This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Katok’s special representation theorem
for multidimensional Borel flows

Konstantin Slutsky Department of Mathematics
Iowa State University
411 Morrill Road
Ames, IA 50011
[email protected]
Abstract.

Katok’s special representation theorem states that any free ergodic measure-preserving d\mathbb{R}^{d}-flow can be realized as a special flow over a d\mathbb{Z}^{d}-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 d\mathbb{R}^{d}-flows emerge from d\mathbb{Z}^{d}-actions through the special flow construction using bi-Lipschitz cocycles.

The author was partially supported by NSF grant DMS-2153981

1. Introduction

1.1. Overview

Theorems of Ambrose and Kakutani [1, 2] established a connection between measure-preserving \mathbb{Z}-actions and \mathbb{R}-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 d\mathbb{R}^{d}-flows for d2d\geq 2. 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 d\mathbb{R}^{d}-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 (X,)(X,\mathcal{B}), i.e., a set XX with a σ\sigma-algebra \mathcal{B} of Borel sets for some Polish topology on XX. 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 \mathbb{R}-flows was proved by Wagh in [30] showing that, just like in ergodic theory, all free Borel \mathbb{R}-flows emerge as “flows under a function” over Borel \mathbb{Z}-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 \mathbb{R}-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 d\mathbb{R}^{d}-flows.

1.2. Structure of the paper

Constructions of orbit equivalent d\mathbb{R}^{d}-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 \mathbb{R}-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 \mathbb{Z}-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 d\mathbb{R}^{d}-flow admits an integer grid—a Borel cross-section invariant under the d\mathbb{Z}^{d}-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 d\mathbb{R}^{d}-flow can be represented as a special flow generated by a bi-Lipschitz cocycle with Lipschitz constants arbitrarily close to 11. 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, XX denotes a standard Borel space.

2.1. Partial actions

Let GG be a standard Borel group, that is a group with a structure of a standard Borel space that makes group operations Borel. A partial GG-action111More precisely, we should call such (E,ϕ)(E,\phi) 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 (E,ϕ)(E,\phi), where EE is a Borel equivalence relation on XX and ϕ:XG\phi:X\to G is a Borel map that is injective on each EE-class: ϕ(x)ϕ(y)\phi(x)\neq\phi(y) whenever xEyxEy. The map ϕ\phi 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

Aϕ={(g,x,y)G×X×X:xEy and gϕ(x)=ϕ(y)}.A_{\phi}=\bigl{\{}(g,x,y)\in G\times X\times X:xEy\textrm{ and }g\phi(x)=\phi(y)\bigr{\}}.

Injectivity of ϕ\phi on EE-classes ensures that for each xXx\in X and gGg\in G there is at most one yXy\in X such that (g,x,y)Aϕ(g,x,y)\in A_{\phi}. When such a yy exists, we say that the action of gg on xx is defined and set gx=ygx=y. Clearly, (e,x,x)Aϕ(e,x,x)\in A_{\phi} for all xXx\in X, thus ex=xex=x; also g2(g1x)=(g2g1)xg_{2}(g_{1}x)=(g_{2}g_{1})x whenever all the terms are defined. The set AϕA_{\phi} is a graph of a total action GXG\curvearrowright X if and only if for each xXx\in X and gGg\in G there does exist some yXy\in X such that (g,x,y)Aϕ(g,x,y)\in A_{\phi}; in this case the orbit equivalence relation generated by the action coincides with EE.

Example 2.1.

An easy way of getting a partial action is by restricting a total one. Suppose we have a free Borel action GXG\curvearrowright X with the corresponding orbit equivalence relation EGE_{G} and suppose that a Borel equivalence sub-relation EEGE\subseteq E_{G} admits a Borel selector—a Borel EE-invariant map π:XX\pi:X\to X such that xEπ(x)xE\pi(x) for all xXx\in X. If ϕ:XG\phi:X\to G is the map specified uniquely by the condition ϕ(x)π(x)=x\phi(x)\pi(x)=x, then (E,ϕ)(E,\phi) is a partial GG-action.

Sub-relations EE 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 GXG\curvearrowright X of a locally compact second countable group. A cross-section of the action is a Borel set 𝒞X\mathcal{C}\subseteq X that intersects every orbit in a countable non-empty set. A cross-section 𝒞X\mathcal{C}\subseteq X is

  • discrete if (Kx)𝒞(Kx)\cap\mathcal{C} is finite for every xXx\in X and compact KGK\subseteq G;

  • UU-lacunary, where UGU\subseteq G is a neighborhood of the identity, if Uc𝒞={c}Uc\cap\mathcal{C}=\{c\} for all c𝒞c\in\mathcal{C};

  • lacunary if it is UU-lacunary for some neighborhood of the identity UU;

  • cocompact if K𝒞=XK\mathcal{C}=X for some compact KGK\subseteq G.

Let 𝒞\mathcal{C} be a lacunary cross-section for GXG\curvearrowright X, which exists by [13, Corollary 1.2]. Any lcsc group GG admits a compatible left-invariant proper metric [28], and any left-invariant metric dd can be transferred to orbits due to freeness of the action via dist(x,y)=d(g,e)\mathrm{dist}(x,y)=d(g,e) for the unique gGg\in G such that gx=ygx=y. One can now define the so-called Voronoi tessellation of orbits by associating with each xXx\in X the closest point π𝒞(x)𝒞\pi_{\mathcal{C}}(x)\in\mathcal{C} of the cross-section 𝒞\mathcal{C} as determined by dist\mathrm{dist}. Properness of the metric ensures that, for a ball BRGB_{R}\subseteq G of radius RR, BR={gG:d(g,e)R}B_{R}=\{g\in G:d(g,e)\leq R\}, and any xXx\in X, the set 𝒞BRx\mathcal{C}\cap B_{R}x is finite. Indeed, there can be at most λ(BR+r)/λ(Br)\lambda(B_{R+r})/\lambda(B_{r}) points in the intersection, where λ\lambda is a Haar measure on the group and r>0r>0 is so small that BrcBrc=B_{r}c\cap B_{r}c^{\prime}=\varnothing whenever c,c𝒞c,c^{\prime}\in\mathcal{C} 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 𝒞\mathcal{C} and associated each xx 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 E𝒞EGE_{\mathcal{C}}\subseteq E_{G} whose equivalence classes are the cells of the Voronoi tessellation: xE𝒞yxE_{\mathcal{C}}y if and only if π𝒞(x)=π𝒞(y)\pi_{\mathcal{C}}(x)=\pi_{\mathcal{C}}(y).

Assumed freeness of the action GXG\curvearrowright X allows for a natural identification of each Voronoi cell with a subset of the acting group via the map π𝒞1(c)xϕ𝒞(x)G\pi_{\mathcal{C}}^{-1}(c)\ni x\mapsto\phi_{\mathcal{C}}(x)\in G such that ϕ𝒞(x)c=x\phi_{\mathcal{C}}(x)c=x, 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 ϕ:XG\phi:X\to G, we can compose it with an arbitrary Borel injection f:GGf:G\to G to get a different partial action fϕf\circ\phi. 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 GG be a standard Borel group. A sequence (En,ϕn)(E_{n},\phi_{n}), nn\in\mathbb{N}, of partial GG-actions on XX is said to be convergent if it satisfies the following properties:

  • monotonicity: equivalence relations EnE_{n} form an increasing sequence, that is EnEn+1E_{n}\subseteq E_{n+1} for all nn;

  • coherence: for each nn the map x(ϕn(x))1ϕn+1(x)x\mapsto(\phi_{n}(x))^{-1}\phi_{n+1}(x) is EnE_{n}-invariant;

  • exhaustiveness: for all xXx\in X and all gGg\in G there exist nn and yXy\in X such that xEnyxE_{n}y and gϕn(x)=ϕn(y)g\phi_{n}(x)=\phi_{n}(y).

With such a sequence one can associate a free Borel (left) action GXG\curvearrowright X, called the limit of (En,ϕn)n(E_{n},\phi_{n})_{n}, whose graph is nAϕn\bigcup_{n}A_{\phi_{n}}. Coherence ensures that the partial action defined by ϕn+1\phi_{n+1} is an extension of the one given by ϕn\phi_{n}. Indeed, if xEnyxE_{n}y are such that gϕn(x)=ϕn(y)g\phi_{n}(x)=\phi_{n}(y), then also xEn+1yxE_{n+1}y by monotonicity and, using coherence,

gϕn+1(x)=gϕn(x)(ϕn(x))1ϕn+1(x)=ϕn(y)(ϕn(y))1ϕn+1(y)=ϕn+1(y),g\phi_{n+1}(x)=g\phi_{n}(x)(\phi_{n}(x))^{-1}\phi_{n+1}(x)=\phi_{n}(y)(\phi_{n}(y))^{-1}\phi_{n+1}(y)=\phi_{n+1}(y),

whence AϕnAϕn+1A_{\phi_{n}}\subseteq A_{\phi_{n+1}}. If CC is an EnE_{n}-class, and s=(ϕn(x))1ϕn+1(x)s=(\phi_{n}(x))^{-1}\phi_{n+1}(x) for some xCx\in C, then ϕn+1(C)=ϕn(C)s\phi_{n+1}(C)=\phi_{n}(C)s, so the image ϕn(C)\phi_{n}(C) gets shifted on the right inside ϕn+1(C)\phi_{n+1}(C). If we want to build a right action of the group, then ϕn(C)\phi_{n}(C) should be shifted on the left instead.

Finally, exhaustiveness guarantees that gxgx gets defined eventually: for all gGg\in G and xXx\in X there are nn and yXy\in X such that (g,x,y)Aϕn(g,x,y)\in A_{\phi_{n}}. It is straightforward to check that nAϕn\bigcup_{n}A_{\phi_{n}} is a graph of a total Borel action GXG\curvearrowright X. Equally easy is to check that the action is free, and its orbits are precisely the equivalence classes of nEn\bigcup_{n}E_{n}.

This framework, general as it is, delegates most of the complexity to the construction of maps ϕn\phi_{n}. 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 EGE_{G} onto the cross-section 𝒞\mathcal{C} is hyperfinite, i.e., there is an increasing sequence of finite Borel equivalence relations FnF_{n} on 𝒞\mathcal{C} such that nFn=EG|𝒞\bigcup_{n}F_{n}=E_{G}|_{\mathcal{C}}. We can use this sequence to define xEnyxE_{n}y whenever π𝒞(x)Fnπ𝒞(y)\pi_{\mathcal{C}}(x)F_{n}\pi_{\mathcal{C}}(y), which yields an increasing sequence of Borel equivalence relations EnE_{n} such that EG=nEnE_{G}=\bigcup_{n}E_{n}.

The equivalence relations FnF_{n} admit Borel transversals, i.e., there are Borel sets 𝒞n\mathcal{C}_{n} that pick exactly one point from each FnF_{n}-class. Just as in Section 2.2, we may define ϕn(x)\phi_{n}(x) to be such an element gGg\in G that gc=xgc=x for the unique c𝒞nc\in\mathcal{C}_{n} satisfying xEncxE_{n}c. This gives a convergent sequence of partial GG-actions (En,ϕn)n(E_{n},\phi_{n})_{n} whose limit is the original action GXG\curvearrowright X.

2.5. Partial actions revisited

In practice, it is often more convenient to allow equivalence relations EnE_{n} to be defined on proper subsets of XX. Let XnXX_{n}\subseteq X, nn\in\mathbb{N}, be Borel subsets, and suppose for each nn, EnE_{n} is a Borel equivalence relation on XnX_{n}. We say that the sequence (En)n(E_{n})_{n} is monotone if the following conditions are satisfied for all mnm\leq n:

  • Em|XmXnEn|XmXnE_{m}|_{X_{m}\cap X_{n}}\subseteq E_{n}|_{X_{m}\cap X_{n}};

  • if xXmXnx\in X_{m}\cap X_{n} then the whole EmE_{m}-class of xx is in XnX_{n}.

Partial action maps ϕn:XnG\phi_{n}:X_{n}\to G, where, as earlier, GG is a standard Borel group, need to satisfy the appropriate versions of coherence and exhaustiveness:

  • coherence: XmXnx(ϕm(x))1ϕn(x)X_{m}\cap X_{n}\ni x\mapsto(\phi_{m}(x))^{-1}\phi_{n}(x) is EmE_{m}-invariant for each m<nm<n;

  • exhaustiveness: for each xXx\in X and gGg\in G there exist nn and yXny\in X_{n} such that xXnx\in X_{n}, xEnyxE_{n}y, and gϕn(x)=ϕn(y)g\phi_{n}(x)=\phi_{n}(y).

A sequence of partial GG-actions (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} will be called convergent if it satisfies the above properties of monotonicity, coherence, and exhaustiveness. Note that the condition nXn=X\bigcup_{n}X_{n}=X follows from exhaustiveness, so sets XnX_{n} must cover all of XX.

Convergent sequences (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} 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 EnE_{n} to the equivalence relation E^n\hat{E}_{n} on all of XX by

xE^nymnxEmy or x=y;x\hat{E}_{n}y\iff\exists m\leq n\ xE_{m}y\textrm{ or }x=y;

and also extend ϕn\phi_{n} to ϕ^n:XG\hat{\phi}_{n}:X\to G by setting ϕ^n(x)=ϕm(x)\hat{\phi}_{n}(x)=\phi_{m}(x) for the maximal mnm\leq n such that xXmx\in X_{m} or ϕ^n(x)=e\hat{\phi}_{n}(x)=e if no such mm exists. It is straightforward to check that (E^n,ϕ^n)n(\hat{E}_{n},\hat{\phi}_{n})_{n} is a convergent sequence of partial GG-actions in the sense of Section 2.3. By the limit of the sequence of partial actions (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} we mean the limit of (E^n,ϕ^n)n(\hat{E}_{n},\hat{\phi}_{n})_{n} as defined earlier.

Remark 2.2.

A variant of this generalized formulation, which we encounter in Proposition 2.4 below, occurs when sets XnX_{n} are nested: X0X1X2X_{0}\subseteq X_{1}\subseteq X_{2}\subseteq\cdots. Monotonicity of equivalence relations then simplifies to E0E1E2E_{0}\subseteq E_{1}\subseteq E_{2}\subseteq\cdots and coherence becomes equivalent to the EnE_{n}-invariant of maps Xnx(ϕn(x))1ϕn+1(x)GX_{n}\ni x\mapsto(\phi_{n}(x))^{-1}\phi_{n+1}(x)\in G.

As was mentioned above, it is easy to create new partial actions simply by composing a partial action ϕ:XG\phi:X\to G with some Borel bijection f:GGf:G\to G (or f:GHf:G\to H 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 ff to different EnE_{n}-classes, which naturally raises concern of how to ensure that construction is performed in a Borel way. In applications, the modification ff applied to an EnE_{n}-class CC, usually depends on the “shape” of CC and the EmE_{m}-classes it contains, but does not depend on other EnE_{n}-classes. If there are only countably many such “configurations” of EnE_{n}-classes, resulting partial actions fϕf\circ\phi 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 (En,ϕn)n(E_{n},\phi_{n})_{n} be a convergent sequence of partial actions on XX. For an EnE_{n}-class CC, let m(C)\mathcal{E}_{m}(C) denote the collection of EmE_{m}-classes contained in CC. Given two EnE_{n}-classes CC and CC^{\prime}, we denote by ϕn(C)ϕn(C)\phi_{n}(C)\equiv\phi_{n}(C^{\prime}) the existence for each mnm\leq n of a bijection m(C)DDm(C)\mathcal{E}_{m}(C)\ni D\mapsto D^{\prime}\in\mathcal{E}_{m}(C^{\prime}) such that ϕn(D)=ϕn(D)\phi_{n}(D)=\phi_{n}(D^{\prime}) for all Dm(C)D\in\mathcal{E}_{m}(C). Collection of images {ϕn(D):Dmnm(C)}\{\phi_{n}(D):D\in\bigcup_{m\leq n}\mathcal{E}_{m}(C)\} constitutes the “configuration” of CC referred to earlier.

We say that the sequence (En,ϕn)n(E_{n},\phi_{n})_{n} of partial actions is rational if for each nn there exists a Borel EnE_{n}-invariant partition X=kYkX=\bigsqcup_{k}Y_{k} such that for each kk one has ϕn(C)ϕn(C)\phi_{n}(C)\equiv\phi_{n}(C^{\prime}) for all EnE_{n}-classes C,CYkC,C^{\prime}\subseteq Y_{k}.

Remark 2.3.

This concept of rationality applies verbatim to convergent sequences of partial actions (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} as described in Section 2.5. One can check that such a sequence is rational if and only if the sequence (E^n,ϕ^n)(\hat{E}_{n},\hat{\phi}_{n}) 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 \mathbb{R}-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 \mathbb{R}-flow as a limit of partial \mathbb{R}-actions.

Proposition 2.4.

Any free Borel \mathbb{R}-flow on XX can be represented as a limit of a convergent rational sequence of partial \mathbb{R}-actions (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} such that

  1. (1)

    both XnX_{n} and EnE_{n} are increasing: X0X1X_{0}\subseteq X_{1}\subseteq\cdots and E0E1E_{0}\subseteq E_{1}\subseteq\cdots; (see Remark 2.2)

  2. (2)

    each En+1E_{n+1}-class contains finitely many EnE_{n}-classes;

  3. (3)

    each E0E_{0}-class has cardinality of continuum;

  4. (4)

    for each En+1E_{n+1}-class CC the set CXnC\setminus X_{n} has cardinality of continuum.

Proof.

Any \mathbb{R}-flow admits a rational222Rationality of the cross-section here means that the distance between any two points of 𝒞\mathcal{C} is a rational number. More generally, rationality of a cross-section 𝒞\mathcal{C} for an d\mathbb{R}^{d}-action means rdr\in\mathbb{Q}^{d} whenever c+r=cc+r=c^{\prime} for some c,c𝒞c,c^{\prime}\in\mathcal{C}. (4,4)(-4,4)-lacunary cross-section (see [24, Section 2]), which we denote by 𝒞\mathcal{C}. Let (E𝒞,ϕ𝒞)(E_{\mathcal{C}},\phi_{\mathcal{C}}) be the partial \mathbb{R}-action as defined in Section 2.2. If DD is an E𝒞E_{\mathcal{C}}-class, then ϕ𝒞(D)\phi_{\mathcal{C}}(D) is an interval. For ϵ>0\epsilon>0, let DϵD^{\epsilon} consist of those xDx\in D such that ϕ𝒞(x)\phi_{\mathcal{C}}(x) is at least ϵ\epsilon away from the boundary points of ϕ𝒞(D)\phi_{\mathcal{C}}(D). In other words, DϵD^{\epsilon} is obtained by shrinking the class DD by ϵ\epsilon from each side.

The restriction of the orbit equivalence relation onto 𝒞\mathcal{C} is hyperfinite. This fact is true in the much wider generality of actions of locally compact Abelian groups [4]. Specifically for \mathbb{R}-flows, E|𝒞E|_{\mathcal{C}} is generated by the first return map—a Borel automorphism of 𝒞\mathcal{C} that sends a point in 𝒞\mathcal{C} to the next one according to the order of the \mathbb{R}-flow. The first return map is well defined and is invertible, except for the orbits, where 𝒞\mathcal{C} 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 \mathbb{Z}-actions are hyperfinite (see, for instance, [6, Theorem 5.1]), and thus so is the restriction E|𝒞E|_{\mathcal{C}}.

In particular, we can represent the \mathbb{R}-flow as the limit of a convergent sequence of partial actions (En,ϕn)n(E^{\prime}_{n},\phi^{\prime}_{n})_{n} as described in Section 2.4. Note that (En,ϕn)n(E^{\prime}_{n},\phi^{\prime}_{n})_{n} is necessarily rational by rationality of 𝒞\mathcal{C}. Such a sequence satisfies items (2) and (3), but fails (4). We fix this by shrinking equivalence classes to achieve proper containment. Let (ϵn)n(\epsilon_{n})_{n} be a strictly decreasing sequence of positive reals such that 1>ϵ01>\epsilon_{0} and limnϵn=0\lim_{n}\epsilon_{n}=0. Put Xn=DϵnX^{\prime}_{n}=\bigcup D^{\epsilon_{n}}, where the union is taken over all E𝒞E_{\mathcal{C}}-classes DD. Note that sets XnX^{\prime}_{n} fail to cover XX, because the boundary points of any E𝒞E_{\mathcal{C}}-class do not belong to any of XnX^{\prime}_{n}. Put Y=XnXnY=X\setminus\bigcup_{n}X^{\prime}_{n} and let Xn=XnYX_{n}=X^{\prime}_{n}\cup Y. Clearly, (Xn)n(X_{n})_{n} is an increasing sequence of Borel sets and nXn=X\bigcup_{n}X_{n}=X.

Finally, set En=En|XnE_{n}=E^{\prime}_{n}|_{X_{n}} and ϕn:Xn\phi_{n}:X_{n}\to\mathbb{R} to be ϕn|Xn\phi^{\prime}_{n}|_{X_{n}}. The sequence (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} of partial \mathbb{R}-actions satisfies the conditions of the proposition. ∎

All non-smooth orbit equivalence relations produced by free Borel \mathbb{R}-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 GG be a group. We say that a set AGA\subseteq G admits infinitely many disjoint right translates if there is a sequence (gn)n(g_{n})_{n} of elements of GG such that AgmAgn=Ag_{m}\cap Ag_{n}=\emptyset for all mnm\neq n.

Lemma 2.5.

Let GG be a non-compact Polish group. There exists a neighborhood of the identity VGV\subseteq G such that for any finite FGF\subseteq G the set VFVF 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 GG is non-compact if and only if there exists a neighborhood of the identity UGU\subseteq G such that F1UF2GF_{1}UF_{2}\neq G for all finite F1,F2GF_{1},F_{2}\subseteq G. Let VGV\subseteq G be a symmetric neighborhood of the identity such that V2UV^{2}\subseteq U. We claim that such a set VV has the desired property. Pick a finite FGF\subseteq G, set g0=eg_{0}=e and choose gng_{n} inductively as follows. Let F1=F1F_{1}=F^{-1} and F2,n=F{gk:k<n}F_{2,n}=F\cdot\{g_{k}:k<n\}. The defining property of UU assures existence of gnF1UF2,ng_{n}\not\in F_{1}UF_{2,n}. Translates (VFgn)n(VFg_{n})_{n} are then pairwise disjoint, for if VFgmVFgnVFg_{m}\cap VFg_{n}\neq\emptyset for m<nm<n, then gnF1V1VFgmF1UF2,ng_{n}\in F^{-1}V^{-1}VFg_{m}\subseteq F_{1}UF_{2,n}, contradicting the construction. ∎

Theorem 2.6.

Let EE be an orbit equivalence relation given by a free Borel \mathbb{R}-flow on XX. Any non-discrete non-compact Polish group GG admits a free Borel action GXG\curvearrowright X such that EG=EE_{G}=E.

Proof.

Let (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} be a convergent sequence of partial \mathbb{R}-actions as in Proposition 2.4 and let VGV\subseteq G be given by Lemma 2.5. Choose a countable dense (hn)n(h_{n})_{n} in GG so that nVhn=G\bigcup_{n}Vh_{n}=G. Since the sequence of partial \mathbb{R}-actions is rational, one may pick for each nn a Borel EnE_{n}-invariant partition Xn=kYn,kX_{n}=\bigsqcup_{k}Y_{n,k} such that ϕn(C)ϕn(C)\phi_{n}(C)\equiv\phi_{n}(C^{\prime}) for all EnE_{n}-classes C,CYn,kC,C^{\prime}\subseteq Y_{n,k}. We construct a convergent sequence of partial GG-actions (Xn,En,ψn)n(X_{n},E_{n},\psi_{n})_{n} such that for each nn and kk there exists a finite set FGF\subseteq G such that {hi:i<n}F\{h_{i}:i<n\}\subseteq F and ψn(C)=VF\psi_{n}(C)=VF for all EnE_{n}-classes CYn,kC\subseteq Y_{n,k}.

For any E0E_{0}-class CC, both ϕ0(C)\phi_{0}(C)\subseteq\mathbb{R} and VGV\subseteq G are Borel sets of the same cardinality. We may therefore pick a Borel bijection fk:ϕ0(C)Vf_{k}:\phi_{0}(C)\to V where CY0,kC\subseteq Y_{0,k}. For the base of the inductive construction we set ψ0|Yk=fkϕ0\psi_{0}|_{Y_{k}}=f_{k}\circ\phi_{0}. Suppose that ψm:XmG\psi_{m}:X_{m}\to G, mnm\leq n, have been constructed.

We now construct ψn+1\psi_{n+1}. Let CC be an En+1E_{n+1}-class and let D1,,DlD_{1},\ldots,D_{l} be a complete list of EnE_{n}-classes contained in CC. By the inductive assumption, there are finite sets F1,,FlGF_{1},\ldots,F_{l}\subseteq G such that ψn(Di)=VFi\psi_{n}(D_{i})=VF_{i}. Let F~=ilFi\tilde{F}=\bigcup_{i\leq l}F_{i}. By the choice of VV, there are elements g1,,glGg_{1},\ldots,g_{l}\in G such that VF~giV\tilde{F}g_{i}, are pairwise disjoint for 1il1\leq i\leq l. Pick a finite FGF\subseteq G large enough that F~giF\tilde{F}g_{i}\subseteq F, {hi:i<n+1}F\{h_{i}:i<n+1\}\subseteq F, and VFilVF~giVF\setminus\bigcup_{i\leq l}V\tilde{F}g_{i} has cardinality of continuum (the latter can be achieved, for instance, by assuring that one more disjoint translate of VF~V\tilde{F} is inside VFVF). Note that ϕn+1(CXn)=ϕn+1(C)ilϕn+1(Di)\phi_{n+1}(C\setminus X_{n})=\phi_{n+1}(C)\setminus\bigcup_{i\leq l}\phi_{n+1}(D_{i}) has cardinality of continuum by the properties guaranteed by Proposition 2.4. Pick any Borel bijection

f:ϕn+1(C)ilϕn+1(Di)VFiψn(Di)gif:\phi_{n+1}(C)\setminus\bigcup_{i\leq l}\phi_{n+1}(D_{i})\to VF\setminus\bigcup_{i}\psi_{n}(D_{i})g_{i}

and define ψn+1\psi_{n+1} by the conditions ψn+1|Di=ψn|Digi\psi_{n+1}|_{D_{i}}=\psi_{n}|_{D_{i}}\cdot g_{i} and ψn+1|CilDi=fϕn+1\psi_{n+1}|_{C\setminus\bigcup_{i\leq l}D_{i}}=f\circ\phi_{n+1}. Just as in the base case, the same modification ff works for all classes En+1E_{n+1}-classes C,CC,C^{\prime} such that ϕn+1(C)ϕn+1(C)\phi_{n+1}(C)\equiv\phi_{n+1}(C^{\prime}), which ensures Borelness of the construction.

It is now easy to check that (Xn,En,ψn)n(X_{n},E_{n},\psi_{n})_{n} is a convergent sequence of partial GG-actions, hence its limit is a free Borel action GXG\curvearrowright X such that EG=EE_{G}=E. ∎

Remark 2.7.

Theorem 2.6 highlights difference with actions of discrete groups, since a free Borel \mathbb{Z}-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 \mathbb{Z}-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 d\mathbb{R}^{d}-flow is bi-Lipschitz orbit equivalent to a flow with an integer grid. Sections 3.13.3 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 f:XYf:X\to Y between metric spaces (X,dY)(X,d_{Y}) and (Y,dY)(Y,d_{Y}) is KK-Lipschitz if dY(f(x1),f(x2))KdX(x1,x2)d_{Y}(f(x_{1}),f(x_{2}))\leq Kd_{X}(x_{1},x_{2}) for all x1,x2Xx_{1},x_{2}\in X, and it is (K1,K2)(K_{1},K_{2})-bi-Lipschitz if ff is injective, K2K_{2}-Lipschitz, and f1f^{-1} is K11K^{-1}_{1}-Lipschitz, which can equivalently be stated as

K1dX(x1,x2)dY(f(x1),f(x2))K2dX(x1,x2) for all x1,x2X.K_{1}d_{X}(x_{1},x_{2})\leq d_{Y}(f(x_{1}),f(x_{2}))\leq K_{2}d_{X}(x_{1},x_{2})\quad\textrm{ for all }x_{1},x_{2}\in X.

The Lipschitz constant of a Lipschitz map ff is the smallest KK with respect to which ff is KK-Lipschitz.

3.1. Linked sets

Given two Lipschitz maps f:AAf:A\to A^{\prime} and g:BBg:B\to B^{\prime} that agree on the intersection ABA\cap B, the map fg:ABABf\cup g:A\cup B\to A^{\prime}\cup B^{\prime}, in general, may not be Lipschitz. The following condition is sufficient to ensure that fgf\cup g is Lipschitz with the Lipschitz constant bounded by the maximum of the constants of ff and gg.

Definition 3.1.

Let (X,d)(X,d) be a metric space and A,BXA,B\subseteq X be its subsets. We say that AA and BB are linked if for all xAx\in A and yBy\in B there exists zABz\in A\cap B such that d(x,y)=d(x,z)+d(z,y)d(x,y)=d(x,z)+d(z,y).

Lemma 3.2.

Let (X,d)(X,d) be a metric space, f:AAf:A\to A^{\prime}, g:BBg:B\to B^{\prime} be KK-Lipschitz maps between subsets of XX and suppose that f|AB=g|ABf|_{A\cap B}=g|_{A\cap B}. If AA and BB are linked, then fg:ABABf\cup g:A\cup B\to A^{\prime}\cup B^{\prime} is KK-Lipschitz.

Recall that a metric space (X,d)(X,d) is geodesic if for all points x,yXx,y\in X there exists a geodesic between them—an isometry τ:[0,d(x,y)]X\tau:\bigl{[}0,d(x,y)\bigr{]}\to X such that τ(0)=x\tau(0)=x and τ(d(x,y))=y\tau\bigl{(}d(x,y)\bigr{)}=y. For geodesic metric spaces, closed sets A,BXA,B\subseteq X are always linked whenever the boundary of one of them is contained in the other. The boundary of a set AA will be denoted by A\partial A, and intA\operatorname*{\mathrm{int}}A will stand for the interior of AA.

Lemma 3.3.

Suppose (X,d)(X,d) is a geodesic metric space. If A,BXA,B\subseteq X are closed and satisfy AB\partial A\subseteq B, then AA and BB 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 (X,d)(X,d) be a geodesic metric space and AXA\subseteq X be a closed set. Suppose (Ai)i=1n(A_{i})_{i=1}^{n} are pairwise disjoint closed subsets of AA and hi:AiAih_{i}:A_{i}\to A_{i} are (K1,K2)(K_{1},K_{2})-bi-Lipschitz maps such that hi|Aih_{i}|_{\partial A_{i}} is the identity map for each 1in1\leq i\leq n. The map g:AAg:A\to A given by

g(x)={hi(x) if xAi,xotherwiseg(x)=\begin{cases}h_{i}(x)&\textrm{ if $x\in A_{i}$},\\ x&\textrm{otherwise}\end{cases}

is (K1,K2)(K_{1},K_{2})-bi-Lipschitz.

3.3. Lipschitz shifts

Let (X,)(X,\left\lVert\cdot\right\rVert) be a normed space and let AXA\subseteq X 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 ξ:AX\xi:A\to X is a KK-Lipschitz map, K<1K<1, then Axx+ξ(x)XA\ni x\mapsto x+\xi(x)\in X is (1K,1+K)(1-K,1+K)-bi-Lipschitz.

For the rest of Section 3.3, we fix a vector vXv\in X and a real K>vK>||v||. Let the function fA,K,v:AXf_{A,K,v}:A\to X be given by

fA,K,v(x)=x+d(x,A)Kv,f_{A,K,v}(x)=x+\frac{d(x,\partial A)}{K}v,

where d(x,A)d(x,\partial A) denotes the distance from xx to the boundary of AA. This function (as well as its variant to be introduced shortly) is (1K1v,1+K1v)(1-K^{-1}||v||,1+K^{-1}||v||)-bi-Lipschitz. To simplify the notation, we set α+=1+K1v\alpha^{+}=1+K^{-1}||v|| and α=1K1v\alpha^{-}=1-K^{-1}||v||.

Lemma 3.6.

The function fA,K,vf_{A,K,v} is an (α,α+)(\alpha^{-},\alpha^{+})-bi-Lipschitz homeomorphism onto AA.

Fix a real L>0L>0 and let AL={xA:d(x,A)L}A^{L}=\{x\in A:d(x,\partial A)\geq L\} be the set of those elements that are at least LL units of distance away from the boundary of AA.

Lemma 3.7.

fA,K,v|AL=fAL,K,v+LK1vf_{A,K,v}|_{A^{L}}=f_{A^{L},K,v}+LK^{-1}v and fA,K,v(AL)=AL+LK1vf_{A,K,v}(A^{L})=A^{L}+LK^{-1}v.

A truncated shift function hA,K,v,L:AXh_{A,K,v,L}:A\to X is defined by

hA,K,v,L(x)={fA,K,v(x)for xAAL,x+LK1vfor xAL.h_{A,K,v,L}(x)=\begin{cases}f_{A,K,v}(x)&\text{for $x\in A\setminus A^{L}$},\\ x+LK^{-1}v&\text{for $x\in A^{L}$}.\par\end{cases}
Lemma 3.8.

The function hA,K,v,Lh_{A,K,v,L} is an (α,α+)(\alpha^{-},\alpha^{+})-bi-Lipschitz homeomorphism onto AA.

3.4. Lipschitz equivalence to grid flows

The maps hA,K,v,Lh_{A,K,v,L} can be used to show that any free Borel d\mathbb{R}^{d}-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 d\mathbb{Z}^{d}-actions, transfers to d\mathbb{R}^{d}-flows.

For the rest of the paper, we fix a norm ||||||\cdot|| on d\mathbb{R}^{d} and let d(x,y)=xyd(x,y)=||x-y|| be the corresponding metric on d\mathbb{R}^{d}. Recall that BR(r)dB_{R}(r)\subseteq\mathbb{R}^{d} denotes a closed ball of radius RR centered at rdr\in\mathbb{R}^{d}.

Lemma 3.9.

Let K>0K>0 be a positive real. Any free d\mathbb{R}^{d}-flow on a standard Borel space XX is a limit of a rational convergent sequence of partial actions (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} (see Section 2.5) such that for each EnE_{n}-class CC

  1. (1)

    ϕn(C)\phi_{n}(C) is a closed and bounded subset of d\mathbb{R}^{d} and BK(0)ϕn(C)B_{K}(0)\subseteq\phi_{n}(C);

  2. (2)

    the set of EmE_{m}-classes, mnm\leq n, contained in CC is finite;

  3. (3)

    d(ϕn(D),ϕn(C))Kd(\phi_{n}(D),\partial\phi_{n}(C))\geq K for any EmE_{m}-class DD such that DCD\subseteq C.

Before outlining the proof, we need to introduce some notation. Let E1,,EnE_{1},\ldots,E_{n} be equivalence relations on X1,,XnX_{1},\ldots,X_{n} respectively. By E1EnE_{1}\vee\cdots\vee E_{n} we mean the equivalence relation EE on inXi\bigcup_{i\leq n}X_{i} generated by EiE_{i}, i.e., xEyxEy whenever there exist x1,,xmx_{1},\ldots,x_{m} and for each 1im1\leq i\leq m there exists 1j(i)n1\leq j(i)\leq n such that x1=xx_{1}=x, xm=yx_{m}=y and xiEj(i)xi+1x_{i}E_{j(i)}x_{i+1} for all 1i<m1\leq i<m.

If EE is an equivalence relation on YXY\subseteq X and K>0K>0, we define the relation E+KE^{+K} on Y+K={xX:dist(x,y)K for some yY}Y^{+K}=\{x\in X:\mathrm{dist}(x,y)\leq K\textrm{ for some }y\in Y\} by x1E+Kx2x_{1}E^{+K}x_{2} if and only if there are y1,y2Yy_{1},y_{2}\in Y such that dist(x1,y1)K\mathrm{dist}(x_{1},y_{1})\leq K, dist(x2,y2)K\mathrm{dist}(x_{2},y_{2})\leq K and y1Ey2y_{1}Ey_{2}. Note that in general, E+KE^{+K} may not be an equivalence relation if two EE-classes get connected after the “fattening”. However, E+KE^{+K} is an equivalence relation if dist(C1,C2)>2K\mathrm{dist}(C_{1},C_{2})>2K holds for all distinct EE-classes C1,C2C_{1},C_{2}.

Proof of Lemma 3.9.

One starts with a sufficiently fast-growing sequence of radii ana_{n} (say, an=K1000n+1a_{n}=K1000^{n+1} is fast enough) and chooses using [3] (see also [18, Lemma A.2]) a sequence of Borel Ban(0)B_{a_{n}}(0)-lacunary cross-sections 𝒞nX\mathcal{C}_{n}\subseteq X such that

(1) xXϵ>0n such that dist(x,𝒞n)<ϵan,\forall x\in X\ \forall\epsilon>0\ \exists^{\infty}n\textrm{ such that }\mathrm{dist}(x,\mathcal{C}_{n})<\epsilon a_{n},

where dist(x,𝒞n)=inf{dist(x,c):c𝒞n}\mathrm{dist}(x,\mathcal{C}_{n})=\inf\{\mathrm{dist}(x,c):c\in\mathcal{C}_{n}\} and \exists^{\infty} stands for “there exist infinitely many”. We may assume without loss of generality that cross-sections 𝒞n\mathcal{C}_{n} are rational in the sense that if c1+r=c2c_{1}+r=c_{2} for some c1,c2n𝒞nc_{1},c_{2}\in\bigcup_{n}\mathcal{C}_{n} then rdr\in\mathbb{Q}^{d}. This can be achieved by moving elements of 𝒞n\mathcal{C}_{n} 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 XnX_{n} and EnE_{n} inductively with the base X0=𝒞0+Ba0/10(0)X_{0}=\mathcal{C}_{0}+B_{a_{0}/10}(0), and xE0yxE_{0}y if and only if there is c𝒞0c\in\mathcal{C}_{0} such that x,yc+Ba0/10(0)x,y\in c+B_{a_{0}/10}(0). For the inductive step, begin with X~n=𝒞n+Ban/10(0)\tilde{X}_{n}=\mathcal{C}_{n}+B_{a_{n}/10}(0) and E~n\tilde{E}_{n} being given analogously to the base case: xE~nyx\tilde{E}_{n}y if and only if there is some c𝒞nc\in\mathcal{C}_{n} such that dist(x,c)an/10\mathrm{dist}(x,c)\leq a_{n}/10 and dist(y,c)an/10\mathrm{dist}(y,c)\leq a_{n}/10. Set En=E~nEn1+KE0+KE^{\prime}_{n}=\tilde{E}_{n}\vee E^{+K}_{n-1}\vee\cdots\vee E_{0}^{+K} and let Xn=X~ni=0n1Xi+KX^{\prime}_{n}=\tilde{X}_{n}\cup\bigcup_{i=0}^{n-1}X_{i}^{+K} be the domain of EnE^{\prime}_{n}. Finally, let XnX_{n} be the EnE^{\prime}_{n}-saturation of X~n\tilde{X}_{n}, i.e., xXnx\in X_{n} if and only if there exists yX~ny\in\tilde{X}_{n} such that xEnyxE^{\prime}_{n}y. Put En=En|XnE_{n}=E^{\prime}_{n}|_{X_{n}}.

An alternative description of an EnE_{n}-class is as follows. One starts with an E~n\tilde{E}_{n}-class CnC_{n} and joins it first with all En1+KE^{+K}_{n-1}-classes DD that intersect CnC_{n}. Let the resulting E~nEn1+K\tilde{E}_{n}\vee E^{+K}_{n-1}-class be denoted by Cn1C_{n-1}. Next we add all En2+KE^{+K}_{n-2}-classes that intersect Cn2C_{n-2} producing an E~nEn1+KEn2+K\tilde{E}_{n}\vee E^{+K}_{n-1}\vee E^{+K}_{n-2}-class Cn2C_{n-2}. The process terminates with an EnE_{n}-class C0C_{0}.

It is easy to check inductively that the diameter of any EnE_{n}-class CC satisfies diam(C)an/3\mathrm{diam}(C)\leq a_{n}/3 and therefore dist(C1,C2)an/32K\mathrm{dist}(C_{1},C_{2})\geq a_{n}/3\gg 2K for all distinct EnE_{n}-classes C1,C2C_{1},C_{2} by the lacunarity of 𝒞n\mathcal{C}_{n}. The latter shows that En+KE^{+K}_{n} is an equivalence relation on Xn+KX_{n}^{+K}.

Monotonicity of the sequence (Xn,En)n(X_{n},E_{n})_{n} is evident from the construction. Eq. (1) is crucial for establishing the fact that nXn=X\bigcup_{n}X_{n}=X. Indeed, for each xXx\in X there exists some nn such that dist(x,𝒞n)<an/10\mathrm{dist}(x,\mathcal{C}_{n})<a_{n}/10 and thus also xX~nXnx\in\tilde{X}_{n}\subseteq X_{n}.

The maps ϕn:Xnd\phi_{n}:X_{n}\to\mathbb{R}^{d}, needed to specify partial d\mathbb{R}^{d}-actions, are defined by the condition ϕn(x)c=x\phi_{n}(x)c=x for the unique c𝒞nc\in\mathcal{C}_{n} such that cEnxcE_{n}x. Note that d(ϕn(D),ϕn(C))Kd(\phi_{n}(D),\partial\phi_{n}(C))\geq K for any EmE_{m}-class DD, m<nm<n, that is contained in an EnE_{n}-class CC is a consequence of the fact that D+KCD^{+K}\subseteq C by the construction. The convergent sequence of partial actions (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} therefore satisfies the desired properties. ∎

Let 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} be free d\mathbb{R}^{d}-flows on XX that generate the same orbit equivalence relation, E𝔉1=E𝔉2E_{\mathfrak{F}_{1}}=E_{\mathfrak{F}_{2}}, and let ρ=ρ𝔉1,𝔉2:d×Xd\rho=\rho_{\mathfrak{F}_{1},\mathfrak{F}_{2}}:\mathbb{R}^{d}\times X\to\mathbb{R}^{d} be the associated cocycle map, defined for xXx\in X and rdr\in\mathbb{R}^{d} by the condition x+2r=x+1ρ(r,x)x+_{2}r=x+_{1}\rho(r,x). We say that the cocycle ρ\rho is (K1,K2)(K_{1},K_{2})-bi-Lipschitz if such is the map ρ(,x):dd\rho(\,\cdot\,,x):\mathbb{R}^{d}\to\mathbb{R}^{d} for all xXx\in X:

(2) K1r2r1ρ(r2,x)ρ(r1,x)K2r2r1.K_{1}||r_{2}-r_{1}||\leq||\rho(r_{2},x)-\rho(r_{1},x)||\leq K_{2}||r_{2}-r_{1}||.

Since ρ(r2,x)ρ(r1,x)=ρ(r2r1,x+1r1)\rho(r_{2},x)-\rho(r_{1},x)=\rho(r_{2}-r_{1},x+_{1}r_{1}), Lipschitz condition (2) for a cocycle can be equivalently and more concisely stated as

(3) K1ρ(r,x)rK2 for all xX and rd{0}.K_{1}\leq\dfrac{||\rho(r,x)||}{||r||}\leq K_{2}\quad\textrm{ for all $x\in X$ and $r\in\mathbb{R}^{d}\setminus\{0\}$}.
Remark 3.10.

Note that cocycles ρ𝔉1,𝔉2\rho_{\mathfrak{F}_{1},\mathfrak{F}_{2}} and ρ𝔉2,𝔉1\rho_{\mathfrak{F}_{2},\mathfrak{F}_{1}} are connected via the identities

ρ𝔉1,𝔉2(ρ𝔉2,𝔉1(r,x),x)=r and ρ𝔉2,𝔉1(ρ𝔉1,𝔉2(r,x),x)=r.\rho_{\mathfrak{F}_{1},\mathfrak{F}_{2}}(\rho_{\mathfrak{F}_{2},\mathfrak{F}_{1}}(r,x),x)=r\qquad\textrm{ and }\qquad\rho_{\mathfrak{F}_{2},\mathfrak{F}_{1}}(\rho_{\mathfrak{F}_{1},\mathfrak{F}_{2}}(r,x),x)=r.

In particular, if ρ𝔉1,𝔉2\rho_{\mathfrak{F}_{1},\mathfrak{F}_{2}} is (K1,K2)(K_{1},K_{2})-bi-Lipschitz, then ρ𝔉2,𝔉1\rho_{\mathfrak{F}_{2},\mathfrak{F}_{1}} is (K21,K11)(K_{2}^{-1},K_{1}^{-1})-bi-Lipschitz.

Definition 3.11.

Let 𝔉\mathfrak{F} be a free d\mathbb{R}^{d}-flow on XX. An integer grid for the flow 𝔉\mathfrak{F} is a d\mathbb{Z}^{d}-invariant Borel subset ZXZ\subseteq X whose intersection with each orbit of the flow is a d\mathbb{Z}^{d}-orbit. In other words, Z+d=XZ+\mathbb{R}^{d}=X, Z+d=ZZ+\mathbb{Z}^{d}=Z, and z1+d=z2+dz_{1}+\mathbb{Z}^{d}=z_{2}+\mathbb{Z}^{d} for all z1,z2Zz_{1},z_{2}\in Z such that z1E𝔉z2z_{1}E_{\mathfrak{F}}z_{2}.

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 𝔉1\mathfrak{F}_{1} be a free Borel d\mathbb{R}^{d}-flow on XX. For any α>1\alpha>1 there exists a free Borel d\mathbb{R}^{d}-flow 𝔉2\mathfrak{F}_{2} on XX that admits an integer grid, induces the sames orbit equivalence as does 𝔉1\mathfrak{F}_{1}, i.e., E𝔉1=E𝔉2E_{\mathfrak{F}_{1}}=E_{\mathfrak{F}_{2}}, and whose associated cocycle ρ𝔉1,𝔉2\rho_{\mathfrak{F}_{1},\mathfrak{F}_{2}} is (α1,α)(\alpha^{-1},\alpha)-bi-Lipschitz.

Proof.

Let RR be so big that the ball BR(0)dB_{R}(0)\subseteq\mathbb{R}^{d} satisfies d+BR(0)=d\mathbb{Z}^{d}+B_{R}(0)=\mathbb{R}^{d}. Choose K>0K>0 large enough to ensure that α=1K1R>α1\alpha^{-}=1-K^{-1}R>\alpha^{-1}, and therefore also α+=1+K1R<α\alpha^{+}=1+K^{-1}R<\alpha. Let (Xn,En,ϕn)n(X_{n},E_{n},\phi_{n})_{n} be a rational convergent sequence of partial actions produced by Lemma 3.9 for the chosen value of KK. For an EnE_{n}-class CC, let CC^{\prime} denote the collection of all xCx\in C that are at least KK-distance away from the boundary of CC:

C={xC:d(ϕn(x),ϕn(C))K}.C^{\prime}=\{x\in C:d(\phi_{n}(x),\partial\phi_{n}(C))\geq K\}.

If DD is an EmE_{m}-class such that DCD\subseteq C, then item (3) of Lemma 3.9 guarantees the inclusion DCD\subseteq C^{\prime}. Let Xn=CX^{\prime}_{n}=\bigcup C^{\prime}, where the union is taken over all EnE_{n}-classes CC, and set En=En|XnE_{n}^{\prime}=E_{n}|_{X^{\prime}_{n}}, ϕn=ϕn|Xn\phi^{\prime}_{n}=\phi_{n}|_{X^{\prime}_{n}}. Note that (Xn,En,ϕn)n(X^{\prime}_{n},E^{\prime}_{n},\phi^{\prime}_{n})_{n} is a rational convergent sequence of partial actions whose limit is the flow 𝔉1\mathfrak{F}_{1}. The flow 𝔉2\mathfrak{F}_{2} will be constructed as the limit of partial actions (Xn,En,ψn)(X^{\prime}_{n},E^{\prime}_{n},\psi_{n}), where maps ψn\psi_{n} will be defined inductively and will satisfy ψn(C)=ϕn(C)\psi_{n}(C^{\prime})=\phi_{n}(C^{\prime}) for all EnE_{n}-classes CC. The sets Zn=ψn1(d)Z_{n}=\psi_{n}^{-1}(\mathbb{Z}^{d}) will satisfy ZmXnZnZ_{m}\cap X^{\prime}_{n}\subseteq Z_{n} for mnm\leq n, and Z=nZnZ=\bigcup_{n}Z_{n} will be an integer grid for 𝔉2\mathfrak{F}_{2}.

KKCCCC^{\prime}D1D_{1}D2D_{2}
Figure 1. Construction of the integer grid

For the base of the construction, set ψ0=ϕ0\psi_{0}=\phi^{\prime}_{0} and Z0=ψ01(d)Z_{0}=\psi_{0}^{-1}(\mathbb{Z}^{d}). Next, consider a typical E1E_{1}-class CC with D1,,DlD_{1},\ldots,D_{l} being a complete list of E0E_{0}-classes contained in it (see Figure 1). Consider the set Z~C=ϕ11(d)C\tilde{Z}_{C^{\prime}}=\phi_{1}^{-1}(\mathbb{Z}^{d})\cap C^{\prime}, which is the integer grid inside CC^{\prime} (marked by dots in Figure 1). Each of the DiD_{i}-classes comes with the grid Z~Di=ψ01(d)Di\tilde{Z}_{D^{\prime}_{i}}=\psi_{0}^{-1}(\mathbb{Z}^{d})\cap D^{\prime}_{i} constructed at the previous stage (depicted by crosses in Figure 1). The coherence condition for partial actions guarantees existence of some sids_{i}\in\mathbb{R}^{d}, ili\leq l, such that

ϕ1(Di)=ϕ0(Di)+si=ψ0(Di)+si.\phi_{1}(D^{\prime}_{i})=\phi_{0}(D^{\prime}_{i})+s_{i}=\psi_{0}(D^{\prime}_{i})+s_{i}.

In general, the grid Z~C\tilde{Z}_{C^{\prime}} does not contain Z~Di\tilde{Z}_{D^{\prime}_{i}}, but for each ili\leq l, we can find a vector vidv_{i}\in\mathbb{R}^{d} of norm viR||v_{i}||\leq R such that Z~Di+1viZ~C\tilde{Z}_{D^{\prime}_{i}}+_{1}v_{i}\subseteq\tilde{Z}_{C^{\prime}}. More precisely, we take for viv_{i} any vector in BR(0)B_{R}(0) such that si+vids_{i}+v_{i}\in\mathbb{Z}^{d}, which exists by the choice of RR. Let hi:ϕ1(Di)ϕ1(Di)h_{i}:\phi_{1}(D_{i})\to\phi_{1}(D_{i}) be the function hϕ1(Di),K,vi,Kh_{\phi_{1}(D_{i}),K,v_{i},K}, which is (α,α+)(\alpha^{-},\alpha^{+})-bi-Lipschitz by Lemma 3.8. Finally, define g1:ϕ1(C)ϕ1(C)g_{1}:\phi_{1}(C^{\prime})\to\phi_{1}(C^{\prime}) to be

g1(r)={hi(r)if rϕ1(Di),rotherwise.g_{1}(r)=\begin{cases}h_{i}(r)&\textrm{if $r\in\phi_{1}(D_{i})$},\\ r&\textrm{otherwise}.\end{cases}

Lemma 3.4 has been tailored specifically to show that g1g_{1} is (α,α+)(\alpha^{-},\alpha^{+})-bi-Lipschitz. We set ψ1|C=g1ϕ1|C\psi_{1}|_{C^{\prime}}=g_{1}\circ\phi_{1}|_{C^{\prime}}. Note that

(4) ψ1(Di)=g1ϕ1(Di)=hiϕ1(Di)=ϕ1(Di)+KK1vi=ϕ0(Di)+si+vi=ψ0(Di)+si+vi,\begin{split}\psi_{1}(D^{\prime}_{i})&=g_{1}\circ\phi_{1}(D^{\prime}_{i})=h_{i}\circ\phi_{1}(D^{\prime}_{i})=\phi_{1}(D^{\prime}_{i})+KK^{-1}v_{i}\\ &=\phi_{0}(D^{\prime}_{i})+s_{i}+v_{i}=\psi_{0}(D^{\prime}_{i})+s_{i}+v_{i},\end{split}

which validates coherence and, in view of si+vids_{i}+v_{i}\in\mathbb{Z}^{d}, gives ψ11(d)Di=ψ01(d)Di\psi^{-1}_{1}(\mathbb{Z}^{d})\cap D^{\prime}_{i}=\psi_{0}^{-1}(\mathbb{Z}^{d})\cap D^{\prime}_{i} for all ili\leq l.

While we have provided the definition of ψ1\psi_{1} on a single E1E_{1}-class CC, the same construction can be done in a Borel way across all E1E_{1}-classes CC using rationality of the sequence of partial actions just like we did in Theorem 2.6. If we let Z1=ψ11(d)Z_{1}=\psi_{1}^{-1}(\mathbb{Z}^{d}), then Z0X1Z1Z_{0}\cap X_{1}\subseteq Z_{1} by Eq. (4).

The general inductive step is analogous. Suppose that we have constructed maps ψk\psi_{k} for knk\leq n. An En+1E_{n+1}-class CC contains finitely many subclasses D1,,DlD_{1},\ldots,D_{l}, where DiD_{i} is an EmiE_{m_{i}}-class, mi<nm_{i}<n, and no DiD_{i} is contained in a bigger EmE_{m}-class for some mi<m<nm_{i}<m<n. By coherence and inductive assumption, there exist sids_{i}\in\mathbb{R}^{d}, ili\leq l, such that

ϕn+1(Di)=ϕmi(Di)+si=ψmi(Di)+si.\phi_{n+1}(D^{\prime}_{i})=\phi_{m_{i}}(D^{\prime}_{i})+s_{i}=\psi_{m_{i}}(D^{\prime}_{i})+s_{i}.

Choose vectors viBR(0)v_{i}\in B_{R}(0) to satisfy si+vids_{i}+v_{i}\in\mathbb{Z}^{d}, set hi:ϕn+1(Di)ϕn+1(Di)h_{i}:\phi_{n+1}(D_{i})\to\phi_{n+1}(D_{i}) to be hϕn+1(Di),K,vi,Kh_{\phi_{n+1}(D_{i}),K,v_{i},K}, and define an (α,α+)(\alpha^{-},\alpha^{+})-bi-Lipschitz function gn+1g_{n+1} by

gn+1(r)={hi(r)if rϕn+1(Di),rotherwise.g_{n+1}(r)=\begin{cases}h_{i}(r)&\textrm{if $r\in\phi_{n+1}(D_{i})$},\\ r&\textrm{otherwise}.\end{cases}

Finally, set ψn+1|C=gn+1ϕn+1|C\psi_{n+1}|_{C^{\prime}}=g_{n+1}\circ\phi_{n+1}|_{C^{\prime}} and extend this definition to a Borel map ψn+1:Xn+1d\psi_{n+1}:X^{\prime}_{n+1}\to\mathbb{R}^{d} using the rationality of the sequence of partial actions. Coherence of the maps (ψk)kn+1(\psi_{k})_{k\leq n+1} and the inclusion ZmXn+1Zn+1Z_{m}\cap X^{\prime}_{n+1}\subseteq Z_{n+1} for mn+1m\leq n+1 follow from the analog of Eq. (4).

It remains to check the bi-Lipschitz condition for the resulting cocycle ρ𝔉1,𝔉2\rho_{\mathfrak{F}_{1},\mathfrak{F}_{2}}. It is easier to work with the cocycle ρ𝔉2,𝔉1\rho_{\mathfrak{F}_{2},\mathfrak{F}_{1}}, which for x,x+rXnx,x+r\in X^{\prime}_{n} satisfies

ρ𝔉2,𝔉1(r,x)=gn(ϕn(x)+r)gn(ϕn(x)),\rho_{\mathfrak{F}_{2},\mathfrak{F}_{1}}(r,x)=g_{n}(\phi_{n}(x)+r)-g_{n}(\phi_{n}(x)),

and is therefore (α,α+)(\alpha^{-},\alpha^{+})-bi-Lipschitz, because so is gng_{n}. Hence, ρ𝔉2,𝔉1\rho_{\mathfrak{F}_{2},\mathfrak{F}_{1}}is also (α1,α)(\alpha^{-1},\alpha)-bi-Lipschitz, because α1<α<α+<α\alpha^{-1}<\alpha^{-}<\alpha^{+}<\alpha by the choice of KK. Finally, we apply Remark 3.10 to conclude that ρ𝔉1,𝔉2\rho_{\mathfrak{F}_{1},\mathfrak{F}_{2}} is also (α1,α)(\alpha^{-1},\alpha)-bi-Lipschitz. ∎

Restricting the action of 𝔉2\mathfrak{F}_{2} onto the integer grid ZZ, we get the following corollary.

Corollary 3.13.

Let 𝔉\mathfrak{F} be a free Borel d\mathbb{R}^{d}-flow on XX. For any α>1\alpha>1 there exist a cross-section ZXZ\subseteq X and a free d\mathbb{Z}^{d}-action TT on ZZ such that the cocycle ρ=ρ𝔉,T:d×Xd\rho=\rho_{\mathfrak{F},T}:\mathbb{Z}^{d}\times X\to\mathbb{R}^{d} given by Tnx=x+ρ(n,x)T_{n}x=x+\rho(n,x) is (α1,α)(\alpha^{-1},\alpha)-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 d\mathbb{R}^{d}-flows with free d\mathbb{Z}^{d}-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 GXG\curvearrowright X, a (Borel) cocycle with values in a group HH is a (Borel) map ρ:G×XH\rho:G\times X\to H that satisfies the cocycle identity:

ρ(g2g1,x)=ρ(g2,g1x)ρ(g1,x)for all g1,g2G and xX.\rho(g_{2}g_{1},x)=\rho(g_{2},g_{1}x)\rho(g_{1},x)\quad\textrm{for all $g_{1},g_{2}\in G$ and $x\in X$}.

We are primarily concerned with the Abelian groups d\mathbb{Z}^{d} and d\mathbb{R}^{d} in this section, so the cocycle identity will be written additively. A cocycle ρ:G×XH\rho:G\times X\to H is said to be injective if ρ(g,x)eH\rho(g,x)\neq e_{H} for all geGg\neq e_{G} and all xXx\in X, where eGe_{G} and eHe_{H} are the identity elements of the corresponding groups. Suppose that furthermore the groups GG and HH are locally compact. We say that ρ\rho escapes to infinity if for all xXx\in X, limgρ(g,x)=+\lim_{g\to\infty}\rho(g,x)=+\infty in the sense that for any compact KHHK_{H}\subseteq H there exists a compact KGGK_{G}\subseteq G such that ρ(g,x)KH\rho(g,x)\not\in K_{H} whenever gKGg\not\in K_{G}.

Example 4.1.

Suppose aH:HXa_{H}:H\curvearrowright X and aG:GYa_{G}:G\curvearrowright Y, YXY\subseteq X, are free actions of groups GG and HH on standard Borel spaces, and suppose that we have containment of orbit equivalence relations EGEHE_{G}\subseteq E_{H}. For each yYy\in Y and gGg\in G, there exists a unique ρaH,aG(g,y)H\rho_{a_{H},a_{G}}(g,y)\in H such that aH(ρaH,aG(g,y),y)=aG(g,y)a_{H}(\rho_{a_{H},a_{G}}(g,y),y)=a_{G}(g,y). The map (g,y)ρaH,aG(g,y)(g,y)\mapsto\rho_{a_{H},a_{G}}(g,y) 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 \mathbb{R}-flows and \mathbb{Z}-actions are tightly connected through the “flow under a function” construction. Let T:ZZT:Z\to Z be a free Borel automorphism of a standard Borel space and f:Z>0f:Z\to\mathbb{R}^{>0} be a positive Borel function. There is a natural definition of a flow 𝔉:X\mathfrak{F}:\mathbb{R}\curvearrowright X on the space X={(z,t):zZ,0t<f(z)}X=\{(z,t):z\in Z,0\leq t<f(z)\} under the graph of ff. The action (z,t)+r(z,t)+r for a positive rr is defined by shifting the point (z,t)(z,t) by rr units upward until the graph of ff is reached, then jumping to the point (Tz,0)(Tz,0), and continuing to flow upward until the graph of ff at TzTz is reached, etc. More formally,

(z,t)+r=(Tkz,t+ri=0k1f(Tiz))(z,t)+r=\bigl{(}T^{k}z,t+r-\sum\limits_{i=0}^{k-1}f(T^{i}z)\bigr{)}

for the unique k0k\geq 0 such that i=0k1f(Tiz)t+r<i=0kf(Tiz)\sum_{i=0}^{k-1}f(T^{i}z)\leq t+r<\sum_{i=0}^{k}f(T^{i}z); for r0r\leq 0 the action is defined by “flowing backward”, i.e.,

(z,t)+r=(Tkz,t+r+i=1kf(Tiz))(z,t)+r=\bigl{(}T^{-k}z,t+r+\sum\limits_{i=1}^{k}f(T^{-i}z)\bigr{)}

for k0k\geq 0 such that 0t+r+i=1kf(Tiz)<f(Tkz)0\leq t+r+\sum\limits_{i=1}^{k}f(T^{-i}z)<f(T^{-k}z). The action is well-defined provided that the fibers within the orbits of TT have infinite cumulative lengths:

(5) i=0f(Tiz)=+ and i=0f(Tiz)=+ for all zZ.\sum_{i=0}^{\infty}f(T^{i}z)=+\infty\quad\textrm{ and }\quad\sum_{i=0}^{\infty}f(T^{-i}z)=+\infty\quad\textrm{ for all }z\in Z.

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 TT be a free d\mathbb{Z}^{d}-action on a standard Borel space ZZ and let ρ:d×Zd\rho:\mathbb{Z}^{d}\times Z\to\mathbb{R}^{d} be a Borel cocycle. One can construct a d\mathbb{Z}^{d}-action T^\hat{T}, the so-called principal d\mathbb{R}^{d}-extension, defined on Z×dZ\times\mathbb{R}^{d} via T^n(z,r)=(Tnz,r+ρ(n,z))\hat{T}_{n}(z,r)=(T_{n}z,r+\rho(n,z)). An easy application of the cocycle identity verifies axioms of the action. While the action TT will typically have complicated dynamics, the action T^\hat{T} admits a Borel transversal as long as the cocycle ρ\rho escapes to infinity.

Lemma 4.2.

If the cocycle ρ\rho satisfies limnρ(n,z)=+\lim_{n\to\infty}||\rho(n,z)||=+\infty for all zZz\in Z, then the action T^\hat{T} has a Borel transversal.

Proof.

Let Yk={(z,r)Z×d:rk}Y_{k}=\{(z,r)\in Z\times\mathbb{R}^{d}:||r||\leq k\}. We claim that each orbit of T^\hat{T} intersects YkY_{k} in a finite (possibly empty) set. Indeed, cocycle values escaping to infinity yield for any (z,r)Z×d(z,r)\in Z\times\mathbb{R}^{d} a number NN so large that r+ρ(n,z)>k||r+\rho(n,z)||>k whenever nN||n||\geq N. In particular, nN||n||\geq N implies T^n(z,r)=(Tnz,r+ρ(n,z))Yk\hat{T}_{n}(z,r)=(T_{n}z,r+\rho(n,z))\not\in Y_{k}. Hence, the intersection of the orbit of (z,r)(z,r) with YkY_{k} is finite.

Set Y=k(YkndT^nYk1)Y=\bigsqcup_{k\in\mathbb{N}}\bigl{(}Y_{k}\setminus\bigcup_{n\in\mathbb{Z}^{d}}\hat{T}_{n}Y_{k-1}\bigr{)}. Each orbit of T^\hat{T} intersects YY in a finite and necessarily non-empty set, so ET^|YE_{\hat{T}}|_{Y} is a finite Borel equivalence relation. A Borel transversal for ET^|YE_{\hat{T}}|_{Y} is also a transversal for the action of T^\hat{T}. ∎

We assume now that the cocycle ρ\rho satisfies the assumptions of Lemma 4.2, and X=(Z×d)/ET^X=(Z\times\mathbb{R}^{d})/E_{\hat{T}} therefore carries the structure of a standard Borel space as a push-forward of the factor map π:Z×dX\pi:Z\times\mathbb{R}^{d}\to X, which sends a point to its ET^E_{\hat{T}}-equivalence class.

There is a natural d\mathbb{R}^{d}-flow 𝔉^\hat{\mathfrak{F}} on Z×dZ\times\mathbb{R}^{d} which acts by shifting the second coordinate: (z,r)+𝔉^s=(z,r+s)(z,r)+_{\hat{\mathfrak{F}}}s=(z,r+s). This flow commutes with the d\mathbb{Z}^{d}-action T^\hat{T} and therefore projects onto the flow 𝔉\mathfrak{F} on XX given by the condition π((z,r)+𝔉^s)=π(z,r)+𝔉s\pi((z,r)+_{\hat{\mathfrak{F}}}s)=\pi(z,r)+_{\mathfrak{F}}s. We say that 𝔉\mathfrak{F} is the special flow over TT generated by the cocycle ρ\rho. Freeness of TT implies freeness of 𝔉\mathfrak{F}.

The construction outlined above, works just as well in the context of ergodic theory, where the space ZZ would be endowed with a finite measure ν\nu preserved by the action TT. The product of ν\nu with the Lebesgue measure on d\mathbb{R}^{d} induces a measure μ\mu on XX, which is preserved by the flow 𝔉\mathfrak{F}. Furthermore, μ\mu is finite provided the cocycle ρ\rho 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 11.

As will be shown shortly, such a representation result continues to hold in the framework of descriptive set theory, and every free Borel d\mathbb{R}^{d}-flow is Borel isomorphic to a special flow over some free Borel d\mathbb{Z}^{d}-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 d\mathbb{Z}^{d}-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 Zz(z,0)Z×{0}Z\ni z\mapsto(z,0)\in Z\times\{0\} be denoted by ι\iota. If the cocycle ρ\rho is injective, then πι:Zπ(Z×{0})=Y\pi\circ\iota:Z\to\pi(Z\times\{0\})=Y is a bijection and YY intersects every orbit of 𝔉\mathfrak{F} in a non-empty countable set. The d\mathbb{Z}^{d}-action TT on ZZ can be transferred via πι\pi\circ\iota to give a free d\mathbb{Z}^{d}-action T=πιTι1π1T^{\prime}=\pi\circ\iota\circ T\circ\iota^{-1}\circ\pi^{-1} on YY. Let ρ=ρT,𝔉:d×Yd\rho^{\prime}=\rho_{T^{\prime},\mathfrak{F}}:\mathbb{Z}^{d}\times Y\to\mathbb{R}^{d} be the cocycle of the action πιTι1π1\pi\circ\iota\circ T\circ\iota^{-1}\circ\pi^{-1}; in other words

(6) Tn(y)=(πιTnι1π1)(y)=y+𝔉ρ(n,y) for all nd and yY.T^{\prime}_{n}(y)=(\pi\circ\iota\circ T_{n}\circ\iota^{-1}\circ\pi^{-1})(y)=y+_{\mathfrak{F}}\rho^{\prime}(n,y)\quad\textrm{ for all $n\in\mathbb{Z}^{d}$ and $y\in Y$}.

If y=(πι)(z)y=(\pi\circ\iota)(z) for zZz\in Z, then Eq. (6) translates into

π(Tnz,0)=π(z,ρ(n,y)).\pi(T_{n}z,0)=\pi(z,\rho^{\prime}(n,y)).

Since π(Tnz,0)=π(z,ρ(n,Tnz))=π(z,ρ(n,z))\pi(T_{n}z,0)=\pi(z,\rho(-n,T_{n}z))=\pi(z,-\rho(n,z)), we conclude that ρ(n,y)=ρ(n,z)\rho^{\prime}(n,y)=-\rho(n,z), where y=(πι)(z)y=(\pi\circ\iota)(z). In particular, YY is a discrete cross-section for the flow 𝔉\mathfrak{F} precisely because ρ\rho escapes to infinity.

Conversely, if 𝔉\mathfrak{F} is any free d\mathbb{R}^{d}-flow on a standard Borel space XX, and ZXZ\subseteq X is a discrete cross-section with a d\mathbb{Z}^{d}-action TT on it, then 𝔉\mathfrak{F} is isomorphic to the special flow over TT generated by the (necessarily injective) cocycle ρT,𝔉-\rho_{T,\mathfrak{F}}.

Let us say that a cocycle ρ\rho 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 d\mathbb{Z}^{d}-action on a discrete cross-section of the flow.

For instance, given any free d\mathbb{Z}^{d}-action TT on ZZ, we may consider the admissible cocycle ρ(n,z)=n\rho(n,z)=-n for all zZz\in Z and ndn\in\mathbb{Z}^{d}. The set Y=π(Z×{0})Y=\pi(Z\times\{0\}) is then an integer grid for the flow 𝔉\mathfrak{F} (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 d\mathbb{R}^{d}-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 d\mathbb{Z}^{d}-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 d\mathbb{Z}^{d}-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 α>1\alpha>1. Any free ergodic measure-preserving d\mathbb{R}^{d}-flow on a standard Lebesgue space is isomorphic to a special flow over a free ergodic measure-preserving d\mathbb{Z}^{d}-action generated by an (α1,α)(\alpha^{-1},\alpha)-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 d\mathbb{R}^{d}-flows and establishes isomorphism on all orbits.

Theorem 4.3.

Pick some α>1\alpha>1. Any free Borel d\mathbb{R}^{d}-flow is isomorphic to a special flow over a free Borel d\mathbb{Z}^{d}-action generated by an (α1,α)(\alpha^{-1},\alpha)-bi-Lipschitz cocycle.

Proof.

Let 𝔉\mathfrak{F} be a free Borel d\mathbb{R}^{d}-flow on XX. Corollary 3.13 gives a cross-section ZXZ\subseteq X and a d\mathbb{Z}^{d}-action TT on it such that the cocycle ρ𝔉,T:d×Xd\rho_{\mathfrak{F},T}:\mathbb{Z}^{d}\times X\to\mathbb{R}^{d} is (α1,α)(\alpha^{-1},\alpha)-bi-Lipschitz. By the discussion in Section 4.4, this gives a representation of the flow as a special flow over TT generated by the cocycle ρ𝔉,T-\rho_{\mathfrak{F},T}, which is also (α1,α)(\alpha^{-1},\alpha)-bi-Lipschitz. ∎

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 n{{\mathbb{Z}}}^{n} 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 K{{K}}-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 d{{\mathbb{R}}}^{d} 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, L1L^{1} 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 n{{\mathbb{R}}}^{n} and free n{{\mathbb{R}}}^{n}-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