from the set Π(S^G(α)A;p,p+)\Pi(\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime};p_{-},p_{+}) of path homotopy classes in S^G(α)A\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime} from pp_{-} to p+p_{+} to the fundamental group π1(SG(α),)\pi_{1}(S^{\circ}_{G}(\alpha),\infty). This map is injective. We do the same (in a GG-equivariant manner) for the preimage of SGαS_{G}\smallsetminus\alpha in SS and thus get GG-covers S(α)SG(α)S(\alpha)\to S_{G}(\alpha), S^(α)S^G(α)\widehat{S}^{\circ}(\alpha)\to\widehat{S}_{G}(\alpha) and GG-orbits PS(α)P_{\infty}\subset S(\alpha)^{\circ}, P±S^(α)P_{\pm}\subset\widehat{S}^{\circ}(\alpha), so that we end up with the diagram below (in which the vertical maps are GG-coverings):

π1α{\pi^{-1}\alpha}P{P_{\infty}}PP+{P_{-}\cup P_{+}}S{S}S(α){S(\alpha)}S^(α){\widehat{S}(\alpha)}α{\alpha}{}{\{\infty\}}{p,p+}{\{p_{-},p_{+}\}}SG{S_{G}}SG(α){S_{G}(\alpha)}S^G(α){\widehat{S}_{G}(\alpha)}

This construction comes with GG-equivariant bijections PPP+P_{-}\cong P_{\infty}\cong P_{+}. Our assumption on the covering SSGS\to S_{G} amounts to the following two properties: (i) S^(α)\widehat{S}(\alpha) is connected and stays so if we remove the preimage of AA^{\prime} and (ii) the three GG-orbits PP_{\infty}, P±P_{\pm} are regular. So the choice of a point in PP_{\infty} (which is equivalent to the choice of a lift α~\tilde{\alpha} of α\alpha) identifies these GG-sets with GG (on which GG acts by left translation). In particular, we thus identify IsoG(P,P+)AutG(P)\operatorname{Iso}_{G}(P_{-},P_{+})\cong\operatorname{Aut}_{G}(P_{\infty}) with GG (where gGg\in G acts on GG by right translation over g1g^{-1}).

Refer to caption

Figure 1. The surface SGS_{G}, its quotient SG(α)S_{G}(\alpha) and the normalization S^G(α)\widehat{S}_{G}(\alpha).

For any path in S^G(α)A\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime} from pp_{-} to p+p_{+}, the GG covering is trivial over it, so that we have an associated GG-bijection PP+P_{-}\cong P_{+}. Since the GG-covering over S^G(α)A\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime} is connected, the resulting map

Π(S^G(α)A;p,p+)IsoG(P,P+)AutG(P)\Pi(\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime};p_{-},p_{+})\to\operatorname{Iso}_{G}(P_{-},P_{+})\cong\operatorname{Aut}_{G}(P_{\infty})

is onto by standard covering theory. We will say that an element of Π(S^G(α)A;p,p+)\Pi(\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime};p_{-},p_{+}) is GG-trivial if its image in AutG(P)\operatorname{Aut}_{G}(P_{\infty}) under the above map is the identity. Such elements make up a coset for the kernel of the natural homomorphism π1(S^G(α)A,p)G\pi_{1}(\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime},p_{-})\to G.

Lemma 5.1.

Every element of Π(S^G(α)A;p,p+)\Pi(\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime};p_{-},p_{+}) is representable by some arc (== embedded unit interval) in S^G(α)A\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime} from pp_{-} to p+p_{+}. We can arrange that this arc lifts to an embedding of the circle /{\mathbb{R}}/{\mathbb{Z}} in SGS^{\circ}_{G} which meets α\alpha in a single point with intersection number 11. In particular, every element of AutG(P)\operatorname{Aut}_{G}(P) is realized by the monodromy along an embedded circle β\beta which does not meet AA^{\prime} and meets α\alpha in one point only and does so transversally with intersection number 11.

Proof.

We first represent the homotopy class by an immersion of the unit interval with only transversal self-intersections. That number of self-intersections is finite and if this number is positive, we lower it by moving the last point of self-intersection towards p+p_{+} and then slide the path over p+p_{+}. By iterating this procedure we obtain a representative which is an embedding. It is clear that we can make this arc lift to an embedded circle. The second assertion then follows. ∎

We choose a lift α~\tilde{\alpha} of α\alpha and write aH1(S)a\in\operatorname{H}_{1}(S) for its homology class. Since RaRa is an an isotropic sublattice of H1(S)\operatorname{H}_{1}(S), we have that a,a=0\langle a,a\rangle=0. We let IH1(S)I\subset\operatorname{H}_{1}(S) be the homology supported by the preimage of AA^{\prime}: this is a free RR-submodule (where as before, R=GR={\mathbb{Z}}G) with a generator for every connected component of AA^{\prime}. It is clear that Ra+IRa+I is isotropic. We saw that the lift α~\tilde{\alpha} identifies AutG(P)\operatorname{Aut}_{G}(P) with the group GG with GG acting on itself by right translations. Lemma 5.1 above shows that all such elements are obtained from a loop of the type described there. From that lemma we also derive:

Corollary 5.2.

Let bIb\in I^{\perp} be such that ab=1a\cdot b=1 and gab=0ga\cdot b=0 for gG{1}g\in G\smallsetminus\{1\}. Then some bb+Rab^{\prime}\in b+Ra can be represented by a lift of an embedded circle β\beta in SGAS^{\circ}_{G}\smallsetminus A^{\prime} which meets α\alpha transversally at a unique point (and for which necessarily αβ=1\alpha\cdot\beta=1). In particular, bb^{\prime} is RR-isotropic and the RR-linear map defined by

2(R)I,𝐞a,𝐟b\mathcal{H}^{2}(R)\to I^{\perp},\quad\mathbf{e}\mapsto a,\;\mathbf{f}\mapsto b^{\prime}

is an embedding of skew-hermitian RR-modules whose orthogonal complement supplements its image: we obtain a basic hyperbolic summand of the RR-module H1(S)\operatorname{H}_{1}(S).

Proof.

It is not difficult to see that the homology class bb is representable by a map from the circle to SS^{\circ} which meets the preimage of α\alpha exactly once (hence in a point of α~\tilde{\alpha}) and does not meet the preimage of AA^{\prime}. We apply Lemma 5.1 to its image in SGAS^{\circ}_{G}\smallsetminus A^{\prime} or rather to the resulting arc in S^G(α)A\widehat{S}_{G}^{\circ}(\alpha)\smallsetminus A^{\prime} which connects pp_{-} with p+p_{+}: this then produces an embedding β\beta of the circle in SGAS^{\circ}_{G}\smallsetminus A^{\prime} which meets α\alpha only once and with intersection number 11 over which the GG-covering is trivial. The lift β~\tilde{\beta} of β\beta which meets α~\tilde{\alpha} defines a homology class bb^{\prime} which differs from by bb by a class supported by the preimage of α\alpha, that is, an element of RaRa.

The proof of the last paragraph is straightforward. ∎

Let us call an ordered pair a,ba,b in H1(S)\operatorname{H}_{1}(S) RR-hyperbolic if a,a=b,b=0\langle a,a\rangle=\langle b,b\rangle=0 and a,b=1\langle a,b\rangle=1. Such a pair defines a basic hyperbolic summand Ra+RbH1(S)Ra+Rb\subset\operatorname{H}_{1}(S) and gives rise to an embedding of Γ(G)\Gamma(G) in Sp(H1(S))G\operatorname{Sp}(\operatorname{H}_{1}(S))^{G}. We shall denote the latter’s image by Γ(a,b)\Gamma(a,b) and the image of Γo(G)\Gamma_{o}(G) by Γo(a,b)\Gamma_{o}(a,b). We write a{\mathscr{B}}_{a} for the set of bH1(S)b\in\operatorname{H}_{1}(S) for which the pair (a,b)(a,b) is RR-hyperbolic and Γo(a)\Gamma_{o}(a) resp.  Γ(a)\Gamma(a) for the subgroup of Sp(H1(S))G\operatorname{Sp}(\operatorname{H}_{1}(S))^{G} generated by its subgroups Γo(a,b)\Gamma_{o}(a,b) resp. Γ(a,b)\Gamma(a,b) with bab\in{\mathscr{B}}_{a}.

Fix a base point pp on α\alpha and write p~\tilde{p} for its preimage in α~\tilde{\alpha}.

Let hGh\in G and regard hh as an element of AutG(P)\operatorname{Aut}_{G}(P). Let β\beta be as in Lemma 5.1, which we may (and will) assume to meet α\alpha in pp such that the lift β~\tilde{\beta} of β\beta (as an arc) which begins in p~\tilde{p} ends in hp~h\tilde{p}. Let SGβS^{\beta}_{G} be a closed regular neighborhood of αβ\alpha\cup\beta in SGAS^{\circ}_{G}\smallsetminus A^{\prime}. This is a compact genus one surface whose boundary SGβ\partial S^{\beta}_{G} is connected. The homotopy class of this boundary (with its natural orientation) is in the free homotopy class of the commutator [β]1[α]1[β][α][\beta]^{-1}[\alpha]^{-1}[\beta][\alpha] (we write path composition functorially, so the order of travel is read from right to left). This commutator has trivial image in GG (since [α][\alpha] has), and so the GG-covering SSGS\to S_{G} is trivial over SGβ\partial S^{\beta}_{G}. The preimage of SGβ\partial S^{\beta}_{G} in SS is the boundary of the preimage SβS^{\beta} of SGβS^{\beta}_{G} and the Dehn twist along SGβ\partial S^{\beta}_{G} lifts in a GG-equivariant manner to a multi-Dehn twist DβD^{\beta} along that boundary. The following lemma generalizes one of the constructions given in [3] for the case when GG is cyclic (in that paper they are depicted as Figs. 2 and 3).

Lemma 5.3.

The multi-Dehn twist DβD^{\beta} acts on H1(S)\operatorname{H}_{1}(S) as Ta(2eheh)T_{a}(2-e_{h}-e^{\dagger}_{h}) (where we use formula (2), noting that 2ehehR+2-e_{h}-e^{\dagger}_{h}\in R_{+}).

Proof.

The lift of the commutator [β]1[α]1[β][α][\beta]^{-1}[\alpha]^{-1}[\beta][\alpha] that passes through p~\tilde{p} first traverses the embedded circle α~\tilde{\alpha}, then traverses β~\tilde{\beta}, then traverses the circle hα~h\tilde{\alpha} in the opposite direction and then returns via the inverse of β~\tilde{\beta} to p~\tilde{p}. So the homology class of this lift of the commutator (and hence of the corresponding lift of SGβ\partial S^{\beta}_{G}) is ahaa-ha. If we replace p~\tilde{p} by gp~g\tilde{p} with gGg\in G, then this replaces aa by gaga and hh by ghg1ghg^{-1}, so that the corresponding class is gagha=g(1h)aga-gha=g(1-h)a. By a standard formula, the resulting action on H1(S)\operatorname{H}_{1}(S) is given by

Dβ(x)=x+gG(xg(1h)a)g(1h)a==x+gG((xga)ga(xgha)ga(xga)gha+(xgha)gha)==x+2x,aax,haax,aha=Ta(2eheh)(x).\textstyle D^{\beta}_{*}(x)=x+\sum_{g\in G}(x\cdot g(1-h)a)g(1-h)a=\\ \textstyle=x+\sum_{g\in G}\Big{(}(x\cdot ga)ga-(x\cdot gha)ga-(x\cdot ga)gha+(x\cdot gha)gha\Big{)}=\\ =x+2\langle x,a\rangle a-\langle x,ha\rangle a-\langle x,a\rangle ha=T_{a}(2-e_{h}-e^{\dagger}_{h})(x).\qed
Proposition 5.4.

Let bH1(S)b\in\operatorname{H}_{1}(S) be such that (a,b)(a,b) is an RR-hyperbolic pair. Then the image of Diff+(S)GSp(H1(S))G\operatorname{Diff}^{+}(S)^{G}\to\operatorname{Sp}(\operatorname{H}_{1}(S))^{G} contains Γ(a,b)\Gamma(a,b).

Proof.

We first show this for Γ0(a,b)\Gamma_{0}(a,b). Let β\beta be as in Corollary 5.2 (and thus represent an element of b+Rab+Ra). The diffeomorphisms of SGS_{G} with support in the interior of SGβS^{\beta}_{G} have as their image in the mapping class group of SGS_{G} a centrally extended copy of SL(2,)\operatorname{SL}(2,{\mathbb{Z}}) with the central subgroup generated by the Dehn twist along the boundary of SGβS^{\beta}_{G}. This Dehn twist acts trivially on H1(SGβ)\operatorname{H}_{1}(S^{\beta}_{G}). These diffeomorphisms lift to diffeomorphisms of SS with support in SβS^{\beta} with the central subgroup acting trivially on H1(S)\operatorname{H}_{1}(S). We thus obtain in the image of Diff+(S)GSp(H1(S))G\operatorname{Diff}^{+}(S)^{G}\to\operatorname{Sp}(\operatorname{H}_{1}(S))^{G} a copy of SL2()\operatorname{SL}_{2}({\mathbb{Z}}).

The multi-Dehn twist associated to α\alpha acts on H1(S)\operatorname{H}_{1}(S) as xx+gG(xga)ga=x+x,aa=Ta(1)(x)x\mapsto x+\sum_{g\in G}(x\cdot ga)ga=x+\langle x,a\rangle a=T_{a}(1)(x). By lemma 5.3 the image of Diff+(S)GSp(H1(S))G\operatorname{Diff}^{+}(S)^{G}\to\operatorname{Sp}(\operatorname{H}_{1}(S))^{G} also contains the transvections Ta(2eheh)T_{a}(2-e_{h}-e_{h}^{\dagger}) for all hGh\in G. Hence that image contains all of Ta(R++)T_{a}(R_{++}). This proves that the image of Diff+(S)GSp(H1(S))G\operatorname{Diff}^{+}(S)^{G}\to\operatorname{Sp}(\operatorname{H}_{1}(S))^{G} contains Γo(a,b)\Gamma_{o}(a,b).

So it remains to show that the image of GG in Γ(a,b)\Gamma(a,b) is realized by Diff+(S)G\operatorname{Diff}^{+}(S)^{G}. For this we use mapping classes of push type. Consider the smooth surface SG/SGβS_{G}/S_{G}^{\beta} that is obtained as a quotient of SGS_{G} by contracting SGβS_{G}^{\beta} to a point (that we shall call qq). If we do the same for the connected components of SβS^{\beta} in GG we get a GG-cover S//SβSG/SGβS/\!\!/S^{\beta}\to S_{G}/S_{G}^{\beta} fitting in the commutative diagram

SS//SβGGSGSG/SGβ\begin{CD}S@>{}>{}>S/\!\!/S^{\beta}\\ @V{G}V{}V@V{G}V{}V\\ S_{G}@>{}>{}>S_{G}/S^{\beta}_{G}\end{CD}

The covering on the right does not branch over qq and so its preimage QQ in S//SβS/\!\!/S^{\beta} is a regular GG-orbit. For every gGg\in G, there is a closed loop γ\gamma of SG/SGβS_{G}/S^{\beta}_{G} based at qq which avoids branch points and induces in QQ right multiplication by gg. The corresponding point-pushing map on SG/SGβS_{G}/S^{\beta}_{G} (chosen to fix branch points) lifts to a GG-equivariant diffeomorphism φ\varphi of S//SβS/\!\!/S^{\beta} that extends this permutation.

Refer to caption

Figure 2. Point pushing (or rather, ‘small circle pushing’) the genus one surface SβS^{\beta} along SG/SGβS_{G}/S^{\beta}_{G}.

Such a point-pushing map is isotopic to the identity on SG/SGβS_{G}/S^{\beta}_{G} and hence the same is true for its lift φ\varphi. In particular, φ\varphi acts trivially on H1(S//Sβ)\operatorname{H}_{1}(S/\!\!/S^{\beta}). It is not difficult to see that the point pushing map and its lift ϕ\phi can be ‘lifted’ to SS and SGS_{G} by ‘small circle pushing’. Since H1(SSβ,(SSβ))H1(S//Sβ)\operatorname{H}_{1}(S\smallsetminus S^{\beta},\partial(S\smallsetminus S^{\beta}))\to\operatorname{H}^{1}(S/\!\!/S^{\beta}) is an isomorphism, the action on H1(SSβ,(SSβ))\operatorname{H}_{1}(S\smallsetminus S^{\beta},\partial(S\smallsetminus S^{\beta})) will be trivial. Clearly the components of SβS_{\beta} will be permuted according to the right action of gg and thus realizes gΓ(a,b)g\in\Gamma(a,b) in the image of Diff+(S)G\operatorname{Diff}^{+}(S)^{G}. ∎

Part (ii) of the corollary below establishes the Putman-Wieland property of Theorem 1.1.

Corollary 5.5 (Hyperbolic generation).

The following properties hold:

  1. (i)

    the subset {a}a\{a\}\cup{\mathscr{B}}_{a} of H1(S;)\operatorname{H}_{1}(S;{\mathbb{Q}}) spans the latter over G{\mathbb{Q}}G,

  2. (ii)

    the subgroup Γo(a)\Gamma_{o}(a) of Sp(H1(S))\operatorname{Sp}(\operatorname{H}_{1}(S)) generated by the subgroups Γo(a,b)\Gamma_{o}(a,b) with bab\in{\mathscr{B}}_{a} (and hence Diff+(S)G\operatorname{Diff}^{+}(S)^{G}) has no nonzero finite orbit in H1(S)\operatorname{H}_{1}(S).

Proof.

Let cH1(S;)c\in\operatorname{H}_{1}(S;{\mathbb{Q}}) be perpendicular to {a}a\{a\}\cup{\mathscr{B}}_{a}. We prove that cc is then perpendicular to every xH1(S)x\in\operatorname{H}_{1}(S); since the intersection form is nondegenerate, this will imply that c=0c=0 and hence that the G{\mathbb{Z}}G-submodule of H1(S)\operatorname{H}_{1}(S) generated {a}a\{a\}\cup{\mathscr{B}}_{a} is of finite index. To this end, let bab\in{\mathscr{B}}_{a}. Then x:=(1+x,a)b+xx^{\prime}:=(1+\langle x,a\rangle)b+x has the property that a,x=1\langle a,x^{\prime}\rangle=1. By Corollary 5.2 is (a,x′′)(a,x^{\prime\prime}) is a hyperbolic pair for some x′′x+Rax^{\prime\prime}\in x^{\prime}+Ra, so that 0=x′′,c=x,c0=\langle x^{\prime\prime},c\rangle=\langle x,c\rangle.

For (ii) it suffices to show that for every finite index subgroup ΓΓo(a)\Gamma\subset\Gamma_{o}(a), the fixed part H1(S)Γ\operatorname{H}_{1}(S)^{\Gamma} is trivial. Note that H1(S)Γo(a,b)\operatorname{H}_{1}(S)^{\Gamma_{o}(a,b)} is the perp of Ra+RbRa+Rb in H1(S)\operatorname{H}_{1}(S) with respect to the intersection pairing. The Γo(a,b)\Gamma_{o}(a,b)-invariant part of H1(S)\operatorname{H}_{1}(S) is not changed if we replace Γo(a,b)\Gamma_{o}(a,b) by the finite index subgroup ΓΓo(a,b)\Gamma\cap\Gamma_{o}(a,b) and hence H1(S)Γ\operatorname{H}_{1}(S)^{\Gamma} is perpendicular to Ra+RbRa+Rb. As this is true for all bab\in{\mathscr{B}}_{a} and {a}a\{a\}\cup{\mathscr{B}}_{a} generates H1(S)\operatorname{H}_{1}(S) as an RR-module, it follows that H1(S)Γ\operatorname{H}_{1}(S)^{\Gamma} must be trivial. ∎

We can now finish the proof of our main theorem.

Proof of the main theorem 1.1.

Let us denote the image of Diff+(M)G\operatorname{Diff}^{+}(M)^{G} in Sp(H1(S;))G\operatorname{Sp}(\operatorname{H}_{1}(S;{\mathbb{Q}}))^{G} by Γ\Gamma and the image of the latter in the factor U(H1(S)[χ])\operatorname{U}(\operatorname{H}_{1}(S)[\chi]) of Sp(H1(S;))G=χU(H1(S)[χ])\operatorname{Sp}(\operatorname{H}_{1}(S;{\mathbb{Q}}))^{G}=\prod_{\chi}\operatorname{U}(\operatorname{H}_{1}(S)[\chi]) by Γχ\Gamma_{\chi}. By combining Corollary 5.5 with Theorem 4.2, we see that under the assumptions of (ii), Γχ\Gamma_{\chi} is an arithmetic subgroup of U(H1(S)[χ])\operatorname{U}(\operatorname{H}_{1}(S)[\chi]).

Note that Γχ:=ΓEU(H1(S)[χ])\Gamma^{\chi}:=\Gamma\cap\operatorname{EU}(\operatorname{H}_{1}(S)[\chi]) is a normal subgroup of Γχ\Gamma_{\chi}. It remains to see that Γχ\Gamma^{\chi} is of finite index in Γχ\Gamma_{\chi}. Since EU(H1(S)[χ])\operatorname{EU}(\operatorname{H}_{1}(S)[\chi]) is almost simple and of real rank 2\geq 2, it follows from a general result of Margulis ([4], Assertion (A), Ch. VIII) that this is the case unless Γχ\Gamma^{\chi} meets EU(H1(S)[χ])\operatorname{EU}(\operatorname{H}_{1}(S)[\chi]) in the center. But Proposition 5.4 shows that Γχ\Gamma^{\chi} contains a subgroup isomorphic to Γo(χ)\Gamma_{o}(\chi) and so this last possibility does not occur. ∎

Remark 5.6.

This argument shows that if we are in the setting of (i) (triviality of the cover over a genus one subsurface of SGS_{G}), then Γχ\Gamma_{\chi} is an arithmetic subgroup of U(H1(S)[χ])\operatorname{U}(\operatorname{H}_{1}(S)[\chi]), unless VχDχV_{\chi}^{\dagger}\cong D_{\chi} and the image of GG in VχV_{\chi} is of the type given in Proposition 3.1. Denote that image by GχG_{\chi} and let GχG^{\chi} stand for the kernel of GGχG\to G_{\chi}. Since H1(S)[χ]\operatorname{H}_{1}(S)[\chi] already arises on the GχG_{\chi}-cover SGχSGS_{G^{\chi}}\to S_{G} in the sense that H1(S)[χ]=H1(SGχ)[χ]\operatorname{H}_{1}(S)[\chi]=\operatorname{H}_{1}(S_{G^{\chi}})[\chi], we may for the arithmeticity question just as well focus on this intermediate cover.

In the cases (i) and (ii) of 3.1, so when DχD_{\chi} equals {\mathbb{Q}}, the Gaussian field or the Eisenstein field, then GχG_{\chi} is a group of roots of unity and hence cyclic. When the genus of SGS_{G} is at least 33, we can always find a closed subsurface of genus 2 over which the covering SGχSGS_{G^{\chi}}\to S_{G} is trivial and so Γχ\Gamma_{\chi} is then arithmetic. In the remaining cases, GχG_{\chi} is a particular kind of Kleinian group. It might well be that a GχG_{\chi}-cover is then also trivial over the complement of a genus two subsurface of the quotient surface when the genus of the latter is 3\geq 3. If true, then it would follow that Γ\Gamma would always be arithmetic if the genus of SGS_{G} is at least 33 and the covering is trivial over a genus one subsurface of SGS_{G}.

References

  • [1] F. Grünewald, M. Larsen, A. Lubotzky, J. Malestein: Arithmetic quotients of the mapping class group. Geom. Funct.  Anal. 25 (2015) 1493–1542, https://arxiv.org/pdf/1307.2593v2
  • [2] A. J. Hahn, O. T. O’Meara: The classical groups and K-theory. Grundlehren Math. Wiss.  291, Springer-Verlag (1989).
  • [3] E. Looijenga: Prym Representations of Mapping Class Groups. Geom. Dedicata 64 (1997), 69–83.
  • [4] G. A. Margulis: Discrete subgroups of semisimple Lie groups. Ergeb. Math. Grenzgeb. 17, Springer-Verlag, Berlin.
  • [5] D. Mumford: Abelian varieties. TIFR Studies in Math. 5, Oxford University Press, London (1970).
  • [6] A. Putman, B. Wieland: Abelian quotients of subgroups of the mappings class group and higher Prym representations, J. Lond. Math. Soc.  88 (2013), 79–96.
  • [7] M. S. Raghunathan: A note on generators for arithmetic subgroups of algebraic groups. Pacific J. Math. 152 (1992), 365–373.
  • [8] T. A. Springer: Linear algebraic groups. Birkhäuser Boston, Inc., Boston, MA (2009).
  • [9] T. N. Venkataramana: On systems of generators of arithmetic subgroups of higher rank groups. Pacific J. Math. 166 (1994), 193–212.
  • [10] G. E. Wall: The structure of a unitary factor group. Inst. Hautes Études Sci. Publ. Math.  1 (1959).