Last time we introduced Poisson boundaries hoping to use them to distinguish between lattices of and
,
. More precisely, we followed Furstenberg to define and construct Poisson boundaries as a tool that would allow to prove the following statement:
Theorem 1 (Furstenberg (1967)) A lattice of
can’t be realized as a lattice in
for
(or, in the language introduced in the previous post,
,
, can’t envelope a discrete group enveloped by
).
Here, we recall that, very roughly speaking, these Poisson boundaries were certain “maximal” topological objects attached to locally compact groups with probability measures
in such a way that the points in the boundary
were almost sure limits of
-random walks on
and the probability measure
was a
-stationary measure giving the distribution at which
-random walks hit the boundary.
For this second (final) post, we will discuss (below the fold) some examples of Poisson boundaries and, after that, we will sketch the proof of Theorem 1.
Remark 1 The basic references for this post are the same ones from the previous post, namely, Furstenberg’s survey, his original articles and A. Furman’s survey.
1. Some examples of Poisson boundaries
1.1. Abelian groups
The Poisson boundary of a locally compact Abelian group with respect to any probability measure
is trivial.
Indeed, in terms of the characterization of Poisson boundaries via -harmonic functions, this amounts to say that bounded
-harmonic functions on locally compact Abelian groups are constant, a fact proved by Choquet-Deny here.
More generally, there is a natural notion of random walk entropy (cf. Furman’s survey) allowing to characterize the triviality of Poisson boundary: more precisely, Kaimanovich-Vershik showed here that a discrete countable group
equipped with a probability measure
has trivial Poisson boundary if and only if
.
Remark 2 In addition to these results, it is worth to mention that:
- Furstenberg showed that any probability measure
on a locally compact non-amenable group
whose support
generates
admits bounded non-constant
-harmonic functions;
- Kaimanovich-Vershik and Rosenblatt showed that given a locally compact amenable group
, there exists a (symmetric) probability measure
with full support
such that all bounded
-harmonic functions on
are constant.
See Furman’s survey for more comments on these results (as well as related ones).
1.2. Free group on 2 generators
Let be the free group on
generators
and
equipped with the Bernoulli probability measure
for
. Last time, during the naive description of the Poisson boundary, we saw that the space
of infinite words on
satisfy the non-cancellation condition
for all
was a natural candidate for a boundary of
(as far as
-random walks were concerned).
1.2.1. Computation of the Poisson boundary of
As the reader can imagine, equipped with an adequate measure
is the Poisson boundary of
. The proof of this fact goes along the following lines. Firstly, let us show that
is a
-space, i.e.,
Lemma 2
acts continuously on
(for the topology of pointwise convergence).
Proof: The action of on
is defined by continuity: given
and
,
, let
. In other words, if
, then
is the infinite word obtained by concatenation of
and
, and then collapsing if necessary: e.g., if
, then
,
and
, and
This proves the lemma.
Next, let us consider a stationary sequence of independent
-valued random variables with distribution
(so that
). We claim that
converges to a point
of
with probability
:
Lemma 3
with probability
.
Proof: The basic idea to get convergence is simple: we write the product , we delete all pairs
canceling out (i.e.,
), and we continue to delete “canceling pairs” until we get a point
in
.
Of course, this procedure fails to produce a limit point only if the product
keeps degenerating indefinitely.
However, this possibility occurs only with zero probability thanks to the transience of the random walk on the free group
(i.e, the fact that this random walk wanders to
with full probability; cf., e.g., this classical paper of Kesten).
Equivalently, letting denote the length of an element
(i.e., the minimal length of products expressing
in terms of the generators
), the transience of
means that
as
with probability
. In particular, with probability
, for each
, we can select
the largest value of
such that
, so that the right multiplication of
by
will not change the first
letters of the product
, and, a fortiori, we get (with probability
) a well-defined limit word that we can denote without any ambiguity
This proves the lemma.
An immediate corollary of the proof of this lemma (and the definition of pointwise convergence) is:
Corollary 4 For any
,
converges to
. In particular, given any probability measure
on
, one has
Another way of phrasing this corollary is:
carries only one
-stationary measure
.
Indeed, from the limit point provided by the previous lemma, we can define a sequence
of
-valued random variables by setting
and it is not hard to check that
is a
-process (as defined in the previous post). Therefore, if
is a
-stationary probability measure on
, the convergence asserted in the corollary and Proposition 7 of the previous post says that
is the distribution of a
-process. Furthermore, Proposition 5 of the previous post asserts that this
-process is exactly
, that is, the unique
-stationary measure
on
is the distribution of the
-process
.
In summary, denoting by the unique
-stationary on
(namely, the distribution of
), our discussion so far shows that
is a boundary of
. In fact, one can make this boundary entirely explicit because it is not hard to guess
:
Lemma 5 Let
be the probability measure on
giving the maximal independence between the coordinates, i.e.,
takes one of the four values
with equal probability,
takes one of the three possible values
with equal probability, etc., so that the
-measure of a cylinder obtained by prescribing the first
-entries is
, i.e.,
Proof: From our previous discussion, it is sufficient to check that is
-stationary, i.e.,
. This fact is not hard to check: we have just to verify that the
-integral and the
-integral of characteristic functions of cylinders coincide, i.e., we have to compute (and show the equality of) the
-measure and the
-measure of any given cylinder
In this direction, note that, by definition,
On the other hand, for , we have that
is a cylinder of size
(i.e., corresponding to the prescription of
entries) unless
where
is a cylinder of size
. In particular,
By plugging this into the previous equation, we deduce that
as desired. This proves the lemma.
At this point, it remains only to check that is the Poisson boundary, i.e., all boundaries of
are equivariant images of
and the Poisson formula for
-harmonic functions.
We will skip the proof of this last fact because it would lead us far from the scope of this post. Instead, we refer to Dynkin-Maljutov paper where it is shown that is the Martin boundary (and, a fortiori, the Poisson boundary).
1.2.2. A law of large numbers for
Using our knowledge of the Poisson boundary of
, let us sketch a proof of the fact that
as with probability
(where
is the length of
).
Firstly, let us observe that is absolutely continuous with respect to
for each
. Indeed, since
is
-stationary, i.e.,
our claim is true for the generators of
, and, a fortiori, it is also true for all
.
Actually, the Radon-Nikodym density is not hard to compute: if
is the limit of the sequence
, then one can check by induction on
that
for all sufficiently large. In other words,
Consider now the quantity
For each , we know that
. From this, it is possible to prove (after some work [related to the fact that we want to let
and
vary…]) that
By Birkhoff’s ergodic theorem, the time-average on the right-hand side of this equality converges to the spatial-average with probability , i.e.,
Using (1), we can compute this integral as follows. By definition of ,
Now, for each and
, we have that
unless
in which case
. In particular, from (1), we get that
for each . By plugging this into the previous equation, we obtain that
and, hence,
i.e., as
with probability
.
1.2.3. A random walk on the free group on
generators
The arguments of the previous subsubsections also apply to the free groups on
generators equipped with the Bernoulli probability assigning equal probabilities to the
singletons consisting of the generators and their inverses.
However, the simple-minded extension of this discussion to the free group on countably many generators fails because we do not have a Bernoulli measure.
Nevertheless, it is possible to equip with some probability measure
such that the Poisson boundary of
coincides with the Poisson boundary
of
.
In fact, this is so because “behaves like a lattice” of
. More concretely, the important fact about
is that it is isomorphic to the commutator subgroup
of
. Using this fact, we can formalize the idea that
is a lattice of
through the following lemma:
Lemma 6
(or, more precisely, the commutator subgroup
) is a recurrence set for the random walk
on
in the sense that
meets
infinitely often with probability
.
Proof: The quotient group is the free Abelian group on
generators, i.e.,
. Now, by projecting to
the random walk
on
, we obtain the simple random walk
on
. In this notation, the assertion that
is a recurrence set of
corresponds to the well-known fact that the simple random walk
on
returns infinitely often to the origin.
From this lemma, we can construct (at least heuristically) a probability measure on
whose Poisson boundary is the same of
(namely,
). In fact, letting
be the random walk on
, we know that (with probability
) there is a subsequence
hitting
. On the other hand,
converges to a limit point
. In particular, it follows that the boundary points
are also limits of the
-valued random variables
. Thus, if one can show that
are independent random variables with a fixed distribution
, then
has Poisson boundary
. Here, the keyword is the strong Markov property of
: more precisely, we notice that from
to
we multiply (to the right)
; since
and
belong to
, we have that
and, in particular, this suggests that
is the distribution of the position of
at the first time that
enters
; of course, to formalize this, we need to know that the the entrance times
of
on
are themselves random variables (or rather that
are still independent random variables) and this is precisely a consequence of the strong Markov property of
.
Alternatively, one can show that has Poisson boundary
by working with harmonic functions. More precisely, our task consists into showing that the restrictions to
of
-harmonic functions on
are
-harmonic and all
-harmonic functions on
arise in this way (by restriction of
-harmonic functions on
). In this direction, one recalls that, for each
, the quantity
is the probability that
takes value
at the first time it enters
, and then one shows the desired facts about
-harmonic functions versus
-harmonic functions with the aid of the following abstract lemma:
Lemma 7 Let
be a discrete,
a probability measure on
and
a stationary sequence of independent random variables with distribution
. Suppose that
is a recurrence set of
and let
be a
-harmonic function. Then,
where
is the distribution of the first point of
hit by
.
We refer to pages 31 and 32 of Furstenberg’s survey for a (short) proof of this lemma and its application to the computation of the Poisson boundary of .
1.3. ,
Let ,
, and denote by
the (maximal compact) subgroup of orthogonal matrices in
and
the (Borel or minimal parabolic) subgroup upper triangular matrices in
. From the Gram-Schmidt orthogonalization process, we know that
.
Definition 8 A probability measure
on
is called spherical if
is absolutely continuous with respect to Haar measure on
and
is
-bi-invariant, i.e.,
for all
, or, equivalently,
for all
.
Let us now consider the homogenous space (of complete flags in
), i.e., a space where
acts continuously and transitively (by left multiplication
on cosets
, of course).
Since acts transitively on
, there exists an unique
-invariant measure on
that we will denote
(philosophically, this is comparable to the fact that the unique translation-invariant measure on the real line is Lebesgue). Now, given any probability measure
on
and any spherical measure
on
, note that
is a
-invariant probability measure on
, and, a fortiori,
. In particular,
is the unique
-stationary measure on
.
In this context, Furstenberg proved the following theorem:
Theorem 9 (Furstenberg)
is the Poisson boundary of
whenever
is a spherical measure.
We will not comment on the proof of this theorem here: instead, we refer the reader to the original article of Furstenberg for more details (and, in fact, a more general result valid for all semi-simple Lie groups ).
An interesting consequence of this theorem is the following fact:
Corollary 10 The class of
-harmonic functions on
is the same for all spherical measures
.
Proof: By the definition of the Poisson boundary, the fact (ensured by Furstenberg’s theorem) that is the Poisson boundary of
means that we have a Poisson formula
giving all -harmonic functions
on
as integrals of bounded measurable functions
on
. Since the right-hand side of this equation does not depend on
(only on
), the corollary follows.
From this corollary, it is natural to define a harmonic function on as a
-harmonic function with respect to any spherical measure
.
For later use, let us describe more geometrically the Poisson boundaries and
(resp.) of
and
(resp.). In fact, we already hinted that, in general,
is the complete flag variety of
, but we will discuss the particular cases of
and
because our plan is to show Theorem 1 in the context of
and
only.
1.3.1. Poisson boundary of
Consider the usual action of on
and the corresponding induced action on the projective space
of directions/lines
in
associated to non-zero vectors
.
By definition, the subgroup of upper triangular matrices in
is the stabilizer of the direction
generated by the vector
. In particular, since
acts transitively on
, we deduce that
.
Let us now try to understand how is attached to
in terms of the measure topology described in the previous post.
By definition, an element is close to
whenever the measure
is close to
.
On the other hand, a “large” element
i.e., a matrix with large operator norm
has at least one large of the column vector, either
or
, and, if both column vectors are large, then their directions
and
are close by unimodularity of
(indeed, if they are both large and they form a rectangle of area
, then their angle must be small). In particular, we conclude that, for large elements
, most directions
are close to the larger of
,
except when
is approximately orthogonal to
,
(where
is the transpose of
).
In summary, we just proved the following lemma:
Lemma 11 Given
, there exists a compact subset
such that, if
, then there are two intervals
of sizes
with
.
An interesting consequence of this lemma concerns the action of large elements of on non-atomic measures on
.
Indeed, given an arbitrary non-atomic probability measure on
, we have that for each
there exists
such that any interval
of size
has
-measure
(because any point has zero
-measure by assumption).
In particular, by combining this information with the previous lemma, we deduce that any large element of
moves most of the mass of
to a small interval. More precisely, denoting by
and
the intervals of sizes
provided by the lemma, we have that
, i.e.,
. Therefore,
(since
), i.e., most of the mass of
is concentrated on
. In other words, we proved the following crucial lemma:
Lemma 12 Let
be a non-atomic probability measure on
. Then, as
,
converges to a Dirac mass.
An alternative way of phrasing this lemma is: as , it approaches a definite point of
. In particular,
is compact, that is, we get a compactification of
by attaching
! (This result is comparable to Corollary 4 in the context of free groups.) As we will see below, this compacteness property is no longer true for
(i.e., in the context of
), and this is one of the main differences between the Poisson boundaries of
and
. As the reader might imagine, this observation will be important for the proof of Theorem 1 (that is, at the moment of distinguishing between the lattices of
and
).
Note that, from this lemma, we can show directly that is a boundary of
whenever
is spherical (without appealing to Furstenberg’s theorem above). Indeed, denoting by
the random walk with respect to
, we have that, by definition,
is a boundary of
if and only if
converges to a Dirac mass with probability
. Since
is a compact space, we know that
converges to some measure with probability
(cf. Corollary 11 of the previous post) and the previous lemma says that this measure is a Dirac mass unless the products
stay bounded (with positive probability). Now, a random product of elements in a topological group remains bounded (with positive probability) only if the distribution is supported on a compact subgroup (by Birkhoff’s ergodic theorem and the fact that a compact semigroup of a group is a subgroup), but this is not the case as
is spherical (and thus absolutely continuous with respect to Haar).
Finally, once we know that is a boundary of
, i.e.,
converges to a Dirac mass
where
is a
-process on
, we can obtain the following interesting consequences. Firstly,
, i.e., the random walk is transient. Moreover, the direction of the larger of the column vectors of
converges with probability
to a direction
, and, in fact, it can be shown that both column vectors become large. In particular, by unimodularity, it follows that, with probability
, both column vectors of
converge to a random point (direction) of
.
1.3.2. Poisson boundary of
By Furstenberg’s theorem, is the space underlying the Poisson boundary of
equipped with any spherical measure
. Here, we recall that
is the subgroup of upper triangular
matrices in
.
We proceed as before, i.e., let us consider the usual action of on
and the induced action on the space
of pairs
of points in the projective space
corresponding to orthogonal directions
and
in
.
Note that acts naturally on
via
(our choice of
is natural because the orthogonality condition is preserved and
is an automorphism of
[in particular, our matrices are multiplied in the “correct order”]). Moreover, this action is transitive, so that
is the quotient of
by the stabilizer of a point of
. For sake of concreteness, let us consider the point
(where
are the vectors of the canonical basis of
) and let us determine its stabilizer in
. By definition, if
, say
stabilizes , then
and
, i.e.,
and
that is, is upper triangular. In other words,
is the stabilizer of
and, thus,
.
Next, let us again try to understand how is attached to
in terms of the measure topology. Once more, by performing the random walk
with distribution
, one can check that the column vectors of
converge to a limit vector
in
. In terms of the boundary
, we can recover
by letting
be the point (
-process) of
obtained as the almost sure limit of
.
Of course, in this description, the role of remains somewhat mysterious, and so let us try to uncover it now. By definition,
is the limit point of the column vectors of
, i.e., the random walk obtained from
by applying the automorphism
. On the other hand, if
are the column vectors of
, then the vectors
,
and
are the column vectors of
(where
is the vector product of
). In particular, the fact that these vectors converge to
means that the perpendicular directions to the 3 planes passing through the vectors
converge to
, i.e., all of these 3 planes converge to the plane perpendicular to
.
Finally, let us make more precise the observation from the previous subsubsection that the boundary behaviors of elements in and
are different as this is one of the main points in the proof of Theorem 1.
Once more, let us point out that Lemma 12 above says that any element approaches the boundary (i.e., becomes large) by getting close to a specific point
. On the other hand, this is no longer true for elements
: for instance, the sequence
goes to infinity without getting close to any specific point of because its larger column vectors
and
do not converge together to a single direction in
. Instead, it is not hard to convince oneself (by playing with the definitions) that the sequence of measures
converges to a measure supported on the great circle
. In general, a great circle is a fiber of one of the two natural fibrations
and it is possible to show that the measures
can approach measures supported on any given great circle.
Anyhow, the basic point is that is compact but
is not compact. For sake of completeness, let us point out what are the measure that we must add to
to get a compactification of
. It can be proven that any convergent sequence
tends to a Dirac mass, a circle measure (i.e., a measure supported on a great circle) or an absolutely continuous measure with respect to
(this last case occurring only if
is bounded in
). Moreover, in the case of getting a circle measure as a limit, we get a very specific object: by identifying the great circle with
and denoting by
the Lebesgue measure, one has that all circle measure have the form
where
(and
acts on its boundary
).
1.3.3. Harmonic functions on
Closing this quick discussion on the Poisson boundaries of , let us quickly comment on the relationship between
-harmonic functions on
and classical harmonic functions on Poincaré’s disk.
Let be a spherical measure on
. For the sake of simplicity of notation, we will say that a random walk
with respect to a spherical measure is a Brownian motion. By definition of sphericity of
, the transition from
to
has the same probability as the transition from
to
, and from
to
for each
. In particular, the Brownian motion
can be transferred to a Brownian motion
on the symmetric space
of
.
Now, we recall that is Poincaré’s disk
/hyperbolic upper-half plane
: more concretely, by letting
act on the upper-half plane
by Möebius transformations, i.e., isometries of
equipped with the hyperbolic metric,
for , we see that
is the stabilizer of
. In particular,
is naturally identified with
via
, and, hence, we can also think of
as the Poincaré’s disk after considering the fractional linear transformation sending
to
in such a way that
is sent to the origin
and
is sent to
.
In summary, we can think of the Brownian motion as a Brownian motion
on Poincaré’s disk. Here, it is worth to point out that the transitions of
are not given by group multiplication as
acts on the left and the
‘s are multiplied from the right. Finally, if we transfer the measure topology on
to
, we get the usual Euclidean topology on the closed unit disk
. Indeed, suppose that
with
. Then,
transfers most of the mass of
to a point of
close to
. In particular, if
, then
, that is,
converges to
, and, a fortiori, the Euclidean topology on
is the topology we get after transferring the measure topology.
Finally, note that the value of a harmonic function (with respect to any spherical measure) depends only on the coset
(by sphericity and the mean value property of harmonic functions). Thus,
induces a function
on
. By Poisson’s formula (in the definition of Poisson boundary), we have that
On the other hand, by computing the density (using that
acts via Möebius transformations and
is the Lebesgue measure), we can show that
where is the classical Poisson kernel on the unit disc. In other words, by letting
, we obtain that
and, hence, the function is harmonic in the classical sense. In summary, the two notions of `harmonic’ are the same. Probabilistically speaking, the formula above says that the value
is obtained by integrating the boundary values
with respect to the `hitting measure
on the boundary starting at
‘ (as
is the distribution of the limit of
[because
is the distribution of the limit of
and by invariance of the Brownian motion under the group]).
1.4. Mapping class group and Teichmüller space
As it is “customary”, the mapping class groups and Teichmüller spaces are very close to lattices in Lie groups and homogenous spaces. In particular, this partly motivates these two articles of Kaimanovich and Masur about the Poisson boundary of the mapping class group and Teichmüller spaces, where it is shown that it is the Thurston compactification (via projective measure foliations) equipped with a natural harmonic measure. Of course, it is out of the scope of this post to comment on this subject and we refer the curious reader to (very well-written) papers of Kaimanovich-Masur.
2. Poisson boundary of lattices of
After this series of examples of Poisson boundaries, let us come back to the proof of Theorem 1. At this stage, we know that equipped with any spherical measure has Poisson boundary
and now we want to distinguish between lattices of
.
As we mentioned in the previous post, the basic idea is that a `nice’ random walk in a lattice of
should see the whole boundary of
. In fact, this statement should be compared with the results in Subsubsection 1.2.3 above where we saw that an adequate random walk in the free group
in
generators sees the whole boundary of the free group
in two generators because
behaved like a lattice in
, or, more accurately, it was a recurrence set for the symmetric random walk in
.
However, this heuristic for free groups can not be applied ipsis-literis to lattices of
because a countable set
can not be a recurrence set for a Brownian motion in
. Nevertheless, one still has the following result:
Theorem 13 (Furstenberg) If
is a lattice of
, then there exists a probability measure
on
such that the Poisson boundary of
coincides with the Poisson boundary
of
(with respect to any spherical measure).
In order to simplify the exposition, we will restrict our attention to the low dimensional cases. More concretely, we will sketch the construction of in the case of cocompact lattices in
and we will show that
has
as a boundary. However, we will not enter into the details of showing that
is the Poisson boundary of
: instead, we refer the reader to Furstenberg’s survey for a proof in the case of
(i.e.,
) and his original article for the general case.
2.1. Construction of in the case
Consider again the symmetric space associated to
. This space has a natural
-invariant metric
and, using this metric, we have a function
measuring the distance to the origin. Note that, in the particular case , the function
has a very simple expression:
for .
Proposition 14 If
is a cocompact subgroup of
, then there exists a probability measure
on
such that
- (a)
, i.e.,
for all
,
- (b)
is
-stationary, i.e.,
,
- (c) the restriction of the function
to
is
-integrable, i.e.,
Remark 3 The condition (b) above implies that the restriction to
of an arbitrary harmonic function on
is
-harmonic. In particular, this means that a harmonic function on
satisfies plenty (at least one per cocompact lattice of
) of discrete mean-value equalities.
For the proof of this proposition, we will focus on the construction of a measure satisfying item (b) and then we will adjust it to satisfy items (a) and (c). Also, we will discuss only the case of cocompact lattices in
.
In this direction, let us re-interpret item (b) in terms of the Brownian motion on Poincaré’s disk (that is, the symmetric space
of
). As we saw above,
is the hitting distribution on
of the Brownian motion starting at
.
In particular, the stationarity condition in item (b) says that the hitting probability on
starting at the origin
is a convex linear combination of the hitting probabilities
on
starting at the points
(for
). This hints how we must show item (b): the measure
will correspond to the weights
used to write
is a linear convex combination of
,
. Keeping this goal in mind, it is clear that the following lemma will help us with our task:
Lemma 15 Let
be a cocompact lattice of
and denote
the hitting probability on
of a Brownian motion starting at
. Then, there are two constants
and
such that, for any
, one has
where
, the points
and
are within a hyperbolic distance
of
, and
is a non-negative measure of total mass
.
Proof: Since is cocompact, it has a compact fundamental domain. In particular, we can find a large constant
such that the hyperbolic ball
of radius
around any
contains in its interior at least one point of the form
with
.
For sake of simplicity of the exposition, during this proof, we will replace the discrete-time Brownian motion by its continuous-time analog for a technical reason that we discuss now.
For each and
in the interior of
, the hitting measure
on
starting at
can be computed by noticing that a continuous-time Brownian motion emanating from
must hit the circle
before heading towards
. Of course, the same is not true for the discrete-time Brownian motion (as we can jump across
), but we could have overcome this small difficulty by considering an annulus around
. However, we will stick to the continuous-time Brownian motion to simplify matters. Anyhow, by combining this observation with the strong Markov property of the Brownian motion, one has that
where is the hitting distribution of
for a Brownian motion starting at
(i.e., for each interval
, the measure
is the probability that a Brownian motion starting at
first hits
at a point in
.
Next, let us consider the points of the form ,
, inside (the interior of)
, choose positive numbers
and consider the measure
By the previous formula for the hitting measures , we can rewrite the measure above as
Pictorially, this integral represents weighted contributions from the following Brownian motions:
At this point, we observe that the measures are absolutely continuous with respect to the Lebesgue measure on
whenever
belongs to the interior of
(as our Brownian motion is guided by a spherical measure [by definition]). Therefore, by taking
small (depending on how close
is to the circle
), we can ensure that the measure
appearing in the right-hand side of (2) is positive. Furthermore, note that this scenario is –invariant: if we replace
by
for
, the circle
is replaced by
and the elements
are replaced by
, but we can keep the same values of
.
In other words, the values of (making the measure in (3) positive), and, a fortiori, the quantity
, depends only on the coset
where
satisfies
. Therefore, by the compactness of
, we can find some
such that, for all
, the values of
(making (3) positive) can be chosen so that
In particular, it follows that is a positive measure of total mass
.
In summary, we managed to write
where is a positive measure of total mass
, as desired.
By taking in this lemma, we know that one can write
as a convex linear combination of a “main contribution” coming from
‘s (where the distance from
‘s to
are
) and a “boundary contribution” coming from an integral of
with respect to a measure
of total mass
.
From this point, the idea to construct a probability measure on
satisfying item (b) of Proposition 14 is very simple: we repeatedly apply the lemma to the
‘s appearing in the “boundary contribution” in order to push it away to infinity; here, the convergence of this procedure is ensured by the fact that the boundary measure
loses a definite factor (of
) of its mass at each step.
More concretely, this idea can be formalized as follows. By induction, assume that, at the th step, we wrote
where only for
such that the distance
between
and
is
and
is a positive measure on the hyperbolic ball
(of radius
and center
) of total mass
.
By applying the lemma to , we also have
where only for
, and
is supported in
and it has total mass
.
By combining these equations, we deduce that
where only for
, and
is a positive measure supported on
whose total mass is
In particular, by setting , we find that
that is, the stationarity condition of item (b) of Proposition 14 is proved.
Now, we claim that item (c) of Proposition 14 is satisfied by . Indeed, by construction, for the elements
with
, the quantity
comes from the
-combination of the contributions of the measures
in the right-hand side of (4). Since the measure
has total mass
, we deduce that
and, therefore,
In particular, given that any element with
appears
times in the sum above, we conclude that
that is, the -integrability condition on
in item (c) of Proposition 14 is verified.
Finally, the full support condition in item (a) of Proposition 14 might not be true for the probability measure constructed above. However, it is not hard to fulfill this condition by slightly changing the construction above. Indeed, it suffices to add all points
at distance
from
in the
th step of the construction of
and then assign to them some tiny but positive probabilities so that the measure
in the right-hand side of (4) is still positive. In this way, we are sure that in the end of the construction of
, all
‘s in
were assigned some non-trivial mass.
This completes the sketch of the proof of Proposition 14 in the case (i.e., cocompact lattices in
).
After constructing , let us show that the Poisson boundary
of
equipped with a spherical measure is a boundary of
.
2.2. is the Poisson boundary of
A reasonably detailed proof that is the Poisson boundary of
is somewhat lengthy because the verification of the maximality property (i.e., any boundary of
is an equivariant image of
) needs a certain amount of computation (in fact, we might come to this point later in a future post, but, for now, let us skip this point). In particular, we will content ourselves with checking only that
is a boundary of
in the cases
and
.
As it turns out, the fact that is a boundary of
(in the case
) is not hard. By definition, we have to show that a stationary sequence
of independent random variables with distribution
has the property that
converges to a Dirac mass with probability
. On the other hand, by Corollary 11 of the previous post and the compactness of
, we know that
converges to some measure with probability
, and, by Lemma 12, this limit measure is a Dirac mass if the elements
are unbounded. Now, it is clear that these elements are unbounded because
has distribution
and, by construction (cf. item (a) of Proposition 14),
is fully supported on a lattice of
(and, thus, its support is not confined in a compact subgroup).
Next, let us show that is a boundary of
(in the case
). In this direction, we will need the following lemma playing the role of an analog of Lemma 12 in the context of
:
Lemma 16 Let
be a probability measure on
such that:
- (i)
has a rich support: the support of
is not confined to a compact or reducible subgroup of
;
- (ii) the norm function is
–
-integrable:
;
- (iii)
is
-stationary:
Then, for any stationary sequence
of independent random variables with distribution
, the sequence of measures
converges to a Dirac mass on
with probability
.
Before proving this lemma, let us see how it allows to prove that is a boundary of
for the measure
constructed in Proposition 14, thus completing our sketch of proof of Furstenberg’s theorem 13. By definition of boundary, it suffices to check that the measure
provided by Proposition 14 fits the assumptions of Lemma 16. Now, by item (a) of Proposition 14,
is fully supported on the lattice
of
. Since a lattice of
is Zariski dense (by Borel’s density theorem),
is not a compact subgroup nor reducible subgroup of
, that is,
satisfies item (i) of the lemma above. Next, we notice that a computation shows that the distance function to origin
in the symmetric space
satisfies
and
. In particular, the integrability condition in item (c) of Proposition 14 implies that
satisfies item (ii) of the lemma above. Finally, the item (b) of Proposition 14 is precisely the stationarity condition in item (iii) of the lemma above.
So, let us complete the discussion in this section by sketching the proof of Lemma 16.
Proof: By the compactness of (and Corollary 11 of the previous post), we know that
converges to some measure with probability
. This allows us to consider the sequence
of measure-valued random variables.
Our task consists into showing that are Dirac masses. Keeping this goal in mind, note that
and
is independent of
for
. Also, let us observe that we can extend the sequence
can be extended to non-positive indices
by relabeling
as
and shifting the remaining variables. Here, by stationarity of
(by definition) and
(by item (iii)), the variables
with positive indices
are probabilistically isomorphic to the original sequence (before shifting). In other words, we can assume that our sequence to
is defined for all integer indices
. (In terms of Dynamical Systems, this is analog to taking the natural extension
,
of the unilateral shift
,
). In any case, we can write
where
is a stationary sequence of independent random variables and
is independent of
‘s.
At this point, let us recall the discussion of Subsubsection 1.3.2 on the Poisson boundary of . In this subsubsection we saw that there are only three possibilities for any limit of the measures
,
such as
: it is either a Dirac mass, a circle measure or an absolutely continuous (w.r.t.
) measure, the latter case occurring only when
stays bounded in
.
By ergodicity (of the shift dynamics underlying the sequence ), we have that only one of these possibilities for
can occur with positive probability.
Now, can not be absolutely continuous w.r.t.
because this would mean that
is bounded (with positive probability) and, a fortiori,
is confined to a compact subgroup of
, a contradiction with our assumption in item (i) about the distribution
of
‘s.
Therefore, our task is reduced to show that ‘s are not circle measures with probability
. For sake of concreteness, let us fix
by assuming that
is a circle measure supported on our “preferred” circle
. In order to show that
are Dirac masses, it suffices to check that the angles between the column vectors of the matrices
converge to
as
. In other terms, denoting by
,
,
the column vectors of
, and by noticing that
,
,
play symmetric roles, we want to check that
as .
The idea to show this is based on the simplicity of the Lyapunov spectra of random products of matrices in with a distribution law
that is not confined to compact or reducible subgroups. More concretely, we will show that the column vectors
and
have a definite exponential growth
with (the top Lyapunov exponent) and, similarly, the column vector
of the matrix
has a definite exponential growth
where (the sum of the two largest Lyapunov exponents) satisfies
(i.e., the top Lyapunov exponent is simple, i.e., it does not coincide with the second largest Lyapunov exponent). Of course, if we show these properties, then
converges exponentially to
as
(since
).
Let us briefly sketch the proof of these exponential growth properties. We write as a “Birkhoff sum”
where for
and
. As it turns out, the sequence
is not stationary, but it is almost stationary, so that Birkhoff’s ergodic theorem says that the time averages converge to the spatial average (with probability
):
where is the rotation-invariant distribution of
‘s. Logically, this application of the ergodic theorem is valid only if we check that the observable
is integrable. However, this is not hard to verify: by definition,
, so that the desired integrability is a consequence of the integrability requirement in item (ii). A similar argument shows that
where . So, it remains only to check the simplicity condition
. For this sake, we combine the two integrals defining
and
, and we transfer it from
to
. In this way, we obtain:
On the other hand, by definition, runs over orthogonal pairs of directions in
. Thus, since
is not confined to an orthogonal subgroup of
, we have that the integral in the right-hand side of the equation above is strictly positive, i.e.,
.
In summary, we showed that, for each fixed , the sequence
converges to a Dirac mass with probability
. Since
are independent of
, this actually proves that
converges to Dirac masses independently of
. Finally, since the sequence
is stationary, we conclude that the
were Dirac masses to begin with.
This completes the proof of Lemma 16.
Closing this post, we will use the fact that is the Poisson boundary of
(where
is the probability measure constructed in Proposition 14) to show Furstenberg’s theorem 1 that the lattices of
are “distinct” from the lattices of
.
3. End of the proof of Furstenberg’s theorem 1
In this section, we will show the following statement (providing a slightly stronger version of Theorem 1) in the case :
Theorem 17 A cocompact lattice of
,
, can not be isomorphic to a subgroup of
.
The proof of this theorem proceeds by contradiction. Suppose that is isomorphic to a cocompact lattice of
and also to a subgroup of
.
By Theorem 14, we can equip with a fully supported probability measure
such that
has Poisson boundary
.
Let us think now of as a subgroup of
. We claim that
can not be confined to a compact subgroup of
: indeed, if this were the case,
would be Abelian; however, we saw that Abelian groups have trivial Poisson boundary, while
.
Next, let us observe that admits some
-stationary probability measure
, i.e.,
. Indeed, this is a consequence of the Krylov-Bogolyubov argument: we fix
an arbitrary probability measure on the compact space
, and we extract a
-stationary probability
as a limit of some convergent subsequence of the sequence
We affirm that is a boundary of
. In fact, given a stationary sequence
of independent random variables with distribution
, we know that the elements
are unbounded as
is a non-compact subgroup of
(as we just saw) and
is fully supported on
. By Lemma 12, it follows that
converges to a Dirac mass, and so, by Proposition 7 of the previous post, we deduce that
is a boundary of
.
By definition of Poisson boundary, the facts that has Poisson boundary
and
is a boundary of
imply that
is an equivariant image of
under some equivariant map
. We will prove that this is not possible (thus completing today’s post). The basic idea is that the sole way of going to infinity in
is to approach
, i.e.,
converges to a Dirac mass as
in
, but, on the other hand, we can go to infinity in
without approaching
, i.e., we can let
in
in such a way that
converges to a circle measure. Then, in the end of the day, these distinct (and incompatible) boundary behaviors (Dirac mass versus circle measure) will lead to the desired contradiction.
In order to get Dirac mass behavior in the context of , our plan is to apply Lemma 12. But, before doing so, we need to know that
is not atomic, a property that we claim to be true. Indeed, suppose to the contrary that
for some point
, and denote by
, a set of
-measure
. Consider the translates
of
under
. On one hand, if
intersects
in a subset of positive measure, then their images
and
under
in
intersect, and, thus, by equivariance of
and the fact that
is a singleton,
, and, a fortiori,
. In particular, the property that
whenever
intersects
with positive measure implies that
does not act ergodically on
(because the
‘s,
, do not get mixed together). However, this is a contradiction with Moore’s ergodicity theorem (implying that a lattice of
acts ergodically on
). This shows that
is non-atomic.
In particular, given and
two disjoint closed subsets of
, if
is any sequence going to
, then
Indeed, this is so because Lemma 11 (and the proof of Lemma 12) says that, for each , the measure
concentrates at least
of its mass in an interval of length
for all
sufficiently large. In particular, for any
smaller than the distance separating the disjoint closed sets
and
(i.e.,
), we obtain that either
or
for
sufficiently large, as desired.
We can rephrase the “concentration property” of the last paragraph in terms of -harmonic functions as follows. Let
,
, be measurable functions supported on
,
, and consider the associated
-harmonic functions
Then, the “concentration property” is
Now, let us “transfer” this picture to the context, i.e., let us think of the
-harmonic
,
, as defined on
. By the Poisson formula (and the fact that
), we can represent the
-harmonic functions
,
, as
where ,
, are bounded measurable functions on
. By replacing the variable
by
in this formula, we see that
can be extended to harmonic functions on
(w.r.t. any spherical measure on
).
In what follows, we will try to understand the boundary behavior of ‘s, ultimately to contradict the concentration property (6). In this direction, let us first “transfer” the concentration property (6) from
to
as follows. Observe that
is a cocompact lattice of
, so that we can select a bounded fundamental domain
, i.e., a bounded set such that any element of
has the form
with
and
. Now, given
, let us compare the values
of
and one of its “neighbors”
,
. Here, we observe that
In particular, since the right-hand side is bounded for in the bounded set
, we conclude that the ratio between
is uniformly bounded away from and
for
and
. Therefore, since the values of a positive
-harmonic function
at
and
can be written as
and
we deduce that the values and
are uniformly comparable for
and
, that is, there exists a constant
such that
for all and
. Hence, given
, there exists
such that
because for some
,
. Furthermore, when letting
, we have that
(as
and
is bounded). Thus, by combining this with the “concentration property” (6), we get the following concentration property in the context of
:
Let us now try to contradict the “transferred concentration property” (8) by analyzing the values when
without approaching
(something that is not possible in
!). More concretely, let us consider the sequences
We want to investigate the “boundary” values of the harmonic functions and
along the sequences
and
(and some adequate translates). For this, we need an analog of Fatou’s theorem saying that harmonic functions have boundary values along almost all radial direction. In our context, the analog of Fatou’s theorem goes as follows.
The limits of the sequences and
are circle measures
and
supported on the great circles
and
Also, it is possible to check that all circle measures have the form or
for
in the orthogonal subgroup
of
. Now, we affirm that, given
a bounded measurable function on
, the integrals
are defined for almost every . Indeed, we recall that
acts transitively on
, so that
where
is a finite group, i.e.,
is a finite cover of
. The great circles
and
correspond to two
-parameter subgroups
and
of
(as any great circle of
passing through the identity coset). Also, the great circles of
are just the cosets
and
,
, modulo
. In particular, given a bounded measurable function
on
, we can lift it to a bounded measurable function
on
and then define
For later use, we observe that
Anyhow, in this setting, Fatou’s theorem implies that
(at least) in measure as .
Therefore, if is the harmonic function associated to
, then
Now let us come back to the harmonic functions and
constructed above satisfying the concentration property (8). Denoting by
and
the “boundary values” of
(along
and
), we see that (10) and the concentration property (8) imply that
We will show that this is not possible by choosing the disjoint closed sets and
of
leading to
and
in a careful way and by using the fact that the
-stationary measure
on
is non-atomic.
More concretely, since is not atomic, we can choose a compact subset
of
with
that we fix once and for all. Set
and construct from
a harmonic function
as above (cf. (5)) and let
the associated function in (7). Denote by
and
the boundary values of
along sequences
and
.
Now, we consider an increasing sequence of compact subsets exhausting
. Again, set
, construct the corresponding harmonic functions
(cf. (5)), and let
and
be the boundary values of
along sequences
and
. By construction, since
is an increasing sequence, the functions
and
form two sequences of non-negative functions uniformly bounded by
that do not decrease as
increases. It follows that we can extract limits
and
. Moreover, from the definition of
and
, we get
Also, since exhausts
, i.e.,
, we obtain from (9) that
Furthermore, the concentration property (11) implies that
for almost every .
Finally, since the harmonic functions have nice integral representations as in (7), i.e.,
we can extract a limit such that
where and
are circle measures in our preferred circles
and
.
Now, we can get a contradiction as follows. By (12), (13), and the concentration property (14), we have that
for almost every . By (15), this means that the functions
and
are constant (
or
) on each great circle (up to some neglectable expectional set of zero measure).
In particular, after lifting the function on
to a function
on
, we have that
for
(where
and
are the lifts of the great circles
and
supporting
and
), that is,
Note that is a subgroup of
. It follows that
: indeed, it is a general fact that the group generated by two distinct
-parameter subgroups of
(such as
and
) is the whole
. This shows that
, and, a fortiori,
is the constant function
or
.
However, can not be
nor
: in fact, by (7) and (5), we know that
a contradiction with the fact that or
by our choice of
with
.
This completes our sketch of the proof of Theorem 17 (and Theorem 1).
Dear Matheus,
Just wanted to let you know that this is fantastic blog post and I will be recommending it to my students to read!
Jayadev
By: Jayadev Athreya on August 1, 2013
at 11:22 am
Great post! It is indeed frustrating that Theorem 13 is (was!) hard to find online. -Alex
By: Alex on August 2, 2013
at 3:43 pm