1 Introduction
In this work we consider products of locally compact second countable (lcsc) groups (hereafter lcsc groups) and their lattices.
Definition 1.1. Let
$T$
be an lcsc group. A discrete subgroup
$\unicode[STIX]{x1D6E4}<T$
is a lattice in
$T$
if the coset space
$T/\unicode[STIX]{x1D6E4}$
carries a finite
$T$
-invariant measure.
Let
$T=T_{1}\times T_{2}\times \cdots \times T_{n}$
be a product of
$n$
lcsc groups and denote by
$\unicode[STIX]{x1D70B}_{i}:T\rightarrow T_{i}$
the projections. A subgroup
$\unicode[STIX]{x1D6E4}<T$
is said to be a lattice with dense projections in
$T$
if
$\unicode[STIX]{x1D6E4}<T$
is a lattice, and
$\unicode[STIX]{x1D70B}_{i}(\unicode[STIX]{x1D6E4})$
is dense in
$T_{i}$
for every
$i\in \{1,2,\ldots ,n\}$
.
In the above definition
$n$
is an arbitrary integer, but the case
$n=1$
is redundant (it forces
$T$
to be discrete and
$\unicode[STIX]{x1D6E4}=T$
), thus we are mostly concerned with
$n\geqslant 2$
. The following theorem extends the case of the classical super-rigidity theorem of Margulis [Reference MargulisMar91, ch. VII, Theorem (5.6)], which concerns irreducible lattices in a semisimple Lie group of higher rank. The case of a lattice in a simple Lie group of higher rank is treated in [Reference Bader and FurmanBF18]. The target group
$\mathbf{G}(k)$
below is defined over a valued field
$k$
that need not be a local field. The notion of being bounded for subsets of such a group
$\mathbf{G}(k)$
, a group bornology as defined by Bruhat and Tits [Reference Bruhat and TitsBT72, (3.1.1)], is discussed in detail in [Reference Bader, Duchesne and LécureuxBDL17, § 4].
Theorem 1.2 (Super-rigidity for lattices in products).
Let
$T=T_{1}\times T_{2}\times \cdots \times T_{n}$
be a product of lcsc groups, and let
$\unicode[STIX]{x1D6E4}<T$
be a lattice with dense projections. Let
$k$
be a field with an absolute value. Assume that, as a metric space,
$k$
is complete. Let
$G=\mathbf{G}(k)$
be the
$k$
-points of a connected, adjoint,
$k$
-simple algebraic group. Let
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6E4}\rightarrow G$
be a homomorphism. Assume
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4})$
is Zariski dense and unbounded in
$G$
. Then there exists a unique continuous homomorphism
$\bar{\unicode[STIX]{x1D70C}}:T\rightarrow G$
satisfying
$\unicode[STIX]{x1D70C}=\bar{\unicode[STIX]{x1D70C}}|_{\unicode[STIX]{x1D6E4}}$
. Moreover, that homomorphism is of the form
$\bar{\unicode[STIX]{x1D70C}}=\bar{\unicode[STIX]{x1D70C}}_{i}\circ \unicode[STIX]{x1D70B}_{i}$
for some (necessarily unique)
$i\in \{1,\ldots ,n\}$
and some (necessarily unique) continuous homomorphism
$\bar{\unicode[STIX]{x1D70C}}_{i}:T_{i}\rightarrow G$
.
The proof of Theorem 1.2 will be given in § 6. Similar super-rigidity theorems for lattices in products were given by Monod [Reference MonodMon06] and Gelander, Karlsson and Margulis [Reference Gelander, Karlsson and MargulisGKM08] under the further assumption that
$\unicode[STIX]{x1D6E4}<T$
is cocompact (or satisfies a non-trivial integrability condition) and by Caprace and Monod [Reference Caprace and MonodCM09, Theorems 5.1, 5.6] under a finite generation assumption on
$\unicode[STIX]{x1D6E4}$
(note also that the first two references consider a wider class of possible targets). No integrability or finite generation assumptions are forced on
$\unicode[STIX]{x1D6E4}$
in Theorem 1.2. This is a hint that the above theorem could be generalized to the setting of a cocycle, which is indeed the case. However, we choose not to elaborate on this point here.
A nice application of the theory of lattices in products is given by Caprace and Monod in the following lemma.
Lemma 1.3 (Caprace and Monod [Reference Caprace and MonodCM09, Lemma 5.15]).
Let
$S$
be a locally compact group,
$\unicode[STIX]{x1D6E4}<S$
a lattice and
$i:\unicode[STIX]{x1D6EC}{\hookrightarrow}S$
a countable dense subgroup containing and commensurating
$\unicode[STIX]{x1D6E4}$
. Then there exist a locally compact group
$S^{\prime }$
and a homomorphism
$\unicode[STIX]{x1D703}:\unicode[STIX]{x1D6EC}\rightarrow S^{\prime }$
such that
$\unicode[STIX]{x1D703}(\unicode[STIX]{x1D6E4})$
is precompact in
$S^{\prime }$
and
$i\times \unicode[STIX]{x1D703}(\unicode[STIX]{x1D6EC})<S\times S^{\prime }$
is a lattice with dense projections.
The group
$S^{\prime }$
appearing in the above lemma is the completion of
$\unicode[STIX]{x1D6EC}$
with respect to the left uniform structure generated by conjugates of
$\unicode[STIX]{x1D6E4}$
. Applying Theorem 1.2 with
$T_{1}=S$
,
$T_{2}=S^{\prime }$
and
$T=T_{1}\times T_{2}$
, we easily obtain the following corollary.
Corollary 1.4 (Super-rigidity for commensurators).
Let
$T$
be an lcsc group, and let
$\unicode[STIX]{x1D6E4}<T$
be a lattice. Let
$\unicode[STIX]{x1D6EC}$
be a countable dense subgroup of
$T$
containing and commensurating
$\unicode[STIX]{x1D6E4}$
. Let
$k$
be a field with an absolute value. Assume that, as a metric space,
$k$
is complete. Let
$G$
be the
$k$
-points of a connected, adjoint,
$k$
-simple algebraic group. Let
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6EC}\rightarrow G$
be a homomorphism. Assume
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6EC})$
is Zariski dense and
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4})$
is unbounded in
$G$
. Then there exists a unique continuous homomorphism
$\bar{\unicode[STIX]{x1D70C}}:T\rightarrow G$
such that
$\unicode[STIX]{x1D70C}=\bar{\unicode[STIX]{x1D70C}}|_{\unicode[STIX]{x1D6EC}}$
.
This corollary is a generalization of another well-known theorem of Margulis [Reference MargulisMar91, ch. VII, Theorem (5.4)]. Here too, similar theorems were proved by Monod [Reference MonodMon06, Theorem A.1], Gelander et al. [Reference Gelander, Karlsson and MargulisGKM08, Theorem 8.1] and Caprace and Monod [Reference Caprace and MonodCM09, Theorems 5.17] under further assumptions on
$\unicode[STIX]{x1D6E4}$
and
$\unicode[STIX]{x1D6EC}$
.
Note that Theorem 1.2 and Corollary 1.4 are new even in the case where
$T$
is a semisimple group over a local field, as
$k$
is not assumed to be a local field. The analogous super-rigidity theorems of Margulis are cornerstones in his celebrated proof of his arithmeticity theorems [Reference MargulisMar91, ch. IX (1.9), Theorems (A) and (B)]. However, due to the standing assumption that the target group is defined over a local field, he is forced to assume in the aforementioned Theorems (A) and (B) that the lattice
$\unicode[STIX]{x1D6E4}$
is finitely generated. Our new versions of these theorems show that these finite generation issues can be circumvented. We thank Venkataramana for pointing out this application to us. We should note that these assumptions were later removed, in case (A) by Raghunathan [Reference RaghunathanRag89], and in case (B) by Lifschitz [Reference LifschitzLif03]. However, using our Theorem 1.2 in case (A) and Corollary 1.4 in case (B) is easier and avoids the detailed discussion of the shape of the cusps in locally symmetric spaces of function fields undertaken in [Reference RaghunathanRag89] and relied upon in [Reference LifschitzLif03]. We will elaborate further on this point in a forthcoming paper.
Theorem 1.2 and Corollary 1.4 have striking applications if the group
$T$
is not linear. We say that a group
$\unicode[STIX]{x1D6E4}$
is solvable by locally finite if there exists a normal solvable subgroup
$S{\vartriangleleft}\unicode[STIX]{x1D6E4}$
such that
$\unicode[STIX]{x1D6E4}/S$
is locally finite, that is, it is a directed union of finite subgroups. We say that topological group
$H$
is amenable if every continuous
$H$
-action on a compact space has an invariant probability measure.
Theorem 1.5. Let
$T=T_{1}\times T_{2}\times \cdots \times T_{n}$
be a product of lcsc groups, and let
$\unicode[STIX]{x1D6E4}<T$
be a lattice with dense projections. Assume that for every continuous homomorphism
$\unicode[STIX]{x1D713}:T\rightarrow \operatorname{GL}_{d}(k)$
, where
$k$
is a complete field with absolute value and
$d$
is an integer, the closure of the image,
$\overline{\unicode[STIX]{x1D713}(T)}$
, is amenable. Then for every integer
$d$
, field
$K$
and linear representation
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6E4}\rightarrow \operatorname{GL}_{d}(K)$
, the image
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is solvable by locally finite. Furthermore, if
$\unicode[STIX]{x1D6E4}$
is assumed to be finitely generated, the class of fields considered in the target of
$\unicode[STIX]{x1D713}$
can be taken to be the class of local fields, and then the image
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is solvable by finite.
Corollary 1.6. Let
$T$
be an lcsc group, and let
$\unicode[STIX]{x1D6E4}<T$
be a lattice. Let
$\unicode[STIX]{x1D6EC}$
be a countable dense subgroup of
$T$
containing and commensurating
$\unicode[STIX]{x1D6E4}$
. Assume that for every continuous homomorphism
$\unicode[STIX]{x1D713}:T\rightarrow \operatorname{GL}_{d}(k)$
, where
$k$
is a complete field with an absolute value and
$d$
is an integer, the closure of the image,
$\overline{\unicode[STIX]{x1D713}(T)}$
, is amenable. Then for every integer
$d$
, field
$K$
and linear representation
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6EC}\rightarrow \operatorname{GL}_{d}(K)$
, the image
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is solvable by locally finite. Furthermore, if
$\unicode[STIX]{x1D6EC}$
is assumed to be finitely generated, the class of fields considered in the target of
$\unicode[STIX]{x1D713}$
can be taken to be the class of local fields, and then the image
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is solvable by finite.
Corollary 1.7. Assume given for every
$i=1,\ldots ,n$
a totally disconnected lcsc group
$T_{i}$
in which for every closed non-trivial normal subgroup the quotient group is amenable. Let
$T=T_{1}\times T_{2}\times \cdots \times T_{n}$
and let
$\unicode[STIX]{x1D6E4}<T$
be a lattice with dense projections such that for each
$i$
,
$\unicode[STIX]{x1D70B}_{i}|_{\unicode[STIX]{x1D6E4}}$
is injective (and dense in
$T_{i}$
). Assume that
$T_{1}$
is non-amenable and for every continuous homomorphism
$\unicode[STIX]{x1D713}:T_{1}\rightarrow \operatorname{GL}_{d}(k)$
, where
$k$
is a complete field with an absolute value and
$d$
is an integer, the closure of the image,
$\overline{\unicode[STIX]{x1D713}(T_{1})}$
, is amenable. Then for every integer
$d$
, field
$K$
and linear representation
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6E4}\rightarrow \operatorname{GL}_{d}(K)$
, the image
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is solvable by locally finite. Furthermore, if
$\unicode[STIX]{x1D6E4}$
is assumed to be finitely generated, the class of fields considered in the target of
$\unicode[STIX]{x1D713}$
could be taken to be the class of local fields, and then the image
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is solvable by finite.
The proofs of the above three theorems will be given in § 8. In their proofs we will use new linearity criteria, namely Theorem 7.1 and Corollary 7.2, which are of independent interest. These will be proven in § 7. Note that, unlike Theorems 1.2, 1.5 and Corollary 1.7 are not tautological in the case
$n=1$
.
The last three results could be further strengthened if we assume that
$\unicode[STIX]{x1D6E4}$
has property (T), or that we are in a situation corresponding to the setting of [Reference Bader and ShalomBS06, Theorem 1.1], where one can deduce further that the image of
$\unicode[STIX]{x1D6E4}$
is finite in any linear representation. We refer the reader to [Reference MonodMon05] where such a theorem is proven, along with a very nice converse: under certain natural assumptions on a finitely generated lattice in a product, the only way it could have a linear representation with an infinite image is if it is an arithmetic lattice to begin with. For a similar conclusion (in characteristic
$0$
) with no finite generation assumption, see [Reference Bader, Furman and SauerBFS19, Theorem 4.5].
1.1 Disclaimers
Theorem 1.2 appeared in our manuscript [Reference Bader and FurmanBF14a], which we do not intend to publish as, in retrospect, we find it hard to read. In writing this paper we have made an effort to improve the presentation and to keep things concise and simple. The paper [Reference Bader and FurmanBF14a] contains further results, regarding higher rank lattices and general cocycle super-rigidity, on which we intend to elaborate in two forthcoming papers.
2 Algebraic varieties as Polish spaces
In this section we fix a field
$k$
with an absolute value
$|\cdot |:k\rightarrow [0,\infty )$
as defined and discussed in [Reference Engler and PrestelEP05]. We assume that the absolute value is non-trivial, and that
$(k,|\cdot |)$
is complete (as a metric space) and separable (in the sense of having a countable dense subset). The theory of manifolds over such fields is developed in [Reference SerreSer06, Part II, ch. I].
We also fix a
$k$
-algebraic group
$\mathbf{G}$
. We will discuss the category of
$k$
-
$\mathbf{G}$
-varieties. A
$k$
-
$\mathbf{G}$
-variety is a
$k$
-variety endowed with an algebraic action of
$\mathbf{G}$
which is defined over
$k$
. A morphism of such varieties is a
$k$
-morphism which commutes with the
$\mathbf{G}$
-action. By a
$k$
-coset variety we mean a variety of the form
$\mathbf{G}/\mathbf{H}$
for some
$k$
-algebraic subgroup
$\mathbf{H}<\mathbf{G}$
(see [Reference BorelBor91, Theorem 6.8]).
Each
$k$
-
$\mathbf{G}$
-variety gives rise to a topological space
$V=\mathbf{V}(k)$
endowed with its
$(k,|\cdot |)$
-topology. Topological notions, unless otherwise stated, will refer to this topology. In particular,
$G=\mathbf{G}(k)$
is a topological group. It is locally compact if
$k$
is a local field.
The norm
$|\cdot |$
on
$k$
allows one to define the associated bornology (see [Reference Bruhat and TitsBT72, (3.1.1)]) on
$k$
, on the affine spaces
$k^{N}$
, and on affine
$k$
-varieties such as
$\mathbf{G}(k)$
. We refer to [Reference Bader, Duchesne and LécureuxBDL17, § 4] for a detailed discussion of the bornology on
$\mathbf{G}(k)$
induced by
$(k,|\cdot |)$
. In the special case where
$k$
is a local field, being a bounded subset of
$\mathbf{G}(k)$
is equivalent to being precompact in the topology on
$\mathbf{G}(k)$
induced by
$k$
.
Recall that a topological space is called Polish if it is separable and completely metrizable. For a good survey on the subject we recommend [Reference KechrisKec95]. We mention that the class of Polish spaces is closed under countable disjoint unions and countable products. A
$G_{\unicode[STIX]{x1D6FF}}$
subset of a Polish spaces is Polish so, in particular, a locally closed subset of a Polish space is Polish. A Hausdorff space which admits a finite open covering by Polish open sets is itself Polish. Indeed, such a space is clearly metrizable (e.g. by the Smirnov metrization theorem) so it is Polish by Sierpiński’s theorem [Reference KechrisKec95, Theorem 8.19] which states that the image of an open map from a Polish space to a separable metrizable space is Polish. Sierpiński’s theorem also implies that for a Polish group
$K$
and a closed subgroup
$L$
, the quotient topology on
$K/L$
is Polish. Effros’s lemma [Reference EffrosEff65, Lemma 2.5] says that the quotient topology on
$K/L$
is the unique
$K$
-invariant Polish topology on this space.
Proposition 2.1. The
$k$
-points of a
$k$
-variety form a Polish space. In particular,
$G=\mathbf{G}(k)$
is a Polish group. If
$\mathbf{V}$
is a
$k$
-
$\mathbf{G}$
-variety then the
$G$
-orbits in
$V$
are locally closed. For
$v\in V$
the orbit
$\mathbf{G}v$
is a
$k$
-subvariety of
$\mathbf{V}$
. The stabilizer
$\mathbf{H}<\mathbf{G}$
is defined over
$k$
and the orbit map
$\mathbf{G}/\mathbf{H}\rightarrow \mathbf{G}v$
is defined over
$k$
. Denoting
$H=\mathbf{H}(k)$
, the induced map
$G/H\rightarrow Gv$
is a homeomorphism when
$G/H$
is endowed with the quotient space topology and
$Gv$
is endowed with the subspace topology.
Proof. Since
$k$
is complete and separable it is Polish and so is the affine space
$\mathbb{A}^{n}(k)\simeq k^{n}$
. The set of
$k$
-points of a
$k$
-affine variety is closed in the
$k$
-points of the affine space, hence it is a Polish subspace. It follows that the set of
$k$
-points of any
$k$
-variety is a Polish space, as this space is a Hausdorff space which admits a finite open covering by Polish open sets, namely the
$k$
-points of its
$k$
-affine charts.
The fact that the
$G$
-orbits in
$V$
are locally closed is proven in the appendix of [Reference Bernšteĭn and ZelevinskiĭBZ76]. Note that in [Reference Bernšteĭn and ZelevinskiĭBZ76] the statement is claimed only for non-archimedean local fields, but the proof is actually correct for any field with a complete non-trivial absolute value, which is the setting of [Reference SerreSer06, Part II, ch. III] on which [Reference Bernšteĭn and ZelevinskiĭBZ76] relies.
For
$v\in V$
the orbit
$\mathbf{G}v$
is a
$k$
-subvariety of
$\mathbf{V}$
by [Reference BorelBor91, Proposition 6.7]. The stabilizer
$\mathbf{H}<\mathbf{G}$
is defined over
$k$
by [Reference BorelBor91, Proposition 1.7] and we obtain an orbit map which is defined over
$k$
by [Reference BorelBor91, Theorem 6.8]. Clearly
$H$
is the stabilizer of
$v$
in
$G$
and the orbit map restricts to a continuous map from
$G/H$
onto
$Gv$
. Since
$Gv$
is a Polish subset of
$V$
, as it is locally closed, we conclude by Effros’ lemma that the latter map is a homeomorphism.◻
3 Ergodic-theoretical preliminaries
In this section we review some notions and facts from ergodic theory. Recall that a standard Borel space is a measurable space which is isomorphic as such to a Polish topological space endowed with its Borel
$\unicode[STIX]{x1D70E}$
-algebra, and a Lebesgue space is a standard Borel space endowed with a
$\unicode[STIX]{x1D70E}$
-finite measure class.
Given a Lebesgue space
$Y$
and a Borel space
$V$
, we denote by
$\operatorname{L}^{0}(Y,V)$
the space of all classes of measurable maps, identified up to almost everywhere equality, from
$Y$
to
$V$
. The elements of
$\operatorname{L}^{0}(Y,V)$
are called Lebesgue maps from
$Y$
to
$V$
. Given a standard Borel space
$V$
, a Lebesgue space
$Y$
and an essentially surjective Borel map
$\unicode[STIX]{x1D70B}:V\rightarrow Y$
(i.e.
$\unicode[STIX]{x1D70B}$
is measurable for a Borel model of
$Y$
and its image has full measure), we denote by
$\operatorname{L}^{0}(\unicode[STIX]{x1D70B})$
the set of all equivalence classes of measurable sections of
$\unicode[STIX]{x1D70B}$
, defined up to almost everywhere equality.
Given two Lebesgue spaces
$X,Y$
, we use the term Lebesgue morphism to denote a Lebesgue map from
$X$
to the underlying Borel space of
$Y$
which is measure class preserving, that is, the preimage of a null set in
$Y$
is null in
$X$
. There is a correspondence between Lebesgue morphisms
$X\rightarrow Y$
and von Neumann algebra morphisms
$L^{\infty }(Y)\rightarrow L^{\infty }(X)$
.
Every lcsc group (i.e. lcsc topological group) gives rise to a Lebesgue space when endowed with its Haar measure class. Accordingly, we will often regard lcsc groups as Lebesgue spaces without further mention. Given an lcsc group
$S$
, a Lebesgue space
$Y$
is called an
$S$
-Lebesgue space when it is endowed with a Lebesgue action of
$S$
. That is a Lebesgue morphism
$S\times Y\rightarrow Y$
which gives rise to an action of
$S$
on
$L^{\infty }(Y)$
by automorphisms of von Neumann algebras. A Borel space
$V$
is called an
$S$
-Borel space when it endowed with a Borel action of
$S$
. That is a Borel morphism
$S\times V\rightarrow V$
which gives rise to an action of
$S$
on
$V$
by automorphisms of a Borel space.
Definition 3.1 (Amenability; see [Reference ZimmerZim84, Definition 4.3.1]).
Given an lcsc group
$S$
, an
$S$
-Lebesgue space
$Y$
is called amenable if for every
$S$
-Borel space
$V$
and an essentially surjective
$S$
-equivariant Borel map
$\unicode[STIX]{x1D70B}:V\rightarrow Y$
with a measurably defined compact convex structure on the fibers, such that the
$S$
-action restricted to the fibers is by continuous affine maps, one has
$L^{0}(\unicode[STIX]{x1D70B})^{S}\neq \emptyset$
. That is, every
$S$
-Borel bundle of convex compact sets over
$S$
admits an invariant measurable section.
Definition 3.2 (Metric ergodicity; see [Reference Bader and FurmanBF14b, § 2]).
Given an lcsc group
$S$
, an
$S$
-Lebesgue space
$Y$
is called metrically ergodic if for every separable metric space
$V$
on which
$S$
acts continuously by isometries, every
$S$
-equivariant Lebesgue map
$Y\rightarrow V$
is almost everywhere constant.
Theorem 3.3 (Kaimanovich and Zimmer).
For every lcsc group
$S$
there exists an
$S$
-Lebesgue space
$Y$
which is both amenable and metrically ergodic.
Remark 3.4. In fact [Reference KaimanovichKai03] shows that there exists a space
$Y$
which is amenable and doubly metrically ergodic. With a slight restriction, this was first proven in [Reference Burger and MonodBM02]. A slightly stronger theorem is proven in [Reference Bader and FurmanBF14b, Theorem 2.7], but we will not discuss these extensions here.
On the proof of Theorem 3.3.
In [Reference KaimanovichKai03] the weaker statement that every group possesses a strong boundary in the sense of [Reference Burger and MonodBM02] is proven, but the same proof actually proves Theorem 3.3, as explained in [Reference Glasner and WeissGW16, Remark 4.3].◻
Lemma 3.5. Given an lcsc group
$S$
and a lattice
$\unicode[STIX]{x1D6E4}<S$
, if
$Y$
is an amenable and metrically ergodic
$S$
-Lebesgue space then it is also amenable and metrically ergodic as a
$\unicode[STIX]{x1D6E4}$
-Lebesgue space.
Proof. The fact that
$Y$
is
$\unicode[STIX]{x1D6E4}$
-amenable follows from [Reference ZimmerZim84, 4.3.5]. To show that
$Y$
is
$\unicode[STIX]{x1D6E4}$
-metrically ergodic, we fix a measurable
$\unicode[STIX]{x1D6E4}$
-equivariant map
$\unicode[STIX]{x1D719}:Y\rightarrow V$
, where
$(V,d)$
is a metric space on which
$\unicode[STIX]{x1D6E4}$
acts isometrically, and argue to show that it is essentially constant. Replacing
$d(x,y)$
by
$\min (d(x,y),1)$
we may assume
$d$
to be bounded. We consider the space
$\operatorname{L}^{0}(S,V)^{\unicode[STIX]{x1D6E4}}$
consisting of Lebesgue maps which are
$\unicode[STIX]{x1D6E4}$
-equivariant with respect to the left action of
$\unicode[STIX]{x1D6E4}$
on
$S$
. Given
$\unicode[STIX]{x1D6FC},\unicode[STIX]{x1D6FD}\in \operatorname{L}^{0}(S,V)^{\unicode[STIX]{x1D6E4}}$
, note that
$d(\unicode[STIX]{x1D6FC}(x),\unicode[STIX]{x1D6FD}(x))$
is a
$\unicode[STIX]{x1D6E4}$
-invariant function on
$S$
that descends to a well-defined function on
$\unicode[STIX]{x1D6E4}\backslash S$
. Using this observation, we define on
$\operatorname{L}^{0}(S,V)^{\unicode[STIX]{x1D6E4}}$
a function
$D$
by setting, for
$\unicode[STIX]{x1D6FC},\unicode[STIX]{x1D6FD}\in \operatorname{L}^{0}(S,V)^{\unicode[STIX]{x1D6E4}}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU1.gif?pub-status=live)
Then
$D$
is a metric on
$\operatorname{L}^{0}(S,V)^{\unicode[STIX]{x1D6E4}}$
, and
$S$
acts continuously and isometrically on this space via its right regular action on the domain. The map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU2.gif?pub-status=live)
defined using Fubini’s theorem (see [Reference MargulisMar91, ch. VII, Lemma 1.3]), is
$S$
-equivariant. By the
$S$
-metric ergodicity of
$Y$
we conclude that
$\unicode[STIX]{x1D6F7}$
is essentially constant, and thus
$\unicode[STIX]{x1D719}$
is essentially constant too.◻
Lemma 3.6. Let
$S_{1},\ldots ,S_{n}$
be lcsc groups, and let
$Y_{i}$
be an amenable and metrically ergodic
$S_{i}$
-Lebesgue space for each
$i=1,\ldots ,n$
. Then
$Y_{1}\times \cdots \times Y_{n}$
is an amenable and metrically ergodic
$S_{1}\times \cdots \times S_{n}$
-Lebesgue space.
Proof. By induction, it suffices to consider the case
$n=2$
. The proof that
$Y=Y_{1}\times Y_{2}$
is
$S$
-amenable for
$S=S_{1}\times S_{2}$
is well known, so we merely sketch it. Let
$\unicode[STIX]{x1D70B}:C\rightarrow Y$
be an
$S$
-Borel bundle of convex compact sets over
$Y$
. For every
$y_{1}\in Y_{1}$
we let
$C_{y_{1}}=\unicode[STIX]{x1D70B}^{-1}(\{y_{1}\}\times Y_{2})$
and
$\operatorname{L}^{0}(\unicode[STIX]{x1D70B}|_{C_{y_{1}}}\!)$
be the corresponding space of sections. We view these spaces as a convex compact bundle over
$Y_{1}$
(using the obvious weak* topology) and denote by
$\operatorname{L}^{0}(Y_{1},\operatorname{L}^{0}(\unicode[STIX]{x1D70B}|_{C_{y_{1}}}\!))$
its space of sections. Its obvious identification with
$\operatorname{L}^{0}(\unicode[STIX]{x1D70B})$
gives an identification
$\operatorname{L}^{0}(\unicode[STIX]{x1D70B})^{S}\simeq \operatorname{L}^{0}(Y_{1},\operatorname{L}^{0}(\unicode[STIX]{x1D70B}|_{C_{y_{1}}}\!)^{S_{2}})^{S_{1}}$
. The right-hand side is non-empty by our amenability assumptions, hence so is the left-hand side.
Let
$V$
be an
$S$
-metric space and
$\unicode[STIX]{x1D719}:Y\rightarrow V$
an
$S$
-equivariant Lebesgue map. For almost every
$y_{1}\in Y_{1}$
,
$\unicode[STIX]{x1D719}_{\{y_{1}\}\times Y_{2}}$
is defined by Fubini’s theorem, and it is essentially constant, as
$Y_{2}$
is
$S_{2}$
-metrically ergodic. Thus
$\unicode[STIX]{x1D719}$
is reduced to a map
$\unicode[STIX]{x1D719}^{\prime }:Y_{1}\rightarrow V$
which is again essentially constant, as
$Y_{1}$
is
$S_{1}$
-metrically ergodic. Thus
$\unicode[STIX]{x1D719}$
is essentially constant.◻
Lemma 3.7. Let
$S$
be an lcsc group and
$X,Y$
be
$S$
-Lebesgue spaces. Assume the action on
$X$
is ergodic and probability measure preserving and the action on
$Y$
is metrically ergodic. Then the diagonal
$S$
-action on
$X\times Y$
is ergodic.
Proof. For
$f\in L^{\infty }(X\times Y)^{S}$
, using Fubini’s theorem, we define
$F:Y\rightarrow L^{\infty }(X)\subset L^{2}(X)$
by
$F(y)(x)=f(x,y)$
. It is easily checked that
$F$
is
$S$
-equivariant. The image of
$F$
must be
$S$
-invariant, by the metric ergodicity of
$Y$
, as the
$S$
-action on
$L^{2}(X)$
is continuous. By ergodicity of
$X$
this image is a constant function, thus
$f$
is constant.◻
Lemma 3.8. Let
$S_{1},\ldots ,S_{n}$
be lcsc groups, let
$\unicode[STIX]{x1D6E4}<S=S_{1}\times \cdots \times S_{n}$
be a lattice with dense projections, and let
$Y_{i}$
be metrically ergodic
$S_{i}$
-Lebesgue spaces for
$i=1,\ldots ,n$
. Then the
$\unicode[STIX]{x1D6E4}$
-action on
$S_{i}\times \prod _{j\neq i}Y_{j}$
is ergodic.
Proof. We write
$S=S_{i}\times S_{i}^{\prime }$
, where
$S_{i}^{\prime }=\prod _{j\neq i}S_{j}$
. Since
$\unicode[STIX]{x1D70B}_{i}(\unicode[STIX]{x1D6E4})$
is dense in
$S_{i}$
,
$\unicode[STIX]{x1D6E4}$
acts ergodically on
$S_{i}=S/S_{i}^{\prime }$
. Thus
$S_{i}^{\prime }$
acts ergodically on
$X=S/\unicode[STIX]{x1D6E4}$
. Note that the latter action preserves a probability measure. The diagonal
$S_{i}^{\prime }$
-action on
$Y_{i}^{\prime }=\prod _{j\neq i}Y_{j}$
is metrically ergodic (Lemma 3.6), and, using Lemma 3.7, we conclude that
$X\times Y_{i}^{\prime }$
is
$S_{i}^{\prime }$
-ergodic. It follows that
$S\times Y_{i}^{\prime }$
is
$\unicode[STIX]{x1D6E4}\times S_{i}^{\prime }$
-ergodic, where
$\unicode[STIX]{x1D6E4}$
acts by its left action on
$S$
, and
$S_{i}^{\prime }$
acts diagonally via it right action on
$S$
and its given action on
$Y_{i}^{\prime }$
. Note that the Lebesgue isomorphism
$S\times Y_{i}^{\prime }\rightarrow S\times Y_{i}^{\prime }$
,
$(s,y)\mapsto (s,sy)$
intertwines the above-described action
$\unicode[STIX]{x1D6E4}\times S_{i}^{\prime }\curvearrowright S\times Y_{i}^{\prime }$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU3.gif?pub-status=live)
Taking the
$S_{i}^{\prime }$
-orbit space, we conclude that
$\unicode[STIX]{x1D6E4}$
acts ergodically on
$S_{i}\times Y_{i}^{\prime }$
.◻
We also make use of the following special situation.
Lemma 3.9. Let
$S$
be an lcsc group,
$X$
a Lebesgue
$S$
-space,
$\unicode[STIX]{x1D6E4}$
a countable group,
$\unicode[STIX]{x1D6E4}\rightarrow S$
a homomorphism,
$Z$
a Lebesgue
$\unicode[STIX]{x1D6E4}$
-space. Let
$V$
be a Borel
$\unicode[STIX]{x1D6E4}$
-space and
$\unicode[STIX]{x1D719}:X\times Z\rightarrow V$
a measurable
$\unicode[STIX]{x1D6E4}$
-map. Then for almost every
$x\in X$
the map
$\hat{\unicode[STIX]{x1D719}}_{x}:S\times Z\rightarrow V$
given by
$\hat{\unicode[STIX]{x1D719}}_{x}(s,z)=\unicode[STIX]{x1D719}(sx,z)$
is a measurable
$\unicode[STIX]{x1D6E4}$
-map.
Proof. Fix a Borel map
$\unicode[STIX]{x1D719}$
in its class of equivalent measurable maps. The Borel set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU4.gif?pub-status=live)
is conull in
$X\times Z$
. So by Fubini’s theorem ([Reference KechrisKec95, Theorem 8.19]), there is a conull set
$X_{0}\subset X$
so that for each
$x\in X_{0}$
for almost every
$z\in Z$
one has
$(x,z)\in A$
. Since the action map
$S\times X\rightarrow X$
is non-singular, the preimage of
$X_{0}$
is conull in
$S\times X$
, and so for a conull set
$X_{1}\subset X$
for
$x\in X_{1}$
the set
$S_{x}=\{s\in S\mid sx\in X_{0}\}$
is conull in
$S$
. Therefore for every
$x\in X_{1}$
,
$s\in S_{x}$
, and almost every
$z\in Z$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU5.gif?pub-status=live)
for all
$\unicode[STIX]{x1D6FE}\in \unicode[STIX]{x1D6E4}$
.◻
4 Algebraic representations of ergodic actions
Throughout this section we fix the following data:
∙ an lcsc group
$S$ ;
∙ an ergodic
$S$ -Lebesgue space
$Y$ ;
∙ a field
$k$ with a non-trivial absolute value which is separable and complete (as a metric space);
∙ a
$k$ -algebraic group
$\mathbf{G}$ ;
∙ a continuous homomorphism
$\unicode[STIX]{x1D70C}:S\rightarrow \mathbf{G}(k)$ , where
$\mathbf{G}(k)$ is regarded as a Polish group (see Proposition 2.1).
Definition 4.1. Given all the data above, an algebraic representation of
$Y$
consists of the following data:
∙ a
$k$ -
$\mathbf{G}$ -algebraic variety
$\mathbf{V}$ ;
∙ an
$S$ -equivariant Lebesgue map
$\unicode[STIX]{x1D719}:Y\rightarrow \mathbf{V}(k)$ , where
$\mathbf{V}(k)$ is regarded as a Polish space (see Proposition 2.1).
Sometimes we abbreviate the notation by saying that
$\mathbf{V}$
is an algebraic representation of
$Y$
, and denote
$\unicode[STIX]{x1D719}$
by
$\unicode[STIX]{x1D719}_{\mathbf{V}}$
for clarity. A morphism from the algebraic representation
$\mathbf{U}$
to the algebraic representation
$\mathbf{V}$
consists of:
∙ a
$k$ -algebraic map
$\unicode[STIX]{x1D713}:\mathbf{U}\rightarrow \mathbf{V}$ which is
$\mathbf{G}$ -equivariant, and such that
$\unicode[STIX]{x1D719}_{\mathbf{V}}$ agrees almost everywhere with
$\unicode[STIX]{x1D713}\circ \unicode[STIX]{x1D719}_{\mathbf{U}}$ .
An algebraic representation
$\mathbf{V}$
of
$Y$
is said to be a coset algebraic representation if in addition
$\mathbf{V}=\mathbf{G}/\mathbf{H}$
for some
$k$
-algebraic subgroup
$\mathbf{H}<\mathbf{G}$
.
Proposition 4.2. Let
$\mathbf{V}$
be an algebraic representation of
$Y$
. Then there exist a coset algebraic representation
$\mathbf{G}/\mathbf{H}$
and a morphism of representations from
$\mathbf{G}/\mathbf{H}$
to
$\mathbf{V}$
, that is, a
$k$
-
$\mathbf{G}$
-algebraic map
$i:\mathbf{G}/\mathbf{H}\rightarrow \mathbf{V}$
such that
$\unicode[STIX]{x1D719}_{\mathbf{V}}=i\circ \unicode[STIX]{x1D719}_{\mathbf{G}/\mathbf{H}}$
.
This follows essentially from Proposition 2.1 together with the well-known argument given in [Reference ZimmerZim84, Lemma 5.2.11]. For the reader’s convenience we reproduce this argument below.
Proof. Denote
$V=\mathbf{V}(k)$
and
$G=\mathbf{G}(k)$
. By Proposition 2.1 we know that every
$G$
-orbit is locally closed in
$G$
. Consider the orbit space
$V/G$
endowed with the quotient topology and Borel structure. The map
$Y\rightarrow V\rightarrow V/G$
is Borel. We push the measure class given on
$Y$
and obtain a measure class on
$V/G$
. Since
$V$
is Polish it has a countable basis, thus so does
$V/G$
. Let
$\{U_{n}~|~n\in \mathbb{N}\}$
be sequence of subsets of
$V/G$
consisting of the elements of a countable basis and their complements. Set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU6.gif?pub-status=live)
Then
$U$
has a full measure and, in particular, is non-empty. We claim that
$U$
is a singleton. Indeed, since every
$G$
-orbit is locally closed in
$V$
, the quotient topology of
$V/G$
is
$T_{0}$
, thus if
$U$
contained two distinct points we could find a basis set
$U_{n}$
which separates them, but by the ergodicity of
$G$
on
$Y$
either it or its complement would be of full measure, which contradicts the definition of
$U$
.
Fixing
$v\in V$
which is in the preimage of
$U$
, we conclude that
$\unicode[STIX]{x1D719}(X)$
is essentially contained in
$Gv$
. Let
$\mathbf{H}<\mathbf{G}$
be the stabilizer of
$v$
and
$H=\mathbf{H}(k)$
. By Proposition 2.1 we obtain a
$k$
-algebraic map
$i:\mathbf{G}/\mathbf{H}\rightarrow \mathbf{V}$
whose restriction to
$G/H$
gives a homeomorphism with the orbit
$Gv$
. We let
$\unicode[STIX]{x1D719}_{\mathbf{G}/\mathbf{H}}=(i|_{G/H})^{-1}\circ \unicode[STIX]{x1D719}$
. By extending the codomain of
$\unicode[STIX]{x1D719}_{\mathbf{G}/\mathbf{H}}$
to
$\mathbf{G}/\mathbf{H}(k)$
via the embedding
$G/H{\hookrightarrow}\mathbf{G}/\mathbf{H}(k)$
, we are done.◻
In [Reference ZimmerZim84, Definition 9.2.2], following Mackey, Zimmer defined the ‘algebraic hull of a cocycle’. We will not discuss this notion here, but we do point out its close relation with the following theorem (to be precise, it coincides with the group
$\mathbf{H}_{0}$
appearing in the proof below).
Theorem 4.3 (Cf. [Reference ZimmerZim84, Proposition 9.2.1]).
The category of algebraic representations of
$Y$
has an initial object. Moreover, there exists an initial object which is a coset algebraic representation.
Proof. We consider the collection
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU7.gif?pub-status=live)
This is a non-empty collection as it contains
$\mathbf{G}$
. By the Neotherian property, this collection contains a minimal element. We choose such a minimal element
$\mathbf{H}_{0}$
and fix corresponding
$\unicode[STIX]{x1D719}_{0}:Y\rightarrow (\mathbf{G}/\mathbf{H}_{0})(k)$
. We argue to show that this coset representation is the required initial object.
Fix any algebraic representation of
$Y$
,
$\mathbf{V}$
. It is clear that, if it exists, a morphism of algebraic representations from
$\mathbf{G}/\mathbf{H}_{0}$
to
$\mathbf{V}$
is unique, as two different
$\mathbf{G}$
-maps
$\mathbf{G}/\mathbf{H}_{0}\rightarrow \mathbf{V}$
agree nowhere. All that remains is to show existence. To this end, we consider the product representation
$\mathbf{V}\times \mathbf{G}/\mathbf{H}_{0}$
given by
$\unicode[STIX]{x1D719}=\unicode[STIX]{x1D719}_{\mathbf{V}}\times \unicode[STIX]{x1D719}_{0}$
. Applying Proposition 4.2 to this product representation, we obtain the commutative diagram
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqn1.gif?pub-status=live)
By the minimality of
$\mathbf{H}_{0}$
, the
$\mathbf{G}$
-morphism
$p_{2}\circ i:\mathbf{G}/\mathbf{H}\rightarrow \mathbf{G}/\mathbf{H}_{0}$
must be a
$k$
-isomorphism. We thus obtain the
$k$
-
$\mathbf{G}$
-morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU8.gif?pub-status=live)
Definition 4.4 (Algebraic gate).
With a slight abuse of terminology (as it is not canonical), we call (a choice of) a coset representation
$Y\rightarrow \mathbf{G}/\mathbf{H}(k)$
which is initial as an algebraic representation of
$Y$
, the algebraic gate of
$Y$
. If
$\mathbf{H}\lneq \mathbf{G}$
we say that the algebraic gate of
$Y$
is non-trivial.
The notion of the algebraic gate and the applications that we will derive below are only interesting if the gate is non-trivial. The following theorem gives a criterion for non-triviality.
Theorem 4.5. Assume the
$S$
-Lebesgue space
$Y$
is both amenable and metrically ergodic. Assume the
$k$
-algebraic group
$\mathbf{G}$
is connected,
$k$
-simple and adjoint and assume that
$\unicode[STIX]{x1D70C}(S)$
is unbounded in
$\mathbf{G}(k)$
. Then the gate of
$Y$
is non-trivial.
Proof. Using the amenability of
$Y$
, it follows from [Reference Bader, Duchesne and LécureuxBDL17, Corollary 1.17] that if the gate of
$Y$
is non-trivial then there exist a separable metric space
$V$
on which
$\mathbf{G}(k)$
acts by isometries and with bounded stabilizers and an
$S$
-equivariant map
$Y\rightarrow V$
. But the latter possibility is ruled out by the assumption that
$Y$
is
$S$
-metrically ergodic and the assumption that
$\unicode[STIX]{x1D70C}(S)<\mathbf{G}(k)$
is unbounded.
Before proceeding to our next theorem, let us state without proof the following proposition which provides an identification of
$\operatorname{Aut}_{\mathbf{G}}(\mathbf{G}/\mathbf{H})$
that we will keep using throughout the paper. The proposition is well known and easy to prove.
Proposition 4.6. Fix a
$k$
-subgroup
$\mathbf{H}<\mathbf{G}$
and denote
$\mathbf{N}=N_{\mathbf{G}}(\mathbf{H})$
. This is again a
$k$
-subgroup. Any element
$n\in \mathbf{N}$
gives a
$\mathbf{G}$
-automorphism of
$\mathbf{G}/\mathbf{H}$
by
$g\mathbf{H}\mapsto gn^{-1}\mathbf{H}$
. The homomorphism
$\mathbf{N}\rightarrow \operatorname{Aut}_{\mathbf{G}}(\mathbf{G}/\mathbf{H})$
thus obtained is onto and its kernel is
$\mathbf{H}$
. Under the obtained identification
$\mathbf{N}/\mathbf{H}\simeq \operatorname{Aut}_{\mathbf{G}}(\mathbf{G}/\mathbf{H})$
, the
$k$
-points of the
$k$
-group
$\mathbf{N}/\mathbf{H}$
are identified with the
$k$
-
$\mathbf{G}$
-automorphisms of
$\mathbf{G}/\mathbf{H}$
.
The following theorem provides a most useful tool in the study of rigidity.
Theorem 4.7. Assume
$\unicode[STIX]{x1D719}:Y\rightarrow \mathbf{G}/\mathbf{H}(k)$
is the algebraic gate of
$Y$
. Let
$S^{\prime }$
be an lcsc group which acts on
$Y$
commuting with the
$S$
-action. Then there exists a unique homomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU9.gif?pub-status=live)
which turns
$\unicode[STIX]{x1D719}$
into an
$S\times S^{\prime }$
-equivariant map, where
$S\times S^{\prime }$
acts on the codomain via
$\unicode[STIX]{x1D70C}\times \unicode[STIX]{x1D70C}^{\prime }$
. This homomorphism
$\unicode[STIX]{x1D70C}^{\prime }$
is continuous.
Proof. For a given
$s^{\prime }\in S^{\prime }$
we consider the diagram
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqn2.gif?pub-status=live)
where we denote by
$\unicode[STIX]{x1D70C}^{\prime }(s^{\prime })$
the dotted arrow, which is the unique
$k$
-algebraic
$\mathbf{G}$
-equivariant morphism
$\mathbf{G}/\mathbf{H}\rightarrow \mathbf{G}/\mathbf{H}$
given by the fact that
$\unicode[STIX]{x1D719}$
is a gate map and
$\unicode[STIX]{x1D719}\circ s^{\prime }$
is an algebraic representation of
$Y$
. By the uniqueness of the dotted arrow, it is easily checked that the correspondence
$s^{\prime }\mapsto \unicode[STIX]{x1D70C}^{\prime }(s^{\prime })$
forms a homomorphism from
$S^{\prime }$
to the group of
$k$
-
$\mathbf{G}$
-automorphisms of
$\mathbf{G}/\mathbf{H}$
which we identify with
$N_{\mathbf{G}}(\mathbf{H})/\mathbf{H}(k)$
using Proposition 4.6. All that remains is to check the continuity of
$\unicode[STIX]{x1D70C}^{\prime }$
.
To simplify the notation we let
$V=\mathbf{G}/\mathbf{H}(k)$
,
$M=N_{\mathbf{G}}(\mathbf{H})/\mathbf{H}(k)$
and
$U=\operatorname{L}^{0}(Y,V)$
. We endow
$U$
with the action of
$M$
by postcomposition, and the action of
$S^{\prime }$
by precomposition. By the fact that
$M$
acts freely on
$V$
, we obtain that
$M$
acts freely on
$U$
as well. Using Proposition 2.1, [Reference ZimmerZim84, Proposition 3.3.1] gives that the
$M$
-action on
$U$
has locally closed orbits.Footnote
1
It follows that the
$M$
-orbit map
$M\rightarrow U$
,
$m\mapsto m\circ \unicode[STIX]{x1D719}$
is a homeomorphism onto its image,
$M\unicode[STIX]{x1D719}$
. We let
$\unicode[STIX]{x1D6FC}:M\unicode[STIX]{x1D719}\rightarrow M$
be its inverse.
By the fact that the map
$S^{\prime }\times Y\rightarrow V$
,
$(s^{\prime },y)\mapsto \unicode[STIX]{x1D719}(s^{\prime }y)$
is almost everywhere defined and measurable, we obtain that the associated map
$\unicode[STIX]{x1D6FD}:S^{\prime }\rightarrow U$
,
$s^{\prime }\mapsto \unicode[STIX]{x1D719}\circ s^{\prime }$
is almost everywhere defined and measurable; see [Reference MargulisMar91, ch. VII, Lemma 1.3]. Since by the definition of
$\unicode[STIX]{x1D70C}^{\prime }$
,
$\unicode[STIX]{x1D719}\circ s^{\prime }=\unicode[STIX]{x1D70C}^{\prime }(s^{\prime })\circ \unicode[STIX]{x1D719}$
, we conclude that
$\unicode[STIX]{x1D70C}^{\prime }$
agrees almost everywhere with
$\unicode[STIX]{x1D6FC}\circ \unicode[STIX]{x1D6FD}$
which is almost everywhere defined and measurable. It follows that
$\unicode[STIX]{x1D70C}^{\prime }$
is measurable. By [Reference RosendalRos09, Lemma 2.1] we conclude that
$\unicode[STIX]{x1D70C}^{\prime }$
is a continuous homomorphism.◻
5 Extension of homomorphisms defined on dense subgroups
Throughout this section we fix:
∙ lcsc groups
$S$ and
$S^{\prime }$ and a continuous homomorphism
$\unicode[STIX]{x1D703}:S\rightarrow S^{\prime }$ such that
$\unicode[STIX]{x1D703}(S)$ is dense in
$S^{\prime }$ ;
∙ a field
$k$ with a non-trivial absolute value which is separable and complete (as a metric space);
∙ a
$k$ -algebraic group
$\mathbf{G}$ ;
∙ a continuous homomorphism
$\unicode[STIX]{x1D70C}:S\rightarrow \mathbf{G}(k)$ , where
$\mathbf{G}(k)$ is regarded as a Polish group (see Proposition 2.1), such that
$\unicode[STIX]{x1D70C}(S)$ is Zariski dense in
$\mathbf{G}$ .
We will explain how under some assumptions the homomorphism
$\unicode[STIX]{x1D70C}$
extends to
$S^{\prime }$
via
$\unicode[STIX]{x1D703}$
. The main result of this section is the following theorem.
Theorem 5.1. Assume
$\mathbf{G}$
is a connected, adjoint,
$k$
-simple group and let
$\mathbf{V}$
be a
$k$
-
$\mathbf{G}$
-variety which has no
$\mathbf{G}$
-fixed point. Let
$Z$
be an
$S^{\prime }$
-Lebesgue space. Consider
$Z$
as an
$S$
-Lebesgue space via
$\unicode[STIX]{x1D703}$
and assume that there exists an
$S$
-equivariant Lebesgue map
$\unicode[STIX]{x1D70E}:Z\rightarrow \mathbf{V}(k)$
. Then there exists a continuous homomorphism
$\bar{\unicode[STIX]{x1D70C}}:S^{\prime }\rightarrow \mathbf{G}(k)$
such that
$\unicode[STIX]{x1D70C}=\bar{\unicode[STIX]{x1D70C}}\circ \unicode[STIX]{x1D703}$
.
The proof will rely on the following lemma.
Lemma 5.2. Assume that in addition to
$\unicode[STIX]{x1D70C}:S\rightarrow \mathbf{G}(k)$
we are given a continuous homomorphism
$\unicode[STIX]{x1D70C}^{\prime }:S^{\prime }\rightarrow \mathbf{G}(k)$
and a Lebesgue map
$\unicode[STIX]{x1D719}:S^{\prime }\rightarrow \mathbf{G}(k)$
which is
$S\times S^{\prime }$
-equivariant with respect to the left
$S$
- and right
$S^{\prime }$
-actions on
$S^{\prime }$
and their corresponding actions on
$\mathbf{G}(k)$
via
$\unicode[STIX]{x1D70C}\times \unicode[STIX]{x1D70C}^{\prime }$
. Then
$\unicode[STIX]{x1D70C}=\operatorname{inn}(g)\circ \unicode[STIX]{x1D70C}^{\prime }\circ \unicode[STIX]{x1D703}$
for some
$g\in \mathbf{G}(k)$
, where
$\operatorname{inn}(g)$
denotes the corresponding inner automorphism of
$\mathbf{G}(k)$
.
Proof of the lemma.
The map
$\unicode[STIX]{x1D713}:S^{\prime }\rightarrow \mathbf{G}(k)$
given by
$s^{\prime }\mapsto \unicode[STIX]{x1D719}(s^{\prime })\unicode[STIX]{x1D70C}^{\prime }(s^{\prime })^{-1}$
is
$S^{\prime }$
-invariant, hence constant almost everywhere. We let
$g$
be its essential value. Given
$s\in S$
, we pick
$s^{\prime }\in S^{\prime }$
such that both
$s^{\prime }$
and
$\unicode[STIX]{x1D703}(s)s^{\prime }$
are both
$\unicode[STIX]{x1D719}$
and
$\unicode[STIX]{x1D713}$
-generic. We obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU10.gif?pub-status=live)
and conclude that
$\unicode[STIX]{x1D70C}(s)=\operatorname{inn}(g)\circ \unicode[STIX]{x1D70C}^{\prime }\circ \unicode[STIX]{x1D703}(s)$
.◻
Proof of Theorem 5.1.
First, we replace the unknown
$S^{\prime }$
-Lebesgue space
$Z$
with the group
$S^{\prime }$
endowed with its Haar measure. We do this by restricting
$\unicode[STIX]{x1D70E}$
to a generic orbit. More pedantically, we argue as follows. We consider the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU11.gif?pub-status=live)
and use Fubini’s theorem in order to fix a generic point
$z\in Z$
such that the map
$s^{\prime }\rightarrow \unicode[STIX]{x1D70E}(s^{\prime }z)$
is almost everywhere defined on
$S^{\prime }$
. We observe that the latter map
$S^{\prime }\rightarrow \mathbf{V}(k)$
is
$S$
-equivariant. Replacing
$\unicode[STIX]{x1D70E}$
with this map, we assume, as we may, that
$Z=S^{\prime }$
and
$\unicode[STIX]{x1D70E}:S^{\prime }\rightarrow \mathbf{V}$
is an
$S$
-equivariant Lebesgue map.
Note that
$S^{\prime }$
is
$S$
-ergodic, as
$\unicode[STIX]{x1D703}(S)$
is dense in
$S^{\prime }$
. Therefore we may use § 4 in the setting
$Y=S^{\prime }$
and view
$\unicode[STIX]{x1D70E}$
as an algebraic representation of the
$S$
-space
$S^{\prime }$
. By Theorem 4.3 we obtain the algebraic gate of
$S^{\prime }$
,
$\unicode[STIX]{x1D719}:S^{\prime }\rightarrow \mathbf{G}/\mathbf{H}(k)$
for some
$k$
-algebraic group
$\mathbf{H}\leqslant \mathbf{G}$
. Since there exists a
$\mathbf{G}$
-equivariant
$k$
-morphism
$\mathbf{G}/\mathbf{H}\rightarrow \mathbf{V}$
and
$\mathbf{V}$
has no
$\mathbf{G}$
-fixed point, we conclude that
$\mathbf{H}\lneq \mathbf{G}$
.
We now consider the right action of the group
$S^{\prime }$
on the space
$S^{\prime }$
and conclude by Theorem 4.7 the existence of a continuous homomorphism
$\unicode[STIX]{x1D70C}^{\prime }:S^{\prime }\rightarrow \mathbf{N}/\mathbf{H}(k)$
, where
$\mathbf{N}=N_{\mathbf{G}}(\mathbf{H})$
, such that
$\unicode[STIX]{x1D719}$
is
$S\times S^{\prime }$
-equivariant via
$\unicode[STIX]{x1D70C}\times \unicode[STIX]{x1D70C}^{\prime }$
.
We claim that
$\mathbf{H}=\{e\}$
. Assume not. Then by the fact that
$\mathbf{G}$
is simple we conclude that
$\mathbf{N}\lneq \mathbf{G}$
. Composing
$\unicode[STIX]{x1D719}$
with the map
$\mathbf{G}/\mathbf{H}(k)\rightarrow \mathbf{G}/\mathbf{N}(k)$
, we obtain a map
$S^{\prime }\rightarrow \mathbf{G}/\mathbf{N}(k)$
which is left
$S$
-equivariant and right
$S^{\prime }$
-invariant. By its right
$S^{\prime }$
-invariance, it is essentially constant. By the left
$S$
-equivariance, its essential image is
$\unicode[STIX]{x1D70C}(S)$
-invariant, hence also
$\mathbf{G}(k)$
-invariant, as
$\unicode[STIX]{x1D70C}(S)$
is Zariski dense in
$\mathbf{G}$
. This is absurd, thus the claim is proven.
It follows from the claim that
$\mathbf{N}=\mathbf{G}$
and thus
$\unicode[STIX]{x1D70C}^{\prime }$
is a continuous homomorphism from
$S^{\prime }$
to
$\mathbf{G}(k)$
and
$\unicode[STIX]{x1D719}$
is a map from
$S^{\prime }$
to
$\mathbf{G}(k)$
which is
$S\times S^{\prime }$
-equivariant via
$\unicode[STIX]{x1D70C}\times \unicode[STIX]{x1D70C}^{\prime }:S\times S^{\prime }\rightarrow \mathbf{G}(k)\times \mathbf{G}(k)$
. We are in a situation to apply Lemma 5.2 and obtain
$g\in \mathbf{G}(k)$
such that
$\unicode[STIX]{x1D70C}=\operatorname{inn}(g)\circ \unicode[STIX]{x1D70C}^{\prime }\circ \unicode[STIX]{x1D703}$
. By setting
$\bar{\unicode[STIX]{x1D70C}}=\operatorname{inn}(g)\circ \unicode[STIX]{x1D70C}^{\prime }$
, we are done.◻
6 Proof of Theorem 1.2
Below we give a proof of the main theorem. A priori, the valued field
$k$
is assumed merely to be a complete field, but by the fact that
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4})$
is unbounded we can assume that the absolute value on
$k$
is non-trivial. Further, by the countability of
$\unicode[STIX]{x1D6E4}$
, we may replace
$k$
with a complete and separable (in the topological sense) subfield
$k^{\prime }$
such that
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4})\subset \mathbf{G}(k^{\prime })$
. Therefore we henceforth assume that the given absolute value on the field
$k$
is non-trivial and that
$k$
is complete and separable as a metric space. We start with the proof of the existence of a continuous homomorphism
$\bar{\unicode[STIX]{x1D70C}}:T\rightarrow \mathbf{G}(k)$
; the uniqueness will follow from the general Lemma 6.1 given below.
Using Theorem 3.3, for each
$T_{i}$
one can choose a Lebesgue
$T_{i}$
-space
$B_{i}$
which is amenable and metrically ergodic. By Lemma 3.6 the product
$B=B_{1}\times \cdots \times B_{n}$
is amenable and metrically ergodic as a
$T$
-Lebesgue space, and Lemma 3.5 implies that
$B$
is also amenable and metrically ergodic as a
$\unicode[STIX]{x1D6E4}$
-Lebesgue space.
We write
$[n]=\{1,\ldots ,n\}$
and for a subset
$I\subset [n]$
denote by
$T_{I}=\prod _{i\in I}T_{i}$
the factor of the group
$T$
, and by
$B_{I}=\prod _{i\in I}B_{i}$
the measurable factor of the Lebesgue space
$B$
. By convention,
$T_{\emptyset }=\{e\}$
and
$B_{\emptyset }=\{\ast \}$
.
Consider pairs
$(i,J)$
where
$i\in [n]$
,
$J\subset [n]\setminus \{i\}$
, and view
$T_{i}\times B_{J}$
as a Lebesgue space with commuting actions of
$\unicode[STIX]{x1D6E4}$
and
$T_{i}$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU12.gif?pub-status=live)
Lemma 3.8 shows that the
$\unicode[STIX]{x1D6E4}$
-action on
$T_{i}\times B_{[n]\setminus \{i\}}$
, and hence the
$\unicode[STIX]{x1D6E4}$
-action on its quotient
$T_{i}\times B_{J}$
, is ergodic. Therefore we can apply the considerations of § 4 to the group
$S=\unicode[STIX]{x1D6E4}$
acting on the space
$Y=T_{i}\times B_{J}$
. Theorem 4.3 provides the corresponding algebraic gate: a minimal
$k$
-algebraic subgroup
$\mathbf{H}_{i,J}<\mathbf{G}$
for which there exists a
$\unicode[STIX]{x1D6E4}$
-map
$\unicode[STIX]{x1D719}_{i,J}:T_{i}\times B_{J}\rightarrow \mathbf{G}/\mathbf{H}_{i,J}(k)$
.
Theorem 4.5 ensures that the algebraic gate of
$B=B_{[n]}$
is non-trivial: there is a proper
$k$
-subgroup
$\mathbf{H}\neq \mathbf{G}$
and a measurable
$\unicode[STIX]{x1D6E4}$
-map
$B\rightarrow \mathbf{G}/\mathbf{H}(k)$
. For any
$i\in [n]$
we can apply Lemma 3.9 in the setting
$S=T_{i}$
,
$X=B_{i}$
and
$Z=B_{[n]\setminus \{i\}}$
to obtain a measurable
$\unicode[STIX]{x1D6E4}$
-map
$T_{i}\times B_{[n]\setminus \{i\}}\rightarrow \mathbf{G}/\mathbf{H}(k)$
. It follows that
$\mathbf{H}_{i,[n]\setminus \{i\}}$
can be conjugated into
$\mathbf{H}$
, and is therefore a proper subgroup of
$\mathbf{G}$
.
Let
$J\subset [n]$
be a subset of minimal size
$|J|$
for which there exists
$i\in [n]\setminus J$
so that the algebraic gate of
$T_{i}\times B_{J}$
is non-trivial:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU13.gif?pub-status=live)
Since
$T_{i}$
acts on
$T_{i}\times B_{J}$
by right translation on the first coordinate, commuting with the ergodic
$\unicode[STIX]{x1D6E4}$
-action, Lemma 4.7 with
$S=\unicode[STIX]{x1D6E4}$
,
$Y=T_{i}\times B_{J}$
and
$S^{\prime }=T_{i}$
yields a continuous homomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU14.gif?pub-status=live)
so that
$\unicode[STIX]{x1D719}_{i,J}(\unicode[STIX]{x1D6FE}xt,\unicode[STIX]{x1D6FE}b)=\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6FE})\unicode[STIX]{x1D719}_{i,J}(x,b)\unicode[STIX]{x1D70E}(t)$
for almost every
$(x,b)\in T_{i}\times B_{J}$
. Let
$\mathbf{L}$
denote the smallest
$k$
-algebraic subgroup in
$\mathbf{N}_{\mathbf{G}}(\mathbf{H}_{i,J})$
for which
$\mathbf{L}/\mathbf{H}_{i,J}(k)$
contains the image
$\unicode[STIX]{x1D70E}(T_{i})$
.
Claim.
$\mathbf{L}=\mathbf{G}$
and
$\mathbf{H}_{i,J}=\{e\}$
.
Once the claim is proven, we obtain that there is a continuous homomorphism
$\unicode[STIX]{x1D70E}:T_{i}\rightarrow \mathbf{G}(k)$
with Zariski dense image, and a measurable map
$\unicode[STIX]{x1D719}:T_{i}\times B_{J}\rightarrow \mathbf{G}(k)$
so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU15.gif?pub-status=live)
for almost every
$(x,b)\in T_{i}\times B_{J}$
.
Proof. Assume
$\mathbf{L}\neq \mathbf{G}$
. The
$T_{i}$
-equivariance property of
$\unicode[STIX]{x1D719}_{i,J}$
implies that this map descends to a measurable
$\unicode[STIX]{x1D6E4}$
-map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU16.gif?pub-status=live)
Note that this is possible only if
$J\neq \emptyset$
. Pick
$j\in J$
and use Lemma 3.9 to obtain a measurable
$\unicode[STIX]{x1D6E4}$
-map
$\unicode[STIX]{x1D713}:T_{j}\times B_{J\setminus \{j\}}\rightarrow \mathbf{G}/\mathbf{L}(k)$
. Under the assumption that
$\mathbf{L}$
is a proper subgroup in
$\mathbf{G}$
, the gate of
$T_{j}\times B_{J\setminus \{j\}}$
is non-trivial, contradicting the minimality property of
$J\subset [n]$
.
Therefore
$\mathbf{L}=\mathbf{G}$
. This implies that
$\mathbf{N}_{\mathbf{G}}(\mathbf{H}_{i,J})=\mathbf{G}$
, meaning that
$\mathbf{H}_{i,J}$
is normal in the simple group
$\mathbf{G}$
. Since
$\mathbf{H}_{i,J}$
is known to be a proper subgroup, we have
$\mathbf{H}=\{e\}$
. Thus the previously constructed continuous homomorphism
$\unicode[STIX]{x1D70E}$
and the measurable map
$\unicode[STIX]{x1D719}=\unicode[STIX]{x1D719}_{i,J}$
range into
$\mathbf{G}(k)$
. Finally,
$\unicode[STIX]{x1D70E}(T_{i})$
is Zariski dense in
$\mathbf{G}$
because its Zariski closure is
$\mathbf{L}$
.◻
Claim. The
$\unicode[STIX]{x1D6E4}$
-map
$\unicode[STIX]{x1D719}:T_{i}\times B_{J}\rightarrow \mathbf{G}(k)$
descends to a
$\unicode[STIX]{x1D6E4}$
-map
$T_{i}\rightarrow \mathbf{G}(k)$
.
Proof. Let us show that for every
$j\in J$
for almost every
$t\in T_{j}$
one has
$\unicode[STIX]{x1D719}(x,b)=\unicode[STIX]{x1D719}(x,tb)$
almost everywhere on
$T_{i}\times B_{J}$
. Since
$T_{J}$
acts ergodically on
$B_{J}$
the claim would follow. The case of
$J=\emptyset$
being trivial, we assume
$J$
is non-empty, fix
$j\in J$
, and set
$K=J\setminus \{j\}$
. Choose almost every
$b\in B_{J}$
to create a
$\unicode[STIX]{x1D6E4}$
-equivariant lift by Lemma 3.9,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU17.gif?pub-status=live)
Consider the setting of § 4 for the group
$S=\unicode[STIX]{x1D6E4}\times T_{i}$
, the homomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU18.gif?pub-status=live)
and the Lebesgue
$S$
-space
$Y=T_{i}\times T_{j}\times B_{K}$
with the action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU19.gif?pub-status=live)
Note that the
$S$
-action on
$Y$
is ergodic, because the
$\unicode[STIX]{x1D6E4}$
-action on
$T_{j}\times B_{K}$
is ergodic by Lemma 3.8. Write
$\unicode[STIX]{x1D6E5}_{\mathbf{G}}<\mathbf{G}\times \mathbf{G}$
for the diagonal subgroup. The measurable map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU20.gif?pub-status=live)
is
$\unicode[STIX]{x1D70C}\times \unicode[STIX]{x1D70E}$
-equivariant. Thus the gate of
$Y$
is given by:
∙ a
$k$ -variety
$\mathbf{V}=\mathbf{G}\times \mathbf{G}/\mathbf{D}$ where
$\mathbf{D}$ is a
$k$ -algebraic subgroup
$\mathbf{D}<\unicode[STIX]{x1D6E5}_{\mathbf{G}}$ ;
∙ a measurable
$\unicode[STIX]{x1D6E4}\times T_{i}$ -map
$\hat{\unicode[STIX]{x1D713}}:Y\rightarrow \mathbf{V}(k)$ for which
$\unicode[STIX]{x1D713}=\unicode[STIX]{x1D70B}\circ \hat{\unicode[STIX]{x1D713}}$ where
$\unicode[STIX]{x1D70B}:\mathbf{V}(k)\rightarrow \mathbf{G}\times \mathbf{G}/\unicode[STIX]{x1D6E5}_{\mathbf{G}}(k)$ is the projection.
In fact, we must have
$\mathbf{D}=\unicode[STIX]{x1D6E5}_{\mathbf{G}}$
. Indeed, otherwise under the projection to the first factor
$\unicode[STIX]{x1D70B}_{1}:\mathbf{G}\times \mathbf{G}\rightarrow \mathbf{G}$
, the image
$\mathbf{M}=\unicode[STIX]{x1D70B}_{1}(\mathbf{D})$
would be a proper subgroup of
$\mathbf{G}$
, and the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20191206043038666-0363:S0010437X19007607:S0010437X19007607_eqnU21.gif?pub-status=live)
would factor through a measurable
$\unicode[STIX]{x1D6E4}$
-map
$T_{j}\times B_{K}\rightarrow \mathbf{G}/\mathbf{M}(k)$
, contrary to our assumption of minimality for
$J\subset [n]$
. Thus
$\mathbf{D}=\unicode[STIX]{x1D6E5}_{\mathbf{G}}$
and
$\hat{\unicode[STIX]{x1D713}}=\unicode[STIX]{x1D713}$
.
Consider the action of
$S^{\prime }=T_{j}$
on
$Y=T_{i}\times T_{j}\times B_{K}$
by
$s:(x,y,b)\mapsto (x,ys^{-1},b)$
. It commutes with the
$S=\unicode[STIX]{x1D6E4}\times T_{i}$
-action. We can apply Lemma 4.7 to obtain a continuous homomorphism
$\unicode[STIX]{x1D70F}:T_{j}\rightarrow N_{\mathbf{G}\times \mathbf{G}}(\unicode[STIX]{x1D6E5}_{\mathbf{G}})/\unicode[STIX]{x1D6E5}_{\mathbf{G}}(k)$
for which
$\unicode[STIX]{x1D713}=\hat{\unicode[STIX]{x1D713}}$
is
$T_{j}$
-equivariant. But as
$\mathbf{G}$
is center-free,
$\unicode[STIX]{x1D6E5}_{\mathbf{G}}$
is its own normalizer in
$\mathbf{G}\times \mathbf{G}$
. So
$\unicode[STIX]{x1D70F}$
is trivial, and therefore
$\unicode[STIX]{x1D713}$
is
$T_{j}$
-invariant. Since
$\unicode[STIX]{x1D713}$
was a lift of
$\unicode[STIX]{x1D719}:T_{i}\times B_{J}\rightarrow \mathbf{G}(k)$
for almost every
$b\in B_{J}$
, we proved that
$\unicode[STIX]{x1D719}$
is essentially
$T_{j}$
-invariant, and this is true for all
$j\in J$
. This completes the proof of the claim.◻
At this point we have a measurable map
$\unicode[STIX]{x1D719}:T_{i}\rightarrow \mathbf{G}(k)$
and a continuous homomorphism
$\unicode[STIX]{x1D70E}:T_{i}\rightarrow \mathbf{G}(k)$
so that
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6FE}t)=\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6FE})\unicode[STIX]{x1D719}(t)$
and
$\unicode[STIX]{x1D719}(tt^{\prime })=\unicode[STIX]{x1D719}(t)\unicode[STIX]{x1D70E}(t^{\prime })$
. Applying Lemma 5.2 with
$\unicode[STIX]{x1D703}:\unicode[STIX]{x1D6E4}\overset{}{\longrightarrow }T\overset{\unicode[STIX]{x1D70B}_{i}}{\longrightarrow }T_{i}$
, we deduce that for some
$g\in \mathbf{G}(k)$
the homomorphism
$\bar{\unicode[STIX]{x1D70C}}_{i}=\operatorname{inn}(g)\circ \unicode[STIX]{x1D70E}:T_{i}\overset{}{\longrightarrow }\mathbf{G}$
extends
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6E4}\rightarrow \mathbf{G}(k)$
in the sense that
$\bar{\unicode[STIX]{x1D70C}}_{i}\circ \unicode[STIX]{x1D703}=\unicode[STIX]{x1D70C}$
. This completes the proof of the existence of the continuous extension homomorphism
$\bar{\unicode[STIX]{x1D70C}}:T\rightarrow \mathbf{G}(k)$
.
The uniqueness part follows from the following general lemma, by which we conclude our proof.
Lemma 6.1. Let
$S$
be an lcsc group and
$\unicode[STIX]{x1D6E4}<S$
a lattice. Let
$k$
be a complete valued field and
$\mathbf{G}$
be a connected
$k$
-simple adjoint algebraic group. Let
$\unicode[STIX]{x1D70C}_{1},\unicode[STIX]{x1D70C}_{2}:S\rightarrow \mathbf{G}(k)$
be continuous homomorphisms, and assume that
$\unicode[STIX]{x1D70C}_{1}$
has an unbounded and Zariski dense image. If
$\unicode[STIX]{x1D70C}_{1}|_{\unicode[STIX]{x1D6E4}}=\unicode[STIX]{x1D70C}_{2}|_{\unicode[STIX]{x1D6E4}}$
then
$\unicode[STIX]{x1D70C}_{1}=\unicode[STIX]{x1D70C}_{2}$
.
Proof. As before, without loss of generality we assume that the absolute value on
$k$
is non-trivial and as a metric space
$k$
is complete and separable. We will prove the lemma by contradiction, assuming there exists
$s_{0}\in S$
with
$\unicode[STIX]{x1D70C}_{1}(s_{0})\neq \unicode[STIX]{x1D70C}_{2}(s_{0})$
.
We will first show that
$\unicode[STIX]{x1D70C}_{1}(\unicode[STIX]{x1D6E4})=\unicode[STIX]{x1D70C}_{2}(\unicode[STIX]{x1D6E4})$
is Zariski dense in
$\mathbf{G}$
. We denote by
$L_{1}$
the closure of
$\unicode[STIX]{x1D70C}_{1}(S)$
in
$\mathbf{G}(k)$
. By [Reference Bader, Duchesne and LécureuxBDL17, Proposition 1.9] there exists a
$k$
-subgroup
$\mathbf{H}_{1}<\mathbf{G}$
which is normalized by
$L_{1}$
such that
$L_{1}$
has a precompact image in
$N_{\mathbf{G}}(\mathbf{H}_{1})/\mathbf{H}_{1}(k)$
and such that for every
$k$
-
$\mathbf{G}$
-variety
$\mathbf{V}$
, any
$L_{1}$
-invariant finite measure is supported on the subvariety of
$\mathbf{H}_{1}$
-fixed points,
$\mathbf{V}^{\mathbf{H}_{1}}\cap \mathbf{V}(k)$
. As
$\unicode[STIX]{x1D70C}_{1}$
has a Zariski dense image,
$L_{1}$
is Zariski dense in
$\mathbf{G}$
and we obtain that
$\mathbf{H}_{1}$
is normal in
$\mathbf{G}$
. As
$\mathbf{G}$
is
$k$
-simple, we conclude that either
$\mathbf{H}_{1}=\{e\}$
or
$\mathbf{H}_{1}=\mathbf{G}$
. The first option is ruled out, as
$\unicode[STIX]{x1D70C}_{1}(S)$
is unbounded while
$L_{1}$
has a precompact image in
$N_{\mathbf{G}}(\mathbf{H}_{1})/\mathbf{H}_{1}(k)$
. Thus
$\mathbf{H}_{1}=\mathbf{G}$
. We denote by
$\mathbf{G}_{1}$
the Zariski closure of
$\unicode[STIX]{x1D70C}_{1}(\unicode[STIX]{x1D6E4})$
and let
$\mathbf{V}_{1}=\mathbf{G}/\mathbf{G}_{1}$
. The map
$\unicode[STIX]{x1D70C}_{1}$
gives rise to an obvious map
$S/\unicode[STIX]{x1D6E4}\rightarrow \mathbf{V}_{1}(k)$
. By pushing the
$S$
-invariant measure of
$S/\unicode[STIX]{x1D6E4}$
we obtain an
$L_{1}$
-invariant finite measure on
$\mathbf{V}_{1}(k)$
. It follows that this measure is supported on the subvariety of
$\mathbf{G}$
-fixed points. In particular, there exists a
$\mathbf{G}$
-fixed point in
$\mathbf{V}_{1}=\mathbf{G}/\mathbf{G}_{1}$
and we conclude that
$\mathbf{G}_{1}=\mathbf{G}$
. Thus, indeed,
$\unicode[STIX]{x1D70C}_{1}(\unicode[STIX]{x1D6E4})$
is Zariski dense in
$\mathbf{G}$
.
Next we consider the homomorphism
$R=\unicode[STIX]{x1D70C}_{1}\times \unicode[STIX]{x1D70C}_{2}:S\rightarrow \mathbf{G}(k)\times \mathbf{G}(k)$
, we let
$L$
be the closure of
$R(S)$
and let
$\mathbf{L}<\mathbf{G}\times \mathbf{G}$
be the Zariski closure of
$L$
. We note that, by the discussion above, the Zariski closure of
$R(\unicode[STIX]{x1D6E4})$
coincides with the diagonal group
$\unicode[STIX]{x1D6E5}_{\mathbf{G}}<\mathbf{G}\times \mathbf{G}$
. In particular,
$\unicode[STIX]{x1D6E5}_{\mathbf{G}}$
is contained in
$\mathbf{L}$
. We argue to show that
$\mathbf{L}=\mathbf{G}\times \mathbf{G}$
. The
$k$
-subgroup
$\mathbf{L}\cap (\mathbf{G}\times \{e\})$
is non-trivial, as the element
$\unicode[STIX]{x1D70C}_{1}(s_{0})\unicode[STIX]{x1D70C}_{2}(s_{0})^{-1}$
is contained in its group of
$k$
-points, since both
$(\unicode[STIX]{x1D70C}_{1}(s_{0}),\unicode[STIX]{x1D70C}_{2}(s_{0}))$
and
$(\unicode[STIX]{x1D70C}_{2}(s_{0}),\unicode[STIX]{x1D70C}_{2}(s_{0}))$
are elements of
$\mathbf{L}(k)$
. As this subgroup is normalized by
$\mathbf{L}$
, hence by
$\unicode[STIX]{x1D6E5}_{\mathbf{G}}$
, we obtain that this is a non-trivial normal
$k$
-subgroup of
$\mathbf{G}\times \{e\}$
. As
$\mathbf{G}\times \{e\}$
is
$k$
-simple, we conclude that
$\mathbf{L}\cap (\mathbf{G}\times \{e\})=\mathbf{G}\times \{e\}$
, that is,
$\mathbf{G}\times \{e\}<\mathbf{L}$
. As
$\mathbf{G}\times \mathbf{G}$
is generated by
$\mathbf{G}\times \{e\}$
and
$\unicode[STIX]{x1D6E5}_{\mathbf{G}}$
, we conclude that indeed
$\mathbf{L}=\mathbf{G}\times \mathbf{G}$
.
We use [Reference Bader, Duchesne and LécureuxBDL17, Proposition 1.9] again, this time for the
$k$
-group
$\mathbf{G}\times \mathbf{G}$
and the Zariski dense closed subgroup
$L<(\mathbf{G}\times \mathbf{G})(k)$
, and conclude having a normal
$k$
-subgroup
$\mathbf{H}<\mathbf{G}\times \mathbf{G}$
such that
$L$
has a precompact image in
$(\mathbf{G}\times \mathbf{G})/\mathbf{H}(k)$
and such that for every
$k$
-
$\mathbf{G}$
-variety
$\mathbf{V}$
, any
$L$
-invariant finite measure is supported on the subvariety of
$\mathbf{H}$
-fixed points,
$\mathbf{V}^{\mathbf{H}}\cap \mathbf{V}(k)$
. We consider
$\mathbf{V}=(\mathbf{G}\times \mathbf{G})/\unicode[STIX]{x1D6E5}_{\mathbf{G}}$
. The map
$R$
gives rise to an obvious map
$S/\unicode[STIX]{x1D6E4}\rightarrow \mathbf{V}(k)$
. By pushing the
$S$
-invariant measure of
$S/\unicode[STIX]{x1D6E4}$
we obtain an
$L$
-invariant finite measure on
$\mathbf{V}(k)$
. It follows that this measure is supported on the subvariety of
$\mathbf{H}$
-fixed points. Since the non-trivial normal
$k$
-subgroups of
$\mathbf{G}\times \mathbf{G}$
, namely
$\mathbf{G}\times \{e\}$
,
$\{e\}\times \mathbf{G}$
and
$\mathbf{G}\times \mathbf{G}$
itself, have no fixed points in
$\mathbf{V}$
, we conclude that
$\mathbf{H}=\{e\}$
. It follows that
$L$
is compact in
$\mathbf{G}\times \mathbf{G}$
, which contradicts the assumption that
$\unicode[STIX]{x1D70C}_{1}$
has an unbounded image.◻
7 Linearity criteria
This section is devoted to the proof of the following theorem.
Theorem 7.1. For any countable group
$\unicode[STIX]{x1D6EC}$
and a commensurated subgroup
$\unicode[STIX]{x1D6E4}<\unicode[STIX]{x1D6EC}$
the following properties are equivalent.
(1) For every field
$K$ , integer
$d$ and a group homomorphism
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6EC}\rightarrow \operatorname{GL}_{d}(K)$ ,
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$ is solvable by locally finite.
(2) For every finite index subgroup
$\unicode[STIX]{x1D6EC}^{\prime }<\unicode[STIX]{x1D6EC}$ , complete field with an absolute value
$k$ , connected adjoint
$k$ -simple algebraic group
$\mathbf{G}$ and Zariski dense group homomorphism
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6EC}^{\prime }\rightarrow \mathbf{G}(k)$ ,
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}\cap \unicode[STIX]{x1D6EC}^{\prime })$ is bounded in
$\mathbf{G}(k)$ .
Furthermore, if
$\unicode[STIX]{x1D6EC}$
is assumed to be finitely generated, the class of fields considered in (2) could be taken to be the class of local fields.
Taking
$\unicode[STIX]{x1D6E4}=\unicode[STIX]{x1D6EC}$
, we obtain the following corollary, which is an extension of the linearity criterion [Reference Bader, Caprace and LécureuxBCL19, Theorem A.1], not adhering to a finite generation assumption.
Corollary 7.2. For any countable group
$\unicode[STIX]{x1D6E4}$
the following properties are equivalent.
(1) For every field
$K$ , integer
$d$ and a group homomorphism
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6E4}\rightarrow \operatorname{GL}_{d}(K)$ ,
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$ is solvable by locally finite.
(2) For every finite index subgroup
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6E4}$ , complete field with an absolute value
$k$ , connected adjoint
$k$ -simple algebraic group
$\mathbf{G}$ and Zariski dense group homomorphism
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6E4}^{\prime }\rightarrow \mathbf{G}(k)$ ,
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}^{\prime })$ is bounded in
$\mathbf{G}(k)$ .
Furthermore, if
$\unicode[STIX]{x1D6E4}$
is assumed to be finitely generated, the class of fields considered in (2) could be taken to be the class of local fields.
Before proving Theorem 7.1 let us make some preliminary lemmas. The following important lemma is essentially due to Breuillard and Gelander, and it is an elaboration of an older lemma of Tits.
Lemma 7.3. Let
$K$
be a countable field,
$R<K$
a finitely generated subring and
$I\subset R$
an infinite subset. Then there exist a complete field with an absolute value
$k$
and an embedding
$K{\hookrightarrow}k$
such that the image of
$I$
in
$k$
is unbounded. Furthermore, if
$K$
is finitely generated as a field then
$k$
could be taken to be a local field.
Proof. The case where
$K$
is the fraction field of
$R$
is dealt with in [Reference Breuillard and GelanderBG07, Lemma 2.1]. If
$K$
is finitely generated, say by a finite set
$A$
, upon replacing
$R$
with the ring
$R^{\prime }$
generated by
$A\cup B$
where
$B$
is a finite set of generators of
$R$
, we are reduced to the previous case. The general case follows from the case where
$K$
is finitely generated as follows. Denote by
$K_{0}$
the fraction field of
$R$
and observe that it is finitely generated as a field. Let
$k_{0}$
be a corresponding local field such that there exists an embedding
$i:K_{0}{\hookrightarrow}k_{0}$
with
$i(I)$
unbounded. Let
$\bar{k}_{0}$
be an algebraic closure of
$k_{0}$
and note that the absolute value of
$k_{0}$
extends uniquely to an absolute value on
$\bar{k}_{0}$
[Reference Bosch, Güntzer and RemmertBGR84, § 3.2.4, Theorem 2]. Let
$k$
be the completion of
$\bar{k}_{0}$
with respect to this absolute value. Note that
$k$
is an algebraically closed field [Reference Bosch, Güntzer and RemmertBGR84, § 3.4.1, Proposition 3] which is of uncountable transcendental degree over its prime field. Denote by
$j$
the composition
$K_{0}{\hookrightarrow}k_{0}{\hookrightarrow}k$
and note that
$j(I)$
is unbounded. As
$K_{0}$
is countable,
$k$
is of uncountable transcendental degree over
$j(K_{0})$
. As
$K$
is countable and
$k$
is algebraically closed and of uncountable transcendental degree over
$j(K_{0})$
,
$j$
extends to an embedding
$K{\hookrightarrow}k$
.◻
Lemma 7.4. Let
$K$
be a field and
$\mathbf{G}$
a
$K$
-algebraic group. Let
$\unicode[STIX]{x1D6E4}$
be a subgroup of
$\mathbf{G}(K)$
which is not locally finite. Then there exists a complete field with an absolute value
$k$
and an embedding
$K{\hookrightarrow}k$
such that the image of
$\unicode[STIX]{x1D6E4}$
is unbounded in
$\mathbf{G}(k)$
. Furthermore, if the field
$K$
is finitely generated then the field
$k$
could be taken to be a local field.
Proof. We let
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6E4}$
be an infinite finitely generated group. We fix an injective
$K$
-representation
$\mathbf{G}\rightarrow \operatorname{GL}_{n}$
, let
$I\subset K$
be the set of matrix coefficients of
$\unicode[STIX]{x1D6E4}^{\prime }$
and let
$R<K$
be the subring generated by
$I$
. Note that
$I$
is infinite and
$R$
is finitely generated. We let
$K{\hookrightarrow}k$
be a field extension as provided by Lemma 7.3 and note that the image of
$\unicode[STIX]{x1D6E4}$
is unbounded in
$\mathbf{G}(k)$
, as the image of
$\unicode[STIX]{x1D6E4}^{\prime }$
is unbounded in
$\operatorname{GL}_{n}(k)$
.◻
Lemma 7.5. Fix a field
$K$
and a connected adjoint
$K$
-simple algebraic group
$\mathbf{G}$
. Let
$\unicode[STIX]{x1D6EC}<\mathbf{G}(K)$
be a Zariski dense subgroup. Let
$\unicode[STIX]{x1D6E4}<\unicode[STIX]{x1D6EC}$
a commensurated subgroup. Assume
$\unicode[STIX]{x1D6E4}$
is infinite. Then
$\unicode[STIX]{x1D6E4}$
is Zariski dense in
$\mathbf{G}$
.
Proof. Consider the Zariski closure
$\mathbf{H}=\overline{\unicode[STIX]{x1D6E4}}^{Z}$
and let
$\mathbf{H}^{0}$
be its identity connected component. By [Reference BorelBor91, AG, Theorem 14.4]
$\mathbf{H}$
is defined over
$K$
and
$\unicode[STIX]{x1D6E4}<\mathbf{H}(K)$
. We set
$\unicode[STIX]{x1D6E4}^{\prime }=\unicode[STIX]{x1D6E4}\cap \mathbf{H}^{0}(K)$
. As
$\mathbf{H}^{0}(K)<\mathbf{H}(K)$
is of finite index,
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6E4}$
is of finite index. It follows that
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6EC}$
is commensurated. Thus
$\mathbf{H}^{0}<\mathbf{G}$
is commensurated. As the connected group
$\mathbf{H}^{0}$
has no Zariski closed subgroups of finite index, we conclude that
$\mathbf{H}^{0}$
is normal in
$\mathbf{G}$
. By the assumption that
$\unicode[STIX]{x1D6E4}$
is infinite we obtain that
$\mathbf{H}^{0}$
is non-trivial. By the simplicity of
$\mathbf{G}$
we conclude that
$\mathbf{H}^{0}=\mathbf{G}$
. In particular,
$\mathbf{H}=\mathbf{G}$
. Thus
$\unicode[STIX]{x1D6E4}$
is indeed Zariski dense in
$\mathbf{G}$
.◻
Lemma 7.6. Let
$k$
be a field with absolute value and
$\mathbf{G}$
a semisimple algebraic group defined over
$k$
. Let
$R$
be a topological group which is separable and amenableFootnote
2
and let
$\unicode[STIX]{x1D70C}:R\rightarrow \mathbf{G}(k)$
be a continuous homomorphism with Zariski dense image. Then
$\unicode[STIX]{x1D70C}(R)$
is bounded in
$\mathbf{G}(k)$
.
This lemma is essentially claimed in [Reference Bader, Duchesne and LécureuxBDL17, Corollary 1.18] and it indeed follows from the considerations of [Reference Bader, Duchesne and LécureuxBDL17]. However, [Reference Bader, Duchesne and LécureuxBDL17, Corollary 1.18] is not a formal corollary of [Reference Bader, Duchesne and LécureuxBDL17, Corollary 1.17] as claimed in [Reference Bader, Duchesne and LécureuxBDL17], as the latter considers merely locally compact groups. We thank the anonymous referee for spotting this gap, which is closed by the proof below.
Proof. We assume, as we may, that
$k$
is non-discrete, complete and separable as a metric space. We argue as in the proof of [Reference Bader, Duchesne and LécureuxBDL17, Theorem 6.1]. By [Reference Bader, Duchesne and LécureuxBDL17, Lemma 4.5] we may assume
$\mathbf{G}$
is adjoint and, upon replacing
$R$
with the product of its projections to the various simple factors, the problem is easily reduced to the case where
$\mathbf{G}$
is
$k$
-simple. We thus assume
$\mathbf{G}$
is indeed
$k$
-simple.
Using [Reference BorelBor91, Proposition 1.10], there exists a
$k$
-closed immersion of
$\mathbf{G}$
into some
$\operatorname{GL}_{n}$
. Fixing a minimum dimensional such representation, using the fact that
$\mathbf{G}$
is
$k$
-simple, we assume, as we may, that this representation is
$k$
-irreducible. By the fact that
$\mathbf{G}$
is adjoint, the associated morphism
$\mathbf{G}\rightarrow \operatorname{PGL}_{n}$
is a closed immersion as well. We will denote for convenience
$E=k^{n}$
. Via this representation,
$\mathbf{G}(k)$
acts continuously and faithfully on the metric space of homothety classes of norms,
$I(E)$
, and on the compact space of homothety classes of seminorms,
$S(E)$
, introduced in [Reference Bader, Duchesne and LécureuxBDL17, § 5].
Using the amenability of
$R$
, there exists an
$R$
-invariant ergodic probability measure
$\unicode[STIX]{x1D707}$
on
$S(E)$
which we now fix. This measure is also fixed by the closure of
$\unicode[STIX]{x1D70C}(R)$
, thus, upon replacing
$R$
with the closure of
$\unicode[STIX]{x1D70C}(R)$
, we assume, as we may, that
$R$
is a closed subgroup of
$\mathbf{G}(k)$
. By [Reference Bader, Duchesne and LécureuxBDL17, Proposition 5.4], there is a
$\mathbf{G}(k)$
-invariant measurable partition
$S(E)=\bigcup _{d=0}^{n-1}S_{d}(E)$
, given by the dimension of the kernels of the seminorms. By ergodicity
$\unicode[STIX]{x1D707}$
is supported on
$S_{d}(E)$
for some
$0\leqslant d\leqslant n-1$
.
We assume now
$d>0$
. We denote by
$\operatorname{Gr}_{d}(E)$
the Grassmannian of
$d$
-dimensional
$k$
-subspaces of
$E$
and consider the map
$S_{d}(E)\rightarrow \operatorname{Gr}_{d}(E)$
taking a seminorm to its kernel. This map is measurable by [Reference Bader, Duchesne and LécureuxBDL17, Proposition 5.4]. Pushing forward the measure
$\unicode[STIX]{x1D707}$
, we obtain an
$R$
-invariant measure
$\unicode[STIX]{x1D708}$
on
$\operatorname{Gr}_{d}(E)$
. By [Reference Bader, Duchesne and LécureuxBDL17, Proposition 1.9] there exists a
$k$
-subgroup
$\mathbf{H}<\mathbf{G}$
which is normalized by
$R$
such that the image of
$R$
is precompact in
$N_{\mathbf{G}}(\mathbf{H})/\mathbf{H}(k)$
and
$\unicode[STIX]{x1D708}$
is supported on the set of
$\mathbf{H}$
-fixed points in
$\operatorname{Gr}_{d}(E)$
. As
$R<\mathbf{G}$
is Zariski dense and
$\mathbf{G}$
is
$k$
-simple, we deduce that either
$\mathbf{H}=\{e\}$
or
$\mathbf{H}=\mathbf{G}$
. By the irreducibility of our chosen representation there are no
$\mathbf{G}$
-fixed points in
$\operatorname{Gr}(E)$
and we conclude that
$\mathbf{H}=\{e\}$
. It follows that
$R$
is compact in
$\mathbf{G}(k)$
and, in particular, is bounded.
Next we assume
$d=0$
, that is,
$\unicode[STIX]{x1D707}$
is supported on
$I(E)$
. By [Reference Bader, Duchesne and LécureuxBDL17, Lemma 5.1]
$I(E)$
is a metric space on which
$\mathbf{G}(k)$
acts continuously and isometrically and the stabilizers of bounded subsets of
$I(E)$
are bounded in
$\mathbf{G}(k)$
. We find a ball
$B\subset I(E)$
such that
$\unicode[STIX]{x1D707}(B)>1/2$
. It follows that for any
$g\in R$
,
$gB$
intersects
$B$
. Thus the set
$RB$
is bounded in
$I(E)$
. It follows that the stabilizer in
$\mathbf{G}(k)$
of the set
$RB$
is bounded. As
$R$
is contained in this stabilizer, we conclude that
$R$
is indeed bounded.◻
Proof of Theorem 7.1.
We first observe the easier implication
$(1)\Rightarrow (2)$
. We assume by contradiction a finite index subgroup
$\unicode[STIX]{x1D6EC}^{\prime }<\unicode[STIX]{x1D6EC}$
, a complete field with absolute value
$k$
, a connected adjoint
$k$
-simple algebraic group
$\mathbf{G}$
and group homomorphism
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6EC}^{\prime }\rightarrow \mathbf{G}(k)$
such that
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6EC}^{\prime })$
is Zariski dense in
$\mathbf{G}$
and
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}\,\cap \,\unicode[STIX]{x1D6EC}^{\prime })$
is unbounded in
$\mathbf{G}(k)$
. We set
$\unicode[STIX]{x1D6E4}^{\prime }=\unicode[STIX]{x1D6E4}\,\cap \,\unicode[STIX]{x1D6EC}^{\prime }$
and observe that
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6E4}$
is of finite index. Note that
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}^{\prime })$
is unbounded in
$\mathbf{G}(k)$
, hence infinite, and it is commensurated by the Zariski dense subgroup
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6EC}^{\prime })$
. By Lemma 7.5 we obtain that
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}^{\prime })$
is Zariski dense in
$\mathbf{G}$
. By Lemma 7.6 we conclude that
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}^{\prime })$
is not amenable. By embedding
$\mathbf{G}(k)$
in
$\operatorname{GL}_{d}(k)$
we consider
$\unicode[STIX]{x1D70C}$
as a linear representation
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6EC}^{\prime }\rightarrow \operatorname{GL}_{d}(k)$
. We induce
$\unicode[STIX]{x1D70C}$
from
$\unicode[STIX]{x1D6EC}^{\prime }$
to
$\unicode[STIX]{x1D6EC}$
and obtain a representation
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6EC}\rightarrow \operatorname{GL}_{d^{\prime }}(k)$
.
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is not amenable as it contains the non-amenable subgroup
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4}^{\prime })\simeq \unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}^{\prime })$
. It follows that
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is not solvable by locally finite. This completes the proof of
$(1)\Rightarrow (2)$
We now focus on the implication
$(2)\Rightarrow (1)$
. We assume by contradiction a field
$K$
, an integer
$d$
and a group homomorphism
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6EC}\rightarrow \operatorname{GL}_{d}(K)$
such that
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is not solvable by locally finite. We assume, as we may, that
$K$
is a countable field, and if
$\unicode[STIX]{x1D6EC}$
is finitely generated we assume also that
$K$
is finitely generated. Tits’s theorem [Reference TitsTit72, Theorems 1 and 2] implies that
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is not amenable. We let
$\mathbf{H}$
be the Zariski closure of
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6EC})$
. By [Reference BorelBor91, AG, Theorem 14.4],
$\mathbf{H}$
is defined over
$K$
and
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6EC})<\mathbf{H}(K)$
. We denote by
$\mathbf{H}^{0}$
the identity connected component in
$\mathbf{H}$
and observe that
$\mathbf{H}^{0}(K)<\mathbf{H}(K)$
is of finite index. We set
$\unicode[STIX]{x1D6EC}^{\prime }=\unicode[STIX]{x1D719}^{-1}(\mathbf{H}^{0}(K))$
. Note that
$\unicode[STIX]{x1D6EC}^{\prime }<\unicode[STIX]{x1D6EC}$
is of finite index and the Zariski closure of
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6EC}^{\prime })$
is
$\mathbf{H}^{0}$
, as it is contained in the connected group
$\mathbf{H}^{0}$
and is of finite index in
$\mathbf{H}$
.
We set
$\unicode[STIX]{x1D6E4}^{\prime }=\unicode[STIX]{x1D6EC}^{\prime }\cap \unicode[STIX]{x1D6E4}$
. Observe that
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6E4}$
is of finite index. In particular,
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6EC}^{\prime }$
is commensurated and
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4}^{\prime })$
is not amenable. We let
$\mathbf{R}$
be the solvable radical of
$\mathbf{H}^{0}$
and set
$\mathbf{L}=\mathbf{H}^{0}/\mathbf{R}$
,
$\unicode[STIX]{x1D70B}:\mathbf{H}^{0}\rightarrow \mathbf{L}$
. As
$\mathbf{R}(K)$
is solvable and
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4}^{\prime })$
is not amenable, we conclude that
$\unicode[STIX]{x1D70B}\circ \unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4}^{\prime })$
is not amenable. In particular,
$\mathbf{L}$
is non-trivial. Upon replacing
$\mathbf{L}$
by its quotient modulo its center, which is finite, we may further assume that
$\mathbf{L}$
is center-free. Thus
$\mathbf{L}$
is a connected, adjoint, semisimple
$K$
-group and
$\unicode[STIX]{x1D70B}\circ \unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4}^{\prime })$
is Zariski dense in
$\mathbf{L}$
. The adjoint group
$\mathbf{L}$
is
$K$
-isomorphic to a direct product of finitely many connected, adjoint,
$K$
-simple
$K$
-groups. The image of the projection of
$\unicode[STIX]{x1D70B}\circ \unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4}^{\prime })$
to at least one of these factors is not amenable and, in particular, not locally finite. We fix such a
$K$
-simple factor of
$\mathbf{L}$
and denote it
$\mathbf{G}$
. We let
$\unicode[STIX]{x1D713}:\unicode[STIX]{x1D6EC}^{\prime }\rightarrow \mathbf{G}(K)$
be the composition of
$\unicode[STIX]{x1D70B}\circ \unicode[STIX]{x1D719}$
with the projection
$\mathbf{L}(K)\rightarrow \mathbf{G}(K)$
and conclude that
$\unicode[STIX]{x1D713}(\unicode[STIX]{x1D6EC}^{\prime })$
is Zariski dense and
$\unicode[STIX]{x1D713}(\unicode[STIX]{x1D6E4}^{\prime })$
is not locally finite in the connected adjoint
$K$
-simple
$K$
-algebraic group
$\mathbf{G}$
. Using Lemma 7.4, we find a complete field with an absolute value
$k$
and an embedding
$K{\hookrightarrow}k$
such that the image of
$\unicode[STIX]{x1D6E4}^{\prime }$
is unbounded in
$\mathbf{G}(k)$
, which is, furthermore, a local field if the field
$K$
is finitely generated. We obtain a contradiction, as the image of
$\unicode[STIX]{x1D6E4}^{\prime }$
is unbounded under the composed map
$\unicode[STIX]{x1D6EC}^{\prime }\rightarrow \mathbf{G}(K)\rightarrow \mathbf{G}(k)$
.◻
8 Proofs of Theorem 1.5 and Corollaries 1.6 and 1.7
Proof of Theorem 1.5.
Assume by contradiction that there exist an integer
$d$
, a field
$K$
and a linear representation
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6E4}\rightarrow \operatorname{GL}_{d}(K)$
such that the image
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is not solvable by locally finite. Then, by Corollary 7.2, there exist a finite index subgroup
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6E4}$
, a complete field with an absolute value
$k$
, a connected adjoint
$k$
-simple algebraic group
$\mathbf{G}$
and group homomorphism
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6E4}^{\prime }\rightarrow \mathbf{G}(k)$
such that
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}^{\prime })$
is Zariski dense and unbounded in
$\mathbf{G}(k)$
. If
$\unicode[STIX]{x1D6E4}$
is finitely generated,
$k$
could be taken to be a local field.
For every
$i=1,\ldots ,n$
let
$T_{i}^{\prime }=\overline{\unicode[STIX]{x1D70B}_{i}(\unicode[STIX]{x1D6E4}^{\prime })}<T_{i}$
be the closure of the projection of
$\unicode[STIX]{x1D6E4}^{\prime }$
in
$T_{i}$
, and
$T^{\prime }=T_{1}^{\prime }\times \cdots \times T_{n}^{\prime }<T$
be a closed subgroup. Since
$\unicode[STIX]{x1D6E4}^{\prime }<\unicode[STIX]{x1D6E4}$
has finite index, each
$T_{i}^{\prime }<T_{i}$
has finite index, and so
$T^{\prime }<T$
has finite index. Note also that
$\unicode[STIX]{x1D6E4}^{\prime }<T^{\prime }$
is a lattice with dense projections. Applying Theorem 1.2, we conclude that there exists a continuous homomorphism
$\bar{\unicode[STIX]{x1D70C}}:T^{\prime }\rightarrow \mathbf{G}(k)$
such that
$\bar{\unicode[STIX]{x1D70C}}|_{\unicode[STIX]{x1D6E4}^{\prime }}=\unicode[STIX]{x1D70C}$
. In particular,
$\bar{\unicode[STIX]{x1D70C}}(T^{\prime })$
is Zariski dense and unbounded and so is its closure,
$\overline{\bar{\unicode[STIX]{x1D70C}}(T^{\prime })}$
. By Lemma 7.6 we conclude that
$\overline{\bar{\unicode[STIX]{x1D70C}}(T^{\prime })}$
is not amenable.
Fixing a
$k$
-representation
$\mathbf{G}(k)\rightarrow \operatorname{GL}_{d^{\prime }}(k)$
we obtain a continuous linear representation
$T^{\prime }\rightarrow \operatorname{GL}_{d^{\prime }}(k)$
. Inducing this representation to
$T$
, we obtain a continuous homomorphism
$\unicode[STIX]{x1D713}:T\rightarrow \operatorname{GL}_{d^{\prime \prime }}(k)$
such that
$\overline{\unicode[STIX]{x1D713}(T)}$
is not amenable. This is a contradiction.◻
Proof of Corollary 1.6.
Assume by contradiction that there exist an integer
$d$
, a field
$K$
and a linear representation
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6EC}\rightarrow \operatorname{GL}_{d}(K)$
such that
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is not solvable by locally finite. By Theorem 7.1 we obtain a finite index subgroup
$\unicode[STIX]{x1D6EC}^{\prime }<\unicode[STIX]{x1D6EC}$
, a complete field with an absolute value
$k$
, a connected adjoint
$k$
-simple algebraic group
$\mathbf{G}$
and a Zariski dense group homomorphism
$\unicode[STIX]{x1D70C}:\unicode[STIX]{x1D6EC}^{\prime }\rightarrow \mathbf{G}(k)$
such that
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}\cap \unicode[STIX]{x1D6EC}^{\prime })$
is unbounded in
$\mathbf{G}(k)$
. If
$\unicode[STIX]{x1D6EC}$
is finitely generated,
$k$
could be taken to be a local field.
We let
$T^{\prime }$
be the closure of
$\unicode[STIX]{x1D6EC}^{\prime }$
in
$T$
and set
$\unicode[STIX]{x1D6E4}^{\prime }=\unicode[STIX]{x1D6E4}\cap \unicode[STIX]{x1D6EC}^{\prime }$
. Observe that
$T^{\prime }<T$
is a closed subgroup of finite index. Thus
$T^{\prime }<T$
is open and
$\unicode[STIX]{x1D6E4}^{\prime }<T^{\prime }$
is a lattice. Furthermore,
$\unicode[STIX]{x1D6EC}^{\prime }$
is a countable dense subgroup of
$T^{\prime }$
containing and commensurating
$\unicode[STIX]{x1D6E4}^{\prime }$
. Applying Corollary 1.4, we obtain a continuous homomorphism
$\bar{\unicode[STIX]{x1D70C}}:T^{\prime }\rightarrow \mathbf{G}(k)$
such that
$\unicode[STIX]{x1D70C}=\bar{\unicode[STIX]{x1D70C}}|_{\unicode[STIX]{x1D6EC}^{\prime }}$
. As
$\unicode[STIX]{x1D70C}(\unicode[STIX]{x1D6E4}^{\prime })$
is unbounded we obtain that
$\bar{\unicode[STIX]{x1D70C}}(T^{\prime })$
is unbounded. By Lemma 7.6 we conclude that
$\overline{\bar{\unicode[STIX]{x1D70C}}(T^{\prime })}$
is not amenable.
Fixing a
$k$
-representation
$\mathbf{G}(k)\rightarrow \operatorname{GL}_{d^{\prime }}(k)$
, we obtain a continuous linear representation
$T^{\prime }\rightarrow \operatorname{GL}_{d^{\prime }}(k)$
. Inducing this representation to
$T$
, we obtain a continuous homomorphism
$\unicode[STIX]{x1D713}:T\rightarrow \operatorname{GL}_{d^{\prime \prime }}(k)$
such that
$\overline{\unicode[STIX]{x1D713}(T)}$
is not amenable. This is a contradiction.◻
Proof of Corollary 1.7.
Assume by contradiction that there exist an integer
$d$
, a field
$K$
and a linear representation
$\unicode[STIX]{x1D719}:\unicode[STIX]{x1D6E4}\rightarrow \operatorname{GL}_{d}(K)$
such that
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is not solvable by locally finite. In particular, by Tits’s theorem [Reference TitsTit72, Theorems 1 and 2],
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})$
is not amenable. We claim that
$\unicode[STIX]{x1D719}$
is injective. Assume by contradiction that
$N=\ker \unicode[STIX]{x1D719}$
is non-trivial. By the assumption that for each
$i$
,
$\unicode[STIX]{x1D70B}_{i}|\unicode[STIX]{x1D6E4}$
is injective and for each non-trivial closed normal subgroup of
$T_{i}$
the corresponding quotient group is amenable, we obtain that
$T_{i}/\overline{\unicode[STIX]{x1D70B}_{i}(N)}$
is amenable. This contradicts [Reference Bader and ShalomBS06, Theorem 3.7(iv)], which guarantees that then
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E4})\simeq \unicode[STIX]{x1D6E4}/N$
is amenable. We conclude that
$\unicode[STIX]{x1D719}$
is indeed injective.
We set
$S=T_{2}\times \cdots \times T_{n}$
, let
$U<S$
be a compact open subgroup and set
$\unicode[STIX]{x1D6E5}=\unicode[STIX]{x1D6E4}\cap (T_{1}\times U)$
. Note that
$\unicode[STIX]{x1D6E5}$
is a lattice in
$T_{1}\times U$
, as
$\unicode[STIX]{x1D6E4}$
is a lattice in
$T$
and
$T_{1}\times U$
is open in
$T$
. As
$U$
is compact, we deduce that
$\unicode[STIX]{x1D70B}_{1}(\unicode[STIX]{x1D6E5})<T_{1}$
is a lattice. Since
$U<S$
is commensurated, we obtain that
$T_{1}\times U<T=T_{1}\times S$
is commensurated, thus
$\unicode[STIX]{x1D6E5}<\unicode[STIX]{x1D6E4}$
is commensurated.
As
$\unicode[STIX]{x1D70B}_{1}$
is injective on
$\unicode[STIX]{x1D6E4}$
, we may identify
$\unicode[STIX]{x1D6E4}$
and
$\unicode[STIX]{x1D6E5}$
with
$\unicode[STIX]{x1D70B}_{1}(\unicode[STIX]{x1D6E4})$
and
$\unicode[STIX]{x1D70B}_{1}(\unicode[STIX]{x1D6E5})$
correspondingly, and consider them as subgroups of
$T_{1}$
. Applying Corollary 1.6 in this setting, we obtain that
$\unicode[STIX]{x1D719}(\unicode[STIX]{x1D6E5})$
is solvable by locally finite and, in particular, is amenable. By the injectivity of
$\unicode[STIX]{x1D719}$
we conclude that
$\unicode[STIX]{x1D6E5}$
is amenable. As
$\unicode[STIX]{x1D6E5}<T_{1}$
is a lattice, it follows that
$T_{1}$
is amenable. This is a contradiction.◻
Acknowledgements
We owe thanks to many friends, too numerous to be listed here, for their constant stimulation and interest in our work. Special thanks go to T. N. Venkataramana, Pierre-Emmanuel Caprace, Tsachik Gelander and the anonymous referee. We are also indebted to Bruno Duchesne and Jean Lécureux for their contribution to this project.