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

spacing=nonfrench

Smooth orbit equivalence
of multidimensional Borel flows

Konstantin Slutsky Department of Mathematics
Iowa State University
411 Morrill Road
Ames, IA 50011
kslutsky@gmail.com
Abstract.

Free Borel d\mathbb{R}^{d}-flows are smoothly equivalent if there is a Borel bijection between the phase spaces that maps orbits onto orbits and is a CC^{\infty}-smooth orientation preserving diffeomorphism between orbits. We show that all free non-tame Borel d\mathbb{R}^{d}-flows are smoothly equivalent in every dimension d2d\geq 2. This answers a question of B. Miller and C. Rosendal.

Key words and phrases:
Orbit equivalence, time-change equivalence, smooth orbit equivalence, Borel flow
Konstantin Slutsky’s research is partially supported by the ANR project AGRUME (ANR-17-CE40-0026).

1. Introduction

Let us begin by defining the notions mentioned in the title as well as the related concepts that are needed to state the main results of our work. A Borel flow is a Borel action dΩ\mathbb{R}^{d}\curvearrowright\Omega of the Euclidean group on a standard Borel space Ω\Omega. An action of rd\vec{r}\in\mathbb{R}^{d} upon xΩx\in\Omega is denoted by x+rx+\vec{r}. An orbit equivalence between two flows dΩ\mathbb{R}^{d}\curvearrowright\Omega and dΩ\mathbb{R}^{d}\curvearrowright\Omega^{\prime} is a Borel bijection ξ:ΩΩ\xi:\Omega\to\Omega^{\prime} that sends orbits onto orbits: ξ(x+d)=ξ(x)+d\xi\bigl{(}x+\mathbb{R}^{d}\bigr{)}=\xi(x)+\mathbb{R}^{d} for all xΩx\in\Omega; when such a map exists, we say that the flows are orbit equivalent. For an action dΩ\mathbb{R}^{d}\curvearrowright\Omega we let EE denote the corresponding orbit equivalence relation: xEyx+d=y+dxEy\iff x+\mathbb{R}^{d}=y+\mathbb{R}^{d}. When the action is moreover free, ρ:Ed\rho:E\to\mathbb{R}^{d} will stand for the associated cocycle, determined uniquely by the condition x+ρ(x,y)=yx+\rho(x,y)=y. Given free Borel flows on phase spaces Ω\Omega and Ω\Omega^{\prime}, any orbit equivalence ξ:ΩΩ\xi:\Omega\to\Omega^{\prime} gives rise to a map αξ:Ω×dd\alpha_{\xi}:\Omega\times\mathbb{R}^{d}\to\mathbb{R}^{d} defined by αξ(x,r)=ρ(ξ(x),ξ(x+r))\alpha_{\xi}(x,\vec{r})=\rho\bigl{(}\xi(x),\xi(x+\vec{r})\bigr{)}. A Borel orbit equivalence ξ\xi is said to be a smooth equivalence if αξ(x,):dd\alpha_{\xi}(x,\,\cdot\,):\mathbb{R}^{d}\to\mathbb{R}^{d} is a CC^{\infty}-smooth orientation preserving diffeomorphism for all points xΩx\in\Omega.

1.1. Prior work

The concept of orbit equivalence originated in ergodic theory, where the set-up differs in two essential aspects. First, one endows phase spaces of flows with probability measures. The flows are then assumed to preserve (or to quasi-preserve) these measures. Likewise, orbit equivalence maps are required to be at least quasi-measure-preserving. Second, all the properties of interest are expected to hold up to a null set. For instance, an orbit equivalence map may mix elements between orbits as long as this behavior is confined to a set of measure zero. The latter is a notable relaxation of the Borel definition.

Smooth equivalence of one-dimensional flows, better known under the name of time-change equivalence, is closely connected to the notion of Kakutani equivalence of automorphisms [Kak43], and has been studied extensively since the pioneering works of J. Feldman [Fel76] and A. Katok[Kat75, Kat77]. An important milestone was the monograph of D. Ornstein, D. Rudolph, and B. Weiss [ORW82], which showed, in particular, that there is a continuum of pairwise time-change inequivalent ergodic measure-preserving flows. The higher-dimensional case was considered by D. Rudolph [Rud79], where he found a striking difference with the one-dimensional situation—all ergodic measure-preserving d\mathbb{R}^{d}-flows, d2d\geq 2, are smoothly equivalent. J. Feldman obtained a similar result for quasi-measure-preserving flows in [Fel91, Fel92].

Note that in the definition of time-change equivalence, it is essential to allow for the orbit equivalence maps to be quasi-measure-preserving even if all the flows are assumed to be measure-preserving (see [Nak88, Remark 4.5] regarding the connection between the integrability of the cocycle as required in [Kat75] and measure class preservation of the orbit equivalence). This underlines the strength of Rudolph’s result, as it is shown in [Rud79, Proposition 1.1] that in the dimensions d2d\geq 2 there is a single class of ergodic measure-preserving flows under measure-preserving smooth equivalence relation. Further discussion on what restrictions on the orbit equivalence maps may produce finer equivalence relations among higher-dimensional flows can also be found in [Rud79].

In this paper, we are interested in the descriptive set-theoretical viewpoint. This means that neither flows are assumed to preserve any measures (thus increasing the pool of flows to consider), nor orbit equivalence maps have to be quasi-measure-preserving (which may potentially collapse previously inequivalent flows into the same class). On the other hand, the necessity to run constructions on all orbits may, in principle, increase the number of equivalence classes, as complicated dynamics of a flow can be contained in a null set. All in all, this framework is in a general position to the one of ergodic theory, and ahead of time, it is not clear how versatile smooth equivalence will turn out to be. The key work that investigated the subject from this purely Borel vantage point is the article by B. Miller and C. Rosendal [MR10], where they studied Kakutani equivalence of Borel automorphisms and classified one-dimensional flows up to descriptive time-change equivalence.

Theorem 1 (Miller–Rosendal [MR10, Theorem B]).

All non-tame111A flow dΩ\mathbb{R}^{d}\curvearrowright\Omega is tame if there is a Borel set SΩS\subseteq\Omega that intersects every orbit of the flow in a single point. The term smooth is often used in the literature instead, but since we also work with diffeomorphisms, this word will be used in the traditional sense of differential geometry. Tame flows should be considered trivial in the context of the questions we are interested in this paper. free Borel \mathbb{R}-flows are smoothly equivalent.

As the one-dimensional case has been settled, they posed [MR10, Problem C] the following problem: “Classify free Borel d\mathbb{R}^{d}-actions on Polish spaces up to (CC^{\infty}-)time-change isomorphism.” In other words, does the analog of D. Rudolph and J. Feldman theorems hold? Are there two (non-tame) inequivalent free Borel d\mathbb{R}^{d}-flows for any d2d\geq 2?

These and related topics were studied in [Slu19], where we showed that any two non-tame free d\mathbb{R}^{d}-flows, d2d\geq 2, are smoothly equivalent up to a compressible set. The method to prove this result was an expansion of the one used in [Fel91], and such a statement is about as far as ergodic-theoretical methods can go, since a compressible set has measure zero relative to any probability measure invariant under the flow.

1.2. Main results

In the present work, we give a complete answer to Problem C of [MR10] by showing that all non-tame free d\mathbb{R}^{d}-flows, d2d\geq 2, are smoothly equivalent (Theorem 21). Table LABEL:number-classes provides a concise summary and compares the number of classes up to smooth equivalence in ergodic theory and Borel dynamics.

\ctable

[ label=number-classes, pos=htb, caption = Number of classes of smooth orbit equivalence., mincapwidth =]ccc Ergodic Theory Borel Dynamics
d=1d=1 𝔠\mathfrak{c}-many [ORW82] one [MR10]
d2d\geq 2 one [Rud79, Fel91] one

Many results in ergodic theory and Borel dynamics of d\mathbb{Z}^{d} and d\mathbb{R}^{d} actions are based on the fact that such actions are (essentially) hyperfinite. In ergodic theory, this is manifested by a group of related theorems that usually go under the name of “Rokhlin Lemma”. The key idea here is that one can find a measurable set that intersects every orbit of the flow in a set of pairwise disjoint rectangles (more precisely, dd-dimensional parallelepipeds). Moreover, one often takes a sequence of such sets, where rectangles cohere and eventually cover all the orbits (at least, up to a null set). The details of the assumptions on such regions vary, but a construction of this form is present in many arguments, including the references above. The direct analog of such a tower of coherent rectangular regions is not possible in Borel dynamics. One, therefore, has to rely on more complicated geometric shapes (see, for instance, [JKL02, Theorem 1.16] and [GJ15]).

Our argument also requires regions witnessing hyperfiniteness. The key property we need is for them to be smooth disks. S. Gao, S. Jackson, E. Krohne, and B. Seward [GJKS] have shown the possibility to construct such regions for low-dimensional flows. Their argument is an elaboration of the orthogonal marker regions technique developed in [GJ15]. We take a different path and build upon the approach presented by A. Marks and S. Unger in [MU17, Appendix A]. Section 2 is devoted to these topics and it leads to Theorem 5 that shows the existence of such disk-shaped regions in all dimensions.

In order to prove that all non-tame free d\mathbb{R}^{d}-flows are smoothly equivalent, we leverage the work of Miller–Rosendal that handles the case of d=1d=1. To this end Section 3 introduces the concept of a special flow, which is a type of an d\mathbb{R}^{d}-flow that is build over a one-dimensional flow in a very primitive way. We show in Theorem 20 that all d\mathbb{R}^{d}-flows are smoothly equivalent to a special flow. This piece is the technical core of this paper.

Finally, in Section 4 we prove the main result on smooth equivalence of free d\mathbb{R}^{d}-flows (Theorem 21) and conclude with some remarks on its potential strengthening.

1.3. Notations

The following notations are used throughout the paper: B(R)dB(R)\subseteq\mathbb{R}^{d} denotes a ball of radius RR centered at the origin; ||||||\,\cdot\,|| stands for the 2\ell^{2}- norm in d\mathbb{R}^{d}; and dist(r,r)\mathrm{dist}(\vec{r},\vec{r}\,^{\prime}) refers to the Euclidean distance in d\mathbb{R}^{d}. By a diffeomorphism we always mean a CC^{\infty}-smooth orientation preserving diffeomorphism. A smooth disk therefore refers to any compact region in d\mathbb{R}^{d} that is diffeomorphic to a ball. Interior of a set FdF\subseteq\mathbb{R}^{d} is denoted by intF\operatorname{\mathrm{int}}F, and F\partial{F} stands for the boundary of FF. Given a Cartesian product X1×X2××XmX_{1}\times X_{2}\times\cdots\times X_{m}, projk:i=1mXiXk\mathrm{proj}_{k}:\prod_{i=1}^{m}X_{i}\to X_{k} denotes the projection onto the kthk^{\textrm{th}} coordinate, and, more generally, proj[k,l]\mathrm{proj}_{[k,l]} will denote the projection onto Xk××XlX_{k}\times\cdots\times X_{l}, for klk\leq l.

1.4. Acknowledgement

The author expresses his appreciation to Todor Tsankov for numerous helpful discussions on the topic of this paper.

2. Disk-Shaped Coherent Regions

We begin by stating the following classical fact from differential topology (see, for instance, [Fel91, Proposition 2.6]), which will be used throughout the paper to justify the existence of diffeomorphisms moving disks in a prescribed fashion.

Lemma 2 (Extension Lemma).

Let FF and FF^{\prime} be smooth disks in d\mathbb{R}^{d}, d2d\geq 2, each containing mm smooth disks in its interior: D1,,DmintFD_{1},\ldots,D_{m}\subset\operatorname{\mathrm{int}}F and D1,,DmintFD_{1}^{\prime},\ldots,D_{m}^{\prime}\subset\operatorname{\mathrm{int}}F^{\prime}. Suppose that disks DiD_{i} are pairwise disjoint and so are the disks DiD_{i}^{\prime}. Any collection of orientation preserving diffeomorphisms ϕi:DiDi\phi_{i}:D_{i}\to D_{i}^{\prime} can be extended to an orientation preserving diffeomorphism ψ:FF\psi:F\to F^{\prime}.

Theorems in Borel dynamics of d\mathbb{R}^{d} and d\mathbb{Z}^{d} actions often rely on variants of the hyperfiniteness construction. Our argument is no exception, and this section gives the specific version to be used later in Section 3. The cases of d=2d=2 and d=3d=3 of Theorem 5 are due to S. Gao, S. Jackson, E. Krohne, and B. Seward; they are announced to appear in [GJKS]. We borrow the structure of our argument from A. Marks and S. Unger [MU17, Appendix A] and supplement it with Lemma 3 to get the desired shape of the regions for all dimensions d2d\geq 2.

Lemma 3 (Separation Lemma).

Let 0<R1<R20<R_{1}<R_{2} be positive reals and let D1,,DndD_{1},\ldots,D_{n}\subset\mathbb{R}^{d}, d2d\geq 2, be smooth pairwise disjoint disks of diameter diam(Di)<(R2R1)/2\mathrm{diam}(D_{i})<(R_{2}-R_{1})/2. There exists a smooth disk FF wedged between the two balls, B(R1)FB(R2)B(R_{1})\subseteq F\subseteq B(R_{2}), such that for each ii either DiintFD_{i}\subseteq\operatorname{\mathrm{int}}F or FDi=F\cap D_{i}=\varnothing.

R1R_{1}R2R_{2}FFRefer to caption
Figure 1. Separation Lemma.

Figure 1 illustrates the statement. Disks DiD_{i} are marked in gray and the required disk FF is dashed.

Proof.

The proof is by induction on the number of disks nn. The base case n=0n=0 is trivial, we argue the step from n1n-1 to nn. If none of the disks DiD_{i} lie inside the open annulus A=intB(R2)B(R1)A=\operatorname{\mathrm{int}}B(R_{2})\setminus B(R_{1}), then the ball F=B((R2+R1)/2)F=B((R_{2}+R_{1})/2) works. Otherwise, select a ball Di0AD_{i_{0}}\subset A. By the inductive assumption there is a disk FF^{\prime} that fulfills the conclusions of the lemma for all disks DiD_{i}, ii0i\neq i_{0}. We are done if also Di0F=D_{i_{0}}\cap F^{\prime}=\varnothing or Di0intFD_{i_{0}}\subset\operatorname{\mathrm{int}}F^{\prime}, so assume otherwise (Figure 2(a)).

Find a smooth disk GAG\subset A that contains Di0intGD_{i_{0}}\subset\operatorname{\mathrm{int}}G in its interior and does not intersect any other disk DiD_{i}. Pick a disk ZintGZ\subset\operatorname{\mathrm{int}}G that is disjoint from F\partial{F}^{\prime} (Figure 2(b)). Such a disk can be found, since the boundary F\partial{F}^{\prime} is nowhere dense. Choose a diffeomorphism ψ\psi supported on GG such that ψ(Di0)=Z\psi(D_{i_{0}})=Z. Lemma 2 may be used to justify the existence of such a diffeomorphism. We have either ψ(Di0)intF\psi(D_{i_{0}})\subset\operatorname{\mathrm{int}}F^{\prime} or ψ(Di0)F=\psi(D_{i_{0}})\cap F^{\prime}=\varnothing. Set F=ψ1(F)F=\psi^{-1}(F^{\prime}) (Figure 2(c)).

Di0D_{i_{0}}FF^{\prime}
(a)
GGZZ
(b)
FF
(c)
Figure 2. Construction of a disk FF that separates disks DiD_{i}.

Since ψ\psi is supported on GG, both conditions FDi=F\cap D_{i}=\varnothing and DiintFD_{i}\subset\operatorname{\mathrm{int}}F, ii0i\neq i_{0}, continue to hold whenever they did so for FF^{\prime} instead of FF. By construction we now also have either FDi0=F\cap D_{i_{0}}=\varnothing or Di0intFD_{i_{0}}\subset\operatorname{\mathrm{int}}F. ∎

Let dΩ\mathbb{R}^{d}\curvearrowright\Omega be a free Borel flow, and let EE be its orbit equivalence relation. A set 𝒞Ω\mathcal{C}\subset\Omega is said to be

  • RR-discrete, where RR is a positive real, if (c+B(R))(c+B(R))=(c+B(R))\cap(c^{\prime}+B(R))=\varnothing for all distinct c,c𝒞c,c^{\prime}\in\mathcal{C};

  • discrete if it is RR-discrete for some R>0R>0;

  • cocompact if there exists R>0R>0 such that 𝒞+B(R)=Ω\mathcal{C}+B(R)=\Omega;

  • complete if it intersects every orbit of the action;

  • a cross section if it is discrete and complete;

  • on a rational grid (or simply rational, for short) if ρ(c,c)d\rho(c,c^{\prime})\in\mathbb{Q}^{d} for all c,c𝒞c,c^{\prime}\in\mathcal{C} such that cEccEc^{\prime}.

We make use of the following result due to C. M. Boykin and S. Jackson.

Lemma 4 (Boykin–Jackson [BJ07], cf. Lemma A.2 of [MU17]).

Let a1<a2<a_{1}<a_{2}<\cdots be an increasing sequence of natural numbers. For any free Borel flow dΩ\mathbb{R}^{d}\curvearrowright\Omega there exists a sequence of aia_{i}-discrete cocompact cross sections 𝒞i\mathcal{C}_{i} such that i𝒞i\bigcup_{i}\mathcal{C}_{i} is rational and for all ϵ>0\epsilon>0, for every xΩx\in\Omega, there are infinitely many ii such that ρ(x,c)<ϵai||\rho(x,c)||<\epsilon a_{i} for some c𝒞ic\in\mathcal{C}_{i}.

Proof.

The direct adaptation to d\mathbb{R}^{d}-flows of the argument [MU17, Lemma A.2] (presented therein for d\mathbb{Z}^{d} actions) produces cross sections 𝒞i\mathcal{C}^{\prime}_{i} that satisfy all the conclusions except possibly for i𝒞i\bigcup_{i}\mathcal{C}^{\prime}_{i} being rational. As shown in [Slu19, Lemma 2.3], there is a rational grid for the flow, i.e., there is a complete rational set QΩQ\subset\Omega invariant under the action of d\mathbb{Q}^{d}. Using Luzin-Novikov Theorem (see [Kec95, 18.14]), one can find cross sections 𝒞iQ\mathcal{C}_{i}\subset Q and Borel bijections ζi:𝒞i𝒞i\zeta_{i}:\mathcal{C}_{i}^{\prime}\to\mathcal{C}_{i} such that i𝒞i\bigcup_{i}\mathcal{C}_{i} is rational and for all c𝒞ic\in\mathcal{C}_{i}^{\prime} one has cEζi(c)cE\zeta_{i}(c) and ρ(c,ζi(c))<1||\rho(c,\zeta_{i}(c))||<1. In other words, every element in 𝒞i\mathcal{C}_{i}^{\prime} can be shifted by distance <1<1 to ensure that all the cross sections are on the same rational grid. This argument is the content of [Slu19, Lemma 2.4].

The cross sections 𝒞i\mathcal{C}_{i} continue to be cocompact and still satisfy the key property that for every xΩx\in\Omega and ϵ>0\epsilon>0 there are infinitely many ii with ρ(c,x)<ϵai||\rho(c,x)||<\epsilon a_{i} for some c𝒞ic\in\mathcal{C}_{i}. The only minor issue is that this modification reduces the discreteness parameter by 11. Therefore if the original cross sections 𝒞i\mathcal{C}_{i}^{\prime} were chosen to be (ai+1)(a_{i}+1)-discrete, then each of 𝒞i\mathcal{C}_{i} is guaranteed to be aia_{i}-discrete. ∎

To formulate the next theorem we need an extra bit of notation. Let dΩ\mathbb{R}^{d}\curvearrowright\Omega be a free Borel flow. For a set WΩ×ΩW\subseteq\Omega\times\Omega and cΩc\in\Omega we let W(c)W(c) denote the slice over cc, i.e., W(c)={xΩ:(c,x)W}W(c)=\{x\in\Omega:(c,x)\in W\}. We also denote by W~\widetilde{W} the set {(c,r)Ω×d:c+rW(c)}\bigl{\{}(c,\vec{r}\,)\in\Omega\times\mathbb{R}^{d}:c+\vec{r}\in W(c)\bigr{\}}. Note that W~(c)\widetilde{W}(c) is the region of d\mathbb{R}^{d} described by W(c)W(c), when cc is taken to be the origin of the coordinate system.

Theorem 5.

Let dΩ\mathbb{R}^{d}\curvearrowright\Omega be a free Borel flow and let EE denote its orbit equivalence relation. There exist cross sections 𝒞n\mathcal{C}_{n} and Borel sets Wn(𝒞n×Ω)EW_{n}\subseteq(\mathcal{C}_{n}\times\Omega)\cap E such that n𝒞n\bigcup_{n}\mathcal{C}_{n} is rational and for all nn\in\mathbb{N}:

  1. (i)

    W~n(c)\widetilde{W}_{n}(c) is a smooth disk for every c𝒞nc\in\mathcal{C}_{n}.

  2. (ii)

    Sets Wn(c)W_{n}(c), c𝒞nc\in\mathcal{C}_{n}, are pairwise disjoint.

  3. (iii)

    For every c𝒞mc^{\prime}\in\mathcal{C}_{m}, m<nm<n, and every c𝒞nc\in\mathcal{C}_{n}, either Wm(c)Wn(c)=W_{m}(c^{\prime})\cap W_{n}(c)=\varnothing or Wm(c)Wn(c)W_{m}(c^{\prime})\subseteq W_{n}(c). Moreover, in the latter case ρ(c,c)+W~m(c)\rho(c,c^{\prime})+\widetilde{W}_{m}(c^{\prime}) is contained in the interior of W~n(c)\widetilde{W}_{n}(c).

  4. (iv)

    For all xΩx\in\Omega and all compact KdK\subset\mathbb{R}^{d} there are mm and c𝒞mc\in\mathcal{C}_{m} such that x+KWm(c)x+K\subseteq W_{m}(c).

  5. (v)

    There are smooth disks An,kdA_{n,k}\subseteq\mathbb{R}^{d}, kk\in\mathbb{N}, and a Borel partition 𝒞n=k𝒞n,k\mathcal{C}_{n}=\bigsqcup_{k\in\mathbb{N}}\mathcal{C}_{n,k} such that

    Wn=k{(c,c+r):c𝒞n,k,rAn,k}andW~n=k{(c,r):c𝒞n,k,rAn,k}.W_{n}=\bigsqcup_{k}\bigl{\{}(c,c+\vec{r}\,):c\in\mathcal{C}_{n,k},\vec{r}\in A_{n,k}\bigr{\}}\quad\textrm{and}\quad\widetilde{W}_{n}=\bigsqcup_{k}\bigl{\{}(c,\vec{r}\,):c\in\mathcal{C}_{n,k},\vec{r}\in A_{n,k}\bigr{\}}.
Proof.

Set an=5na_{n}=5^{n}, and let 𝒞n\mathcal{C}_{n} be a sequence of cross sections produced by Lemma 4. Note that n𝒞n\bigcup_{n}\mathcal{C}_{n} is guaranteed to be rational. We construct a sequence of Borel sets Wn(𝒞n×Ω)EW_{n}\subseteq(\mathcal{C}_{n}\times\Omega)\cap E, which will also satisfy

(1) B(an/2)W~n(c)B(an)for all c𝒞n.B(a_{n}/2)\subseteq\widetilde{W}_{n}(c)\subseteq B(a_{n})\quad\textrm{for all }c\in\mathcal{C}_{n}.

This property will later be helpful in establishing item (iv).

For the base of the argument set W1={(c,c+r):c𝒞1,rB(a1)}W_{1}=\bigl{\{}(c,c+\vec{r}\,):c\in\mathcal{C}_{1},\vec{r}\in B(a_{1})\bigr{\}}. Note that item (v) holds with a trivial partition 𝒞1,1=𝒞1\mathcal{C}_{1,1}=\mathcal{C}_{1}, 𝒞1,j=\mathcal{C}_{1,j}=\varnothing for j2j\geq 2, and A1,1=B(a1)A_{1,1}=B(a_{1}). Suppose now that WiW_{i} have been constructed for i<ni<n and satisfy all the items of the theorem. Cross section 𝒞n\mathcal{C}_{n} is ana_{n}-discrete, so regions c+B(an)c+B(a_{n}) are pairwise disjoint as cc ranges over 𝒞n\mathcal{C}_{n}.

For a given c𝒞nc\in\mathcal{\mathcal{C}}_{n} we consider regions Wi(c)W_{i}(c^{\prime}), i<ni<n, that intersect c+B(an)c+B(a_{n}) and that are not contained in a bigger such region. More formally, begin by choosing all the elements c1n1,,cln1n1𝒞n1c_{1}^{n-1},\ldots,c_{l_{n-1}}^{n-1}\in\mathcal{C}_{n-1} such that Wn1(cjn1)(c+B(an))W_{n-1}(c_{j}^{n-1})\cap(c+B(a_{n}))\neq\varnothing; next, pick all c1n2,,cln2n2𝒞n2c_{1}^{n-2},\ldots,c_{l_{n-2}}^{n-2}\in\mathcal{C}_{n-2} such that Wn2(cjn2)(c+B(an))W_{n-2}(c_{j}^{n-2})\cap(c+B(a_{n}))\neq\varnothing and Wn2(cjn2)Wn1(cin1)=W_{n-2}(c_{j}^{n-2})\cap W_{n-1}(c_{i}^{n-1})=\varnothing for all 1iln11\leq i\leq l_{n-1}; continue in the same fashion, terminating in a collection c11,,cl11𝒞1c^{1}_{1},\ldots,c^{1}_{l_{1}}\in\mathcal{C}_{1} such that W1(cj1)(c+B(an))W_{1}(c_{j}^{1})\cap(c+B(a_{n}))\neq\varnothing and W1(cj1)Wk(cik)=W_{1}(c_{j}^{1})\cap W_{k}(c_{i}^{k})=\varnothing for all 2k<n2\leq k<n, and all 1ilk1\leq i\leq l_{k}. Note that in view of Eq. (1), there can only be finitely many points cikc_{i}^{k} at each step. Let c1,,cli<n𝒞ic_{1},\ldots,c_{l}\in\bigcup_{i<n}\mathcal{C}_{i} be an enumeration of the elements cikc_{i}^{k}, 1k<n1\leq k<n, 1ilk1\leq i\leq l_{k}, and let for 1jl1\leq j\leq l, the number i(j)i(j) be such that cj𝒞i(j)c_{j}\in\mathcal{C}_{i(j)}.

Sets Wi(j)(cj)W_{i(j)}(c_{j}) are pairwise disjoint, and we therefore find ourselves in the set up of Lemma 3, where the ball B(an)B(a_{n}) interacts with a number of pairwise disjoint smooth disks ρ(c,cj)+W~i(j)(cj)\rho(c,c_{j})+\widetilde{W}_{i(j)}(c_{j}), each having diameter 2an1<an/2\leq 2*a_{n-1}<a_{n}/2. Lemma 3 claims that we can find a smooth disk FF squeezed according to B(an/2)FB(an)B(a_{n}/2)\subseteq F\subseteq B(a_{n}), and such that every region ρ(c,cj)+W~i(j)(cj)\rho(c,c_{j})+\widetilde{W}_{i(j)}(c_{j}) is either contained in the interior of FF or is disjoint from it. Set Wn(c)={c+r:rF}W_{n}(c)=\{c+\vec{r}:\vec{r}\in F\} and note that W~n(c)\widetilde{W}_{n}(c) fulfills Eq. (1).

We claim that this construction can be done in such a way that only countably many distinct shapes for FF are used. Indeed, the input to Lemma 3, which produced FF, is determined by the number ll of regions Wi(j)(cj)W_{i(j)}(c_{j}) intersecting c+B(an)c+B(a_{n}), by the shape of these regions, and by their location relative to cc. Since the union k𝒞k\bigcup_{k}\mathcal{C}_{k} is rational, the vector (ρ(c,c1),,ρ(c,cl))(\rho(c,c_{1}),\ldots,\rho(c,c_{l})) is in l\mathbb{Q}^{l}. By inductive assumption, for each cj𝒞i(j)=k𝒞i(j),kc_{j}\in\mathcal{C}_{i(j)}=\bigsqcup_{k}\mathcal{C}_{i(j),k}, there is some k(j)k(j)\in\mathbb{N} such that Wi(j)(cj)=cj+Ai(j),k(j)W_{i(j)}(c_{j})=c_{j}+A_{i(j),k(j)} for a smooth disk Ai(j),k(j)dA_{i(j),k(j)}\subseteq\mathbb{R}^{d}. Thus, the input to Lemma 3 is uniquely determined by the tuple

(l,ρ(c,c1),,ρ(c,cl),i(1),k(1),,i(l),k(l)).\Bigl{(}l,\ \rho(c,c_{1}),\ldots,\rho(c,c_{l}),\ i(1),k(1),\ldots,i(l),k(l)\Bigr{)}.

There are only countably many such tuples and we can assume that the same disk FF is used whenever the input tuple is the same. This guarantees compliance with item (v). Note also that such regions WnW_{n} are automatically Borel.

It remains to verify that sets WnW_{n} satisfy the rest of the conclusions of the theorem. Item (i) is fulfilled by the choice of FF. Item (ii) holds since 𝒞n\mathcal{C}_{n} is ana_{n}-discrete and FB(an)F\subseteq B(a_{n}). Compliance with item (iii) is the key property of the disk FF produced by Lemma 3.

We argue that item (iv) holds. Pick a point xΩx\in\Omega and a compact KdK\subset\mathbb{R}^{d}. Let n0n_{0} be so large that KB(an0/4)K\subseteq B(a_{n_{0}}/4). According to the property of cross sections 𝒞n\mathcal{C}_{n} guaranteed by Lemma 4, for ϵ=1/4\epsilon=1/4 there exists n1n0n_{1}\geq n_{0} such that ρ(x,c)<an1/4||\rho(x,c)||<a_{n_{1}}/4 for some c𝒞n1c\in\mathcal{C}_{n_{1}}, i.e., xc+B(an1/4)x\in c+B(a_{n_{1}}/4). We therefore have

x+Kx+B(an0/4)c+B(an1/4)+B(an0/4)c+B(an1/2)Wn1(c),x+K\subseteq x+B(a_{n_{0}}/4)\subseteq c+B(a_{n_{1}}/4)+B(a_{n_{0}}/4)\subseteq c+B(a_{n_{1}}/2)\subseteq W_{n_{1}}(c),

where the last inclusion follows from Eq. (1). ∎

In the proof above we chose a family of pairwise disjoint regions Wi(j)(cj)W_{i(j)}(c_{j}) that intersect c+B(an)c+B(a_{n}). In the sequel, we will need a similar family of subregions of a region Wn(c)W_{n}(c). The following lemma and definition isolate the relevant notion.

Lemma 6.

Let 𝒞n\mathcal{C}_{n} and WnW_{n}, nn\in\mathbb{N}, be as in Theorem 5. For each nn and each c𝒞nc\in\mathcal{C}_{n} there exists a family c1,,cli<n𝒞ic_{1},\ldots,c_{l}\in\bigcup_{i<n}\mathcal{C}_{i} such that for i(j)i(j) given by the condition cj𝒞i(j)c_{j}\in\mathcal{C}_{i(j)} one has

  1. (i)

    Wi(j)(cj)Wn(c)W_{i(j)}(c_{j})\subseteq W_{n}(c) for all 1jl1\leq j\leq l;

  2. (ii)

    sets Wi(j)(cj)W_{i(j)}(c_{j}) are pairwise disjoint for 1jl1\leq j\leq l;

  3. (iii)

    for any m<nm<n and c𝒞mc^{\prime}\in\mathcal{C}_{m} such that Wm(c)Wn(c)W_{m}(c^{\prime})\subseteq W_{n}(c) there exists 1jl1\leq j\leq l such that Wm(c)Wi(j)(cj)W_{m}(c^{\prime})\subseteq W_{i(j)}(c_{j}).

Proof.

Just like in the proof of Theorem 5, let c1n1,,cln1n1𝒞n1c_{1}^{n-1},\ldots,c_{l_{n-1}}^{n-1}\in\mathcal{C}_{n-1} be all the elements (if any) such that Wn1(cjn1)Wn(c)W_{n-1}(c_{j}^{n-1})\subseteq W_{n}(c). In view of 5(ii), sets Wn1(cjn1)W_{n-1}(c_{j}^{n-1}) are pairwise disjoint. Pick all the elements c1n2,,cln2n2𝒞n2c_{1}^{n-2},\ldots,c_{l_{n-2}}^{n-2}\in\mathcal{C}_{n-2} satisfying Wn2(cjn2)Wn(c)W_{n-2}(c_{j}^{n-2})\subseteq W_{n}(c), but Wn2(cjn2)W_{n-2}(c_{j}^{n-2}) is disjoint from all Wn1(cin1)W_{n-1}(c_{i}^{n-1}), 1iln11\leq i\leq l_{n-1}. Note that by 5(iii) the latter is equivalent to saying that Wn2(cjn2)W_{n-2}(c_{j}^{n-2}) is not contained in any of Wn1(ckn1)W_{n-1}(c_{k}^{n-1}), 1kln11\leq k\leq l_{n-1}.

One continues in the same fashion. At step kk we pick elements c1nk,,clnknk𝒞nkc_{1}^{n-k},\ldots,c_{l_{n-k}}^{n-k}\in\mathcal{C}_{n-k} that are contained in Wn(c)W_{n}(c) and are disjoint from all the sets Wnj(cinj)W_{n-j}(c_{i}^{n-j}), 1j<k1\leq j<k, 1ilnj1\leq i\leq l_{n-j}, constructed at the previous steps. The process terminates with the selection of elements c11,,cl11𝒞1c_{1}^{1},\ldots,c_{l_{1}}^{1}\in\mathcal{C}_{1}.

The points cjkc_{j}^{k}, 1k<n1\leq k<n, 1jlk1\leq j\leq l_{k}, satisfy the conditions of this lemma. Items (i) and (ii) are evident, and (iii) follows from the observation that if Wm(c)Wn(c)W_{m}(c^{\prime})\subseteq W_{n}(c) was not picked during the construction, then it had to intersect some set Wk(cjk)W_{k}(c_{j}^{k}) for an element cjkc_{j}^{k}, k>mk>m, picked earlier. By the condition 5(iii) this means Wm(c)Wk(cjk)W_{m}(c^{\prime})\subseteq W_{k}(c_{j}^{k}) as desired. ∎

Definition 7.

A family of regions Wi(j)(cj)W_{i(j)}(c_{j}) satisfying the conclusions of Lemma 6 is called a maximal family of subregions of Wn(c)W_{n}(c).

Remark 8.

It is easy to check that the maximal family of subregions of any Wn(c)W_{n}(c) is necessarily unique, but this will not play a role in our arguments.

Lemma 9.

Let 𝒞n\mathcal{C}_{n} and WnW_{n}, nn\in\mathbb{N}, be as in Theorem 5. For every m1m_{1}\in\mathbb{N} and c1𝒞m1c_{1}\in\mathcal{C}_{m_{1}} there exist a sequence of integers m1<m2<m3<m_{1}<m_{2}<m_{3}<\cdots and elements cj𝒞mjc_{j}\in\mathcal{C}_{m_{j}} such that the regions Wmj(cj)W_{m_{j}}(c_{j}) satisfy the inclusions Wmj(cj)Wmj+1(cj+1)W_{m_{j}}(c_{j})\subseteq W_{m_{j+1}}(c_{j+1}) for all 1j<1\leq j<\infty.

Proof.

The set W~m1(c1)\widetilde{W}_{m_{1}}(c_{1}) is a disk by 5(i), and in particular it is a compact region in d\mathbb{R}^{d}. We may therefore pick a compact KdK\subseteq\mathbb{R}^{d} such that the inclusion W~m1(c1)K\widetilde{W}_{m_{1}}(c_{1})\subset K is proper. By 5(iv) there exists some m2m_{2} and c2𝒞m2c_{2}\in\mathcal{C}_{m_{2}} such that Wm1(c1)c1+KWm2(c2)W_{m_{1}}(c_{1})\subset c_{1}+K\subseteq W_{m_{2}}(c_{2}). Items 5(iii) and 5(ii) guarantee that m2>m1m_{2}>m_{1}. The same choice can now be iterated to construct the desired sequence m1<m2<m3<m_{1}<m_{2}<m_{3}<\cdots and elements cj𝒞mjc_{j}\in\mathcal{C}_{m_{j}}. ∎

3. Equivalence to Special Flows

One of the simplest ways to construct an d\mathbb{R}^{d}-flow is to start with an \mathbb{R}-flow on some standard Borel space Ω1\Omega_{1} and define the action dΩ1×d1\mathbb{R}^{d}\curvearrowright\Omega_{1}\times\mathbb{R}^{d-1} by

Ω1×d1(y,q)+(r1,,rd)=(y+r1,q+(r2,,rd)).\Omega_{1}\times\mathbb{R}^{d-1}\ni(y,\vec{q}\,)+(r_{1},\ldots,r_{d})=(y+r_{1},\vec{q}+(r_{2},\ldots,r_{d})).

We say that a flow dΩ\mathbb{R}^{d}\curvearrowright\Omega is special if it is isomorphic to a flow of the form above. This is an ad hoc notion, which we use to reduce smooth equivalence of multidimensional flows to the one dimensional situation. Our goal in this section is to show that every free Borel d\mathbb{R}^{d}-flow is smoothly equivalent to a special one. The argument goes through a sequence of lemmas, and we begin by establishing some common notation.

Throughout the section we fix a free Borel d\mathbb{R}^{d}-flow 𝔉\mathfrak{F}, d2d\geq 2, let 𝒞n\mathcal{C}_{n} be the cross sections and Wn𝒞n×ΩW_{n}\subseteq\mathcal{C}_{n}\times\Omega be the corresponding regions produced by Theorem 5. Let VnΩV_{n}\subseteq\Omega denote the projection of WnW_{n} onto the second coordinate, and let πn:Vn𝒞n\pi_{n}:V_{n}\to\mathcal{C}_{n} be defined by the condition (πn(x),x)Wn(\pi_{n}(x),x)\in W_{n} for all xVnx\in V_{n}. Note that VnV_{n} is Borel as proj2:WnVn\mathrm{proj}_{2}:W_{n}\to V_{n} is injective by 5(ii) and πn\pi_{n} is Borel since its graph is the flip of WnW_{n}. Define for m<nm<n sets

Pm,n={(c,c)𝒞m×𝒞n:Wm(c)Wn(c)},P_{m,n}=\bigl{\{}(c^{\prime},c)\in\mathcal{C}_{m}\times\mathcal{C}_{n}:W_{m}(c^{\prime})\subseteq W_{n}(c)\bigr{\}},

which encode regions of the level mm inside a given region of the level nn. Sets W~n(c)\widetilde{W}_{n}(c) are smooth disks, and our first lemma shows that specific diffeomorphisms onto balls can be chosen to cohere across levels.

Lemma 10.

There exist radius maps tn:𝒞n>0t_{n}:\mathcal{C}_{n}\to\mathbb{R}^{>0}, “diffeomorphism” functions ϕn:Vnd\phi_{n}:V_{n}\to\mathbb{R}^{d}, and shift maps sm,n:Pm,n0s_{m,n}:P_{m,n}\to\mathbb{R}^{\geq 0} subject to the following conditions to be valid for all m<nm<n, and all (c,c)Pm,n(c^{\prime},c)\in P_{m,n} :

  1. (i)

    tm(c)mt_{m}(c^{\prime})\geq m;

  2. (ii)

    W~m(c)rϕm(c+r)B(tm(c))\widetilde{W}_{m}(c^{\prime})\ni\vec{r}\mapsto\phi_{m}(c^{\prime}+\vec{r})\in B(t_{m}(c^{\prime})) is a CC^{\infty} orientation preserving diffeomorphism onto the ball B(tm(c))B(t_{m}(c^{\prime}));

  3. (iii)

    ϕn(x)=ϕm(x)+sm,n(c,c)\phi_{n}(x)=\phi_{m}(x)+\vec{s}_{m,n}(c^{\prime},c) for all xWm(c)x\in W_{m}(c^{\prime}), where sm,n(c,c)=sm,n(c,c)×0d1\vec{s}_{m,n}(c^{\prime},c)=s_{m,n}(c^{\prime},c)\times\vec{0}^{d-1};

  4. (iv)

    tm(c)+sm,n(c,c)tn(c)1t_{m}(c^{\prime})+s_{m,n}(c^{\prime},c)\leq t_{n}(c)-1;

  5. (v)

    there is a Borel partition 𝒞m=k𝒞m,k\mathcal{C}_{m}=\bigsqcup_{k}\mathcal{C}_{m,k} such that W~m(c1)=W~m(c2)\widetilde{W}_{m}(c_{1})=\widetilde{W}_{m}(c_{2}), tm(c1)=tm(c2)t_{m}(c_{1})=t_{m}(c_{2}), and ϕm(c1+r)=ϕm(c2+r)\phi_{m}(c_{1}+\vec{r})=\phi_{m}(c_{2}+\vec{r}) for all c1,c2𝒞m,kc_{1},c_{2}\in\mathcal{C}_{m,k} and all rW~m(c1)\vec{r}\in\widetilde{W}_{m}(c_{1}).

The meaning of these conditions is as follows. Item (i) ensures that radii go to infinity as mm\to\infty. According to item (ii), each map ϕm\phi_{m} encodes a family of diffeomorphisms, one for each c𝒞mc^{\prime}\in\mathcal{C}_{m}. Formally speaking, these diffeomorphisms are maps from W~m(c)\widetilde{W}_{m}(c^{\prime}) onto B(tm(c))B(t_{m}(c^{\prime})). However, we will occasionally abuse the language by calling the map ϕm|Wm(c):Wm(c)B(tm(c))\phi_{m}|_{W_{m}(c^{\prime})}:W_{m}(c^{\prime})\to B(t_{m}(c^{\prime})) a diffeomorphism.

Item (iii) postulates that ϕn\phi_{n} extends ϕm\phi_{m} up to a translation of the range along the xx-axis, where the translation value is constant over each Wm(c)W_{m}(c^{\prime}) region and is equal to sm,n(c,c)s_{m,n}(c^{\prime},c). Condition (iv) is a reformulation of the inequality

dist(ϕm(Wm(c))+sm,n(c,c),B(tn(c)))1.\mathrm{dist}\bigl{(}\phi_{m}(W_{m}(c^{\prime}))+\vec{s}_{m,n}(c^{\prime},c),\ \partial{B}(t_{n}(c))\bigr{)}\geq 1.

It means that disks ϕm(Wm(c))+sm,n(c,c)\phi_{m}(W_{m}(c^{\prime}))+\vec{s}_{m,n}(c^{\prime},c) are at least one unit of distance away from the boundary of B(tn(c))B(t_{n}(c)) (see Figure 3). Similarly to Theorem 5(v), item (v) says that we need to consider only countably many different diffeomorphisms ϕm|Wm(c)\phi_{m}|_{W_{m}(c^{\prime})}. The only purpose of this property is to make it easy for us to argue that the flow 𝔉\mathfrak{F}^{\prime}, which will be constructed later in this section, is Borel.

Proof of Lemma 10.

The construction goes by induction on nn, and we begin with its base. Set t1(c)=1t_{1}(c)=1 for all c𝒞1c\in\mathcal{C}_{1}. By item (i) of Theorem 5, each region W~1(c)\widetilde{W}_{1}(c), c𝒞1c\in\mathcal{C}_{1}, is a smooth disk. So for ϕ1:V1B(1)\phi_{1}:V_{1}\to B(1) we pick any map satisfying (ii) and (v), which can be done since there are only countably many shapes W~1(c)\widetilde{W}_{1}(c) by 5(v).

For the inductive step consider a region Wn(c)W_{n}(c). We pick points c1,,cli<n𝒞ic_{1},\ldots,c_{l}\in\bigcup_{i<n}\mathcal{C}_{i}, and integers i(j)i(j) defined by cj𝒞i(j)c_{j}\in\mathcal{C}_{i(j)}, that correspond to a maximal family of subregions of Wn(c)W_{n}(c) as per Lemma 6. For such points cjc_{j} we have ρ(c,cj)+W~i(j)(cj)intW~n(c)\rho(c,c_{j})+\widetilde{W}_{i(j)}(c_{j})\subset\operatorname{\mathrm{int}}\widetilde{W}_{n}(c), as guaranteed by 5(iii). An example of such a region Wn(c)W_{n}(c) is shown in Figure 4 on page 4. By inductive assumption regions W~i(j)(cj)\widetilde{W}_{i(j)}(c_{j}) are diffeomorphic to balls B(ti(j)(cj))B(t_{i(j)}(c_{j})) via the diffeomorphisms rϕi(j)(cj+r)\vec{r}\mapsto\phi_{i(j)}(c_{j}+\vec{r}). We shift these balls along the xx-axis to make them disjoint, and view them inside a sufficiently large ball in d\mathbb{R}^{d} (see Figure 3). More specifically, let

Wi(j),n={(c,x)Wi(j):cproj1(Pi(j),n) and cproj1(Pi(j),k) for all i(j)<k<n},W_{i(j),n}^{\prime}=\{(c^{\prime},x)\in W_{i(j)}:c^{\prime}\in\mathrm{proj}_{1}(P_{i(j),n})\textrm{ and }c^{\prime}\not\in\mathrm{proj}_{1}(P_{i(j),k})\textrm{ for all }i(j)<k<n\},

and put Vi(j),n=proj2(Wi(j),n)V_{i(j),n}^{\prime}=\mathrm{proj}_{2}(W^{\prime}_{i(j),n}); in other words, Wi(j),nW_{i(j),n}^{\prime} consists of those regions Wi(j)(c)W_{i(j)}(c^{\prime}) for which nn is the smallest index to satisfy Wi(j)(c)Wn(c)W_{i(j)}(c^{\prime})\subset W_{n}(c) for some c𝒞nc\in\mathcal{C}_{n}. Set for 1jl1\leq j\leq l

si(j),n(cj,c)=(j1)+ti(j)(cj)+21k<jti(k)(ck),s_{i(j),n}(c_{j},c)=(j-1)+t_{i(j)}(c_{j})+2\sum_{1\leq k<j}t_{i(k)}(c_{k}),

and consider the map ϕi(j):Vi(j),nd\phi^{\prime}_{i(j)}:V_{i(j),n}^{\prime}\to\mathbb{R}^{d} to be given for xWi(j)(cj)x\in W_{i(j)}(c_{j}) by ϕi(j)(x)=ϕi(j)(x)+si(j),n(cj,c)\phi^{\prime}_{i(j)}(x)=\phi_{i(j)}(x)+\vec{s}_{i(j),n}(c_{j},c). Note that restrictions ϕi(j)|Wi(j)(cj)\phi^{\prime}_{i(j)}|_{W_{i(j)}(c_{j})} are diffeomorphisms onto disks B(ti(j)(cj))+si(j),n(cj,c)B(t_{i(j)}(c_{j}))+\vec{s}_{i(j),n}(c_{j},c). The radius tn(c)t_{n}(c) is taken to be sufficiently large to contain these disks: tn(c)=max{n,si(l),n(cl,c)+ti(l)(cl)+1}t_{n}(c)=\max\bigl{\{}n,\ s_{i(l),n}(c_{l},c)+t_{i(l)}(c_{l})+1\bigr{\}}. This ensures that images ϕi(j)(Wi(j)(cj))\phi^{\prime}_{i(j)}(W_{i(j)}(c_{j})) are inside B(tn(c))B(t_{n}(c)), and are furthermore at least 11 unit of distance away from its boundary, which yields item (iv). Item (i) also continues to be satisfied by this choice of tn(c)t_{n}(c).

0\vec{0}\cdotsradius is tn(c)t_{n}(c) 11si(2),n(c2,c)s_{i(2),n}(c_{2},c)si(l),n(cl,c)s_{i(l),n}(c_{l},c)ti(3)(c3)t_{i(3)}(c_{3})1\geq 1
Figure 3. Alignment of disks ϕi(j)(Wi(j)(cj))+si(j),n(cj,c)\phi_{i(j)}(W_{i(j)}(c_{j}))+\vec{s}_{i(j),n}(c_{j},c).

Extension Lemma 2 can now be applied to diffeomorphisms ρ(c,cj)+W~i(j)(cj)rϕi(j)(c+r)\rho(c,c_{j})+\widetilde{W}_{i(j)}(c_{j})\ni\vec{r}\mapsto\phi^{\prime}_{i(j)}(c+\vec{r}), since they are defined on disjoint disks ρ(c,cj)+W~i(j)(cj)\rho(c,c_{j})+\widetilde{W}_{i(j)}(c_{j}) and have disjoint images B(ti(j)(cj))+si(j),n(cj,c)B(t_{i(j)}(c_{j}))+\vec{s}_{i(j),n}(c_{j},c), 1jl1\leq j\leq l. All the domains of these maps lie in the interior of the disk W~n(c)\widetilde{W}_{n}(c), while the images are subsets of intB(tn(c))\operatorname{\mathrm{int}}B(t_{n}(c)). We therefore can find a common extension to a diffeomorphism ϕn|Wn(c):Wn(c)B(tn(c))\phi_{n}|_{W_{n}(c)}:W_{n}(c)\to B(t_{n}(c)) that satisfies

ϕn(x)=ϕi(j)(x)=ϕi(j)(x)+si(j),n(cj,c)\phi_{n}(x)=\phi_{i(j)}^{\prime}(x)=\phi_{i(j)}(x)+\vec{s}_{i(j),n}(c_{j},c)

for all xWi(j)(cj) and all 1jlx\in W_{i(j)}(c_{j})\textrm{ and all }1\leq j\leq l, thus implying (iii).

We are not quite done yet though. The construction above defined diffeomorphisms ϕn\phi_{n} and radii tn(c)t_{n}(c) for all c𝒞nc\in\mathcal{C}_{n}, but shifts sm,n(c,c)s_{m,n}(c^{\prime},c) are currently defined only for those c𝒞mc^{\prime}\in\mathcal{C}_{m} that belong to the maximal family of subregions of Wn(c)W_{n}(c). Nonetheless, values sm,n(c,c)s_{m,n}(c^{\prime},c) satisfying item (iii), are uniquely specified for all (c,c)Pm,n(c^{\prime},c)\in P_{m,n} based on the following observation. Pick any c𝒞mc^{\prime}\in\mathcal{C}_{m}, m<nm<n, such that Wm(c)Wn(c)W_{m}(c^{\prime})\subseteq W_{n}(c), let jj be the unique index 1jl1\leq j\leq l such that Wm(c)Wi(j)(cj)W_{m}(c^{\prime})\subseteq W_{i(j)}(c_{j}). Suppose cc^{\prime} does not belong to the maximal family of subregions of Wn(c)W_{n}(c), hence m<i(j)m<i(j). Using the inductive assumption (iii), we find that for any xWm(c)x\in W_{m}(c^{\prime})

ϕn(x)\displaystyle\phi_{n}(x) =ϕi(j)(x)+si(j),n(cj,c)\displaystyle=\phi_{i(j)}(x)+\vec{s}_{i(j),n}(c_{j},c)
=ϕm(x)+sm,i(j)(c,cj)+si(j),n(cj,c).\displaystyle=\phi_{m}(x)+\vec{s}_{m,i(j)}(c^{\prime},c_{j})+\vec{s}_{i(j),n}(c_{j},c).

Thus, for sm,n(c,c)=sm,i(j)(c,cj)+si(j),n(cj,c)s_{m,n}(c^{\prime},c)=s_{m,i(j)}(c^{\prime},c_{j})+s_{i(j),n}(c_{j},c) item (iii) holds for all m<nm<n and all c𝒞mc^{\prime}\in\mathcal{C}_{m} such that Wm(c)Wn(c)W_{m}(c^{\prime})\subseteq W_{n}(c).

We check that (iv) continues to hold. Let us assume that this property has been verified for regions at levels below nn, and by construction we have also established

(2) tn(c)si(j),n(cj,c)ti(j)(cj)+1.t_{n}(c)-s_{i(j),n}(c_{j},c)\geq t_{i(j)}(c_{j})+1.

Using the additivity of values sm,n(c,c)s_{m,n}(c^{\prime},c) shown above we get

tn(c)tm(c)sm,n(c,c)\displaystyle t_{n}(c)-t_{m}(c^{\prime})-s_{m,n}(c^{\prime},c) =tn(c)si(j),n(cj,c)sm,i(j)(c,cj)tm(c)\displaystyle=t_{n}(c)-s_{i(j),n}(c_{j},c)-s_{m,i(j)}(c^{\prime},c_{j})-t_{m}(c^{\prime})
Eq. (2) ti(j)(cj)+1sm,i(j)(c,cj)tm(c)\displaystyle\geq t_{i(j)}(c_{j})+1-s_{m,i(j)}(c^{\prime},c_{j})-t_{m}(c^{\prime})
inductive assumption tm(c)+2tm(c)=2.\displaystyle\geq t_{m}(c^{\prime})+2-t_{m}(c^{\prime})=2.

Therefore (iv) holds for all m<nm<n and all (c,c)Pm,n(c^{\prime},c)\in P_{m,n}.

Finally, to guarantee item (v) note that diffeomorphisms ϕn\phi_{n} have been chosen using Lemma 2 based on the shapes of regions W~n(c)\widetilde{W}_{n}(c), as well as shapes of subregions W~i(j)(cj)\widetilde{W}_{i(j)}(c_{j}), and their locations inside W~n(c)\widetilde{W}_{n}(c) specified by values ρ(c,cj)\rho(c,c_{j}). By Theorem 5, the union k𝒞k\bigcup_{k}\mathcal{C}_{k} is on a rational grid, so all the values ρ(c,cj)\rho(c,c_{j}) are rational. Also, by 5(v), there are only countably many possible shapes for regions Wi(j)(cj)W_{i(j)}(c_{j}). We may therefore choose the same diffeomorphism ϕn|Wn(c)\phi_{n}|_{W_{n}(c)} whenever the inputs to Lemma 2 are the same, which guarantees fulfillment of item (v). ∎

The item 10(iv) above guarantees that the image of Wm(c)W_{m}(c^{\prime}) under ϕn\phi_{n} is at least 11 unit of distance away from the boundary of B(tn(c))B(t_{n}(c)) whenever Wm(c)Wn(c)W_{m}(c^{\prime})\subseteq W_{n}(c). The following two lemmas show that we can find such nn and c𝒞nc\in\mathcal{C}_{n} for which the set ϕn(Wm(c))\phi_{n}(W_{m}(c^{\prime})) is as far from the boundary of B(tn(c))B(t_{n}(c)) as we desire.

Lemma 11.

Let m1<m2<m3<m_{1}<m_{2}<m_{3}<\cdots be an increasing sequence of integers and cj𝒞mjc_{j}\in\mathcal{C}_{m_{j}} be elements such that Wmj(cj)Wmj+1(cj+1)W_{m_{j}}(c_{j})\subseteq W_{m_{j+1}}(c_{j+1}) (such a sequence is produced by Lemma 9). For all j2j\geq 2 one has

tmj(cj)1k<jsmk,mk+1(ck,ck+1)tm1(c1)j1.t_{m_{j}}(c_{j})-\sum_{1\leq k<j}s_{m_{k},m_{k+1}}(c_{k},c_{k+1})-t_{m_{1}}(c_{1})\geq j-1.
Proof.

The argument is a simple induction coupled with item 10(iv), which, in particular, gives the base

tm2(c2)sm1,m2(c1,c2)tm1(c1)1.t_{m_{2}}(c_{2})-s_{m_{1},m_{2}}(c_{1},c_{2})-t_{m_{1}}(c_{1})\geq 1.

Suppose the statement has been established for j1j-1:

(3) tmj1(cj1)1k<j1smk,mk+1(ck,ck+1)tm1(c1)j2.t_{m_{j-1}}(c_{j-1})-\mkern-16.0mu\sum_{1\leq k<j-1}\mkern-16.0mus_{m_{k},m_{k+1}}(c_{k},c_{k+1})-t_{m_{1}}(c_{1})\geq j-2.

By item 10(iv) we have tmj(cj)smj1,mj(cj1,cj)1+tmj1(cj1)t_{m_{j}}(c_{j})-s_{m_{j-1},m_{j}}(c_{j-1},c_{j})\geq 1+t_{m_{j-1}}(c_{j-1}), and therefore

tmj(cj)\displaystyle t_{m_{j}}(c_{j}) 1k<jsmk,mk+1(ck,ck+1)tm1(c1)\displaystyle-\mkern-10.0mu\sum_{1\leq k<j}\mkern-10.0mus_{m_{k},m_{k+1}}(c_{k},c_{k+1})-t_{m_{1}}(c_{1})
=tmj(cj)smj1,mj(cj1,cj)1k<j1smk,mk+1(ck,ck+1)tm1(c1)\displaystyle=t_{m_{j}}(c_{j})-s_{m_{j-1},m_{j}}(c_{j-1},c_{j})-\mkern-16.0mu\sum_{1\leq k<j-1}\mkern-16.0mus_{m_{k},m_{k+1}}(c_{k},c_{k+1})-t_{m_{1}}(c_{1})
1+tmj1(cj1)1k<j1smk,mk+1(ck,ck+1)tm1(c1)\displaystyle\geq 1+t_{m_{j-1}}(c_{j-1})-\mkern-16.0mu\sum_{1\leq k<j-1}\mkern-16.0mus_{m_{k},m_{k+1}}(c_{k},c_{k+1})-t_{m_{1}}(c_{1})
Eq. (3) 1+j2=j1,\displaystyle\geq 1+j-2=j-1,

which yields the step of induction. ∎

Lemma 12.

For any xΩx\in\Omega and any R0R\in\mathbb{R}^{\geq 0} there exist nn and c𝒞nc\in\mathcal{C}_{n} such that xWn(c)x\in W_{n}(c) and

ϕn(x)+Rtn(c).||\phi_{n}(x)||+R\leq t_{n}(c).
Proof.

In view of 5(iv) there is some m1m_{1} and c1𝒞m1c_{1}\in\mathcal{C}_{m_{1}} such that xWm1(c1)x\in W_{m_{1}}(c_{1}). By Lemma 9 there exist levels m1<m2<m_{1}<m_{2}<\cdots and points cj𝒞mjc_{j}\in\mathcal{C}_{m_{j}} such that Wmj(cj)Wmj+1(cj+1)W_{m_{j}}(c_{j})\subseteq W_{m_{j+1}}(c_{j+1}). For each jj we have in view of 10(iii)

tmj(cj)ϕmj(x)\displaystyle t_{m_{j}}(c_{j})-||\phi_{m_{j}}(x)|| =tmj(cj)ϕmj1(x)+smj1,mj(cj1,cj)\displaystyle=t_{m_{j}}(c_{j})-||\phi_{m_{j-1}}(x)+\vec{s}_{m_{j-1},m_{j}}(c_{j-1},c_{j})||
=j2 further applications of 10(iii)\displaystyle=\cdots\quad\textrm{$j-2$ further applications of~{}\ref{lem:coherent-diffs}\eqref{item:coherence-of-steps}}
=tmj(cj)||ϕm1(x)+1k<jsmk,mk+1(ck,ck+1)||\displaystyle=t_{m_{j}}(c_{j})-\Bigl{|}\Bigl{|}\phi_{m_{1}}(x)+\mkern-10.0mu\sum_{1\leq k<j}\mkern-7.0mu\vec{s}_{m_{k},m_{k+1}}(c_{k},c_{k+1})\Bigr{|}\Bigr{|}
tmj(cj)tm1(c1)1k<jsmk,mk+1(ck,ck+1)\displaystyle\geq t_{m_{j}}(c_{j})-t_{m_{1}}(c_{1})-\mkern-10.0mu\sum_{1\leq k<j}\mkern-7.0mus_{m_{k},m_{k+1}}(c_{k},c_{k+1})
Lemma 11 j1.\displaystyle\geq j-1.

Therefore n=mjn=m_{j} and c=cjc=c_{j} satisfy the conclusions of the lemma as long as j1Rj-1\geq R. ∎

We now define a new flow 𝔉\mathfrak{F}^{\prime} on the same phase space Ω\Omega. Notation xrx\oplus\vec{r} will be used to distinguish the action of 𝔉\mathfrak{F}^{\prime} from the action of the original flow 𝔉\mathfrak{F}. For xΩx\in\Omega and rd\vec{r}\in\mathbb{R}^{d} let c𝒞nc\in\mathcal{C}_{n} be such that xWn(c)x\in W_{n}(c) and ϕn(x)+rB(tn(c))\phi_{n}(x)+\vec{r}\in B(t_{n}(c)) (such nn and cc exist by Lemma 12). The 𝔉\mathfrak{F}^{\prime} action of r\vec{r} upon xx is defined by

xr=(ϕn|Wn(c))1(ϕn(x)+r),x\oplus\vec{r}=(\phi_{n}|_{W_{n}(c)})^{-1}(\phi_{n}(x)+\vec{r}),

or, equivalently, xrx\oplus\vec{r} is the element of Wn(c)W_{n}(c) for which ϕn(xr)=ϕn(x)+r\phi_{n}(x\oplus\vec{r})=\phi_{n}(x)+\vec{r}. The geometric interpretation of the action is as follows. We use the diffeomorphism ϕn\phi_{n} to identify Wn(c)W_{n}(c) with B(tn(c))B(t_{n}(c)). One acts upon ϕn(x)B(tn(c))d\phi_{n}(x)\in B(t_{n}(c))\subseteq\mathbb{R}^{d} by translation. Assuming the image lies within the same ball B(tn(c))B(t_{n}(c)), we can pull it back to an element of Wn(c)W_{n}(c), which is what xrx\oplus\vec{r} is defined to be. As we argue below, this definition does not depend on the choice of nn and cc due to the coherence of diffeomorphisms ϕn\phi_{n} provided by item 10(iii). Having this simple picture of the action in their mind will make it easy for the reader to follow the somewhat tedious but elementary computations that constitute a large portion of the remainder of this section.

Lemma 13.

The definition of xrx\oplus\vec{r} does not depend on the choice of nn and cc.

Proof.

Let c𝒞mc^{\prime}\in\mathcal{C}_{m} be another element that can be used in the definition of xrx\oplus\vec{r}, i.e., xWm(c)x\in W_{m}(c^{\prime}) and ϕm(x)+rB(tm(c))\phi_{m}(x)+\vec{r}\in B(t_{m}(c^{\prime})). Item 5(ii) implies mnm\neq n, and 5(iii) guarantees that either Wm(c)Wn(c)W_{m}(c^{\prime})\subseteq W_{n}(c) or Wn(c)Wm(c)W_{n}(c)\subseteq W_{m}(c^{\prime}). Since roles of mm and nn are symmetric, we may assume without loss of generality that the former is the case. Consider the chain of equalities

ϕn(x)+r\displaystyle\phi_{n}(x)+\vec{r} =ϕm(x)+sm,n(c,c)+r\displaystyle=\phi_{m}(x)+\vec{s}_{m,n}(c^{\prime},c)+\vec{r}
=ϕm((ϕm|Wm(c))1(ϕm(x)+r))+sm,n(c,c)\displaystyle=\phi_{m}\Bigl{(}(\phi_{m}|_{W_{m}(c^{\prime})})^{-1}\bigl{(}\phi_{m}(x)+\vec{r}\bigr{)}\Bigr{)}+\vec{s}_{m,n}(c^{\prime},c)
item 10(iii) =ϕn((ϕm|Wm(c))1(ϕm(x)+r)).\displaystyle=\phi_{n}\Bigl{(}(\phi_{m}|_{W_{m}(c^{\prime})})^{-1}\bigl{(}\phi_{m}(x)+\vec{r}\bigr{)}\Bigr{)}.

Applying (ϕn|Wn(c))1(\phi_{n}|_{W_{n}(c)})^{-1} to the first and the last expressions above yields

(ϕn|Wn(c))1(ϕn(x)+r)=(ϕm|Wm(c))1(ϕm(x)+r),(\phi_{n}|_{W_{n}(c)})^{-1}(\phi_{n}(x)+\vec{r})=(\phi_{m}|_{W_{m}(c^{\prime})})^{-1}(\phi_{m}(x)+\vec{r}),

which finishes the proof of the lemma. ∎

Having established that xrx\oplus\vec{r} is well-defined, we can now verify it to be a free Borel flow.

Lemma 14.

The map Ω×d(x,r)xrΩ\Omega\times\mathbb{R}^{d}\ni(x,\vec{r})\mapsto x\oplus\vec{r}\in\Omega defines a free Borel action of d\mathbb{R}^{d} on Ω\Omega. This flow is smoothly equivalent to 𝔉\mathfrak{F}. Moreover, the identity map id:ΩΩ\mathrm{id}:\Omega\to\Omega is a smooth equivalence between 𝔉\mathfrak{F} and 𝔉\mathfrak{F}^{\prime}.

Proof.

Pick r1,r2d\vec{r}_{1},\vec{r}_{2}\in\mathbb{R}^{d}, and use Lemma 12 to choose nn, c𝒞nc\in\mathcal{C}_{n}, such that

ϕn(x)+r1+r2tn(c).||\phi_{n}(x)||+||\vec{r}_{1}||+||\vec{r}_{2}||\leq t_{n}(c).

This inequality guarantees that (ϕn|Wn(c))1(\phi_{n}|_{W_{n}(c)})^{-1} is defined in the following terms:

(xr1)r2\displaystyle(x\oplus\vec{r}_{1})\oplus\vec{r}_{2} =(ϕn|Wn(c))1(ϕn(xr1)+r2)\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\phi_{n}(x\oplus\vec{r}_{1})+\vec{r}_{2}\bigr{)}
=(ϕn|Wn(c))1(ϕn(x)+r1+r2)=x(r1+r2).\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\phi_{n}(x)+\vec{r}_{1}+\vec{r}_{2}\bigr{)}=x\oplus(\vec{r}_{1}+\vec{r}_{2}).

Coupled with the straightforward x0=xx\oplus\vec{0}=x, these computations show that \oplus defines a flow on Ω\Omega. This flow is free, because the maps ϕn|Wn(c)\phi_{n}|_{W_{n}(c)} are injective.

It is easy to see that orbits of 𝔉\mathfrak{F}^{\prime} coincide with those of 𝔉\mathfrak{F}. The inclusion E𝔉E𝔉E_{\mathfrak{F}^{\prime}}\subseteq E_{\mathfrak{F}} is guaranteed by the condition WnE𝔉W_{n}\subseteq E_{\mathfrak{F}}. For the inverse direction, let x,yΩx,y\in\Omega be such that xE𝔉yxE_{\mathfrak{F}}y. By 5(iv), there are nn and c𝒞nc\in\mathcal{C}_{n} such that x,yWn(c)x,y\in W_{n}(c). One has

x(ϕn(y)ϕn(x))\displaystyle x\oplus\bigl{(}\phi_{n}(y)-\phi_{n}(x)\bigr{)} =(ϕn|Wn(c))1(ϕn(x)+ϕn(y)ϕn(x))\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\phi_{n}(x)+\phi_{n}(y)-\phi_{n}(x)\bigr{)}
=(ϕn|Wn(c))1(ϕn(y))=y,\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}(\phi_{n}(y))=y,

and thus E𝔉=E𝔉E_{\mathfrak{F}}=E_{\mathfrak{F}^{\prime}}.

The flow 𝔉\mathfrak{F}^{\prime} is Borel. To justify this set α:Ω×dd\alpha:\Omega\times\mathbb{R}^{d}\to\mathbb{R}^{d} to be defined by xr=x+α(x,r)x\oplus\vec{r}=x+\alpha(x,\vec{r}); it suffices to show that α\alpha is Borel. In general, we would have to verify that the value (ϕn|Wn(c))1(ϕn(x)+r)(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\phi_{n}(x)+\vec{r}\,\bigr{)} depends in a Borel way on xΩx\in\Omega and rd\vec{r}\in\mathbb{R}^{d}, which requires going into the details of the way maps ϕn|Wn(c)\phi_{n}|_{W_{n}(c)} are constructed based on nn and c𝒞nc\in\mathcal{C}_{n}. However, we ensured in item 10(v) that there is a countable Borel partition of each cross section 𝒞n=k𝒞n,k\mathcal{C}_{n}=\bigsqcup_{k}\mathcal{C}_{n,k} such that W~n(c1)=W~n(c2)\widetilde{W}_{n}(c_{1})=\widetilde{W}_{n}(c_{2}), tn(c1)=tn(c2)t_{n}(c_{1})=t_{n}(c_{2}), and ϕn|Wn(c1)(c1+r)=ϕn|Wn(c2)(c2+r)\phi_{n}|_{W_{n}(c_{1})}(c_{1}+\vec{r})=\phi_{n}|_{W_{n}(c_{2})}(c_{2}+\vec{r}) for all c1,c2𝒞n,kc_{1},c_{2}\in\mathcal{C}_{n,k} and all rW~(c1)\vec{r}\in\widetilde{W}(c_{1}). Let W~n,k\widetilde{W}_{n,k} denote the common shape of W~n(c)\widetilde{W}_{n}(c), and similarly let tn,kt_{n,k} be the radius tn(c)t_{n}(c) common for all c𝒞n,kc\in\mathcal{C}_{n,k}. Likewise, maps ϕn|Wn(c)(c+)\phi_{n}|_{W_{n}(c)}(c+\cdot) produce the same diffeomorphism φn,k:W~n,kB(tn,k)\varphi_{n,k}:\widetilde{W}_{n,k}\to B(t_{n,k}) regardless of the choice of c𝒞n,kc\in\mathcal{C}_{n,k}.

Recall that πn:Vn𝒞n\pi_{n}:V_{n}\to\mathcal{C}_{n} is the map that associates the distinguished point cc to every xx in Wn(c)W_{n}(c). Let

Xn,k={(x,r)Ω×d:πn(x)𝒞n,k,xVn, and φn,k(ρ(πn(x),x))+rtn,k}X_{n,k}=\{(x,\vec{r})\in\Omega\times\mathbb{R}^{d}:\pi_{n}(x)\in\mathcal{C}_{n,k},\ x\in V_{n},\textrm{ and }\varphi_{n,k}(\rho(\pi_{n}(x),x))+\vec{r}\leq t_{n,k}\}

denote the set of those pairs (x,r)(x,\vec{r}) for which xrx\oplus\vec{r} can be defined using Wn(c)W_{n}(c) for some c𝒞n,kc\in\mathcal{C}_{n,k}. To conclude that α\alpha is Borel, we observe that for (x,r)Xn,k(x,\vec{r})\in X_{n,k} its value equals

(4) α(x,r)=ρ(x,πn(x))+φn,k1(φn,k(ρ(πn(x),x))+r),\alpha(x,\vec{r})=\rho(x,\pi_{n}(x))+\varphi_{n,k}^{-1}\bigl{(}\varphi_{n,k}(\rho(\pi_{n}(x),x))+\vec{r}\bigr{)},

which is a composition of Borel functions. It is at this point that our efforts in ensuring stability of the construction in 5(v) and 10(v) yield their fruits. We have only one diffeomorphism φn,k\varphi_{n,k} in the definition of α\alpha which applies to all arguments (x,r)Xn,k(x,\vec{r})\in X_{n,k}. Thus α|Xn,k\alpha|_{X_{n,k}} is Borel regardless of how this diffeomorphism was chosen. Since n,kXn,k=Ω×d\bigcup_{n,k}X_{n,k}=\Omega\times\mathbb{R}^{d}, we may conclude that α\alpha is Borel on all of Ω×d\Omega\times\mathbb{R}^{d}.

We have already established that id:ΩΩ\mathrm{id}:\Omega\to\Omega is an orbit equivalence, and it remains to verify that it is smooth, which amounts to showing that α(x,):dd\alpha(x,\cdot):\mathbb{R}^{d}\to\mathbb{R}^{d} is a diffeomorphism for each xΩx\in\Omega. It is a bijection, since 𝔉,𝔉\mathfrak{F},\ \mathfrak{F}^{\prime} are free and E𝔉=E𝔉E_{\mathfrak{F}}=E_{\mathfrak{F}^{\prime}}, and it is smooth and orientation preserving, since according to Eq. (4) for any fixed xΩx\in\Omega the function α(x,)\alpha(x,\cdot) is a composition of φn,k1\varphi_{n,k}^{-1} with translation maps. ∎

The goal of this section is to show that every free flow is smoothly equivalent to a special one. So far starting with a free flow 𝔉\mathfrak{F} we have constructed a smoothly equivalent flow 𝔉\mathfrak{F}^{\prime}, and it remains to verify that the latter is special. We need a subset Ω1Ω\Omega_{1}\subseteq\Omega invariant under the shifts Ω1s×0d1=Ω1\Omega_{1}\oplus s\times\vec{0}^{d-1}=\Omega_{1} for all ss\in\mathbb{R}. Set LnWnL_{n}\subseteq W_{n} to correspond to the preimages of the xx-axis inside B(tn(c))B(t_{n}(c)) under the diffeomorphisms ϕn\phi_{n}:

Ln={(c,x)Wn:ϕn(x)[tn(c),tn(c)]×0d1}.L_{n}=\bigl{\{}(c,x)\in W_{n}:\phi_{n}(x)\in\bigl{[}-t_{n}(c),t_{n}(c)\bigr{]}\times\vec{0}^{d-1}\bigr{\}}.

Sets Ln(c)L_{n}(c) represent line segments inside regions Wn(c)W_{n}(c) (see Figure 4). Set L=nLnL=\bigcup_{n}L_{n}, and let Ω1\Omega_{1} be the projection of LL onto the second coordinate:

Ω1={x:(c,x)L for some cn𝒞n}.\Omega_{1}=\Bigl{\{}x:(c,x)\in L\textrm{ for some }c\in\bigcup_{n}\mathcal{C}_{n}\Bigr{\}}.
Wi(j)(cj)W_{i(j)}(c_{j})Li(j)(cj)L_{i(j)}(c_{j})Ln(c)L_{n}(c)Wn(c)W_{n}(c)
Figure 4. Region Wn(c)W_{n}(c), containing four subregions Wi(j)(cj)W_{i(j)}(c_{j}) marked in gray. Each of the subregions has a line segment Li(j)(cj)L_{i(j)}(c_{j}), which are all contained inside Ln(c)L_{n}(c).
Lemma 15.

The set Ω1Ω\Omega_{1}\subseteq\Omega is Borel and Ω1s×0d1=Ω1\Omega_{1}\oplus s\times\vec{0}^{d-1}=\Omega_{1} for all ss\in\mathbb{R}.

Proof.

The set Ω1=nproj2(Ln)\Omega_{1}=\bigcup_{n}\mathrm{proj}_{2}(L_{n}) is Borel, since projections proj2:LnΩ1\mathrm{proj}_{2}:L_{n}\to\Omega_{1} are injective, hence have Borel images (see [Kec95, Corollary 15.2]). To check shift invariance, pick xΩ1x\in\Omega_{1} and ss\in\mathbb{R}; let s\vec{s} denote the vector s×0d1s\times\vec{0}^{d-1}. There has to exist some mm and c𝒞mc^{\prime}\in\mathcal{C}_{m} such that (c,x)Lm(c^{\prime},x)\in L_{m}. Pick n>mn>m and c𝒞nc\in\mathcal{C}_{n} to satisfy x,xsWn(c)x,x\oplus\vec{s}\in W_{n}(c), which exist by 5(iv). Note that (c,x)Ln(c,x)\in L_{n} as according to 10(iii) ϕn(x)=ϕm(x)+sm,n(c,c),\phi_{n}(x)=\phi_{m}(x)+\vec{s}_{m,n}(c^{\prime},c), and therefore ϕm(x)[tm(c),tm(c)]×0d1\phi_{m}(x)\in[-t_{m}(c^{\prime}),t_{m}(c^{\prime})]\times\vec{0}^{d-1} implies

ϕn(x)\displaystyle\phi_{n}(x) [tm(c)sm,n(c,c),tm(c)+sm,n(c,c)]×0d1\displaystyle\in\bigl{[}-t_{m}(c^{\prime})-s_{m,n}(c^{\prime},c),\ t_{m}(c^{\prime})+s_{m,n}(c^{\prime},c)\bigr{]}\times\vec{0}^{d-1}
item 10(iv) [tn(c),tn(c)]×0d1.\displaystyle\subseteq[-t_{n}(c),t_{n}(c)]\times\vec{0}^{d-1}.

Since xsWn(c)x\oplus\vec{s}\in W_{n}(c), we have ϕn(x)+s[tn(c),tn(c)]×0d1\phi_{n}(x)+\vec{s}\in[-t_{n}(c),t_{n}(c)]\times\vec{0}^{d-1}, thus xsΩ1x\oplus\vec{s}\in\Omega_{1} as claimed. ∎

The following lemma will be helpful in establishing that 𝔉\mathfrak{F}^{\prime} is special.

Lemma 16.

For any xΩx\in\Omega there exists qd1\vec{q}\in\mathbb{R}^{d-1} such that x0×qΩ1x\oplus 0\times\vec{q}\in\Omega_{1}.

Proof.

Pick some xΩx\in\Omega and, as usual, let nn, c𝒞nc\in\mathcal{C}_{n} be chosen to satisfy xWn(c)x\in W_{n}(c). Set qd1\vec{q}\in\mathbb{R}^{d-1} to be the negative of the projection of ϕn(x)\phi_{n}(x) onto the last (d1)(d-1)-many coordinates: q=proj[2,d](ϕn(x))\vec{q}=-\mathrm{proj}_{[2,d]}(\phi_{n}(x)). By the definition of the action,

x0×q\displaystyle x\oplus 0\times\vec{q} =(ϕn|Wn(c))1(ϕn(x)+0×q)\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}(\phi_{n}(x)+0\times\vec{q}\,)
=(ϕn|Wn(c))1(ϕn(x)0×proj[2,d](ϕn(x)))\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\phi_{n}(x)-0\times\mathrm{proj}_{[2,d]}(\phi_{n}(x))\bigr{)}
=(ϕn|Wn(c))1(proj1(ϕn(x))×0d1)Ln(c)Ω1,\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\mathrm{proj}_{1}(\phi_{n}(x))\times\vec{0}^{d-1}\bigr{)}\in L_{n}(c)\subseteq\Omega_{1},

whence q\vec{q} satisfies the conclusion of the lemma. ∎

We have established that Ω1\Omega_{1} is invariant under the \mathbb{R}-flow corresponding to the actions by vectors s×0s\times\vec{0}, and we may therefore naturally define a special flow 𝔉¯\overline{\mathfrak{F}}^{\prime} on Ω1×d1\Omega_{1}\times\mathbb{R}^{d-1} by

(x,q)(r1,r2,,rd)=(xr1×0d1,q+(r2,,rd)).(x,\vec{q}\,)\boxplus(r_{1},r_{2},\ldots,r_{d})=\bigl{(}x\oplus r_{1}\times\vec{0}^{d-1},\ \vec{q}+(r_{2},\ldots,r_{d})\bigr{)}.

Note that \boxplus is used for the action to distinguish it from both the actions given by 𝔉\mathfrak{F} and 𝔉\mathfrak{F}^{\prime}. We are going to verify that 𝔉¯\overline{\mathfrak{F}}^{\prime} is isomorphic to 𝔉\mathfrak{F}^{\prime}, and to this end we define two Borel maps μ:ΩΩ1\mu:\Omega\to\Omega_{1} and ν:Ωd1\nu:\Omega\to\mathbb{R}^{d-1} such that Ωx(μ(x),ν(x))Ω1×d1\Omega\ni x\mapsto(\mu(x),\nu(x))\in\Omega_{1}\times\mathbb{R}^{d-1} will be the desired isomorphism. For xΩx\in\Omega, nn and c𝒞nc\in\mathcal{C}_{n}, xWn(c)x\in W_{n}(c), set

μ(x)\displaystyle\mu(x) =(ϕn|Wn(c))1(proj1(ϕn(x))×0d1)Ω1\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\mathrm{proj}_{1}(\phi_{n}(x))\times\vec{0}^{d-1}\bigr{)}\in\Omega_{1}
ν(x)\displaystyle\nu(x) =proj[2,d](ϕn(x))d1.\displaystyle=\mathrm{proj}_{[2,d]}(\phi_{n}(x))\in\mathbb{R}^{d-1}.
Lemma 17.

Maps μ\mu and ν\nu are well-defined in the sense that their values do not depend on the choice of nn and c𝒞nc\in\mathcal{C}_{n}.

Proof.

Let m<nm<n and c𝒞mc^{\prime}\in\mathcal{C}_{m} be other elements such that xWm(c)x\in W_{m}(c). Once again, ϕn(x)=ϕm(x)+sm,n(c,c)\phi_{n}(x)=\phi_{m}(x)+\vec{s}_{m,n}(c^{\prime},c), by item 10(iii). In particular, the projections of ϕn(x)\phi_{n}(x) and ϕm(x)\phi_{m}(x) onto the last (d1)(d-1)-many coordinates are equal, because sm,n(c,c)=sm,n(c,c)×0d1\vec{s}_{m,n}(c^{\prime},c)=s_{m,n}(c^{\prime},c)\times\vec{0}^{d-1} This shows that ν(x)\nu(x) is well-defined.

For s=proj1(ϕn(x))s=\mathrm{proj}_{1}(\phi_{n}(x)), s=proj1(ϕm(x))s^{\prime}=\mathrm{proj}_{1}(\phi_{m}(x)), and s=s×0d1\vec{s}=s\times\vec{0}^{d-1}, s=s×0d1\vec{s}\,^{\prime}=s^{\prime}\times\vec{0}^{d-1}, we have

s\displaystyle\vec{s} =s+sm,n(c,c)\displaystyle=\vec{s}\,^{\prime}+\vec{s}_{m,n}(c^{\prime},c)
=ϕm((ϕm|Wm(c))1(s))+sm,n(c,c)\displaystyle=\phi_{m}\bigl{(}(\phi_{m}|_{W_{m}(c^{\prime})})^{-1}(\vec{s}\,^{\prime})\bigr{)}+\vec{s}_{m,n}(c^{\prime},c)
item 10(iii) =ϕn((ϕm|Wm(c))1(s)).\displaystyle=\phi_{n}\bigl{(}(\phi_{m}|_{W_{m}(c^{\prime})})^{-1}(\vec{s}\,^{\prime})\bigr{)}.

Applying (ϕn|Wn(c))1(\phi_{n}|_{W_{n}(c)})^{-1} to both sides yields

(ϕn|Wn(c))1(s)=(ϕm|Wm(c))1(s),\displaystyle(\phi_{n}|_{W_{n}(c)})^{-1}(\vec{s})=(\phi_{m}|_{W_{m}(c^{\prime})})^{-1}(\vec{s}\,^{\prime}),

which shows that the value μ(x)Ω1\mu(x)\in\Omega_{1} does not depend on the choice of nn and c𝒞nc\in\mathcal{C}_{n}. ∎

Lemma 18.

The map Ωx(μ(x),ν(x))Ω1×d1\Omega\ni x\mapsto(\mu(x),\nu(x))\in\Omega_{1}\times\mathbb{R}^{d-1} is a bijection.

Proof.

For injectivity, let x,yΩx,y\in\Omega be distinct; recall that the orbit equivalence relations of the flows 𝔉\mathfrak{F} and 𝔉\mathfrak{F}^{\prime} coincide by Lemma 14, and we denote it by EE. Note that xEμ(x)xE\mu(x) and yEμ(y)yE\mu(y), so if ¬xEy\neg xEy, then μ(x)μ(y)\mu(x)\neq\mu(y). Thus we need to consider the case xEyxEy and by 5(iv) there is some nn and c𝒞nc\in\mathcal{C}_{n} such that x,yWn(c)x,y\in W_{n}(c). Since ϕn|Wn(c):Wn(c)B(tn(c))\phi_{n}|_{W_{n}(c)}:W_{n}(c)\to B(t_{n}(c)) is injective, ϕn(x)ϕn(y)\phi_{n}(x)\neq\phi_{n}(y), and thus either

proj1(ϕn(x))\displaystyle\mathrm{proj}_{1}(\phi_{n}(x)) proj1(ϕn(y))or\displaystyle\neq\mathrm{proj}_{1}(\phi_{n}(y))\qquad\textrm{or}
proj[2,d](ϕn(x))\displaystyle\mathrm{proj}_{[2,d]}(\phi_{n}(x)) proj[2,d](ϕn(y)),\displaystyle\neq\mathrm{proj}_{[2,d]}(\phi_{n}(y)),

and hence either μ(x)μ(y)\mu(x)\neq\mu(y) or ν(x)ν(y)\nu(x)\neq\nu(y).

For surjectivity, pick (x,q)Ω1×d1(x,\vec{q}\,)\in\Omega_{1}\times\mathbb{R}^{d-1}, and nn, c𝒞nc\in\mathcal{C}_{n}, with xWn(c)x\in W_{n}(c) and tn(c)ϕn(x)qt_{n}(c)-||\phi_{n}(x)||\geq||\vec{q}||, which exist by Lemma 12. Note that by the coherence property of LnL_{n} established in the proof of Lemma 15, we have (c,x)Ln(c,x)\in L_{n}, which is equivalent to ϕn(x)×0d1\phi_{n}(x)\in\mathbb{R}\times\vec{0}^{d-1}. These conditions ensure that (ϕn|Wn(c))1(\phi_{n}|_{W_{n}(c)})^{-1} is defined on ϕn(x)+0×q\phi_{n}(x)+0\times\vec{q} and

(μ,ν)((ϕn|Wn(c))1(ϕn(x)+0×q))=(x,q),(\mu,\nu)\bigl{(}(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\phi_{n}(x)+0\times\vec{q}\,\bigr{)}\bigr{)}=(x,\vec{q}\,),

witnessing surjectivity. ∎

At last, we can check that the flows 𝔉\mathfrak{F}^{\prime} and 𝔉¯\overline{\mathfrak{F}}^{\prime} are isomorphic.

Lemma 19.

The map Ωx(μ(x),ν(x))Ω1×d1\Omega\ni x\mapsto(\mu(x),\nu(x))\in\Omega_{1}\times\mathbb{R}^{d-1} is an isomorphism of flows 𝔉\mathfrak{F}^{\prime} and 𝔉¯\overline{\mathfrak{F}}^{\prime}.

Proof.

Once Lemma 18 is available, our remaining goal is to show that (μ,ν)(xr)=(μ(x),ν(x))r(\mu,\nu)(x\oplus\vec{r}\,)=(\mu(x),\nu(x))\boxplus\vec{r} for all xΩx\in\Omega and all rd\vec{r}\in\mathbb{R}^{d}. We first verify this for those r\vec{r} that satisfy proj1(r)=0\mathrm{proj}_{1}(\vec{r}\,)=0, i.e., r=0×q\vec{r}=0\times\vec{q}, for some qd1\vec{q}\in\mathbb{R}^{d-1}. One has

ν(x0×q)\displaystyle\nu(x\oplus 0\times\vec{q}\,) =proj[2,d](ϕn(x)+0×q)\displaystyle=\mathrm{proj}_{[2,d]}\bigl{(}\phi_{n}(x)+0\times\vec{q}\,\bigr{)}
=proj[2,d](ϕn(x))+q=ν(x)+q,\displaystyle=\mathrm{proj}_{[2,d]}\bigl{(}\phi_{n}(x)\bigr{)}+\vec{q}=\nu(x)+\vec{q},
μ(x0×q)\displaystyle\mu(x\oplus 0\times\vec{q}\,) =(ϕn|Wn(c))1(proj1(ϕn(x)+0×q)×0d1)\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\mathrm{proj}_{1}\bigl{(}\phi_{n}(x)+0\times\vec{q}\,\bigr{)}\times\vec{0}^{d-1}\bigr{)}
=(ϕn|Wn(c))1(proj1(ϕn(x))×0d1)=μ(x).\displaystyle=(\phi_{n}|_{W_{n}(c)})^{-1}\bigl{(}\mathrm{proj}_{1}(\phi_{n}(x))\times\vec{0}^{d-1}\bigr{)}=\mu(x).

We have thus shown that

(5) (μ,ν)(x0×q)=((μ,ν)(x))0×qfor all qd1 and all xΩ.(\mu,\nu)(x\oplus 0\times\vec{q}\,)=\bigl{(}(\mu,\nu)(x)\bigr{)}\boxplus 0\times\vec{q}\quad\textrm{for all $\vec{q}\in\mathbb{R}^{d-1}$ and all $x\in\Omega$.}

Note also that (μ,ν)(y)=(y,0)(\mu,\nu)(y)=(y,\vec{0}) for all yΩ1Ωy\in\Omega_{1}\subseteq\Omega, and therefore by Lemma 15 for all ss\in\mathbb{R} and all yΩ1y\in\Omega_{1}

(6) (μ,ν)(ys×0d1)=(ys×0d1,0)=(μ,ν)(y)(s×0d1).(\mu,\nu)(y\oplus s\times\vec{0}^{d-1})=(y\oplus s\times\vec{0}^{d-1},\vec{0})=(\mu,\nu)(y)\boxplus(s\times\vec{0}^{d-1}).

For any xΩx\in\Omega, Lemma 16 gives an element q0d1\vec{q}_{0}\in\mathbb{R}^{d-1} such that x0×q0Ω1x\oplus 0\times\vec{q}_{0}\in\Omega_{1}. Let rd\vec{r}\in\mathbb{R}^{d} be arbitrary, and write it as r=s×0d1+0×q\vec{r}=s\times\vec{0}^{d-1}+0\times\vec{q} for some ss\in\mathbb{R} and qd1\vec{q}\in\mathbb{R}^{d-1}. We have

(μ,ν)(xr)\displaystyle(\mu,\nu)(x\oplus\vec{r}\,) =(μ,ν)(x0×q0(0×q0)s×0d10×q)\displaystyle=(\mu,\nu)\bigl{(}x\oplus 0\times\vec{q}_{0}\oplus(-0\times\vec{q}_{0})\oplus s\times\vec{0}^{d-1}\oplus 0\times\vec{q}\,\bigr{)}
=(μ,ν)(x0×q0s×0d10×(qq0))\displaystyle=(\mu,\nu)\bigl{(}x\oplus 0\times\vec{q}_{0}\oplus s\times\vec{0}^{d-1}\oplus 0\times(\vec{q}-\vec{q}_{0})\bigr{)}
Eq. (5) =(μ,ν)(x0×q0s×0d1)0×(qq0)\displaystyle=(\mu,\nu)\bigl{(}x\oplus 0\times\vec{q}_{0}\oplus s\times\vec{0}^{d-1}\bigr{)}\boxplus 0\times(\vec{q}-\vec{q}_{0})
Eq. (6) =(μ,ν)(x0×q0)s×0d10×(qq0)\displaystyle=(\mu,\nu)(x\oplus 0\times\vec{q}_{0})\boxplus s\times\vec{0}^{d-1}\boxplus 0\times(\vec{q}-\vec{q}_{0})
Eq. (5) =(μ,ν)(x)0×q0s×0d10×(qq0)\displaystyle=(\mu,\nu)(x)\boxplus 0\times\vec{q}_{0}\boxplus s\times\vec{0}^{d-1}\boxplus 0\times(\vec{q}-\vec{q}_{0})
=(μ,ν)(x)(s×q)=(μ,ν)(x)r.\displaystyle=(\mu,\nu)(x)\boxplus(s\times\vec{q}\,)=(\mu,\nu)(x)\boxplus\vec{r}.

Thus (μ,ν)(\mu,\nu) is an isomorphism between flows 𝔉\mathfrak{F}^{\prime} and 𝔉¯\overline{\mathfrak{F}}^{\prime}. ∎

The following theorem summarizes the analysis that has been conducted in Lemmas 10 through 19.

Theorem 20.

Every free Borel d\mathbb{R}^{d}-flow, d2d\geq 2, is smoothly equivalent to a special flow.

4. Smooth Equivalence of Flows

We are finally ready for the proof of the main result of this article—smooth equivalence of all non-tame Borel d\mathbb{R}^{d}-flows. For this we just need to combine Theorem 20 with the result of Miller–Rosendal on one-dimensional flows.

Theorem 21.

All non-tame free Borel d\mathbb{R}^{d}-flows, d2d\geq 2, are smoothly equivalent.

Proof.

Let 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} be non-tame free Borel d\mathbb{R}^{d}-flows. By Theorem 20, each of them is smoothly equivalent to a special flow dΩi×d1\mathbb{R}^{d}\curvearrowright\Omega_{i}\times\mathbb{R}^{d-1}, i=1,2i=1,2. Note that neither of the \mathbb{R}-flows Ωi\mathbb{R}\curvearrowright\Omega_{i} can be tame, for otherwise the corresponding d\mathbb{R}^{d}-flow would also be tame. By the Miller–Rosendal result (see Theorem 1), there is a smooth equivalence ξ¯:Ω1Ω2\bar{\xi}:\Omega_{1}\to\Omega_{2} between the \mathbb{R}-flows. Let αξ¯:Ω1×\alpha_{\bar{\xi}}:\Omega_{1}\times\mathbb{R}\to\mathbb{R} be the corresponding family of diffeomorphisms defined by αξ¯(x,s)=ρ(ξ¯(x),ξ¯(x+s))\alpha_{\bar{\xi}}(x,s)=\rho(\bar{\xi}(x),\bar{\xi}(x+s)). The map ξ:Ω1×d1Ω2×d1\xi:\Omega_{1}\times\mathbb{R}^{d-1}\to\Omega_{2}\times\mathbb{R}^{d-1} defined by ξ(x,q)=(ξ¯(x),q)\xi(x,\vec{q}\,)=(\bar{\xi}(x),\vec{q}\,) is a smooth equivalence between the special flows, because

αξ((x,q),(r1,,rd))=(αξ¯(x,r1),r2,,rd)d\alpha_{\xi}\bigl{(}(x,\vec{q}\,),(r_{1},\ldots,r_{d})\bigr{)}=(\alpha_{\bar{\xi}}(x,r_{1}),r_{2},\ldots,r_{d})\in\mathbb{R}^{d}

is a CC^{\infty}-smooth orientation preserving diffeomorphism for all xΩ1x\in\Omega_{1} and all qd1\vec{q}\in\mathbb{R}^{d-1}. Flows 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} are smoothly equivalent by transitivity of the smooth equivalence relation. ∎

Our approach to the construction of smooth equivalence between multidimensional flows is different from those taken in the ergodic theoretical antecedents. Of particular interest is the technique used in [Fel91]. Their strategy is to start with an orbit equivalence between cross sections of flows222The conditions of when such an orbit equivalence exists are well understood and follow from Dye’s Theorem [Dye59, Dye63] and Dougherty–Jackson–Kechris classification of the hyperfinite equivalence relations [DJK94]., and extend such a map to a smooth equivalence by a back-and-forth argument. It would be interesting to establish the possibility of such an extension in the descriptive set-theoretical context. In [Slu19], such an extension was constructed up to a compressible set. It was also proven therein that if a complete extension is always possible, then it will imply smooth equivalence of all flows (Theorem 21 above).

Conjecture 22.

Let dΩ1\mathbb{R}^{d}\curvearrowright\Omega_{1} and dΩ2\mathbb{R}^{d}\curvearrowright\Omega_{2} be free Borel flows, d2d\geq 2, let 𝒞iΩi\mathcal{C}_{i}\subseteq\Omega_{i} be cocompact cross sections, and let ζ:𝒞1𝒞2\zeta:\mathcal{C}_{1}\to\mathcal{C}_{2} be an orbit equivalence (i.e., a Borel bijection such that cE1cζ(c)E2ζ(c)cE_{1}c^{\prime}\iff\zeta(c)E_{2}\zeta(c^{\prime})). There exists a smooth equivalence ξ:Ω1Ω2\xi:\Omega_{1}\to\Omega_{2} that extends ζ\zeta.

Rudolph’s Theorem [Rud79, Proposition 1.1] produces a smooth equivalence of d\mathbb{R}^{d}-flows which is also a Lebesgue orbit equivalence, i.e., a map that preserves the Lebesgue measure between orbits (see [Slu17] for the discussion of the notion of Lebesgue orbit equivalence). This is a significant strengthening of the smooth equivalence, and it is therefore interesting to see if such a strengthening is possible in the descriptive set-theoretical context as well. Any Lebesgue orbit equivalence produces an isomorphism between the spaces of invariant measures of the flows [Slu17, Theorem 4.5], so we need to restrict ourselves to flows that have the same cardinalities of sets of ergodic invariant probability measures. According to [Slu17, Theorem 9.1], this would ensure the existence of a Lebesgue orbit equivalence between the flows and, of course, Theorem 21 guarantees that there is a smooth equivalence as well. The question is whether the two can be combined.333We would like to thank the referee for suggesting this question.

Question 23.

Let 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} be free non-tame Borel d\mathbb{R}^{d}-flows, d2d\geq 2, having the same numbers of invariant ergodic probability measures. Is there a smooth equivalence between these flows which is also a Lebesgue orbit equivalence?

References

  • [BJ07] Charles M. Boykin and Steve Jackson. Borel boundedness and the lattice rounding property. In Advances in logic, volume 425 of Contemp. Math., pages 113–126. Amer. Math. Soc., Providence, RI, 2007.
  • [DJK94] Randall Dougherty, Steve Jackson, and Alexander S. Kechris. The structure of hyperfinite Borel equivalence relations. Trans. Amer. Math. Soc., 341(1):193–225, 1994.
  • [Dye59] Henry A. Dye. On groups of measure preserving transformations. I. Amer. J. Math., 81:119–159, 1959.
  • [Dye63] Henry A. Dye. On groups of measure preserving transformations. II. Amer. J. Math., 85:551–576, 1963.
  • [Fel76] Jacob Feldman. New KK-automorphisms and a problem of Kakutani. Israel J. Math., 24(1):16–38, 1976.
  • [Fel91] Jacob Feldman. Changing orbit equivalences of 𝐑d{\bf R}^{d} actions, d2d\geq 2, to be C{C}^{\infty} on orbits. Internat. J. Math., 2(4):409–427, 1991.
  • [Fel92] Jacob Feldman. Correction to: “Changing orbit equivalences of 𝐑d{\bf R}^{d} actions, d2d\geq 2, to be CC^{\infty} on orbits”. Internat. J. Math., 3(3):349–350, 1992.
  • [GJ15] Su Gao and Steve Jackson. Countable abelian group actions and hyperfinite equivalence relations. Invent. Math., 201(1):309–383, 2015.
  • [GJKS] Su Gao, Steve Jackson, Edward Krohne, and Brandon Seward. Borel combinatorics of abelian group actions. In preparation.
  • [JKL02] Steve Jackson, Alexander S. Kechris, and Alain Louveau. Countable Borel equivalence relations. J. Math. Log., 2(1):1–80, 2002.
  • [Kak43] Shizuo Kakutani. Induced measure preserving transformations. Proc. Imp. Acad. Tokyo, 19:635–641, 1943.
  • [Kat75] Anatole B. Katok. Time change, monotone equivalence, and standard dynamical systems. Dokl. Akad. Nauk SSSR, 223(4):789–792, 1975.
  • [Kat77] Anatole B. Katok. Monotone equivalence in ergodic theory. Izv. Akad. Nauk SSSR Ser. Mat., 41(1):104–157, 231, 1977.
  • [Kec95] Alexander S. Kechris. Classical descriptive set theory, volume 156 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1995.
  • [MR10] Benjamin D. Miller and Christian Rosendal. Descriptive Kakutani equivalence. J. Eur. Math. Soc. (JEMS), 12(1):179–219, 2010.
  • [MU17] Andrew S. Marks and Spencer T. Unger. Borel circle squaring. Ann. of Math. (2), 186(2):581–605, 2017.
  • [Nak88] Munetaka Nakamura. Time change and orbit equivalence in ergodic theory. Hiroshima Math. J., 18(2):399–412, 1988.
  • [ORW82] Donald S. Ornstein, Daniel Rudolph, and Benjamin Weiss. Equivalence of measure preserving transformations. Mem. Amer. Math. Soc., 37(262):xii+116, 1982.
  • [Rud79] Daniel Rudolph. Smooth orbit equivalence of ergodic 𝐑d{\bf R}^{d} actions, d2d\geq 2. Trans. Amer. Math. Soc., 253:291–302, 1979.
  • [Slu17] Konstantin Slutsky. Lebesgue orbit equivalence of multidimensional Borel flows: a picturebook of tilings. Ergodic Theory Dynam. Systems, 37(6):1966–1996, 2017.
  • [Slu19] Konstantin Slutsky. On time change equivalence of Borel flows. Fund. Math., 247(1):1–24, 2019.