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

Dynamical obstructions to classification by (co)homology and other TSI-group invariants

Shaun Allison Department of Mathematical Sciences, Carnegie Mellon University, Pittsburgh, PA 15213 [email protected] https://www.math.cmu.edu/ sallison/  and  Aristotelis Panagiotopoulos Mathematics Department, Caltech, 1200 E. California Blvd, Pasadena, CA 91125 [email protected] http://www.its.caltech.edu/ panagio/
Abstract.

In the spirit of Hjorth’s turbulence theory, we introduce “unbalancedness”: a new dynamical obstruction to classifying orbit equivalence relations by actions of Polish groups which admit a two side invariant metric (TSI). Since abelian groups are TSI, unbalancedness can be used for identifying which classification problems cannot be solved by classical homology and cohomology theories.

In terms of applications, we show that Morita equivalence of continuous-trace CC^{*}-algebras, as well as isomorphism of Hermitian line bundles, are not classifiable by actions of TSI groups. In the process, we show that the Wreath product of any two non-compact subgroups of SS_{\infty} admits an action whose orbit equivalence relation is generically ergodic against any action of a TSI group and we deduce that there is an orbit equivalence relation of a CLI group which is not classifiable by actions of TSI groups.

Key words and phrases:
Polish group, invariant metric, generically ergodic, turbulence, TSI, CLI, Borel reduction, continuous-trace CC^{*}-algebra, Morita equivalence, Hermitian line bundle
2000 Mathematics Subject Classification:
Primary 54H05, 37B02, 54H11; Secondary 46L35, 55R15
We are grateful to A. Shani, M. Lupini, J. Bergfalk, and A.S. Kechris for all the useful and inspiring discussions, as well as to S. Coskey and J.D. Clemens for sharing an early draft of [CC] with us. We would also like to thank the anonymous referee for their valuable comments and for raising our attention to several subtle errors in an earlier version of this paper. Finally, we want to acknowledge the hospitality and financial support of the California Institute of Technology during the visit of S.A. in the winter of 2020

1. Introduction

1.1. Classification by (co)homological invariants

One of the leading questions in many mathematical research programs is whether a certain classification problem admits a “satisfactory” solution. What constitutes a satisfactory solution depends of course on the context, and it is often subject to change when the original goals are deemed hopeless. Invariant descriptive set theory provides a formal framework for measuring the complexity of classification problems and for showing which types of invariants are inadequate for complete classification.

Hjorth’s turbulence theory is one of the biggest accomplishments in this direction as it provides obstructions to classification by countable structures, i.e., for classification using only isomorphism types of countable structures as invariants. Historically, this type of classification played important role in ergodic theory since von Neumann’s seminal paper [vN32], where he shows that ergodic measure-preserving transformations of discrete spectrum are completely classified up to isomorphism by their (countable) spectrum. This led to the hope that, by adding further structure on these countable invariants, one could classify all ergodic measure-preserving transformations via countable structures. However, 70 years after the publication of [vN32], the newly developed theory of turbulence [Hjo00] was used to establish that this hope was too optimistic; see [Hjo01, FW04].

Another very common type of classification that one encounters in mathematical practice is classification by (co)homological invariants, i.e., classification using elements of homology or cohomology groups of some appropriate chain complex. For example:

  1. (1)

    there is an assignment pc(p)p\mapsto c(p), from Hermitian line bundles p:EBp\colon E\to B over a locally compact metrizable space BB, to the Čech cohomology group H2(B)\mathrm{H}^{2}(B) of BB, so that pp and pp^{\prime} are isomorphic over BB iff c(p)=c(p)c(p)=c(p^{\prime}); see [RW98].

  2. (2)

    there is an assignment 𝒜c(𝒜)\mathcal{A}\mapsto c(\mathcal{A}), from continuous-trace CC^{*}-algebras 𝒜\mathcal{A} with locally compact metrizable spectrum SS, to the Čech cohomology group H3(S)\mathrm{H}^{3}(S) of SS, so that 𝒜\mathcal{A} and 𝒜\mathcal{A}^{\prime} are Morita equivalent over SS, iff c(𝒜)=c(𝒜)c(\mathcal{A})=c(\mathcal{A}^{\prime}); see [Bla09].

  3. (3)

    there is an assignment ec(e)e\mapsto c(e), from group extensions e:E2e\colon E\to\mathbb{Q}_{2} of the dyadic rationals 2\mathbb{Q}_{2} by \mathbb{Z}, to the Steenrod homology group H0(Σ)\mathrm{H}_{0}(\Sigma) of the character Σ\Sigma of 2\mathbb{Q}_{2}, so that ee and ee^{\prime} are isomorphic iff c(e)=c(e)c(e)=c(e^{\prime}); see [EM42].

Up until recently, classification by (co)homological invariants had not been considered from the perspective of invariant descriptive set theory. However, as a consequence of the results in [BLP19], classification by the above (co)homological invariants forms a well defined complexity class in the standard setup of invariant descriptive set theory. This complexity class is entirely contained within the class of all classification problems which are classifiable by abelian group actions; see Problem 1. As a consequence, Corollary 1.4 below provides a new anti-classification criterion, in the spirit of Hjorth’s turbulence theory, which can be used as an obstruction to classification by (co)homological invariants. To illustrate its use, we will apply it to show that the coordinate free versions of Hermitian line bundle isomorphism and of Morita equivalence between continuous-trace CC^{*}-algebras cannot be classified by (co)homological invariants as in the examples (1) and (2) above.

1.2. Classification problems

The formal framework that is often used for measuring the complexity of classification problems is the Borel reduction hierarchy. Formally, a classification problem is a pair (X,E)(X,E), where XX is a Polish space and EE is an analytic equivalence relation. A classification problem (X,E)(X,E) is considered to be of “less or equal complexity” to the classification problem (Y,F)(Y,F), if there is a Borel reduction from EE to FF, i.e., a Borel map f:XYf\colon X\to Y, so that for all x,xXx,x^{\prime}\in X we have:

xExf(x)Ff(x).xEx^{\prime}\iff f(x)Ff(x^{\prime}).

At the lower end of this complexity hierarchy we have—in increasing complexity—the classification problems which are: concretely classifiable; essentially countable; and classifiable by countable structures. A classification problem (X,E)(X,E) is concrete classifiable if EE Borel reduces to the equality relation of some Polish space. It is essentially countable if it Borel reduces to a Borel equivalence relation FF which has equivalence classes of countable size. We finally say that (X,E)(X,E) is classifiable by countable structures if (X,E)(X,E) Borel reduces to the problem (X,iso)(X_{\mathcal{L}},\simeq_{\mathrm{iso}}), of classifying, up to isomorphism, all countable \mathcal{L}-structures (graphs, groups, rings, etc.) of some fixed language \mathcal{L}.

In order to show that two classification problems differ in complexity, it is imperative to have a basic obstruction theory for Borel reductions. These obstructions often come from dynamics and they directly apply to classification problems of the form (X,EXG)(X,E^{G}_{X}), where EXGE^{G}_{X} is the orbit equivalence relation of the continuous action of a Polish group GG on a Polish space XX. A classical dynamical obstruction to concrete classification is generic ergodicity; see [Gao09]. Similarly, Hjorth’s turbulence theory [Hjo00] provides obstructions to classifiability by countable structures. Finally, storminess [Hjo05], as well as local approximability [KMPZ20], are both dynamical obstruction to being essentially countable.

These dynamical obstructons can be seen to answer the following general problem that was considered in [LP18]: if 𝒞\mathcal{C} is a class of Polish groups, then we say that (X,E)(X,E) is classifiable by 𝒞\mathcal{C}-group actions if it is Borel reducible to an orbit equivalence relation (Y,EYH)(Y,E^{H}_{Y}), where HH is a group from 𝒞\mathcal{C}.

Problem 1.

Given a class 𝒞\mathcal{C} of Polish groups, which dynamical conditions on a Polish GG-space XX ensure that (X,EXG)(X,E^{G}_{X}) is not classifiable by 𝒞\mathcal{C}-group actions?

Indeed, generic ergodicity provides an answer for 𝒞={compact Polish groups}\mathcal{C}=\{\text{compact Polish groups}\}; turbulence for 𝒞={non-Archimedean Polish groups}\mathcal{C}=\{\text{non-Archimedean Polish groups}\}; while storminess and local-approximability for the class 𝒞={locally-compact Polish groups}\mathcal{C}=\{\text{locally-compact Polish groups}\}. More recently, an answer to this problem for the case where 𝒞\mathcal{C} is the class of all Polish CLI groups has been given [LP18]. Recall that a Polish group is CLI if it admits a complete and left-invariant metric.

PolishCLITSInon-Archimedeanlocally compactcompact
Figure 1. Classification by 𝒞\mathcal{C}-group actions for various group classes 𝒞\mathcal{C}.

The main goal of this paper is to provide an answer to the above problem for the case where 𝒞\mathcal{C} is the class of all Polish groups which admit a two side invariant (TSI) metric. Since every abelian group is TSI the obstructions we introduce can be used for showing when a classification problem is not classifiable by (co)homological invariants.

1.3. Definitions and main results

A Polish space XX is a separable, completely metrizable topological space. A Polish group GG is a topological group whose topology is Polish. Let dd be a metric on GG that is compatible with the topology. We say that dd is left invariant if d(gh,gh)=d(h,h)d(gh,gh^{\prime})=d(h,h^{\prime}), for all g,h,hGg,h,h^{\prime}\in G and right invariant if d(hg,hg)=d(h,h)d(hg,h^{\prime}g)=d(h,h^{\prime}), for all g,h,hGg,h,h^{\prime}\in G. We say that dd is two side invariant, if it is both left and right invariant. We say that a Polish group is TSI, if it admits a two sided invariant metric which is compatible with the topology. Such groups are often called balanced since, by a theorem of Klee, they are precisely the Polish groups which admit a neighborhood basis of the identity consisting of conjugation-invariant open sets. We say that GG is non-Archimedean, if it admits a basis of open neighborhoods of the identity consisting of open subgroups. A Polish GG-space is a Polish space XX together with a continuous left action of a Polish group GG on XX. If xXx\in X, we write [x][x] to denote the orbit GxGx of xx. We denote by EXGE^{G}_{X} the associated orbit equivalence relation:

xEXGy[x]=[y].xE^{G}_{X}y\iff[x]=[y].
Definition 1.1.

Let XX be a Polish GG-space and let x,yXx,y\in X. We write xyx\leftrightsquigarrow y, if for every open neighborhood VV of the identity of GG and every open set UXU\subseteq X having nonempty intersection with the orbit of xx or yy, there exist gx,gyGg^{x},g^{y}\in G with gxxUg^{x}x\in U and gyyUg^{y}y\in U, so that:

(gyy)V(gxx)¯ and (gxx)V(gyy)¯.(g^{y}y)\in\overline{V(g^{x}x)}\text{ and }(g^{x}x)\in\overline{V(g^{y}y)}.

It is clear that xyx\leftrightsquigarrow y implies that Gx¯=Gy¯\overline{Gx}=\overline{Gy}. It is also clear that this is a symmetric relation that is invariant under the action of GG; that is, if g,hGg,h\in G, we have that xyx\leftrightsquigarrow y if and only if gxhygx\leftrightsquigarrow hy. As a consequence we can write [x][y][x]\leftrightsquigarrow[y] whenever xyx\leftrightsquigarrow y, without ambiguity.

Definition 1.2.

Let XX be a Polish GG-space and let CXC\subseteq X be a GG-invariant set. The unbalanced graph associated to the action of GG on CC is the graph (C/G,)(C/G,\leftrightsquigarrow), where C/G:={[x]xC}C/G:=\{[x]\mid x\in C\}, and [x][y][x]\leftrightsquigarrow[y] if and only if xyx\leftrightsquigarrow y.

We say that (C/G,)(C/G,\leftrightsquigarrow) is connected, if for every x,yCx,y\in C there is a path in C/GC/G from [x][x] to [y][y]; that is, a sequence x0,,xn1Cx_{0},\ldots,x_{n-1}\in C so that x=x0,xn1=yx=x_{0},x_{n-1}=y and xi1xix_{i-1}\leftrightsquigarrow x_{i}, for all 0<i<n0<i<n. We say that (X/G,)(X/G,\leftrightsquigarrow) is generically semi-connected, if for every GG-invariant comeager CXC\subseteq X, there is a comeager DCD\subseteq C so that for every x,yDx,y\in D there is a path between [x][x] and [y][y] in (C/G,)(C/G,\leftrightsquigarrow). We say that a Polish GG-space XX is generically unbalanced, if it has meager orbits and (X/G,)(X/G,\leftrightsquigarrow) is generically semi-connected.

Let (X,E),(Y,F)(X,E),(Y,F) be two classification problems. A Baire-measurable homomorphism from EE to FF is a a Baire-measurable map f:XYf:X\rightarrow Y so that xExf(x)Ff(x)xEx^{\prime}\implies f(x)Ff(x^{\prime}). It is a Baire-measurable reduction, if we additionally have f(x)Ff(x)xExf(x)Ff(x^{\prime})\implies xEx^{\prime}. We say that (X,E)(X,E) is generically FF-ergodic, if for every Baire-measurable homomorphism from EE to FF there is a comeager subset CXC\subseteq X so that f(x)F(x)f(x)F(x^{\prime}) for all x,xCx,x^{\prime}\in C. The following theorem and its corollary are the main results of this paper.

Theorem 1.3.

Let XX be a Polish GG-space and let YY be a Polish HH-space, where HH is TSI. If (X/G,)(X/G,\leftrightsquigarrow) is generically semi-connected, then EXGE^{G}_{X} is generically EYHE^{H}_{Y}-ergodic.

Corollary 1.4 (Obstruction to classification by TSI).

If the Polish GG-space XX is generically unbalanced, then the orbit equivalence relation EXGE^{G}_{X} is not classifiable by TSI-group actions.

We now turn to applications. In [CC], a new family of jump operators EE[Γ]E\mapsto E^{[\Gamma]} was introduced which are similar to the Friedman–Stanley jump EE+E\mapsto E^{+}: for every countable group Γ\Gamma, the Γ\Gamma-jump of the classification problem (X,E)(X,E) is the classification problem

(XΓ,E[Γ]),withxE[Γ]x(γΓ)(αΓ)x(γ1α)Ex(α).(X^{\Gamma},E^{[\Gamma]}),\quad\text{with}\quad xE^{[{\Gamma}]}x^{\prime}\iff(\exists\gamma\in\Gamma)\;(\forall\alpha\in\Gamma)\;x(\gamma^{-1}\alpha)\;E\;x^{\prime}(\alpha).

In the same paper they showed that the \mathbb{Z}-jump E0[]E^{[\mathbb{Z}]}_{0} of E0E_{0} is generically ergodic with respect to the countable product E0ωE^{\omega}_{0} of E0E_{0}. In [All20], the stronger result was shown that, in fact, E0[]E^{[\mathbb{Z}]}_{0} is generically EYHE^{H}_{Y}-ergodic, whenever HH is both a non-Archimedean and TSI Polish group. As a consequence of Theorem 1.3, we now have that E0[]E^{[\mathbb{Z}]}_{0} is generically EYHE^{H}_{Y}-ergodic for every TSI Polish group HH and therefore E0[]E^{[\mathbb{Z}]}_{0} is not classifiable by any TSI-group action. In fact, in Section 2 we define a PP-jump operator EE[P]E\mapsto E^{[P]}, for every Polish permutation group PP, which is a common generalization of the Friedman–Stanley jump and the jump operators defined in [CC]. It turns out that PP-jumps are particularly natural in the context of the generalized Bernoulli shifts from [KMPZ20]: if EE is the orbit equivalence relation of the generalized Bernoulli shift of the Polish permutation group QQ, then E[P]E^{[P]} is the orbit equivalence relation of the generalized Bernoulli shift of (P Wr Q)(P\text{ Wr }Q). The anti-classification result for E0[]E^{[\mathbb{Z}]}_{0} is a particular instance of the next corollary. Here, (P Wr G)(P\text{ Wr }G) is just the Wreath product of PP and GG; see Section 2.

Theorem 1.5.

Let XX be a Polish GG-space which has a dense orbit, and let PP be Polish group of permutations of a countable set NN. If all PP-orbits of NN are infinite, then the unbalanced graph of the Polish (P Wr G)(P\text{ Wr }G)-space XNX^{N} is generically semi-connected.

Corollary 1.6.

If P,QSP,Q\leq S_{\infty} are both non-compact Polish permutation groups then the Bernoulli shift of (P Wr Q)(P\text{ Wr }Q) has a generically unbalanced closed subshift.

Corollary 1.6 provides many examples of orbit equivalence relations of CLI groups which are not classifiable by TSI-group actions. However, all such examples are classifiable by countable structures. The next corollary shows that the complexity class within the CLI region but outside of the TSI and the non-Archimedean region in Figure 1 is non-empty:

Corollary 1.7.

There is a Polish GG-space XX of a CLI Polish group GG which is turbulent and generically unbalanced.

We finally illustrate how our results apply to natural classification problems from topology and operator algebras. For any locally compact metrizable space TT, consider the problem (CTr(T),)(\mathrm{CTr}^{*}(T),\equiv_{\mathcal{M}}), of classifying all separable continuous-trace CC^{*}-algebras with spectrum TT up to Morita equivalence; and the problem (Bun(T),iso)(\mathrm{Bun}_{\mathbb{C}}(T),\simeq_{\mathrm{iso}}), of classifying all Hermitian line bundles over TT up to isomorphism. In Section 6 we show that both problems are in general not classifiable by actions of TSI-groups, even when TT is a CW-complex. In contrast, recall that the base-preserving versions T\equiv_{\mathcal{M}}^{T}, and isoT\simeq_{\mathrm{iso}}^{T}, of the above problems are always classifiable by TSI—in fact by abelian—group actions; [BLP19].

1.4. Structure of the paper

Theorem 1.3, which is the main result of this paper, is proved in Section 5. The necessary background is developed independently in Section 3 and Section 4. In Section 2 we assume Theorem 1.3 and provide “in vitro” applications, such as Theorem 1.5, Corollary 1.6, and Corollary 1.7. In Section 6 we discuss applications in topology and operator algebras. Finally, in Section 7 we informally announce some extensions of this work beyond the TSI dividing line which is currently work in progress.

2. Wreath products, PP-jumps, and Bernoulli shifts

In this section, for every Polish permutation group PP of a countable set we introduce a PP-jump operator for GG-spaces and equivalence relations, in the spirit of [CC]. We then draw some connections with the theory of generalized Bernoulli shifts from [KMPZ20]. We also prove Theorem 1.5, and derive Corollary 1.6 and Corollary 1.7.

Let NN be a countable set. We denote by S(N)S(N) the group of all permutations of NN. This is a Polish group, when endowed with the pointwise convergence topology. By a Polish group of permutations of NN we mean any closed subgroup PP of S(N)S(N). Since we are considering left actions, the pertinent action P×NNP\times N\to N is given by

(p,n)p(n), for any map p:NN in P.(p,n)\mapsto p(n),\text{ for any map }p\colon N\to N\text{ in }P.

Let PP be as above and let GG be an arbitrary Polish group. The group

GN:={𝒈:NG}, where (𝒈2𝒈1)(n):=𝒈2(n)𝒈1(n),G^{N}:=\{\bm{g}\colon N\to G\},\text{ where }(\bm{g}_{2}\bm{g}_{1})(n):=\bm{g}_{2}(n)\bm{g}_{1}(n),

is Polish and there is a natural PP-action φ:P×GNGN\varphi\colon P\times G^{N}\to G^{N} on GNG^{N} by automorphisms:

φ(p,𝒈)=𝒈p, where 𝒈p(n):=𝒈(p1(n)).\varphi(p,\bm{g})=\bm{g}^{p},\text{ where }\bm{g}^{p}(n):=\bm{g}(p^{-1}(n)).

The Wreath product (P Wr NG)(P\text{ Wr }_{\mkern-6.0muN}\,G) of PP and GG, or simply (P Wr G)(P\text{ Wr }G), is the group

P Wr G:=PφGN.P\text{ Wr }G:=P\rtimes_{\varphi}G^{N}.

Concretely, elements of P Wr GP\text{ Wr }G are all pairs (p,𝒈)(p,\bm{g}), where pPp\in P, 𝒈GN\bm{g}\in G^{N}, and

(p2,𝒈2)(p1,𝒈1):=(p2p1,φ(p1,𝒈2)𝒈1), where (φ(p1,𝒈2)𝒈1)(n)=𝒈2(p11(n))𝒈1(n).(p_{2},\bm{g}_{2})\cdot(p_{1},\bm{g}_{1}):=(p_{2}p_{1},\varphi(p_{1},\bm{g}_{2})\bm{g}_{1}),\text{ where }\big{(}\varphi(p_{1},\bm{g}_{2})\bm{g}_{1}\big{)}(n)=\bm{g}_{2}(p_{1}^{-1}(n))\bm{g}_{1}(n).

If XX is a Polish GG-space, then the PP-jump of GXG\curvearrowright X is the (P Wr G)(P\text{ Wr }G)-space XNX^{N}:

(p,𝒈)𝒙:=𝒙𝒈,p, where 𝒙𝒈,p(n):=𝒈p(n)𝒙p(n)=𝒈(p1(n))𝒙(p1(n)).(p,\bm{g})\cdot\bm{x}:=\bm{x}^{\bm{g},p},\text{ where }\bm{x}^{\bm{g},p}(n):=\bm{g}^{p}(n)\cdot\bm{x}^{p}(n)=\bm{g}(p^{-1}(n))\cdot\bm{x}(p^{-1}(n)).

Similarly, let (XN,E[P])(X^{N},E^{[P]}) be the PP-jump of a classification problem (X,E)(X,E), where:

𝒙E[P]𝒙(pP)(nN)𝒙(p1(n))E𝒙(n)\bm{x}E^{[P]}\bm{x}^{\prime}\iff(\exists p\in P)\;(\forall n\in N)\;\;\bm{x}(p^{-1}(n))E\bm{x}^{\prime}(n)

Notice that if P=S(N)P=S(N), then this is simply the Friedman–Stanley jump EE+E\mapsto E^{+}, and if PP is the left regular representation of a countable group Γ\Gamma as a subgroup of S(Γ)S(\Gamma), then this is the Γ\Gamma-jump EE[Γ]E\mapsto E^{[\Gamma]}, introduced in [CC]. Clearly, the orbit equivalence relation of the PP-jump of a GG-space XX is the PP-jump of the orbit equivalence relation of the same space. We may now proceed to the proof of Theorem 1.5.

Proof of Theorem 1.5..

Let XX be a Polish GG-space which has a dense orbit and let PS(N)P\leq S(N) be a Polish permutation group on a countable set NN with infinite orbits.

Fix any comeager set CXNC\subseteq X^{N}. For any fixed nNn\in N consider “column” space X{n}X^{\{n\}} of all maps from the singleton {n}\{n\} to XX. This is naturally isomorphic to the Polish GG-space XX. By intersecting CC with the appropriate comeager set we may assume without loss of generality that for all 𝒙C\bm{x}\in C we have that:

(1) for all nN, the orbit of 𝒙(n) is dense in X{n}.\text{for all }n\in N,\text{ the orbit of }\bm{x}(n)\text{ is dense in }X^{\{n\}}.

Here we use that {xX[x] is dense}\{x\in X\mid[x]\text{ is dense}\} is a GδG_{\delta} subset of XX, and since by assumption it contains a dense orbit, it is comeager.

Let EE be the collection of all bijective maps from NN to the space N×{0,1}N\times\{0,1\}, of two disjoint copies of NN. Then EE is a Polish space with the pointwise convergence topology. Since every PP-orbit of NN is infinite, by the Neumann’s lemma [Neu76, Lemma 2.3] we have that for every finite A,BNA,B\subseteq N, there is pPp\in P so that (pA)B=(p\cdot A)\cap B=\emptyset. As a consequence, for the generic eEe\in E and for every i{0,1}i\in\{0,1\}:

(2) if AN is finite, then there is pP so that e(p1(n))=(p1(n),i), for all nA.\text{if }A\subseteq N\text{ is finite, then there is }p\in P\text{ so that }e(p^{-1}(n))=(p^{-1}(n),i),\text{ for all }n\in A.

Fix some eEe\in E satisfying the above and consider the map φ:XN×XNXN\varphi\colon X^{N}\times X^{N}\to X^{N}, where φ(𝒙0,𝒙1)\varphi(\bm{x}_{0},\bm{x}_{1}) is the function 𝒙:NX\bm{x}\colon N\to X, with 𝒙(n)=𝒙0(m)\bm{x}(n)=\bm{x}_{0}(m), if e(n)=(m,0)e(n)=(m,0); and 𝒙(n)=𝒙1(m)\bm{x}(n)=\bm{x}_{1}(m), if e(n)=(m,1)e(n)=(m,1).

Claim 2.1.

The map φ:XN×XNXN\varphi\colon X^{N}\times X^{N}\to X^{N} is a homeomorphism of topological spaces.

Proof of Claim..

Since ee is bijective, every 𝒙XN\bm{x}\in X^{N} pulls back to some (𝒙0,𝒙1)(\bm{x}_{0},\bm{x}_{1}) via φ\varphi and every (𝒙0,𝒙1)X×X(\bm{x}_{0},\bm{x}_{1})\in X\times X pushes forward to some 𝒙X\bm{x}\in X, i.e., φ\varphi is a bijection. Continuity of φ\varphi and φ1\varphi^{-1} is straightforward. ∎

Consider the comeager subset Cφ:=φ1(C)(C×C)C^{\varphi}:=\varphi^{-1}(C)\bigcap(C\times C) of XN×XNX^{N}\times X^{N}. By Kuratowski-Ulam, there is some 𝒛C\bm{z}\in C so that D:={𝒚X(𝒛,𝒚)Cφ}D:=\{\bm{y}\in X\mid(\bm{z},\bm{y})\in C^{\varphi}\} is a comeager subet of XX. We are now done with the proof, since by the next claim, for every 𝒙,𝒚D\bm{x},\bm{y}\in D we have the following path between 𝒙\bm{x} and 𝒚\bm{y} in CC:

𝒙φ(𝒙,𝒛)𝒛φ(𝒛,𝒚)𝒚.\bm{x}\;\leftrightsquigarrow\;\varphi(\bm{x},\bm{z})\;\leftrightsquigarrow\;\bm{z}\;\leftrightsquigarrow\;\varphi(\bm{z},\bm{y})\;\leftrightsquigarrow\;\bm{y}.
Claim 2.2.

For all 𝐱0,𝐱1C\bm{x}_{0},\bm{x}_{1}\in C we have that φ(𝐱0,𝐱1)𝐱0\varphi(\bm{x}_{0},\bm{x}_{1})\leftrightsquigarrow\bm{x}_{0} and φ(𝐱0,𝐱1)𝐱1\varphi(\bm{x}_{0},\bm{x}_{1})\leftrightsquigarrow\bm{x}_{1}.

Proof of Claim..

It suffices to prove that φ(𝒙0,𝒙1)𝒙0\varphi(\bm{x}_{0},\bm{x}_{1})\leftrightsquigarrow\bm{x}_{0} since the other case is symmetric. Set 𝒙=𝒙0\bm{x}=\bm{x}_{0} and 𝒚=φ(𝒙0,𝒙1)\bm{y}=\varphi(\bm{x}_{0},\bm{x}_{1}). Let V(P Wr G)V\subseteq(P\text{ Wr }G) be an open neighborhood of the identity and let UXU\subseteq X be any non-empty open set. By shrinking both sets if necessary, we assume that there is a finite set ANA\subseteq N, a map u:AXu\colon A\to X, an open neighborhood WGW\subseteq G of the identity of GG, and some ε>0\varepsilon>0, so that

V={(p,𝒈)(P Wr G)𝒈(n)W, for all nA, and p fixes every aA}, and V=\{(p,\bm{g})\in(P\text{ Wr }G)\mid\bm{g}(n)\in W,\text{ for all }n\in A,\text{ and }p\text{ fixes every }a\in A\},\text{ and }
U={𝒙XNd(𝒙(n),u(n))<ε, for all nA},U=\{\bm{x}\in X^{N}\mid d\big{(}\bm{x}(n),u(n)\big{)}<\varepsilon,\text{ for all }n\in A\},

where dd is any metric on XX that is compatible with the topology.

By (2) there is pPp\in P so that ((p,𝟏𝑮)𝒙)(n)=((p,𝟏𝑮)𝒚)(n)\big{(}(p,\bm{1_{G}})\cdot\bm{x}\big{)}(n)=\big{(}(p,\bm{1_{G}})\cdot\bm{y}\big{)}(n), for all nAn\in A. By (1) we may set gx=gy:=(p,𝒈)g^{x}=g^{y}:=(p,\bm{g}) for some 𝒈GN\bm{g}\in G^{N} so that in addition to

(3) ((p,𝒈)𝒙)(n)=((p,𝒈)𝒚)(n), for all nA,\big{(}(p,\bm{g})\cdot\bm{x}\big{)}(n)=\big{(}(p,\bm{g})\cdot\bm{y}\big{)}(n),\text{ for all }n\in A,

we also have gx𝒙,gy𝒚Ug^{x}\cdot\bm{x},g^{y}\cdot\bm{y}\in U. But VV contains the following subgroup of (P Wr G)(P\text{ Wr }{}G):

{(1P,𝒉)𝒉(n)=1G, for all nA}.\{(1_{P},\bm{h})\mid\bm{h}(n)=1_{G},\text{ for all }n\in A\}.

By (1) and (3), it now follows that (gy𝒚)V(gx𝒙)¯(g^{y}\bm{y})\in\overline{V\cdot(g^{x}\cdot\bm{x})} and (gx𝒙)V(gy𝒚)¯(g^{x}\bm{x})\in\overline{V\cdot(g^{y}\cdot\bm{y})}. ∎

Let PP be a Polish permutation group of a countable set NN. The generalized Bernoulli shift of PP is the Polish PP-space N\mathbb{R}^{N} where (p,x)xp(p,x)\mapsto x^{p} with xp(n)=x(p1(n))x^{p}(n)=x(p^{-1}(n)), for every nNn\in N. This indeed generalizes the classical Bernoulli shift ΓΓ\Gamma\curvearrowright\mathbb{R}^{\Gamma} for countable disrete groups, where (g,x)gx(g,x)\mapsto g\cdot x with (gx)(γ)=x(g1γ)(g\cdot x)(\gamma)=x(g^{-1}\gamma). In [KMPZ20] it was shown that the Borel reduction complexity of the orbit equivalence relation of PNP\curvearrowright\mathbb{R}^{N} is often a reflection of the dynamical properties of PP.

In addition to PP above, consider a Polish permutation group QQ on some set MM. Notice that P Wr NQP\text{ Wr }_{\mkern-6.0muN}Q may be realized as a Polish permutation group on the set N×MN\times M via the faithful action P Wr NQN×MP\text{ Wr }_{\mkern-6.0muN}Q\curvearrowright N\times M with

(p,𝒒)(n,m)(p(n),(𝒒(n))(m)).(p,\bm{q})\cdot(n,m)\mapsto(p(n),(\bm{q}(n))(m)).

Moreover, its associated generalized Bernoulli shift P Wr NQN×MP\text{ Wr }_{\mkern-6.0muN}Q\curvearrowright\mathbb{R}^{N\times M} is naturally isomorphic to the PP-jump P Wr NQ(M)NP\text{ Wr }_{\mkern-6.0muN}Q\curvearrowright(\mathbb{R}^{M})^{N} of the generalized Bernoulli shift QMQ\curvearrowright\mathbb{R}^{M} of QQ. In particular, since E0E_{0} is Borel isomorphic to the orbit equivalence of the (classical) Bernoulli shift of \mathbb{Z}, we see that the \mathbb{Z}-jump E0[]E^{[\mathbb{Z}]}_{0} of E0E_{0} can be identified with the orbit equivalence relation of the generalized Bernoulli shift

 Wr ×.\mathbb{Z}\text{ Wr }_{\mkern-6.0mu\mathbb{Z}}\mathbb{Z}\curvearrowright\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}}.

In [CC] it was shown that E0[]E_{0}^{[\mathbb{Z}]} is generically ergodic with respect to E0ωE_{0}^{\omega}. This was later generalized in [All20], where it was shown that E0[]E_{0}^{[\mathbb{Z}]} is generically ergodic with respect to any orbit equivalence relation of any non-Archimedean TSI Polish group action. The following is a special case of Corollary 1.6 (in the form of Lemma 2.4 below). Notice that since E0[]E_{0}^{[\mathbb{Z}]} has meager equivalence classes this implies that it is not classifiable by TSI-group actions.

Corollary 2.3.

E0[]E_{0}^{[\mathbb{Z}]} is generically ergodic with respect actions of TSI-groups.

We proceed to the proof of Corollary 1.6 after restating it in a more precise way:

Lemma 2.4.

Let PS(N)P\leq S(N) and QS(M)Q\leq S(M) be Polish permutation groups on countable sets N,MN,M. If P,QP,Q are non-compact then the closed invariant subspace:

{xM×Nx(m,n)0 both orbits [n]PN and [m]QM are infinite}\{x\in\mathbb{R}^{M\times N}\mid x(m,n)\neq 0\implies\text{ both orbits }[n]_{P}\subseteq N\text{ and }[m]_{Q}\subseteq M\text{ are infinite}\}

of the Bernoulli shift of (P Wr NQ)(P\text{ Wr }_{\mkern-6.0muN}Q), has meager orbits and is generically ergodic with respect to actions of TSI Polish groups.

Proof of Corollary 1.6 in the form of Lemma 2.4.

It is easy to see that the orbits are meager given that the range(x)\mathrm{range}(x) is a countable subset of \mathbb{R}, for all xN×Mx\in\mathbb{R}^{N\times M}.

Let MMM^{\infty}\subseteq M and NNN^{\infty}\subseteq N be the collection of all points whose QQ-orbit and PP-orbit, respectively, is infinite. Let also QS(M)Q^{*}\leq S(M^{\infty}) and PS(N)P^{*}\leq S(N^{\infty}) be the image of QQ and PP respectively in their new representation. It is clear that the closed invariant subspace in the statement of the corollary is isomorphic to action:

(P Wr Q)N×M,(P^{*}\text{ Wr }_{\mkern-6.0mu\mathbb{N}}Q^{*})\curvearrowright\mathbb{R}^{N^{\infty}\times M^{\infty}},

and as in the previous paragraph this is just the PP^{*}-jump of the Benroulli shift of QQ^{*}. But the Bernoulli shift of QQ^{*} is generically ergodic, since QQ^{*} is non-compact (see [KMPZ20]), and all orbits of PP^{*} are infinite. The rest follows from Theorem 1.5

Corollary 1.6 gives many examples of classification problems which are classifiable by countable structures but not by actions of TSI groups. We may similarly find orbit equivalence relations which are classifiable neither by countable structures nor by TSI group actions. In fact we may do so while acting with a CLI group.

Proof of Corollary 1.7.

Let (l2,+)(l_{2},+) be the group of all sequences (an)n(a_{n})_{n} of reals which are square-summable. The action of l2l_{2} on \mathbb{R}^{\mathbb{N}} with (an)n(xn)n:=(an+xn)n(a_{n})_{n}\cdot(x_{n})_{n}:=(a_{n}+x_{n})_{n} is turbulent, see [Gao09]. In particular, it has meager orbits all of which happen to be dense. Let GXG\curvearrowright X be the \mathbb{Z}-jump of l2l_{2}\curvearrowright\mathbb{R}^{\mathbb{N}}. It is easy to check that this space is turbulent and still has meager orbits. The rest follows from Theorem 1.5 and the fact that the wreath product of CLI groups is CLI. ∎

3. Definable classification induces homomorphism between unbalanced graphs

The following Theorem is the main result of this section.

Theorem 3.1.

Suppose XX is a Polish GG-space and YY is a Polish HH for Polish groups GG and HH. For any Baire-measurable homomorphism f:EXGEYHf:E^{G}_{X}\rightarrow E^{H}_{Y}, there is a GG-invariant comeager set CXC\subseteq X such that for any x0,y0Cx_{0},y_{0}\in C, if x0y0x_{0}\leftrightsquigarrow y_{0} then f(x0)f(y0)f(x_{0})\leftrightsquigarrow f(y_{0}).

For the proof of this Theorem we will rely on two lemmas. The following “orbit-continuity” lemma is essentially [Hjo00, Lemma 3.17] modified as in the beginning of the proof of [Hjo00, Theorem 3.18]. For a direct proof see [LP18], noting that the the set CC produced in that proof happens to be invariant.

Lemma 3.2.

Suppose XX is a Polish GG-space and YY is a Polish HH-space for Polish groups GG and HH. For any Baire-measurable homomorphism f:EXGEYHf:E^{G}_{X}\rightarrow E^{H}_{Y}, there is an invariant comeager set CXC\subseteq X such that:

  1. (1)

    ff restricted to CC is continuous; and

  2. (2)

    for any x0Cx_{0}\in C and any open neighborhood WW of the identity of HH, there is an open neighborhood UU of x0x_{0} and an open neighborhood VV of the identity of GG such that for any xUCx\in U\cap C and for a comeager set of gVg\in V, we have f(gx)Wf(x)f(gx)\in Wf(x).

The next lemma says that the witnesses gxg^{x} and gyg^{y} in the definition of xyx\leftrightsquigarrow y can be taken to be locally generic.

Lemma 3.3.

For any Polish GG-space XX and any x,yXx,y\in X, if xyx\leftrightsquigarrow y, then for any open neighborhood VV of the identity of GG and any nonempty open neighborhood UXU\subseteq X of xx or yy, there is a nonempty open set of gxGg^{x}\in G and a nonempty open set of gyGg^{y}\in G such that gxx,gyyUg^{x}x,g^{y}y\in U, gyyV(gxx)¯g^{y}y\in\overline{V(g^{x}x)}, and gxx(V(gyy)¯g^{x}x\in\overline{(V(g^{y}y)}.

Proof.

Let UXU\subseteq X be an open set intersecting the orbits of xx or yy, and let VV an open neighborhood of the identity of GG. Choose another open neighborhood V0V_{0} of the identity such that V03VV_{0}^{3}\subseteq V. By the definition we can find some hx,hyGh^{x},h^{y}\in G such that hx,hyUh^{x},h^{y}\in U and hxxV0(hyy)¯h^{x}x\in\overline{V_{0}(h^{y}y)} and hyyV0(hxx)¯h^{y}y\in\overline{V_{0}(h^{x}x)}.

However, observe that for any gxV0hxg^{x}\in V_{0}h^{x} and gyV0hyg^{y}\in V_{0}h^{y}, we have gxxV(gyy)¯g^{x}x\in\overline{V(g^{y}y)} and gyyV(gxx)¯g^{y}y\in\overline{V(g^{x}x)}. Also, the set of gxV0hxg^{x}\in V_{0}h^{x} such that gxxUg^{x}x\in U and the set of gyV0hyg^{y}\in V_{0}h^{y} such that gyyUg^{y}y\in U are both open and nonempty. Thus the sets of gxg^{x} and gyg^{y} satisfying the conditions contains nonempty open sets and thus are nonmeager. ∎

We turn now to the proof of Theorem 3.1.

Proof of Theorem 3.1.

Fix an arbitrary open neighborhood WW of the identity of HH, and an open set UYU\subseteq Y intersecting the orbit of f(y0)f(y_{0}) (the case that UU intersects the orbit of f(x0)f(x_{0}) is similar). By the invariance of \leftrightsquigarrow, it would suffice to prove the claim for the Baire-measurable homomorphism xhf(x)x\mapsto h\cdot f(x) for any hHh\in H. Thus without loss of generality we may assume that UU is a neighborhood of f(y0)f(y_{0}). By the orbit-continuity lemma, we can find open neighborhoods UU^{\prime} of y0y_{0} and WW of the identity of GG such that for every xUCx\in U^{\prime}\cap C, there is a comeager set of vVv\in V such that f(vx)Wf(x)f(vx)\in Wf(x). By shrinking UU^{\prime}, we may assume that f[U]Uf[U^{\prime}]\subseteq U.

Since x0y0x_{0}\leftrightsquigarrow y_{0}, by the previous lemma, we may find some group elements gx,gyGg^{x},g^{y}\in G such that gxx0,gyy0UCg^{x}x_{0},g^{y}y_{0}\in U^{\prime}\cap C, gxx0V(gxx)¯g^{x}x_{0}\in\overline{V(g^{x}x)} and gyy0V(gyy)¯g^{y}y_{0}\in\overline{V(g^{y}y)}. Since ff is a homomorphism, we may fix group elements hx,hyHh^{x},h^{y}\in H such that hxf(x0)=f(gxx0)h^{x}f(x_{0})=f(g^{x}x_{0}) and hyf(y0)=f(gyy0)h^{y}f(y_{0})=f(g^{y}y_{0}), which are both elements of UU.

To see that hxf(x0)W(hyf(y0))¯h^{x}f(x_{0})\in\overline{W(h^{y}f(y_{0}))}, fix an open neighborhood U0U_{0} of hxf(x0)h^{x}f(x_{0}). Notice gxx0f1[U0]Cg^{x}x_{0}\in f^{-1}[U_{0}]\cap C and the set vVv\in V such that v(gyy0)f1[U0]v(g^{y}y_{0})\in f^{-1}[U_{0}] is nonempty open, thus we can choose one such that f(v(gyy0))Wf(gyy0)=W(hyf(y0))f(v(g^{y}y_{0}))\in Wf(g^{y}y_{0})=W(h^{y}f(y_{0})). Checking that hyf(y0)W(hxf(x0))¯h^{y}f(y_{0})\in\overline{W(h^{x}f(x_{0}))} is the same. ∎

4. Strong ergodicity properties and dynamical back and forth

Let XX be a Polish GG-space. In this section we define binary relations Gα\precsim^{\alpha}_{G} and Gα\sim^{\alpha}_{G} on XX, for every α<ω1\alpha<\omega_{1}. Intuitively, two points x,yXx,y\in X satisfy xGαyx\precsim^{\alpha}_{G}y, iff Player II has a non-losing strategy of rank α\alpha in a dynamical analogue of the classical Ehrenfeucht–Fraïssé game, where Player I is the “spoiler” and starts by partially specifying xx. In Proposition 4.2, which is the main result of this section, we derive some strong ergodicity properties under the assumption xGαyx\precsim^{\alpha}_{G}y. Notice that the ideas developed in this section are similar in spirit to the content of [Hjo00, Section 6.4].

Definition 4.1.

Let XX be a Polish GG-space, let VV be an open neighborhood of the identity of GG and let x,yXx,y\in X. By recursion on α<ω1\alpha<\omega_{1}, we simultaneously define relations xVαyx\precsim^{\alpha}_{V}y and xVαyx\sim^{\alpha}_{V}y as follows:

  1. (1)

    Let xV0yx\precsim_{V}^{0}y holds exactly when xVy¯x\in\overline{Vy};

  2. (2)

    If Vα\precsim_{V}^{\alpha} is defined, then xVαyx\sim^{\alpha}_{V}y holds exactly when xVαyx\precsim_{V}^{\alpha}y and yVαxy\precsim_{V}^{\alpha}x;

  3. (3)

    Assume for some ordinal α\alpha that Wβ\sim^{\beta}_{W} is defined for every ordinal β<α\beta<\alpha and every open neighborhood WW of the identity of GG. Then, xVαyx\precsim^{\alpha}_{V}y holds exactly when for every open neighborhood WW of the identity of GG there exists some vVv\in V, such that for every β<α\beta<\alpha we have that vyWβxvy\sim^{\beta}_{W}x.

Note that the relations Vα\precsim^{\alpha}_{V} are not necessarily symmetric or transitive. The relations Vα\sim^{\alpha}_{V} are symmetric by definition, but they are also not necessarily transitive. It’s also worth noting that by an easy argument V1V0\sim^{1}_{V}\subseteq\sim^{0}_{V}, and then it follows that VαVβ\sim^{\alpha}_{V}\subseteq\sim^{\beta}_{V} and VαVβ\precsim^{\alpha}_{V}\subseteq\precsim^{\beta}_{V} for every βα\beta\leq\alpha. Similarly, whenever WVW\subseteq V are basic open neighborhoods of the identity of GG, it is easy to see that WαVα\sim^{\alpha}_{W}\subseteq\sim^{\alpha}_{V}.

Before we state the main result of this section, recall the notation associated with the Vaught transforms. Let AXA\subseteq X be a Baire-measurable set and let VGV\subseteq G be any open set. If xXx\in X, we write xAVx\in A^{*V} if the set {vVvxA}\{v\in V\mid vx\in A\} is comeager in VV. We write xAΔVx\in A^{\Delta V} if the set {vVvxA}\{v\in V\mid vx\in A\} is non-meager in VV. For basic properties of the Vaught transforms one may consult [Gao09].

Proposition 4.2.

Let GG an arbitrary Polish group and let XX be a Polish GG-space. If xGαyx\precsim^{\alpha}_{G}y and α1\alpha\geq 1, then for every 𝚷α0\mathbf{\Pi}^{0}_{\alpha}-set AXA\subseteq X we have that

xAΔGyAΔG.x\in A^{\Delta G}\Rightarrow y\in A^{\Delta G}.

We start by recording some useful basic properties of the relations Vα\precsim^{\alpha}_{V}.

Lemma 4.3.

Let XX be a Polish GG-space and let V,WV,W be open neighborhoods of the identity of GG. For every α<ω1\alpha<\omega_{1} and every x,y,zXx,y,z\in X we have that:

  1. (1)

    If xVαyx\precsim^{\alpha}_{V}y and gGg\in G, then gxgVg1αgygx\precsim^{\alpha}_{gVg^{-1}}gy; and

  2. (2)

    If xVαyx\precsim^{\alpha}_{V}y and yWαzy\precsim^{\alpha}_{W}z, then xVWαzx\precsim^{\alpha}_{VW}z.

Proof.

For (1), if xV0yx\precsim^{0}_{V}y, then x=limnhnyx=\lim_{n}h_{n}y for some hnVh_{n}\in V. But then, by continuity of the action we have gx=limnghny=limnghng1gygx=\lim_{n}gh_{n}y=\lim_{n}gh_{n}g^{-1}gy; that is, gxgVg10gygx\precsim^{0}_{gVg^{-1}}gy. Assume now that xVαyx\precsim^{\alpha}_{V}y and let WGW\subseteq G be an open neighborhood of the identity of GG. Since xVαyx\precsim^{\alpha}_{V}y and g1Wgg^{-1}Wg is an open neighborhood of the identity of GG, there is some vVv\in V such that for every β<α\beta<\alpha, vyg1Wgβxvy\sim^{\beta}_{g^{-1}Wg}x. By the inductive assumption, for every β<α\beta<\alpha we have that gvyWβgxgvy\sim^{\beta}_{W}gx. Hence (gvg1)gyWβgx(gvg^{-1})gy\sim^{\beta}_{W}gx for every β<α\beta<\alpha, with gvg1gVg1gvg^{-1}\in gVg^{-1}, as desired.

For (2), suppose first that xV0yx\precsim^{0}_{V}y and yW0zy\precsim^{0}_{W}z. For an arbitrary open neighborhood UxU\ni x, we can find an open neighborhood UyU^{\prime}\ni y and some vVv\in V such that vUUvU^{\prime}\subseteq U. Then we can find some wWw\in W such that wzUwz\in U^{\prime}, in which case vwzUvwz\in U, where vwVWvw\in VW. Thus xVWz¯x\in\overline{VWz}.

Now suppose xVαyx\precsim^{\alpha}_{V}y and yWαzy\precsim^{\alpha}_{W}z for some α1\alpha\geq 1. Fix an open neighborhood OO of the identity of GG, with the goal of showing for some gVWg\in VW that for every β<α\beta<\alpha, gzOβxgz\sim^{\beta}_{O}x. Let O1O_{1} be an open neighborhood of the identity of GG so that O12OO_{1}^{2}\subseteq O. Since xVαyx\precsim^{\alpha}_{V}y, there is vVv\in V such that for every β<α\beta<\alpha, we have vyO1βxvy\sim^{\beta}_{O_{1}}x. Now find some wWw\in W such that for every β<α\beta<\alpha, wzv1O1vβywz\sim_{v^{-1}O_{1}v}^{\beta}y. By Lemma 4.3.(1), for every β<α\beta<\alpha we get vwzO1βvyvwz\sim_{O_{1}}^{\beta}vy. Thus for every β<α\beta<\alpha, we have vwzO12βxvwz\sim^{\beta}_{O^{2}_{1}}x by the induction hypothesis, and therefore vwzOβxvwz\sim_{O}^{\beta}x as desired. ∎

We may now proceed to the proof of Proposition 4.2.

Proof of Proposition 4.2.

Notice that if xAΔGx\in A^{\Delta G}, one can find an open neighborhood VV of the identity of GG and a group element gGg\in G such that gxAVgx\in A^{*V}. By Lemma 4.3.(1), we have gxGαgygx\precsim_{G}^{\alpha}gy. Then there is some hGh\in G such that for every β<α\beta<\alpha, hgyWβgxhgy\precsim_{W}^{\beta}gx, where WW is chosen to be some open neighborhood of the identity of GG such that W2VW^{2}\subseteq V. Thus it suffices to prove the following claim, which tells us that hgyAWhgy\in A^{*W} and thus yAΔGy\in A^{\Delta G}.

Claim 4.4.

Let V,WV,W be open neighborhoods of the identity of GG so that W2VW^{2}\subseteq V. If for some α1\alpha\geq 1 we have A𝚷α0(X)A\in\mathbf{\Pi}^{0}_{\alpha}(X) and xWβyx\precsim^{\beta}_{W}y for every β<α\beta<\alpha, then yAVy\in A^{*V} implies xAWx\in A^{*W}.

Proof of Claim..

We proceed by induction on α1\alpha\geq 1. First, suppose xW0yx\precsim^{0}_{W}y and that yAVy\in A^{*V} for some closed set AXA\subseteq X. Assuming for the sake of contradiction that xAWx\not\in A^{*W}, one can pick some w0Ww_{0}\in W such that w0xUw_{0}x\in U, where U:=AcU:=A^{c}. By Lemma 4.3.(1), we get w0xw0Ww010w0yw_{0}x\precsim^{0}_{w_{0}Ww_{0}^{-1}}w_{0}y, in which case one can pick some w1Ww_{1}\in W such that w0w1yUw_{0}w_{1}y\in U. So there is an open neighborhood of w0w1w_{0}w_{1} of elements ww such that wyUwy\in U. Since W2VW^{2}\subseteq V, this contradicts yAVy\in A^{*V}.

Suppose now that, for some ordinal α>1\alpha>1 the claim is true below α\alpha, that xWβyx\precsim^{\beta}_{W}y for every β<α\beta<\alpha, and that yAVy\in A^{*V} for some 𝚷α0\mathbf{\Pi}^{0}_{\alpha} set AA. Write A=nωBnA=\bigcap_{n\in\omega}B_{n} where each BnXB_{n}\subseteq X is Σβn0\Sigma^{0}_{\beta_{n}} for some βn<α\beta_{n}<\alpha. Assume for the sake of contradiction that xAWx\not\in A^{*W}. Then there is some w0Ww_{0}\in W and an open neighborhood W0W_{0} of the identity of GG, as well as some n0ωn_{0}\in\omega, such that W0w0WW_{0}w_{0}\subseteq W and w0x(XBn0)W0w_{0}x\in(X\setminus B_{n_{0}})^{*W_{0}}. Choose an open neighborhood W1W_{1} of the identity of GG such that W12W0W_{1}^{2}\subseteq W_{0}. Then we can find some wWw\in W such that for every β<α\beta<\alpha, wyw01W1w0βxwy\precsim_{w_{0}^{-1}W_{1}w_{0}}^{\beta}x. By Lemma 4.3.(1) we have w0wyW1βw0xw_{0}wy\precsim^{\beta}_{W_{1}}w_{0}x. Thus by the induction hypothesis applied to w0x(XBn0)W0w_{0}x\in(X\setminus B_{n_{0}})^{*W_{0}}, we have w0wy(XBn0)W1w_{0}wy\in(X\setminus B_{n_{0}})^{*W_{1}}. But W1w0wVW_{1}w_{0}w\cap V\neq\emptyset, contradicting that yAVy\in A^{*V}. ∎

5. Dynamics of TSI Polish groups

In this section, we derive some consequences for the relations \leftrightsquigarrow and Hα\precsim^{\alpha}_{H}, when these relations have been defined on a Polish HH-space YY, where HH is a TSI Polish group. We then conclude with the proof of Theorem 1.3.

Throughout this section HH is a TSI Polish group and YY is a Polish HH-space. We also fix a countable basis \mathcal{B} of open, symmetric, and conjugation-invariant neighborhoods of the identity of HH. We assume that GG\in\mathcal{B} and that for any VV\in\mathcal{B}, we have V2V^{2}\in\mathcal{B}.

Lemma 5.1.

Let HH be a Polish TSI group and YY a Polish HH-space. We have:

  1. (1)

    if xyx\leftrightsquigarrow y, then xH1yx\sim^{1}_{H}y;

  2. (2)

    Vα\precsim^{\alpha}_{V} and Vα\sim^{\alpha}_{V} coincide for all VV\in\mathcal{B} and all ordinals α>0\alpha>0; and

  3. (3)

    Hα\sim^{\alpha}_{H} is an equivalence relation for all ordinals α>0\alpha>0.

Proof.

For (1), by taking UU to be XX in the definition of xyx\leftrightsquigarrow y, we can find gx,gyg^{x},g^{y} such that gyyV0gxxg^{y}y\sim^{0}_{V}g^{x}x. By Lemma 4.3(1) and the conjugation-invariance of VV, we get gyV0xgy\sim^{0}_{V}x for g=(gx)1gyg=(g^{x})^{-1}g^{y} as desired.

For (2), it follows immediately from the definitions that VαVα\sim_{V}^{\alpha}\subseteq\precsim_{V}^{\alpha}, so it suffices to show that VαVα\precsim_{V}^{\alpha}\subseteq\sim_{V}^{\alpha}. This amounts to showing that Vα\precsim_{V}^{\alpha} is symmetric. If xVαyx\precsim^{\alpha}_{V}y and WW\in\mathcal{B}, then there is some vVv\in V such that for every β<α\beta<\alpha, we have that vyWβxvy\sim^{\beta}_{W}x. By Lemma 4.3(1) it follows by the conjugation-invariance of WW that v1xWβyv^{-1}x\sim^{\beta}_{W}y for all β<α\beta<\alpha. Since VV is symmetric, v1Vv^{-1}\in V. Hence, yVαxy\precsim^{\alpha}_{V}x.

For (3), transitivity follows from Lemma 4.3(2), and symmetry by (2) above.

In the rest of this section we will want to refer to the relations Vα\precsim^{\alpha}_{V} and Vα\sim^{\alpha}_{V} computed according to multiple Polish topologies on the same space. For any topology σ\sigma making (Y,σ)(Y,\sigma) a Polish HH-space, we will use the notation Vα,σ\precsim^{\alpha,\sigma}_{V} and Vα,σ\sim^{\alpha,\sigma}_{V} to refer to the relations Vα\precsim^{\alpha}_{V} and Vα\sim^{\alpha}_{V} as computed in that space. We will use τ\tau to refer to the original topology on YY, but keep denoting Vα,τ\precsim^{\alpha,\tau}_{V} and Vα,τ\sim^{\alpha,\tau}_{V} simply by Vα\precsim^{\alpha}_{V} and Vα\sim^{\alpha}_{V}. For every VV\in\mathcal{B}, cYc\in Y, and ordinal β>0\beta>0, let

AVβ(c):={dYγ<β,dVγc}.A^{\beta}_{V}(c):=\{d\in Y\mid\forall\gamma<\beta,d\sim^{\gamma}_{V}c\}.

The following technical lemma will be useful.

Lemma 5.2.

Let σ\sigma be an additional topology on YY so that both YY and (Y,σ)(Y,\sigma) are Polish HH-spaces. Let a,b,cYa,b,c\in Y and let α2\alpha\geq 2 be an ordinal so that:

  1. (1)

    for every β<α\beta<\alpha we have aHβca\sim_{H}^{\beta}c and bHβcb\sim_{H}^{\beta}c; and

  2. (2)

    for every V,WV,W\in\mathcal{B}, the set (AVα(c))ΔW(A^{\alpha}_{V}(c))^{\Delta W} is in σ\sigma.

Then

aH1,σbaHαb.a\sim_{H}^{1,\sigma}b\implies a\sim_{H}^{\alpha}b.
Proof.

Let VV be an arbitrary open neighborhood of the identity of HH. Our goal is to show that there is some hHh\in H such that for every β<α\beta<\alpha, hbVβahb\sim_{V}^{\beta}a. To that end, let V0V_{0}\in\mathcal{B} with V01VV_{0}^{-1}\subseteq V, and fix hHh\in H such that hbV00,σahb\sim^{0,\sigma}_{V_{0}}a. Fix any ordinal β\beta such that β<α\beta<\alpha. We claim that hbVβahb\sim^{\beta}_{V}a. To see this, let WW\in\mathcal{B}. We will find some vVv\in V such that vaWγhbva\sim^{\gamma}_{W}hb, for all γ<β\gamma<\beta.

Let W0W_{0}\in\mathcal{B} so that W04WW_{0}^{4}\subseteq W. Because aHβca\sim^{\beta}_{H}c we may choose some gHg\in H such that for every γ<β\gamma<\beta, gaW0γcga\sim_{W_{0}}^{\gamma}c, in which case gaAW0β(c)ga\in A^{\beta}_{W_{0}}(c). By Lemma 4.3(2), we can see that wgaAW02β(c)wga\in A^{\beta}_{W_{0}^{2}}(c) for any wW0w\in W_{0}. Thus a(AW02β(c))ΔW0ga\in(A^{\beta}_{W_{0}^{2}}(c))^{\Delta W_{0}g}. Since (AW02β(c))ΔW0g(A^{\beta}_{W_{0}^{2}}(c))^{\Delta W_{0}g} is an open neighborhood of aa in σ\sigma, there is vV0v\in V_{0} such that vhb(AW02β(c))ΔW0gvhb\in(A^{\beta}_{W_{0}^{2}}(c))^{\Delta W_{0}g}. By definition, for some wW0w\in W_{0}, we have γ<β,wgvhbW02γc\forall\gamma<\beta,wgvhb\sim_{W_{0}^{2}}^{\gamma}c. For every γ<β\gamma<\beta, since gaW0γcga\sim^{\gamma}_{W_{0}}c and wgvhbW02γcwgvhb\sim_{W_{0}^{2}}^{\gamma}c, by Lemma 4.3(2), we have that gaW03γwgvhbga\sim^{\gamma}_{W_{0}^{3}}wgvhb, and thus gaW04γgvhbga\sim^{\gamma}_{W_{0}^{4}}gvhb. By Lemma 4.3(1), we get v1aW04γhbv^{-1}a\sim^{\gamma}_{W_{0}^{4}}hb and thus v1aWγhbv^{-1}a\sim^{\gamma}_{W}hb for every γ<β\gamma<\beta as desired. ∎

We may now proceed to the proof of Theorem 1.3.

Proof Theorem 1.3.

Let f:XYf:X\rightarrow Y be a Baire-measurable homomorphism from EXGE^{G}_{X} to EYHE^{H}_{Y}, for some Polish HH-space YY, where HH is a TSI Polish group.

Claim 5.3.

For all 1α<ω11\leq\alpha<\omega_{1} there is comeager CαXC_{\alpha}\subseteq X so that for all x,yCαx,y\in C_{\alpha},

f(x)Hαf(y).f(x)\sim^{\alpha}_{H}f(y).
Proof.

For α=1\alpha=1, by Lemma 3.1 we have a comeager set DD such that for any x,yDx,y\in D, if xyx\leftrightsquigarrow y, then f(x)f(y)f(x)\leftrightsquigarrow f(y). Since (X/G,)(X/G,\leftrightsquigarrow) is generically semi-connected, we can find a comeager set CDC\subseteq D such that for any x,yCx,y\in C, there is a \leftrightsquigarrow-path between xx and yy through DD. In particular, by Lemma 5.1(3), for any x,yCx,y\in C, we have f(x)H1f(y)f(x)\sim^{1}_{H}f(y). So we may set C1:=CC_{1}:=C.

Assume now that for some countable α2\alpha\geq 2 we have that defined CβC_{\beta} for all β<α\beta<\alpha as in the claim. Fix some zβ<αCβz\in\bigcap_{\beta<\alpha}C_{\beta}. Observe that the set AVα(f(z))={aYβ<α,aVβf(z)}A^{\alpha}_{V}(f(z))=\{a\in Y\mid\forall\beta<\alpha,a\sim^{\beta}_{V}f(z)\} is Borel, and f(x)AVα(f(z))f(x)\in A^{\alpha}_{V}(f(z)) for every xβ<αCβx\in\bigcap_{\beta<\alpha}C_{\beta}. Find a new topology σ\sigma on YY such that: (AWα)ΔV(A^{\alpha}_{W})^{\Delta V} is open for every V,WV,W\in\mathcal{B}; (Y,σ)(Y,\sigma) is a Polish HH-space; and σ\sigma generates the same Borel sets as τ\tau (see [Gao09, Lemma 4.4.3]). Applying Lemma 3.1 and generic semi-connectedness of (X/G,)(X/G,\leftrightsquigarrow) as in the previous paragraph with the space (Y,σ)(Y,\sigma) in place of YY, we can find a comeager set Cα<βCαC\subseteq\bigcap_{\alpha<\beta}C_{\alpha} so that for every x,yCx,y\in C, f(x)Hσ,1f(y)f(x)\sim_{H}^{\sigma,1}f(y). By Lemma 5.2, taking cc to be f(z)f(z), we have that f(x)Hαf(y)f(x)\sim_{H}^{\alpha}f(y), for every x,yCx,y\in C. Set Cα:=CC_{\alpha}:=C. ∎

Claim 5.4.

There is a comeager set CXC\subseteq X and a countable ordinal λ\lambda such that for every xCx\in C, [f(x)][f(x)] is 𝚷λ0\mathbf{\Pi}^{0}_{\lambda}

Proof.

By [BK96, Theorem 7.3.1], there is a Baire-measurable function g:Y2ω×ωg:Y\rightarrow 2^{\omega\times\omega} and a sequence {Aζ}ζω1\{A_{\zeta}\}_{\zeta\in\omega_{1}} of pairwise-disjoint invariant Borel sets such that EYHAζE^{H}_{Y}\upharpoonright A_{\zeta} is Borel for every ζω1\zeta\in\omega_{1}, and g(y)g(y) is a well-order such that yA|g(y)|y\in A_{|g(y)|} for every yYy\in Y. As gg is Baire-measurable and thus so is gfg\circ f, we may find a dense GδG_{\delta} subset CXC\subseteq X such that (gf)C(g\circ f)\upharpoonright C is continuous. Applying 𝚺11\mathbf{\Sigma}^{1}_{1}-boundedness (see [Kec95, Theorem 31.2]) to (gf)C(g\circ f)\upharpoonright C, we can find a countable ordinal γ\gamma such that f[C]Aγf[C]\subseteq A_{\gamma}. Then since EYHAγE^{H}_{Y}\upharpoonright A_{\gamma} is Borel, we can find the countable ordinal λ\lambda with the property that [f(x)][f(x)] is 𝚷λ0\mathbf{\Pi}^{0}_{\lambda} for every xCx\in C. ∎

Fix now λ\lambda and CXC\subseteq X as in the claim such that for any xCx\in C, [f(x)][f(x)] is 𝚷λ0\mathbf{\Pi}^{0}_{\lambda}. By the previous claim, the set D:=CCλD:=C\cap C_{\lambda} is comeager and for any x,yDx,y\in D, f(x)Hλf(y)f(x)\sim^{\lambda}_{H}f(y). By Lemma 4.2, this means that for any 𝚷λ0\mathbf{\Pi}^{0}_{\lambda} set AA, f(x)AΔHf(x)\in A^{\Delta H} iff f(y)AΔHf(y)\in A^{\Delta H}. Since [f(x)][f(x)] is 𝚷λ0\mathbf{\Pi}^{0}_{\lambda} for all xDx\in D, we have that every xDx\in D maps to the same HH-orbit in YY. ∎

6. Applications

In this section we illustrate how the “in vitro” results we have developed so far apply to natural classification problems from topology and operator algebras. We start by reviewing some definitions regarding fibre bundles. We then show that coordinate free isomorphism between Hermitian line bundles and Morita equivalence between continuous-trace CC^{*}-algebras are not classifiable by TSI-group actions.

6.1. The Polish space of locally trivial fibre bundles

Let BB be a locally compact metrizable topological space, and let FF be a Polish GG-space, for some Polish group GG. A locally trivial fibre bundle over BB with fibre FF and structure group GG, or simply a fibre bundle over BB consists of a Polish space EE; a continuous map p:EBp\colon E\to B; a locally finite open cover 𝒰\mathcal{U} of BB; and a homeomorphism hU:p1(U)U×Fh_{U}\colon p^{-1}(U)\to U\times F, for each U𝒰U\in\mathcal{U}; so that:

  1. (1)

    if bU𝒰b\in U\in\mathcal{U}, then hUh_{U} restricts to a homeomorphism from p1(b)p^{-1}(b) to {b}×F\{b\}\times F;

  2. (2)

    if U,V𝒰U,V\in\mathcal{U}, there is a contiunous t(U,V):UVGt_{(U,V)}\colon U\cap V\to G, so that for all bUVb\in U\cap V,

    (hVhU1)(b,f)=(b,t(U,V)(b)f)(h_{V}\circ h_{U}^{-1})(b,f)=(b,t_{(U,V)}(b)f)

The maps hUh_{U} above are called charts and t(U,V)t_{(U,V)} are called the transition maps. Notice that we can always choose 𝒰\mathcal{U} to be a subset of some fixed countable basis \mathcal{B} of the topology of BB, and we can recover EE as the colimit of the above separable data (together with a 11-cocycle condition). Hence, we may form the Polish space Bun(B,G,F)\mathrm{Bun}(B,G,F), of all locally trivial fibre bundle over BB, with fibre FF, and structure group GG, as a GδG_{\delta} subset of the Polish space

2×(U,V)2C(UV,G)2^{\mathcal{B}}\times\prod_{(U,V)\in\mathcal{B}^{2}}C(U\cap V,G)

There are two natural classification problems on Bun(B,G,F)\mathrm{Bun}(B,G,F): the isomorphism relation iso\simeq_{\mathrm{iso}}; and the isomorphism over BB relation isoB\simeq_{\mathrm{iso}}^{B}. First, notice that If p,q:EBp,q\colon E\to B are elements of Bun(B,G,F)\mathrm{Bun}(B,G,F), then we may always choose a common open cover 𝒰\mathcal{U} of BB so that pp and qq are locally trivialized by some (hU),(t(U,V))(h_{U}),(t_{(U,V)}) and (kU),(s(U,V))(k_{U}),(s_{(U,V)}), respectively, with U,V𝒰U,V\in\mathcal{U}. We write pisoqp\simeq_{\mathrm{iso}}q, if there are homeomorphisms π:EE\pi\colon E\to E, ρ:BB\rho\colon B\to B, and a continuous e(U,V:Uρ1(V)Ge_{(U,V}\colon U\cap\rho^{-1}(V)\to G, so that qπ=ρpq\circ\pi=\rho\circ p, and for all U,V𝒰U,V\in\mathcal{U}, for all bUρ1(V)b\in U\cap\rho^{-1}(V), and for all fFf\in F we have that

(hVπhU1)(b,f)=(ρ(b),e(U,V)(b)f).(h_{V}\circ\pi\circ h^{-1}_{U})(b,f)=\big{(}\rho(b),e_{(U,V)}(b)f\big{)}.

We write pisoBqp\simeq_{\mathrm{iso}}^{B}q if ρ\rho above can be taken to be idB\mathrm{id}_{B}.

6.2. Isomorphism of Hermitian line bundles

Let BB be a locally compact metrizable space. By a Hermitian line bundle over BB we mean any locally trivial fibre bundle over BB with fibre F:=F:=\mathbb{C} and structure group G:=U()=𝕋G:=\mathrm{U}(\mathbb{C})=\mathbb{T} being the unitary group of \mathbb{C} acting on \mathbb{C} with rotations. Let Bun(B)\mathrm{Bun}_{\mathbb{C}}(B) be the standard Borel space of all Hermitian line bundles over BB. By a result of [BLP19], isoB\simeq_{\mathrm{iso}}^{B} is classifiable by TSI group actions:

Proposition 6.1 ([BLP19](Corollary 5.12.)).

The problem (Bun(B),isoB)(\mathrm{Bun}_{\mathbb{C}}(B),\simeq_{\mathrm{iso}}^{B}) is classifiable by non-Archimedean, abelian group actions.

In contrast, for the relation iso\simeq_{\mathrm{iso}} we have the following result.

Corollary 6.2.

There exists a locally compact metrizable topological space BB, so that (Bun(B),iso)(\mathrm{Bun}_{\mathbb{C}}(B),\simeq_{\mathrm{iso}}) is not classifiable by TSI group actions. In fact, BB can be taken to be the geometric realization of a countable, locally-finite, CW-complex.

Proof.

The dyadic solenoid Σ\Sigma is the inverse limit of the inverse system (𝕋i,fij)(\mathbb{T}_{i},f^{j}_{i}) where 𝕋i:=𝕋\mathbb{T}_{i}:=\mathbb{T} is the unit circle, viewed as a multiplicative subgroup of \mathbb{C}, and fij:𝕋j𝕋if^{j}_{i}\colon\mathbb{T}_{j}\to\mathbb{T}_{i} is the two-fold cover zz2z\mapsto z^{2}. Let CC be the homotopy limit of the same inverse system. This is formed by taking the disjoint union of the spaces:

𝕋0×[0,1],𝕋1×[0,1],𝕋2×[0,1],\mathbb{T}_{0}\times[0,1],\mathbb{T}_{1}\times[0,1],\mathbb{T}_{2}\times[0,1],\ldots

and identifying the point (z,0)𝕋i+1×[0,1](z,0)\in\mathbb{T}_{i+1}\times[0,1] with the point (z2,1)𝕋i×[0,1](z^{2},1)\in\mathbb{T}_{i}\times[0,1], for each i0i\geq 0. Clearly CC is a locally finite CW-complex.

Recall now that the quotient Bun(C)/isoC\mathrm{Bun}_{\mathbb{C}}(C)/\simeq_{\mathrm{iso}}^{C} is in bijective correspondence with the first Čech cohomology group H1(C,𝕋)H^{1}(C,\mathbb{T}) of CC with coefficients from 𝕋\mathbb{T}; [RW98, Proposition 4.53]. Utilizing the short exact sequence 0𝕋00\to\mathbb{Z}\to\mathbb{R}\to\mathbb{T}\to 0 associated to the universal covering of 𝕋\mathbb{T} the later is isomorphic to the second Čech cohomology group H2(C,)H^{2}(C,\mathbb{Z}) of CC with coefficients from \mathbb{Z} [RW98, Theorem 4.42]. By Steenrod duality [Ste40], and since CC is homotopy equivalent to a solenoid complement S3ΣS^{3}\setminus\Sigma, H2(C,)H^{2}(C,\mathbb{Z}) isomorphic to 0-th Steenrod homology group H0(C,)H_{0}(C,\mathbb{Z}).

In [BLP19] it was shown that the Čech cohomology groups for locally compact metrizable spaces, as well as the Steenrod homology groups for compact metrizable spaces, are quotients of Polish GG-spaces. Moreover all the computations described in the previous paragraph lift to Borel reductions on the level of Polish spaces; see [BLP19, Lemma 2.14, Theorem 3.12, and Section 5.5]. By [BLP19, Proposition 4.2] we have (Bun(C),isoC)(\mathrm{Bun}_{\mathbb{C}}(C),\simeq_{\mathrm{iso}}^{C}) is Borel bireducible with the orbit equivalence relation of the action of \mathbb{Z} on its dyadic profinite completion 𝟐\mathbb{Z}_{\bm{2}} by left-translation, which is Borel bireducible to the equivalence relation (2,E0)(2^{\mathbb{N}},E_{0}) of eventual equality of binary sequences. It turns out that (Bun(C),iso)(\mathrm{Bun}_{\mathbb{C}}(C),\simeq_{\mathrm{iso}}) is also Borel bireducible to (2,E0)(2^{\mathbb{N}},E_{0}). Indeed, by Borel functoriality of the definable Čech cohomology—this is proved in [BLP], but for CW-complexes it can be checked by hand—the action of Homeo(C)\mathrm{Homeo}(C) on CC induces definable endomorphisms of

0𝟐𝟐/0,0\to\mathbb{Z}\to\mathbb{Z}_{\bm{2}}\to\mathbb{Z}_{\bm{2}}/\mathbb{Z}\to 0,

as in [BLP19, Section 5.3]. It follows by [BLP19, Proposition 5.6] that (Bun(C),iso)(\mathrm{Bun}_{\mathbb{C}}(C),\simeq_{\mathrm{iso}}) is also Borel bireducible to (2,E0)(2^{\mathbb{N}},E_{0}).

Fix some point pp in CC, say the one corresponding to (1,0)𝕋0×[0,1](1,0)\in\mathbb{T}_{0}\times[0,1], and let BB be the CW-complex which is attained by taking the disjoint union of \mathbb{Z}-many copies (Ck)(C_{k}) of CC and connecting the point pp of CkC_{k} to the point pp of Ck+1C_{k+1} by gluing on them the endpoints of a homeomorphic copy of the interval [0,1][0,1]. Hence, BB is a \mathbb{Z}-line of intervals. Every homeomorphism of CC acts on the indexing copy of \mathbb{Z} by the group Γ:=(/2)\Gamma:=(\mathbb{Z}/2\mathbb{Z})\rtimes\mathbb{Z} in the obvious way. It is easy to see that the Γ\Gamma-jump E0[Γ]E^{[\Gamma]}_{0} of E0E_{0} reduces to (Bun(B),iso)(\mathrm{Bun}_{\mathbb{C}}(B),\simeq_{\mathrm{iso}}). By Theorem 1.3 and Theorem 1.5 we have that E0ΓE^{\Gamma}_{0} is not classifiable by TSI-group actions. ∎

6.3. Morita equivalence of continuous-trace CC^{*}-algebras

In what follows we will only consider separable CC^{*}-algebras AA whose spectrum A^\widehat{A} is Hausdorff. This implies that A^\widehat{A} is a locally compact metrizable space. By the Gelfand-Naimark theorem, the subclass of all such commutative CC^{*}-algebras is “locally-concretely” classified via the assignment AA^A\mapsto\widehat{A}: every two commutative CC^{*}-algebras with homeomorphic spectrum are isomorphic. The unique up to isomorphism such commutative CC^{*}-algebra of spectrum SS is simply the algebra C0(S,)C_{0}(S,\mathbb{C}), of all continuous maps from SS to \mathbb{C} which vanish at infinity. It turns out the Borel complexity of similar “local” classification problems increases drastically even in the case of continuous-trace CC^{*}-algebras, which is the closest it gets to being commutative. For more on the general theory of CC^{*}-algebras than we provide here, see [RW98, Bla09].

Let SS be a locally compact metrizable space and let 𝒦()\mathcal{K}(\mathcal{H}) be the CC^{*}-algebra of all compact operators on the separable Hilbert space. Two CC^{*}-algebra A,BA,B with spectrum SS are Morita equivalent if A𝒦()A\otimes\mathcal{K}(\mathcal{H}) and B𝒦()B\otimes\mathcal{K}(\mathcal{H}) are isomorphic as CC^{*}-algebras. In general, this isomorphism may only preserve the spectrum up to homeomorphism. When this induced homeomorphism can be taken to be idS\mathrm{id}_{S} then we say that AA and BB are Morita equivalent over SS. Any CC^{*}-algebra AA with Hausdorff spectrum can be endowed with C0(S)C_{0}(S)-module structure, where C0(S)C_{0}(S) is the collection of all continuous f:Sf\colon S\to\mathbb{C} which vanish at infinity. As a consequence, AA and BB are Morita equivalent over SS if and only if they are isomorphic via a C0(S)C_{0}(S)-linear map.

Let CTr(S)\mathrm{CTr}^{*}(S) be the space of all continuous-trace CC^{*}-algebras with spectrum SS. These are all CC^{*}-algebras AA, for which there is an open cover 𝒰\mathcal{U} of SS consisting of relatively compact sets so that, for all U𝒰U\in\mathcal{U}, if AU¯A^{\overline{U}} is the quotient algebra induced by U¯S\overline{U}\subseteq S, then AU¯A^{\overline{U}} is Morita equivalent to C(U¯,)C(\overline{U},\mathbb{C}) over U¯\overline{U}; [RW98, Proposition 5.15]. In other words continuous-trace CC^{*}-algebras are precisely the algebras which are locally Morita equivalent to commutative. It turns out that any algebra ACTr(S)A\in\mathrm{CTr}^{*}(S) can be identified with a locally trivial fibre bundle over SS, whose fibre FF is the CC^{*}-algebra 𝒦(H)\mathcal{K}(H) of all compact operators of a separable (potentially finite dimensional) Hilbert space, and the strcture group GG is Aut(𝒦(H))\mathrm{Aut}(\mathcal{K}(H)); [Bla09, IV.1.7.7, IV.1.7.8]. Hence, similarly to subsection 6.1, we may view CTr(S)\mathrm{CTr}^{*}(S) as a Polish space. The space CTrStable(S)\mathrm{CTr}_{Stable}^{*}(S) of all stable continuous-trace CC^{*}-algebras with spectrum SS is the subspace Bun(B,Aut(𝒦()),𝒦())\mathrm{Bun}(B,\mathrm{Aut}(\mathcal{K}(\mathcal{H})),\mathcal{K}(\mathcal{H})) of CTr(S)\mathrm{CTr}^{*}(S)—where \mathcal{H} is the separable infinite dimensional Hilbert space.

We consider the following two classification problems: let (CTr(S),)(\mathrm{CTr}^{*}(S),\equiv_{\mathcal{M}}) be the problem of classifying all elements of CTr(S)\mathrm{CTr}^{*}(S) up to Morita equivalence; and let (CTr(S),S)(\mathrm{CTr}^{*}(S),\equiv_{\mathcal{M}}^{S}) be the problem of classifying all elements of CTr(S)\mathrm{CTr}^{*}(S) up to Morita equivalence over SS. By a result of [BLP19], S\equiv_{\mathcal{M}}^{S} is classifiable by TSI group actions:

Proposition 6.3 ([BLP19](Corollary 5.14.)).

The problem (CTr(S),S)(\mathrm{CTr}^{*}(S),\equiv_{\mathcal{M}}^{S}) is classifiable by non-Archimedean, abelian group actions.

In contrast, for the relation \equiv_{\mathcal{M}} we have the following result.

Corollary 6.4.

There exists a locally compact metrizable topological space SS, so that (CTr(S),)(\mathrm{CTr}^{*}(S),\equiv_{\mathcal{M}}) is not classifiable by TSI group actions. In fact, SS can be taken to be the geometric realization of a countable, locally-finite, CW-complex.

Proof.

First notice that the Borel map implementing AA𝒦()A\mapsto A\otimes\mathcal{K}(\mathcal{H}), is a Borel reduction, inducing a bijection from CTr(S)/\mathrm{CTr}^{*}(S)/\equiv_{\mathcal{M}} to CTrStable(S)/\mathrm{CTr}_{Stable}^{*}(S)/\equiv_{\mathcal{M}}. Hence, it sufffices to consider the problem (CTrStable(S),)(\mathrm{CTr}_{Stable}^{*}(S),\equiv_{\mathcal{M}}) instead.

By the Dixmier-Douady classification theorem we have that CTrStable(S)/S\mathrm{CTr}_{Stable}^{*}(S)/\equiv_{\mathcal{M}}^{S} is in bijective correspondence with the third Čech cohomology group H3(S,)H^{3}(S,\mathbb{Z}) of SS with coefficients from \mathbb{Z}; see [RW98, Theorem 5.29]. It follows that CTrStable(S)/\mathrm{CTr}_{Stable}^{*}(S)/\equiv_{\mathcal{M}} is in bijective correspondence with H3(S,)/ΓH^{3}(S,\mathbb{Z})/\Gamma, where Γ\Gamma is the group of all automorphism of H3(S,)H^{3}(S,\mathbb{Z}) induced by the action of Homeo(S)\mathrm{Homeo}(S) on H3(S,)H^{3}(S,\mathbb{Z}); see [Bla09, IV.1.7.15]. Similarly to the proof of Corollary 6.2, the Dixmier-Douady correspondence and all the cohomological manipulations lift to Borel reductions on the level of a appropriate Polish spaces; see [BLP19]. The rest of the proof follows as in Corollary 6.2: let DD be the suspension C×[0,1]/C\times[0,1]/\sim of the space CC which we defined in the proof of Corollary 6.2, and let SS be attained from DD in the same way that BB was attained from CC in the same proof. Notice that by properties of the suspension we have that H3(D,)=H2(C,)H^{3}(D,\mathbb{Z})=H^{2}(C,\mathbb{Z}). ∎

Remark 6.5.

The complexity of the (CTr(S),)(\mathrm{CTr}^{*}(S),\equiv_{\mathcal{M}}) has been studied in [BLP] for several spaces SS. For example, it is shown that, when SS is the homotopy limit coming from the defining inverse system of a dd-dimensional solenoid, then (CTr(S),Mor)(\mathrm{CTr}^{*}(S),\simeq^{\mathrm{Mor}}) is always essentially countable but, when d2d\geq 2, it is not essentially treeable.

7. The space between TSI and CLI

With Corollary 1.7 we established that the class of CLI Polish groups can produce strictly more complicated orbit equivalence relations than the class of TSI groups from the point of view of Borel (or even Baire-measurable) reductions. The obvious question is how many different complexity classes lie between the class of all classification problems which are classifiable by TSI-group actions and the ones which are classifiable by CLI-group actions. In this final section we illustrate how the methods we developed here can be adapted to show that there is an ω1\omega_{1}-sequence of strictly increasing complexity classes. The discussion here will be informal since the details will be provided in an upcoming paper.

Let XX be a Polish GG and recall the unbalanced graph relation \leftrightsquigarrow that we defined between pairs of points of XX. In the context of the next definition we may refer to it as the 11-unbalanced relation and we denote it by G1\leftrightsquigarrow^{1}_{G}.

Definition 7.1.

Let XX be a Polish GG-space, let VV be an open neighborhood of the identity of GG, and let α<ω1\alpha<\omega_{1}. We define the relation Vα\leftrightsquigarrow^{\alpha}_{V} on XX by induction. Let

  1. (1)

    xV0yx\leftrightsquigarrow^{0}_{V}y, if yVx¯y\in\overline{Vx} and xVy¯x\in\overline{Vy};

  2. (2)

    xVαyx\leftrightsquigarrow^{\alpha}_{V}y, if for every open neighborhood WW of the identity of GG, and every open set UXU\subseteq X having non-empty intersection with the orbit of xx or yy, there exist vx,vyVv^{x},v^{y}\in V with vxxUv^{x}x\in U and vyyUv^{y}y\in U, so that vyyWβvxxv^{y}y\leftrightsquigarrow^{\beta}_{W}v^{x}x, for all β<α\beta<\alpha.

The α\alpha-unbalanced graph associated to GXG\curvearrowright X is the graph (X/G,Gα)(X/G,\leftrightsquigarrow^{\alpha}_{G}).

CLITSIα\alpha-balanced

In [Mal11], Malicki used iterated Wreath products to define an ω1\omega_{1}-sequence (Pαα<ω1)(P_{\alpha}\mid\alpha<\omega_{1}) of Polish permutation groups. Using this sequence Malicki established that the collection of all CLI group forms a coanalytic non-Borel subset of the standard Borel space of all Polish groups. Using the techniques we developed here one may show that the α\alpha-unbalanced graph of the Bernoulli shift of PαP_{\alpha} is generically semi-connected (see Section 1.3) and that any orbit equivalence relation with generically semi-connected β\beta-unbalanced graph is generically ergodic for actions of PαP_{\alpha}, when α<β\alpha<\beta. Moreover, in a certain weak sense, these complexity classes are cofinal in the class of all orbit equivalence relations of CLI groups: if for any pair x,yx,y of elements of a Polish GG-space XX we have xGαyx\leftrightsquigarrow^{\alpha}_{G}y for all countable ordinals α\alpha, yet yGxy\not\in Gx, then GG cannot be CLI.

References

  • [All20] S. Allison. Non-Archimedian TSI polish groups and their potential complexity spectrum. 2020.
  • [BK96] H. Becker and A. S. Kechris. The descriptive set theory of Polish group actions, volume 232 of London Mathematical Society Lecture Note Series. Cambridge University Press, 1996.
  • [Bla09] B. Blackadar. Operator algebras. Theory of C*-algebras and von Neumann algebras, volume 122 of Encyclopaedia of Mathematical Sciences. Springer, 2009.
  • [BLP] J. Bergfalk, M. Lupini, and A. Panagiotopoulos. Definable homology and cohomology and classification problems. In preparation.
  • [BLP19] J. Bergfalk, M. Lupini, and A. Panagiotopoulos. Definable (co)homology, pro-torus rigidity, and (co)homological classification. https://arxiv.org/abs/1909.03641, 2019.
  • [CC] J. Clemens and S. Coskey. New jump operators on equivalence relations, and scattered linear orders. In preparation.
  • [EM42] S. Eilenberg and S. MacLane. Group Extensions and Homology. Annals of Mathematics, 43(4):757–831, 1942.
  • [FW04] M. Foreman and B. Weiss. An anti-classification theorem for ergodic measure preserving transformations. Journal of the European Mathematical Society, 6:277–292, 2004.
  • [Gao09] S. Gao. Invariant descriptive set theory, volume 293 of Pure and Applied Mathematics. CRC Press, 2009.
  • [Hjo00] G. Hjorth. Classification and orbit equivalence relations, volume 75 of Mathematical Surveys and Monographs. American Mathematical Society, 2000.
  • [Hjo01] G. Hjorth. On invariants for measure preserving transformations. Fundamenta Mathematicae, 169:51–84, 2001.
  • [Hjo05] G. Hjorth. A dichotomy theorem for being essentially countable. In Logic and its applications, volume 380 of Contemp. Math., pages 109–127. Amer. Math. Soc., 2005.
  • [Kec95] A. S. Kechris. Classical descriptive set theory, volume 156 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1995.
  • [KMPZ20] A. S. Kechris, M. Malicki, A. Panagiotopoulos, and J. Zielinski. On Polish groups admitting non-essentially countable actions. Ergodic Theory and Dynamical Systems, 10.1017/etds.2020.133, pages 1–15, 2020.
  • [LP18] M. Lupini and A. Panagiotopoulos. Games orbits play and obstructions to Borel reducibility. Groups Geom. Dyn., 12:1461–1483, 2018.
  • [Mal11] M. Malicki. On Polish groups admitting a compatible complete left-invariant metric. J. Symbolic Logic, 76:437–447, 2011.
  • [Neu76] P. M. Neumann. The structure of finitary permutation groups. Arch. Math. (Basel), 27:3–17, 1976.
  • [RW98] I. Raeburn and D. P. Williams. Morita equivalence and continuous-trace C*-algebras, volume 60 of Mathematical Surveys and Monographs. American Mathematical Society, 1998.
  • [Ste40] N. E. Steenrod. Regular cycles of compact metric spaces. Annals of Mathematics, 41(4):833–851, 1940.
  • [vN32] John von Neumann. Zur operatorenmethode in der klassischen mechanik. Annals of Mathematics, 41(33):587–642, 1932.