Full Groups of Flows
Abstract.
We introduce the concept of an full group associated with a measure-preserving action of a Polish normed group on a standard probability space. Such groups are shown to carry a natural separable complete metric, and are thus Polish. Our construction generalizes full groups of actions of discrete groups, which have been studied recently by the first author.
We show that under minor assumptions on the actions, topological derived subgroups of full groups are topologically simple and — when the acting group is locally compact and amenable — are whirly amenable and generically two-generated. full groups of actions of compactly generated locally compact Polish groups are shown to remember the orbit equivalence class of the action.
For measure-preserving actions of the real line (also often called measure-preserving flows), the topological derived subgroup of an full groups is shown to coincide with the kernel of the index map, which implies that full groups of free measure-preserving flows are topologically finitely generated if and only if the flow admits finitely many ergodic components. The latter is in a striking contrast to the case of -actions, where the number of topological generators is controlled by the entropy of the action.
We also study the coarse geometry of the full groups. The norm on the derived subgroup of the full group of an aperiodic action of a locally compact amenable group is proved to be maximal in the sense of Rosendal. For measure-preserving flows, this holds for the norm on all of the full group.
2020 Mathematics Subject Classification:
Primary 37A10, 37A15; Secondary 37A05, 37A20.Chapter 1 Introduction
Full groups were introduced by H. Dye [Dye59] in the framework of measure-preserving actions of countable groups as measurable analogues of unitary groups of von Neumann algebras, by mimicking the fact that the latter are stable under countable cutting and pasting of partial isometries. These Polish groups have since been recognized as important invariants as they encode the induced partition of the space into orbits. A similar viewpoint applies in the setup of minimal homeomorphisms on the Cantor space [GPS99], where likewise the full groups are responsible for the orbit equivalence class of the action.
Full groups are defined to consist of transformations which act by a permutation on each orbit. When the action is free, one can associate with an element of the full group a cocycle defined by the equation . From the point of view of topological dynamics, it is natural to consider the subgroup of those for which the cocycle map is continuous, which is the defining condition for the so-called topological full groups. The latter has a much tighter control of the action, and encodes minimal homeomorphisms of the Cantor space up to flip-conjugacy (see [GPS99]).
A celebrated result of H. Dye states that all ergodic -actions produce the same partition up to isomorphism, and hence the associated full groups are all isomorphic. The first named author has been motivated by the above to seek for the analog of topological full groups in the context of ergodic theory, which was achieved in [LM18] by imposing integrability conditions on the cocycle. In particular, he introduced full groups of measure-preserving ergodic transformations, and showed based on the result of R. M. Belinskaja [Bel68] that they also determine the action up to flip-conjugacy. Unlike in the context of Cantor dynamics, these full groups are uncountable, but they carry a natural Polish topology.
In this work, we widen the concept of an full group and associate such an object with any measure-preserving Borel action of a Polish normed group (the reader may consult Appendix A for a concise reminder about group norms). Quasi-isometric compatible norms will result in the same full groups, so actions of Polish boundedly generated groups have canonical full groups associated with them based on to the work of C. Rosendal [Ros22]. Our study also parallels the generalization of the full group construction introduced by A. Carderi and the first named author in [CLM16], where full groups were defined for Borel measure-preserving actions of Polish groups.
1.1. Main results
Let be a Polish group with a compatible norm and consider a Borel measure-preserving action on a standard probability space . The group action defines an orbit equivalence relation by declaring points equivalent whenever . The norm induces a metric onto each -class via . Following [CLM16], a full group of the action is denoted by and is defined as the collection of all measure-preserving that satisfy for all . The full group is given by those for which the map is integrable. This defines a subgroup of , and we show in Theorem 2.10 that these groups are Polish in the topology of the norm . The strategy of establishing this statement is analogous to that of [CLM18], where the Polish topology for full groups was defined.
It is a general and well-known phenomenon in the study of all kinds and variants of full groups that their structure is usually best understood through the derived subgroups. Our setup is no exception.
Theorem 1.
The topological derived group of any aperiodic full group is equal to the closed subgroup generated by involutions.
The argument needed for Theorem 1 is quite robust. We extract the idea used in [LM18], isolate the class of finitely full groups, and show that under mild assumptions on the action, Theorem 1 holds for such groups. We provide these arguments in Section 3 and in Corollary 3.15 in particular. Alongside we mention Corollary 3.21 which implies that full groups of ergodic actions are topologically simple.
For the rest of our results we narrow down the generality of the acting groups, and consider locally compact Polish normed groups. In Chapter 4, we show that if is a dense subgroup of a locally compact Polish normed group then is dense in . In fact, we prove a considerably stronger statement by showing that for each and there is such that .
Recall that a topological group is amenable if all of its continuous actions on compact spaces preserve some Radon probability measure, and that it is whirly amenable if it is amenable and moreover every invariant Radon measure is supported on the set of fixed points. The following is a combination of Theorem 5.8 and Corollary 5.10.
Theorem 2.
Let be a measure-preserving action of a locally compact Polish normed group. Consider the following three statements:
-
(1)
is amenable;
-
(2)
the topological derived group is whirly amenable.
-
(3)
the full group is amenable;
The implications (1) (2) (3) always hold. If is unimodular and the action is free, then the three statements above are all equivalent.
When the acting group is amenable and orbits of the action are uncountable, we are able to compute the topological rank of the derived full groups — that is, the minimal number of elements needed to generate a dense subgroup. Theorem 5.19 contains a stronger version of the following.
Theorem 3.
Let be a measure-preserving action of an amenable locally compact Polish normed group on a standard probability space . If all orbits of the action are uncountable, then the topological rank of the derived full group is equal to .
It is instructive to contrast the situation with the actions of finitely generated groups, where finiteness of the topological rank of the derived full group is equivalent to finiteness of the Rokhlin entropy of the action [LM21].
Our most refined understanding of full groups is achieved for free actions of , which are known as flows. All the results we described so far are valid for all compatible norms on the acting group. When it comes to the actions of , however, we consider only the standard Euclidean norm on it. Just like the actions of , flows give rise to an important homomorphism, known as the index map. Assuming the flow is ergodic, the index map can be described most easily as , where is the cocycle of . Chapter 6 is devoted to the analysis of the index map for general -flows.
The most technically challenging result of our work is summarized in Theorem 10.1, which identifies the derived full group of a flow with the kernel of the index map, and describes the abelianization of .
Theorem 4.
Let be a measure-preserving flow on . The kernel of the index map is equal to the derived full group of the flow, and the topological abelianization of is .
Theorem 4 parallels the known results for -actions from [LM18]. The structure of its proof, however, has an important difference. We rely crucially on the fact that each element of the full group acts in a measure-preserving manner on each orbit. This allows us to use Hopf’s decomposition (described in Appendix B) in order to separate any given element into two parts — recurrent and dissipative. If the acting group were discrete, the recurrent part would reduce to periodic orbits only. This is not at all the case for non-discrete groups, hence we need a new machinery to understand non-periodic recurrent transformations. To cope with this, we introduce the concept of an intermitted transformation, which plays the central role in Chapter 8, and which we hope will find other applications.
Theorems 3 and 4 can be combined to obtain estimates for the topological rank of the whole full groups of flows, which is the content of Proposition 10.3.
Theorem 5.
Let be a free measure-preserving flow on a standard probability space . The topological rank is finite if and only if the flow has finitely many ergodic components. Moreover, if has exactly ergodic components then
In particular, the topological rank of the full group of an ergodic flow is equal to either , or . We conjecture that it is always equal to , and more generally that the topological rank of the full group of any measure-preserving flow is equal to where is the number of ergodic components.
Our work connects to the notion of orbit equivalence, an intermediate notion between orbit equivalence and conjugacy. It can be traced back to the work of R. M. Belinskaja[Bel68] but recently attracted more attention. Stated in our framework, two flows are orbit equivalent if they can be conjugated so that the first flow is contained in the full group of the second and vice versa. A symmetric version of Belinskaja’s theorem is that ergodic -actions are orbit equivalent if and only if they are flip conjugate. It is very natural to wonder whether this amazing result has a version for flows. Our Theorem 10.14 implies the following.
Theorem 6.
If two measure-preserving ergodic flows are orbit equivalent, then they admit some cross-sections whose induced transformations111We refer the reader to Definition 10.11 and the paragraph that follows it for details on the measure-preserving transformation one associates to a cross-section. are flip-conjugate.
We do not know whether the above result is optimal, that is, whether having flip-conjugate cross-sections implies orbit equivalence, but it seems unlikely. It is tempting to think that the correct analogue of Belinskaja’s theorem would be a positive answer to the following question.
Question 1.1.
Let and be free ergodic measure-preserving flows which are orbit equivalent. Is it true that there is such that and are isomorphic, where denotes the multiplication by ?
Let us also mention that Theorem 6 implies that there are uncountably many full groups of ergodic free measure-preserving flows up to (topological) group isomorphism (see Corollary 10.16 and the paragraph right after its proof).
Finally, we also investigate the coarse geometry of the full groups. We establish that the norm is maximal (in the sense of C. Rosendal[Ros22], see also Appendix A.2) on the derived subgroup of an full group of an aperiodic measure-preserving action of any locally compact amenable Polish group (Theorem 5.5). For the measure-preserving flows, the norm is, in fact, maximal on the whole full group (Theorem 10.18).
1.2. Preliminaries
1.2.1. Ergodic theory
Our work belongs to the field of ergodic theory, which means that all the constructions are defined and results are proven up to null sets. On a number of occasions, we allow ourselves to deviate from the pedantic accuracy and write “for all …” when we really ought to say “for almost all …”, etc. The only part where certain care needs to be exercised in this regard appears in Chapter 2. We define full groups for Borel measure-preserving actions of Polish normed groups, and we need a genuine action on the space for these to make sense just as in [CLM16]. Boolean actions (also called near actions) of general Polish groups do not admit realizations in general [GTW05], and even when they do, it could happen that different realizations yield different full groups. This subtlety disappears once we move our attention to locally compact group actions, which is the case for Chapter 4 and onwards. All Boolean actions of locally compact Polish groups admit Borel realizations which are all conjugate up to measure zero (and hence have the same full group), so null sets can be neglected just as they always are in ergodic theory.
By a standard probability space we mean the unique (up to isomorphism) separable atomless measure space with , i.e., the unit interval with the Lebesgue measure. A few times in Chapter 5 and Appendix C.1 we refer to a standard Lebesgue space, by which we mean a separable finite measure space, , thus in contrast to the notion of the standard probability space allowing atoms and omitting the normalization requirement. We denote by the group of all measure-preserving bijections of up to measure zero. This is a Polish group for the weak topology, defined by if and only if for all Borel, . The weak topology is a Polish group topology, see [Kec10, Sec. 1]. Given , its support is the set
A measure-preserving bijection is called periodic when almost all its orbits are finite. Periodicity implies the existence of a fundamental domain for , namely a measurable set which intersects every -orbit at exactly one point. Since the ambient measure is finite, the existence of a fundamental domain actually characterizes periodicity.
1.2.2. Orbit equivalence relations
Any group action induces the orbit equivalence relation , where two points are -equivalent whenever . We will usually write this equivalence relation simply as for brevity. For the actions generated by an automorphism , we denote the corresponding orbit equivalence relation by . For clarity, we may sometimes want to name a measure-preserving action as and write . Then for all we denote by the measure-preserving transformation of induced by the action of .
We encounter various equivalence relations throughout this monograph. An equivalence class of a point under the relation is denoted by and the saturation of a set is denoted by and is defined to be the union of -equivalence classes of the elements of : . In particular, is the orbit of under the action of . The reader may notice that the notation for a saturation resembles that for the full group of an action (see Chapter 2). Both notations are standard, and we hope that confusion will not arise, as it applies to objects of different nature — sets and actions, respectively.
1.2.3. Actions of locally compact groups
Consider a measure-preserving action of a locally compact Polish (equivalently, second-countable) group on a standard Lebesgue space . A complete section for the action is a measurable set that intersects almost every orbit, i.e., . A cross-section is a complete section such that for some non-empty neighborhood of the identity we have whenever are distinct. When the need to mention such a neighborhood explicitly arises, we say that is a -lacunary cross-section.
With any cross-section one associates a decomposition of the phase space known as the Voronoi tessellation. Slightly more generally, Appendix C.2 defines the concept of a tessellation over a cross-section, which corresponds to a set for which the fibers , , partition the phase space. Every tessellation gives rise to an equivalence relation , where points are deemed equivalent whenever they belong to the same fiber , and to the projection map that associates with each the unique which fiber the point belongs to, and is therefore defined by the condition for all .
When the action is free, each orbit can be identified with the acting group. Such a correspondence depends on the choice of the anchor point within the orbit, but suffices to transfer structures invariant under right translations from the group onto the orbits of the action. For instance, if the acting group is locally compact, then a right-invariant Haar measure can be pushed onto orbits by setting as discussed in Section 4.2. Freeness of the action gives rise to the cocycle map which is well-defined by the condition . Elements of the full group are characterized as measure-preserving transformations such that for all . With each one may therefore associate the map , also known as the cocycle map, and defined by . Both the context and the notation will clarify which cocycle map is being referred to.
All these concepts appear prominently in the chapters which deal with free measure-preserving flows, that is actions of on the standard probability space. We use the additive notation for such actions: . The group carries a natural linear order which is invariant under the group operation and can therefore be transferred onto orbits. More specifically, given a free measure-preserving flow we use the notation whenever and belong to the same orbit and for some . Every cross-section of a free flow intersects each orbit in a bi-infinite fashion — each has a unique successor and a unique predecessor in the order of the orbit. One therefore has a bijection , called the the first return map or the induced map, which sends to the next element of the cross-section within the same orbit. We also make use of the gap function that measures the lengths of intervals of the cross-section, i.e., .
There is also a canonical tessellation associated with a cross-section which partitions each orbit into intervals between adjacent points of and is given by . The associated equivalence relation is denoted simply by and groups points which belong to the same interval of the tessellation, . The -equivalence class of is equal to .
Often enough we need to restrict sets and functions to an -class. Since such a need arises very frequently, especially in Chapter 9, we adopt the following shorthand notations. Given a set and a point the intersection is denoted simply by . Likewise, stands for and corresponds to the Lebesgue measure of the set . Moreover, will usually be shortened to , when the tessellation is clear from the context.
Chapter 2 full groups of Polish group actions
We begin by defining the key notion of interest for our work, namely the full groups of measure-preserving Borel actions of Polish normed groups on a standard probability space. Admittedly, the overall focus will be on actions of locally compact groups, and flows in particular. Nonetheless, the concept of an full group can be introduced for actions of arbitrary Polish normed groups, and we therefore begin with this level of generality.
2.1. spaces with values in metric spaces
By a Polish metric space we mean a separable complete metric space.
Definition 2.1.
Let be a standard probability space, let be a Polish metric space, and let be a measurable function. We define the -pointed space as the metric space of measurable functions such that , equipped with the metric
which is finite by the triangle inequality using the function as the middle point.
Proposition 2.2.
Let be a standard probability space and be a Polish metric space. is a Polish metric space for any measurable function .
Proof.
The argument follows closely the classical proof that is a Polish metric space. To check completeness, let us pick a Cauchy sequence in . Without loss of generality we may assume that , . Consider the sets , . Chebyshev’s inequality shows that , whence . The Borel–Cantelli lemma implies that is pointwise Cauchy for almost every . Since is complete, the pointwise limit of exists, and we denote it by . Define functions by
and note that by Fatou’s lemma. Finally, we conclude that
where the last convergence follows from Lebesgue’s dominated convergence theorem.
To verify separability, pick a countable dense set and note that the subspace of maps taking values in is -dense (in fact, this subspace is dense in the much stronger sup metric). It then follows that the set of functions that take only finitely many values (all of which are elements of ) is still dense. Finally, one uses a dense countable subalgebra of the measure algebra on and further restricts this subspace to the functions that are measurable with respect to the chosen subalgebra. The resulting countable collection is dense in . ∎
The group of measure-preserving automorphisms has a natural action by composition on , i.e., . Note that every automorphism acts by an isometry.
Proposition 2.3.
Let be a standard probability space, be a Polish metric space, and be a measurable function. The action of on is continuous.
Proof.
The argument mirrors the one in [CLM16, Prop. 2.9(1)]. Given sequences and we need to show that . Since the action is by isometries,
It therefore suffices to show that for any and any convergent sequence of automorphisms one has as . The latter is enough to check for functions that take only finitely many values since those are dense in . Suppose is such a step function over a partition . Convergence implies for all , which easily yields . ∎
When is a Polish group, there is a natural choice of the function , namely the constant function , where is the identity element of the group. We therefore simplify the notation in this case and write , omitting the subscript .
Recall that a Polish normed group is a Polish group together with a compatible norm on it (see Appendix A.1). In particular, if is a Polish normed group, there is a canonical choice of a complete metric on , namely
The corresponding space is Polish by Proposition 2.2. Moreover, it is a Polish group under pointwise operations.
Proposition 2.4.
Let be a Polish normed group, and let be a Borel measure-preserving action on a standard probability space. The space is a Polish normed group under the pointwise operations, , , and the norm .
Proof.
The space can equivalently be defined as the collection of all measurable functions with finite norm, . Using the properties of the norm on ,
Hence, is closed under the group operations and is a group norm on it.
To show that group operations are continuous, it suffices to check that for any and any sequence , , converging to zero, , there is a subsequence such that as (see, for instance, [BO10, Thm 3.4 and Lem. 3.5]).
Since converges to in , we may pass to a subsequence such that for almost all . Let and note that for all
It remains to apply Lebesgue’s dominated convergence theorem to the sequence , , concluding that . ∎
2.2. full groups of Polish normed group actions
Let be a Polish normed group, and let be a measure-preserving Borel action on a standard probability space . Let also denote the equivalence relation induced by this action, namely
The norm induces a metric on each -equivalence class via
(2.1) |
Properties of the metric are straightforward except, possibly, for the implication . To justify the latter, let , , be a sequence such that and . Elements , , belong to the stabilizer of . By Miller’s theorem [Mil77], stabilizers of all points are closed, whence fixes . Thus , and as claimed.
A. Carderi and the first named author introduced in [CLM16] orbit full groups of Borel measure-preserving Polish group actions on standard probability spaces, which we will simply call full groups. Given such an action , they define the full group of the action to consist of those measure-preserving transformations that preserve the equivalence classes of :
They showed that full groups are Polish with respect to the natural topology of convergence in measure.
Suppose that the acting group is furthermore endowed with a compatible norm, which therefore induces a metric on the equivalence classes of . We define a subgroup of that consists of those automorphisms for which the map is integrable. Such a subgroup, we argue in this section, also carries a natural Polish topology.
Definition 2.5.
Let be a Borel measure-preserving action of a Polish normed group on a standard probability space ; let be the associated metric on the orbits of the action. The norm of an automorphism is denoted by and is defined by the integral . In general, many elements of the full group will have an infinite norm, and the full group of the action consists of the automorphisms for which the norm is finite: .
Elements of form a group under the composition, as can readily be verified using the triangle inequality for and the fact that transformations are measure-preserving. Likewise, it is straightforward to check that is indeed a norm on . Our goal is to prove that the topology of the norm on is a Polish topology. Mimicking the approach taken in [CLM16], we provide a different definition of the full group, where Polishness of the topology will be readily obtainable, and then argue that the two constructions are isometrically isomorphic.
Remark 2.6.
The notion of full groups, discussed here, encompasses full groups from [CLM16], since the latter corresponds to the case when is equipped with a compatible bounded norm.
We recall some basic facts from [CLM16]. denotes the space of measurable functions ; this space is Polish with respect to the topology of convergence in measure. One can endow the with a Polish topology such that the evaluation map , given by , becomes continuous.
Remark 2.7.
In [CLM16], the possibility of making continuous is obtained by appealing to the remarkable but difficult result of H. Becker and A. S. Kechris, which states that every Borel -action has a continuous model [BK96, Thm. 5.2.1]. Let us point out that one can also derive this from the easier fact that every Borel -action can be Borel embedded into a continuous -action on a Polish compact space (see, for instance, [BK96, Thm. 2.6.6]), as we can endow the latter with the push-forward measure and work with it instead.
Let the set be the preimage of under :
Since is a subset of (see [CLM16, Prop. 2.9] and the remark after it), is in , hence Polish in the induced topology. The group operations can be pulled from onto (cf. [CLM16, p. 91]) as follows: for and define the multiplication via and the inverse111The symbol has already been used in the definition of the pointwise inverse on all of . We introduce a different operation here, hence the slightly unusual choice of the symbol to denote the inverse operation. by . These operations turn into a Polish group and into a continuous homomorphism.
The space admits a natural inclusion , which is continuous, as can be seen by noting that the equivalent metric on generates the convergence in measure topology on (see [CLM16, Prop. 2.7]), and for all . Set , which we endow with the topology induced form . Since is a subset of , we may omit the inclusion map when convenient.
Proposition 2.8.
is a Polish group with the multiplication and the inverse . The function is a compatible group norm on and is a continuous homomorphism.
Proof.
First of all, we need to show that these operations are well-defined in the sense that functions and belong to whenever so do their arguments. To this end observe that for
Now note that since is measure-preserving, we have
and therefore
In particular, , and thus is closed under the multiplication. Similarly, implies
Thus is also closed under taking inverses. Since these operations define a group structure on , it follows that is an (abstract) subgroup of . Note that we have also established that is a group norm on . The multiplication and the operation of taking the inverse are continuous in the topology of , which is a consequence of the continuity of coupled with Propositions 2.3 and 2.4. Since is a subset of , we conclude that it is a Polish group in the topology induced by the norm . ∎
Let denote the kernel of , and let denote the quotient norm induced by (see Proposition A.3 regarding the properties of the quotient norm). The factor group is evidently a Polish normed group, and it turns out to be isometrically isomorphic to the full group introduced in Definition 2.5 as we will now see. Let denote the homomorphism induced by onto the factor group.
Proposition 2.9.
The homomorphism establishes an isometric isomorphism between and .
Proof.
We begin by showing that holds for any . By the definition of the quotient norm,
For any fixed , we have , and therefore
This readily implies the inequality . For the other direction, let and consider the set
Using Jankov-von Neumann uniformization theorem, one may pick a measurable map that satisfies and for almost all . Since , we have
As is an arbitrary positive real, we conclude that .
It remains to check that is surjective. For an automorphism , consider the set
Applying the Jankov-von Neumann uniformization theorem once again we get a map such that and . The latter inequality together with the assumption that easily imply that and thus is the preimage of under . ∎
Results discussed thus far can be summarized as follows.
Theorem 2.10.
Let be a Borel measure-preserving action of a Polish normed group on a standard probability space. The full group is a Polish normed group relative to the norm .
Remark 2.11.
When the acting group is finitely generated and equipped with the word length metric with respect to the finite generating set, it can be shown that the left-invariant metric induced by the norm on the full group is complete (see [LM18, Prop. 3.4 and 3.5] and the remark thereafter for a more general statement). Nevertheless, generally full groups do not admit compatible complete left-invariant metrics, i.e., they are not necessarily CLI groups. For instance, if is acting by rotation on the circle, the full group of the action is all of , which is not CLI.
Let us point out a possibility to generalize our framework. Given a standard probability space , consider an extended Borel metric on , i.e., a Borel metric that is allowed to take the value (Eq. (2.1) provides such an example). Note that the relation is an equivalence relation. One can now define the full group of in complete analogy with Definition 2.5 as the group of all whose norm is finite.
Question 2.12.
Suppose that restricts to a complete separable metric on each equivalence class , . Is the full group of Polish in the topology of the norm ?
2.3. full groups and quasi-metric structures
When viewed as a normed group, the full group depends on the choice of a compatible norm on . The topological structure on , however, depends only on the quasi-metric structure of the acting group. Recall that two norms and on a Polish group are quasi-isometric if there exists a constant such that for all ,
Lemma 2.13.
Let and be two quasi-isometric compatible norms on a Polish group , and let be a Borel measure-preserving action on a standard probability space. The full groups associated with the two norms are equal as topological groups.
Proof.
The quasi-isometry condition implies that a function satisfies if and only if . In particular, the full groups associated with these norms are equal as abstract groups.
Both topologies make the inclusion of into continuous by Proposition 2.8, and, in particular, the inclusion map is Borel. Since injective images of Borel sets by Borel maps are Borel (see, for example, [Kec95, Thm. 15.1]), it follows that both topologies induce the same Borel structure on , which also coincides with the one induced by the weak topology on . A standard automatic continuity result (originally due to S. Banach [Ban32, Thm. 4 p. 23]) then yields equality of the two topologies (see also the second paragraph following [BK96, Lem. 1.2.6]). ∎
When a Polish group admits a canonical choice of the quasi-metric structure, full groups are unambiguously defined as topological groups without the need to choose any particular norm on . This is the case for boundedly generated Polish groups—the class of groups identified and studied by C. Rosendal in his treatise [Ros22]. Appendix A.2 provides a succinct review of the concept of maximal norms on boundedly generated Polish groups.
An example of this situation is given by , where the usual Euclidean norm is maximal in the sense of Definition A.5.
Remark 2.14.
We will see in the last chapter that the natural norm on the full groups of -actions is maximal so that it defines a quasi-metric structure which is canonically associated with the topological group structure.
2.4. Embedding isometrically in full groups
We now show a general result on the geometry of full groups endowed with the norm , which says that they are quite big.
Given a -finite measured space , denote by the space of all finite measure subsets identified up to measure zero and endowed with the metric .
Proposition 2.15.
Let be a Polish normed group acting by measure-preserving transformations on a standard probability space . If
then the metric space embeds isometrically into the full group of endowed with its metric, and hence so does .
Proof.
Since is a full group, any of its elements can be written as a product of three involutions belonging to by [Ryz85]. By assumption, so there must be an involution which does not belong to . Denote by the -algebra on consisting of -invariant sets, endowed with the measure given by . Since , the measure is -finite. Also, is non-atomic, because so is , and infinite, because . There is only one -finite standard atomless infinite measured space up to isomorphism (namely ) so we conclude that is isometric to . Composing this isometry with , we get the desired isometric embedding .
Finally, we observe that can be embedded into by taking a function to its epigraph, namely the set of all such that or . Since there is again only one infinite -finite standard atomless measured space and is such a space, we get an isometric embedding as wanted. ∎
Remark 2.16.
Full groups of actions of Polish groups are always coarsely bounded. In fact, they are coarsely bounded even as discrete groups222Being coarsely bounded as a discrete group is also called the Bergman property., which is a result due to M. Droste, W. C. Holland and G. Ulbrich [DHU08] (see also [Mil04, Section I.8] for a more general statement which encompasses the non-ergodic case). In particular, the above result is actually a sharp dichotomy: every full group of a Polish normed group action is either coarsely bounded, or it contains an isometric copy of .
2.5. Stability under the first return map
Some of the basic properties of full groups are discussed—in the wider generality of induction friendly finitely full groups—in Chapter 3. The often-used fundamental fact is the closure of full groups under taking the induced maps, which is a generalization of [LM18, Prop. 3.6]. We formulate this in Proposition 2.18.
Let be a measure-preserving transformation. Recall that for a measurable subset , the induced map is supported on and is defined to be for where is the smallest integer such that . By the Poincaré recurrence theorem, such a map yields a well-defined measure-preserving transformation.
Proposition 2.18.
Let be a Borel measure-preserving action of a Polish normed group . For any element and any measurable set , the induced transformation belongs to and moreover
Proof.
For , let be the set of elements of whose return time is equal to ; note that . Let as before be the metric induced by the group norm on the orbits of the action. To estimate the value of , observe that
Using the triangle inequality, we get
Thus and as claimed. ∎
Chapter 3 Polish finitely full groups
The main object of our investigation in this work are full groups of Borel measure-preserving actions of Polish normed groups. Some results, however, are valid in the more general context of what we call Polish finitely full groups. It encompasses full groups and allows us to put some of the proofs on topological simplicity and on maximal norms from [LM18, LM21] in a unified and broadened context.
Starting with a Polish finitely full group as defined in Section 3.1, we construct in Section 3.2 a natural closed subgroup of the latter which we call the symmetric subgroup, analogous to V. Nekrashevych’s symmetric and alternating topological full groups [Nek19]. We show that this closed subgroup coincides with the closure of the derived group under a mild hypothesis, satisfied by full groups, which we call induction friendliness. Section 3.3 is devoted to the study of closed normal subgroups of the symmetric subgroup: we show that they correspond to invariant sets, a fact which easily yields topological simplicity when the ambient Polish finitely full group is ergodic. Finally, in Section 3.4 we provide a condition normed induction friendly Polish finitely full groups which guarantees maximality on the symmetric subgroup in the sense of C. Rosendal (a brief reminder of the relevant notions is given in Appendix A.2).
3.1. Polish full and finitely full groups
H. Dye defined a subgroup as being full when it is stable under the cutting and pasting of its elements along a countable partition: given any partition of and any sequence such that the family also partitions , the element obtained as the reunion over of the restrictions belongs to . In particular, the group itself is full.
Given any , the group obtained by cutting and pasting elements of along countable partitions is the smallest full subgroup containing . We denote it by and call it the full group generated by .
Recall that the uniform topology on is the topology induced by the uniform metric defined by
The following can essentially be traced back to H. Dye [Dye59, Lem. 5.4].
Proposition 3.1.
The metric is complete on any full group , and it is separable if and only if the full group is generated by a countable group.
Proof.
Suppose that is a Cauchy sequence in the full group . Taking a subsequence, we may assume that for all . By the Borel-Cantelli lemma, for almost every there is some such that for all . Let for such , and note that is a measure-preserving bijection111 This also follows from the fact due to P. Halmos [Hal17] that is -complete. and . By construction, is obtained by cutting and pasting the elements of along a countable partition so , since is full.
Suppose is separable and let be a countable dense subgroup. The group is a countably generated full group which is dense in , so by completeness. The converse is obtained by noting that if generates , then one can view as the full group of the equivalence relation generated by a realization of the action of on , which is -separable by [Kec10, Prop. 3.2]. ∎
The full groups that we are considering are not full in the sense of H. Dye unless the norm on the acting Polish group is bounded, a case which was considered earlier in [CLM16]. They nevertheless satisfy the following weaker property.
Definition 3.2.
A group of measure-preserving transformations is finitely full if for any partition and such that the sets also partition , the element obtained as the reunion over of the restrictions belongs to .
We have the following useful relationship between fullness and finite fullness.
Proposition 3.3.
The -closure of any finitely full group is equal to the full group generated by . Moreover, every element is a limit of elements of whose support is contained in the support of .
Proof.
Since full groups are -closed and using the definition of fullness, it suffices to show that every element is a limit of elements of that belong to the full group generated by .
Since every is a product of three involutions in 222In fact, we only need the much easier fact that every element is a limit of products of two involutions from its full group, which follows by combining Theorem 3.3 and Sublemma 4.3 from [Kec10]. [Ryz85], it suffices to show that every involution in is a limit of elements of whose support is contained in the support of that involution. Let be such an involution, let be a partition of , and let in be such that for all . Pick a fundamental domain for , i.e., and . If , then for all , and, since is an involution, for all . Let
Clearly , since is finitely full. Furthermore, uniformly and by construction, which finishes the proof. ∎
Consider a finitely full group which is a Borel subset of and therefore inherits the structure of a standard Borel space. If is Polishable, i.e., if it admits a Polish group topology compatible with the Borel structure, then such topology is necessarily unique and must refine the weak topology inherited from (standard automatic continuity results can be found, for instance, in [BK96, Sec. 1.6]). We refer to such Polishable groups endowed with their unique Polish group topology refining the weak topology as Polish finitely full groups. In this monograph, our motivating example for introducing this class is of course full groups.
For any subgroup , there is the smallest finitely full group containing . Note that if is a finite group, then the finitely full group it generates coincides with the full group it generates. This, in particular, applies to the group generated by a periodic transformation with bounded periods.
Proposition 3.4.
Suppose is a Polish finitely full group, and is a periodic transformation with bounded periods. The topology induced by on the full group of is equal to the uniform topology.
Proof.
The weak and the uniform topologies on coincide since is periodic. We already mentioned that the topology of refines the weak topology. Since is Polish in the uniform topology, by the automatic continuity [BK96, Thm. 1.2.6], the topology induced by on the full group of is refined by the uniform topology. Hence the uniform topology and the topology induced from onto must coincide. ∎
We conclude this preliminary discussion with a definition of aperiodicity which applies to arbitrary subgroups of . Such a notion was already worked out by H. Dye [Dye59, Sec. 2] when he introduced type II subgroups. An equivalent version which suffices for our purposes is as follows.
Definition 3.5.
A subgroup is aperiodic it it admits a countable weakly dense subgroup whose action on has no finite orbits.
It can be checked that for an aperiodic , every countable weakly dense subgroup has infinite orbits almost surely. Further discussion of aperiodicity can be found in Appendix D.4.
3.2. Derived subgroup and symmetric subgroup
Our goal in this section is to identify when the closed derived subgroup of a Polish finitely full group is topologically generated by involutions. We start by noting that aperiodic finitely full groups admit many involutions in the sense of [Fre04, p. 384]:
Lemma 3.6.
Let be a finitely full aperiodic group. For every measurable nontrivial , there is a nontrivial involution whose support is contained in .
Proof.
The first and the second items of the following definition constitute analogues of V. Nekrashevych’s symmetric and alternating topological full groups [Nek19], respectively. In the setup of full groups, however, these notions coincide, as we will see shortly.
Definition 3.7.
Given a Polish finitely full group , we let
-
•
be the closed subgroup of generated by involutions, which we call the symmetric subgroup of .
-
•
be the closed subgroup of generated by -cycles, i.e., generated by periodic transformations whose non-trivial orbits have size .
-
•
be the closed subgroup generated by commutators (also known as the topological derived subgroup).
All these groups are closed normal subgroups of , and because every -cycle is a commutator of two involutions from its full group.
Proposition 3.8.
for any aperiodic finitely full group .
Proof.
We need to show that every involution is a limit of products of -cycles. Let be an involution, and let denote its fundamental domain; thus . By Lemma D.13, one can find an involution whose support is equal to . Since is finitely full, we may write as an increasing union , , where each is -invariant, and for every the transformation induced by on belongs to the group itself. Let denote the restriction of onto and note that in the uniform topology, and hence also in the topology of by Proposition 3.4. Our plan is to use the following permutation identity
(3.1) |
where corresponds to , to , and corresponds to . To this end, let be a fundamental domain for , put (which corresponds to the involution ), and, at last, set (corresponding to ). Figure 3.1 illustrates the relations between these sets and transformations.
Equation (3.1) translates into , so is a product of two -cycles, hence it belongs to . Since by construction , we conclude that . ∎
We do not know whether holds for all finitely full groups, but here is a convenient sufficient condition.
Definition 3.9.
A Polish finitely full group is called induction friendly if it is stable under taking induced transformations and, furthermore, whenever and is an increasing sequence of -invariant sets such that , then .
In the above definition, we require stability under taking the induced transformations and so always belongs to . However, for -invariant , is already a consequence of being finitely full.
Observe that full groups of measure-preserving actions of Polish normed groups are finitely full and also induction friendly. Indeed, finite fullness follows from a straightforward computation, while induction friendliness is a direct consequence of Proposition 2.18 and Lebesgue dominated convergence theorem.
Lemma 3.10.
In an induction friendly Polish finitely full group , every periodic element belongs to .
Proof.
Suppose is periodic. For , let be the set of whose -orbit has cardinality at most . Each is -invariant and . Moreover, is periodic, so it can be written as a product of two involutions from its full group, and since is finitely full and the periods of are bounded, these two involutions belong to . The conclusion follows from induction friendliness and convergence . ∎
Lemma 3.11.
Let be an induction friendly Polish finitely full group, and be the aperiodic part of , i.e.,
For any such that one has .
Proof.
Since , the transformation is periodic and therefore belongs to by Lemma 3.10. Hence
Remark 3.12.
Usefulness of the above lemma stems from the following simple observation. If satisfy and , then if and only if . In particular, for as in Lemma 3.11, whenever . This fact is used in the proof of the next lemma.
Lemma 3.13.
Suppose is an induction friendly Polish finitely full group. If are aperiodic on their supports, then .
Proof.
Let be a cross-section for the restriction of onto . In other words, is a measurable set satisfying . The induced transformation commutes with , since their supports are disjoint. We would be done if . Indeed, in this case , by Lemma 3.11 and is trivial, hence .
Motivated by this observation, we argue as follows. Pick a vanishing nested sequence of cross-sections for , i.e., , for all , and (see also Lemma D.11). Such a sequence of cross-sections exists since is assumed to be aperiodic on its support. Define inductively sets , , by setting , and letting be the part of that does not belong to the -saturation of any , ,
By construction, saturations under of the sets are pairwise disjoint, and the saturation of their union is the whole space, , because sets vanish.
Let , , and note that , and by the induction friendliness of . By construction, transformations and have disjoint supports for each and, therefore, commute. Since all sets are cross-sections for , one has by Lemma 3.11 and Remark 3.12. Taking the limit as , this yields . Finally, the -saturation of is all of , we use Lemma 3.11 and Remark 3.12 once again to conclude that , as claimed. ∎
Proposition 3.14.
If is an aperiodic induction friendly Polish finitely full group, then .
Proof.
Inclusion holds for any Polish finitely full group and Proposition 3.8 gives . We therefore concentrate on proving the reverse inclusion: given , we need to check that . Let and be the aperiodic parts of and respectively, so that , by Lemma 3.11. By construction, and are aperiodic on their supports and therefore by Lemma 3.13. It remains to use Remark 3.12 to conclude that necessarily , as needed. ∎
Corollary 3.15.
Let be a Polish normed group, and let be an aperiodic Borel measure-preserving action on a standard probability space . The three subgroups of introduced in Definition 3.7 coincide:
Moreover, they are all equal to the closure of the group generated by periodic elements of .
3.3. Topological simplicity of the symmetric group
We now move on to showing that symmetric subgroups of ergodic Polish finitely full groups are always topologically simple. Our argument abstracts from [LM18, Sec. 3.4]. In particular, we rely on conditional measures associated with subgroups of , whose construction and basic properties are recalled in Appendix D.
Lemma 3.16.
Let be an aperiodic Polish finitely full group, let be two involutions whose supports are disjoint and have the same -conditional measure. Then and are approximately conjugate in , i.e., there are such that .
Proof.
Let (resp. ) be a fundamental domain of the restriction of (resp. ) to its support. Then and there is an involution such that .
Since is finitely full, there is an increasing sequence of subsets of such that the involutions induced by on belong to , and . Let and define involutions which almost conjugate to as follows. For , let
For all and all , an easy calculation yields that:
-
•
if , then ;
-
•
if , then ;
-
•
and in all other cases.
In particular, and Proposition 3.4, applied to the full group of the involution (which contains both and ), guarantees that . ∎
Lemma 3.17.
Let be an aperiodic Polish finitely full group, let be an involution, and let be a -invariant subset contained in . Suppose that there exists an involution such that is disjoint from . Then for all -invariant functions , there is an involution such that is an involution whose support has -conditional measure .
Proof.
Let be a fundamental domain for the restriction of to and note that . By Maharam’s lemma (Theorem D.12), there is such that . The set is -invariant and satisfies . Consider the involution defined by
A straightforward computation shows that is an involution which coincides with on , with on , and is trivial elsewhere. Hence the support of is equal to , and has -conditional measure . ∎
Proposition 3.18.
Let be an aperiodic Polish finitely full group, let , and let denote the -saturation of . The closed subgroup of generated by the -conjugates of contains .
Proof.
Let the closed subgroup of generated by the -conjugates of be denoted by . We can find whose -translates cover and which satisfies . Since -translates of cover , we conclude that the -translates of cover , and so for all . By Maharam’s lemma (Theorem D.12), we can find whose -conditional measure is everywhere less than , and is strictly positive on . Let and take to be an involution such that is disjoint from .
Let be an involution such that . Using the facts that is finitely full, that and that , one can find an increasing sequence of -invariant subsets of such that and for each both and . Transformations belong to , and are, in fact, involutions whose support is equal to and has conditional measure at most . Let us define for brevity
For every , let denote the -saturation of . Note that and the union is increasing. Every involution supported on is thus the uniform limit of the involutions it induces on ’s. By Proposition 3.4, it therefore suffices to show that contains all the involutions which are supported on some .
Let be an involution supported on some . Let be a fundamental domain for the restriction of to its support. Using Maharam’s lemma repeatedly, we can partition into a countable family such that
(3.2) |
If we let , the sequence forms a partition of into -invariant sets. In particular, in the uniform topology and therefore in the topology of as well by Proposition 3.4. Moreover, the support of has -conditional measure at most by Eq. (3.2). The set is disjoint from by construction. Lemma 3.17 applies and provides an involution in whose support has the same conditional measure as that of . Lemma 3.16 shows that each belongs to and therefore also , as needed. ∎
Theorem 3.19.
Let be an aperiodic Polish finitely full group. For any closed normal subgroup there is a unique -invariant set such that .
Proof.
First, observe that for -invariant and , any involution supported in decomposes into the product of one involution supported in , and one supported in . It follows that the closed group generated by is equal to . Also, by Proposition 3.4, whenever is an increasing sequence of -invariant sets, one has
The set is thus directed and is closed under the countable unions. It therefore admits a unique maximum element, which is the set we seek. Indeed, , and the reverse inclusion is a direct consequence of Proposition 3.18.
It remains to argue that the set satisfying is unique. Suppose towards a contradiction that for . By symmetry, we may assume that . Lemma 3.6 provides an involution whose support is nontrivial and is contained in , thus but , contradicting . ∎
Corollary 3.20.
Let be an aperiodic Polish finitely full group. The group is topologically simple if and only if is ergodic.
Proof.
Specifying the corollary above to full groups and using Corollary 3.15, we obtain the following result.
Corollary 3.21.
Let be a Polish normed group, and let be an aperiodic Borel measure-preserving action on a standard probability space . The topological derived subgroup of the full group of the action is topologically simple if and only if the action is ergodic.
3.4. Maximal norms on the derived subgroup
The purpose of this section is to establish sufficient conditions for a norm on the derived subgroup of an induction friendly Polish finitely full group to be maximal in the sense of Section 2.3. Our argument follows closely the one given in [LM21, Sec. 6.2] for amenable graphings. The main application of Proposition 3.24 will be given in Theorem 5.5, but we hope that the setup of this section can be useful in other contexts, such as -integrable full groups [CJMT22].
Definition 3.22.
A norm on a subgroup is additive if for all with disjoint supports.
The following lemma parallels [LM21, Lem. 6.4] and is the key to showing that the norm on the derived subgroup is both coarsely proper and large-scale geodesic.
Lemma 3.23.
Let be a finitely full Polish group, and suppose that is a compatible additive norm on . For any periodic with bounded periods and for every , there are periodic elements such that
Proof.
Let and be a fundamental domain for . We may identify with the interval endowed with the Lebesgue measure. Put , , and let be the -saturation of . Note that for all since is -invariant and is finitely full, and that is continuous.
The map is thus continuous with respect to the uniform topology on , and therefore also with respect to the topology of by Proposition 3.4. Whence the function given by is also continuous.
We have and , so the intermediate value theorem yields existence of reals such that for all . Set for . By construction, each is -invariant and . Putting , we get . Finally for each the equality and additivity of the norm gives
hence for all , as needed. ∎
Proposition 3.24.
Let be an induction friendly Polish finitely full group and let be a compatible additive norm on it. If the set of periodic elements is dense in , then is a maximal norm on .
Proof.
In view of Proposition A.10, it suffices to show that is both large-scale geodesic (see Definition A.8) and coarsely proper (see Definition A.9). Note that induction friendliness yields density in of periodic automorphisms with bounded periods.
To see that is large-scale geodesic (with constant ), let us take a non-trivial and pick a periodic with bounded periods such that . Note that
(3.3) |
Fix large enough to ensure . By Lemma 3.23, we may decompose into a product of elements each of norm at most . Therefore
where and each of , , has norm at most and, in view of Eq. (3.3),
thus concluding the proof that is large-scale geodesic.
We now show that is coarsely proper. Fix and . Let be so large that . Then every element of norm at most is a product of elements of norm at most , namely one element of norm at most , where is periodic with bounded periods as provided by density, and , where each has norm at most as per Lemma 3.23. Thus is both coarsely proper and large-scale geodesic, and hence is maximal by Proposition A.10. ∎
Remark 3.25.
We do not have an example of an induction friendly Polish finitely full group such that the periodic elements are not dense in . We suspect that such groups do exist, for instance when is the full group of a free action of the free group on generators.
Chapter 4 Full groups of locally compact group actions
In this chapter, we narrow down the generality of the narrative and focus on actions of locally compact Polish groups, or equivalently, of locally compact second-countable groups. Such restrictions enlarge our toolbox in a number of ways. For instance, all locally compact Polish group actions admit cross-sections to which the so-called Voronoi tessellations can be associated. We use this to show in Section 4.1 a natural density result for subsets of full groups defined from dense subsets of the acting group (Theorem 4.2 and Corollary 4.3). For reader’s convenience, Appendix C.2 contains a concise reminder of the needed facts about tessellations.
Another key property of free111Motivated by our focus on -flows, this monograph primarily concentrates on free actions. We note, however, that each orbit of a Borel action of a locally compact Polish group is a homogeneous space, since point stabilizers are necessarily closed. In particular, orbits can be endowed with the Haar measure even without the freeness assumption. actions of locally compact groups is the existence of a Haar measure on each individual orbit. As we discuss in Section 4.2, elements of the full group act by non-singular transformations and, in particular, admit the Hopf decomposition (see Appendix B). Section 4.3 explains how these orbitwise decompositions can be understood globally, yielding a natural generalization of the periodic/aperiodic partition for elements of the full group of a measure-preserving action of a discrete group. The periodic part in the later case corresponds to the conservative piece of the Hopf decomposition, which generally exhibits a much more complicated dynamical behavior. We will get back to this in Chapters 7 and 8.
In the final Section 4.4, we connect full groups to the notion of orbit equivalence for actions of locally compact compactly generated Polish groups.
4.1. Dense subgroups in full groups
Our goal in this section is to prove that any element of the full group can be approximated arbitrarily well by an automorphism that piecewise acts by elements of a given dense subset of .
Definition 4.1.
A measure-preserving transformation between two measurable sets is said to be -decomposable, where , if there exist a measurable partition and elements such that for all .
The property of being -decomposable is similar to being an element of the full group generated by except that we do not require the transformation to be defined on all of .
Theorem 4.2.
Let be a measure-preserving action of a locally compact Polish group, let be a compatible norm on with the associated metric on the orbits , and let be a dense set. For any and any there exists an -decomposable transformation such that .
Theorem 4.2 establishes density of -decomposable transformations in the very strong uniform topology given by . In particular, it pertains to the topology.
Corollary 4.3.
Let be a measure-preserving action of a locally compact Polish group, let be a compatible norm on , and let be a dense subgroup. The full group is dense in .
Remark 4.4.
Theorem 4.2 is an improvement upon the conclusion of [CLM18, Thm. 2.1], which shows that is dense in whenever is a dense subgroup of . While the proof, which we present below, establishes density in a much stronger topology through more elementary means, we note that, as already mentioned in [CLM18, Thm. 2.3], their methods apply to all suitable (in the sense of [Bec13]; see also Definition 4.7) actions of Polish groups, whereas our approach here crucially uses local compactness of the acting group to guarantee existence of various cross-sections.
Let be a cross-section for a measure-preserving action and let be a tessellation over (in the sense of Appendix C.2). Let also be the push-forward measure on the cross-section and be the disintegration of over (see Appendix C.1 and Theorem C.1, specifically). Without loss of generality, we assume, whenever convenient, that the set in the statement of Theorem 4.2 is countable.
Definition 4.5.
Two Borel sets are said to be
-
•
-proportionate if the equivalence holds for -almost all ;
-
•
-equimeasurable if for -almost all .
For the context of Lemmas 4.6 through 4.10, we let denote an open symmetric neighborhood of the identity of , and stands for an -lacunary tessellation.
Lemma 4.6.
If are -proportionate Borel sets then
Proof.
By the defining property of the disintegration,
and so we need to check that for -almost all . Since and are -proportionate, it suffices to show that whenever . For any satisfying the latter, one necessarily has (because and is -lacunary, by assumption), and thus . In particular, . It remains to use the inclusion , which together with -lacunarity of , guarantees that
For the proof of the next lemma, we need the notion of a suitable action, introduced by H. Becker [Bec13, Def. 1.2.7].
Definition 4.7.
A measure-preserving Borel action of a Polish group is suitable if for all Borel sets one of the two options holds:
-
(1)
for any open neighborhood of the identity there exists such that ;
-
(2)
there exist Borel sets , such that and an open neighborhood of the identity such that .
All measure-preserving actions of locally compact Polish groups are known to be suitable (see [Bec13, Thm. 1.2.9]).
Lemma 4.8.
For all non-negligible -proportionate Borel sets , there exists an open set such that for all .
Proof.
Let , which is dense in , and put . We apply the dichotomy in the definition of a suitable action to the sets and show that item (2) cannot hold.
Indeed, suppose there exist , satisfying
and an open neighborhood of the identity such that . Set , where is an enumeration of , and note that and , simply because . Since is dense in , we have and thus . Lemma 4.6, applied to and , guarantees that , which is possible only when , contradicting the assumption that is non-negligible.
We are left with the alternative of the item (1), and so there has to exist some such that . Since , there exists such that . It remains to note that and that is an open condition on , since the homomorphism associated to the measure-preserving action of on is continuous (see for instance [CLM18, Lem. 1.2]). ∎
Lemma 4.9.
For any non-empty open and for any non-negligible Borel set , there exists such that
Proof.
Let be the Borel bijection and consider the push-forward measure , which for can be expressed as . Let be an enumeration of and set
We claim that . Indeed, for each the set of such that is non-empty and open, hence there is such that .
Finally, is non-negligible by assumption, i.e., , so there exists such that , which translates into the required
Lemma 4.10.
For all non-negligible -proportionate Borel sets , there exists such that
Proof.
The plan is to reduce the setup of this lemma to that of Lemma 4.8. Let be a symmetric neighborhood of the identity that is furthermore small enough to guarantee that is -lacunary. Apply Lemma 4.9 to find such that for
one has . Set , and note that and are non-negligible -proportionate sets. Moreover, by construction.
Repeat the same steps for and find such that for
we have . Set and . Once again, sets and are non-negligible, -proportionate and are both contained in .
We now apply Lemma 4.8 to sets and , viewed as a -lacunary tessellation, yielding an open such that for all . Note that since and is, in fact, -lacunary, the equality holds for all and . We conclude that for all and hence any satisfies the conclusion of the lemma. ∎
A measure-preserving map is -coherent if -almost surely one has .
Lemma 4.11.
For all -equimeasurable Borel sets , there exists a -coherent -decomposable measure-preserving bijection .
Proof.
Let be an enumeration of . Consider the set
and let . Note that the sets and are -equimeasurable, so we may continue in the same fashion and construct sets such that
We define by the condition for .
Sets and are -equimeasurable. If either one of them (and thus necessarily both of them) were non-negligible, Lemma 4.10 would yields an element that moves a portion of into , contradicting the construction. We conclude that
and is therefore as required. ∎
Lemma 4.12.
Suppose is a cocompact tessellation and let be -equimeasurable Borel sets. For any and any -coherent measure-preserving there exists is a -coherent -decomposable such that .
Proof.
Let be a -cocompact tessellation over some cross-section such that the diameter of each region in is less than . Suppose is -cocompact. By Lemma C.9, we can find a finite partition of such that each is -lacunary, which guarantees that, for each , every intersects at most one class , . For each set and . We re-enumerate sets and as a sequence , , and note that for all one has
Moreover, sets and are -equimeasurable, so Lemma 4.11 yields -coherent -decomposable measure-preserving maps . The transformation can now be defined by the condition whenever . It is easy to check that is as claimed. ∎
Proof of Theorem 4.2.
Fix a cocompact cross-section , and let be a nested and exhaustive sequence of compact neighborhoods of the identity in . For all , select based on Lemma C.9 a finite sequence of cocompact cross-sections such that each is -lacunary and . Re-enumerate cross-sections , , , into a sequence and let be the Voronoi tessellation over .
Let , and use Lemma 4.12 to find an -decomposable measure-preserving map that satisfies the inequality . Set
and observe that , , form a partition of . Find transformations by repeated applications of Lemma 4.12 applied to the tessellations . The element defined by satisfies the conclusion of the theorem. ∎
4.2. Orbital transformations
Let be a free measure-preserving action of a locally compact Polish group on a standard probability space. Fix a right-invariant Haar measure on . Any orbit can be identified with the group itself via the map , and can be pushed via this identification onto orbits resulting in a collection of measures on defined by . Right invariance of the measure ensures that depends only on the orbit and is independent of the choice of the base point, i.e., whenever .
This section focuses on two main facts: the so-called mass-transport principle, given in Eq. (4.1) below, and non-singularity of the transformations induced by elements of onto orbits of the action, formulated in Proposition 4.13. Both of these topics have been discussed in the literature in many related contexts including, for instance, [CLM18, Appen. A] and the treatise [ADR00]. We are, however, not aware of any specific reference from which Eq. (4.1) and Proposition 4.13 can be readily deduced. The following derivations are therefore included for reader’s convenience, with the disclaimer that these results are likely to be known to experts.
Freeness of the action allows us to identify the equivalence relation with via , . The push-forward of the product measure is denoted by and can equivalently be defined by
where and .
In general, the flip transformation , , is not -invariant. Set to be , which amounts to . Following the computation as in [CLM18, Prop. A.11], one can easily check that , where is the associated left-invariant measure, . If we define the measure on to be
then . In particular, is -invariant if and only if , i.e., is unimodular.
Let be the left Haar modulus function given for by . Recall that is a continuous homomorphism (see, for instance, [Nac65, Prop. 7]), measures and belong to the same measure class and for all (see [Nac65, p. 79]).
A function is -integrable if and only if is -integrable, which together with the expression for the Radon-Nikodym derivative and Fubini’s theorem yields the following identity:
(4.1) |
When the group is unimodular, this expression attains a very symmetric form and is known as the mass-transport principle:
(4.2) |
Any automorphism induces for each a transformation of the -finite measure space . In general, does not preserve , however, it is always non-singular, and the Radon-Nikodym derivative can be described explicitly. Note that the full group admits two natural actions on the equivalence relation : the left action is given by , and the right action is defined as . A straightforward verification (see [CLM18, Lem. A.9]) shows that is always -invariant. Since , for all we have
Let , i.e., . The equality is equivalent to = . The latter implies that each Borel and all measurable we have
Fubini’s theorem | |||
which is possible only if for -almost all . Passing to the measures on the orbits, this translates for each into . If is a countable algebra of Borel sets in that generates the whole Borel -algebra, then for each , is an algebra of Borel subsets of the orbit , which generates the Borel -algebra on it. We have established that for -almost all the two measures, and , coincide on each , , thus -almost surely .
Equality translates into and the Radon-Nikodym derivative can now be computed as follows.
preservers | |||
We summarize the content of this section into a proposition.
Proposition 4.13.
Let be a locally compact Polish group acting freely on a standard probability space . Let be a right Haar measure on , be the corresponding Haar modulus, and let be the family of measures obtained by pushing onto orbits via the action map. Each induces a non-singular transformation of for almost every , and moreover one has for all Borel sets . If is unimodular, then for -almost all .
For future reference, we isolate a simple lemma, which is an immediate consequence of Fubini’s theorem.
Lemma 4.14.
Let be a locally compact Polish group acting freely on a standard probability space . Let , , , and be as above. For any Borel set the following are equivalent:
-
(1)
;
-
(2)
for -almost all ;
-
(3)
for -almost all .
4.3. The Hopf decomposition of elements of the full group
Fix an element of the full group of a free measure-preserving action of a locally compact Polish group . As explained in Section 4.2, acts naturally in a non-singular manner on each -orbit. This action thus has a Hopf decomposition (see Appendix B). We will now explain how to understand globally this decomposition, obtaining a generalization of the fact that when is discrete, any element of the full group decomposes the space into a periodic and an aperiodic part.
Let us pick a cocompact cross-section and let be the associated Voronoi tessellation (see Appendix C.2). Set to be the projection map given by the condition for all . Define the dissipative and conservative sets as follows:
In plain words, the dissipative set consists of those points whose orbit has a finite intersection with the Voronoi region of . The conservative set , on the other hand, collects all the points whose orbit is bi-recurrent in the region. We argue in Proposition 4.16 that sets and induce the Hopf decomposition for for almost every ; in particular, is a partition of , which is independent of the choice of the cross-section .
Lemma 4.15.
Sets and partition the phase space: .
Proof.
Define sets and according to
and note that . To show that it is enough to verify that .
This is done by noting that these sets admit pairwise disjoint copies using piecewise translations by powers of . In view of the fact that is measure-preserving, this implies that and are null. To be more precise, set and define inductively , where is the smallest natural number such that . Note that is well-defined, for otherwise would belong to . Sets , , are pairwise disjoint, and have the same measure since is measure-preserving. We conclude that . The argument for is similar. ∎
Proposition 4.16 (Hopf decomposition).
Let be a free measure-preserving action of a locally compact Polish group on a standard probability space . Let be a right Haar measure on and be the push-forward of onto the orbits as described in Section 4.2. For any element , the measurable -invariant partition defined above satisfies that for -almost all the partition is the Hopf decomposition for on . Moreover, there is only one partition satisfying this property up to null sets.
Proof.
According to Proposition 4.13, we may assume that for all the map is a non-singular transformation with respect to and satisfies for all Borel .
Let , , denote the Hopf’s decomposition for . For any , the set
is a wandering set and therefore up to a null set. If satisfies , , then is finite, and therefore , whence also
Claim.
We have for each .
Proof of the claim.
Otherwise we can find and a wandering set of positive measure, . Construct a sequence of sets by setting and
where the value of is well-defined for each and , since all points in return to their Voronoi domain infinitely often. Define a transformation as , and note that for all one has . The region is precompact, since is cocompact, and therefore using continuity of the Haar modulus one can pick such that for all and all .
Since is composed of powers of , Proposition 4.13 ensures that
whence for each . We now arrive at a contradiction, as , , form a pairwise disjoint infinite family of subsets of whose measure is uniformly bounded away from zero by , which is impossible, since by cocompactness of . This finishes the proof of the claim. ∎
We have established by now that and, up to a null set, by the claim above. Finally, implies via Lemma 4.14 for -almost all , and therefore -almost surely. Sets and thus satisfy the conclusion of the proposition.
For the uniqueness part of the proposition, suppose and are two partitions of such that
for -almost all . One therefore also has , and hence by Lemma 4.14. ∎
We end this section with a natural definition which will be useful for analyzing elements of the full group.
Definition 4.17.
Let be a free measure-preserving action of a locally compact Polish group on a standard probability space , and let . Consider the -invariant partition provided by the Hopf decomposition of as per the previous proposition. We say that is dissipative when and that is conservative when .
When is discrete, observe that is dissipative if and only if it is aperiodic (all its orbits are infinite), and that is conservative if and only if it is periodic (all its orbits are finite).
Example 4.18.
Let us give a general example of dissipative elements of the full group. Let be a free measure-preserving action of a locally compact Polish group on a standard probability space . If generates a discrete infinite subgroup, then the element of the full group is dissipative. Indeed, the action of on each orbit is isomorphic to the -action by left translation on endowed with its right Haar measure, which is dissipative since it admits a Borel fundamental domain and has only infinite orbits. For instance, if , such a domain is given by the interval (or , if is negative).
In Chapter 7, we build an interesting example of a conservative element in the full group of any free measure-preserving flow: its action on each orbit is actually ergodic, and its cocycle is bounded.
4.4. full groups and orbit equivalence
We now restrict ourselves to the setup where the acting group is locally compact Polish and compactly generated, endowed with a maximal compatible norm (the existence of such a norm for locally compact Polish group is equivalent to being compactly generated, see [Ros21, Cor. 2.8 and Thm. 2.53]). For such a group, as explained in Section 2.3, it makes sense to talk about the associated full group by endowing the group with a maximal norm.
The following definition is the natural extension of the notion of orbit equivalence to the locally compact case, stated in terms of full groups.
Definition 4.19.
Let and be the respective measure-preserving actions of two locally compact Polish compactly generated groups and on a standard probability space . We say that and are orbit equivalent when there is a measure-preserving transformation such that for all and all ,
In other words, up to conjugating by , we have that the image of is contained in the full group of , and the image of is contained in the full group of .
We now show that full groups do remember actions up to orbit equivalence as abstract groups. This is done by finding a spacial realization of the isomorphism between the full groups. Such techniques originated in the work of H. Dye [Dye59] and have been greatly generalized by D. H. Fremlin [Fre04, 384D]. We recall that a subgroup of is said to have many involutions if for any non-trivial measurable there exists a non-trivial involution such that . The group of quasi-measure-preserving transformations of is denoted by .
Theorem 4.20 (Fremlin).
Let be subgroups of with many involutions. For any isomorphism there exists such that for all .
Proposition 4.21.
If two ergodic measure-preserving actions of locally compact compactly generated Polish groups have isomorphic full groups, then they are also orbit equivalent.
Proof.
Denote by and the two actions on the same standard probability space . Since the full groups of ergodic actions have many involutions (see, for example, Lemma 3.6), any isomorphism admits a spatial realization by some . The Radon–Nikodym derivative of with respect to is easily seen to be preserved by every element of , and hence must be constant by ergodicity. We conclude that , and therefore by the definition the actions and are orbit equivalent. ∎
Remark 4.22.
Similarly to the finitely generated case [LM21, Sec. 8.1], one could define full orbit equivalence between actions as equality of full groups up to conjugacy, which is a strengthening of orbit equivalence (indeed the latter only requires inclusion of each acting group in the full group of the other acting group). It would be interesting to have examples of actions which are orbit equivalent, but not fully orbit equivalent.
We end this section by showing that orbit equivalence is equivalent to a stronger definition where we ask that, up to conjugating by , we moreover have that, on a full measure set , the and orbits coincide. This will be a direct consequence of the following proposition. The proof of this proposition is the same as that of [CLM16, Prop. 3.8] which was not stated in the level of generality we need. Since it is short, we reproduce it here.
Proposition 4.23.
Let and be two locally compact Polish groups acting in a Borel measure-preserving manner on a standard probability space , denote by the -action and suppose that . Then there is a full measure Borel subset such that
Proof.
Let be the Haar measure on . Since , for all and almost all , we have . By Fubini’s theorem, this implies that the Borel set
has full measure. Now let , and let be such that . We want to show that .
Since and are in , the sets
have full measure and so has full measure. Take , and note that , so the two orbits and intersect, hence . ∎
Corollary 4.24.
Two measure-preserving actions of locally compact compact compactly generated Polish groups and on a standard probability space are orbit equivalent if and only if they can be conjugated so as to share the same orbits on a full measure Borel subset , and for all and there are Borel maps
such that for all ,
and finally, if we denote by and maximal norms on and respectively, then
Remark 4.25.
Note that both and are actually Borel globally (as maps and ) as a consequence of the Arsenin selection theorem for Borel sets with sections and the fact that point stabilizers are closed, a result of D. Miller.
Proof of Corollary 4.24.
It is clear from the definition of full groups that the conditions in the corollary are sufficient for orbit equivalence. Observe that up to conjugating the two actions, they do share the same full group. Since full groups contain the acting groups, we can apply Proposition 4.23 twice and get the desired full measure Borel subset restricted to which orbits coincide. The remaining statements are then direct consequences of the definition of full groups. ∎
We will see in the final chapter that there are free ergodic -flows which are not orbit equivalent. This will be done by relating orbit equivalence to flip-Kakutani equivalence. In the discrete amenable case, an important result of Austin shows that entropy is preserved by orbit equivalence [Aus16]. We wonder what happens in the general locally compact setup.
Question 4.26.
Let be an amenable non-discrete non-compact compactly generated locally compact Polish group. Are there two measure-preserving ergodic actions of which are not orbit equivalent?
Chapter 5 Derived full groups for locally compact amenable groups
Given a measure-preserving action of a normed Polish group on , the derived full group of the action is by definition the closure in of the group generated by commutators, i.e., elements of the form , where . Provided the -action is aperiodic, the latter can be described in three different ways using the fact that is induction friendly, as explained in the Section 3.2 (see Corollary 3.15):
-
•
is the closure of the group generated by involutions;
-
•
is the closure of the group generated by -cycles;
-
•
is the closure of the group generated by periodic elements.
In particular, all periodic elements of actually belong to .
Compared to the previous chapter, we impose one further restriction on the acting group, and consider actions of a locally compact amenable Polish normed group . Appendix G of [BdlHV08] contains an excellent review of amenability for locally compact Polish groups. As before, we fix a measure-preserving action on a standard probability space , and let denote the family of metrics induced onto the orbits by the norm.
In Section 5.1, we will first exhibit a dense increasing chain of subgroups in . This dense chain is used in the two remaining sections. In Section 5.2, we show that amenability of the group is reflected in whirly amenability of , while in Section 5.3 we prove by a Baire category argument that has a dense -generated subgroup.
5.1. Dense chain of subgroups
An equivalence relation is said to be uniformly bounded if there is and such that and , where .
Lemma 5.1.
Let be a locally compact amenable Polish normed group acting on a standard probability space . There exists a sequence of cross-sections , , and tessellations over such that for all
-
(1)
and (up to a null set);
-
(2)
is uniformly bounded.
Proof.
Let be a cocompact cross-section, be the Voronoi tessellation over , be the associated reduction, and be the push-forward measure on . Recall that is uniformly bounded, since is cocompact. Let be the equivalence relation obtained by restricting onto . By a theorem of A. Connes, J. Feldman, and B. Weiss [CFW81], is hyperfinite on an invariant set of -full measure. Throwing away a -invariant null set, we may write , where is an increasing sequence of Borel equivalence relations with finite classes. For , define to be the set of points in the cross-section whose -class is bounded in diameter by :
Note that the sets are -invariant, nested, and for every . Pick so large as to ensure and let . The sets are -invariant, increasing, and . Define equivalence relations on by setting whenever or and . Note that whenever . Let be a Borel transversal for and define . It is straightforward to check that each is a tessellation over , and equivalence relations satisfy the conclusions of the lemma. ∎
The equivalence relations produced by Lemma 5.1 give rise to a nested chain of groups . Some basic facts about such groups can be found in Appendix C.2. The following lemma establishes that such a chain is dense in the derived full group.
Lemma 5.2.
Let be a locally compact amenable Polish normed group acting on a standard probability space and let be a sequence of equivalence relations as in Lemma 5.1. If the action is aperiodic, then the union is contained in the derived full group and is dense therein.
Proof.
By definition, is a subgroup of . Since equivalence relations are uniformly bounded, we actually have , and the topology induced by the metric on coincides with the topology induced from . Moreover, in view of Proposition C.7, is topologically generated by periodic transformations, so we actually have as a consequence of Lemma 3.10 and Corollary 3.15.
It remains to verify that the union is dense in . To this end, recall that by Corollary 3.15 the derived full group is topologically generated by involutions. So let be an involution and set , . Note that is -invariant since is an involution. Moreover, as , and thus the induced transformations converge to in the topology of . We conclude that is dense in the derived full group. ∎
Corollary 5.3.
Let be a locally compact amenable Polish normed group acting on a standard probability space . Suppose that almost every orbit of the action is uncountable. There exists a chain of closed subgroups such that is dense in , and each is isomorphic to for some standard Lebesgue space . If moreover each orbit of the action has measure zero, then one can assume that all are atomless and each is isomorphic to .
Proof.
Corollary 5.4.
Let be a locally compact amenable Polish normed group acting on a standard probability space . If the action is aperiodic, then the set of periodic elements is dense in the derived full group .
Proof.
Consider a chain of subgroups given by Lemma 5.2. Periodic elements are dense in these groups for their natural topology (see Proposition C.7 and the discussion preceding it). These topologies are compatible with the standard Borel structure of induced by the weak topology and therefore must refine the topology by the standard automatic continuity arguments [BK96, Sec. 1.6]. Hence periodic elements are dense in all of , as claimed. ∎
Corollary 5.4 together with Proposition 3.24 show that the norm is maximal on derived full groups of aperiodic measure-preserving actions of locally compact amenable Polish normed groups (see Section 2.3 for a short reminder on maximality of norms). In particular, such groups are boundedly generated by [Ros22, Thm. 2.53].
Theorem 5.5.
Let be a locally compact amenable Polish normed group acting on a standard probability space . If the action is aperiodic, then the norm is maximal on the derived full group .
We do not know if the amenability hypothesis can be removed, even when is discrete and the action is free.
5.2. Whirly amenability
Lemma 5.2 is a powerful tool to deduce various dynamical properties of derived full groups. Recall that a Polish group is said to be whirly amenable if it is amenable and for any continuous action of on a compact space any invariant measure is supported on the set of fixed points of the action. In particular, each such action has to have some fixed points, so whirly amenable groups are extremely amenable.
Proposition 5.6.
Let be a smooth measurable equivalence relation on a standard Lebesgue space . If is atomless, then the full group is whirly amenable.
Proof.
In view of Proposition C.6, the full group is isomorphic to
where is the group of permutations of an -element set, and , . Since a product of whirly amenable groups is whirly amenable, it suffices to show that the groups appearing in the decomposition above, namely , , and , , are whirly amenable.
Remark 5.7.
The assumption of being atomless cannot be omitted in the proposition above. Indeed, will factor onto for some , as long as an -class contains at least atoms of of the same measure. However, if all -atoms within each -class have distinct measures, then the restriction of onto the atomic part of is trivial, which suffices to conclude the whirly amenability of the group .
Theorem 5.8.
Let be a measure-preserving action of an amenable locally compact Polish normed group on a standard probability space . If the action is aperiodic, then the derived full group is whirly amenable. In particular, is amenable.
Proof.
Lemma 5.2 shows that has an increasing dense chain of subgroups of the form , where are smooth measurable equivalence relations on . Proposition 5.6 applies and shows that groups are whirly amenable. The latter is sufficient to conclude whirly amenability of , as any invariant measure for the action of the derived group is also invariant for the induced actions, hence it has to be supported on the intersection of fixed points of all , which coincides with the set of fixed points for the action of .
The fact that is amenable now follows from the fact that every abelian group is amenable, and every amenable extension of an amenable group must itself be amenable (for instance, see [BdlHV08, Prop. G.2.2]). ∎
Remark 5.9.
Note that in general is not extremely amenable. For flows, it factors onto via the index map (see Chapter 6) and admits continuous actions on compact spaces without fixed points, so is not extremely amenable (and in particular, it is not whirly amenable) for any free measure-preserving flow.
Corollary 5.10.
Let be a free measure-preserving action of a unimodular locally compact Polish group on a standard probability space . The following are equivalent:
-
(1)
is amenable.
-
(2)
is amenable.
-
(3)
The derived full group is amenable.
-
(4)
The derived full group is extremely amenable.
-
(5)
The derived full group is whirly amenable.
Proof.
We established the implication (1)(5) in Theorem 5.8. The chain of implications (5)(4)(3) is straightforward, and (3)(2) follows from the stability of amenability under group extensions, which was already discussed in Theorem 5.8.
For the last implication (2) (1), first recall that the orbit full group of the action is generated by involutions. It follows that the orbit full group is topologically generated by involutions whose cocycles are integrable (actually, one can even ask that the cocycles are bounded). In particular, the full group is dense in the orbit full group, and so assuming (2) we conclude that the orbit full group is amenable. The amenability of then follows from [CLM18, Thm. 5.1]. ∎
Remark 5.11.
We have to require unimodularity in order to be able to apply [CLM18, Thm. 5.1]. It seems likely that the unimodularity hypothesis can be dropped in this result, but we do not pursue this question further.
5.3. Topological generators
We now concern ourselves with the question of determining the topological rank of derived full groups. Our approach will be based on the dense chain of subgroups established in Corollary 5.3, and the first step is to study the topological rank of the group .
Let and be standard Lebesgue spaces. Consider the product space equipped with the product measure and let be the product of the discrete equivalence relation on and the anti-discrete on ; in other words, if and only if . As discussed in Appendix C.1, the following two groups are one and the same:
-
(1)
the full group ;
-
(2)
the topological group .
Suppose that is atomless. Pick a hyperfinite ergodic equivalence relation on so that is dense in , where stands for the collection of aperiodic automorphisms of . Set to be the equivalence relation on given by whenever and . A standard application of the Jankov-von Neumann uniformization theorem yields the following lemma.
Lemma 5.12.
is dense in .
Our first goal is to establish that the topological rank of is . We do so by first verifying this under the assumption that is atomless, and then deducing the general case.
We say that a topological group is generically -generated, , if the set of -tuples that generate a dense subgroup of is dense in . Note that the set of such tuples is always a set, so if is generically -generated, then a comeager set of -tuples generates a dense subgroup of .
Proposition 5.13.
Suppose that is atomless. The group is generically -generated.
Proof.
Lemma 5.14.
For all topological groups and one has
If is generically -generated, then so are and as well.
Proof.
The inequality on ranks is immediate from the trivial observation that if is dense in , then is dense in and is dense in .
Suppose is generically -generated, pick an open set and note that naturally corresponds to an open subset of via the isomorphism . Since is generically -generated, there is a tuple that generates a dense subgroup and . We conclude that generates a dense subgroup of and the lemma follows. ∎
Lemma 5.15.
For any separable topological group
If is generically -generated for some , then so is the group .
Proof.
In view of Lemma 5.14, , and, since the group is separable, we only need to consider the case when the rank is finite.
It is notationally convenient to shrink the interval and work with the group instead as it can naturally be viewed as a closed subgroup of via the identification , where
Pick families dense in , and dense in .
Let us call a function a multi-index if for all but finitely many . We use to denote the set of all multi-indices. Given , let be an element of . Note that is dense in and thus is a dense family in .
Pick a tuple that generates a dense subgroup. For each pair , there exists a sequence of reduced words in the free group on generators such that converges to in measure. By passing to a subsequence, we may assume that pointwise almost surely. In other words, the set
has Lebesgue measure for each , and hence so does the set
Pick some , , and set
Elements naturally belong to , and we claim that they generate a dense subgroup therein, witnessing . To this end recall that pointwise almost surely. In particular,
in measure and, for each ,
is guaranteed by choosing . We conclude that
in , and therefore
Finally, suppose that is generically -generated. Choose open sets ,. Shrinking them if necessary, we may assume that all have the form , where is open in , and , , are open in .
Pick , , to consist of those functions satisfying and for all , . Note that .
Since is assumed to be generically -generated, there is a tuple generating a dense subgroup in such that for each . Running the above construction, we get a tuple such that , , whence is generically -generated. ∎
Lemma 5.15 remains valid if we take the product with a finite power of , which follows from Lemma 5.14.
Corollary 5.16.
For any separable topological group and any one has
If is generically -generated for some , then so is the group .
We may now strengthen Proposition 5.13 by dropping the assumption on being atomless.
Proposition 5.17.
Let be a standard Lebesgue space and be a standard probability space. The topological group is generically -generated.
Proof.
Proposition 5.18.
Let be a Polish group and let be a dense chain of Polish subgroups, . If each is generically -generated, then is generically -generated.
Proof.
We need to show that for any open and any open there is a tuple such that . Since groups are nested and is dense in , there is so large that and . It remains to use the fact that is generically -generated to find the required tuple. ∎
Theorem 5.19.
Let be a measure-preserving action of a locally compact amenable Polish normed group on a standard probability space . If almost every orbit of the action is uncountable, then the derived full group is generically -generated.
Proof.
Corollary 5.20.
Let be a measure-preserving action of a locally compact amenable Polish normed group on a standard probability space . If almost every orbit of the action is uncountable, then the derived full group has topological rank .
Proof.
The assumption for orbits to be uncountable is essential, and Corollary 5.20 is in a striking difference with the dynamical interpretation of the topological rank of derived full groups for actions of discrete groups. As shown in [LM21, Thm. 4.3], given an aperiodic measure-preserving action of a finitely generated group , the topological rank of is finite if and only if the action has finite Rokhlin entropy.
Chapter 6 The index map for full groups of flows
We now turn our attention to flows, i.e., measure-preserving actions of . Since the group of reals is locally compact, amenable, unimodular, and, of course, Polish, all of the results in the previous chapters apply to -flows. A much more in-depth understanding of full groups of flows is possible and is based on the existence of the so-called index map, which we define and investigate in this chapter. This map is a continuous homomorphism from the full group of the flow to the additive group of reals, which can be thought of measuring the average shift distance. When the flow is ergodic, such averages are the same across orbits. By taking the ergodic decomposition of the flow , we can adopt a slightly more general vantage point and view the index map as a homomorphism into the space of functions on the space of invariant measures , .
Understanding the kernel of the index map is the task of fundamental importance. We will subsequently identify with the derived topological subgroup of (Theorem 10.1). This will allow us to describe abelianizations of full groups of flows and estimate the number of their topological generators.
It has already been mentioned that any element of a full group of a flow induces Lebesgue measure-preserving transformations on orbits (Section 4.2). When furthermore belongs to the full group, these transformations are special—they leave “half-lines” invariant up to a set of finite measure. Such transformations form the so-called commensurating group. Let us therefore begin with a more formal treatment of this group, which has already appeared in the literature before, for instance in [RS98].
6.1. Self commensurating automorphisms of a subset
Consider an infinite measure space . We say that two measurable sets are commensurate if the measure of their symmetric difference is finite, . The relation of being commensurate is an equivalence relation, and all sets of finite measure fall into a single class. Note also that if and are both commensurate to some , then so is the intersection ; in other words, all equivalence classes of commensurability are closed under finite intersections.
Let denote the collection of all measurable that are commensurate to . Fix some and consider the semigroup of measure-preserving transformations between elements of . More precisely, let be the collection of measure-preserving maps between sets , which we call the self commensurating semigroup of .
We use the notation and to refer to the domain and the range of , respectively. As usual, we identify two maps that differ on a null set. Since classes of commensurability are closed under finite intersections, the set forms a semigroup under the composition.
This semigroup carries a natural equivalence relation: whenever the transformations disagree on a set of finite measure, . This equivalence is, moreover, a congruence, i.e., if and , then . One may therefore push the semigroup structure from onto the set of equivalence classes, which we denote by . An important observation is that is a group. Indeed, the identity corresponds to the map on , and for a representative , its inverse inside is, naturally, given by . We call the self commensurating automorphism group of .
The self commensurating semigroup admits an important homomorphism into the reals, , called the index map and defined by
Lemma 6.1.
For all , the index map satisfies the following:
-
(1)
if is such that and , then
-
(2)
if is a restriction of , that is , then .
Proof.
(1) If is commensurate to and , , then
Proposition 6.2.
The index map is a homomorphism. Moreover, if are equivalent, , then .
Proof.
The proposition above implies that the index map respects the relation , and hence gives rise to a map from to the reals.
Corollary 6.3.
The index map factors to a group homomorphism
6.2. The commensurating automorphism group
Let us again consider an infinite measure space and a measurable subset. We now define the commensurating automorphism group of in as the group of all measure-preserving transformations such that . We denote this group by .
Every naturally gives rise to an element of by considering its restriction . The following lemma shows that in this case we may use any other set commensurate to instead without changing the corresponding element of the commensurating group.
Lemma 6.4.
Let be a measure-preserving automorphism. If for some , then and for all .
Proof.
Since commensuration is an equivalence relation and is commensurate to , the assumption is equivalent to . Moreover, given , we only need to show that is finite in order to conclude that . So we compute
Thus the measure is finite, hence for all . Finally, , since these transformations agree on . ∎
To summarize, if for some , then all restrictions , , are pairwise equivalent, hence correspond to the same element . According to Proposition 6.2, the index of this element can be computed as for any .
6.3. Index map on full groups of -flows
Let be a free measure-preserving Borel flow, let be the associated full group, where we endow with the standard Euclidean norm, and let . The action of upon is denoted additively by . Recall that the cocycle of is denoted by and is defined by the equality for all . We are going to argue that, on every orbit, induces a measure-preserving transformation that belongs to the commensurate group of , when the orbit is identified with the real line.
Consider the function defined by
One can think of as a “charge function” that spreads charge over each interval and over . Note that . Since belongs to the full group, its cocycle is integrable, which means that is -integrable (see Section 4.2). We apply the mass-transport principle, which shows that
Let denote the transformation induced by onto the orbit of obtained by identifying the origin of the real line with . The following two quantities are therefore finite:
In particular, belongs to the commensurating group of . The second quantity, on the other hand, is equal to the index of . By Section 6.2, whenever . For any , we therefore have an orbit invariant measurable map given by . Note that for any -invariant set , we have
(6.1) |
Let , , be the ergodic decomposition of (see Appendix C.3). Since the map is -invariant, it produces a map via for any such that or, equivalently, via
Note also that
thus . We can now define the index map of a (possibly non-ergodic) flow as a function .
Definition 6.5.
Let be a free measure-preserving flow on a standard probability space ; let also be the space of -invariant ergodic probability measures, where is the probability measure yielding the disintegration of . The index map is the function given by .
Proposition 6.6.
For any free measure-preserving flow , the index map is a continuous and surjective homomorphism.
Proof.
The index map is a homomorphism, since, as we have discussed earlier, is equal to the index of . Continuity follows from the fact that is a Borel homomorphism between Polish groups. To see surjectivity, pick any , view it as a map via the identification . Define the automorphism by . It is straightforward to check that and . ∎
The quotient group naturally inherits the quotient norm given by
By Proposition 6.6, the index map induces an isomorphism between and . We argue that this isomorphism is, in fact, an isometry.
Proposition 6.7.
The index map induces an isometric isomorphism between and , where the former is endowed with the quotient norm and the latter bears the usual norm.
Proof.
Since , it suffices to show that for all
Let . We first show the inequality .
Pick any . For any -invariant measurable , and
where we rely on Eq. (6.1) and being measure-preserving. Consider the -invariant sets
The norm can be estimated from below as follows.
We conclude that
For the other direction, consider a transformation defined by ; note that , for all , and . Therefore
and the desired equality of norms follows. ∎
Using a similar reasoning, we get the following characterization of the full group and the index map, where for all we let be the measure-preserving transformation of given by (see Section 4.2).
Proposition 6.8.
Let be a free measure-preserving -flow. Consider the set . Then for every , we have
In particular, the full group of can be seen as the commensurating group of inside the full group of . Moreover, in the ergodic case, the index of as defined above is equal to its index as a commensurating transformation of the set in the sense of Section 6.1.
Proof.
Through the identification , the measure-preserving transformation is acting on as , and the set becomes . We then have
by the mass-transport principle, which yields the conclusion, since by the definition of the norm .
The moreover part follows from a similar computation. ∎
Remark 6.9.
The full group of embeds via into the group of measure-preserving transformations of . One could use this and the fact that the commensurating automorphism group of is a Polish group in order to give another proof that full groups of measure-preserving -flows are themselves Polish.
Chapter 7 Orbitwise ergodic bounded elements of full groups
The purpose of this chapter is to contrast some of the differences in the dynamics of the elements of full groups of -actions and those arising from -flows. Let be an element of the full group of a measure-preserving aperiodic transformation and let be the cocycle associated with for . Since is a discrete group, the conservative part in the Hopf’s decomposition for (see Appendix B) reduces to the set of periodic orbits. In particular, an aperiodic has to be dissipative, hence as . When belongs to the full group of the action, a theorem of R. M. Belinskaja [Bel68, Thm. 3.2] strengthens this conclusion and asserts that for almost all in the dissipative component of either or .
Given an arbitrary free measure-preserving flow , we build an example of an aperiodic for which the signs in keep alternating indefinitely for almost all . In fact, we present a transformation that acts ergodically on each orbit of the flow (in particular, it is conservative and globally ergodic as soon as the flow is ergodic). Moreover, we ensure it has a uniformly bounded cocycle. Our argument uses a variant of the well-known cutting and stacking construction adapted for infinite measure spaces. Additional technical difficulties arise from the necessity to work across all orbits of the flow simultaneously. The transformation will arise as a limit of special partial maps we call castles, which we now define.
The pseudo full group of the flow is the set of injective Borel maps between Borel sets , , for which there exists a countable Borel partition of the domain and a countable family of reals such that for every . Such maps are measure-preserving isomorphisms between and . The support of is the set
Given in the pseudo full group and a Borel set , we let
In particular, if is disjoint from . A castle is an element of the pseudo full group of the flow such that for the sequence consists of pairwise disjoint subsets which cover its support. Since is measure-preserving, for almost every there is such that . It follows that is also a castle. The set is called the basis of the castle, and the basis of its inverse is called its ceiling, which is equal to . Observe that if two castles have disjoint supports, then their union is also a castle. We denote by the element of the pseudo full group which takes every element of the basis of to the corresponding element of the ceiling.
Remark 7.1.
Equivalently, one could define a castle as an element of the pseudo full group which induces a graphing consisting of finite segments only (see [KM04, Sec. 17] for the definition of a graphing). It induces a partial order defined by if and only if there is such that . The basis of the castle is the set of minimal elements, while the ceiling is the set of maximal ones. Finally, is the map which takes a minimal element to the unique maximal element above it.
Theorem 7.2.
Let be a free measure-preserving flow. There exists that acts ergodically on every orbit of the flow and whose cocycle is bounded by . Moreover, the signs in keep changing indefinitely for almost all .
Proof.
Fix a free measure-preserving flow , and let be a cross-section. Since is lacunary, for any there exists ; we denote this value by . This gives the first return map via , which is Borel. There is also a natural bijective correspondence between and the set . Let be the “Lebesgue measure” on given by
The measure on can be disintegrated as for some finite (but not necessarily probability) measure on (see, for instance, [Slu17, Sec. 4] and Appendix C.1).
Let be a vanishing sequence of markers—a sequence of nested cross-sections with the empty intersection: . We may arrange to be such that for all . Put
and . Note that . Our first goal is to define an element of the pseudo full group with domain and range equal to such that for almost every , the action of on the intersection of the orbit of with is ergodic, and which has cocycle bounded by . It will then be easy to modify to an element of the full group whose action on each orbit of the flow is ergodic at the cost of increasing the cocycle bound to .
Our first transformation will arise as the limit of a sequence of castles , with each belonging to the pseudo full group of . We also use another family of castles which allows us to extend by “going back” from its ceiling to its basis while keeping the cocycle bound (this is our main adjustment compared to the usual cutting and stacking procedure). Both sequences of castles will have their cocycles bounded by . Here are the basic constraints that these sequences have to satisfy:
-
(1)
for all , ;
-
(2)
for all , extends ;
-
(3)
tends to as tends to .
Bases and ceilings of and will satisfy additional constraints which will enable us to make the induction work and ensure ergodicity on each orbit of the flow. In order to specify these constraints properly, we introduce the following notation.
Each orbit of the flow comes with the linear order inherited from via if and only if for some . Set to be the minimum of the intersection of with the cone .
Let and be the set of those which are maximal in among points of for some ; in other words,
Note that by construction the distance between and is less than for each . Let be the map which assigns to the -least element of which is greater than .
The bases and ceilings of and are as follows.
-
•
the basis of is ;
-
•
the ceiling of is ;
-
•
the basis of is ;
-
•
the ceiling of is .
Furthermore, we impose two translation conditions, which help us to preserve the above concrete definitions of the bases and ceilings at the inductive step when we construct and :
-
•
for all and all .
-
•
for all and all .
The first step of the construction consists of the castle , which has the basis and ceiling , and the castle defined for with ceiling .
We now concentrate on the induction step: suppose and have been built for some , let us construct and .
The strategy is to split the basis of and into two equal intervals and “interleave” the “two halves” of with “one half” of followed by “gluing” adjacent ceilings and basis within the same segment (see Figure 7.2). To this end, we introduce two intermediate castles and which will ensure that “wiggles” more than , yielding ergodicity of the final transformation.
Define two new half measure subsets of the bases and respectively:
-
•
;
-
•
;
and let
and
where the two equalities are consequences of the translation conditions. Let be the -saturation of , and note that the restriction of to is a castle with support , whose basis is and whose ceiling is . Finally, let
We define the partial measure-preserving transformation to be used for “gluing together” and the restriction of to :
-
•
for all and
-
•
for all .
Set , whereas is simply the restriction of onto the complement of . Observe that has basis and ceiling
while has basis
and ceiling
We continue to have , but the support of is half the support of , and .
The ceiling of is equal to , whereas we need the ceiling of to be equal to . We obtain the required and out of and respectively by “passing through each element of ”.
Note that is equal to the set of such that . Each can be written uniquely as where and . Set
and note that belongs to , hence is a measure-preserving bijection from onto .
The transformation is set to be , and we claim that it is a castle with basis and ceiling . This amounts to showing that for all , there is such that is not defined. Pick and write it as for some and . Let be the successor of in , which we suppose not to be an element of . By the construction of and , there is such that , which means that . Iterating this argument, we eventually find such that for some such that the successor of in belongs to . By the definition of we must have some such that , whereas is not defined, thus is indeed a castle.
Extension of is defined similarly by connecting adjacent segments of and by a translation. More specifically, each can be written uniquely as for some and . The restriction of to is a bijection , we denote its inverse by and let . The map can be checked to be a castle with basis and ceiling as desired. It also follows that the translation conditions continue to be satisfied by both of and .
Transformations extend each other, so is an element of the pseudo full group supported on . Note also that
and therefore . We claim that , seen as a measure-preserving transformation of , induces an ergodic measure-preserving transformation on for almost all , where is endowed with the Lebesgue measure. This follows from the fact that induces a rank-one transformation of the infinite measure space : for all Borel of finite Lebesgue measure and all , there are , , and a subset such that are pairwise disjoint and
Indeed, at each step for every , the iterates of by the restriction of to the interval are disjoint “intervals of size ”, i.e., sets of the form , and these iterates cover a proportion of (the rest of this interval being ).
It remains to extend supported on to a measure-preserving transformation with . Let be the leftover set,
and put
Figure 7.3 illustrates an interval between and . Within this gap, corresponds to , and is an interval of the exact same length adjacent to it on the left. Note that by construction. Let be the natural translation map, for all satisfying . Observe that is a measure-preserving bijection and its cocycle is bounded by .
We now rewire the orbits of and define as follows (see Figure 7.3):
It is straightforward to verify that is a free measure-preserving transformation, and the distance for all , because and for all in their domains. Note that the transformation induced by on is equal to , so since the latter is ergodic on every orbit of the flow intersected with and since , it follows that is ergodic on every orbit of the flow and satisfies the conclusion of the theorem. ∎
Remark 7.3.
The bound in the formulation of Theorem 7.2 is of no significance as by rescaling the flow it can be replaced with any .
Chapter 8 Conservative and intermitted transformations
Interesting dynamics of conservative transformations is present only in the non-discrete case, as it reduces to periodicity for countable group actions. Chapter 7 provides an illustrative construction of a conservative automorphism, and shows that they exist in full groups of all free flows. The present chapter is devoted to the study of such elements. The central role is played by the concept of an intermitted transformation, which is related to the notion of induced transformation. Using this tool we show that all conservative elements of can be approximated by periodic automorphisms, and hence belong to the derived full group of ; see Corollary 8.8.
Throughout the chapter, we fix a free measure-preserving flow on a standard Lebesgue space . Given a cross-section , recall that we defined an equivalence relation by declaring whenever there is such that both and belong to the gap between and . More formally, if there is such that , and , . Such an equivalence relation is smooth.
Now let be a conservative transformation. Under the action of , almost every point returns to its -class infinitely often, which suggests the idea of the first return map.
Definition 8.1.
The intermitted transformation is defined by
The map is well-defined, since is conservative, and it preserves the measure , since belongs to the full group of .
Remark 8.2.
The concept of an intermitted transformation makes sense for any equivalence relation for which intersection of any orbit of with any -class is either empty or infinite. In particular, intermitted transformations can be considered for any conservative in a full group of a locally compact group action. For instance, with a cocompact cross-section we can associate an equivalence relation of lying in same cell of the Voronoi tessellation (see Appendix C.2). Such an equivalence relation does have the aforementioned transversal property, and hence intermitted transformation is well-defined.
Note also the following connection with the more familiar construction of the induced transformation. Let , let be a set of positive measure, and define to be the equivalence relation with two classes: and . Induced transformations and commute and satisfy .
The next lemma forms the core of this chapter. It shows that the operation of taking an intermitted transformation does not increase the norm. As we discuss later in Remark 8.5, the analog of this statement is false even for -flows, which perhaps justifies the technical nature of the argument.
Lemma 8.3.
Let be a conservative automorphism and let be a cross-section. Let also be the set of points where and differ: . One has .
Proof.
By the definition of , for any the arc from to jumps over at least one point of . We may therefore represent as the sum of the distance from to the first point of along the arc plus the rest of the arc. More formally, for let be the unique such that . Define by
Note that , and set , so that
For instance, in the context of Figure 8.1, and . Let us partition , where
and consists of those for which the signs of and are different. For example, referring to the same figure, , while .
To prove the lemma it is enough to show two inequalities:
(8.1) |
(8.2) |
Eq. (8.1) is straightforward, since equality of signs of and implies that is closer than to the point over which goes the arc from to . For example, the point in Figure 8.1 satisfies
Thus for all and so
which gives (8.1). The other inequality will take us a bit more work.
For , let be the smallest integer such that the sign of is opposite to that of . In less formal words, is the smallest integer such that the arc from to jumps over . In particular, points , , are all on the same side relative to , while is on the other side of it. We consider the map given by . Properties of this map will be crucial for establishing the inequality (8.2), so let us provide some explanations first.
Consider once again Figure 8.1, which shows a partial orbit of a point for up to and several points . First, as we have already noted before, , since ; moreover, , since is to the left of , while is to the right of it, so and have the opposite signs. Also, , because is the first point in the orbit to left of , thus . In particular, generally , but is the case for whenever and are -equivalent.
The next point in the orbit , whereas but , because and both and are positive. The point belongs to and has with . Points , but whether any of them are elements of is not clear from Figure 8.1, as the orbit segment is too short to clarify the values of , . However, if happen to lie in , then with , and , , . In particular, the function is not necessarily one-to-one, but we are going to argue that it is always finite-to-one.
Claim 1.
If are distinct points such that , then .
Proof of the claim.
Suppose satisfy . The definition of implies that and must belong to the same orbit of , and we may assume without loss of generality that for some . If the orbit of and is aperiodic, it implies that that and , . However, even if the orbit is periodic, either for the smallest positive integer such that or for the smallest positive integer such that . Interchanging the roles of and if necessary, we may therefore assume that holds for some , , regardless of the type of orbit we consider.
Suppose and are -equivalent. Let be the smallest natural number for which and are -equivalent. By the assumption and the choice of we have . By the definition of , all points , , are on the same side of . In particular, this applies to and , which shows that and have the same sign, thus . ∎
The above claim implies that the function is finite-to-one for the arc from to intersects only finitely many -equivalence classes, and the preimage of picks at most one point from each such class. Note also that for all , but may not be an element of . Among the -equivalence classes that the arc from to goes over, two are special—the intervals that contain and , respectively. Our goal will be to bound the sum of over the points with the same value by (see Claim 3 below). For a typical point we can bound simply by the length of the interval of its -class. For example, Figure 8.1 does not specify , but we can be sure that . In view of Claim 1, such an estimate comes close to showing that the sum of over with the same image is bounded by . It merely comes close, due to the two special -classes mentioned above, where our estimate needs to be improved. The next claim shows that one of these special cases is of no concern as is never -equivalent to .
Claim 2.
For all we have .
Proof of the claim.
Suppose towards the contradiction that , and let be the smallest integer for which ; in particular, . Note that by the assumption, and by the definition of , has the same sign as , whence . ∎
Pick some with non-empty preimage , and let be all the elements in . For instance, in the situation depicted in Figure 8.1, we may have and , , , and . The following claim unlocks the path towards the inequality (8.2).
Claim 3.
In the above notation, .
Proof of the claim.
Recall that the arc from to crosses at least one point in . If is the first such point, then is defined to be . For instance, in the notation of Figure 8.1, . Each point is located under the arc from to , and by Claim 2, no point belongs to the interval from to . In the language of our concrete example, no point can be between and . As discussed before, is always bounded by the length of the gap to which belongs. This is sufficient to prove the claim if no is equivalent to , as in this case the whole -equivalence class of every is fully contained under the interval between and , and distinct represent distinct -classes by Claim 1. This is the situation depicted in Figure 8.1, and our argument boils down to the inequalities
Suppose there is some such that . By Claim 1 such must be unique, and we assume without loss of generality that . For example, this situation would occur if in Figure 8.1 were equal to . Let be the first element of over which goes the arc from to (it would be the point in Figure 8.1). It is enough to show that , as we can use the previous estimate for all other , . Note that , and by assumption, which implies that the signs of and are different. The latter is equivalent to saying that is between and , i.e., , and the claim follows. ∎
We are now ready to finish the proof of this lemma. We have already shown that is finite-to-one, so let , , be such that is -to-one on . Let , and recall that . Sets are pairwise disjoint. Let , , be Borel bijections that pick the th point in the preimage: . Note that maps are measure-preserving, since they belong to the pseudo full group of , and for all by Claim 3. One now has
Summing these inequalities over we get
where the last inequality is based on the fact that sets are pairwise disjoint. This finishes the proof of the inequality (8.2) as well as the lemma. ∎
Several important facts follow easily from Lemma 8.3. For one, it implies that for any cross-section the intermitted transformation belongs to . In fact, we have the following inequality on the norms.
Corollary 8.4.
For any intermitted transformation one has .
Proof.
Remark 8.5.
As we discussed in Remark 8.2, the concept of an intermitted transformation applies wider than the case of one-dimensional flows. We mention, however, that the analog of Lemma 8.3 and Corollary 8.4 does not hold even for free measure-preserving -flows. Consider an annulus depicted in Figure 8.2(a) and let be the rotation by an angle around the center of this annulus. Let the equivalence relation consist of two classes, each composing half of the ring. For a point such that , will be close to the other side of the class. It is easy to arrange the parameters (the angle and the radii of the annulus) so that for all such that .
Every free measure-preserving flow admits a tiling of its orbits by rectangles. The transformation can be defined similarly to Figure 8.2(a) on each rectangle of the tiling by splitting each tile into two equivalence classes as in 8.2(b). The resulting transformation will have bounded orbits and satisfy relative to the equivalence relation whose classes are the half tiles.
When the gaps in a cross-section are large, and will often be -equivalent, and it therefore natural to expect that will be close to . This intuition is indeed valid, and the following approximation result is the most important consequence of Lemma 8.3.
Lemma 8.6.
Let be a conservative transformation. For any there exists such for any cross-section with for all one has .
Proof.
Let , , be the set of points whose cocycle is at least in the absolute value. Since , we may pick is so large that . Pick any real such that . We claim that it satisfies the conclusion of the lemma. To verify this we pick a cross-section with all gaps having size at least . Set as before . Since
our task is to estimate this integral. This can be done in a rather crude way. We can simply use the triangle inequality , and deduce
where the last inequality is based on Lemma 8.3.
It remains to show that . Let be the region that leaves out intervals of length on both sides of each point in . Note that for any one has and thus for such points. Therefore, , where , and thus
Lemma 8.7.
Let be a conservative transformation. For any there exists a periodic transformation such that .
Proof.
By Lemma 8.6, we can find a cocompact cross-section such that . Let be an upper bound for gaps in . Recall that the cocycle is uniformly bounded by , and, in fact, the same is true for any element in the full group of . In particular, we may use Rokhlin’s lemma to find a periodic such that , and conclude that . We therefore have
Corollary 8.8.
If is conservative then belongs to the derived full group , in particular its index satisfies .
Chapter 9 Dissipative and monotone transformations
The previous chapter studied conservative transformations, whereas this one concentrates on dissipative ones. Our goal will be to show that any dissipative of index belongs to the derived subgroup . We begin however by describing some general aspects of dynamics of dissipative automorphisms.
Recall that according to Proposition 4.16, any transformation induces a -invariant partition of the phase space such that is conservative and is dissipative. Formally speaking, a transformation is said to be dissipative if the partition trivializes to . For the purpose of this chapter it is however convenient to widen this notion just a bit by allowing to have fixed points.
Definition 9.1.
A transformation is said to be dissipative if , where is the dissipative element of the Hopf’s decomposition for .
9.1. Orbit limits and monotone transformations
We begin by showing that dynamics of dissipative transformations in full groups of -flows is similar to those in full groups of actions. We do so by establishing an analog of R. M. Belinskaja’s result [Bel68, Thm. 3.2]. Recall that a sequence of reals is said to have an almost constant sign if all but finitely many elements of the sequence have the same sign.
Proposition 9.2.
Let be a measure-preserving transformation of the real line which commensurates the set , suppose that is dissipative. Then for almost all , the sequence of reals has an almost constant sign.
Proof.
Let be the set of reals such that does not have an almost constant sign, and suppose by contradiction that has positive measure. Since is dissipative, we can find a Borel wandering set for which non-trivially intersects . All the translates of are disjoint, and, for all , does not have an almost constant sign.
Since is dissipative, for almost all , the sequence of absolute values tends to (see Proposition B.4). In particular, there are infinitely many points in the -orbit of such that but . Since the map which maps to is measure-preserving, this yields that the set of such that has infinite measure, contradicting the fact that commensurates the set . ∎
Corollary 9.3.
Let be a dissipative transformation. For almost all , the sequence , has an almost constant sign.
Proof.
Let . For all , denote by the measure-preserving transformation of induced by on the -orbit of . By the proof of Proposition 6.8, the integral
is finite. In particular, for almost every , the transformation commensurates the set . The conclusion now follows directly from the previous proposition. ∎
For any dissipative transformation in an full group of a free locally compact Polish group action and for almost every , as , in the sense that eventually escapes any compact subset of the acting group. In the context of flows, Corollary 9.3 strengthens this statement and implies that must converge to either or .
Corollary 9.4.
If is dissipative, then for almost every point either or . ∎
In view of this corollary, there is a canonical -invariant decomposition of into “positive” and “negative” orbits.
Definition 9.5.
Let be a dissipative automorphism. Its support is partitioned into , where
The set is said to be positive evasive and is negative evasive.
According to Corollary 9.3, for almost every , eventually either all are to the right of or all are to the left of it. There are points for which the adverb “eventually” can, in fact, be dropped.
Corollary 9.6.
Let be a dissipative transformation and let
The set is a complete section for .
Proof.
We need to show that almost every orbit of intersects . Let and suppose for definiteness that . Since , there is , and therefore . ∎
Definition 9.7.
A dissipative transformation is monotone if for almost all , and for almost all .
Corollary 9.8.
Let be a dissipative transformation. There is a complete section and a periodic transformation such that and is monotone.
Proof.
Take to be as in Corollary 9.6 and note that is periodic and satisfies the conclusions of the corollary. ∎
As we discussed at the beginning of the chapter, our goal is to show that the index of the kernel map coincides with the derived subgroup of . Note that if is as above, then , and, coupled with the results of Chapter 8, it will suffice to show that all monotone transformations of index zero belong to . This will be the focus of the rest of this chapter and will take some effort to achieve, but the main strategy is to show that such automorphisms can be approximated by periodic maps, which is the content of Theorem 9.15 below.
9.2. Arrival and departure sets
Throughout the rest of this chapter, we fix a cross-section and a monotone transformation . The arrival set is the set of the first visitors to classes: . Analogously, the departure set is defined to be . We also let denote and ; likewise for and . Note that , and thus . There is, however, another useful map from onto .
We define the transfer value by the condition
and the transfer function is defined to be . Note that is measure-preserving. The transfer value introduces a partition of the arrival set , where ; by applying the transfer function, it also produces a partition for the departure set: , where .
In plain words, is the number of points in . Therefore if for some then also . In Sections 9.3 and 9.4 we modify the transformation on the arrival and departure sets and we want to do this in a way that affects as many orbits as possible as measured by . This amounts to using sets (and ) with as high values of as possible. The next lemma will be helpful in conducting such a selection in a measurable way across all of .
Lemma 9.9.
Let be a measurable set with a measurable partition and let be a measurable function such that for all . There are measurable and such that for any for which one has
Proof.
For such that set
Note that one necessarily has . Set
These functions and satisfy the conclusions of the lemma. ∎
Definition 9.10.
Consider the partition of the positive arrival set and let , , and be as in Lemma 9.9. The set defined by the condition
is said to be the positive -copious arrival set. The positive -copious departure set is given by . The definitions of the negative -copious arrival and departure sets use the partition of the negative arrival set and are analogous.
Copious sets maximize measure of their saturation under the action of . In other words, among all subsets for which , the measure is maximal when . In particular, if is close to , then we expect to be close to . The following lemma quantifies this intuition.
Lemma 9.11.
Let be such that for all , and let be the -copious arrival set constructed in Lemma 9.9. If there exists such that for all , then
and therefore also .
An analogous statement is valid for the negative arrival set .
Proof.
Let be as in Lemma 9.9 and note that
whenever satisfies . Recall that for we have for all and sets are pairwise disjoint. In particular,
(9.1) |
Note also that implies
(9.2) |
For any we have
The inequality for the measure follows by disintegrating into .
The argument for the negative arrival set is completely analogous. ∎
9.3. Coherent modifications
We remind the reader that our goal is to show that any dissipative transformation of index can be approximated by periodic transformations. One approach to “loop” the orbits of is by mapping to and to (cf. Figure 9.6). For such a modification to work, measures and have to be equal. Recall that implies that for almost every , the measure of points such that equals the measure of those for which . If one could guarantee that , then the aforementioned modification would indeed work. In the case of actions, discreteness of the acting group allows one to find a cross-section for which this condition does hold. Whereas for the flows, we have to deal with the possibility that can be “scattered” (see Figure 9.4) along the orbit and be unbounded, which is the key reason for the increased complexity compared to the argument for actions.
Since we can’t hope to “loop” all the orbits of , we will do the next best thing, and apply the modification of Figure 9.6 on “most” orbits as measured by . Copious sets discussed in Section 9.2 have large saturations under , but, generally speaking, fail to satisfy for the same reason as do the sets . Our plan is to use the “ of room” provided by the difference in order to modify into some with the same arrival and departure sets as , but for which also holds. In this section, we describe two abstract modifications of dissipative transformations, and the approximation strategy outlined above will later be implemented in Section 9.4.
Since we are about to consider arrival and departure sets of different transformations, we use the notation to denote the positive arrival set constructed for a transformation ; likewise for negative arrival and departure sets, etc.
Lemma 9.12.
Let and be measure-preserving transformations on subject to the following conditions:
-
(1)
, ;
-
(2)
, , and , ;
-
(3)
and for all .
The transformation is monotone, for all , and the sets , remain the same:
Moreover, the integral of lengths of “departing arcs” remains unchanged:
and the following estimate on is available:
Proof.
Figure 9.2 illustrates the definition of the transformation . Equality of the arrival and departure sets is straightforward to verify. Note that for all , and therefore . In fact, the following four integrals vanish:
(9.3) |
Observe that is positive on and negative on , thus
Finally, note that for any , the arc from to intersects the arc from to (both arcs go over the same point of ), and therefore
Integration over yields
Lemma 9.13.
Let be a monotone transformation, let be such that for all and the function is -invariant (i.e., whenever and belong to the same orbit of the flow). Let be the arrival subset that corresponds to , i.e., . Let and be any measure-preserving transformations such that and for all in the corresponding domains. Define by the following formula:
The transformation is a measure-preserving automorphism from the full group and for all . The integral of distances can be estimated as follows:
The following figure illustrates the notions of Lemma 9.13.
Proof.
It is straightforward to verify that is a measure-preserving transformation. For the integral inequality note that for any one has
and therefore
A similar inequality holds for , and the lemma follows. ∎
9.4. Periodic approximations
We now have all the ingredients necessary to prove that monotone transformations can be approximated by periodic automorphisms. Our arguments follow the approach outlined at the beginning of Section 9.3.
In the following lemma, we assume that the Lebesgue measure of those that jump over any given is bounded from above by some , and that most of such jumps — of measure at least — are between adjacent -classes. We are going to construct a periodic approximation of the transformation with an explicit bound on , which can be made small for a sufficiently sparse cross-section . When the flow is ergodic, this lemma alone suffices to conclude that . Theorem 9.15 builds upon Lemma 9.14 and treats the general case.
Lemma 9.14.
Let be a monotone transformation, let be a positive real, and let . Let be a cross-section such that for all . Let be reals such that for all :
There exists a periodic transformation such that and
Proof.
Let and be the departure and the arrival sets of . Figure 9.4 depicts the arrival and the departure sets for an element of the cross-section . Note that preimages may come from different (possibly, infinitely many) -equivalence classes; likewise, images of the departure set may visit several -equivalence classes.
Set to be the constant function; in view of the assumptions on , we may apply Lemma 9.9 to get positive and negative -copious arrival sets , , as well as the corresponding departure sets and . Set and . We have for all . Let
be the set of arcs that jump from/to the next -equivalence class. By the assumptions of the lemma, we have and for all . Let be any measure-preserving transformation such that:
-
•
is supported on ;
-
•
and ;
-
•
for all ;
and moreover
(9.4) |
Select a transformation such that
-
•
is supported on ;
-
•
and ;
-
•
for all ;
and moreover
(9.5) |
Figure 9.5 illustrates these maps. Note that while in general , one has for all by the definition of the -copious departure set.
Let be the transformation obtained by applying Lemma 9.12 to , and . The automorphism satisfies and for all . Choose a measure-preserving transformation such that for all in the domain of . Set . Let be the transformation that is produced by Lemma 9.13 applied to , , and (see Figure 9.6).
Finally, set to be
We claim that satisfies the conclusions of the lemma. It is periodic, since the transformation is the identity map and by construction. It remains to estimate .
We concentrate on estimating . Recall that for all , hence for . Set and note that for , and therefore using the conclusion of Lemma 9.12 we have
(9.6) |
The integral can now be estimated as follows.
Finally, we consider the integral and partition its domain as , which yields
where the last inequality follows from Lemma 9.11 with . Combining all the inequalities together, we get
Lemma 9.14 allows us to approximate with a periodic transformation a monotone for which the Lebesgue measure of points jumping over any given is roughly constant across orbits. To deal with the general case, we simply need to split the phase space into countably many segments invariant under the flow, and apply Lemma 9.14 on each of them separately. Small care needs to be taken to ensure that values , which appear in the formulation of Lemma 9.14, remain uniformly small across the partition of . Details are presented in the following theorem.
Theorem 9.15.
Let be a monotone transformation that belongs to the kernel of the index map, hence
for almost all . For any there exists a periodic transformation such that and .
Proof.
Let be such that for the set
one has . Pick a cross-section with gaps so large that
for all , which ensures
(9.7) |
Note also that Eq. (9.7) holds for any cross-section , since and for all .
For any positive real there exists a positive so small that and . We may therefore pick countably many positive reals , , such that and
(9.8) |
Define intervals , .
Let be the map that measures the set of forward arcs over its argument:
Set and construct inductively . Sets are pairwise disjoint, and moreover, for all , , . Let , , be the function defined by
Set and define inductively . Let denote the saturated set . Finally, for all , let be such that for all . Sets and satisfy the following conditions:
-
(1)
is a cross-section for the restriction of the flow onto ;
-
(2)
sets , , are pairwise disjoint.
-
(3)
and for all .
Let denote the restriction of onto . Apply Lemma 9.14 to the transformation , cross-section , which has gaps at least , and , . Let be the resulting periodic transformation on . Set . We claim that satisfies conclusions of the theorem. Set and note that , whence .
and the theorem follows. ∎
Corollary 9.16.
Let be a measure-preserving flow and be a dissipative transformation. If , then .
Chapter 10 Conclusions
Our objective in this last chapter is to draw several conclusions regarding the structure of the full groups of measure-preserving flows. The analysis conducted in Chapters 8 and 9 leads to the most technically challenging result of our work, which is the following theorem.
Theorem 10.1.
Let be a free measure-preserving flow on a standard probability space. The kernel of the index map coincides with the closed derived subgroup .
Proof.
Inclusion is automatic since the image of is abelian. For the other direction, pick a transformation and consider its Hopf’s decomposition provided by Proposition 4.16. We have , where is conservative and is dissipative. According to Corollary 8.8, and , whence . Therefore, the dissipative part satisfies the assumptions of Corollary 9.16, which yields , and hence as desired. ∎
10.1. Topological ranks of full groups
Empowered with the result above and Corollary 5.20, we can estimate the topological ranks of full groups of flows. We recall the following well-known inequalities.
Proposition 10.2.
Let be a surjective continuous homomorphism of Polish groups. The topological rank satisfies
Proposition 10.3.
Let be a free measure-preserving flow on a standard probability space . The topological rank is finite if and only if the flow has finitely many ergodic components. Moreover, if has exactly ergodic components then
Proof.
Let be the space of probability invariant ergodic measures of the flow, and let be the probability measure on such that (see Appendix C.3). Proposition 6.6 shows that the index map is continuous and surjective. An application of Proposition 10.2 yields
(10.1) |
where the last equality is based on Theorem 10.1 and Corollary 5.20. Since is a Banach space, its topological rank is finite if and only if its dimension is finite, which is equivalent to being purely atomic with finitely many atoms. We have shown that is finite if and only if the flow has only finitely many ergodic components. The moreover part of the proposition follows from the inequality (10.1) and the observation that . ∎
As already mentioned in the introduction, we conjecture that the topological rank completely remembers the number of ergodic components.
Conjecture 10.4.
Let be a measure-preserving flow. If it has exactly ergodic components, then .
Provided the conjecture holds, we have a priori no way of distinguishing full groups of ergodic flows as topological groups. For -actions, it is a consequence of Belinskaya’s theorem that there are many full groups. The next two sections are devoted to analogues of her result for flows, yielding that there are many full groups of free ergodic flows, although we don’t have a concrete way of distinguishing them (we will discuss in the last section their geometry, which might help there).
10.2. Katznelson’s conjugation theorem
R. M. Belinskaja [Bel68] showed that if measure-preserving transformations generate the same orbit equivalence relation, i.e., , and , then and are conjugated. Y. Katznelson found a different argument and isolated a sufficient condition for conjugacy of measure-preserving transformations (see [CJMT22, Theorem A.1]). In the following, for , , and we let denote the set .
Theorem 10.5 (Katznelson).
Suppose are measure-preserving transformations that generate the same orbit equivalence relation, . If the symmetric difference is finite for almost all , then and are conjugated by an element from the full group .
The analog of this result for free measure-preserving flows will be proved shortly in Theorem 10.9. But first we discuss an important application of Theorem 10.5. Consider a free measure-preserving flow . Given a dissipative transformation (in the sense of Definition 9.1), Proposition B.4 implies that almost every non-trivial -orbit is a discrete subset of unbounded both from below and from above. The order induced on by the flow may disagree with the -order of points. One may therefore define the -reordering of to be the first return transformation induced by the ordering of the flow on the orbits of :
Note that and generate the same orbit equivalence relation, .
If belongs to the full group of the flow, either or is finite, depending on whether or (cf. Corollary 9.4). Which symmetric difference is finite may depend on the point , and Theorem 10.5 can be used to show that and its reordering are flip conjugated.
Definition 10.6.
Let and be standard probability spaces, and let , . Measure-preserving transformations and are flip conjugate if there exist an isomorphism of measure spaces and measurable partitions , such that
-
(1)
and ;
-
(2)
are -invariant and are -invariant;
-
(3)
and .
Note that when one of the ’s is ergodic, our definition of flip-conjugacy coincides with the standard one, which requires or to have full measure.
Proposition 10.7.
Any dissipative and its -reordering are flip conjugated by an element from the full group .
Proof.
Consider the decomposition into the positive and negative orbits as in Definition 9.5. In particular, and are finite for and , respectively. Theorem 10.5 implies that there exist automorphisms and such that and . The transformation given by
belongs to the full group and witnesses flip conjugacy of and . ∎
The transformation conjugating and in Theorem 10.5 can be written fairly explicitly. This is done in terms of the function defined as follows. Suppose is a (possibly infinite) measure space, and let be measurable sets such that . We set . This function satisfies a few properties which the reader can easily verify.
Proposition 10.8.
Suppose is a measure space. For all such that , the following holds:
-
(1)
;
-
(2)
and ;
-
(3)
.
Any orbit of a measure-preserving transformation can be endowed with a counting measure. Given and as in the statement of Theorem 10.5, set and define . One can verify that and (further details can be found in [CJMT22, Theorem A.1]).
Let now and be measure-preserving flows on a standard probability space ; we denote the actions of upon by and , respectively. Suppose that their full groups coincide, , and so the flows share the same orbits, . For , let , , denote the “right half-orbit” of . A natural analog of the condition from Theorem 10.5 would be to require finiteness of the Lebesgue measure of for all . This condition alone, however, is not sufficient for conjugacy of and .
Each flow induces a copy of the Lebesgue measure onto orbits via
Since we assume , and so , is a translation invariant measure relative to the action of , and therefore must differ from by a constant: there is an orbit invariant measurable function such that . Any element in preserves , , and therefore cannot conjugate into unless is constantly equal to .
When the flows are ergodic, is a constant, and one may renormalize the flows without changing the full groups. Let be the rescaling of given by . It is straightforward to check that and flows and induce the same measure onto orbits.
For flows that do induce the same measures on the orbits, finiteness of the measure for all is indeed sufficient to conclude conjugacy of the flows.
Theorem 10.9.
Let , , be free measure-preserving flows that share the same orbits, , and induce the same measures onto orbit. If for , then the flows are conjugate by a measure-preserving transformation .
Proof.
Let be the -cocycle defined by . Since and induce the same measure on the orbits, is a Lebesgue measure-preserving automorphism:
For and let
In particular, . Note that
(10.2) |
Also, considering the cases and separately, one can easily verify that for all and
and, in particular,
(10.3) |
Put , then
(10.4) |
The required transformation is given by .
Thus conjugates and . It therefore remains to check that is a measure-preserving bijection. First, note that satisfies . Indeed, (by the analog of Eq. (10.2)), and therefore
(10.5) |
by Proposition 10.8.
To show injectivity, suppose that . In view of Eq. (10.5) and Proposition 10.8,
However, if , then and so . One concludes that and . We have already established that , which shows that the range of is orbit invariant, yielding surjectivity.
Finally, to show that is measure-preserving, it suffices to check that preserves the Lebesgue measure on all the orbits. To this end, let be the -cocycle (i.e., ). For all , one has
hence is the required conjugation between and . ∎
In the case, the above result is the key to Belinskaja’s flip conjugacy result for orbit equivalence. Unfortunately, here we don’t know if it can be useful towards proving an analogous result. In the next section, we nevertheless obtain a weaker result which yields that there are many full groups. We leave the following question open.
Question 10.10.
Given two ergodic flows with equal full groups, do they have to satisfy the hypothesis of the above theorem after appropriate rescaling?
10.3. orbit equivalence implies flip Kakutani equivalence
A measure-preserving action of a compactly generated locally compact Polish group can always be twisted by a continuous automorphism of the group without affecting the full group.
In the case of -actions, this takes a particularly simple form, since the only non-trivial automorphism of is given by . It follows from the results of R. M. Belinskaja [Bel68] that this is up to conjugacy the only way to get an orbit equivalence for ergodic -actions [LM18, Theorem 4.2]: if are two ergodic measure-preserving transformations which are orbit equivalent, then they are flip-conjugate: is conjugate to either or .
As mentioned before, we do not know whether a variant of such rigidity holds when we replace by (see Question 10.17 below), but, as shown in Theorem 10.15, orbit equivalent free measure-preserving flows must at least be flip Kakutani equivalent. In particular, there are uncountably many full groups of free ergodic flows up to abstract group isomorphism.
Let us first define the notion of (flip) Kakutani equivalence of flows. For the main results about this concept, the reader may consults [Kat75, Kat77], where it is called monotone equivalence of flows. Given a measure-preserving automorphism and a positive integrable function , one can define the so-called suspension flow or flow under a function of on the space
under the graph of . For , the action is given by
where is defined uniquely by the condition ; similarly for the action is
where satisfies . Such a flow preserves the restriction onto of the product measure . The space
is a standard probability space. The automorphism in the suspension flow construction is called the base automorphism.
Definition 10.11.
Two flows are (flip) Kakutani equivalent if they are isomorphic to suspension flows over flip conjugate base automorphisms.
It is important to note that the construction of suspension flows can be reversed through the use of cross-sections111In full generality, the definition of a cross-section should actually be relaxed, replacing lacunarity by discreteness in each orbit, and only requiring the gap function of the cross-section to be integrable.. If we have a fixed free flow on and is a co-compact cross-section which is -lacunary where is precompact, then there is a unique probability measure on such that the map taking to is measure-preserving (see e.g. [KPV15, Prop. 4.3] for the general construction). It is then clear that the first-return map is measure-preserving, and our initial flow can be seen as the flow built under the gap function with base transformation .
We need the following important result, which is due to D. Rudolph [Rud76]. Keeping in mind the previous paragraph, it can be reformulated as the fact that every measure-preserving flow is conjugate to a suspension flow over a two-valued function.
Theorem 10.12 (Rudolph).
Let be a free measure-preserving flow on a standard probability space , let , then admits a cross-section whose gap function only takes the values and almost surely.
Remark 10.13.
The second named author has obtained a generalization of this to the purely Borel context, see [Slu19].
Theorem 10.14.
Let be free measure-preserving flows on that share the same orbits, . If , then and are flip Kakutani equivalent.
Proof.
We denote the flow using our usual notation, . As explained right after Definition 10.11, it suffices to find cross-sections for and such that the corresponding first return automorphisms are flip conjugated.
Pick Borel realizations of the flows and let be a Borel cross-section for such that for all , as provided by Theorem 10.12. Define the automorphism by
The transformation is obtained by gluing together the identity map, and , and since all these belong to , which is finitely full, we have that as well. Note that is dissipative and is therefore flip conjugated to its -reordering by Proposition 10.7. In other words, there is a -invariant Borel set of full measure, , and a -invariant Borel partition such that is conjugated to and is conjugated to .
Let be the measure on given for a Borel by . The measure is naturally isomorphic to , where is the Lebesgue measure on , and therefore we have
By Fubini’s theorem, this is equivalent to
Therefore there exists some such that . Note that is the first return map on in the order of the flow , whereas is the first return map in the order induced on the orbits by . Since and are flip conjugated, the flows are flip Kakutani equivalent. ∎
Theorem 10.14 has the following straightforward consequences.
Corollary 10.15.
If two free ergodic measure-preserving flows are orbit equivalent, then they are also flip Kakutani equivalent.
Proof.
This now follows from the definition of orbit equivalence, see Definition 4.19 and the paragraph thereafter. ∎
Corollary 10.16.
If two free ergodic measure-preserving flows have abstractly isomorphic full groups, then they are also flip Kakutani equivalent.
Proof.
We have seen in Proposition 4.21 that isomorphism of full groups of ergodic flows implies orbit equivalence, so the result follows from the previous corollary. ∎
Kakutani equivalence is a highly non-trivial equivalence relation (see, for instance, [ORW82] or [GK21, Kun23]). It seems likely, however, that full groups of flows contain even more information about the action. The only continuous automorphisms of are multiplications by nonzero scalars, and we ask whether isomorphism of full groups necessarily recovers the action up to such an automorphism.
Question 10.17.
Let and be free ergodic measure-preserving flows with isomorphic full groups. Is it true that there is such that and are isomorphic, where denotes the multiplication by ?
Note that a positive answer to Question 10.10 would imply a positive answer to the above question.
10.4. Maximality of the norm and geometry
In this last section, we show that the norm is maximal on full groups of flows. In particular, it defines their quasi-isometry type. Exploring this quasi-isometry type further thus might lead to topological group invariants which distinguish some ergodic flows.
Theorem 10.18.
Let be a free measure-preserving flow. The norm on is maximal.
Proof.
We have already shown that the norm on the derived full group is maximal (see Theorem 5.5). Denote by the space of -invariant ergodic probability measures, where is the probability measure arising from the disintegration of which we write as (see Section C.3). The derived full group is equal to the kernel of the surjective index map and the quotient norm on is equal to the norm on by Proposition 6.7. The latter norm is maximal, as is any Banach space norm.
Given a function , let be given by . The cocycle is constant on each ergodic component and . Furthermore, . We show that is both large-scale geodesic and coarsely proper (see Appendix A.2 and Proposition A.10, in particular).
Any can be written as , where the transformation , and . In particular, we have .
Since the norm is maximal on , it is large-scale geodesic. In fact, Proposition 3.24 establishes that it is large-scale geodesic with constant . We may therefore express as a product of elements , where each has norm at most and
The transformation can, for any , also be expressed as a product
Taking sufficiently large, we can ensure that . Therefore, , and
We conclude that the norm on is large-scale geodesic with .
It remains to prove coarse properness. Let and be positive reals. By Theorem 5.5, there is so large that every element in the derived full group of norm at most is a product of elements of norm at most . Let be any integer greater than . We argue that every element of of norm at most is a product of elements of norm at most .
Indeed, if has norm at most , then
and can therefore by written as a product of elements of each of norm . Also, and by the choice of . The conclusion follows. ∎
Remark 10.19.
While the proposition above states that full groups of flows are quite big, one can use Proposition 6.8 to show that they satisfy the Haagerup property. In other words, such groups admit a coarsely proper affine action on a Hilbert space (namely, the affine Hilbert space ).
Corollary 10.16 along with [ORW82, Sec. 12] implies that there are uncountably many full groups of ergodic free flows up to topological group isomorphism. It would be great if their geometry allowed us to distinguish these groups. However, we don’t even know the answer to the following question.
Question 10.20.
Are there two free ergodic measure-preserving flows with non quasi-isometric full groups?
Appendix A Normed groups
We chose to present our work in the framework of groups equipped with compatible norms rather than metrics. These two frameworks are equivalent, but the former has some stylistic advantages, in our opinion. In Appendix A, we remind the reader the concept of a norm on a group (Section A.1) and state C. Rosendal’s results on maximal norms (Section A.2).
A.1. Norms on groups
Definition A.1.
A norm on a group is a map such that for all
-
(1)
if and only if ;
-
(2)
;
-
(3)
.
If is moreover a topological group, a norm on is called compatible if the balls , , form a basis of neighborhoods of the identity.
There is a correspondence between (compatible) left-invariant metrics on a group and (compatible) norms on it. Indeed, given a left-invariant metric on , the function is a norm. Conversely, from a norm one can recover the left-invariant metric via . Analogously, there is a correspondence between norms and right invariant metrics given by .
The language of group norms thus contains the same information as the formalism of left-invariant (or right-invariant) metrics, but it has the stylistic advantage of removing the need of making a choice between the invariant side, when such a choice is immaterial.
Remark A.2.
Note, however, that there are metrics that are neither left- nor right-invariant, which nonetheless induce a group norm via the same formula . Consider for example a Polish group with a compatible left-invariant metric on it. If is not a CLI group, the metric is not complete, but the metric
is complete. Since , we see that induces the same norm as does the left-invariant metric .
There is a canonical way to push a norm onto a factor group.
Proposition A.3 (see [Fre04, Thm. 2.2.10]).
Let be a Polish normed group, and let be a closed normal subgroup of . The function
is a norm on which is compatible with the quotient topology. In particular, is a Polish normed group.
Definition A.4.
A compatible norm on a locally compact Polish group is proper if all balls are compact.
R. A. Struble [Str74] showed that all locally compact Polish groups admit a compatible proper norm.
A.2. Maximal norms
As was noted in Lemma 2.13, quasi-isometric norms yield the same full groups. C. Rosendal identified the class of Polish groups that admit maximal norms, which are unique up to quasi-isometry. In this section, we state some of results from C. Rosendal’s treatise [Ros22], which are relevant to our work. For reader’s convenience, we formulate the following definitions and propositions in the language of group norms as opposed to left-invariant metrics or écartes, as in the original reference.
Definition A.5 ([Ros22, Def. 2.68]).
A compatible norm on a Polish group is said to be maximal if for any compatible norm there is a constant such that for all .
Definition A.6 ([Ros22, Prop. 2.15]).
Let be a Polish group. A subset is coarsely bounded if for every continuous isometric action of on a metric space , the set is bounded for all , i.e., there is such that for all . A Polish group is boundedly generated if it is generated by a coarsely bounded set.
Theorem A.7 ([Ros22, Thm. 2.73]).
A Polish group admits a maximal compatible norm if and only if it is boundedly generated.
The following characterization is available to establish maximality of a given norm.
Definition A.8 ([Ros22, Def. 2.62]).
A norm on a group is called large-scale geodesic if there is such that for any , there are of norm , , such that and
Definition A.9 ([Ros22, Lem. 2.39(2) and Prop. 2.7(5)]).
A compatible norm on a topological group is called coarsely proper if for every and every , there are a finite subset and such that every element of norm at most can be written as a product
where and each has norm at most .
Proposition A.10 ([Ros22, Prop. 2.72]).
A compatible norm on a Polish group is maximal if and only if it is both large-scale geodesic and coarsely proper.
Appendix B Hopf decomposition
An important tool in the theory of non-singular transformations on -finite measure spaces is the Hopf decomposition, which partitions the phase space into the so-called dissipative and recurrent parts reflecting different dynamics of the transformation. In this appendix, we recall the relevant definitions and indicate what happens for measure-preserving transformations of a -finite space. The reader may consult [Kre85, Sec. 1.3] for further details on the following definitions.
Definition B.1.
Let be a non-singular transformation of a -finite measure space . A measurable set is said to be:
-
•
wandering if for all ;
-
•
recurrent if ;
-
•
infinitely recurrent if .
The inclusions above are understood to hold up to a null set. The transformation is:
-
•
dissipative if the phase space is a countable union of wandering sets;
-
•
conservative if there are no wandering sets of positive measure;
-
•
recurrent if every set of positive measure is recurrent;
-
•
infinitely recurrent if every set of positive measure is infinitely recurrent.
It turns out that the properties of being conservative, recurrent, and infinitely recurrent are all mutually equivalent.
Proposition B.2.
Let be a non-singular transformation of a -finite measure space . The following are equivalent:
-
(1)
is conservative;
-
(2)
is recurrent;
-
(3)
is infinitely recurrent.
Among the properties introduced in Definition B.1, only recurrence and dissipativity are therefore different. In fact, any non-singular transformation admits a canonical decomposition, known as the Hopf decomposition, into these two types of action.
Proposition B.3 (Hopf decomposition).
Let be a non-singular transformation of a -finite measure space . There exists an -invariant partition such that is dissipative and is recurrent (equivalently, conservative). Moreover, if is another partition with this property then and .
We also note the following consequence of dissipativity in case the measure is preserved.
Proposition B.4.
Let be a measure-preserving transformation of a -finite measure space and let be its Hopf decomposition. For every set of finite measure, almost every point in eventually escapes :
Proof.
We may as well assume . Let have finite measure. Let be a wandering set whose translates cover . Consider the map which maps to , and observe that is measure-preserving if we endow with the product of the measure induced by on and the counting measure on .
So if there is a positive measure set of such that for infinitely many , by Fubini’s theorem we would have that has infinite measure, a contradiction. The same conclusion is true if we replace by any of its -translates, and since these translates cover the proof is finished. ∎
Appendix C Actions of locally compact Polish groups
In this chapter of the appendix, we collect some well-known facts related to actions of locally compact Polish groups. This exposition is provided for reader’s convenience and completeness. We recall that by a result of G. W. Mackey [Mac62], any Boolean measure-preserving action of a locally compact Polish group can be realized as a spatial Borel action, so we may switch to pointwise formulations, whenever convenient for the exposition.
C.1. Disintegration of measure
Let be a smooth measurable equivalence relation on a standard Lebesgue space , and let be a measurable reduction to the identity relation on some standard Lebesgue space , if and only if . Suppose that is a -finite measure on that is equivalent to the push-forward . A disintegration of relative to is a collection of measures on such that for all Borel sets
-
(1)
for -almost all ;
-
(2)
the map is measurable;
-
(3)
.
A theorem of D. Maharam [Mah50] asserts that can be disintegrated over any as above. In fact, existence of a disintegration can be proved in a setup considerably more general (see, for example, D. H. Fremlin [Fre06, Thm. 452I]), but in the framework of standard Lebesgue spaces, disintegration is essentially unique. While the context of our work is purely ergodic theoretical, we note that the disintegration result holds in the descriptive set theoretical setting as well, as discussed in [Mah84] and [GM89]. Without striving for generality, we formulate here a specific version, which suits our needs.
Theorem C.1 (Disintegration of Measure).
Let be a standard Lebesgue space, be a -finite standard Lebesgue space, and let be a measurable function. If is equivalent to , then there exists a disintegration of over . Moreover, such a disintegration is essentially unique in the sense that if is another disintegration, then for -almost all .
Remark C.2.
It is more common to formulate the disintegration theorem with the assumption that , when one can additionally ensure that for -almost all . Weakening the equality to mere equivalence is a simple consequence, for if is a disintegration of over , then is a disintegration of over .
Let be the set of atoms of the disintegration, i.e., , and let be the equivalence relation on , where two atoms within the same fiber are equivalent whenever they have the same measure: if and only if and . The equivalence relation is measurable and has finite classes -almost surely. Let , , be the union of -equivalence classes of size exactly , thus . Set also to be the atomless part of the disintegration and let denote the restriction of onto .
Consider the group of measure-preserving automorphisms for which holds -almost surely. Every preserves -almost all measures , since is a disintegration of , which has to coincide with by uniqueness of the disintegration. In particular, the partition is invariant under the full group , and for any the restriction for every . Conversely, for a sequence , , one has . We therefore have an isomorphism of (abstract) groups .
The groups can be described quite explicitly. First, let us consider the case , thus . All equivalence classes of the restriction of onto have size . Let be a measurable transversal, i.e., a measurable set intersecting every -class in a single point, and let . Every produces a permutation of -almost every -class, so we can view it as an element of , where is the group of permutations of an -element set. This identification works in both directions and produces an isomorphism . Note also that all are atomless if so is . We allow for , in which case is the trivial group.
Let us now go back to the equivalence relation , and recall that measures are atomless. Let be the encoding of fibers with non-trivial atomless components and put . In particular, for every the space is isomorphic to the interval endowed with the Lebesgue measure. In fact, one can select such isomorphisms in a measurable way across all . More precisely, there is a measurable isomorphism such that for all
-
•
;
-
•
coincides with the Lebesgue measure on .
The reader may find further details in [GM89, Thm. 2.3], where the same construction is discussed in a more refined setting of Borel disintegrations.
Using the isomorphism , we can identify each , , with . Since every preserves -almost every , we may rescale these intervals and view any as an element of . Conversely, every gives rise to via the notationally convoluted but natural
which, in plain words, simply applies upon the corresponding fiber identified with using . This map is an isomorphism between the groups and .
Let us say that has atomless classes if is atomless -almost surely or, equivalently, in the notation above. We may summarize the discussion so far into the following proposition.
Proposition C.3.
Let be a smooth measurable equivalence relation on a standard Lebesgue space . There are (possibly empty) standard Lebesgue spaces , , such that the full group is (abstractly) isomorphic to
where is the group of permutations of a -element set. If is atomless, then so are the spaces , . If has atomless classes, then all , , are negligible and is isomorphic to .
We can further refine the product in Proposition C.3 by decomposing the spaces into individual atoms and the atomless remainders. More specifically, let be a standard Lebesgue space and be a Polish group. Given a measurable partition , every function can be associated with a pair , and , which is an isomorphism of the topological groups. The same consideration applies to finite or countably infinite partitions.
Proposition C.4.
Let be a standard Lebesgue space and be a Polish group. For any finite or countably infinite measurable partition , there is an isomorphism of topological groups and , where is the restriction of onto .
Applying Proposition C.4 to the partition of into the atomless part and individual atoms (if any), and noting that for a singleton the group is naturally isomorphic to , we get the following corollary.
Corollary C.5.
Let be a standard Lebesgue space and be a Polish group. Let be the set of atoms of and be the atomless part. The group is isomorphic to .
Combining the discussion above with Proposition C.3, we obtain a very concrete representation for . In the formulation below, is understood to be the trivial group.
Proposition C.6.
Let be a smooth measurable equivalence relation on a standard Lebesgue space . There are cardinals and such that
If is atomless, then for all ; if has atomless classes, then for all .
So far we viewed as an abstract group. This is because neither of the two natural topologies on play well with the full group construction— is generally not closed in the weak topology, and not separable in the uniform topology whenever . Nonetheless, the isomorphism given in Proposition C.3 shows that there is a natural Polish topology on , which arises when we view groups and as Polish groups in the topology of convergence in measure. It is with respect to this topology we formulate Proposition C.7.
Proposition C.7.
Let be a smooth measurable equivalence relation on a standard Lebesgue space . The set of periodic elements is dense in .
Proof.
Rokhlin’s Lemma implies that any can be approximated in the uniform topology by periodic elements from . Since the uniform topology is stronger than the Polish topology on , the proposition follows. ∎
C.2. Tessellations
An important feature of locally compact group actions is the fact that they all admit measurable cross-sections. This was proved by J. Feldman, P. Hahn, and C. Moore in [FHM78], whereas a Borel version of the result was obtained by A. S. Kechris in [Kec92].
Definition C.8.
Let be a Borel action of a locally compact Polish group. A cross-section is a Borel set which is both
-
•
a complete section for : it intersects every orbit of the action and
-
•
lacunary: for some neighborhood of the identity one has for all distinct .
A cross-section is -cocompact, where is a compact set, if ; a cross-section is cocompact if it is -cocompact for some compact .
Any action admits a -cocompact cross-section, whenever is a compact neighborhood of the identity (see [Slu17, Thm. 2.4]). We also remind the following well-known lemma on the possibility to partition a cross-section into pieces with a prescribed lacunarity parameter.
Lemma C.9.
Let be a Borel action of a locally compact Polish group and be a cross-section for the action. For any compact neighborhood of the identity , there exists a finite Borel partition such that each is -lacunary.
Proof.
Set and let be a compact neighborhood of the identity small enough for to be -lacunary. Define a binary relation on by declaring whenever and . Note that is symmetric since so is . We view as a Borel graph on and claim that it is locally finite. More specifically, if is a right Haar measure, then the degree of each is at most .
Indeed, let be distinct elements such that for all ; in particular for . Let be such that . Lacunarity of asserts that sets are supposed to be pairwise disjoint, which necessitates to be pairwise disjoint for . Clearly as . Using the right-invariance of , we have , and thus , as claimed.
We may now use [KST99, Prop. 4.6] to deduce existence of a finite partition such that no two points in are adjacent. In other words, if are distinct, then , and therefore , which shows that are -lacunary. ∎
Every cross-section gives rise to a smooth subrelation of by associating to “the closest point” of in the same orbit. Such a subrelation is known as the Voronoi tessellation. For the purposes of Chapter 5, we need a slightly more abstract concept of a tessellation which may not correspond to Voronoi domains. While far from being the most general, the following treatment is sufficient for our needs.
Definition C.10.
Let be a Borel action of a locally compact Polish group on a standard Borel space and let be a cross-section. A tessellation over is a Borel set such that
-
(1)
all fibers are pairwise disjoint for ;
-
(2)
for all elements of are -equivalent to , i.e., ;
-
(3)
fibers cover the phase space, .
A tessellation over is -lacunary for an open if
It is -cocompact, , if .
Any tessellation can be viewed as a (flipped) graph of a function, since for any there is a unique such that . We denote such by , which produces a Borel map . There is a natural equivalence relation associated with the tessellation. Namely, and are -equivalent whenever they belong to the same fiber, i.e., . In view of the item (2), and moreover, every -class consists of countably many -classes.
Voronoi tessellations provide a specific way of constructing tessellations over a given cross-section. Suppose that the group is endowed with a compatible proper norm . Let be the associated metric on the orbits of the action (as in Section 2.2) and let be a Borel linear order on . The Voronoi tessellation over the cross-section relative to a proper norm is the set defined by
Properness of the norm ensures that for each there are only finitely many candidates which minimize , and hence each is associated with a unique .
For the sake of Chapter 5, we need a definition of the Voronoi tessellation for norms that may not be proper. The set specified as above may in this case fail to satisfy item (3) of the definition of a tessellation, as for some there may be infinitely many that minimize , none of which are -minimal. We therefore need a different way to resolve the points on the “boundary” between the regions, which can be done, for example, by delegating this task to a proper norm.
Definition C.11.
Let be a compatible norm on and let be a cross-section. Pick a compatible proper norm on and a Borel linear order on . Let and be the metrics on orbits of the action associated with the norms and respectively. The Voronoi tessellation over the cross-section relative to the norm is the set defined by
The definition of the Voronoi tessellation does depend on the choice of the norm and the linear order on the cross-section, but its key properties are the same regardless of these choices. We therefore often do not specify explicitly which and are picked. Note also that if the cross-section is cocompact, then every region of the Voronoi tessellation is bounded, i.e., .
Our goal is to show that equivalence relations are atomless in the sense of Section C.1 as long as each orbit of the action is uncountable. To this end we first need the following lemma.
Lemma C.12.
Let be a locally compact Polish group acting on a standard Lebesgue space by measure-preserving automorphisms. Suppose that almost every orbit of the action is uncountable. If is a measurable set such that the intersection of with almost every orbit is countable, then .
Proof.
Pick a proper norm on , let be a cross-section for the action, let be an open ball around the identity of sufficiently small radius such that whenever are distinct, and let be the Voronoi tessellation over relative to . Note that is fully contained in the -class of and set . Let also be a countable dense subset of .
We claim that it is enough to consider the case when intersects each -class in at most one point. Indeed, the restriction of onto is a smooth countable equivalence relation, so one can write , where each intersects each -class in at most one point. To simplify notations, we assume that already possesses this property.
Let be defined by . Let and note that sets partition . It is therefore enough to show that for any . Pick . The action is measure-preserving and therefore . Set and note that for any and one has . If the action were free, we could easily conclude that , since sets , , would be pairwise disjoint. In general, we need to exhibit a little more care and construct a countable family of pairwise disjoint sets as follows.
For let . The value is well-defined because the stabilizer of is closed and must be nowhere dense in due to the orbit being uncountable. Put and note that . We get a pairwise disjoint infinite family of sets all having the same measure. Since is finite, we conclude that and the lemma follows. ∎
Corollary C.13.
Let be a locally compact Polish group acting on a standard Lebesgue space by measure-preserving automorphisms, let be a cross-section for the action and let be a tessellation. If -almost every orbit of is uncountable, then is atomless.
Proof.
Consider the disintegration of relative to , where and . Let be the set of atoms of the disintegration. Since -almost every is finite, fibers have countably many atoms. Since every tessellation has only countably many tiles within each orbit, we conclude that has countable intersection with almost every orbit of the action. Lemma C.12 applies and shows that , hence is atomless as required. ∎
Consider the full group which by Proposition C.3 and Corollary C.13 is isomorphic to for some standard Lebesgue space . This full group can naturally be viewed as a subgroup of and the topology induced on from the full group coincides with the topology of converges in measure on (see Section 3 of [CLM16]). We therefore have the following corollary.
Corollary C.14.
Let be a locally compact Polish group acting on a standard Lebesgue space by measure-preserving automorphisms, let be a cross-section for the action and let be a tessellation and be the corresponding reduction. If -almost every orbit of is uncountable, then the subgroup is isomorphic as a topological group to . If moreover all orbits of the action have measure zero, then is non-atomic and is isomorphic to .
C.3. Ergodic decomposition
Let be a free measure-preserving action of a locally compact group on a standard probability space . The space of invariant ergodic probability measures of this action possesses a structure of a standard Borel space. The Ergodic Decomposition theorem of V. S. Varadarajan [Var63, Thm. 4.2] asserts that there is an essentially unique Borel -invariant surjection and a probability measure on such that in the sense that for all Borel one has .
There is a one-to-one correspondence between measurable -invariant functions and measurable functions given by . For measures and as above, this correspondence gives an isometric isomorphism between and the subspace of that consists of -invariant functions.
Appendix D Conditional measures
The ergodic decomposition theorem, as formulated in Section C.3, is not available for general probability measure-preserving actions of Polish groups. Conditional measures provide a useful framework to remedy this. As before, stands for the group of measure-preserving automorphisms of a standard probability space. It is more useful, however, to view as the group of measure-preserving automorphisms of the measure algebra of , i.e., is the Boolean algebra of equivalence classes of Borel subsets of , identified up to measure zero. The measure algebra is endowed with a natural metric given by . Completeness of in this metric is a standard and well-known fact (see, for instance, [Kec95, Exer. 17.43]), which we include for reader’s convenience.
Proposition D.1.
The metric space is complete.
Proof.
It suffices to show that a Cauchy sequence admits a converging subsequence. Passing to a subsequence if necessary, we may assume that holds for all , and therefore . The set
is the limit we seek. Indeed, given an and an index chosen so large that , for all and all outside of the set of measure at most , we have if and only if . ∎
Note that closed (or equivalently, metrically complete) subalgebras of are in a natural one-to-one correspondence with complete (in the measure-theoretical sense) -subalgebras of the -algebra of Lebesgue measurable sets.
D.1. Conditional expectations
We review here how conditional expectations can easily be defined without appealing to disintegration.
Let be a closed subalgebra of and let denote the space of real-valued -measurable functions. Note that is a closed subspace of . The -conditional expectation is the orthogonal projection . It is also uniquely defined by the condition
(D.1) |
By the density of step functions in , the conditional expectation can equivalently be defined as the linear contraction satisfying
(D.2) |
Positive functions are exactly those whose dot product with any characteristic function is positive. Letting in Eq. (D.1) range through the collection of all characteristic functions of sets in shows that the conditional expectation is positivity-preserving.
Proposition D.2.
If is non-negative, , then .
While we defined conditional expectations as operators on , their domain can be extended to all of , making a contraction from to . This is justified by the following proposition.
Proposition D.3.
The conditional expectation is a contraction when the domain and the range are endowed with the norms.
Proof.
If is non-negative, , then Eq. D.1 yields
Since by Proposition D.2, we conclude that for all non-negative .
For an arbitrary , set and . Note that functions are non-negative and belong to . Furthermore, and . We therefore have
but the latter term is equal to , which finishes the proof. ∎
Remark D.4.
By the previous proposition, admits a (necessarily unique) extension to a contraction
Moreover, since every non-negative integrable function can be written as an increasing limit of bounded non-negative functions, the analog of Proposition D.2 continues to hold for .
D.2. Conditional measures
Throughout this section, we let denote the characteristic function of .
Definition D.5.
Let be a closed subalgebra of . The -conditional measure of , denoted by , is the conditional expectation of the characteristic function of , i.e., .
In particular, the conditional measure is an -measurable function. It enjoys the following natural properties.
Proposition D.6.
Let be a closed subalgebra. The following properties hold for all :
-
(1)
and , where and denote the constant maps;
-
(2)
takes values in and ;
-
(3)
is -additive: if , , is a partition then
where the convergence holds in ;
-
(4)
if fixes every element of , then .
Proof.
The first item is clear from the fact that both and belong to , so their characteristic functions are fixed by . The second item follows from the first and positivity of the conditional expectation; the equality is a direct consequence of Eq. D.2. The third one is a consequence of the continuity of and its linearity, noting that in .
Finally, the last item follows from the uniqueness of conditional expectation given by Eq. D.2. Indeed, if an automorphism fixes every element of , then
so . Taking for , we conclude that . ∎
D.3. Conditional measures and full groups
Conditional measures, as defined in Section D.2, are associated with closed subalgebras of . Each subgroup gives rise to the subalgebra of -invariant sets, and we may therefore associate a conditional measure with the group itself.
Definition D.7.
Let be a subgroup of . The closed subalgebra of -invariant sets is denoted by and consists of all such that for all .
By definition, is ergodic if . In this case, the -conditional measure is the measure itself. The following lemma is an easy consequence of the definitions of the full group generated by a subgroup (Section 3.1) and the weak topology on .
Lemma D.8.
Let be a group.
-
(1)
If is the full group generated by , then .
-
(2)
If is dense in the weak topology, then .
Given a subgroup , we denote the -conditional measure simply by . Note that is ergodic if and only if .
Recall that a partial measure-preserving automorphism of is a measure-preserving bijection between measurable subsets of , called the domain and the range of , respectively. The pseudo full group generated by a group is denoted by and consists of all partial automorphisms for which there exists a partition and elements such that for all . Elements of automatically preserve the conditional measure in view of item (4) of Proposition D.6.
Lemma D.9.
Let be a group and let satisfy . There exists an element such that and .
Proof.
Let be a countable weakly dense subgroup of . Note that by Lemma D.8, and also clearly .
We define inductively sequences and of subsets of and respectively by setting and , and then putting for
By construction, the sets are pairwise disjoint subsets of , , and the sets are pairwise disjoint subsets of . We claim that is the desired element of .
Suppose towards a contradiction that either or . Since preserves and , the sets and have the same -conditional measure, which is not constantly equal to zero. The set is -invariant and non zero. Its conditional measure is therefore the characteristic function , which must be greater than or equal to . We conclude that . In particular, there is the first such that is non zero. By construction, this set should be a subset of , yielding the desired contradiction. ∎
Proposition D.10.
Let be a full subgroup of . The following conditions are equivalent for all :
-
(1)
;
-
(2)
there is such that .
-
(3)
there is an involution such that and .
D.4. Aperiodicity
A countable subgroup is called aperiodic if almost all the orbits of some (equivalently, any) realization of its action on are infinite. The so-called Maharam’s lemma provides a characterization of aperiodicity in a purely measure-algebraic way. We begin by formulating a variant of the standard Marker Lemma for countable Borel equivalence relations (see, for instance, [KM04, Lemma 6.7]).
Lemma D.11.
Let be a Borel action of a countable group on a standard Borel space . For every Borel , there is a decreasing sequence of Borel subsets of such that for each , and the set intersects the -orbit of every in at most one point. Furthermore, if all orbits of are infinite, sets can be chosen to have the empty intersection, .
The following result is essentially due to H. Dye [Dye59], where it is called Maharam’s lemma.
Theorem D.12 (Maharam’s lemma).
Let be a countable subgroup. The following are equivalent:
-
(1)
is aperiodic;
-
(2)
for any and any -measurable function satisfying , there is , , such that .
Proof.
Let us begin with the easier (2)(1), which is proved by the contrapositive. Assume that (1) does not hold and is not aperiodic. Let be such that the -invariant set has non-zero measure. We may assume that bears a Borel total order (for instance, by identifying with ). Let be the set of maximal points of the -element -orbits and set , , to be the element of the pseudo full group that takes every to its -successor in the orbit . Given any , the set is -invariant, hence takes values in . Also
where the last equality is a consequence of Proposition D.6. We conclude that necessarily takes values in , which contradicts (2).
We now assume that is aperiodic and prove the direct implication (1)(2). The argument is based on the following crucial claim.
Claim.
For every , for every -measurable not almost surely zero such that , there is a non zero satisfying .
Proof of the claim.
Let be a vanishing sequence of subsets of given by Lemma D.11. Note that in , since and the ’s are decreasing. Passing to a subsequence, we may assume that convergence holds pointwise. Set and note that and therefore .
Pointwise convergence guarantees existence of an index such that , and so the set is as required. ∎
The conclusion of the theorem now follows from a standard application of Zorn’s lemma111A more constructive version of the whole argument can be found in [LM14, Prop. D.1].. Indeed, the latter provides a maximal family of pairwise disjoint positive measure elements of contained in and satisfying . The index set has to be countable, and if then . Assume towards a contradiction that is not equal to almost everywhere, and use the previous claim to get a non null with , contradicting the maximality of . Therefore as claimed. ∎
We conclude this appendix with a useful consequence of aperiodicity.
Lemma D.13.
Let be an aperiodic full group. For each set , there is an involution whose support is equal to .
Proof.
References
- [ADR00] C. Anantharaman-Delaroche and J. Renault, Amenable groupoids, Monographies de L’Enseignement Mathématique [Monographs of L’Enseignement Mathématique], vol. 36, L’Enseignement Mathématique, Geneva, 2000, With a foreword by Georges Skandalis and Appendix B by E. Germain. MR 1799683
- [Aus16] Tim Austin, Behaviour of Entropy Under Bounded and Integrable Orbit Equivalence, Geometric and Functional Analysis 26 (2016), no. 6, 1483–1525.
- [Ban32] Stefan Banach, Théorie des opérations linéaires, Warsaw, 1932.
- [BdlHV08] Bachir Bekka, Pierre de la Harpe, and Alain Valette, Kazhdan’s property (T), vol. 11, Cambridge University Press, Cambridge, 2008 (English).
- [Bec13] Howard Becker, Cocycles and continuity, Trans. Amer. Math. Soc. 365 (2013), no. 2, 671–719. MR 2995370
- [Bel68] R. M. Belinskaja, Partitionings of a Lebesgue space into trajectories which may be defined by ergodic automorphisms, Funkcional. Anal. i Priložen. 2 (1968), no. 3, 4–16. MR 0245756
- [BK96] Howard Becker and Alexander S. Kechris, The descriptive set theory of Polish group actions, London Mathematical Society Lecture Note Series, vol. 232, Cambridge University Press, Cambridge, 1996. MR 1425877
- [BO10] Nicholas H. Bingham and Adam J. Ostaszewski, Normed versus topological groups: dichotomy and duality, Dissertationes Math. 472 (2010), 138. MR 2743093
- [CFW81] Alain Connes, Jacob Feldman, and Benjamin Weiss, An amenable equivalence relation is generated by a single transformation, Ergodic Theory Dynam. Systems 1 (1981), no. 4, 431–450 (1982). MR 662736
- [CJMT22] Alessandro Carderi, Matthieu Joseph, François Le Maître, and Romain Tessera, Belinskaya’s theorem is optimal, 2022.
- [CLM16] Alessandro Carderi and François Le Maître, More Polish full groups, Topology Appl. 202 (2016), 80–105. MR 3464151
- [CLM18] by same author, Orbit full groups for locally compact groups, Trans. Amer. Math. Soc. 370 (2018), no. 4, 2321–2349. MR 3748570
- [DHU08] Manfred Droste, W. Charles Holland, and Georg Ulbrich, On full groups of measure-preserving and ergodic transformations with uncountable cofinalities, Bulletin of the London Mathematical Society 40 (2008), no. 3, 463–472.
- [Dye59] Henry A. Dye, On Groups of Measure Preserving Transformations. I, American Journal of Mathematics 81 (1959), no. 1, 119–159.
- [FHM78] Jacob Feldman, Peter Hahn, and Calvin C. Moore, Orbit structure and countable sections for actions of continuous groups, Adv. in Math. 28 (1978), no. 3, 186–230. MR 492061
- [Fre04] David H. Fremlin, Measure theory. Vol. 3, Torres Fremlin, Colchester, 2004, Measure algebras, Corrected second printing of the 2002 original. MR 2459668 (2011a:28003)
- [Fre06] by same author, Measure theory. Vol. 4, Torres Fremlin, Colchester, 2006, Topological measure spaces. Part I, II, Corrected second printing of the 2003 original. MR 2462372
- [GK21] Marlies Gerber and Philipp Kunde, Anti-classification results for the Kakutani equivalence relation, September 2021.
- [GM89] Siegfried Graf and R. Daniel Mauldin, A classification of disintegrations of measures, Measure and measurable dynamics (Rochester, NY, 1987), Contemp. Math., vol. 94, Amer. Math. Soc., Providence, RI, 1989, pp. 147–158. MR 1012985
- [GP02] Thierry Giordano and Vladimir Pestov, Some extremely amenable groups, C. R. Math. Acad. Sci. Paris 334 (2002), no. 4, 273–278. MR 1891002
- [GPS99] Thierry Giordano, Ian F. Putnam, and Christian F. Skau, Full groups of Cantor minimal systems, Israel J. Math. 111 (1999), 285–320. MR 1710743
- [GTW05] Eli Glasner, Boris Tsirelson, and Benjamin Weiss, The automorphism group of the Gaussian measure cannot act pointwise, Israel Journal of Mathematics 148 (2005), no. 1, 305–329.
- [Hal17] Paul R. Halmos, Lectures on ergodic theory, Dover Publications, Mineola, NY, 2017.
- [Kat75] Anatole B. Katok, Time change, monotone equivalence, and standard dynamical systems, Doklady Akademii Nauk SSSR 223 (1975), no. 4, 789–792. MR 0412383
- [Kat77] by same author, Monotone equivalence in ergodic theory, Izvestiya Akademii Nauk SSSR. Seriya Matematicheskaya 41 (1977), no. 1, 104–157, 231. MR 0442195
- [Kec92] Alexander S. Kechris, Countable sections for locally compact group actions, Ergodic Theory Dynam. Systems 12 (1992), no. 2, 283–295. MR 1176624
- [Kec95] by same author, Classical descriptive set theory, Graduate Texts in Mathematics, vol. 156, Springer-Verlag, New York, 1995.
- [Kec10] by same author, Global aspects of ergodic group actions, Mathematical Surveys and Monographs, vol. 160, American Mathematical Society, Providence, RI, 2010. MR 2583950
- [KM04] Alexander S. Kechris and Benjamin D. Miller, Topics in orbit equivalence, Lecture Notes in Mathematics, no. 1852, Springer, Berlin ; New York, 2004.
- [KPV15] David Kyed, Henrik Petersen, and Stefaan Vaes, L2-Betti numbers of locally compact groups and their cross section equivalence relations, Trans. Amer. Math. Soc. 367 (2015), no. 7, 4917–4956.
- [Kre85] Ulrich Krengel, Ergodic theorems, De Gruyter Studies in Mathematics, vol. 6, Walter de Gruyter & Co., Berlin, 1985, With a supplement by Antoine Brunel. MR 797411
- [KST99] Alexander S. Kechris, Sławomir Solecki, and Stevo Todorcevic, Borel chromatic numbers, Adv. Math. 141 (1999), no. 1, 1–44. MR 1667145
- [Kun23] Philipp Kunde, Anti-classification results for weakly mixing diffeomorphisms, March 2023.
- [LM14] François Le Maître, Sur les groupes pleins préservant une mesure de probabilité, Ph.D. thesis, ENS Lyon, 2014.
- [LM16] François Le Maître, On full groups of non-ergodic probability-measure-preserving equivalence relations, Ergodic Theory Dynam. Systems 36 (2016), no. 7, 2218–2245. MR 3568978
- [LM18] by same author, On a measurable analogue of small topological full groups, Adv. Math. 332 (2018), 235–286. MR 3810253
- [LM21] by same author, On a measurable analogue of small topological full groups II, Ann. Inst. Fourier (Grenoble) 71 (2021), no. 5, 1885–1927. MR 4398251
- [Mac62] George W. Mackey, Point realizations of transformation groups, Illinois J. Math. 6 (1962), 327–335. MR 143874
- [Mah50] Dorothy Maharam, Decompositions of measure algebras and spaces, Trans. Amer. Math. Soc. 69 (1950), 142–160. MR 36817
- [Mah84] by same author, On the planar representation of a measurable subfield, Measure theory, Oberwolfach 1983 (Oberwolfach, 1983), Lecture Notes in Math., vol. 1089, Springer, Berlin, 1984, pp. 47–57. MR 786682
- [Mil77] Douglas E. Miller, On the measurability of orbits in Borel actions, Proc. Amer. Math. Soc. 63 (1977), no. 1, 165–170. MR 440519
- [Mil04] Benjamin D. Miller, Full groups, classification, and equivalence relations, Ph.D. thesis, University of California, Berkeley, 2004.
- [Nac65] Leopoldo Nachbin, The Haar integral, D. Van Nostrand Co., Inc., Princeton, N.J.-Toronto-London, 1965. MR 0175995
- [Nek19] Volodymyr Nekrashevych, Simple groups of dynamical origin, Ergodic Theory Dynam. Systems 39 (2019), no. 3, 707–732. MR 3904185
- [ORW82] Donald S. Ornstein, Daniel J. Rudolph, and Benjamin Weiss, Equivalence of measure preserving transformations, Memoirs of the American Mathematical Society 37 (1982), no. 262, xii+116. MR 653094
- [PS17] Vladimir G. Pestov and Friedrich Martin Schneider, On amenability and groups of measurable maps, J. Funct. Anal. 273 (2017), no. 12, 3859–3874. MR 3711882
- [Ros21] Christian Rosendal, Coarse Geometry of Topological Groups, Cambridge Tracts in Mathematics, Cambridge University Press, Cambridge, 2021.
- [Ros22] by same author, Coarse geometry of topological groups, Cambridge Tracts in Mathematics, vol. 223, Cambridge University Press, Cambridge, 2022. MR 4327092
- [RS98] Guyan Robertson and Tim Steger, Negative definite kernels and a dynamical characterization of property (T) for countable groups, Ergodic Theory and Dynamical Systems 18 (1998), no. 1, 247–253 (en).
- [Rud76] Daniel Rudolph, A two-valued step coding for ergodic flows, Mathematische Zeitschrift 150 (1976), no. 3, 201–220.
- [Ryz85] V. V. Ryzhikov, Representation of transformations preserving the lebesgue measure in the form of periodic transformations, Mathematical notes of the Academy of Sciences of the USSR 38 (1985), no. 6, 978–981.
- [Slu17] Konstantin Slutsky, Lebesgue orbit equivalence of multidimensional Borel flows: a picturebook of tilings, Ergodic Theory Dynam. Systems 37 (2017), no. 6, 1966–1996. MR 3681992
- [Slu19] by same author, Regular cross sections of Borel flows, Journal of the European Mathematical Society 21 (2019), no. 7, 1985–2050.
- [Str74] Raimond A. Struble, Metrics in locally compact groups, Compositio Math. 28 (1974), 217–222. MR 348037
- [Var63] Veeravalli S. Varadarajan, Groups of automorphisms of Borel spaces, Trans. Amer. Math. Soc. 109 (1963), 191–220. MR 159923