Superrigidity for irreducible lattices and geometric splitting
Nicolas Monod
University of Chicago
-
Abstract.
We prove general superrigidity results for actions of irreducible lattices on
spaces; first, in terms of the ideal boundary, and then for the intrinsic geometry (including for infinite-dimensional spaces). In particular, one obtains a new and self-contained proof of Margulis' superrigidity theorem for uniform irreducible lattices in non-simple groups.
The proofs rely on simple geometric arguments, including a splitting theorem which can be viewed as an infinite-dimensional (and singular) generalization of the Lawson-Yau/Gromoll-Wolf theorem. Appendix A gives a very elementary proof of commensurator superrigidity; Appendix B proves that all our results also hold for certain non-uniform lattices.
1 Introduction
1.1 Superrigidity
In the early seventies, Margulis proved his celebrated superrigidity theorem for irreducible lattices in semi-simple Lie and algebraic groups of higher rank. One of the motivations for this result is that it implies arithmeticity : a complete classification of higher rank lattices.
In the case where the semi-simple group is not almost simple, superrigidity reads as follows (see [M1] , page 2):
1. Theorem (Margulis).
Let
be an irreducible lattice in
, where
are local fields and
are connected simply connected semi-simple
-groups without
-anisotropic factors and
. Let
be a local field,
a connected adjoint
-simple
-group and
a homomorphism with Zariski-dense unbounded image.
Then
extends to a continuous homomorphism
.
Our goal is to abandon completely the realm of algebraic (or Lie) groups and to establish a generalization of this theorem for uniform lattices in products. There will be no assumptions on the product group
. Instead of the algebraic group
for the target we shall consider general isometry groups of
spaces. An additional feature of our proof is that it is fully self-contained; this gives in particular a new, rather elementary, and purely geometric proof of Margulis' result for uniform lattices. For instance, the idiosyncrasies of positive characteristics (see Venkataramana [V] ) vanish.
Here is our setting: (i) Recall that a lattice
in a product
of arbitrary locally compact groups is said irreducible if the projection of
to each factor
is dense. In the classical semi-simple case of Theorem 1 , this follows from the stronger notion of algebraic irreducibility assumed therein.
Remark.
The irreducibility assumption is not a restriction whatsoever: One verifies that any lattice
is an irreducible lattice in the product
of the closures
of its projections to
. In particular, for any discrete cocompact subgroup
, the theorems below apply and extend the
-actions to
.
(ii) We shall replace
with an isometry group
of an arbitrary complete
space
; in the classical case,
is the symmetric space of
when
is Archimedean and the associated Bruhat-Tits building otherwise. The Zariski-density is a necessary assumption for Theorem 1 as stated; before proposing two replacements for that assumption, we first state a result without any further assumption, which is possible for proper spaces upon passing to the geometric boundary:
2. Theorem.
Let
be an irreducible uniform lattice in a product
of locally compact
-compact groups. Let
act by isometries on a proper
space
without global fixed point.
Then there is a non-empty closed
-invariant set
on which the
-action extends continuously to a
-action. Moreover this action factors through
for some
.
(See the text for a more precise statement; e.g.
arises as the boundary of a
-space lying in
.) We turn now to extending homomorphisms
as in Margulis' statement.
We propose first, still for
proper, the following substitute for Zariski-density in simple adjoint groups:
3. Definition.
A subgroup
is indecomposable if for every non-empty
-invariant closed subset
, the closure
acts faithfully on
and
is closed in
(for the topology of uniform convergence).
This always holds in the setting of Theorem 1 . Indecomposability turns out to be very natural, see Section 6.2 . In particular, a non-trivial indecomposable subgroup
cannot fix any point at infinity. In negative curvature, indecomposability is essentially automatic upon passing to invariant subspaces.
4. Corollary.
Let
be an irreducible uniform lattice in a product
of locally compact
-compact groups. Let
be a closed subgroup, where
is a proper
space, and let
be a homomorphism with indecomposable unbounded image.
Then
extends to a continuous homomorphism
.
This result immediately implies Margulis' Theorem 1 for uniform lattices.
Further, if we keep
, it shows similarly that for an irreducible uniform lattice in a general product group
, all completely reducible linear representations in finite-dimensional vector spaces over all local fields are completely determined by the continuous linear representations of
. Specialising in the other direction: Even when
is an algebraic group, the above theorem yields a new family of superrigidity results.
Remark.
It is easy to verify (and inherent in the proofs) that the extended map
factors through some
. Similarly for Margulis' Theorem 1 and Theorem 6 below.
Although the above results are set in the context of locally compact spaces, our proof involves in an essential way infinite-dimensional
spaces. Indeed, the overall strategy is to induce the
-action to a
-action on a space of
maps
and then to prove a splitting theorem for the latter. It turns out that we can prove superrigidity also when
itself is infinite-dimensional; to this end, we propose our second substitute for Zariski-density in intrinsic geometric terms for any
space
:
5. Definition.
A subgroup
is reduced if there is no unbounded closed convex subset
such that
is at finite (Hausdorff) distance from
for all
.
6. Theorem.
Let
be an irreducible uniform lattice in a product
of locally compact
-compact groups. Let
be a closed subgroup, where
is any complete
space not isometric to a finite-dimensional Euclidean space
. Let
be a homomorphism with reduced unbounded image.
Then
extends to a continuous homomorphism
.
(Continuous homomorphism and closed subgroup are to be defined suitably when
is not proper.)
Remark.
As stated, the above theorem does not hold for the very special case
(see Section 6.5 ). Although Hilbert spaces, in particular
, are special examples of complete
spaces, their linear structure allows a more detailed analysis; see Shalom [
Sh]
.
We have so far considered only uniform lattices. The non-uniform case raises some difficulties, which can however be overcome under certain assumptions:
For a discussion of these concepts, more precise statements and proofs, see Appendix B.
1.2 Splitting
The geometry of infinite-dimensional
spaces is in some regards very different from their classical analogues; a first glimpse into the richness of phenomena arising there is given by the study of unitary representations and their cocycles – Kazhdan's property (T) is perhaps one of the first instances. A central issue is that the boundary
does not reflect sufficiently the structure of the isometries of
; indeed
may even be empty. Therefore, we introduce the following notion, which coincides simply with the existence of a fixed point in
when
is a proper
space:
8. Definition.
Let
be a topological group with a continuous action by isometries on a metric space
. The
-action on
is evanescent if there is an unbounded set
such that for every compact set
the set
is bounded.
We now can turn to our splitting theorem. Observe that there is no assumption whatsoever on the topology of
or
:
9. Theorem.
Let
be a complete
space with a continuous
-action by isometries, where
is any product of topological groups
.
Either the
-action is evanescent or there is a canonical non-empty closed convex
-invariant subspace
which splits
-equivariantly isometrically as a product
of
-spaces
.
(See Proposition 34 for the importance of evanescence.) The special case where
is locally compact, though it is not representative of the above, is essentially known:
10. Corollary.
Let
be a proper
space with a
-action by isometries, where
is any product of groups
.
Either there is a
-fixed point in
, or there is a non-empty closed convex
-invariant subspace
which splits
-equivariantly isometrically as a product of
-spaces
. □
When
is in fact a Riemannian manifold, this result is Schroeder's generalization [Sch] of the splitting theorems proved by Lawson-Yau [LY] and Gromoll-Wolf [GW] around 1970.
The latter are of differential nature – and Schroeder's proof relies in an essential way on a Riemannian argument of Eberlein [E] using currents. However, there is a purely
statement in [BH] (II.6.21 and II.6.25(3)); since the argument therein requires to extend geodesics indefinitely within every invariant subspace, it is assumed in that reference that
has this property and that
is cocompact or at least has full limit set. Similar particular cases are obtained by Jost-Yau [JY3] using harmonic maps. In our situation, it is impossible to assume anything on the space
, but it turns out that no assumption is needed.
1.3 Comments
One of the tools allowing us to deal with the infinite-dimensional spaces is a weakened topology
that we introduce; it is a common generalization of the weak topology in Hilbert spaces and the topology introduced in [MS1] for trees. Another more technical aspect is the behaviour of evanescence under induction (Theorem 53 and Appendix B). Other applications of our methods to rigidity theory will be exposed elsewhere; in particular, the special case of homomorphism to algebraic groups
leads, following Margulis' ideas, to an Arithmeticity vs. Non-Linearity Alternative for irreducible lattices in suitable product groups [Md] .
Related results. Margulis' Theorem 1 has been followed by numerous related results.
We refer e.g. to [Bu] ,[GP] and references therein; to Zimmer's non-linear superrigidity [Z] ; to the geometric superrigidity of Jost-Yau [JY1] ,[JY2] ,[J] and Mok-Siu-Yeung [MSY] . Moreover, in the
setting, the flexibility of bounded cohomology allows for very general conclusions [MS1] ,[MS2] , including cocycle superrigidity in the spirit of Zimmer. By contrast, the present approach is more simple-minded: No boundary maps, no cohomology, no harmonic maps, and of course no theory of algebraic groups.
We also point out that an intermediate result in our strategy (Theorem 55 ) provides a totally geodesic
-map in complete generality (from an auxiliary
-space), which is often enough to deduce geometric superrigidity.
On the exposition and use of previous work. We seek a self-contained presentation; in order to illustrate the new concepts that we introduce, we bring a number of examples or simple propositions which are not needed for the proof of the main results. We are inspired by various sources: Our first motivation is of course Margulis' work. The idea of inducing actions is classical in rigidity. Spaces of
maps appear notably in [KS] . Our general strategy is analogous to Shalom's work for unitary representations and cohomology [Sh] . The proof of Theorem 9 borrows some arguments from [Sch] and some others from [BH,II.6] .
Acknowledgments. I am grateful to J.-F. Quint for providing the elegant proof of Lemma 59 ; I thank D. Fisher and T. Gelander for helpful remarks on an earlier version.
This work was partially supported by FNS grant 8220-067641 and NSF grant DMS 0204601.
2 Informal Outline of the Reasoning
Since our proof is accompanied with many geometric generalities, it might be useful to outline the general argument in order to highlight its simplicity:
2.1 Superrigidity
Consider a lattice
acting on a complete
space
.
We construct an associated
-space
by considering the (right-)
-equivariant maps
. The distance between
is
and the
-action is
.
The motivation for this new “induced”
space
is this: On the one hand, there is a correspondence between properties of the
-action on
and of the
-action on
; to wit, non-evanescence is preserved. On the other hand, even though
is a priori a more unwieldy space (typically not locally compact), the
-action on it is subject to the splitting theorem.
We obtain a splitting
into
-spaces. If
, then for
(
) we have a rectangle
. It follows that
for almost all
. If the projection of
to
is dense, this implies that
is a.e. constant. We thus obtain isometric maps
compatible with the
-actions. This is the main step in extending the
-action on
or on
to
.
2.2 Splitting
Consider a
-action on a complete
space
; we are motivated by the above “induced” spaces but work in complete generality.
Reduce to the case
. We show that either the
-action is evanescent or there is a minimal non-empty closed convex
-invariant subset
. The main point here is a geometric analogue of the Banach-Alaoğlu theorem: We introduce a weakened topology for which bounded closed convex sets are compact. A compactness argument then produces the minimal set.
The “sandwich lemma” implies as in [Sch] and [BH,II.6] that the collection of all such sets
has a foliated structure which is preserved by the
-action. An addition to their arguments here is that even though we lack any further assumptions, this foliation is a global isometric splitting. This follows by showing that the holonomy consists of Clifford translations, hence is trivial.
3 Geometric Preliminaries
3.1
Our general background reference is [BH] (see also [Ba] ,[J] ). Let
be a metric space with metric
. A map
, where
is any interval, is geodesic if it is isometric. The space
is geodesic if every pair of points is joined by some geodesic segment, and
is convex if it contains any geodesic segment joining any two of its points. For any subsets
of a metric space
, we denote by
the closed convex hull of
and
.
We write
for the closed ball of radius
around
; if
is a geodesic space and
, then this is the closure of the open ball
.
3.2
A CAT(0) space is a geodesic space
such that for any triangle
the midpoint
of any geodesic between
satisfies the courbure négative inequality of Bruhat-Tits [BH,II.1.9] :
|
(1)
|
Equivalently, the distance between any points in the sides of any geodesic triangle are bounded by the corresponding distances in Euclidean triangles. In particular, geodesics are unique: the segment from
to
is
. Examples include all symmetric spaces and Bruhat-Tits buildings; all simply connected manifolds of non-positive sectional curvature; Hilbert spaces; simply connected Euclidean or hyperbolic simplicial complexes satisfying certain local link conditions [BH,II.5.4] .
Here is an example of application of the
inequalities. Let
. For
let
be the points of
at distance
of
, respectively of
. Then
|
(2)
|
This type of inequalities was first proved by Reshetnyak [Rk] (we refer to [KS,2.1.3] for a derivation of ( 2 ) from ( 1 )).
3.3
The circumradius of a bounded set
is the infimum
of all
such that there is
with
. If
is
and complete, then this infimum is achieved and (for
) there is a unique point
such that
; this point is called the circumcentre of
. If in addition
is convex and closed, then
.
11. Lemma.
Let
be a complete
space,
two non-empty bounded closed convex sets,
the corresponding circumcentres and
the circumradii.
Then
.
-
Proof.
If
, there is nothing to prove. Otherwise, the midpoint
of
is not the circumcentre of
and thus there is
with
. Replacing this in ( 1 ) gives
, as required. □
3.4
If
a non-empty closed convex subset of the complete
space
, then there is a nearest point projection map
which does not increase distances.
Another consequence of ( 1 ) is the Sandwich Lemma (see [BH,II.2.12] ): Assume that
are two non-empty closed convex subsets such that the function
is constant on
and likewise
on
; denote this (common) constant by
. Then there is a canonical isometry
such that
In the particular case where
are geodesic segments, we say that they determine a Euclidean rectangle.
Similar arguments apply to the following setting: An isometry
such that
is independent of
is called a Clifford translation [BH,II.6.14] . If
, then there is an isometry
intertwining
with the translation by
along
.
More generally, any complete
space
splits canonically as
, where
is a (possibly trivial or finite-dimensional) Hilbert space and
does not admit any Clifford translation. Moreover,
preserves this splitting. For all this, see [BH,II.6.15] .
3.5
Let
be a topological group acting on a metric space
by isometries. The following are equivalent:
(i) For all
the map
,
is continuous at
.
(ii) The action map
is continuous.
When this happens, we say that the action is continuous. When
is proper, we always endow its isometry group
with the compact-open topology; in that case,
is a locally compact second countable topological group and the above conditions are equivalent to the continuity of the homomorphism
. For
general, we do not topologize
but still call a homomorphism
continuous when the
-action is so, and a subgroup
is said closed if its orbits in
are so. A standard argument implies:
12. Lemma.
Let
be a locally compact second countable group with an action by isometries on a complete separable metric space
. If for all
the map
,
is measurable, then the action is continuous. □
We call a subset
bounded if for some (or equivalently any)
the set
is bounded in
; when
is proper, this coincides with the usual definition in which a subset of a locally compact group
is bounded when it has compact closure.
3.6
The boundary at infinity
of a complete
space
can be defined as the set of equivalence classes of geodesic rays in
, where two rays are equivalent (asymptotic) if they remain at bounded distance from each other. The bordification
can be identified with the inverse limit of the closed balls
as
under the maps
, wherein
is any basepoint; the resulting topology is called the cone topology [BH,II.8] .
When
is proper,
is a second countable compact space (hence metrizable) and the action map
is continuous.
3.7
Let
be a metric space. Unless otherwise stated, every topological epithet will always refer to the topology
induced by the metric
. One can define a weaker topology
by letting
be the weakest topology on
such that for all
the map
is continuous. This topology is always Hausdorff; we shall however be more interested in the following:
13. Definition.
Let
be a metric space. We define the topology
to be the weakest topology on
for which all
-closed convex sets are
-closed.
Here is the main property of
.
14. Theorem.
Let
be a complete
space and
a bounded closed convex subset. Then
is compact
for the topology
.
This is a common generalization of the two quite different cases of Hilbert spaces in the weak topology and of trees in the topology considered in [MS1] ; see the examples below.
-
Proof of the theorem.
We need to prove that for any family
of
-closed sets
such that
has the finite intersection property, the intersection of all
is non-empty. By Alexander's sub-base theorem, it is enough to consider a family
consisting of closed convex sets
. We may assume
upon replacing each
by
. Let
be the set of non-empty finite subsets
, ordered by inclusion. To any
we associate the circumradius
of
and its circumcentre
. Notice that
is a non-increasing net since
implies
and thus
. On the other hand this net is non-negative, therefore converges and thus is a Cauchy net. Applying Lemma 11 to the sets
we deduce that
is a Cauchy net and hence converges to a limit
. For every
, all points
with
belong to
; therefore the limit
is still in
. It follows that
is in
as was to be shown. □
15. Remark.
Lemma 11 can also be used to explain the following (probably well-known) fact: Let
be a non-expanding map of a complete
space
. If
has bounded orbits, then its (closed convex) fixed set is non-empty.
Indeed, pick any
; let
and let
be its circumcentre,
its circumradius. Since
is decreasing, Lemma 11 implies that
has a limit
. Since
, we have
. Applying again Lemma 11 , it follows that
converges to
, which is thus fixed.
3.8
This Section 3.8 serves only to illustrate
.
16. Lemma.
For any complete
space
we have
.
-
Proof.
We only need to show
. Let
be closed convex. The set
is
-closed. We claim that
. On the one hand,
because for all
we have
regardless of
. On the other hand, if
then
is not in
for
. □
17. Lemma.
If
is
-compact, then the restriction of
and
to
coincide.
-
Proof.
It is enough to show that for all
and all
there is a
-neighbourhood
of
such that
. That ball, together with all
for every
with
, covers
. Therefore there is a finite set
of points
with
such that
covers
. The set
has the sought properties. □
The topology
is familiar in some particular cases:
18. Example.
If
is a Hilbert space, then
and
both coincide with the weak topology.
19. Example.
If
is the standard infinite-dimensional (separable) real hyperbolic space
(see e.g. [
BIM]
), then
and
coincide; moreover, they are induced from the weak topology if we realize
with the ball model in a Hilbert space.
20. Example.
If
is a simplicial tree, then
and
coincide; moreover, they coincide with the weak topology
on trees introduced in [
MS1]
.
However, the topology
is not all that straightforward in general. It seems not to be Hausdorff for higher rank Bruhat-Tits buildings or symmetric spaces, and it is unclear to us what happens for complex hyperbolic spaces (of finite or infinite dimension).
Moreover, whilst in a Hilbert space all weakly compact sets are bounded, this is not so even in the simplest examples of
spaces:
21. Example.
Let
be a simplicial tree consisting of countably many rays of finite but unbounded length, all meeting at one vertex. Then the space
is
-compact even though unbounded. Notice in addition that
.
Let
be any complete
space. We extend the topology
to
by declaring that for any
-closed convex set
the (usual) closure
of
in
is
-closed.
22. Remark.
The (compact) topology that
determines on
through the realization of
as inverse limit of closed balls is in general coarser than
, even when restricted to
: Already for Hilbert spaces, one obtains the weaker bounded weak topology. These topologies coincide however when
is a tree or a real hyperbolic space as in Example 19 .
Whilst
is not
-compact when
is e.g. an infinite-dimensional Hilbert space, we have:
23. Proposition.
If the complete
space
is Gromov-hyperbolic, then
is
-compact.
-
Proof.
It suffices to show that for any nested family
of non-empty closed convex sets
the intersection
is non-empty. Fix
. If
is bounded, we are done by Theorem 14 . Otherwise, we claim that
is a single point of
.
Indeed, in view of the sequential model of
[GH,p. 120] , it is enough to show that for any choices
,
, the Gromov product
tends to infinity along
. This follows since
is comparable to
, which is bounded below by
or
. □
3.9
We will now analyse evanescence (Definition 8 ); let
be a topological group with a continuous action by isometries on a metric space
.
24. Definition.
Let
be a subset of
. Then a subset
such that
is bounded will be said
-evanescent.
Thus, by Definition 8 , the action is evanescent if and only if there is an unbounded set
which is
-evanescent for every compact set
; we then call
itself evanescent.
25. Lemma.
Suppose that
is a
space, let
be compact in
and
. Then there exists no unbounded
-evanescent set if and only if there is
and
such that
-
Proof.
Sufficiency is obvious. Suppose conversely that the condition fails. Then for every
there is
such that
for all
. Let
be the point at distance
from
on
. A comparison argument shows that
and thus
is a
-evanescent sequence. □
26. Remark.
It follows from this lemma that the
-action on
is evanescent if and only if every compactly generated subgroup of
has a non-trivial fixed point for its diagonal action on some (or equivalently any) asymptotic cone of
along a free ultrafiltre (the base-point of an asymptotic cone is a trivial fixed point). We shall not use this characterization.
27. Proposition.
Suppose that
is a complete
space.
(i) If there exists a
-fixed point in
then the
-action on
is evanescent.
(ii) The converse holds if
is proper.
-
Proof.
If there exists a point
fixed by
, then any ray pointing to
is an evanescent set. In case (ii),
is compact and it follows that any unbounded evanescent sequence has a subsequence converging to some
; the definition of the cone topology on
shows that
is
-fixed. □
Another natural definition is as follows:
28. Definition.
A continuous
-action on a metric space
is weakly evanescent if for every compact
there is an unbounded
-evanescent set in
.
We point out that for both Definitions 8 and 28 it is enough to consider unbounded evanescent sequences. Whilst evanescence is in general stronger than weak evanescence, we have:
29. Proposition.
Suppose that
is a complete
space. Then weak evanescence implies evanescence if either (i)
is proper; or (ii) if there is a countable cofinal chain in the directed set of compact subsets of
.
30. Remark.
The assumption in (ii) is satisfied if
is locally compact
-compact (since
is then a countable union of compact sets of non-empty interior); but it also holds in other cases (e.g. if
is the additive group of a dual Banach space endowed with the weak-* topology).
-
Proof of the proposition.
For (i), observe that the family
has the finite intersection property so that we can use compactness of
and Proposition 27 .
In case (ii), one can choose an increasing sequence of compact sets
such that any compact set in contained in
for
big enough. Let
be unbounded and
-evanescent and let
be a constant bounding the numbers
of Definition 24 . Fix any point
and define a diagonal sequence
by picking first
with
and then letting
be the point at distance
of
on
. A comparison argument as in the proof of Lemma 25 shows that
is evanescent. □
Evanescence behaves in a simple way with respect to direct products:
31. Proposition.
Let
be a product of topological groups and suppose that
is a product of two unbounded
spaces
with continuous
-action. Endow
with the product
-action.
(i) Both
-actions on
are evanescent.
(ii) The
-action on
is (weakly) evanescent if and only if at least one of the
-actions on
is (weakly) evanescent.
Observe that (i) stands in contrast to Shalom's notion of (non-)uniformity [Sh] .
-
Proof.
For the first point, fix any
and let
. This set is evanescent for
, since for any compact
the set
is bounded and
acts trivially on
; likewise for
. A similar argument shows that if, say, the
-action on
is (weakly) evanescent, then the
-action on
is so too. As for the converse, assume
is an unbounded
-evanescent sequence in
for some compact
. Fix
and set
,
. Since
is
-equivariant and does not increase distances (and similarly for
), it is enough to show that either
or
is unbounded. But otherwise
would itself be bounded. □
3.10
We recall that following Kazhdan a unitary or orthogonal representation
is said almost to have non-zero invariant vectors if for every compact
and every
there is a unit vector
such that
|
(3)
|
32. Proposition.
Suppose that
is a Hilbert space. Then the
-action on
is weakly evanescent if and only if the associated orthogonal representation almost has non-zero invariant vectors.
-
Proof.
A straightforward verification using e.g. sequences of the form
, where
satisfies ( 3 ) for
. □
33. Remark.
In particular, one has the following well known characterization when
is as in Remark 30 :
almost has non-zero invariant vectors if and only if there is an asymptotically invariant sequence, i.e. a sequence
with
and
uniformly for
on compact sets.
4 A General Splitting Theorem
This section addresses Theorem 9 .
4.1
Consider the very simplest product group:
, and the very simplest non-proper
space: a Hilbert space. Then, already, evanescence is the right replacement for fixed points at infinity in the sense that Theorem 9 fails if we formulate it in terms of
. We give a more general counter-example:
34. Proposition.
Let
be any two locally compact separable groups without Kazhdan's property (T). Then there is an action of
by isometries on a Hilbert space
such that:
(i) The
-action on
has no fixed point.
(ii) No non-empty closed convex
-invariant subspace
splits as product of
-spaces.
-
Proof.
By Theorem 1 in [BV] , each
admits an orthogonal representation
such that
almost has non-zero invariant vectors but
is weakly mixing, that is, has no finite-dimensional subrepresentation. Let
be the
-representation
and
the associated Hilbert space. Notice that
itself is weakly mixing; this implies
, see [BR] . On the other hand, one deduces immediately with the definition ( 3 ) that
almost has non-zero invariant vectors. As a well-known consequence of the closed graph theorem (observed by Guichardet [G,Théorème 1] ), it follows that there is a non-trivial
-cocycle
in the closure of the space of coboundaries. We endow
with the corresponding continuous
-action by (affine) isometries. For (i), observe that a fixed point in
would give a non-zero
-fixed vector for
. For (ii), assume for a contradiction that
splits as
. For every
, the vector
is fixed by
for
, which in view of
shows that both
are reduced to a single point. This point being fixed by the affine
-action,
is trivial, contradicting the assumption. □
Notice that the action here is weakly evanescent by Proposition 32 , hence evanescent by Proposition 29 . Also, the action is non-uniform in Shalom's sense [Sh] , and indeed the above situation stands in contrast to the results of [Sh] for uniform actions.
4.2
We undertake now the proof of Theorem 9 . First observe that it is enough to consider the case
. Indeed, the case
is tautological; further, assume that for
we have a subspace
splitting equivariantly as
, where
is a
-space and
a
-space. In order to apply induction, we just need to observe that the
action on
cannot be evanescent unless the
-action on
is so, see Proposition 31 .
We can from now on assume that the
-action on
is not evanescent. Theorem 14 allows us to get started with the following:
35. Proposition.
Let
be a complete
space with a continuous
-action by isometries, where
is any product of two topological groups.
If the
-action is not evanescent then there is a minimal non-empty closed convex
-invariant set in
.
-
Proof.
Choose any point
and let
be the set of non-empty closed convex
-invariant subsets of
, ordered by reverse inclusion. By Hausdorff 's maximal principle, there is a (non-empty) maximal totally ordered subset
. If the intersection
is non-empty, we are done; we assume for a contradiction that it is empty and will show that the
-action on
is evanescent. The net
indexed by
is non-decreasing; we shall prove that it converges to infinity.
Indeed, if not, it would have a limit
and the family
would be a nested collection of non-empty closed convex bounded sets with empty intersection.
This contradicts
-compactness established in Theorem 14 . It follows that the set
is unbounded. It is now enough to show that for any compact sets
the set
|
(4)
|
is bounded. On the one hand,
is
-equivariant and thus
implies that the family
is bounded. On the other hand, for any
the function
is constant on
-orbits, so that by convexity and continuity of the metric it is bounded by
on
. It follows now from
that the collection ( 4 ) is bounded. □
36. Remark.
In particular, any non-evanescent
-action admits a minimal non-empty closed convex
-invariant set and actually any nested family of non-empty closed convex
-invariant sets has non-empty intersection. See also Remarks 39 .
4.3
We now know that the set
of minimal non-empty closed convex
-invariant sets in
is non-empty. We shall use repeatedly the following obvious
37. Lemma.
If
is a minimal non-empty closed convex
-invariant set, then any convex
-invariant continuous (or lower semi-continuous) function on
is constant.
-
Proof.
If
were to assume two distinct values
, then
would be a strictly smaller non-empty closed convex
-invariant set. □
Fix
and let
. Observe that each
has a natural
-action. We can now consider the following setup, borrowed from [BH,pp. 239–241] : For every
, the function
is constant on
by Lemma 37 since it is
-invariant; likewise with
interchanged. The Sandwich Lemma (Section 3.4 ) yields a canonical isometry
|
(5)
|
such that
|
(6)
|
In particular, the distance
defines indeed a metric on the set
and this metric is geodesic. Furthermore we have
and hence obtain a well-defined bijection
by setting
, wherein
is the unique element of
containing
.
4.4
At this point, the main remaining steps are to show that
is actually an isometry and that it intertwines the
-actions; we need the following key fact:
38. Proposition.
For all
we have
.
-
Proof.
First we point out that if all three sets
were just geodesic lines, then this proposition is a well known general fact holding for any three parallel lines in any metric space, see [BH,II.2.15] . However, in our case, it is not even necessary that
should contain any line. Thus, denote by
the above map
. The properties of the isometry
in ( 5 ) imply that
is an isometry for all
; moreover, this isometry
is
-equivariant because
is
-invariant. Therefore,
is a
-equivariant isometry. It follows that the function
is a convex continuous
-invariant function of
; Lemma 37 implies that it is constant. Thus
is a Clifford translation of
. We need to show that
is trivial. But a non-trivial Clifford translation preserves the image of a geodesic line
; indeed, recall that in fact in that case
would split for
(Section 3.4 ; see [BH,II.6.15] – though we will not need this. We may now apply the general fact mentioned earlier ([BH,II.2.15] ) to the three lines
,
and
. We deduce that
translates
trivially. Therefore, the constant
vanishes and
as was to be shown. □
4.5
It follows now that
is isometric; for completeness (and because it contains a misprint), we give the calculation of [BH,p. 241] : Let
; using twice that ( 6 ) defines an isometry to the Cartesian product
for all
, we have
Applying now Proposition 38 to
and
we deduce
, so that
, and
is isometric as claimed; it is onto by ( 6 ).
39. Remarks.
Let
be a topological group with a continuous action by isometries on a complete
space
. (1) The above arguments show that the (possibly empty) union of all minimal non-empty closed convex
-invariant subspaces
splits as a product
; we call
the space of components. (2) If the action is non-evanescent, then there is a canonical minimal non-empty closed convex
-invariant subspace
. Indeed,
is non-empty by Remark 36 and bounded by non-evanescence; hence it has a circumcentre
and we let
.
4.6
It remains to check that
intertwines the
-action on
with the product action on
, and actually for the
factor this immediately follows from the
-equivariance of
. However, a priori, the bijection
transports the
-action on
to a
-action on
of the form
where we only know that the assignment
defined by
determines a well-defined
-action on
, which moreover commutes with the
-action.
We need to show that
for all
. As in the proof of Proposition 38 (for
), we deduce from Lemma 37 that
is a Clifford translation of
; so if it were non-trivial it would preserve a line
in
and in particular fix a point
. Moreover, we would have
for some
and all
. Thus for all
and all
we have
which implies that
fixes
. On the other hand,
(not just
) fixes
since
lies in
. Now there is a
-fixed point in
, contradicting the assumption according to Proposition 27 .
Being isometric to a product of geodesic spaces,
is itself geodesic and hence convex in
. Likewise, it is closed in
because it is a product of complete spaces (indeed, the uniform structure on
is complete because it is in fact a product uniform structure; alternatively, apply the hands-on argument in [BH,p. 240] ).
40. Remark.
It does not follow a priori from the above proof that
is minimal. However, we may preface the whole proof by replacing
with the canonical component provided by Remarks 39 . Therefore, in the non-evanescent case of Theorem 9 , we obtain in addition that each
is minimal.
This concludes the proof of Theorem 9 . □
5 Induction and its Properties
5.1
We begin by defining general “Pythagorean integrals” of metric spaces; this is not the only natural integral of metric spaces, see Remark 48 .
41. Definition.
Let
be a standard Borel space with a probability measure
and let
be a metric space. We denote by
the space of all measurable maps (up to null-sets)
with separable range and such that for some (and hence any)
the function
is in
.
We endow
with the metric defined by
42. Remarks.
(i) The
condition is independent of
by the triangle inequality in
since
is finite. (ii) The triangle inequality in
follows by combining the Cauchy-Schwarz inequality with the triangle inequality in
.
Such spaces were considered e.g. in [KS] . The following is straightforward:
43. Lemma.
Suppose that
is complete, respectively separable. Then so is
. □
We now describe geodesic segments in
; compare [BH,I.5.3] .
44. Proposition.
Let
be a complete metric space,
as in Definition 41 and
any interval. A continuous map
is a geodesic if and only if there is a mesurable map
and a collection
of geodesics
such that
for all
and
-a.e.
(i.e.
is a semi-density).
-
Proof.
The condition is sufficient; conversely, suppose that
is a geodesic. It suffices to show that there is a dense subset
such that for a.e.
, the map
coincides on
with a linearly reparametrized geodesic
. Indeed, if
denotes the reparametrization factor and
the corresponding geodesic, we obtain (sufficiency) a geodesic
which coincides with
on
, hence equals
by continuity. In order to exhibit
, it is enough to prove that for every
in
and
we have
|
(7)
|
for
-a.e
. The triangle inequality and
give
-a.e., with equality if and only if ( 7 ) holds. Integrating, we find
with equality if and only if ( 7 ) holds
-a.e. But equality does hold here since
is geodesic. □
45. Remark.
The above definitions and the general facts of this section hold more generally for the integral of a measurable field of metric spaces
over
. The only difference is that one must fix a choice of a section of base-points in order to define the
condition. With this addition, one may also consider infinite measure spaces
. None of this will be needed or used, therefore we leave details to the reader.
5.2
We now specialize to the
setting.
46. Lemma.
Suppose that
is a complete
space. Then
is also a complete
space.
-
Proof.
The space
is geodesic by (the trivial part of ) Proposition 44 . We need to check inequality ( 1 ) for
in
and
the midpoint of any geodesic line from
to
. By Proposition 44 ,
is the midpoint of a geodesic from
to
for a.e.
.
Therefore, the inequality holds pointwise and thus we can integrate it since it is linear in the squares of the distances. □
47. Example.
Let
be a Riemannian manifold of finite volume, and denote by
the associated volume form. For any
, the space of all positive definite symmetric bilinear forms on the tangent space
which also induce
is a
space, since it is isomorphic to the symmetric space
associated to
, where
. Thus, if
denotes the (normalized) probability space underlying
, then the space of “
Riemannian metrics” on
inducing
is isomorphic to
. Observe that it is endowed with a natural isometric action of the space of volume-preserving diffeomorphisms of
when
is compact.
48. Remark.
It follows from Proposition 44 that for
complete
the boundary of
can be easily described as what we call a join integral
where the right hand side stands for the set of pairs
consisting of a measurable map
and a semi-density
on
; we identify
with
when
-a.e. and
-a.e. We point out to the interested reader that this join integral behaves well with respect to Tits geometry ; indeed, our definition of the right hand side
makes sense for more general (fields of) spaces replacing
and can be endowed with a natural metric by means of integrating the spherical cosine law. We shall not use any of this since the boundary of non-proper spaces contains too little information for our purposes.
The next proposition establishes that for a
space
, Euclidean rectangles (cf.
Section 3.4 ) in
decompose as a “field of parallelograms” in
over
(of course, the latter need not be rectangles, as shows even the simplest possible example of the decomposition
, which corresponds to an atomic
).
49. Proposition.
Suppose that
is complete
. Let
be an interval and
two geodesics determining a Euclidean rectangle. Let
and
be as in Proposition 44 for
,
. Then for almost every
the function
is constant on
. Moreover,
-a.e.
-
Proof.
The function
is constant on
. On the other hand, the functions
are convex and non-negative. Thus the first part of the proposition follows from the general fact that an integral of the squares of a family of convex non-negative functions on
parametrized by a finite measure space is constant if and only if almost every function in the family is constant.
Since
bound a Euclidean rectangle in
, we may for any two
apply this first part of the proposition to the two geodesics
,
. The conclusion is precisely that
holds for a.e.
. □
5.3
For any
, where
is complete
, there is a unique point
minimising
; this point is called the barycentre of
. Indeed, we may embed
isometrically into
by
,
. Since the image is a closed convex subspace, the barycentre can be defined by the nearest point projection of
to that image. In particular, it follows from this definition that for the barycentres
of
we have
The special case
yields
Actually, the first inequality can be strengthened to
|
(8)
|
(though we will only use this in Appendix A and with
replaced by a finite set). Indeed, for
define
as in Section 3.2 . Integrating ( 2 ) with
,
and using
yields ( 8 ) when
goes to zero. Notice that ( 8 ) also reads
|
(9)
|
More refined inequalities can be found in [KS,2.5.2] .
5.4
A particular case of Definition 41 arises as follows. Let
be a locally compact second countable group,
a uniform lattice and
a metric space with a
-action by isometries.
Since
is cocompact and
second countable, one can find a relatively compact Borel right fundamental domain
with the property that for any compact
the set
|
(10)
|
is finite, see Exercice 12 of [Bo3,VII§2] (just mind that the scholar of Nancago considers left fundamental domains). We shall from now on agree to consider only such domains; we further endow
with a Haar measure
normalized by
and write
for
.
50. Remark.
We assumed
second countable so that its Borel structure is standard; in addition, we will often assume
separable. We do however prove the theorems of the introduction in the full generality stated there by showing in due time how to reduce to the current assumptions.
51. Definition.
We denote by
the space of all measurable maps (up to null-sets)
with separable range and such that (i) for all
,
one has
and (ii) for some (hence any)
, the function
is in
. We endow
with the metric defined by
|
(11)
|
There is a canonical isometry
|
(12)
|
given by restriction to
. Moreover, the choice of
is equivalent to the choice of a Borel map
such that
|
(13)
|
The isomorphism
induces a
-action on
which is described by the rule
(the dot notation emphasizes the difference between the two actions). Moreover, this action is measure-preserving since the existence of
forces
to be unimodular. Observe that the inverse to the restriction map in ( 12 ) consists in extending
to a map
|
(14)
|
We will abuse notation in omitting the subscript `ext'.
52. Lemma.
Assume that
is complete and separable. Then there is a well-defined continuous
-action by isometries on
defined by
, where
.
In other words, ( 14 ) and ( 13 ) show that the corresponding
-action on
is
|
(15)
|
and that the latter is well-defined. We call this the induced
-action.
-
Proof of the lemma.
First we need to show that
is still in
. Since
is relatively compact, the finiteness of ( 10 ) guarantees that there are
in
such that the union of the
covers
. Fix a base-point
. Then
But in view of
, each term
is finite since
is in
. This action preserves the distance ( 11 ). Ad continuity: By Lemma 12 , it is enough to show that the map
is measurable for all
in
. This follows from the fact that the map
,
is measurable. (Alternatively, for
, one can also show that continuous functions are dense using barycentres weighted by continuous approximate units on
.) □
5.5
Some properties of the
-action on
are trivially equivalent to the corresponding property for the
-action on
; for instance, the existence of fixed points. Evanescence is more subtle; a very simple instance is when
is a Hilbert space, in which case one has the following standard theorem: if a unitary representation of the cocompact lattice
does not weakly contain the trivial representation, then the induced
-representation does not either (this follows e.g. from the topological Frobenius reciprocity of [Bl] ). The following is a geometric generalization:
53. Theorem.
Let
be a locally compact second countable group,
a uniform lattice and
a complete separable
space with a
-action by isometries. If the
-action on
is evanescent, then the
-action on
is evanescent.
Observe that one cannot reduce this to a statement about asymptotic cones through Remark 26 ; indeed, even when
, an asymptotic cone on
is already a so-called non-standard hull of the Hilbert space
– whilst
is its own asymptotic cone.
Instead, the general idea is to project an evanescent sequence of functions to the space constant functions. This does not quite work since one needs to spread out the domain of these functions beyond any fundamental domain in order to capture generators of
. The following proof is particularly simple thanks to Lemma 54 ; another argument is given in Appendix B for certain non-uniform lattices.
-
Proof of Theorem 53 .
The finiteness of the set in ( 10 ) implies in particular that for every relatively compact Borel set
of positive measure and any
in
the restriction
is in
(as in Lemma 52 ). Therefore we may define a point
by taking the barycentre of
; that is,
minimises
For every
, define the element
of
by
(extending the notation of Section 5.3 ); finally, write
.
Fix a finite set
. We shall produce an unbounded
-evanescent set
for every unbounded
-evanescent set
in
with
a suitable compact set. This shows that weak evanescence of the
-action on
implies weak evanescence for the
-action on
; the statement for evanescence follows by Proposition 29 and Remark 30 .
Define a relatively compact set
and choose a compact set
with
54. Lemma.
The function
is bounded on each
-evanescent set in
.
-
Proof of the lemma.
Let
be a
-evanescent set. There is
such that
for all
and all
. Let
; we have
by the choice of
. After the change of variable
this is
which is just
This is bounded in terms of
and the measure of
. □
Choose now an unbounded
-evanescent set
in
and consider the subset
of
. We contend that
is
-evanescent. First, we check that
is unbounded: Indeed, since
, for all
|
(16)
|
is bounded by Lemma 54 , and therefore this first claim follows from the estimate
for
since
is unbounded. We now need to estimate
uniformly over
and
. Write
The first term is taken care of by ( 16 ) and Lemma 54 , whilst the second is bounded since
is
-evanescent and
. As for the last term, we have
so we are done by Lemma 54 . □
6 Superrigidity
Throughout this section, when considering irreducible lattices
, we shall always assume
. This is indeed not a restriction, because for
the definition of irreducibility implies
and there is nothing to prove.
6.1
We now give, in a slightly cumbersome formulation, a key intermediate statement to which various superrigidity statements will be reduced. A map is called totally geodesic if it takes geodesic segments to (possibly reparametrized) geodesic segments. Recall that the separability assumptions, made as a matter of convenience, will be disposed of in due time.
55. Theorem.
Let
be an irreducible uniform lattice in a product
of locally compact second countable groups and let
be a complete separable
space with a non-evanescent
-action by isometries.
Then there is a (canonical and minimal) non-empty closed convex
-invariant subspace
which splits isometrically and
-equivariantly as a product
of minimal
-spaces
. Moreover,
consists of continuous functions and the evaluation map
at
is a totally geodesic
-equivariant Lipschitz map. The restriction of
to every copy of each
is isometric. In particular, if for some
the set
is non-empty, then
preserves
and the
-action on
extends continuously to a
-action which factors through
.
As usual,
is really a function class and is said continuous if it contains a continuous representative. We shall use the following criterion, which is readily checked using Fubini-Lebesgue.
56. Lemma.
Let
be a measurable function from a locally compact second countable group
to a separable complete metric space
. Assume that for all
and all
there is a neighbourhood
of
such that
. Then
agrees almost everywhere with a continuous function. □
-
Proof of Theorem 55 .
Recall that
is the complete separable
space with continuous
-action granted by Definition 51 and Lemmata 43 , 46 and 52 . Theorem 53 shows that the
-action is not evanescent, and thus the splitting Theorem 9 implies that there is a closed convex
-invariant subspace
which splits isometrically and
-equivariantly as a product
of
-spaces
endowed with the product action.
Recall from Remark 40 that
can be assumed to be a minimal non-empty closed convex
-space.
We may, and shall from now on, identify each
with a subspace of
. For every
, there is a unique copy of
in
containing
; we denote by
this
-invariant closed convex subset
. We claim
|
(17)
|
Indeed, fix a Borel fundamental domain
for
as in Section 5.4 . We shall make use of the identification ( 12 ) and of the notation ( 15 ). It suffices to prove the claim for a.e.
.
Let
and let
be the geodesic from
to
. If
is any element of
for
, the splitting of
shows that
and
determine a Euclidean rectangle in
. Let
and
be as in Proposition 44 for
,
. Then Proposition 49 shows that
a.e. In other words, since
, the function (class)
is
-invariant for all
. Since the projection of
to
is dense, the subproduct
acts ergodically on
. Therefore,
(and hence also
) is constant; this constant is one by
. The claim now follows since
and
are the endpoints of
. Next, we recall (for any
) the inequality
; this allows us to bound the distance between any
by applying ( 17 ) to each factor
, obtaining
|
(18)
|
Since
is
-invariant, it follows that for any compact neighbourhood
of
which goes to zero as
by continuity of the
-action (Lemma 52 ). This shows that every
is continuous (Lemma 56 ).
We may now define a map
by
. This map is
-equivariant by definition. For every
and each
, the restriction of
to
is isometric because of ( 17 ). Further,
is
-Lipschitz by ( 18 ) and totally geodesic by Proposition 44 . Assume, for some
, that
; set
. Notice that
, being isometric to
, is complete, hence closed, and convex. In particular,
is closed in
[BH,p. 266] and the isometry
induces a homeomorphism
. The
-action on
being the product action,
is
-invariant and the (continuous)
-action on it factors through
. Summing up, we obtained a
-equivariant homeomorphism
, finishing the proof. □
6.2
In this subsection we analyse the notion of indecomposability introduced in Definition 3 (for proper spaces). The two obvious obstructions to indecomposability of a group
are (i) a
-fixed point at infinity (when
) and (ii) a
-invariant splitting
with unbounded factors; indeed in both cases we obtain closed invariant sets at infinity on which
does not act faithfully. It is not clear to us to what extent the topological part of Definition 3 is really an additional restriction in our setting.
Let
be a proper
space,
a subgroup and
a non-empty closed
-invariant subset; write
for the stabiliser of
in
. Recall that the topology of uniform convergence coincides on
with the compact-open topology and turns
into a topological group; the natural homomorphism
is continuous (compare Section 3.6 ). Moreover
is polish (hence Baire) because
is metrizable. Notice that
is closed in
and hence contains
.
57. Remark.
The conditions of Definition 3 hold if and only if
is a topological isomorphism onto its image. Indeed, indecomposability implies that
is a continuous group isomorphism from a locally compact second countable group onto a Baire group and thus is a topological isomorphism by the usual Baire category argument. The converse (which we do not use) holds since any locally compact subgroup of a topological group is closed (Corollaire 2 in [
Bo2,II§3no3]
).
In the two general instances where we verify indecomposability (Lemmata 59 and 60 ), we obtain the a priori stronger statement for the coarser topology of pointwise convergence by the following criterion (mind that
is not a topological group for this topology).
58. Lemma.
Let
be a locally compact second countable group with a faithful continuous action
on a compact Hausdorff topological space
. The associated injective homomorphism
is a topological isomorphism onto its image endowed with the topology of pointwise convergence if and only if for any sequence
tending to infinity in
there is
with
.
-
Proof.
Let
be any net of
converging to
for the pointwise topology on
but not in the
-topology. Since
is locally compact second countable and its action on
continuous, we obtain a sequence
converging to infinity in
such that
for all
. The criterion applied to
yields a contradiction. The converse follows from the continuity of the action. □
We now verify that indecomposability generalizes indeed Zariski-density for subgroups of adjoint simple algebraic groups.
59. Lemma.
Let
be a local field and
a connected adjoint
-simple
-group. Let
be the symmetric space, respectively the Bruhat-Tits building, associated to
according to whether
is Archimedean or not. Then any Zariski-dense subgroup
is indecomposable.
The following proof was kindly provided by J.-F. Quint.
-
Proof.
Let
be a non-empty
-invariant closed subset and
its pointwise stabiliser in
. Since
is an intersection of parabolic subgroups, it is an algebraic subgroup; but
normalizes it because it preserves
, so by Zariski-density
is normal in
and hence trivial.
Let now
be a sequence tending to infinity in
. Since parabolic subgroups are cocompact, there is no loss of generality in supposing that
is contained in a single
-orbit; thus
can be considered as a Zariski-dense subset of
for some parabolic
. We may fix an irreducible representation
of
on a
-vector space
(of dimension, say,
) such that
is the stabiliser of some line
, see [T] . Suppose for now that
; the Cartan decomposition gives
for some
in a positive Weyl chamber and
. Upon passing to a subsequence, we may assume that
converge to some
. Since
tends to infinity in
, the sequence
(where
is the operator norm) converges, after possibly extracting a subsequence, to a linear map onto a proper subspace
, with kernel
. Since
is irreducible,
so that by Zariski-density there is
not in
. However, for any line
,
tends to a line in
. Thus
. For general
, the same argument applies to the corresponding Cartan decomposition of
over
. □
For the sake of illustration, we now take a brief look at the context of negative curvature (even though the methods proposed in [MS1] ,[MS2] are more powerful in that case). Recall that if the proper
space
is Gromov-hyperbolic, e.g.
, an action on
is said elementary if it fixes a point in
or stabilises a pair of points in
.
60. Lemma.
Let
be a group with a non-elementary action by isometries on a Gromov-hyperbolic proper
space
. Then, upon possibly replacing
by a non-empty closed convex
-invariant subspace, the
-action on
is indecomposable.
(The
-action is called indecomposable if the image of
in
is an indecomposable subgroup.)
-
Proof.
By Remark 36 , there is a minimal non-empty closed convex
-invariant subspace
; observe that the
-action on
is still non-elementary. Thus we may assume
minimal. Recall that there is a continuous
-equivariant map
from distinct triples in
to
(indeed, for any distinct
, the sum
of Busemann functions is bounded below and proper as shown by a comparison argument e.g. using the arbre approximatif of Théorème 12 (ii) in [GH,2§2] ; therefore, one can take for
the circumcentre of its min-set). Let now
be a non-empty closed
-invariant set. By non-elementarity,
contains at least three distinct points
. The pointwise stabiliser
of
in
is therefore compact since it fixes
; thus, the closed convex subspace
is non-empty. Since
normalizes
, it preserves
; hence,
and thus
is trivial.
Let now
be any sequence in the stabiliser
of
such that
for all
. It remains only to show that the sequence
is bounded in
, or equivalently in
. This follows since
tends to
. □
6.3
We now proceed to prove Theorem 2 and Corollary 4 .
61. Proposition.
It is enough for Theorem 2 to consider the case where
is second countable.
-
End of proof of Theorem 2 .
Suppose first that the
-action on
is evanescent. Since
is proper, it follows from Proposition 27 that there is a
-fixed point
. In that case the conclusion of the theorem holds trivially for
.
Thus we may suppose that the action is non-evanescent; in addition, we may assume
second countable (Proposition 61 ) and
is separable since it is a proper metric space.
Therefore Theorem 55 applies; we shall use its notation. If every
is bounded, then
is reduced to a point by minimality. The image of this point under
is a
-fixed point in
, contradicting unboundedness. Hence we may assume that there is an index
such that
is unbounded. Since
is isometric to the unbounded closed convex subspace
of the proper space
, we deduce that its boundary is non-empty. Now the conclusion of Theorem 55 is exactly what was to be shown. □
-
Proof of Corollary 4 .
Keep the notation of the corollary, set
and let
be as in Theorem 2 . The conclusion of that theorem implies that the composed map
with
topologized as in Section 6.2 , extends to a continuous homomorphism
factoring through one of the canonical projections
. Since the image of
in
is dense and
closed,
. By Remark 57 ,
lifts to a continuous homomorphism
factoring through
and extending
. This proves the theorem since
. □
-
Proof of Margulis' Theorem 1 for
cocompact.
By Lemma 59 , we may apply Corollary 4 . □
6.4
With Theorem 6 in view, we now turn to complete
spaces
that are not assumed proper and analyse reduced subgroups of
following Definition 5 . In the beginning of Section 6.2 we mentioned two immediate restrictions following from indecomposability; the analogous two restrictions hold in the present setting aswell, as shown by (i) and (ii) in the following lemma.
63. Lemma.
Let
be a complete
space and
an unbounded reduced subgroup. Then: (i) The action is not evanescent.
(ii) There is no
-invariant splitting
with non-trivial factors.
(iii) If
is countable, then
is separable.
(iv)
has no non-trivial Clifford translation unless it is isometric to a Hilbert space.
-
Proof.
(ii) Assume that the splitting
is preserved by
. Since
is unbounded, one of the factors, say
, is unbounded. For any
the set
enjoys the property of Definition 5 . Therefore
and hence
is trivial.
(iv) This follows from (ii) in view of the canonical splitting of
recalled at the end of Section 3.4 .
(i) Suppose for a contradiction that
is an unbounded evanescent set. Since for every
the displacement function
is continuous and convex, there is no loss of generality in assuming
closed and convex upon replacing it by
. Then Definition 5 implies
. It follows now from this same condition that the unbounded
space
contains no unbounded closed convex subset
at all. In particular,
cannot be a Hilbert space and thus, by (iv),
does not admit any non-trivial Clifford translation. Since
acts non-trivially, there is
whose displacement function is not identically zero; however, since
, the supremum
is finite. If the displacement length were constant,
would be a non-trivial Clifford translation. Thus for some
the closed convex set
is non-empty, hence bounded. Since
, the circumcentre
is
-fixed. Let
be the circumradius of
. Since we assumed
unbounded, there is a sequence
in
with
. Let
be the point of
at distance
of
; then a comparison argument shows that
tends to zero since
. Thus, for
large enough,
is in
, a contradiction proving (i).
(iii) Let
be the closed convex hull of some
-orbit. Since
is unbounded and
-invariant,
. Thus it remains to show that
is separable. This is indeed the case: If
is any countable set (such as a
-orbit) and
denotes the set obtained by adjoining to
all midpoints of pairs in
, then the union
contains the midpoint of any of its pairs, and thus is dense in
. □
We can now complete the main case of Theorem 6 .
-
End of proof of Theorem 6 , non-Hilbertian case.
Observe first that
is countable since it is a lattice in a locally compact
-compact group. The argument of Proposition 61 shows that we may assume
second countable. Moreover, since
is unbounded and reduced, Lemma 63 implies that its action is non-evanescent and that
is separable. Thus, we may apply Theorem 55 ; we keep its notation and denote by
the point common to all
. We know, as in the proof of Theorem 2 , that at least some
is unbounded; say
. Recall that
is a closed convex subset of
since it is the isometric image of a complete convex set. Fix
and write
for the decomposition along
, where
. Then, for all
,
since
and we have a product action on
. Thus, for each
, the distance between
and
is bounded independently of
. Since the action is reduced, it follows that
. At this point, in order to conclude the proof and exhibit
via
, it only remains to show that
fixes
(equivalently, that
is trivial) for all
. Indeed, the continuity of
as defined in Section 3.5 follows from the continuity of
on
, and then
must range in
by irreducibility of
.
To that end, we note that the above discussion did not depend on the choice of the “copy”
; hence
also holds for all
. Now pick
for some
. Since
and
are isometries onto
, the map
defined by
is an isometry. Since
and
belong to a copy of
, we have
independently of
. Thus
is a Clifford translation; since we assume that
is not isometric to a Hilbert space,
is trivial by (iv) in Lemma 63 . This proves that
fixes
and therefore also
, finishing the proof of Theorem 6 in this case. □
6.5
We still have to prove Theorem 6 in the case where
is a Hilbert space. We provide a proof just for the sake of completeness; this is a very special case for which our methods are rather coarse — indeed it is not so natural to assume the action reduced here since the linear structure allows for stronger results, see [Sh] . Moreover, Theorem 6 does not hold as stated when
even in the simplest cases:
64. Example.
Consider the affine groups
and set
. Consider the irreducible unifom lattice
defined by
The
-action on
defined by
is unbounded and reduced. However it does not extend continuously to
. Indeed, if it did, the linear part of the
-action would have to factor through one of the
, say
, because the two canonical projections are the only homomorphisms
that are non-trivial on the diagonal. It then follows that
acts trivially altogether, which is impossible since the homomorphism
does not extend continuously to
.
A yet simpler example is
,
. Similar examples can be constructed in higher dimension.
The above example is rather typical for counter-examples to the statement of Theorem 6 :
65. Theorem.
The statement of Theorem 6 holds for any complete
space
unless (i)
is isometric to
for some
and (ii) the linear part of the
-action is irreducible and is the restriction of a
-subrepresentation of
endowed with the quasi-regular
-representation. Moreover, (ii) holds for at least two distinct indices
.
In particular, this theorem completes the proof of Theorem 6 since it is assumed in the latter that
.
We begin with a general observation about lattices in products:
66. Lemma.
Let
be a lattice in a product
of locally compact groups and let
be any homomorphism to a topological group
.
Suppose that for two distinct indices
there are continuous homomorphisms
such that the composed homomorphisms
extend
.
Then
is relatively compact in
.
-
Proof.
Let
be the two indices and define a continuous map
by
. Since
descends to a map
we obtain a probability Radon measure
on
from the (normalized) invariant measure on
. Since
for all
and
, the measure
is invariant under
, hence under
.
It remains to observe that the stabilizer
of any Radon probability measure on
is compact. Since
is Radon there is a compact set
with
. Therefore
for all
. It follows
. □
(The proof is slightly shorter when
is cocompact.)
-
Proof of Theorem 65 .
In view of the proof of Theorem 6 given in Section 6.4 under the assumption that
was not a Hilbert space, we may now assume that
is a Hilbert space.
Recall that a
-action by isometries on
is given by an orthogonal representation
and a cocycle
. We claim that
is irreducible. Indeed, if
were a non-trivial orthogonal decomposition preserved by
, then
would contradict the condition of Definition 5 since
is at finite distance of
for all
.
We adopt now the notations of the proof of Theorem 6 given in Section 6.4 ; the non-Hilbertian assumption on
was only used at the very end in order to prove that
fixes
(equivalently, that
is trivial) for all
. Therefore, we assume now that
is non-trivial for some
and need to characterize
and its
-action as in Theorem 65 . We obtained in the proof of Section 6.4 an isometry
such that for each
, the distance between
and
is bounded independently of
. Therefore, the new
-action on
transported via
(and
) from the
-action on
(i.e.
) differs from the original
-action by its translation cocycle only. In other words,
extends continuously to a homomorphism
where the orthogonal group
is endowed with the strong operator topology. Applying the same argument to
we are in position to use Lemma 66 and conclude that
ranges in a compact subgroup of
. The Peter-Weyl theorem implies that
is finite-dimensional since
is irreducible. It remains only to prove that
for
. The
-representation
induced from
is canonically isomorphic to
since
and hence
. Likewise,
and hence
. Since
is irreducible, the claim follows for
and is esteablished in the same way for
. □
Appendix A: Commensurator Superrigidity
Let
be a locally compact
-compact group,
a cocompact (or square-integrable, weakly cocompact) lattice and
a dense subgroup commensurating
, i.e.
has finite index in
for all
; equivalently, all
-orbits in
are finite.
In an unpublished manuscript [M2] (see also [Bu] ), Margulis proves the following theorem (under a more relaxed non-positive curvature assumption, assuming
compactly generated,
cocompact, and assuming there are no parallel orbits).
A1. Theorem.
Suppose
acts by isometries on a complete
space
such that the resulting
-action is non-evanescent. Then, upon possibly passing to a non-empty
-invariant closed convex subspace, the
-action extends continuously to a
-action.
Margulis' proof uses generalized harmonic maps. We give an elementary proof illustrating the techniques introduced above; in spirit, this is a non-linear analogue of [Sh] . In the particular case where both
and
are simple algebraic groups, this result leads to Margulis' arithmeticity criterion, see [M1] ,[AB] .
-
Proof.
Using Lemma 62 , we may assume
second countable; we can assume
countable and thus
separable upon passing to the closed convex hull of a
-orbit. Let
be the induced
-space, non-evanescent by Theorem 53 (resp. Appendix B). It is enough to show that there is a non-empty
-invariant closed convex subspace
such that for all
the function
on
is right
-invariant, since then it is essentially constant and we get a
-equivariant isometric map
by evaluation(s) (compare 6.1 ). Let
be the net of finite
-invariant sets
. For any
and a.e.
let
be the unique
minimising
; this is the barycentre construction (Section 5.3 ) for the uniform measure on
. We thus obtain a well-defined
-equivariant map
.
The barycentre inequality ( 9 ) of Section 5.3 yields
(19)
d
2
(
F
A
f
(
g
)
,
F
A
f
′
(
g
)
)
≤
|
A
|
−
1
∑
a
∈
A
d
2
(
f
(
g
a
)
,
f
′
(
g
a
)
)
−
|
A
|
−
1
∑
a
∈
A
(
d
(
f
(
g
a
)
,
f
′
(
g
a
)
)
−
d
(
F
A
f
(
g
)
,
F
A
f
′
(
g
)
)
)
2
The first term already implies that
is non-expanding (by integrating over
); likewise, ( 19 ) shows that if
are
-fixed, then
is invariant under the group
generated by the preimage of
in
. Every
-orbit being evanescent, it is bounded; thus
by Remark 15 – and we are done if
is finitely generated, taking
large enough and
.
For general
, let
be the set of
-components
(Remarks 39 ) such that
is
-invariant
. It follows from the preceding that
;
is convex (use e.g. Proposition 49 ) and closed since
-convergence of functions implies a.e.
subconvergence. Since
is bounded by non-evanescence, the directed family
has non-empty intersection by Theorem 14 and yields a component
as sought. □
Appendix B: Induction for Certain Non-Uniform Lattices
This appendix discusses non-uniform lattices. The cocompactness assumption was only used in Sections 5.4 and 5.5 , and it was needed only in defining the action on the induced space (Lemma 52 ) and in proving Theorem 53 . Thus, replacing them respectively with Lemma B2 and Theorem B5 below, we conclude that all our results (including Appendix A) hold as claimed in Theorem 7 . We insist however that for classical lattices the integrability condition discussed below is dependent on Margulis' arithmeticity theorem (interestingly, Margulis originally proved arithmeticity of non-uniform lattices without – and before – his superrigidity).
Let
be a locally compact second countable group,
a lattice. We use the notation of Sections 5.4 and 5.5 , e.g. for
,
and normalizing the covolume of
to one.
B1. Definition (See [Sh,1.II] ).
The lattice
is square-integrable if it is finitely generated and if, for the length function
associated to some (or equivalently any) finite generating set, there is a fundamental domain
(with null boundary) such that
(We note that when dealing with uniform lattices we never imposed finite generation, thus allowing for lattices in groups that are not compactly generated.)
Y. Shalom explains in [Sh,§2] why the condition of Definition B1 always holds for lattices as in Theorem 1 ; B. Rémy proves in [R2] that it holds for all Kac-Moody lattices. We refer to [R*] ,[R1] for general Kac-Moody groups, in particular for the following result of Rémy: Any Kac-Moody group over
is an irreducible lattice in the product of its associated twin building groups (modulo its finite centre), when
is large enough.
The following parallels [Sh,1.II] :
B2. Lemma.
Let
be as in Definition B1 . Then Lemma 52 and formula ( 15 ) provide a well-defined continuous
-action by isometries on
.
-
Proof.
The only additional verification we need to do is that for
,
and
the integral
is finite. The latter is
in view of ( 15 ); since
is finite, it is enough to show that
is finite. Let
be a finite generating set and
the associated length function; since
is bounded by
, we conclude by square-integrability of
. □
The evanescence question (Theorem 53 ) is more difficult; we shall establish a geometric generalization of an argument given in the linear setting by Margulis [M1,III.1] .
B3. Definition ([M1,III.1.8] ).
The lattice
is weakly cocompact if the
-representation
(i.e. the orthogonal complement of the trivial representation in
) does not almost have non-zero invariant vectors (compare 3.10 ).
The definition is obviously satisfied whenever
, or equivalently
, has Kazhdan's property (T); this disposes right away with most higher rank groups. According to Margulis [M1,II.1.12] , it also holds for connected semisimple Lie groups
even when they are not Kazhdan, see [Be] for a proof. Any Kac-Moody group over
whose Cartan matrix has finite entries is Kazhdan whenever
is large enough by a general result of Dymara-Januszkiewicz [DJ] .
B4. Remark.
If there is an asymptotically invariant sequence
(see Remark 33 ) of non-negative functions in
such that for every relatively compact
the integral
tends to zero, then
is not weakly cocompact.
Indeed, in that case
, thus the norm of the projection of
on
tends to one, yielding (after renormalization) an asymptotically invariant sequence in
.
B5. Theorem.
Assume that
is square-integrable and weakly cocompact.
Let
be a complete separable
space with a non-evanescent
-action by isometries. Then the
-action on
is non-evanescent.
-
Proof.
Fix
. Suppose for a contradiction that there is an evanescent sequence
in
such that
. Let
be a non-negative continuous function on
of integral one; we may assume that
has compact support
with
null.
For
, we want to define
as the barycentre of
with respect to the measure
; we thus have to prove that for (some, hence any
) the integral
is finite for a.e.
(hence a.e.
). This follows from Tonelli's theorem applied to
(20)
∫
F
∫
F
d
2
(
f
n
(
h
−
1
g
)
,
x
)
η
(
h
)
d
h
d
g
=
∫
F
η
(
h
)
∫
F
d
2
(
f
n
(
h
−
1
g
)
,
x
)
d
g
d
h
=
∫
K
η
(
h
)
d
2
(
h
f
n
,
ψ
x
)
d
h
≤
sup
{
d
2
(
h
f
n
,
ψ
x
)
:
h
∈
K
}
<
∞
.
By the definition of barycentres (Section 5.3 ),
|
(21)
|
Thus,
is square-integrable because we apply ( 20 ) to
Since in addition
is
-equivariant by definition,
. Setting
in ( 21 ) yields
|
(22)
|
which is bounded independently of
by evanescence of
. It follows that
is also an evanescent sequence with
. Define now
by
.
We claim that
is an evanescent sequence in the linear
-space
. Indeed, since
, the triangle inequality gives
so that by Minkowski's inequality and the definition of
The first term is bounded over compact sets by evanescence of
and the second by continuity of the
-action on
; the claim follows. In particular, since
, the sequence
is asymptotically invariant.
The goal now is to contradict weak cocompactness by applying Remark B4 to
.
Therefore it suffices to show that for any relatively compact
the integral
is bounded independently of
.
Since the
-action on
is not weakly evanescent (Proposition 29 ), there is by Lemma 25 a finite set
,
and
such that
for all
. Thus,
Therefore,
(23)
λ
∫
C
φ
n
(
g
)
d
g
≤
∑
γ
∈
F
∫
C
d
(
f
¯
n
(
g
γ
)
,
f
¯
n
(
g
)
)
d
g
+
d
0
≤
∑
γ
∈
F
(
∫
C
d
2
(
f
¯
n
(
g
γ
)
,
f
¯
n
(
g
)
)
d
g
)
1
/
2
+
d
0
.
It is now enough to prove that for all
the integral
is bounded independently of
. To that end, set
,
for any
,
. By definition of
and change of variable,
is the minimiser
of
. Therefore, the inequality corresponding to ( 21 ) yields
Thus,
where the integral over
is finite because we bound the above double integral by
which is finite by relative compactness of
. Moreover, the latter term is bounded independently of
in view of
since the first summand here is bounded by evanescence and the second has been treated previously with ( 22 ). This concludes the proof. □
References
-
Norbert A'Campo and Marc Burger, Réseaux arithmétiques et commensurateur d'après G. A. Margulis, Invent. Math. 116 (1994), no. 1-3, 1–25.
-
Werner Ballmann, Lectures on spaces of nonpositive curvature, DMV Seminar, vol. 25, Birkhäuser Verlag, Basel, 1995, With an appendix by Misha Brin.
-
Mohammed El Bachir Bekka, On uniqueness of invariant means, Proc. Amer. Math. Soc. 126 (1998), no. 2, 507–514.
-
Mohammed E. B. Bekka and Alain Valette, Kazhdan's property
and amenable representations, Math. Z. 212 (1993), no. 2, 293–299.
-
Vitaly Bergelson and Joseph Rosenblatt, Mixing actions of groups, Illinois J. Math. 32 (1988), no. 1, 65–80.
-
Philippe Blanc, Sur la cohomologie continue des groupes localement compacts, Ann. Sci. École Norm. Sup. (4) 12 (1979), no. 2, 137–168.
-
Nicolas Bourbaki, Éléments de mathématique. Fascicule II. Livre III, Topologie générale, Chapitres 1 et 2. Quatrième édition. Actualités Scientifiques et Industrielles, No 1142, Hermann, Paris, 1965.
-
Nicolas Bourbaki, Éléments de mathématique. Fascicule III. Livre III, Topologie générale, Chapitres 3 et 4. Troisième édition. Actualités Scientifiques et Industrielles, No 1143, Hermann, Paris, 1960.
-
Nicolas Bourbaki, Éléments de mathématique, Fascicule XXIX. Livre IV, Intégration, Chapitres 7 et 8, Actualités Scientifiques et Industrielles, No 1306, Hermann, Paris, 1963.
-
Martin R. Bridson and André Haefliger, Metric spaces of non-positive curvature, Grundlehren der Mathematischen Wissenschaften 319, Springer, Berlin, 1999.
-
Marc Burger, Rigidity properties of group actions on
-spaces, Proceedings of the International Congress of Mathematicians, Vol. 1, 2 (Zürich, 1994) (Basel), Birkhäuser, 1995, pp. 761–769.
-
Marc Burger, Alessandra Iozzi, and Nicolas Monod, Equivariant embeddings of trees into hyperbolic spaces, To appear in Int. Math. Res. Not. (2004).
-
Jan Dymara and Tadeusz Januszkiewicz, Cohomology of buildings and their automorphism groups, Invent. Math. 150 (2002), no. 3, 579–627.
-
Patrick Eberlein, Isometry groups of simply connected manifolds of nonpositive curvature. II, Acta Math. 149 (1982), no. 1-2, 41–69.
-
Étienne Ghys and Pierre de la Harpe (eds.), Sur les groupes hyperboliques d'après Mikhael Gromov, Birkhäuser Verlag, Basel, 1990, Papers from the Swiss Seminar on Hyperbolic Groups held in Bern, 1988.
-
Detlef Gromoll and Joseph A. Wolf, Some relations between the metric structure and the algebraic structure of the fundamental group in manifolds of nonpositive curvature, Bull. Amer. Math. Soc. 77 (1971), 545–552.
-
Mikhaïl Gromov and Pierre Pansu, Rigidity of lattices: an introduction, Geometric topology: recent developments (Montecatini Terme, 1990) (Berlin), Lecture Notes in Math., vol. 1504, Springer, Berlin, 1991, pp. 39–137.
-
Alain Guichardet, Sur la cohomologie des groupes topologiques, II, Bull. Sci. Math. (2) 96 (1972), 305–332.
-
Jürgen Jost, Nonpositive curvature: geometric and analytic aspects, Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 1997.
-
Jürgen Jost and Shing-Tung Yau, Applications of quasilinear PDE to algebraic geometry and arithmetic lattices , Conf. Proc. Lecture Notes Algebraic Geom., I, Internat. Press, Cambridge, MA, 1993, pp. 169–193.
-
, Harmonic maps and superrigidity, Tsing Hua lectures on geometry & analysis (Hsinchu, 1990–1991), Internat. Press, Cambridge, MA, 1997, pp. 213–246.
-
, Harmonic maps and rigidity theorems for spaces of nonpositive curvature, Comm. Anal. Geom. 7 (1999), no. 4, 681–694.
-
Shizuo Kakutani and Kunihiko Kodaira, Über das Haarsche Maß in der lokal bikompakten Gruppe, Proc. Imp. Acad. Tokyo 20 (1944), 444–450.
-
Nicholas J. Korevaar and Richard M. Schoen, Sobolev spaces and harmonic maps for metric space targets, Comm. Anal. Geom. 1 (1993), no. 3-4, 561–659.
-
H. Blaine Lawson, Jr. and Shing-Tung Yau, Compact manifolds of nonpositive curvature, J. Differential Geometry 7 (1972), 211–228.
-
, Discrete subgroups of semisimple Lie groups, Springer-Verlag, Berlin, 1991.
-
, Superrigidity for commensurability subgroups and generalized harmonic maps, unpublished manuscript.
-
Ngaiming Mok, Yum Tong Siu, and Sai-Kee Yeung, Geometric superrigidity, Invent. Math. 113 (1993), no. 1, 57–83.
-
Nicolas Monod, Arithmeticity vs. non-linearity for irreducible lattices, Preprint, 2004.
-
Nicolas Monod and Yehuda Shalom, Cocycle superrigidity and bounded cohomology for negatively curved spaces, J. Differential Geometry 67 (2004), 1–61.
-
, Negative curvature from a cohomological viewpoint and cocycle superrigidity, C. R. Acad. Sci. Paris Sér. I Math. 337 (2003), no. 10, 635–638.
-
Bertrand Rémy, Groupes de Kac-Moody déployés et presque déployés, Astérisque (2002), no. 277, viii+348.
-
, Kac-Moody groups as discrete groups, Preprint, 2003.
-
, Integrability of induction cocycles for Kac-Moody groups, Preprint, 2004.
-
Yuriĭ G. Reshetnyak, Non-expansive maps in a space of curvature no greater than
, Sibirsk. Mat. Ž. 9 (1968), 918–927.
-
Viktor Schroeder, A splitting theorem for spaces of nonpositive curvature, Invent. Math. 79 (1985), no. 2, 323–327.
-
Yehuda Shalom, Rigidity of commensurators and irreducible lattices, Invent. Math. 141 (2000), no. 1, 1–54.
-
Jacques Tits, Représentations linéaires irréductibles d'un groupe réductif sur un corps quelconque, J. Reine Angew. Math. 247 (1971), 196–220.
-
Tyakal N. Venkataramana, On superrigidity and arithmeticity of lattices in semisimple groups over local fields of arbitrary characteristic, Invent. Math. 92 (1988), no. 2, 255–306.
-
Robert J. Zimmer, Ergodic theory and semisimple groups, Birkhäuser Verlag, Basel, 1984.
University of Chicago