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

𝐋𝟏\mathbf{L}^{\mathbf{1}} Full Groups of Flows

François Le Maître Université Paris Cité and Sorbonne Université, CNRS, IMJ-PRG, F-75013 Paris, France francois.le-maitre@imj-prg.fr  and  Konstantin Slutsky Department of Mathematics
Iowa State University
411 Morrill Road
Ames, IA 50011
kslutsky@gmail.com
Abstract.

We introduce the concept of an L1\mathrm{L}^{1} full group associated with a measure-preserving action of a Polish normed group on a standard probability space. Such groups are shown to carry a natural separable complete metric, and are thus Polish. Our construction generalizes L1\mathrm{L}^{1} full groups of actions of discrete groups, which have been studied recently by the first author.

We show that under minor assumptions on the actions, topological derived subgroups of L1\mathrm{L}^{1} full groups are topologically simple and — when the acting group is locally compact and amenable — are whirly amenable and generically two-generated. L1\mathrm{L}^{1} full groups of actions of compactly generated locally compact Polish groups are shown to remember the L1\mathrm{L}^{1} orbit equivalence class of the action.

For measure-preserving actions of the real line (also often called measure-preserving flows), the topological derived subgroup of an L1\mathrm{L}^{1} full groups is shown to coincide with the kernel of the index map, which implies that L1\mathrm{L}^{1} full groups of free measure-preserving flows are topologically finitely generated if and only if the flow admits finitely many ergodic components. The latter is in a striking contrast to the case of \mathbb{Z}-actions, where the number of topological generators is controlled by the entropy of the action.

We also study the coarse geometry of the L1\mathrm{L}^{1} full groups. The L1\mathrm{L}^{1} norm on the derived subgroup of the L1\mathrm{L}^{1} full group of an aperiodic action of a locally compact amenable group is proved to be maximal in the sense of Rosendal. For measure-preserving flows, this holds for the L1\mathrm{L}^{1} norm on all of the L1\mathrm{L}^{1} full group.

2020 Mathematics Subject Classification:
Primary 37A10, 37A15; Secondary 37A05, 37A20.
François Le Maître’s research was partially supported by the ANR project AGRUME (ANR-17-CE40-0026), the ANR Project AODynG (ANR-19-CE40-0008) and the IdEx Université de Paris (ANR-18-IDEX-0001).
Konstantin Slutsky’s research was partially supported by the ANR project AGRUME (ANR-17-CE40-0026) and NSF Grant DMS-2153981.

Chapter 1 Introduction

Full groups were introduced by H. Dye [Dye59] in the framework of measure-preserving actions of countable groups as measurable analogues of unitary groups of von Neumann algebras, by mimicking the fact that the latter are stable under countable cutting and pasting of partial isometries. These Polish groups have since been recognized as important invariants as they encode the induced partition of the space into orbits. A similar viewpoint applies in the setup of minimal homeomorphisms on the Cantor space [GPS99], where likewise the full groups are responsible for the orbit equivalence class of the action.

Full groups are defined to consist of transformations which act by a permutation on each orbit. When the action is free, one can associate with an element hh of the full group a cocycle defined by the equation h(x)=ρh(x)xh(x)=\rho_{h}(x)\cdot x. From the point of view of topological dynamics, it is natural to consider the subgroup of those hh for which the cocycle map is continuous, which is the defining condition for the so-called topological full groups. The latter has a much tighter control of the action, and encodes minimal homeomorphisms of the Cantor space up to flip-conjugacy (see [GPS99]).

A celebrated result of H. Dye states that all ergodic \mathbb{Z}-actions produce the same partition up to isomorphism, and hence the associated full groups are all isomorphic. The first named author has been motivated by the above to seek for the analog of topological full groups in the context of ergodic theory, which was achieved in [LM18] by imposing integrability conditions on the cocycle. In particular, he introduced L1\mathrm{L}^{1} full groups of measure-preserving ergodic transformations, and showed based on the result of R. M. Belinskaja [Bel68] that they also determine the action up to flip-conjugacy. Unlike in the context of Cantor dynamics, these L1\mathrm{L}^{1} full groups are uncountable, but they carry a natural Polish topology.

In this work, we widen the concept of an L1\mathrm{L}^{1} full group and associate such an object with any measure-preserving Borel action of a Polish normed group (the reader may consult Appendix A for a concise reminder about group norms). Quasi-isometric compatible norms will result in the same L1\mathrm{L}^{1} full groups, so actions of Polish boundedly generated groups have canonical L1\mathrm{L}^{1} full groups associated with them based on to the work of C. Rosendal [Ros22]. Our study also parallels the generalization of the full group construction introduced by A. Carderi and the first named author in [CLM16], where full groups were defined for Borel measure-preserving actions of Polish groups.

1.1. Main results

Let GG be a Polish group with a compatible norm \lVert\cdot\rVert and consider a Borel measure-preserving action GXG\curvearrowright X on a standard probability space (X,μ)(X,\mu). The group action defines an orbit equivalence relation G\mathcal{R}_{G} by declaring points x1,x2Xx_{1},x_{2}\in X equivalent whenever Gx1=Gx2G\cdot x_{1}=G\cdot x_{2}. The norm induces a metric onto each G\mathcal{R}_{G}-class via D(x1,x2)=infgG{g:gx1=x2}D(x_{1},x_{2})=\inf_{g\in G}\{\lVert g\rVert:gx_{1}=x_{2}\}. Following [CLM16], a full group of the action is denoted by [G][\mathcal{R}_{G}\mkern 1.5mu] and is defined as the collection of all measure-preserving TAut(X,μ)T\in\mathrm{Aut}(X,\mu) that satisfy xGTxx\mathcal{R}_{G}Tx for all xXx\in X. The L1\mathrm{L}^{1} full group [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is given by those T[G]T\in[\mathcal{R}_{G}\mkern 1.5mu] for which the map XxD(x,Tx)X\ni x\mapsto D(x,Tx) is integrable. This defines a subgroup of [G][\mathcal{R}_{G}\mkern 1.5mu], and we show in Theorem 2.10 that these groups are Polish in the topology of the norm T=XD(x,Tx)𝑑μ(x)\lVert T\rVert=\int_{X}D(x,Tx)\,d\mu(x). The strategy of establishing this statement is analogous to that of [CLM18], where the Polish topology for full groups [G][\mathcal{R}_{G}\mkern 1.5mu] was defined.

It is a general and well-known phenomenon in the study of all kinds and variants of full groups that their structure is usually best understood through the derived subgroups. Our setup is no exception.

Theorem 1.

The topological derived group of any aperiodic L1\mathrm{L}^{1} full group is equal to the closed subgroup generated by involutions.

The argument needed for Theorem 1 is quite robust. We extract the idea used in [LM18], isolate the class of finitely full groups, and show that under mild assumptions on the action, Theorem 1 holds for such groups. We provide these arguments in Section 3 and in Corollary 3.15 in particular. Alongside we mention Corollary 3.21 which implies that L1\mathrm{L}^{1} full groups of ergodic actions are topologically simple.

For the rest of our results we narrow down the generality of the acting groups, and consider locally compact Polish normed groups. In Chapter 4, we show that if H<GH<G is a dense subgroup of a locally compact Polish normed group GG then [HX]1[H\curvearrowright X\mkern 1.5mu]_{1} is dense in [GX]1[G\curvearrowright X\mkern 1.5mu]_{1}. In fact, we prove a considerably stronger statement by showing that for each T[GX]T\in[G\curvearrowright X\mkern 1.5mu] and ϵ>0\epsilon>0 there is S[HX]S\in[H\curvearrowright X\mkern 1.5mu] such that esssupxXD(Tx,Sx)<ϵ\operatorname*{ess\,sup}_{x\in X}D(Tx,Sx)<\epsilon.

Recall that a topological group is amenable if all of its continuous actions on compact spaces preserve some Radon probability measure, and that it is whirly amenable if it is amenable and moreover every invariant Radon measure is supported on the set of fixed points. The following is a combination of Theorem 5.8 and Corollary 5.10.

Theorem 2.

Let GXG\curvearrowright X be a measure-preserving action of a locally compact Polish normed group. Consider the following three statements:

  1. (1)

    GG is amenable;

  2. (2)

    the topological derived group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is whirly amenable.

  3. (3)

    the L1\mathrm{L}^{1} full group [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is amenable;

The implications (1)\implies (2)\implies (3) always hold. If GG is unimodular and the action is free, then the three statements above are all equivalent.

When the acting group is amenable and orbits of the action are uncountable, we are able to compute the topological rank of the derived L1\mathrm{L}^{1} full groups — that is, the minimal number of elements needed to generate a dense subgroup. Theorem 5.19 contains a stronger version of the following.

Theorem 3.

Let GXG\curvearrowright X be a measure-preserving action of an amenable locally compact Polish normed group on a standard probability space (X,μ)(X,\mu). If all orbits of the action are uncountable, then the topological rank of the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is equal to 22.

It is instructive to contrast the situation with the actions of finitely generated groups, where finiteness of the topological rank of the derived L1\mathrm{L}^{1} full group is equivalent to finiteness of the Rokhlin entropy of the action [LM21].

Our most refined understanding of L1\mathrm{L}^{1} full groups is achieved for free actions of \mathbb{R}, which are known as flows. All the results we described so far are valid for all compatible norms on the acting group. When it comes to the actions of \mathbb{R}, however, we consider only the standard Euclidean norm on it. Just like the actions of \mathbb{Z}, flows give rise to an important homomorphism, known as the index map. Assuming the flow is ergodic, the index map can be described most easily as [X]1TX|ρT|𝑑μ[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}\ni T\mapsto\int_{X}|\rho_{T}|\,d\mu, where ρT\rho_{T} is the cocycle of TT. Chapter 6 is devoted to the analysis of the index map for general \mathbb{R}-flows.

The most technically challenging result of our work is summarized in Theorem 10.1, which identifies the derived L1\mathrm{L}^{1} full group of a flow with the kernel of the index map, and describes the abelianization of [X]1[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}.

Theorem 4.

Let \mathcal{F} be a measure-preserving flow on (X,μ)(X,\mu). The kernel of the index map is equal to the derived L1\mathrm{L}^{1} full group of the flow, and the topological abelianization of []1[\mathcal{F}\mkern 1.5mu]_{1} is \mathbb{R}.

Theorem 4 parallels the known results for \mathbb{Z}-actions from [LM18]. The structure of its proof, however, has an important difference. We rely crucially on the fact that each element of the full group acts in a measure-preserving manner on each orbit. This allows us to use Hopf’s decomposition (described in Appendix B) in order to separate any given element T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} into two parts — recurrent and dissipative. If the acting group were discrete, the recurrent part would reduce to periodic orbits only. This is not at all the case for non-discrete groups, hence we need a new machinery to understand non-periodic recurrent transformations. To cope with this, we introduce the concept of an intermitted transformation, which plays the central role in Chapter 8, and which we hope will find other applications.

Theorems 3 and 4 can be combined to obtain estimates for the topological rank of the whole L1\mathrm{L}^{1} full groups of flows, which is the content of Proposition 10.3.

Theorem 5.

Let \mathcal{F} be a free measure-preserving flow on a standard probability space (X,μ)(X,\mu). The topological rank rk([]1)\mathrm{rk}([\mathcal{F}\mkern 1.5mu]_{1}) is finite if and only if the flow has finitely many ergodic components. Moreover, if \mathcal{F} has exactly nn ergodic components then

n+1rk([]1)n+3.n+1\leq\mathrm{rk}([\mathcal{F}\mkern 1.5mu]_{1})\leq n+3.

In particular, the topological rank of the L1\mathrm{L}^{1} full group of an ergodic flow is equal to either 22, 33 or 44. We conjecture that it is always equal to 22, and more generally that the topological rank of the L1\mathrm{L}^{1} full group of any measure-preserving flow is equal to n+1n+1 where nn is the number of ergodic components.

Our work connects to the notion of L1\mathrm{L}^{1} orbit equivalence, an intermediate notion between orbit equivalence and conjugacy. It can be traced back to the work of R. M. Belinskaja[Bel68] but recently attracted more attention. Stated in our framework, two flows are L1\mathrm{L}^{1} orbit equivalent if they can be conjugated so that the first flow is contained in the L1\mathrm{L}^{1} full group of the second and vice versa. A symmetric version of Belinskaja’s theorem is that ergodic \mathbb{Z}-actions are L1\mathrm{L}^{1} orbit equivalent if and only if they are flip conjugate. It is very natural to wonder whether this amazing result has a version for flows. Our Theorem 10.14 implies the following.

Theorem 6.

If two measure-preserving ergodic flows are L1\mathrm{L}^{1} orbit equivalent, then they admit some cross-sections whose induced transformations111We refer the reader to Definition 10.11 and the paragraph that follows it for details on the measure-preserving transformation one associates to a cross-section. are flip-conjugate.

We do not know whether the above result is optimal, that is, whether having flip-conjugate cross-sections implies L1\mathrm{L}^{1} orbit equivalence, but it seems unlikely. It is tempting to think that the correct analogue of Belinskaja’s theorem would be a positive answer to the following question.

Question 1.1.

Let 1\mathcal{F}_{1} and 2\mathcal{F}_{2} be free ergodic measure-preserving flows which are L1\mathrm{L}^{1} orbit equivalent. Is it true that there is α\alpha\in\mathbb{R}^{*} such that 1\mathcal{F}_{1} and 2mα\mathcal{F}_{2}\circ m_{\alpha} are isomorphic, where mαm_{\alpha} denotes the multiplication by α\alpha?

Let us also mention that Theorem 6 implies that there are uncountably many L1\mathrm{L}^{1} full groups of ergodic free measure-preserving flows up to (topological) group isomorphism (see Corollary 10.16 and the paragraph right after its proof).

Finally, we also investigate the coarse geometry of the L1\mathrm{L}^{1} full groups. We establish that the L1\mathrm{L}^{1} norm is maximal (in the sense of C. Rosendal[Ros22], see also Appendix A.2) on the derived subgroup of an L1\mathrm{L}^{1} full group of an aperiodic measure-preserving action of any locally compact amenable Polish group (Theorem 5.5). For the measure-preserving flows, the L1\mathrm{L}^{1} norm is, in fact, maximal on the whole full group (Theorem 10.18).

1.2. Preliminaries

1.2.1. Ergodic theory

Our work belongs to the field of ergodic theory, which means that all the constructions are defined and results are proven up to null sets. On a number of occasions, we allow ourselves to deviate from the pedantic accuracy and write “for all xx…” when we really ought to say “for almost all xx…”, etc. The only part where certain care needs to be exercised in this regard appears in Chapter 2. We define L1\mathrm{L}^{1} full groups for Borel measure-preserving actions of Polish normed groups, and we need a genuine action on the space XX for these to make sense just as in [CLM16]. Boolean actions (also called near actions) of general Polish groups do not admit realizations in general [GTW05], and even when they do, it could happen that different realizations yield different full groups. This subtlety disappears once we move our attention to locally compact group actions, which is the case for Chapter 4 and onwards. All Boolean actions of locally compact Polish groups admit Borel realizations which are all conjugate up to measure zero (and hence have the same full group), so null sets can be neglected just as they always are in ergodic theory.

By a standard probability space we mean the unique (up to isomorphism) separable atomless measure space (X,μ)(X,\mu) with μ(X)=1\mu(X)=1, i.e., the unit interval [0,1][0,1] with the Lebesgue measure. A few times in Chapter 5 and Appendix C.1 we refer to a standard Lebesgue space, by which we mean a separable finite measure space, μ(X)<\mu(X)<\infty, thus in contrast to the notion of the standard probability space allowing atoms and omitting the normalization requirement. We denote by Aut(X,μ)\mathrm{Aut}(X,\mu) the group of all measure-preserving bijections of (X,μ)(X,\mu) up to measure zero. This is a Polish group for the weak topology, defined by TnTT_{n}\to T if and only if for all AXA\subseteq X Borel, μ(Tn(A)T(A))0\mu(T_{n}(A)\bigtriangleup T(A))\to 0. The weak topology is a Polish group topology, see [Kec10, Sec. 1]. Given TAut(X,μ)T\in\mathrm{Aut}(X,\mu), its support is the set

suppT={xX:T(x)x}.\operatorname*{supp}T=\{x\in X:T(x)\neq x\}.

A measure-preserving bijection TT is called periodic when almost all its orbits are finite. Periodicity implies the existence of a fundamental domain AA for TT, namely a measurable set which intersects every TT-orbit at exactly one point. Since the ambient measure μ\mu is finite, the existence of a fundamental domain actually characterizes periodicity.

1.2.2. Orbit equivalence relations

Any group action GXG\curvearrowright X induces the orbit equivalence relation GX\mathcal{R}_{G\curvearrowright X}, where two points x,yXx,y\in X are GX\mathcal{R}_{G\curvearrowright X}-equivalent whenever Gx=GyG\cdot x=G\cdot y. We will usually write this equivalence relation simply as G\mathcal{R}_{G} for brevity. For the actions X\mathbb{Z}\curvearrowright X generated by an automorphism TAut(X,μ)T\in\mathrm{Aut}(X,\mu), we denote the corresponding orbit equivalence relation by T\mathcal{R}_{T}. For clarity, we may sometimes want to name a measure-preserving action as α\alpha and write G𝛼XG\overset{\alpha}{\curvearrowright}X. Then for all gGg\in G we denote by α(g)\alpha(g) the measure-preserving transformation of (X,μ)(X,\mu) induced by the action of gg.

We encounter various equivalence relations throughout this monograph. An equivalence class of a point xXx\in X under the relation \mathcal{R} is denoted by [x][x]_{\mathcal{R}} and the saturation of a set AXA\subseteq X is denoted by [A][A]_{\mathcal{R}} and is defined to be the union of \mathcal{R}-equivalence classes of the elements of AA: [A]=xA[x][A]_{\mathcal{R}}=\bigcup_{x\in A}[x]_{\mathcal{R}}. In particular, [x]T[x]_{\mathcal{R}_{T}} is the orbit of xx under the action of TT. The reader may notice that the notation for a saturation [A][A]_{\mathcal{R}} resembles that for the full group of an action [GX][G\curvearrowright X\mkern 1.5mu] (see Chapter 2). Both notations are standard, and we hope that confusion will not arise, as it applies to objects of different nature — sets and actions, respectively.

1.2.3. Actions of locally compact groups

Consider a measure-preserving action of a locally compact Polish (equivalently, second-countable) group GG on a standard Lebesgue space (X,μ)(X,\mu). A complete section for the action is a measurable set 𝒞X\mathcal{C}\subseteq X that intersects almost every orbit, i.e., μ(XG𝒞)=0\mu(X\setminus G\cdot\mathcal{C})=0. A cross-section is a complete section 𝒞X\mathcal{C}\subseteq X such that for some non-empty neighborhood of the identity UGU\subseteq G we have UcUc=Uc\cap Uc^{\prime}=\varnothing whenever c,c𝒞c,c^{\prime}\in\mathcal{C} are distinct. When the need to mention such a neighborhood UU explicitly arises, we say that 𝒞\mathcal{C} is a UU-lacunary cross-section.

With any cross-section 𝒞\mathcal{C} one associates a decomposition of the phase space known as the Voronoi tessellation. Slightly more generally, Appendix C.2 defines the concept of a tessellation over a cross-section, which corresponds to a set 𝒲𝒞×X\mathcal{W}\subseteq\mathcal{C}\times X for which the fibers 𝒲c={xX:(c,x)𝒲}\mathcal{W}_{c}=\{x\in X:(c,x)\in\mathcal{W}\}, c𝒞c\in\mathcal{C}, partition the phase space. Every tessellation 𝒲\mathcal{W} gives rise to an equivalence relation 𝒲\mathcal{R}_{\mathcal{W}}, where points x,yXx,y\in X are deemed equivalent whenever they belong to the same fiber 𝒲c\mathcal{W}_{c}, and to the projection map π𝒲:X𝒞\pi_{\mathcal{W}}:X\to\mathcal{C} that associates with each xXx\in X the unique c𝒞c\in\mathcal{C} which fiber 𝒲c\mathcal{W}_{c} the point xx belongs to, and is therefore defined by the condition (π𝒲(x),x)𝒲(\pi_{\mathcal{W}}(x),x)\in\mathcal{W} for all xXx\in X.

When the action GXG\curvearrowright X is free, each orbit GxG\cdot x can be identified with the acting group. Such a correspondence ggxg\mapsto gx depends on the choice of the anchor point xx within the orbit, but suffices to transfer structures invariant under right translations from the group GG onto the orbits of the action. For instance, if the acting group is locally compact, then a right-invariant Haar measure λ\lambda can be pushed onto orbits by setting λx(A)={gG:gxA}\lambda_{x}(A)=\{g\in G:gx\in A\} as discussed in Section 4.2. Freeness of the action GXG\curvearrowright X gives rise to the cocycle map ρ:GXG\rho:\mathcal{R}_{G\curvearrowright X}\to G which is well-defined by the condition ρ(x,y)x=y\rho(x,y)\cdot x=y. Elements of the full group [GX][G\curvearrowright X\mkern 1.5mu] are characterized as measure-preserving transformations TAut(X,μ)T\in\mathrm{Aut}(X,\mu) such that (T(x),x)GX(T(x),x)\in\mathcal{R}_{G\curvearrowright X} for all xXx\in X. With each T[GX]T\in[G\curvearrowright X\mkern 1.5mu] one may therefore associate the map ρT:XG\rho_{T}:X\to G, also known as the cocycle map, and defined by ρT(x)=ρ(x,Tx)\rho_{T}(x)=\rho(x,Tx). Both the context and the notation will clarify which cocycle map is being referred to.

All these concepts appear prominently in the chapters which deal with free measure-preserving flows, that is actions of \mathbb{R} on the standard probability space. We use the additive notation for such actions: ×X(x,r)x+rX\mathbb{R}\times X\ni(x,r)\mapsto x+r\in X. The group \mathbb{R} carries a natural linear order which is invariant under the group operation and can therefore be transferred onto orbits. More specifically, given a free measure-preserving flow X\mathbb{R}\curvearrowright X we use the notation x<yx<y whenever xx and yy belong to the same orbit and y=x+ry=x+r for some r>0r>0. Every cross-section 𝒞\mathcal{C} of a free flow intersects each orbit in a bi-infinite fashion — each c𝒞c\in\mathcal{C} has a unique successor and a unique predecessor in the order of the orbit. One therefore has a bijection σ𝒞:𝒞𝒞\sigma_{\mathcal{C}}:\mathcal{C}\to\mathcal{C}, called the the first return map or the induced map, which sends c𝒞c\in\mathcal{C} to the next element of the cross-section within the same orbit. We also make use of the gap function that measures the lengths of intervals of the cross-section, i.e., gap𝒞(c)=ρ(c,σ𝒞(c))\mathrm{gap}_{\mathcal{C}}(c)=\rho(c,\sigma_{\mathcal{C}}(c)).

There is also a canonical tessellation associated with a cross-section 𝒞\mathcal{C} which partitions each orbit into intervals between adjacent points of 𝒞\mathcal{C} and is given by 𝒲𝒞={(c,x)𝒞×X:cx<σ𝒞(c)}\mathcal{W}_{\mathcal{C}}=\{(c,x)\in\mathcal{C}\times X:c\leq x<\sigma_{\mathcal{C}}(c)\}. The associated equivalence relation 𝒲𝒞\mathcal{R}_{\mathcal{W}_{\mathcal{C}}} is denoted simply by 𝒞\mathcal{R}_{\mathcal{C}} and groups points (x,y)X(x,y)\in\mathcal{R}_{\mathbb{R}\curvearrowright X} which belong to the same interval of the tessellation, π𝒞(x)=π𝒞(y)\pi_{\mathcal{C}}(x)=\pi_{\mathcal{C}}(y). The 𝒞\mathcal{R}_{\mathcal{C}}-equivalence class of xXx\in X is equal to [x]𝒞=π𝒞(x)+[0,gap𝒞(π𝒞(x)))[x]_{\mathcal{R}_{\mathcal{C}}}=\pi_{\mathcal{C}}(x)+\bigl{[}0,\mathrm{gap}_{\mathcal{C}}(\pi_{\mathcal{C}}(x))\bigr{)}.

Often enough we need to restrict sets and functions to an 𝒞\mathcal{R}_{\mathcal{C}}-class. Since such a need arises very frequently, especially in Chapter 9, we adopt the following shorthand notations. Given a set AXA\subseteq X and a point c𝒞c\in\mathcal{C} the intersection A[c]𝒞A\cap[c]_{\mathcal{R}_{\mathcal{C}}} is denoted simply by A(c)A(c). Likewise, λc𝒲𝒞(A)\lambda_{c}^{\mathcal{W}_{\mathcal{C}}}(A) stands for λ({r:c+rA[c]𝒞})\lambda(\{r\in\mathbb{R}:c+r\in A\cap[c]_{\mathcal{R}_{\mathcal{C}}}\}) and corresponds to the Lebesgue measure of the set A[c]𝒞A\cap[c]_{\mathcal{R}_{\mathcal{C}}}. Moreover, λc𝒲𝒞(A)\lambda_{c}^{\mathcal{W}_{\mathcal{C}}}(A) will usually be shortened to λc𝒞(A)\lambda_{c}^{\mathcal{C}}(A), when the tessellation is clear from the context.

Chapter 2 L1\mathrm{L}^{1} full groups of Polish group actions

We begin by defining the key notion of interest for our work, namely the L1\mathrm{L}^{1} full groups of measure-preserving Borel actions of Polish normed groups on a standard probability space. Admittedly, the overall focus will be on actions of locally compact groups, and flows in particular. Nonetheless, the concept of an L1\mathrm{L}^{1} full group can be introduced for actions of arbitrary Polish normed groups, and we therefore begin with this level of generality.

2.1. L1\mathrm{L}^{1} spaces with values in metric spaces

By a Polish metric space we mean a separable complete metric space.

Definition 2.1.

Let (X,μ)(X,\mu) be a standard probability space, let (Y,dY)(Y,d_{Y}) be a Polish metric space, and let e^:XY\hat{e}:X\to Y be a measurable function. We define the e^\hat{e}-pointed L1\mathrm{L}^{1} space Le^1(X,Y)\mathrm{L}_{\hat{e}}^{1}(X,Y) as the metric space of measurable functions f:XYf:X\to Y such that XdY(e^(x),f(x))𝑑μ(x)<+\int_{X}d_{Y}\bigl{(}\hat{e}(x),f(x)\bigr{)}\,d\mu(x)<+\infty, equipped with the metric

d~Y(f1,f2)=XdY(f1(x),f2(x))𝑑μ(x),\tilde{d}_{Y}(f_{1},f_{2})=\int_{X}d_{Y}(f_{1}(x),f_{2}(x))\,d\mu(x),

which is finite by the triangle inequality using the function e^\hat{e} as the middle point.

Proposition 2.2.

Let (X,μ)(X,\mu) be a standard probability space and (Y,dY)(Y,d_{Y}) be a Polish metric space. (Le^1(X,Y),d~Y)(\mathrm{L}_{\hat{e}}^{1}(X,Y),\tilde{d}_{Y}) is a Polish metric space for any measurable function e^:XY\hat{e}:X\to Y.

Proof.

The argument follows closely the classical proof that (L1(X,),d~)(\mathrm{L}^{1}(X,\mathbb{R}),\tilde{d}_{\mathbb{R}}) is a Polish metric space. To check completeness, let us pick a Cauchy sequence (fn)n(f_{n})_{n\in\mathbb{N}} in Le^1(X,Y)\mathrm{L}_{\hat{e}}^{1}(X,Y). Without loss of generality we may assume that d~Y(fn,fn+1)<2n\tilde{d}_{Y}(f_{n},f_{n+1})<2^{-n}, nn\in\mathbb{N}. Consider the sets An={xX:dY(fn(x),fn+1(x))1/n2}A_{n}=\{x\in X:d_{Y}(f_{n}(x),f_{n+1}(x))\geq 1/n^{2}\}, n1n\geq 1. Chebyshev’s inequality shows that μ(An)n22n\mu(A_{n})\leq n^{2}2^{-n}, whence nμ(An)<\sum_{n}\mu(A_{n})<\infty. The Borel–Cantelli lemma implies that (fn(x))n(f_{n}(x))_{n\in\mathbb{N}} is pointwise Cauchy for almost every xXx\in X. Since (Y,dY)(Y,d_{Y}) is complete, the pointwise limit of (fn)n(f_{n})_{n\in\mathbb{N}} exists, and we denote it by f:XYf:X\to Y. Define functions hn,h:X0h_{n},h:X\to\mathbb{R}^{\geq 0} by

hn(x)=i<ndY(fi(x),fi+1(x)),h(x)=idY(fi(x),fi+1(x))=limnhn(x),h_{n}(x)=\sum_{i<n}d_{Y}(f_{i}(x),f_{i+1}(x)),\quad h(x)=\sum_{i\in\mathbb{N}}d_{Y}(f_{i}(x),f_{i+1}(x))=\lim_{n\to\infty}h_{n}(x),

and note that hL1(X,)h\in\mathrm{L}^{1}(X,\mathbb{R}) by Fatou’s lemma. Finally, we conclude that

d~Y(fn,f)\displaystyle\tilde{d}_{Y}(f_{n},f) =XdY(fn(x),f(x))𝑑μ(x)Xk=ndY(fk(x),fk+1(x))dμ(x)\displaystyle=\int_{X}d_{Y}(f_{n}(x),f(x))\,d\mu(x)\leq\int_{X}\sum_{k=n}^{\infty}d_{Y}(f_{k}(x),f_{k+1}(x))\,d\mu(x)
=X(h(x)hn(x))𝑑μ(x)0,\displaystyle=\int_{X}(h(x)-h_{n}(x))\,d\mu(x)\to 0,

where the last convergence follows from Lebesgue’s dominated convergence theorem.

To verify separability, pick a countable dense set DYD\subseteq Y and note that the subspace of maps taking values in DD is d~Y\tilde{d}_{Y}-dense (in fact, this subspace is dense in the much stronger sup metric). It then follows that the set of functions that take only finitely many values (all of which are elements of DD) is still dense. Finally, one uses a dense countable subalgebra of the measure algebra on XX and further restricts this subspace to the functions that are measurable with respect to the chosen subalgebra. The resulting countable collection is dense in Le^1(X,Y)\mathrm{L}^{1}_{\hat{e}}(X,Y). ∎

The group of measure-preserving automorphisms Aut(X,μ)\mathrm{Aut}(X,\mu) has a natural action by composition on Le^1(X,Y)\mathrm{L}^{1}_{\hat{e}}(X,Y), i.e., (Tf)(x)=f(T1x)(T\cdot f)(x)=f(T^{-1}x). Note that every automorphism acts by an isometry.

Proposition 2.3.

Let (X,μ)(X,\mu) be a standard probability space, (Y,dY)(Y,d_{Y}) be a Polish metric space, and e^:XY\hat{e}:X\to Y be a measurable function. The action of Aut(X,μ)\mathrm{Aut}(X,\mu) on Le^1(X,Y)\mathrm{L}_{\hat{e}}^{1}(X,Y) is continuous.

Proof.

The argument mirrors the one in [CLM16, Prop. 2.9(1)]. Given sequences TnTT_{n}\to T and fnff_{n}\to f we need to show that TnfnTfT_{n}\cdot f_{n}\to T\cdot f. Since the action is by isometries,

d~Y(Tnfn,Tf)=d~Y(fn,Tn1Tf)d~Y(fn,f)+d~Y(f,Tn1Tf).\tilde{d}_{Y}(T_{n}\cdot f_{n},T\cdot f)=\tilde{d}_{Y}(f_{n},T_{n}^{-1}T\cdot f)\leq\tilde{d}_{Y}(f_{n},f)+\tilde{d}_{Y}(f,T_{n}^{-1}T\cdot f).

It therefore suffices to show that for any fLe^1(X,Y)f\in\mathrm{L}^{1}_{\hat{e}}(X,Y) and any convergent sequence of automorphisms TnTT_{n}\to T one has d~Y(f,Tn1Tf)0\tilde{d}_{Y}(f,T_{n}^{-1}\circ T\cdot f)\to 0 as nn\to\infty. The latter is enough to check for functions that take only finitely many values since those are dense in Le^1(X,Y)\mathrm{L}^{1}_{\hat{e}}(X,Y). Suppose ff is such a step function over a partition X=i=1mAiX=\bigsqcup_{i=1}^{m}A_{i}. Convergence TnTT_{n}\to T implies μ(Tn1T(Ai)Ai)0\mu(T_{n}^{-1}T(A_{i})\triangle A_{i})\to 0 for all 1im1\leq i\leq m, which easily yields d~Y(f,Tn1Tf)0\tilde{d}_{Y}(f,T_{n}^{-1}T\cdot f)\to 0. ∎

When YY is a Polish group, there is a natural choice of the function e^\hat{e}, namely the constant function e^(x)=e\hat{e}(x)=e, where ee is the identity element of the group. We therefore simplify the notation in this case and write L1(X,Y)\mathrm{L}^{1}(X,Y), omitting the subscript e^\hat{e}.

Recall that a Polish normed group is a Polish group together with a compatible norm on it (see Appendix A.1). In particular, if (G,)(G,\lVert\cdot\rVert) is a Polish normed group, there is a canonical choice of a complete metric on GG, namely

dG(u,v)=(u1v+vu1)/2.d_{G}(u,v)=(\lVert u^{-1}v\rVert+\lVert vu^{-1}\rVert)/2.

The corresponding space L1(X,G)\mathrm{L}^{1}(X,G) is Polish by Proposition 2.2. Moreover, it is a Polish group under pointwise operations.

Proposition 2.4.

Let (G,)(G,\lVert\cdot\rVert) be a Polish normed group, and let GXG\curvearrowright X be a Borel measure-preserving action on a standard probability space. The space L1(X,G)\mathrm{L}^{1}(X,G) is a Polish normed group under the pointwise operations, (fg)(x)=f(x)g(x)(f\cdot g)(x)=f(x)g(x), f1(x)=f(x)1f^{-1}(x)=f(x)^{-1}, and the norm f1L1(X,G)=Xf(x)𝑑μ(x)\lVert f\rVert_{1}^{\mathrm{L}^{1}(X,G)}=\int_{X}\lVert f(x)\rVert\,d\mu(x).

Proof.

The space L1(X,G)\mathrm{L}^{1}(X,G) can equivalently be defined as the collection of all measurable functions f:XGf:X\to G with finite norm, f1L1(X,G)<\lVert f\rVert_{1}^{\mathrm{L}^{1}(X,G)}<\infty. Using the properties of the norm \lVert\cdot\rVert on GG,

fg1L1(X,G)\displaystyle\lVert fg\rVert_{1}^{\mathrm{L}^{1}(X,G)} =Xf(x)g(x)𝑑μ(x)X(f(x)+g(x))𝑑μ(x)\displaystyle=\int_{X}\lVert f(x)g(x)\rVert\,d\mu(x)\leq\int_{X}(\lVert f(x)\rVert+\lVert g(x)\rVert)\,d\mu(x)
=f1L1(X,G)+g1L1(X,G),\displaystyle=\lVert f\rVert_{1}^{\mathrm{L}^{1}(X,G)}+\lVert g\rVert_{1}^{\mathrm{L}^{1}(X,G)},
f11L1(X,G)\displaystyle\lVert f^{-1}\rVert_{1}^{\mathrm{L}^{1}(X,G)} =Xf(x)1𝑑μ(x)Xf(x)𝑑μ(x)=f1L1(X,G).\displaystyle=\int_{X}\lVert f(x)^{-1}\rVert\,d\mu(x)\leq\int_{X}\lVert f(x)\rVert\,d\mu(x)=\lVert f\rVert_{1}^{\mathrm{L}^{1}(X,G)}.

Hence, L1(X,G)\mathrm{L}^{1}(X,G) is closed under the group operations and 1L1(X,G)\lVert\cdot\rVert_{1}^{\mathrm{L}^{1}(X,G)} is a group norm on it.

To show that group operations are continuous, it suffices to check that for any gL1(X,G)g\in\mathrm{L}^{1}(X,G) and any sequence fnL1(X,G)f_{n}\in\mathrm{L}^{1}(X,G), nn\in\mathbb{N}, converging to zero, fn1L1(X,G)0\lVert f_{n}\rVert_{1}^{\mathrm{L}^{1}(X,G)}\to 0, there is a subsequence (fnk)k(f_{n_{k}})_{k} such that gfnkg11L1(X,G)0\lVert gf_{n_{k}}g^{-1}\rVert_{1}^{\mathrm{L}^{1}(X,G)}\to 0 as kk\to\infty (see, for instance, [BO10, Thm 3.4 and Lem. 3.5]).

Since fnf_{n} converges to 0 in L1(X,G)\mathrm{L}^{1}(X,G), we may pass to a subsequence (fnk)k(f_{n_{k}})_{k} such that fnk(x)0\lVert f_{n_{k}}(x)\rVert\to 0 for almost all xXx\in X. Let M=maxk{fnk1L1(X,G)}M=\max_{k}\{\lVert f_{n_{k}}\rVert_{1}^{\mathrm{L}^{1}(X,G)}\} and note that for all kk

Xg(x)fnk(x)g(x)1𝑑μ(x)X(2g(x)+fnk(x))𝑑μ(x)2g1L1(X,G)+M.\int_{X}\lVert g(x)f_{n_{k}}(x)g(x)^{-1}\rVert\,d\mu(x)\ \leq\int_{X}(2\lVert g(x)\rVert+\lVert f_{n_{k}}(x)\rVert)\,d\mu(x)\leq 2\lVert g\rVert_{1}^{\mathrm{L}^{1}(X,G)}+M.

It remains to apply Lebesgue’s dominated convergence theorem to the sequence gfnkg1gf_{n_{k}}g^{-1}, kk\in\mathbb{N}, concluding that gfnkg11L1(X,G)0\lVert gf_{n_{k}}g^{-1}\rVert_{1}^{\mathrm{L}^{1}(X,G)}\to 0. ∎

2.2. L1\mathrm{L}^{1} full groups of Polish normed group actions

Let (G,)(G,\left\lVert\cdot\right\rVert) be a Polish normed group, and let GXG\curvearrowright X be a measure-preserving Borel action on a standard probability space (X,μ)(X,\mu). Let also GX×X\mathcal{R}_{G}\subseteq X\times X denote the equivalence relation induced by this action, namely

G={(x,gx):xX,gG}.\mathcal{R}_{G}=\{(x,g\cdot x):x\in X,g\in G\}.

The norm induces a metric on each G\mathcal{R}_{G}-equivalence class via

(2.1) D(x,y)=infuG{u:ux=y} for (x,y)G.D(x,y)=\inf_{u\in G}\{\,\left\lVert u\right\rVert:ux=y\,\}\text{ for }(x,y)\in\mathcal{R}_{G}.

Properties of the metric are straightforward except, possibly, for the implication D(x,y)=0x=yD(x,y)=0\implies x=y. To justify the latter, let unGu_{n}\in G, nn\in\mathbb{N}, be a sequence such that uneu_{n}\to e and unx=yu_{n}x=y. Elements un1u0u^{-1}_{n}u_{0}, nn\in\mathbb{N}, belong to the stabilizer of xx. By Miller’s theorem [Mil77], stabilizers of all points are closed, whence u0=limnun1u0u_{0}=\lim_{n}u_{n}^{-1}u_{0} fixes xx. Thus u0x=xu_{0}x=x, and x=yx=y as claimed.

A. Carderi and the first named author introduced in [CLM16] orbit full groups of Borel measure-preserving Polish group actions on standard probability spaces, which we will simply call full groups. Given such an action GXG\curvearrowright X, they define the full group of the action [GX][G\curvearrowright X\mkern 1.5mu] to consist of those measure-preserving transformations TAut(X,μ)T\in\mathrm{Aut}(X,\mu) that preserve the equivalence classes of G\mathcal{R}_{G}:

[GX]={TAut(X,μ):xX(x,T(x))G}.[G\curvearrowright X]=\bigl{\{}T\in\mathrm{Aut}(X,\mu):\forall x\in X\ (x,T(x))\in\mathcal{R}_{G}\bigr{\}}.

They showed that full groups are Polish with respect to the natural topology of convergence in measure.

Suppose that the acting group GG is furthermore endowed with a compatible norm, which therefore induces a metric DD on the equivalence classes of G\mathcal{R}_{G}. We define a subgroup of [GX][G\curvearrowright X\mkern 1.5mu] that consists of those automorphisms TT for which the map xD(x,Tx)x\mapsto D(x,Tx) is integrable. Such a subgroup, we argue in this section, also carries a natural Polish topology.

Definition 2.5.

Let GXG\curvearrowright X be a Borel measure-preserving action of a Polish normed group (G,)(G,\left\lVert\cdot\right\rVert) on a standard probability space XX; let D:G0D:\mathcal{R}_{G}\to\mathbb{R}^{\geq 0} be the associated metric on the orbits of the action. The L1\mathrm{L}^{1} norm of an automorphism T[GX]T\in[G\curvearrowright X\mkern 1.5mu] is denoted by T1\left\lVert T\right\rVert_{1} and is defined by the integral T1=XD(x,Tx)𝑑μ(x)\left\lVert T\right\rVert_{1}=\int_{X}D(x,Tx)\,d\mu(x). In general, many elements of the full group will have an infinite norm, and the L1\mathrm{L}^{1} full group of the action consists of the automorphisms for which the norm is finite: [GX]1={T[GX]:T1<}[G\curvearrowright X\mkern 1.5mu]_{1}=\{T\in[G\curvearrowright X\mkern 1.5mu]:\left\lVert T\right\rVert_{1}<\infty\}.

Elements of [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} form a group under the composition, as can readily be verified using the triangle inequality for DD and the fact that transformations are measure-preserving. Likewise, it is straightforward to check that \left\lVert\cdot\right\rVert is indeed a norm on [GX]1[G\curvearrowright X\mkern 1.5mu]_{1}. Our goal is to prove that the topology of the norm 1\left\lVert\cdot\right\rVert_{1} on [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is a Polish topology. Mimicking the approach taken in [CLM16], we provide a different definition of the L1\mathrm{L}^{1} full group, where Polishness of the topology will be readily obtainable, and then argue that the two constructions are isometrically isomorphic.

Remark 2.6.

The notion of L1\mathrm{L}^{1} full groups, discussed here, encompasses full groups from [CLM16], since the latter corresponds to the case when GG is equipped with a compatible bounded norm.

We recall some basic facts from [CLM16]. L0(X,G)\mathrm{L}^{0}(X,G) denotes the space of measurable functions f:XGf:X\to G; this space is Polish with respect to the topology of convergence in measure. One can endow the XX with a Polish topology such that the evaluation map Φ:L0(X,G)L0(X,X)\Phi:\mathrm{L}^{0}(X,G)\to\mathrm{L}^{0}(X,X), given by Φ(f)(x)=f(x)x\Phi(f)(x)=f(x)\cdot x, becomes continuous.

Remark 2.7.

In [CLM16], the possibility of making Φ\Phi continuous is obtained by appealing to the remarkable but difficult result of H. Becker and A. S. Kechris, which states that every Borel GG-action has a continuous model [BK96, Thm. 5.2.1]. Let us point out that one can also derive this from the easier fact that every Borel GG-action can be Borel embedded into a continuous GG-action on a Polish compact space (see, for instance, [BK96, Thm. 2.6.6]), as we can endow the latter with the push-forward measure and work with it instead.

Let the set PFL0(X,G)\mathrm{PF}\subseteq\mathrm{L}^{0}(X,G) be the preimage of Aut(X,μ)\mathrm{Aut}(X,\mu) under Φ\Phi:

PF={fL0(X,G):Φ(f)Aut(X,μ)}.\mathrm{PF}=\bigl{\{}f\in\mathrm{L}^{0}(X,G):\Phi(f)\in\mathrm{Aut}(X,\mu)\bigr{\}}.

Since Aut(X,μ)\mathrm{Aut}(X,\mu) is a GδG_{\delta} subset of L0(X,X)\mathrm{L}^{0}(X,X) (see [CLM16, Prop. 2.9] and the remark after it), PF\mathrm{PF} is GδG_{\delta} in L0(X,G)\mathrm{L}^{0}(X,G), hence Polish in the induced topology. The group operations can be pulled from Aut(X,μ)\mathrm{Aut}(X,\mu) onto PF\mathrm{PF} (cf. [CLM16, p. 91]) as follows: for f,gPFf,g\in\mathrm{PF} and xXx\in X define the multiplication via (fg)(x)=f(Φ(g)(x))g(x)(f*g)(x)=f(\Phi(g)(x))g(x) and the inverse111The symbol f1f^{-1} has already been used in the definition of the pointwise inverse on all of L1(X,G)\mathrm{L}^{1}(X,G). We introduce a different operation here, hence the slightly unusual choice of the symbol to denote the inverse operation. by inv(f)(x)=f(Φ(f)1(x))1\mathrm{inv}(f)(x)=f(\Phi(f)^{-1}(x))^{-1}. These operations turn PF\mathrm{PF} into a Polish group and Φ:PFAut(X,μ)\Phi:\mathrm{PF}\to\mathrm{Aut}(X,\mu) into a continuous homomorphism.

The space L1(X,G)\mathrm{L}^{1}(X,G) admits a natural inclusion ι:L1(X,G)L0(X,G)\iota:\mathrm{L}^{1}(X,G)\hookrightarrow\mathrm{L}^{0}(X,G), which is continuous, as can be seen by noting that the equivalent metric dG=min{1,dG}d^{\prime}_{G}=\min\{1,d_{G}\} on GG generates the convergence in measure topology on L0(X,G)\mathrm{L}^{0}(X,G) (see [CLM16, Prop. 2.7]), and d~G(f,g)d~G(f,g)\tilde{d}_{G}(f,g)\geq\tilde{d}^{\prime}_{G}(f,g) for all f,gL1(X,G)f,g\in\mathrm{L}^{1}(X,G). Set PF1=ι1(PF)\mathrm{PF}^{1}=\iota^{-1}(\mathrm{PF}), which we endow with the topology induced form L1(X,G)\mathrm{L}^{1}(X,G). Since L1(X,G)\mathrm{L}^{1}(X,G) is a subset of L0(X,G)\mathrm{L}^{0}(X,G), we may omit the inclusion map ι\iota when convenient.

Proposition 2.8.

PF1\mathrm{PF}^{1} is a Polish group with the multiplication (f,g)(fg)(f,g)\mapsto(f*g) and the inverse finv(f)f\mapsto\mathrm{inv}(f). The function ff1L1(X,G)f\mapsto\left\lVert f\right\rVert_{1}^{\mathrm{L}^{1}(X,G)} is a compatible group norm on PF1\mathrm{PF}^{1} and ΦιPF1:PF1Aut(X,μ)\Phi\circ\iota\restriction_{\mathrm{PF}^{1}}:\mathrm{PF}^{1}\to\mathrm{Aut}(X,\mu) is a continuous homomorphism.

Proof.

First of all, we need to show that these operations are well-defined in the sense that functions fgf*g and inv(f)\mathrm{inv}(f) belong to L1(X,G)\mathrm{L}^{1}(X,G) whenever so do their arguments. To this end observe that for f,gPF1f,g\in\mathrm{PF}^{1}

fg1L1(X,G)\displaystyle\lVert f*g\rVert_{1}^{\mathrm{L}^{1}(X,G)} =Xf(Φ(g)(x))g(x)𝑑μ(x)\displaystyle=\int_{X}\left\lVert f(\Phi(g)(x))g(x)\right\rVert\,d\mu(x)
Xf(Φ(g)(x))𝑑μ(x)+Xg(x)𝑑μ(x).\displaystyle\leq\int_{X}\left\lVert f(\Phi(g)(x))\right\rVert\,d\mu(x)+\int_{X}\left\lVert g(x)\right\rVert\,d\mu(x).

Now note that since Φ(g)\Phi(g) is measure-preserving, we have

Xf(Φ(g)(x))𝑑μ(x)=Xf(x)𝑑μ(x),\displaystyle\int_{X}\left\lVert f(\Phi(g)(x))\right\rVert\,d\mu(x)=\int_{X}\left\lVert f(x)\right\rVert\,d\mu(x),

and therefore

fg1L1(X,G)Xf(x)𝑑μ(x)+Xg(x)𝑑μ(x)=f1L1(X,G)+g1L1(X,G).\left\lVert f*g\right\rVert_{1}^{\mathrm{L}^{1}(X,G)}\leq\int_{X}\left\lVert f(x)\right\rVert\,d\mu(x)+\int_{X}\left\lVert g(x)\right\rVert\,d\mu(x)=\left\lVert f\right\rVert_{1}^{\mathrm{L}^{1}(X,G)}+\left\lVert g\right\rVert_{1}^{\mathrm{L}^{1}(X,G)}.

In particular, fgL1(X,G)f*g\in\mathrm{L}^{1}(X,G), and thus PF1\mathrm{PF}^{1} is closed under the multiplication. Similarly, Φ(f)Aut(X,μ)\Phi(f)\in\mathrm{Aut}(X,\mu) implies

inv(f)1L1(X,G)\displaystyle\left\lVert\mathrm{inv}(f)\right\rVert^{\mathrm{L}^{1}(X,G)}_{1} =Xf(Φ(f)1(x))1𝑑μ(x)\displaystyle=\int_{X}\lVert f(\Phi(f)^{-1}(x))^{-1}\rVert\,d\mu(x)
=Xf(x)1𝑑μ(x)=f1L1(X,G).\displaystyle=\int_{X}\lVert f(x)^{-1}\rVert\,d\mu(x)=\left\lVert f\right\rVert^{\mathrm{L}^{1}(X,G)}_{1}.

Thus PF1\mathrm{PF}^{1} is also closed under taking inverses. Since these operations define a group structure on PF\mathrm{PF}, it follows that PF1\mathrm{PF}^{1} is an (abstract) subgroup of PF\mathrm{PF}. Note that we have also established that L1(X,G)\left\lVert\cdot\right\rVert^{\mathrm{L}^{1}(X,G)} is a group norm on PF1\mathrm{PF}^{1}. The multiplication and the operation of taking the inverse are continuous in the topology of L1(X,G)\mathrm{L}^{1}(X,G), which is a consequence of the continuity of Φι\Phi\circ\iota coupled with Propositions 2.3 and 2.4. Since PF1\mathrm{PF}^{1} is a GδG_{\delta} subset of L1(X,G)\mathrm{L}^{1}(X,G), we conclude that it is a Polish group in the topology induced by the norm 1L1(X,G)\lVert\cdot\rVert^{\mathrm{L}^{1}(X,G)}_{1}. ∎

Let KPF1K\trianglelefteq\mathrm{PF}^{1} denote the kernel of ΦιPF1\Phi\circ\iota\restriction_{\mathrm{PF}^{1}}, and let 1PF1/K\left\lVert\cdot\right\rVert^{\mathrm{PF}^{1}/K}_{1} denote the quotient norm induced by 1L1(X,G)\left\lVert\cdot\right\rVert^{\mathrm{L}^{1}(X,G)}_{1} (see Proposition A.3 regarding the properties of the quotient norm). The factor group (PF1/K,1PF1/K)(\mathrm{PF}^{1}/K,\left\lVert\cdot\right\rVert^{\mathrm{PF}^{1}/K}_{1}) is evidently a Polish normed group, and it turns out to be isometrically isomorphic to the L1\mathrm{L}^{1} full group introduced in Definition 2.5 as we will now see. Let Φ~:PF1/KAut(X,μ)\tilde{\Phi}:\mathrm{PF}^{1}/K\to\mathrm{Aut}(X,\mu) denote the homomorphism induced by ΦιPF1\Phi\circ\iota\restriction_{\mathrm{PF}^{1}} onto the factor group.

Proposition 2.9.

The homomorphism Φ~:PF1/KAut(X,μ)\tilde{\Phi}:\mathrm{PF}^{1}/K\to\mathrm{Aut}(X,\mu) establishes an isometric isomorphism between (PF1/K,1PF1/K)(\mathrm{PF}^{1}/K,\left\lVert\cdot\right\rVert^{\mathrm{PF}^{1}/K}_{1}) and ([GX]1,1)([G\curvearrowright X\mkern 1.5mu]_{1},\left\lVert\cdot\right\rVert_{1}).

Proof.

We begin by showing that gK1PF/K=Φ~(gK)1\lVert gK\rVert^{\mathrm{PF}/K}_{1}=\lVert\tilde{\Phi}(gK)\rVert_{1} holds for any gKPF1/KgK\in\mathrm{PF}^{1}/K. By the definition of the quotient norm,

gK1PF1/K=infkKXg(x)k(x)𝑑μ(x).\left\lVert gK\right\rVert^{\mathrm{PF^{1}/K}}_{1}=\inf_{k\in K}\int_{X}\left\lVert g(x)k(x)\right\rVert\,d\mu(x).

For any fixed kKk\in K, we have g(x)k(x)x=g(x)xg(x)k(x)\cdot x=g(x)\cdot x, and therefore

D(x,g(x)x)g(x)k(x)for almost every xX.D(x,g(x)\cdot x)\leq\left\lVert g(x)k(x)\right\rVert\quad\textrm{for almost every $x\in X$}.

This readily implies the inequality Φ~(gK)1gK1PF1/K\lVert\tilde{\Phi}(gK)\rVert_{1}\leq\left\lVert gK\right\rVert^{\mathrm{PF}^{1}/K}_{1}. For the other direction, let ϵ>0\epsilon>0 and consider the set

{(x,u)X×G:g(x)x=ux and uD(x,g(x)x)+ϵ}.\{(x,u)\in X\times G:g(x)\cdot x=u\cdot x\textrm{ and }\left\lVert u\right\rVert\leq D(x,g(x)\cdot x)+\epsilon\}.

Using Jankov-von Neumann uniformization theorem, one may pick a measurable map g0:XGg_{0}:X\to G that satisfies g0(x)x=g(x)xg_{0}(x)\cdot x=g(x)\cdot x and g0(x)D(x,g(x)x)+ϵ\left\lVert g_{0}(x)\right\rVert\leq D(x,g(x)\cdot x)+\epsilon for almost all xXx\in X. Since xg(x)1g0(x)Kx\mapsto g(x)^{-1}g_{0}(x)\in K, we have

Φ~(gK)1\displaystyle\lVert\tilde{\Phi}(gK)\rVert_{1} =XD(x,g(x)x)𝑑μ(x)\displaystyle=\int_{X}D(x,g(x)\cdot x)\,d\mu(x)
Xg(x)g(x)1g0(x)𝑑μ(x)ϵ\displaystyle\geq\int_{X}\left\lVert g(x)g(x)^{-1}g_{0}(x)\right\rVert\,d\mu(x)-\epsilon
gK1PF1/Kϵ.\displaystyle\geq\left\lVert gK\right\rVert_{1}^{\mathrm{PF}^{1}/K}-\epsilon.

As ϵ\epsilon is an arbitrary positive real, we conclude that gK1PF1/K=Φ~(gK)1\left\lVert gK\right\rVert_{1}^{\mathrm{PF}^{1}/K}=\lVert\tilde{\Phi}(gK)\rVert_{1}.

It remains to check that Φ~\tilde{\Phi} is surjective. For an automorphism T[GX]1T\in[G\curvearrowright X\mkern 1.5mu]_{1}, consider the set

{(x,u)X×G:Tx=ux and uD(x,Tx)+1}.\{(x,u)\in X\times G:Tx=u\cdot x\textrm{ and }\left\lVert u\right\rVert\leq D(x,Tx)+1\}.

Applying the Jankov-von Neumann uniformization theorem once again we get a map gL0(X,G)g\in\mathrm{L}^{0}(X,G) such that Φ(g)=T\Phi(g)=T and g(x)D(x,Tx)+1\left\lVert g(x)\right\rVert\leq D(x,Tx)+1. The latter inequality together with the assumption that T[GX]1T\in[G\curvearrowright X\mkern 1.5mu]_{1} easily imply that gL1(X,G)g\in\mathrm{L}^{1}(X,G) and thus gKPF1/KgK\in\mathrm{PF}^{1}/K is the preimage of TT under Φ~\tilde{\Phi}. ∎

Results discussed thus far can be summarized as follows.

Theorem 2.10.

Let GXG\curvearrowright X be a Borel measure-preserving action of a Polish normed group (G,)(G,\left\lVert\cdot\right\rVert) on a standard probability space. The L1\mathrm{L}^{1} full group [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is a Polish normed group relative to the norm T1=XD(x,Tx)𝑑μ(x)\left\lVert T\right\rVert_{1}=\int_{X}D(x,Tx)\,d\mu(x).

Remark 2.11.

When the acting group is finitely generated and equipped with the word length metric with respect to the finite generating set, it can be shown that the left-invariant metric induced by the norm on the L1\mathrm{L}^{1} full group is complete (see [LM18, Prop. 3.4 and 3.5] and the remark thereafter for a more general statement). Nevertheless, generally L1\mathrm{L}^{1} full groups do not admit compatible complete left-invariant metrics, i.e., they are not necessarily CLI groups. For instance, if G=G=\mathbb{R} is acting by rotation on the circle, the L1\mathrm{L}^{1} full group of the action is all of Aut(𝕊1,λ)\mathrm{Aut}(\mathbb{S}^{1},\lambda), which is not CLI.

Let us point out a possibility to generalize our framework. Given a standard probability space (X,μ)(X,\mu), consider an extended Borel metric DD on XX, i.e., a Borel metric that is allowed to take the value ++\infty (Eq. (2.1) provides such an example). Note that the relation D(x,y)<+D(x,y)<+\infty is an equivalence relation. One can now define the L1\mathrm{L}^{1} full group of DD in complete analogy with Definition 2.5 as the group of all TAut(X,μ)T\in\mathrm{Aut}(X,\mu) whose norm TD=XD(x,T(x))𝑑μ(x)\left\lVert T\right\rVert_{D}=\int_{X}D(x,T(x))\,d\mu(x) is finite.

Question 2.12.

Suppose that DD restricts to a complete separable metric on each equivalence class {yX:D(x,y)<+}\{y\in X:D(x,y)<+\infty\}, xXx\in X. Is the L1\mathrm{L}^{1} full group of DD Polish in the topology of the norm D\left\lVert\cdot\right\rVert_{D}?

2.3. L1\mathrm{L}^{1} full groups and quasi-metric structures

When viewed as a normed group, the L1\mathrm{L}^{1} full group [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} depends on the choice of a compatible norm on GG. The topological structure on [GX]1[G\curvearrowright X\mkern 1.5mu]_{1}, however, depends only on the quasi-metric structure of the acting group. Recall that two norms \lVert\cdot\rVert and \lVert\cdot\rVert^{\prime} on a Polish group GG are quasi-isometric if there exists a constant C>0C>0 such that for all gGg\in G,

1CgCgCg+C.\frac{1}{C}\lVert g\rVert-C\leq\lVert g\rVert^{\prime}\leq C\lVert g\rVert+C.
Lemma 2.13.

Let \lVert\cdot\rVert and \lVert\cdot\rVert^{\prime} be two quasi-isometric compatible norms on a Polish group GG, and let G(X,μ)G\curvearrowright(X,\mu) be a Borel measure-preserving action on a standard probability space. The L1\mathrm{L}^{1} full groups associated with the two norms are equal as topological groups.

Proof.

The quasi-isometry condition implies that a function f:XGf:X\to G satisfies Xf(x)𝑑μ(x)<+\int_{X}\lVert f(x)\rVert\,d\mu(x)<+\infty if and only if Xf(x)𝑑μ(x)<+\int_{X}\lVert f(x)\rVert^{\prime}\,d\mu(x)<+\infty. In particular, the L1\mathrm{L}^{1} full groups associated with these norms are equal as abstract groups.

Both topologies make the inclusion of [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} into Aut(X,μ)\mathrm{Aut}(X,\mu) continuous by Proposition 2.8, and, in particular, the inclusion map is Borel. Since injective images of Borel sets by Borel maps are Borel (see, for example, [Kec95, Thm. 15.1]), it follows that both topologies induce the same Borel structure on [GX]1[G\curvearrowright X\mkern 1.5mu]_{1}, which also coincides with the one induced by the weak topology on Aut(X,μ)\mathrm{Aut}(X,\mu). A standard automatic continuity result (originally due to S. Banach [Ban32, Thm. 4 p. 23]) then yields equality of the two topologies (see also the second paragraph following [BK96, Lem. 1.2.6]). ∎

When a Polish group GG admits a canonical choice of the quasi-metric structure, L1\mathrm{L}^{1} full groups [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} are unambiguously defined as topological groups without the need to choose any particular norm on GG. This is the case for boundedly generated Polish groups—the class of groups identified and studied by C. Rosendal in his treatise [Ros22]. Appendix A.2 provides a succinct review of the concept of maximal norms on boundedly generated Polish groups.

An example of this situation is given by G=G=\mathbb{R}, where the usual Euclidean norm is maximal in the sense of Definition A.5.

Remark 2.14.

We will see in the last chapter that the natural L1\mathrm{L}^{1} norm on the L1\mathrm{L}^{1} full groups of \mathbb{R}-actions is maximal so that it defines a quasi-metric structure which is canonically associated with the topological group structure.

2.4. Embedding L1\mathrm{L}^{1} isometrically in L1\mathrm{L}^{1} full groups

We now show a general result on the geometry of L1\mathrm{L}^{1} full groups endowed with the L1\mathrm{L}^{1} norm 1\left\lVert\cdot\right\rVert_{1}, which says that they are quite big.

Given a σ\sigma-finite measured space (X,,λ)(X,\mathcal{B},\lambda), denote by MAlgf(X,λ)\mathrm{MAlg}_{f}(X,\lambda) the space of all finite measure subsets BB\in\mathcal{B} identified up to measure zero and endowed with the metric dλ(B1,B2)=λ(B1B2)d_{\lambda}(B_{1},B_{2})=\lambda(B_{1}\bigtriangleup B_{2}).

Proposition 2.15.

Let (G,)(G,\left\lVert\cdot\right\rVert) be a Polish normed group acting by measure-preserving transformations on a standard probability space (X,μ)(X,\mu). If

[GX]1[GX],[G\curvearrowright X\mkern 1.5mu]_{1}\neq[G\curvearrowright X],

then the metric space (MAlgf(,λ),dλ)(\mathrm{MAlg}_{f}(\mathbb{R},\lambda),d_{\lambda}) embeds isometrically into the L1\mathrm{L}^{1} full group of GXG\curvearrowright X endowed with its L1\mathrm{L}^{1} metric, and hence so does L1(X,μ,)\mathrm{L}^{1}(X,\mu,\mathbb{R}).

Proof.

Since [GX][G\curvearrowright X] is a full group, any of its elements can be written as a product of three involutions belonging to [GX][G\curvearrowright X] by [Ryz85]. By assumption, [GX]1[GX][G\curvearrowright X\mkern 1.5mu]_{1}\neq[G\curvearrowright X] so there must be an involution U[GX]U\in[G\curvearrowright X] which does not belong to [GX]1[G\curvearrowright X\mkern 1.5mu]_{1}. Denote by U\mathcal{B}_{U} the σ\sigma-algebra on suppU\operatorname*{supp}U consisting of UU-invariant sets, endowed with the measure given by λU(A)=UA1\lambda_{U}(A)=\left\lVert U_{A}\right\rVert_{1}. Since suppU=n{xsuppU:D(x,U(x))n}\operatorname*{supp}U=\bigcup_{n}\{x\in\operatorname*{supp}U:D(x,U(x))\leq n\}, the measure λU\lambda_{U} is σ\sigma-finite. Also, λU\lambda_{U} is non-atomic, because so is μ\mu, and infinite, because U[GX]1U\not\in[G\curvearrowright X\mkern 1.5mu]_{1}. There is only one σ\sigma-finite standard atomless infinite measured space up to isomorphism (namely (,(),λ)(\mathbb{R},\mathcal{B}(\mathbb{R}),\lambda)) so we conclude that (MAlgf(suppU,λU),dλU)(\mathrm{MAlg}_{f}(\operatorname*{supp}U,\lambda_{U}),d_{\lambda_{U}}) is isometric to (MAlgf(,λ),dλ)(\mathrm{MAlg}_{f}(\mathbb{R},\lambda),d_{\lambda}). Composing this isometry with AUAA\mapsto U_{A}, we get the desired isometric embedding (MAlgf(,λ),dλ)[GX]1(\mathrm{MAlg}_{f}(\mathbb{R},\lambda),d_{\lambda})\to[G\curvearrowright X\mkern 1.5mu]_{1}.

Finally, we observe that L1(X,μ,)\mathrm{L}^{1}(X,\mu,\mathbb{R}) can be embedded into MAlgf(X×,μλ)\mathrm{MAlg}_{f}(X\times\mathbb{R},\mu\otimes\lambda) by taking a function ff to its epigraph, namely the set of all (x,y)X×(x,y)\in X\times\mathbb{R} such that f(x)y0f(x)\leq y\leq 0 or 0yf(x)0\leq y\leq f(x). Since there is again only one infinite σ\sigma-finite standard atomless measured space and (X×,μλ)(X\times\mathbb{R},\mu\otimes\lambda) is such a space, we get an isometric embedding L1(X,μ,)MAlgf(,λ)\mathrm{L}^{1}(X,\mu,\mathbb{R})\to\mathrm{MAlg}_{f}(\mathbb{R},\lambda) as wanted. ∎

Remark 2.16.

Full groups of actions of Polish groups are always coarsely bounded. In fact, they are coarsely bounded even as discrete groups222Being coarsely bounded as a discrete group is also called the Bergman property., which is a result due to M. Droste, W. C. Holland and G. Ulbrich [DHU08] (see also [Mil04, Section I.8] for a more general statement which encompasses the non-ergodic case). In particular, the above result is actually a sharp dichotomy: every L1\mathrm{L}^{1} full group of a Polish normed group action is either coarsely bounded, or it contains an isometric copy of L1(X,μ,)\mathrm{L}^{1}(X,\mu,\mathbb{R}).

Remark 2.17.

Since n\mathbb{R}^{n} endowed with the 1\ell^{1} norm embeds isometrically into L1(X,μ,)\mathrm{L}^{1}(X,\mu,\mathbb{R}), Proposition 2.15 significantly improves [LM21, Prop. 6.9].

2.5. Stability under the first return map

Some of the basic properties of L1\mathrm{L}^{1} full groups are discussed—in the wider generality of induction friendly finitely full groups—in Chapter 3. The often-used fundamental fact is the closure of L1\mathrm{L}^{1} full groups under taking the induced maps, which is a generalization of [LM18, Prop. 3.6]. We formulate this in Proposition 2.18.

Let TAut(X,μ)T\in\mathrm{Aut}(X,\mu) be a measure-preserving transformation. Recall that for a measurable subset AXA\subseteq X, the induced map TAT_{A} is supported on AA and is defined to be Tn(x)T^{n}(x) for xAx\in A where n1n\geq 1 is the smallest integer such that Tn(x)AT^{n}(x)\in A. By the Poincaré recurrence theorem, such a map yields a well-defined measure-preserving transformation.

Proposition 2.18.

Let GXG\curvearrowright X be a Borel measure-preserving action of a Polish normed group (G,)(G,\left\lVert\cdot\right\rVert). For any element T[GX]1T\in[G\curvearrowright X\mkern 1.5mu]_{1} and any measurable set AXA\subseteq X, the induced transformation TAT_{A} belongs to [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} and moreover TA1T1.\left\lVert T_{A}\right\rVert_{1}\leq\left\lVert T\right\rVert_{1}.

Proof.

For n1n\geq 1, let AnA_{n} be the set of elements of AA whose return time is equal to nn; note that X=n1i=0n1Ti(An)X=\bigsqcup_{n\geq 1}\bigsqcup_{i=0}^{n-1}T^{i}(A_{n}). Let as before D:G0D:\mathcal{R}_{G}\to\mathbb{R}^{\geq 0} be the metric induced by the group norm \left\lVert\cdot\right\rVert on the orbits of the action. To estimate the value of TA1\left\lVert T_{A}\right\rVert_{1}, observe that

TA1\displaystyle\left\lVert T_{A}\right\rVert_{1} =XD(x,TAx)𝑑μ(x)=n=1AnD(x,TAx)𝑑μ(x)\displaystyle=\int_{X}D(x,T_{A}x)\,d\mu(x)=\sum_{n=1}^{\infty}\int_{A_{n}}\mkern-10.0muD(x,T_{A}x)\,d\mu(x)
=n=1AnD(x,Tnx)𝑑μ(x).\displaystyle=\sum_{n=1}^{\infty}\int_{A_{n}}\mkern-10.0muD(x,T^{n}x)\,d\mu(x).

Using the triangle inequality, we get

TA1\displaystyle\left\lVert T_{A}\right\rVert_{1} n=1i=0n1AnD(Tix,Ti+1x)𝑑μ(x)\displaystyle\leq\sum_{n=1}^{\infty}\sum_{i=0}^{n-1}\int_{A_{n}}\mkern-10.0muD(T^{i}x,T^{i+1}x)\,d\mu(x)
=n=1i=0n1Ti(An)D(x,Tx)d(μTi)(x)\displaystyle=\sum_{n=1}^{\infty}\sum_{i=0}^{n-1}\int_{T^{i}(A_{n})}\mkern-35.0muD(x,Tx)\,d(\mu\circ T^{-i})(x)
T preserves μ\displaystyle\because\ T\textrm{ preserves }\mu =n=1i=0n1Ti(An)D(x,Tx)𝑑μ(x)=XD(x,Tx)𝑑μ(x)=T1.\displaystyle=\sum_{n=1}^{\infty}\sum_{i=0}^{n-1}\int_{T^{i}(A_{n})}\mkern-35.0muD(x,Tx)\,d\mu(x)=\int_{X}D(x,Tx)\,d\mu(x)=\left\lVert T\right\rVert_{1}.

Thus TA[GX]1T_{A}\in[G\curvearrowright X\mkern 1.5mu]_{1} and TA1T1\left\lVert T_{A}\right\rVert_{1}\leq\left\lVert T\right\rVert_{1} as claimed. ∎

Chapter 3 Polish finitely full groups

The main object of our investigation in this work are L1\mathrm{L}^{1} full groups of Borel measure-preserving actions of Polish normed groups. Some results, however, are valid in the more general context of what we call Polish finitely full groups. It encompasses L1\mathrm{L}^{1} full groups and allows us to put some of the proofs on topological simplicity and on maximal norms from [LM18, LM21] in a unified and broadened context.

Starting with a Polish finitely full group as defined in Section 3.1, we construct in Section 3.2 a natural closed subgroup of the latter which we call the symmetric subgroup, analogous to V. Nekrashevych’s symmetric and alternating topological full groups [Nek19]. We show that this closed subgroup coincides with the closure of the derived group under a mild hypothesis, satisfied by L1\mathrm{L}^{1} full groups, which we call induction friendliness. Section 3.3 is devoted to the study of closed normal subgroups of the symmetric subgroup: we show that they correspond to invariant sets, a fact which easily yields topological simplicity when the ambient Polish finitely full group is ergodic. Finally, in Section 3.4 we provide a condition normed induction friendly Polish finitely full groups which guarantees maximality on the symmetric subgroup in the sense of C. Rosendal (a brief reminder of the relevant notions is given in Appendix A.2).

3.1. Polish full and finitely full groups

H. Dye defined a subgroup 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) as being full when it is stable under the cutting and pasting of its elements along a countable partition: given any partition (An)n(A_{n})_{n} of XX and any sequence (gn)n(g_{n})_{n} such that the family (gn(An))n(g_{n}(A_{n}))_{n} also partitions XX, the element TAut(X,μ)T\in\mathrm{Aut}(X,\mu) obtained as the reunion over nn\in\mathbb{N} of the restrictions gnAng_{n}\restriction_{A_{n}} belongs to 𝔾\mathbb{G}. In particular, the group Aut(X,μ)\mathrm{Aut}(X,\mu) itself is full.

Given any 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu), the group obtained by cutting and pasting elements of 𝔾\mathbb{G} along countable partitions is the smallest full subgroup containing 𝔾\mathbb{G}. We denote it by [𝔾][\mathbb{G}] and call it the full group generated by 𝔾\mathbb{G}.

Recall that the uniform topology on Aut(X,μ)\mathrm{Aut}(X,\mu) is the topology induced by the uniform metric dud_{u} defined by

du(T1,T2)=μ({xX:T1xT2x}).d_{u}(T_{1},T_{2})=\mu(\{x\in X:T_{1}x\neq T_{2}x\}).

The following can essentially be traced back to H. Dye [Dye59, Lem. 5.4].

Proposition 3.1.

The metric dud_{u} is complete on any full group 𝔾\mathbb{G}, and it is separable if and only if the full group is generated by a countable group.

Proof.

Suppose that (Tn)n(T_{n})_{n} is a Cauchy sequence in the full group 𝔾\mathbb{G}. Taking a subsequence, we may assume that du(Tn,Tn+1)<2nd_{u}(T_{n},T_{n+1})<2^{-n} for all nn. By the Borel-Cantelli lemma, for almost every xXx\in X there is some NN\in\mathbb{N} such that Tnx=TNxT_{n}x=T_{N}x for all nNn\geq N. Let Tx=TNxTx=T_{N}x for such N=N(x)N=N(x), and note that TT is a measure-preserving bijection111 This also follows from the fact due to P. Halmos [Hal17] that Aut(X,μ)\mathrm{Aut}(X,\mu) is dud_{u}-complete. and du(Tn,T)2n+1d_{u}(T_{n},T)\leq 2^{-n+1}. By construction, TT is obtained by cutting and pasting the elements TnT_{n} of 𝔾\mathbb{G} along a countable partition so T𝔾T\in\mathbb{G}, since 𝔾\mathbb{G} is full.

Suppose 𝔾\mathbb{G} is separable and let Γ\Gamma be a countable dense subgroup. The group [Γ][\Gamma] is a countably generated full group which is dense in 𝔾\mathbb{G}, so 𝔾=[Γ]\mathbb{G}=[\Gamma] by completeness. The converse is obtained by noting that if Γ\Gamma generates 𝔾\mathbb{G}, then one can view 𝔾\mathbb{G} as the full group of the equivalence relation generated by a realization of the action of Γ\Gamma on (X,μ)(X,\mu), which is dud_{u}-separable by [Kec10, Prop. 3.2]. ∎

The L1\mathrm{L}^{1} full groups that we are considering are not full in the sense of H. Dye unless the norm on the acting Polish group is bounded, a case which was considered earlier in [CLM16]. They nevertheless satisfy the following weaker property.

Definition 3.2.

A group 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) of measure-preserving transformations is finitely full if for any partition X=A1AnX=A_{1}\sqcup\cdots\sqcup A_{n} and g1,,gn𝔾g_{1},\ldots,g_{n}\in\mathbb{G} such that the sets g1A1,,gnAng_{1}A_{1},\ldots,g_{n}A_{n} also partition XX, the element TAut(X,μ)T\in\mathrm{Aut}(X,\mu) obtained as the reunion over i{1,,n}i\in\{1,\ldots,n\} of the restrictions giAig_{i}\restriction_{A_{i}} belongs to 𝔾\mathbb{G}.

We have the following useful relationship between fullness and finite fullness.

Proposition 3.3.

The dud_{u}-closure of any finitely full group 𝔾\mathbb{G} is equal to the full group [𝔾][\mathbb{G}] generated by 𝔾\mathbb{G}. Moreover, every element T[𝔾]T\in[\mathbb{G}] is a limit of elements of 𝔾\mathbb{G} whose support is contained in the support of TT.

Proof.

Since full groups are dud_{u}-closed and using the definition of fullness, it suffices to show that every element T[𝔾]T\in[\mathbb{G}] is a limit of elements of 𝔾\mathbb{G} that belong to the full group generated by TT.

Since every T[𝔾]T\in[\mathbb{G}] is a product of three involutions in [T][T]222In fact, we only need the much easier fact that every element is a limit of products of two involutions from its full group, which follows by combining Theorem 3.3 and Sublemma 4.3 from [Kec10]. [Ryz85], it suffices to show that every involution in [𝔾][\mathbb{G}] is a limit of elements of 𝔾\mathbb{G} whose support is contained in the support of that involution. Let UU be such an involution, let (An)n(A_{n})_{n} be a partition of XX, and let (gn)n(g_{n})_{n} in 𝔾\mathbb{G} be such that Ux=gnxUx=g_{n}x for all xAnx\in A_{n}. Pick a fundamental domain BB for UU, i.e., BU(B)=B\cap U(B)=\varnothing and suppU=BU(B)\operatorname*{supp}U=B\cup U(B). If Bn=AnBB_{n}=A_{n}\cap B, then Ux=gnxUx=g_{n}x for all xBnx\in B_{n}, and, since UU is an involution, Ux=gn1xUx=g_{n}^{-1}x for all xU(Bn)x\in U(B_{n}). Let

Unx={Ux if xmn(BmU(Bm)),x otherwise.U_{n}x=\left\{\begin{array}[]{cl}Ux&\text{ if }x\in\bigcup_{m\leq n}\left(B_{m}\cup U(B_{m})\right),\\ x&\text{ otherwise.}\end{array}\right.

Clearly Un𝔾U_{n}\in\mathbb{G}, since 𝔾\mathbb{G} is finitely full. Furthermore, UnUU_{n}\to U uniformly and suppUnsuppU\operatorname*{supp}U_{n}\subseteq\operatorname*{supp}U by construction, which finishes the proof. ∎

Consider a finitely full group 𝔾\mathbb{G} which is a Borel subset of Aut(X,μ)\mathrm{Aut}(X,\mu) and therefore inherits the structure of a standard Borel space. If 𝔾\mathbb{G} is Polishable, i.e., if it admits a Polish group topology compatible with the Borel structure, then such topology is necessarily unique and must refine the weak topology inherited from Aut(X,μ)\mathrm{Aut}(X,\mu) (standard automatic continuity results can be found, for instance, in [BK96, Sec. 1.6]). We refer to such Polishable groups 𝔾\mathbb{G} endowed with their unique Polish group topology refining the weak topology as Polish finitely full groups. In this monograph, our motivating example for introducing this class is of course L1\mathrm{L}^{1} full groups.

For any subgroup GAut(X,μ)G\leq\mathrm{Aut}(X,\mu), there is the smallest finitely full group containing GG. Note that if HAut(X,μ)H\leq\mathrm{Aut}(X,\mu) is a finite group, then the finitely full group it generates coincides with the full group it generates. This, in particular, applies to the group generated by a periodic transformation with bounded periods.

Proposition 3.4.

Suppose 𝔾\mathbb{G} is a Polish finitely full group, and U𝔾U\in\mathbb{G} is a periodic transformation with bounded periods. The topology induced by 𝔾\mathbb{G} on the full group of UU is equal to the uniform topology.

Proof.

The weak and the uniform topologies on [U][U\mkern 1.5mu] coincide since UU is periodic. We already mentioned that the topology of 𝔾\mathbb{G} refines the weak topology. Since [U][U\mkern 1.5mu] is Polish in the uniform topology, by the automatic continuity [BK96, Thm. 1.2.6], the topology induced by 𝔾\mathbb{G} on the full group of UU is refined by the uniform topology. Hence the uniform topology and the topology induced from 𝔾\mathbb{G} onto [U][U\mkern 1.5mu] must coincide. ∎

We conclude this preliminary discussion with a definition of aperiodicity which applies to arbitrary subgroups of Aut(X,μ)\mathrm{Aut}(X,\mu). Such a notion was already worked out by H. Dye [Dye59, Sec. 2] when he introduced type II subgroups. An equivalent version which suffices for our purposes is as follows.

Definition 3.5.

A subgroup GAut(X,μ)G\leq\mathrm{Aut}(X,\mu) is aperiodic it it admits a countable weakly dense subgroup whose action on (X,μ)(X,\mu) has no finite orbits.

It can be checked that for an aperiodic GAut(X,μ)G\leq\mathrm{Aut}(X,\mu), every countable weakly dense subgroup has infinite orbits almost surely. Further discussion of aperiodicity can be found in Appendix D.4.

3.2. Derived subgroup and symmetric subgroup

Our goal in this section is to identify when the closed derived subgroup of a Polish finitely full group is topologically generated by involutions. We start by noting that aperiodic finitely full groups admit many involutions in the sense of [Fre04, p. 384]:

Lemma 3.6.

Let 𝔾\mathbb{G} be a finitely full aperiodic group. For every measurable nontrivial AXA\subseteq X, there is a nontrivial involution g𝔾g\in\mathbb{G} whose support is contained in AA.

Proof.

By Lemma D.13, there is an involution T[𝔾]T\in[\mathbb{G}] whose support is equal to AA. By the moreover part of Proposition 3.3, TT is the dud_{u}-limit of gn𝔾g_{n}\in\mathbb{G} supported in AA. In particular, one of the gng_{n}’s is nontrivial and g=gng=g_{n} satisfies the statement of the lemma. ∎

The first and the second items of the following definition constitute analogues of V. Nekrashevych’s symmetric and alternating topological full groups [Nek19], respectively. In the setup of L1\mathrm{L}^{1} full groups, however, these notions coincide, as we will see shortly.

Definition 3.7.

Given a Polish finitely full group 𝔾\mathbb{G}, we let

  • 𝔖(𝔾)\mathfrak{S}(\mathbb{G}) be the closed subgroup of 𝔾\mathbb{G} generated by involutions, which we call the symmetric subgroup of 𝔾\mathbb{G}.

  • 𝔄(𝔾)\mathfrak{A}(\mathbb{G}) be the closed subgroup of 𝔾\mathbb{G} generated by 33-cycles, i.e., generated by periodic transformations whose non-trivial orbits have size 33.

  • D(𝔾)D(\mathbb{G}) be the closed subgroup generated by commutators (also known as the topological derived subgroup).

All these groups are closed normal subgroups of 𝔾\mathbb{G}, and 𝔄(𝔾)𝔖(𝔾)D(𝔾)\mathfrak{A}(\mathbb{G})\leq\mathfrak{S}(\mathbb{G})\cap D(\mathbb{G}) because every 33-cycle is a commutator of two involutions from its full group.

Proposition 3.8.

𝔄(𝔾)=𝔖(𝔾)\mathfrak{A}(\mathbb{G})=\mathfrak{S}(\mathbb{G}) for any aperiodic finitely full group 𝔾\mathbb{G}.

Proof.

We need to show that every involution is a limit of products of 33-cycles. Let U𝔾U\in\mathbb{G} be an involution, and let DD denote its fundamental domain; thus suppU=DU(D)\operatorname*{supp}U=D\sqcup U(D). By Lemma D.13, one can find an involution V[𝔾]V\in[\mathbb{G}\mkern 1.5mu] whose support is equal to DD. Since 𝔾\mathbb{G} is finitely full, we may write DD as an increasing union D=nDnD=\bigcup_{n}D_{n}, DnDn+1D_{n}\subseteq D_{n+1}, where each DnD_{n} is VV-invariant, and for every nn\in\mathbb{N} the transformation VnV_{n} induced by VV on DnD_{n} belongs to the group 𝔾\mathbb{G} itself. Let UnU_{n} denote the restriction of UU onto DnU(Dn)D_{n}\sqcup U(D_{n}) and note that UnUU_{n}\to U in the uniform topology, and hence also in the topology of 𝔾\mathbb{G} by Proposition 3.4. Our plan is to use the following permutation identity

(3.1) (12)(34)=(12)(23)(24)(23)=(123)(423),(12)(34)=(12)(23)(24)(23)=(123)(423),

where UnU_{n} corresponds to (12)(34)(12)(34), VnV_{n} to (13)(13), and UnVnUnU_{n}V_{n}U_{n} corresponds to (24)(24). To this end, let CnC_{n} be a fundamental domain for VnV_{n}, put Wn=UCnU(Cn)W_{n}=U\restriction_{C_{n}\sqcup U(C_{n})} (which corresponds to the involution (12)(12)), and, at last, set Sn=WnVnWnS_{n}=W_{n}V_{n}W_{n} (corresponding to (23)=(12)(13)(12)(23)=(12)(13)(12)). Figure 3.1 illustrates the relations between these sets and transformations.

1CnC_{n}WnW_{n}2Un(Cn)U_{n}(C_{n})VnV_{n}3Vn(Cn)V_{n}(C_{n})UnVnUnU_{n}V_{n}U_{n}4UnVn(Cn)U_{n}V_{n}(C_{n})SnS_{n}Un(Dn)U_{n}(D_{n})DnD_{n}
Figure 3.1. Involution UnU_{n} is a products of 33-cycles via (12)(34)=(123)(234)(12)(34)=(123)(234).

Equation (3.1) translates into Un=(WnSn)((UnVnUn)Sn)U_{n}=\bigl{(}W_{n}S_{n}\bigr{)}\bigl{(}(U_{n}V_{n}U_{n})S_{n}\bigr{)}, so UnU_{n} is a product of two 33-cycles, hence it belongs to 𝔄(𝔾)\mathfrak{A}(\mathbb{G}). Since by construction UnUU_{n}\to U, we conclude that U𝔄(𝔾)U\in\mathfrak{A}(\mathbb{G}). ∎

We do not know whether 𝔄(𝔾)=D(𝔾)\mathfrak{A}(\mathbb{G})=D(\mathbb{G}) holds for all finitely full groups, but here is a convenient sufficient condition.

Definition 3.9.

A Polish finitely full group 𝔾\mathbb{G} is called induction friendly if it is stable under taking induced transformations and, furthermore, whenever T𝔾T\in\mathbb{G} and (An)n(A_{n})_{n} is an increasing sequence of TT-invariant sets such that nAn=A\bigcup_{n}A_{n}=A, then TAnTAT_{A_{n}}\to T_{A}.

In the above definition, we require stability under taking the induced transformations and so TAnT_{A_{n}} always belongs to 𝔾\mathbb{G}. However, for TT-invariant AnA_{n}, TAn𝔾T_{A_{n}}\in\mathbb{G} is already a consequence of 𝔾\mathbb{G} being finitely full.

Observe that L1\mathrm{L}^{1} full groups of measure-preserving actions of Polish normed groups are finitely full and also induction friendly. Indeed, finite fullness follows from a straightforward computation, while induction friendliness is a direct consequence of Proposition 2.18 and Lebesgue dominated convergence theorem.

Lemma 3.10.

In an induction friendly Polish finitely full group 𝔾\mathbb{G}, every periodic element belongs to 𝔖(𝔾)\mathfrak{S}(\mathbb{G}).

Proof.

Suppose TT is periodic. For nn\in\mathbb{N}, let AnA_{n} be the set of xXx\in X whose TT-orbit has cardinality at most nn. Each AnA_{n} is TT-invariant and nAn=X\bigcup_{n}A_{n}=X. Moreover, TAnT_{A_{n}} is periodic, so it can be written as a product of two involutions from its full group, and since 𝔾\mathbb{G} is finitely full and the periods of TAnT_{A_{n}} are bounded, these two involutions belong to 𝔾\mathbb{G}. The conclusion follows from induction friendliness and convergence TAnTT_{A_{n}}\to T. ∎

Lemma 3.11.

Let 𝔾\mathbb{G} be an induction friendly Polish finitely full group, T𝔾T\in\mathbb{G} and FXF\subseteq X be the aperiodic part of TT, i.e.,

F={xX:Tkxx for all k0}.F=\{x\in X:T^{k}x\neq x\textrm{ for all }k\neq 0\}.

For any AXA\subseteq X such that FkTk(A)F\subseteq\bigcup_{k\in\mathbb{Z}}T^{k}(A) one has TA𝔖(𝔾)=T𝔖(𝔾)T_{A}\mathfrak{S}(\mathbb{G})=T\mathfrak{S}(\mathbb{G}).

Proof.

Since FkTk(A)F\subseteq\bigcup_{k\in\mathbb{Z}}T^{k}(A), the transformation T1TAT^{-1}T_{A} is periodic and therefore belongs to 𝔖(𝔾)\mathfrak{S}(\mathbb{G}) by Lemma 3.10. Hence

T𝔖(𝔾)=TT1TA𝔖(𝔾)=TA𝔖(𝔾).T\mathfrak{S}(\mathbb{G})=TT^{-1}T_{A}\mathfrak{S}(\mathbb{G})=T_{A}\mathfrak{S}(\mathbb{G}).\qed
Remark 3.12.

Usefulness of the above lemma stems from the following simple observation. If T,T,U,UT,T^{\prime},U,U^{\prime} satisfy T𝔖(𝔾)=T𝔖(𝔾)T\mathfrak{S}(\mathbb{G})=T^{\prime}\mathfrak{S}(\mathbb{G}) and U𝔖(𝔾)=U𝔖(𝔾)U\mathfrak{S}(\mathbb{G})=U^{\prime}\mathfrak{S}(\mathbb{G}), then [T,U]𝔖(𝔾)[T,U]\in\mathfrak{S}(\mathbb{G}) if and only if [T,U]𝔖(𝔾)[T^{\prime},U^{\prime}]\in\mathfrak{S}(\mathbb{G}). In particular, for AA as in Lemma 3.11, [T,U]𝔖(𝔾)[T,U]\in\mathfrak{S}(\mathbb{G}) whenever [TA,U]𝔖(𝔾)[T_{A},U]\in\mathfrak{S}(\mathbb{G}). This fact is used in the proof of the next lemma.

Lemma 3.13.

Suppose 𝔾\mathbb{G} is an induction friendly Polish finitely full group. If T,U𝔾T,U\in\mathbb{G} are aperiodic on their supports, then [T,U]𝔖(𝔾)[T,U]\in\mathfrak{S}(\mathbb{G}).

Proof.

Let CC be a cross-section for the restriction of T\mathcal{R}_{T} onto suppT\operatorname*{supp}T. In other words, CXC\subseteq X is a measurable set satisfying iTi(C)=suppT\bigcup_{i\in\mathbb{Z}}T^{i}(C)=\operatorname*{supp}T. The induced transformation UXCU_{X\setminus C} commutes with TCT_{C}, since their supports are disjoint. We would be done if suppUiUi(XC)\operatorname*{supp}U\subseteq\bigcup_{i\in\mathbb{Z}}U^{i}(X\setminus C). Indeed, in this case T𝔖(𝔾)=TC𝔖(𝔾)T\mathfrak{S}(\mathbb{G})=T_{C}\mathfrak{S}(\mathbb{G}), U𝔖(𝔾)=UXC𝔖(𝔾)U\mathfrak{S}(\mathbb{G})=U_{X\setminus C}\mathfrak{S}(\mathbb{G}) by Lemma 3.11 and [TC,UXC][T_{C},U_{X\setminus C}] is trivial, hence [T,U]𝔖(𝔾)[T,U]\in\mathfrak{S}(\mathbb{G}).

Motivated by this observation, we argue as follows. Pick a vanishing nested sequence (Cn)n(C_{n})_{n\in\mathbb{N}} of cross-sections for TsuppT\mathcal{R}_{T}\restriction_{\operatorname*{supp}T}, i.e., CnCn+1C_{n}\supseteq C_{n+1}, kTk(Cn)=suppT\bigcup_{k\in\mathbb{Z}}T^{k}(C_{n})=\operatorname*{supp}T for all nn\in\mathbb{N}, and nCn=\bigcap_{n\in\mathbb{N}}C_{n}=\varnothing (see also Lemma D.11). Such a sequence of cross-sections exists since TT is assumed to be aperiodic on its support. Define inductively sets BnB^{\prime}_{n}, nn\in\mathbb{N}, by setting B0=XC0B^{\prime}_{0}=X\setminus C_{0}, and letting BnB^{\prime}_{n} be the part of XCnX\setminus C_{n} that does not belong to the UU-saturation of any BkB^{\prime}_{k}, k<nk<n,

Bn=(XCn)k<niUi(Bk).B^{\prime}_{n}=(X\setminus C_{n})\setminus\bigcup_{k<n}\bigcup_{i\in\mathbb{Z}}U^{i}(B^{\prime}_{k}).

By construction, saturations under UU of the sets BnB^{\prime}_{n} are pairwise disjoint, and the saturation of their union is the whole space, iUi(nBn)=X\bigcup_{i\in\mathbb{Z}}U^{i}\bigl{(}\bigcup_{n\in\mathbb{N}}B^{\prime}_{n}\bigr{)}=X, because sets CnC_{n} vanish.

Let Bn=k<nBkB_{n}=\bigsqcup_{k<n}B^{\prime}_{k}, B=kBkB=\bigsqcup_{k\in\mathbb{N}}B^{\prime}_{k}, and note that UBn,UB𝔾U_{B_{n}},U_{B}\in\mathbb{G}, and UBkUBU_{B_{k}}\to U_{B} by the induction friendliness of 𝔾\mathbb{G}. By construction, transformations TCnT_{C_{n}} and UBnU_{B_{n}} have disjoint supports for each nn and, therefore, commute. Since all sets CnC_{n} are cross-sections for TsuppT\mathcal{R}_{T}\restriction_{\operatorname*{supp}T}, one has [T,UBn]𝔖(𝔾)[T,U_{B_{n}}]\in\mathfrak{S}(\mathbb{G}) by Lemma 3.11 and Remark 3.12. Taking the limit as nn\to\infty, this yields [T,UB]𝔖(𝔾)[T,U_{B}]\in\mathfrak{S}(\mathbb{G}). Finally, the UU-saturation of BB is all of XX, we use Lemma 3.11 and Remark 3.12 once again to conclude that [T,U]𝔖(𝔾)[T,U]\in\mathfrak{S}(\mathbb{G}), as claimed. ∎

Proposition 3.14.

If 𝔾\mathbb{G} is an aperiodic induction friendly Polish finitely full group, then 𝔖(𝔾)=D(𝔾)\mathfrak{S}(\mathbb{G})=D(\mathbb{G}).

Proof.

Inclusion 𝔄(𝔾)D(𝔾)\mathfrak{A}(\mathbb{G})\leq D(\mathbb{G}) holds for any Polish finitely full group and Proposition 3.8 gives 𝔖(𝔾)D(𝔾)\mathfrak{S}(\mathbb{G})\leq D(\mathbb{G}). We therefore concentrate on proving the reverse inclusion: given T,U𝔾T,U\in\mathbb{G}, we need to check that [T,U]𝔖(𝔾)[T,U]\in\mathfrak{S}(\mathbb{G}). Let FTF_{T} and FUF_{U} be the aperiodic parts of TT and UU respectively, so that T𝔖(𝔾)=TFT𝔖(𝔾)T\mathfrak{S}(\mathbb{G})=T_{F_{T}}\mathfrak{S}(\mathbb{G}), U𝔖(𝔾)=UFU𝔖(𝔾)U\mathfrak{S}(\mathbb{G})=U_{F_{U}}\mathfrak{S}(\mathbb{G}) by Lemma 3.11. By construction, TFTT_{F_{T}} and UFUU_{F_{U}} are aperiodic on their supports and therefore [TFT,UFU]𝔖(𝔾)[T_{F_{T}},U_{F_{U}}]\in\mathfrak{S}(\mathbb{G}) by Lemma 3.13. It remains to use Remark 3.12 to conclude that necessarily [T,U]𝔖(𝔾)[T,U]\in\mathfrak{S}(\mathbb{G}), as needed. ∎

Corollary 3.15.

Let GG be a Polish normed group, and let GXG\curvearrowright X be an aperiodic Borel measure-preserving action on a standard probability space (X,μ)(X,\mu). The three subgroups of [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} introduced in Definition 3.7 coincide:

D([GX]1)=𝔄([GX]1)=𝔖([GX]1).D([G\curvearrowright X\mkern 1.5mu]_{1})=\mathfrak{A}([G\curvearrowright X\mkern 1.5mu]_{1})=\mathfrak{S}([G\curvearrowright X\mkern 1.5mu]_{1}).

Moreover, they are all equal to the closure of the group generated by periodic elements of [GX]1[G\curvearrowright X]_{1}.

Proof.

The equality D([GX]1)=𝔄([GX]1)=𝔖([GX]1)D([G\curvearrowright X]_{1})=\mathfrak{A}([G\curvearrowright X]_{1})=\mathfrak{S}([G\curvearrowright X]_{1}) follows immediately from Propositions 3.8 and 3.14, since [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is both finitely full and induction friendly. All these groups are equal to the closure of the group generated by periodic elements of [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} in view of Lemma 3.10 and the fact that this group obviously contains 𝔖([GX]1)\mathfrak{S}([G\curvearrowright X\mkern 1.5mu]_{1}). ∎

3.3. Topological simplicity of the symmetric group

We now move on to showing that symmetric subgroups of ergodic Polish finitely full groups are always topologically simple. Our argument abstracts from [LM18, Sec. 3.4]. In particular, we rely on conditional measures associated with subgroups of Aut(X,μ)\mathrm{Aut}(X,\mu), whose construction and basic properties are recalled in Appendix D.

Lemma 3.16.

Let 𝔾\mathbb{G} be an aperiodic Polish finitely full group, let U,V𝔾U,V\in\mathbb{G} be two involutions whose supports are disjoint and have the same 𝔾\mathbb{G}-conditional measure. Then UU and VV are approximately conjugate in 𝔖(𝔾)\mathfrak{S}(\mathbb{G}), i.e., there are Tn𝔖(𝔾)T_{n}\in\mathfrak{S}(\mathbb{G}) such that TnUTn1VT_{n}UT_{n}^{-1}\to V.

Proof.

Let AA (resp. BB) be a fundamental domain of the restriction of UU (resp. VV) to its support. Then μ𝔾(A)=μ𝔾(B)\mu_{\mathbb{G}}(A)=\mu_{\mathbb{G}}(B) and there is an involution T[𝔾]T\in[\mathbb{G}] such that T(A)=BT(A)=B.

Since 𝔾\mathbb{G} is finitely full, there is an increasing sequence (An)n(A_{n})_{n} of subsets of AA such that the involutions TnT^{\prime}_{n} induced by TT on AnU(An)A_{n}\cup U(A_{n}) belong to 𝔾\mathbb{G}, and nAn=A\bigcup_{n}A_{n}=A. Let Bn=T(An)=Tn(An)B_{n}=T(A_{n})=T^{\prime}_{n}(A_{n}) and define involutions Tn𝔾T_{n}\in\mathbb{G} which almost conjugate UU to VV as follows. For xXx\in X, let

Tnx={Txif xAnBnVTUxif xU(An)UTVxif xV(Bn)xotherwise.T_{n}x=\left\{\begin{array}[]{cl}Tx&\text{if }x\in A_{n}\sqcup B_{n}\\ VTUx&\text{if }x\in U(A_{n})\\ UTVx&\text{if }x\in V(B_{n})\\ x&\text{otherwise.}\end{array}\right.

For all nn\in\mathbb{N} and all xXx\in X, an easy calculation yields that:

  • if x(AU(A))(AnU(An))x\in(A\cup U(A))\setminus(A_{n}\cup U(A_{n})), then TnUTnx=UxT_{n}UT_{n}x=Ux;

  • if xBnV(Bn)x\in B_{n}\cup V(B_{n}), then TnUTnx=VxT_{n}UT_{n}x=Vx;

  • and TnUTnx=xT_{n}UT_{n}x=x in all other cases.

In particular, du(TnUTn,V)0d_{u}(T_{n}UT_{n},V)\to 0 and Proposition 3.4, applied to the full group of the involution UVUV (which contains both UU and VV), guarantees that TnUTnVT_{n}UT_{n}\to V. ∎

Lemma 3.17.

Let 𝔾\mathbb{G} be an aperiodic Polish finitely full group, let U𝔾U\in\mathbb{G} be an involution, and let AA be a UU-invariant subset contained in suppU\operatorname*{supp}U. Suppose that there exists an involution V𝔾V\in\mathbb{G} such that V(A)V(A) is disjoint from suppU\operatorname*{supp}U. Then for all 𝔾\mathbb{G}-invariant functions f2μ𝔾(A)f\leq 2\mu_{\mathbb{G}}(A), there is an involution W𝔾W\in\mathbb{G} such that UWUWUWUW is an involution whose support has 𝔾\mathbb{G}-conditional measure ff.

Proof.

Let BAB\subseteq A be a fundamental domain for the restriction of UU to AA and note that μ𝔾(B)=μ𝔾(A)/2\mu_{\mathbb{G}}(B)=\mu_{\mathbb{G}}(A)/2. By Maharam’s lemma (Theorem D.12), there is CBC\subseteq B such that μ𝔾(C)=f/4\mu_{\mathbb{G}}(C)=f/4 . The set D=CU(C)D=C\sqcup U(C) is UU-invariant and satisfies μ𝔾(D)=f/2\mu_{\mathbb{G}}(D)=f/2. Consider the involution W𝔾W\in\mathbb{G} defined by

Wx={Vxif xDV(D)xotherwise.Wx=\left\{\begin{array}[]{cl}Vx&\text{if }x\in D\sqcup V(D)\\ x&\text{otherwise.}\end{array}\right.

A straightforward computation shows that UWUWUWUW is an involution which coincides with UU on DD, with VUVVUV on V(D)V(D), and is trivial elsewhere. Hence the support of UWUWUWUW is equal to DV(D)D\sqcup V(D), and has 𝔾\mathbb{G}-conditional measure ff. ∎

Proposition 3.18.

Let 𝔾\mathbb{G} be an aperiodic Polish finitely full group, let T𝔾T\in\mathbb{G}, and let AA denote the 𝔾\mathbb{G}-saturation of suppT\operatorname*{supp}T. The closed subgroup of 𝔾\mathbb{G} generated by the 𝔖(𝔾)\mathfrak{S}(\mathbb{G})-conjugates of TT contains 𝔖(𝔾)A\mathfrak{S}(\mathbb{G})_{A}.

Proof.

Let the closed subgroup of 𝔾\mathbb{G} generated by the 𝔖(𝔾)\mathfrak{S}(\mathbb{G})-conjugates of TT be denoted by GG. We can find BsuppTB\subseteq\operatorname*{supp}T whose TT-translates cover suppT\operatorname*{supp}T and which satisfies BT(B)=B\cap T(B)=\varnothing. Since TT-translates of BB cover suppT\operatorname*{supp}T, we conclude that the 𝔾\mathbb{G}-translates of BB cover AA, and so μ𝔾(B)(x)>0\mu_{\mathbb{G}}(B)(x)>0 for all xAx\in A. By Maharam’s lemma (Theorem D.12), we can find CBC\subseteq B whose 𝔾\mathbb{G}-conditional measure is everywhere less than 1/41/4, and is strictly positive on AA. Let D=CT(C)D=C\sqcup T(C) and take V[𝔾]V\in[\mathbb{G}\mkern 1.5mu] to be an involution such that V(CT(C))V(C\sqcup T(C)) is disjoint from CT(C)C\sqcup T(C).

Let W[𝔾]W\in[\mathbb{G}] be an involution such that suppW=C\operatorname*{supp}W=C. Using the facts that 𝔾\mathbb{G} is finitely full, that T𝔾T\in\mathbb{G} and that V,W[𝔾]V,W\in[\mathbb{G}], one can find an increasing sequence (Cn)n(C_{n})_{n} of WW-invariant subsets of CC such that nCn=C\bigcup_{n}C_{n}=C and for each nn\in\mathbb{N} both WCn𝔾W_{C_{n}}\in\mathbb{G} and VCnT(Cn)V(CnT(Cn))𝔾V_{C_{n}\sqcup T(C_{n})\sqcup V(C_{n}\sqcup T(C_{n}))}\in\mathbb{G}. Transformations WCnTWCnT1W_{C_{n}}TW_{C_{n}}T^{-1} belong to GG, and are, in fact, involutions whose support is equal to CnT(Cn)C_{n}\sqcup T(C_{n}) and has conditional measure at most 2μ𝔾(C)1/22\mu_{\mathbb{G}}(C)\leq 1/2. Let us define for brevity

U~n=WCnTWCnT1G and V~n=VCnT(Cn)V(CnT(Cn))𝔾.\tilde{U}_{n}=W_{C_{n}}TW_{C_{n}}T^{-1}\in G\text{ and }\tilde{V}_{n}=V_{C_{n}\sqcup T(C_{n})\sqcup V(C_{n}\sqcup T(C_{n}))}\in\mathbb{G}.

For every nn\in\mathbb{N}, let AnA_{n} denote the 𝔾\mathbb{G}-saturation of CnC_{n}. Note that A=nAnA=\bigcup_{n}A_{n} and the union is increasing. Every involution supported on AA is thus the uniform limit of the involutions it induces on AnA_{n}’s. By Proposition 3.4, it therefore suffices to show that GG contains all the involutions which are supported on some AnA_{n}.

Let UU be an involution UU supported on some AnA_{n}. Let DD be a fundamental domain for the restriction of UU to its support. Using Maharam’s lemma repeatedly, we can partition DD into a countable family (Dk)k(D_{k})_{k} such that

(3.2) μ𝔾(Dk)μ𝔾(suppU~n)/2 for all k.\mu_{\mathbb{G}}(D_{k})\leq\mu_{\mathbb{G}}(\operatorname*{supp}\tilde{U}_{n})/2\quad\textrm{ for all $k\in\mathbb{N}$}.

If we let Ek=DkU(Dk)E_{k}=D_{k}\sqcup U(D_{k}), the sequence (Ek)k(E_{k})_{k} forms a partition of suppU\operatorname*{supp}U into UU-invariant sets. In particular, U=limki=0kUEkU=\lim_{k}\prod_{i=0}^{k}U_{E_{k}} in the uniform topology and therefore in the topology of 𝔾\mathbb{G} as well by Proposition 3.4. Moreover, the support of UEkU_{E_{k}} has 𝔾\mathbb{G}-conditional measure at most μ𝔾(suppU~n)\mu_{\mathbb{G}}(\operatorname*{supp}\tilde{U}_{n}) by Eq. (3.2). The set V~n(suppU~n)\tilde{V}_{n}(\operatorname*{supp}\tilde{U}_{n}) is disjoint from suppU~n\operatorname*{supp}\tilde{U}_{n} by construction. Lemma 3.17 applies and provides an involution in GG whose support has the same conditional measure as that of UEkU_{E_{k}}. Lemma 3.16 shows that each UEkU_{E_{k}} belongs to GG and therefore also UGU\in G, as needed. ∎

Theorem 3.19.

Let 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) be an aperiodic Polish finitely full group. For any closed normal subgroup N𝔖(𝔾)N\leq\mathfrak{S}(\mathbb{G}) there is a unique 𝔾\mathbb{G}-invariant set AA such that N=𝔖(𝔾)AN=\mathfrak{S}(\mathbb{G})_{A}.

Proof.

First, observe that for 𝔾\mathbb{G}-invariant A1A_{1} and A2A_{2}, any involution U𝔾U\in\mathbb{G} supported in A1A2A_{1}\cup A_{2} decomposes into the product of one involution supported in A1A_{1}, and one supported in A2A_{2}. It follows that the closed group generated by 𝔖(𝔊)A1𝔖(𝔊)A2\mathfrak{S}(\mathfrak{G})_{A_{1}}\cup\mathfrak{S}(\mathfrak{G})_{A_{2}} is equal to 𝔖(𝔊)A1A2\mathfrak{S}(\mathfrak{G})_{A_{1}\cup A_{2}}. Also, by Proposition 3.4, whenever (An)n(A_{n})_{n} is an increasing sequence of 𝔾\mathbb{G}-invariant sets, one has

n𝔖(𝔾)An¯=𝔖(𝔾)nAn.\overline{\bigcup_{n}\mathfrak{S}(\mathbb{G})_{A_{n}}}=\mathfrak{S}(\mathbb{G})_{\bigcup_{n}A_{n}}.

The set {AMAlg(X,μ):A is 𝔾-invariant and 𝔖(𝔾)AN}\{A\in\mathrm{MAlg}(X,\mu):A\text{ is }\mathbb{G}\text{-invariant and }\mathfrak{S}(\mathbb{G})_{A}\leq N\} is thus directed and is closed under the countable unions. It therefore admits a unique maximum element, which is the set AA we seek. Indeed, 𝔖(𝔾)AN\mathfrak{S}(\mathbb{G})_{A}\leq N, and the reverse inclusion is a direct consequence of Proposition 3.18.

It remains to argue that the set AA satisfying N=𝔖(𝔾)AN=\mathfrak{S}(\mathbb{G})_{A} is unique. Suppose towards a contradiction that 𝔖(𝔾)A1=𝔖(𝔾)A2\mathfrak{S}(\mathbb{G})_{A_{1}}=\mathfrak{S}(\mathbb{G})_{A_{2}}for A1A2A_{1}\neq A_{2}. By symmetry, we may assume that μ(A1A2)>0\mu(A_{1}\setminus A_{2})>0. Lemma 3.6 provides an involution V𝔾V\in\mathbb{G} whose support is nontrivial and is contained in A1A2A_{1}\setminus A_{2}, thus V𝔖(𝔾)A1V\in\mathfrak{S}(\mathbb{G})_{A_{1}} but V𝔖(𝔾)A2V\not\in\mathfrak{S}(\mathbb{G})_{A_{2}}, contradicting 𝔖(𝔾)A1=𝔖(𝔾)A2\mathfrak{S}(\mathbb{G})_{A_{1}}=\mathfrak{S}(\mathbb{G})_{A_{2}}. ∎

Corollary 3.20.

Let 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) be an aperiodic Polish finitely full group. The group 𝔖(𝔾)\mathfrak{S}(\mathbb{G}) is topologically simple if and only if 𝔾\mathbb{G} is ergodic.

Proof.

If 𝔾\mathbb{G} is ergodic, then 𝔖(𝔾)\mathfrak{S}(\mathbb{G}) is topologically simple by Theorem 3.19. Conversely, suppose that 𝔾\mathbb{G} is not ergodic and let AXA\subseteq X be a 𝔾\mathbb{G}-invariant set with μ(A){0,1}\mu(A)\not\in\{0,1\}. Then 𝔖(𝔾)A\mathfrak{S}(\mathbb{G})_{A} is a normal subgroup of 𝔾\mathbb{G} which is neither trivial nor equal to 𝔖(𝔾)\mathfrak{S}(\mathbb{G}) as a consequence of Lemma 3.6 applied to AA and its complement. ∎

Specifying the corollary above to L1\mathrm{L}^{1} full groups and using Corollary 3.15, we obtain the following result.

Corollary 3.21.

Let GG be a Polish normed group, and let GXG\curvearrowright X be an aperiodic Borel measure-preserving action on a standard probability space (X,μ)(X,\mu). The topological derived subgroup of the L1\mathrm{L}^{1} full group of the action is topologically simple if and only if the action is ergodic.

3.4. Maximal norms on the derived subgroup

The purpose of this section is to establish sufficient conditions for a norm on the derived subgroup of an induction friendly Polish finitely full group to be maximal in the sense of Section 2.3. Our argument follows closely the one given in [LM21, Sec. 6.2] for amenable graphings. The main application of Proposition 3.24 will be given in Theorem 5.5, but we hope that the setup of this section can be useful in other contexts, such as φ\varphi-integrable full groups [CJMT22].

Definition 3.22.

A norm \left\lVert\cdot\right\rVert on a subgroup 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) is additive if TS=T+S\left\lVert TS\right\rVert=\left\lVert T\right\rVert+\left\lVert S\right\rVert for all T,S𝔾T,S\in\mathbb{G} with disjoint supports.

The following lemma parallels [LM21, Lem. 6.4] and is the key to showing that the norm on the derived subgroup is both coarsely proper and large-scale geodesic.

Lemma 3.23.

Let 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) be a finitely full Polish group, and suppose that \left\lVert\cdot\right\rVert is a compatible additive norm on 𝔾\mathbb{G}. For any periodic U𝔾U\in\mathbb{G} with bounded periods and for every nn\in\mathbb{N}, there are periodic elements U1,,Un𝔾U_{1},...,U_{n}\in\mathbb{G} such that

U=U1Un and Ui=Un for every 1in.U=U_{1}\cdots U_{n}\text{ and }\left\lVert U_{i}\right\rVert=\frac{\left\lVert U\right\rVert}{n}\text{ for every }1\leq i\leq n.
Proof.

Let M=UM=\left\lVert U\right\rVert and AXA\subseteq X be a fundamental domain for UU. We may identify AA with the interval [0,μ(A)][0,\mu(A)] endowed with the Lebesgue measure. Put At=[0,t]AA_{t}=[0,t]\cap A, 0tμ(A)0\leq t\leq\mu(A), and let Bt=nUn(At)B_{t}=\bigcup_{n\in\mathbb{Z}}U^{n}(A_{t}) be the UU-saturation of AtA_{t}. Note that UBt𝔾U_{B_{t}}\in\mathbb{G} for all t[0,μ(A)]t\in[0,\mu(A)] since BtB_{t} is UU-invariant and 𝔾\mathbb{G} is finitely full, and that tBtt\mapsto B_{t} is continuous.

The map [0,μ(A)]tUBt[U]𝔾[0,\mu(A)]\ni t\mapsto U_{B_{t}}\in[U\mkern 1.5mu]\subseteq\mathbb{G} is thus continuous with respect to the uniform topology on [U][U\mkern 1.5mu], and therefore also with respect to the topology of 𝔾\mathbb{G} by Proposition 3.4. Whence the function ψ:[0,μ(A)]\psi:[0,\mu(A)]\to\mathbb{R} given by ψ(t)=UBt\psi(t)=\left\lVert U_{B_{t}}\right\rVert is also continuous.

We have ψ(0)=0\psi(0)=0 and ψ(μ(A))=M\psi(\mu(A))=M, so the intermediate value theorem yields existence of reals 0=t0<t1<<tn1<tn=μ(A)0=t_{0}<t_{1}<\cdots<t_{n-1}<t_{n}=\mu(A) such that ψ(ti)=iMn\psi(t_{i})=\frac{iM}{n} for all i{0,,n}i\in\{0,...,n\}. Set Ci=BtiBti1C_{i}=B_{t_{i}}\setminus B_{t_{i-1}} for i{1,,n}i\in\{1,...,n\}. By construction, each CiC_{i} is UU-invariant and X=i=1nCiX=\bigsqcup_{i=1}^{n}C_{i}. Putting Ui=UAiU_{i}=U_{A_{i}}, we get U=i=1nUiU=\prod_{i=1}^{n}U_{i}. Finally for each i{1,,n}i\in\{1,...,n\} the equality Ci=BtiBti1C_{i}=B_{t_{i}}\setminus B_{t_{i-1}} and additivity of the norm gives

ψ(ti)=UBti=UiUBti1=Ui+UBti1=Ui+ψ(ti1),\psi(t_{i})=\lVert U_{B_{t_{i}}}\rVert=\lVert U_{i}U_{B_{t_{i-1}}}\rVert=\lVert U_{i}\rVert+\lVert U_{B_{t_{i-1}}}\rVert=\lVert U_{i}\rVert+\psi(t_{i-1}),

hence Ui=Un\left\lVert U_{i}\right\rVert=\frac{\left\lVert U\right\rVert}{n} for all ini\leq n, as needed. ∎

Proposition 3.24.

Let 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) be an induction friendly Polish finitely full group and let \left\lVert\cdot\right\rVert be a compatible additive norm on it. If the set of periodic elements is dense in D(𝔾)D(\mathbb{G}), then \left\lVert\cdot\right\rVert is a maximal norm on D(𝔾)D(\mathbb{G}).

Proof.

In view of Proposition A.10, it suffices to show that \left\lVert\cdot\right\rVert is both large-scale geodesic (see Definition A.8) and coarsely proper (see Definition A.9). Note that induction friendliness yields density in D(𝔾)D(\mathbb{G}) of periodic automorphisms with bounded periods.

To see that \left\lVert\cdot\right\rVert is large-scale geodesic (with constant K=2K=2), let us take a non-trivial TD(𝔾)T\in D(\mathbb{G}) and pick a periodic UD(𝔾)U\in D(\mathbb{G}) with bounded periods such that TU1<min{2,T/2}\left\lVert TU^{-1}\right\rVert<\min\{2,\left\lVert T\right\rVert/2\}. Note that

(3.3) U=U1=T1TU1T1+TU13T/2\lVert U\rVert=\lVert U^{-1}\rVert=\lVert T^{-1}TU^{-1}\rVert\leq\lVert T^{-1}\rVert+\lVert TU^{-1}\rVert\leq 3\lVert T\rVert/2

Fix nn\in\mathbb{N} large enough to ensure 3T2n2\frac{3\left\lVert T\right\rVert}{2n}\leq 2. By Lemma 3.23, we may decompose UU into a product of nn elements U1,,UnU_{1},\dots,U_{n} each of norm at most 3T2n2\frac{3\left\lVert T\right\rVert}{2n}\leq 2. Therefore

T=(TU1)U1Un,T=(TU^{-1})\cdot U_{1}\cdots U_{n},

where TU1TU^{-1} and each of UiU_{i}, 1in1\leq i\leq n, has norm at most 22 and, in view of Eq. (3.3),

TU1+i=1nUiT2+U2T,\lVert TU^{-1}\rVert+\sum_{i=1}^{n}\left\lVert U_{i}\right\rVert\leq\frac{\left\lVert T\right\rVert}{2}+\lVert U\rVert\leq 2\lVert T\rVert,

thus concluding the proof that \left\lVert\cdot\right\rVert is large-scale geodesic.

We now show that \left\lVert\cdot\right\rVert is coarsely proper. Fix ϵ>0\epsilon>0 and R>0R>0. Let nn\in\mathbb{N} be so large that nϵR+ϵn\epsilon\geq R+\epsilon. Then every element TD(𝔾)T\in D(\mathbb{G}) of norm at most RR is a product of n+1n+1 elements of norm at most ϵ\epsilon, namely one element TU1TU^{-1} of norm at most ϵ\epsilon, where UU is periodic with bounded periods as provided by density, and U=U1UnU=U_{1}\cdots U_{n}, where each UiU_{i} has norm at most R+ϵnϵ\frac{R+\epsilon}{n}\leq\epsilon as per Lemma 3.23. Thus \left\lVert\cdot\right\rVert is both coarsely proper and large-scale geodesic, and hence is maximal by Proposition A.10. ∎

Remark 3.25.

We do not have an example of an induction friendly Polish finitely full group 𝔾\mathbb{G} such that the periodic elements are not dense in D(𝔾)D(\mathbb{G}). We suspect that such groups do exist, for instance when 𝔾\mathbb{G} is the L1\mathrm{L}^{1} full group of a free action of the free group on 22 generators.

Chapter 4 Full groups of locally compact group actions

In this chapter, we narrow down the generality of the narrative and focus on actions of locally compact Polish groups, or equivalently, of locally compact second-countable groups. Such restrictions enlarge our toolbox in a number of ways. For instance, all locally compact Polish group actions admit cross-sections to which the so-called Voronoi tessellations can be associated. We use this to show in Section 4.1 a natural density result for subsets of L1\mathrm{L}^{1} full groups defined from dense subsets of the acting group (Theorem 4.2 and Corollary 4.3). For reader’s convenience, Appendix C.2 contains a concise reminder of the needed facts about tessellations.

Another key property of free111Motivated by our focus on \mathbb{R}-flows, this monograph primarily concentrates on free actions. We note, however, that each orbit of a Borel action of a locally compact Polish group is a homogeneous space, since point stabilizers are necessarily closed. In particular, orbits can be endowed with the Haar measure even without the freeness assumption. actions of locally compact groups is the existence of a Haar measure on each individual orbit. As we discuss in Section 4.2, elements of the full group act by non-singular transformations and, in particular, admit the Hopf decomposition (see Appendix B). Section 4.3 explains how these orbitwise decompositions can be understood globally, yielding a natural generalization of the periodic/aperiodic partition for elements of the full group of a measure-preserving action of a discrete group. The periodic part in the later case corresponds to the conservative piece of the Hopf decomposition, which generally exhibits a much more complicated dynamical behavior. We will get back to this in Chapters 7 and 8.

In the final Section 4.4, we connect L1\mathrm{L}^{1} full groups to the notion of L1\mathrm{L}^{1} orbit equivalence for actions of locally compact compactly generated Polish groups.

4.1. Dense subgroups in L1\mathrm{L}^{1} full groups

Our goal in this section is to prove that any element of the full group [GX][G\curvearrowright X\mkern 1.5mu] can be approximated arbitrarily well by an automorphism that piecewise acts by elements of a given dense subset of GG.

Definition 4.1.

A measure-preserving transformation T:ABT:A\to B between two measurable sets A,BXA,B\subseteq X is said to be HH-decomposable, where HAut(X,μ)H\subseteq\mathrm{Aut}(X,\mu), if there exist a measurable partition A=kAkA=\bigsqcup_{k\in\mathbb{N}}A_{k} and elements hkHh_{k}\in H such that TAk=hkT\restriction_{A_{k}}=h_{k} for all kk\in\mathbb{N}.

The property of being HH-decomposable is similar to being an element of the full group generated by HH except that we do not require the transformation to be defined on all of XX.

Theorem 4.2.

Let GXG\curvearrowright X be a measure-preserving action of a locally compact Polish group, let \left\lVert\cdot\right\rVert be a compatible norm on GG with the associated metric on the orbits D:G0D:\mathcal{R}_{G}\to\mathbb{R}^{\geq 0}, and let HGH\subseteq G be a dense set. For any T[GX]T\in[G\curvearrowright X\mkern 1.5mu] and any ϵ>0\epsilon>0 there exists an HH-decomposable transformation S[GX]S\in[G\curvearrowright X\mkern 1.5mu] such that esssupxXD(Tx,Sx)<ϵ\operatorname*{ess\,sup}_{x\in X}D(Tx,Sx)<\epsilon.

Theorem 4.2 establishes density of HH-decomposable transformations in the very strong uniform topology given by esssup\operatorname*{ess\,sup}. In particular, it pertains to the L1\mathrm{L}^{1} topology.

Corollary 4.3.

Let GXG\curvearrowright X be a measure-preserving action of a locally compact Polish group, let \lVert\cdot\rVert be a compatible norm on GG, and let HGH\subseteq G be a dense subgroup. The L1\mathrm{L}^{1} full group [HX]1[H\curvearrowright X\mkern 1.5mu]_{1} is dense in [GX]1[G\curvearrowright X\mkern 1.5mu]_{1}.

Remark 4.4.

Theorem 4.2 is an improvement upon the conclusion of [CLM18, Thm. 2.1], which shows that [HX][H\curvearrowright X\mkern 1.5mu] is dense in [GX][G\curvearrowright X\mkern 1.5mu] whenever HH is a dense subgroup of GG. While the proof, which we present below, establishes density in a much stronger topology through more elementary means, we note that, as already mentioned in [CLM18, Thm. 2.3], their methods apply to all suitable (in the sense of [Bec13]; see also Definition 4.7) actions of Polish groups, whereas our approach here crucially uses local compactness of the acting group to guarantee existence of various cross-sections.

Let 𝒞\mathcal{C} be a cross-section for a measure-preserving action GXG\curvearrowright X and let 𝒲\mathcal{W} be a tessellation over 𝒞\mathcal{C} (in the sense of Appendix C.2). Let also ν𝒲\nu_{\mathcal{W}} be the push-forward measure (π𝒲)μ(\pi_{\mathcal{W}})_{*}\mu on the cross-section and (μc)c𝒞(\mu_{c})_{c\in\mathcal{C}} be the disintegration of μ\mu over (π𝒲,ν𝒲)(\pi_{\mathcal{W}},\nu_{\mathcal{W}}) (see Appendix C.1 and Theorem C.1, specifically). Without loss of generality, we assume, whenever convenient, that the set HH in the statement of Theorem 4.2 is countable.

Definition 4.5.

Two Borel sets A,BXA,B\subseteq X are said to be

  • 𝒲\mathcal{W}-proportionate if the equivalence μc(A)=0μc(B)=0\mu_{c}(A)=0\iff\mu_{c}(B)=0 holds for ν𝒲\nu_{\mathcal{W}}-almost all c𝒞c\in\mathcal{C};

  • 𝒲\mathcal{W}-equimeasurable if μc(A)=μc(B)\mu_{c}(A)=\mu_{c}(B) for ν𝒲\nu_{\mathcal{W}}-almost all c𝒞c\in\mathcal{C}.

For the context of Lemmas 4.6 through 4.10, we let NN denote an open symmetric neighborhood of the identity of GG, and 𝒲\mathcal{W} stands for an NN-lacunary tessellation.

Lemma 4.6.

If A,BN𝒞A,B\subseteq N\cdot\mathcal{C} are 𝒲\mathcal{W}-proportionate Borel sets then

μ(BN2A)=0.\mu(B\setminus N^{2}\cdot A)=0.
Proof.

By the defining property of the disintegration,

μ(BN2A)=𝒞μc(BN2A)𝑑ν𝒲(c),\mu(B\setminus N^{2}\cdot A)=\int_{\mathcal{C}}\mu_{c}(B\setminus N^{2}\cdot A)\,d\nu_{\mathcal{W}}(c),

and so we need to check that μc(BN2A)=0\mu_{c}(B\setminus N^{2}\cdot A)=0 for ν𝒲\nu_{\mathcal{W}}-almost all cc. Since AA and BB are 𝒲\mathcal{W}-proportionate, it suffices to show that μc(BN2A)=0\mu_{c}(B\setminus N^{2}\cdot A)=0 whenever μc(A)0\mu_{c}(A)\neq 0. For any c𝒞c\in\mathcal{C} satisfying the latter, one necessarily has cNAc\in N\cdot A (because AN𝒞A\subseteq N\cdot\mathcal{C} and 𝒲\mathcal{W} is NN-lacunary, by assumption), and thus NcN2AN\cdot c\subseteq N^{2}\cdot A. In particular, (BN2A)Nc=(B\setminus N^{2}\cdot A)\cap N\cdot c=\varnothing. It remains to use the inclusion BN𝒞B\subseteq N\cdot\mathcal{C}, which together with NN-lacunarity of 𝒲\mathcal{W}, guarantees that

μc(BN2A)=μc((BN2A)Nc)=0.\mu_{c}(B\setminus N^{2}\cdot A)=\mu_{c}((B\setminus N^{2}\cdot A)\cap N\cdot c)=0.\qed

For the proof of the next lemma, we need the notion of a suitable action, introduced by H. Becker [Bec13, Def. 1.2.7].

Definition 4.7.

A measure-preserving Borel action GXG\curvearrowright X of a Polish group GG is suitable if for all Borel sets A,BXA,B\subseteq X one of the two options holds:

  1. (1)

    for any open neighborhood of the identity MGM\subseteq G there exists gMg\in M such that μ(gAB)>0\mu(gA\cap B)>0;

  2. (2)

    there exist Borel sets AAA^{\prime}\subseteq A, BBB^{\prime}\subseteq B such that μ(AA)=0=μ(BB)\mu(A\setminus A^{\prime})=0=\mu(B\setminus B^{\prime}) and an open neighborhood of the identity MGM\subseteq G such that MAB=M\cdot A^{\prime}\cap B^{\prime}=\varnothing.

All measure-preserving actions of locally compact Polish groups are known to be suitable (see [Bec13, Thm. 1.2.9]).

Lemma 4.8.

For all non-negligible 𝒲\mathcal{W}-proportionate Borel sets A,BN𝒞A,B\subseteq N\cdot\mathcal{C}, there exists an open set UN3U\subseteq N^{3} such that μ(gAB)>0\mu(gA\cap B)>0 for all gUg\in U.

Proof.

Let H1=HN2H_{1}=H\cap N^{2}, which is dense in N2=NN1N^{2}=NN^{-1}, and put A1=H1AA_{1}=H_{1}\cdot A. We apply the dichotomy in the definition of a suitable action to the sets A1,BA_{1},B and show that item (2) cannot hold.

Indeed, suppose there exist A1A1A_{1}^{\prime}\subseteq A_{1}, BBB^{\prime}\subseteq B satisfying

μ(A1A1)=0=μ(BB),\mu(A_{1}\setminus A^{\prime}_{1})=0=\mu(B\setminus B^{\prime}),

and an open neighborhood of the identity MGM\subseteq G such that (MA1)B=(M\cdot A^{\prime}_{1})\cap B^{\prime}=\varnothing. Set A=n(hn1A1A)A^{\prime}=\bigcap_{n}(h_{n}^{-1}A^{\prime}_{1}\cap A), where (hn)n(h_{n})_{n\in\mathbb{N}} is an enumeration of H1H_{1}, and note that μ(AA)=0\mu(A\setminus A^{\prime})=0 and (MH1A)B=(MH_{1}\cdot A^{\prime})\cap B^{\prime}=\varnothing, simply because H1AA1H_{1}\cdot A^{\prime}\subseteq A^{\prime}_{1}. Since H1H_{1} is dense in N2N^{2}, we have N2MH1N^{2}\subseteq MH_{1} and thus (N2A)B=(N^{2}\cdot A^{\prime})\cap B^{\prime}=\varnothing. Lemma 4.6, applied to AA^{\prime} and BB^{\prime}, guarantees that μ(BN2A)=0\mu(B^{\prime}\setminus N^{2}\cdot A^{\prime})=0, which is possible only when μ(B)=0\mu(B^{\prime})=0, contradicting the assumption that BB is non-negligible.

We are left with the alternative of the item (1), and so there has to exist some gNg\in N such that μ(gA1B)>0\mu(gA_{1}\cap B)>0. Since A1=H1AA_{1}=H_{1}\cdot A, there exists hH1h\in H_{1} such that μ(ghAB)>0\mu(ghA\cap B)>0. It remains to note that ghN3gh\in N^{3} and that μ(gAB)>0\mu(g^{\prime}A\cap B)>0 is an open condition on gg^{\prime}, since the homomorphism GAut(X,μ)G\to\mathrm{Aut}(X,\mu) associated to the measure-preserving action of GG on (X,μ)(X,\mu) is continuous (see for instance [CLM18, Lem. 1.2]). ∎

Lemma 4.9.

For any non-empty open VNV\subseteq N and for any non-negligible Borel set AXA\subseteq X, there exists hHh\in H such that

μ({xA:hxV𝒞 and π𝒲(x)=π𝒲(hx)})>0.\mu(\{x\in A:hx\in V\cdot\mathcal{C}\textrm{ and }\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(hx)\})>0.
Proof.

Let ζ:X𝒲\zeta:X\to\mathcal{W} be the Borel bijection ζ(x)=(π𝒲(x),x)\zeta(x)=(\pi_{\mathcal{W}}(x),x) and consider the push-forward measure ζμ\zeta_{*}\mu, which for Z𝒲Z\subseteq\mathcal{W} can be expressed as ζμ(Z)=𝒞μc(Zc)𝑑ν𝒲(c)\zeta_{*}\mu(Z)=\int_{\mathcal{C}}\mu_{c}(Z_{c})\,d\nu_{\mathcal{W}}(c). Let (hn)n(h_{n})_{n\in\mathbb{N}} be an enumeration of HH and set

Wn={(c,x)𝒲:π𝒲(x)=π𝒲(hnx) and hnxV𝒞}.W_{n}=\{(c,x)\in\mathcal{W}:\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(h_{n}x)\textrm{ and }h_{n}x\in V\cdot\mathcal{C}\}.

We claim that nWn=𝒲\bigcup_{n}W_{n}=\mathcal{W}. Indeed, for each (c,x)𝒲(c,x)\in\mathcal{W} the set of gGg\in G such that gxVcgx\in V\cdot c is non-empty and open, hence there is hnHh_{n}\in H such that hnxVch_{n}x\in V\cdot c.

Finally, AA is non-negligible by assumption, i.e., 0<μ(A)=ζμ(ζ(A))0<\mu(A)=\zeta_{*}\mu(\zeta(A)), so there exists WnW_{n} such that ζμ(ζ(A)Wn)>0\zeta_{*}\mu(\zeta(A)\cap W_{n})>0, which translates into the required

μ({xA:hnxV𝒞 and π𝒲(x)=π𝒲(hnx)})>0.\mu(\{x\in A:h_{n}x\in V\cdot\mathcal{C}\textrm{ and }\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(h_{n}x)\})>0.\qed
Lemma 4.10.

For all non-negligible 𝒲\mathcal{W}-proportionate Borel sets A,BXA,B\subseteq X, there exists hHh\in H such that

μ({xA:hxB and π𝒲(x)=π𝒲(hx)})>0.\mu(\{x\in A:hx\in B\textrm{ and }\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(hx)\})>0.
Proof.

The plan is to reduce the setup of this lemma to that of Lemma 4.8. Let VNV\subseteq N be a symmetric neighborhood of the identity that is furthermore small enough to guarantee that 𝒲\mathcal{W} is V4V^{4}-lacunary. Apply Lemma 4.9 to find h1Hh_{1}\in H such that for

A={xA:h1xV𝒞 and π𝒲(x)=π𝒲(h1x)}A^{\prime}=\{x\in A:h_{1}x\in V\cdot\mathcal{C}\textrm{ and }\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(h_{1}x)\}

one has μ(A)>0\mu(A^{\prime})>0. Set A1=h1AA_{1}=h_{1}A^{\prime}, B1=π𝒲1({c𝒞:μc(A1)>0})BB_{1}=\pi_{\mathcal{W}}^{-1}(\{c\in\mathcal{C}:\mu_{c}(A_{1})>0\})\cap B and note that A1A_{1} and B1B_{1} are non-negligible 𝒲\mathcal{W}-proportionate sets. Moreover, A1V𝒞A_{1}\subseteq V\cdot\mathcal{C} by construction.

Repeat the same steps for B1B_{1} and find h2Hh_{2}\in H such that for

B1={xB1:h2xV𝒞 and π𝒲(x)=π𝒲(h2x)}B_{1}^{\prime}=\{x\in B_{1}:h_{2}x\in V\cdot\mathcal{C}\textrm{ and }\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(h_{2}x)\}

we have μ(B1)>0\mu(B_{1}^{\prime})>0. Set B2=h2B1B_{2}=h_{2}B_{1}^{\prime} and A2=A1π𝒲1({c𝒞:μc(B2)>0})A_{2}=A_{1}\cap\pi_{\mathcal{W}}^{-1}(\{c\in\mathcal{C}:\mu_{c}(B_{2})>0\}). Once again, sets A2A_{2} and B2B_{2} are non-negligible, 𝒲\mathcal{W}-proportionate and are both contained in V𝒞V\cdot\mathcal{C}.

We now apply Lemma 4.8 to sets A2,B2A_{2},B_{2} and 𝒲\mathcal{W}, viewed as a VV-lacunary tessellation, yielding an open UV3U\subseteq V^{3} such that μ(gA2B2)>0\mu(gA_{2}\cap B_{2})>0 for all gUg\in U. Note that since UV3U\subseteq V^{3} and 𝒲\mathcal{W} is, in fact, V4V^{4}-lacunary, the equality π𝒲(x)=π𝒲(gx)\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(gx) holds for all xV𝒞x\in V\cdot\mathcal{C} and gUg\in U. We conclude that μ(h21gh1AB)>0\mu(h_{2}^{-1}gh_{1}A\cap B)>0 for all gUg\in U and hence any hh21Uh1Hh\in h_{2}^{-1}Uh_{1}\cap H satisfies the conclusion of the lemma. ∎

A measure-preserving map T:ABT:A\to B is 𝒲\mathcal{W}-coherent if μ\mu-almost surely one has π𝒲(x)=π𝒲(Tx)\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(Tx).

Lemma 4.11.

For all 𝒲\mathcal{W}-equimeasurable Borel sets A,BXA,B\subseteq X, there exists a 𝒲\mathcal{W}-coherent HH-decomposable measure-preserving bijection T:ABT:A\to B.

Proof.

Let (hn)n(h_{n})_{n\in\mathbb{N}} be an enumeration of HH. Consider the set

A0={xA:h0xB and π𝒲(x)=π𝒲(h0x)},A_{0}=\bigl{\{}x\in A:h_{0}x\in B\textrm{ and }\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(h_{0}x)\bigr{\}},

and let B0=h0A0B_{0}=h_{0}A_{0}. Note that the sets AA0A\setminus A_{0} and BB0B\setminus B_{0} are 𝒲\mathcal{W}-equimeasurable, so we may continue in the same fashion and construct sets AkA_{k} such that

Ak={xAi<kAi:hkxBi<kBi and π𝒲(x)=π𝒲(hkx)}.A_{k}=\Bigl{\{}x\in A\setminus\bigsqcup_{i<k}A_{i}:h_{k}x\in B\setminus\bigsqcup_{i<k}B_{i}\textrm{ and }\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(h_{k}x)\Bigr{\}}.

We define T:kAkkBkT:\bigsqcup_{k\in\mathbb{N}}A_{k}\to\bigsqcup_{k\in\mathbb{N}}B_{k} by the condition Tx=hkxTx=h_{k}x for xAkx\in A_{k}.

Sets AkAkA\setminus\bigsqcup_{k\in\mathbb{N}}A_{k} and BkBkB\setminus\bigsqcup_{k\in\mathbb{N}}B_{k} are 𝒲\mathcal{W}-equimeasurable. If either one of them (and thus necessarily both of them) were non-negligible, Lemma 4.10 would yields an element hHh\in H that moves a portion of AkAkA\setminus\bigsqcup_{k\in\mathbb{N}}A_{k} into BkBkB\setminus\bigsqcup_{k\in\mathbb{N}}B_{k}, contradicting the construction. We conclude that

μ(AkAk)=0=μ(BkBk)\mu(A\setminus\bigsqcup_{k\in\mathbb{N}}A_{k})=0=\mu(B\setminus\bigsqcup_{k\in\mathbb{N}}B_{k})

and TT is therefore as required. ∎

Lemma 4.12.

Suppose 𝒲\mathcal{W} is a cocompact tessellation and let A,BXA,B\subseteq X be 𝒲\mathcal{W}-equimeasurable Borel sets. For any ϵ>0\epsilon>0 and any 𝒲\mathcal{W}-coherent measure-preserving T:ABT:A\to B there exists is a 𝒲\mathcal{W}-coherent HH-decomposable T~:AB\tilde{T}:A\to B such that esssupxAD(Tx,T~x)<ϵ\operatorname*{ess\,sup}_{x\in A}D(Tx,\tilde{T}x)<\epsilon.

Proof.

Let 𝒱\mathcal{V} be a KK^{\prime}-cocompact tessellation over some cross-section 𝒞\mathcal{C}^{\prime} such that the diameter of each region in 𝒱\mathcal{V} is less than ϵ\epsilon. Suppose 𝒲\mathcal{W} is KK-cocompact. By Lemma C.9, we can find a finite partition of 𝒞=in𝒞i\mathcal{C}^{\prime}=\bigsqcup_{i\leq n}\mathcal{C}^{\prime}_{i} such that each 𝒞i\mathcal{C}^{\prime}_{i} is KK2KK^{\prime}K^{2}K^{\prime}-lacunary, which guarantees that, for each ii, every 𝒲c\mathcal{W}_{c} intersects at most one class 𝒱c\mathcal{V}_{c^{\prime}}, c𝒞ic^{\prime}\in\mathcal{C}^{\prime}_{i}. For each i,j<ni,j<n set A(i,j)={xA:π𝒱(x)𝒞i,π𝒱(Tx)𝒞j}A_{(i,j)}=\{x\in A:\pi_{\mathcal{V}}(x)\in\mathcal{C}^{\prime}_{i},\pi_{\mathcal{V}}(Tx)\in\mathcal{C}_{j}^{\prime}\} and B(i,j)=TA(i,j)B_{(i,j)}=TA_{(i,j)}. We re-enumerate sets A(i,j)A_{(i,j)} and B(i,j)B_{(i,j)} as a sequence AkA_{k}, BkB_{k}, kn2k\leq n^{2} and note that for all x,yAkx,y\in A_{k} one has

π𝒲(x)=π𝒲(y)(π𝒱(x)=π𝒱(y) and π𝒱(Tx)=π𝒱(Ty)).\pi_{\mathcal{W}}(x)=\pi_{\mathcal{W}}(y)\implies\bigl{(}\pi_{\mathcal{V}}(x)=\pi_{\mathcal{V}}(y)\textrm{ and }\pi_{\mathcal{V}}(Tx)=\pi_{\mathcal{V}}(Ty)\bigr{)}.

Moreover, sets AkA_{k} and T(Ak)T(A_{k}) are 𝒲\mathcal{W}-equimeasurable, so Lemma 4.11 yields 𝒲\mathcal{W}-coherent HH-decomposable measure-preserving maps Tk:AkT(Ak)T_{k}:A_{k}\to T(A_{k}). The transformation T~:AB\tilde{T}:A\to B can now be defined by the condition T~x=Tkx\tilde{T}x=T_{k}x whenever xAkx\in A_{k}. It is easy to check that T~\tilde{T} is as claimed. ∎

Proof of Theorem 4.2.

Fix a cocompact cross-section 𝒞\mathcal{C}, and let (Un)n(U_{n})_{n} be a nested and exhaustive sequence of compact neighborhoods of the identity in GG. For all nn\in\mathbb{N}, select based on Lemma C.9 a finite sequence of cocompact cross-sections 𝒞1n,,𝒞knn\mathcal{C}^{n}_{1},\ldots,\mathcal{C}^{n}_{k_{n}} such that each 𝒞in\mathcal{C}^{n}_{i} is UnU_{n}-lacunary and 𝒞=i=1kn𝒞in\mathcal{C}=\bigsqcup_{i=1}^{k_{n}}\mathcal{C}^{n}_{i}. Re-enumerate cross-sections 𝒞in\mathcal{C}^{n}_{i}, nn\in\mathbb{N}, 1ikn1\leq i\leq k_{n}, into a sequence (𝒞k)k=0(\mathcal{C}_{k})_{k=0}^{\infty} and let 𝒱k\mathcal{V}_{k} be the Voronoi tessellation over 𝒞k\mathcal{C}_{k}.

Let A0={xX:π𝒱0(x)=π𝒱0(Tx)}A_{0}=\{x\in X:\pi_{\mathcal{V}_{0}}(x)=\pi_{\mathcal{V}_{0}}(Tx)\}, and use Lemma 4.12 to find an HH-decomposable measure-preserving map T0:A0T(A0)T_{0}:A_{0}\to T(A_{0}) that satisfies the inequality esssupxA0D(T0x,Tx)<ϵ\operatorname*{ess\,sup}_{x\in A_{0}}D(T_{0}x,Tx)<\epsilon. Set

Ak={xX:π𝒱k(x)=π𝒱k(Tx) and xl<kAl}A_{k}=\bigl{\{}x\in X:\pi_{\mathcal{V}_{k}}(x)=\pi_{\mathcal{V}_{k}}(Tx)\textrm{ and }x\not\in\bigsqcup_{l<k}A_{l}\bigr{\}}

and observe that AkA_{k}, kk\in\mathbb{N}, form a partition of XX. Find transformations Tk:AkT(Ak)T_{k}:A_{k}\to T(A_{k}) by repeated applications of Lemma 4.12 applied to the tessellations 𝒱k\mathcal{V}_{k}. The element S[GX]S\in[G\curvearrowright X\mkern 1.5mu] defined by Sx=TkxSx=T_{k}x satisfies the conclusion of the theorem. ∎

4.2. Orbital transformations

Let GXG\curvearrowright X be a free measure-preserving action of a locally compact Polish group on a standard probability space. Fix a right-invariant Haar measure λ\lambda on GG. Any orbit [x]G[x]_{\mathcal{R}_{G}} can be identified with the group itself via the map Gggx[x]GG\ni g\mapsto gx\in[x]_{\mathcal{R}_{G}}, and λ\lambda can be pushed via this identification onto orbits resulting in a collection (λx)xX(\lambda_{x})_{x\in X} of measures on XX defined by λx(A)=λ({gG:gxA})\lambda_{x}(A)=\lambda(\{g\in G:gx\in A\}). Right invariance of the measure ensures that λx\lambda_{x} depends only on the orbit [x]G[x]_{\mathcal{R}_{G}} and is independent of the choice of the base point, i.e., λx=λy\lambda_{x}=\lambda_{y} whenever xGyx\mathcal{R}_{G}y.

This section focuses on two main facts: the so-called mass-transport principle, given in Eq. (4.1) below, and non-singularity of the transformations induced by elements of [GX][G\curvearrowright X\mkern 1.5mu] onto orbits of the action, formulated in Proposition 4.13. Both of these topics have been discussed in the literature in many related contexts including, for instance, [CLM18, Appen. A] and the treatise [ADR00]. We are, however, not aware of any specific reference from which Eq. (4.1) and Proposition 4.13 can be readily deduced. The following derivations are therefore included for reader’s convenience, with the disclaimer that these results are likely to be known to experts.

Freeness of the action allows us to identify the equivalence relation G\mathcal{R}_{G} with X×GX\times G via Φ:X×GG\Phi:X\times G\to\mathcal{R}_{G}, Φ(x,g)=(x,gx)\Phi(x,g)=(x,gx). The push-forward Φ(μ×λ)\Phi_{*}(\mu\times\lambda) of the product measure is denoted by MM and can equivalently be defined by

M(A)=Xλx(Ax)𝑑μ(x),M(A)=\int_{X}\lambda_{x}(A_{x})\,d\mu(x),

where AGA\subseteq\mathcal{R}_{G} and Ax={yX:(x,y)A}A_{x}=\{y\in X:(x,y)\in A\}.

In general, the flip transformation σ:GG\sigma:\mathcal{R}_{G}\to\mathcal{R}_{G}, σ(x,y)=(y,x)\sigma(x,y)=(y,x), is not MM-invariant. Set Ψ:X×GX×G\Psi:X\times G\to X\times G to be Ψ=Φ1σΦ\Psi=\Phi^{-1}\circ\sigma\circ\Phi, which amounts to Ψ(x,g)=(gx,g1)\Psi(x,g)=(gx,g^{-1}). Following the computation as in [CLM18, Prop. A.11], one can easily check that Ψ(μ×λ)=μ×λ^\Psi_{*}(\mu\times\lambda)=\mu\times\widehat{\lambda}, where λ^\widehat{\lambda} is the associated left-invariant measure, λ^(A)=λ(A1)\widehat{\lambda}(A)=\lambda(A^{-1}). If we define the measure M^\widehat{M} on G\mathcal{R}_{G} to be

M^(A)=Φ(μ×λ^)=Xλ^x(Ax)𝑑μ(x),\widehat{M}(A)=\Phi_{*}(\mu\times\widehat{\lambda})=\int_{X}\widehat{\lambda}_{x}(A_{x})\,d\mu(x),

then σM=M^\sigma_{*}M=\widehat{M}. In particular, σ\sigma is MM-invariant if and only if λ=λ^\lambda=\widehat{\lambda}, i.e., GG is unimodular.

Let Δ:G>0\Delta:G\to\mathbb{R}^{>0} be the left Haar modulus function given for gGg\in G by λ(gA)=Δ(g)λ(A)\lambda(gA)=\Delta(g)\lambda(A). Recall that Δ:G>0\Delta:G\to\mathbb{R}^{>0} is a continuous homomorphism (see, for instance, [Nac65, Prop. 7]), measures λ\lambda and λ^\widehat{\lambda} belong to the same measure class and dλ^dλ(g)=Δ(g1)\frac{d\widehat{\lambda}}{d\lambda}(g)=\Delta(g^{-1}) for all gGg\in G (see [Nac65, p. 79]).

A function f:Gf:\mathcal{R}_{G}\to\mathbb{R} is MM-integrable if and only if X×G(x,g)f(x,gx)X\times G\ni(x,g)\mapsto f(x,gx) is (μ×λ)(\mu\times\lambda)-integrable, which together with the expression for the Radon-Nikodym derivative d(μ×λ^)d(μ×λ)=dλ^dλ\frac{d(\mu\times\widehat{\lambda})}{d(\mu\times\lambda)}=\frac{d\widehat{\lambda}}{d\lambda} and Fubini’s theorem yields the following identity:

(4.1) XGf(x,gx)𝑑λ(g)𝑑μ(x)=XGΔ(g)f(gx,x)𝑑λ(g)𝑑μ(x).\int_{X}\int_{G}f(x,g\cdot x)\,d\lambda(g)d\mu(x)=\int_{X}\int_{G}\Delta(g)f(g\cdot x,x)\,d\lambda(g)d\mu(x).

When the group GG is unimodular, this expression attains a very symmetric form and is known as the mass-transport principle:

(4.2) XGf(x,gx)𝑑λ(g)𝑑μ(x)=XGf(gx,x)𝑑λ(g)𝑑μ(x).\int_{X}\int_{G}f(x,g\cdot x)\,d\lambda(g)d\mu(x)=\int_{X}\int_{G}f(g\cdot x,x)\,d\lambda(g)d\mu(x).

Any automorphism T[GX]T\in[G\curvearrowright X\mkern 1.5mu] induces for each xXx\in X a transformation of the σ\sigma-finite measure space (X,λx)(X,\lambda_{x}). In general, TT does not preserve λx\lambda_{x}, however, it is always non-singular, and the Radon-Nikodym derivative dTλxdλx\frac{dT_{*}\lambda_{x}}{d\lambda_{x}} can be described explicitly. Note that the full group [GX][G\curvearrowright X\mkern 1.5mu] admits two natural actions on the equivalence relation G\mathcal{R}_{G}: the left action ll is given by lT(x,y)=(Tx,y)l_{T}(x,y)=(Tx,y), and the right action rr is defined as rT(x,y)=(x,Ty)r_{T}(x,y)=(x,Ty). A straightforward verification (see [CLM18, Lem. A.9]) shows that ll is always MM-invariant. Since rTσ=σlTr_{T}\circ\sigma=\sigma\circ l_{T}, for all T[GX]T\in[G\curvearrowright X\mkern 1.5mu] we have

(rT)M^=(rTσ)M=(σlT)M=σM=M^.(r_{T})_{*}\widehat{M}=(r_{T}\circ\sigma)_{*}M=(\sigma\circ l_{T})_{*}M=\sigma_{*}M=\widehat{M}.

Let Θ=Φ1rTΦ\Theta=\Phi^{-1}\circ r_{T}\circ\Phi, i.e., Θ(x,g)=(x,ρTg(x))\Theta(x,g)=(x,\rho_{Tg}(x)). The equality (rT)M^=M^(r_{T})_{*}\widehat{M}=\widehat{M} is equivalent to Θ(μ×λ^)\Theta_{*}(\mu\times\widehat{\lambda}) = μ×λ^\mu\times\widehat{\lambda}. The latter implies that each Borel BGB\subseteq G and all measurable AXA\subseteq X we have

Aλ^(B)𝑑μ\displaystyle\int_{A}\widehat{\lambda}(B)\,d\mu =(μ×λ^)(A×B)=Θ(μ×λ^)(A×B)\displaystyle=(\mu\times\widehat{\lambda})(A\times B)=\Theta_{*}(\mu\times\widehat{\lambda})(A\times B)
=(μ×λ^)({(x,g)X×G:(x,ρTg(x))A×B})\displaystyle=(\mu\times\widehat{\lambda})\bigl{(}\{(x,g)\in X\times G:(x,\rho_{Tg}(x))\in A\times B\}\bigr{)}
Fubini’s theorem =Aλ^({gG:ρTg(x)B})𝑑μ(x)\displaystyle=\int_{A}\widehat{\lambda}(\{g\in G:\rho_{Tg}(x)\in B\})\,d\mu(x)
=Aλ^({gG:gxT1Bx})𝑑μ(x),\displaystyle=\int_{A}\widehat{\lambda}(\{g\in G:gx\in T^{-1}Bx\})\,d\mu(x),

which is possible only if λ^({gG:gxT1Bx})=λ^(B)\widehat{\lambda}(\{g\in G:gx\in T^{-1}Bx\})=\widehat{\lambda}(B) for μ\mu-almost all xx. Passing to the measures on the orbits, this translates for each BB into λ^x(T1Bx)=λ^x(Bx)\widehat{\lambda}_{x}(T^{-1}Bx)=\widehat{\lambda}_{x}(Bx). If (Bn)n(B_{n})_{n\in\mathbb{N}} is a countable algebra of Borel sets in GG that generates the whole Borel σ\sigma-algebra, then for each xXx\in X, (Bnx)n(B_{n}x)_{n\in\mathbb{N}} is an algebra of Borel subsets of the orbit [x]G[x]_{\mathcal{R}_{G}}, which generates the Borel σ\sigma-algebra on it. We have established that for μ\mu-almost all xXx\in X the two measures, λ^x\widehat{\lambda}_{x} and Tλ^xT_{*}\widehat{\lambda}_{x}, coincide on each BnxB_{n}x, nn\in\mathbb{N}, thus μ\mu-almost surely λ^x=Tλ^x\widehat{\lambda}_{x}=T_{*}\widehat{\lambda}_{x}.

Equality dλ^dλ(g)=Δ(g1)\frac{d\widehat{\lambda}}{d\lambda}(g)=\Delta(g^{-1}) translates into dλ^xdλx(y)=Δ(ρ(x,y)1)=Δ(ρ(y,x))\frac{d\widehat{\lambda}_{x}}{d\lambda_{x}}(y)=\Delta(\rho(x,y)^{-1})=\Delta(\rho(y,x)) and the Radon-Nikodym derivative dTλxdλx\frac{dT_{*}\lambda_{x}}{d\lambda_{x}} can now be computed as follows.

dTλxdλx(y)\displaystyle\frac{dT_{*}\lambda_{x}}{d\lambda_{x}}(y) =dTλxdTλ^x(y)dTλ^xdλ^x(y)dλ^xdλx(y)\displaystyle=\frac{dT_{*}\lambda_{x}}{dT_{*}\widehat{\lambda}_{x}}(y)\cdot\frac{dT_{*}\widehat{\lambda}_{x}}{d\widehat{\lambda}_{x}}(y)\cdot\frac{d\widehat{\lambda}_{x}}{d\lambda_{x}}(y)
TT preservers λ^x\widehat{\lambda}_{x} =dTλxdTλ^x(y)dλ^xdλx(y)=dλxdλ^x(T1y)dλ^xdλx(y)\displaystyle=\frac{dT_{*}\lambda_{x}}{dT_{*}\widehat{\lambda}_{x}}(y)\cdot\frac{d\widehat{\lambda}_{x}}{d\lambda_{x}}(y)=\frac{d\lambda_{x}}{d\widehat{\lambda}_{x}}(T^{-1}y)\cdot\frac{d\widehat{\lambda}_{x}}{d\lambda_{x}}(y)
=(dλ^xdλx(T1y))1dλ^xdλx(y)\displaystyle=\Bigl{(}\frac{d\widehat{\lambda}_{x}}{d\lambda_{x}}(T^{-1}y)\Bigr{)}^{-1}\cdot\frac{d\widehat{\lambda}_{x}}{d\lambda_{x}}(y)
=Δ(ρ(x,T1y)1)1Δ(ρ(x,y)1)\displaystyle=\Delta(\rho(x,T^{-1}y)^{-1})^{-1}\Delta(\rho(x,y)^{-1})
=Δ(ρ(x,T1y)ρ(y,x))=Δ(ρT1(y)).\displaystyle=\Delta\bigl{(}\rho(x,T^{-1}y)\cdot\rho(y,x)\bigr{)}=\Delta(\rho_{T^{-1}}(y)).

We summarize the content of this section into a proposition.

Proposition 4.13.

Let GG be a locally compact Polish group acting freely GXG\curvearrowright X on a standard probability space (X,μ)(X,\mu). Let λ\lambda be a right Haar measure on GG, Δ:G>0\Delta:G\to\mathbb{R}^{>0} be the corresponding Haar modulus, and let (λx)xX(\lambda_{x})_{x\in X} be the family of measures obtained by pushing λ\lambda onto orbits via the action map. Each T[GX]T\in[G\curvearrowright X\mkern 1.5mu] induces a non-singular transformation of (X,λx)(X,\lambda_{x}) for almost every xXx\in X, and moreover one has λx(T1A)=AΔ(ρT1(y))𝑑λx(y)\lambda_{x}(T^{-1}A)=\int_{A}\Delta(\rho_{T^{-1}}(y))\,d\lambda_{x}(y) for all Borel sets AXA\subseteq X. If GG is unimodular, then Tλx=λxT_{*}\lambda_{x}=\lambda_{x} for μ\mu-almost all xXx\in X.

For future reference, we isolate a simple lemma, which is an immediate consequence of Fubini’s theorem.

Lemma 4.14.

Let GG be a locally compact Polish group acting freely on a standard probability space (X,μ)(X,\mu). Let λ\lambda, λ^\widehat{\lambda}, (λx)xX(\lambda_{x})_{x\in X}, and (λ^)xX(\widehat{\lambda})_{x\in X} be as above. For any Borel set AXA\subseteq X the following are equivalent:

  1. (1)

    μ(A)=0\mu(A)=0;

  2. (2)

    λx(A)=0\lambda_{x}(A)=0 for μ\mu-almost all xXx\in X;

  3. (3)

    λ^x(A)=0\widehat{\lambda}_{x}(A)=0 for μ\mu-almost all xXx\in X.

Proof.

(1) \iff  (2) Using Fubini’s Theorem on (X×G,μ×λ)(X\times G,\mu\times\lambda) to rearrange the order of quantifiers, one has:

μ(A)=0gGμxXgxAμxXλgGgxAμxXλx(A)=0.\mu(A)=0\iff\forall g\in G\ \forall^{\mu}x\in X\ gx\not\in A\\ \iff\forall^{\mu}x\in X\ \forall^{\lambda}g\in G\ gx\not\in A\iff\forall^{\mu}x\in X\ \lambda_{x}(A)=0.

(2) \iff  (3) is evident, since λ\lambda and λ^\widehat{\lambda} are equivalent measures, hence so are λx\lambda_{x} and λ^x\widehat{\lambda}_{x} for all xXx\in X. ∎

4.3. The Hopf decomposition of elements of the full group

Fix an element T[GX]T\in[G\curvearrowright X\mkern 1.5mu] of the full group of a free measure-preserving action of a locally compact Polish group GG. As explained in Section 4.2, TT acts naturally in a non-singular manner on each GG-orbit. This action thus has a Hopf decomposition (see Appendix B). We will now explain how to understand globally this decomposition, obtaining a generalization of the fact that when GG is discrete, any element of the full group decomposes the space into a periodic and an aperiodic part.

Let us pick a cocompact cross-section 𝒞\mathcal{C} and let 𝒱𝒞\mathcal{V}_{\mathcal{C}} be the associated Voronoi tessellation (see Appendix C.2). Set π𝒞:X𝒞\pi_{\mathcal{C}}:X\to\mathcal{C} to be the projection map given by the condition (π𝒞(x),x)𝒱𝒞(\pi_{\mathcal{C}}(x),x)\in\mathcal{V}_{\mathcal{C}} for all xXx\in X. Define the dissipative and conservative sets as follows:

D\displaystyle D ={xX:nk such that |k|n one has π𝒞(x)π𝒞(Tkx)},\displaystyle=\bigl{\{}x\in X:\exists n\in\mathbb{N}\ \forall k\in\mathbb{Z}\textrm{ such that }|k|\geq n\textrm{ one has }\pi_{\mathcal{C}}(x)\neq\pi_{\mathcal{C}}(T^{k}x)\bigr{\}},
C\displaystyle C ={xX:nk1,k2 such that\displaystyle=\bigl{\{}x\in X:\forall n\in\mathbb{N}\ \exists k_{1},k_{2}\in\mathbb{Z}\textrm{ such that }
k1n,nk2 and π𝒞(Tk1x)=π𝒞(x)=π𝒞(Tk2x)}.\displaystyle\qquad\qquad k_{1}\leq-n,n\leq k_{2}\textrm{ and }\pi_{\mathcal{C}}(T^{k_{1}}x)=\pi_{\mathcal{C}}(x)=\pi_{\mathcal{C}}(T^{k_{2}}x)\bigr{\}}.

In plain words, the dissipative set DD consists of those points xx whose orbit has a finite intersection with the Voronoi region of xx. The conservative set CC, on the other hand, collects all the points whose orbit is bi-recurrent in the region. We argue in Proposition 4.16 that sets DD and CC induce the Hopf decomposition for T[x]TT\restriction_{[x]_{\mathcal{R}_{T}}} for almost every xXx\in X; in particular, DCD\sqcup C is a partition of XX, which is independent of the choice of the cross-section 𝒞\mathcal{C}.

Lemma 4.15.

Sets DD and CC partition the phase space: X=DCX=D\sqcup C.

Proof.

Define sets N+N_{+} and NN_{-} according to

N+\displaystyle N_{+} ={xX(DC):k1π𝒞(Tkx)π𝒞(x)},\displaystyle=\{x\in X\setminus(D\sqcup C):\forall k\geq 1\ \pi_{\mathcal{C}}(T^{k}x)\neq\pi_{\mathcal{C}}(x)\},
N\displaystyle N_{-} ={xX(DC):k1π𝒞(Tkx)π𝒞(x)},\displaystyle=\{x\in X\setminus(D\sqcup C):\forall k\geq 1\ \pi_{\mathcal{C}}(T^{-k}x)\neq\pi_{\mathcal{C}}(x)\},

and note that X(DC)kTk(N+N)X\setminus(D\sqcup C)\subseteq\bigcup_{k\in\mathbb{Z}}T^{k}(N_{+}\cup N_{-}). To show that X=DCX=D\sqcup C it is enough to verify that μ(N+)=0=μ(N)\mu(N_{+})=0=\mu(N_{-}).

This is done by noting that these sets admit pairwise disjoint copies using piecewise translations by powers of TT. In view of the fact that TT is measure-preserving, this implies that N+N_{+} and NN_{-} are null. To be more precise, set N0=NN_{-}^{0}=N_{-} and define inductively Nn={Tk(x)x:xNn1}N^{n}_{-}=\{T^{k(x)}x:x\in N^{n-1}_{-}\}, where k(x)1k(x)\geq 1 is the smallest natural number such that π𝒞(Tk(x)x)=π𝒞(x)\pi_{\mathcal{C}}(T^{k(x)}x)=\pi_{\mathcal{C}}(x). Note that k(x)k(x) is well-defined, for otherwise xx would belong to DD. Sets NnN^{n}_{-}, nn\in\mathbb{N}, are pairwise disjoint, and have the same measure since TT is measure-preserving. We conclude that μ(N)=0\mu(N_{-})=0. The argument for μ(N+)=0\mu(N_{+})=0 is similar. ∎

Proposition 4.16 (Hopf decomposition).

Let GXG\curvearrowright X be a free measure-preserving action of a locally compact Polish group on a standard probability space (X,μ)(X,\mu). Let λ\lambda be a right Haar measure on GG and (λx)xX(\lambda_{x})_{x\in X} be the push-forward of λ\lambda onto the orbits as described in Section 4.2. For any element T[GX]T\in[G\curvearrowright X\mkern 1.5mu], the measurable TT-invariant partition X=DCX=D\sqcup C defined above satisfies that for μ\mu-almost all xXx\in X the partition [x]G=([x]GD)([x]GC)[x]_{\mathcal{R}_{G}}=([x]_{\mathcal{R}_{G}}\cap D)\sqcup([x]_{\mathcal{R}_{G}}\cap C) is the Hopf decomposition for T[x]GT\restriction_{[x]_{\mathcal{R}_{G}}} on ([x]G,λx)([x]_{\mathcal{R}_{G}},\lambda_{x}). Moreover, there is only one partition X=DCX=D\sqcup C satisfying this property up to null sets.

Proof.

According to Proposition 4.13, we may assume that for all xXx\in X the map T[x]G:[x]G[x]GT\restriction_{[x]_{\mathcal{R}_{G}}}:[x]_{\mathcal{R}_{G}}\to[x]_{\mathcal{R}_{G}} is a non-singular transformation with respect to λx\lambda_{x} and satisfies λx(TA)=AΔ(ρT(y))𝑑λx(y)\lambda_{x}(TA)=\int_{A}\Delta(\rho_{T}(y))\,d\lambda_{x}(y) for all Borel AXA\subseteq X.

Let [x]G=DxCx[x]_{\mathcal{R}_{G}}=D_{x}\sqcup C_{x}, xXx\in X, denote the Hopf’s decomposition for T[x]GT\restriction_{[x]_{\mathcal{R}_{G}}}. For any c𝒞c\in\mathcal{C}, the set

W~c={x(𝒱𝒞)c:Tkx(𝒱𝒞)c for all k1}\widetilde{W}_{c}=\bigl{\{}x\in(\mathcal{V}_{\mathcal{C}})_{c}:T^{k}x\not\in(\mathcal{V}_{\mathcal{C}})_{c}\textrm{ for all }k\geq 1\bigr{\}}

is a wandering set and therefore W~cDx\widetilde{W}_{c}\subseteq D_{x} up to a null set. If xDx\in D satisfies x(𝒱𝒞)cx\in(\mathcal{V}_{\mathcal{C}})_{c}, c𝒞c\in\mathcal{C}, then [x]G(𝒱𝒞)c[x]_{\mathcal{R}_{G}}\cap(\mathcal{V}_{\mathcal{C}})_{c} is finite, and therefore [x]G(𝒱𝒞)ckTkW~c[x]_{\mathcal{R}_{G}}\cap(\mathcal{V}_{\mathcal{C}})_{c}\subseteq\bigcup_{k\in\mathbb{Z}}T^{k}\widetilde{W}_{c}, whence also

[x]GDc𝒞[x]GkTkW~cDx.[x]_{\mathcal{R}_{G}}\cap D\subseteq\bigcup_{c\in\mathcal{C}\cap[x]_{\mathcal{R}_{G}}}\bigcup_{k\in\mathbb{Z}}T^{k}\widetilde{W}_{c}\subseteq D_{x}.
Claim.

We have λx([x]GCDx)=0\lambda_{x}([x]_{\mathcal{R}_{G}}\cap C\cap D_{x})=0 for each xXx\in X.

Proof of the claim.

Otherwise we can find c𝒞[x]Gc\in\mathcal{C}\cap[x]_{\mathcal{R}_{G}} and a wandering set W[x]G(𝒱𝒞)cCW\subseteq[x]_{\mathcal{R}_{G}}\cap(\mathcal{V}_{\mathcal{C}})_{c}\cap C of positive measure, λx(W)>0\lambda_{x}(W)>0. Construct a sequence of sets WnW_{n} by setting W0=WW_{0}=W and

Wn={Tkn(y)y:yW0 and kn(y) is minimal such that π𝒞(Tkn(y))=π𝒞(y) and Tkn(y)yk<nWk},W_{n}=\bigl{\{}T^{k_{n}(y)}y:y\in W_{0}\textrm{ and }k_{n}(y)\textrm{ is minimal such that }\\ \pi_{\mathcal{C}}(T^{k_{n}(y)})=\pi_{\mathcal{C}}(y)\textrm{ and }T^{k_{n}(y)}y\not\in\bigcup_{k<n}W_{k}\bigr{\}},

where the value of kn(y)k_{n}(y) is well-defined for each yW0y\in W_{0} and nn\in\mathbb{N}, since all points in CC return to their Voronoi domain infinitely often. Define a transformation Sn:W0WnS_{n}:W_{0}\to W_{n} as Sn(y)=Tkn(y)yS_{n}(y)=T^{k_{n}(y)}y, and note that for all nn\in\mathbb{N} one has ρSn(y)ρ((𝒱𝒞)c,(𝒱𝒞)c)\rho_{S_{n}}(y)\in\rho((\mathcal{V}_{\mathcal{C}})_{c},(\mathcal{V}_{\mathcal{C}})_{c}). The region ρ((𝒱𝒞)c,(𝒱𝒞)c)\rho((\mathcal{V}_{\mathcal{C}})_{c},(\mathcal{V}_{\mathcal{C}})_{c}) is precompact, since 𝒞\mathcal{C} is cocompact, and therefore using continuity of the Haar modulus Δ:G>0\Delta:G\to\mathbb{R}^{>0} one can pick ϵ>0\epsilon>0 such that Δ(ρSn(y))>ϵ\Delta(\rho_{S_{n}}(y))>\epsilon for all yW0y\in W_{0} and all nn\in\mathbb{N}.

Since SnS_{n} is composed of powers of TT, Proposition 4.13 ensures that

λx(SnW0)=W0Δ(ρSn(y))𝑑λx(y),\lambda_{x}(S_{n}W_{0})=\int_{W_{0}}\Delta(\rho_{S_{n}}(y))\,d\lambda_{x}(y),

whence λx(SnW0)ϵλx(W0)\lambda_{x}(S_{n}W_{0})\geq\epsilon\lambda_{x}(W_{0}) for each nn\in\mathbb{N}. We now arrive at a contradiction, as WnW_{n}, nn\in\mathbb{N}, form a pairwise disjoint infinite family of subsets of (𝒱𝒞)c(\mathcal{V}_{\mathcal{C}})_{c} whose measure is uniformly bounded away from zero by ϵλx(W0)\epsilon\lambda_{x}(W_{0}), which is impossible, since λx((𝒱𝒞)c)<\lambda_{x}((\mathcal{V}_{\mathcal{C}})_{c})<\infty by cocompactness of 𝒞\mathcal{C}. This finishes the proof of the claim. ∎

We have established by now that D[x]GDxD\cap[x]_{\mathcal{R}_{G}}\subseteq D_{x} and, up to a null set, C[x]GCxC\cap[x]_{\mathcal{R}_{G}}\subseteq C_{x} by the claim above. Finally, μ(X(DC))=0\mu(X\setminus(D\sqcup C))=0 implies via Lemma 4.14 λx((D[x]G)(C[x]G))=0\lambda_{x}((D\cap[x]_{\mathcal{R}_{G}})\sqcup(C\cap[x]_{\mathcal{R}_{G}}))=0 for μ\mu-almost all xXx\in X, and therefore λx((D[x]G)Dx)=0=λx((C[x]G)Cx)\lambda_{x}((D\cap[x]_{\mathcal{R}_{G}})\triangle D_{x})=0=\lambda_{x}((C\cap[x]_{\mathcal{R}_{G}})\triangle C_{x}) μ\mu-almost surely. Sets DD and CC thus satisfy the conclusion of the proposition.

For the uniqueness part of the proposition, suppose D,CD,C and D,CD^{\prime},C^{\prime} are two partitions of XX such that

λx(DDx)=0=λx(DDx) and λx(CCx)=0=λx(CCx)\lambda_{x}(D\triangle D_{x})=0=\lambda_{x}(D^{\prime}\triangle D_{x})\textrm{ and }\lambda_{x}(C\triangle C_{x})=0=\lambda_{x}(C^{\prime}\triangle C_{x})

for μ\mu-almost all xXx\in X. One therefore also has μxXλx(DD)=0=λx(CC)\forall^{\mu}x\in X\ \lambda_{x}(D\triangle D^{\prime})=0=\lambda_{x}(C\triangle C^{\prime}), and hence μ(DD)=0\mu(D\triangle D^{\prime})=0 by Lemma 4.14. ∎

We end this section with a natural definition which will be useful for analyzing elements of the full group.

Definition 4.17.

Let GXG\curvearrowright X be a free measure-preserving action of a locally compact Polish group on a standard probability space (X,μ)(X,\mu), and let T[GX]T\in[G\curvearrowright X]. Consider the TT-invariant partition X=DCX=D\sqcup C provided by the Hopf decomposition of TT as per the previous proposition. We say that TT is dissipative when D=XD=X and that TT is conservative when C=XC=X.

When GG is discrete, observe that TT is dissipative if and only if it is aperiodic (all its orbits are infinite), and that TT is conservative if and only if it is periodic (all its orbits are finite).

Example 4.18.

Let us give a general example of dissipative elements of the full group. Let G𝛼XG\overset{\alpha}{\curvearrowright}X be a free measure-preserving action of a locally compact Polish group on a standard probability space (X,μ)(X,\mu). If gGg\in G generates a discrete infinite subgroup, then the element of the full group α(g)\alpha(g) is dissipative. Indeed, the action of α(g)\alpha(g) on each orbit is isomorphic to the gg-action by left translation on GG endowed with its right Haar measure, which is dissipative since it admits a Borel fundamental domain and has only infinite orbits. For instance, if G=G=\mathbb{R}, such a domain is given by the interval [0,g)[0,g) (or (g,0](g,0], if gg is negative).

In Chapter 7, we build an interesting example of a conservative element in the full group of any free measure-preserving flow: its action on each orbit is actually ergodic, and its cocycle is bounded.

4.4. L1\mathrm{L}^{1} full groups and L1\mathrm{L}^{1} orbit equivalence

We now restrict ourselves to the setup where the acting group GG is locally compact Polish and compactly generated, endowed with a maximal compatible norm \left\lVert\cdot\right\rVert (the existence of such a norm for locally compact Polish group is equivalent to being compactly generated, see [Ros21, Cor. 2.8 and Thm. 2.53]). For such a group, as explained in Section 2.3, it makes sense to talk about the associated L1\mathrm{L}^{1} full group by endowing the group with a maximal norm.

The following definition is the natural extension of the notion of L1\mathrm{L}^{1} orbit equivalence to the locally compact case, stated in terms of full groups.

Definition 4.19.

Let α\alpha and β\beta be the respective measure-preserving actions of two locally compact Polish compactly generated groups GG and HH on a standard probability space (X,μ)(X,\mu). We say that α\alpha and β\beta are L1\mathrm{L}^{1} orbit equivalent when there is a measure-preserving transformation SAut(X,μ)S\in\mathrm{Aut}(X,\mu) such that for all gGg\in G and all hHh\in H,

Sα(g)S1[H𝛽X]1 and S1β(h)S[G𝛼X]1.S\alpha(g)S^{-1}\in[H\overset{\beta}{\curvearrowright}X]_{1}\text{ and }S^{-1}\beta(h)S\in[G\overset{\alpha}{\curvearrowright}X]_{1}.

In other words, up to conjugating α\alpha by SS, we have that the image of α\alpha is contained in the L1\mathrm{L}^{1} full group of β\beta, and the image of β\beta is contained in the L1\mathrm{L}^{1} full group of α\alpha.

We now show that L1\mathrm{L}^{1} full groups do remember actions up to L1\mathrm{L}^{1} orbit equivalence as abstract groups. This is done by finding a spacial realization of the isomorphism between the full groups. Such techniques originated in the work of H. Dye [Dye59] and have been greatly generalized by D. H. Fremlin [Fre04, 384D]. We recall that a subgroup GG of Aut(X,μ)\mathrm{Aut}(X,\mu) is said to have many involutions if for any non-trivial measurable AXA\subseteq X there exists a non-trivial involution UGU\in G such that suppUA\operatorname*{supp}U\subseteq A. The group of quasi-measure-preserving transformations of (X,μ)(X,\mu) is denoted by Aut(X,μ)\mathrm{Aut}^{*}(X,\mu).

Theorem 4.20 (Fremlin).

Let G,HG,H be subgroups of Aut(X,μ)\mathrm{Aut}(X,\mu) with many involutions. For any isomorphism ψ:GH\psi:G\to H there exists SAut(X,μ)S\in\mathrm{Aut}^{*}(X,\mu) such that ψ(T)=STS1\psi(T)=STS^{-1} for all TGT\in G.

Proposition 4.21.

If two ergodic measure-preserving actions of locally compact compactly generated Polish groups have isomorphic L1\mathrm{L}^{1} full groups, then they are also L1\mathrm{L}^{1} orbit equivalent.

Proof.

Denote by G𝛼G\overset{\alpha}{\curvearrowright} and H𝛽H\overset{\beta}{\curvearrowright} the two actions on the same standard probability space (X,μ)(X,\mu). Since the L1\mathrm{L}^{1} full groups of ergodic actions have many involutions (see, for example, Lemma 3.6), any isomorphism ψ:[G𝛼X]1[H𝛽X]1\psi:[G\overset{\alpha}{\curvearrowright}X\mkern 1.5mu]_{1}\to[H\overset{\beta}{\curvearrowright}X\mkern 1.5mu]_{1} admits a spatial realization by some SAut(X,μ)S\in\mathrm{Aut}^{*}(X,\mu). The Radon–Nikodym derivative of SμS_{*}\mu with respect to μ\mu is easily seen to be preserved by every element of [H𝛽X]1[H\overset{\beta}{\curvearrowright}X\mkern 1.5mu]_{1}, and hence must be constant by ergodicity. We conclude that SAut(X,μ)S\in\mathrm{Aut}(X,\mu), and therefore by the definition the actions α\alpha and β\beta are L1\mathrm{L}^{1} orbit equivalent. ∎

Remark 4.22.

Similarly to the finitely generated case [LM21, Sec. 8.1], one could define L1\mathrm{L}^{1} full orbit equivalence between actions as equality of L1\mathrm{L}^{1} full groups up to conjugacy, which is a strengthening of L1\mathrm{L}^{1} orbit equivalence (indeed the latter only requires inclusion of each acting group in the L1\mathrm{L}^{1} full group of the other acting group). It would be interesting to have examples of actions which are L1\mathrm{L}^{1} orbit equivalent, but not L1\mathrm{L}^{1} fully orbit equivalent.

We end this section by showing that L1\mathrm{L}^{1} orbit equivalence is equivalent to a stronger definition where we ask that, up to conjugating α\alpha by SS, we moreover have that, on a full measure set X0XX_{0}\subseteq X, the α\alpha and β\beta orbits coincide. This will be a direct consequence of the following proposition. The proof of this proposition is the same as that of [CLM16, Prop. 3.8] which was not stated in the level of generality we need. Since it is short, we reproduce it here.

Proposition 4.23.

Let GG and HH be two locally compact Polish groups acting in a Borel measure-preserving manner on a standard probability space (X,μ)(X,\mu), denote by α\alpha the GG-action and suppose that α(G)[HX]\alpha(G)\leq[H\curvearrowright X]. Then there is a full measure Borel subset X0XX_{0}\subseteq X such that

G(X0×X0)H.\mathcal{R}_{G}\cap\left(X_{0}\times X_{0}\right)\subseteq\mathcal{R}_{H}.
Proof.

Let λ\lambda be the Haar measure on GG. Since α(G)[HX]\alpha(G)\leq[H\curvearrowright X], for all gGg\in G and almost all xXx\in X, we have gxHxgx\in Hx. By Fubini’s theorem, this implies that the Borel set

X0={xX: for λ-almost all gG, we have gxHx}X_{0}=\{x\in X:\text{ for }\lambda\text{-almost all }g\in G,\text{ we have }gx\in Hx\}

has full measure. Now let xX0x\in X_{0}, and let g1Gg_{1}\in G be such that g1xX0g_{1}x\in X_{0}. We want to show that g1xHxg_{1}x\in Hx.

Since xx and g1xg_{1}x are in X0X_{0}, the sets

A={gG:gxHx}andB={gG:gxHg1x}A=\{g\in G:gx\in Hx\}\quad\textrm{and}\quad B=\{g\in G:gx\in Hg_{1}x\}

have full measure and so ABA\cap B has full measure. Take gABg\in A\cap B, and note that gxHxHg1xgx\in Hx\cap Hg_{1}x, so the two orbits HxHx and Hg1xHg_{1}x intersect, hence g1xHxg_{1}x\in Hx. ∎

Corollary 4.24.

Two measure-preserving actions of locally compact compact compactly generated Polish groups GG and HH on a standard probability space (X,μ)(X,\mu) are L1\mathrm{L}^{1} orbit equivalent if and only if they can be conjugated so as to share the same orbits on a full measure Borel subset X0XX_{0}\subseteq X, and for all gGg\in G and hHh\in H there are Borel maps

ρG(g,):X0H and ρH(h,):X0G\rho_{G}(g,\,\cdot\,):X_{0}\to H\text{ and }\rho_{H}(h,\,\cdot\,):X_{0}\to G

such that for all xX0x\in X_{0},

gx=ρG(g,x)x and hx=ρH(h,x)x,g\cdot x=\rho_{G}(g,x)\cdot x\text{ and }h\cdot x=\rho_{H}(h,x)\cdot x,

and finally, if we denote by G\left\lVert\cdot\right\rVert_{G} and H\left\lVert\cdot\right\rVert_{H} maximal norms on GG and HH respectively, then

XρG(g,x)H𝑑μ(x)<+ and XρH(h,x)G𝑑μ(x)<+.\int_{X}\left\lVert\rho_{G}(g,x)\right\rVert_{H}d\mu(x)<+\infty\text{ and }\int_{X}\left\lVert\rho_{H}(h,x)\right\rVert_{G}d\mu(x)<+\infty.
Remark 4.25.

Note that both ρG\rho_{G} and ρH\rho_{H} are actually Borel globally (as maps ρG:G×X0H\rho_{G}:G\times X_{0}\to H and ρH:H×X0G\rho_{H}:H\times X_{0}\to G) as a consequence of the Arsenin selection theorem for Borel sets with KσK_{\sigma} sections and the fact that point stabilizers are closed, a result of D. Miller.

Proof of Corollary 4.24.

It is clear from the definition of L1\mathrm{L}^{1} full groups that the conditions in the corollary are sufficient for L1\mathrm{L}^{1} orbit equivalence. Observe that up to conjugating the two actions, they do share the same full group. Since L1\mathrm{L}^{1} full groups contain the acting groups, we can apply Proposition 4.23 twice and get the desired full measure Borel subset X0X_{0} restricted to which orbits coincide. The remaining statements are then direct consequences of the definition of L1\mathrm{L}^{1} full groups. ∎

We will see in the final chapter that there are free ergodic \mathbb{R}-flows which are not L1\mathrm{L}^{1} orbit equivalent. This will be done by relating L1\mathrm{L}^{1} orbit equivalence to flip-Kakutani equivalence. In the discrete amenable case, an important result of Austin shows that entropy is preserved by L1\mathrm{L}^{1} orbit equivalence [Aus16]. We wonder what happens in the general locally compact setup.

Question 4.26.

Let GG be an amenable non-discrete non-compact compactly generated locally compact Polish group. Are there two measure-preserving ergodic actions of GG which are not L1\mathrm{L}^{1} orbit equivalent?

Chapter 5 Derived L1\mathrm{L}^{1} full groups for locally compact amenable groups

Given a measure-preserving action of a normed Polish group (G,)(G,\left\lVert\cdot\right\rVert) on (X,μ)(X,\mu), the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X]_{1}) of the action is by definition the closure in [GX]1[G\curvearrowright X]_{1} of the group generated by commutators, i.e., elements of the form TUT1U1TUT^{-1}U^{-1}, where T,U[GX]1T,U\in[G\curvearrowright X]_{1}. Provided the GG-action is aperiodic, the latter can be described in three different ways using the fact that [GX]1[G\curvearrowright X]_{1} is induction friendly, as explained in the Section 3.2 (see Corollary 3.15):

  • D([GX]1)D([G\curvearrowright X]_{1}) is the closure of the group generated by involutions;

  • D([GX]1)D([G\curvearrowright X]_{1}) is the closure of the group generated by 33-cycles;

  • D([GX]1)D([G\curvearrowright X]_{1}) is the closure of the group generated by periodic elements.

In particular, all periodic elements of [GX]1[G\curvearrowright X]_{1} actually belong to D([GX]1)D([G\curvearrowright X]_{1}).

Compared to the previous chapter, we impose one further restriction on the acting group, and consider actions of a locally compact amenable Polish normed group (G,)(G,\lVert\cdot\rVert). Appendix G of [BdlHV08] contains an excellent review of amenability for locally compact Polish groups. As before, we fix a measure-preserving action GXG\curvearrowright X on a standard probability space (X,μ)(X,\mu), and let D:G0D:\mathcal{R}_{G}\to\mathbb{R}^{\geq 0} denote the family of metrics induced onto the orbits by the norm.

In Section 5.1, we will first exhibit a dense increasing chain of subgroups in D([GX]1)D([G\curvearrowright X]_{1}). This dense chain is used in the two remaining sections. In Section 5.2, we show that amenability of the group is reflected in whirly amenability of D([GX]1)D([G\curvearrowright X]_{1}), while in Section 5.3 we prove by a Baire category argument that D([GX]1)D([G\curvearrowright X]_{1}) has a dense 22-generated subgroup.

5.1. Dense chain of subgroups

An equivalence relation G\mathcal{R}\subseteq\mathcal{R}_{G} is said to be uniformly bounded if there is M>0M>0 and XXX^{\prime}\subseteq X such that μ(XX)=0\mu(X\setminus X^{\prime})=0 and sup(x1,x2)D(x1,x2)M\sup_{(x_{1},x_{2})\in\mathcal{R}^{\prime}}D(x_{1},x_{2})\leq M, where =X×X\mathcal{R}^{\prime}=\mathcal{R}\cap X^{\prime}\times X^{\prime}.

Lemma 5.1.

Let (G,)(G,\lVert\cdot\rVert) be a locally compact amenable Polish normed group acting on a standard probability space (X,μ)(X,\mu). There exists a sequence of cross-sections 𝒞n\mathcal{C}_{n}, nn\in\mathbb{N}, and tessellations 𝒲n\mathcal{W}_{n} over 𝒞n\mathcal{C}_{n} such that for all nn\in\mathbb{N}

  1. (1)

    𝒲n𝒲n+1\mathcal{R}_{\mathcal{W}_{n}}\subseteq\mathcal{R}_{\mathcal{W}_{n+1}} and k𝒲k=G\bigcup_{k\in\mathbb{N}}\mathcal{R}_{\mathcal{W}_{k}}=\mathcal{R}_{G} (up to a null set);

  2. (2)

    𝒲n\mathcal{R}_{\mathcal{W}_{n}} is uniformly bounded.

Proof.

Let 𝒞\mathcal{C} be a cocompact cross-section, 𝒱𝒞\mathcal{V}_{\mathcal{C}} be the Voronoi tessellation over 𝒞\mathcal{C}, π𝒱𝒞:X𝒞\pi_{\mathcal{V}_{\mathcal{C}}}:X\to\mathcal{C} be the associated reduction, and ν=(π𝒱𝒞)μ\nu=(\pi_{\mathcal{V}_{\mathcal{C}}})_{*}\mu be the push-forward measure on 𝒞\mathcal{C}. Recall that 𝒱𝒞\mathcal{R}_{\mathcal{V}_{\mathcal{C}}} is uniformly bounded, since 𝒞\mathcal{C} is cocompact. Let EE be the equivalence relation obtained by restricting G\mathcal{R}_{G} onto 𝒞\mathcal{C}. By a theorem of A. Connes, J. Feldman, and B. Weiss [CFW81], EE is hyperfinite on an invariant set of ν\nu-full measure. Throwing away a GG-invariant null set, we may write E=nEnE=\bigcup_{n}E_{n}, where (En)n(E_{n})_{n\in\mathbb{N}} is an increasing sequence of Borel equivalence relations with finite classes. For m,nm,n\in\mathbb{N}, define An,mA_{n,m} to be the set of points in the cross-section whose EnE_{n}-class is bounded in diameter by mm:

An,m={c𝒞:D(c1,c2)m for all c1,c2𝒞 such that c1Enc and c2Enc}.A_{n,m}=\bigl{\{}c\in\mathcal{C}:D(c_{1},c_{2})\leq m\textrm{ for all }c_{1},c_{2}\in\mathcal{C}\textrm{ such that }c_{1}E_{n}c\textrm{ and }c_{2}E_{n}c\bigr{\}}.

Note that the sets An,mA_{n,m} are EnE_{n}-invariant, nested, and mAn,m=𝒞\bigcup_{m}A_{n,m}=\mathcal{C} for every nn\in\mathbb{N}. Pick mnm_{n} so large as to ensure ν(𝒞An,mn)<2n\nu(\mathcal{C}\setminus A_{n,m_{n}})<2^{-n} and let Bn=knAk,mkB_{n}=\bigcap_{k\geq n}A_{k,m_{k}}. The sets BnB_{n} are EnE_{n}-invariant, increasing, and limnν(Bn)=ν(𝒞)\lim_{n}\nu(B_{n})=\nu(\mathcal{C}). Define equivalence relations FnF_{n} on 𝒞\mathcal{C} by setting c1Fnc2c_{1}F_{n}c_{2} whenever c1=c2c_{1}=c_{2} or c1,c2Bnc_{1},c_{2}\in B_{n} and c1Enc2c_{1}E_{n}c_{2}. Note that D(c1,c2)mnD(c_{1},c_{2})\leq m_{n} whenever c1Fnc2c_{1}F_{n}c_{2}. Let 𝒞n𝒞\mathcal{C}_{n}\subseteq\mathcal{C} be a Borel transversal for FnF_{n} and define 𝒲n={(c,x)𝒞n×X:cFnπ𝒱𝒞(x)}\mathcal{W}_{n}=\{(c,x)\in\mathcal{C}_{n}\times X:cF_{n}\pi_{\mathcal{V}_{\mathcal{C}}}(x)\}. It is straightforward to check that each 𝒲n\mathcal{W}_{n} is a tessellation over 𝒞n\mathcal{C}_{n}, and equivalence relations 𝒲n\mathcal{R}_{\mathcal{W}_{n}} satisfy the conclusions of the lemma. ∎

The equivalence relations 𝒲n\mathcal{R}_{\mathcal{W}_{n}} produced by Lemma 5.1 give rise to a nested chain of groups [𝒲0][𝒲1][\mathcal{R}_{\mathcal{W}_{0}}\mkern 1.5mu]\leq[\mathcal{R}_{\mathcal{W}_{1}}\mkern 1.5mu]\leq\cdots. Some basic facts about such groups can be found in Appendix C.2. The following lemma establishes that such a chain is dense in the derived L1\mathrm{L}^{1} full group.

Lemma 5.2.

Let (G,)(G,\lVert\cdot\rVert) be a locally compact amenable Polish normed group acting on a standard probability space (X,μ)(X,\mu) and let (n)n(\mathcal{R}_{n})_{n\in\mathbb{N}} be a sequence of equivalence relations as in Lemma 5.1. If the action is aperiodic, then the union n[n]\bigcup_{n}[\mathcal{R}_{n}] is contained in the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) and is dense therein.

Proof.

By definition, [n][\mathcal{R}_{n}\mkern 1.5mu] is a subgroup of [G][\mathcal{R}_{G}\mkern 1.5mu]. Since equivalence relations n\mathcal{R}_{n} are uniformly bounded, we actually have [n][GX]1[\mathcal{R}_{n}\mkern 1.5mu]\leq[G\curvearrowright X\mkern 1.5mu]_{1}, and the topology induced by the L1\mathrm{L}^{1} metric on [n][\mathcal{R}_{n}\mkern 1.5mu] coincides with the topology induced from [G][\mathcal{R}_{G}\mkern 1.5mu]. Moreover, in view of Proposition C.7, [n][\mathcal{R}_{n}\mkern 1.5mu] is topologically generated by periodic transformations, so we actually have [n]D([GX]1)[\mathcal{R}_{n}\mkern 1.5mu]\leq D([G\curvearrowright X\mkern 1.5mu]_{1}) as a consequence of Lemma 3.10 and Corollary 3.15.

It remains to verify that the union n[n]\bigcup_{n}[\mathcal{R}_{n}\mkern 1.5mu] is dense in D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}). To this end, recall that by Corollary 3.15 the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is topologically generated by involutions. So let UD([GX]1)U\in D([G\curvearrowright X\mkern 1.5mu]_{1}) be an involution and set Xn={xX:(x,U(x))n}X_{n}=\{x\in X:(x,U(x))\in\mathcal{R}_{n}\}, nn\in\mathbb{N}. Note that XnX_{n} is UU-invariant since UU is an involution. Moreover, μ(Xn)1\mu(X_{n})\to 1 as nn=G\bigcup_{n}\mathcal{R}_{n}=\mathcal{R}_{G}, and thus the induced transformations UXn[n]U_{X_{n}}\in[\mathcal{R}_{n}\mkern 1.5mu] converge to UU in the topology of [GX]1[G\curvearrowright X\mkern 1.5mu]_{1}. We conclude that n[n]\bigcup_{n}[\mathcal{R}_{n}\mkern 1.5mu] is dense in the derived L1\mathrm{L}^{1} full group. ∎

Corollary 5.3.

Let (G,)(G,\lVert\cdot\rVert) be a locally compact amenable Polish normed group acting on a standard probability space (X,μ)(X,\mu). Suppose that almost every orbit of the action is uncountable. There exists a chain H0H1D([GX]1)H_{0}\leq H_{1}\leq\cdots\leq D([G\curvearrowright X\mkern 1.5mu]_{1}) of closed subgroups such that nHn\bigcup_{n}H_{n} is dense in D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}), and each HnH_{n} is isomorphic to L0(Yn,νn,Aut([0,1],λ))\mathrm{L}^{0}(Y_{n},\nu_{n},\mathrm{Aut}([0,1],\lambda)) for some standard Lebesgue space (Yn,νn)(Y_{n},\nu_{n}). If moreover each orbit of the action has measure zero, then one can assume that all (Yn,νn)(Y_{n},\nu_{n}) are atomless and each HnH_{n} is isomorphic to L0([0,1],λ,Aut([0,1],λ))\mathrm{L}^{0}([0,1],\lambda,\mathrm{Aut}([0,1],\lambda)).

Proof.

Apply Lemmas 5.1 and 5.2 to get a dense chain of subgroups [0][1]D([GX]1)[\mathcal{R}_{0}\mkern 1.5mu]\leq[\mathcal{R}_{1}\mkern 1.5mu]\leq\cdots\leq D([G\curvearrowright X\mkern 1.5mu]_{1}) and use Corollary C.14 to deduce that each [n][\mathcal{R}_{n}\mkern 1.5mu] has the desired form. ∎

Corollary 5.4.

Let (G,)(G,\lVert\cdot\rVert) be a locally compact amenable Polish normed group acting on a standard probability space (X,μ)(X,\mu). If the action is aperiodic, then the set of periodic elements is dense in the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X]_{1}).

Proof.

Consider a chain of subgroups [n][\mathcal{R}_{n}\mkern 1.5mu] given by Lemma 5.2. Periodic elements are dense in these groups for their natural topology (see Proposition C.7 and the discussion preceding it). These topologies are compatible with the standard Borel structure of Aut(X,μ)\mathrm{Aut}(X,\mu) induced by the weak topology and therefore must refine the L1\mathrm{L}^{1} topology by the standard automatic continuity arguments [BK96, Sec. 1.6]. Hence periodic elements are dense in all of D([GX]1)D([G\curvearrowright X]_{1}), as claimed. ∎

Corollary 5.4 together with Proposition 3.24 show that the L1\mathrm{L}^{1} norm is maximal on derived L1\mathrm{L}^{1} full groups of aperiodic measure-preserving actions of locally compact amenable Polish normed groups (see Section 2.3 for a short reminder on maximality of norms). In particular, such groups are boundedly generated by [Ros22, Thm.  2.53].

Theorem 5.5.

Let (G,)(G,\lVert\cdot\rVert) be a locally compact amenable Polish normed group acting on a standard probability space (X,μ)(X,\mu). If the action is aperiodic, then the L1\mathrm{L}^{1} norm is maximal on the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X]_{1}).

We do not know if the amenability hypothesis can be removed, even when GG is discrete and the action is free.

5.2. Whirly amenability

Lemma 5.2 is a powerful tool to deduce various dynamical properties of derived L1\mathrm{L}^{1} full groups. Recall that a Polish group GG is said to be whirly amenable if it is amenable and for any continuous action of GG on a compact space any invariant measure is supported on the set of fixed points of the action. In particular, each such action has to have some fixed points, so whirly amenable groups are extremely amenable.

Proposition 5.6.

Let \mathcal{R} be a smooth measurable equivalence relation on a standard Lebesgue space (X,μ)(X,\mu). If μ\mu is atomless, then the full group [][\mathcal{R}\mkern 1.5mu] is whirly amenable.

Proof.

In view of Proposition C.6, the full group [][\mathcal{R}\mkern 1.5mu] is isomorphic to

L0([0,1],λ,Aut([0,1],λ))ϵ0×Aut([0,1],λ)κ0×n1L0([0,1],λ,𝔖n)ϵ0,\mathrm{L}^{0}([0,1],\lambda,\mathrm{Aut}([0,1],\lambda))^{\epsilon_{0}}\times\mathrm{Aut}([0,1],\lambda)^{\kappa_{0}}\times\prod_{n\geq 1}\mathrm{L}^{0}([0,1],\lambda,\mathfrak{S}_{n})^{\epsilon_{0}},

where 𝔖n\mathfrak{S}_{n} is the group of permutations of an nn-element set, and ϵn{0,1}\epsilon_{n}\in\{0,1\}, κ00\kappa_{0}\leq\aleph_{0}. Since a product of whirly amenable groups is whirly amenable, it suffices to show that the groups appearing in the decomposition above, namely L0([0,1],λ,Aut([0,1],λ))\mathrm{L}^{0}([0,1],\lambda,\mathrm{Aut}([0,1],\lambda)), Aut([0,1],λ)\mathrm{Aut}([0,1],\lambda), and L0([0,1],λ,𝔖n)\mathrm{L}^{0}([0,1],\lambda,\mathfrak{S}_{n}), n1n\geq 1, are whirly amenable.

The group Aut([0,1],λ)\mathrm{Aut}([0,1],\lambda) is whirly amenable by [GP02] (it is, in fact, a so-called Levy group). Finally, we apply a theorem of V. Pestov and F. M. Schneider [PS17], which asserts that a group L0([0,1],λ,G)\mathrm{L}^{0}([0,1],\lambda,G) is whirly amenable if and only if GG is amenable. This readily implies whirly amenability of L0([0,1],λ,𝔖n)\mathrm{L}^{0}([0,1],\lambda,\mathfrak{S}_{n}) and L0([0,1],λ,Aut([0,1],λ))\mathrm{L}^{0}([0,1],\lambda,\mathrm{Aut}([0,1],\lambda)). ∎

Remark 5.7.

The assumption of μ\mu being atomless cannot be omitted in the proposition above. Indeed, [][\mathcal{R}\mkern 1.5mu] will factor onto 𝔖n\mathfrak{S}_{n} for some n2n\geq 2, as long as an \mathcal{R}-class contains at least 22 atoms of μ\mu of the same measure. However, if all μ\mu-atoms within each \mathcal{R}-class have distinct measures, then the restriction of [][\mathcal{R}\mkern 1.5mu] onto the atomic part of XX is trivial, which suffices to conclude the whirly amenability of the group [][\mathcal{R}\mkern 1.5mu].

Theorem 5.8.

Let GXG\curvearrowright X be a measure-preserving action of an amenable locally compact Polish normed group on a standard probability space (X,μ)(X,\mu). If the action is aperiodic, then the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is whirly amenable. In particular, [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is amenable.

Proof.

Lemma 5.2 shows that D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) has an increasing dense chain of subgroups HnH_{n} of the form [n][\mathcal{R}_{n}\mkern 1.5mu], where n\mathcal{R}_{n} are smooth measurable equivalence relations on XX. Proposition 5.6 applies and shows that groups HnH_{n} are whirly amenable. The latter is sufficient to conclude whirly amenability of D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}), as any invariant measure for the action of the derived group is also invariant for the induced HnH_{n} actions, hence it has to be supported on the intersection of fixed points of all HnH_{n}, which coincides with the set of fixed points for the action of D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}).

The fact that [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is amenable now follows from the fact that every abelian group is amenable, and every amenable extension of an amenable group must itself be amenable (for instance, see [BdlHV08, Prop. G.2.2]). ∎

Remark 5.9.

Note that in general [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is not extremely amenable. For flows, it factors onto \mathbb{R} via the index map (see Chapter 6) and \mathbb{R} admits continuous actions on compact spaces without fixed points, so [X]1[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} is not extremely amenable (and in particular, it is not whirly amenable) for any free measure-preserving flow.

Corollary 5.10.

Let GXG\curvearrowright X be a free measure-preserving action of a unimodular locally compact Polish group on a standard probability space (X,μ)(X,\mu). The following are equivalent:

  1. (1)

    GG is amenable.

  2. (2)

    [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is amenable.

  3. (3)

    The derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is amenable.

  4. (4)

    The derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is extremely amenable.

  5. (5)

    The derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is whirly amenable.

Proof.

We established the implication (1)\implies(5) in Theorem 5.8. The chain of implications (5)\implies(4)\implies(3) is straightforward, and (3)\implies(2) follows from the stability of amenability under group extensions, which was already discussed in Theorem 5.8.

For the last implication (2)\implies (1), first recall that the orbit full group of the action is generated by involutions. It follows that the orbit full group is topologically generated by involutions whose cocycles are integrable (actually, one can even ask that the cocycles are bounded). In particular, the L1\mathrm{L}^{1} full group [GX]1[G\curvearrowright X\mkern 1.5mu]_{1} is dense in the orbit full group, and so assuming (2) we conclude that the orbit full group [GX][G\curvearrowright X\mkern 1.5mu] is amenable. The amenability of GG then follows from [CLM18, Thm. 5.1]. ∎

Remark 5.11.

We have to require unimodularity in order to be able to apply [CLM18, Thm. 5.1]. It seems likely that the unimodularity hypothesis can be dropped in this result, but we do not pursue this question further.

5.3. Topological generators

We now concern ourselves with the question of determining the topological rank of derived L1\mathrm{L}^{1} full groups. Our approach will be based on the dense chain of subgroups established in Corollary 5.3, and the first step is to study the topological rank of the group L0([0,1],Aut([0,1]))\mathrm{L}^{0}([0,1],\mathrm{Aut}([0,1])).

Let (Y,ν)(Y,\nu) and (Z,λ)(Z,\lambda) be standard Lebesgue spaces. Consider the product space Y×ZY\times Z equipped with the product measure ν×λ\nu\times\lambda and let \mathcal{R} be the product of the discrete equivalence relation on YY and the anti-discrete on ZZ; in other words, (y1,z1)(y2,z2)(y_{1},z_{1})\mathcal{R}(y_{2},z_{2}) if and only if y1=y2y_{1}=y_{2}. As discussed in Appendix C.1, the following two groups are one and the same:

  1. (1)

    the full group [][\mathcal{R}\mkern 1.5mu];

  2. (2)

    the topological group L0(Y,ν,Aut(Z,λ))\mathrm{L}^{0}(Y,\nu,\mathrm{Aut}(Z,\lambda)).

Suppose that (Z,λ)(Z,\lambda) is atomless. Pick a hyperfinite ergodic equivalence relation EE on ZZ so that APER(Z)[E]\mathrm{APER}(Z)\cap[E\mkern 1.5mu] is dense in Aut(Z,λ)\mathrm{Aut}(Z,\lambda), where APER(Z)\mathrm{APER}(Z) stands for the collection of aperiodic automorphisms of ZZ. Set 0=idY×E\mathcal{R}_{0}=\mathrm{id}_{Y}\times E to be the equivalence relation on Y×ZY\times Z given by (y1,z1)0(y2,z2)(y_{1},z_{1})\mathcal{R}_{0}(y_{2},z_{2}) whenever y1=y2y_{1}=y_{2} and z1Ez2z_{1}Ez_{2}. A standard application of the Jankov-von Neumann uniformization theorem yields the following lemma.

Lemma 5.12.

APER(Y×Z)[0]\mathrm{APER}(Y\times Z)\cap[\mathcal{R}_{0}\mkern 1.5mu] is dense in []L0(Y,ν,Aut([0,1],λ))[\mathcal{R}\mkern 1.5mu]\simeq\mathrm{L}^{0}(Y,\nu,\mathrm{Aut}([0,1],\lambda)).

Our first goal is to establish that the topological rank of [][\mathcal{R}\mkern 1.5mu] is 22. We do so by first verifying this under the assumption that (Y,ν)(Y,\nu) is atomless, and then deducing the general case.

We say that a topological group GG is generically kk-generated, kk\in\mathbb{N}, if the set of kk-tuples (g1,,gk)Gk(g_{1},\ldots,g_{k})\in G^{k} that generate a dense subgroup of GG is dense in GkG^{k}. Note that the set of such tuples is always a GδG_{\delta} set, so if GG is generically kk-generated, then a comeager set of kk-tuples generates a dense subgroup of GG.

Proposition 5.13.

Suppose that (Y,ν)(Y,\nu) is atomless. The group [][\mathcal{R}\mkern 1.5mu] is generically 22-generated.

Proof.

By [LM16, Thm 5.1], the set of pairs

(S,T)(APER(Y×Z)[0])×[0](S,T)\in(\mathrm{APER}(Y\times Z)\cap[\mathcal{R}_{0}\mkern 1.5mu])\times[\mathcal{R}_{0}\mkern 1.5mu]

such that S,T¯=[0]\overline{\langle S,T\rangle}=[\mathcal{R}_{0}\mkern 1.5mu] is dense GδG_{\delta} for the uniform topology. In view of Lemma 5.12, this implies that [][\mathcal{R}\mkern 1.5mu] is generically 22-generated. ∎

Lemma 5.14.

For all topological groups GG and HH one has

rk(G×H)max{rk(G),rk(H)}.\mathrm{rk}(G\times H)\geq\max\{\mathrm{rk}(G),\mathrm{rk}(H)\}.

If G×HG\times H is generically kk-generated, then so are GG and HH as well.

Proof.

The inequality on ranks is immediate from the trivial observation that if (g1,h1),,(gk,hk)\langle(g_{1},h_{1}),\ldots,(g_{k},h_{k})\rangle is dense in G×HG\times H, then g1,,gk\langle g_{1},\ldots,g_{k}\rangle is dense in GG and h1,,hk\langle h_{1},\ldots,h_{k}\rangle is dense in HH.

Suppose G×HG\times H is generically kk-generated, pick an open set UGkU\subseteq G^{k} and note that U×HkU\times H^{k} naturally corresponds to an open subset of (G×H)k(G\times H)^{k} via the isomorphism (G×H)kGk×Hk(G\times H)^{k}\simeq G^{k}\times H^{k}. Since G×HG\times H is generically kk-generated, there is a tuple (gi,hi)i=1k(G×H)k(g_{i},h_{i})_{i=1}^{k}\in(G\times H)^{k} that generates a dense subgroup and (gi,hi)i=1kU×Hk(g_{i},h_{i})_{i=1}^{k}\in U\times H^{k}. We conclude that (gi)i=1kU(g_{i})_{i=1}^{k}\in U generates a dense subgroup of GG and the lemma follows. ∎

Lemma 5.15.

For any separable topological group GG

rk(L0([0,1],λ,G))=rk(L0([0,1],λ,G)×G).\mathrm{rk}(\mathrm{L}^{0}([0,1],\lambda,G))=\mathrm{rk}\bigl{(}\mathrm{L}^{0}([0,1],\lambda,G)\times G^{\mathbb{N}}\bigr{)}.

If L0([0,1],λ,G)\mathrm{L}^{0}([0,1],\lambda,G) is generically kk-generated for some kk\in\mathbb{N}, then so is the group L0([0,1],λ,G)×G\mathrm{L}^{0}([0,1],\lambda,G)\times G^{\mathbb{N}}.

Proof.

In view of Lemma 5.14, rk(L0([0,1],λ,G))rk(L0([0,1],λ,G)×G)\mathrm{rk}(\mathrm{L}^{0}([0,1],\lambda,G))\leq\mathrm{rk}\bigl{(}\mathrm{L}^{0}([0,1],\lambda,G)\times G^{\mathbb{N}}\bigr{)}, and, since the group GG is separable, we only need to consider the case when the rank rk(L0([0,1],λ,G))\mathrm{rk}(\mathrm{L}^{0}([0,1],\lambda,G)) is finite.

It is notationally convenient to shrink the interval and work with the group L0([0,1/2],λ,G)×G\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}} instead as it can naturally be viewed as a closed subgroup of L0([0,1],λ,G)\mathrm{L}^{0}([0,1],\lambda,G) via the identification f×(gi)iζf\times(g_{i})_{i\in\mathbb{N}}\mapsto\zeta, where

ζ(t)={f(t)if 0t<1/2,giif 12i1t<12i2 for i.\zeta(t)=\begin{cases}f(t)&\textrm{if $0\leq t<1/2$},\\ g_{i}&\textrm{if $1-2^{-i-1}\leq t<1-2^{-i-2}$ for $i\in\mathbb{N}$}.\end{cases}

Pick families (ξl)l(\xi_{l})_{l\in\mathbb{N}} dense in L0([0,1/2],λ,G)\mathrm{L}^{0}([0,1/2],\lambda,G), and (hm)m(h_{m})_{m\in\mathbb{N}} dense in GG.

Let us call a function α:\alpha:\mathbb{N}\to\mathbb{N} a multi-index if α(i)=0\alpha(i)=0 for all but finitely many ii\in\mathbb{N}. We use <\mathbb{N}^{<\mathbb{N}} to denote the set of all multi-indices. Given α<\alpha\in\mathbb{N}^{<\mathbb{N}}, let hα=(hα(i))ih_{\alpha}=(h_{\alpha(i)})_{i\in\mathbb{N}} be an element of GG^{\mathbb{N}}. Note that {hα:α<}\{h_{\alpha}:\alpha\in\mathbb{N}^{<\mathbb{N}}\} is dense in GG^{\mathbb{N}} and thus {ξl×hα:l,α<}\{\xi_{l}\times h_{\alpha}:l\in\mathbb{N},\alpha\in\mathbb{N}^{<\mathbb{N}}\} is a dense family in L0([0,1/2],λ,G)×G\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}}.

Pick a tuple f1,,fkL0([0,1],λ,G)f_{1},\ldots,f_{k}\in\mathrm{L}^{0}([0,1],\lambda,G) that generates a dense subgroup. For each pair (l,α)×<(l,\alpha)\in\mathbb{N}\times\mathbb{N}^{<\mathbb{N}}, there exists a sequence of reduced words (wnl,α)n(w_{n}^{l,\alpha})_{n\in\mathbb{N}} in the free group on kk generators such that wnl,α(f1,,fk)w^{l,\alpha}_{n}(f_{1},\ldots,f_{k}) converges to ξl×hα\xi_{l}\times h_{\alpha} in measure. By passing to a subsequence, we may assume that wnl,α(f1,,fk)ξl×hαw^{l,\alpha}_{n}(f_{1},\ldots,f_{k})\to\xi_{l}\times h_{\alpha} pointwise almost surely. In other words, the set

Pl,α={t[0,1]:wnl,α(f1,,fk)(t)(ξl×hα)(t)}P_{l,\alpha}=\bigl{\{}t\in[0,1]:w^{l,\alpha}_{n}(f_{1},\ldots,f_{k})(t)\to(\xi_{l}\times h_{\alpha})(t)\bigr{\}}

has Lebesgue measure 11 for each (l,α)×<(l,\alpha)\in\mathbb{N}\times\mathbb{N}^{<\mathbb{N}}, and hence so does the set

P=lα<Pl,α.P=\bigcap_{l\in\mathbb{N}}\bigcap_{\alpha\in\mathbb{N}^{<\mathbb{N}}}P_{l,\alpha}.

Pick some tjP[12j1,12j2)t_{j}\in P\cap[1-2^{-j-1},1-2^{-j-2}), jj\in\mathbb{N}, and set

f~i(t)={fi(t)for 0t<1/2,fi(tj)for 12j1t<12j2 for j.\tilde{f}_{i}(t)=\begin{cases}f_{i}(t)&\textrm{for $0\leq t<1/2$},\\ f_{i}(t_{j})&\textrm{for $1-2^{-j-1}\leq t<1-2^{-j-2}$ for $j\in\mathbb{N}$}.\end{cases}

Elements f~i\tilde{f}_{i} naturally belong to L0([0,1/2],λ,G)×G\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}}, and we claim that they generate a dense subgroup therein, witnessing rk(L0([0,1/2],λ,G)×G)k\mathrm{rk}(\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}})\leq k. To this end recall that wnl,α(f1,,fk)ξl×hαw_{n}^{l,\alpha}(f_{1},\ldots,f_{k})\to\xi_{l}\times h_{\alpha} pointwise almost surely. In particular,

wnl,α(f1,,fk)[0,1/2]ξl×hα[0,1/2]w_{n}^{l,\alpha}(f_{1},\ldots,f_{k})\restriction_{[0,1/2]}\to\xi_{l}\times h_{\alpha}\restriction_{[0,1/2]}

in measure and, for each jj\in\mathbb{N},

wnl,α(f1,,fk)(tj)(ξl×hα)(tj)=hα(j)w_{n}^{l,\alpha}(f_{1},\ldots,f_{k})(t_{j})\to(\xi_{l}\times h_{\alpha})(t_{j})=h_{\alpha(j)}

is guaranteed by choosing tjPt_{j}\in P. We conclude that

wnl,α(f~1,,f~k)ξl×hαw_{n}^{l,\alpha}(\tilde{f}_{1},\ldots,\tilde{f}_{k})\to\xi_{l}\times h_{\alpha}

in L0([0,1/2],λ,G)×G\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}}, and therefore

rk(L0([0,1/2],λ,G)×G)k.\mathrm{rk}(\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}})\leq k.

Finally, suppose that L0([0,1],λ,G)\mathrm{L}^{0}([0,1],\lambda,G) is generically kk-generated. Choose open sets UiL0([0,1/2],λ,G)×GU_{i}\subseteq\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}},1ik1\leq i\leq k. Shrinking them if necessary, we may assume that all UiU_{i} have the form Ui=A0i×A1i××Ani×GU_{i}=A_{0}^{i}\times A_{1}^{i}\times\cdots\times A_{n}^{i}\times G^{\mathbb{N}}, where A0iA_{0}^{i} is open in L0([0,1/2],λ,G)\mathrm{L}^{0}([0,1/2],\lambda,G), and AjiA_{j}^{i}, j1j\geq 1, are open in GG.

Pick ViL0([0,1],λ,G)V_{i}\subseteq\mathrm{L}^{0}([0,1],\lambda,G), 1ik1\leq i\leq k, to consist of those functions ff satisfying f|[0,1/2]A0f|_{[0,1/2]}\in A_{0} and f(t)Ajf(t)\in A_{j} for all t[12j1,12j2)t\in[1-2^{-j-1},1-2^{-j-2}), 1jn1\leq j\leq n. Note that ViL0([0,1/2],λ,G)×G=UiV_{i}\cap\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}}=U_{i}.

Since L0([0,1],λ,G)\mathrm{L}^{0}([0,1],\lambda,G) is assumed to be generically kk-generated, there is a tuple (f1,,fk)(f_{1},\ldots,f_{k}) generating a dense subgroup in L0([0,1],λ,G)\mathrm{L}^{0}([0,1],\lambda,G) such that fiVif_{i}\in V_{i} for each ii. Running the above construction, we get a tuple (f~1,,f~k)L0([0,1/2],λ,G)×G(\tilde{f}_{1},\ldots,\tilde{f}_{k})\in\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}} such that f~iUi\tilde{f}_{i}\in U_{i}, 1ik1\leq i\leq k, whence L0([0,1/2],λ,G)×G\mathrm{L}^{0}([0,1/2],\lambda,G)\times G^{\mathbb{N}} is generically kk-generated. ∎

Lemma 5.15 remains valid if we take the product with a finite power of GG, which follows from Lemma 5.14.

Corollary 5.16.

For any separable topological group GG and any mm\in\mathbb{N} one has

rk(L0([0,1],λ,G))=rk(L0([0,1],λ,G))×Gm.\mathrm{rk}(\mathrm{L}^{0}([0,1],\lambda,G))=\mathrm{rk}(\mathrm{L}^{0}([0,1],\lambda,G))\times G^{m}.

If rk(L0([0,1],λ,G))\mathrm{rk}(\mathrm{L}^{0}([0,1],\lambda,G)) is generically kk-generated for some kk\in\mathbb{N}, then so is the group L0([0,1],λ,G)×Gm\mathrm{L}^{0}([0,1],\lambda,G)\times G^{m}.

We may now strengthen Proposition 5.13 by dropping the assumption on (Y,ν)(Y,\nu) being atomless.

Proposition 5.17.

Let (Y,ν)(Y,\nu) be a standard Lebesgue space and (Z,λ)(Z,\lambda) be a standard probability space. The topological group L0(Y,ν,Aut(Z,λ))\mathrm{L}^{0}(Y,\nu,\mathrm{Aut}(Z,\lambda)) is generically 22-generated.

Proof.

Let YaY_{a} be the set of atoms of YY, put Y0=YYaY_{0}=Y\setminus Y_{a} and ν0=νY0\nu_{0}=\nu\restriction_{Y_{0}}. The group L0(Y,ν,Aut(Z,λ))\mathrm{L}^{0}(Y,\nu,\mathrm{Aut}(Z,\lambda)) is naturally isomorphic to

L0(Y0,ν0,Aut(Z,λ))×Aut(Z,λ)|Ya|.\mathrm{L}^{0}(Y_{0},\nu_{0},\mathrm{Aut}(Z,\lambda))\times\mathrm{Aut}(Z,\lambda)^{|Y_{a}|}.

An application of Proposition 5.13 together with Lemma 5.15 or Corollary 5.16 (depending on whether YaY_{a} is infinite or not) finishes the proof. ∎

Proposition 5.18.

Let GG be a Polish group and let H0H1GH_{0}\leq H_{1}\leq\cdots\leq G be a dense chain of Polish subgroups, nHn¯=G\overline{\bigcup_{n}H_{n}}=G. If each HnH_{n} is generically kk-generated, then GG is generically kk-generated.

Proof.

We need to show that for any open UGkU\subseteq G^{k} and any open VGV\subseteq G there is a tuple (g1,,gk)U(g_{1},\ldots,g_{k})\in U such that g1,,gkV\langle g_{1},\ldots,g_{k}\rangle\cap V\neq\varnothing. Since groups HnH_{n} are nested and nHn\bigcup_{n}H_{n} is dense in GG, there is nn so large that UHnkU\cap H_{n}^{k}\neq\varnothing and VHnV\cap H_{n}\neq\varnothing. It remains to use the fact that HnH_{n} is generically kk-generated to find the required tuple. ∎

Theorem 5.19.

Let GXG\curvearrowright X be a measure-preserving action of a locally compact amenable Polish normed group on a standard probability space (X,μ)(X,\mu). If almost every orbit of the action is uncountable, then the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is generically 22-generated.

Proof.

In view of Corollary 5.3, there is a chain of subgroups

H0H1D([GX]1),nHn¯=D([GX]1),H_{0}\leq H_{1}\leq\cdots\leq D([G\curvearrowright X\mkern 1.5mu]_{1}),\quad\overline{\bigcup_{n}H_{n}}=D([G\curvearrowright X\mkern 1.5mu]_{1}),

where each HnH_{n} is isomorphic to L0(Yn,νn,Aut([0,1],λ))\mathrm{L}^{0}(Y_{n},\nu_{n},\mathrm{Aut}([0,1],\lambda)) for some standard Lebesgue space (Yn,νn)(Y_{n},\nu_{n}). By Proposition 5.17, every HnH_{n} is generically 22-generated and we may apply Proposition 5.18. ∎

Corollary 5.20.

Let GXG\curvearrowright X be a measure-preserving action of a locally compact amenable Polish normed group on a standard probability space (X,μ)(X,\mu). If almost every orbit of the action is uncountable, then the derived L1\mathrm{L}^{1} full group D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) has topological rank 22.

Proof.

Theorem 5.19 implies that the topological rank is at most two. To see that it is actually equal to 22, simply note that D([GX]1)D([G\curvearrowright X\mkern 1.5mu]_{1}) is not abelian (e.g. by the proof of Proposition 3.8). ∎

The assumption for orbits to be uncountable is essential, and Corollary 5.20 is in a striking difference with the dynamical interpretation of the topological rank of derived L1\mathrm{L}^{1} full groups for actions of discrete groups. As shown in [LM21, Thm. 4.3], given an aperiodic measure-preserving action of a finitely generated group ΓX\Gamma\curvearrowright X, the topological rank of D([ΓX]1)D([\Gamma\curvearrowright X\mkern 1.5mu]_{1}) is finite if and only if the action has finite Rokhlin entropy.

Chapter 6 The index map for L1\mathrm{L}^{1} full groups of flows

We now turn our attention to flows, i.e., measure-preserving actions of \mathbb{R}. Since the group of reals is locally compact, amenable, unimodular, and, of course, Polish, all of the results in the previous chapters apply to \mathbb{R}-flows. A much more in-depth understanding of L1\mathrm{L}^{1} full groups of flows is possible and is based on the existence of the so-called index map, which we define and investigate in this chapter. This map is a continuous homomorphism from the L1\mathrm{L}^{1} full group of the flow to the additive group of reals, which can be thought of measuring the average shift distance. When the flow is ergodic, such averages are the same across orbits. By taking the ergodic decomposition of the flow \mathcal{F}, we can adopt a slightly more general vantage point and view the index map \mathcal{I} as a homomorphism into the L1\mathrm{L}^{1} space of functions on the space of invariant measures (,p)(\mathcal{E},p), :[]1L1(,p)\mathcal{I}:[\mathcal{F}\mkern 1.5mu]_{1}\to\mathrm{L}^{1}(\mathcal{E},p).

Understanding the kernel of the index map is the task of fundamental importance. We will subsequently identify ker\ker\mathcal{I} with the derived topological subgroup of []1[\mathcal{F}\mkern 1.5mu]_{1} (Theorem 10.1). This will allow us to describe abelianizations of L1\mathrm{L}^{1} full groups of flows and estimate the number of their topological generators.

It has already been mentioned that any element TT of a full group of a flow induces Lebesgue measure-preserving transformations on orbits (Section 4.2). When TT furthermore belongs to the L1\mathrm{L}^{1} full group, these transformations are special—they leave “half-lines” invariant up to a set of finite measure. Such transformations form the so-called commensurating group. Let us therefore begin with a more formal treatment of this group, which has already appeared in the literature before, for instance in [RS98].

6.1. Self commensurating automorphisms of a subset

Consider an infinite measure space (Z,λ)(Z,\lambda). We say that two measurable sets A,BZA,B\subseteq Z are commensurate if the measure of their symmetric difference is finite, λ(AB)<\lambda(A\triangle B)<\infty. The relation of being commensurate is an equivalence relation, and all sets of finite measure fall into a single class. Note also that if AA and BB are both commensurate to some CC, then so is the intersection ABA\cap B; in other words, all equivalence classes of commensurability are closed under finite intersections.

Let (B)\mathfrak{C}(B) denote the collection of all measurable AZA\subseteq Z that are commensurate to BB. Fix some YZY\subseteq Z and consider the semigroup of measure-preserving transformations between elements of (Y)\mathfrak{C}(Y). More precisely, let Iso(Y,λ)\mathrm{Iso}^{\star}(Y,\lambda) be the collection of measure-preserving maps T:ABT:A\to B between sets A,B(Y)A,B\in\mathfrak{C}(Y), which we call the self commensurating semigroup of (Y,λ)(Y,\lambda).

We use the notation domT=A\operatorname{\mathrm{dom}}T=A and rngT=B\operatorname{\mathrm{rng}}T=B to refer to the domain and the range of TT, respectively. As usual, we identify two maps that differ on a null set. Since classes of commensurability are closed under finite intersections, the set Iso(Y,λ)\mathrm{Iso}^{\star}(Y,\lambda) forms a semigroup under the composition.

This semigroup carries a natural equivalence relation: TST\sim S whenever the transformations disagree on a set of finite measure, λ({x:TxSx})<\lambda(\{x:Tx\neq Sx\})<\infty. This equivalence is, moreover, a congruence, i.e., if T1S1T_{1}\sim S_{1} and T2S2T_{2}\sim S_{2}, then T1T2S1S2T_{1}\circ T_{2}\sim S_{1}\circ S_{2}. One may therefore push the semigroup structure from Iso(Y,λ)\mathrm{Iso}^{\star}(Y,\lambda) onto the set of equivalence classes, which we denote by Aut(Y,λ)\mathrm{Aut}^{\star}(Y,\lambda). An important observation is that Aut(Y,λ)\mathrm{Aut}^{\star}(Y,\lambda) is a group. Indeed, the identity corresponds to the map xxx\mapsto x on YY, and for a representative TIso(Y,λ)T\in\mathrm{Iso}^{\star}(Y,\lambda), its inverse inside Aut(Y,λ)\mathrm{Aut}^{\star}(Y,\lambda) is, naturally, given by T1:rngTdomTT^{-1}:\operatorname{\mathrm{rng}}T\to\operatorname{\mathrm{dom}}T. We call Aut(Y,λ)\mathrm{Aut}^{\star}(Y,\lambda) the self commensurating automorphism group of YY.

The self commensurating semigroup admits an important homomorphism into the reals, :Iso(Y,λ)\mathcal{I}:\mathrm{Iso}^{\star}(Y,\lambda)\to\mathbb{R}, called the index map and defined by

(T)=λ(domTrngT)λ(rngTdomT).\mathcal{I}(T)=\lambda(\operatorname{\mathrm{dom}}T\setminus\operatorname{\mathrm{rng}}T)-\lambda(\operatorname{\mathrm{rng}}T\setminus\operatorname{\mathrm{dom}}T).
Lemma 6.1.

For all TIso(Y,λ)T\in\mathrm{Iso}^{\star}(Y,\lambda), the index map satisfies the following:

  1. (1)

    if A(Y)A\in\mathfrak{C}(Y) is such that domTA\operatorname{\mathrm{dom}}T\subseteq A and rngTA\operatorname{\mathrm{rng}}T\subseteq A, then

    (T)=λ(ArngT)λ(AdomT);\mathcal{I}(T)=\lambda(A\setminus\operatorname{\mathrm{rng}}T)-\lambda(A\setminus\operatorname{\mathrm{dom}}T);
  2. (2)

    if TIso(Y,λ)T^{\prime}\in\mathrm{Iso}^{\star}(Y,\lambda) is a restriction of TT^{\prime}, that is T=TdomTT^{\prime}=T\restriction_{\operatorname{\mathrm{dom}}T^{\prime}}, then (T)=(T)\mathcal{I}(T^{\prime})=\mathcal{I}(T).

Proof.

(1) If AZA\subseteq Z is commensurate to YY and domTA\operatorname{\mathrm{dom}}T\subseteq A, rngTA\operatorname{\mathrm{rng}}T\subseteq A, then

(T)\displaystyle\mathcal{I}(T) =λ(domTrngT)λ(rngTdomT)\displaystyle=\lambda(\operatorname{\mathrm{dom}}T\setminus\operatorname{\mathrm{rng}}T)-\lambda(\operatorname{\mathrm{rng}}T\setminus\operatorname{\mathrm{dom}}T)
=λ(ArngT)λ(A(domTrngT))\displaystyle=\lambda(A\setminus\operatorname{\mathrm{rng}}T)-\lambda(A\setminus(\operatorname{\mathrm{dom}}T\cup\operatorname{\mathrm{rng}}T))
(λ(AdomT)λ(A(domTrngT)))\displaystyle\qquad-(\lambda(A\setminus\operatorname{\mathrm{dom}}T)-\lambda(A\setminus(\operatorname{\mathrm{dom}}T\cup\operatorname{\mathrm{rng}}T)))
=λ(ArngT)λ(AdomT).\displaystyle=\lambda(A\setminus\operatorname{\mathrm{rng}}T)-\lambda(A\setminus\operatorname{\mathrm{dom}}T).

(2) If TIso(Y,λ)T^{\prime}\in\mathrm{Iso}^{\star}(Y,\lambda) is a restriction of TT, then

T(domTdomT)=rngTrngT.T(\operatorname{\mathrm{dom}}T\setminus\operatorname{\mathrm{dom}}T^{\prime})=\operatorname{\mathrm{rng}}T\setminus\operatorname{\mathrm{rng}}T^{\prime}.

Thus for any A(Y)A\in\mathfrak{C}(Y) containing both domT\operatorname{\mathrm{dom}}T and rngT\operatorname{\mathrm{rng}}T, item (1) implies

(T)\displaystyle\mathcal{I}(T) =λ(AdomT)λ(ArngT)\displaystyle=\lambda(A\setminus\operatorname{\mathrm{dom}}T)-\lambda(A\setminus\operatorname{\mathrm{rng}}T)
=λ(AdomT)λ(domTdomT)(λ(BrngT)λ(rngTrngT))\displaystyle=\lambda(A\setminus\operatorname{\mathrm{dom}}T^{\prime})-\lambda(\operatorname{\mathrm{dom}}T\setminus\operatorname{\mathrm{dom}}T^{\prime})-(\lambda(B\setminus\operatorname{\mathrm{rng}}T^{\prime})-\lambda(\operatorname{\mathrm{rng}}T\setminus\operatorname{\mathrm{rng}}T^{\prime}))
=λ(AdomT)λ(ArngT)=(T),\displaystyle=\lambda(A\setminus\operatorname{\mathrm{dom}}T^{\prime})-\lambda(A\setminus\operatorname{\mathrm{rng}}T^{\prime})=\mathcal{I}(T^{\prime}),

where the equality λ(domTdomT)=λ(rngTrngT)\lambda(\operatorname{\mathrm{dom}}T\setminus\operatorname{\mathrm{dom}}T^{\prime})=\lambda(\operatorname{\mathrm{rng}}T\setminus\operatorname{\mathrm{rng}}T^{\prime}) is based on TT being measure-preserving. ∎

Proposition 6.2.

The index map :Iso(Y,λ)\mathcal{I}:\mathrm{Iso}^{\star}(Y,\lambda)\to\mathbb{R} is a homomorphism. Moreover, if T,SIso(Y,λ)T,S\in\mathrm{Iso}^{\star}(Y,\lambda) are equivalent, TST\sim S, then (T)=(S)\mathcal{I}(T)=\mathcal{I}(S).

Proof.

In view of Lemma 6.1(2), to check that (T1T2)=(T1)+(T2)\mathcal{I}(T_{1}\circ T_{2})=\mathcal{I}(T_{1})+\mathcal{I}(T_{2}) we may pass to restrictions of these transformations and assume that rngT2=domT1\operatorname{\mathrm{rng}}T_{2}=\operatorname{\mathrm{dom}}T_{1}. Pick a set A(Y)A\in\mathfrak{C}(Y) large enough to contain the domains and ranges of T1T_{1} and T2T_{2}; by Lemma 6.1(1)

(T1T2)\displaystyle\mathcal{I}(T_{1}\circ T_{2}) =λ(ArngT1)λ(AdomT2)\displaystyle=\lambda(A\setminus\operatorname{\mathrm{rng}}T_{1})-\lambda(A\setminus\operatorname{\mathrm{dom}}T_{2})
=λ(ArngT1)λ(AdomT1)+λ(ArngT2)λ(AdomT2)\displaystyle=\lambda(A\setminus\operatorname{\mathrm{rng}}T_{1})-\lambda(A\setminus\operatorname{\mathrm{dom}}T_{1})+\lambda(A\setminus\operatorname{\mathrm{rng}}T_{2})-\lambda(A\setminus\operatorname{\mathrm{dom}}T_{2})
=(T1)+(T2).\displaystyle=\mathcal{I}(T_{1})+\mathcal{I}(T_{2}).

For the moreover part, suppose that T,SIsoY(Y,λ)T,S\in\mathrm{Iso}^{\star}_{Y}(Y,\lambda) are equivalent. Let UU be the restriction of TT and SS onto the set {x:Tx=Sx}\{x:Tx=Sx\}. Using Lemma 6.1(2) once again, we get (T)=(U)=(S)\mathcal{I}(T)=\mathcal{I}(U)=\mathcal{I}(S), hence the index map is invariant under the equivalence relation \sim. ∎

The proposition above implies that the index map respects the relation \sim, and hence gives rise to a map from Aut(Y,λ)\mathrm{Aut}^{\star}(Y,\lambda) to the reals.

Corollary 6.3.

The index map factors to a group homomorphism

:Aut(Y,λ).\mathcal{I}:\mathrm{Aut}^{\star}(Y,\lambda)\to\mathbb{R}.

6.2. The commensurating automorphism group

Let us again consider an infinite measure space (Z,λ)(Z,\lambda) and YZY\subseteq Z a measurable subset. We now define the commensurating automorphism group of YY in ZZ as the group of all measure-preserving transformations TAut(Z,λ)T\in\mathrm{Aut}(Z,\lambda) such that λ(YT(Y))<\lambda(Y\triangle T(Y))<\infty. We denote this group by AutY(Z,λ)\mathrm{Aut}_{Y}(Z,\lambda).

Every TAutY(Z,λ)T\in\mathrm{Aut}_{Y}(Z,\lambda) naturally gives rise to an element of Aut(Y,λ)\mathrm{Aut}^{\star}(Y,\lambda) by considering its restriction TYT\restriction_{Y}. The following lemma shows that in this case we may use any other set AA commensurate to YY instead without changing the corresponding element of the commensurating group.

Lemma 6.4.

Let TAut(Z,λ)T\in\mathrm{Aut}(Z,\lambda) be a measure-preserving automorphism. If TAIso(Y,λ)T\restriction_{A}\in\mathrm{Iso}^{\star}(Y,\lambda) for some A(Y)A\in\mathfrak{C}(Y), then TBIso(Y,λ)T\restriction_{B}\in\mathrm{Iso}^{\star}(Y,\lambda) and TBTAT\restriction_{B}\sim T\restriction_{A} for all B(Y)B\in\mathfrak{C}(Y).

Proof.

Since commensuration is an equivalence relation and AA is commensurate to YY, the assumption TAIso(Y,λ)T\restriction_{A}\in\mathrm{Iso}^{\star}(Y,\lambda) is equivalent to λ(AT(A))<\lambda(A\triangle T(A))<\infty. Moreover, given B(Y)B\in\mathfrak{C}(Y), we only need to show that λ(BT(B))\lambda(B\triangle T(B)) is finite in order to conclude that TBIso(Y,λ)T\restriction_{B}\in\mathrm{Iso}^{\star}(Y,\lambda). So we compute

λ(BT(B))=\displaystyle\lambda(B\triangle T(B))= λ(BT(B))+λ(T(B)B)\displaystyle\lambda(B\setminus T(B))+\lambda(T(B)\setminus B)
\displaystyle\leq λ(AT(A))+λ(BA)+λ(T(AB))\displaystyle\lambda(A\setminus T(A))+\lambda(B\setminus A)+\lambda(T(A\setminus B))
+λ(T(A)A)+λ(AB)+λ(T(BA))\displaystyle\qquad+\lambda(T(A)\setminus A)+\lambda(A\setminus B)+\lambda(T(B\setminus A))
=\displaystyle= λ(AT(A))+2λ(AB)<.\displaystyle\lambda(A\triangle T(A))+2\lambda(A\triangle B)<\infty.

Thus the measure λ(BT(B))\lambda(B\triangle T(B)) is finite, hence TBIso(Y,λ)T\restriction_{B}\in\mathrm{Iso}^{\star}(Y,\lambda) for all B(Y)B\in\mathfrak{C}(Y). Finally, TATBT\restriction_{A}\sim T\restriction_{B}, since these transformations agree on ABA\cap B. ∎

To summarize, if TAIso(Y,λ)T\restriction_{A}\in\mathrm{Iso}^{\star}(Y,\lambda) for some A(Y)A\in\mathfrak{C}(Y), then all restrictions TBT\restriction_{B}, B(Y)B\in\mathfrak{C}(Y), are pairwise equivalent, hence correspond to the same element TYAut(Y,λ)T\restriction_{Y}\in\mathrm{Aut}^{\star}(Y,\lambda). According to Proposition 6.2, the index (TY)\mathcal{I}(T\restriction_{Y}) of this element can be computed as (TY)=λ(BT(B))λ(BT1(B))\mathcal{I}(T\restriction_{Y})=\lambda(B\setminus T(B))-\lambda(B\setminus T^{-1}(B)) for any B(Y)B\in\mathfrak{C}(Y).

6.3. Index map on L1\mathrm{L}^{1} full groups of \mathbb{R}-flows

Let =X\mathcal{F}=\mathbb{R}\curvearrowright X be a free measure-preserving Borel flow, let []1[\mathcal{F}\mkern 1.5mu]_{1} be the associated L1\mathrm{L}^{1} full group, where we endow \mathbb{R} with the standard Euclidean norm, and let T[]1T\in[\mathcal{F}\mkern 1.5mu]_{1}. The action of rr\in\mathbb{R} upon xXx\in X is denoted additively by x+rx+r. Recall that the cocycle of TT is denoted by ρT:X\rho_{T}:X\to\mathbb{R} and is defined by the equality T(x)=x+ρT(x)T(x)=x+\rho_{T}(x) for all xXx\in X. We are going to argue that, on every orbit, TT induces a measure-preserving transformation that belongs to the commensurate group of 0\mathbb{R}^{\geq 0}, when the orbit is identified with the real line.

Consider the function f:{1,0,1}f:\mathcal{R}_{\mathcal{F}}\to\{-1,0,1\} defined by

f(x,y)={1if x<y<T(x),1if T(x)<y<x,0otherwise.f(x,y)=\begin{cases}1&\textrm{if }x<y<T(x),\\ -1&\textrm{if }T(x)<y<x,\\ 0&\textrm{otherwise}.\end{cases}

One can think of ff as a “charge function” that spreads charge +1+1 over each interval (x,T(x))(x,T(x)) and 1-1 over (T(x),x)(T(x),x). Note that f(x,x+r)𝑑λ(r)=ρT(x)\int_{\mathbb{R}}f(x,x+r)\,d\lambda(r)=\rho_{T}(x). Since TT belongs to the L1\mathrm{L}^{1} full group, its cocycle is integrable, which means that ff is MM-integrable (see Section 4.2). We apply the mass-transport principle, which shows that

Xf(x,x+r)𝑑λ(r)𝑑μ(x)=Xf(x+r,x)𝑑λ(r)𝑑μ(x).\int_{X}\int_{\mathbb{R}}f(x,x+r)\,d\lambda(r)d\mu(x)=\int_{X}\int_{\mathbb{R}}f(x+r,x)\,d\lambda(r)d\mu(x).

Let TxAut(,λ)T_{x}\in\mathrm{Aut}(\mathbb{R},\lambda) denote the transformation induced by TT onto the orbit of xx obtained by identifying the origin of the real line with xx. The following two quantities are therefore finite:

|f(x+r,x)|𝑑λ(r)\displaystyle\int_{\mathbb{R}}|f(x+r,x)|\,d\lambda(r) =λ(0Tx(0))+λ(Tx(0)0),\displaystyle=\lambda\bigl{(}\mathbb{R}^{\geq 0}\setminus T_{x}(\mathbb{R}^{\geq 0})\bigr{)}+\lambda\bigl{(}T_{x}(\mathbb{R}^{\geq 0})\setminus\mathbb{R}^{\geq 0}\bigr{)},
f(x+r,x)𝑑λ(r)\displaystyle\int_{\mathbb{R}}f(x+r,x)\,d\lambda(r) =λ(0Tx(0))λ(Tx(0)0).\displaystyle=\lambda\bigl{(}\mathbb{R}^{\geq 0}\setminus T_{x}(\mathbb{R}^{\geq 0})\bigr{)}-\lambda\bigl{(}T_{x}(\mathbb{R}^{\geq 0})\setminus\mathbb{R}^{\geq 0}\bigr{)}.

In particular, Tx0T_{x}\restriction_{\mathbb{R}^{\geq 0}} belongs to the commensurating group of 0\mathbb{R}^{\geq 0}. The second quantity, on the other hand, is equal to the index of Tx0T_{x}\restriction_{\mathbb{R}^{\geq 0}}. By Section 6.2, (Tx0)=(Ty0)\mathcal{I}(T_{x}\restriction_{\mathbb{R}^{\geq 0}})=\mathcal{I}(T_{y}\restriction_{\mathbb{R}^{\geq 0}}) whenever xyx\mathcal{R}_{\mathcal{F}}y. For any T[]1T\in[\mathcal{F}\mkern 1.5mu]_{1}, we therefore have an orbit invariant measurable map hT:Xh_{T}:X\to\mathbb{R} given by hT(x)=f(x+r,x)𝑑λ(r)h_{T}(x)=\int_{\mathbb{R}}f(x+r,x)\,d\lambda(r). Note that for any \mathcal{F}-invariant set YXY\subseteq X, we have

(6.1) YρT(x)𝑑μ(x)=YhT(x)𝑑μ(x).\int_{Y}\rho_{T}(x)\,d\mu(x)=\int_{Y}h_{T}(x)\,d\mu(x).

Let (,p)(\mathcal{E},p), XxνxX\ni x\mapsto\nu_{x}\in\mathcal{E}, be the ergodic decomposition of (X,μ,)(X,\mu,\mathcal{F}) (see Appendix C.3). Since the map hTh_{T} is \mathcal{R}_{\mathcal{F}}-invariant, it produces a map h~T:\tilde{h}_{T}:\mathcal{E}\to\mathbb{R} via h~(ν)=h(x)\tilde{h}(\nu)=h(x) for any xx such that ν=νx\nu=\nu_{x} or, equivalently, via

h~T(ν)=Xf(x+r,x)𝑑λ(r)𝑑ν(x).\tilde{h}_{T}(\nu)=\int_{X}\int_{\mathbb{R}}f(x+r,x)\,d\lambda(r)d\nu(x).

Note also that

XhT(x)𝑑μ(x)=Xf(x+r,x)𝑑λ(r)𝑑μ(x)=h~T(ν)𝑑p(ν),\int_{X}h_{T}(x)\,d\mu(x)=\int_{X}\int_{\mathbb{R}}f(x+r,x)\,d\lambda(r)d\mu(x)=\int_{\mathcal{E}}\tilde{h}_{T}(\nu)\,dp(\nu),

thus h~TL1(,)\tilde{h}_{T}\in\mathrm{L}^{1}(\mathcal{E},\mathbb{R}). We can now define the index map of a (possibly non-ergodic) flow as a function :[]1L1(,)\mathcal{I}:[\mathcal{F}\mkern 1.5mu]_{1}\to\mathrm{L}^{1}(\mathcal{E},\mathbb{R}).

Definition 6.5.

Let =X\mathcal{F}=\mathbb{R}\curvearrowright X be a free measure-preserving flow on a standard probability space (X,μ)(X,\mu); let also (,p)(\mathcal{E},p) be the space of \mathcal{F}-invariant ergodic probability measures, where pp is the probability measure yielding the disintegration of μ\mu. The index map is the function :[]1L1(,)\mathcal{I}:[\mathcal{F}\mkern 1.5mu]_{1}\to\mathrm{L}^{1}(\mathcal{E},\mathbb{R}) given by (T)(ν)=h~T(ν)=Xf(x+r,x)𝑑λ(r)𝑑ν(x)\mathcal{I}(T)(\nu)=\tilde{h}_{T}(\nu)=\int_{X}\int_{\mathbb{R}}f(x+r,x)\,d\lambda(r)d\nu(x).

Proposition 6.6.

For any free measure-preserving flow =X\mathcal{F}=\mathbb{R}\curvearrowright X, the index map :[]1L1(,)\mathcal{I}:[\mathcal{F}\mkern 1.5mu]_{1}\to\mathrm{L}^{1}(\mathcal{E},\mathbb{R}) is a continuous and surjective homomorphism.

Proof.

The index map is a homomorphism, since, as we have discussed earlier, hT(x)h_{T}(x) is equal to the index of Tx0T_{x}\restriction_{\mathbb{R}^{\geq 0}}. Continuity follows from the fact that \mathcal{I} is a Borel homomorphism between Polish groups. To see surjectivity, pick any h~L1(,)\tilde{h}\in\mathrm{L}^{1}(\mathcal{E},\mathbb{R}), view it as a map h:Xh:X\to\mathbb{R} via the identification h(x)=h~(νx)h(x)=\tilde{h}(\nu_{x}). Define the automorphism TAut(X,μ)T\in\mathrm{Aut}(X,\mu) by T(x)=x+h(x)T(x)=x+h(x). It is straightforward to check that T[]1T\in[\mathcal{F}\mkern 1.5mu]_{1} and (T)=h\mathcal{I}(T)=h. ∎

The quotient group []1/ker[\mathcal{F}\mkern 1.5mu]_{1}/\ker{\mathcal{I}} naturally inherits the quotient norm given by

Tker1=infSkerTS1.\left\lVert T\ker{\mathcal{I}}\right\rVert_{1}=\inf_{S\in\ker\mathcal{I}}\left\lVert TS\right\rVert_{1}.

By Proposition 6.6, the index map induces an isomorphism between []1/ker[\mathcal{F}\mkern 1.5mu]_{1}/\ker{\mathcal{I}} and L1(,)\mathrm{L}^{1}(\mathcal{E},\mathbb{R}). We argue that this isomorphism is, in fact, an isometry.

Proposition 6.7.

The index map \mathcal{I} induces an isometric isomorphism between []1/ker[\mathcal{F}\mkern 1.5mu]_{1}/\ker{\mathcal{I}} and L1(,)\mathrm{L}^{1}(\mathcal{E},\mathbb{R}), where the former is endowed with the quotient norm and the latter bears the usual L1\mathrm{L}^{1} norm.

Proof.

Since X|hT(x)|𝑑μ(x)=|h~T(ν)|𝑑p(ν)\int_{X}|h_{T}(x)|\,d\mu(x)=\int_{\mathcal{E}}|\tilde{h}_{T}(\nu)|\,dp(\nu), it suffices to show that for all T[]1T\in[\mathcal{F}\mkern 1.5mu]_{1}

infSkerTS1=X|hT|𝑑μ.\inf_{S\in\ker\mathcal{I}}\left\lVert TS\right\rVert_{1}=\int_{X}|h_{T}|\,d\mu.

Let T[]1T\in[\mathcal{F}\mkern 1.5mu]_{1}. We first show the inequality infSkerTS1X|hT|𝑑μ\displaystyle\inf_{S\in\ker\mathcal{I}}\left\lVert TS\right\rVert_{1}\geq\int_{X}|h_{T}|\,d\mu.

Pick any SkerS\in\ker\mathcal{I}. For any \mathcal{F}-invariant measurable YXY\subseteq X, YρS𝑑μ=0\int_{Y}\rho_{S}\,d\mu=0 and

YρTS𝑑μ=YρT(S(x))𝑑μ(x)+YρS(x)𝑑μ(x)=YρT𝑑μ=YhT𝑑μ,\int_{Y}\rho_{TS}\,d\mu=\int_{Y}\rho_{T}(S(x))\,d\mu(x)+\int_{Y}\rho_{S}(x)\,d\mu(x)=\int_{Y}\rho_{T}\,d\mu=\int_{Y}h_{T}\,d\mu,

where we rely on Eq. (6.1) and SS being measure-preserving. Consider the \mathcal{F}-invariant sets

Y<0={xX:hT(x)<0}andY0={xX:hT(x)0}.Y^{<0}=\{x\in X:h_{T}(x)<0\}\quad\textrm{and}\quad Y^{\geq 0}=\{x\in X:h_{T}(x)\geq 0\}.

The norm TS1\left\lVert TS\right\rVert_{1} can be estimated from below as follows.

TS1\displaystyle\left\lVert TS\right\rVert_{1} =X|ρTS|𝑑μ=Y<0|ρTS|𝑑μ+Y0|ρTS|𝑑μ\displaystyle=\int_{X}|\rho_{TS}|\,d\mu=\int_{Y^{<0}}|\rho_{TS}|\,d\mu+\int_{Y^{\geq 0}}|\rho_{TS}|\,d\mu
|Y<0ρTS𝑑μ|+|Y0ρTS𝑑μ|\displaystyle\geq\left\lvert\int_{Y^{<0}}\rho_{TS}\,d\mu\right\rvert+\left\lvert\int_{Y^{\geq 0}}\rho_{TS}\,d\mu\right\rvert
=|Y<0hT𝑑μ|+|Y0hT𝑑μ|\displaystyle=\left\lvert\int_{Y^{<0}}h_{T}\,d\mu\right\rvert+\left\lvert\int_{Y^{\geq 0}}h_{T}\,d\mu\right\rvert
=Y<0hT𝑑μ+Y0hT𝑑μ=X|hT|𝑑μ.\displaystyle=-\int_{Y^{<0}}h_{T}\,d\mu+\int_{Y^{\geq 0}}h_{T}\,d\mu=\int_{X}|h_{T}|\,d\mu.

We conclude that

infSkerTS1X|hT|𝑑μ.\inf_{S\in\ker{\mathcal{I}}}\left\lVert TS\right\rVert_{1}\geq\int_{X}|h_{T}|\,d\mu.

For the other direction, consider a transformation TT^{\prime} defined by T(x)=x+hT(x)T^{\prime}(x)=x+h_{T}(x); note that T[]1T^{\prime}\in[\mathcal{F}\mkern 1.5mu]_{1}, ρT(x)=hT(x)=hT(x)\rho_{T^{\prime}}(x)=h_{T^{\prime}}(x)=h_{T}(x) for all xXx\in X, and T1TkerT^{-1}T^{\prime}\in\ker\mathcal{I}. Therefore

infSkerTS1TT1T1=T1=X|hT|𝑑μ=X|hT|𝑑μ,\inf_{S\in\ker{\mathcal{I}}}\left\lVert TS\right\rVert_{1}\leq\left\lVert TT^{-1}T^{\prime}\right\rVert_{1}=\left\lVert T^{\prime}\right\rVert_{1}=\int_{X}|h_{T^{\prime}}|\,d\mu=\int_{X}|h_{T}|\,d\mu,

and the desired equality of norms follows. ∎

Using a similar reasoning, we get the following characterization of the L1\mathrm{L}^{1} full group and the index map, where for all T[]T\in[\mathcal{R}_{\mathcal{F}}] we let rTr_{T} be the measure-preserving transformation of (,M)(\mathcal{R}_{\mathcal{F}},M) given by rT(x,y)=(x,T(y))r_{T}(x,y)=(x,T(y)) (see Section 4.2).

Proposition 6.8.

Let =X\mathcal{F}=\mathbb{R}\curvearrowright X be a free measure-preserving \mathbb{R}-flow. Consider the set 0={(x,y):xy}\mathcal{R}^{\geq 0}=\{(x,y)\in\mathcal{R}_{\mathcal{F}}:x\geq y\}. Then for every T[X]T\in[\mathbb{R}\curvearrowright X], we have

T1=M(0rT(0)).\left\lVert T\right\rVert_{1}=M\left(\mathcal{R}^{\geq 0}\bigtriangleup r_{T}(\mathcal{R}^{\geq 0})\right).

In particular, the L1\mathrm{L}^{1} full group of \mathcal{F} can be seen as the commensurating group of 0\mathcal{R}^{\geq 0} inside the full group of \mathcal{R}. Moreover, in the ergodic case, the index of TT as defined above is equal to its index as a commensurating transformation of the set 0\mathcal{R}^{\geq 0} in the sense of Section 6.1.

Proof.

Through the identification (x,t)(x,x+t)(x,t)\mapsto(x,x+t), the measure-preserving transformation rTr_{T} is acting on X×X\times\mathbb{R} as idX×Tx\mathrm{id_{X}}\times T_{x}, and the set 0\mathcal{R}^{\geq 0} becomes X×0X\times\mathbb{R}^{\geq 0}. We then have

M(0rT(0))\displaystyle M(\mathcal{R}^{\geq 0}\bigtriangleup r_{T}(\mathcal{R}^{\geq 0})) =Xλ(0(Tx(0)))𝑑μ(x)\displaystyle=\int_{X}\lambda(\mathbb{R}^{\geq 0}\bigtriangleup(T_{x}(\mathbb{R}^{\geq 0})))\,d\mu(x)
=X|ρT|𝑑μ\displaystyle=\int_{X}\left\lvert\rho_{T}\right\rvert\,d\mu

by the mass-transport principle, which yields the conclusion, since by the definition of the norm T1=X|ρT|𝑑μ\left\lVert T\right\rVert_{1}=\int_{X}\left\lvert\rho_{T}\right\rvert\,d\mu.

The moreover part follows from a similar computation. ∎

Remark 6.9.

The full group of \mathcal{R} embeds via TrTT\mapsto r_{T} into the group of measure-preserving transformations of (,M)(\mathcal{R},M). One could use this and the fact that the commensurating automorphism group of 0\mathcal{R}^{\geq 0} is a Polish group in order to give another proof that L1\mathrm{L}^{1} full groups of measure-preserving \mathbb{R}-flows are themselves Polish.

Chapter 7 Orbitwise ergodic bounded elements of full groups

The purpose of this chapter is to contrast some of the differences in the dynamics of the elements of full groups of \mathbb{Z}-actions and those arising from \mathbb{R}-flows. Let S[X]S\in[\mathbb{Z}\curvearrowright X\mkern 1.5mu] be an element of the full group of a measure-preserving aperiodic transformation and let ρSk:X\rho_{S^{k}}:X\to\mathbb{Z} be the cocycle associated with SkS^{k} for kk\in\mathbb{Z}. Since \mathbb{Z} is a discrete group, the conservative part in the Hopf’s decomposition for SS (see Appendix B) reduces to the set of periodic orbits. In particular, an aperiodic S[X]S\in[\mathbb{Z}\curvearrowright X\mkern 1.5mu] has to be dissipative, hence |ρSk(x)||\rho_{S^{k}}(x)|\to\infty as kk\to\infty. When SS belongs to the L1\mathrm{L}^{1} full group of the action, a theorem of R. M. Belinskaja [Bel68, Thm. 3.2] strengthens this conclusion and asserts that for almost all xx in the dissipative component of SS either ρSk(x)+\rho_{S^{k}}(x)\to+\infty or ρSk(x)\rho_{S^{k}}(x)\to-\infty.

Given an arbitrary free measure-preserving flow X\mathbb{R}\curvearrowright X, we build an example of an aperiodic S[X]1S\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} for which the signs in {ρSk(x):k}\{\rho_{S^{k}}(x):k\in\mathbb{N}\} keep alternating indefinitely for almost all xXx\in X. In fact, we present a transformation that acts ergodically on each orbit of the flow (in particular, it is conservative and globally ergodic as soon as the flow is ergodic). Moreover, we ensure it has a uniformly bounded cocycle. Our argument uses a variant of the well-known cutting and stacking construction adapted for infinite measure spaces. Additional technical difficulties arise from the necessity to work across all orbits of the flow simultaneously. The transformation will arise as a limit of special partial maps we call castles, which we now define.

The pseudo full group of the flow is the set of injective Borel maps φ:domφrngφ\varphi:\operatorname{\mathrm{dom}}\varphi\to\operatorname{\mathrm{rng}}\varphi between Borel sets domφX\operatorname{\mathrm{dom}}\varphi\subseteq X, rngφX\operatorname{\mathrm{rng}}\varphi\subseteq X, for which there exists a countable Borel partition (An)n(A_{n})_{n\in\mathbb{N}} of the domain domφ\operatorname{\mathrm{dom}}\varphi and a countable family of reals (tn)n(t_{n})_{n\in\mathbb{N}} such that φ(x)=x+tn\varphi(x)=x+t_{n} for every xAnx\in A_{n}. Such maps are measure-preserving isomorphisms between (domφ,μdomφ)(\operatorname{\mathrm{dom}}\varphi,\mu\restriction_{\operatorname{\mathrm{dom}}\varphi}) and (rngφ,μrngφ)(\operatorname{\mathrm{rng}}\varphi,\mu\restriction_{\operatorname{\mathrm{rng}}\varphi}). The support of φ\varphi is the set

suppφ={xdomφ:φ(x)x}{xrngφ:φ1(x)x}.\operatorname*{supp}\varphi=\{x\in\operatorname{\mathrm{dom}}\varphi:\varphi(x)\neq x\}\cup\{x\in\operatorname{\mathrm{rng}}\varphi:\varphi^{-1}(x)\neq x\}.

Given φ\varphi in the pseudo full group and a Borel set AXA\subseteq X, we let

φ(A)={φ(x):xAdomφ}.\varphi(A)=\{\varphi(x):x\in A\cap\operatorname{\mathrm{dom}}\varphi\}.

In particular, φ(A)=\varphi(A)=\varnothing if AA is disjoint from domφ\operatorname{\mathrm{dom}}\varphi. A castle is an element φ\varphi of the pseudo full group of the flow such that for B=domφrngφB=\operatorname{\mathrm{dom}}\varphi\setminus\operatorname{\mathrm{rng}}\varphi the sequence (φk(B))k(\varphi^{k}(B))_{k\in\mathbb{N}} consists of pairwise disjoint subsets which cover its support. Since φ\varphi is measure-preserving, for almost every xBx\in B there is kk\in\mathbb{N} such that φk(x)domφ\varphi^{k}(x)\not\in\operatorname{\mathrm{dom}}\varphi. It follows that φ1\varphi^{-1} is also a castle. The set BB is called the basis of the castle, and the basis of its inverse CC is called its ceiling, which is equal to rngφdomφ\operatorname{\mathrm{rng}}\varphi\setminus\operatorname{\mathrm{dom}}\varphi. Observe that if two castles have disjoint supports, then their union is also a castle. We denote by φ:BC\vec{\varphi}:B\to C the element of the pseudo full group which takes every element of the basis of φ\varphi to the corresponding element of the ceiling.

Remark 7.1.

Equivalently, one could define a castle as an element φ\varphi of the pseudo full group which induces a graphing consisting of finite segments only (see [KM04, Sec. 17] for the definition of a graphing). It induces a partial order φ\leq_{\varphi} defined by xφyx\leq_{\varphi}y if and only if there is kk\in\mathbb{N} such that y=φk(x)y=\varphi^{k}(x). The basis of the castle is the set of minimal elements, while the ceiling is the set of maximal ones. Finally, φ\vec{\varphi} is the map which takes a minimal element to the unique maximal element above it.

Theorem 7.2.

Let X\mathbb{R}\curvearrowright X be a free measure-preserving flow. There exists S[X]S\in[\mathbb{R}\curvearrowright X\mkern 1.5mu] that acts ergodically on every orbit of the flow and whose cocycle is bounded by 44. Moreover, the signs in {ρSk(x):k}\{\rho_{S^{k}}(x):k\in\mathbb{N}\} keep changing indefinitely for almost all xXx\in X.

Proof.

Fix a free measure-preserving flow X\mathbb{R}\curvearrowright X, and let 𝒞X\mathcal{C}\subset X be a cross-section. Since 𝒞\mathcal{C} is lacunary, for any c𝒞c\in\mathcal{C} there exists min{r>0:c+r𝒞}\min\{r>0:c+r\in\mathcal{C}\}; we denote this value by gap𝒞(c)\mathrm{gap}_{\mathcal{C}}(c). This gives the first return map σ𝒞:𝒞𝒞\sigma_{\mathcal{C}}:\mathcal{C}\to\mathcal{C} via σ𝒞(c)=c+gap𝒞(c)\sigma_{\mathcal{C}}(c)=c+\mathrm{gap}_{\mathcal{C}}(c), which is Borel. There is also a natural bijective correspondence between XX and the set {(c,r)𝒞×0:c𝒞,0r<gap𝒞(c)}\{(c,r)\in\mathcal{C}\times\mathbb{R}^{\geq 0}:c\in\mathcal{C},0\leq r<\mathrm{gap}_{\mathcal{C}}(c)\}. Let λc𝒞\lambda_{c}^{\mathcal{C}} be the “Lebesgue measure” on c+[0,gap𝒞(c))c+[0,\mathrm{gap}_{\mathcal{C}}(c)) given by

λc𝒞(A)=λ({r:0r<gap𝒞(c),c+rA}).\lambda_{c}^{\mathcal{C}}(A)=\lambda(\{r\in\mathbb{R}:0\leq r<\mathrm{gap}_{\mathcal{C}}(c),c+r\in A\}).

The measure μ\mu on XX can be disintegrated as μ(A)=𝒞λc𝒞(A)𝑑ν(c)\mu(A)=\int_{\mathcal{C}}\lambda_{c}^{\mathcal{C}}(A)\,d\nu(c) for some finite (but not necessarily probability) measure ν\nu on 𝒞\mathcal{C} (see, for instance, [Slu17, Sec. 4] and Appendix C.1).

Let (𝒞n)n(\mathcal{C}_{n})_{n\in\mathbb{N}} be a vanishing sequence of markers—a sequence of nested cross-sections 𝒞1𝒞2𝒞3\mathcal{C}_{1}\supset\mathcal{C}_{2}\supset\mathcal{C}_{3}\cdots with the empty intersection: n𝒞n=\bigcap_{n\in\mathbb{N}}\mathcal{C}_{n}=\varnothing. We may arrange 𝒞1\mathcal{C}_{1} to be such that gap𝒞1(c)(2,3)\mathrm{gap}_{\mathcal{C}_{1}}(c)\in\mathopen{(}2,3\mathclose{)} for all c𝒞1c\in\mathcal{C}_{1}. Put

𝒞0={c+k:c𝒞1,k{0,1,2}}\mathcal{C}_{0}=\{c+k:c\in\mathcal{C}_{1},k\in\{0,1,2\}\}

and Y=𝒞1+[0,2)Y=\mathcal{C}_{1}+[0,2). Note that μ(XY)13\mu(X\setminus Y)\leq\frac{1}{3}. Our first goal is to define an element φ\varphi of the pseudo full group with domain and range equal to YY such that for almost every xYx\in Y, the action of φ\varphi on the intersection of the orbit of xx with YY is ergodic, and which has cocycle bounded by 33. It will then be easy to modify φ\varphi to an element of the full group whose action on each orbit of the flow is ergodic at the cost of increasing the cocycle bound to 44.

Our first transformation φ\varphi will arise as the limit of a sequence of castles (φn)n(\varphi_{n})_{n\in\mathbb{N}}, with each φn\varphi_{n} belonging to the pseudo full group of 𝒞n\mathcal{R}_{\mathcal{C}_{n}}. We also use another family of castles (ψn)n(\psi_{n})_{n\in\mathbb{N}} which allows us to extend φn\varphi_{n} by “going back” from its ceiling to its basis while keeping the cocycle bound (this is our main adjustment compared to the usual cutting and stacking procedure). Both sequences of castles will have their cocycles bounded by 33. Here are the basic constraints that these sequences have to satisfy:

  1. (1)

    for all n1n\geq 1, Y=suppφnsuppψnY=\operatorname*{supp}\varphi_{n}\sqcup\operatorname*{supp}\psi_{n};

  2. (2)

    for all n1n\geq 1, φn+1\varphi_{n+1} extends φn\varphi_{n};

  3. (3)

    μ(suppψn)\mu(\operatorname*{supp}\psi_{n}) tends to 0 as nn tends to ++\infty.

Bases and ceilings of (φn)n(\varphi_{n})_{n\in\mathbb{N}} and (ψn)n(\psi_{n})_{n\in\mathbb{N}} will satisfy additional constraints which will enable us to make the induction work and ensure ergodicity on each orbit of the flow. In order to specify these constraints properly, we introduce the following notation.

Each orbit of the flow comes with the linear order << inherited from \mathbb{R} via x<yx<y if and only if y=x+ty=x+t for some t>0t>0. Set κ𝒞n(x)\kappa_{\mathcal{C}_{n}}(x) to be the minimum of the intersection of 𝒞n\mathcal{C}_{n} with the cone {yX:yx}\{y\in X:y\geq x\}.

Let 𝒟1=𝒞1+2𝒞0\mathcal{D}_{1}=\mathcal{C}_{1}+2\subseteq\mathcal{C}_{0} and 𝒟n\mathcal{D}_{n} be the set of those x𝒟1x\in\mathcal{D}_{1} which are maximal in κ𝒞n1(c)\kappa_{\mathcal{C}_{n}}^{-1}(c) among points of 𝒟1\mathcal{D}_{1} for some c𝒞nc\in\mathcal{C}_{n}; in other words,

𝒟n={x𝒟1:(x,κ𝒞n(x))𝒞0=}.\mathcal{D}_{n}=\{x\in\mathcal{D}_{1}:(x,\kappa_{\mathcal{C}_{n}}(x))\cap\mathcal{C}_{0}=\varnothing\}.

Note that by construction the distance between xx and κ𝒞n(x)\kappa_{\mathcal{C}_{n}}(x) is less than 11 for each x𝒟nx\in\mathcal{D}_{n}. Let ιn\iota_{n} be the map 𝒞n𝒟n\mathcal{C}_{n}\to\mathcal{D}_{n} which assigns to c𝒞nc\in\mathcal{C}_{n} the <<-least element of 𝒟n\mathcal{D}_{n} which is greater than cc.

𝒞1\mathcal{C}_{1}𝒞2\mathcal{C}_{2}𝒟1\mathcal{D}_{1}𝒞1\mathcal{C}_{1}𝒟1\mathcal{D}_{1}𝒞1\mathcal{C}_{1}𝒟1\mathcal{D}_{1}𝒟2\mathcal{D}_{2}𝒞1\mathcal{C}_{1}𝒞2\mathcal{C}_{2}𝒟1\mathcal{D}_{1}𝒞1\mathcal{C}_{1}𝒟2\mathcal{D}_{2}𝒟1\mathcal{D}_{1}𝒞1\mathcal{C}_{1}𝒞2\mathcal{C}_{2}
Figure 7.1. An example of cross-sections 𝒞0\mathcal{C}_{0} (all points), 𝒞1\mathcal{C}_{1} (dots of size and above), 𝒞2\mathcal{C}_{2} (marked as ) and 𝒟1\mathcal{D}_{1}, 𝒟2\mathcal{D}_{2}.

The bases and ceilings of φn\varphi_{n} and ψn\psi_{n} are as follows.

  • the basis of φn\varphi_{n} is An=𝒞n+[0,12n)A_{n}=\mathcal{C}_{n}+\left[0,\frac{1}{2^{n}}\right);

  • the ceiling of φn\varphi_{n} is Bn=𝒟n+[1212n,12)B_{n}=\mathcal{D}_{n}+\left[-\frac{1}{2}-\frac{1}{2^{n}},-\frac{1}{2}\right);

  • the basis of ψn\psi_{n} is Cn=𝒟n+[12,12+12n)C_{n}=\mathcal{D}_{n}+\left[-\frac{1}{2},-\frac{1}{2}+\frac{1}{2^{n}}\right);

  • the ceiling of ψn\psi_{n} is Dn=𝒞n+[12,12+12n)D_{n}=\mathcal{C}_{n}+\left[\frac{1}{2},\frac{1}{2}+\frac{1}{2^{n}}\right).

Furthermore, we impose two translation conditions, which help us to preserve the above concrete definitions of the bases and ceilings at the inductive step when we construct φn+1\varphi_{n+1} and ψn+1\psi_{n+1}:

  • φn(c+t)=ιn(c)+t1212n\vec{\varphi}_{n}(c+t)=\iota_{n}(c)+t-\frac{1}{2}-\frac{1}{2^{n}} for all c𝒞nc\in\mathcal{C}_{n} and all t[0,12n)t\in\left[0,\frac{1}{2^{n}}\right).

  • ψn(d+t)=ιn1(d)+t+1\vec{\psi}_{n}(d+t)=\iota_{n}^{-1}(d)+t+1 for all d𝒟nd\in\mathcal{D}_{n} and all t[12,12+12n)t\in\left[-\frac{1}{2},-\frac{1}{2}+\frac{1}{2^{n}}\right).

The first step of the construction consists of the castle φ1:xx+1\varphi_{1}:x\mapsto x+1, which has the basis A1=𝒞1+[0,12)A_{1}=\mathcal{C}_{1}+[0,\frac{1}{2}) and ceiling B1=𝒟1+[1,12)B_{1}=\mathcal{D}_{1}+[-1,-\frac{1}{2}), and the castle ψ1:xx1\psi_{1}:x\mapsto x-1 defined for xC1x\in C_{1} with ceiling D1=𝒞1+[12,1)D_{1}=\mathcal{C}_{1}+[\frac{1}{2},1).

We now concentrate on the induction step: suppose φn\varphi_{n} and ψn\psi_{n} have been built for some n1n\geq 1, let us construct φn+1\varphi_{n+1} and ψn+1\psi_{n+1}.

The strategy is to split the basis of φn\varphi_{n} and ψn\psi_{n} into two equal intervals and “interleave” the “two halves” of φn\varphi_{n} with “one half” of ψn\psi_{n} followed by “gluing” adjacent ceilings and basis within the same 𝒞n+1\mathcal{C}_{n+1} segment (see Figure 7.2). To this end, we introduce two intermediate castles φ~n\tilde{\varphi}_{n} and ψ~n\tilde{\psi}_{n} which will ensure that φn+1\varphi_{n+1} “wiggles” more than φn\varphi_{n}, yielding ergodicity of the final transformation.

Define two new half measure subsets of the bases AnA_{n} and CnC_{n} respectively:

  • An1=𝒞n+[0,12n+1)A_{n}^{1}=\mathcal{C}_{n}+\left[0,\frac{1}{2^{n+1}}\right);

  • Cn0=𝒟n+[12+12n+1,12+12n)C_{n}^{0}=\mathcal{D}_{n}+\left[-\frac{1}{2}+\frac{1}{2^{n+1}},-\frac{1}{2}+\frac{1}{2^{n}}\right);

and let

Bn0=φn(An1)=𝒟n+[1212n,1212n+1),B_{n}^{0}=\vec{\varphi}_{n}(A_{n}^{1})=\mathcal{D}_{n}+\left[-\frac{1}{2}-\frac{1}{2^{n}},-\frac{1}{2}-\frac{1}{2^{n+1}}\right),

and

Dn0=ψn(Cn0)=𝒞n+[12+12n+1,12+12n),D_{n}^{0}=\vec{\psi}_{n}(C_{n}^{0})=\mathcal{C}_{n}+\left[\frac{1}{2}+\frac{1}{2^{n+1}},\frac{1}{2}+\frac{1}{2^{n}}\right),

where the two equalities are consequences of the translation conditions. Let EnE_{n} be the ψn\psi_{n}-saturation of Cn0C_{n}^{0}, and note that the restriction of ψn\psi_{n} to EnE_{n} is a castle with support EnE_{n}, whose basis is Cn0C^{0}_{n} and whose ceiling is Dn0D_{n}^{0}. Finally, let

An0=AnAn1=𝒞n+[12n+1,12n).A_{n}^{0}=A_{n}\setminus A_{n}^{1}=\mathcal{C}_{n}+\left[\frac{1}{2^{n+1}},\frac{1}{2^{n}}\right).

We define the partial measure-preserving transformation ξn:Bn0Dn0Cn0An0\xi_{n}:B_{n}^{0}\sqcup D_{n}^{0}\to C_{n}^{0}\sqcup A_{n}^{0} to be used for “gluing together” φn\varphi_{n} and the restriction of ψn\psi_{n} to EnE_{n}:

  • ξn(b)=b+32n+1Cn0\xi_{n}(b)=b+\frac{3}{2^{n+1}}\in C_{n}^{0} for all bBn0b\in B_{n}^{0} and

  • ξn(d)=d12An0\xi_{n}(d)=d-\frac{1}{2}\in A_{n}^{0} for all dDn0d\in D_{n}^{0}.

Set φ~n=φnξnψnEn\tilde{\varphi}_{n}=\varphi_{n}\sqcup\xi_{n}\sqcup\psi_{n\restriction E_{n}}, whereas ψ~n\tilde{\psi}_{n} is simply the restriction of ψn\psi_{n} onto the complement of EnE_{n}. Observe that φ~n\tilde{\varphi}_{n} has basis An1A_{n}^{1} and ceiling

Bn1=BnBn0=𝒟n+[1212n+1,12),B_{n}^{1}=B_{n}\setminus B_{n}^{0}=\mathcal{D}_{n}+\left[-\frac{1}{2}-\frac{1}{2^{n+1}},-\frac{1}{2}\right),

while ψ~n\tilde{\psi}_{n} has basis

Cn1=CnCn0=𝒟n+[12,12+12n+1)C^{1}_{n}=C_{n}\setminus C_{n}^{0}=\mathcal{D}_{n}+\left[-\frac{1}{2},-\frac{1}{2}+\frac{1}{2^{n+1}}\right)

and ceiling

Dn1=DnDn0=𝒞n+[12,12+12n+1).D_{n}^{1}=D_{n}\setminus D_{n}^{0}=\mathcal{C}_{n}+\left[\frac{1}{2},\frac{1}{2}+\frac{1}{2^{n+1}}\right).

We continue to have Y=suppφ~nsuppψ~nY=\operatorname*{supp}\tilde{\varphi}_{n}\sqcup\operatorname*{supp}\tilde{\psi}_{n}, but the support of ψ~n\tilde{\psi}_{n} is half the support of ψn\psi_{n}, and μ(suppψ~n)=12μ(suppψn)\mu(\operatorname*{supp}\tilde{\psi}_{n})=\frac{1}{2}\mu(\operatorname*{supp}\psi_{n}).

AnA_{n}BnB_{n}Bn0B^{0}_{n}Bn1B^{1}_{n}\longrightarrowφn\varphi_{n}DnD_{n}\longleftarrowψn\psi_{n}CnC_{n}AnA_{n}BnB_{n}\longrightarrowφn\varphi_{n}DnD_{n}\longleftarrowψn\psi_{n}CnC_{n}An1A^{1}_{n}An0A^{0}_{n}Dn0D_{n}^{0}Dn1D_{n}^{1}Cn0C_{n}^{0}Cn1C_{n}^{1}ξn\xi_{n}ξn\xi_{n}An1A^{1}_{n}An0A^{0}_{n}Bn0B^{0}_{n}Bn1B^{1}_{n}Dn0D_{n}^{0}Dn1D_{n}^{1}Cn0C_{n}^{0}Cn1C_{n}^{1}ξn\xi_{n}ξn\xi_{n}ξn\xi^{\prime}_{n}ξn′′\xi^{\prime\prime}_{n}
Figure 7.2. Inductive step.

The ceiling of φ~n\tilde{\varphi}_{n} is equal to Bn1=𝒟n+[1212n+1,12)B_{n}^{1}=\mathcal{D}_{n}+\left[-\frac{1}{2}-\frac{1}{2^{n+1}},-\frac{1}{2}\right), whereas we need the ceiling of φn+1\varphi_{n+1} to be equal to Bn+1=𝒟n+1+[1212n+1,12)B_{n+1}=\mathcal{D}_{n+1}+\left[-\frac{1}{2}-\frac{1}{2^{n+1}},-\frac{1}{2}\right). We obtain the required φn+1\varphi_{n+1} and ψn+1\psi_{n+1} out of φ~n\tilde{\varphi}_{n} and ψ~n\tilde{\psi}_{n} respectively by “passing through each element of 𝒞n𝒞n+1\mathcal{C}_{n}\setminus\mathcal{C}_{n+1} ”.

Note that 𝒟n+1\mathcal{D}_{n+1} is equal to the set of d𝒟nd\in\mathcal{D}_{n} such that κ𝒞n(d)𝒞n+1\kappa_{\mathcal{C}_{n}}(d)\in\mathcal{C}_{n+1}. Each xBn1Bn+1x\in B_{n}^{1}\setminus B_{n+1} can be written uniquely as x=d+tx=d+t where d𝒟n𝒟n+1d\in\mathcal{D}_{n}\setminus\mathcal{D}_{n+1} and t[1212n+1,12)t\in\left[-\frac{1}{2}-\frac{1}{2^{n+1}},-\frac{1}{2}\right). Set

ξn(x)=κ𝒞n(d)+t+12+12n+1,\xi^{\prime}_{n}(x)=\kappa_{\mathcal{C}_{n}}(d)+t+\frac{1}{2}+\frac{1}{2^{n+1}},

and note that ξn(x)\xi^{\prime}_{n}(x) belongs to (𝒞n𝒞n+1)+[0,12n+1)=An1An+1(\mathcal{C}_{n}\setminus\mathcal{C}_{n+1})+\left[0,\frac{1}{2^{n+1}}\right)=A_{n}^{1}\setminus A_{n+1}, hence ξn\xi^{\prime}_{n} is a measure-preserving bijection from Bn1Bn+1B_{n}^{1}\setminus B_{n+1} onto An1An+1A_{n}^{1}\setminus A_{n+1}.

The transformation φn+1\varphi_{n+1} is set to be φ~nξn\tilde{\varphi}_{n}\sqcup\xi^{\prime}_{n}, and we claim that it is a castle with basis An+1A_{n+1} and ceiling Bn+1B_{n+1}. This amounts to showing that for all xAn+1x\in A_{n+1}, there is kk\in\mathbb{N} such that φn+1k(x)\varphi_{n+1}^{k}(x) is not defined. Pick xAn+1x\in A_{n+1} and write it as c0+tc_{0}+t for some c0𝒞n+1c_{0}\in\mathcal{C}_{n+1} and t[0,12n+1)t\in[0,\frac{1}{2^{n+1}}). Let c1c_{1} be the successor of c0c_{0} in 𝒞n\mathcal{C}_{n}, which we suppose not to be an element of 𝒞n+1\mathcal{C}_{n+1}. By the construction of φ~n\tilde{\varphi}_{n} and ξn\xi^{\prime}_{n}, there is kk\in\mathbb{N} such that ξn(φ~nk(x))c+[0,12n+1)\xi^{\prime}_{n}(\tilde{\varphi}_{n}^{k}(x))\in c^{\prime}+[0,\frac{1}{2^{n+1}}), which means that φn+1k+1(x)c+[0,12n+1)\varphi_{n+1}^{k+1}(x)\in c^{\prime}+[0,\frac{1}{2^{n+1}}). Iterating this argument, we eventually find k0,pk_{0},p\in\mathbb{N} such that φn+1k0(x)cp+[0,12n+1)\varphi_{n+1}^{k_{0}}(x)\in c_{p}+[0,\frac{1}{2^{n+1}}) for some cp𝒞nc_{p}\in\mathcal{C}_{n} such that the successor cp+1c_{p+1} of cpc_{p} in 𝒞n\mathcal{C}_{n} belongs to 𝒞n+1\mathcal{C}_{n+1}. By the definition of φ~n\tilde{\varphi}_{n} we must have some ll\in\mathbb{N} such that φn+1k0+l(x)=φ~nl(φn+1k0(x))Bn+1\varphi_{n+1}^{k_{0}+l}(x)=\tilde{\varphi}_{n}^{l}(\varphi_{n+1}^{k_{0}}(x))\in B_{n+1}, whereas φn+1k0+l+1(x)\varphi_{n+1}^{k_{0}+l+1}(x) is not defined, thus φn+1\varphi_{n+1} is indeed a castle.

Extension ψn+1\psi_{n+1} of ψ~n\tilde{\psi}_{n} is defined similarly by connecting adjacent segments of Dn1D^{1}_{n} and Cn1C_{n}^{1} by a translation. More specifically, each xDn1Dn+1x\in D^{1}_{n}\setminus D_{n+1} can be written uniquely as x=c+tx=c+t for some c𝒞n𝒞n+1c\in\mathcal{C}_{n}\setminus\mathcal{C}_{n+1} and t[12,12+12n+1)t\in[\frac{1}{2},\frac{1}{2}+\frac{1}{2^{n+1}}). The restriction of κ𝒞n\kappa_{\mathcal{C}_{n}} to 𝒟n\mathcal{D}_{n} is a bijection 𝒟n𝒞n\mathcal{D}_{n}\to\mathcal{C}_{n}, we denote its inverse by pnp_{n} and let ξn′′(x)=pn(c)+t1\xi^{\prime\prime}_{n}(x)=p_{n}(c)+t-1. The map ψn+1=ψ~nξn′′\psi_{n+1}=\tilde{\psi}_{n}\sqcup\xi^{\prime\prime}_{n} can be checked to be a castle with basis Cn+1C_{n+1} and ceiling Dn+1D_{n+1} as desired. It also follows that the translation conditions continue to be satisfied by both of φn+1\varphi_{n+1} and ψn+1\psi_{n+1}.

Transformations φn\varphi_{n} extend each other, so φ=nφn\varphi=\bigcup_{n}\varphi_{n} is an element of the pseudo full group supported on Y=suppφnsuppψnY=\operatorname*{supp}\varphi_{n}\sqcup\operatorname*{supp}\psi_{n}. Note also that

μ(suppψn+1)=μ(suppψn)/2,\mu(\operatorname*{supp}\psi_{n+1})=\mu(\operatorname*{supp}\psi_{n})/2,

and therefore domφ=Y=rngφ\operatorname{\mathrm{dom}}\varphi=Y=\operatorname{\mathrm{rng}}\varphi. We claim that φ\varphi, seen as a measure-preserving transformation of YY, induces an ergodic measure-preserving transformation on (y+)Y(y+\mathbb{R})\cap Y for almost all yYy\in Y, where y+y+\mathbb{R} is endowed with the Lebesgue measure. This follows from the fact that φ\varphi induces a rank-one transformation of the infinite measure space (y+)Y(y+\mathbb{R})\cap Y: for all Borel A(y+)YA\subseteq(y+\mathbb{R})\cap Y of finite Lebesgue measure and all ϵ>0\epsilon>0, there are B(y+)YB\subseteq(y+\mathbb{R})\cap Y, kk\in\mathbb{N}, and a subset F{0,,k}F\subseteq\{0,\dots,k\} such that B,φ(B),,φk(B)B,\varphi(B),\dots,\varphi^{k}(B) are pairwise disjoint and

λ(A(fFφf(B)))<ϵ.\lambda(A\bigtriangleup(\bigsqcup_{f\in F}\varphi^{f}(B)))<\epsilon.

Indeed, at each step nn for every c𝒞nc\in\mathcal{C}_{n}, the iterates of c+[0,12n)c+[0,\frac{1}{2^{n}}) by the restriction of φn\varphi_{n} to the interval [c,ιn(c))[c,\iota_{n}(c)) are disjoint “intervals of size 2n2^{-n}”, i.e., sets of the form t+[0,12n)t+[0,\frac{1}{2^{n}}), and these iterates cover a proportion 112n1-\frac{1}{2^{n}} of [c,ιn(c))[c,\iota_{n}(c)) (the rest of this interval being [c,ιn(c))suppψn[c,\iota_{n}(c))\cap\operatorname*{supp}\psi_{n}).

It remains to extend φ\varphi supported on YY to a measure-preserving transformation SS with suppS=X\operatorname*{supp}S=X. Let Z=XYZ=X\setminus Y be the leftover set,

Z={c+t:c𝒞1:2t<gap𝒞1(c)},Z=\{c+t:c\in\mathcal{C}_{1}:2\leq t<\mathrm{gap}_{\mathcal{C}_{1}}(c)\},

and put

Z={c+t:c𝒞1, 2gap𝒞1(c)t<2}.Z^{\prime}=\{c+t:c\in\mathcal{C}_{1},\ 2-\mathrm{gap}_{\mathcal{C}_{1}}(c)\leq t<2\}.

Figure 7.3 illustrates an interval between c𝒞1c\in\mathcal{C}_{1} and c=σ𝒞1(c)c^{\prime}=\sigma_{\mathcal{C}_{1}}(c). Within this gap, ZZ corresponds to [c+2,c+2+gap𝒞1(c))[c+2,c+2+\mathrm{gap}_{\mathcal{C}_{1}}(c)), and ZZ^{\prime} is an interval of the exact same length adjacent to it on the left. Note that ZYZ^{\prime}\subseteq Y by construction. Let η:ZZ\eta:Z^{\prime}\to Z be the natural translation map, η(x)=x+gap𝒞1(c)\eta(x)=x+\mathrm{gap}_{\mathcal{C}_{1}}(c) for all xZx\in Z^{\prime} satisfying xc+[0,gap𝒞1(c))x\in c+[0,\mathrm{gap}_{\mathcal{C}_{1}}(c)). Observe that η\eta is a measure-preserving bijection and its cocycle is bounded by 11.

\cdotscccc^{\prime}ZZ ZZ^{\prime} φ\varphiSS
Figure 7.3. Construction of the transformation SS.

We now rewire the orbits of φ\varphi and define S:XXS:X\to X as follows (see Figure 7.3):

S(x)={φ(x)if xZZ;η(x)if xZ;φ(η1(x))if xZ.S(x)=\begin{cases}\varphi(x)&\textrm{if $x\not\in Z\cup Z^{\prime}$};\\ \eta(x)&\textrm{if $x\in Z^{\prime}$};\\ \varphi(\eta^{-1}(x))&\textrm{if $x\in Z$}.\end{cases}

It is straightforward to verify that SS is a free measure-preserving transformation, and the distance D(x,Sx)4D(x,Sx)\leq 4 for all xXx\in X, because |ρφ(x)|3|\rho_{\varphi}(x)|\leq 3 and |ρη(x)|1|\rho_{\eta}(x)|\leq 1 for all xx in their domains. Note that the transformation induced by SS on YY is equal to φ\varphi, so since the latter is ergodic on every orbit of the flow intersected with YY and since X=YZX=Y\sqcup Z, it follows that SS is ergodic on every orbit of the flow and satisfies the conclusion of the theorem. ∎

Remark 7.3.

The bound 44 in the formulation of Theorem 7.2 is of no significance as by rescaling the flow it can be replaced with any ϵ>0\epsilon>0.

Chapter 8 Conservative and intermitted transformations

Interesting dynamics of conservative transformations is present only in the non-discrete case, as it reduces to periodicity for countable group actions. Chapter 7 provides an illustrative construction of a conservative automorphism, and shows that they exist in L1\mathrm{L}^{1} full groups of all free flows. The present chapter is devoted to the study of such elements. The central role is played by the concept of an intermitted transformation, which is related to the notion of induced transformation. Using this tool we show that all conservative elements of [X]1[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} can be approximated by periodic automorphisms, and hence belong to the derived L1\mathrm{L}^{1} full group of X\mathbb{R}\curvearrowright X; see Corollary 8.8.

Throughout the chapter, we fix a free measure-preserving flow X\mathbb{R}\curvearrowright X on a standard Lebesgue space (X,μ)(X,\mu). Given a cross-section 𝒞X\mathcal{C}\subset X, recall that we defined an equivalence relation 𝒞\mathcal{R}_{\mathcal{C}} by declaring x𝒞yx\mathcal{R}_{\mathcal{C}}y whenever there is c𝒞c\in\mathcal{C} such that both xx and yy belong to the gap between cc and σ𝒞(c)\sigma_{\mathcal{C}}(c). More formally, x𝒞yx\mathcal{R}_{\mathcal{C}}y if there is c𝒞c\in\mathcal{C} such that ρ(c,x)0\rho(c,x)\geq 0, ρ(c,y)0\rho(c,y)\geq 0 and ρ(x,σ𝒞(c))>0\rho(x,\sigma_{\mathcal{C}}(c))>0, ρ(y,σ𝒞(c))>0\rho(y,\sigma_{\mathcal{C}}(c))>0. Such an equivalence relation is smooth.

Now let T[X]T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu] be a conservative transformation. Under the action of TT, almost every point returns to its 𝒞\mathcal{R}_{\mathcal{C}}-class infinitely often, which suggests the idea of the first return map.

Definition 8.1.

The intermitted transformation T𝒞:XXT_{\mathcal{R}_{\mathcal{C}}}:X\to X is defined by

T𝒞x=Tn(x)x,where n(x)=min{n1:x𝒞Tn(x)x}.T_{\mathcal{R}_{\mathcal{C}}}x=T^{n(x)}x,\quad\textrm{where }n(x)=\min\{n\geq 1:x\mathcal{R}_{\mathcal{C}}T^{n(x)}x\}.

The map T𝒞T_{\mathcal{R}_{\mathcal{C}}} is well-defined, since TT is conservative, and it preserves the measure μ\mu, since T𝒞T_{\mathcal{R}_{\mathcal{C}}} belongs to the full group of TT.

Remark 8.2.

The concept of an intermitted transformation TET_{E} makes sense for any equivalence relation EE for which intersection of any orbit of TT with any EE-class is either empty or infinite. In particular, intermitted transformations can be considered for any conservative T[GX]T\in[G\curvearrowright X\mkern 1.5mu] in a full group of a locally compact group action. For instance, with a cocompact cross-section 𝒞\mathcal{C} we can associate an equivalence relation of lying in same cell of the Voronoi tessellation (see Appendix C.2). Such an equivalence relation does have the aforementioned transversal property, and hence intermitted transformation is well-defined.

Note also the following connection with the more familiar construction of the induced transformation. Let TAut(X,μ)T\in\mathrm{Aut}(X,\mu), let AXA\subseteq X be a set of positive measure, and define 𝒜\mathcal{A} to be the equivalence relation with two classes: AA and XAX\setminus A. Induced transformations TAT_{A} and TXAT_{X\setminus A} commute and satisfy TATXA=T𝒜T_{A}\circ T_{X\setminus A}=T_{\mathcal{A}}.

The next lemma forms the core of this chapter. It shows that the operation of taking an intermitted transformation does not increase the norm. As we discuss later in Remark 8.5, the analog of this statement is false even for 2\mathbb{R}^{2}-flows, which perhaps justifies the technical nature of the argument.

Lemma 8.3.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a conservative automorphism and let 𝒞\mathcal{C} be a cross-section. Let also YY be the set of points where TT and T𝒞T_{\mathcal{R}_{\mathcal{C}}} differ: Y={xX:TxT𝒞x}Y=\{x\in X:Tx\neq T_{\mathcal{R}_{\mathcal{C}}}x\}. One has Y|ρT𝒞|𝑑μY|ρT|𝑑μ\int_{Y}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}|\,d\mu\leq\int_{Y}|\rho_{T}|\,d\mu.

Proof.

By the definition of YY, for any xYx\in Y the arc from xx to TxTx jumps over at least one point of 𝒞\mathcal{C}. We may therefore represent |ρT(x)||\rho_{T}(x)| as the sum of the distance from xx to the first point of 𝒞\mathcal{C} along the arc plus the rest of the arc. More formally, for xXx\in X let π𝒞(x)\pi_{\mathcal{C}}(x) be the unique c𝒞c\in\mathcal{C} such that xc+[0,gap𝒞(c))x\in c+[0,\mathrm{gap}_{\mathcal{C}}(c)). Define α:Y0\alpha:Y\to\mathbb{R}^{\geq 0} by

α(x)={|ρ(x,σ𝒞(π𝒞(x)))|,if ρ(x,Tx)>0,|ρ(x,π𝒞(x))|if ρ(x,Tx)<0.\alpha(x)=\begin{cases}|\rho(x,\sigma_{\mathcal{C}}(\pi_{\mathcal{C}}(x)))|,&\textrm{if $\rho(x,Tx)>0$},\\ |\rho(x,\pi_{\mathcal{C}}(x))|&\textrm{if $\rho(x,Tx)<0$}.\end{cases}

Note that α(x)|ρT(x)|\alpha(x)\leq|\rho_{T}(x)|, and set β(x)=|ρT(x)|α(x)\beta(x)=|\rho_{T}(x)|-\alpha(x), so that

Y|ρT|𝑑μ=Yα𝑑μ+Yβ𝑑μ.\int_{Y}|\rho_{T}|\,d\mu=\int_{Y}\alpha\,d\mu+\int_{Y}\beta\,d\mu.

For instance, in the context of Figure 8.1, α(x4)=ρ(x4,c2)\alpha(x_{4})=\rho(x_{4},c_{2}) and β(x4)=ρ(c2,x5)\beta(x_{4})=\rho(c_{2},x_{5}). Let us partition Y=YY′′Y=Y^{\prime}\sqcup Y^{\prime\prime}, where

Y={xY:ρ(x,Tx) and ρ(x,T𝒞x) have the same sign or T𝒞x=x},Y^{\prime}=\bigl{\{}x\in Y:\rho(x,Tx)\textrm{ and }\rho(x,T_{\mathcal{R}_{\mathcal{C}}}x)\textrm{ have the same sign or $T_{\mathcal{R}_{\mathcal{C}}}x=x$}\,\bigr{\}},

and Y′′=YYY^{\prime\prime}=Y\setminus Y^{\prime} consists of those xYx\in Y for which the signs of ρ(x,Tx)\rho(x,Tx) and ρ(x,T𝒞x)\rho(x,T_{\mathcal{R}_{\mathcal{C}}}x) are different. For example, referring to the same figure, x0Y′′x_{0}\in Y^{\prime\prime}, while x2Yx_{2}\in Y^{\prime}.

To prove the lemma it is enough to show two inequalities:

(8.1) Y|ρT𝒞(x)|𝑑μ(x)Yα(x)𝑑μ(x),\int_{Y^{\prime}}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)|\,d\mu(x)\leq\int_{Y}\alpha(x)\,d\mu(x),
(8.2) Y′′|ρT𝒞(x)|𝑑μ(x)Yβ(x)𝑑μ(x).\int_{Y^{\prime\prime}}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)|\,d\mu(x)\leq\int_{Y}\beta(x)\,d\mu(x).

Eq. (8.1) is straightforward, since equality of signs of ρ(x,Tx)\rho(x,Tx) and ρ(x,T𝒞x)\rho(x,T_{\mathcal{R}_{\mathcal{C}}}x) implies that T𝒞xT_{\mathcal{R}_{\mathcal{C}}}x is closer than xx to the point c𝒞c\in\mathcal{C} over which goes the arc from xx to TxTx. For example, the point x2x_{2} in Figure 8.1 satisfies

|ρT𝒞(x2)|=ρ(x2,x4)ρ(x2,c2)=α(x2).|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x_{2})|=\rho(x_{2},x_{4})\leq\rho(x_{2},c_{2})=\alpha(x_{2}).

Thus |ρT𝒞(x)|α(x)|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)|\leq\alpha(x) for all xYx\in Y^{\prime} and so

Y|ρT𝒞|𝑑μYα𝑑μYα𝑑μ,\int_{Y^{\prime}}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}|\,d\mu\leq\int_{Y^{\prime}}\alpha\,d\mu\leq\int_{Y}\alpha\,d\mu,

which gives (8.1). The other inequality will take us a bit more work.

For xY′′x\in Y^{\prime\prime}, let N(x)1N(x)\geq 1 be the smallest integer such that the sign of ρ(x,TN(x)+1x)\rho(x,T^{N(x)+1}x) is opposite to that of ρT(x)\rho_{T}(x). In less formal words, N(x)N(x) is the smallest integer such that the arc from TN(x)xT^{N(x)}x to TN(x)+1xT^{N(x)+1}x jumps over xx. In particular, points TkxT^{k}x, 1kN(x)1\leq k\leq N(x), are all on the same side relative to xx, while TN(x)+1xT^{N(x)+1}x is on the other side of it. We consider the map η:Y′′X\eta:Y^{\prime\prime}\to X given by η(x)=TN(x)x\eta(x)=T^{N(x)}x. Properties of this map will be crucial for establishing the inequality (8.2), so let us provide some explanations first.

c0c_{0}c1c_{1}c2c_{2}c3c_{3}c4c_{4}x0x_{0}x1x_{1}x2x_{2}x3x_{3}x4x_{4}x5x_{5}x6x_{6}x7x_{7}x8x_{8}x9x_{9}
Figure 8.1. Dynamics of a conservative orbit.

Consider once again Figure 8.1, which shows a partial orbit of a point x0x_{0} for xi=Tix0x_{i}=T^{i}x_{0} up to i9i\leq 9 and several points ci𝒞c_{i}\in\mathcal{C}. First, as we have already noted before, x0Yx_{0}\in Y, since ¬x0𝒞x1\neg x_{0}\mathcal{R}_{\mathcal{C}}x_{1}; moreover, x0Y′′x_{0}\in Y^{\prime\prime}, since x9=T𝒞x0x_{9}=T_{\mathcal{R}_{\mathcal{C}}}x_{0} is to the left of x0x_{0}, while x1x_{1} is to the right of it, so ρ(x0,x1)\rho(x_{0},x_{1}) and ρ(x0,x9)\rho(x_{0},x_{9}) have the opposite signs. Also, N(x0)=7N(x_{0})=7, because x8x_{8} is the first point in the orbit to left of x0x_{0}, thus η(x0)=x7\eta(x_{0})=x_{7}. In particular, generally TN(x)+1xT𝒞xT^{N(x)+1}x\neq T_{\mathcal{R}_{\mathcal{C}}}x, but TN(x)+1x=T𝒞xT^{N(x)+1}x=T_{\mathcal{R}_{\mathcal{C}}}x is the case for xY′′x\in Y^{\prime\prime} whenever TN(x)+1xT^{N(x)+1}x and xx are 𝒞\mathcal{R}_{\mathcal{C}}-equivalent.

The next point in the orbit x1Yx_{1}\not\in Y, whereas x2Yx_{2}\in Y but x2Y′′x_{2}\not\in Y^{\prime\prime}, because T𝒞x2=x4T_{\mathcal{R}_{\mathcal{C}}}x_{2}=x_{4} and both ρ(x2,x3)\rho(x_{2},x_{3}) and ρ(x2,x4)\rho(x_{2},x_{4}) are positive. The point x3x_{3} belongs to Y′′Y^{\prime\prime} and has N(x3)=1N(x_{3})=1 with η(x3)=x4\eta(x_{3})=x_{4}. Points x4,x5,x6Yx_{4},x_{5},x_{6}\in Y, but whether any of them are elements of Y′′Y^{\prime\prime} is not clear from Figure 8.1, as the orbit segment is too short to clarify the values of T𝒞xiT_{\mathcal{R}_{\mathcal{C}}}x_{i}, i=4,5,6i=4,5,6. However, if x4,x5,x6x_{4},x_{5},x_{6} happen to lie in Y′′Y^{\prime\prime}, then N(x5)=1N(x_{5})=1 with η(x5)=x6\eta(x_{5})=x_{6}, and N(x4)=3N(x_{4})=3, N(x6)=1N(x_{6})=1, η(x4)=η(x6)=x7=η(x0)\eta(x_{4})=\eta(x_{6})=x_{7}=\eta(x_{0}). In particular, the function xη(x)x\mapsto\eta(x) is not necessarily one-to-one, but we are going to argue that it is always finite-to-one.

Claim 1.

If x,yY′′x,y\in Y^{\prime\prime} are distinct points such that η(x)=η(y)\eta(x)=\eta(y), then ¬x𝒞y\neg x\mathcal{R}_{\mathcal{C}}y.

Proof of the claim.

Suppose x,yY′′x,y\in Y^{\prime\prime} satisfy η(x)=η(y)\eta(x)=\eta(y). The definition of η\eta implies that xx and yy must belong to the same orbit of TT, and we may assume without loss of generality that y=Tk0xy=T^{k_{0}}x for some k01k_{0}\geq 1. If the orbit of xx and yy is aperiodic, it implies that that N(x)>k0N(x)>k_{0} and N(y)+k0=N(x)N(y)+k_{0}=N(x), N(y)1N(y)\geq 1. However, even if the orbit is periodic, either N(y)+k0=N(x)N(y)+k_{0}=N(x) for the smallest positive integer k0k_{0} such that y=Tk0xy=T^{k_{0}}x or N(x)+k0=N(y)N(x)+k_{0}^{\prime}=N(y) for the smallest positive integer k0k_{0}^{\prime} such that x=Tk0yx=T^{k_{0}^{\prime}}y. Interchanging the roles of xx and yy if necessary, we may therefore assume that N(y)+k0=N(x)N(y)+k_{0}=N(x) holds for some k01k_{0}\geq 1, Tk0x=yT^{k_{0}}x=y, regardless of the type of orbit we consider.

Suppose xx and yy are 𝒞\mathcal{R}_{\mathcal{C}}-equivalent. Let k1k\geq 1 be the smallest natural number for which xx and TkxT^{k}x are 𝒞\mathcal{R}_{\mathcal{C}}-equivalent. By the assumption x𝒞yx\mathcal{R}_{\mathcal{C}}y and the choice of k0k_{0} we have kk0<N(x)k\leq k_{0}<N(x). By the definition of N(x)N(x), all points TixT^{i}x, 1iN(x)1\leq i\leq N(x), are on the same side of xx. In particular, this applies to TxTx and TkxT^{k}x, which shows that ρ(x,Tx)\rho(x,Tx) and ρ(x,T𝒞x)\rho(x,T_{\mathcal{R}_{\mathcal{C}}}x) have the same sign, thus xY′′x\not\in Y^{\prime\prime}. ∎

The above claim implies that the function xη(x)x\mapsto\eta(x) is finite-to-one for the arc from η(x)\eta(x) to Tη(x)T\eta(x) intersects only finitely many 𝒞\mathcal{R}_{\mathcal{C}}-equivalence classes, and the preimage of η(x)\eta(x) picks at most one point from each such class. Note also that η(x)Y\eta(x)\in Y for all xY′′x\in Y^{\prime\prime}, but η(x)\eta(x) may not be an element of Y′′Y^{\prime\prime}. Among the 𝒞\mathcal{R}_{\mathcal{C}}-equivalence classes that the arc from η(x)\eta(x) to Tη(x)T\eta(x) goes over, two are special—the intervals that contain Tη(x)T\eta(x) and η(x)\eta(x), respectively. Our goal will be to bound the sum of |ρT𝒞(x)||\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)| over the points xx with the same η(x)\eta(x) value by β(η(x))\beta(\eta(x)) (see Claim 3 below). For a typical point xx we can bound |ρT𝒞(x)||\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)| simply by the length of the interval of its 𝒞\mathcal{R}_{\mathcal{C}}-class. For example, Figure 8.1 does not specify T𝒞x4T_{\mathcal{R}_{\mathcal{C}}}x_{4}, but we can be sure that |ρT𝒞(x4)|ρ(c1,c2)|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x_{4})|\leq\rho(c_{1},c_{2}). In view of Claim 1, such an estimate comes close to showing that the sum of |ρT𝒞(x)||\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)| over xx with the same image η(x)\eta(x) is bounded by |ρ(η(x),Tη(x))||\rho(\eta(x),T\eta(x))|. It merely comes close, due to the two special 𝒞\mathcal{R}_{\mathcal{C}}-classes mentioned above, where our estimate needs to be improved. The next claim shows that one of these special cases is of no concern as xx is never 𝒞\mathcal{R}_{\mathcal{C}}-equivalent to η(x)\eta(x).

Claim 2.

For all xY′′x\in Y^{\prime\prime} we have ¬x𝒞η(x)\neg x\mathcal{R}_{\mathcal{C}}\eta(x).

Proof of the claim.

Suppose towards the contradiction that x𝒞η(x)x\mathcal{R}_{\mathcal{C}}\eta(x), and let k1k\geq 1 be the smallest integer for which x𝒞Tk(x)x\mathcal{R}_{\mathcal{C}}T^{k}(x); in particular, T𝒞x=TkxT_{\mathcal{R}_{\mathcal{C}}}x=T^{k}x. Note that kN(x)k\leq N(x) by the assumption, and by the definition of N(x)N(x), ρ(Tkx,x)\rho(T^{k}x,x) has the same sign as ρT(x)\rho_{T}(x), whence xY′′x\not\in Y^{\prime\prime}. ∎

Pick some yYy\in Y with non-empty preimage η1(y)\eta^{-1}(y), and let z1,,znY′′z_{1},\ldots,z_{n}\in Y^{\prime\prime} be all the elements in η1(y)\eta^{-1}(y). For instance, in the situation depicted in Figure 8.1, we may have n=3n=3 and z1=x0z_{1}=x_{0}, z2=x4z_{2}=x_{4}, z3=x6z_{3}=x_{6}, and y=x7y=x_{7}. The following claim unlocks the path towards the inequality (8.2).

Claim 3.

In the above notation, i=1n|ρT𝒞(zi)|β(y)\sum_{i=1}^{n}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(z_{i})|\leq\beta(y).

Proof of the claim.

Recall that the arc from yy to TyTy crosses at least one point in 𝒞\mathcal{C}. If c𝒞c\in\mathcal{C} is the first such point, then β(y)\beta(y) is defined to be |ρ(c,Ty)||\rho(c,Ty)|. For instance, in the notation of Figure 8.1, β(x7)=|ρ(c4,x8)|\beta(x_{7})=|\rho(c_{4},x_{8})|. Each point ziz_{i} is located under the arc from yy to TyTy, and by Claim 2, no point ziz_{i} belongs to the interval from cc to yy. In the language of our concrete example, no point ziz_{i} can be between c4c_{4} and x7x_{7}. As discussed before, |ρT𝒞(x)||\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)| is always bounded by the length of the gap to which xx belongs. This is sufficient to prove the claim if no ziz_{i} is equivalent to TyTy, as in this case the whole 𝒞\mathcal{R}_{\mathcal{C}}-equivalence class of every ziz_{i} is fully contained under the interval between cc and TyTy, and distinct ziz_{i} represent distinct 𝒞\mathcal{R}_{\mathcal{C}}-classes by Claim 1. This is the situation depicted in Figure 8.1, and our argument boils down to the inequalities

|ρT𝒞(x0)|+|ρT𝒞(x4)|+|ρT𝒞(x5)|\displaystyle|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x_{0})|+|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x_{4})|+|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x_{5})| |ρ(c0,c1)|+|ρ(c1,c2)|+|ρ(c2,c3)|\displaystyle\leq|\rho(c_{0},c_{1})|+|\rho(c_{1},c_{2})|+|\rho(c_{2},c_{3})|
|ρ(c0,c4)|β(x7).\displaystyle\leq|\rho(c_{0},c_{4})|\leq\beta(x_{7}).

Suppose there is some ziz_{i} such that zi𝒞Tyz_{i}\mathcal{R}_{\mathcal{C}}Ty. By Claim 1 such ziz_{i} must be unique, and we assume without loss of generality that z1𝒞Tyz_{1}\mathcal{R}_{\mathcal{C}}Ty. For example, this situation would occur if in Figure 8.1 Tx7Tx_{7} were equal to x9x_{9}. Let cc^{\prime} be the first element of 𝒞\mathcal{C} over which goes the arc from z1z_{1} to Tz1Tz_{1} (it would be the point c1c_{1} in Figure 8.1). It is enough to show that |ρT𝒞(z1)||ρ(T𝒞z1,c)||\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(z_{1})|\leq|\rho(T_{\mathcal{R}_{\mathcal{C}}}z_{1},c^{\prime})|, as we can use the previous estimate for all other |ρT𝒞(zi)||\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(z_{i})|, i2i\geq 2. Note that T𝒞z1=TyT_{\mathcal{R}_{\mathcal{C}}}z_{1}=Ty, and z1Y′′z_{1}\in Y^{\prime\prime} by assumption, which implies that the signs of ρ(z1,T𝒞z1)\rho(z_{1},T_{\mathcal{R}_{\mathcal{C}}}z_{1}) and ρ(z1,c)\rho(z_{1},c^{\prime}) are different. The latter is equivalent to saying that z1z_{1} is between T𝒞z1T_{\mathcal{R}_{\mathcal{C}}}z_{1} and cc^{\prime}, i.e., |ρ(T𝒞z1,c)|=|ρT𝒞(z1)|+|ρ(z1,c)||\rho(T_{\mathcal{R}_{\mathcal{C}}}z_{1},c^{\prime})|=|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(z_{1})|+|\rho(z_{1},c^{\prime})|, and the claim follows. ∎

We are now ready to finish the proof of this lemma. We have already shown that η\eta is finite-to-one, so let Yn′′Y′′Y_{n}^{\prime\prime}\subseteq Y^{\prime\prime}, n1n\geq 1, be such that xη(x)x\mapsto\eta(x) is nn-to-one on Yn′′Y_{n}^{\prime\prime}. Let Rn=η(Yn′′)R_{n}=\eta(Y_{n}^{\prime\prime}), and recall that RnYR_{n}\subseteq Y. Sets RnR_{n} are pairwise disjoint. Let ϕk,n:RnYn′′\phi_{k,n}:R_{n}\to Y_{n}^{\prime\prime}, 1kn1\leq k\leq n, be Borel bijections that pick the kkth point in the preimage: Yn′′=i=1nϕk,n(Rn)Y_{n}^{\prime\prime}=\bigsqcup_{i=1}^{n}\phi_{k,n}(R_{n}). Note that maps ϕk,n:Rnϕk,n(Rn)\phi_{k,n}:R_{n}\to\phi_{k,n}(R_{n}) are measure-preserving, since they belong to the pseudo full group of TT, and k=1n|ρT𝒞(ϕk,n(x))|β(x)\sum_{k=1}^{n}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(\phi_{k,n}(x))|\leq\beta(x) for all xRnx\in R_{n} by Claim 3. One now has

Yn′′|ρT𝒞(x)|𝑑μ(x)\displaystyle\int_{Y_{n}^{\prime\prime}}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)|\,d\mu(x) =k=1nϕk,n(Rn)|ρT𝒞(x)|𝑑μ(x)\displaystyle=\sum_{k=1}^{n}\int_{\phi_{k,n}(R_{n})}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)|\,d\mu(x)
ϕn,k are measure-preserving\displaystyle\because\textrm{$\phi_{n,k}$ are measure-preserving} =Rnk=1n|ρT𝒞(ϕk,n1(x))|dμ(x)\displaystyle=\int_{R_{n}}\sum_{k=1}^{n}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(\phi_{k,n}^{-1}(x))|\,d\mu(x)
Claim 3\displaystyle\because\textrm{Claim 3} Rnβ(x)𝑑μ(x).\displaystyle\leq\int_{R_{n}}\beta(x)\,d\mu(x).

Summing these inequalities over nn we get

Y′′|ρT𝒞(x)|𝑑μ(x)\displaystyle\int_{Y^{\prime\prime}}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)|\,d\mu(x) =n=1Yn′′|ρT𝒞(x)|𝑑μ(x)\displaystyle=\sum_{n=1}^{\infty}\int_{Y_{n}^{\prime\prime}}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)|\,d\mu(x)
n=1Rnβ(x)𝑑μ(x)Yβ(x)𝑑μ(x),\displaystyle\leq\sum_{n=1}^{\infty}\int_{R_{n}}\beta(x)\,d\mu(x)\leq\int_{Y}\beta(x)\,d\mu(x),

where the last inequality is based on the fact that sets RnR_{n} are pairwise disjoint. This finishes the proof of the inequality (8.2) as well as the lemma. ∎

Several important facts follow easily from Lemma 8.3. For one, it implies that for any cross-section 𝒞\mathcal{C} the intermitted transformation T𝒞T_{\mathcal{R}_{\mathcal{C}}} belongs to [X]1[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}. In fact, we have the following inequality on the norms.

Corollary 8.4.

For any intermitted transformation T𝒞T_{\mathcal{R}_{\mathcal{C}}} one has T𝒞1T1\lVert T_{\mathcal{R}_{\mathcal{C}}}\rVert_{1}\leq\left\lVert T\right\rVert_{1}.

Proof.

By the definition of the set YY in Lemma 8.3, ρT𝒞(x)=ρT(x)\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)=\rho_{T}(x) for all xYx\not\in Y, hence

X|ρT𝒞|𝑑μ\displaystyle\int_{X}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}|\,d\mu =XY|ρT𝒞|𝑑μ+Y|ρT𝒞|𝑑μ\displaystyle=\int_{X\setminus Y}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}|\,d\mu+\int_{Y}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}|\,d\mu
Lemma 8.3\displaystyle\because\textrm{Lemma~{}\ref{lem:TEC-estimate}} XY|ρT|𝑑μ+Y|ρT|𝑑μ=X|ρT|𝑑μ,\displaystyle\leq\int_{X\setminus Y}|\rho_{T}|\,d\mu+\int_{Y}|\rho_{T}|\,d\mu=\int_{X}|\rho_{T}|\,d\mu,

which shows T𝒞1T1\lVert T_{\mathcal{R}_{\mathcal{C}}}\rVert_{1}\leq\lVert T\rVert_{1}. ∎

Remark 8.5.

As we discussed in Remark 8.2, the concept of an intermitted transformation applies wider than the case of one-dimensional flows. We mention, however, that the analog of Lemma 8.3 and Corollary 8.4 does not hold even for free measure-preserving 2\mathbb{R}^{2}-flows. Consider an annulus depicted in Figure 8.2(a) and let TT be the rotation by an angle α\alpha around the center of this annulus. Let the equivalence relation EE consist of two classes, each composing half of the ring. For a point xx such that ¬xETx\neg xETx, TExT_{E}x will be close to the other side of the class. It is easy to arrange the parameters (the angle α\alpha and the radii of the annulus) so that ρTE(x)>ρT(x)\lVert\rho_{T_{E}}(x)\rVert>\lVert\rho_{T}(x)\rVert for all xx such that TxTExTx\neq T_{E}x.

\vdotsα\alphaTTxxTExT_{E}x
(a)
(b)
Figure 8.2. Construction of a conservative transformation TT with TE1>T1\lVert T_{E}\rVert_{1}>\lVert T\rVert_{1}.

Every free measure-preserving flow 2X\mathbb{R}^{2}\curvearrowright X admits a tiling of its orbits by rectangles. The transformation T[2X]1T\in[\mathbb{R}^{2}\curvearrowright X\mkern 1.5mu]_{1} can be defined similarly to Figure 8.2(a) on each rectangle of the tiling by splitting each tile into two equivalence classes as in 8.2(b). The resulting transformation TT will have bounded orbits and satisfy TE1>T1\lVert T_{E}\rVert_{1}>\lVert T\rVert_{1} relative to the equivalence relation EE whose classes are the half tiles.

When the gaps in a cross-section 𝒞\mathcal{C} are large, xx and TxTx will often be 𝒞\mathcal{R}_{\mathcal{C}}-equivalent, and it therefore natural to expect that T𝒞T_{\mathcal{R}_{\mathcal{C}}} will be close to TT. This intuition is indeed valid, and the following approximation result is the most important consequence of Lemma 8.3.

Lemma 8.6.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a conservative transformation. For any ϵ>0\epsilon>0 there exists MM such for any cross-section 𝒞\mathcal{C} with gap𝒞(c)M\mathrm{gap}_{\mathcal{C}}(c)\geq M for all c𝒞c\in\mathcal{C} one has TT𝒞11<ϵ\lVert T\circ T^{-1}_{\mathcal{R}_{\mathcal{C}}}\rVert_{1}<\epsilon.

Proof.

Let AK={xX:|ρT(x)|K}A_{K}=\{x\in X:|\rho_{T}(x)|\geq K\}, K0K\in\mathbb{R}^{\geq 0}, be the set of points whose cocycle is at least KK in the absolute value. Since T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}, we may pick K1K\geq 1 is so large that AK|ρT|𝑑μ<ϵ/4\int_{A_{K}}|\rho_{T}|\,d\mu<\epsilon/4. Pick any real MM such that 2K2/M<ϵ/42K^{2}/M<\epsilon/4. We claim that it satisfies the conclusion of the lemma. To verify this we pick a cross-section 𝒞\mathcal{C} with all gaps having size at least MM. Set as before Y={xX:TxT𝒞x}Y=\{x\in X:Tx\neq T_{\mathcal{R}_{\mathcal{C}}}x\}. Since

TT𝒞11=YD(Tx,T𝒞x)𝑑μ(x),\bigl{\lVert}T\circ T^{-1}_{\mathcal{R}_{\mathcal{C}}}\bigr{\rVert}_{1}=\int_{Y}D(Tx,T_{\mathcal{R}_{\mathcal{C}}}x)\,d\mu(x),

our task is to estimate this integral. This can be done in a rather crude way. We can simply use the triangle inequality D(Tx,T𝒞x)|ρT(x)|+|ρT𝒞(x)|D(Tx,T_{\mathcal{R}_{\mathcal{C}}}x)\leq|\rho_{T}(x)|+|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)|, and deduce

YD(Tx,T𝒞x)𝑑μ(x)Y|ρT|𝑑μ+Y|ρT𝒞|𝑑μ2Y|ρT|𝑑μ,\int_{Y}D(Tx,T_{\mathcal{R}_{\mathcal{C}}}x)\,d\mu(x)\leq\int_{Y}|\rho_{T}|\,d\mu+\int_{Y}|\rho_{T_{\mathcal{R}_{\mathcal{C}}}}|\,d\mu\leq 2\int_{Y}|\rho_{T}|\,d\mu,

where the last inequality is based on Lemma 8.3.

It remains to show that Y|ρT|𝑑μ<ϵ/2\int_{Y}|\rho_{T}|\,d\mu<\epsilon/2. Let X~={c+[K,gap𝒞(c)K]:c𝒞}\widetilde{X}=\{c+[K,\mathrm{gap}_{\mathcal{C}}(c)-K]:c\in\mathcal{C}\} be the region that leaves out intervals of length KK on both sides of each point in 𝒞\mathcal{C}. Note that for any xX~AKx\in\widetilde{X}\setminus A_{K} one has x𝒞Txx\mathcal{R}_{\mathcal{C}}Tx and thus T𝒞x=TxT_{\mathcal{R}_{\mathcal{C}}}x=Tx for such points. Therefore, YAKBKY\subseteq A_{K}\sqcup B_{K}, where BK=X(X~AK)B_{K}=X\setminus(\widetilde{X}\cup A_{K}), and thus

Y|ρT|𝑑μAK|ρT|𝑑μ+BK|ρT|𝑑μ<ϵ/4+K2K/M<ϵ/2.\int\limits_{Y}|\rho_{T}|\,d\mu\leq\int\limits_{A_{K}}|\rho_{T}|\,d\mu+\int\limits_{B_{K}}|\rho_{T}|\,d\mu<\epsilon/4+K\cdot 2K/M<\epsilon/2.\qed
Lemma 8.7.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a conservative transformation. For any ϵ>0\epsilon>0 there exists a periodic transformation P[T]P\in[T\mkern 1.5mu] such that TP11<ϵ\lVert T\circ P^{-1}\rVert_{1}<\epsilon.

Proof.

By Lemma 8.6, we can find a cocompact cross-section 𝒞\mathcal{C} such that TT𝒞1<ϵ/2\lVert T\circ T^{-1}_{\mathcal{R}_{\mathcal{C}}}\rVert<\epsilon/2. Let M~\tilde{M} be an upper bound for gaps in 𝒞\mathcal{C}. Recall that the cocycle |ρT𝒞(x)||\rho_{T_{\mathcal{R}_{\mathcal{C}}}}(x)| is uniformly bounded by M~\tilde{M}, and, in fact, the same is true for any element in the full group of T𝒞T_{\mathcal{R}_{\mathcal{C}}}. In particular, we may use Rokhlin’s lemma to find a periodic P[T𝒞]P\in[T_{\mathcal{R}_{\mathcal{C}}}\mkern 1.5mu] such that T𝒞P1<ϵ/2M~\lVert T_{\mathcal{R}_{\mathcal{C}}}\circ P^{-1}\rVert<\epsilon/2\tilde{M}, and conclude that T𝒞P11<ϵ/2\lVert T_{\mathcal{R}_{\mathcal{C}}}\circ P^{-1}\rVert_{1}<\epsilon/2. We therefore have

TP11TT𝒞11+T𝒞P11<ϵ.\lVert T\circ P^{-1}\rVert_{1}\leq\lVert T\circ T^{-1}_{\mathcal{R}_{\mathcal{C}}}\rVert_{1}+\lVert T_{\mathcal{R}_{\mathcal{C}}}\circ P^{-1}\rVert_{1}<\epsilon.\qed
Corollary 8.8.

If T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} is conservative then TT belongs to the derived full group D([X]1)D([\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}), in particular its index satisfies (T)=0\mathcal{I}(T)=0.

Proof.

Follows directly from Lemma 8.7 and Corollary 3.15. ∎

Chapter 9 Dissipative and monotone transformations

The previous chapter studied conservative transformations, whereas this one concentrates on dissipative ones. Our goal will be to show that any dissipative T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} of index (T)=0\mathcal{I}(T)=0 belongs to the derived subgroup D([X]1)D([\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}). We begin however by describing some general aspects of dynamics of dissipative automorphisms.

Recall that according to Proposition 4.16, any transformation T[X]T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu] induces a TT-invariant partition of the phase space X=DCX=D\sqcup C such that T|CT|_{C} is conservative and T|DT|_{D} is dissipative. Formally speaking, a transformation is said to be dissipative if the partition trivializes to D=XD=X. For the purpose of this chapter it is however convenient to widen this notion just a bit by allowing TT to have fixed points.

Definition 9.1.

A transformation T[X]T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu] is said to be dissipative if D=suppTD=\operatorname*{supp}T, where DD is the dissipative element of the Hopf’s decomposition for TT.

9.1. Orbit limits and monotone transformations

We begin by showing that dynamics of dissipative transformations in L1\mathrm{L}^{1} full groups of \mathbb{R}-flows is similar to those in L1\mathrm{L}^{1} full groups of \mathbb{Z} actions. We do so by establishing an analog of R. M. Belinskaja’s result [Bel68, Thm. 3.2]. Recall that a sequence of reals is said to have an almost constant sign if all but finitely many elements of the sequence have the same sign.

Proposition 9.2.

Let SS be a measure-preserving transformation of the real line which commensurates the set \mathbb{R}^{-}, suppose that SS is dissipative. Then for almost all xx\in\mathbb{R}, the sequence of reals (Sk(x)x)k(S^{k}(x)-x)_{k\in\mathbb{N}} has an almost constant sign.

Proof.

Let QQ be the set of reals xx such that (Sk(x)x)k(S^{k}(x)-x)_{k\in\mathbb{N}} does not have an almost constant sign, and suppose by contradiction that QQ has positive measure. Since SS is dissipative, we can find a Borel wandering set AA\subseteq\mathbb{R} for SS which non-trivially intersects QQ. All the translates of Q=QAQ^{\prime}=Q\cap A are disjoint, and, for all xQx\in Q^{\prime}, (Sk(x)x)k(S^{k}(x)-x)_{k\in\mathbb{N}} does not have an almost constant sign.

Since SS is dissipative, for almost all xQx\in Q^{\prime}, the sequence of absolute values (|Sk(x)|)k(\left\lvert S^{k}(x)\right\rvert)_{k\in\mathbb{N}} tends to ++\infty (see Proposition B.4). In particular, there are infinitely many points yy in the SS-orbit of xx such that y<0y<0 but S(y)>0S(y)>0. Since the map Q×Q^{\prime}\times\mathbb{Z}\to\mathbb{R} which maps (x,k)(x,k) to Sk(x)S^{k}(x) is measure-preserving, this yields that the set of y<0y<0 such that S(y)>0S(y)>0 has infinite measure, contradicting the fact that SS commensurates the set \mathbb{R}^{-}. ∎

Corollary 9.3.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a dissipative transformation. For almost all xsuppTx\in\operatorname*{supp}T, the sequence (ρ(x,Tk(x)))k(\rho(x,T^{k}(x)))_{k\in\mathbb{N}}, has an almost constant sign.

Proof.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}. For all xXx\in X, denote by TxT_{x} the measure-preserving transformation of \mathbb{R} induced by TT on the \mathbb{R}-orbit of xx. By the proof of Proposition 6.8, the integral

Xλ(0(Tx(0)))𝑑μ(x)\int_{X}\lambda(\mathbb{R}^{\geq 0}\bigtriangleup(T_{x}(\mathbb{R}^{\geq 0})))d\mu(x)

is finite. In particular, for almost every xXx\in X, the transformation TxT_{x} commensurates the set 0\mathbb{R}^{\geq 0}. The conclusion now follows directly from the previous proposition. ∎

For any dissipative transformation in an L1\mathrm{L}^{1} full group of a free locally compact Polish group action and for almost every xXx\in X, ρ(x,Tnx)\rho(x,T^{n}x)\to\infty as nn\to\infty, in the sense that ρ(x,Tnx)\rho(x,T^{n}x) eventually escapes any compact subset of the acting group. In the context of flows, Corollary 9.3 strengthens this statement and implies that ρ(x,Tnx)\rho(x,T^{n}x) must converge to either ++\infty or -\infty.

Corollary 9.4.

If T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} is dissipative, then for almost every point xsuppTx\in\operatorname*{supp}T either limnρ(x,Tnx)=+\lim\limits_{n\to\infty}\rho(x,T^{n}x)=+\infty or limnρ(x,Tnx)=\lim\limits_{n\to\infty}\rho(x,T^{n}x)=-\infty. ∎

In view of this corollary, there is a canonical TT-invariant decomposition of suppT\operatorname*{supp}T into “positive” and “negative” orbits.

Definition 9.5.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a dissipative automorphism. Its support is partitioned into XX\vec{X}\sqcup\reflectbox{$\vec{\reflectbox{$X$}}$}, where

X\displaystyle\vec{X} ={xsuppT:limnρ(x,Tnx)=+},\displaystyle=\bigl{\{}x\in\operatorname*{supp}T:\lim\limits_{n\to\infty}\rho(x,T^{n}x)=+\infty\bigr{\}},
X\displaystyle\reflectbox{$\vec{\reflectbox{$X$}}$} ={xsuppT:limnρ(x,Tnx)=}.\displaystyle=\bigl{\{}x\in\operatorname*{supp}T:\lim\limits_{n\to\infty}\rho(x,T^{n}x)=-\infty\bigr{\}}.

The set X\vec{X} is said to be positive evasive and X\vec{\reflectbox{$X$}} is negative evasive.

According to Corollary 9.3, for almost every xsuppTx\in\operatorname*{supp}T, eventually either all TnxT^{n}x are to the right of xx or all are to the left of it. There are points xx for which the adverb “eventually” can, in fact, be dropped.

Corollary 9.6.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a dissipative transformation and let

A\displaystyle\vec{A} ={xX:ρ(x,Tnx)>0 for all n1},\displaystyle=\{x\in\vec{X}:\rho(x,T^{n}x)>0\textrm{ for all }n\geq 1\},
A\displaystyle\reflectbox{$\vec{\reflectbox{$A$}}$} ={xX:ρ(x,Tnx)<0 for all n1}.\displaystyle=\{x\in\reflectbox{$\vec{\reflectbox{$X$}}$}:\rho(x,T^{n}x)<0\textrm{ for all }n\geq 1\}.

The set A=AAA=\vec{A}\sqcup\reflectbox{$\vec{\reflectbox{$A$}}$} is a complete section for T|suppTT|_{\operatorname*{supp}T}.

Proof.

We need to show that almost every orbit of TT intersects AA. Let xsuppTx\in\operatorname*{supp}T and suppose for definiteness that xXx\in\vec{X}. Since limnρ(x,Tnx)=+\lim_{n\to\infty}\rho(x,T^{n}x)=+\infty, there is n0=max{n:ρ(x,Tnx)0}n_{0}=\max\{n\in\mathbb{N}:\rho(x,T^{n}x)\leq 0\}, and therefore Tn0xAT^{n_{0}}x\in\vec{A}. ∎

Definition 9.7.

A dissipative transformation T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} is monotone if ρ(x,Tx)>0\rho(x,Tx)>0 for almost all xXx\in\vec{X}, and ρ(x,Tx)<0\rho(x,Tx)<0 for almost all xXx\in\reflectbox{$\vec{\reflectbox{$X$}}$}.

Corollary 9.8.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a dissipative transformation. There is a complete section AsuppTA\subseteq\operatorname*{supp}T and a periodic transformation P[X]1[T]P\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}\cap[T\mkern 1.5mu] such that T=PTAT=P\circ T_{A} and TAT_{A} is monotone.

Proof.

Take AA to be as in Corollary 9.6 and note that P=TTA1P=T\circ T_{A}^{-1} is periodic and satisfies the conclusions of the corollary. ∎

As we discussed at the beginning of the chapter, our goal is to show that the index of the kernel map coincides with the derived subgroup of [X]1[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}. Note that if T=PTAT=P\circ T_{A} is as above, then (T)=(TA)\mathcal{I}(T)=\mathcal{I}(T_{A}), and, coupled with the results of Chapter 8, it will suffice to show that all monotone transformations of index zero belong to D([X]1)D([\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}). This will be the focus of the rest of this chapter and will take some effort to achieve, but the main strategy is to show that such automorphisms can be approximated by periodic maps, which is the content of Theorem 9.15 below.

9.2. Arrival and departure sets

Throughout the rest of this chapter, we fix a cross-section 𝒞X\mathcal{C}\subset X and a monotone transformation T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}. The arrival set A𝒞A_{\mathcal{C}} is the set of the first visitors to E𝒞E_{\mathcal{C}} classes: A𝒞={xsuppT:¬xE𝒞T1x}A_{\mathcal{C}}=\{x\in\operatorname*{supp}T:\neg xE_{\mathcal{C}}T^{-1}x\}. Analogously, the departure set D𝒞D_{\mathcal{C}} is defined to be D𝒞={xsuppT:¬xE𝒞Tx}D_{\mathcal{C}}=\{x\in\operatorname*{supp}T:\neg xE_{\mathcal{C}}Tx\}. We also let A𝒞\vec{A}_{\mathcal{C}} denote A𝒞XA_{\mathcal{C}}\cap\vec{X} and A𝒞=A𝒞X\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}=A_{\mathcal{C}}\cap\reflectbox{$\vec{\reflectbox{$X$}}$}; likewise for D𝒞\vec{D}_{\mathcal{C}} and D𝒞\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}. Note that T(D𝒞)=A𝒞T(D_{\mathcal{C}})=A_{\mathcal{C}}, and thus T1(A𝒞)=D𝒞T^{-1}(A_{\mathcal{C}})=D_{\mathcal{C}}. There is, however, another useful map from A𝒞A_{\mathcal{C}} onto D𝒞D_{\mathcal{C}}.

cccc^{\prime}xA𝒞x\in\vec{A}_{\mathcal{C}}T4xD𝒞T^{4}x\in\vec{D}_{\mathcal{C}}t𝒞(x)=4t_{\mathcal{C}}(x)=4
Figure 9.1. Arrival and Departure sets.

We define the transfer value t𝒞:A𝒞t_{\mathcal{C}}:A_{\mathcal{C}}\to\mathbb{N} by the condition

t𝒞(x)=min{n0:TnxD𝒞}t_{\mathcal{C}}(x)=\min\{n\geq 0:T^{n}x\in D_{\mathcal{C}}\}

and the transfer function τ𝒞:A𝒞D𝒞\tau_{\mathcal{C}}:A_{\mathcal{C}}\to D_{\mathcal{C}} is defined to be τ𝒞(x)=Tt𝒞(x)x\tau_{\mathcal{C}}(x)=T^{t_{\mathcal{C}}(x)}x. Note that τ𝒞\tau_{\mathcal{C}} is measure-preserving. The transfer value introduces a partition of the arrival set A𝒞=nA𝒞nA_{\mathcal{C}}=\bigsqcup_{n\in\mathbb{N}}A_{\mathcal{C}}^{n}, where A𝒞n=t𝒞1(n)A_{\mathcal{C}}^{n}=t_{\mathcal{C}}^{-1}(n); by applying the transfer function, it also produces a partition for the departure set: D𝒞=nD𝒞nD_{\mathcal{C}}=\bigsqcup_{n\in\mathbb{N}}D_{\mathcal{C}}^{n}, where D𝒞n=τ𝒞(A𝒞n)D_{\mathcal{C}}^{n}=\tau_{\mathcal{C}}(A_{\mathcal{C}}^{n}).

In plain words, t𝒞(x)+1t_{\mathcal{C}}(x)+1 is the number of points in [x]ET[x]E𝒞[x]_{E_{T}}\cap[x]_{E_{\mathcal{C}}}. Therefore if λc𝒞(A𝒞n)λc𝒞(A𝒞m)\lambda_{c}^{\mathcal{C}}(A_{\mathcal{C}}^{n})\geq\lambda_{c}^{\mathcal{C}}(A_{\mathcal{C}}^{m}) for some nmn\geq m then also λc𝒞([A𝒞n]ET)λc𝒞([A𝒞m]ET)\lambda_{c}^{\mathcal{C}}([A_{\mathcal{C}}^{n}]_{E_{T}})\geq\lambda_{c}^{\mathcal{C}}([A_{\mathcal{C}}^{m}]_{E_{T}}). In Sections 9.3 and 9.4 we modify the transformation TT on the arrival and departure sets and we want to do this in a way that affects as many orbits as possible as measured by λc𝒞\lambda_{c}^{\mathcal{C}}. This amounts to using sets A𝒞nA_{\mathcal{C}}^{n} (and D𝒞nD_{\mathcal{C}}^{n}) with as high values of nn as possible. The next lemma will be helpful in conducting such a selection in a measurable way across all of c𝒞c\in\mathcal{C}.

Lemma 9.9.

Let AXA\subseteq X be a measurable set with a measurable partition A=nAnA=\bigsqcup_{n}A_{n} and let ξ:𝒞0\xi:\mathcal{C}\to\mathbb{R}^{\geq 0} be a measurable function such that ξ(c)λc𝒞(A)\xi(c)\leq\lambda_{c}^{\mathcal{C}}(A) for all c𝒞c\in\mathcal{C}. There are measurable ν:𝒞\nu:\mathcal{C}\to\mathbb{N} and r:𝒞0r:\mathcal{C}\to\mathbb{R}^{\geq 0} such that for any c𝒞c\in\mathcal{C} for which ξ(c)>0\xi(c)>0 one has

λc𝒞((n>ν(c)An)(Aν(c)(c+[0,r(c)])))=ξ(c).\lambda_{c}^{\mathcal{C}}\bigl{(}\bigl{(}\mkern-6.0mu\bigsqcup_{n>\nu(c)}\mkern-6.0muA_{n}\bigr{)}\cup\bigl{(}A_{\nu(c)}\cap(c+[0,r(c)])\bigr{)}\bigr{)}=\xi(c).
Proof.

For c𝒞c\in\mathcal{C} such that ξ(c)>0\xi(c)>0 set

ν(c)=min{n:λc𝒞(k>nAk)<ξ(c)}.\nu(c)=\min\bigl{\{}n\in\mathbb{N}:\lambda_{c}^{\mathcal{C}}\bigl{(}\bigsqcup_{k>n}A_{k}\bigr{)}<\xi(c)\bigr{\}}.

Note that one necessarily has λc𝒞(Aν(c))ξ(c)λc𝒞(n>ν(c)An)\lambda_{c}^{\mathcal{C}}\bigl{(}A_{\nu(c)}\bigr{)}\geq\xi(c)-\lambda_{c}^{\mathcal{C}}\bigl{(}\bigsqcup_{n>\nu(c)}A_{n}\bigr{)}. Set

r(c)=min{a0:λc𝒞(Aν(c)(c+[0,a]))=ξ(c)λc𝒞(n>ν(c)An)}.r(c)=\min\bigl{\{}a\geq 0:\lambda_{c}^{\mathcal{C}}\bigl{(}A_{\nu(c)}\cap(c+[0,a])\bigr{)}=\xi(c)-\lambda_{c}^{\mathcal{C}}\bigl{(}\mkern-6.0mu\bigsqcup_{n>\nu(c)}\mkern-8.0muA_{n}\bigr{)}\bigr{\}}.

These functions ν\nu and rr satisfy the conclusions of the lemma. ∎

Definition 9.10.

Consider the partition of the positive arrival set A𝒞=nA𝒞n\vec{A}_{\mathcal{C}}=\bigsqcup_{n}\vec{A}_{\mathcal{C}}^{n} and let ξ:𝒞0\xi:\mathcal{C}\to\mathbb{R}^{\geq 0}, r:𝒞0r:\mathcal{C}\to\mathbb{R}^{\geq 0}, and ν:𝒞\nu:\mathcal{C}\to\mathbb{N} be as in Lemma 9.9. The set A𝒞\vec{A}^{\bullet}_{\mathcal{C}} defined by the condition

A𝒞(c)=n>ν(c)A𝒞n(c)(A𝒞ν(c)(c+[0,r(c)]))for all c𝒞\vec{A}^{\bullet}_{\mathcal{C}}(c)=\mkern-6.0mu\bigsqcup_{n>\vec{\nu}(c)}\mkern-6.0mu\vec{A}_{\mathcal{C}}^{n}(c)\cup\bigl{(}A_{\mathcal{C}}^{\vec{\nu}(c)}\cap(c+[0,\vec{r}(c)])\bigr{)}\quad\textrm{for all }c\in\mathcal{C}

is said to be the positive ξ\xi-copious arrival set. The positive ξ\xi-copious departure set is given by D𝒞=τ𝒞(A𝒞)\vec{D}_{\mathcal{C}}^{\bullet}=\tau_{\mathcal{C}}(\vec{A}_{\mathcal{C}}^{\bullet}). The definitions of the negative ξ\xi-copious arrival and departure sets use the partition A𝒞=nA𝒞n\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}=\bigsqcup_{n}\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}^{n} of the negative arrival set and are analogous.

Copious sets maximize measure λc𝒞\lambda_{c}^{\mathcal{C}} of their saturation under the action of TT. In other words, among all subsets AA𝒞A^{\prime}\subseteq\vec{A}_{\mathcal{C}} for which λc𝒞(A)=ξ(c)\lambda_{c}^{\mathcal{C}}(A^{\prime})=\xi(c), the measure λc𝒞([A]ET)\lambda_{c}^{\mathcal{C}}([A^{\prime}]_{E_{T}}) is maximal when A(c)=A𝒞(c)A^{\prime}(c)=\vec{A}_{\mathcal{C}}^{\bullet}(c). In particular, if λc𝒞(A𝒞)\lambda_{c}^{\mathcal{C}}(\vec{A}_{\mathcal{C}}^{\bullet}) is close to λc𝒞(A𝒞)\lambda_{c}^{\mathcal{C}}(\vec{A}_{\mathcal{C}}), then we expect λc𝒞([A𝒞]ET)\lambda_{c}^{\mathcal{C}}([\vec{A}_{\mathcal{C}}^{\bullet}]_{E_{T}}) to be close to λc𝒞([A𝒞]ET)\lambda_{c}^{\mathcal{C}}([\vec{A}_{\mathcal{C}}]_{E_{T}}). The following lemma quantifies this intuition.

Lemma 9.11.

Let ξ:𝒞0\xi:\mathcal{C}\to\mathbb{R}^{\geq 0} be such that ξ(c)λc𝒞(A𝒞)\xi(c)\leq\lambda_{c}^{\mathcal{C}}(\vec{A}_{\mathcal{C}}) for all c𝒞c\in\mathcal{C}, and let A𝒞\vec{A}_{\mathcal{C}}^{\bullet} be the ξ\xi-copious arrival set constructed in Lemma 9.9. If there exists 1/2>δ>01/2>\delta>0 such that ξ(c)(1δ)λc𝒞(A𝒞)\xi(c)\geq(1-\delta)\lambda_{c}^{\mathcal{C}}(\vec{A}_{\mathcal{C}}) for all c𝒞c\in\mathcal{C}, then

λc𝒞([A𝒞(c)A𝒞(c)]ET)δ1δλc𝒞(X)for all c𝒞,\lambda_{c}^{\mathcal{C}}([\vec{A}_{\mathcal{C}}(c)\setminus\vec{A}^{\bullet}_{\mathcal{C}}(c)]_{E_{T}})\leq\frac{\delta}{1-\delta}\lambda_{c}^{\mathcal{C}}(\vec{X})\quad\textrm{for all }c\in\mathcal{C},

and therefore also μ([A𝒞A𝒞]ET)δ1δμ(X)\mu([\vec{A}_{\mathcal{C}}\setminus\vec{A}^{\bullet}_{\mathcal{C}}]_{E_{T}})\leq\frac{\delta}{1-\delta}\mu(\vec{X}).

An analogous statement is valid for the negative arrival set A𝒞\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}.

Proof.

Let ν\nu be as in Lemma 9.9 and note that

k>ν(c)A𝒞k(c)A𝒞(c)kν(c)A𝒞k(c)\bigsqcup_{k>\nu(c)}\vec{A}_{\mathcal{C}}^{k}(c)\subseteq\vec{A}_{\mathcal{C}}^{\bullet}(c)\subseteq\bigsqcup_{k\geq\nu(c)}\vec{A}_{\mathcal{C}}^{k}(c)

whenever c𝒞c\in\mathcal{C} satisfies ξ(c)>0\xi(c)>0. Recall that for xA𝒞nx\in\vec{A}_{\mathcal{C}}^{n} we have xE𝒞TkxE_{\mathcal{C}}T^{k} for all 0kn0\leq k\leq n and sets Tk(A𝒞n)T^{k}(\vec{A}_{\mathcal{C}}^{n}) are pairwise disjoint. In particular,

(9.1) λc𝒞(X)λc𝒞([kν(c)A𝒞k(c)]ET)(ν(c)+1)λc𝒞(kν(c)A𝒞k(c))(ν(c)+1)λc𝒞(A𝒞)=(ν(c)+1)ξ(c).\begin{split}\lambda_{c}^{\mathcal{C}}(\vec{X})&\geq\lambda_{c}^{\mathcal{C}}\bigl{(}\bigl{[}\mkern-6.0mu\bigsqcup_{k\geq\nu(c)}\mkern-6.0mu\vec{A}_{\mathcal{C}}^{k}(c)\bigr{]}_{E_{T}}\bigr{)}\geq(\nu(c)+1)\lambda_{c}^{\mathcal{C}}\bigl{(}\mkern-6.0mu\bigsqcup_{k\geq\nu(c)}\mkern-6.0mu\vec{A}_{\mathcal{C}}^{k}(c)\bigr{)}\\ &\geq(\nu(c)+1)\lambda_{c}^{\mathcal{C}}(\vec{A}_{\mathcal{C}}^{\bullet})=(\nu(c)+1)\xi(c).\end{split}

Note also that ξ(c)(1δ)λc𝒞(A𝒞)\xi(c)\geq(1-\delta)\lambda_{c}^{\mathcal{C}}(\vec{A}_{\mathcal{C}}) implies

(9.2) λc𝒞(A𝒞A𝒞)ξ(c)δ/(1δ).\lambda_{c}^{\mathcal{C}}(\vec{A}_{\mathcal{C}}\setminus\vec{A}^{\bullet}_{\mathcal{C}})\leq\xi(c)\delta/(1-\delta).

For any c𝒞c\in\mathcal{C} we have

λc𝒞([A𝒞(c)A𝒞(c)]ET)\displaystyle\lambda_{c}^{\mathcal{C}}([\vec{A}_{\mathcal{C}}(c)\setminus\vec{A}^{\bullet}_{\mathcal{C}}(c)]_{E_{T}}) λc𝒞({Tkx:xA𝒞(c)A𝒞(c),0kν(c)})\displaystyle\leq\lambda_{c}^{\mathcal{C}}(\{T^{k}x:x\in\vec{A}_{\mathcal{C}}(c)\setminus\vec{A}_{\mathcal{C}}^{\bullet}(c),0\leq k\leq\vec{\nu}(c)\})
(ν(c)+1)λc𝒞(A𝒞A𝒞)\displaystyle\leq(\vec{\nu}(c)+1)\lambda_{c}^{\mathcal{C}}(\vec{A}_{\mathcal{C}}\setminus\vec{A}_{\mathcal{C}}^{\bullet})
(9.2)\displaystyle\because~{}\eqref{eq:complement-copiuous-estimate} (ν(c)+1)ξ(c)δ/(1δ)\displaystyle\leq(\vec{\nu}(c)+1)\vec{\xi}(c)\delta/(1-\delta)
(9.1)\displaystyle\because~{}\eqref{eq:xi-nu-bound} λc𝒞(X)δ/(1δ).\displaystyle\leq\lambda_{c}^{\mathcal{C}}(\vec{X})\delta/(1-\delta).

The inequality for the measure μ\mu follows by disintegrating μ\mu into 𝒞λc𝒞()\int_{\mathcal{C}}\lambda_{c}^{\mathcal{C}}(\,\cdot\,).

The argument for the negative arrival set is completely analogous. ∎

9.3. Coherent modifications

We remind the reader that our goal is to show that any dissipative transformation T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} of index (T)=0\mathcal{I}(T)=0 can be approximated by periodic transformations. One approach to “loop” the orbits of TT is by mapping D𝒞(c)\vec{D}_{\mathcal{C}}(c) to A𝒞(c)\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}(c) and D𝒞(c)\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}(c) to A𝒞(c)\vec{A}_{\mathcal{C}}(c) (cf. Figure 9.6). For such a modification to work, measures λc𝒞(D𝒞(c))\lambda_{c}^{\mathcal{C}}(\vec{D}_{\mathcal{C}}(c)) and λc𝒞(A𝒞(c))\lambda_{c}^{\mathcal{C}}(\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}(c)) have to be equal. Recall that (T)=0\mathcal{I}(T)=0 implies that for almost every c𝒞c\in\mathcal{C}, the measure of points xx such that xc<Txx\leq c<Tx equals the measure of those yy for which Ty<cyTy<c\leq y. If one could guarantee that T(D𝒞(c))=A𝒞(σ𝒞(c))T(\vec{D}_{\mathcal{C}}(c))=\vec{A}_{\mathcal{C}}(\sigma_{\mathcal{C}}(c)), then the aforementioned modification would indeed work. In the case of \mathbb{Z} actions, discreteness of the acting group allows one to find a cross-section 𝒞\mathcal{C} for which this condition does hold. Whereas for the flows, we have to deal with the possibility that T(D𝒞(c))T(\vec{D}_{\mathcal{C}}(c)) can be “scattered” (see Figure 9.4) along the orbit and be unbounded, which is the key reason for the increased complexity compared to the argument for \mathbb{Z} actions.

Since we can’t hope to “loop” all the orbits of TT, we will do the next best thing, and apply the modification of Figure 9.6 on “most” orbits as measured by λc𝒞\lambda_{c}^{\mathcal{C}}. Copious sets discussed in Section 9.2 have large saturations under TT, but, generally speaking, fail to satisfy T(D𝒞(c))=A𝒞(σ𝒞(c))T(\vec{D}_{\mathcal{C}}^{\bullet}(c))=\vec{A}_{\mathcal{C}}^{\bullet}(\sigma_{\mathcal{C}}(c)) for the same reason as do the sets D𝒞(c)\vec{D}_{\mathcal{C}}(c). Our plan is to use the “ϵ\epsilon of room” provided by the difference D𝒞(c)D𝒞(c)\vec{D}_{\mathcal{C}}(c)\setminus\vec{D}_{\mathcal{C}}^{\bullet}(c) in order to modify TT into some TT^{\prime} with the same arrival and departure sets as TT, but for which also T(D𝒞(c))=A𝒞(σ𝒞(c))T^{\prime}(\vec{D}_{\mathcal{C}}^{\bullet}(c))=\vec{A}_{\mathcal{C}}^{\bullet}(\sigma_{\mathcal{C}}(c)) holds. In this section, we describe two abstract modifications of dissipative transformations, and the approximation strategy outlined above will later be implemented in Section 9.4.

Since we are about to consider arrival and departure sets of different transformations, we use the notation A𝒞[U]\vec{A}_{\mathcal{C}}[U] to denote the positive arrival set constructed for a transformation UU; likewise for negative arrival and departure sets, etc.

Lemma 9.12.

Let ϕ\phi and ϕ\phi^{\prime} be measure-preserving transformations on XX subject to the following conditions:

  1. (1)

    supp(ϕ)D𝒞\operatorname*{supp}(\phi)\subseteq D_{\mathcal{C}}, supp(ϕ)A𝒞\operatorname*{supp}(\phi^{\prime})\subseteq A_{\mathcal{C}};

  2. (2)

    ϕ(D𝒞)=D𝒞\phi(\vec{D}_{\mathcal{C}})=\vec{D}_{\mathcal{C}}, ϕ(D𝒞)=D𝒞\phi(\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}})=\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}, and ϕ(A𝒞)=A𝒞\phi^{\prime}(\vec{A}_{\mathcal{C}})=\vec{A}_{\mathcal{C}}, ϕ(A𝒞)=A𝒞\phi^{\prime}(\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}})=\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}};

  3. (3)

    xE𝒞ϕ(x)xE_{\mathcal{C}}\phi(x) and xE𝒞ϕ(x)xE_{\mathcal{C}}\phi^{\prime}(x) for all xsuppTx\in\operatorname*{supp}T.

The transformation Ux=ϕTϕ(x)Ux=\phi^{\prime}T\phi(x) is monotone, Ux=TxUx=Tx for all xD𝒞x\not\in D_{\mathcal{C}}, and the sets D𝒞D_{\mathcal{C}}, A𝒞A_{\mathcal{C}} remain the same:

X[U]\displaystyle\vec{X}[U] =X\displaystyle=\vec{X} X[U]\displaystyle\reflectbox{$\vec{\reflectbox{$X$}}$}[U] =X,\displaystyle=\reflectbox{$\vec{\reflectbox{$X$}}$},
D𝒞[U]\displaystyle\vec{D}_{\mathcal{C}}[U] =D𝒞\displaystyle=\vec{D}_{\mathcal{C}} D𝒞[U]\displaystyle\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}[U] =D𝒞,\displaystyle=\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}},
A𝒞[U]\displaystyle\vec{A}_{\mathcal{C}}[U] =A𝒞\displaystyle=\vec{A}_{\mathcal{C}} A𝒞[U]\displaystyle\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}[U] =A𝒞.\displaystyle=\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}.

Moreover, the integral of lengths of “departing arcs” remains unchanged:

D𝒞|ρU|𝑑μ=D𝒞|ρT|𝑑μ,\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{U}|\,d\mu=\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{T}|\,d\mu,\\

and the following estimate on XD(Tx,Ux)𝑑μ(x)\int_{X}D(Tx,Ux)\,d\mu(x) is available:

XD(Tx,Ux)𝑑μ(x)2D𝒞|ρT(x)|𝑑μ(x).\int_{X}D(Tx,Ux)\,d\mu(x)\leq 2\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{T}(x)|\,d\mu(x).
X\vec{X}

X\vec{\reflectbox{$X$}}

σ𝒞(c)\sigma_{\mathcal{C}}(c)ϕ\phiϕ\phi^{\prime}ϕ\phi^{\prime}ϕ\phiD𝒞(c)\vec{D}_{\mathcal{C}}(c)A𝒞(c)\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}(c)A𝒞(σ𝒞(c))\vec{A}_{\mathcal{C}}(\sigma_{\mathcal{C}}(c))D𝒞(σ𝒞(c))\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}(\sigma_{\mathcal{C}}(c))
Figure 9.2. The transformation U=ϕTϕU=\phi^{\prime}T\phi defined in Lemma 9.12.
Proof.

Figure 9.2 illustrates the definition of the transformation UU. Equality of the arrival and departure sets is straightforward to verify. Note that ϕ(D𝒞(c))=D𝒞(c)\phi(\vec{D}_{\mathcal{C}}(c))=\vec{D}_{\mathcal{C}}(c) for all c𝒞c\in\mathcal{C}, and therefore D𝒞ρϕ𝑑μ=0\int_{\vec{D}_{\mathcal{C}}}\rho_{\phi}\,d\mu=0. In fact, the following four integrals vanish:

(9.3) D𝒞ρϕ𝑑μ=D𝒞ρϕ𝑑μ=A𝒞ρϕ𝑑μ=A𝒞ρϕ𝑑μ=0.\int_{\vec{D}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi}\,d\mu=\int_{\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi}\,d\mu=\int_{\vec{A}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi^{\prime}}\,d\mu=\int_{\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi^{\prime}}\,d\mu=0.

Observe that ρU\rho_{U} is positive on D𝒞\vec{D}_{\mathcal{C}} and negative on D𝒞\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}, thus

D𝒞|ρU|𝑑μ=\displaystyle\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{U}|\,d\mu= D𝒞ρϕTϕ𝑑μD𝒞ρϕTϕ𝑑μ\displaystyle\int_{\vec{D}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi^{\prime}T\phi}\,d\mu-\int_{\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi^{\prime}T\phi}\,d\mu
cocycle identity=\displaystyle\because\textrm{cocycle identity}= D𝒞ρϕ𝑑μ(x)+D𝒞ρT(ϕ(x))𝑑μ(x)+D𝒞ρϕ(Tϕ(x))𝑑μ(x)\displaystyle\int_{\vec{D}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi}\,d\mu(x)+\int_{\vec{D}_{\mathcal{C}}}\mkern-10.0mu\rho_{T}(\phi(x))\,d\mu(x)+\int_{\vec{D}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi^{\prime}}(T\phi(x))\,d\mu(x)
D𝒞ρϕ𝑑μD𝒞ρT(ϕ(x))𝑑μ(x)D𝒞ρϕ(Tϕ(x))𝑑μ(x)\displaystyle-\int_{\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi}\,d\mu-\int_{\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{T}(\phi(x))\,d\mu(x)-\int_{\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi^{\prime}}(T\phi(x))\,d\mu(x)
=\displaystyle= D𝒞ρϕ𝑑μ+D𝒞ρT𝑑μ+A𝒞ρϕ𝑑μ\displaystyle\int_{\vec{D}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi}\,d\mu+\int_{\vec{D}_{\mathcal{C}}}\mkern-10.0mu\rho_{T}\,d\mu+\int_{\vec{A}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi^{\prime}}\,d\mu
D𝒞ρϕ𝑑μD𝒞ρT𝑑μA𝒞ρϕ𝑑μ\displaystyle-\int_{\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi}\,d\mu-\int_{\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{T}\,d\mu-\int_{\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{\phi^{\prime}}\,d\mu
Eq. (9.3)=\displaystyle\because~{}\textrm{Eq.~{}\eqref{eq:vanishing-cocycles}}= D𝒞ρT𝑑μD𝒞ρT𝑑μ=D𝒞|ρT|𝑑μ.\displaystyle\int_{\vec{D}_{\mathcal{C}}}\mkern-10.0mu\rho_{T}\,d\mu-\int_{\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}}\mkern-10.0mu\rho_{T}\,d\mu=\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{T}|\,d\mu.

Finally, note that for any xD𝒞x\in D_{\mathcal{C}}, the arc from xx to TxTx intersects the arc from T1ϕTϕ(x)T^{-1}\phi^{\prime}T\phi(x) to ϕTϕ(x)\phi^{\prime}T\phi(x) (both arcs go over the same point of 𝒞\mathcal{C}), and therefore

D(Tx,Ux)|ρT(x)|+|ρT(T1ϕTϕ(x))|.\displaystyle D(Tx,Ux)\leq|\rho_{T}(x)|+|\rho_{T}(T^{-1}\phi^{\prime}T\phi(x))|.

Integration over D𝒞D_{\mathcal{C}} yields

XD(Tx,Ux)𝑑μ(x)=D𝒞D(Tx,Ux)𝑑μ(x)2D𝒞|ρT(x)|𝑑μ(x).\int_{X}D(Tx,Ux)\,d\mu(x)=\int_{D_{\mathcal{C}}}D(Tx,Ux)\,d\mu(x)\leq 2\int_{D_{\mathcal{C}}}|\rho_{T}(x)|\,d\mu(x).\qed
Lemma 9.13.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a monotone transformation, let FD𝒞F\subseteq D_{\mathcal{C}} be such that λc𝒞(F)=λc𝒞(F)\lambda_{c}^{\mathcal{C}}(\vec{F})=\lambda_{c}^{\mathcal{C}}(\reflectbox{$\vec{\reflectbox{$F$}}$}) for all c𝒞c\in\mathcal{C} and the function 𝒞cλc𝒞(F)\mathcal{C}\ni c\mapsto\lambda_{c}^{\mathcal{C}}(F) is EE-invariant (i.e., λc𝒞(F)=λc(F)\lambda_{c}^{\mathcal{C}}(F)=\lambda_{c^{\prime}}(F) whenever cc and cc^{\prime} belong to the same orbit of the flow). Let ZA𝒞Z\subseteq A_{\mathcal{C}} be the arrival subset that corresponds to FF, i.e., Z=T(F)Z=T(F). Let ψ:FZ\psi:\vec{F}\to\reflectbox{$\vec{\reflectbox{$Z$}}$} and ψ:FZ\psi^{\prime}:\reflectbox{$\vec{\reflectbox{$F$}}$}\to\vec{Z} be any measure-preserving transformations such that ψ(x)E𝒞x\psi(x)E_{\mathcal{C}}x and ψ(x)E𝒞x\psi^{\prime}(x)E_{\mathcal{C}}x for all xx in the corresponding domains. Define V:XXV:X\to X by the following formula:

Vx={ψ(x)if xF,ψ(x)if xF,Txotherwise.Vx=\begin{cases}\psi(x)&\textrm{if }x\in\vec{F},\\ \psi^{\prime}(x)&\textrm{if }x\in\reflectbox{$\vec{\reflectbox{$F$}}$},\\ Tx&\textrm{otherwise.}\end{cases}

The transformation VV is a measure-preserving automorphism from the full group [X][\mathbb{R}\curvearrowright X\mkern 1.5mu] and Vx=TxVx=Tx for all xFx\not\in F. The integral of distances D(Tx,Vx)D(Tx,Vx) can be estimated as follows:

XD(Tx,Vx)𝑑μ(x)2D𝒞|ρT(x)|𝑑μ(x).\int_{X}D(Tx,Vx)\,d\mu(x)\leq 2\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{T}(x)|\,d\mu(x).

The following figure illustrates the notions of Lemma 9.13.

X\vec{X}

X\vec{\reflectbox{$X$}}

ccTTTTψ\psiψ\psi^{\prime}F\vec{F}

Z\vec{\reflectbox{$Z$}}

Z\vec{Z}

F\vec{\reflectbox{$F$}}

Figure 9.3. The transformation VV defined in Lemma 9.13.
Proof.

It is straightforward to verify that VV is a measure-preserving transformation. For the integral inequality note that for any xFx\in\vec{F} one has

D(Tx,Vx)|ρT(x)|+|ρT(T1x)|,D(Tx,Vx)\leq|\rho_{T}(x)|+|\rho_{T}(T^{-1}x)|,

and therefore

FD(Tx,Vx)𝑑μ(x)F|ρT|𝑑μ+F|ρT|𝑑μ=F|ρT|𝑑μD𝒞|ρT|𝑑μ.\displaystyle\int_{\vec{F}}D(Tx,Vx)\,d\mu(x)\leq\int_{\vec{F}}|\rho_{T}|\,d\mu+\int_{\reflectbox{$\vec{\reflectbox{$F$}}$}}|\rho_{T}|\,d\mu=\int_{F}|\rho_{T}|\,d\mu\leq\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{T}|\,d\mu.

A similar inequality holds for Fd(Tx,Vx)𝑑μ\int_{\reflectbox{$\vec{\reflectbox{$F$}}$}}d(Tx,Vx)\,d\mu, and the lemma follows. ∎

9.4. Periodic approximations

We now have all the ingredients necessary to prove that monotone transformations can be approximated by periodic automorphisms. Our arguments follow the approach outlined at the beginning of Section 9.3.

In the following lemma, we assume that the Lebesgue measure of those xXx\in\vec{X} that jump over any given c𝒞c\in\mathcal{C} is bounded from above by some β\beta, and that most of such jumps — of measure at least γ\gamma — are between adjacent E𝒞E_{\mathcal{C}}-classes. We are going to construct a periodic approximation PP of the transformation TT with an explicit bound on XD(Tx,Px)𝑑μ(x)\int_{X}D(Tx,Px)\,d\mu(x), which can be made small for a sufficiently sparse cross-section 𝒞\mathcal{C}. When the flow is ergodic, this lemma alone suffices to conclude that TD([X]1)T\in D([\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}). Theorem 9.15 builds upon Lemma 9.14 and treats the general case.

Lemma 9.14.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a monotone transformation, let K>0K>0 be a positive real, and let J={xsuppT:|ρT(x)|K}J=\{x\in\operatorname*{supp}T:|\rho_{T}(x)|\geq K\}. Let 𝒞\mathcal{C} be a cross-section such that gap𝒞(c)>K\mathrm{gap}_{\mathcal{C}}(c)>K for all c𝒞c\in\mathcal{C}. Let 0<γ<β0<\gamma<\beta be reals such that for all c𝒞c\in\mathcal{C}:

λc𝒞\displaystyle\lambda_{c}^{\mathcal{C}} ({xX:x<σ𝒞(c)Tx,TxE𝒞σ𝒞(c)})>γ,\displaystyle(\{x\in\vec{X}:x<\sigma_{\mathcal{C}}(c)\leq Tx,TxE_{\mathcal{C}}\sigma_{\mathcal{C}}(c)\})>\gamma,
λc𝒞\displaystyle\lambda_{c}^{\mathcal{C}} ({xX:Tx<cx,TxE𝒞σ𝒞1(c)})>γ,\displaystyle(\{x\in\reflectbox{$\vec{\reflectbox{$X$}}$}:Tx<c\leq x,TxE_{\mathcal{C}}\sigma^{-1}_{\mathcal{C}}(c)\})>\gamma,
λ\displaystyle\lambda ({xX:x<σ𝒞(c)Tx})<β.\displaystyle(\{x\in\vec{X}:x<\sigma_{\mathcal{C}}(c)\leq Tx\})<\beta.
λ\displaystyle\lambda ({xX:Tx<cx})<β.\displaystyle(\{x\in\reflectbox{$\vec{\reflectbox{$X$}}$}:Tx<c\leq x\})<\beta.

There exists a periodic transformation P[X]1P\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} such that suppPsuppT\operatorname*{supp}P\subseteq\operatorname*{supp}T and

XD(Tx,Px)𝑑μ(x)5D𝒞|ρT|𝑑μ+J|ρT|𝑑μ+K(βγ)γμ(suppT).\int_{X}D(Tx,Px)\,d\mu(x)\leq 5\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{T}|\,d\mu+\int_{J}|\rho_{T}|\,d\mu+\frac{K(\beta-\gamma)}{\gamma}\mu(\operatorname*{supp}T).
Proof.

Let D𝒞D_{\mathcal{C}} and A𝒞A_{\mathcal{C}} be the departure and the arrival sets of TT. Figure 9.4 depicts the arrival A𝒞(c)\vec{A}_{\mathcal{C}}(c) and the departure D𝒞(c)\vec{D}_{\mathcal{C}}(c) sets for an element cc of the cross-section 𝒞\mathcal{C}. Note that preimages T1(A𝒞(c))T^{-1}(\vec{A}_{\mathcal{C}}(c)) may come from different (possibly, infinitely many) E𝒞E_{\mathcal{C}}-equivalence classes; likewise, images T(D𝒞(c))T(\vec{D}_{\mathcal{C}}(c)) of the departure set may visit several E𝒞E_{\mathcal{C}}-equivalence classes.

X\vec{X}τ𝒞\tau_{\mathcal{C}}\ldots\ldots\ldotsA𝒞(c)\vec{A}_{\mathcal{C}}(c) D𝒞(c)\vec{D}_{\mathcal{C}}(c) ccσ𝒞(c)\sigma_{\mathcal{C}}(c)
Figure 9.4. The arrival A𝒞(c)\vec{A}_{\mathcal{C}}(c) and the departure D𝒞(c)\vec{D}_{\mathcal{C}}(c) sets for some c𝒞c\in\mathcal{C}.

Set ξ(c)=γ\xi(c)=\gamma to be the constant function; in view of the assumptions on γ\gamma, we may apply Lemma 9.9 to get positive and negative ξ\xi-copious arrival sets A𝒞A𝒞\vec{A}^{\bullet}_{\mathcal{C}}\subseteq A_{\mathcal{C}}, A𝒞A𝒞\reflectbox{$\vec{\reflectbox{$A$}}$}^{\bullet}_{\mathcal{C}}\subseteq A_{\mathcal{C}}, as well as the corresponding departure sets D𝒞=τ𝒞(A𝒞)\vec{D}^{\bullet}_{\mathcal{C}}=\tau_{\mathcal{C}}(\vec{A}^{\bullet}_{\mathcal{C}}) and D𝒞=τ𝒞(A𝒞)\reflectbox{$\vec{\reflectbox{$D$}}$}^{\bullet}_{\mathcal{C}}=\tau_{\mathcal{C}}(\reflectbox{$\vec{\reflectbox{$A$}}$}^{\bullet}_{\mathcal{C}}). Set A𝒞=A𝒞A𝒞A^{\bullet}_{\mathcal{C}}=\vec{A}^{\bullet}_{\mathcal{C}}\sqcup\reflectbox{$\vec{\reflectbox{$A$}}$}^{\bullet}_{\mathcal{C}} and D𝒞=D𝒞D𝒞D^{\bullet}_{\mathcal{C}}=\vec{D}^{\bullet}_{\mathcal{C}}\sqcup\reflectbox{$\vec{\reflectbox{$D$}}$}^{\bullet}_{\mathcal{C}}. We have λ(A𝒞(c))=2γ=λ(D𝒞(c))\lambda(A^{\bullet}_{\mathcal{C}}(c))=2\gamma=\lambda(D^{\bullet}_{\mathcal{C}}(c)) for all c𝒞c\in\mathcal{C}. Let

A𝒞\displaystyle A^{\circ}_{\mathcal{C}} ={xA𝒞:T1xE𝒞σ𝒞1(π𝒞(x))}{xA𝒞:T1xE𝒞σ𝒞(π𝒞(x))},\displaystyle=\bigl{\{}x\in\vec{A}_{\mathcal{C}}:T^{-1}xE_{\mathcal{C}}\sigma_{\mathcal{C}}^{-1}(\pi_{\mathcal{C}}(x))\bigr{\}}\cup\bigl{\{}x\in\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}}:T^{-1}xE_{\mathcal{C}}\sigma_{\mathcal{C}}(\pi_{\mathcal{C}}(x))\bigr{\}},
D𝒞\displaystyle D^{\circ}_{\mathcal{C}} ={xD𝒞:TxE𝒞σ𝒞(π𝒞(x))}{xD𝒞:TxE𝒞σ𝒞1(π𝒞(x))},\displaystyle=\bigl{\{}x\in\vec{D}_{\mathcal{C}}:TxE_{\mathcal{C}}\sigma_{\mathcal{C}}(\pi_{\mathcal{C}}(x))\bigr{\}}\cup\bigl{\{}x\in\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}}:TxE_{\mathcal{C}}\sigma_{\mathcal{C}}^{-1}(\pi_{\mathcal{C}}(x))\bigr{\}},

be the set of arcs that jump from/to the next E𝒞E_{\mathcal{C}}-equivalence class. By the assumptions of the lemma, we have λ(D𝒞(c))γ\lambda(\vec{D}^{\circ}_{\mathcal{C}}(c))\geq\gamma and λ(A𝒞(c))γ\lambda(\vec{A}^{\circ}_{\mathcal{C}}(c))\geq\gamma for all c𝒞c\in\mathcal{C}. Let ϕ\phi be any measure-preserving transformation such that:

  • ϕ\phi is supported on D𝒞D_{\mathcal{C}};

  • ϕ(D𝒞)=D𝒞\phi(\vec{D}_{\mathcal{C}})=\vec{D}_{\mathcal{C}} and ϕ(D𝒞)=D𝒞\phi(\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}})=\reflectbox{$\vec{\reflectbox{$D$}}$}_{\mathcal{C}};

  • ϕ(x)E𝒞x\phi(x)E_{\mathcal{C}}x for all xXx\in X;

and moreover

(9.4) ϕ(D𝒞)D𝒞.\phi(D^{\bullet}_{\mathcal{C}})\subseteq D^{\circ}_{\mathcal{C}}.

Select a transformation ϕ\phi^{\prime} such that

  • ϕ\phi^{\prime} is supported on A𝒞A_{\mathcal{C}};

  • ϕ(A𝒞)=A𝒞\phi^{\prime}(\vec{A}_{\mathcal{C}})=\vec{A}_{\mathcal{C}} and ϕ(A𝒞)=A𝒞\phi^{\prime}(\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}})=\reflectbox{$\vec{\reflectbox{$A$}}$}_{\mathcal{C}};

  • ϕ(x)E𝒞x\phi^{\prime}(x)E_{\mathcal{C}}x for all xXx\in X;

and moreover

(9.5) ϕ(Tϕ(D𝒞))=A𝒞.\phi^{\prime}(T\circ\phi(D^{\bullet}_{\mathcal{C}}))=A^{\bullet}_{\mathcal{C}}.

Figure 9.5 illustrates these maps. Note that while in general τ𝒞(A(c))D(c)\tau_{\mathcal{C}}\bigl{(}\vec{A}^{\circ}(c)\bigr{)}\neq\vec{D}^{\circ}(c), one has τ𝒞(A(c))=D(c)\tau_{\mathcal{C}}\bigl{(}\vec{A}^{\bullet}(c)\bigr{)}=\vec{D}^{\bullet}(c) for all c𝒞c\in\mathcal{C} by the definition of the ξ\xi-copious departure set.

X\vec{X}ccA𝒞(c)\vec{A}^{\circ}_{\mathcal{C}}(c) D𝒞(c)\vec{D}^{\circ}_{\mathcal{C}}(c) A𝒞(c)\vec{A}^{\bullet}_{\mathcal{C}}(c) D𝒞(c)\vec{D}^{\bullet}_{\mathcal{C}}(c) τ𝒞\tau_{\mathcal{C}}ϕ\phi^{\prime}ϕ\phi^{\prime}ϕ\phi
Figure 9.5. Automorphism ϕ\phi maps D𝒞(c)D^{\bullet}_{\mathcal{C}}(c) into D𝒞(c)D^{\circ}_{\mathcal{C}}(c) and (ϕ)1(\phi^{\prime})^{-1} sends A𝒞(c)A^{\bullet}_{\mathcal{C}}(c) into A𝒞(c)A^{\circ}_{\mathcal{C}}(c).

Let UU be the transformation obtained by applying Lemma 9.12 to TT, ϕ\phi and ϕ\phi^{\prime}. The automorphism UU satisfies U(D𝒞(c))=A𝒞(σ𝒞(c))U(\vec{D}^{\bullet}_{\mathcal{C}}(c))=\vec{A}^{\bullet}_{\mathcal{C}}(\sigma_{\mathcal{C}}(c)) and U(D𝒞(c))=A𝒞(σ𝒞1(c))U(\reflectbox{$\vec{\reflectbox{$D$}}$}^{\bullet}_{\mathcal{C}}(c))=\reflectbox{$\vec{\reflectbox{$A$}}$}^{\bullet}_{\mathcal{C}}(\sigma^{-1}_{\mathcal{C}}(c)) for all c𝒞c\in\mathcal{C}. Choose a measure-preserving transformation ψ:D𝒞A𝒞\psi:\vec{D}^{\bullet}_{\mathcal{C}}\to\reflectbox{$\vec{\reflectbox{$A$}}$}^{\bullet}_{\mathcal{C}} such that xE𝒞ψ(x)xE_{\mathcal{C}}\psi(x) for all xx in the domain of ψ\psi. Set ψ=τ𝒞1ψ1τ𝒞1:D𝒞A𝒞\psi^{\prime}=\tau_{\mathcal{C}}^{-1}\circ\psi^{-1}\circ\tau_{\mathcal{C}}^{-1}:\reflectbox{$\vec{\reflectbox{$D$}}$}^{\bullet}_{\mathcal{C}}\to\vec{A}^{\bullet}_{\mathcal{C}}. Let VV be the transformation that is produced by Lemma 9.13 applied to UU, ψ\psi, and ψ\psi^{\prime} (see Figure 9.6).

X\vec{X}

X\vec{\reflectbox{$X$}}

ccA𝒞(c)\vec{A}^{\bullet}_{\mathcal{C}}(c) D𝒞(c)\vec{D}^{\bullet}_{\mathcal{C}}(c) τ𝒞\tau_{\mathcal{C}}D𝒞(c)\reflectbox{$\vec{\reflectbox{$D$}}$}^{\bullet}_{\mathcal{C}}(c) A𝒞(c)\reflectbox{$\vec{\reflectbox{$A$}}$}^{\bullet}_{\mathcal{C}}(c) τ𝒞\tau_{\mathcal{C}}ψ\psiψ=τ𝒞1ψ1τ𝒞1\psi^{\prime}=\tau_{\mathcal{C}}^{-1}\circ\psi^{-1}\circ\tau_{\mathcal{C}}^{-1}
Figure 9.6. Construction of the automorphism VV from UU, ψ\psi, and ψ\psi^{\prime}.

Finally, set P:XXP:X\to X to be

Px={Vxif x[D𝒞]EV,xotherwise.Px=\begin{cases}Vx&\textrm{if }x\in[D^{\bullet}_{\mathcal{C}}]_{E_{V}},\\ x&\textrm{otherwise}.\\ \end{cases}

We claim that PP satisfies the conclusions of the lemma. It is periodic, since the transformation ψτ𝒞ψτ𝒞\psi^{\prime}\circ\tau_{\mathcal{C}}\circ\psi\circ\tau_{\mathcal{C}} is the identity map and suppPsuppT\operatorname*{supp}P\subseteq\operatorname*{supp}T by construction. It remains to estimate XD(Tx,Px)𝑑μ(x)\int_{X}D(Tx,Px)\,d\mu(x).

XD(Tx,Px)𝑑μ(x)\displaystyle\int_{X}D(Tx,Px)\,d\mu(x) XD(Tx,Ux)𝑑μ(x)+XD(Ux,Vx)𝑑μ(x)\displaystyle\leq\int_{X}D(Tx,Ux)\,d\mu(x)+\int_{X}D(Ux,Vx)\,d\mu(x)
+XD(Vx,Px)𝑑μ(x)\displaystyle\qquad+\int_{X}D(Vx,Px)\,d\mu(x)
[Estimates of Lemma 9.12 and Lemma 9.13]\displaystyle\leq[\textrm{Estimates of Lemma~{}\ref{lem:departure-set-transformation} and Lemma~{}\ref{lem:gluing-arriving-and-departure-sets}}]
4D𝒞|ρT|𝑑μ+XD(Vx,Px)𝑑μ(x).\displaystyle\leq 4\int_{D_{\mathcal{C}}}\mkern-10.0mu|\rho_{T}|\,d\mu+\int_{X}D(Vx,Px)\,d\mu(x).

We concentrate on estimating XD(Vx,Px)𝑑μ(x)\int_{X}D(Vx,Px)\,d\mu(x). Recall that Tx=Ux=VxTx=Ux=Vx for all xD𝒞x\not\in D_{\mathcal{C}}, hence ρT(x)=ρV(x)\rho_{T}(x)=\rho_{V}(x) for xD𝒞x\not\in D_{\mathcal{C}}. Set Ψ=(XX)[D𝒞]EV\Psi=(\vec{X}\cup\reflectbox{$\vec{\reflectbox{$X$}}$})\setminus[D^{\bullet}_{\mathcal{C}}]_{E_{V}} and note that Vx=UxVx=Ux for xΨx\in\Psi, and therefore using the conclusion of Lemma 9.12 we have

(9.6) D𝒞Ψ|ρV|𝑑μ=D𝒞Ψ|ρU|𝑑μD𝒞|ρU|𝑑μ=D𝒞|ρT|𝑑μ.\int_{D_{\mathcal{C}}\cap\Psi}|\rho_{V}|\,d\mu=\int_{D_{\mathcal{C}}\cap\Psi}|\rho_{U}|\,d\mu\leq\int_{D_{\mathcal{C}}}|\rho_{U}|\,d\mu=\int_{D_{\mathcal{C}}}|\rho_{T}|\,d\mu.

The integral XD(Vx,Px)𝑑μ(x)\int_{X}D(Vx,Px)\,d\mu(x) can now be estimated as follows.

XD(Vx,Px)𝑑μ(x)\displaystyle\int_{X}D(Vx,Px)\,d\mu(x) =Ψ|ρV|𝑑μ\displaystyle=\int_{\Psi}|\rho_{V}|\,d\mu
ΨD𝒞|ρV|𝑑μ+D𝒞Ψ|ρV|𝑑μ\displaystyle\leq\int_{\Psi\setminus D_{\mathcal{C}}}|\rho_{V}|\,d\mu+\int_{D_{\mathcal{C}}\cap\Psi}|\rho_{V}|\,d\mu
Tx=Vx for xD𝒞 and Eq. (9.6)\displaystyle\because\textrm{$Tx=Vx$ for $x\not\in D_{\mathcal{C}}$ and Eq.~{}\eqref{eq:cocycle-bound-T-and-V}} ΨD𝒞|ρT|𝑑μ+D𝒞|ρT|𝑑μ.\displaystyle\leq\int_{\Psi\setminus D_{\mathcal{C}}}|\rho_{T}|\,d\mu+\int_{D_{\mathcal{C}}}|\rho_{T}|\,d\mu.

Finally, we consider the integral ΨD𝒞|ρT|𝑑μ\int_{\Psi\setminus D_{\mathcal{C}}}|\rho_{T}|\,d\mu and partition its domain ΨD𝒞\Psi\setminus D_{\mathcal{C}} as (J(ΨD𝒞))((ΨD𝒞)J)(J\cap(\Psi\setminus D_{\mathcal{C}}))\sqcup((\Psi\setminus D_{\mathcal{C}})\setminus J), which yields

ΨD𝒞|ρT|𝑑μ\displaystyle\int_{\Psi\setminus D_{\mathcal{C}}}|\rho_{T}|\,d\mu J|ρT|𝑑μ+Kμ(Ψ)\displaystyle\leq\int_{J}|\rho_{T}|\,d\mu+K\mu(\Psi)
J|ρT|𝑑μ+K(βγ)γμ(suppT),\displaystyle\leq\int_{J}|\rho_{T}|\,d\mu+\frac{K(\beta-\gamma)}{\gamma}\mu(\operatorname*{supp}T),

where the last inequality follows from Lemma 9.11 with δ=1γ/β\delta=1-\gamma/\beta. Combining all the inequalities together, we get

XD(Tx,Px)𝑑μ(x)5D𝒞|ρT|𝑑μ+J|ρT|𝑑μ+K(βγ)γμ(suppT).\int_{X}D(Tx,Px)\,d\mu(x)\leq 5\int_{D_{\mathcal{C}}}|\rho_{T}|\,d\mu+\int_{J}|\rho_{T}|\,d\mu+\frac{K(\beta-\gamma)}{\gamma}\mu(\operatorname*{supp}T).\qed

Lemma 9.14 allows us to approximate with a periodic transformation a monotone TT for which the Lebesgue measure of points jumping over any given cXc\in X is roughly constant across orbits. To deal with the general case, we simply need to split the phase space XX into countably many segments invariant under the flow, and apply Lemma 9.14 on each of them separately. Small care needs to be taken to ensure that values (βγ)/γ(\beta-\gamma)/\gamma, which appear in the formulation of Lemma 9.14, remain uniformly small across the partition of XX. Details are presented in the following theorem.

Theorem 9.15.

Let T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a monotone transformation that belongs to the kernel of the index map, hence

λ({xsuppT:x<cTx})=λ({ysuppT:Ty<cy})\lambda(\{x\in\operatorname*{supp}T:x<c\leq Tx\})=\lambda(\{y\in\operatorname*{supp}T:Ty<c\leq y\})

for almost all cXc\in X. For any ϵ>0\epsilon>0 there exists a periodic transformation P[X]1P\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} such that suppPsuppT\operatorname*{supp}P\subseteq\operatorname*{supp}T and XD(Tx,Px)𝑑μ(x)<ϵ\int_{X}D(Tx,Px)\,d\mu(x)<\epsilon.

Proof.

Let Kϵ1K_{\epsilon}\geq 1 be such that for the set

Jϵ={xsuppT:|ρT(x)|Kϵ}J_{\epsilon}=\{x\in\operatorname*{supp}T:|\rho_{T}(x)|\geq K_{\epsilon}\}

one has Jϵ|ρT|𝑑μ<ϵ/18\int_{J_{\epsilon}}|\rho_{T}|\,d\mu<\epsilon/18. Pick a cross-section 𝒞\mathcal{C} with gaps so large that

2Kϵ2/gap(c)<ϵ/152K_{\epsilon}^{2}/\mathrm{gap}(c)<\epsilon/15

for all c𝒞c\in\mathcal{C}, which ensures

(9.7) Kϵμ(D𝒞Jϵ)ϵ/15.K_{\epsilon}\cdot\mu(D_{\mathcal{C}}\setminus J_{\epsilon})\leq\epsilon/15.

Note also that Eq. (9.7) holds for any cross-section 𝒞𝒞\mathcal{C}^{\prime}\subseteq\mathcal{C}, since D𝒞D𝒞D_{\mathcal{C}^{\prime}}\subseteq D_{\mathcal{C}} and gap𝒞(c)gap𝒞(c)\mathrm{gap}_{\mathcal{C}^{\prime}}(c)\geq\mathrm{gap}_{\mathcal{C}}(c) for all c𝒞c\in\mathcal{C}^{\prime}.

For any positive real α>0\alpha>0 there exists a positive δ(α)=δ>0\delta(\alpha)=\delta>0 so small that δ<α\delta<\alpha and 2δ/(αδ)<ϵ/3Kϵ2\delta/(\alpha-\delta)<\epsilon/3K_{\epsilon}. We may therefore pick countably many positive reals αn>0\alpha_{n}>0, n1n\geq 1, such that >0=n(αnδn/2,αn+δn/2)\mathbb{R}^{>0}=\bigcup_{n}(\alpha_{n}-\delta_{n}/2,\alpha_{n}+\delta_{n}/2) and

(9.8) (2δnαnδn)<ϵ3Kϵn1.\Bigl{(}\frac{2\delta_{n}}{\alpha_{n}-\delta_{n}}\Bigr{)}<\frac{\epsilon}{3K_{\epsilon}}\quad\forall n\geq 1.

Define intervals In=(αnδn/2,αn+δn/2)I_{n}=(\alpha_{n}-\delta_{n}/2,\alpha_{n}+\delta_{n}/2), n1n\geq 1.

Let ζ:𝒞R0\zeta:\mathcal{C}\to R^{\geq 0} be the map that measures the set of forward arcs over its argument:

ζ(c)=λ({xX:x<cTx}).\zeta(c)=\lambda\bigl{(}\{x\in\vec{X}:x<c\leq Tx\}\bigr{)}.

Set 𝒞1=ζ1(I1)\mathcal{C}_{1}=\zeta^{-1}(I_{1}) and construct inductively 𝒞n=ζ1(In)[k<n𝒞k]E\mathcal{C}_{n}=\zeta^{-1}(I_{n})\setminus\bigl{[}\bigcup_{k<n}\mathcal{C}_{k}\bigr{]}_{E}. Sets 𝒞n\mathcal{C}_{n} are pairwise disjoint, and moreover, ¬c1Ec2\neg c_{1}Ec_{2} for all c1𝒞n1c_{1}\in\mathcal{C}_{n_{1}}, c2𝒞n2c_{2}\in\mathcal{C}_{n_{2}}, n1n2n_{1}\neq n_{2}. Let χn:𝒞n\chi_{n}:\mathcal{C}_{n}\to\mathbb{N}, n1n\geq 1, be the function defined by

χn(c)=min{m:λ({xX:x<cTx,D(x,c)m,D(Tx,c)m})>ζ(c)δn/2}.\chi_{n}(c)=\min\Bigl{\{}m\in\mathbb{N}:\\ \lambda\bigl{(}\bigl{\{}x\in\vec{X}:x<c\leq Tx,D(x,c)\leq m,D(Tx,c)\leq m\bigr{\}}\bigr{)}>\zeta(c)-\delta_{n}/2\Bigr{\}}.

Set 𝒞n,1=χn1(1)\mathcal{C}^{\prime}_{n,1}=\chi^{-1}_{n}(1) and define inductively 𝒞n,m=χ1(m)[k<m𝒞n,k]E\mathcal{C}^{\prime}_{n,m}=\chi^{-1}(m)\setminus\bigl{[}\bigcup_{k<m}\mathcal{C}^{\prime}_{n,k}\bigr{]}_{E}. Let Xn,mX_{n,m} denote the saturated set [𝒞n,m]E[\mathcal{C}^{\prime}_{n,m}]_{E}. Finally, for all m,n1m,n\geq 1, let 𝒞n,m𝒞n,m\mathcal{C}_{n,m}\subseteq\mathcal{C}_{n,m}^{\prime} be such that gap𝒞n,m(c)>m\mathrm{gap}_{\mathcal{C}_{n,m}}(c)>m for all c𝒞n,mc\in\mathcal{C}_{n,m}. Sets 𝒞n,m\mathcal{C}_{n,m} and Xn,mX_{n,m} satisfy the following conditions:

  1. (1)

    𝒞n,m\mathcal{C}_{n,m} is a cross-section for the restriction of the flow onto Xn,mX_{n,m};

  2. (2)

    sets Xn,mX_{n,m}, m,n1m,n\geq 1, are pairwise disjoint.

  3. (3)

    ζ(c)In\zeta(c)\in I_{n} and λc𝒞({xX:x<σ𝒞n,m(c)Tx})>αnδn\lambda_{c}^{\mathcal{C}}\bigl{(}\{x\in\vec{X}:x<\sigma_{\mathcal{C}_{n,m}}(c)\leq Tx\}\bigr{)}>\alpha_{n}-\delta_{n} for all c𝒞n,mc\in\mathcal{C}_{n,m}.

Let Tn,mT_{n,m} denote the restriction of TT onto Xn,mX_{n,m}. Apply Lemma 9.14 to the transformation Tn,mT_{n,m}, cross-section 𝒞n,m\mathcal{C}_{n,m}, which has gaps at least KϵK_{\epsilon}, and β=αn+δn\beta=\alpha_{n}+\delta_{n}, γ=αnδn\gamma=\alpha_{n}-\delta_{n}. Let Pn,mP_{n,m} be the resulting periodic transformation on Xn,mX_{n,m}. Set P=n,mPn,mP=\bigsqcup_{n,m}P_{n,m}. We claim that PP satisfies conclusions of the theorem. Set 𝒞=n,m𝒞n,m\mathcal{C}^{\prime}=\bigsqcup_{n,m}\mathcal{C}_{n,m} and note that 𝒞𝒞\mathcal{C}^{\prime}\subseteq\mathcal{C}, whence D𝒞D𝒞D_{\mathcal{C}^{\prime}}\subseteq D_{\mathcal{C}}.

XD(Tx,Px)𝑑μ(x)\displaystyle\int_{X}D(Tx,Px)\,d\mu(x) =n,mXn,mD(Tn,mx,Pn,mx)𝑑μ(x)\displaystyle=\sum_{n,m}\int_{X_{n,m}}D(T_{n,m}x,P_{n,m}x)\,d\mu(x)
Lemma 9.14\displaystyle\because\textrm{Lemma~{}\ref{lem:periodic-construction-one-cross-section}} 5n,mD𝒞n,m|ρT|𝑑μ+n,mJϵXn,m|ρT|𝑑μ\displaystyle\leq 5\sum_{n,m}\int_{D_{\mathcal{C}_{n,m}}}|\rho_{T}|\,d\mu+\sum_{n,m}\int_{J_{\epsilon}\cap X_{n,m}}|\rho_{T}|\,d\mu
+n,mKϵ(2δnαnδn)μ(Xn,m)\displaystyle\qquad\qquad+\sum_{n,m}K_{\epsilon}\Bigl{(}\frac{2\delta_{n}}{\alpha_{n}-\delta_{n}}\Bigr{)}\mu(X_{n,m})
Eq. (9.8)\displaystyle\because\textrm{Eq.~{}\eqref{eq:alpha-delta-condition}} 5D𝒞|ρT|𝑑μ+Jϵ|ρT|𝑑μ+(ϵ/3)μ(X)\displaystyle\leq 5\int_{D_{\mathcal{C}}}|\rho_{T}|\,d\mu+\int_{J_{\epsilon}}|\rho_{T}|\,d\mu+(\epsilon/3)\mu(X)
5D𝒞Jϵ|ρT|𝑑μ+6Jϵ|ρT|𝑑μ+ϵ/3\displaystyle\leq 5\int_{D_{\mathcal{C}}\setminus J_{\epsilon}}|\rho_{T}|\,d\mu+6\int_{J_{\epsilon}}|\rho_{T}|\,d\mu+\epsilon/3
choice of Kϵ\displaystyle\because\textrm{choice of $K_{\epsilon}$} <5Kϵμ(D𝒞Jϵ)+ϵ/3+ϵ/3\displaystyle<5K_{\epsilon}\mu(D_{\mathcal{C}}\setminus J_{\epsilon})+\epsilon/3+\epsilon/3
Eq. (9.7)\displaystyle\because\textrm{Eq.~{}\eqref{eq:gaps-in-C-are-large}} ϵ,\displaystyle\leq\epsilon,

and the theorem follows. ∎

Corollary 9.16.

Let X\mathbb{R}\curvearrowright X be a measure-preserving flow and T[X]1T\in[\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1} be a dissipative transformation. If (T)=0\mathcal{I}(T)=0, then TD([X]1)T\in D([\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}).

Proof.

By Corollary 9.8, there is a monotone transformation UU and a periodic transformation PP such that T=UPT=U\circ P. Since PD([X]1)P\in D([\mathbb{R}\curvearrowright X\mkern 1.5mu]_{1}) by Corollary 3.15, it remains to show that UU belongs to the derived subgroup. The latter follows from Theorem 9.15, since (U)=(T)(P)=0\mathcal{I}(U)=\mathcal{I}(T)-\mathcal{I}(P)=0. ∎

Chapter 10 Conclusions

Our objective in this last chapter is to draw several conclusions regarding the structure of the L1\mathrm{L}^{1} full groups of measure-preserving flows. The analysis conducted in Chapters 8 and 9 leads to the most technically challenging result of our work, which is the following theorem.

Theorem 10.1.

Let :X\mathcal{F}:\mathbb{R}\curvearrowright X be a free measure-preserving flow on a standard probability space. The kernel of the index map coincides with the closed derived subgroup D([]1)D([\mathcal{F}\mkern 1.5mu]_{1}).

Proof.

Inclusion D([]1)kerD([\mathcal{F}\mkern 1.5mu]_{1})\subseteq\ker{\mathcal{I}} is automatic since the image of \mathcal{I} is abelian. For the other direction, pick a transformation TkerT\in\ker{\mathcal{I}} and consider its Hopf’s decomposition X=CDX=C\sqcup D provided by Proposition 4.16. We have T=TCTDT=T_{C}\circ T_{D}, where TC[]1T_{C}\in[\mathcal{F}\mkern 1.5mu]_{1} is conservative and TD[]1T_{D}\in[\mathcal{F}\mkern 1.5mu]_{1} is dissipative. According to Corollary 8.8, (TC)=0\mathcal{I}(T_{C})=0 and TCD([]1)T_{C}\in D([\mathcal{F}\mkern 1.5mu]_{1}), whence (TD)=(T)(TC)=0\mathcal{I}(T_{D})=\mathcal{I}(T)-\mathcal{I}(T_{C})=0. Therefore, the dissipative part TDT_{D} satisfies the assumptions of Corollary 9.16, which yields TDD([]1)T_{D}\in D([\mathcal{F}\mkern 1.5mu]_{1}), and hence TD([]1)T\in D([\mathcal{F}\mkern 1.5mu]_{1}) as desired. ∎

10.1. Topological ranks of L1\mathrm{L}^{1} full groups

Empowered with the result above and Corollary 5.20, we can estimate the topological ranks of L1\mathrm{L}^{1} full groups of flows. We recall the following well-known inequalities.

Proposition 10.2.

Let ϕ:GH\phi:G\to H be a surjective continuous homomorphism of Polish groups. The topological rank rk(G)\mathrm{rk}(G) satisfies

rk(H)rk(G)rk(H)+rk(kerϕ).\mathrm{rk}(H)\leq\mathrm{rk}(G)\leq\mathrm{rk}(H)+\mathrm{rk}(\ker{\phi}).
Proposition 10.3.

Let :X\mathcal{F}:\mathbb{R}\curvearrowright X be a free measure-preserving flow on a standard probability space (X,μ)(X,\mu). The topological rank rk([]1)\mathrm{rk}([\mathcal{F}\mkern 1.5mu]_{1}) is finite if and only if the flow has finitely many ergodic components. Moreover, if \mathcal{F} has exactly nn ergodic components then

n+1rk([]1)n+3.n+1\leq\mathrm{rk}([\mathcal{F}\mkern 1.5mu]_{1})\leq n+3.
Proof.

Let \mathcal{E} be the space of probability invariant ergodic measures of the flow, and let pp be the probability measure on \mathcal{E} such that μ=ν𝑑p(ν)\mu=\int_{\mathcal{E}}\nu\,dp(\nu) (see Appendix C.3). Proposition 6.6 shows that the index map :[]1L1(,p)\mathcal{I}:[\mathcal{F}\mkern 1.5mu]_{1}\to\mathrm{L}^{1}(\mathcal{E},p) is continuous and surjective. An application of Proposition 10.2 yields

(10.1) rk(L1(,p))rk([]1)rk(L1(,p))+rk(ker)=rk(L1(,p))+2.\mathrm{rk}(\mathrm{L}^{1}(\mathcal{E},p))\leq\mathrm{rk}([\mathcal{F}\mkern 1.5mu]_{1})\leq\mathrm{rk}(\mathrm{L}^{1}(\mathcal{E},p))+\mathrm{rk}(\ker\mathcal{I})=\mathrm{rk}(\mathrm{L}^{1}(\mathcal{E},p))+2.

where the last equality is based on Theorem 10.1 and Corollary 5.20. Since L1(,ν)\mathrm{L}^{1}(\mathcal{E},\nu) is a Banach space, its topological rank is finite if and only if its dimension is finite, which is equivalent to (,p)(\mathcal{E},p) being purely atomic with finitely many atoms. We have shown that rk([]1)\mathrm{rk}([\mathcal{F}\mkern 1.5mu]_{1}) is finite if and only if the flow has only finitely many ergodic components. The moreover part of the proposition follows from the inequality (10.1) and the observation that rk(L1(,p))=dim(L1(,p))+1\mathrm{rk}(\mathrm{L}^{1}(\mathcal{E},p))=\dim(\mathrm{L}^{1}(\mathcal{E},p))+1. ∎

As already mentioned in the introduction, we conjecture that the topological rank completely remembers the number of ergodic components.

Conjecture 10.4.

Let \mathcal{F} be a measure-preserving flow. If it has exactly nn ergodic components, then rk([]1)=n+1\mathrm{rk}([\mathcal{F}\mkern 1.5mu]_{1})=n+1.

Provided the conjecture holds, we have a priori no way of distinguishing L1\mathrm{L}^{1} full groups of ergodic flows as topological groups. For \mathbb{Z}-actions, it is a consequence of Belinskaya’s theorem that there are many L1\mathrm{L}^{1} full groups. The next two sections are devoted to analogues of her result for flows, yielding that there are many L1\mathrm{L}^{1} full groups of free ergodic flows, although we don’t have a concrete way of distinguishing them (we will discuss in the last section their geometry, which might help there).

10.2. Katznelson’s conjugation theorem

R. M. Belinskaja [Bel68] showed that if measure-preserving transformations T,UAut(X,μ)T,U\in\mathrm{Aut}(X,\mu) generate the same orbit equivalence relation, i.e., T=U\mathcal{R}_{T}=\mathcal{R}_{U}, and U[T]1U\in[T\mkern 1.5mu]_{1}, then TT and UU are conjugated. Y. Katznelson found a different argument and isolated a sufficient condition for conjugacy of measure-preserving transformations (see [CJMT22, Theorem A.1]). In the following, for TAut(X,μ)T\in\mathrm{Aut}(X,\mu), xXx\in X, and AA\subseteq\mathbb{Z} we let TAxT^{A}x denote the set {Tkx:kA}\{T^{k}x:k\in A\}.

Theorem 10.5 (Katznelson).

Suppose T,UAut(X,μ)T,U\in\mathrm{Aut}(X,\mu) are measure-preserving transformations that generate the same orbit equivalence relation, T=U\mathcal{R}_{T}=\mathcal{R}_{U}. If the symmetric difference TxUxT^{\mathbb{N}}x\bigtriangleup U^{\mathbb{N}}x is finite for almost all xx, then TT and UU are conjugated by an element from the full group [T]=[U][T\mkern 1.5mu]=[U\mkern 1.5mu].

The analog of this result for free measure-preserving flows will be proved shortly in Theorem 10.9. But first we discuss an important application of Theorem 10.5. Consider a free measure-preserving flow :X\mathcal{F}:\mathbb{R}\curvearrowright X. Given a dissipative transformation T[]T\in[\mathcal{F}\mkern 1.5mu] (in the sense of Definition 9.1), Proposition B.4 implies that almost every non-trivial TT-orbit [x]T[x]_{\mathcal{R}_{T}} is a discrete subset of [x][x]_{\mathcal{R}} unbounded both from below and from above. The order induced on [x]T[x]_{\mathcal{R}_{T}} by the flow may disagree with the TT-order of points. One may therefore define the \mathcal{F}-reordering of TT to be the first return transformation T~\tilde{T} induced by the ordering of the flow on the orbits of TT:

T~x=x+min{r>0:x+r[x]T}for xsuppT.\tilde{T}x=x+\min\{r>0:x+r\in[x]_{\mathcal{R}_{T}}\}\qquad\textrm{for }x\in\operatorname*{supp}T.

Note that TT and T~\tilde{T} generate the same orbit equivalence relation, T=T~\mathcal{R}_{T}=\mathcal{R}_{\tilde{T}}.

If TT belongs to the L1\mathrm{L}^{1} full group of the flow, either TxT~xT^{\mathbb{N}}x\bigtriangleup\tilde{T}^{\mathbb{N}}x or TxT~xT^{\mathbb{N}}x\bigtriangleup\tilde{T}^{-\mathbb{N}}x is finite, depending on whether limnρ(x,Tnx)=+\lim_{n}\rho(x,T^{n}x)=+\infty or limnρ(x,Tnx)=\lim_{n}\rho(x,T^{n}x)=-\infty (cf. Corollary 9.4). Which symmetric difference is finite may depend on the point xXx\in X, and Theorem 10.5 can be used to show that TT and its reordering T~\tilde{T} are flip conjugated.

Definition 10.6.

Let (X1,μ1)(X_{1},\mu_{1}) and (X2,μ2)(X_{2},\mu_{2}) be standard probability spaces, and let TiAut(Xi,μi)T_{i}\in\mathrm{Aut}(X_{i},\mu_{i}), i=1,2i=1,2. Measure-preserving transformations T1T_{1} and T2T_{2} are flip conjugate if there exist an isomorphism of measure spaces S:X1X2S:X_{1}\to X_{2} and measurable partitions X1=X1X1+X_{1}=X_{1}^{-}\sqcup X_{1}^{+}, X2=X2X2+X_{2}=X_{2}^{-}\sqcup X_{2}^{+} such that

  1. (1)

    S(X1)=X2S(X_{1}^{-})=X_{2}^{-} and S(X1+)=X2+S(X_{1}^{+})=X_{2}^{+};

  2. (2)

    X1,X1+X_{1}^{-},X_{1}^{+} are T1T_{1}-invariant and X2,X2+X_{2}^{-},X_{2}^{+} are T2T_{2}-invariant;

  3. (3)

    ST1X1+S1=T2X2+ST_{1}\restriction_{X_{1}^{+}}S^{-1}=T_{2}\restriction_{X_{2}^{+}} and ST1X1S1=T21X2ST_{1}\restriction_{X_{1}^{-}}S^{-1}=T^{-1}_{2}\restriction_{X_{2}^{-}}.

Note that when one of the TiT_{i}’s is ergodic, our definition of flip-conjugacy coincides with the standard one, which requires XiX_{i}^{-} or Xi+X_{i}^{+} to have full measure.

Proposition 10.7.

Any dissipative T[]1T\in[\mathcal{F}\mkern 1.5mu]_{1} and its \mathcal{F}-reordering T~\tilde{T} are flip conjugated by an element from the full group [T]=[T~][T\mkern 1.5mu]=[\tilde{T}\mkern 1.5mu].

Proof.

Consider the decomposition suppT=XX\operatorname*{supp}T=\reflectbox{$\vec{\reflectbox{$X$}}$}\sqcup\vec{X} into the positive and negative orbits as in Definition 9.5. In particular, TxT~xT^{\mathbb{N}}x\triangle\tilde{T}^{\mathbb{N}}x and TxT~xT^{\mathbb{N}}x\triangle\tilde{T}^{-\mathbb{N}}x are finite for xXx\in\vec{X} and xXx\in\reflectbox{$\vec{\reflectbox{$X$}}$}, respectively. Theorem 10.5 implies that there exist automorphisms S1[TX]S_{1}\in[T\restriction_{{\vec{X}}}\mkern 1.5mu] and S2[TX]S_{2}\in[T\restriction_{{\reflectbox{$\vec{\reflectbox{$X$}}$}}}\mkern 1.5mu] such that S1TXS11=T~XS_{1}T\restriction_{\vec{X}}S_{1}^{-1}=\tilde{T}\restriction_{\vec{X}} and S2TXS21=T~1XS_{2}T\restriction_{\reflectbox{$\vec{\reflectbox{$X$}}$}}S_{2}^{-1}=\tilde{T}^{-1}\restriction_{\reflectbox{$\vec{\reflectbox{$X$}}$}}. The transformation SS given by

Sx={S1xif xX,S2xif xX,xotherwiseSx=\begin{cases}S_{1}x&\textrm{if $x\in\vec{X}$},\\ S_{2}x&\textrm{if $x\in\reflectbox{$\vec{\reflectbox{$X$}}$}$},\\ x&\textrm{otherwise}\\ \end{cases}

belongs to the full group [T][T\mkern 1.5mu] and witnesses flip conjugacy of TT and T~\tilde{T}. ∎

The transformation conjugating TT and UU in Theorem 10.5 can be written fairly explicitly. This is done in terms of the function δ\delta defined as follows. Suppose (Ω,λ)(\Omega,\lambda) is a (possibly infinite) measure space, and let A,BΩA,B\subseteq\Omega be measurable sets such that λ(AB)<+\lambda(A\bigtriangleup B)<+\infty. We set δ(A,B)=λ(AB)λ(BA)\delta(A,B)=\lambda(A\setminus B)-\lambda(B\setminus A). This function satisfies a few properties which the reader can easily verify.

Proposition 10.8.

Suppose (Ω,λ)(\Omega,\lambda) is a measure space. For all A,B,C,a,ΩA,B,C,a,\subseteq\Omega such that λ(AB),λ(BC),λ(AC),λ(a)<+\lambda(A\bigtriangleup B),\lambda(B\bigtriangleup C),\lambda(A\bigtriangleup C),\lambda(a)<+\infty, the following holds:

  1. (1)

    δ(A,C)=δ(A,B)+δ(B,C)\delta(A,C)=\delta(A,B)+\delta(B,C);

  2. (2)

    δ(A,A)=0\delta(A,A)=0 and δ(A,B)=δ(B,A)\delta(A,B)=-\delta(B,A);

  3. (3)

    δ(Aa,B)=δ(A,B)+(λ(a)2λ(aA))\delta(A\bigtriangleup a,B)=\delta(A,B)+(\lambda(a)-2\lambda(a\cap A)).

Any orbit of a measure-preserving transformation can be endowed with a counting measure. Given TT and UU as in the statement of Theorem 10.5, set τ(x)=δ(Ux,Tx)\tau(x)=\delta(U^{\mathbb{N}}x,T^{\mathbb{N}}x) and define Sx=Uτ(x)xSx=U^{\tau(x)}x. One can verify that S[U]=[T]S\in[U\mkern 1.5mu]=[T\mkern 1.5mu] and STS1=USTS^{-1}=U (further details can be found in [CJMT22, Theorem A.1]).

Let now 1\mathcal{F}_{1} and 2\mathcal{F}_{2} be measure-preserving flows on a standard probability space (X,μ)(X,\mu); we denote the actions of rr\in\mathbb{R} upon xXx\in X by x+1rx+_{1}r and x+2rx+_{2}r, respectively. Suppose that their full groups coincide, [1]=[2][\mathcal{F}_{1}\mkern 1.5mu]=[\mathcal{F}_{2}\mkern 1.5mu], and so the flows share the same orbits, 1=2\mathcal{R}_{\mathcal{F}_{1}}=\mathcal{R}_{\mathcal{F}_{2}}. For xXx\in X, let si(x)=x+i[0,)s_{i}(x)=x+_{i}[0,\infty), i=1,2i=1,2, denote the “right half-orbit” of xx. A natural analog of the condition |TxUx|<|T^{\mathbb{N}}x\bigtriangleup U^{\mathbb{N}}x|<\infty from Theorem 10.5 would be to require finiteness of the Lebesgue measure of s1(x)s2(x)s_{1}(x)\bigtriangleup s_{2}(x) for all xXx\in X. This condition alone, however, is not sufficient for conjugacy of 1\mathcal{F}_{1} and 2\mathcal{F}_{2}.

Each flow induces a copy of the Lebesgue measure onto orbits via

λi,x(A)=λ({r:x+irA}).\lambda_{i,x}(A)=\lambda(\{r\in\mathbb{R}:x+_{i}r\in A\}).

Since we assume [1]=[2][\mathcal{F}_{1}\mkern 1.5mu]=[\mathcal{F}_{2}\mkern 1.5mu], and so 2[1]\mathcal{F}_{2}\subseteq[\mathcal{F}_{1}\mkern 1.5mu], λ1,x\lambda_{1,x} is a translation invariant measure relative to the action of 2\mathcal{F}_{2}, and therefore must differ from λ2,x\lambda_{2,x} by a constant: there is an orbit invariant measurable function c:X>0c:X\to\mathbb{R}^{>0} such that λ2,x=c(x)λ1,x\lambda_{2,x}=c(x)\lambda_{1,x}. Any element in [1]=[2][\mathcal{F}_{1}\mkern 1.5mu]=[\mathcal{F}_{2}\mkern 1.5mu] preserves λi,x\lambda_{i,x}, i=1,2i=1,2, and therefore cannot conjugate 1\mathcal{F}_{1} into 2\mathcal{F}_{2} unless c(x)c(x) is constantly equal to 11.

When the flows are ergodic, c(x)=cc(x)=c is a constant, and one may renormalize the flows without changing the full groups. Let 2\mathcal{F}_{2}^{\prime} be the rescaling of 2\mathcal{F}_{2} given by x+2r=x+2crx+_{2}^{\prime}r=x+_{2}cr. It is straightforward to check that λ2,x(A)=c1λ2,x(A)=λ1,x(A)\lambda_{2,x}^{\prime}(A)=c^{-1}\lambda_{2,x}(A)=\lambda_{1,x}(A) and flows 1\mathcal{F}_{1} and 2\mathcal{F}_{2}^{\prime} induce the same measure onto orbits.

For flows that do induce the same measures on the orbits, finiteness of the measure s1(x)s2(x)s_{1}(x)\bigtriangleup s_{2}(x) for all xXx\in X is indeed sufficient to conclude conjugacy of the flows.

Theorem 10.9.

Let i\mathcal{F}_{i}, i=1,2i=1,2, be free measure-preserving flows that share the same orbits, 1=2\mathcal{R}_{\mathcal{F}_{1}}=\mathcal{R}_{\mathcal{F}_{2}}, and induce the same measures (λx)xX(\lambda_{x})_{x\in X} onto orbit. If λx(s1(x)s2(x))<+\lambda_{x}(s_{1}(x)\bigtriangleup s_{2}(x))<+\infty for xXx\in X, then the flows are conjugate by a measure-preserving transformation S[1]S\in[\mathcal{F}_{1}\mkern 1.5mu].

Proof.

Let n:X×n:X\times\mathbb{R}\to\mathbb{R} be the 1,2\mathcal{F}_{1},\mathcal{F}_{2}-cocycle defined by x+2r=x+1n(x,r)x+_{2}r=x+_{1}n(x,r). Since 1\mathcal{F}_{1} and 2\mathcal{F}_{2} induce the same measure on the orbits, n(x,):n(x,\cdot):\mathbb{R}\to\mathbb{R} is a Lebesgue measure-preserving automorphism:

λ(n(x,A))\displaystyle\lambda(n(x,A)) =λ1,x({x+1n(x,r):rA})\displaystyle=\lambda_{1,x}(\{x+_{1}n(x,r):r\in A\})
=λ2,x({x+2r:rA})=λ(A).\displaystyle=\lambda_{2,x}(\{x+_{2}r:r\in A\})=\lambda(A).

For xXx\in X and r{+}r\in\mathbb{R}\cup\{+\infty\} let

si,r(x)={x+i[0,r)if r0,x+i[r,0)if r<0.s_{i,r}(x)=\begin{cases}x+_{i}[0,r)&\textrm{if }r\geq 0,\\ x+_{i}[r,0)&\textrm{if }r<0.\end{cases}

In particular, si(x)=si,+(x)s_{i}(x)=s_{i,+\infty}(x). Note that

(10.2) s1(x+2r)=s1(x)s1,n(x,r)(x),s2(x+2r)=s2(x)s2,r(x).\begin{split}s_{1}(x+_{2}r)&=s_{1}(x)\bigtriangleup s_{1,n(x,r)}(x),\\ s_{2}(x+_{2}r)&=s_{2}(x)\bigtriangleup s_{2,r}(x).\\ \end{split}

Also, considering the cases r<0r<0 and r0r\geq 0 separately, one can easily verify that for all rr\in\mathbb{R} and i=1,2i=1,2

λi,x(si,r(x))2λi,x(s2(x)s2,r(x))=r.\lambda_{i,x}(s_{i,r}(x))-2\lambda_{i,x}(s_{2}(x)\cap s_{2,r}(x))=-r.

and, in particular,

(10.3) λ1,x(s1,n(x,r)(x))2λ1,x(s1(x)s1,n(x,r)(x))=n(x,r),λ2,x(s2,r(x))2λ2,x(s2(x)s2,r(x))=r.\begin{split}\lambda_{1,x}(s_{1,n(x,r)}(x))-2\lambda_{1,x}(s_{1}(x)\cap s_{1,n(x,r)}(x))&=-n(x,r),\\ \lambda_{2,x}(s_{2,r}(x))-2\lambda_{2,x}(s_{2}(x)\cap s_{2,r}(x))&=-r.\end{split}

Put τ(x)=δ(s1(x),s2(x))\tau(x)=\delta(s_{1}(x),s_{2}(x)), then

(10.4) τ(x+2r)=δ(s1(x+2r),s2(x+2r))Eq.(10.2)=δ(s1(x)s1,n(x,r)(x),s2(x+2r))Prop.10.8=δ(s1(x),s2(x+2r))+λ1,x(s1,n(x,r)(x))2λ1,x(s1(x)s1,n(x,r)(x))Eq.(10.3)=δ(s1(x),s2(x+2r))n(x,r)Prop.10.8=δ(s2(x+2r),s1(x))n(x,r)=δ(s1(x),s2(x+2r))(λ2,x(s2,r(x))2λ2,x(s2(x)s2,r(x)))n(x,r)Eq.(10.3)=δ(s1(x),s2(x))n(x,r)+r.\begin{split}\tau(x+_{2}r)&=\delta(s_{1}(x+_{2}r),s_{2}(x+_{2}r))\\ \because\textrm{Eq.}~{}\eqref{eq:shift-half-orbit}&=\delta(s_{1}(x)\bigtriangleup s_{1,n(x,r)}(x),s_{2}(x+_{2}r))\\ \because\textrm{Prop.}~{}\ref{prop:nabla-properties}&=\delta(s_{1}(x),s_{2}(x+_{2}r))+\\ &\qquad\lambda_{1,x}(s_{1,n(x,r)}(x))-2\lambda_{1,x}(s_{1}(x)\cap s_{1,n(x,r)}(x))\\ \because\textrm{Eq.}~{}\eqref{eq:cocycle-correction}&=\delta(s_{1}(x),s_{2}(x+_{2}r))-n(x,r)\\ \because\textrm{Prop.}~{}\ref{prop:nabla-properties}&=-\delta(s_{2}(x+_{2}r),s_{1}(x))-n(x,r)=\delta(s_{1}(x),s_{2}(x+_{2}r))-\\ &\qquad(\lambda_{2,x}(s_{2,r}(x))-2\lambda_{2,x}(s_{2}(x)\cap s_{2,r}(x)))-n(x,r)\\ \because\textrm{Eq.}~{}\eqref{eq:cocycle-correction}&=\delta(s_{1}(x),s_{2}(x))-n(x,r)+r.\\ \end{split}

The required transformation S:XXS:X\to X is given by Sx=x+1τ(x)Sx=x+_{1}\tau(x).

S(x+2r)\displaystyle S(x+_{2}r) =(x+2r)+1τ(x+2r)=(x+1n(x,r))+1τ(x+2r)\displaystyle=(x+_{2}r)+_{1}\tau(x+_{2}r)=(x+_{1}n(x,r))+_{1}\tau(x+_{2}r)
Eq. (10.4)\displaystyle\because\textrm{Eq.~{}\eqref{eq:tau-shift}} =x+1(n(x,r)+τ(x)n(x,r)+r)=Sx+1r.\displaystyle=x+_{1}(n(x,r)+\tau(x)-n(x,r)+r)=Sx+_{1}r.

Thus SS conjugates 1\mathcal{F}_{1} and 2\mathcal{F}_{2}. It therefore remains to check that SS is a measure-preserving bijection. First, note that SxSx satisfies δ(s1(Sx),s2(x))=0\delta(s_{1}(Sx),s_{2}(x))=0. Indeed, s1(Sx)=s1(x)s1,τ(x)(x)s_{1}(Sx)=s_{1}(x)\bigtriangleup s_{1,\tau(x)}(x) (by the analog of Eq. (10.2)), and therefore

(10.5) δ(s1(Sx),s2(x))=τ(x)τ(x)=0\delta(s_{1}(Sx),s_{2}(x))=\tau(x)-\tau(x)=0

by Proposition 10.8.

To show injectivity, suppose that Sx=SySx=Sy. In view of Eq. (10.5) and Proposition 10.8,

δ(s2(x),s2(y))=δ(s2(x),s1(Sx))+δ(s1(Sy),s2(y))=0.\delta(s_{2}(x),s_{2}(y))=\delta(s_{2}(x),s_{1}(Sx))+\delta(s_{1}(Sy),s_{2}(y))=0.

However, if y=x+2ry=x+_{2}r, then s2(y)=s2(x)s2,r(x)s_{2}(y)=s_{2}(x)\bigtriangleup s_{2,r}(x) and so δ(s2(x),s2(y))=r\delta(s_{2}(x),s_{2}(y))=r. One concludes that r=0r=0 and x=yx=y. We have already established that S(x+2r)=Sx+1rS(x+_{2}r)=Sx+_{1}r, which shows that the range of SS is orbit invariant, yielding surjectivity.

Finally, to show that SS is measure-preserving, it suffices to check that SS preserves the Lebesgue measure λ1,x=λ2,x\lambda_{1,x}=\lambda_{2,x} on all the orbits. To this end, let n:X×n^{\prime}:X\times\mathbb{R}\to\mathbb{R} be the 1\mathcal{F}_{1}-cocycle (i.e., x+1r=x+2n(x,r)x+_{1}r=x+_{2}n^{\prime}(x,r)). For all rr^{\prime}\in\mathbb{R}, one has

λ1,x(Ss1,r(x))\displaystyle\lambda_{1,x}(Ss_{1,r^{\prime}}(x)) =λ1,x({y+1τ(y):ys1,r(x)})\displaystyle=\lambda_{1,x}(\{y+_{1}\tau(y):y\in s_{1,r^{\prime}}(x)\})
=λ1,x({(x+1r)+1τ(x+1r):0r<r})\displaystyle=\lambda_{1,x}(\{(x+_{1}r)+_{1}\tau(x+_{1}r):0\leq r<r^{\prime}\})
=λ({r+(τ(x)r+n(x,r)):0r<r})\displaystyle=\lambda(\{r+(\tau(x)-r+n^{\prime}(x,r)):0\leq r<r^{\prime}\})
=λ({n(x,r):0r<r})=λ(n(x,[0,r)))=r,\displaystyle=\lambda(\{n^{\prime}(x,r):0\leq r<r^{\prime}\})=\lambda(n^{\prime}(x,[0,r^{\prime})))=r^{\prime},

hence SAut(X,μ)S\in\mathrm{Aut}(X,\mu) is the required conjugation between 1\mathcal{F}_{1} and 2\mathcal{F}_{2}. ∎

In the \mathbb{Z} case, the above result is the key to Belinskaja’s flip conjugacy result for L1\mathrm{L}^{1} orbit equivalence. Unfortunately, here we don’t know if it can be useful towards proving an analogous result. In the next section, we nevertheless obtain a weaker result which yields that there are many L1\mathrm{L}^{1} full groups. We leave the following question open.

Question 10.10.

Given two ergodic flows with equal L1\mathrm{L}^{1} full groups, do they have to satisfy the hypothesis of the above theorem after appropriate rescaling?

10.3. L1\mathrm{L}^{1} orbit equivalence implies flip Kakutani equivalence

A measure-preserving action of a compactly generated locally compact Polish group can always be twisted by a continuous automorphism of the group without affecting the L1\mathrm{L}^{1} full group.

In the case of \mathbb{Z}-actions, this takes a particularly simple form, since the only non-trivial automorphism of \mathbb{Z} is given by nnn\mapsto-n. It follows from the results of R. M. Belinskaja [Bel68] that this is up to conjugacy the only way to get an L1\mathrm{L}^{1} orbit equivalence for ergodic \mathbb{Z}-actions [LM18, Theorem 4.2]: if T1,T2T_{1},T_{2} are two ergodic measure-preserving transformations which are L1\mathrm{L}^{1} orbit equivalent, then they are flip-conjugate: T1T_{1} is conjugate to either T2T_{2} or T21T_{2}^{-1}.

As mentioned before, we do not know whether a variant of such rigidity holds when we replace \mathbb{Z} by \mathbb{R} (see Question 10.17 below), but, as shown in Theorem 10.15, L1\mathrm{L}^{1} orbit equivalent free measure-preserving flows must at least be flip Kakutani equivalent. In particular, there are uncountably many L1\mathrm{L}^{1} full groups of free ergodic flows up to abstract group isomorphism.

Let us first define the notion of (flip) Kakutani equivalence of flows. For the main results about this concept, the reader may consults [Kat75, Kat77], where it is called monotone equivalence of flows. Given a measure-preserving automorphism TAut(Z,ν)T\in\mathrm{Aut}(Z,\nu) and a positive integrable function fL1(Z,ν)f\in\mathrm{L}^{1}(Z,\nu), one can define the so-called suspension flow or flow under a function of TT 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. For r0r\geq 0, the action (z,t)+r(z,t)+r is given by

(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{)}

where k0k\geq 0 is defined uniquely by the condition 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); similarly for r0r\leq 0 the action is

(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{)},

where k0k\geq 0 satisfies 0t+r+i=1kf(Tiz)<f(Tkz)0\leq t+r+\sum\limits_{i=1}^{k}f(T^{-i}z)<f(T^{-k}z). Such a flow preserves the restriction onto XX of the product measure ν×λ\nu\times\lambda. The space

(X,μ), where μ=ν×λZfdνX(X,\mu),\text{ where }\mu=\dfrac{\nu\times\lambda}{\int_{Z}f\,d\nu}\restriction_{X}

is a standard probability space. The automorphism TT in the suspension flow construction is called the base automorphism.

Definition 10.11.

Two flows are (flip) Kakutani equivalent if they are isomorphic to suspension flows over flip conjugate base automorphisms.

It is important to note that the construction of suspension flows can be reversed through the use of cross-sections111In full generality, the definition of a cross-section should actually be relaxed, replacing lacunarity by discreteness in each orbit, and only requiring the gap function of the cross-section to be integrable.. If we have a fixed free flow on (X,μ)(X,\mu) and 𝒞X\mathcal{C}\subseteq X is a co-compact cross-section which is UU-lacunary where UU is precompact, then there is a unique probability measure ν\nu on 𝒞\mathcal{C} such that the map U×𝒞𝒞+UXU\times\mathcal{C}\to\mathcal{C}+U\subseteq X taking (t,c)(t,c) to c+tc+t is measure-preserving (see e.g. [KPV15, Prop. 4.3] for the general construction). It is then clear that the first-return map σ𝒞:𝒞𝒞\sigma_{\mathcal{C}}:\mathcal{C}\to\mathcal{C} is measure-preserving, and our initial flow can be seen as the flow built under the gap function gap𝒞\mathrm{gap}_{\mathcal{C}} with base transformation σ𝒞\sigma_{\mathcal{C}}.

We need the following important result, which is due to D. Rudolph [Rud76]. Keeping in mind the previous paragraph, it can be reformulated as the fact that every measure-preserving flow is conjugate to a suspension flow over a two-valued function.

Theorem 10.12 (Rudolph).

Let \mathcal{F} be a free measure-preserving flow on a standard probability space (X,μ)(X,\mu), let t0t_{0}\in\mathbb{R}\setminus\mathbb{Q}, then \mathcal{F} admits a cross-section whose gap function only takes the values 11 and t0t_{0} almost surely.

Remark 10.13.

The second named author has obtained a generalization of this to the purely Borel context, see [Slu19].

Theorem 10.14.

Let ,\mathcal{F},\mathcal{F}^{\prime} be free measure-preserving flows on (X,μ)(X,\mu) that share the same orbits, =\mathcal{R}_{\mathcal{F}}=\mathcal{R}_{\mathcal{F}^{\prime}}. If []1\mathcal{F^{\prime}}\leq[\mathcal{F}\mkern 1.5mu]_{1}, then \mathcal{F} and \mathcal{F}^{\prime} are flip Kakutani equivalent.

Proof.

We denote the flow \mathcal{F} using our usual notation, (x,t)x+t(x,t)\mapsto x+t. As explained right after Definition 10.11, it suffices to find cross-sections for \mathcal{F} and \mathcal{F}^{\prime} such that the corresponding first return automorphisms are flip conjugated.

Pick Borel realizations of the flows and let 𝒞X\mathcal{C}\subseteq X be a Borel cross-section for \mathcal{F} such that gap𝒞(c){1,t0}\mathrm{gap}_{\mathcal{C}}(c)\in\{1,t_{0}\} for all c𝒞c\in\mathcal{C}, as provided by Theorem 10.12. Define the automorphism T:XXT:X\to X by

Tx={σ𝒞(c)+αif x=c+α for some c𝒞α[0,1],xotherwise.Tx=\begin{cases}\sigma_{\mathcal{C}}(c)+\alpha&\textrm{if $x=c+\alpha$ for some $c\in\mathcal{C}$, $\alpha\in[0,1]$},\\ x&\textrm{otherwise}.\end{cases}

The transformation TT is obtained by gluing together the identity map, xx+1x\mapsto x+1 and xx+t0x\mapsto x+t_{0}, and since all these belong to []1[\mathcal{F}]_{1}, which is finitely full, we have that T[]1T\in[\mathcal{F}]_{1} as well. Note that TT is dissipative and is therefore flip conjugated to its \mathcal{F}^{\prime}-reordering T~\tilde{T} by Proposition 10.7. In other words, there is a TT-invariant Borel set ZXZ\subseteq X of full measure, μ(Z)=1\mu(Z)=1, and a TT-invariant Borel partition Z=Z+ZZ=Z^{+}\sqcup Z^{-} such that TZ+T\restriction_{Z^{+}} is conjugated to T~Z+\tilde{T}\restriction_{Z^{+}} and TZT\restriction_{Z^{-}} is conjugated to T~1Z\tilde{T}^{-1}\restriction_{Z^{-}}.

Let ν\nu be the measure on 𝒞\mathcal{C} given for a Borel A𝒞A\subseteq\mathcal{C} by ν(A)=μ(A+[0,1))\nu(A)=\mu(A+[0,1)). The measure μ𝒞+[0,1)\mu\restriction_{\mathcal{C}+[0,1)} is naturally isomorphic to (ν×λ)𝒞+[0,1)(\nu\times\lambda)\restriction_{\mathcal{C}+[0,1)}, where λ\lambda is the Lebesgue measure on [0,1][0,1], and therefore we have

ν×λ(c,λ)𝒞×[0,1)c+αZ.\forall^{\nu\times\lambda}(c,\lambda)\in\mathcal{C}\times[0,1)\quad c+\alpha\in Z.

By Fubini’s theorem, this is equivalent to

λα[0,1)νc𝒞(c+αZ).\forall^{\lambda}\alpha\in[0,1)\ \forall^{\nu}c\in\mathcal{C}\quad(c+\alpha\in Z).

Therefore there exists some α0[0,1)\alpha_{0}\in[0,1) such that ν({c𝒞:c+α0Z})=1\nu(\{c\in\mathcal{C}:c+\alpha_{0}\in Z\})=1. Note that T𝒞+α0T\restriction_{\mathcal{C}+\alpha_{0}} is the first return map on 𝒞+α0\mathcal{C}+\alpha_{0} in the order of the flow \mathcal{F}, whereas T~𝒞+α0\tilde{T}\restriction_{\mathcal{C}+\alpha_{0}} is the first return map in the order induced on the orbits by \mathcal{F}^{\prime}. Since T𝒞+α0T\restriction_{\mathcal{C}+\alpha_{0}} and T~𝒞+α0\tilde{T}\restriction_{\mathcal{C}+\alpha_{0}} are flip conjugated, the flows are flip Kakutani equivalent. ∎

Theorem 10.14 has the following straightforward consequences.

Corollary 10.15.

If two free ergodic measure-preserving flows are L1\mathrm{L}^{1} orbit equivalent, then they are also flip Kakutani equivalent.

Proof.

This now follows from the definition of L1\mathrm{L}^{1} orbit equivalence, see Definition 4.19 and the paragraph thereafter. ∎

Corollary 10.16.

If two free ergodic measure-preserving flows have abstractly isomorphic L1\mathrm{L}^{1} full groups, then they are also flip Kakutani equivalent.

Proof.

We have seen in Proposition 4.21 that isomorphism of L1\mathrm{L}^{1} full groups of ergodic flows implies L1\mathrm{L}^{1} orbit equivalence, so the result follows from the previous corollary. ∎

Kakutani equivalence is a highly non-trivial equivalence relation (see, for instance, [ORW82] or [GK21, Kun23]). It seems likely, however, that L1\mathrm{L}^{1} full groups of flows contain even more information about the action. The only continuous automorphisms of \mathbb{R} are multiplications by nonzero scalars, and we ask whether isomorphism of L1\mathrm{L}^{1} full groups necessarily recovers the action up to such an automorphism.

Question 10.17.

Let 1\mathcal{F}_{1} and 2\mathcal{F}_{2} be free ergodic measure-preserving flows with isomorphic L1\mathrm{L}^{1} full groups. Is it true that there is α\alpha\in\mathbb{R}^{*} such that 1\mathcal{F}_{1} and 2mα\mathcal{F}_{2}\circ m_{\alpha} are isomorphic, where mαm_{\alpha} denotes the multiplication by α\alpha?

Note that a positive answer to Question 10.10 would imply a positive answer to the above question.

10.4. Maximality of the L1\mathrm{L}^{1} norm and geometry

In this last section, we show that the L1\mathrm{L}^{1} norm is maximal on L1\mathrm{L}^{1} full groups of flows. In particular, it defines their quasi-isometry type. Exploring this quasi-isometry type further thus might lead to topological group invariants which distinguish some ergodic flows.

Theorem 10.18.

Let \mathcal{F} be a free measure-preserving flow. The L1\mathrm{L}^{1} norm on []1[\mathcal{F}]_{1} is maximal.

Proof.

We have already shown that the L1\mathrm{L}^{1} norm on the derived L1\mathrm{L}^{1} full group is maximal (see Theorem 5.5). Denote by (,p)(\mathcal{E},p) the space of \mathcal{F}-invariant ergodic probability measures, where pp is the probability measure arising from the disintegration of μ\mu which we write as xνxx\mapsto\nu_{x} (see Section C.3). The derived L1\mathrm{L}^{1} full group is equal to the kernel of the surjective index map :[]1]L1(,)\mathcal{I}:[\mathcal{F}\mkern 1.5mu]_{1}]\to\mathrm{L}^{1}(\mathcal{E},\mathbb{R}) and the quotient norm on []1/kerI[\mathcal{F}]_{1}/\ker I is equal to the L1\mathrm{L}^{1} norm on L1(,p)\mathrm{L}^{1}(\mathcal{E},p) by Proposition 6.7. The latter norm is maximal, as is any Banach space norm.

Given a function fL1(,p)f\in\mathrm{L}^{1}(\mathcal{E},p), let Uf[]1U_{f}\in[\mathcal{F}\mkern 1.5mu]_{1} be given by Uf(x)=x+f(νx)U_{f}(x)=x+f(\nu_{x}). The cocycle ρUf(x)=f(νx)\rho_{U_{f}}(x)=f(\nu_{x}) is constant on each ergodic component and Uf1=f1\lVert U_{f}\rVert_{1}=\lVert f\rVert_{1}. Furthermore, (Uf)=f\mathcal{I}(U_{f})=f. We show that \lVert\,\cdot\,\rVert is both large-scale geodesic and coarsely proper (see Appendix A.2 and Proposition A.10, in particular).

Any T[]1T\in[\mathcal{F}\mkern 1.5mu]_{1} can be written as T=(TU(T)1)U(T)T=(TU_{\mathcal{I}(T)}^{-1})U_{\mathcal{I}(T)}, where the transformation TU(T)1ker=D([]1)TU_{\mathcal{I}(T)}^{-1}\in\ker\mathcal{I}=D([\mathcal{F}\mkern 1.5mu]_{1}), and U(T)1T1\lVert U_{\mathcal{I}(T)}\rVert_{1}\leq\lVert T\rVert_{1}. In particular, we have TU(T)112T1\lVert TU_{\mathcal{I}(T)}^{-1}\rVert_{1}\leq 2\lVert T\rVert_{1}.

Since the L1\mathrm{L}^{1} norm is maximal on D([]1)D([\mathcal{F}\mkern 1.5mu]_{1}), it is large-scale geodesic. In fact, Proposition 3.24 establishes that it is large-scale geodesic with constant K=2K=2. We may therefore express TU(T)1TU_{\mathcal{I}(T)}^{-1} as a product V1VnV_{1}\cdots V_{n} of elements ViD([]1)V_{i}\in D([\mathcal{F}\mkern 1.5mu]_{1}), where each ViV_{i} has norm at most KK and

i=1nVi1KTU(T)112KT1.\sum_{i=1}^{n}\lVert V_{i}\rVert_{1}\leq K\lVert TU_{\mathcal{I}(T)}^{-1}\rVert_{1}\leq 2K\lVert T\rVert_{1}.

The transformation U(T)U_{\mathcal{I}(T)} can, for any m1m\geq 1, also be expressed as a product

U(T)=U(T)/mU(T)/m=U(T)/mm.U_{\mathcal{I}(T)}=U_{\mathcal{I}(T)/m}\cdots U_{\mathcal{I}(T)/m}=U_{\mathcal{I}(T)/m}^{m}.

Taking mm sufficiently large, we can ensure that U(T)/m1=(T)/m1K\lVert U_{\mathcal{I}(T)/m}\rVert_{1}=\lVert\mathcal{I}(T)/m\rVert_{1}\leq K. Therefore, T=(V1Vn)(U(T)/mU(T)/m)T=(V_{1}\cdots V_{n})(U_{\mathcal{I}(T)/m}\cdots U_{\mathcal{I}(T)/m}), and

i=1nVi1+j=1mU(T)/m12KT1+U(T)13KT1.\sum_{i=1}^{n}\lVert V_{i}\rVert_{1}+\sum_{j=1}^{m}\lVert U_{\mathcal{I}(T)/m}\rVert_{1}\leq 2K\lVert T\rVert_{1}+\lVert U_{\mathcal{I}(T)}\rVert_{1}\leq 3K\lVert T\rVert_{1}.

We conclude that the norm \lVert\,\cdot\,\rVert on []1[\mathcal{F}\mkern 1.5mu]_{1} is large-scale geodesic with K=3K=6K^{\prime}=3K=6.

It remains to prove coarse properness. Let ϵ>0\epsilon>0 and R>0R>0 be positive reals. By Theorem 5.5, there is nn\in\mathbb{N} so large that every element in the derived L1\mathrm{L}^{1} full group of norm at most 2R2R is a product of nn elements of norm at most ϵ\epsilon. Let NN be any integer greater than R/ϵR/\epsilon. We argue that every element of []1[\mathcal{F}\mkern 1.5mu]_{1} of norm at most RR is a product of 2n+N2n+N elements of norm at most ϵ\epsilon.

Indeed, if T=(TU(T)1)U(T)T=(TU_{\mathcal{I}(T)}^{-1})U_{\mathcal{I}(T)} has norm at most RR, then

TU(T)112T12R,\lVert TU_{\mathcal{I}(T)}^{-1}\rVert_{1}\leq 2\left\lVert T\right\rVert_{1}\leq 2R,

and TU(T)1TU_{\mathcal{I}(T)}^{-1} can therefore by written as a product of nn elements of D([F]1)D([F\mkern 1.5mu]_{1}) each of norm ϵ\leq\epsilon. Also, U(T)=U(T)/NNU_{\mathcal{I}(T)}=U_{\mathcal{I}(T)/N}^{N} and U(T)/N1ϵ\lVert U_{\mathcal{I}(T)/N}\rVert_{1}\leq\epsilon by the choice of NN. The conclusion follows. ∎

Remark 10.19.

While the proposition above states that L1\mathrm{L}^{1} full groups of flows are quite big, one can use Proposition 6.8 to show that they satisfy the Haagerup property. In other words, such groups admit a coarsely proper affine action on a Hilbert space (namely, the affine Hilbert space χ0+L2(,M)\chi_{\mathcal{R}^{\geq 0}}+\mathrm{L}^{2}(\mathcal{R},M)).

Corollary 10.16 along with [ORW82, Sec. 12] implies that there are uncountably many L1\mathrm{L}^{1} full groups of ergodic free flows up to topological group isomorphism. It would be great if their geometry allowed us to distinguish these groups. However, we don’t even know the answer to the following question.

Question 10.20.

Are there two free ergodic measure-preserving flows with non quasi-isometric L1\mathrm{L}^{1} full groups?

Appendix A Normed groups

We chose to present our work in the framework of groups equipped with compatible norms rather than metrics. These two frameworks are equivalent, but the former has some stylistic advantages, in our opinion. In Appendix A, we remind the reader the concept of a norm on a group (Section A.1) and state C. Rosendal’s results on maximal norms (Section A.2).

A.1. Norms on groups

Definition A.1.

A norm on a group GG is a map :G0\lVert\cdot\rVert:G\to\mathbb{R}^{\geq 0} such that for all g,hGg,h\in G

  1. (1)

    g=0\lVert g\rVert=0 if and only if g=eg=e;

  2. (2)

    g=g1\lVert g\rVert=\lVert g^{-1}\rVert;

  3. (3)

    ghg+h\lVert gh\rVert\leq\lVert g\rVert+\lVert h\rVert.

If GG is moreover a topological group, a norm \left\lVert\cdot\right\rVert on GG is called compatible if the balls {gG:||g||<r}\{g\in G:||g||<r\}, r>0r>0, form a basis of neighborhoods of the identity.

There is a correspondence between (compatible) left-invariant metrics on a group and (compatible) norms on it. Indeed, given a left-invariant metric dd on GG, the function g=d(e,g)\lVert g\rVert=d(e,g) is a norm. Conversely, from a norm \lVert\cdot\rVert one can recover the left-invariant metric dd via d(g,h)=g1hd(g,h)=\lVert g^{-1}h\rVert. Analogously, there is a correspondence between norms and right invariant metrics given by d(g,h)=hg1d(g,h)=\lVert hg^{-1}\rVert.

The language of group norms thus contains the same information as the formalism of left-invariant (or right-invariant) metrics, but it has the stylistic advantage of removing the need of making a choice between the invariant side, when such a choice is immaterial.

Remark A.2.

Note, however, that there are metrics that are neither left- nor right-invariant, which nonetheless induce a group norm via the same formula g=d(g,e)\lVert g\rVert=d(g,e). Consider for example a Polish group GG with a compatible left-invariant metric dd^{\prime} on it. If GG is not a CLI group, the metric dd^{\prime} is not complete, but the metric

d(f,g)=d(f,g)+d(f1,g1)2d(f,g)=\frac{d^{\prime}(f,g)+d^{\prime}(f^{-1},g^{-1})}{2}

is complete. Since d(g,e)=d(g,e)d(g,e)=d^{\prime}(g,e), we see that dd induces the same norm \lVert\cdot\rVert as does the left-invariant metric dd^{\prime}.

There is a canonical way to push a norm onto a factor group.

Proposition A.3 (see [Fre04, Thm. 2.2.10]).

Let (G,)(G,\lVert\cdot\rVert) be a Polish normed group, and let HGH\trianglelefteq G be a closed normal subgroup of GG. The function

gHG/H=inf{gh:hH}\lVert gH\rVert^{G/H}=\inf\{\lVert gh\rVert:h\in H\}

is a norm on G/HG/H which is compatible with the quotient topology. In particular, (G/H,G/H)(G/H,\lVert\cdot\rVert^{G/H}) is a Polish normed group.

Definition A.4.

A compatible norm \lVert\cdot\rVert on a locally compact Polish group GG is proper if all balls {gG:gr}\{g\in G:\lVert g\rVert\leq r\} are compact.

R. A. Struble [Str74] showed that all locally compact Polish groups admit a compatible proper norm.

A.2. Maximal norms

As was noted in Lemma 2.13, quasi-isometric norms yield the same L1\mathrm{L}^{1} full groups. C. Rosendal identified the class of Polish groups that admit maximal norms, which are unique up to quasi-isometry. In this section, we state some of results from C. Rosendal’s treatise [Ros22], which are relevant to our work. For reader’s convenience, we formulate the following definitions and propositions in the language of group norms as opposed to left-invariant metrics or écartes, as in the original reference.

Definition A.5 ([Ros22, Def. 2.68]).

A compatible norm \left\lVert\cdot\right\rVert on a Polish group GG is said to be maximal if for any compatible norm \left\lVert\cdot\right\rVert^{\prime} there is a constant C>0C>0 such that gCg+C\lVert g\rVert^{\prime}\leq C\lVert g\rVert+C for all gGg\in G.

Definition A.6 ([Ros22, Prop. 2.15]).

Let GG be a Polish group. A subset AGA\subseteq G is coarsely bounded if for every continuous isometric action of GG on a metric space (M,dM)(M,d_{M}), the set AmA\cdot m is bounded for all mMm\in M, i.e., there is K>0K>0 such that dM(a1m,a2m)Kd_{M}(a_{1}\cdot m,a_{2}\cdot m)\leq K for all a1,a2Aa_{1},a_{2}\in A. A Polish group GG is boundedly generated if it is generated by a coarsely bounded set.

Theorem A.7 ([Ros22, Thm. 2.73]).

A Polish group admits a maximal compatible norm if and only if it is boundedly generated.

The following characterization is available to establish maximality of a given norm.

Definition A.8 ([Ros22, Def. 2.62]).

A norm \left\lVert\cdot\right\rVert on a group GG is called large-scale geodesic if there is K>0K>0 such that for any gGg\in G, there are g1,,gnGg_{1},...,g_{n}\in G of norm giK\lVert g_{i}\rVert\leq K, 1in1\leq i\leq n, such that g=g1gng=g_{1}\cdots g_{n} and

i=1ngiKg.\sum_{i=1}^{n}\left\lVert g_{i}\right\rVert\leq K\left\lVert g\right\rVert.
Definition A.9 ([Ros22, Lem. 2.39(2) and Prop. 2.7(5)]).

A compatible norm \lVert\cdot\rVert on a topological group GG is called coarsely proper if for every ϵ>0\epsilon>0 and every R>0R>0, there are a finite subset FGF\subseteq G and nn\in\mathbb{N} such that every element gGg\in G of norm at most RR can be written as a product

g=f1g1fngn,g=f_{1}g_{1}\cdots f_{n}g_{n},

where f1,,fnFf_{1},\dots,f_{n}\in F and each gig_{i} has norm at most ϵ\epsilon.

Proposition A.10 ([Ros22, Prop. 2.72]).

A compatible norm \left\lVert\cdot\right\rVert on a Polish group GG is maximal if and only if it is both large-scale geodesic and coarsely proper.

Appendix B Hopf decomposition

An important tool in the theory of non-singular transformations on σ\sigma-finite measure spaces is the Hopf decomposition, which partitions the phase space into the so-called dissipative and recurrent parts reflecting different dynamics of the transformation. In this appendix, we recall the relevant definitions and indicate what happens for measure-preserving transformations of a σ\sigma-finite space. The reader may consult [Kre85, Sec. 1.3] for further details on the following definitions.

Definition B.1.

Let SS be a non-singular transformation of a σ\sigma-finite measure space (Ω,λ)(\Omega,\lambda). A measurable set AΩA\subseteq\Omega is said to be:

  • wandering if ASk(A)=A\cap S^{k}(A)=\varnothing for all k1k\geq 1;

  • recurrent if Ak1Sk(A)A\subseteq\bigcup_{k\geq 1}S^{k}(A);

  • infinitely recurrent if An1knSk(A)A\subseteq\bigcap_{n\geq 1}\bigcup_{k\geq n}S^{k}(A).

The inclusions above are understood to hold up to a null set. The transformation SS is:

  • dissipative if the phase space Ω\Omega is a countable union of wandering sets;

  • conservative if there are no wandering sets of positive measure;

  • recurrent if every set of positive measure is recurrent;

  • infinitely recurrent if every set of positive measure is infinitely recurrent.

It turns out that the properties of being conservative, recurrent, and infinitely recurrent are all mutually equivalent.

Proposition B.2.

Let SS be a non-singular transformation of a σ\sigma-finite measure space (Ω,λ)(\Omega,\lambda). The following are equivalent:

  1. (1)

    SS is conservative;

  2. (2)

    SS is recurrent;

  3. (3)

    SS is infinitely recurrent.

Among the properties introduced in Definition B.1, only recurrence and dissipativity are therefore different. In fact, any non-singular transformation admits a canonical decomposition, known as the Hopf decomposition, into these two types of action.

Proposition B.3 (Hopf decomposition).

Let SS be a non-singular transformation of a σ\sigma-finite measure space (Ω,λ)(\Omega,\lambda). There exists an SS-invariant partition Ω=DC\Omega=D\sqcup C such that SDS\restriction_{D} is dissipative and SCS\restriction_{C} is recurrent (equivalently, conservative). Moreover, if Ω=DC\Omega=D^{\prime}\sqcup C^{\prime} is another partition with this property then λ(DD)=0\lambda(D\triangle D^{\prime})=0 and λ(CC)=0\lambda(C\triangle C^{\prime})=0.

We also note the following consequence of dissipativity in case the measure is preserved.

Proposition B.4.

Let SS be a measure-preserving transformation of a σ\sigma-finite measure space (Ω,λ)(\Omega,\lambda) and let Ω=DC\Omega=D\sqcup C be its Hopf decomposition. For every set AΩA\subseteq\Omega of finite measure, almost every point in DD eventually escapes AA:

λxDNnNTnxA.\forall^{\lambda}x\in D\ \exists N\ \forall n\geq N\ T^{n}x\not\in A.
Proof.

We may as well assume D=ΩD=\Omega. Let AΩA\subseteq\Omega have finite measure. Let QQ be a wandering set whose translates cover Ω\Omega. Consider the map Φ:Q×Ω\Phi:Q\times\mathbb{Z}\to\Omega which maps (x,n)(x,n) to Tn(x)T^{n}(x), and observe that Φ\Phi is measure-preserving if we endow Q×Q\times\mathbb{Z} with the product of the measure induced by λ\lambda on QQ and the counting measure on \mathbb{Z}.

So if there is a positive measure set of xQx\in Q such that Sn(x)AS^{n}(x)\in A for infinitely many nn\in\mathbb{N}, by Fubini’s theorem we would have that AA has infinite measure, a contradiction. The same conclusion is true if we replace QQ by any of its SS-translates, and since these translates cover Ω\Omega the proof is finished. ∎

Appendix C Actions of locally compact Polish groups

In this chapter of the appendix, we collect some well-known facts related to actions of locally compact Polish groups. This exposition is provided for reader’s convenience and completeness. We recall that by a result of G. W. Mackey [Mac62], any Boolean measure-preserving action of a locally compact Polish group can be realized as a spatial Borel action, so we may switch to pointwise formulations, whenever convenient for the exposition.

C.1. Disintegration of measure

Let \mathcal{R} be a smooth measurable equivalence relation on a standard Lebesgue space (X,μ)(X,\mu), and let π:XY\pi:X\to Y be a measurable reduction to the identity relation on some standard Lebesgue space (Y,ν)(Y,\nu), π(x)=π(y)\pi(x)=\pi(y) if and only if xyx\mathcal{R}y. Suppose that ν\nu is a σ\sigma-finite measure on YY that is equivalent to the push-forward πμ\pi_{*}\mu. A disintegration of μ\mu relative to (π,ν)(\pi,\nu) is a collection of measures (μy)yY(\mu_{y})_{y\in Y} on XX such that for all Borel sets AXA\subseteq X

  1. (1)

    μy(Xπ1(y))=0\mu_{y}(X\setminus\pi^{-1}(y))=0 for ν\nu-almost all yYy\in Y;

  2. (2)

    the map Yyμy(A)Y\ni y\mapsto\mu_{y}(A)\in\mathbb{R} is measurable;

  3. (3)

    μ(A)=Yμy(A)dν(y)\mu(A)=\int_{Y}\mu_{y}(A)\,d\nu(y).

A theorem of D. Maharam [Mah50] asserts that μ\mu can be disintegrated over any (π,ν)(\pi,\nu) as above. In fact, existence of a disintegration can be proved in a setup considerably more general (see, for example, D. H. Fremlin [Fre06, Thm. 452I]), but in the framework of standard Lebesgue spaces, disintegration is essentially unique. While the context of our work is purely ergodic theoretical, we note that the disintegration result holds in the descriptive set theoretical setting as well, as discussed in [Mah84] and [GM89]. Without striving for generality, we formulate here a specific version, which suits our needs.

Theorem C.1 (Disintegration of Measure).

Let (X,μ)(X,\mu) be a standard Lebesgue space, (Y,ν)(Y,\nu) be a σ\sigma-finite standard Lebesgue space, and let π:XY\pi:X\to Y be a measurable function. If πμ\pi_{*}\mu is equivalent to ν\nu, then there exists a disintegration (μy)yY(\mu_{y})_{y\in Y} of μ\mu over (π,ν)(\pi,\nu). Moreover, such a disintegration is essentially unique in the sense that if (μy)yY(\mu^{\prime}_{y})_{y\in Y} is another disintegration, then μy=μy\mu_{y}=\mu_{y}^{\prime} for ν\nu-almost all yYy\in Y.

Remark C.2.

It is more common to formulate the disintegration theorem with the assumption that πμ=ν\pi_{*}\mu=\nu, when one can additionally ensure that μy(X)=μ(X)\mu_{y}(X)=\mu(X) for ν\nu-almost all yy. Weakening the equality πμ=ν\pi_{*}\mu=\nu to mere equivalence is a simple consequence, for if (μy)yY(\mu_{y})_{y\in Y} is a disintegration of μ\mu over (π,πμ)(\pi,\pi_{*}\mu), then (dπμdν(y)μy)yY\bigl{(}\frac{d\pi_{*}\mu}{d\nu}(y)\cdot\mu_{y}\bigr{)}_{y\in Y} is a disintegration of μ\mu over (π,ν)(\pi,\nu).

Let XaXX_{a}\subseteq X be the set of atoms of the disintegration, i.e., Xa={xX:μy(x)>0 for some yY}X_{a}=\{x\in X:\mu_{y}(x)>0\textrm{ for some }y\in Y\}, and let FF be the equivalence relation on XaX_{a}, where two atoms within the same fiber are equivalent whenever they have the same measure: x1Fx2x_{1}Fx_{2} if and only if μπ(x1)(x1)=μπ(x2)(x2)\mu_{\pi(x_{1})}(x_{1})=\mu_{\pi(x_{2})}(x_{2}) and π(x1)=π(x2)\pi(x_{1})=\pi(x_{2}). The equivalence relation FF is measurable and has finite classes μ\mu-almost surely. Let XnX_{n}, n1n\geq 1, be the union of FF-equivalence classes of size exactly nn, thus Xa=n1XnX_{a}=\bigsqcup_{n\geq 1}X_{n}. Set also X0=XXaX_{0}=X\setminus X_{a} to be the atomless part of the disintegration and let n\mathcal{R}_{n} denote the restriction of \mathcal{R} onto XnX_{n}.

Consider the group []Aut(X,μ)[\mathcal{R}\mkern 1.5mu]\leq\mathrm{Aut}(X,\mu) of measure-preserving automorphisms for which xTxx\mathcal{R}Tx holds μ\mu-almost surely. Every T[]T\in[\mathcal{R}\mkern 1.5mu] preserves ν\nu-almost all measures μy\mu_{y}, since (Tμy)yY(T_{*}\mu_{y})_{y\in Y} is a disintegration of Tμ=μT_{*}\mu=\mu, which has to coincide with (μy)yY(\mu_{y})_{y\in Y} by uniqueness of the disintegration. In particular, the partition X=nXnX=\bigsqcup_{n\in\mathbb{N}}X_{n} is invariant under the full group [][\mathcal{R}\mkern 1.5mu], and for any T[]T\in[\mathcal{R}\mkern 1.5mu] the restriction TXn[n]T\restriction_{X_{n}}\in[\mathcal{R}_{n}\mkern 1.5mu] for every nn\in\mathbb{N}. Conversely, for a sequence Tn[n]T_{n}\in[\mathcal{R}_{n}\mkern 1.5mu], nn\in\mathbb{N}, one has T=nTn[]T=\bigsqcup_{n}T_{n}\in[\mathcal{R}\mkern 1.5mu]. We therefore have an isomorphism of (abstract) groups []n[n][\mathcal{R}\mkern 1.5mu]\cong\prod_{n\in\mathbb{N}}[\mathcal{R}_{n}\mkern 1.5mu].

The groups [n][\mathcal{R}_{n}\mkern 1.5mu] can be described quite explicitly. First, let us consider the case n1n\geq 1, thus XnXaX_{n}\subseteq X_{a}. All equivalence classes of the restriction of FF onto XnX_{n} have size nn. Let YnXnY_{n}\subseteq X_{n} be a measurable transversal, i.e., a measurable set intersecting every FF-class in a single point, and let νn=μYn\nu_{n}=\mu\restriction_{Y_{n}}. Every T[n]T\in[\mathcal{R}_{n}\mkern 1.5mu] produces a permutation of μ\mu-almost every FF-class, so we can view it as an element of L0(Yn,νn,𝔖n)\mathrm{L}^{0}(Y_{n},\nu_{n},\mathfrak{S}_{n}), where 𝔖n\mathfrak{S}_{n} is the group of permutations of an nn-element set. This identification works in both directions and produces an isomorphism [n]L0(Yn,νn,𝔖n)[\mathcal{R}_{n}\mkern 1.5mu]\cong\mathrm{L}^{0}(Y_{n},\nu_{n},\mathfrak{S}_{n}). Note also that all νn\nu_{n} are atomless if so is μ\mu. We allow for μ(Xn)=0\mu(X_{n})=0, in which case L0(Yn,νn,𝔖n)\mathrm{L}^{0}(Y_{n},\nu_{n},\mathfrak{S}_{n}) is the trivial group.

Let us now go back to the equivalence relation 0=X0×X0\mathcal{R}_{0}=\mathcal{R}\cap X_{0}\times X_{0}, and recall that measures μyX0\mu_{y}\restriction_{X_{0}} are atomless. Let Y0={y:μy(X0)>0}Y_{0}=\{y:\mu_{y}(X_{0})>0\} be the encoding of fibers with non-trivial atomless components and put ν0=νY0\nu_{0}=\nu\restriction_{Y_{0}}. In particular, for every yY0y\in Y_{0} the space (X0,μy)(X_{0},\mu_{y}) is isomorphic to the interval [0,μy(X0)][0,\mu_{y}(X_{0})] endowed with the Lebesgue measure. In fact, one can select such isomorphisms in a measurable way across all yY0y\in Y_{0}. More precisely, there is a measurable isomorphism ψ:X0{(y,r)Y0×:0rμy(X0)}\psi:X_{0}\to\{(y,r)\in Y_{0}\times\mathbb{R}:0\leq r\leq\mu_{y}(X_{0})\} such that for all yY0y\in Y_{0}

  • ψ(π1(y)X0)={y}×[0,μy(X0)]\psi(\pi^{-1}(y)\cap X_{0})=\{y\}\times[0,\mu_{y}(X_{0})];

  • ψμyX0\psi_{*}\mu_{y}\restriction_{X_{0}} coincides with the Lebesgue measure on {y}×[0,μy(X0)]\{y\}\times[0,\mu_{y}(X_{0})].

The reader may find further details in [GM89, Thm. 2.3], where the same construction is discussed in a more refined setting of Borel disintegrations.

Using the isomorphism ψ\psi, we can identify each π1(y)X0\pi^{-1}(y)\cap X_{0}, yY0y\in Y_{0}, with [0,μy(X0)][0,\mu_{y}(X_{0})]. Since every T[0]T\in[\mathcal{R}_{0}\mkern 1.5mu] preserves ν\nu-almost every μy\mu_{y}, we may rescale these intervals and view any T[0]T\in[\mathcal{R}_{0}\mkern 1.5mu] as an element of L0(Y0,ν0,Aut([0,1],λ))\mathrm{L}^{0}(Y_{0},\nu_{0},\mathrm{Aut}([0,1],\lambda)). Conversely, every fL0(Y0,ν0,Aut([0,1],λ))f\in\mathrm{L}^{0}(Y_{0},\nu_{0},\mathrm{Aut}([0,1],\lambda)) gives rise to Tf[0]T_{f}\in[\mathcal{R}_{0}\mkern 1.5mu] via the notationally convoluted but natural

Tf(x)=ψ1(π(x),(f(π(x))proj2(ψ(x))/μπ(x)(X0))μπ(x)(X0)),T_{f}(x)=\psi^{-1}\bigl{(}\pi(x),\bigl{(}f(\pi(x))\cdot\mathrm{proj}_{2}(\psi(x))/\mu_{\pi(x)}(X_{0})\bigr{)}\mu_{\pi(x)}(X_{0})\bigr{)},

which, in plain words, simply applies f(π(x))f(\pi(x)) upon the corresponding fiber identified with [0,1][0,1] using ψ\psi. This map is an isomorphism between the groups [0][\mathcal{R}_{0}\mkern 1.5mu] and L0(Y0,ν0,Aut([0,1],λ))\mathrm{L}^{0}(Y_{0},\nu_{0},\mathrm{Aut}([0,1],\lambda)).

Let us say that \mathcal{R} has atomless classes if μy\mu_{y} is atomless ν\nu-almost surely or, equivalently, μ(Xa)=0\mu(X_{a})=0 in the notation above. We may summarize the discussion so far into the following proposition.

Proposition C.3.

Let \mathcal{R} be a smooth measurable equivalence relation on a standard Lebesgue space (X,μ)(X,\mu). There are (possibly empty) standard Lebesgue spaces (Yk,νk)(Y_{k},\nu_{k}), kk\in\mathbb{N}, such that the full group []Aut(X,μ)[\mathcal{R}\mkern 1.5mu]\leq\mathrm{Aut}(X,\mu) is (abstractly) isomorphic to

L0(Y0,ν0,Aut([0,1],λ))×n1L0(Yn,νn,𝔖n),\mathrm{L}^{0}(Y_{0},\nu_{0},\mathrm{Aut}([0,1],\lambda))\times\prod_{n\geq 1}\mathrm{L}^{0}(Y_{n},\nu_{n},\mathfrak{S}_{n}),

where 𝔖n\mathfrak{S}_{n} is the group of permutations of a nn-element set. If μ\mu is atomless, then so are the spaces (Yn,νn)(Y_{n},\nu_{n}), n1n\geq 1. If \mathcal{R} has atomless classes, then all (Yn,νn)(Y_{n},\nu_{n}), n1n\geq 1, are negligible and [][\mathcal{R}\mkern 1.5mu] is isomorphic to L0(Y0,ν0,Aut([0,1],λ))\mathrm{L}^{0}(Y_{0},\nu_{0},\mathrm{Aut}([0,1],\lambda)).

We can further refine the product in Proposition C.3 by decomposing the spaces (Yn,νn)(Y_{n},\nu_{n}) into individual atoms and the atomless remainders. More specifically, let (Z,νZ)(Z,\nu_{Z}) be a standard Lebesgue space and GG be a Polish group. Given a measurable partition Z=Z0Z1Z=Z_{0}\sqcup Z_{1}, every function fL0(Z,νZ,G)f\in\mathrm{L}^{0}(Z,\nu_{Z},G) can be associated with a pair (f0,f1)L0(Z0,νZ,0,G)×L0(Z1,νZ,1,G)(f_{0},f_{1})\in\mathrm{L}^{0}(Z_{0},\nu_{Z,0},G)\times\mathrm{L}^{0}(Z_{1},\nu_{Z,1},G), νZ,i=νZZi\nu_{Z,i}=\nu_{Z}\restriction_{Z_{i}} and fi=fZif_{i}=f\restriction_{Z_{i}}, which is an isomorphism of the topological groups. The same consideration applies to finite or countably infinite partitions.

Proposition C.4.

Let (Z,νZ)(Z,\nu_{Z}) be a standard Lebesgue space and GG be a Polish group. For any finite or countably infinite measurable partition Z=nIZnZ=\bigsqcup_{n\in I}Z_{n}, there is an isomorphism of topological groups L0(Z,νZ,G)\mathrm{L}^{0}(Z,\nu_{Z},G) and nIL0(Zn,νZ,n,G)\prod_{n\in I}\mathrm{L}^{0}(Z_{n},\nu_{Z,n},G), where νZ,n\nu_{Z,n} is the restriction of νZ\nu_{Z} onto ZnZ_{n}.

Applying Proposition C.4 to the partition of (Z,νZ)(Z,\nu_{Z}) into the atomless part Z0Z_{0} and individual atoms Zk={zk}Z_{k}=\{z_{k}\} (if any), and noting that for a singleton ZkZ_{k} the group L0(Zk,νZ,k,G)\mathrm{L}^{0}(Z_{k},\nu_{Z,k},G) is naturally isomorphic to GG, we get the following corollary.

Corollary C.5.

Let (Z,νZ)(Z,\nu_{Z}) be a standard Lebesgue space and GG be a Polish group. Let ZaZZ_{a}\subseteq Z be the set of atoms of ZZ and Z0=ZZaZ_{0}=Z\setminus Z_{a} be the atomless part. The group L0(Z,νZ,G)\mathrm{L}^{0}(Z,\nu_{Z},G) is isomorphic to L0(Z0,νZZ0,G)×G|Za|\mathrm{L}^{0}(Z_{0},\nu_{Z}\restriction_{Z_{0}},G)\times G^{|Z_{a}|}.

Combining the discussion above with Proposition C.3, we obtain a very concrete representation for [][\mathcal{R}\mkern 1.5mu]. In the formulation below, G0G^{0} is understood to be the trivial group.

Proposition C.6.

Let \mathcal{R} be a smooth measurable equivalence relation on a standard Lebesgue space (X,μ)(X,\mu). There are cardinals κn0\kappa_{n}\leq\aleph_{0} and ϵn{0,1}\epsilon_{n}\in\{0,1\} such that

[]L0([0,1],λ,Aut([0,1],λ))ϵ0×Aut([0,1],λ)κ0×(n1L0([0,1],λ,𝔖n)ϵn×𝔖nκn).[\mathcal{R}\mkern 1.5mu]\cong\mathrm{L}^{0}([0,1],\lambda,\mathrm{Aut}([0,1],\lambda))^{\epsilon_{0}}\times\mathrm{Aut}([0,1],\lambda)^{\kappa_{0}}\times\Bigl{(}\prod_{n\geq 1}\mathrm{L}^{0}([0,1],\lambda,\mathfrak{S}_{n})^{\epsilon_{n}}\times\mathfrak{S}_{n}^{\kappa_{n}}\Bigr{)}.

If μ\mu is atomless, then κn=0\kappa_{n}=0 for all n1n\geq 1; if \mathcal{R} has atomless classes, then ϵn=0\epsilon_{n}=0 for all n1n\geq 1.

So far we viewed [][\mathcal{R}\mkern 1.5mu] as an abstract group. This is because neither of the two natural topologies on Aut(X,μ)\mathrm{Aut}(X,\mu) play well with the full group construction—[][\mathcal{R}\mkern 1.5mu] is generally not closed in the weak topology, and not separable in the uniform topology whenever μ(X0)>0\mu(X_{0})>0. Nonetheless, the isomorphism given in Proposition C.3 shows that there is a natural Polish topology on [][\mathcal{R}\mkern 1.5mu], which arises when we view groups L0(Y0,ν0,Aut([0,1],λ))\mathrm{L}^{0}(Y_{0},\nu_{0},\mathrm{Aut}([0,1],\lambda)) and L0(Yn,νn,𝔖n)\mathrm{L}^{0}(Y_{n},\nu_{n},\mathfrak{S}_{n}) as Polish groups in the topology of convergence in measure. It is with respect to this topology we formulate Proposition C.7.

Proposition C.7.

Let \mathcal{R} be a smooth measurable equivalence relation on a standard Lebesgue space (X,μ)(X,\mu). The set of periodic elements is dense in [][\mathcal{R}\mkern 1.5mu].

Proof.

Rokhlin’s Lemma implies that any T[]T\in[\mathcal{R}\mkern 1.5mu] can be approximated in the uniform topology by periodic elements from [T][][T\mkern 1.5mu]\subseteq[\mathcal{R}\mkern 1.5mu]. Since the uniform topology is stronger than the Polish topology on [][\mathcal{R}\mkern 1.5mu], the proposition follows. ∎

C.2. Tessellations

An important feature of locally compact group actions is the fact that they all admit measurable cross-sections. This was proved by J. Feldman, P. Hahn, and C. Moore in [FHM78], whereas a Borel version of the result was obtained by A. S. Kechris in [Kec92].

Definition C.8.

Let GXG\curvearrowright X be a Borel action of a locally compact Polish group. A cross-section is a Borel set 𝒞X\mathcal{C}\subseteq X which is both

  • a complete section for G\mathcal{R}_{G}: it intersects every orbit of the action and

  • lacunary: for some neighborhood of the identity 1GUG1_{G}\in U\subseteq G one has UcUc=U\cdot c\cap U\cdot c^{\prime}=\varnothing for all distinct c,c𝒞c,c^{\prime}\in\mathcal{C}.

A cross-section 𝒞\mathcal{C} is KK-cocompact, where KGK\subseteq G is a compact set, if K𝒞=XK\cdot\mathcal{C}=X; a cross-section is cocompact if it is KK-cocompact for some compact KGK\subseteq G.

Any action GXG\curvearrowright X admits a KK-cocompact cross-section, whenever KGK\subseteq G is a compact neighborhood of the identity (see [Slu17, Thm. 2.4]). We also remind the following well-known lemma on the possibility to partition a cross-section into pieces with a prescribed lacunarity parameter.

Lemma C.9.

Let GXG\curvearrowright X be a Borel action of a locally compact Polish group and 𝒞\mathcal{C} be a cross-section for the action. For any compact neighborhood of the identity VGV\subseteq G, there exists a finite Borel partition 𝒞=i𝒞i\mathcal{C}=\bigsqcup_{i}\mathcal{C}_{i} such that each 𝒞i\mathcal{C}_{i} is VV-lacunary.

Proof.

Set K=(VV1)2K=(V\cup V^{-1})^{2} and let UGU\subseteq G be a compact neighborhood of the identity small enough for 𝒞\mathcal{C} to be UU-lacunary. Define a binary relation 𝒢\mathcal{G} on 𝒞\mathcal{C} by declaring (c,c)𝒢(c,c^{\prime})\in\mathcal{G} whenever cKcc\in K\cdot c^{\prime} and ccc\neq c^{\prime}. Note that 𝒢\mathcal{G} is symmetric since so is KK. We view 𝒢\mathcal{G} as a Borel graph on 𝒞\mathcal{C} and claim that it is locally finite. More specifically, if λ\lambda is a right Haar measure, then the degree of each c𝒞c\in\mathcal{C} is at most λ(UK)λ(U)1\bigl{\lfloor}\frac{\lambda(U\cdot K)}{\lambda(U)}\bigr{\rfloor}-1.

Indeed, let c0,,cN𝒞c_{0},\ldots,c_{N}\in\mathcal{C} be distinct elements such that ciKc0c_{i}\in K\cdot c_{0} for all iNi\leq N; in particular (ci,c0)𝒢(c_{i},c_{0})\in\mathcal{G} for i1i\geq 1. Let kiKk_{i}\in K be such that kic0=cik_{i}\cdot c_{0}=c_{i}. Lacunarity of 𝒞\mathcal{C} asserts that sets Uci=Ukic0U\cdot c_{i}=Uk_{i}\cdot c_{0} are supposed to be pairwise disjoint, which necessitates UkiUk_{i} to be pairwise disjoint for 0iN0\leq i\leq N. Clearly UkiUKUk_{i}\subseteq UK as kiKk_{i}\in K. Using the right-invariance of λ\lambda, we have λ(UK)λ(iNUki)=(N+1)λ(U)\lambda(UK)\geq\lambda\bigl{(}\bigsqcup_{i\leq N}Uk_{i}\bigr{)}=(N+1)\lambda(U), and thus N+1λ(UK)λ(U)N+1\leq\frac{\lambda(UK)}{\lambda(U)}, as claimed.

We may now use [KST99, Prop. 4.6] to deduce existence of a finite partition 𝒞=i𝒞i\mathcal{C}=\bigsqcup_{i}\mathcal{C}_{i} such that no two points in 𝒞i\mathcal{C}_{i} are adjacent. In other words, if c,c𝒞ic,c^{\prime}\in\mathcal{C}_{i} are distinct, then cKcc\not\in K\cdot c^{\prime}, and therefore VcVc=V\cdot c\cap V\cdot c^{\prime}=\varnothing, which shows that 𝒞i\mathcal{C}_{i} are VV-lacunary. ∎

Every cross-section 𝒞\mathcal{C} gives rise to a smooth subrelation of G\mathcal{R}_{G} by associating to xXx\in X “the closest point” of 𝒞\mathcal{C} in the same orbit. Such a subrelation is known as the Voronoi tessellation. For the purposes of Chapter 5, we need a slightly more abstract concept of a tessellation which may not correspond to Voronoi domains. While far from being the most general, the following treatment is sufficient for our needs.

Definition C.10.

Let GXG\curvearrowright X be a Borel action of a locally compact Polish group on a standard Borel space and let 𝒞X\mathcal{C}\subseteq X be a cross-section. A tessellation over 𝒞\mathcal{C} is a Borel set 𝒲𝒞×X\mathcal{W}\subseteq\mathcal{C}\times X such that

  1. (1)

    all fibers 𝒲c={xX:(c,x)𝒲}\mathcal{W}_{c}=\{x\in X:(c,x)\in\mathcal{W}\} are pairwise disjoint for c𝒞c\in\mathcal{C};

  2. (2)

    for all c𝒞c\in\mathcal{C} elements of 𝒲c\mathcal{W}_{c} are G\mathcal{R}_{G}-equivalent to cc, i.e., {c}×𝒲cG\{c\}\times\mathcal{W}_{c}\subseteq\mathcal{R}_{G};

  3. (3)

    fibers cover the phase space, X=c𝒞𝒲cX=\bigsqcup_{c\in\mathcal{C}}\mathcal{W}_{c}.

A tessellation 𝒲\mathcal{W} over 𝒞\mathcal{C} is NN-lacunary for an open NGN\subseteq G if

{(c,Nc):c𝒞}𝒲.\{(c,N\cdot c):c\in\mathcal{C}\}\subseteq\mathcal{W}.

It is KK-cocompact, KGK\subseteq G, if 𝒲{(c,Kc):c𝒞}\mathcal{W}\subseteq\{(c,K\cdot c):c\in\mathcal{C}\}.

Any tessellation 𝒲\mathcal{W} can be viewed as a (flipped) graph of a function, since for any xXx\in X there is a unique c𝒞c\in\mathcal{C} such that (c,x)𝒲(c,x)\in\mathcal{W}. We denote such cc by π𝒲(x)\pi_{\mathcal{W}}(x), which produces a Borel map π𝒲:X𝒞\pi_{\mathcal{W}}:X\to\mathcal{C}. There is a natural equivalence relation 𝒲\mathcal{R}_{\mathcal{W}} associated with the tessellation. Namely, x1x_{1} and x2x_{2} are 𝒲\mathcal{R}_{\mathcal{W}}-equivalent whenever they belong to the same fiber, i.e., π𝒲(x1)=π𝒲(x2)\pi_{\mathcal{W}}(x_{1})=\pi_{\mathcal{W}}(x_{2}). In view of the item (2), 𝒲G\mathcal{R}_{\mathcal{W}}\subseteq\mathcal{R}_{G} and moreover, every G\mathcal{R}_{G}-class consists of countably many 𝒲\mathcal{R}_{\mathcal{W}}-classes.

Voronoi tessellations provide a specific way of constructing tessellations over a given cross-section. Suppose that the group GG is endowed with a compatible proper norm \lVert\cdot\rVert. Let D:G0D:\mathcal{R}_{G}\to\mathbb{R}^{\geq 0} be the associated metric on the orbits of the action (as in Section 2.2) and let 𝒞\preceq_{\mathcal{C}} be a Borel linear order on 𝒞\mathcal{C}. The Voronoi tessellation over the cross-section 𝒞\mathcal{C} relative to a proper norm \left\lVert\cdot\right\rVert is the set 𝒱𝒞𝒞×X\mathcal{V}_{\mathcal{C}}\subseteq\mathcal{C}\times X defined by

𝒱𝒞={(c,x)𝒞×X:\displaystyle\mathcal{V}_{\mathcal{C}}=\bigl{\{}(c,x)\in\mathcal{C}\times X\ : cGx and for all c𝒞 such that cGx either\displaystyle c\mathcal{R}_{G}x\textrm{ and for all }c^{\prime}\in\mathcal{C}\textrm{ such that }c^{\prime}\mathcal{R}_{G}x\textrm{ either }
D(c,x)<D(c,x) or\displaystyle D(c,x)<D(c^{\prime},x)\textrm{ or }
(D(c,x)=D(c,x) and c𝒞c)}.\displaystyle(\,D(c,x)=D(c^{\prime},x)\textrm{ and }c\preceq_{\mathcal{C}}c^{\prime}\,)\bigr{\}}.

Properness of the norm ensures that for each xXx\in X there are only finitely many candidates cc which minimize D(c,x)D(c,x), and hence each xXx\in X is associated with a unique c𝒞c\in\mathcal{C}.

For the sake of Chapter 5, we need a definition of the Voronoi tessellation for norms that may not be proper. The set 𝒱𝒞\mathcal{V}_{\mathcal{C}} specified as above may in this case fail to satisfy item (3) of the definition of a tessellation, as for some xXx\in X there may be infinitely many c𝒞c\in\mathcal{C} that minimize D(c,x)D(c,x), none of which are 𝒞\preceq_{\mathcal{C}}-minimal. We therefore need a different way to resolve the points on the “boundary” between the regions, which can be done, for example, by delegating this task to a proper norm.

Definition C.11.

Let \left\lVert\cdot\right\rVert be a compatible norm on GG and let 𝒞\mathcal{C} be a cross-section. Pick a compatible proper norm \left\lVert\cdot\right\rVert^{\prime} on GG and a Borel linear order 𝒞\preceq_{\mathcal{C}} on 𝒞\mathcal{C}. Let DD and DD^{\prime} be the metrics on orbits of the action associated with the norms \left\lVert\cdot\right\rVert and \left\lVert\cdot\right\rVert^{\prime} respectively. The Voronoi tessellation over the cross-section 𝒞\mathcal{C} relative to the norm \left\lVert\cdot\right\rVert is the set 𝒱𝒞𝒞×X\mathcal{V}_{\mathcal{C}}\subseteq\mathcal{C}\times X defined by

𝒱𝒞={(c,x)𝒞×X:\displaystyle\mathcal{V}_{\mathcal{C}}=\bigl{\{}(c,x)\in\mathcal{C}\times X\ : cGx and for all c𝒞 such that cGx either\displaystyle c\mathcal{R}_{G}x\textrm{ and for all }c^{\prime}\in\mathcal{C}\textrm{ such that }c^{\prime}\mathcal{R}_{G}x\textrm{ either }
D(c,x)<D(c,x) or\displaystyle D(c,x)<D(c^{\prime},x)\textrm{ or }
(D(c,x)=D(c,x) and D(c,x)<D(c,x)) or\displaystyle(\,D(c,x)=D(c^{\prime},x)\textrm{ and }D^{\prime}(c,x)<D^{\prime}(c^{\prime},x)\,)\textrm{ or }
(D(c,x)=D(c,x) and D(c,x)=D(c,x) and c𝒞c)}.\displaystyle(\,D(c,x)=D(c^{\prime},x)\textrm{ and }D^{\prime}(c,x)=D^{\prime}(c^{\prime},x)\textrm{ and }c\preceq_{\mathcal{C}}c^{\prime}\,)\bigr{\}}.

The definition of the Voronoi tessellation does depend on the choice of the norm \left\lVert\cdot\right\rVert^{\prime} and the linear order 𝒞\preceq_{\mathcal{C}} on the cross-section, but its key properties are the same regardless of these choices. We therefore often do not specify explicitly which \left\lVert\cdot\right\rVert^{\prime} and 𝒞\preceq_{\mathcal{C}} are picked. Note also that if the cross-section is cocompact, then every region of the Voronoi tessellation is bounded, i.e., supxXD(x,π𝒱𝒞(x))<+\sup_{x\in X}D(x,\pi_{\mathcal{V}_{\mathcal{C}}}(x))<+\infty.

Our goal is to show that equivalence relations 𝒲\mathcal{R}_{\mathcal{W}} are atomless in the sense of Section C.1 as long as each orbit of the action is uncountable. To this end we first need the following lemma.

Lemma C.12.

Let GG be a locally compact Polish group acting on a standard Lebesgue space (X,μ)(X,\mu) by measure-preserving automorphisms. Suppose that almost every orbit of the action is uncountable. If 𝒜X\mathcal{A}\subseteq X is a measurable set such that the intersection of 𝒜\mathcal{A} with almost every orbit is countable, then μ(𝒜)=0\mu(\mathcal{A})=0.

Proof.

Pick a proper norm \left\lVert\cdot\right\rVert on GG, let 𝒞\mathcal{C} be a cross-section for the action, let B2rGB_{2r}\subseteq G be an open ball around the identity of sufficiently small radius 2r>02r>0 such that B2rcB2rc=B_{2r}\cdot c\cap B_{2r}\cdot c^{\prime}=\varnothing whenever c,c𝒞c,c^{\prime}\in\mathcal{C} are distinct, and let 𝒱𝒞\mathcal{V}_{\mathcal{C}} be the Voronoi tessellation over 𝒞\mathcal{C} relative to \left\lVert\cdot\right\rVert. Note that B2rcB_{2r}\cdot c is fully contained in the 𝒱𝒞\mathcal{R}_{\mathcal{V}_{\mathcal{C}}}-class of cc and set X=Br𝒞X=B_{r}\cdot\mathcal{C}. Let also (gn)n(g_{n})_{n\in\mathbb{N}} be a countable dense subset of GG.

We claim that it is enough to consider the case when 𝒜\mathcal{A} intersects each 𝒱𝒞\mathcal{R}_{\mathcal{V}_{\mathcal{C}}}-class in at most one point. Indeed, the restriction of 𝒱𝒞\mathcal{R}_{\mathcal{V}_{\mathcal{C}}} onto 𝒜\mathcal{A} is a smooth countable equivalence relation, so one can write 𝒜=n𝒜n\mathcal{A}=\bigsqcup_{n\in\mathbb{N}}\mathcal{A}_{n}^{\prime}, where each 𝒜n\mathcal{A}_{n}^{\prime} intersects each 𝒱𝒞\mathcal{R}_{\mathcal{V}_{\mathcal{C}}}-class in at most one point. To simplify notations, we assume that 𝒜\mathcal{A} already possesses this property.

Let γ:X\gamma:X\to\mathbb{N} be defined by γ(x)=min{n:x𝒱𝒞gnx and gnxX}\gamma(x)=\min\{n\in\mathbb{N}:x\mathcal{R}_{\mathcal{V}_{\mathcal{C}}}g_{n}x\textrm{ and }g_{n}x\in X\}. Let 𝒜n=𝒜γ1(n)\mathcal{A}_{n}=\mathcal{A}\cap\gamma^{-1}(n) and note that sets 𝒜n\mathcal{A}_{n} partition 𝒜\mathcal{A}. It is therefore enough to show that μ(𝒜n)=0\mu(\mathcal{A}_{n})=0 for any nn\in\mathbb{N}. Pick n0n_{0}\in\mathbb{N}. The action is measure-preserving and therefore μ(𝒜n0)=μ(gn0𝒜n0)\mu(\mathcal{A}_{n_{0}})=\mu(g_{n_{0}}\mathcal{A}_{n_{0}}). Set 0=gn0𝒜n0\mathcal{B}_{0}=g_{n_{0}}\mathcal{A}_{n_{0}} and note that for any x0x\in\mathcal{B}_{0} and gBrGg\in B_{r}\subseteq G one has gx𝒱𝒞xgx\mathcal{R}_{\mathcal{V}_{\mathcal{C}}}x. If the action were free, we could easily conclude that μ(0)=0\mu(\mathcal{B}_{0})=0, since sets g0g\mathcal{B}_{0}, gBrg\in B_{r}, would be pairwise disjoint. In general, we need to exhibit a little more care and construct a countable family of pairwise disjoint sets n\mathcal{B}_{n} as follows.

For x0x\in\mathcal{B}_{0} let τn(x)=min{m:x𝒱𝒞gmx and gmxknn}\tau_{n}(x)=\min\{m\in\mathbb{N}:x\mathcal{R}_{\mathcal{V}_{\mathcal{C}}}g_{m}x\textrm{ and }g_{m}x\not\in\bigcup_{k\leq n}\mathcal{B}_{n}\}. The value τn(x)\tau_{n}(x) is well-defined because the stabilizer of xx is closed and must be nowhere dense in BrB_{r} due to the orbit GxG\cdot x being uncountable. Put n+1={gτn(x)x:x0}\mathcal{B}_{n+1}=\{g_{\tau_{n}(x)}x:x\in\mathcal{B}_{0}\} and note that μ(n)=μ(0)\mu(\mathcal{B}_{n})=\mu(\mathcal{B}_{0}). We get a pairwise disjoint infinite family of sets n\mathcal{B}_{n} all having the same measure. Since μ\mu is finite, we conclude that μ(0)=0\mu(\mathcal{B}_{0})=0 and the lemma follows. ∎

Corollary C.13.

Let GG be a locally compact Polish group acting on a standard Lebesgue space (X,μ)(X,\mu) by measure-preserving automorphisms, let 𝒞\mathcal{C} be a cross-section for the action and let 𝒲𝒞×X\mathcal{W}\subseteq\mathcal{C}\times X be a tessellation. If μ\mu-almost every orbit of GG is uncountable, then 𝒲\mathcal{R}_{\mathcal{W}} is atomless.

Proof.

Consider the disintegration (μc)c𝒞(\mu_{c})_{c\in\mathcal{C}} of 𝒲\mathcal{R}_{\mathcal{W}} relative to (π𝒲,ν)(\pi_{\mathcal{W}},\nu), where π𝒲:X𝒞\pi_{\mathcal{W}}:X\to\mathcal{C} and ν=(π𝒲)μ\nu=(\pi_{\mathcal{W}})_{*}\mu. Let XaXX_{a}\subseteq X be the set of atoms of the disintegration. Since ν\nu-almost every μc\mu_{c} is finite, fibers π1𝒲(c)\pi^{-1}_{\mathcal{W}}(c) have countably many atoms. Since every tessellation has only countably many tiles within each orbit, we conclude that XaX_{a} has countable intersection with almost every orbit of the action. Lemma C.12 applies and shows that μ(Xa)=0\mu(X_{a})=0, hence 𝒲\mathcal{R}_{\mathcal{W}} is atomless as required. ∎

Consider the full group [𝒲][\mathcal{R}_{\mathcal{W}}\mkern 1.5mu] which by Proposition C.3 and Corollary C.13 is isomorphic to L0(Y,ν,Aut([0,1],λ))\mathrm{L}^{0}(Y,\nu,\mathrm{Aut}([0,1],\lambda)) for some standard Lebesgue space (Y,ν)(Y,\nu). This full group can naturally be viewed as a subgroup of [G][\mathcal{R}_{G}\mkern 1.5mu] and the topology induced on [𝒲][\mathcal{R}_{\mathcal{W}}\mkern 1.5mu] from the full group [G][\mathcal{R}_{G}\mkern 1.5mu] coincides with the topology of converges in measure on L0(Y,ν,Aut([0,1],λ))\mathrm{L}^{0}(Y,\nu,\mathrm{Aut}([0,1],\lambda)) (see Section 3 of [CLM16]). We therefore have the following corollary.

Corollary C.14.

Let GG be a locally compact Polish group acting on a standard Lebesgue space (X,μ)(X,\mu) by measure-preserving automorphisms, let 𝒞\mathcal{C} be a cross-section for the action and let 𝒲𝒞×X\mathcal{W}\subseteq\mathcal{C}\times X be a tessellation and π𝒲:X𝒞\pi_{\mathcal{W}}:X\to\mathcal{C} be the corresponding reduction. If μ\mu-almost every orbit of GG is uncountable, then the subgroup [𝒲][G][\mathcal{R}_{\mathcal{W}}\mkern 1.5mu]\leq[\mathcal{R}_{G}\mkern 1.5mu] is isomorphic as a topological group to L0(𝒞,(π𝒲)μ,Aut([0,1],λ))\mathrm{L}^{0}(\mathcal{C},(\pi_{\mathcal{W}})_{*}\mu,\mathrm{Aut}([0,1],\lambda)). If moreover all orbits of the action have measure zero, then (π𝒲)μ(\pi_{\mathcal{W}})_{*}\mu is non-atomic and [𝒲][\mathcal{R}_{\mathcal{W}}\mkern 1.5mu] is isomorphic to L0([0,1],λ,Aut([0,1],λ))\mathrm{L}^{0}([0,1],\lambda,\mathrm{Aut}([0,1],\lambda)).

C.3. Ergodic decomposition

Let GXG\curvearrowright X be a free measure-preserving action of a locally compact group on a standard probability space (X,μ)(X,\mu). The space =EINV(GX)\mathcal{E}=\mathrm{EINV}(G\curvearrowright X) of invariant ergodic probability measures of this action possesses a structure of a standard Borel space. The Ergodic Decomposition theorem of V. S. Varadarajan [Var63, Thm. 4.2] asserts that there is an essentially unique Borel G\mathcal{R}_{G}-invariant surjection XxνxX\ni x\mapsto\nu_{x}\in\mathcal{E} and a probability measure pp on \mathcal{E} such that μ=νdp(ν)\mu=\int_{\mathcal{E}}\nu\,dp(\nu) in the sense that for all Borel AXA\subseteq X one has μ(A)=ν(A)dp(ν)\mu(A)=\int_{\mathcal{E}}\nu(A)\,dp(\nu).

There is a one-to-one correspondence between measurable G\mathcal{R}_{G}-invariant functions h:Xh:X\to\mathbb{R} and measurable functions h~:\tilde{h}:\mathcal{E}\to\mathbb{R} given by h~(νx)=h(x)\tilde{h}(\nu_{x})=h(x). For measures μ\mu and pp as above, this correspondence gives an isometric isomorphism between L1(,)\mathrm{L}^{1}(\mathcal{E},\mathbb{R}) and the subspace of L1(X,)\mathrm{L}^{1}(X,\mathbb{R}) that consists of G\mathcal{R}_{G}-invariant functions.

Appendix D Conditional measures

The ergodic decomposition theorem, as formulated in Section C.3, is not available for general probability measure-preserving actions of Polish groups. Conditional measures provide a useful framework to remedy this. As before, Aut(X,μ)\mathrm{Aut}(X,\mu) stands for the group of measure-preserving automorphisms of a standard probability space. It is more useful, however, to view Aut(X,μ)\mathrm{Aut}(X,\mu) as the group of measure-preserving automorphisms of the measure algebra MAlg(X,μ)\mathrm{MAlg}(X,\mu) of (X,μ)(X,\mu), i.e., is the Boolean algebra of equivalence classes of Borel subsets of XX, identified up to measure zero. The measure algebra is endowed with a natural metric dμd_{\mu} given by dμ(A,B)=μ(AB)d_{\mu}(A,B)=\mu(A\bigtriangleup B). Completeness of (MAlg(X,μ))(\mathrm{MAlg}(X,\mu)) in this metric is a standard and well-known fact (see, for instance, [Kec95, Exer. 17.43]), which we include for reader’s convenience.

Proposition D.1.

The metric space (MAlg(X,μ),dμ)(\mathrm{MAlg}(X,\mu),d_{\mu}) is complete.

Proof.

It suffices to show that a Cauchy sequence (An)n(A_{n})_{n} admits a converging subsequence. Passing to a subsequence if necessary, we may assume that dμ(An,An+1)<2nd_{\mu}(A_{n},A_{n+1})<2^{-n} holds for all nn, and therefore nμ(AnAn+1)<+\sum_{n\in\mathbb{N}}\mu(A_{n}\bigtriangleup A_{n+1})<+\infty. The set

A={xX:xAn for all but finitely many n}A=\{x\in X:x\in A_{n}\text{ for all but finitely many }n\in\mathbb{N}\}

is the limit we seek. Indeed, given an ϵ>0\epsilon>0 and an index NN chosen so large that nNμ(AnAn+1)<ϵ\sum_{n\geq N}\mu(A_{n}\bigtriangleup A_{n+1})<\epsilon, for all nNn\geq N and all xx outside of the set nNAnAn+1\bigcup_{n\geq N}A_{n}\bigtriangleup A_{n+1} of measure at most ϵ\epsilon, we have xAnx\in A_{n} if and only if xAx\in A. ∎

Note that closed (or equivalently, metrically complete) subalgebras of MAlg(X,μ)\mathrm{MAlg}(X,\mu) are in a natural one-to-one correspondence with complete (in the measure-theoretical sense) σ\sigma-subalgebras of the σ\sigma-algebra of Lebesgue measurable sets.

D.1. Conditional expectations

We review here how conditional expectations can easily be defined without appealing to disintegration.

Let MM be a closed subalgebra of MAlg(X,μ)\mathrm{MAlg}(X,\mu) and let L2(M,μ)\mathrm{L}^{2}(M,\mu) denote the L2\mathrm{L}^{2} space of real-valued MM-measurable functions. Note that L2(M,μ)\mathrm{L}^{2}(M,\mu) is a closed subspace of L2(X,μ)=L2(MAlg(X,μ),μ)\mathrm{L}^{2}(X,\mu)=\mathrm{L}^{2}(\mathrm{MAlg}(X,\mu),\mu). The MM-conditional expectation is the orthogonal projection 𝔼M:L2(X,μ)L2(M,μ)\mathbb{E}_{M}:\mathrm{L}^{2}(X,\mu)\to\mathrm{L}^{2}(M,\mu). It is also uniquely defined by the condition

(D.1) Xfgdμ=X𝔼M(f)gdμfor all fL2(X,μ) and all gL2(M,μ).\int_{X}fg\,d\mu=\int_{X}\mathbb{E}_{M}(f)g\,d\mu\quad\textrm{for all $f\in\mathrm{L}^{2}(X,\mu)$ and all $g\in\mathrm{L}^{2}(M,\mu)$}.

By the density of step functions in L2(M,μ)\mathrm{L}^{2}(M,\mu), the conditional expectation can equivalently be defined as the linear contraction L2(X,μ)L2(M,μ)\mathrm{L}^{2}(X,\mu)\to\mathrm{L}^{2}(M,\mu) satisfying

(D.2) Afdμ=A𝔼M(f)dμfor all AM and all fL2(X,μ).\int_{A}f\,d\mu=\int_{A}\mathbb{E}_{M}(f)\,d\mu\quad\textrm{for all $A\in M$ and all $f\in\mathrm{L}^{2}(X,\mu)$}.

Positive functions are exactly those whose dot product with any characteristic function is positive. Letting gg in Eq. (D.1) range through the collection of all characteristic functions of sets in MM shows that the conditional expectation 𝔼M\mathbb{E}_{M} is positivity-preserving.

Proposition D.2.

If fL2(X,μ)f\in\mathrm{L}^{2}(X,\mu) is non-negative, f0f\geq 0, then 𝔼M(f)0\mathbb{E}_{M}(f)\geq 0.

While we defined conditional expectations as operators on L2(X,μ)\mathrm{L}^{2}(X,\mu), their domain can be extended to all of L1(X,μ)\mathrm{L}^{1}(X,\mu), making 𝔼M\mathbb{E}_{M} a contraction from L1(X,μ)\mathrm{L}^{1}(X,\mu) to L1(M,μ)\mathrm{L}^{1}(M,\mu). This is justified by the following proposition.

Proposition D.3.

The conditional expectation 𝔼M:L2(X,μ)L2(M,μ)\mathbb{E}_{M}:\mathrm{L}^{2}(X,\mu)\to\mathrm{L}^{2}(M,\mu) is a contraction when the domain and the range are endowed with the L1\mathrm{L}^{1} norms.

Proof.

If fL2(X,μ)f\in\mathrm{L}^{2}(X,\mu) is non-negative, f0f\geq 0, then Eq. D.1 yields

f1=Xfdμ=Xf1dμ=X𝔼M(f)1dμ=X𝔼M(f)dμ.\left\lVert f\right\rVert_{1}=\int_{X}f\,d\mu=\int_{X}f\cdot 1\,d\mu=\int_{X}\mathbb{E}_{M}(f)\cdot 1\,d\mu=\int_{X}\mathbb{E}_{M}(f)\,d\mu.

Since 𝔼M(f)0\mathbb{E}_{M}(f)\geq 0 by Proposition D.2, we conclude that 𝔼M(f)1=f1\left\lVert\mathbb{E}_{M}(f)\right\rVert_{1}=\left\lVert f\right\rVert_{1} for all non-negative fL2(X,μ)f\in\mathrm{L}^{2}(X,\mu).

For an arbitrary fL2(X,μ)f\in\mathrm{L}^{2}(X,\mu), set f+=max{f,0}f^{+}=\max\{f,0\} and f=max{f,0}f^{-}=\max\{-f,0\}. Note that functions f+,ff^{+},f^{-} are non-negative and belong to L2(X,μ)\mathrm{L}^{2}(X,\mu). Furthermore, f+f=ff^{+}-f^{-}=f and f+1+f1=f1\left\lVert f^{+}\right\rVert_{1}+\left\lVert f^{-}\right\rVert_{1}=\left\lVert f\right\rVert_{1}. We therefore have

𝔼M(f)1=𝔼M(f+f)1𝔼M(f+)1+𝔼M(f)1,\left\lVert\mathbb{E}_{M}(f)\right\rVert_{1}=\left\lVert\mathbb{E}_{M}(f^{+}-f^{-})\right\rVert_{1}\leq\left\lVert\mathbb{E}_{M}(f^{+})\right\rVert_{1}+\left\lVert\mathbb{E}_{M}(f^{-})\right\rVert_{1},

but the latter term is equal to f+1+f1=f1\left\lVert f^{+}\right\rVert_{1}+\left\lVert f^{-}\right\rVert_{1}=\left\lVert f\right\rVert_{1}, which finishes the proof. ∎

Remark D.4.

By the previous proposition, 𝔼M\mathbb{E}_{M} admits a (necessarily unique) extension to a contraction

𝔼M:L1(X,μ)L1(M,μ).\mathbb{E}_{M}:\mathrm{L}^{1}(X,\mu)\to\mathrm{L}^{1}(M,\mu).

Moreover, since every non-negative integrable function can be written as an increasing limit of bounded non-negative functions, the analog of Proposition D.2 continues to hold for fL1(X,μ)f\in\mathrm{L}^{1}(X,\mu).

D.2. Conditional measures

Throughout this section, we let χA:X{0,1}\chi_{A}:X\to\{0,1\} denote the characteristic function of AXA\subseteq X.

Definition D.5.

Let MM be a closed subalgebra of MAlg(X,μ)\mathrm{MAlg}(X,\mu). The MM-conditional measure of AMAlg(X,μ)A\in\mathrm{MAlg}(X,\mu), denoted by μM(A)\mu_{M}(A), is the conditional expectation of the characteristic function of AA, i.e., μM(A)=𝔼M(χA)\mu_{M}(A)=\mathbb{E}_{M}(\chi_{A}).

In particular, the conditional measure μM(A)\mu_{M}(A) is an MM-measurable function. It enjoys the following natural properties.

Proposition D.6.

Let MMAlg(X,μ)M\subseteq\mathrm{MAlg}(X,\mu) be a closed subalgebra. The following properties hold for all AMAlg(X,μ)A\in\mathrm{MAlg}(X,\mu):

  1. (1)

    μM()=0\mu_{M}(\varnothing)=0 and μM(X)=1\mu_{M}(X)=1, where 0 and 11 denote the constant maps;

  2. (2)

    μM(A)\mu_{M}(A) takes values in [0,1][0,1] and XμM(A)=μ(A)\int_{X}\mu_{M}(A)=\mu(A);

  3. (3)

    μM\mu_{M} is σ\sigma-additive: if A=nAnA=\bigsqcup_{n}A_{n}, AnMAlg(X,μ)A_{n}\in\mathrm{MAlg}(X,\mu), is a partition then

    μM(A)=nμM(An),\mu_{M}(A)=\sum_{n\in\mathbb{N}}\mu_{M}(A_{n}),

    where the convergence holds in L1(M,μ)\mathrm{L}^{1}(M,\mu);

  4. (4)

    if TAut(X,μ)T\in\mathrm{Aut}(X,\mu) fixes every element of MM, then μM(A)=μM(T(A))\mu_{M}(A)=\mu_{M}(T(A)).

Proof.

The first item is clear from the fact that both \varnothing and XX belong to MM, so their characteristic functions are fixed by 𝔼M\mathbb{E}_{M}. The second item follows from the first and positivity of the conditional expectation; the equality is a direct consequence of Eq. D.2. The third one is a consequence of the L1\mathrm{L}^{1} continuity of 𝔼M\mathbb{E}_{M} and its linearity, noting that χA=nχAn\chi_{A}=\sum_{n}\chi_{A_{n}} in L1(M,μ)\mathrm{L}^{1}(M,\mu).

Finally, the last item follows from the uniqueness of conditional expectation given by Eq. D.2. Indeed, if an automorphism TT fixes every element of MM, then

BfT1dμ=T(B)fdμ=Bfdμ for all BMAlg(X,μ),\int_{B}f\circ T^{-1}\,d\mu=\int_{T(B)}f\,d\mu=\int_{B}f\,d\mu\quad\textrm{ for all $B\in\mathrm{MAlg}(X,\mu)$},

so 𝔼M(fT1)=𝔼M(f)\mathbb{E}_{M}(f\circ T^{-1})=\mathbb{E}_{M}(f). Taking f=χAf=\chi_{A} for AMAlg(X,μ)A\in\mathrm{MAlg}(X,\mu), we conclude that μM(T(A))=μM(A)\mu_{M}(T(A))=\mu_{M}(A). ∎

D.3. Conditional measures and full groups

Conditional measures, as defined in Section D.2, are associated with closed subalgebras of MAlg(X,μ)\mathrm{MAlg}(X,\mu). Each subgroup 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) gives rise to the subalgebra of 𝔾\mathbb{G}-invariant sets, and we may therefore associate a conditional measure with the group 𝔾\mathbb{G} itself.

Definition D.7.

Let 𝔾\mathbb{G} be a subgroup of Aut(X,μ)\mathrm{Aut}(X,\mu). The closed subalgebra of 𝔾\mathbb{G}-invariant sets is denoted by M𝔾M_{\mathbb{G}} and consists of all AMAlg(X,μ)A\in\mathrm{MAlg}(X,\mu) such that gA=AgA=A for all g𝔾g\in\mathbb{G}.

By definition, 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) is ergodic if M𝔾={,X}M_{\mathbb{G}}=\{\varnothing,X\}. In this case, the M𝔾M_{\mathbb{G}}-conditional measure is the measure μ\mu itself. The following lemma is an easy consequence of the definitions of the full group generated by a subgroup (Section 3.1) and the weak topology on Aut(X,μ)\mathrm{Aut}(X,\mu).

Lemma D.8.

Let 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) be a group.

  1. (1)

    If [𝔾][\mathbb{G}] is the full group generated by 𝔾\mathbb{G}, then M𝔾=M[𝔾]M_{\mathbb{G}}=M_{[\mathbb{G}]}.

  2. (2)

    If Γ𝔾\Gamma\leq\mathbb{G} is dense in the weak topology, then MΓ=M𝔾M_{\Gamma}=M_{\mathbb{G}}.

Given a subgroup 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu), we denote the M𝔾M_{\mathbb{G}}-conditional measure simply by μ𝔾\mu_{\mathbb{G}}. Note that 𝔾\mathbb{G} is ergodic if and only if μ𝔾=μ\mu_{\mathbb{G}}=\mu.

Recall that a partial measure-preserving automorphism of (X,μ)(X,\mu) is a measure-preserving bijection φ:domφrngφ\varphi:\operatorname{\mathrm{dom}}\varphi\to\operatorname{\mathrm{rng}}\varphi between measurable subsets of XX, called the domain and the range of φ\varphi, respectively. The pseudo full group generated by a group ΓAut(X,μ)\Gamma\leq\mathrm{Aut}(X,\mu) is denoted by [[Γ]][[\Gamma\mkern 1.5mu]] and consists of all partial automorphisms φ:domφrngφ\varphi:\operatorname{\mathrm{dom}}\varphi\to\operatorname{\mathrm{rng}}\varphi for which there exists a partition domφ=nAn\operatorname{\mathrm{dom}}\varphi=\bigsqcup_{n}A_{n} and elements γnΓ\gamma_{n}\in\Gamma such that φAn=γnAn\varphi\restriction_{A_{n}}=\gamma_{n}\restriction_{A_{n}} for all nn. Elements of [[Γ]][[\Gamma\mkern 1.5mu]] automatically preserve the conditional measure μΓ\mu_{\Gamma} in view of item (4) of Proposition D.6.

Lemma D.9.

Let 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) be a group and let A,BMAlg(X,μ)A,B\in\mathrm{MAlg}(X,\mu) satisfy μ𝔾(A)=μ𝔾(B)\mu_{\mathbb{G}}(A)=\mu_{\mathbb{G}}(B). There exists an element φ[[𝔾]]\varphi\in[[\mathbb{G}\mkern 1.5mu]] such that domφ=A\operatorname{\mathrm{dom}}\varphi=A and rngφ=B\operatorname{\mathrm{rng}}\varphi=B.

Proof.

Let Γ={γn:n}\Gamma=\{\gamma_{n}:n\in\mathbb{N}\} be a countable weakly dense subgroup of 𝔾\mathbb{G}. Note that μΓ(A)=μ𝔾(A)=μ𝔾(B)=μΓ(B)\mu_{\Gamma}(A)=\mu_{\mathbb{G}}(A)=\mu_{\mathbb{G}}(B)=\mu_{\Gamma}(B) by Lemma D.8, and also clearly [[Γ]][[𝔾]][[\Gamma\mkern 1.5mu]]\leq[[\mathbb{G}\mkern 1.5mu]].

We define inductively sequences (An)n(A_{n})_{n} and (Bn)n(B_{n})_{n} of subsets of AA and BB respectively by setting A0=Aγ01BA_{0}=A\cap\gamma_{0}^{-1}B and B0=γ0A0B_{0}=\gamma_{0}A_{0}, and then putting for n1n\geq 1

An=(Am<nAm)γn1(Bm<nBm) and Bn=γnAn.A_{n}=\Big{(}A\setminus\bigcup_{m<n}A_{m}\Big{)}\cap\gamma_{n}^{-1}\Big{(}B\setminus\bigcup_{m<n}B_{m}\Big{)}\quad\text{ and }\quad B_{n}=\gamma_{n}A_{n}.

By construction, the sets AnA_{n} are pairwise disjoint subsets of AA, γnAn=Bn\gamma_{n}A_{n}=B_{n}, and the sets BnB_{n} are pairwise disjoint subsets of BB. We claim that φ=nγnAn\varphi=\bigsqcup_{n}\gamma_{n}\restriction_{A_{n}} is the desired element of [[𝔾]][[\mathbb{G}\mkern 1.5mu]].

Suppose towards a contradiction that either domφA\operatorname{\mathrm{dom}}\varphi\neq A or rngφB\operatorname{\mathrm{rng}}\varphi\neq B. Since Γ\Gamma preserves μΓ\mu_{\Gamma} and μΓ(A)=μΓ(B)\mu_{\Gamma}(A)=\mu_{\Gamma}(B), the sets AdomφA\setminus\operatorname{\mathrm{dom}}\varphi and BrngφB\setminus\operatorname{\mathrm{rng}}\varphi have the same MΓM_{\Gamma}-conditional measure, which is not constantly equal to zero. The set A~=γΓγ(Adomφ)\tilde{A}=\bigcup_{\gamma\in\Gamma}\gamma(A\setminus\operatorname{\mathrm{dom}}\varphi) is Γ\Gamma-invariant and non zero. Its conditional measure is therefore the characteristic function χA~\chi_{\tilde{A}}, which must be greater than or equal to μΓ(Adomφ)=μΓ(Brngφ)\mu_{\Gamma}(A\setminus\operatorname{\mathrm{dom}}\varphi)=\mu_{\Gamma}(B\setminus\operatorname{\mathrm{rng}}\varphi). We conclude that BrngφγΓγ(Adomφ)B\setminus\operatorname{\mathrm{rng}}\varphi\subseteq\bigcup_{\gamma\in\Gamma}\gamma(A\setminus\operatorname{\mathrm{dom}}\varphi). In particular, there is the first nn\in\mathbb{N} such that (Adomφ)γn1(Brngφ)(A\setminus\operatorname{\mathrm{dom}}\varphi)\cap\gamma_{n}^{-1}(B\setminus\operatorname{\mathrm{rng}}\varphi) is non zero. By construction, this set should be a subset of AnA_{n}, yielding the desired contradiction. ∎

Proposition D.10.

Let 𝔾\mathbb{G} be a full subgroup of Aut(X,μ)\mathrm{Aut}(X,\mu). The following conditions are equivalent for all A,BMAlg(X,μ)A,B\in\mathrm{MAlg}(X,\mu):

  1. (1)

    μ𝔾(A)=μ𝔾(B)\mu_{\mathbb{G}}(A)=\mu_{\mathbb{G}}(B);

  2. (2)

    there is T𝔾T\in\mathbb{G} such that T(A)=BT(A)=B.

  3. (3)

    there is an involution T𝔾T\in\mathbb{G} such that T(A)=BT(A)=B and suppT=AB\operatorname*{supp}T=A\bigtriangleup B.

Proof.

The implication (2)\Rightarrow(1) is a direct consequence of the definition of M𝔾M_{\mathbb{G}} along with the item (4) of Proposition D.6. Also (3)\Rightarrow(2) is evident.

We now prove the implication (1)\Rightarrow(3). The assumption μ𝔾(A)=μ𝔾(A)\mu_{\mathbb{G}}(A)=\mu_{\mathbb{G}}(A) guarantees that μ𝔾(AB)=μ𝔾(BA)\mu_{\mathbb{G}}(A\setminus B)=\mu_{\mathbb{G}}(B\setminus A). Lemma D.9 applies and produces an element φ[[𝔾]]\varphi\in[[\mathbb{G}\mkern 1.5mu]] such that φ(AB)=BA\varphi(A\setminus B)=B\setminus A. The required involution TT is then given by φφ1idX(AB)\varphi\sqcup\varphi^{-1}\sqcup\mathrm{id}_{X\setminus(A\bigtriangleup B)}. ∎

D.4. Aperiodicity

A countable subgroup ΓAut(X,μ)\Gamma\leq\mathrm{Aut}(X,\mu) is called aperiodic if almost all the orbits of some (equivalently, any) realization of its action on (X,μ)(X,\mu) are infinite. The so-called Maharam’s lemma provides a characterization of aperiodicity in a purely measure-algebraic way. We begin by formulating a variant of the standard Marker Lemma for countable Borel equivalence relations (see, for instance, [KM04, Lemma 6.7]).

Lemma D.11.

Let ΓX\Gamma\curvearrowright X be a Borel action of a countable group on a standard Borel space XX. For every Borel CXC\subseteq X, there is a decreasing sequence (Cn)n(C_{n})_{n} of Borel subsets of CC such that CΓCnC\subseteq\Gamma\cdot C_{n} for each nn, and the set nCn\bigcap_{n}C_{n} intersects the Γ\Gamma-orbit of every xXx\in X in at most one point. Furthermore, if all orbits of Γ\Gamma are infinite, sets CnC_{n} can be chosen to have the empty intersection, nCn=\bigcap_{n}C_{n}=\varnothing.

The following result is essentially due to H. Dye [Dye59], where it is called Maharam’s lemma.

Theorem D.12 (Maharam’s lemma).

Let ΓAut(X,μ)\Gamma\leq\mathrm{Aut}(X,\mu) be a countable subgroup. The following are equivalent:

  1. (1)

    Γ\Gamma is aperiodic;

  2. (2)

    for any AMAlg(X,μ)A\in\mathrm{MAlg}(X,\mu) and any MΓM_{\Gamma}-measurable function f:X[0,1]f:X\to[0,1] satisfying fμΓ(A)f\leq\mu_{\Gamma}(A), there is BAB\subseteq A, BMAlg(X,μ)B\in\mathrm{MAlg}(X,\mu), such that μΓ(B)=f\mu_{\Gamma}(B)=f.

Proof.

Let us begin with the easier (2)\Rightarrow(1), which is proved by the contrapositive. Assume that (1) does not hold and Γ\Gamma is not aperiodic. Let nn\in\mathbb{N} be such that the Γ\Gamma-invariant set Xn={xX:|Γx|=n}X_{n}=\{x\in X:\left\lvert\Gamma\cdot x\right\rvert=n\} has non-zero measure. We may assume that XX bears a Borel total order (for instance, by identifying XX with [0,1][0,1]). Let A={xXn:x=max{Γx}}A=\{x\in X_{n}:x=\max\{\Gamma\cdot x\}\} be the set of maximal points of the nn-element Γ\Gamma-orbits and set φ\varphi, domφ=XnA\operatorname{\mathrm{dom}}\varphi=X_{n}\setminus A, to be the element of the pseudo full group [[Γ]][[\Gamma\mkern 1.5mu]] that takes every xXnAx\in X_{n}\setminus A to its <<-successor in the orbit Γx\Gamma\cdot x. Given any BAB\subseteq A, the set k=0n1φk(B)\bigsqcup_{k=0}^{n-1}\varphi^{-k}(B) is Γ\Gamma-invariant, hence μΓ(k=0n1φk(B))\mu_{\Gamma}(\bigsqcup_{k=0}^{n-1}\varphi^{-k}(B)) takes values in {0,1}\{0,1\}. Also

μΓ(k=0n1φk(B))=k=0n1μΓ(φk(B))=nμΓ(B),\mu_{\Gamma}\big{(}\bigsqcup_{k=0}^{n-1}\varphi^{-k}(B)\big{)}=\sum_{k=0}^{n-1}\mu_{\Gamma}\big{(}\varphi^{-k}(B)\big{)}=n\mu_{\Gamma}(B),

where the last equality is a consequence of Proposition D.6. We conclude that μΓ(B)\mu_{\Gamma}(B) necessarily takes values in {0,1n}\{0,\frac{1}{n}\}, which contradicts (2).

We now assume that Γ\Gamma is aperiodic and prove the direct implication (1)\Rightarrow(2). The argument is based on the following crucial claim.

Claim.

For every CMAlg(X,μ)C\in\mathrm{MAlg}(X,\mu), for every MΓM_{\Gamma}-measurable not almost surely zero f:X[0,1]f:X\to[0,1] such that fμΓ(C)f\leq\mu_{\Gamma}(C), there is a non zero BCB\subseteq C satisfying μΓ(B)f\mu_{\Gamma}(B)\leq f.

Proof of the claim.

Let (Cn)n(C_{n})_{n} be a vanishing sequence of subsets of CC given by Lemma D.11. Note that μΓ(Cn)0\mu_{\Gamma}(C_{n})\to 0 in L1\mathrm{L}^{1}, since nCn=\bigcap_{n}C_{n}=\varnothing and the CnC_{n}’s are decreasing. Passing to a subsequence, we may assume that convergence μΓ(Cn)0\mu_{\Gamma}(C_{n})\to 0 holds pointwise. Set Bn={xCn:μΓ(Cn)(x)f(x)}B_{n}=\{x\in C_{n}:\mu_{\Gamma}(C_{n})(x)\leq f(x)\} and note that μΓ(Bn)μΓ(Cn)\mu_{\Gamma}(B_{n})\leq\mu_{\Gamma}(C_{n}) and therefore μΓ(Bn)f\mu_{\Gamma}(B_{n})\leq f.

Pointwise convergence μΓ(Cn)0\mu_{\Gamma}(C_{n})\to 0 guarantees existence of an index nn such that μ(Bn)>0\mu(B_{n})>0, and so the set B=BnB=B_{n} is as required. ∎

The conclusion of the theorem now follows from a standard application of Zorn’s lemma111A more constructive version of the whole argument can be found in [LM14, Prop. D.1].. Indeed, the latter provides a maximal family (Bi)iI(B_{i})_{i\in I} of pairwise disjoint positive measure elements of MAlg(X,μ)\mathrm{MAlg}(X,\mu) contained in AA and satisfying iIμΓ(Bi)f\sum_{i\in I}\mu_{\Gamma}(B_{i})\leq f. The index set II has to be countable, and if B=iIBiB=\bigsqcup_{i\in I}B_{i} then μΓ(B)=iIμΓ(Bi)f\mu_{\Gamma}(B)=\sum_{i\in I}\mu_{\Gamma}(B_{i})\leq f. Assume towards a contradiction that μΓ(B)\mu_{\Gamma}(B) is not equal to ff almost everywhere, and use the previous claim to get a non null BABB^{\prime}\subseteq A\setminus B with μΓ(B)fμΓ(B)\mu_{\Gamma}(B^{\prime})\leq f-\mu_{\Gamma}(B), contradicting the maximality of (Bi)iI(B_{i})_{i\in I}. Therefore μΓ(B)=f\mu_{\Gamma}(B)=f as claimed. ∎

We conclude this appendix with a useful consequence of aperiodicity.

Lemma D.13.

Let 𝔾Aut(X,μ)\mathbb{G}\leq\mathrm{Aut}(X,\mu) be an aperiodic full group. For each set BMAlg(X,μ)B\in\mathrm{MAlg}(X,\mu), there is an involution U𝔾U\in\mathbb{G} whose support is equal to BB.

Proof.

Theorem D.12 gives ABA\subseteq B such that μ𝔾(A)=μ𝔾(B)/2\mu_{\mathbb{G}}(A)=\mu_{\mathbb{G}}(B)/2. We then have μ𝔾(BA)=μ𝔾(B)μ𝔾(B)/2=μ𝔾(A)\mu_{\mathbb{G}}(B\setminus A)=\mu_{\mathbb{G}}(B)-\mu_{\mathbb{G}}(B)/2=\mu_{\mathbb{G}}(A), and item (3) of Proposition D.10 provides an involution T𝔾T\in\mathbb{G} satisfying T(BA)=AT(B\setminus A)=A and suppT=(BA)A=B\operatorname*{supp}T=(B\setminus A)\bigtriangleup A=B. ∎

Remark D.14.

Lemma D.13, in fact, characterizes aperiodicity of full groups: if 𝔾\mathbb{G} is not aperiodic, then there is some BMAlg(X,μ)B\in\mathrm{MAlg}(X,\mu) which is not the support of any involution since its M𝔾M_{\mathbb{G}}-conditional measure cannot be split in half (see the proof of the direct implication in Theorem D.12).

References

  • [ADR00] C. Anantharaman-Delaroche and J. Renault, Amenable groupoids, Monographies de L’Enseignement Mathématique [Monographs of L’Enseignement Mathématique], vol. 36, L’Enseignement Mathématique, Geneva, 2000, With a foreword by Georges Skandalis and Appendix B by E. Germain. MR 1799683
  • [Aus16] Tim Austin, Behaviour of Entropy Under Bounded and Integrable Orbit Equivalence, Geometric and Functional Analysis 26 (2016), no. 6, 1483–1525.
  • [Ban32] Stefan Banach, Théorie des opérations linéaires, Warsaw, 1932.
  • [BdlHV08] Bachir Bekka, Pierre de la Harpe, and Alain Valette, Kazhdan’s property (T), vol. 11, Cambridge University Press, Cambridge, 2008 (English).
  • [Bec13] Howard Becker, Cocycles and continuity, Trans. Amer. Math. Soc. 365 (2013), no. 2, 671–719. MR 2995370
  • [Bel68] R. M. Belinskaja, Partitionings of a Lebesgue space into trajectories which may be defined by ergodic automorphisms, Funkcional. Anal. i Priložen. 2 (1968), no. 3, 4–16. MR 0245756
  • [BK96] Howard Becker and Alexander S. Kechris, The descriptive set theory of Polish group actions, London Mathematical Society Lecture Note Series, vol. 232, Cambridge University Press, Cambridge, 1996. MR 1425877
  • [BO10] Nicholas H. Bingham and Adam J. Ostaszewski, Normed versus topological groups: dichotomy and duality, Dissertationes Math. 472 (2010), 138. MR 2743093
  • [CFW81] Alain Connes, Jacob Feldman, and Benjamin Weiss, An amenable equivalence relation is generated by a single transformation, Ergodic Theory Dynam. Systems 1 (1981), no. 4, 431–450 (1982). MR 662736
  • [CJMT22] Alessandro Carderi, Matthieu Joseph, François Le Maître, and Romain Tessera, Belinskaya’s theorem is optimal, 2022.
  • [CLM16] Alessandro Carderi and François Le Maître, More Polish full groups, Topology Appl. 202 (2016), 80–105. MR 3464151
  • [CLM18] by same author, Orbit full groups for locally compact groups, Trans. Amer. Math. Soc. 370 (2018), no. 4, 2321–2349. MR 3748570
  • [DHU08] Manfred Droste, W. Charles Holland, and Georg Ulbrich, On full groups of measure-preserving and ergodic transformations with uncountable cofinalities, Bulletin of the London Mathematical Society 40 (2008), no. 3, 463–472.
  • [Dye59] Henry A. Dye, On Groups of Measure Preserving Transformations. I, American Journal of Mathematics 81 (1959), no. 1, 119–159.
  • [FHM78] Jacob Feldman, Peter Hahn, and Calvin C. Moore, Orbit structure and countable sections for actions of continuous groups, Adv. in Math. 28 (1978), no. 3, 186–230. MR 492061
  • [Fre04] David H. Fremlin, Measure theory. Vol. 3, Torres Fremlin, Colchester, 2004, Measure algebras, Corrected second printing of the 2002 original. MR 2459668 (2011a:28003)
  • [Fre06] by same author, Measure theory. Vol. 4, Torres Fremlin, Colchester, 2006, Topological measure spaces. Part I, II, Corrected second printing of the 2003 original. MR 2462372
  • [GK21] Marlies Gerber and Philipp Kunde, Anti-classification results for the Kakutani equivalence relation, September 2021.
  • [GM89] Siegfried Graf and R. Daniel Mauldin, A classification of disintegrations of measures, Measure and measurable dynamics (Rochester, NY, 1987), Contemp. Math., vol. 94, Amer. Math. Soc., Providence, RI, 1989, pp. 147–158. MR 1012985
  • [GP02] Thierry Giordano and Vladimir Pestov, Some extremely amenable groups, C. R. Math. Acad. Sci. Paris 334 (2002), no. 4, 273–278. MR 1891002
  • [GPS99] Thierry Giordano, Ian F. Putnam, and Christian F. Skau, Full groups of Cantor minimal systems, Israel J. Math. 111 (1999), 285–320. MR 1710743
  • [GTW05] Eli Glasner, Boris Tsirelson, and Benjamin Weiss, The automorphism group of the Gaussian measure cannot act pointwise, Israel Journal of Mathematics 148 (2005), no. 1, 305–329.
  • [Hal17] Paul R. Halmos, Lectures on ergodic theory, Dover Publications, Mineola, NY, 2017.
  • [Kat75] Anatole B. Katok, Time change, monotone equivalence, and standard dynamical systems, Doklady Akademii Nauk SSSR 223 (1975), no. 4, 789–792. MR 0412383
  • [Kat77] by same author, Monotone equivalence in ergodic theory, Izvestiya Akademii Nauk SSSR. Seriya Matematicheskaya 41 (1977), no. 1, 104–157, 231. MR 0442195
  • [Kec92] Alexander S. Kechris, Countable sections for locally compact group actions, Ergodic Theory Dynam. Systems 12 (1992), no. 2, 283–295. MR 1176624
  • [Kec95] by same author, Classical descriptive set theory, Graduate Texts in Mathematics, vol. 156, Springer-Verlag, New York, 1995.
  • [Kec10] by same author, Global aspects of ergodic group actions, Mathematical Surveys and Monographs, vol. 160, American Mathematical Society, Providence, RI, 2010. MR 2583950
  • [KM04] Alexander S. Kechris and Benjamin D. Miller, Topics in orbit equivalence, Lecture Notes in Mathematics, no. 1852, Springer, Berlin ; New York, 2004.
  • [KPV15] David Kyed, Henrik Petersen, and Stefaan Vaes, L2-Betti numbers of locally compact groups and their cross section equivalence relations, Trans. Amer. Math. Soc. 367 (2015), no. 7, 4917–4956.
  • [Kre85] Ulrich Krengel, Ergodic theorems, De Gruyter Studies in Mathematics, vol. 6, Walter de Gruyter & Co., Berlin, 1985, With a supplement by Antoine Brunel. MR 797411
  • [KST99] Alexander S. Kechris, Sławomir Solecki, and Stevo Todorcevic, Borel chromatic numbers, Adv. Math. 141 (1999), no. 1, 1–44. MR 1667145
  • [Kun23] Philipp Kunde, Anti-classification results for weakly mixing diffeomorphisms, March 2023.
  • [LM14] François Le Maître, Sur les groupes pleins préservant une mesure de probabilité, Ph.D. thesis, ENS Lyon, 2014.
  • [LM16] François Le Maître, On full groups of non-ergodic probability-measure-preserving equivalence relations, Ergodic Theory Dynam. Systems 36 (2016), no. 7, 2218–2245. MR 3568978
  • [LM18] by same author, On a measurable analogue of small topological full groups, Adv. Math. 332 (2018), 235–286. MR 3810253
  • [LM21] by same author, On a measurable analogue of small topological full groups II, Ann. Inst. Fourier (Grenoble) 71 (2021), no. 5, 1885–1927. MR 4398251
  • [Mac62] George W. Mackey, Point realizations of transformation groups, Illinois J. Math. 6 (1962), 327–335. MR 143874
  • [Mah50] Dorothy Maharam, Decompositions of measure algebras and spaces, Trans. Amer. Math. Soc. 69 (1950), 142–160. MR 36817
  • [Mah84] by same author, On the planar representation of a measurable subfield, Measure theory, Oberwolfach 1983 (Oberwolfach, 1983), Lecture Notes in Math., vol. 1089, Springer, Berlin, 1984, pp. 47–57. MR 786682
  • [Mil77] Douglas E. Miller, On the measurability of orbits in Borel actions, Proc. Amer. Math. Soc. 63 (1977), no. 1, 165–170. MR 440519
  • [Mil04] Benjamin D. Miller, Full groups, classification, and equivalence relations, Ph.D. thesis, University of California, Berkeley, 2004.
  • [Nac65] Leopoldo Nachbin, The Haar integral, D. Van Nostrand Co., Inc., Princeton, N.J.-Toronto-London, 1965. MR 0175995
  • [Nek19] Volodymyr Nekrashevych, Simple groups of dynamical origin, Ergodic Theory Dynam. Systems 39 (2019), no. 3, 707–732. MR 3904185
  • [ORW82] Donald S. Ornstein, Daniel J. Rudolph, and Benjamin Weiss, Equivalence of measure preserving transformations, Memoirs of the American Mathematical Society 37 (1982), no. 262, xii+116. MR 653094
  • [PS17] Vladimir G. Pestov and Friedrich Martin Schneider, On amenability and groups of measurable maps, J. Funct. Anal. 273 (2017), no. 12, 3859–3874. MR 3711882
  • [Ros21] Christian Rosendal, Coarse Geometry of Topological Groups, Cambridge Tracts in Mathematics, Cambridge University Press, Cambridge, 2021.
  • [Ros22] by same author, Coarse geometry of topological groups, Cambridge Tracts in Mathematics, vol. 223, Cambridge University Press, Cambridge, 2022. MR 4327092
  • [RS98] Guyan Robertson and Tim Steger, Negative definite kernels and a dynamical characterization of property (T) for countable groups, Ergodic Theory and Dynamical Systems 18 (1998), no. 1, 247–253 (en).
  • [Rud76] Daniel Rudolph, A two-valued step coding for ergodic flows, Mathematische Zeitschrift 150 (1976), no. 3, 201–220.
  • [Ryz85] V. V. Ryzhikov, Representation of transformations preserving the lebesgue measure in the form of periodic transformations, Mathematical notes of the Academy of Sciences of the USSR 38 (1985), no. 6, 978–981.
  • [Slu17] Konstantin Slutsky, Lebesgue orbit equivalence of multidimensional Borel flows: a picturebook of tilings, Ergodic Theory Dynam. Systems 37 (2017), no. 6, 1966–1996. MR 3681992
  • [Slu19] by same author, Regular cross sections of Borel flows, Journal of the European Mathematical Society 21 (2019), no. 7, 1985–2050.
  • [Str74] Raimond A. Struble, Metrics in locally compact groups, Compositio Math. 28 (1974), 217–222. MR 348037
  • [Var63] Veeravalli S. Varadarajan, Groups of automorphisms of Borel spaces, Trans. Amer. Math. Soc. 109 (1963), 191–220. MR 159923