1 Introduction
When G is a countable infinite group and
$(X_0,\mu _0)$
is a non-trivial standard probability space, the probability measure-preserving (pmp) action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu1.png?pub-status=live)
is called a Bernoulli action. Probability measure-preserving Bernoulli actions are among the best-studied objects in ergodic theory and they play an important role in operator algebras [Reference IoanaIoa10, Reference PopaPop03, Reference PopaPop06]. When we consider a family of probability measures
$(\mu _g)_{g\in G}$
on the base space
$X_0$
that need not all be equal, the Bernoulli action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn1.png?pub-status=live)
is in general no longer measure-preserving. Instead, we are interested in the case where
$G\curvearrowright (X,\mu )$
is non-singular, that is, the group G preserves the measure class of
$\mu $
. By Kakutani’s criterion for equivalence of infinite product measures the Bernoulli action (1.1) is non-singular if and only if
$\mu _h\sim \mu _g$
for every
$h,g\in G$
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn2.png?pub-status=live)
Here
$H^2(\mu _h,\mu _{gh})$
denotes the Hellinger distance between
$\mu _h$
and
$\mu _{gh}$
(see (2.2)).
It is well known that a pmp Bernoulli action
$G\curvearrowright (X_0,\mu _0)^{G}$
is mixing. In particular, it is ergodic and conservative. However, for non-singular Bernoulli actions, determining conservativeness and ergodicity is much more difficult (see, for instance, [Reference Berendschot and VaesBKV19, Reference DanilenkoDan18, Reference KosloffKos18, Reference Vaes and WahlVW17]).
Besides non-singular Bernoulli actions, another interesting class of non-singular group actions comes from the Gaussian construction, as introduced in [Reference Arano, Isono and MarrakchiAIM19]. If
${\pi \colon G\rightarrow \mathcal {O}(\mathcal {H})}$
is an orthogonal representation of a locally compact second countable (lcsc) group on a real Hilbert space
$\mathcal {H}$
, and if
$c\colon G\rightarrow \mathcal {H}$
is a 1-cocycle for the representation
$\pi $
, then the assignment
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn3.png?pub-status=live)
defines an affine isometric action
$\alpha \colon G\curvearrowright \mathcal {H}$
. To any affine isometric action
$\alpha \colon G\curvearrowright \mathcal {H}$
Arano, Isono and Marrakchi associated a non-singular group action
$\widehat {\alpha }\colon G\curvearrowright \widehat {\mathcal {H}}$
, where
$\widehat {\mathcal {H}}$
is the Gaussian probability space associated to
$\mathcal {H}$
. When
$\alpha \colon G\curvearrowright \mathcal {H}$
is actually an orthogonal representation, this construction is well established and the resulting Gaussian action is pmp. As explained below [Reference Björklund, Kosloff and VaesBV20, Theorem D], if G is a countable infinite group and
$\pi \colon G\rightarrow \ell ^2(G)$
is the left regular representation, the affine isometric representation (1.3) gives rise to a non-singular action that is conjugate with the Bernoulli action
$G\curvearrowright \prod _{g\in G}(\mathbb {R},\nu _{F(g)})$
, where
$F\colon G\rightarrow \mathbb {R}$
is such that
$c_g(h)=F(g^{-1}h)-F(h)$
, and
$\nu _{F(g)}$
denotes the Gaussian probability measure with mean
$F(g)$
and variance
$1$
.
By scaling the 1-cocycle
$c\colon G\rightarrow \mathcal {H}$
with a parameter
$t\in [0,+\infty )$
we get a one-parameter family of non-singular actions
$\widehat {\alpha }^{t}\colon G\curvearrowright \widehat {\mathcal {H}}^{t}$
associated to the affine isometric actions
$\alpha ^{t}\colon G\curvearrowright \mathcal {H}$
, given by
$\alpha ^t_g(\xi )=\pi _g(\xi )+tc(g)$
. Arano, Isono and Marrakchi showed that there exists a
$t_{\mathrm {diss}}\in [0,+\infty )$
such that
$\widehat {\alpha }^t$
is dissipative up to compact stabilizers for every
$t>t_{\mathrm {diss}}$
and infinitely recurrent for every
$t<t_{\mathrm {diss}}$
(see §2 for terminology).
Inspired by the results obtained in [Reference Arano, Isono and MarrakchiAIM19], we study a similar phase transition framework, but in the setting of non-singular Bernoulli actions. Such a phase transition framework for non-singular Bernoulli actions was already considered by Kosloff and Soo in [Reference Kosloff and SooKS20]. They showed the following phase transition result for the family of non-singular Bernoulli actions of
$G=\mathbb {Z}$
with base space
$X_0=\{0,1\}$
that was introduced in [Reference Vaes and WahlVW17, Corollary 6.3]. For every
$t\in [0,+\infty )$
consider the family of measures
$(\mu _n^t)_{n\in \mathbb {Z}}$
given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu2.png?pub-status=live)
Then
$\mathbb {Z}\curvearrowright (X,\mu _t)=\prod _{n\in \mathbb {Z}}(\{0,1\},\mu _n^t)$
is non-singular for every
$t\in [0,+\infty )$
. Kosloff and Soo showed that there exists a
$t_1\in (1/6,+\infty )$
such that
$ \mathbb {Z}\curvearrowright (X,\mu _t)$
is conservative for every
$t<t_1$
and dissipative for every
$t>t_1$
[Reference Kosloff and SooKS20, Theorem 3]. In [Reference Danilenko, Kosloff and RoyDKR20, Example D] the authors describe a family of non-singular Poisson suspensions for which a similar phase transition occurs. These examples arise from dissipative essentially free actions of
$\mathbb {Z}$
, and thus they are non-singular Bernoulli actions. We generalize the phase transition result from [Reference Kosloff and SooKS20] to arbitrary non-singular Bernoulli actions as follows.
Suppose that G is a countable infinite group and let
$(\mu _g)_{g\in G}$
be a family of equivalent probability measure on a standard Borel space
$X_0$
. Let
$\nu $
also be a probability measure on
$X_0$
. For every
$t\in [0,1]$
we consider the family of equivalent probability measures
$(\mu _g^t)_{g\in G}$
that are defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn4.png?pub-status=live)
Our first main result is that in this setting there is a phase transition phenomenon.
Theorem A. Let G be a countable infinite group and assume that the Bernoulli action
$G\curvearrowright (X,\mu _1)=\prod _{g\in G}(X_0,\mu _g)$
is non-singular. Let
$\nu \sim \mu _e$
be a probability measure on
$X_0$
and for every
$t\in [0,1]$
consider the family
$(\mu _g^t)_{g\in G}$
of equivalent probability measures given by (1.4). Then the Bernoulli action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu3.png?pub-status=live)
is non-singular for every
$t\in [0,1]$
and there exists a
$t_1\in [0,1]$
such that
$G\curvearrowright (X,\mu _t)$
is weakly mixing for every
$t<t_1$
and dissipative for every
$t>t_1$
.
Suppose that G is a non-amenable countable infinite group. Recall that for any standard probability space
$(X_0,\mu _0)$
, the pmp Bernoulli action
$G\curvearrowright (X_0,\mu _0)^{G}$
is strongly ergodic. Consider again the family of probability measures
$(\mu _g^t)_{g\in G}$
given by (1.4). In Theorem B below we prove that for t close enough to
$0$
, the resulting non-singular Bernoulli action is strongly ergodic. This is inspired by [Reference Arano, Isono and MarrakchiAIM19, Theorem 7.20] and [Reference Marrakchi and VaesMV20, Theorem 5.1], which state similar results for non-singular Gaussian actions.
Theorem B. Let G be a countable infinite non-amenable group and suppose that the Bernoulli action
$G\curvearrowright (X,\mu _1)=\prod _{g\in G}(X_0,\mu _g)$
is non-singular. Let
$\nu \sim \mu _e$
be a probability measure on
$X_0$
and for every
$t\in [0,1]$
consider the family
$(\mu _g^t)_{g\in G}$
of equivalent probability measures given by (1.4). Then there exists a
$t_0\in (0,1]$
such that
$G\curvearrowright (X,\mu _t)=\prod _{g\in G}(X_0,\mu _g^t)$
is strongly ergodic for every
$t<t_0$
.
Although we can prove a phase transition result in large generality, it remains very challenging to compute the critical value
$t_1$
. However, when
$G\subset \operatorname {Aut}(T)$
, for some locally finite tree T, following [Reference Arano, Isono and MarrakchiAIM19, §10], we can construct generalized Bernoulli actions of which we can determine the conservativeness behaviour very precisely. To put this result into perspective, let us first explain briefly the construction from [Reference Arano, Isono and MarrakchiAIM19, §10].
For a locally finite tree T, let
$\Omega (T)$
denote the set of orientations on T. Let
$p\in (0,1)$
and fix a root
$\rho \in T$
. Define a probability measure
$\mu _p$
on
$\Omega (T)$
by orienting an edge towards
$\rho $
with probability p and away from
$\rho $
with probability
$1-p$
. If
$G\subset \operatorname {Aut}(T)$
is a subgroup, then we naturally obtain a non-singular action
$G\curvearrowright (\Omega (T),\mu _p)$
. Up to equivalence of measures, the measure
$\mu _p$
does not depend on the choice of root
$\rho \in T$
. The Poincaré exponent of
$G\subset \operatorname {Aut}(T)$
is defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn5.png?pub-status=live)
where
$v\in V(T)$
is any vertex of T. In [Reference Arano, Isono and MarrakchiAIM19, Theorem 10.4] Arano, Isono and Marrakchi showed that if
$G\subset \operatorname {Aut}(T)$
is a closed non-elementary subgroup, the action
$G\curvearrowright (\Omega (T),\mu _p)$
is dissipative up to compact stabilizers if
$2\sqrt {p(1-p)}<\exp (-\delta )$
and weakly mixing if
$2\sqrt {p(1-p)}>\exp (-\delta )$
. This motivates the following similar construction.
Let
$E(T)\subset V(T)\times V(T)$
denote the set of oriented edges, so that vertices v and w are adjacent if and only if
$(v,w),(w,v)\in E(T)$
. Suppose that
$X_0$
is a standard Borel space and that
$\mu _0,\mu _1$
are equivalent probability measures on
$X_0$
. Fix a root
$\rho \in T$
and define a family of probability measures
$(\mu _e)_{e\in E(T)}$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn6.png?pub-status=live)
Suppose that
$G\subset \operatorname {Aut}(T)$
is a subgroup. Then the generalized Bernoulli action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn7.png?pub-status=live)
is non-singular and up to conjugacy it does not depend on the choice of root
$\rho \in T$
. In our next main result we generalize [Reference Arano, Isono and MarrakchiAIM19, Theorem 10.4] to non-singular actions of the form (1.7).
Theorem C. Let T be a locally finite tree with root
$\rho \in T$
and let
$G\subset \operatorname {Aut}(T)$
be a non-elementary closed subgroup with Poincaré exponent
$\delta =\delta (G\curvearrowright T)$
. Let
$\mu _0$
and
$\mu _1$
be equivalent probability measures on a standard Borel space
$X_0$
and define a family of equivalent probability measures
$(\mu _e)_{e\in E(T)}$
by (1.6). Then the generalized Bernoulli action (1.7) is dissipative up to compact stabilizers if
$1-H^2(\mu _0,\mu _1)<\exp (-\delta /2)$
and weakly mixing if
$1-H^2(\mu _0,\mu _1)>\exp (-\delta /2)$
.
2 Preliminaries
2.1 Non-singular group actions
Let
$(X,\mu ), (Y,\nu )$
be standard measure spaces. A Borel map
$\varphi \colon X\rightarrow Y$
is called non-singular if the pushforward measure
$\varphi _*\mu $
is equivalent to
$\nu $
. If in addition there exist conull Borel sets
$X_0\subset X$
and
$Y_0\subset Y$
such that
$\varphi \colon X_0\rightarrow Y_0$
is a bijection we say that
$\varphi $
is a non-singular isomorphism. We write
$\operatorname {Aut}(X,\mu )$
for the group of all non-singular automorphisms
$\varphi \colon X\rightarrow X$
, where we identify two elements if they agree almost everywhere. The group
$\operatorname {Aut}(X,\mu )$
carries a canonical Polish topology.
A non-singular group action
$G\curvearrowright (X,\mu )$
of an lcsc group G on a standard measure space
$(X,\mu )$
is a continuous group homomorphism
$G\rightarrow \operatorname {Aut}(X,\mu )$
. A non-singular group action
$G\curvearrowright (X,\mu )$
is called essentially free if the stabilizer subgroup
$G_x=\{g\in G:g\cdot x=x\}$
is trivial for almost every (a.e.)
$x\in X$
. When G is countable this is the same as the condition that
$\mu (\{x\in X:g\cdot x=x\})=0$
for every
$g\in G\setminus \{e\}$
. We say that
$G\curvearrowright (X,\mu )$
is ergodic if every G-invariant Borel set
$A\subset X$
satisfies
$\mu (A)=0$
or
$\mu (X\setminus A)=0$
. A non-singular action
$G\curvearrowright (X,\mu )$
is called weakly mixing if for any ergodic pmp action
$G\curvearrowright (Y,\nu )$
the diagonal product action
$G\curvearrowright X\times Y$
is ergodic. If G is not compact and
$G\curvearrowright (X,\mu )$
is pmp, we say that
$G\curvearrowright X$
is mixing if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu4.png?pub-status=live)
Suppose that
$G\curvearrowright (X,\mu )$
is a non-singular action and that
$\mu $
is a probability measure. A sequence of Borel subsets
$A_n\subset X$
is called almost invariant if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu5.png?pub-status=live)
The action
$G\curvearrowright (X,\mu )$
is called strongly ergodic if every almost invariant sequence
$A_n\subset X$
is trivial, that is,
$\mu (A_n)(1-\mu (A_n))\rightarrow 0$
. The strong ergodicity of
$G\curvearrowright (X,\mu )$
only depends on the measure class of
$\mu $
. When
$(Y,\nu )$
is a standard measure space and
$\nu $
is infinite, a non-singular action
$G\curvearrowright (Y,\nu )$
is called strongly ergodic if
$G\curvearrowright (Y,\nu ')$
is strongly ergodic, where
$\nu '$
is a probability measure that is equivalent to
$\nu $
.
Following [Reference Arano, Isono and MarrakchiAIM19, Definition A.16], we say that a non-singular action
$G\curvearrowright (X,\mu )$
is dissipative up to compact stabilizers if each ergodic component is of the form
${G\curvearrowright G/ K}$
, for a compact subgroup
$K\subset G$
. By [Reference Arano, Isono and MarrakchiAIM19, Theorem A.29] a non-singular action
${G\curvearrowright (X,\mu )}$
, with
$\mu (X)=1$
, is dissipative up to compact stabilizers if and only if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu6.png?pub-status=live)
where
$\unicode{x3bb} $
denotes the left invariant Haar measure on G. We say that
$G\curvearrowright (X,\mu )$
is infinitely recurrent if for every non-negligible subset
$A\subset X$
and every compact subset
$K\subset G$
there exists
$g\in G\setminus K$
such that
$\mu (g\cdot A\cap A)>0$
. By [Reference Arano, Isono and MarrakchiAIM19, Proposition A.28] and Lemma 2.1 below, a non-singular action
$G\curvearrowright (X,\mu )$
, with
$\mu (X)=1$
, is infinitely recurrent if and only if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu7.png?pub-status=live)
A non-singular action
$G\curvearrowright (X,\mu )$
is called dissipative if it is essentially free and dissipative up to compact stabilizers. In that case there exists a standard measure space
$(X_0,\mu _0)$
such that
$G\curvearrowright X$
is conjugate with the action
$G\curvearrowright G\times X_0: \;g\cdot (h,x)=(gh,x)$
. A non-singular action
$G\curvearrowright (X,\mu )$
decomposes, uniquely up to a null set, as
${G\curvearrowright D\sqcup C}$
, where
$G\curvearrowright D$
is dissipative up to compact stabilizers and
$G\curvearrowright C$
is infinitely recurrent. When G is a countable group and
$G\curvearrowright (X,\mu )$
is essentially free, we say that
$G\curvearrowright X$
is conservative if it is infinitely recurrent.
Lemma 2.1. Suppose that G is an lcsc group with left invariant Haar measure
$\unicode{x3bb} $
and that
$(X,\mu )$
is a standard probability space. Assume that
$G\curvearrowright (X,\mu )$
is a non-singular action that is infinitely recurrent. Then we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu8.png?pub-status=live)
Proof. Note that the set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu9.png?pub-status=live)
is G-invariant. Therefore, it suffices to show that
$G\curvearrowright X$
is not infinitely recurrent under the assumption that D has full measure.
Let
$\pi \colon (X,\mu )\rightarrow (Y,\nu )$
be the projection onto the space of ergodic components of
$G\curvearrowright X$
. Then there exist a conull Borel subset
$Y_0\subset Y$
and a Borel map
$\theta \colon Y_0\rightarrow X$
such that
$(\pi \circ \theta )(y)=y$
for every
$y\in Y_0$
.
Write
$X_y=\pi ^{-1}(\{y\})$
. By [Reference Arano, Isono and MarrakchiAIM19, Theorem A.29], for a.e.
$y\in Y$
there exists a compact subgroup
$K_y\subset G$
such that
$G\curvearrowright X_y$
is conjugate with
$G\curvearrowright G/ K_y$
. Let
$G_n\subset G$
be an increasing sequence of compact subsets of G such that
$\bigcup _{n\geq 1}\overset {\circ }{G}_n=G$
. For every
$x\in X$
, write
$G_x=\{g\in G:g\cdot x=x\}$
for the stabilizer subgroup of x. Using an argument as in [Reference Meesschaert, Raum and VaesMRV11, Lemma 10], one shows that for each
$n\geq 1$
the set
$\{x\in X:G_x\subset G_n\}$
is Borel. Thus, for every
$n\geq 1$
the set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu10.png?pub-status=live)
is a Borel subset of Y and we have that
$\nu (\bigcup _{n\geq 1 } U_n)=1$
. Therefore, the sets
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu11.png?pub-status=live)
are analytic and exhaust X up to a set of measure zero. So there exist an
$n_0\in \mathbb {N}$
and a non-negligible Borel set
$B\subset A_{n_0}$
. Suppose that
$h\in G$
is such that
$h\cdot B\cap B\neq \emptyset $
. Then there exist
$y\in U_{n_0}$
and
$g_1,g_2\in G_{n_0}$
such that
$hg_1\cdot \theta (y)=g_2\cdot \theta (y)$
, and we get that
$h\in G_{n_0}K_yG_{n_0}^{-1}\subset G_{n_0}G_{n_0}G_{n_0}^{-1}$
. In other words, for
$h\in G$
outside the compact set
$G_{n_0}G_{n_0}G_{n_0}^{-1}$
we have that
$\mu (h\cdot B \cap B)=0$
, so that
$G\curvearrowright X$
is not infinitely recurrent.
We will frequently use the following result of Schmidt and Walters. Suppose that
${G\curvearrowright (X,\mu )}$
is a non-singular action that is infinitely recurrent and suppose that
${G\curvearrowright (Y,\nu )}$
is pmp and mixing. Then by [Reference Schmidt and WaltersSW81, Theorem 2.3] we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu12.png?pub-status=live)
where
$G\curvearrowright X\times Y$
acts diagonally. Although [Reference Schmidt and WaltersSW81, Theorem 2.3] demands proper ergodicity of the action
$G\curvearrowright (X,\mu )$
, the infinite recurrence assumption is sufficient as remarked in [Reference Arano, Isono and MarrakchiAIM19, Remark 7.4].
2.2 The Maharam extension and crossed products
Let
$(X,\mu )$
be a standard measure space. For any non-singular automorphism
$\varphi \in \operatorname {Aut}(X,\mu )$
, we define its Maharam extension by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu13.png?pub-status=live)
Then
$\widetilde {\varphi }$
preserves the infinite measure
$\mu \times \exp (-t)dt$
. The assignment
$\varphi \mapsto \widetilde {\varphi }$
is a continuous group homomorphism from
$\operatorname {Aut}(X)$
to
$\operatorname {Aut}(X\times \mathbb {R})$
. Thus, for each non-singular group action
$G\curvearrowright (X,\mu )$
, by composing with this map, we obtain a non-singular group action
$G\curvearrowright X\times \mathbb {R}$
, which we call the Maharam extension of
$G\curvearrowright X$
. If
$G\curvearrowright X$
is a non-singular group action, the translation action
$\mathbb {R}\curvearrowright X\times \mathbb {R}$
in the second component commutes with the Maharam extension
$G\curvearrowright X\times \mathbb {R}$
. Therefore, we get a well-defined action
$\mathbb {R}\curvearrowright L^{\infty }(X\times \mathbb {R})^{G}$
, which is the Krieger flow associated to the action
$G\curvearrowright X$
. The Krieger flow is given by
$\mathbb {R}\curvearrowright \mathbb {R}$
if and only if there exists a G-invariant
$\sigma $
-finite measure
$\nu $
on X that is equivalent to
$\mu $
.
Suppose that
$M\subset B(\mathcal {H})$
is a von Neumann algebra represented on the Hilbert space
$\mathcal {H}$
and that
$\alpha \colon G\curvearrowright M$
is a continuous action on M of an lcsc group G. Then the crossed product von Neumann algebra
$M\rtimes _{\alpha } G\subset B(L^2(G,\mathcal {H}))$
is the von Neumann algebra generated by the operators
$\{\pi (x)\}_{x\in M}$
and
$\{u_h\}_{h\in G}$
acting on
$\xi \in L^2(G,\mathcal {H})$
as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu14.png?pub-status=live)
In particular, if
$G\curvearrowright (X,\mu )$
is a non-singular group action, the crossed product
$L^{\infty }(X)\rtimes G\subset B(L^2(G\times X))$
is the von Neumann algebra generated by the operators
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu15.png?pub-status=live)
for
$H\in L^{\infty }(X)$
and
$h\in G$
. If
$G\curvearrowright X$
is non-singular essentially free and ergodic, then
$L^{\infty }(X)\rtimes G$
is a factor. Moreover, when G is a unimodular group, the Krieger flow of
${G\curvearrowright X}$
equals the flow of weights of the crossed product von Neumann algebra
$L^{\infty }(X)\rtimes G$
. For non-unimodular groups this is not necessarily true, motivating the following definition.
Definition 2.2. Let G be an lcsc group with modular function
$\Delta \colon G\rightarrow \mathbb {R}_{>0}$
. Let
$\unicode{x3bb} $
denote the Lebesgue measure on
$\mathbb {R}$
. Suppose that
$\alpha \colon G\curvearrowright (X,\mu )$
is a non-singular action. We define the modular Maharam extension of
$G\curvearrowright X$
as the non-singular action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu16.png?pub-status=live)
Let
$L^{\infty }(X\times \mathbb {R})^{\beta }$
denote the subalgebra of
$\beta $
-invariant elements. We define the flow of weights associated to
$G{\kern-1pt}\curvearrowright{\kern-1pt} X$
as the translation action
$\mathbb {R}{\kern-1pt}\curvearrowright{\kern-1pt} L^{\infty }(X{\kern-1pt}\times{\kern-1pt} \mathbb {R})^{\beta }: (t\cdot H)(x,s)= H(x,s-t)$
.
As we explain below, the flow of weights associated to an essentially free ergodic non-singular action
$G\curvearrowright X$
equals the flow of weights of the crossed product factor
$L^{\infty }(X)\rtimes G$
, justifying the terminology. See also [Reference SauvageotSa74, Proposition 4.1].
Let
$\alpha \colon G\curvearrowright X$
be an essentially free ergodic non-singular group action with modular Maharam extension
$\beta \colon G\curvearrowright X\times \mathbb {R}$
. By [Reference SauvageotSa74, Proposition 1.1] there is a canonical normal semifinite faithful weight
$\varphi $
on
$L^{\infty }(X)\rtimes _{\alpha } G$
such that the modular automorphism group
$\sigma ^{\varphi }$
is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu17.png?pub-status=live)
where
$\Delta \colon G\rightarrow \mathbb {R}_{>0}$
denotes the modular function of G.
For an element
$\xi \in L^2(\mathbb {R}, L^2(G\times X))$
and
$(g,x)\in G\times X$
, write
$\xi _{g,x}$
for the map given by
$\xi _{g,x}(s)=\xi (s,g,x)$
. Then by Fubini’s theorem
$\xi _{g,x}\in L^2(\mathbb {R})$
for a.e.
${(g,x)\in G\times X}$
. Let
$U\colon L^2(\mathbb {R}, L^2(G\times X))\rightarrow L^2(G,L^2(X\times \mathbb {R}))$
be the unitary given on
${\xi \in L^2(\mathbb {R}, L^2(G\times X))}$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu18.png?pub-status=live)
where
$\mathcal {F}^{-1}\colon L^2(\mathbb {R})\rightarrow L^2(\mathbb {R})$
denotes the inverse Fourier transform. One can check that conjugation by U induces an isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu19.png?pub-status=live)
Let
$\kappa \colon L^{\infty }(X\times \mathbb {R})\rightarrow L^{\infty }(X\times \mathbb {R})\rtimes _{\beta }G$
be the inclusion map and let
$\gamma \colon \mathbb {R}\curvearrowright L^{\infty }(X\times \mathbb {R})\rtimes _{\beta }G$
be the action given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu20.png?pub-status=live)
Then one can verify that
$\Psi $
conjugates the dual action
$\widehat {\sigma ^{\varphi }}\colon \mathbb {R}\curvearrowright (L^{\infty }(X)\rtimes _{\alpha } G)\rtimes _{\sigma ^{\varphi }}\mathbb {R}$
and
$\gamma $
. Therefore, we can identify the flow of weights
$\mathbb {R}\curvearrowright \mathcal {Z}((L^{\infty }(X)\rtimes _{\alpha } G)\rtimes _{\sigma ^{\varphi }}\mathbb {R})$
with
$\mathbb {R}\curvearrowright \mathcal {Z}(L^{\infty }(X\times \mathbb {R})\rtimes _{\beta } G)\cong L^{\infty }(X\times \mathbb {R})^{\beta }$
: the flow of weights associated to
${G\curvearrowright X}$
.
Remark 2.3. It will be useful to speak about the Krieger type of a non-singular ergodic action
$G\curvearrowright X$
. In light of the discussion above, we will only use this terminology for countable groups G, so that no confusion arises with the type of the crossed product von Neumann algebra
$L^{\infty }(X)\rtimes G$
. So assume that G is countable and that
$G\curvearrowright (X,\mu )$
is a non-singular ergodic action. Then the Krieger flow is ergodic and we distinguish several cases. If
$\nu $
is atomic, we say that
$G\curvearrowright X$
is of type I. If
$\nu $
is non-atomic and finite, we say that
$G\curvearrowright X$
is of type II
$_{1}$
. If
$\nu $
is non-atomic and infinite, we say that
$G\curvearrowright X$
is of type II
$_{\infty }$
. If the Krieger flow is given by
$\mathbb {R}\curvearrowright \mathbb {R}/\log (\unicode{x3bb} )\mathbb {Z}$
with
$\unicode{x3bb} \in (0,1)$
, we say that
$G\curvearrowright X$
is of type III
$_{\unicode{x3bb} }$
. If the Krieger flow is the trivial flow
$\mathbb {R}\curvearrowright \{\ast \}$
, we say that
$G\curvearrowright X$
is of type III
$_{1}$
. If the Krieger flow is properly ergodic (that is, every orbit has measure zero), we say that
$G\curvearrowright X$
is of type III
$_{0}$
.
2.3 Non-singular Bernoulli actions
Suppose that G is a countable infinite group and that
$(\mu _g)_{g\in G}$
is a family of equivalent probability measures on a standard Borel space
$X_0$
. The action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn8.png?pub-status=live)
is called the Bernoulli action. For two probability measures
$\nu ,\eta $
on a standard Borel space Y, the Hellinger distance
$H^2(\nu ,\eta )$
is defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn9.png?pub-status=live)
where
$\zeta $
is any probability measure on Y such that
$\nu ,\eta \prec \zeta $
. By Kakutani’s criterion for equivalence of infinite product measures [Reference KakutaniKak48] the Bernoulli action (2.1) is non-singular if and only if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu21.png?pub-status=live)
If
$(X,\mu )$
is non-atomic and the Bernoulli action (2.1) is non-singular, then it is essentially free by [Reference Berendschot and VaesBKV19, Lemma 2.2].
Suppose that I is a countable infinite set and that
$(\mu _i)_{i\in I}$
is a family of equivalent probability measures on a standard Borel space
$X_0$
. If G is an lcsc group that acts on I, the action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn10.png?pub-status=live)
is called the generalized Bernoulli action and it is non-singular if and only if
$\sum _{i\in I}H^2(\mu _i,\mu _{g\cdot i})<+\infty $
for every
$g\in G$
. When
$\nu $
is a probability measure on
$X_0$
such that
$\mu _i=\nu $
for every
$i\in I$
, the generalized Bernoulli action (2.3) is pmp and it is mixing if and only if the stabilizer subgroup
$G_i=\{g\in G:g\cdot i=i\}$
is compact for every
$i\in I$
. In particular, if G is countable infinite, the pmp Bernoulli action
$G\curvearrowright (X_0,\mu _0)^{G}$
is mixing.
2.4 Groups acting on trees
Let
$T=(V(T),E(T))$
be a locally finite tree, so that the edge set
$E(T)$
is a symmetric subset of
$ V(T)\times V(T)$
with the property that vertices
$v,w\in V(T)$
are adjacent if and only if
$(v,w),(w,v)\in E(T)$
. When T is clear from the context, we will write E instead of
$E(T)$
. Also we will often write T instead of
$V(T)$
for the vertex set. For any two vertices
$v,w\in T$
let
$[v,w]$
denote the smallest subtree of T that contains v and w. The distance between vertices
$v,w\in T$
is defined as
${d(v,w)=|V([v,w])|-1}$
. Fixing a root
$\rho \in T$
, we define the boundary
$\partial T$
of T as the collection of all infinite line segments starting at
$\rho $
. We equip
$\partial T$
with a metric
$d_\rho $
as follows. If
$\omega ,\omega '\in \partial T$
, let
$v\in T$
be the unique vertex such that
$d(\rho ,v)=\sup _{v\in \omega \cap \omega '}d(\rho , v)$
and define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu22.png?pub-status=live)
Then, up to homeomorphism, the space
$(\partial T, d_{\rho })$
does not depend on the chosen root
$\rho \in T$
. Furthermore, the Hausdorff dimension
$\dim _H \partial T$
of
$(\partial T, d_\rho )$
is also independent of the choice of
$\rho \in T$
.
Let
$\operatorname {Aut}(T)$
denote the group of automorphisms of T. By [Reference TitsTit70, Proposition 3.2], if
$g\in \operatorname {Aut}(T)$
, then either:
-
• g fixes a vertex or interchanges a pair of vertices (in this case we say that g is elliptic);
-
• or there exists a bi-infinite line segment
$L\subset T$ , called the axis of g, such that g acts on L by non-trivial translation (in this case we say that g is hyperbolic).
We equip
$\operatorname {Aut}(T)$
with the topology of pointwise convergence. A subgroup
$G\subset \operatorname {Aut}(T)$
is closed with respect to this topology if and only if for every
$v\in T$
the stabilizer subgroup
$G_v=\{g\in G:g\cdot v= v\}$
is compact. An action of an lcsc group G on T is a continuous homomorphism
$G\rightarrow \operatorname {Aut}(T)$
. We say that the action
$G\curvearrowright T$
is cocompact if there is a finite set
$F\subset E(T)$
such that
$G\cdot F=E(T)$
. A subgroup
$G\subset \operatorname {Aut}(T)$
is called non-elementary if it does not fix any point in
$T\cup \partial T$
and does not interchange any pair of points in
$T\cup \partial T$
. Equivalently,
$G\subset \operatorname {Aut}(T)$
is non-elementary if there exist hyperbolic elements
$h,g\in G$
with axes
$L_h$
and
$L_g$
such that
$L_h\cap L_g$
is finite. If
$G\subset \operatorname {Aut}(T)$
is a non-elementary closed subgroup, there exists a unique minimal G-invariant subtree
$S\subset T$
and G is compactly generated if and only if
$G\curvearrowright S$
is cocompact (see [Reference Caprace and de MedtsCM11, §2]). Recall from (1.5) the definition of the Poincaré exponent
$\delta (G\curvearrowright T)$
of a subgroup
$G\subset \operatorname {Aut}(T)$
. If
$G\subset \operatorname {Aut}(T)$
is a closed subgroup such that
$G\curvearrowright T$
is cocompact, then we have that
$\delta (G\curvearrowright T)=\dim _{H}\partial T$
.
3 Phase transitions of non-singular Bernoulli actions: proof of Theorems A and B
Let G be a countable infinite group and let
$(\mu _g)_{g\in G}$
be a family of equivalent probability measures on a standard Borel space
$X_0$
. Let
$\nu $
also be a probability measure on
$X_0$
. For
$t\in [0,1]$
we define the family of probability measures
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn11.png?pub-status=live)
We write
$\mu _t$
for the infinite product measure
$\mu _t=\prod _{g\in G}\mu _g^t$
on
$X=\prod _{g\in G}X_0$
. We prove Theorem 3.1 below, which is slightly more general than Theorem A.
Theorem 3.1. Let G be a countable infinite group and let
$(\mu _g)_{g\in G}$
be a family of equivalent probability measures on a standard probability space
$X_0$
, which is not supported on a single atom. Assume that the Bernoulli action
$G\curvearrowright \prod _{g\in G}(X_0,\mu _g)$
is non-singular. Let
$\nu $
also be a probability measure on
$X_0$
. Then for every
$t\in [0,1]$
the Bernoulli action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn12.png?pub-status=live)
is non-singular. Assume, in addition, that one of the following conditions holds.
-
(1)
$\nu \sim \mu _e$ .
-
(2)
$\nu \prec \mu _e $ and
$\sup _{g\in G}|{\log}\ d\mu _g/d\mu _e(x)|<+\infty $ for a.e
$x\in X_0$ .
Then there exists a
$t_1\in [0,1]$
such that
$G\curvearrowright (X,\mu _t)$
is dissipative for every
$t>t_1$
and weakly mixing for every
$t<t_1$
.
Remark 3.2. One might hope to prove a completely general phase transition result that only requires
$\nu \prec \mu _e$
, and not the additional assumption that
$\sup _{g\in G}|{\log}\ d\mu _g/d\mu _e(x)|<+\infty $
for a.e.
$x\in X_0$
. However, the following example shows that this is not possible.
Let G be any countable infinite group and let
$G\curvearrowright \prod _{g\in G}(C_0,\eta _g)$
be a conservative non-singular Bernoulli action. Note that Theorem 3.1 implies that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu23.png?pub-status=live)
is conservative for every
$t<1$
. Let
$C_1$
be a standard Borel space and let
$(\mu _g)_{g\in G}$
be a family of equivalent probability measures on
$X_0\kern1.2pt{=}\kern1.2pt C_0\kern1.2pt{\sqcup}\kern1.2pt C_1$
such that
${0\kern1.2pt{<}\kern1pt\sum _{g\in G}\mu _g(C_1)\kern1.2pt{<}\kern1.2pt{+}\kern0.5pt\infty }$
and such that
$\mu _g |_{C_0}\kern1.2pt{=}\kern1.2pt\mu _g(C_0)\eta _g$
. Then the Bernoulli action
$G\kern1.2pt{\curvearrowright}\kern1.2pt (X,\mu )\kern1.2pt{=}\kern1.2pt\prod _{g\in G}(X_0,\mu _g)$
is non-singular with non-negligible conservative part
$C_0^{G}\subset G$
and dissipative part
$X\setminus C_0^G$
. Taking
$\nu =\eta _e\prec \mu _e$
, for each
$t<1$
the Bernoulli action
$G\curvearrowright (X,\mu _t)= \prod _{g\in G}(X_0,(1-t)\eta _e+t\mu _g)$
is constructed in the same way, by starting with the conservative Bernoulli action
$G\curvearrowright \prod _{g\in G}(C_0,(1-t)\eta _e+t\eta _g)$
. So for every
$t\in (0,1)$
the Bernoulli action
$G\curvearrowright (X,\mu _t)$
has non-negligible conservative part and non-negligible dissipative part.
We can also prove a version of Theorem B in the more general setting of Theorem 3.1.
Theorem 3.3. Let G be a countable infinite non-amenable group. Make the same assumptions as in Theorem 3.1 and consider the non-singular Bernoulli actions
${G\curvearrowright (X,\mu _t)}$
given by (3.2). Assume, moreover, that:
-
(1)
$\nu \sim \mu _e$ , or
-
(2)
$\nu \prec \mu _e$ and
$\sup _{g\in G}|{\log}\ d\mu _g/d\mu _e(x)|<+\infty $ for a.e.
$x\in X_0$ .
Then there exists a
$t_0>0$
such that
$G\curvearrowright (X,\mu _t)$
is strongly ergodic for every
$t<t_0$
.
Proof of Theorem 3.1
Assume that
$G\curvearrowright (X,\mu _1)=\prod _{g\in G}(X_0,\mu _g)$
is non-singular. For every
$t\in [0,1]$
we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu24.png?pub-status=live)
so that
$G\curvearrowright (X,\mu _t)$
is non-singular for every
$t\in [0,1]$
. The rest of the proof we divide into two steps.
Claim 1. If
$G\curvearrowright (X,\mu _t)$
is conservative, then
$G\curvearrowright (X,\mu _s)$
is weakly mixing for every
$s<t$
.
Proof of Claim 1
Note that for every
$g\in G$
we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu25.png?pub-status=live)
so that
$(\mu _s)_r=\mu _{sr}$
. Therefore, it suffices to prove that
$G\curvearrowright (X,\mu _s)$
is weakly mixing for every
$s<1$
, assuming that
$G\curvearrowright (X,\mu _1)$
is conservative.
The claim is trivially true for
$s=0$
. So assume that
$G\curvearrowright (X,\mu _1)$
is conservative and fix
$s\in (0,1)$
. Let
$G\curvearrowright (Y,\eta )$
be an ergodic pmp action. Define
$Y_0=X_0\times X_0\times \{0,1\}$
and define the probability measures
$\unicode{x3bb} $
on
$\{0,1\}$
by
$\unicode{x3bb} (0)=s$
. Define the map
$\theta \colon Y_0\rightarrow X_0$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn13.png?pub-status=live)
Then for every
$g\in G$
we have that
$\theta _*(\mu _g\times \nu \times \unicode{x3bb} )=\mu _g^s$
. Write
$Z=\{0,1\}^G$
and equip Z with the probability measure
$\unicode{x3bb} ^{G}$
. We identify the Bernoulli action
$G\curvearrowright Y_0^{G}$
with the diagonal action
$G\curvearrowright X\times X\times Z$
. By applying
$\theta $
in each coordinate we obtain a G-equivariant factor map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn14.png?pub-status=live)
Then the map
$\mathord {\textrm {id}}_Y\times \Psi \colon Y\times X\times X\times Z\rightarrow Y\times X$
is G-equivariant and we have that
$(\mathord {\textrm { id}}_Y\times \Psi )_*(\eta \times \mu _1\times \mu _0\times \unicode{x3bb} ^G)=\eta \times \mu _s$
. The construction above is similar to [Reference Kosloff and SooKS20, §4].
Take
$F\in L^{\infty }(Y\times X,\eta \times \mu _s)^{G}$
. Note that the diagonal action
$G\curvearrowright (Y\times X,\eta \times \mu _1)$
is conservative, since
$G\curvearrowright (Y,\eta )$
is pmp. The action
$G\curvearrowright (X\times Z,\mu _0\times \unicode{x3bb} ^{G})$
can be identified with a pmp Bernoulli action with base space
$(X_0\times \{0,1\},\nu \times \unicode{x3bb} )$
, so that it is mixing. By [Reference Schmidt and WaltersSW81, Theorem 2.3] we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu26.png?pub-status=live)
which implies that the assignment
$(y,x,x',z)\mapsto F(y, \Psi (x,x',z))$
is essentially independent of
$x'$
and z. Choosing a finite set of coordinates
$\mathcal {F}\subset G$
and changing, for
$g\in \mathcal {F}$
, the value
$z_g$
between
$0$
and
$1$
, we see that F is essentially independent of the
$x_g$
-coordinates for
$g\in \mathcal {F}$
. As this is true for any finite set
$\mathcal {F}\subset G$
, we have that
$F\in L^{\infty }(Y)^{G}\mathbin {\overline {\otimes }} 1$
. The action
$G\curvearrowright (Y,\eta )$
is ergodic and therefore F is essentially constant. We conclude that
$G\curvearrowright (X,\mu _s)$
is weakly mixing.
Claim 2. If
$\nu \sim \mu _e$
and if
$G\curvearrowright (X,\mu _t)$
is not dissipative, then
$G\curvearrowright (X,\mu _s)$
is conservative for every
$s<t$
.
Proof of Claim 2
Again it suffices to assume that
$G\curvearrowright (X,\mu _1)$
is not dissipative and to show that
$G\curvearrowright (X,\mu _s)$
is conservative for every
$s<1$
.
When
$s=0$
, the statement is trivial, so assume that
$G\curvearrowright (X,\mu _1)$
is not dissipative and fix
$s\in (0,1)$
. Let
$C\subset X$
denote the non-negligible conservative part of
$G\curvearrowright (X,\mu _1)$
. As in the proof of Claim 1, write
$Z=\{0,1\}^{G}$
and let
$\unicode{x3bb} $
be the probability measure on
$\{0,1\}$
given by
$\unicode{x3bb} (0)=s$
. Writing
$\Psi \colon X\times X\times Z\rightarrow X$
for the G-equivariant map (3.4). We claim that
$\Psi _*((\mu _1\times \mu _0\times \unicode{x3bb} ^{G})|_{C\times X\times Z})\sim \mu _s$
, so that
$G\curvearrowright (X,\mu _s)$
is a factor of a conservative non-singular action, and therefore must be conservative itself.
As
$\Psi _*(\mu _1\times \mu _0\times \unicode{x3bb} ^{G})=\mu _s$
, we have that
$\Psi _*((\mu _1\times \mu _0\times \unicode{x3bb} ^{G}) |_{C\times X\times Z})\prec \mu _s$
. Let
$\mathcal {U}\subset X$
be the Borel set, uniquely determined up to a set of measure zero, such that
$\Psi _*((\mu _1\times \mu _0\times \unicode{x3bb} ^{G}) |_{C\times X\times Z})\sim \mu _s |_{\mathcal {U}}$
. We have to show that
$\mu _s(X\setminus \mathcal {U})=0$
. Fix a finite subset
$\mathcal {F}\subset G$
. For every
$t\in [0,1]$
define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu27.png?pub-status=live)
We shall write
$\gamma _1=\gamma _1^1, \gamma _2=\gamma _2^1$
. Also define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu28.png?pub-status=live)
By applying the map (3.3) in every coordinate, we get factor maps
$\Psi _j\colon Y_j\rightarrow X_j$
that satisfy
$(\Psi _j)_*(\zeta _j)=\gamma _j^{s}$
for
$j=1,2$
. Identify
$X_1\times Y_2\cong X\times (X_0\times \{0,1\})^{G\setminus \mathcal {F}}$
and define the subset
$C'\subset X_1\times Y_2$
by
$C'=C\times (X_0\times \{0,1\})^{G\setminus \mathcal {F}}$
. Let
$\mathcal {U}'\subset X$
be Borel such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu29.png?pub-status=live)
Identify
$Y_1\times X_2\cong X\times (X_0\times \{0,1\})^{\mathcal {F}}$
and define
$V\subset Y_1\times X_2$
by
$V=\mathcal {U}'\times (X_0\times \{0,1\})^{\mathcal {F}}$
. Then we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu30.png?pub-status=live)
Let
$\pi \colon X_1\times X_2\rightarrow X_2$
and
$\pi '\colon Y_1\times X_2\rightarrow X_2$
denote the coordinate projections. Note that by construction we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn15.png?pub-status=live)
Let
$W\subset X_2$
be Borel such that
$\pi _*(\mu _s |_{\mathcal {U}})\sim \gamma _2^s |_{W}$
. For every
$y\in X_2$
define the Borel sets
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu31.png?pub-status=live)
As
$\pi _*((\gamma _1\times \gamma _2^s) |_{\mathcal {U}'})\sim \gamma _2^s |_{W}$
, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu32.png?pub-status=live)
The disintegration of
$(\gamma _1\times \gamma _2^s) |_{\mathcal {U}'}$
along
$\pi $
is given by
$(\gamma _1 |_{\mathcal {U}^{\prime }_y})_{y\in W}$
. Therefore, the disintegration of
$(\zeta _1\times \gamma _2^s) |_{V}$
along
$\pi '$
is given by
$(\gamma _1 |_{\mathcal {U}^{\prime }_y}\times \nu ^{\mathcal {F}}\times \unicode{x3bb} ^{\mathcal {F}})_{y\in W}$
. We conclude that the disintegration of
$(\Psi _1\times \mathord {\textrm {id}}_{X_2})_*((\zeta _1\times \gamma _2^s) |_V)$
along
$\pi $
is given by
$((\Psi _1)_*(\gamma _1 |_{\mathcal {U}^{\prime }_y}\times \nu ^{\mathcal {F}}\times \unicode{x3bb} ^{\mathcal {F}}))_{y\in W}$
. The disintegration of
$\mu _s |_{\mathcal {U}}$
along
$\pi $
is given by
$(\gamma _2^s |_{\mathcal {U}_y})_{y\in W}$
. Since
$\mu _s |_{\mathcal {U}}\sim (\Psi _1\times \mathord {\textrm { id}}_{X_2})_*((\zeta _1\times \gamma _2^s) |_V)$
, we conclude that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu33.png?pub-status=live)
As
$\gamma _1(\mathcal {U}^{\prime }_y)>0$
for
$\gamma _2^s$
-a.e.
$y\in W$
, and using that
$\nu \sim \mu _e$
, we see that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu34.png?pub-status=live)
for
$\gamma _2^{s}$
-a.e.
$y\in W$
. It is clear that also
$(\Psi _1)_*(\gamma _1 |_{\mathcal {U}^{\prime }_y}\times \nu ^{\mathcal {F}}\times \unicode{x3bb} ^{\mathcal {F}})\prec \gamma _1^{s}$
, so that
$\gamma _1^{s} |_{\mathcal {U}_y}\sim \gamma _1^{s}$
for
$\gamma _2^s$
-a.e.
$y\in W$
. Therefore, we have that
$\gamma _1^s(X_1\setminus \mathcal {U}_y)=0$
for
$\gamma _2^s$
-a.e.
$y\in W$
, so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu35.png?pub-status=live)
Since this is true for every finite subset
$\mathcal {F}\subset G$
, we conclude that
$\mu _s(X\setminus \mathcal {U})=0$
.
The conclusion of the proof now follows by combining both claims. Assume that
${G\curvearrowright (X,\mu _t)}$
is not dissipative and fix
$s<t$
. Choose r such that
$s<r<t$
.
$\nu \sim \mu _e$
. By Claim 2 we have that
$G\curvearrowright (X,\mu _r)$
is conservative. Then by Claim 1 we see that
$G\curvearrowright (X,\mu _s)$
is weakly mixing.
$\nu \prec \mu _e$
. As
$\nu \prec \mu _e$
, the measures
$\mu _e^{t}$
and
$\mu _e$
are equivalent. We have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu36.png?pub-status=live)
So if
$\sup _{g\in G}|{\log}\ d\mu _g/d\mu _e(x)|<+\infty $
for a.e
$x\in X_0$
, we also have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu37.png?pub-status=live)
It follows from [Reference Björklund, Kosloff and VaesBV20, Proposition 4.3] that
$G\curvearrowright (X,\mu _t)$
is conservative. Then by Claim 1 we have that
$G\curvearrowright (X,\mu _s)$
is weakly mixing.
Remark 3.4. Let I be a countably infinite set and suppose that we are given a family of equivalent probability measures
$(\mu _i)_{i\in I}$
on a standard Borel space
$X_0$
. Let
$\nu $
be a probability measure on
$X_0$
that is equivalent to all the
$\mu _i$
. If G is an lcsc group that acts on I such that for each
$i \in I$
the stabilizer subgroup
$G_i=\{g\in G:g\cdot i=i\}$
is compact, then the pmp generalized Bernoulli action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu38.png?pub-status=live)
is mixing. For
$t\in [0,1]$
write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu39.png?pub-status=live)
and assume that the generalized Bernoulli action
$G\curvearrowright (X,\mu _1)$
is non-singular.
Since [Reference Schmidt and WaltersSW81, Theorem 2.3] still applies to infinitely recurrent actions of lcsc groups (see [Reference Arano, Isono and MarrakchiAIM19, Remark 7.4]), it is straightforward to adapt the proof of Claim 1 in the proof of Theorem 3.1 to prove that if
$G\curvearrowright (X,\mu _t)$
is infinitely recurrent, then
$G\curvearrowright (X,\mu _s)$
is weakly mixing for every
$s<t$
. Similarly, we can adapt the proof of Claim 2, using that a factor of an infinitely recurrent action is again infinitely recurrent. Together, this leads to the following phase transition result in the lcsc setting.
Assume that
$G_i=\{g\in G:g\cdot i=i\}$
is compact for every
$i\in I$
and that
$\nu \sim \mu _e$
. Then there exists a
$t_1\in [0,1]$
such that
$G\curvearrowright (X,\mu _t)$
is dissipative up to compact stabilizers for every
$t>t_1$
and weakly mixing for every
$t<t_1$
.
Recall the following definition from [Reference Berendschot and VaesBKV19, Definition 4.2]. When G is a countable infinite group and
$G\curvearrowright (X,\mu )$
is a non-singular action on a standard probability space, a sequence
$(\eta _n)$
of probability measures on G is called strongly recurrent for the action
$G\curvearrowright (X,\mu )$
if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu40.png?pub-status=live)
We say that
$G\curvearrowright (X,\mu )$
is strongly conservative if there exists a sequence
$(\eta _n)$
of probability measures on G that is strongly recurrent for
$G\curvearrowright (X,\mu )$
.
Lemma 3.5. Let
$G\curvearrowright (X,\mu )$
and
$G\curvearrowright (Y,\nu )$
be non-singular actions of a countable infinite group G on standard probability spaces
$(X,\mu )$
and
$(Y,\nu )$
. Suppose that
$\psi \colon (X,\mu )\rightarrow (Y,\nu )$
is a measure-preserving G-equivariant factor map and that
$\eta _n$
is a sequence of probability measures on G that is strongly recurrent for the action
${G\curvearrowright (X,\mu )}$
. Then
$\eta _n$
is strongly recurrent for the action
$G\curvearrowright (Y,\nu )$
.
Proof. Let
$E\colon L^0(X,[0,+\infty ))\rightarrow L^0(Y,[0,+\infty ))$
denote the conditional expectation map that is uniquely determined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu41.png?pub-status=live)
for all positive measurable functions
$F\colon X\rightarrow [0,+\infty )$
and
$H\colon Y\rightarrow [0,+\infty )$
. Since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu42.png?pub-status=live)
for every
$k\in G$
, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn16.png?pub-status=live)
By Jensen’s inequality for conditional expectations, applied to the convex function
${t\mapsto 1/t}$
, we also have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn17.png?pub-status=live)
Combining (3.6) and (3.7), we see that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu43.png?pub-status=live)
which converges to
$0$
as
$\eta _n$
is strongly recurrent for
$G\curvearrowright (X,\mu )$
.
We say that a non-singular group action
$G\curvearrowright (X,\mu )$
has an invariant mean if there exists a G-invariant linear functional
$\varphi \in L^{\infty }(X)^*$
. We say that
$G\curvearrowright (X,\mu )$
is amenable (in the sense of Zimmer) if there exists a G-equivariant conditional expectation
$E\colon L^{\infty }(G\times X)\rightarrow L^{\infty }(X)$
, where the action
$G\curvearrowright G\times X$
is given by
$g\cdot (h,x)=(gh,g\cdot x)$
.
Proposition 3.6. Let G be a countable infinite group and let
$(\mu _g)_{g\in G }$
be a family of equivalent probability measures on a standard Borel space
$X_0$
that is not supported on a single atom. Let
$\nu $
be a probability measure on
$X_0$
and for each
$t\in [0,1]$
consider the Bernoulli action (3.2). Assume that
$G\curvearrowright (X,\mu _1)$
is non-singular.
-
(1) If
$G\curvearrowright (X,\mu _t)$ has an invariant mean, then
$G\curvearrowright (X,\mu _s)$ has an invariant mean for every
$s<t$ .
-
(2) If
$G\curvearrowright (X,\mu _t)$ is amenable, then
$G\curvearrowright (X,\mu _s)$ is amenable for every
$s>t$ .
-
(3) If
$G\curvearrowright (X,\mu _t)$ is strongly conservative, then
$G\curvearrowright (X,\mu _s)$ is strongly conservative for every
$s<t$ .
Proof. (1) We may assume that
$t=1$
. So suppose that
$G\curvearrowright (X,\mu _1)$
has an invariant mean and fix
$s<1$
. Let
$\unicode{x3bb} $
be the probability measure on
$\{0,1\}$
that is given by
${\unicode{x3bb} (0)=s}$
. Then by [Reference Arano, Isono and MarrakchiAIM19, Proposition A.9] the diagonal action
$G\curvearrowright (X\times X\times \{0,1\}^{G}, \mu _1\times \mu _0\times \unicode{x3bb} ^{G})$
has an invariant mean. Since
$G\curvearrowright (X,\mu _s)$
is a factor of this diagonal action, it admits a G-invariant mean as well.
(2) It suffices to show that
$G\curvearrowright (X,\mu _1)$
is amenable whenever there exists a
${t\in (0,1)}$
such that
$G\curvearrowright (X,\mu _t)$
is amenable. Write
$\unicode{x3bb} $
for the probability measure on
$\{0,1\}$
given by
$\unicode{x3bb} (0)=t$
. Then
$G\curvearrowright (X,\mu _t)$
is a factor of the diagonal action
$G\curvearrowright (X\times X\times \{0,1\}^{G},\mu _1\times \mu _0\times \unicode{x3bb} ^{G})$
, so by [Reference ZimmerZim78, Theorem 2.4] also the latter action is amenable. Since
$G\curvearrowright (X\times \{0,1\}^{G},\mu _0\times \unicode{x3bb} ^{G})$
is pmp, we have that
$G\curvearrowright (X,\mu _1)$
is amenable.
(3) We may again assume that
$t=1$
. Suppose that
$(\eta _n)$
is a strongly recurrent sequence of probability measures on G for the action
$G\curvearrowright (X,\mu _1)$
. Fix
$s<1$
and let
$\unicode{x3bb} $
be the probability measure on
$\{0,1\}$
defined by
$\unicode{x3bb} (0)=s$
. As the diagonal action
$G\curvearrowright (X\times \{0,1\}^{G},\mu _0\times \unicode{x3bb} ^{G})$
is pmp, the sequence
$\eta _n$
is also strongly recurrent for the diagonal action
$G\curvearrowright (X\times X\times \{0,1\},\mu _1\times \mu _0\times \unicode{x3bb} ^{G})$
. Since
$G\curvearrowright (X,\mu _t)$
is a factor of
$G\curvearrowright (X\times X\times \{0,1\}^{G},\mu _1\times \mu _0\times \unicode{x3bb} ^{G})$
, it follows from Lemma 3.5 that the sequence
$\eta _n$
is strongly recurrent for
$G\curvearrowright (X,\mu _t)$
.
We finally prove Theorem 3.3. The proof relies heavily upon the techniques developed in [Reference Marrakchi and VaesMV20, §5].
Proof of Theorem 3.3
For every
$t\in (0,1]$
write
$\rho ^t$
for the Koopman representation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu44.png?pub-status=live)
Fix
$s\in (0,1)$
and let
$C>0$
be such that
$\log (1-x)\geq -C x$
for every
$x\in [0,s)$
. Then for every
$t<s$
and every
$g\in G$
we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu45.png?pub-status=live)
Because
$G\curvearrowright (X,\mu _1)$
is non-singular we get that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn18.png?pub-status=live)
We claim that there exists a
$t'>0$
such that
$G\curvearrowright (X,\mu _t)$
is non-amenable for every
${t<t'}$
. Suppose, to the contrary, that
$t_n$
is a sequence that converges to zero such that
${G\curvearrowright (X,\mu _{t_n})}$
is amenable for every
$n\in \mathbb {N}$
. Then it follows from [Reference NevoNev03, Theorem 3.7] that
$\rho ^{t_n}$
is weakly contained in the left regular representation
$\unicode{x3bb} _G$
for every
$n\in \mathbb {N}$
. Write
$1_G$
for the trivial representation of G. It follows from (3.8) that
$\bigoplus _{n\in \mathbb {N}}\rho ^{t_n}$
has almost invariant vectors, so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu46.png?pub-status=live)
which is in contradiction to the non-amenability of G. By Theorem 3.1 there exists a
${t_1\in [0,1]}$
such that
$G\curvearrowright (X,\mu _t)$
is weakly mixing for every
$t<t_1$
. Since every dissipative action is amenable (see, for example, [Reference Arano, Isono and MarrakchiAIM19, Theorem A.29]) it follows that
$t_1\geq t'>0$
.
Write
$Z_0=[0,1)$
and let
$\unicode{x3bb} $
denote the Lebesgue probability measure on
$Z_0$
. Let
$\rho ^0$
denote the reduced Koopman representation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu47.png?pub-status=live)
As G is non-amenable,
$\rho ^{0}$
has stable spectral gap. Suppose that for every
$s>0$
we can find
$0<s'<s$
such that
$\rho ^{s'}$
is weakly contained in
$\rho ^{s'}\otimes \rho ^{0}$
. Then there exists a sequence
$s_n$
that converges to zero, such that
$\rho ^{s_n}$
is weakly contained in
$\rho ^{s_n}\otimes \rho ^{0}$
for every
$n\in \mathbb {N}$
. This implies that
$\bigoplus _{n\in \mathbb {N}}\rho ^{s_n}$
is weakly contained in
$(\bigoplus _{n\in \mathbb {N}}\rho ^{s_n})\otimes \rho ^{0}$
. But by (3.8), the representation
$\bigoplus _{n\in \mathbb {N}}\rho ^{s_n}$
has almost invariant vectors, so that
$(\bigoplus _{n\in \mathbb {N}}\rho ^{s_n})\otimes \rho ^{0}$
weakly contains the trivial representation. This is in contradiction to
$\rho ^{0}$
having stable spectral gap. We conclude that there exists an
$s>0$
such that
$\rho ^t$
is not weakly contained in
$\rho ^t\otimes \rho ^0$
for every
$t<s$
.
We prove that
$G\curvearrowright (X,\mu _t)$
is strongly ergodic for every
$t<\min \{t',s\}$
, in which case we can apply [Reference Marrakchi and VaesMV20, Lemma 5.2] to the non-singular action
$G\curvearrowright (X,\mu _t)$
and the pmp action
$G\curvearrowright (X\times Z_0^{G},\mu _0\times \unicode{x3bb} ^G)$
by our choice of
$t'$
and s. After rescaling, we may assume that
$G\curvearrowright (X,\mu _1)$
is ergodic and that
$\rho ^{t}$
is not weakly contained in
$\rho ^{t}\otimes \rho ^{0}$
for every
$t\in (0,1)$
.
Let
$t\in (0,1)$
be arbitrary and define the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu48.png?pub-status=live)
Then
$\Psi $
is G-equivariant and we have that
$\Psi (\mu _1\times \mu _0\times \unicode{x3bb} ^{G})=\mu _t$
. Suppose that
${G\curvearrowright (X,\mu _t)}$
is not strongly ergodic. Then we can find a bounded almost invariant sequence
$f_n\in L^{\infty }(X,\mu _t)$
such that
$\|f_n\|_2=1$
and
$\mu _t(f_n)=0$
for every
$n\in \mathbb {N}$
. Therefore,
$\Psi _*(f_n)$
is a bounded almost invariant sequence for
$G\curvearrowright (X\times X\times Z_0^{G},\mu _1\times \mu _0\times \unicode{x3bb} ^{G})$
. Let
$E\colon L^{\infty }(X\times X\times Z_0^{G})\rightarrow L^{\infty }(X)$
be the conditional expectation that is uniquely determined by
$\mu _1\circ E=\mu _1\times \mu _0\times \unicode{x3bb} ^{G}$
. By [Reference Marrakchi and VaesMV20, Lemma 5.2] we have that
$\lim _{n\rightarrow \infty }\|(E\circ \Psi _*)(f_n)-\Psi _*(f_n)\|_2=0$
. As
$\Psi $
is measure-preserving we get, in particular, that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn19.png?pub-status=live)
Note that if
$\mu _t(f)=0$
for some
$f\in L^{2}(X,\mu _t)$
, we have that
$\mu _1((E\circ \Psi _*)(f))=0$
. So we can view
$E\circ \Psi _*$
as a bounded operator
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu49.png?pub-status=live)
Claim. The bounded operator
$E\circ \Psi _*\colon L^2(X,\mu _t)\ominus \mathbb {C} 1\rightarrow L^2(X,\mu _1)\ominus \mathbb {C} 1$
has norm strictly less than
$1$
.
The claim is in direct contradiction to (3.9), so we conclude that
$G\curvearrowright (X,\mu _t)$
is strongly ergodic.
Proof of claim
For every
$g\in G$
, let
$\varphi _g$
be the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu50.png?pub-status=live)
Then
$E\circ \Psi _*\colon L^2(X_0,\mu _t)\rightarrow L^2(X,\mu _1)$
is given by the infinite product
$\bigotimes _{g\in G}\varphi _g$
. For every
$g\in G$
we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu51.png?pub-status=live)
so that the inclusion map
$\iota _g \colon L^2(X_0,\mu _g^t)\hookrightarrow L^2(X_0,\mu _g)$
satisfies
$\|\iota _g\|\leq t^{-1/2}$
for every
$g\in G$
. We have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu52.png?pub-status=live)
So if we write
$P_g^t$
for the projection map onto
$L^2(X_0,\mu _g^t)\ominus \mathbb {C} 1$
, and
$P_g$
for the projection map onto
$L^2(X_0,\mu _g)\ominus \mathbb {C} 1$
, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn20.png?pub-status=live)
For a non-empty finite subset
$\mathcal {F}\subset G$
let
$V(\mathcal {F})$
be the linear subspace of
$L^2(X,\mu _t)\ominus \mathbb {C} 1$
spanned by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu53.png?pub-status=live)
Then, using (3.10), we see that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu54.png?pub-status=live)
Since
$\bigoplus _{\mathcal {F}\neq \emptyset }V(\mathcal {F})$
is dense inside
$L^2(X,\mu _t)\ominus \mathbb {C} 1$
, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu55.png?pub-status=live)
This also concludes the proof of Theorem 3.3.
4 Non-singular Bernoulli actions arising from groups acting on trees: proof of Theorem C
Let T be a locally finite tree and choose a root
$\rho \in T$
. Let
$\mu _0$
and
$\mu _1$
be equivalent probability measures on a standard Borel space
$X_0$
. Following [Reference Arano, Isono and MarrakchiAIM19, §10], we define a family of equivalent probability measures
$(\mu _e)_{e\in E}$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn21.png?pub-status=live)
Let
$G\subset \operatorname {Aut}(T)$
be a subgroup. When
$g\in G$
and
$e\in E$
, the edges e and
$g\cdot e$
are simultaneously oriented towards, or away from
$\rho $
, unless
$e\in E([\rho ,g\cdot \rho ])$
. As
$E([\rho ,g\cdot \rho ])$
is finite for every
$g\in G$
, the generalized Bernoulli action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn22.png?pub-status=live)
is non-singular. If we start with a different root
$\rho '\in T$
, let
$(\mu ^{\prime }_e)_{e\in E}$
denote the corresponding family of probability measures on
$X_0$
. Then we have that
$\mu _e=\mu ^{\prime }_e$
for all but finitely many
$e\in E$
, so that the measures
$\prod _{e\in E}\mu _e$
and
$\prod _{e\in E}\mu ^{\prime }_e$
are equivalent. Therefore, up to conjugacy, the action (4.2) is independent of the choice of root
$\rho \in T$
.
Lemma 4.1. Let T be a locally finite tree such that each vertex
$v\in V(T)$
has degree at least
$2$
. Suppose that
$G\subset \operatorname {Aut}(T)$
is a countable subgroup. Let
$\mu _0$
and
$\mu _1$
be equivalent probability measures on a standard Borel space
$X_0$
and fix a root
$\rho \in T$
. Then the action
$\alpha \colon G\curvearrowright (X,\mu )$
given by (4.2) is essentially free.
Proof. Take
$g\in G\setminus \{e\}$
. It suffices to show that
$\mu (\{x\in X:g\cdot x=x\})=0$
. If g is elliptic, there exist disjoint infinite subtrees
$T_1,T_2\subset T$
such that
$g\cdot T_1=T_2$
. Note that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu56.png?pub-status=live)
are non-atomic and that g induces a non-singular isomorphism
$\varphi \colon (X_1,\mu _1)\rightarrow (X_2,\mu _2): \varphi (x)_e=x_{g^{-1}\cdot e}$
. We get that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu57.png?pub-status=live)
A fortiori
$\mu (\{x\in X:g\cdot x=x\})=0$
. If g is hyperbolic, let
$L_g\subset T$
denote its axis on which it acts by non-trivial translation. Then
$\prod _{e\in E(L_g)}(X_0,\mu _e)$
is non-atomic and by [Reference Berendschot and VaesBKV19, Lemma 2.2] the action
$g^{\mathbb {Z}}\curvearrowright \prod _{e\in E(L_g)}(X_0,\mu _e)$
is essentially free. This implies that also
$\mu (\{x\in X:g\cdot x=x\})=0$
.
We prove Theorem 4.2 below, which implies Theorem C and also describes the stable type when the action is weakly mixing.
Theorem 4.2. Let T be a locally finite tree with root
$\rho \in T$
. Let
$G\subset \operatorname {Aut}(T)$
be a closed non-elementary subgroup with Poincaré exponent
$\delta =\delta (G\curvearrowright T)$
given by (1.5). Let
$\mu _0$
and
$\mu _1$
be non-trivial equivalent probability measures on a standard Borel space
$X_0$
. Consider the generalized non-singular Bernoulli action
$\alpha \colon G\curvearrowright (X,\mu )$
given by (4.2). Then
$\alpha $
is:
-
• weakly mixing if
$1-H^2(\mu _0,\mu _1)>\exp (-\delta /2)$ ;
-
• dissipative up to compact stabilizers if
$1-H^2(\mu _0,\mu _1)<\exp (-\delta /2)$ .
Let
$G\curvearrowright (Y,\nu )$
be an ergodic pmp action and let
$\Lambda \subset \mathbb {R}$
be the smallest closed subgroup that contains the essential range of the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu58.png?pub-status=live)
Let
$\Delta \colon G\rightarrow \mathbb {R}_{>0}$
denote the modular function and let
$\Sigma $
be the smallest subgroup generated by
$\Lambda $
and
$\log (\Delta (G))$
.
Suppose that
$1-H^2(\mu _0,\mu _1)>\exp (-\delta /2)$
. Then the Krieger flow and the flow of weights of
$\beta \colon G\curvearrowright X\times Y$
are determined by
$\Lambda $
and
$\Sigma $
as follows.
-
(1) If
$\Lambda $ (respectively,
$\Sigma $ ) is trivial, then the Krieger flow (respectively, flow of weights) is given by
$\mathbb {R}\curvearrowright \mathbb {R}$ .
-
(2) If
$\Lambda $ (respectively,
$\Sigma $ ) is dense, then the Krieger flow (respectively, flow of weights) is trivial.
-
(3) If
$\Lambda $ (respectively,
$\Sigma $ ) equals
$a\mathbb {Z}$ , with
$a>0$ , then the Krieger flow (respectively, flow of weights) is given by
$\mathbb {R}\curvearrowright \mathbb {R}/a\mathbb {Z}$ .
In general, we do not know the behaviour of the action (4.2) in the critical situation
${1-H^2(\mu _0,\mu _1)=\exp (-\delta /2)}$
. However, if T is a regular tree and
$G\curvearrowright T$
has full Poincaré exponent, we prove in Proposition 4.3 below that the action is dissipative up to compact stabilizers. This is similar to [Reference Arano, Isono and MarrakchiAIM19, Theorems 8.4 and 9.10].
Proposition 4.3. Let T be a q-regular tree with root
$\rho \in T$
and let
$G\subset \operatorname {Aut}(T)$
be a closed subgroup with Poincaré exponent
$\delta =\delta (G\curvearrowright T)=\log (q-1)$
. Let
$\mu _0$
and
$\mu _1$
be equivalent probability measures on a standard Borel space
$X_0$
.
If
$1-H^2(\mu _0,\mu _1)=(q-1)^{-1/2}$
, then the action (4.2) is dissipative up to compact stabilizers.
Interesting examples of actions of the form (4.2) arise when
$G\subset \operatorname {Aut}(T)$
is the free group on a finite set of generators acting on its Cayley tree. In that case, following [Reference Arano, Isono and MarrakchiAIM19, §6] and [Reference Marrakchi and VaesMV20, Remark 5.3], we can also give a sufficient criterion for strong ergodicity.
Proposition 4.4. Let the free group
$\mathbb {F}_d$
on
$d\geq 2$
generators act on its Cayley tree T. Let
$\mu _0$
and
$\mu _1$
be equivalent probability measures on a standard Borel space
$X_0$
. Then the action (4.2) dissipative if
$1-H^2(\mu _0,\mu _1)\leq (2d-1)^{-1/2}$
and weakly mixing and non-amenable if
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/2}$
. Furthermore, the action (4.2) is strongly ergodic when
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/4}$
.
The proof of Theorem 4.2 below is similar to that of [Reference Lyons and PemantleLP92, Theorem 4] and [Reference Arano, Isono and MarrakchiAIM19, Theorems 10.3 and 10.4]
Proof of Theorem 4.2
Define a family
$(X_e)_{e\in E}$
of independent random variables on
$(X,\mu )=\prod _{e\in E}(X_0,\mu _e)$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn23.png?pub-status=live)
For
$v\in T$
we write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu59.png?pub-status=live)
Then we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu60.png?pub-status=live)
Since
$G\subset \operatorname {Aut}(T)$
is a closed subgroup, for each
$v\in T$
the stabilizer subgroup
$G_v=\{g\in G:g\cdot v= v\}$
is a compact open subgroup of G.
Suppose that
$1-H^2(\mu _0,\mu _1)<\exp (-\delta /2)$
. Then we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu61.png?pub-status=live)
by definition of the Poincaré exponent. Therefore, we have that
$\sum _{v\in G\cdot \rho }\exp (S_v(x)/2)<+\infty $
for a.e.
$x\in X$
. Let
$\unicode{x3bb} $
denote the left invariant Haar measure on G and define
${L=\unicode{x3bb} (G_\rho )}$
, where
$G_\rho =\{g\in G:g\cdot \rho =\rho \}$
. Then we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu62.png?pub-status=live)
We conclude that
$G\curvearrowright (X,\mu )$
is dissipative up to compact stabilizers.
Now assume that
$1-H^2(\mu _0,\mu _1)>\exp (-\delta /2)$
. We start by proving that
$G\curvearrowright (X,\mu )$
is infinitely recurrent. By [Reference Arano, Isono and MarrakchiAIM19, Theorem 8.17] we can find a non-elementary closed compactly generated subgroup
$G'\subset G$
such that
$1-H^2(\mu _0,\mu _1)>\exp (-\delta (G')/2)$
. Let
$T'\subset T$
be the unique minimal
$G'$
-invariant subtree. Then
$G'$
acts cocompactly on
$T'$
and we have that
$\delta (G')=\dim _{H}\partial T'$
. Let X and Y be independent random variables with distributions
$(\log d\mu _1/d\mu _0)_*\mu _0$
and
$(\log d\mu _0/d\mu _1)_*\mu _1$
, respectively. Set
$Z=X+Y$
and write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu63.png?pub-status=live)
The assignment
$t\mapsto \varphi (t)$
is convex,
$\varphi (t)=\varphi (1-t)$
for every t and
$\varphi (1/2)= (1-H^2(\mu _0,\mu _1))^2$
. We conclude that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu64.png?pub-status=live)
Write
$R_k$
for the sum of k independent copies of Z. By the Chernoff–Cramér theorem, as stated in [Reference Lyons and PemantleLP92], there exists an
$M\in \mathbb {N}$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn24.png?pub-status=live)
Below we define a new unoriented tree S. This means that the edge set of S consists of subsets
$\{v,w\}\subset V(S)$
. Fix a vertex
$\rho '\in T'$
and define the unoriented tree S as follows.
-
• S has vertices
$v\in T'$ so that
$d_{T'}(\rho ', v)$ is divisible by M.
-
• There is an edge
$\{v,w\}\in E(S)$ between two vertices
$v,w\in S$ if
$d_{T'}(v,w)=M$ and
$[\rho ',v]_{T'}\subset [\rho ',w]_{T'}$ .
Here the notation
$[\rho ',v]_{T'}$
means that we consider the line segment
$[\rho ',v]$
as a subtree of
$T'$
. We have that
$\dim _H\partial S=M\dim _H \partial T'= M\delta (G')$
. Form a random subgraph
$S(x)$
of S by deleting those edges
$\{v,w\}\in E(S)$
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu65.png?pub-status=live)
This is an edge percolation on S, where each edge remains with probability
${p=\mathbb {P}(R_M\geq 0)}$
. So by (4.4) we have that
$p\exp (\dim _H S)>1$
. Furthermore, if
$\{v,w\}$
and
$\{v',w'\}$
are edges of S so that
$E([v,w]_{T'})\cap E([v',w']_{T'})=\emptyset $
, their presence in
$S(x)$
constitutes independent events. So the percolation process is a quasi-Bernoulli percolation as introduced in [Reference LyonsLyo89]. Taking
$w\in (1,p\exp (\dim _H S))$
and setting
$w_n=w^{-n}$
, it follows from [Reference LyonsLyo89, Theorem 3.1] that percolation occurs almost surely, that is,
$S(x)$
contains an infinite connected component for a.e.
$x\in X$
. Writing
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu66.png?pub-status=live)
this means that for a.e.
$x\in (X,\mu )$
we can find a constant
$a_x>-\infty $
such that
$S^{\prime }_v(x)>a_x$
for infinitely many
$v\in T'$
. As
$T'/G'$
is finite, there exists a vertex
$w\in T'$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn25.png?pub-status=live)
Therefore, by Kolmogorov’s zero–one law, we have that
$\sum _{v\in G'\cdot w}\exp (S^{\prime }_v(x))=+\infty $
almost surely. Since a change of root results in a conjugate action, we may assume that
$\rho =w$
. Then (4.5) implies that
$\sum _{v\in G\cdot \rho }\exp (S_v(x))=+\infty $
for a.e.
$x\in X$
. Writing again L for the Haar measure of the stabilizer subgroup
$G_\rho =\{g\in G:g\cdot \rho = \rho \}$
, we see that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu67.png?pub-status=live)
We conclude that
$G\curvearrowright (X,\mu )$
is infinitely recurrent. We prove that
$G\curvearrowright (X,\mu )$
is weakly mixing using a phase transition result from the previous section. Define the measurable map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu68.png?pub-status=live)
Let
$\nu $
be the probability measure on
$X_0$
determined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu69.png?pub-status=live)
Then we have that
$\nu \sim \mu _0$
and for every
$s>1-\rho $
the probability measures
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu70.png?pub-status=live)
are well defined. We consider the non-singular actions
$G\curvearrowright (X,\eta _s)=\prod _{e\in E}(X_0,\eta _e^s)$
, where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu71.png?pub-status=live)
By the dominated convergence theorem we have that
$H^2(\eta _0^s,\eta _1^s)\rightarrow H^2(\mu _0,\mu _1)$
as
$s\rightarrow 1$
. So we can choose s close enough to
$1$
, but not equal to
$1$
, such that
$1-H^2(\eta _0^s,\eta _1^s)>\exp (-\delta /2)$
. By the first part of the proof we have that
$G\curvearrowright (X,\eta _s)$
is infinitely recurrent. Note that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu72.png?pub-status=live)
Since we assumed that
$G\subset \operatorname {Aut}(T)$
is closed, all the stabilizer subgroups
$G_{v}=\{g\in G:g\cdot v=v\}$
are compact. By Remark 3.4 we conclude that
$G\curvearrowright (X,\mu )$
is weakly mixing.
Let
$G\curvearrowright (Y,\nu )$
be an ergodic pmp action. To determine the Krieger flow and the flow of weights of
$\beta \colon G\curvearrowright X\times Y$
we use a similar approach to [Reference Arano, Isono and MarrakchiAIM19, Theorem 10.4] and [Reference Vaes and WahlVW17, Proposition 7.3]. First we determine the Krieger flow and then we deal with the flow of weights.
As before, let
$G'\subset G$
be a non-elementary compactly generated subgroup such that
${1-H^2(\mu _0,\mu _1)>\exp (-\delta (G')/2)}$
. By [Reference Arano, Isono and MarrakchiAIM19, Theorem 8.7] we may assume that
$G/G'$
is not compact. Let
$T'\subset T$
be the minimal
$G'$
-invariant subtree. Let
$v\in T'$
be as in Lemma 4.5 below so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn26.png?pub-status=live)
Since changing the root yields a conjugate action, we may assume that
$\rho =v$
. Let
$(Z_0,\zeta _0)$
be a standard probability space such that there exist measurable maps
$\theta _0,\theta _1\colon Z_0\rightarrow X_0$
that satisfy
$(\theta _0)_*\zeta _0=\mu _0$
and
$(\theta _1)_*\zeta _0=\mu _1$
. Write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu73.png?pub-status=live)
By the first part of the proof we have that
$G'\curvearrowright (X_2,\rho _2)$
is infinitely recurrent. Define the pmp map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu74.png?pub-status=live)
Consider
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu75.png?pub-status=live)
Since
$gU\triangle U=E(T)([\rho ,g\cdot \rho ])\subset E(T')$
for any
$g\in G'$
, the set
$(E(T)\setminus E(T'))\cap U$
is
$G'$
-invariant. Therefore,
$\Psi $
is a
$G'$
-equivariant factor map. Consider the Maharam extensions
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu76.png?pub-status=live)
of the diagonal actions
$G'\curvearrowright Z\times X_2\times Y$
and
$G'\curvearrowright X\times Y\times \mathbb {R}$
, respectively. Identifying
$(X,\mu )=(X_1,\rho _1)\times (X_2,\rho _2)$
, we obtain a
$G'$
-equivariant factor map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu77.png?pub-status=live)
Take
$F\in L^{\infty }(X\times Y\times \mathbb {R})^{G}$
. By [Reference Arano, Isono and MarrakchiAIM19, Proposition A.33] the Maharam extension
$G'\curvearrowright X_2\times Y\times \mathbb {R} $
is infinitely recurrent. Since
$G'\curvearrowright Z$
is a mixing pmp generalized Bernoulli action we have that
$F\circ \Phi \in L^{\infty }(Z\times X_2\times Y\times \mathbb {R})^{G}\subset 1\mathbin {\overline {\otimes }} L^{\infty }(X_2\times Y\times \mathbb {R})^{G}$
by [Reference Schmidt and WaltersSW81, Theorem 2.3]. Therefore, F is essentially independent of the
$E(T)\setminus E(T')$
-coordinates. Thus, for any
$g\in G$
the assignment
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu78.png?pub-status=live)
is essentially independent of the
$E(T)\setminus E(gT')$
-coordinates. Since
$\log (dg^{-1}\mu /d\mu )$
only depends on the
$E([\rho ,g^{-1}\cdot \rho ])$
-coordinates, we deduce that F is essentially independent of the
$E(T)\setminus (E(gT')\cup E([\rho ,g^{-1}\cdot \rho ]))$
-coordinates, for every
$g\in G$
. Therefore, by (4.6), we have that
$F\in 1\mathbin {\overline {\otimes }} L^{\infty }(Y\times \mathbb {R})$
.
So we have proven that any G-invariant function
$F\in L^{\infty }(X\times Y\times \mathbb {R})$
is of the form
$F(x,y,t)=H(y,t)$
, for some
$H\in L^{\infty }(Y\times \mathbb {R})$
that satisfies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu79.png?pub-status=live)
Since
$0$
is in the essential range of the maps
$\log (dg\mu /d\mu )$
, for every
$g\in G$
, we see that
$H(g\cdot y,t)=H(y,t)$
for a.e.
$(y,t)\in Y\times \mathbb {R}$
. By ergodicity of
$G\curvearrowright Y$
, we conclude that H is of the form
$H(y,t)=P(t)$
, for some
$P\in L^{\infty }(\mathbb {R})$
that satisfies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn27.png?pub-status=live)
Let
$\Gamma \subset \mathbb {R}$
be the subgroup generated by the essential ranges of the maps
$\log (dg\mu /d\mu )$
, for
$g\in G$
. If
$\Gamma =\{0\}$
we can identify
$L^{\infty }(X\times Y\times \mathbb {R})^{G}\cong L^{\infty }(\mathbb {R})$
. If
$\Gamma \subset \mathbb {R}$
is dense, then it follows that P is essentially constant so that the Maharam extension
$G\curvearrowright X\times Y\times \mathbb {R}$
is ergodic, that is, the Krieger flow of
$G\curvearrowright X\times Y$
is trivial. If
$\Gamma =a\mathbb {Z}$
, with
$a>0$
, we conclude by (4.7) that we can identify
$L^{\infty }(X\times Y\times \mathbb {R})^{G}\cong L^{\infty }(\mathbb {R}/a\mathbb {Z})$
, so that the Krieger flow of
$G\curvearrowright X\times Y$
is given by
$\mathbb {R}\curvearrowright \mathbb {R}/a\mathbb {Z}$
. Finally, note that the closure of
$\Gamma $
equals the closure of the subgroup generated by the essential range of the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu80.png?pub-status=live)
So we have calculated the Krieger flow in every case, concluding the proof of the theorem in the case where G is unimodular.
When G is not unimodular, let
$G_0=\ker \Delta $
be the kernel of the modular function. Let
$G\curvearrowright X\times Y\times \mathbb {R}$
be the modular Maharam extension and let
$\alpha \colon G_0\curvearrowright X\times Y\times \mathbb {R}$
be its restriction to the subgroup
$G_0$
. Then we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu81.png?pub-status=live)
By [Reference Arano, Isono and MarrakchiAIM19, Theorem 8.16] we have that
$\delta (G_0)=\delta $
, and we can apply the argument above to conclude that
$L^{\infty }(X\times Y\times \mathbb {R})^{\alpha }\subset 1\mathbin {\overline {\otimes }} 1\mathbin {\overline {\otimes }} L^{\infty }(\mathbb {R})$
. So for every
$F\in L^{\infty }(X\times Y\times \mathbb {R})^{G}$
there exists a
$P\in L^{\infty }(\mathbb {R})$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn28.png?pub-status=live)
Let
$\Pi $
be the subgroup of
$\mathbb {R}$
generated by the essential range of the maps
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu82.png?pub-status=live)
As
$0$
is contained in the essential range of
$\log (dg^{-1}\mu /d\mu )$
, for every
$g\in G$
, we get that
$\log (\Delta (G))\subset \Pi $
. Therefore,
$\Pi $
also contains the subgroup
$\Gamma \subset \mathbb {R}$
defined above. Thus, the closure of
$\Pi $
equals the closure of
$\Sigma $
, where
$\Sigma \subset \mathbb {R}$
is the subgroup as in the statement of the theorem. From (4.8) we conclude that we may identify
$L^{\infty }(X\times Y\times \mathbb {R})^{G}\cong L^{\infty }(\mathbb {R})^{\Sigma }$
, so that the flow of weights of
$G\curvearrowright X\times Y$
is as stated in the theorem.
Lemma 4.5. Let T be a locally finite tree and let
$G\subset \operatorname {Aut}(T)$
be a closed subgroup. Suppose that
$H\subset G$
is a closed compactly generated subgroup that contains a hyperbolic element and assume that
$G/H$
is not compact. Let
$S\subset T$
be the unique minimal H-invariant subtree. Then there exists a vertex
$v\in S$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn29.png?pub-status=live)
Proof. Let
$k\in H$
be a hyperbolic element and let
$L\subset T$
be its axis, on which k acts by a non-trivial translation. Then
$L\subset S$
, as one can show for instance as in the proof of [Reference Caprace and de MedtsCM11, Proposition 3.8]. Pick any vertex
$v\in L$
. We claim that this vertex will satisfy (4.9). Take any
$w\in V(T)\setminus \{v\}$
. As
$G/H$
is not compact, one can show as in [Reference Arano, Isono and MarrakchiAIM19, Theorem 9.7] that there exists a
$g\in G$
such that
$g\cdot w\notin S$
. Since k acts by translation on L, there exists an
$n\in \mathbb {N}$
large enough such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu83.png?pub-status=live)
so that in particular we have that
$w\notin [v,k^ng\cdot v]\cap [v,k^{-n}g\cdot v]=\{v\}$
. Since S is H-invariant, we also have that
$k^ng\cdot w\notin S$
and
$k^{-n}g\cdot w\notin S$
and we conclude that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu84.png?pub-status=live)
Proof of Proposition 4.3
Define the family
$(X_e)_{e\in E}$
of independent random variables on
$(X,\mu )$
by (4.3) and write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu85.png?pub-status=live)
Claim. There exists a
$\delta>0$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu86.png?pub-status=live)
Proof of claim
Note that
$\mathbb {E}(\exp (X_e/2))=1-H^2(\mu _0,\mu _1)$
for every
$e\in E$
. Define a family of random variables
$(W_n)_{n\geq 0}$
on
$(X,\mu )$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu87.png?pub-status=live)
Using that
$1-H^2(\mu _0,\mu _1)=(q-1)^{-1/2}$
, one computes that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu88.png?pub-status=live)
So the sequence
$(W_n)_{n\geq 0}$
is a martingale, and since it is positive it converges almost surely to a finite limit when
$n\rightarrow +\infty $
. Write
$\Sigma _n=\{v\in T:d(v,\rho )=n\}$
. As
${W_n\geq \max _{v\in \Sigma _n}\exp (S_v/2)}$
we conclude that there exists a positive constant
$C<+\infty $
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu89.png?pub-status=live)
For any vertex
$w\in T$
, write
$T_w=\{v\in T:[\rho ,w]\subset [\rho ,v]\}$
: the set of children of w, including w itself. Using the symmetry of the tree and changing the root from
$\rho $
to
$w\in T$
, we also have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn30.png?pub-status=live)
Set
$\nu _0=(\log d\mu _1/d\mu _0)_*\mu _0$
and
$\nu _1=(\log d\mu _0/d\mu _1)_*\mu _1$
. Because
$1-H^2(\mu _0,\mu _1)\neq 0$
we have that
$\mu _0\neq \mu _1$
, so that there exists a
$\delta>0$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu90.png?pub-status=live)
Here
$\nu _0*\nu _1$
denotes the convolution product of
$\nu _0$
with
$\nu _1$
. Therefore, there exists
$N\in \mathbb {N}$
large enough such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn31.png?pub-status=live)
Since for any
$w\in \Sigma _N$
and
$w'\in \Sigma _n$
with
$n\leq N$
, we have that
$S_v-S_w$
is independent of
$S_{w'}$
for every
$v\in T_w$
, and since
$\Sigma _N$
is a finite set, it follows from (4.10) and (4.11) that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu91.png?pub-status=live)
This concludes the proof of the claim.
Let
$\delta>0$
be as in the claim and define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu92.png?pub-status=live)
so that
$\mu (\mathcal {U})>0$
. Let
$G_\rho $
be the stabilizer subgroup of
$\rho $
. Note that for every
$g,h\in G$
we have that
$S_{hg\cdot \rho }(x)=S_{g\cdot \rho }(h^{-1}\cdot x)+S_{h\cdot \rho }(x)$
for a.e.
$x\in X$
, so that for
$h\in G$
we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu93.png?pub-status=live)
It follows that if
$h\notin G_{\rho }$
, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu94.png?pub-status=live)
Since
$G\subset \operatorname {Aut}(T)$
is closed, we have that
$G_\rho $
is compact. So the action
$G\curvearrowright (X,\mu )$
is not infinitely recurrent. Let
$\unicode{x3bb} $
denote the left invariant Haar measure on G. By an adaptation of the proof of [Reference Björklund, Kosloff and VaesBV20, Proposition 4.3], the set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu95.png?pub-status=live)
satisfies
$\mu (D)\in \{0,1\}$
. Since
$G\curvearrowright (X,\mu )$
is not infinitely recurrent, it follows from [Reference Arano, Isono and MarrakchiAIM19, Proposition A.28] that
$\mu (D)>0$
, so that we must have that
$\mu (D)=1$
. By [Reference Arano, Isono and MarrakchiAIM19, Theorem A.29] the action
$G\curvearrowright (X,\mu )$
is dissipative up to compact stabilizers.
We use a similar approach to [Reference Marrakchi and VaesMV20, §6] in the proof of Proposition 4.4.
Proof of Proposition 4.4
It follows from Theorem 4.2 and Proposition 4.3 that the action
$G\curvearrowright (X,\mu )$
, given by (4.2), is dissipative when
$1-H^2(\mu _0,\mu _1)\leq (2d-1)^{-1/2}$
and weakly mixing when
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/2}$
. So it remains to show that
$G\curvearrowright (X,\mu )$
is non-amenable when
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/2}$
and strongly ergodic when
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/4}$
.
Assume first that
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/2}$
. By taking the kernel of a surjective homomorphism
$\mathbb {F}_d\rightarrow \mathbb {Z}$
we find a normal subgroup
$H_1\subset \mathbb {F}_d$
that is free on infinitely many generators. By [Reference Roblin and TapieRT13, Théorème 0.1] we have that
$\delta (H_1)=(2d-1)^{-1/2}$
. Then, using [Reference SullivanSul79, Corollary 6], we can find a finitely generated free subgroup
$H_2\subset H_1$
such that
$H_1=H_2*H_3$
for some free subgroup
$H_3\subset H_1$
and such that
$1-H^2(\mu _0,\mu _1)>\exp (-\delta (H_2)/2)$
. Let
$\psi \colon H_1\rightarrow H_3$
be the surjective group homomorphism uniquely determined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu96.png?pub-status=live)
We set
$N=\ker \psi $
, so that
$H_2\subset N$
and we get that
$1-H^2(\mu _0,\mu _1)>\exp (-\delta (N)/2)$
. Therefore,
$N\curvearrowright (X,\mu )$
is ergodic by Theorem 4.2. Also we have that
$H_1/N\cong H_3$
, which is a free group on infinitely many generators. Therefore,
$H_1\curvearrowright (X,\mu )$
is non-amenable by [Reference Marrakchi and VaesMV20, Lemma 6.4]. A posteriori also
$\mathbb {F}_d\curvearrowright (X,\mu )$
is non-amenable.
Let
$\pi $
be the Koopman representation of the action
$\mathbb {F}_d\curvearrowright (X,\mu )$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu97.png?pub-status=live)
Claim. If
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/4}$
, then
$\pi $
is not weakly contained in the left regular representation.
Proof of claim
Let
$\eta $
denote the canonical symmetric measure on the generator set of
$\mathbb {F}_d$
and define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu98.png?pub-status=live)
The
$\eta $
-spectral radius of
$\alpha \colon \mathbb {F}_d\curvearrowright (X,\mu )$
, which we denote by
$\rho _\eta (\alpha )$
, is by definition the norm of P, as a bounded operator on
$L^2(X,\mu )$
. By [Reference Arano, Isono and MarrakchiAIM19, Proposition A.11] we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu99.png?pub-status=live)
where
$|g|$
denotes the word length of a group element
$g\in \mathbb {F}_d$
. By [Reference Arano, Isono and MarrakchiAIM19, Theorem 6.10] we then have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu100.png?pub-status=live)
if
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/4}$
, and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu101.png?pub-status=live)
if
$1-H^2(\mu _0,\mu _1)\leq (2d-1)^{-1/4}$
. Therefore, if
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/4}$
, we have that
$\rho _\eta (\alpha )>\rho _\eta (\mathbb {F}_d)$
, where
$\rho _\eta (\mathbb {F}_d)$
denotes the
$\eta $
-spectral radius of the left regular representation. This implies that
$\alpha $
is not weakly contained in the left regular representation (see, for instance, [Reference Anantharaman-DelarocheAD03, §3.2]).
Now assume that
$1-H^2(\mu _0,\mu _1)>(2d-1)^{-1/4}$
. As in the proof of Theorem 4.2 there exist probability measures
$\nu , \eta _0$
and
$\eta _1$
on
$X_0$
that are equivalent to
$\mu _0$
and a number
$s\in (0,1)$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu102.png?pub-status=live)
and such that
$1-H^2(\eta _0,\eta _1)>(2d-1)^{-1/4}$
. Consider the non-singular action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu103.png?pub-status=live)
By Theorem 4.2 the action
$\mathbb {F}_d\curvearrowright (X,\eta )$
is ergodic. Write
$\rho $
for the Koopman representation associated to
$\mathbb {F}_d\curvearrowright (X,\eta )$
. By the claim,
$\rho $
is not weakly contained in the left regular representation. Let
$\unicode{x3bb} $
be the probability measure on
$\{0,1\}$
given by
${\unicode{x3bb} (0)=s}$
. Let
$\rho ^0$
be the reduced Koopman representation of the pmp generalized Bernoulli action
${\mathbb {F}_d\curvearrowright (X\times \{0,1\}^{E(T)},\nu ^{E(T)}\times \unicode{x3bb} ^{E(T)})}$
. Then
$\rho ^0$
is contained in a multiple of the left regular representation. Therefore, as
$\rho $
is not weakly contained in the left regular representation,
$\rho $
is not weakly contained in
$\rho \otimes \rho ^{0}$
.
Define the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu104.png?pub-status=live)
Then
$\Psi $
is
$\mathbb {F}_d$
-equivariant and we have that
$\Psi _*(\eta \times \nu ^{E(T)}\times \unicode{x3bb} ^{E(T)})=\mu $
. Suppose that
$\mathbb {F}_d\curvearrowright (X,\mu )$
is not strongly ergodic. Then there exists a bounded almost invariant sequence
$f_n\in L^{\infty }(X,\mu )$
such that
$\|f_n\|_2=1$
and
$\mu (f_n)=0$
for every
$n\in \mathbb {N}$
. Therefore,
$\Psi _*(f_n)$
is a bounded almost invariant sequence for the diagonal action
${\mathbb {F}_d\curvearrowright (X\times X\times \{0,1\}^{E(T)},\eta \times \nu ^{E(T)}\times \unicode{x3bb} ^{E(T)})}$
. Let
$E\colon L^{\infty }(X\times X\times \{0,1\}^{E(T)})\rightarrow L^{\infty }(X)$
be the conditional expectation that is uniquely determined by
$\mu \circ E=\eta \times \nu ^{E(T)}\times \unicode{x3bb} ^{E(T)}$
. By [Reference Marrakchi and VaesMV20, Lemma 5.2] we have that
$\lim _{n\rightarrow \infty }\|(E\circ \Psi _*)(f_n)-\Psi _{*}(f_n)\|_2=0$
, and in particular we get that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqn32.png?pub-status=live)
But just as in the proof of Theorem 3.3 we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu105.png?pub-status=live)
which is in contradiction with (4.12). We conclude that
$\mathbb {F}_d\curvearrowright (X,\mu )$
is strongly ergodic.
Proposition 4.6 below complements Theorem 4.2 by considering groups
$G\subset \operatorname {Aut}(T)$
that are not closed. This is similar to [Reference Arano, Isono and MarrakchiAIM19, Theorem 10.5].
Proposition 4.6. Let T be a locally finite tree with root
$\rho \in T$
. Let
$G\subset \operatorname {Aut}(T)$
be an lcsc group such that the inclusion map
$G\rightarrow \operatorname {Aut}(T)$
is continuous and such that
${G\subset \operatorname {Aut}(T)}$
is not closed. Write
$\delta =\delta (G\curvearrowright T)$
for the Poincaré exponent given by (1.5). Let
$\mu _0$
and
$\mu _1$
be non-trivial equivalent probability measures on a standard Borel space
$X_0$
. Consider the generalized non-singular Bernoulli action
$\alpha \colon G\curvearrowright (X,\mu )$
given by (4.2). Let
$H\subset \operatorname {Aut}(T)$
be the closure of G. Then the following assertions hold.
-
• If
$1-H^2(\mu _0,\mu _1)>\exp (-\delta /2)$ , then
$\alpha $ is ergodic and its Krieger flow is determined by the essential range of the map
(4.13)as in Theorem 4.2.$$ \begin{align} X_0\times X_0\rightarrow \mathbb{R}:\quad (x,x')\mapsto\log(d\mu_0/d\mu_1)(x)-\log(d\mu_0/d\mu_1)(x') \end{align} $$
-
• If
$1-H^2(\mu _0,\mu _1)<\exp (-\delta /2)$ , then each ergodic component of
$\alpha $ is of the form
$G\curvearrowright H/K$ , where K is a compact subgroup of H. In particular, there exists a G-invariant
$\sigma $ -finite measure on X that is equivalent to
$\mu $ .
Proof. Let
$H\subset \operatorname {Aut}(T)$
be the closure of G. Then
$\delta (H)=\delta $
and we can apply Theorem 4.2 to the non-singular action
$H\curvearrowright (X,\mu )$
.
If
$1-H^2(\mu _0,\mu _1)>\exp (-\delta /2)$
, then
$H\curvearrowright X$
is ergodic. As
$G\subset H$
is dense, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu106.png?pub-status=live)
so that
$G\curvearrowright X$
is ergodic. Let
$H\curvearrowright X\times \mathbb {R}$
be the Maharam extension associated to
${H\curvearrowright X}$
. Again, as
$G\subset H$
is dense, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20240107132950335-0805:S014338572300024X:S014338572300024X_eqnu107.png?pub-status=live)
Note that the subgroup generated by the essential ranges of the maps
$\log (dg^{-1}\mu /d\mu )$
, with
$g\in G$
, is the same as the subgroup generated by the essential ranges of the maps
$\log (dh^{-1}\mu /d\mu )$
, with
$h\in H$
. Then one determines the Krieger flow of
$G\curvearrowright X$
as in the proof of Theorem 4.2.
If
$1-H^2(\mu _0,\mu _1)<\exp (-\delta /2)$
, the action
$H\curvearrowright (X,\mu )$
is dissipative up to compact stabilizers. By [Reference Arano, Isono and MarrakchiAIM19, Theorem A.29] each ergodic component is of the form
$H\curvearrowright H/K$
for a compact subgroup
$K\subset H$
. Therefore, each ergodic component of
$G\curvearrowright (X,\mu )$
is of the form
$G\curvearrowright H/K$
, for some compact subgroup
$K\subset H$
.
Acknowledgements
T.B. thanks Stefaan Vaes for his valuable feedback during the process of writing this paper. T.B. is supported by a PhD fellowship fundamental research of the Research Foundation Flanders.