Katok’s special representation theorem
for multidimensional Borel flows
Abstract.
Katok’s special representation theorem states that any free ergodic measure-preserving -flow can be realized as a special flow over a -action. It provides a multidimensional generalization of the “flow under a function” construction. We prove the analog of Katok’s theorem in the framework of Borel dynamics and show that, likewise, all free Borel -flows emerge from -actions through the special flow construction using bi-Lipschitz cocycles.
1. Introduction
1.1. Overview
Theorems of Ambrose and Kakutani [1, 2] established a connection between measure-preserving -actions and -flows by showing that any flow admits a cross-section and can be represented as a “flow under a function”. Their construction provides a foundation for the theory of Kakutani equivalence (also called monotone equivalence) [7, 11] on the one hand and study of the possible ceiling functions in the “flow under a function” representation [21, 16] on the other.
The intuitive geometric picture of a “flow under a function” does not generalize to -flows for . However, Katok [12] re-interpreted it in a way that can readily be adapted to the multidimensional setup, calling flows appearing in this construction special flows. Despite their name, they aren’t so special, since, as showed in the same paper, every free ergodic measure-preserving -flow is metrically isomorphic to a special flow. Just like the works of Ambrose and Kakutani, it opened gates for the study of multidimensional concepts of Kakutani equivalence [5] and stimulated research on tilings of flows[22, 15].
Borel dynamics as a separate field goes back to the work of Weiss [31] and has blossomed into a versatile branch of dynamical systems. The phase space here is a standard Borel space , i.e., a set with a -algebra of Borel sets for some Polish topology on . Some of the key ergodic theoretical results have their counterparts in Borel dynamics, while others do not generalize. For example, Borel version of the Ambrose–Kakutani Theorem on the existence of cross-sections in -flows was proved by Wagh in [30] showing that, just like in ergodic theory, all free Borel -flows emerge as “flows under a function” over Borel -actions. Likewise, Rudolph’s two-valued theorem [21] generalizes to the Borel framework [25]. The theory of Kakutani equivalence, on the other hand, exhibits a different phenomenon. While being a highly non-trivial equivalence relation among measure-preserving flows [20, 9], descriptive set theoretical version of Kakutani equivalence collapses entirely [19].
Considerable work has been done to understand the Borel dynamics of -flows, but relatively few things are known about multidimensional actions. This paper makes a contribution in this direction by showing that the analog of Katok’s special representation theorem does hold for free Borel -flows.
1.2. Structure of the paper
Constructions of orbit equivalent -actions often rely on (essential) hyperfiniteness and use covers of orbits of the flow by coherent and exhaustive regions. This is the case for the aforementioned paper of Katok [12], and related approaches have been used in the descriptive set theoretical setup as well (e.g., [26]). Particular assumptions on such coherent regions, however, depend on the specific application. Section 2 distills a general language of partial actions, in which many of the aforementioned constructions can be formulated. As an application we show that the orbit equivalence relation generated by a free -flow can also be generated by a free action of any non-discrete and non-compact Polish group (see Theorem 2.6). This is in a striking contrast with the actions of discrete groups, where a probability measure-preserving free -action can be generated only by a free action of an amenable group.
Section 3 does the technical work of constructing Lipschitz maps that are needed for Theorem 3.12, which shows, roughly speaking, that up to an arbitrarily small bi-Lipschitz perturbation, any free -flow admits an integer grid—a Borel cross-section invariant under the -action.
Finally, Section 4 discusses the descriptive set theoretical version of Katok’s special flow construction and shows in Theorem 4.3 that, indeed, any free -flow can be represented as a special flow generated by a bi-Lipschitz cocycle with Lipschitz constants arbitrarily close to . This provides a Borel version of Katok’s special representation theorem.
2. Sequences of partial actions
We begin by discussing the framework of partial actions suitable for constructing orbit equivalent actions. Throughout this section, denotes a standard Borel space.
2.1. Partial actions
Let be a standard Borel group, that is a group with a structure of a standard Borel space that makes group operations Borel. A partial -action111More precisely, we should call such a partial free action. Since we are mainly concerned with free actions in what follows, we choose to omit the adjective “free” in the definition. is a pair , where is a Borel equivalence relation on and is a Borel map that is injective on each -class: whenever . The map itself may occasionally be refer to as a partial action when the equivalence relation is clear from the context.
The motivation for the name comes from the following observation. Consider the set
Injectivity of on -classes ensures that for each and there is at most one such that . When such a exists, we say that the action of on is defined and set . Clearly, for all , thus ; also whenever all the terms are defined. The set is a graph of a total action if and only if for each and there does exist some such that ; in this case the orbit equivalence relation generated by the action coincides with .
Example 2.1.
An easy way of getting a partial action is by restricting a total one. Suppose we have a free Borel action with the corresponding orbit equivalence relation and suppose that a Borel equivalence sub-relation admits a Borel selector—a Borel -invariant map such that for all . If is the map specified uniquely by the condition , then is a partial -action.
Sub-relations as in Example 2.1 are often associated with cross-sections of actions of locally compact second countable (lcsc) groups.
2.2. Tessellations of lcsc group actions
Consider a free Borel action of a locally compact second countable group. A cross-section of the action is a Borel set that intersects every orbit in a countable non-empty set. A cross-section is
-
•
discrete if is finite for every and compact ;
-
•
-lacunary, where is a neighborhood of the identity, if for all ;
-
•
lacunary if it is -lacunary for some neighborhood of the identity ;
-
•
cocompact if for some compact .
Let be a lacunary cross-section for , which exists by [13, Corollary 1.2]. Any lcsc group admits a compatible left-invariant proper metric [28], and any left-invariant metric can be transferred to orbits due to freeness of the action via for the unique such that . One can now define the so-called Voronoi tessellation of orbits by associating with each the closest point of the cross-section as determined by . Properness of the metric ensures that, for a ball of radius , , and any , the set is finite. Indeed, there can be at most points in the intersection, where is a Haar measure on the group and is so small that whenever are distinct.
Small care needs to be taken to address the possibility of having several closest points. For example, one may pick a Borel linear order on and associated each with the smallest closest point in the cross-section (see [23, Section 4] or [17, Section B.2] for the specifics). This way we get a Borel equivalence relation whose equivalence classes are the cells of the Voronoi tessellation: if and only if .
Assumed freeness of the action allows for a natural identification of each Voronoi cell with a subset of the acting group via the map such that , which is exactly what the corresponding partial action from Example 2.1 does.
Our intention is to use partial actions to define total actions, and the example above may seem like going “in the wrong direction”. The point, however, is that once we have a partial action , we can compose it with an arbitrary Borel injection to get a different partial action . This pattern is typical in the sense that new partial actions are often constructed by modifying those obtained as restrictions of total actions.
2.3. Convergent sequences of partial actions
A total action can be defined whenever we have a sequence of partial actions that cohere in the appropriate sense. Let be a standard Borel group. A sequence , , of partial -actions on is said to be convergent if it satisfies the following properties:
-
•
monotonicity: equivalence relations form an increasing sequence, that is for all ;
-
•
coherence: for each the map is -invariant;
-
•
exhaustiveness: for all and all there exist and such that and .
With such a sequence one can associate a free Borel (left) action , called the limit of , whose graph is . Coherence ensures that the partial action defined by is an extension of the one given by . Indeed, if are such that , then also by monotonicity and, using coherence,
whence . If is an -class, and for some , then , so the image gets shifted on the right inside . If we want to build a right action of the group, then should be shifted on the left instead.
Finally, exhaustiveness guarantees that gets defined eventually: for all and there are and such that . It is straightforward to check that is a graph of a total Borel action . Equally easy is to check that the action is free, and its orbits are precisely the equivalence classes of .
This framework, general as it is, delegates most of the complexity to the construction of maps . Let us illustrate these concepts on essentially hyperfinite actions of lcsc groups.
2.4. Hyperfinite tessellations of lcsc group actions
In the context of Section 2.2, suppose that, furthermore, the restriction of the orbit equivalence relation onto the cross-section is hyperfinite, i.e., there is an increasing sequence of finite Borel equivalence relations on such that . We can use this sequence to define whenever , which yields an increasing sequence of Borel equivalence relations such that .
The equivalence relations admit Borel transversals, i.e., there are Borel sets that pick exactly one point from each -class. Just as in Section 2.2, we may define to be such an element that for the unique satisfying . This gives a convergent sequence of partial -actions whose limit is the original action .
2.5. Partial actions revisited
In practice, it is often more convenient to allow equivalence relations to be defined on proper subsets of . Let , , be Borel subsets, and suppose for each , is a Borel equivalence relation on . We say that the sequence is monotone if the following conditions are satisfied for all :
-
•
;
-
•
if then the whole -class of is in .
Partial action maps , where, as earlier, is a standard Borel group, need to satisfy the appropriate versions of coherence and exhaustiveness:
-
•
coherence: is -invariant for each ;
-
•
exhaustiveness: for each and there exist and such that , , and .
A sequence of partial -actions will be called convergent if it satisfies the above properties of monotonicity, coherence, and exhaustiveness. Note that the condition follows from exhaustiveness, so sets must cover all of .
Convergent sequences define total actions, which can be most easily seen by reducing this setup to the notationally simpler one given in Section 2.3. To this end, extend to the equivalence relation on all of by
and also extend to by setting for the maximal such that or if no such exists. It is straightforward to check that is a convergent sequence of partial -actions in the sense of Section 2.3. By the limit of the sequence of partial actions we mean the limit of as defined earlier.
Remark 2.2.
A variant of this generalized formulation, which we encounter in Proposition 2.4 below, occurs when sets are nested: . Monotonicity of equivalence relations then simplifies to and coherence becomes equivalent to the -invariant of maps .
As was mentioned above, it is easy to create new partial actions simply by composing a partial action with some Borel bijection (or if we choose to have values in a different group). However, an arbitrary bijection has no reasons to preserve coherence and extra care is necessary to maintain it.
Furthermore, in general we need to apply different modifications to different -classes, which naturally raises concern of how to ensure that construction is performed in a Borel way. In applications, the modification applied to an -class , usually depends on the “shape” of and the -classes it contains, but does not depend on other -classes. If there are only countably many such “configurations” of -classes, resulting partial actions will be Borel as long as we consistently apply the same modification whenever “configurations” are the same. This idea can be formalized as follows.
2.6. Rational sequences of partial actions
Let be a convergent sequence of partial actions on . For an -class , let denote the collection of -classes contained in . Given two -classes and , we denote by the existence for each of a bijection such that for all . Collection of images constitutes the “configuration” of referred to earlier.
We say that the sequence of partial actions is rational if for each there exists a Borel -invariant partition such that for each one has for all -classes .
Remark 2.3.
This concept of rationality applies verbatim to convergent sequences of partial actions as described in Section 2.5. One can check that such a sequence is rational if and only if the sequence is rational.
2.7. Generating the flow equivalence relation
As an application of the partial actions formalism, we show that any orbit equivalence relation given by a free Borel -flow can also be generated by a free action of any non-discrete and non-compact Polish group. For this we need the following representation of an -flow as a limit of partial -actions.
Proposition 2.4.
Any free Borel -flow on can be represented as a limit of a convergent rational sequence of partial -actions such that
-
(1)
both and are increasing: and ; (see Remark 2.2)
-
(2)
each -class contains finitely many -classes;
-
(3)
each -class has cardinality of continuum;
-
(4)
for each -class the set has cardinality of continuum.
Proof.
Any -flow admits a rational222Rationality of the cross-section here means that the distance between any two points of is a rational number. More generally, rationality of a cross-section for an -action means whenever for some . -lacunary cross-section (see [24, Section 2]), which we denote by . Let be the partial -action as defined in Section 2.2. If is an -class, then is an interval. For , let consist of those such that is at least away from the boundary points of . In other words, is obtained by shrinking the class by from each side.
The restriction of the orbit equivalence relation onto is hyperfinite. This fact is true in the much wider generality of actions of locally compact Abelian groups [4]. Specifically for -flows, is generated by the first return map—a Borel automorphism of that sends a point in to the next one according to the order of the -flow. The first return map is well defined and is invertible, except for the orbits, where happens to have the maximal or the minimal point. The latter part of the space evidently admits a Borel selector and is therefore smooth, hence won’t affect hyperfiniteness of the equivalence relation. It remains to recall the standard fact that orbit equivalence relations of -actions are hyperfinite (see, for instance, [6, Theorem 5.1]), and thus so is the restriction .
In particular, we can represent the -flow as the limit of a convergent sequence of partial actions as described in Section 2.4. Note that is necessarily rational by rationality of . Such a sequence satisfies items (2) and (3), but fails (4). We fix this by shrinking equivalence classes to achieve proper containment. Let be a strictly decreasing sequence of positive reals such that and . Put , where the union is taken over all -classes . Note that sets fail to cover , because the boundary points of any -class do not belong to any of . Put and let . Clearly, is an increasing sequence of Borel sets and .
Finally, set and to be . The sequence of partial -actions satisfies the conditions of the proposition. ∎
All non-smooth orbit equivalence relations produced by free Borel -flows are Borel isomorphic to each other [14, Theorem 3]. Theorem 2.6 will show that this orbit equivalence relation can also be generated by a free action of any non-compact and non-discrete Polish group.
Let be a group. We say that a set admits infinitely many disjoint right translates if there is a sequence of elements of such that for all .
Lemma 2.5.
Let be a non-compact Polish group. There exists a neighborhood of the identity such that for any finite the set admits infinitely many disjoint right translates.
Proof.
We begin with the following characterization of compactness established independently by Solecki [27, Lemma 1.2] and Uspenskij [29]: a Polish group is non-compact if and only if there exists a neighborhood of the identity such that for all finite . Let be a symmetric neighborhood of the identity such that . We claim that such a set has the desired property. Pick a finite , set and choose inductively as follows. Let and . The defining property of assures existence of . Translates are then pairwise disjoint, for if for , then , contradicting the construction. ∎
Theorem 2.6.
Let be an orbit equivalence relation given by a free Borel -flow on . Any non-discrete non-compact Polish group admits a free Borel action such that .
Proof.
Let be a convergent sequence of partial -actions as in Proposition 2.4 and let be given by Lemma 2.5. Choose a countable dense in so that . Since the sequence of partial -actions is rational, one may pick for each a Borel -invariant partition such that for all -classes . We construct a convergent sequence of partial -actions such that for each and there exists a finite set such that and for all -classes .
For any -class , both and are Borel sets of the same cardinality. We may therefore pick a Borel bijection where . For the base of the inductive construction we set . Suppose that , , have been constructed.
We now construct . Let be an -class and let be a complete list of -classes contained in . By the inductive assumption, there are finite sets such that . Let . By the choice of , there are elements such that , are pairwise disjoint for . Pick a finite large enough that , , and has cardinality of continuum (the latter can be achieved, for instance, by assuring that one more disjoint translate of is inside ). Note that has cardinality of continuum by the properties guaranteed by Proposition 2.4. Pick any Borel bijection
and define by the conditions and . Just as in the base case, the same modification works for all classes -classes such that , which ensures Borelness of the construction.
It is now easy to check that is a convergent sequence of partial -actions, hence its limit is a free Borel action such that . ∎
Remark 2.7.
Theorem 2.6 highlights difference with actions of discrete groups, since a free Borel -action that preserves a finite measure cannot be generated by a free Borel action of a non-amenable group (see, for instance, [32, Proposition 4.3.3] or [10, Proposition 2.5(ii)]).
However, if we consider hyperfinite equivalence relations without any finite invariant measures, then we do have the analog for -actions. There exists a unique up to isomorphism non-smooth hyperfinite Borel equivalence relation without any finite invariant measures and it can be realized as an orbit equivalence relation of a free action of any infinite countable group [6, Proposition 11.2].
3. Lipschitz Maps
Our goal in this section is to prove Theorem 3.12, which shows that any free Borel -flow is bi-Lipschitz orbit equivalent to a flow with an integer grid. Sections 3.1–3.3 build the necessary tools to construct such an orbit equivalence. Verification of the Lipschitz conditions stated in the lemmas within these sections is straightforward and routine. We therefore omit the arguments in the interest of brevity.
Recall that a map between metric spaces and is -Lipschitz if for all , and it is -bi-Lipschitz if is injective, -Lipschitz, and is -Lipschitz, which can equivalently be stated as
The Lipschitz constant of a Lipschitz map is the smallest with respect to which is -Lipschitz.
3.1. Linked sets
Given two Lipschitz maps and that agree on the intersection , the map , in general, may not be Lipschitz. The following condition is sufficient to ensure that is Lipschitz with the Lipschitz constant bounded by the maximum of the constants of and .
Definition 3.1.
Let be a metric space and be its subsets. We say that and are linked if for all and there exists such that .
Lemma 3.2.
Let be a metric space, , be -Lipschitz maps between subsets of and suppose that . If and are linked, then is -Lipschitz.
Recall that a metric space is geodesic if for all points there exists a geodesic between them—an isometry such that and . For geodesic metric spaces, closed sets are always linked whenever the boundary of one of them is contained in the other. The boundary of a set will be denoted by , and will stand for the interior of .
Lemma 3.3.
Suppose is a geodesic metric space. If are closed and satisfy , then and are linked.
3.2. Inductive step
The following lemma encompasses the inductive step in the construction of the forthcoming Theorem 3.12.
Lemma 3.4.
Let be a geodesic metric space and be a closed set. Suppose are pairwise disjoint closed subsets of and are -bi-Lipschitz maps such that is the identity map for each . The map given by
is -bi-Lipschitz.
3.3. Lipschitz shifts
Let be a normed space and let be a closed bounded subset. We begin with the following elementary and well-known observation regarding Lipschitz perturbations of the identity map.
Lemma 3.5.
If is a -Lipschitz map, , then is -bi-Lipschitz.
For the rest of Section 3.3, we fix a vector and a real . Let the function be given by
where denotes the distance from to the boundary of . This function (as well as its variant to be introduced shortly) is -bi-Lipschitz. To simplify the notation, we set and .
Lemma 3.6.
The function is an -bi-Lipschitz homeomorphism onto .
Fix a real and let be the set of those elements that are at least units of distance away from the boundary of .
Lemma 3.7.
and .
A truncated shift function is defined by
Lemma 3.8.
The function is an -bi-Lipschitz homeomorphism onto .
3.4. Lipschitz equivalence to grid flows
The maps can be used to show that any free Borel -flow is bi-Lipschitz equivalent to a flow admitting an integer grid. This is the content of Theorem 3.12, but first we formulate the properties of partial actions needed for the construction. This is an adaption of the so-called unlayered toast construction announced in [8]. The proof given in [18, Appendix A] for -actions, transfers to -flows.
For the rest of the paper, we fix a norm on and let be the corresponding metric on . Recall that denotes a closed ball of radius centered at .
Lemma 3.9.
Let be a positive real. Any free -flow on a standard Borel space is a limit of a rational convergent sequence of partial actions (see Section 2.5) such that for each -class
-
(1)
is a closed and bounded subset of and ;
-
(2)
the set of -classes, , contained in is finite;
-
(3)
for any -class such that .
Before outlining the proof, we need to introduce some notation. Let be equivalence relations on respectively. By we mean the equivalence relation on generated by , i.e., whenever there exist and for each there exists such that , and for all .
If is an equivalence relation on and , we define the relation on by if and only if there are such that , and . Note that in general, may not be an equivalence relation if two -classes get connected after the “fattening”. However, is an equivalence relation if holds for all distinct -classes .
Proof of Lemma 3.9.
One starts with a sufficiently fast-growing sequence of radii (say, is fast enough) and chooses using [3] (see also [18, Lemma A.2]) a sequence of Borel -lacunary cross-sections such that
(1) |
where and stands for “there exist infinitely many”. We may assume without loss of generality that cross-sections are rational in the sense that if for some then . This can be achieved by moving elements of by an arbitrarily small amount (see [24, Lemma 2.4]) which maintains the property given in Eq. (1). Rationality of cross-sections guarantees that the sequence of partial actions constructed below is rational.
One now defines and inductively with the base , and if and only if there is such that . For the inductive step, begin with and being given analogously to the base case: if and only if there is some such that and . Set and let be the domain of . Finally, let be the -saturation of , i.e., if and only if there exists such that . Put .
An alternative description of an -class is as follows. One starts with an -class and joins it first with all -classes that intersect . Let the resulting -class be denoted by . Next we add all -classes that intersect producing an -class . The process terminates with an -class .
It is easy to check inductively that the diameter of any -class satisfies and therefore for all distinct -classes by the lacunarity of . The latter shows that is an equivalence relation on .
Monotonicity of the sequence is evident from the construction. Eq. (1) is crucial for establishing the fact that . Indeed, for each there exists some such that and thus also .
The maps , needed to specify partial -actions, are defined by the condition for the unique such that . Note that for any -class , , that is contained in an -class is a consequence of the fact that by the construction. The convergent sequence of partial actions therefore satisfies the desired properties. ∎
Let and be free -flows on that generate the same orbit equivalence relation, , and let be the associated cocycle map, defined for and by the condition . We say that the cocycle is -bi-Lipschitz if such is the map for all :
(2) |
Since , Lipschitz condition (2) for a cocycle can be equivalently and more concisely stated as
(3) |
Remark 3.10.
Note that cocycles and are connected via the identities
In particular, if is -bi-Lipschitz, then is -bi-Lipschitz.
Definition 3.11.
Let be a free -flow on . An integer grid for the flow is a -invariant Borel subset whose intersection with each orbit of the flow is a -orbit. In other words, , , and for all such that .
Not every flow admits an integer grid, but, as the following theorem shows, each flow is bi-Lipschitz equivalent to the one that does.
Theorem 3.12.
Let be a free Borel -flow on . For any there exists a free Borel -flow on that admits an integer grid, induces the sames orbit equivalence as does , i.e., , and whose associated cocycle is -bi-Lipschitz.
Proof.
Let be so big that the ball satisfies . Choose large enough to ensure that , and therefore also . Let be a rational convergent sequence of partial actions produced by Lemma 3.9 for the chosen value of . For an -class , let denote the collection of all that are at least -distance away from the boundary of :
If is an -class such that , then item (3) of Lemma 3.9 guarantees the inclusion . Let , where the union is taken over all -classes , and set , . Note that is a rational convergent sequence of partial actions whose limit is the flow . The flow will be constructed as the limit of partial actions , where maps will be defined inductively and will satisfy for all -classes . The sets will satisfy for , and will be an integer grid for .
For the base of the construction, set and . Next, consider a typical -class with being a complete list of -classes contained in it (see Figure 1). Consider the set , which is the integer grid inside (marked by dots in Figure 1). Each of the -classes comes with the grid constructed at the previous stage (depicted by crosses in Figure 1). The coherence condition for partial actions guarantees existence of some , , such that
In general, the grid does not contain , but for each , we can find a vector of norm such that . More precisely, we take for any vector in such that , which exists by the choice of . Let be the function , which is -bi-Lipschitz by Lemma 3.8. Finally, define to be
Lemma 3.4 has been tailored specifically to show that is -bi-Lipschitz. We set . Note that
(4) |
which validates coherence and, in view of , gives for all .
While we have provided the definition of on a single -class , the same construction can be done in a Borel way across all -classes using rationality of the sequence of partial actions just like we did in Theorem 2.6. If we let , then by Eq. (4).
The general inductive step is analogous. Suppose that we have constructed maps for . An -class contains finitely many subclasses , where is an -class, , and no is contained in a bigger -class for some . By coherence and inductive assumption, there exist , , such that
Choose vectors to satisfy , set to be , and define an -bi-Lipschitz function by
Finally, set and extend this definition to a Borel map using the rationality of the sequence of partial actions. Coherence of the maps and the inclusion for follow from the analog of Eq. (4).
It remains to check the bi-Lipschitz condition for the resulting cocycle . It is easier to work with the cocycle , which for satisfies
and is therefore -bi-Lipschitz, because so is . Hence, is also -bi-Lipschitz, because by the choice of . Finally, we apply Remark 3.10 to conclude that is also -bi-Lipschitz. ∎
Restricting the action of onto the integer grid , we get the following corollary.
Corollary 3.13.
Let be a free Borel -flow on . For any there exist a cross-section and a free -action on such that the cocycle given by is -bi-Lipschitz.
4. Special representation theorem
The main goal of this section is to formulate and prove a Borel version of Katok’s special representation theorem [12] that connects free -flows with free -actions. We have already done most of the work in proving Theorem 3.12, and it is now a matter of defining special representations in the Borel context and connecting them to our earlier setup.
4.1. Cocycles
Given a Borel action , a (Borel) cocycle with values in a group is a (Borel) map that satisfies the cocycle identity:
We are primarily concerned with the Abelian groups and in this section, so the cocycle identity will be written additively. A cocycle is said to be injective if for all and all , where and are the identity elements of the corresponding groups. Suppose that furthermore the groups and are locally compact. We say that escapes to infinity if for all , in the sense that for any compact there exists a compact such that whenever .
Example 4.1.
Suppose and , , are free actions of groups and on standard Borel spaces, and suppose that we have containment of orbit equivalence relations . For each and , there exists a unique such that . The map is an injective Borel cocycle. We have already encountered two instances of this idea in Section 3.4.
4.2. Flow under a function
Borel -flows and -actions are tightly connected through the “flow under a function” construction. Let be a free Borel automorphism of a standard Borel space and be a positive Borel function. There is a natural definition of a flow on the space under the graph of . The action for a positive is defined by shifting the point by units upward until the graph of is reached, then jumping to the point , and continuing to flow upward until the graph of at is reached, etc. More formally,
for the unique such that ; for the action is defined by “flowing backward”, i.e.,
for such that . The action is well-defined provided that the fibers within the orbits of have infinite cumulative lengths:
(5) |
The appealing geometric picture of the “flow under a function” does not generalize to higher dimensions, but admits an interpretation as the so-called special flow construction suggested in [12].
4.3. Special flows
Let be a free -action on a standard Borel space and let be a Borel cocycle. One can construct a -action , the so-called principal -extension, defined on via . An easy application of the cocycle identity verifies axioms of the action. While the action will typically have complicated dynamics, the action admits a Borel transversal as long as the cocycle escapes to infinity.
Lemma 4.2.
If the cocycle satisfies for all , then the action has a Borel transversal.
Proof.
Let . We claim that each orbit of intersects in a finite (possibly empty) set. Indeed, cocycle values escaping to infinity yield for any a number so large that whenever . In particular, implies . Hence, the intersection of the orbit of with is finite.
Set . Each orbit of intersects in a finite and necessarily non-empty set, so is a finite Borel equivalence relation. A Borel transversal for is also a transversal for the action of . ∎
We assume now that the cocycle satisfies the assumptions of Lemma 4.2, and therefore carries the structure of a standard Borel space as a push-forward of the factor map , which sends a point to its -equivalence class.
There is a natural -flow on which acts by shifting the second coordinate: . This flow commutes with the -action and therefore projects onto the flow on given by the condition . We say that is the special flow over generated by the cocycle . Freeness of implies freeness of .
The construction outlined above, works just as well in the context of ergodic theory, where the space would be endowed with a finite measure preserved by the action . The product of with the Lebesgue measure on induces a measure on , which is preserved by the flow . Furthermore, is finite provided the cocycle is integrable in the sense of [12, Condition (J), p. 122]. Katok’s special representation theorem asserts that, up to a null set, any free ergodic measure-preserving flow can be obtained via this process. Furthermore, the cocycle can be picked to be bi-Lipschitz with Lipschitz constants arbitrarily close to .
As will be shown shortly, such a representation result continues to hold in the framework of descriptive set theory, and every free Borel -flow is Borel isomorphic to a special flow over some free Borel -action. Moreover, just as in Katok’s original work, Theorem 4.3 provides some significant control on the cocycle that generates the flow, tightly coupling the dynamics of the -action with the dynamics of the flow it produces. But first, we re-interpret the construction in different terms.
4.4. Flows generated by admissible cocycles
Let the map be denoted by . If the cocycle is injective, then is a bijection and intersects every orbit of in a non-empty countable set. The -action on can be transferred via to give a free -action on . Let be the cocycle of the action ; in other words
(6) |
If for , then Eq. (6) translates into
Since , we conclude that , where . In particular, is a discrete cross-section for the flow precisely because escapes to infinity.
Conversely, if is any free -flow on a standard Borel space , and is a discrete cross-section with a -action on it, then is isomorphic to the special flow over generated by the (necessarily injective) cocycle .
Let us say that a cocycle is admissible if it is both injective and escapes to infinity. The discussion of the above two paragraphs can be summarized by saying that, up to a change of sign in the cocycles, representing a flow as a special flow generated by an admissible cocycle is the same thing as finding a free -action on a discrete cross-section of the flow.
For instance, given any free -action on , we may consider the admissible cocycle for all and . The set is then an integer grid for the flow (in the sense of Definition 3.11). Conversely, any flow that admits an integer grid is isomorphic to a special flow generated by such a cocycle.
4.5. Special representation theorem
Restriction of the orbit equivalence relation of any -flow onto a cross-section gives a hyperfinite equivalence relation [10, Theorem 1.16], and therefore can be realized as an orbit equivalence relation by a free Borel -action (as long as the restricted equivalence relation is aperiodic). Since any free flow admits a discrete (in fact, lacunary) aperiodic cross-section, it is isomorphic to a special flow over some action generated by some cocycle. In general, however, the structure of the -orbit and the corresponding orbit of the flow have little to do with each other. Theorem 3.12 and Corollary 3.13 allow us to improve on this and find a special representation generated by a bi-Lipschitz cocycle.
For comparison, Katok’s theorem [12] can be formulated in the parlance of this section as follows.
Theorem (Katok).
Pick some . Any free ergodic measure-preserving -flow on a standard Lebesgue space is isomorphic to a special flow over a free ergodic measure-preserving -action generated by an -bi-Lipschitz cocycle.
As is the case with all ergodic theoretical results, isomorphism is understood to hold up to a set of measure zero. We may now conclude with a Borel version of Katok’s special representation theorem, which holds for all free Borel -flows and establishes isomorphism on all orbits.
Theorem 4.3.
Pick some . Any free Borel -flow is isomorphic to a special flow over a free Borel -action generated by an -bi-Lipschitz cocycle.
References
- [1] Warren Ambrose, Representation of ergodic flows, Annals of Mathematics. Second Series 42 (1941), 723–739.
- [2] Warren Ambrose and Shizuo Kakutani, Structure and continuity of measurable flows, Duke Mathematical Journal 9 (1942), 25–42.
- [3] Charles M. Boykin and Steve Jackson, Borel boundedness and the lattice rounding property, Contemporary Mathematics (Su Gao, Steve Jackson, and Yi Zhang, eds.), vol. 425, American Mathematical Society, Providence, Rhode Island, 2007, pp. 113–126.
- [4] Michael R. Cotton, Abelian group actions and hypersmooth equivalence relations, Annals of Pure and Applied Logic 173 (2022), no. 8, 103122.
- [5] Andrés del Junco and Daniel J. Rudolph, Kakutani equivalence of ergodic actions, Ergodic Theory and Dynamical Systems 4 (1984), no. 1, 89–104.
- [6] Randall Dougherty, Steve Jackson, and Alexander S. Kechris, The structure of hyperfinite Borel equivalence relations, Transactions of the American Mathematical Society 341 (1994), no. 1, 193–225.
- [7] Jacob Feldman, New -automorphisms and a problem of Kakutani, Israel Journal of Mathematics 24 (1976), no. 1, 16–38. MR 0409763
- [8] Su Gao, Steve Jackson, Edward Krohne, and Brandon Seward, Forcing constructions and countable Borel equivalence relations, The Journal of Symbolic Logic 87 (2022), no. 3, 873–893. MR 4472517
- [9] Marlies Gerber and Philipp Kunde, Anti-classification results for the Kakutani equivalence relation, September 2021, Comment: 72 pages, 3 figures.
- [10] Steve Jackson, Alexander S. Kechris, and Alain Louveau, Countable Borel equivalence relations, Journal of Mathematical Logic 2 (2002), no. 1, 1–80.
- [11] Anatole B. Katok, Monotone equivalence in ergodic theory, Izvestiya Akademii Nauk SSSR. Seriya Matematicheskaya 41 (1977), no. 1, 104–157, 231. MR 0442195
- [12] by same author, The special representation theorem for multi-dimensional group actions, Dynamical Systems, Vol. I—Warsaw, 1977, pp. 117–140. Astérisque, No. 49. MR 0492181
- [13] Alexander S. Kechris, Countable sections for locally compact group actions, Ergodic Theory and Dynamical Systems 12 (1992), no. 2, 283–295.
- [14] by same author, Countable sections for locally compact group actions. II, Proceedings of the American Mathematical Society 120 (1994), no. 1, 241–247.
- [15] Bryna Kra, Anthony Quas, and Ayşe Şahin, Rudolph’s two step coding theorem and Alpern’s lemma for actions, Transactions of the American Mathematical Society 367 (2015), no. 6, 4253–4285. MR 3324927
- [16] Ulrich Krengel, On Rudolph’s representation of aperiodic flows, Ann. Inst. H. Poincaré Sect. B (N.S.) 12 (1976), no. 4, 319–338.
- [17] François Le Maître and Konstantin Slutsky, full groups of flows, August 2021.
- [18] Andrew S. Marks and Spencer T. Unger, Borel circle squaring, Annals of Mathematics 186 (2017), no. 2, 581–605.
- [19] Benjamin D. Miller and Christian Rosendal, Descriptive Kakutani equivalence, Journal of the European Mathematical Society (JEMS) 12 (2010), no. 1, 179–219.
- [20] 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
- [21] Daniel J. Rudolph, A two-valued step coding for ergodic flows, Mathematische Zeitschrift 150 (1976), no. 3, 201–220.
- [22] by same author, Rectangular tilings of and free -actions, Dynamical Systems (College Park, MD, 1986–87), Lecture Notes in Math., vol. 1342, Springer, Berlin, 1988, pp. 653–688.
- [23] Konstantin Slutsky, Lebesgue orbit equivalence of multidimensional Borel flows: A picturebook of tilings, Ergodic Theory and Dynamical Systems 37 (2017), no. 6, 1966–1996. MR 3681992
- [24] by same author, On time change equivalence of Borel flows, Fundamenta Mathematicae 247 (2019), no. 1, 1–24. MR 3984276
- [25] by same author, Regular cross sections of Borel flows, Journal of the European Mathematical Society (JEMS) 21 (2019), no. 7, 1985–2050. MR 3959857
- [26] by same author, Smooth orbit equivalence of multidimensional Borel flows, Advances in Mathematics 381 (2021), Paper No. 107626, 22. MR 4215747
- [27] Sławomir Solecki, Actions of non-compact and non-locally compact Polish groups, The Journal of Symbolic Logic 65 (2000), no. 4, 1881–1894. MR 1812189
- [28] Raimond A. Struble, Metrics in locally compact groups, Compositio Mathematica 28 (1974), 217–222.
- [29] Vladimir V. Uspenskij, On subgroups of minimal topological groups, Topology and its Applications 155 (2008), no. 14, 1580–1606. MR 2435151
- [30] Vivek M. Wagh, A descriptive version of Ambrose’s representation theorem for flows, Indian Academy of Sciences. Proceedings. Mathematical Sciences 98 (1988), no. 2-3, 101–108.
- [31] Benjamin Weiss, Measurable dynamics, Conference in Modern Analysis and Probability (New Haven, Conn., 1982), Contemp. Math., vol. 26, Amer. Math. Soc., Providence, RI, 1984, pp. 395–421.
- [32] Robert J. Zimmer, Ergodic theory and semisimple groups, Monographs in Mathematics, vol. 81, Birkhäuser Verlag, Basel, 1984. MR 776417