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

First order rigidity of homeomorphism groups of manifolds

Sang-hyun Kim School of Mathematics, Korea Institute for Advanced Study (KIAS), Seoul, 02455, Korea [email protected] https://www.kimsh.kr Thomas Koberda Department of Mathematics, University of Virginia, Charlottesville, VA 22904-4137, USA [email protected] https://sites.google.com/view/koberdat  and  J. de la Nuez González School of Mathematics, Korea Institute for Advanced Study (KIAS), Seoul, 02455, Korea [email protected]
Abstract.

For every compact, connected manifold MM, we prove the existence of a sentence ϕM\phi_{M} in the language of groups such that the homeomorphism group of another compact manifold NN satisfies ϕM\phi_{M} if and only if NN is homeomorphic to MM. We prove an analogous statement for groups of homeomorphisms preserving Oxtoby–Ulam probability measures.

Key words and phrases:
homeomorphism group, manifold, first order theory, elementary equivalence
2020 Mathematics Subject Classification:
Primary: 20A15, 57S05, ; Secondary: 03C07, 57S25, 57M60

1. Introduction

This article relates topological manifolds, homeomorphism groups, and first order theories. For us, a manifold will mean a second countable, metrizable topological space, each point of which has a closed neighborhood homeomorphic to a fixed closed Euclidean ball. In particular, a manifold is allowed to have boundary. The first order theory (or elementary theory) of a group is the collection of the first order sentences (i.e. sentences that do not involve quantification of subsets) which are valid in the group; see Section 2.3 for details.

We begin by introducing the main objects of study. For a manifold MM (possibly equipped with a probability measure μ\mu), we let Homeo(M)\operatorname{Homeo}(M) and Homeoμ(M)\operatorname{Homeo}_{\mu}(M) denote the homeomorphism group of MM and its μ\mu–preserving subgroup, respectively. We denote by Homeo0(M)\operatorname{Homeo}_{0}(M) and Homeo0,μ(M)\operatorname{Homeo}_{0,\mu}(M) the identity components of Homeo(M)\operatorname{Homeo}(M) and Homeoμ(M)\operatorname{Homeo}_{\mu}(M), respectively. For general topological spaces XX and YY, we write XYX\cong Y if XX and YY are homeomorphic.

We denote by \mathscr{M} the class of all pairs (M,G)(M,G), where MM is a compact, connected manifold and GG is a group satisfying

Homeo0(M)GHomeo(M).\operatorname{Homeo}_{0}(M)\leq G\leq\operatorname{Homeo}(M).

We also let vol\mathscr{M}_{\operatorname{vol}} denote the class of all (M,G)(M,G) where MM is further assumed to be equipped with some Oxtoby–Ulam measure μ\mu (that is, a nonatomic Borel probability measure having full support and assigning measure zero to the boundary), and GG is a group satisfying

Homeo0,μ(M)GHomeoμ(M).\operatorname{Homeo}_{0,\mu}(M)\leq G\leq\operatorname{Homeo}_{\mu}(M).

Note that in this case, we have

Homeo0,μ(M)=Homeo0(M)Homeoμ(M);\operatorname{Homeo}_{0,\mu}(M)=\operatorname{Homeo}_{0}(M)\cap\operatorname{Homeo}_{\mu}(M);

cf. [17].

Remark 1.1.

In statements that apply to both of the classes \mathscr{M} and vol\mathscr{M}_{\operatorname{vol}}, we will often use the notation (vol)\mathscr{M}_{(\operatorname{vol})}; in such a statement, the choices of formulae may differ, even when the formulae share the same names.

We will later modify the definitions of the classes (vol)\mathscr{M}_{(\operatorname{vol})} slightly so that only manifolds of dimension at least two are considered; see the remark at the end of Section 3.3.

To motivate the discussion in this article, we consider the general reconstruction problem of an object from its group of automorphisms. For a general object XX in some category, it is natural to ask the degree to which the group of automorphisms Aut(X)\operatorname{Aut}(X) determines the object XX. This question is not completely precise, since the terms “degree” and “determine” do not have a mathematical meaning here. In our context, the object XX will always be a compact manifold, possibly with boundary, and the group of automorphisms will be one of the groups of homeomorphisms we have defined already.

The precise meaning of “degree” will be “the information encoded in the first order theory”, and “determine” will precisely mean “reconstruct the homeomorphism type”. That is, the goal of this paper is to investigate, under the assumption that (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}, the extent to which the first order theory of GG can be used to reconstruct the homeomorphism type of MM.

Of course, the first order theory of the homeomorphism group of a manifold is not the only data one can investigate for the reconstruction of the homeomorphism type of the underlying manifold. Perhaps the most basic invariant of the group of homeomorphisms of a manifold MM is its isomorphism type.

It is a classical result of Whittaker that the isomorphism type of the homeomorphism group of a compact manifold determines the homeomorphism type of the underlying manifold in the following sense:

Theorem 1.2 (See [49]).

Let MM and NN be compact manifolds, and suppose

ϕ:Homeo(M)Homeo(N)\phi\colon\operatorname{Homeo}(M)\longrightarrow\operatorname{Homeo}(N)

is an isomorphism of groups. Then there exists a homeomorphism

ψ:MN\psi\colon M\longrightarrow N

such that for all fHomeo(M)f\in\operatorname{Homeo}(M), we have ϕ(f)=ψfψ1\phi(f)=\psi\circ f\circ\psi^{-1}.

Whittaker’s result has been generalized by a number of authors; see Chapter 3 of [24] for a survey. For instance, combining the work of Bochner–Mongomery [5] on Hilbert’s fifth problem and of Takens on smooth conjugation between diffeomorphisms [45] (cf. [18]), one obtains that if MM and NN are smooth and closed, and if the diffeomorphism groups Diffk(M)\operatorname{Diff}^{k}(M) and Diff(N)\operatorname{Diff}^{\ell}(N) are isomorphic as groups, then k=k=\ell and each isomorphism between the groups is induced by some CkC^{k}–diffeomorphism between MM and NN.

In the continuous category, a different generalization was given by Rubin. We say that a topological action of a group GG on a topological space XX is locally dense if for each pair (x,U)(x,U) of a point xXx\in X and a neighborhood UXU\subseteq X of xx, the orbit ZZ of xx by the action of the group

G[U]:={gGg(y)=y for all yU}G[U]:=\{g\in G\mid g(y)=y\text{ for all }y\not\in U\}

is somewhere dense; that is, the closure of ZZ has nonempty interior. Rubin’s Theorem can be stated as followsL

Theorem 1.3 ([41]).

Let X1X_{1} and X2X_{2} be perfect, locally compact, Hausdorff topological spaces, and let GiHomeo(Xi)G_{i}\leq\operatorname{Homeo}(X_{i}) be locally dense subgroups for i{1,2}i\in\{1,2\}. If there exists an isomorphism if groups

ϕ:G1G2,\phi\colon G_{1}\longrightarrow G_{2},

then there exists a homeomorphism

ψ:X1X2\psi\colon X_{1}\longrightarrow X_{2}

such that for all gG1g\in G_{1}, we have ϕ(g)=ψgψ1\phi(g)=\psi\circ g\circ\psi^{-1}.

The reason for considering the (a priori much weaker) first order theory of a homeomorphism group instead of the full isomorphism type of the homeomorphism group is because an isomorphism between two groups of homeomorphisms is a rather unwieldy (and frankly unnatural) piece of data. Homeomorphism groups of manifolds are generally much too large to write down, and directly accessing homomorphisms between them is practically impossible. Therefore, we will be interested in more finitary ways of investigating homeomorphism groups of manifolds, namely through their elementary theories.

With this goal in mind, we consider the language of groups, which consists of a binary operation (interpreted as the group operation) and a constant (interpreted as the identity element). Models of the theory of groups are just sets with interpretations of the group operation and identity element which satisfy the axioms of groups. We say that two groups G1G_{1} and G2G_{2} are elementarily equivalent, written G1G2G_{1}\equiv G_{2}, if a first order sentence in the language of groups holds in G1G_{1} if and only if it holds in G2G_{2}; this is sometimes expressed as saying that the theories of G1G_{1} and G2G_{2} agree, i.e.

Th(G1)=Th(G2).\operatorname{{Th}}(G_{1})=\operatorname{{Th}}(G_{2}).

Here, first order refers to the scope of quantification, which is allowed to range over elements (as opposed to subsets, relations, or functions).

Philosophically, the reason for considering first order theories as opposed to second (or higher) order theories is that, whereas it is typically not controversial what “elements” in a structure refer to, the objects which are admitted as “subsets” of a structure depend on the underlying choice of set theory; there is generally no agreement on acceptable axioms for set theory. A further “constructive” benefit of the first order theory of a structure is that it is a syntactic invariant, in the sense that it records a list of “true statements” about the structure which can, in principle, be recorded.

First order rigidity in a class of structures refers to the phenomenon where two elementarily equivalent structures are automatically isomorphic. Of course, a class of structures may or may not enjoy first order rigidity, and a priori elementary equivalence is a much coarser equivalence relation than isomorphism. Because of general model-theoretic phenomena such as the upward Löwenheim–Skolem Theorem (which says roughly that once one has an infinite model of a theory then one has elementarily equivalent models of arbitrarily high cardinality), one should restrict one’s attention to models of the same cardinality; even so, for countable groups, it is not the case that elementary equivalence implies isomorphism. A typical example is the class of nonabelian free groups, wherein any two such groups are elementarily equivalent [22, 44].

The content of this paper fits within a tradition of results establishing that certain classes of structures do enjoy first order rigidity, such as lattices in higher rank [2], function fields [13, 14, 47, 36], rings [29, 20], finite–by–abelian groups [32], and linear groups [35], cf. [43]. Moreover, the themes of this paper are compatible with the philosophy that one should like to distinguish between objects that are difficult to access directly via finite syntactic proxies.

1.1. Elementary equivalence implies homeomorphism

Our main result says precisely that two compact, connected manifolds have elementarily equivalent homeomorphism groups if and only if the underlying manifolds are homeomorphic to each other. More strongly, for each compact connected manifold MM we prove the existence of a group theoretic sentence that asserts “I am homeomorphic to MM”:

Theorem 1.4.

For each compact, connected manifold MM, there exists a sentence ϕM(vol)\phi_{M}^{(\operatorname{vol})} in the language of groups such that when (N,H)(vol)(N,H)\in\mathscr{M}_{(\operatorname{vol})}, we have that

ϕM(vol)Th(H)if and only ifNM.\phi_{M}^{(\operatorname{vol})}\in\operatorname{{Th}}(H)\qquad\text{if and only if}\qquad N\cong M.

In other words, the theories of homeomorphism groups of manifolds are quasi-finitely axiomatizable within the class (vol)\mathscr{M}_{(\operatorname{vol})}, a property that is stronger than first order rigidity.

In Theorem 1.4, we emphasize that MM and NN are not assumed to have any further structure, such as a smooth or piecewise-linear structure. We thus generalize Whittaker’s result without relying on it, and produce for each manifold a finite, group–theoretic sentence that certifies homeomorphism or non–homeomorphism with the manifold. The sentences ϕM\phi_{M} and ϕMvol\phi_{M}^{\operatorname{vol}} are produced explicitly insofar as is possible, though in practice it would be a rather tedious task to record them. We also note that the connectedness hypothesis for NN can also be dropped from the theorem, thus justifying the claim in the abstract; see Corollary 3.7, for instance.

A further motivation for Theorem 1.4 that does not arise from philosophical or foundational considerations centers around the following dynamical question; a number of other related questions are enumerated in Section 9.

Question 1.5.

Let MM be a compact, connected manifold. Under what conditions is there a finitely generated (or countable) group GMHomeo(M)G_{M}\leq\operatorname{Homeo}(M) such that whenever NN is a compact manifold with dimM=dimN\dim M=\dim N on which GMG_{M} acts faithfully with a dense orbit, we have MNM\cong N?

Related results for actions of the full homeomorphism group of MM are given by Chen–Mann [10]. They show that if the identity component of Homeo(M)\operatorname{Homeo}(M) acts transitively on a connected manifold or CW–complex NN, then NN is homeomorphic to a cover of a configuration space of points of MM. In our context, we have the following immediate consequence of the downward Löwenheim–Skolem Theorem:

Corollary 1.6.

To each compact connected manifold MM one can associate a countable group GMHomeo(M)G_{M}\leq\operatorname{Homeo}(M) which is elementarily equivalent to Homeo(M)\operatorname{Homeo}(M), such that for two compact, connected manifolds MM and NN we have

GMGNif and only ifMN.G_{M}\equiv G_{N}\quad\textrm{if and only if}\quad M\cong N.

In particular GMGNG_{M}\cong G_{N} if and only if MNM\cong N.

Remark 1.7.

In [41], there is a cryptic announcement of a version of Theorem 1.4. In particular, Rubin claims that under the assumption V=LV=L (i.e. Gödel constructibility) that two arbitrary connected manifolds are homeomorphic if and only if their homeomorphism groups are elementarily equivalent; it is likely that he implicitly made a few other assumptions (e.g. excluding manifolds with boundary) to avoid trivial counterexamples such as Homeo(0,1)Homeo[0,1]\operatorname{Homeo}(0,1)\cong\operatorname{Homeo}[0,1]. To the knowledge of the authors, the paper bearing the title announced in [41] never appeared, and neither did any result (of any authors whatsoever) proving first order rigidity of homeomorphism groups of manifolds; cf. a related MathOverflow post [38]. We note that we only establish results for compact manifolds, in contrast to Rubin’s original announcement.

Rubin’s original reason for considering the assumption V=LV=L remains unclear, and perhaps the goal was to promote first order equivalence to second order equivalence, using the assumption V=LV=L to conclude the resulting second order equivalent structures are isomorphic; cf. [1]. In work that is ongoing at the time of this writing, the second and third author, together with J. Hanson and C. Rosendal have established that first order rigidity for homeomorphism groups of noncompact manifolds depends on the choice of set theory used.

Our proof of Theorem 1.4 largely consists of two parts. The first part is constructing an expansion of the language of group theory to a seemingly more powerful language, called LAGAPEL_{\operatorname{AGAPE}}. The universe of an LAGAPEL_{\operatorname{AGAPE}} structure corresponding to (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} will contain the group GG, the regular open sets RO(M)\operatorname{RO}(M) of MM, the real numbers \mathbb{R}, the set of continuous maps C0(k,)C^{0}(\mathbb{R}^{k},\mathbb{R}^{\ell}) for

k,ω={0,1,2,}k,\ell\in\omega=\{0,1,2,\ldots\}

and the discrete subsets of MM. Since the expansion is specified by first order definitions, we deduce the following, which roughly means that each sentence in the theory of AGAPE(M,G)\operatorname{AGAPE}(M,G) can be interpreted (in a way that is uniform in (M,G)(M,G)) as a sentence in the theory of the group GG; see Section 2 for a precise definition of uniform interpretation:

Theorem 1.8.

For (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}, the structure AGAPE(M,G)\operatorname{AGAPE}(M,G) is uniformly interpretable in the group structure GG.

The second part of the proof consists in showing that the AGAPE\operatorname{AGAPE} language can express the sentence that “I am homeomorphic to MM”:

Theorem 1.9.

For each (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}, there exists an LAGAPEL_{\operatorname{AGAPE}}–sentence ψM,G(vol)\psi^{(\operatorname{vol})}_{M,G} such that for all (N,H)(vol)(N,H)\in\mathscr{M}_{(\operatorname{vol})}, we have

ψM,G(vol)ThAGAPE(N,H)if and only ifNM.\psi^{(\operatorname{vol})}_{M,G}\in\operatorname{{Th}}\operatorname{AGAPE}(N,H)\quad\text{if and only if}\quad N\cong M.

By Theorem 1.8, we can interpret LAGAPEL_{\operatorname{AGAPE}}–sentences

ψM,Homeo(M)andψM,Homeoμ(M)\psi_{M,\operatorname{Homeo}(M)}\quad\textrm{and}\quad\psi_{M,\operatorname{Homeo}_{\mu}(M)}

as group theoretic sentences ϕM\phi_{M} and ϕMvol\phi_{M}^{\operatorname{vol}} respectively, which distinguish MM from all the other non-homeomorphic manifolds NN; see Lemma 2.11 for a more formal explanation. We thus obtain a proof of Theorem 1.4.

Remark 1.10.

A few of the first order rigidity results obtained in this paper can be obtained for a substantially larger class of groups of homeomorphisms which are much smaller than the full group of homeomorphisms of MM; see the recent paper [27]. For certain groups of homeomorphisms that are “sufficiently dense” in the full group of homeomorphisms (called locally approximating groups), one can prove that the first order theory of these groups determines the underlying manifold up to homotopy equivalence. The first order theory of locally approximating groups of homeomorphisms is substantially weaker than the theory of the full homeomorphism group; indeed, in [27] we can only recover the main theorem of this paper for closed, triangulated manifolds for which the Borel Conjecture holds (i.e. homotopy equivalence implies homeomorphism, which is false in general); among the main technical difficulties of this paper are constructing methods which work for all manifolds (including ones admitting no triangulation), and including manifolds with boundary. We are able to prove the main result because of the remarkable expressivity of the first order theory of the full group of homeomorphisms of a manifold.

We note finally that the first order theory of the full homeomorphism group of a manifold MM is expressive enough to interpret the full second order theory of countable subsets of Homeo(M)\operatorname{Homeo}(M), something which is not possible in a general locally approximating group of homeomorphisms; see the recent paper [26].

1.2. Outline of the paper

The paper is devoted to proving Theorem 1.4 in several steps, each of which builds on the previous. Section 2 gathers basic results from geometric topology and model theory, and fixes notation. In Section 3, we introduce the language and the structure of primary interest for us, called AGAPE\operatorname{AGAPE}. In this structure, we interpret the regular open sets in GG, and construct formulae that encode various topological properties of regular open sets. Section 4 interprets second order arithmetic using regular open sets and actions of homeomorphisms on them. Section 5 encodes individual points of a manifold, together with the exponentiation map. Section 6 interprets the dimension of a manifold, as well as certain definably parametrized embedded Euclidean balls. Section 7 definably parametrizes collar neighborhoods of the boundary of a compact manifold. Section 8 proves Theorem 1.4 by interpreting a result of Cheeger–Kister [9] and by encoding embeddings of manifolds into Euclidean spaces that are “sufficiently near” to a fixed embedding. We conclude with some questions in Section 9.

2. Preliminaries

In this section, we gather some notation, background, and generalities.

2.1. Transitivity of balls in manifolds

The high degree of transitivity of the action of homeomorphism groups on balls in manifolds is crucial for this paper. We begin with the following fundamental fact about Oxtoby–Ulam measures.

Theorem 2.1 (von Neumann [48], Oxtoby–Ulam [34]).

If μ\mu and ν\nu are Oxtoby–Ulam measures on a compact connected manifold MM, then there exists a homeomorphism hh of MM isotopic to the identity and fixing M\partial M such that hμ=νh_{*}\mu=\nu.

Thus, for Oxtoby–Ulam measures, the groups of measure-preserving homeomorphisms of MM are all conjugate to each other. In particular, each (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}} corresponds to a measure that is unique up to topological conjugacy. We will therefore refer to groups of measure-preserving homeomorphisms without specifying a particular Oxtoby–Ulam measure. We refer the reader to [17, 3] for generalities about measure-preserving homeomorphisms.

A group theoretic interpretation of (certain) balls in a manifold will be another crucial step in this paper. The importance of being able to identify regular open sets which are homeomorphic to balls comes from the following lemma, which is originally due to Brown [7] for the first two parts, and to Fathi [17] for the remainder. One may view this as a natural generalization of the fact that every compact connected 2-manifold can be obtained by gluing up the boundary of a polygon in a suitable way. See also [11, 28] for more details. In the statement of the lemma, Bn(r)B^{n}(r) means the compact ball of radius r>0r>0 with the center at the origin in n\mathbb{R}^{n}.

Lemma 2.2 ([7, 17]; cf. [11, Chapter 17]).

For each compact connected nn–manifold MM, there exists a continuous surjection

f:Bn(1)Mf\colon B^{n}(1)\longrightarrow M

such that the following hold:

  1. (i)

    the restriction of ff on intBn(1)\operatorname{int}B^{n}(1) is an embedding onto an open dense subset of MM;

  2. (ii)

    we have f(intBn(1))f(Bn(1))=f(\operatorname{int}B^{n}(1))\cap f(\partial B^{n}(1))=\varnothing, and in particular, Mf(Bn(1))\partial M\subseteq f(\partial B^{n}(1));

If MM is equipped with an Oxtoby–Ulam measure μ\mu, we can further require the following:

  1. (iii)

    we have μ(f(Bn(1)))=0\mu(f(\partial B^{n}(1)))=0;

  2. (iv)

    the measure μ\mu is the pushforward of Lebesgue measure by ff.

The conditions (i) and (ii) already imply that an Oxtoby–Ulam measure on MM exists. For instance, one can pull back the Lebesgue measure on a ball using the surjection

Bn(1)MB^{n}(1)\longrightarrow M

from Lemma 2.2. The condition (iv) is also easy to be obtained from the previous conditions and Theorem 2.1; see also [19].

For a possbily non-compact manifold, we have the following variation also due to Fathi, which loosens the condition on the surjectivity of the map.

Lemma 2.3 ([17]).

If a connected nn–manifold MM has nonempty boundary and if MM is equipped with a nonatomic, fully supported Radon measure μ\mu that assigns zero measure to M\partial M, then there exists an open embedding

f:+n:={(x1,,xn)nxn0}Mf\colon\mathbb{H}^{n}_{+}:=\{(x_{1},\ldots,x_{n})\in\mathbb{R}^{n}\mid x_{n}\geq 0\}\longrightarrow M

such that the following hold:

  1. (i)

    f(int+n)intMf(\operatorname{int}\mathbb{H}^{n}_{+})\subseteq\operatorname{int}M and f(+n)Mf(\partial\mathbb{H}^{n}_{+})\subseteq\partial M;

  2. (ii)

    Mf(+n)M\setminus f(\mathbb{H}^{n}_{+}) is closed and of measure zero.

We call a topologically embedded image of Bn(1)B^{n}(1) in a manifold MnM^{n} a ball. The same goes for an open ball in MM. If there exists an embedding

h:Bn(2)M,h\colon B^{n}(2)\longrightarrow M,

then the image h(Bn(1))h(B^{n}(1)) is called a collared ball [11, Chapter 17]. The same goes for a collared open ball. In the case when MM is equipped with an Oxtoby–Ulam measure μ\mu, we say a collared ball BB is μ\mugood (or, simply good) if B\partial B has measure zero. There exists an arbitrarily small covering of MM by μ\mu–good balls [17]. For brevity of exposition, by a good ball, we mean both a collared ball in the context of (M,G)(M,G)\in\mathscr{M} and a μ\mu–good ball in the context of (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}}. The same goes for a good open ball. Note that a good ball is always contained in the interior of MM.

Recall the topological action of a group GG on intM\operatorname{int}M is path–transitive if for all paths

γ:IintM\gamma\colon I\longrightarrow\operatorname{int}M

and for all neighborhoods UU of γ(I)\gamma(I) there exists hG[U]h\in G[U] such that h(γ(0))=γ(1)h(\gamma(0))=\gamma(1). We say the action of GG on intM\operatorname{int}M is kk–transitive if it induces a transitive action on the configuration space of kk distinct points in intM\operatorname{int}M. A path-transitive action on intM\operatorname{int}M is always kk–transitive whenever dimM>1\dim M>1; see [3, Lemma 7.4.1]. Let us note the following fundamental facts on various notions of transitivity in manifolds.

Lemma 2.4.

[28, Corollaries 2.1 and 2.2] For (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} with dimM>1\dim M>1, we have the following.

  1. (1)

    The action of GG on intM\operatorname{int}M is path–transitive and kk–transitive for all k>0k>0.

  2. (2)

    If B1B_{1} and B2B_{2} are good balls of the same measure in an open connected set UMU\subseteq M, then there exists gG[U]g\in G[U] such that g(B1)=B2g(B_{1})=B_{2}.

Proof.

The path–transitivity of part (1) is well-known; see [3, Section 7.7] for G=Homeo0(M)G=\operatorname{Homeo}_{0}(M), and [17, p. 85] for G=Homeo0,μ(M)G=\operatorname{Homeo}_{0,\mu}(M). The kk–transitivity follows immediately.

The case when U=MU=M in part (2) is precisely given in [28, Corollary 2.2] by Le Roux, based on the Annulus Theorem of Kirby [25] and Quinn [39] as well as the Oxtoby–Ulam theorem. In general, we can exhaust the topological manifold UU by a sequence of compact bounded manifolds {Mi}\{M_{i}\} so that some MiM_{i} contains B1B_{1} and B2B_{2} in its interior; this can be seen from [40], as explained in [37]. We can further require that Mi\partial M_{i} has measure zero by countable additivity. Applying Le Roux’s argument for MiM_{i}, we obtain the desired transitivity.∎

Lemma 2.5.

Let MM be a compact, connected nn–manifold with n2n\geq 2, equipped with an Oxtoby–Ulam measure μ\mu. If UMU\subseteq M is an open connected subset, then for each positive real number r<μ(U)r<\mu(U), there exists a good ball of measure precisely rr inside UU. Moreover, we may require that UBU\setminus B is connected.

Proof.

Note the general fact that for a connected open subset UU of MM and for a collared ball BB in UintMU\cap\operatorname{int}M, the set UBU\setminus B is connected; this can be seen from the fact that a collared ball is cellular, and that each celluar set is pointlike [11, Chapter 17].

Pick sufficiently small good ball QUQ\subseteq U such that the connected nn–manifold

M:=UintQM^{\prime}:=U\setminus\operatorname{int}Q

has measure larger than rr and has nonempty boundary. Applying Lemma 2.3 to MM^{\prime}, we have an open embedding

f:+nMf\colon\mathbb{H}^{n}_{+}\longrightarrow M^{\prime}

such that

f(+n)M=(MU)Q.f(\partial\mathbb{H}^{n}_{+})\subseteq\partial M^{\prime}=(\partial M\cap U)\cup\partial Q.

Since int+n\operatorname{int}\mathbb{H}^{n}_{+} is a countable increasing union of collared balls, we can find a collared ball B^\hat{B} in MM^{\prime} having measure larger than rr; moreover, we can further require that B^\hat{B} is good by countable additivity of μ\mu. Applying Theorem 2.1 to B^\hat{B}, we see that the restriction of B^\hat{B} is conjugate to a Lebesgue measure on a cube. It is then trivial to find a good ball BB^B\subseteq\hat{B} with measure precisely rr. ∎

2.2. Regular open sets and homeomorphism groups

Let XX be a topological space. If AXA\subseteq X is a subset then we write clA\operatorname{cl}A and intA\operatorname{int}A for its closure and interior, respectively, and

fr(A):=clAintA\operatorname{fr}(A):=\operatorname{cl}A\setminus\operatorname{int}A

for the frontier of AA.

A set UXU\subseteq X is regular open if U=intclUU=\operatorname{int}\operatorname{cl}U. For instance, a good ball is always regular open. The set of regular open subsets of XX forms a Boolean algebra, denoted as RO(M)\operatorname{RO}(M). In this Boolean structure, the minimal and maximal elements are the empty set and XX respectively. The meet is the intersection, and the join of two regular open sets UU and VV is given by

UV:=intcl(UV).U\oplus V:=\operatorname{int}\operatorname{cl}(U\cup V).

We write

U1U2=VU_{1}\sqcup U_{2}=V

when VV is the disjoint union of two sets U1U_{1} and U2U_{2}.

The complement coincides with the exterior:

U:=XclU.U^{\perp}:=X\setminus\operatorname{cl}U.

Consequently, the Boolean partial order UVU\leq V coincides with the inclusion UVU\subseteq V for U,VRO(X)U,V\in\operatorname{RO}(X). For each subcollection RO(X)\mathscr{F}\subseteq\operatorname{RO}(X) of regular open sets we can define its supremum as

sup:=intcl()RO(X).\sup\mathscr{F}:=\operatorname{int}\operatorname{cl}\left(\bigcup\mathscr{F}\right)\in\operatorname{RO}(X).

In particular, RO(X)\operatorname{RO}(X) is a complete Boolean algebra. We remark that the collection of open sets of a manifold (or indeed of an arbitrary topological space) is not a Boolean algebra in a natural way, but rather a Heyting algebra, since it is possible that UUU\subsetneq U^{\perp\perp}.

By a regular open cover of a space, we mean a cover consisting of regular open sets. We will repeatedly use the following straightforward fact, which implies that every finite open cover of a normal space can be refined by an open cover which consists of regular open sets.

Lemma 2.6.

If 𝒰={U1,,Um}\mathscr{U}=\{U_{1},\ldots,U_{m}\} is an open cover of a normal space, then there exists a regular open cover 𝒱={V1,,Vm}\mathscr{V}=\{V_{1},\ldots,V_{m}\} such that clViUi\operatorname{cl}V_{i}\subseteq U_{i} for each ii.

Proof.

Under the given hypothesis, one can find an open cover {Wi}\{W_{i}\} satisfying clWiUi\operatorname{cl}W_{i}\subseteq U_{i} for each ii; see [12, Corollary 1.6.4]. It then suffices for us to take Vi:=intclWiV_{i}:=\operatorname{int}\operatorname{cl}W_{i}, which is clearly a regular open set. ∎

Let gHomeo(X)g\in\operatorname{Homeo}(X). We denote its fixed point set by fixg\operatorname{fix}g, and define its (open) support as suppg:=Xfixg\operatorname{supp}g:=X\setminus\operatorname{fix}g. We then define its extended support as

suppeg:=intclsuppg=intcl(Xfixg).\operatorname{{supp}^{\mathrm{e}}}g:=\operatorname{int}\operatorname{cl}\operatorname{supp}g=\operatorname{int}\operatorname{cl}(X\setminus\operatorname{fix}g).

Let GHomeo(X)G\leq\operatorname{Homeo}(X). We define the rigid stabilizer (group) of AXA\subseteq X as

G[A]:={gGsuppgA}.G[A]:=\{g\in G\mid\operatorname{supp}g\subseteq A\}.

If UU is regular open in XX, we note that

G[U]={gGsuppegU}.G[U]=\{g\in G\mid\operatorname{{supp}^{\mathrm{e}}}g\subseteq U\}.

Recall from the introduction that the group GHomeo(X)G\leq\operatorname{Homeo}(X) is locally dense if for each nonempty open set UU and for each pUp\in U we have hat

intcl(G[U].p).\operatorname{int}\operatorname{cl}\left(G[U].p\right)\neq\varnothing.

More weakly, we say GG is locally moving if the rigid stabilizer of each nonempty open set is nontrivial.

If GG is a locally moving group of homeomorphisms of XX then RO(X)\operatorname{RO}(X) has no atoms, and the set of extended supports

{suppeggG}\{\operatorname{{supp}^{\mathrm{e}}}g\mid g\in G\}

is dense in the complete Boolean algebra RO(X)\operatorname{RO}(X), i.e.  for all URO(X)U\in\operatorname{RO}(X) there exists gGg\in G such that suppegU\operatorname{{supp}^{\mathrm{e}}}g\subseteq U; see [42] and [24, Theorem 3.6.11]. When the ambient space is a manifold, the fundamental observation is that every regular open set can actually be represented as the extended support of some homeomorphism.

Proposition 2.7.

Suppose that (M,G)(M,G)\in\mathscr{M} with dimM1\dim M\geq 1, or that (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}} with dimM>1\dim M>1. Then each regular open set of MM is the extended support of some element of GG.

Proof.

Pick a countable dense subset {xi}iω\{x_{i}\}_{i\in\omega} of UintMU\cap\operatorname{int}M. Set j1:=1j_{1}:=1, and pick a good ball B1B_{1} containing xj1=x1x_{j_{1}}=x_{1} such that diamB1<1\operatorname{diam}B_{1}<1 and such that B1UB_{1}\subseteq U. Suppose we have constructed a sequence

j1<j2<jk,j_{1}<j_{2}<\cdots j_{k},

and a disjoint collection of good balls B1,,BkB_{1},\ldots,B_{k} such that xjiBix_{j_{i}}\in B_{i} and such that

diamBi<1/i\operatorname{diam}B_{i}<1/i

for each ii; furthermore, we require that

{x1,x2,,xjk}B1BkU.\{x_{1},x_{2},\ldots,x_{j_{k}}\}\subseteq B_{1}\cup\cdots\cup B_{k}\subseteq U.

If

U=i=1kBi,U=\bigcup_{i=1}^{k}B_{i},

then we terminate the procedure; otherwise, we let jk+1j_{k+1} be the minimal index jj such that

xjW:=Ui=1kBi.x_{j}\in W:=U\setminus\bigcup_{i=1}^{k}B_{i}.

Pick a good ball Bk+1WB_{k+1}\subseteq W containing xjk+1x_{j_{k+1}} such that

diamBk+1<1/(k+1).\operatorname{diam}B_{k+1}<1/(k+1).

Thus, we build an infinite disjoint collection of good balls {Bi}iω\{B_{i}\}_{i\in\omega} in UU such that

{xi}iωiBi.\{x_{i}\}_{i\in\omega}\subseteq\bigcup_{i}B_{i}.

We claim that there exists hiGh_{i}\in G for each ii such that suppehi=intBi\operatorname{{supp}^{\mathrm{e}}}h_{i}=\operatorname{int}B_{i}. In the case where there is no measure under consideration, this is clear from the definition of a good ball. In the case when a measure μ\mu is part of the data, we first pick a homeomorphism hh in Homeo0,Leb(Bn(1),Bn(1))\operatorname{Homeo}_{0,\operatorname{Leb}}(B^{n}(1),\partial B^{n}(1)) whose fixed point set has empty interior; here, the condition that dimM>1\dim M>1 is used. Let us also pick a homeomorphism

ui:Bn(1)Bi.u_{i}\colon B^{n}(1)\longrightarrow B_{i}.

We see from Theorem 2.1 that the pullback measure of μ\mu on Bn(1)B^{n}(1) under the map uiu_{i} is conjugate to (a rescaling of) the Lebesgue measure by a homeomorphism. Hence, by conjugation and extension by the identity, we obtain a homeomorphism hiHomeo0,μ(M)h_{i}\in\operatorname{Homeo}_{0,\mu}(M) satisfying

fixhi=(MintBi)Qi\operatorname{fix}h_{i}=(M\setminus\operatorname{int}B_{i})\sqcup Q_{i}

for some closed set QiBiQ_{i}\subseteq B_{i} with empty interior. This proves the claim.

Since we have

supxd(x,hi(x))diamBi<1/i\sup_{x}d(x,h_{i}(x))\leq\operatorname{diam}B_{i}<1/i

for all ii, we see from the uniform convergence theorem that the infinite product g:=ihig:=\prod_{i}h_{i} converges in Homeo(M)\operatorname{Homeo}(M), and is isotopic to the identity. By definition,

suppeg=intcl(iintBi)=U.\operatorname{{supp}^{\mathrm{e}}}g=\operatorname{int}\operatorname{cl}\left(\bigcup_{i}\operatorname{int}B_{i}\right)=U.

Hence, this map gg satisfies the conclusion. ∎

Note that measure-preserving homeomorphism groups of compact one–manifolds are highly restricted.

Proposition 2.8.

For each compact connected one–manifold MM, there exist a group theoretic formula ϕMvol\phi_{M}^{\operatorname{vol}} such that when (N,H)vol(N,H)\in\mathscr{M}_{\operatorname{vol}}, we have that

HϕMvolH\models\phi_{M}^{\operatorname{vol}}

if and only if NN and MM are homeomorphic.

Proof.

Since Homeoμ(I)/2\operatorname{Homeo}_{\mu}(I)\cong\mathbb{Z}/2\mathbb{Z}, the group theoretic sentence ϕIvol\phi_{I}^{\operatorname{vol}} stating that there are at most two elements in the group is satisfied by a pair (N,H)vol(N,H)\in\mathscr{M}_{\operatorname{vol}} if and only if NIN\cong I. Since Homeoμ(S1)\operatorname{Homeo}_{\mu}(S^{1}) contains the abelian group Homeo0,μ(S1)SO(2,)\operatorname{Homeo}_{0,\mu}(S^{1})\cong\operatorname{SO}(2,\mathbb{R}) as the index–two subgroup, a pair of the form (S1,G)vol(S^{1},G)\in\mathscr{M}_{\operatorname{vol}} satisfies the sentence

ϕS1vol:=(γ1,γ2)[γ12γ22γ12γ22=1]¬ϕIvol.\phi_{S^{1}}^{\operatorname{vol}}:=(\forall\gamma_{1},\gamma_{2})[\gamma_{1}^{2}\gamma_{2}^{2}\gamma_{1}^{-2}\gamma_{2}^{-2}=1]\wedge\neg\phi_{I}^{\operatorname{vol}}.

Finally, if (N,H)vol(N,H)\in\mathscr{M}_{\operatorname{vol}} with dimN>1\dim N>1, then HH is not virtually abelian and hence HH does not satisfy the above formulae. ∎

2.3. First order logic

Proposition 2.8 establishes the measure-preserving case of the main theorem with dimM=1\dim M=1. Our strategy for all the other cases is to build a new language, one which is powerful enough that it can distinguish a given manifold from the other ones, but which can still be “interpreted” to the language of groups. In order to do this, let us begin with a brief review of the basic terminology from multi-sorted first order logic. Details can be found in [31, 46] and also succinctly in [4].

On the syntactic side, a (multi-sorted, first order) language \mathscr{L} is specified by logical symbols and a signature. Logical symbols include quantifiers (\forall, \exists), logical connectives (\wedge, \vee, ¬\neg, \to), the equality (==) and a countable set of variables. We often write auxiliary symbols such as parentheses or commas for the convenience of the reader.

A signature consists of sort symbols, relation symbols (also called as predicate symbols), function symbols and constant symbols. For the brevity of exposition we often regard a function or constant symbol as a special case of a relation symbol. An arity function is also in the signature, which assigns a finite tuple of sort symbols to each relation symbol. The arity function for each constant symbol is further required to assign only a single (i.e. 1–tuple of) sort symbol.

A (well-formed) \mathscr{L}–formula is a juxtaposition of the above symbols which is “valid”; the precise meaning of this validity requires a recursive definition [31], although it is intuitively clear. For instance, if PP is a relation symbol with the arity value (s,t)(s,t) for some sort symbols ss and tt, and if xx and yy are variables with sort values ss and tt, respectively, then PxyPxy is a formula. We write P(x,y)P(x,y) instead of PxyPxy for the ease of reading. The language \mathscr{L} specified by the above information is the collection of all formulae. Unquantified variables in a formula are called free, and a sentence is a formula with no free variables.

On the semantic side, we have an \mathscr{L}structure (or, an \mathscr{L}model) 𝒳\mathscr{X}, which is specified by a set |𝒳||\mathscr{X}| called the universe, a sort function σ\sigma from |𝒳||\mathscr{X}| to the set of sort symbols, and an assignment that is a correspondence from each relation symbol PP to an actual relation P𝒳P^{\mathscr{X}} among tuples of the elements in the universe. For each sort symbol ss, we call s𝒳:=σ1(s)s^{\mathscr{X}}:=\sigma^{-1}(s) the domain of ss in 𝒳\mathscr{X}. It is required that the relation P𝒳P^{\mathscr{X}} respects the arity value of PP. For instance, if PP is as in the previous paragraph, then P𝒳P^{\mathscr{X}} will be a subset of s𝒳×t𝒳s^{\mathscr{X}}\times t^{\mathscr{X}}. A function symbol is assigned the graph of some function, and often written as a function notation such as f(x)=yf(x)=y. A constant symbol is fixed as an element in the universe by an assignment. An assignment (for relations) naturally extends to an assignment ϕ𝒳\phi^{\mathscr{X}} for each formula ϕ\phi. We sometimes omit 𝒳\mathscr{X} from ϕ𝒳\phi^{\mathscr{X}} when the meaning is clear.

For an \mathscr{L}–formula ϕ\phi with a tuple of free variables x¯\underline{x}, and for a tuple a¯\underline{a} of elements in |𝒳||\mathscr{X}|, we write 𝒳ϕ(a¯)\mathscr{X}\models\phi(\underline{a}) if ϕ𝒳\phi^{\mathscr{X}} holds after a¯\underline{a} has been substituted for x¯\underline{x}. We define Th(𝒳)\operatorname{{Th}}(\mathscr{X}) as the set of all \mathscr{L}–sentences ϕ\phi such that 𝒳ϕ\mathscr{X}\models\phi.

Let p,q0p,q\geq 0, and let b¯\underline{b} be a qq–tuple of elements of |𝒳||\mathscr{X}|. A subset AA of XpX^{p} is definable (by ϕ\phi) with parameters b¯=(b1,,bq)\underline{b}=(b_{1},\ldots,b_{q}) if for some formula ϕ\phi with p+qp+q free variables, the set AA coincides with the set

𝒳(ϕ;b¯):={a¯|𝒳|p:ϕ(a¯,b¯)}.\mathscr{X}(\phi;\underline{b}):=\{\underline{a}\in|\mathscr{X}|^{p}\colon\phi(\underline{a},\underline{b})\}.

If q=0q=0 we simply say AA is definable, in which case we denote the above set as 𝒳(ϕ)\mathscr{X}(\phi). We now formalize the concept of “interpreting” a new language.

Definition 2.9.

Let 1\mathscr{L}_{1} and 2\mathscr{L}_{2} be languages. Suppose we have a class 𝒳\mathscr{X} of ordered pairs in the form (X1,X2)(X_{1},X_{2}) with XiX_{i} being an i\mathscr{L}_{i}–structure. We say X2X_{2} is interpretable in X1X_{1} uniformly for (X1,X2)(X_{1},X_{2}) in 𝒳\mathscr{X} if there exist some 1\mathscr{L}_{1}–formulae ϕdom\phi_{\mathrm{dom}} and ϕeq\phi_{\mathrm{eq}}, and there also exists a map α\alpha from the set of 2\mathscr{L}_{2}–formulae to the set of 1\mathscr{L}_{1}–formulae such that the following hold.

for each (X1,X2)𝒳(X_{1},X_{2})\in\mathscr{X}, we have a surjection

ρ:X1(ϕdom)|X2|\rho\colon X_{1}(\phi_{\operatorname{dom}})\longrightarrow|X_{2}|

with its fiber uniformly defined by ϕeq\phi_{\operatorname{eq}} in the sense that

X1(ϕeq)={(x,y)X1(ϕdom)×X1(ϕdom)ρ(x)=ρ(y)}.X_{1}(\phi_{\operatorname{eq}})=\{(x,y)\in X_{1}(\phi_{\operatorname{dom}})\times X_{1}(\phi_{\operatorname{dom}})\mid\rho(x)=\rho(y)\}.

Furthermore, it is required for each 2\mathscr{L}_{2}–formula ψ\psi that

ρ1(X2(ψ))=X1(α(ψ)).\rho^{-1}(X_{2}(\psi))=X_{1}(\alpha(\psi)).

The bijection

ρ1:|X2|X1(ϕdom)/X1(ϕeq)\rho^{-1}\colon|X_{2}|\longrightarrow X_{1}(\phi_{\operatorname{dom}})/X_{1}(\phi_{\operatorname{eq}})

along with the map α\alpha is called a uniform interpretation of X2X_{2} in X1X_{1}.

Remark 2.10.
  1. (1)

    In the above, if ψ\psi is mm–ary (as a relation) and ϕdom\phi_{\operatorname{dom}} is nn–ary, then α(ψ)\alpha(\psi) is mnmn–ary. In practice, we only need to consider relation symbols (in a broad sense, including function and constant symbols) ψ\psi rather than all possible 2\mathscr{L}_{2}–formulae.

  2. (2)

    In various instances of this paper, it will be the case that 12\mathscr{L}_{1}\subseteq\mathscr{L}_{2} and that the interpretation restricts to the identity on 1\mathscr{L}_{1}. As a consequence of such interpretability, we will have that Th(X2)\operatorname{{Th}}(X_{2}) is a conservative extension of Th(X1)\operatorname{{Th}}(X_{1}) for each (X1,X2)𝒳(X_{1},X_{2})\in\mathscr{X}. Also, we will often add a function symbol in 2\mathscr{L}_{2} corresponding to the surjection ρ\rho, which is clearly justified.

The following lemma explains how the combination of Theorems 1.8 and 1.9 implies Theorem 1.4.

Lemma 2.11.

Suppose 1,2\mathscr{L}_{1},\mathscr{L}_{2} and 𝒳\mathscr{X} are as in Definition 2.9 so that X2X_{2} is interpretable in X1X_{1} uniformly for (X1,X2)𝒳(X_{1},X_{2})\in\mathscr{X}. Let (X1,X2)(X_{1},X_{2}) and (Y1,Y2)(Y_{1},Y_{2}) be in 𝒳\mathscr{X}. Then for each sentence ψ\psi belonging to Th(X2)Th(Y2)\operatorname{{Th}}(X_{2})\setminus\operatorname{{Th}}(Y_{2}), the interpretation α(ψ)\alpha(\psi) belongs to Th(X1)Th(Y1)\operatorname{{Th}}(X_{1})\setminus\operatorname{{Th}}(Y_{1}). In particular, if X1Y1X_{1}\equiv Y_{1}, then X2Y2X_{2}\equiv Y_{2}.

3. The AGAPE structure and basic observations

The fundamental universe that we work in will be the group of homeomorphisms of a manifold. Objects such as regular open sets, real numbers, points in the manifold, continuous functions, etc. will all be constructed as definable equivalence classes of definable subsets of finite tuples of homeomorphisms.

3.1. The langauge LAGAPEL_{\operatorname{AGAPE}} and the structure AGAPE(M,G)\operatorname{AGAPE}(M,G)

The ultimate language we will work in will be called AGAPE\operatorname{AGAPE}, which stands for “Action of a Group, Analysis, Points and Exponentiation”. This language is denoted as LAGAPEL_{\operatorname{AGAPE}} and contains the following different sort symbols for k,ωk,\ell\in\omega:

𝐆,RO,,𝒫(),,𝐌,𝐌discint,Contk,.\mathbf{G},\operatorname{RO},\mathbb{N},\mathscr{P}(\mathbb{N}),\mathbb{R},\mathbf{M},\mathbf{M}^{\operatorname{disc-int}},\operatorname{Cont}_{k,\ell}.

The above sorts come with some symbols that are intrinsic to the sort (such as a group operation), and others which relate the sorts to each other, as we spell out below. There will be a countable set of variables for each sort, as is typically required. We also describe an AGAPE\operatorname{AGAPE} structure assigned to each pair (M,G)(M,G) in the class \mathscr{M} or vol\mathscr{M}_{\operatorname{vol}}. In this structure, we give the “intended” choice of the domain of each sort symbol.

The group sort. The domain of the sort symbol 𝐆\mathbf{G} will be the group GG, under our standing assumption that (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}. The signatures only relevant for this sort are

1,,,11,\circ,{}^{-1},

which are respectively assigned with the natural meanings in the group theory. These symbols, along with variables, form the language of groups LAct0=L𝐆L^{0}_{\operatorname{{Act}}}=L_{\mathbf{G}}. The group GG is regarded an L𝐆L_{\mathbf{G}}-structure Act0(M,G)=Act𝐆(M,G)\operatorname{{Act}}^{0}(M,G)=\operatorname{{Act}}_{\mathbf{G}}(M,G). We will usually not write the \circ symbol.

The sort of regular open sets. The domain of the sort symbol RO\operatorname{RO} is the set RO(M)\operatorname{RO}(M) of the regular open sets in MM. The newly introduced signatures for this sort are

,,,,,𝐌,suppe,appl.\subseteq,\cap,\,^{\perp},\oplus,\varnothing,\mathbf{M},\operatorname{{supp}^{\mathrm{e}}},\operatorname{appl}.

The symbol 𝐌\mathbf{M} means the manifold MM in the structure. By the natural assignment as before, we have Boolean symbols

,,,,𝐌,\subseteq,\cap,\oplus,\,^{\perp},\mathbf{M},\varnothing

for the Boolean algebra RO(M)\operatorname{RO}(M). We let the function symbol suppe\operatorname{{supp}^{\mathrm{e}}} mean the map GRO(M)G\longrightarrow\operatorname{RO}(M) defined as

gsuppeg.g\mapsto\operatorname{{supp}^{\mathrm{e}}}g.

We have an assignment for appl\operatorname{appl} so that

appl(g,U)=g(U)\operatorname{appl}(g,U)=g(U)

with gGg\in G and URO(M)U\in\operatorname{RO}(M). The symbols introduce so far (along with countably many variables for each sort) form the language of a group action on a Boolean algebra LAct1=L𝐆,ROL^{1}_{\operatorname{{Act}}}=L_{\mathbf{G},\operatorname{RO}}. The LAct1L^{1}_{\operatorname{{Act}}}–structure described above on the universe GRO(M)G\sqcup\operatorname{RO}(M) is denoted as Act1(M,G)=Act𝐆,RO(M,G)\operatorname{{Act}}^{1}(M,G)=\operatorname{{Act}}_{\bf G,\operatorname{RO}}(M,G).

The sorts from the analysis We then introduce new sort symbols, which are ,𝒫(),\mathbb{N},\mathscr{P}(\mathbb{N}),\mathbb{R} and k,\cong_{k,\ell} for k,ωk,\ell\in\omega. The signatures introduced here are

0,1,<,+,×,,,#π0,norm.0,1,<,+,\times,\in,\subseteq,\#\pi_{0},\operatorname{norm}.

Standard second order arithmetic

Arith2=(,𝒫(),0,1,<,+,×,,)\operatorname{Arith}_{2}=(\mathbb{N},\mathscr{P}(\mathbb{N}),0,1,<,+,\times,\in,\subseteq)

is given the sort symbols \mathbb{N} and 𝒫()\mathscr{P}(\mathbb{N}), as well as with relevant non-logical symbols. We note the ambiguity of our notation that the sort symbols \mathbb{N} and 𝒫()\mathscr{P}(\mathbb{N}) will be assigned with the set of the natural numbers \mathbb{N} and its power set 𝒫()\mathscr{P}(\mathbb{N}), respectively. The symbol #π0\#\pi_{0} is interpreted so that

#π0(U)=k\#\pi_{0}(U)=k

means URO(M)U\in\operatorname{RO}(M) has kk connected components. See Section 4 for details. The ordered ring of the real numbers

{0,1,+,×,<,=}\{0,1,+,\times,<,=\}

is assigned with the sort symbol \mathbb{R} and the signatures above. Note that, as is usual, \mathbb{N} is considered as a subsort of \mathbb{R}, by identifying each integer as a real number.

The domain of the sort symbol Contk,\operatorname{Cont}_{k,\ell} will be the set C(k,)C(\mathbb{R}^{k},\mathbb{R}^{\ell}) of continuous functions. We also have a formula appl(f,x)=y\operatorname{appl}(f,x)=y when the sort value of ff is Contk,\operatorname{Cont}_{k,\ell}, and when xx and yy are tuples of variables assigned with the sort symbol \mathbb{R}. We have the C0C^{0}–norm

fnorm(f):=f,f\mapsto\operatorname{norm}(f):=\|f\|,

which will be also a part of the language. Combining these symbols with LAct1L^{1}_{\operatorname{{Act}}}, we obtain the language LAct2=L𝐆,RO,L^{2}_{\operatorname{{Act}}}=L_{\mathbf{G},\operatorname{RO},\mathbb{R}}. An LAct2L^{2}_{\operatorname{{Act}}}–structure Act2(M,G)=Act𝐆,RO,\operatorname{{Act}}^{2}(M,G)=\operatorname{{Act}}_{\mathbf{G},\operatorname{RO},\mathbb{R}} is assigned to each (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} having the universe

GRO(M)𝒫()k,C(k,).G\sqcup\operatorname{RO}(M)\sqcup\mathscr{P}(\mathbb{N})\sqcup\mathbb{R}\sqcup\bigsqcup_{k,\ell}C(\mathbb{R}^{k},\mathbb{R}^{\ell}).

The point and the discrete subset sorts 𝐌\mathbf{M} and 𝐌discint\mathbf{M}^{\operatorname{disc-int}}. The domain of the sort symbol 𝐌\mathbf{M} will be the set of the points in a manifold. We also introduce the sort symbol 𝐌discint\mathbf{M}^{\operatorname{disc-int}} to mean a subset AA of intM\operatorname{int}M every point of which is isolated in AA. By abuse of notation, the symbols \in and \subseteq introduced above will have multiple meanings (depending on the context), so that they have the arity values (𝐌,RO)(\mathbf{M},\operatorname{RO}), (𝐌,𝐌discint)(\mathbf{M},\mathbf{M}^{\operatorname{disc-int}}) and (𝐌discint,RO)(\mathbf{M}^{\operatorname{disc-int}},\operatorname{RO}).

We also have a cardinality function

#A=m\#A=m

meaning that the cardinality of AintMA\subseteq\operatorname{int}M is mm, assuming that every point in AA is isolated.

The interpretation of points of the manifold will allow us to include symbols such as cl\operatorname{cl} and fr\operatorname{fr}, the closure and frontier of a regular open set, together with membership relations into these sets. These symbols will simply be abbreviations for formulae which impose the intended meaning. We will be able to separate out boundary points of MM from the interior ones, and hence justified to use the notations

π𝐌,πint𝐌\pi\in\partial\mathbf{M},\pi^{\prime}\in\operatorname{int}\mathbf{M}

for point sort variables π\pi and π\pi^{\prime}. The function symbol appl\operatorname{appl} has a natural additional meaning as below:

appl:G×MM.\operatorname{appl}\colon G\times M\longrightarrow M.

In all contexts, we abbreviate appl(γ,x)\operatorname{appl}(\gamma,x) by γ(x)\gamma(x) when the sort of γ\gamma is either 𝐆\mathbf{G} or Contk,\operatorname{Cont}_{k,\ell} and when the sort of xx is (tuples of) \mathbb{R}, 𝐌,𝐌discint\mathbf{M},\mathbf{M}^{\operatorname{disc-int}} or RO\operatorname{RO}.

The omnibus language, combining all of the previous sorts and relevant symbols, is denoted by

LAct3=L𝐆,RO,,𝐌=LAGAPE,L^{3}_{\operatorname{{Act}}}=L_{\mathbf{G},\operatorname{RO},\mathbb{R},\mathbf{M}}=L_{\operatorname{AGAPE}},

or simply as AGAPE\operatorname{AGAPE}. We have so far described the LAGAPEL_{\operatorname{AGAPE}}–structure Act3(M,G)=AGAPE(M,G)\operatorname{{Act}}^{3}(M,G)=\operatorname{AGAPE}(M,G) corresponding to (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}.

Dealing with these structures, we often make use of functions or relations defined by fixed formulae that are not explicitly specified. The following terminology will be handy when we need to avoid ambiguity in such situations:

Definition 3.1.

Let i=0,1,2,3i=0,1,2,3, and let ϕ(vol)\phi_{(\operatorname{vol})} be a formula in LActiL^{i}_{\operatorname{{Act}}}. Suppose for each (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} that a function or relation fM,Gf_{M,G} is defined by ϕ\phi in Acti(M,G)\operatorname{{Act}}^{i}(M,G). Then the collection

{fM,G(M,G)(vol)}\{f_{M,G}\mid(M,G)\in\mathscr{M}_{(\operatorname{vol})}\}

is said to be uniformly defined over (vol)\mathscr{M}_{(\operatorname{vol})}.

Remark 3.2.

In dealing with the sorts in Subsection 2.3, we will distinguish notationally between variables referring to a particular sort and elements of that sort. For the convenience of the reader, we will record a table summarizing the notation. In general, we will write an underline to denote an arbitrary (or simply unspecified) finite tuple of variables or objects.

Sort variable object
Group elements γ\gamma, δ\delta, γ¯\underline{\gamma}, δ¯\underline{\delta} gg, hh
Regular open sets uu, vv, ww, u¯\underline{u}, v¯\underline{v}, w¯\underline{w} UU, VV, WW
Natural numbers α\alpha, β\beta, α¯\underline{\alpha}, β¯\underline{\beta} k,m,nk,m,n
Sets of natural numbers Λ\Lambda, Λ¯\underline{\Lambda} AA
Real numbers ρ\rho, σ\sigma, ρ¯\underline{\rho}, σ¯\underline{\sigma} rr, ss
Sets of points π\pi, π¯\underline{\pi}, τ\tau, τ¯\underline{\tau} pp, qq, TT
Functions χ\chi, θ\theta, χ¯\underline{\chi}, θ¯\underline{\theta} ff

From now on, we will reserve the letters in this table for exclusive use as variables or objects of a particular sort, unless specified otherwise. In the ambient metalanguage, we will use i,jωi,j\in\omega to denote indices. The symbols MM and NN will be reserved for manifolds.

3.2. Interpreting action structures in homeomorphism groups

Since the uniform interpretability (Definition 2.9) is transitive, the following proposition would trivially imply Theorem 1.8.

Proposition 3.3.

For each i=0,1,2i=0,1,2, and uniformly for (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}, the LActi+1L^{i+1}_{\operatorname{{Act}}}–structure Acti+1(M,G)\operatorname{{Act}}^{i+1}(M,G) is interpretable in the LActiL^{i}_{\operatorname{{Act}}}–structure Acti(M,G)\operatorname{{Act}}^{i}(M,G).

The proof of this proposition will require the construction of LActiL_{\operatorname{{Act}}}^{i}–formulae ϕdomi\phi_{\operatorname{dom}}^{i} and ϕeqi\phi_{\operatorname{eq}}^{i}, and a surjection

ρi:Acti(M,G)(ϕdomi)|Acti+1(M,G)|\rho_{i}\colon\operatorname{{Act}}^{i}(M,G)(\phi_{\operatorname{dom}}^{i})\longrightarrow|\operatorname{{Act}}^{i+1}(M,G)|

for all (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} satisfying the conditions of Definition 2.9. Our construction will occupy Sections 4 and 5, as well as most of this section.

Rubin’s Theorem [41, 42] stated in the introduction can be used to prove various reconstruction theorems, by which we mean that group isomorphism types greatly restrict the homeomorphism types of spaces on which groups can act nicely. See [24] for comprehensive references on this, especially regarding diffeomorphism groups.

A key step in the proof of Rubin’s theorem can be rephrased as follows. We emphasize that the formulae below are independent of the choice of the group GG or the space XX.

Theorem 3.4 (Rubin’s Expressibility Theorem, cf. [42]).

There exist first order formulae

(γ1,γ2),appl(γ1,γ2,γ3),(γ1,γ2,γ3),(γ1,γ2,γ3),(γ1,γ2)\subseteq(\gamma_{1},\gamma_{2}),\quad\operatorname{appl}(\gamma_{1},\gamma_{2},\gamma_{3}),\quad\cap(\gamma_{1},\gamma_{2},\gamma_{3}),\quad\oplus(\gamma_{1},\gamma_{2},\gamma_{3}),\quad\perp(\gamma_{1},\gamma_{2})

in the language of groups such that if GG be a locally moving group of homeomorphisms of a Hausdorff topological space XX, then the following hold for all g1,g2,g3Gg_{1},g_{2},g_{3}\in G.

  1. (1)

    G(g1,g2)suppeg1suppeg2g1G[suppeg2]\displaystyle G\models\subseteq(g_{1},g_{2})\Longleftrightarrow\operatorname{{supp}^{\mathrm{e}}}g_{1}\subseteq\operatorname{{supp}^{\mathrm{e}}}g_{2}\Longleftrightarrow g_{1}\in G[\operatorname{{supp}^{\mathrm{e}}}g_{2}].

  2. (2)

    Gappl(g1,g2,g3)appl(g1,suppeg2)=suppeg3\displaystyle G\models\mathrm{appl}(g_{1},g_{2},g_{3})\Longleftrightarrow\operatorname{appl}(g_{1},\operatorname{{supp}^{\mathrm{e}}}g_{2})=\operatorname{{supp}^{\mathrm{e}}}g_{3}.

  3. (3)

    G(g1,g2,g3)suppeg1suppeg2=suppeg3\displaystyle G\models\cap(g_{1},g_{2},g_{3})\Longleftrightarrow\operatorname{{supp}^{\mathrm{e}}}g_{1}\cap\operatorname{{supp}^{\mathrm{e}}}g_{2}=\operatorname{{supp}^{\mathrm{e}}}g_{3}.

  4. (4)

    G(g1,g2,g3)suppeg1suppeg2=suppeg3\displaystyle G\models\oplus(g_{1},g_{2},g_{3})\Longleftrightarrow\operatorname{{supp}^{\mathrm{e}}}g_{1}\oplus\operatorname{{supp}^{\mathrm{e}}}g_{2}=\operatorname{{supp}^{\mathrm{e}}}g_{3}.

  5. (5)

    Gext(g1,g2)suppeg1=(suppeg2)\displaystyle G\models\operatorname{ext}(g_{1},g_{2})\Longleftrightarrow\operatorname{{supp}^{\mathrm{e}}}g_{1}=(\operatorname{{supp}^{\mathrm{e}}}g_{2})^{\perp}.

Proof.

Parts (1) and (2) are given as Theorem 2.5 in [41]; see also [24, Corollary 3.6.9] for a concrete formula. The remaining items are clear from the fact that the supremum in RO(M)\operatorname{RO}(M) is first order expressible in terms of the inclusion relation. ∎

Let (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}}. By Proposition 2.7, we have a surjection

ρ0:GRO(M)\rho_{0}\colon G\longrightarrow\operatorname{RO}(M)

defined as gsuppegg\mapsto\operatorname{{supp}^{\mathrm{e}}}g. Since GG is locally moving on MM, Rubin’s expressibility theorem implies that the fiber

{(g,h)suppeg=suppeh}\{(g,h)\mid\operatorname{{supp}^{\mathrm{e}}}g=\operatorname{{supp}^{\mathrm{e}}}h\}

of ρ0\rho_{0} is definable, and that the Boolean symbols and the function symbols appl\operatorname{appl} and suppe\operatorname{{supp}^{\mathrm{e}}} have group theoretic interpretations; see also parts (1) and (2) of Remark 2.10. We conclude the following, which shows that Proposition 3.3 holds for the case i=0i=0.

Corollary 3.5.

Uniformly for (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}, the L𝐆,ROL_{\bf G,\operatorname{RO}}–structure Act𝐆,RO(M,G)\operatorname{{Act}}_{\bf G,\operatorname{RO}}(M,G) is interpretable in the group structure GG.

Corollary 3.5 can be summarized as saying that GG interprets the group action structure of GG on the algebra of regular open sets, in a way that preserves the meaning of GG. This interpretation is uniform in the underlying pair (M,G)(M,G), and any formula in the language of GG and RO\operatorname{RO} can be expressed entirely in GG, since the formulae in Theorem 3.4 are independent of MM. Henceforth, we will assume that we work in the expanded language LAct1=L𝐆,ROL_{\operatorname{{Act}}}^{1}=L_{\mathbf{G},\operatorname{RO}}.

3.3. First order descriptions of basic topological properties

Recall that whenever the expression UVU\sqcup V is used it is assumed that UU and VV are disjoint.

We now produce first order expressions for some standard point-set–topological properties.

Lemma 3.6.

The following hold for (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}.

  1. (1)

    For U,VRO(M)U,V\in\operatorname{RO}(M), we have that G[U]=G[V]G[U]=G[V] if and only if U=VU=V.

  2. (2)

    For each URO(M)U\in\operatorname{RO}(M), we have that

    G[U]={gGg(V)=V for all regular open set VU}.G[U]=\{g\in G\mid g(V)=V\text{ for all regular open set }V\subseteq U^{\perp}\}.
  3. (3)

    An open subset is path-connected if and only if it is connected.

  4. (4)

    An arbitrary union of connected components of a regular open set is necessarily regular open. More specifically, if a regular open set WW can be written as W=UVW=U\sqcup V for some disjoint pair of open sets UU and VV, then UU and VV are regular open and W=UVW=U\oplus V. Moreover, we have V=WUV=W\cap U^{\perp}.

  5. (5)

    For disjoint pair U,VU,V of regular open sets, we have (i)\Rightarrow(ii)\Rightarrow(iii).

    1. (i)

      UU is connected, and UV=UVU\oplus V=U\sqcup V;

    2. (ii)

      Every gG[UV]g\in G[U\oplus V] satisfies either g(U)=Ug(U)=U or g(U)U=g(U)\cap U=\varnothing;

    3. (iii)

      UV=UVU\oplus V=U\sqcup V

  6. (6)

    Let WW and UU are regular open sets such that UU is connected and such that UWU\subseteq W. Then UU is a connected component of WW if and only if W=UVW=U\oplus V for some regular open VV that is disjoint from UU, and every gG[W]g\in G[W] satisfies either g(U)=Ug(U)=U or g(U)U=g(U)\cap U=\varnothing.

  7. (7)

    The following are all equivalent for a regular open set WW.

    1. (i)

      WW is disconnected;

    2. (ii)

      W=UVW=U\sqcup V for some disjoint pair of nonempty regular open sets UU and VV such that UU is connected;

    3. (iii)

      W=UVW=U\oplus V for some disjoint pair of nonempty regular open sets UU and VV, and every gG[W]g\in G[W] satisfies either g(U)=Ug(U)=U or g(U)U=g(U)\cap U=\varnothing;

    4. (iv)

      W=UV=UVW=U\oplus V=U\sqcup V for some disjoint pair of nonempty regular open sets UU and VV.

  8. (8)

    For two regular open subsets UU and VV satisfying UV=U\cap V=\varnothing, we have that UV=UVU\sqcup V=U\oplus V if and only if each connected component of UVU\oplus V is contained either in UU or in VV.

Proof.

(1) If xUVx\in U\setminus V, then there exists some hG[U]h\in G[U] satisfying h(x)xh(x)\neq x; see [24, Lemma 3.2.3] for instance. In particular, we have

hG[U]G[V].h\in G[U]\setminus G[V]\neq\varnothing.

This proves the nontrivial part of the given implication. We remark that the same statement holds without the assumption that UU and VV are regular open, under the extra hypothesis that M≇IM\not\cong I. Part (2) is similar.

(3) This part is clear from the fact that every manifold is locally path-connected.

(4) Whenever two open sets UU and VV are disjoint we have that UU^{\perp\perp} and VV^{\perp\perp} are also disjoint; see [24, Lemma 3.6.4 (4)], for instance. From

W=UVUVUVW=W,W=U\sqcup V\subseteq U^{\perp\perp}\sqcup V^{\perp\perp}\subseteq U^{\perp\perp}\oplus V^{\perp\perp}\subseteq W^{\perp\perp}=W,

we see that UU and VV are actually regular open and W=UV=UVW=U\oplus V=U\sqcup V. It is clear that VfrU=V\cap\operatorname{fr}U=\varnothing, which implies V=WUV=W\cap U^{\perp}.

(5) The implication (i)\Rightarrow(ii) is clear from that every setwise stabilizer of gG[UV]g\in G[U\oplus V] permutes connected components of UVU\oplus V.

For the implication (ii)\Rightarrow(iii), assume we have a point

p1(UV)(UV).p_{1}\in(U\oplus V)\setminus(U\cup V).

Take a sufficiently small open ball BB around p1p_{1} so that

BUV=intcl(UV)clUclV.B\subseteq U\oplus V=\operatorname{int}\operatorname{cl}(U\cup V)\subseteq\operatorname{cl}U\cup\operatorname{cl}V.

Note also that because

p1U=intclU,p_{1}\not\in U=\operatorname{int}\operatorname{cl}U,

it follows that BclUB\not\subseteq\operatorname{cl}U. Similarly, BclVB\not\subseteq\operatorname{cl}V. This implies that we can choose distinct points

p2,p3BUintMp_{2},p_{3}\in B\cap U\cap\operatorname{int}M

and p4BVintMp_{4}\in B\cap V\cap\operatorname{int}M. Since GG is kk–transitive on BintMB\cap\operatorname{int}M for all kk, we can find a gGg\in G supported in BB satisfying g(p2)=p3g(p_{2})=p_{3} and g(p3)=p4g(p_{3})=p_{4}; see also Lemma 2.4. Then g(U)g(U) is neither UU nor disjoint from UU.

Parts (6) and (7) are clear from the previous parts.

(8) The forward direction comes from the observation that (U,V)(U,V) is a disconnection of UVU\oplus V. The backward direction is trivial since the hypothesis implies that UVUVU\oplus V\subseteq U\cup V. ∎

Let us note the following consequences of Lemma 3.6.

Corollary 3.7.

There exist first order formulae in the language L𝐆,ROL_{\bf{G},\operatorname{RO}} as follows:

  1. (1)

    A formula contained(γ,u)\operatorname{contained}(\gamma,u), also abbreviated as γ𝐆[u]\gamma\in\mathbf{G}[u] such that

    contained(g,U)if and only if suppegU.\models\operatorname{contained}(g,U)\quad\textrm{if and only if }\operatorname{{supp}^{\mathrm{e}}}g\subseteq U.
  2. (2)

    A formula conn(u)\operatorname{conn}(u) such that

    conn(U)if and only if U is connected.\models\operatorname{conn}(U)\quad\textrm{if and only if $U$ is connected.}
  3. (3)

    A formula cc(u,v)\operatorname{cc}(u,v), also abbreviated as uπ0(v)u\in\pi_{0}(v) such that

    cc(U,V)if and only if U is a connected component of V.\models\operatorname{cc}(U,V)\quad\textrm{if and only if $U$ is a connected component of $V$.}
  4. (4)

    A formula ucc(u,v)\operatorname{ucc}(u,v) such that

    ucc(U,V)if and only if U is a union of connected component of V.\models\operatorname{ucc}(U,V)\quad\textrm{if and only if }U\textrm{ is a union of connected component of }V.
  5. (5)

    For all kωk\in\omega, a formula #cck(u)\#\operatorname{cc}_{k}(u) such that

    #cck(U)if and only if U has exactly k connected components.\models\#\operatorname{cc}_{k}(U)\quad\textrm{if and only if $U$ has exactly $k$ connected components.}
  6. (6)

    A formula disj(u,v)\operatorname{disj}(u,v) such that

    disj(U,V) if and only if UV=UV.\models\operatorname{disj}(U,V)\textrm{ if and only if }U\oplus V=U\sqcup V.
  7. (7)

    A formula ccpartition(u,v,w)\operatorname{ccpartition}(u,v,w) such that

    ccpartition(U,V,W)\displaystyle\models\operatorname{ccpartition}(U,V,W) if and only if ucc(U,W)ucc(V,W)W=UV.\displaystyle\quad\text{if and only if }\operatorname{ucc}(U,W)\wedge\operatorname{ucc}(V,W)\wedge W=U\sqcup V.
  8. (8)

    A formula #=(u,v)\#_{=}(u,v) such that for all regular open sets UU and VV having finitely many connected components, we have

    #=(U,V)\models\#_{=}(U,V)

    if and only if UU and VV have the same number of connected components.

Proof.

The existence of the formula contained(γ,u)\operatorname{contained}(\gamma,u) is trivial since suppe\operatorname{supp}^{e} and \subseteq belong to the signature of L𝐆,ROL_{\mathbf{G},\operatorname{RO}}. The formulae conn(u)\operatorname{conn}(u) and cc(u,v)\operatorname{cc}(u,v) exist by parts (6) and (7) of Lemma 3.6. We can then set

ucc(u,u)(w)[cc(w,u)cc(w,u)].\operatorname{ucc}(u^{\prime},u)\equiv(\forall w)[\operatorname{cc}(w,u^{\prime})\rightarrow\operatorname{cc}(w,u)].

The construction of the formulae cck(u)\operatorname{cc}_{k}(u) and disj(u,v)\operatorname{disj}(u,v) follows from the same lemma, which also implies that the formula

ccpartition(u,v,w)[ucc(u,w)ucc(v,w)w=uvuv=]\operatorname{ccpartition}(u,v,w)\equiv[\operatorname{ucc}(u,w)\wedge\operatorname{ucc}(v,w)\wedge w=u\oplus v\wedge u\cap v=\varnothing]

has the meaning required in part (7). Finally, we set

#=(u,v)\displaystyle\#_{=}(u,v)\equiv (uu,vv)[(u^π0(u))[conn(uu^)uu^]\displaystyle(\exists u^{\prime}\subseteq u,\exists v^{\prime}\subseteq v)[(\forall\hat{u}\in\pi_{0}(u))[\operatorname{conn}(u^{\prime}\cap\hat{u})\wedge u^{\prime}\cap\hat{u}\neq\varnothing]\wedge
(v^π0(v))[conn(vv^)vv^](γ)[γ(u)=v]].\displaystyle(\forall\hat{v}\in\pi_{0}(v))[\operatorname{conn}(v^{\prime}\cap\hat{v})\wedge v^{\prime}\cap\hat{v}\neq\varnothing]\wedge(\exists\gamma)[\gamma(u^{\prime})=v^{\prime}]].

From the transitivity on good balls (of equal measure, in the measure preserving case) as in Lemma 2.4, we see #=(u,v)\#_{=}(u,v) has the intended meaning. ∎

Using the above formula, we can distinguish the case that dimM=1\dim M=1 among all compact connected manifolds.

Corollary 3.8.

For each compact connected one–manifold MM, there exist L𝐆,ROL_{\bf G,\operatorname{RO}}–formulae ϕM\phi_{M} such that when (N,H)(N,H)\in\mathscr{M}, we have that

Act𝐆,RO(N,H)ϕM\operatorname{{Act}}_{\bf G,\operatorname{RO}}(N,H)\models\phi_{M}

if and only if NN and MM are homeomorphic.

Proof.

We let ϕI\phi_{I} be the L𝐆,ROL_{\bf G,\operatorname{RO}}–formula expressing that for all pairwise disjoint, proper, nonempty regular open sets U1,U2U_{1},U_{2} and U3U_{3} the exterior of UiU_{i} is disconnected for some ii. This formula holds for MIM\cong I since at least one of clUi\operatorname{cl}U_{i} does not intersect M\partial M, and hence UiU_{i}^{\perp} separates the two endpoints of MM. It is clear that ϕI\phi_{I} is never satisfied by other compact connected manifolds.

We now suppose that MM is a compact connected manifold not homeomorphic to II. Then for all disjoint, proper, non-empty regular open sets UU and VV satisfying UV=UVU\oplus V=U\sqcup V, the set M(UV)M\setminus(U\oplus V) is disconnected. From Corollary 3.7, we obtain the formula ϕS1\phi_{S^{1}} expressing that MM and S1S^{1} are homeomorphic. ∎

By Proposition 2.8 and Corollary 3.8, we establish Theorem 1.4 for the case when MM is one–dimensional. Henceforth, we modify the definitions of \mathscr{M} and vol\mathscr{M}_{\operatorname{vol}}, replacing these classes by subclasses where all the manifolds in consideration are of dimension at least two.

3.4. Further topological properties

We will need several more general first order formulae to express basic topological properties of regular open sets. One of primary importance will be a formula which implies that a particular regular open set UU is contained in a collared ball inside of another regular open set VV. This is not particularly difficult to state and prove in the class \mathscr{M}, but is substantially harder in vol\mathscr{M}_{\operatorname{vol}}. For the rest of this section, we will make a standing assumption that (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}, and that the underlying structure is Act𝐆,RO(M,G)\operatorname{{Act}}_{\mathbf{G},\operatorname{RO}}(M,G).

3.4.1. Relative-compactness regarding good balls

We use the preceding results to find first order formulae that compare measures of regular open sets. For the remainder of this subsection, we assume that MM is a connected, compact nn–manifold with n>1n>1, equipped with an Oxtoby–Ulam measure μ\mu.

Lemma 3.9.

There exists a formula vol(u1,u2,v)\operatorname{vol}_{\leq}(u_{1},u_{2},v) in the language L𝐆,ROL_{\bf{G},\operatorname{RO}} such that for all (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}} with an Oxtoby–Ulam measure μ\mu on MM, and for all any triple (U1,U2,V)(U_{1},U_{2},V) with U1U_{1} and U2U_{2} connected and U1,U2VU_{1},U_{2}\subseteq V, we have the following:

  1. (1)

    If μ(U1)μ(U2)\mu(U_{1})\leq\mu(U_{2}) then vol(U1,U2,V)\operatorname{vol}_{\leq}(U_{1},U_{2},V) holds.

  2. (2)

    If vol(U1,U2,V)\operatorname{vol}_{\leq}(U_{1},U_{2},V) holds then μ(U1)μ((clU2)V)\mu(U_{1})\leq\mu((\operatorname{cl}U_{2})\cap V).

Proof.

Suppose first that μ(U1)μ(U2)\mu(U_{1})\leq\mu(U_{2}), and let W0RO(M)\varnothing\neq W_{0}\in\operatorname{RO}(M) be arbitrary. By Lemma 2.5, we can find a good ball BU1B\subseteq U_{1} such that

μ(U1)μ(W0)<μ(B)<μ(U1)μ(U2),\mu(U_{1})-\mu(W_{0})<\mu(B)<\mu(U_{1})\leq\mu(U_{2}),

and such that U1BU_{1}\setminus B connected. Lemma 2.4 furnishes gHomeo0,μ(M)[V]g\in\operatorname{Homeo}_{0,\mu}(M)[V] such that g(B)U2g(B)\subseteq U_{2}, but clearly there is no μ\mu-preserving hh such that

h(W0)U1B=U1(intB).h(W_{0})\subseteq U_{1}\setminus B=U_{1}\cap(\operatorname{int}B)^{\perp}.

We have just established that vol(U1,U2,V)\operatorname{vol}_{\leq}(U_{1},U_{2},V) holds with W1:=intBW_{1}:=\operatorname{int}B, where

vol(u1,u2,v)\displaystyle\operatorname{vol}_{\leq}(u_{1},u_{2},v)\equiv (w0,w1)[w1u1conn(u1w1)\displaystyle(\forall w_{0}\neq\varnothing,\exists w_{1})[w_{1}\subseteq u_{1}\wedge\operatorname{conn}(u_{1}\cap w_{1}^{\perp})\wedge
(γ𝐆)[γ(w0)u1w1](δ𝐆[v])[δ(w1)u2]].\displaystyle(\forall\gamma\in\mathbf{G})[\gamma(w_{0})\not\subseteq u_{1}\cap w_{1}^{\perp}]\wedge(\exists\delta\in\mathbf{G}[v])[\delta(w_{1})\subseteq u_{2}]].

Let us now suppose for a contradiction that vol(U1,U2,V)\operatorname{vol}_{\leq}(U_{1},U_{2},V) holds but that

μ(U1)>μ(VclU2).\mu(U_{1})>\mu(V\cap\operatorname{cl}U_{2}).

Let W0W_{0} be the interior of a good ball in MM with measure r0<μ(U1)μ(VclU2)r_{0}<\mu(U_{1})-\mu(V\cap\operatorname{cl}U_{2}). It suffices to show that there is no witness W1W_{1} as required by vol\operatorname{vol}_{\leq}.

If such a W1W_{1} exists then by the condition on γ\gamma, we see again from Lemmas 2.4 and 2.5 that μ(U1W1)r0\mu(U_{1}\cap W_{1}^{\perp})\leq r_{0}. Moreover, there is a group element gG[V]g\in G[V] such that g(W1)U2g(W_{1})\subseteq U_{2}, so that in fact

g((clW1)V)(clU2)V.g((\operatorname{cl}W_{1})\cap V)\subseteq(\operatorname{cl}U_{2})\cap V.

We then obtain

r0=μ(W0)μ(U1W1)μ(U1)μ(VclW1)μ(U1)μ(VclU2).r_{0}=\mu(W_{0})\geq\mu(U_{1}\cap W_{1}^{\perp})\geq\mu(U_{1})-\mu(V\cap\operatorname{cl}W_{1})\geq\mu(U_{1})-\mu(V\cap\operatorname{cl}U_{2}).

This violates the choice of r0r_{0}. ∎

The foregoing discussion allows us to characterize when a regular open set UU is contained in a collared ball BB inside a regular open set VV. There are separate formulae which apply in the measure-preserving case, and in the general case.

Lemma 3.10.

There exists a first order formula RCB(vol)(u,v)\operatorname{RCB}_{(\operatorname{vol})}(u,v) such that for each (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}, we have that

RCB(vol)(U,V)\models\operatorname{RCB}_{(\operatorname{vol})}(U,V)

if and only if UU is relatively compact in some good ball contained in VV.

Recall our convention that this lemma actually claims to produce two formulae, namely RCB(u,v)\operatorname{RCB}(u,v) and RCBvol(u,v)\operatorname{RCB}_{\operatorname{vol}}(u,v).

Proof of Lemma 3.10.

Let us consider the formula RCB(u,v)\operatorname{RCB}(u,v), which expresses that there exists some component v^\hat{v} of vv satisfying the following two conditions:

  • v^\hat{v} contains uu;

  • for each nonempty, regular open set ww contained in v^\hat{v}, there exists some element γ𝐆[v^]\gamma\in\mathbf{G}[\hat{v}] that moves uu into ww.

We first claim that this formula satisfies the conclusion for (M,G)(M,G)\in\mathscr{M}. Indeed, if UU is relatively compact in a collared ball BVB\subseteq V, then there exists a unique V^π0(V)\hat{V}\in\pi_{0}(V) containing BB, and hence UU. For each nonempty regular open WV^W\subseteq\hat{V}, we see from Lemma 2.4 that some gG[V^]g\in G[\hat{V}] satisfies

g(U)g(B)W,g(U)\subseteq g(B)\subseteq W,

as desired. Conversely, suppose RCB(U,V)\models\operatorname{RCB}(U,V) holds and let V^\hat{V} be the connected component of VV containing UU. Let us fix a collared ball BB in V^\hat{V} and set W:=intBW:=\operatorname{int}B. By assumption, we can find gG[V^]g\in G[\hat{V}] such that g(U)Wg(U)\subseteq W. Then UU is relatively compact in the collared ball g1(B)g^{-1}(B) in VV.

For the case when (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}}, we set

RCBvol(u,v)(u,v^)[(v^π0(v)uuv^conn(u))\displaystyle\operatorname{RCB}_{\operatorname{vol}}(u,v)\equiv(\exists u^{\prime},\hat{v})[(\hat{v}\in\pi_{0}(v)\wedge u\subseteq u^{\prime}\subsetneq\hat{v}\wedge\operatorname{conn}(u^{\prime}))\wedge
(w)[(conn(w)wv^¬vol(w,u,v^))(γ𝐆[v^])[γ(u)w]]].\displaystyle(\forall w)[(\operatorname{conn}(w)\wedge w\subseteq\hat{v}\wedge\neg\operatorname{vol}_{\leq}(w,u^{\prime},\hat{v}))\longrightarrow(\exists\gamma\in\mathbf{G}[\hat{v}])[\gamma(u^{\prime})\subseteq w]]].

In order to prove the forward direction, assume that RCBvol(U,V)\operatorname{RCB}_{\operatorname{vol}}(U,V) holds for some nonempty U,VRO(M)U,V\in\operatorname{RO}(M). Let UU^{\prime} and V^\hat{V} be witnesses for the existentially quantified variables uu^{\prime} and v^\hat{v}. Since UV^U^{\prime}\subsetneq\hat{V}, the Boolean subtraction V^(U)\hat{V}\cap(U^{\prime})^{\perp} is nonempty. We now see that

μ(V^)=μ(V^clU)+μ(V^(U))>μ(V^clU).\mu(\hat{V})=\mu(\hat{V}\cap\operatorname{cl}U^{\prime})+\mu(\hat{V}\cap(U^{\prime})^{\perp})>\mu(\hat{V}\cap\operatorname{cl}U^{\prime}).

So, Lemma 2.5 furnishes a good ball BV^B\subseteq\hat{V} satisfying

μ(B)>μ(V^clU).\mu(B)>\mu(\hat{V}\cap\operatorname{cl}U^{\prime}).

By Lemma 3.9, we have that ¬vol(intB,U,V^)\neg\operatorname{vol}_{\leq}(\operatorname{int}B,U^{\prime},\hat{V}), and that some gG[V^]g\in G[\hat{V}] satisfies that g(U)intBg(U^{\prime})\subseteq\operatorname{int}B. It follows that

clUclUg1(B)V^V,\operatorname{cl}U\subseteq\operatorname{cl}U^{\prime}\subseteq g^{-1}(B)\subseteq\hat{V}\subseteq V,

as desired.

For the backward direction, we pick a good ball BB satisfying UBV^U\subseteq B\subseteq\hat{V} for a suitable V^π0(V)\hat{V}\in\pi_{0}(V) and set U:=intBU^{\prime}:=\operatorname{int}B. Consider an arbitrary connected regular open set WV^W\subseteq\hat{V} satisfying ¬vol(W,U,V^)\neg\operatorname{vol}_{\leq}(W,U^{\prime},\hat{V}). From Lemma 3.9 again, we see that

μ(W)>μ(U)=μ(B).\mu(W)>\mu(U^{\prime})=\mu(B).

We may therefore find some gG[V^]g\in G[\hat{V}] such that g(B)Wg(B)\subseteq W. This shows that RCBvol(U,V)\operatorname{RCB}_{\operatorname{vol}}(U,V) holds. ∎

When using Lemma 3.10, we will write RCB\operatorname{RCB} both in the case of the full homeomorphism group and the measure-preserving homeomorphism group, suppressing the symbol vol\operatorname{vol} from the notation.

Many of the formulae below will actually have different meanings for \mathscr{M} and for vol\mathscr{M}_{\operatorname{vol}}, though sometimes coincide in their implications; we record the fact that RCB(vol)(U,M)\operatorname{RCB}_{(\operatorname{vol})}(U,M) implies that clUintM\operatorname{cl}U\subseteq\operatorname{int}M.

3.4.2. Detecting finiteness of components

From part (5) of Corollary 3.7, we can detect whether or not a given regular open set has exactly kk connected component in the theory of GG for each fixed kωk\in\omega. It is not obvious a priori how to express the infinitude of the connected components of URO(M)U\in\operatorname{RO}(M), as such an infinitude would be equivalent to the infinite conjunction

¬cc0(U)¬cc1(U)¬cc2(U).\neg\mathrm{cc}_{0}(U)\wedge\neg\mathrm{cc}_{1}(U)\wedge\neg\mathrm{cc}_{2}(U)\wedge\cdots.

However, one can express such an infinitude in a single formula.

Definition 3.11.

Let us set

dispersed(u)(u^π0(u))[RCB(u^,uu^)].\operatorname{dispersed}(u)\equiv(\forall\hat{u}\in\pi_{0}(u))[\operatorname{RCB}(\hat{u},u^{\perp}\oplus\hat{u})].

We say a regular open set UU is dispersed if dispersed(U)\operatorname{dispersed}(U) holds.

Note that dispersed(U)\operatorname{dispersed}(U) implies that

clU^cl(UU^)=\operatorname{cl}\hat{U}\cap\operatorname{cl}(U\setminus\hat{U})=\varnothing

for each connected component U^\hat{U} of UU. Let us introduce another formula in the lemma below that will play crucial roles in several places of this paper; the proof is straightforward and we omit it.

Lemma 3.12.

There exists an L𝐆,ROL_{\mathbf{G},\operatorname{RO}}–formula seq(u,v,γ)\operatorname{seq}(u,v,\gamma) such that

Act𝐆,RO(M,G)seq(U,V,g)\operatorname{{Act}}_{\mathbf{G},\operatorname{RO}}(M,G)\models\operatorname{seq}(U,V,g)

for U,VRO(M)U,V\in\operatorname{RO}(M) and gGg\in G if and only if the following conditions hold for a unique Uπ0(U)U^{\prime}\in\pi_{0}(U):

  1. (i)

    the set UU is dispersed;

  2. (ii)

    we have that Vg(V)UV\cup g(V)\subseteq U;

  3. (iii)

    for all U^π0(U)\hat{U}\in\pi_{0}(U), the set U^V\hat{U}\cap V is nonempty and connected;

  4. (iv)

    for all U^π0(U){U}\hat{U}\in\pi_{0}(U)\setminus\{U^{\prime}\}, the set U^g(V)\hat{U}\cap g(V) is nonempty and connected;

  5. (v)

    we have that Ug(V)=U^{\prime}\cap g(V)=\varnothing;

  6. (vi)

    if a union WW of connected components of UU satisfies that UWU^{\prime}\subseteq W and that g(VW)Wg(V\cap W)\subseteq W, then W=UW=U.

In a situation as in Lemma 3.12, we can enumerate the components of UU as

U^0=U,U^1,\hat{U}_{0}=U^{\prime},\hat{U}_{1},\ldots

so that g(VU^i)U^i+1g(V\cap\hat{U}_{i})\subseteq\hat{U}_{i+1} for each i0i\geq 0. Furthermore, we have an injection

σ=σU,V,g:π0(U)π0(U){U}\sigma=\sigma_{U,V,g}\colon\pi_{0}(U)\longrightarrow\pi_{0}(U)\setminus\{U^{\prime}\}

sending U^i\hat{U}_{i} to U^i+1\hat{U}_{i+1} for each iωi\in\omega. We also note that for each iωi\in\omega there exists a uniformly definable function seqi(u,v,g)\operatorname{seq}_{i}(u,v,g) such that

seqi(U,V,g)=U^i.\operatorname{seq}_{i}(U,V,g)=\hat{U}_{i}.

We can now establish the main result of this subsection.

Lemma 3.13.

There exists a formula infcomp(w)\operatorname{infcomp}(w) such that

infcomp(W)if and only if W has infinitely many connected components.\models\operatorname{infcomp}(W)\quad\textrm{if and only if }W\text{ has infinitely many connected components}.
Proof.

Let us define

infcomp(w)(u,v,γ)[uw(w^π0(w))[conn(uw^)]seq(u,v,γ)].\operatorname{infcomp}(w)\equiv(\exists u,v,\gamma)[\varnothing\neq u\subseteq w\wedge(\forall\hat{w}\in\pi_{0}(w))[\operatorname{conn}(u\cap\hat{w})]\wedge\operatorname{seq}(u,v,\gamma)].

In order to prove the forward direction, suppose we have seq(U,V,g)\operatorname{seq}(U,V,g) for some nonempty UWU\subseteq W, such that each connected component W^\hat{W} of WW satisfies conn(UW^)\operatorname{conn}(U\cap\hat{W}). In particular, we have |π0(U)||π0(W)||\pi_{0}(U)|\leq|\pi_{0}(W)|. The injection σU,V,g\sigma_{U,V,g} above certifies that π0(U)\pi_{0}(U) is an infinite set. Hence, π0(W)\pi_{0}(W) is infinite as well.

For the backward direction, suppose that WW has infinitely many components. We will establish infcomp(W)\models\operatorname{infcomp}(W) only in the case of (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}}, since the case (M,G)(M,G)\in\mathscr{M} is strictly easier. We use an idea similar to the proof of Proposition 2.7. We first find distinct components {W^i}iω\{\hat{W}_{i}\}_{i\in\omega} of WW such that some sequence {xi}iω\{x_{i}\}_{i\in\omega} satisfying xiW^ix_{i}\in\hat{W}_{i} converges to some point xMx^{*}\in M. We consider a sufficiently small compact chart neighborhood BB of xx^{*}, which still intersects infinitely many components of WW. Let n=dimMn=\dim M. By the Oxtoby–Ulam theorem, we can simply identify BB with Bn(1)B^{n}(1) or Bn(1)+nB^{n}(1)\cap\mathbb{H}^{n}_{+} equipped with the Lebesgue measure. The point xx^{*} is then identified with the origin OO. By shrinking each W^i\hat{W}_{i} to U^iintB\hat{U}_{i}\subseteq\operatorname{int}B and passing to a subsequence, we can further require the following for all i0i\geq 0.

  • The open set U^i\hat{U}_{i} is an open Euclidean ball, converging to x=Ox^{*}=O;

  • We have dist(x,U^i+1)+diam(U^i+1)<dist(x,U^i)\operatorname{dist}(x^{*},\hat{U}_{i+1})+\operatorname{diam}(\hat{U}_{i+1})<\operatorname{dist}(x^{*},\hat{U}_{i}).

We set

U:=iU^i=iU^iU:=\bigsqcup_{i}\hat{U}_{i}=\oplus_{i}\hat{U}_{i}

and U=U^0U^{\prime}=\hat{U}_{0}. We can find a disjoint collection of compact topological balls {Di}\{D_{i}\} such that intDi\operatorname{int}D_{i} intersects both U^i\hat{U}_{i} and U^i+1\hat{U}_{i+1}, and no other U^j\hat{U}_{j}’s. Using the path–transitivity as in Lemma 2.4, we can inductively find a

giG[intDi]g_{i}\in G[\operatorname{int}D_{i}]

sending some good ball CiU^iC_{i}\subseteq\hat{U}_{i} onto another good ball inside U^i+1\hat{U}_{i+1}. We will set

V=iωintCi.V=\bigcup_{i\in\omega}\operatorname{int}C_{i}.

By the uniform convergence theorem, the sequence

{i=0kgi}k0\left\{\prod_{i=0}^{k}g_{i}\right\}_{k\geq 0}

converges to a homeomorphism gHomeo0,μ(M)Gg\in\operatorname{Homeo}_{0,\mu}(M)\leq G, which witnesses the properties that the formula infcomp(U)\operatorname{infcomp}(U) requires. ∎

It follows immediately that we may also test whether a regular open set has finitely many components, and write

fincomp(u)¬infcomp(u).\operatorname{fincomp}(u)\equiv\neg\operatorname{infcomp}(u).

3.4.3. Touching and containing the boundary

By a collar (embedding) of the boundary in a manifold MM, we mean an embedding

h:M×[0,1)Mh\colon\partial M\times[0,1)\longrightarrow M

that extends the identity map

M×{0}M;\partial M\times\{0\}\longrightarrow\partial M;

we sometimes allow hh to be an embedding of M×[0,1]\partial M\times[0,1]. The image of a collar embedding is called a collar neighborhood. A fundamental result due to Brown [6, Theorem 2] says that the boundary of a topological manifold admits a collar. We now produce several formulae regarding the boundary of a given manifold.

Lemma 3.14.

There exist L𝐆,ROL_{\mathbf{G},\operatorname{RO}}–formulae as follows:

  1. (1)

    A formula touch(u)\operatorname{touch}_{\partial}(u) such that

    touch(U)if and only if the closure of U nontrivially intersects M.\models\operatorname{touch}_{\partial}(U)\quad\textrm{if and only if the closure of }U\text{ nontrivially intersects }\partial M.
  2. (2)

    A formula stab(γ)\operatorname{stab}_{\partial}(\gamma) such that

    stab(g)if and only if g setwise stabilizes each boundary component of M.\models\operatorname{stab}_{\partial}(g)\quad\textrm{if and only if }g\text{ setwise stabilizes each boundary component of }M.
Proof.

(1) Let us define the formula

finint(u,w)(u)[fincomp(u)(u^π0(u))[u^wu^π0(u)]].\operatorname{finint}(u,w)\equiv(\exists u^{\prime})[\operatorname{fincomp}(u^{\prime})\wedge(\forall\hat{u}\in\pi_{0}(u))[\hat{u}\cap w\neq\varnothing\rightarrow\hat{u}\in\pi_{0}(u^{\prime})]].

It is clear from the formulation that

finint(U,W)\models\operatorname{finint}(U,W)

if and only if UU meets WW in finitely many components on UU. We now set

touch(u)\displaystyle\operatorname{touch}_{\partial}(u)\equiv (u)[uuinfcomp(u)(w)[RCB(w,M)finint(u,w)]].\displaystyle(\exists u^{\prime})[u^{\prime}\subseteq u\wedge\operatorname{infcomp}(u^{\prime})\wedge(\forall w)[\operatorname{RCB}(w,M)\longrightarrow\operatorname{finint}(u^{\prime},w)]].

Suppose that clUM\operatorname{cl}U\cap\partial M\neq\varnothing. Choose a sequence of points {pi}iω\{p_{i}\}_{i\in\omega} in UU converging to a point in M\partial M, and choose small open balls UipiU_{i}\ni p_{i} in UU with pairwise disjoint closures and with radii tending to zero. Let UU^{\prime} be the union of these balls. Now, if WW fails to satisfy finint(U,W)\operatorname{finint}(U^{\prime},W), then WW must meet infinitely many of the balls UiU_{i}; thus clWM\operatorname{cl}W\cap\partial M\neq\varnothing. In particular, ¬RCB(W,M)\neg\operatorname{RCB}(W,M).

Conversely, suppose that clUM=\operatorname{cl}U\cap\partial M=\varnothing, and let UUU^{\prime}\subseteq U have infinitely many components {Ui}iω\{U_{i}\}_{i\in\omega}. As in Lemma 3.13, by shrinking components of UU^{\prime} and passing to a subsequence, we may assume that each UiU_{i} is an open ball, that the sequence has shrinking radii, and converges monotonically to the origin in an open chart in n\mathbb{R}^{n}. Moreover, the origin in this chart lies in the interior of MM, by assumption.

We may take WW to be a neighborhood of the origin in this chart, which then satisfies RCB(W,M)\operatorname{RCB}(W,M) and meets infinitely many components of UU^{\prime}. Thus, UU^{\prime} fails to witness touch(U)\operatorname{touch}_{\partial}(U), and so touch(U)\operatorname{touch}_{\partial}(U) does not hold.

(2) Setting

contain(u)¬touch(u),\operatorname{contain}_{\partial}(u)\equiv\neg\operatorname{touch}_{\partial}(u^{\perp}),

we see that contain(U)\operatorname{contain}_{\partial}(U) holds if and only if MU\partial M\subseteq U. We now define

stab(γ)(u,u^)[(u^π0(u)contain(u)touch(u^))u^γ(u^)].\operatorname{stab}_{\partial}(\gamma)\equiv(\forall u,\hat{u})[(\hat{u}\in\pi_{0}(u)\wedge\operatorname{contain}_{\partial}(u)\wedge\operatorname{touch}_{\partial}(\hat{u}))\rightarrow\hat{u}\cap\gamma(\hat{u})\neq\varnothing].

We claim that stab(g)\operatorname{stab}_{\partial}(g) holds for gGg\in G if and only if gg setwise stabilizes each component of

M=12k.\partial M=\partial_{1}\sqcup\partial_{2}\sqcup\cdots\sqcup\partial_{k}.

For the forward direction, suppose we have stab(g)\operatorname{\operatorname{stab}}_{\partial}(g). By the aforementioned result of Brown, we can pick a closure–disjoint collection of collar neighborhoods {V1,V2,,Vk}\{V_{1},V_{2},\ldots,V_{k}\} of the components of M\partial M. Defining

U:=i=1kVi,U:=\bigsqcup_{i=1}^{k}V_{i},

we see from the hypothesis that g(Vi)Vig(V_{i})\cap V_{i}\neq\varnothing for each ii, which trivially implies g(i)=ig(\partial_{i})=\partial_{i}. The backward direction is clear after observing that the hypothesis of stab(g)\operatorname{stab}_{\partial}(g) simply says that U^\hat{U} contains at least one boundary component.∎

4. Interpretation of second-order arithmetic

The goal of this section is to prove that the group GG interprets second order arithmetic and analysis uniformly for (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}, establishing the case of i=1i=1 in Proposition 3.3.

4.1. An example of an interpretation of first order arithmetic

As a warm-up, let us interpret first order arithmetic

Arith1=(,+,×,0,1)\operatorname{Arith}_{1}=(\mathbb{N},+,\times,0,1)

in the structure Act𝐆,RO(M,G)\operatorname{{Act}}_{\mathbf{G},\operatorname{RO}}(M,G). For this, we consider the surjection

#π0:{URO(M)fincomp(U)}\#\pi_{0}\colon\{U\in\operatorname{RO}(M)\mid\operatorname{fincomp}(U)\}\longrightarrow\mathbb{N}

sending each UU to #π0(U)\#\pi_{0}(U), namely the cardinality of π0(U)\pi_{0}(U). The domain of this surjection is clearly definable, and so is the fiber by the formula #=(u,v)\#_{=}(u,v) in Corollary 3.7. To complete an interpretation of Arith1\operatorname{Arith}_{1}, it suffices to establish the following:

Lemma 4.1.

There exist L𝐆,ROL_{\mathbf{G},\operatorname{RO}}–formulae #+\#_{+} and #×\#_{\times} such that the following hold for all U,V,WU,V,W having finitely many connected components.

  1. (1)

    We have #+(U,V,W)if and only if#π0(W)=#π0(U)+#π0(V)\models\#_{+}(U,V,W)\quad\textrm{if and only if}\quad\#\pi_{0}(W)=\#\pi_{0}(U)+\#\pi_{0}(V).

  2. (2)

    We have #×(U,V,W)if and only if#π0(W)=#π0(U)#π0(V)\models\#_{\times}(U,V,W)\quad\textrm{if and only if}\quad\#\pi_{0}(W)=\#\pi_{0}(U)\cdot\#\pi_{0}(V).

Proof.

Recall the meaning of the formula ccpartition\operatorname{ccpartition} from Corollary 3.7. Let us make the following definitions.

#+(u,v,w)\displaystyle\#_{+}(u,v,w) (w0,w1)[ccpartition(w0,w1,w)#=(w0,u)#=(w1,v)],\displaystyle\equiv(\exists w_{0},w_{1})[\operatorname{ccpartition}(w_{0},w_{1},w)\wedge\#_{=}(w_{0},u)\wedge\#_{=}(w_{1},v)],
#×(u,v,w)\displaystyle\#_{\times}(u,v,w) (w)[(wu)#=(w,w)(u^π0(u))[#=(u^w,v)]].\displaystyle\equiv(\exists w^{\prime})[(w^{\prime}\subseteq u)\wedge\#_{=}(w,w^{\prime})\wedge(\forall\hat{u}\in\pi_{0}(u))[\#_{=}(\hat{u}\cap w^{\prime},v)]].

It is straightforward to check that these formulae have the intended meanings.∎

4.2. Our interpretation of second order arithmetic

We now describe an interpretation of second order arithmetic

Arith2=(,𝒫(),0,1,+,×,),\operatorname{Arith}_{2}=(\mathbb{N},\mathscr{P}(\mathbb{N}),0,1,+,\times,\in),

which has two sorts, namely \mathbb{N} and 𝒫()\mathscr{P}(\mathbb{N}). In particular, we will have to be able to quantify over subsets of \mathbb{N}.

In order to achieve this, we will consider more restricted class of regular open sets UU, the components of which admit a linear order as described by the formula seq(U,V,g)\operatorname{seq}(U,V,g); see Section 3.4.2. In this linear order of UU, the kk–th component U^k\hat{U}_{k} will interpret the integer kωk\in\omega, and a union of the connected components WW will interpret a subset in a natural way. We will utilize Lemma 4.1, but not the actual interpretation itself from the previous subsection.

To be more concrete, let us first note the following.

Lemma 4.2.

There exists a uniformly defined function seq(u,v,γ,u^)\operatorname{seq}_{\restriction}(u,v,\gamma,\hat{u}) such that if

seq(U,V,g)U^π0(U),\models\operatorname{seq}(U,V,g)\wedge\hat{U}\in\pi_{0}(U),

then for the unique kωk\in\omega satisfying U^=seqk(U,V,g)\hat{U}=\operatorname{seq}_{k}(U,V,g), we have that

seq(U,V,g,U^)=0ikseqi(U,V,g).\operatorname{seq}_{\restriction}(U,V,g,\hat{U})=\oplus_{0\leq i\leq k}\operatorname{seq}_{i}(U,V,g).
Proof.

It is routine to check that the following has the intended meaning:

(w=seq(u,v,γ,u^))\displaystyle(w=\operatorname{seq}_{\restriction}(u,v,\gamma,\hat{u}))\equiv ucc(w,u)(seq0(u,v,γ)u^)w\displaystyle\operatorname{ucc}(w,u)\wedge(\operatorname{seq}_{0}(u,v,\gamma)\oplus\hat{u})\subseteq w\wedge
γ(vu^)w=γ1(wγ(v))w.\displaystyle\gamma(v\cap\hat{u})\cap w=\varnothing\wedge\gamma^{-1}(w\cap\gamma(v))\subseteq w.\qed

Let us consider the set

X1:={(U,V,g,U^)seq(U,V,g)U^π0(U)]},X_{1}:=\{(U,V,g,\hat{U})\mid\operatorname{seq}(U,V,g)\wedge\hat{U}\in\pi_{0}(U)]\},

which is definable in Act𝐆,RO(M,G)\operatorname{{Act}}_{\mathbf{G},\operatorname{RO}}(M,G) uniformly for (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})}. We have a surjection

ρ1:=#π0seq:X1.\rho_{1}:=\#\pi_{0}\circ\operatorname{seq}_{\restriction}\colon X_{1}\longrightarrow\mathbb{N}.

This surjection satisfies

k=ρ1(U,V,g,U^)if and only ifU^=seqk(U,V,g).k=\rho_{1}(U,V,g,\hat{U})\quad\text{if and only if}\quad\hat{U}=\operatorname{seq}_{k}(U,V,g).

The fiber of ρ1\rho_{1} is

{(y¯,z¯)X1×X1#=(seq(y¯),seq(z¯))},\{(\underline{y},\underline{z})\in X_{1}\times X_{1}\mid\#_{=}(\operatorname{seq}_{\restriction}(\underline{y}),\operatorname{seq}_{\restriction}(\underline{z}))\},

and hence uniformly definable. It is trivial to check that ρ1\rho_{1} produces a uniform interpretation of Arith1\operatorname{Arith}_{1} to Act𝐆,RO(M,G)\operatorname{{Act}}_{\mathbf{G},\operatorname{RO}}(M,G). For instance, we have

ρ1(y¯)+ρ1(y¯)=ρ1(y¯′′)\rho_{1}(\underline{y})+\rho_{1}(\underline{y}^{\prime})=\rho_{1}(\underline{y}^{\prime\prime})

if and only if

#+(seq(y¯),seq(y¯),seq(y¯′′)).\models\#_{+}(\operatorname{seq}_{\restriction}(\underline{y}),\operatorname{seq}_{\restriction}(\underline{y}^{\prime}),\operatorname{seq}_{\restriction}(\underline{y}^{\prime\prime})).

After this interpretation of Arith1\operatorname{Arith}_{1}, the symbol #\# has an intended meaning as a function from RO(M)\operatorname{RO}(M) to \mathbb{N}. We have uniformly defined functions seqcomp(u,v,γ,α)\operatorname{seqcomp}(u,v,\gamma,\alpha) and seqcomp(u,v,γ,α)\operatorname{seqcomp}_{\restriction}(u,v,\gamma,\alpha) satisfying

seqcomp(U,V,g,k)\displaystyle\operatorname{seqcomp}(U,V,g,k) =seqk(U,V,g),\displaystyle=\operatorname{seq}_{k}(U,V,g),
seqcomp(U,V,g,k)\displaystyle\operatorname{seqcomp}_{\restriction}(U,V,g,k) =i=1kseqcomp(U,V,g,i).\displaystyle=\oplus_{i=1}^{k}\operatorname{seqcomp}(U,V,g,i).

Similarly, we consider another uniformly definable set

X1:={(U,V,g,W)seq(U,V,g)ucc(W,U)]}.X_{1}^{\prime}:=\{(U,V,g,W)\mid\operatorname{seq}(U,V,g)\wedge\operatorname{ucc}(W,U)]\}.

We have a surjection

ρ1:X1𝒫()\rho_{1}^{\prime}\colon X_{1}^{\prime}\longrightarrow\mathscr{P}(\mathbb{N})

defined by the condition

ρ1(U,V,g,W):={ρ1(U,V,g,W^)W^π0(W)}.\rho_{1}^{\prime}(U,V,g,W):=\{\rho_{1}(U,V,g,\hat{W})\mid\hat{W}\in\pi_{0}(W)\}.

Since the fiber of ρ1\rho_{1} is uniformly definable, so is that of ρ1\rho_{1}^{\prime}. We will introduce the function symbol 𝒫#\mathscr{P}_{\#} in LAct2L^{2}_{\operatorname{{Act}}} interpreted as ρ1\rho_{1}^{\prime}.

Finally, we have

ρ1(U1,V1,g1,U^)ρ1(U2,V2,g2,W2)\rho_{1}(U_{1},V_{1},g_{1},\hat{U})\in\rho_{1}^{\prime}(U_{2},V_{2},g_{2},W_{2})

if and only if

#π0seq(U1,V1,g1,U^)=#π0seq(U2,V2,g2,W^)\#\pi_{0}\operatorname{seq}_{\restriction}(U_{1},V_{1},g_{1},\hat{U})=\#\pi_{0}\operatorname{seq}_{\restriction}(U_{2},V_{2},g_{2},\hat{W})

for some W^π0(W2)\hat{W}\in\pi_{0}(W_{2}). Hence, the pair of surjections (ρ1,ρ1)(\rho_{1},\rho_{1}^{\prime}) produces the desired interpretation of the two–sorted structure Arith2\operatorname{Arith}_{2}. We note that the order relation symbol <<, the successor symbol SS, and the inclusion symbol \subseteq are naturally interpreted as a consequence.

4.3. Analysis

The interpretation of \mathbb{R} is now standard. From \mathbb{N}, we interpret \mathbb{Z}, together with addition, multiplication, and order, by imposing a suitable definable equivalence relation on a suitable definable subset of 2\mathbb{N}^{2}. We similarly interpret \mathbb{Q} by imposing a suitable definable equivalence relation on a definable subset of ×\mathbb{Z}\times\mathbb{Z}.

We define \mathbb{R} together with addition, multiplication, and order via Dedekind cuts of \mathbb{Q}; all this is interpretable because of our access to 𝒫()\mathscr{P}(\mathbb{N}). Finally, we have canonical identifications of

,\mathbb{N}\subseteq\mathbb{Z}\subseteq\mathbb{Q}\subseteq\mathbb{R},

wherein we set == to be the relation identifying natural numbers with their images under this sequence of inclusions. In the sequel, we will simply talk about natural numbers, integers, or rationals as elements of \mathbb{R} without further comment. We further may assume to have k\mathbb{R}^{k} in the universe of the structure for all kωk\in\omega.

In order to justify the introduction of the sort symbol Contk,\operatorname{Cont}_{k,\ell} in the structure, let us first note that each function in C(,)C(\mathbb{R},\mathbb{R}) is uniquely determined by its restriction on \mathbb{Q}. Since

||=(2ω)ω=2ω=||,|\mathbb{R}^{\mathbb{Q}}|=(2^{\omega})^{\omega}=2^{\omega}=|\mathbb{R}|,

we have an interpretation of C(,)C(\mathbb{R},\mathbb{R}) by \mathbb{R}, and hence, that of

C(k,).C(\mathbb{R}^{k},\mathbb{R}^{\ell}).

This latter set is the domain of Contk,\operatorname{Cont}_{k,\ell}, and the function symbols

appl(χ,ρ)=σ,norm(χ)=ρ\operatorname{appl}(\chi,\rho)=\sigma,\quad\operatorname{norm}(\chi)=\rho

are interpreted accordingly. In practice, we write

f(r)=s,f=rf(r)=s,\quad\|f\|=r

for the above formulae. The expanded language containing Act𝐆,RO\operatorname{{Act}}_{\mathbf{G},\operatorname{RO}} structure, second order arithmetic, and analysis will be written Act2=Act𝐆,RO,\operatorname{{Act}}^{2}=\operatorname{{Act}}_{\mathbf{G},\operatorname{RO},\mathbb{R}}. This establishes the uniform interpretability of Act𝐆,RO,(M,G)\operatorname{{Act}}_{\mathbf{G},\operatorname{RO},\mathbb{R}}(M,G) to Act𝐆,RO(M,G)\operatorname{{Act}}_{\mathbf{G},\operatorname{RO}}(M,G), namely Proposition 3.3 for the case i=1i=1.

5. Interpretation of points

We now wish to be able to talk about points of MM more directly, and prove Proposition 3.3 for the case i=2i=2. This will complete the proof of Theorem 1.8.

Rubin [41] accesses points in a space with a locally dense action via a certain collection of ultrafilters consisting of regular open sets; in his approach, the intersection of the closures of all the open sets in each ultrafilter corresponds to a single point of the space. We cannot follow this approach directly, as we need to stay within the first order theory of groups and Boolean algebras. Instead, we consider a certain collection of regular open sets such that the components in each of those open sets converge to a single point of the manifold. We continue to make the standing assumption that (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} with dimM>1\dim M>1, unless stated otherwise.

5.1. Encoding points of a manifold

Using the L𝐆,RO,L_{\mathbf{G},\operatorname{RO},\mathbb{R}}–formulae introduced in the preceding sections, we define the following new formulae:

cof(w,u)\displaystyle\operatorname{cof}(w,u)\equiv infcomp(w)wu(u^π0(u))[conn(wu^)],\displaystyle\operatorname{infcomp}(w)\wedge w\subseteq u\wedge(\forall\hat{u}\in\pi_{0}(u))[\operatorname{conn}(w\cap\hat{u})],
cofcontain(w,u)\displaystyle\operatorname{cofcontain}(w,u)\equiv (w)[wwcof(w,u)],\displaystyle(\exists w^{\prime})[w^{\prime}\subseteq w\wedge\operatorname{cof}(w^{\prime},u)],
cofmove(γ,u0,u1)\displaystyle\operatorname{cofmove}(\gamma,u_{0},u_{1})\equiv (w)[cof(w,u0)cof(γ(w),u1)(w^π0(w))[RCB(w^,u0)]]\displaystyle(\exists w)[\operatorname{cof}(w,u_{0})\wedge\operatorname{cof}(\gamma(w),u_{1})\wedge(\forall\hat{w}\in\pi_{0}(w))[\operatorname{RCB}(\hat{w},u_{0})]]

Note that when cofmove(g,U0,U1)\models\operatorname{cofmove}(g,U_{0},U_{1}), we can find some WW whose connected components can be written as

W=iωWi,W=\bigsqcup_{i\in\omega}W_{i},

with the property that each WiW_{i} is contained in some relatively compact ball inside U0U_{0}; moreover, no two components of WW belong to the same component of U0U_{0}, and similarly for g(W)g(W) and U1U_{1}.

We consider the definable set

StGlim(U):={G[W]WRO(M) and cof(W,U)},\operatorname{St}^{\lim}_{G}(U):=\bigcup\{G[W^{\perp}]\mid W\in\operatorname{RO}(M)\text{ and }\operatorname{cof}(W,U)\},

which we call as the limit stabilizer of UU in GG. Intuitively, each element of this set fixes some open set that is arbitrarily close to a certain limit point of the components of UU. We will write γStlim(u)\gamma\in\operatorname{St}^{\lim}(u) for the formula corresponding to gStGlim(U)g\in\operatorname{St}^{\lim}_{G}(U).

Remark 5.1.

One can rephrase Rubin’s interpretation of points in second order logic [41] as follows, as summarized in [24, Theorem 3.6.17]. Rubin allowed certain collections (called, good ultrafilters) of regular open sets to interpret a single point in the space, by taking the intersection of the closures of those open sets. He then proved that two good ultrafilters PP and QQ interpret different points pp and qq if and only if the group

G{Q}:={G[W]WQ}G\{Q^{\perp}\}:=\bigcup\{G[W^{\perp}]\mid W\in Q\}

acts sufficiently transitively, in the sense that for some UPU\in P, every VRO(M)V\in\operatorname{RO}(M) satisfying VU\varnothing\neq V\subseteq U is an element of the set

G{Q}(P).G\{Q^{\perp}\}(P).

In our approach, we will utilize the sufficient transitivity of the limit stabilizer characterized in terms of the formula cofmove(γ,w0,w1)\operatorname{cofmove}(\gamma,w_{0},w_{1}).

Consider the set X2:=Act2(M,G)(ϕdom2)X_{2}:=\operatorname{{Act}}^{2}(M,G)(\phi_{\operatorname{dom}}^{2}), defined by the following formula:

ϕdom2(u,v,γ)\displaystyle\phi_{\operatorname{dom}}^{2}(u,v,\gamma)\equiv seq(u,v,γ)(w0,w1)[cof(w0,u)cof(w1,u)\displaystyle\operatorname{seq}(u,v,\gamma)\wedge(\forall w_{0},w_{1})[\operatorname{cof}(w_{0},u)\wedge\operatorname{cof}(w_{1},u)\longrightarrow
(δStlim(w0))[cofmove(δ,w0,w1)]].\displaystyle(\exists\delta\in\operatorname{St}^{\lim}(w_{0}))[\operatorname{cofmove}(\delta,w_{0},w_{1})]].

The following lemma furnishes an interpretation of the points.

Lemma 5.2.

For each (U,V,g)X2(U,V,g)\in X_{2} and for an arbitrary sequence {xi}iω\{x_{i}\}_{i\in\omega} satisfying

xiseqcomp(U,V,g,i)x_{i}\in\operatorname{seqcomp}(U,V,g,i)

for all iωi\in\omega, the limit

ρ2(U,V,g):=limixi\rho_{2}(U,V,g):=\lim_{i\to\infty}x_{i}

exists in MM, and is independent of the choice of {xi}iω\{x_{i}\}_{i\in\omega}. Moreover, the following conclusions hold:

  1. (1)

    The map ρ2:X2M\rho_{2}\colon X_{2}\longrightarrow M is surjective.

  2. (2)

    We have

    ρ2(U0,V0,g0)=ρ2(U1,V1,g1)\rho_{2}(U_{0},V_{0},g_{0})=\rho_{2}(U_{1},V_{1},g_{1})

    if and only if some gStGlim(U0)g\in\operatorname{St}^{\lim}_{G}(U_{0}) satisfies

    cofmove(g,U0,U1).\operatorname{cofmove}(g,U_{0},U_{1}).
  3. (3)

    We have

    h(ρ2(U,V,g))=ρ2(U,V,g)h(\rho_{2}(U,V,g))=\rho_{2}(U^{\prime},V^{\prime},g^{\prime})

    if and only if

    ρ2(h(U),h(V),hgh1)=ρ2(U,V,g).\rho_{2}(h(U),h(V),hgh^{-1})=\rho_{2}(U^{\prime},V^{\prime},g^{\prime}).
  4. (4)

    We have ρ2(U,V,g)W\rho_{2}(U,V,g)\not\in W if and only if some (U,V,g)X2(U^{\prime},V^{\prime},g^{\prime})\in X_{2} satisfies

    UW=(ρ2(U,V,g)=ρ2(U,V,g)).U^{\prime}\cap W=\varnothing\wedge\left(\rho_{2}(U,V,g)=\rho_{2}(U^{\prime},V^{\prime},g^{\prime})\right).
  5. (5)

    We have ρ2(U,V,g)intM\rho_{2}(U,V,g)\in\operatorname{int}M if and only if there exists some WW such that RCB(W,M)\operatorname{RCB}(W,M) and such that cofcontain(W,U)\operatorname{cofcontain}(W,U).

Proof.

Let (U,V,g)X2(U,V,g)\in X_{2}, and let

{xiseqcomp(U,V,g,i)}iω\{x_{i}\in\operatorname{seqcomp}(U,V,g,i)\}_{i\in\omega}

be a sequence. In particular, we have xiintMx_{i}\in\operatorname{int}M. Suppose two subsequences

{y0,j}jω,{y1,j}jω{xi}iω\{y_{0,j}\}_{j\in\omega},\{y_{1,j}\}_{j\in\omega}\subseteq\{x_{i}\}_{i\in\omega}

converge to two distinct points y0y_{0} and y1y_{1}. For i=0i=0 and i=1i=1, we let WiW_{i} be the union of sufficiently small good open balls Wi,jW_{i,j} centered at yi,jy_{i,j}. In particular, we may assume that cof(Wi,U)\operatorname{cof}(W_{i},U), and that

limjWi,j={yi}\lim_{j}W_{i,j}=\{y_{i}\}

in the Hausdorff sense. By hypothesis, we have some hStlim(W0)h\in\operatorname{St}^{\lim}(W_{0}) such that

cofmove(h,W0,W1).\models\operatorname{cofmove}(h,W_{0},W_{1}).

Since hh fixes points arbitrarily close to y0y_{0}, we have h(y0)=y0h(y_{0})=y_{0}. It follows that

y0=h(y0)=limh(y0,j)=y1.y_{0}=h(y_{0})=\lim h(y_{0,j})=y_{1}.

This proves the existence of the claimed limit. The same argument also implies the independence of the limit from the choice of {xi}iω\{x_{i}\}_{i\in\omega}, and also the backward direction of part (2). The surjectivity of ρ2\rho_{2} in part (1) is clear, after choosing UU to be a suitable sequence of good open balls converging to a given point in the Hausdorff sense.

We now verify the forward direction of part (2). By hypothesis, we can find two sequences {x0,j}iω\{x_{0,j}\}_{i\in\omega} and {x1,j}iω\{x_{1,j}\}_{i\in\omega} such that

xi,jseqcomp(Ui,Vi,gi,j).x_{i,j}\in\operatorname{seqcomp}(U_{i},V_{i},g_{i},j).

As in the proof of Lemma 3.13, we can find a disjoint collection of good balls {Di}\{D_{i}\} of decreasing sizes such that each DiD_{i} contains x0,jx_{0,j} and x1,jx_{1,j}, after passing to a subsequence if necessary. By the uniform convergence theorem, we have some hGh\in G such that h(x0,j)=x1,jh(x_{0,j})=x_{1,j} for all jj, and such that hh pointwise fixes some nonempty open set inside

seqcomp(U0,V0,g0,j)Dj.\operatorname{seqcomp}(U_{0},V_{0},g_{0},j)\cap D_{j}^{\perp}.

In particular, we have that hStGlim(U0)h\in\operatorname{St}^{\lim}_{G}(U_{0}) and that cofmove(h,U0,U1)\operatorname{cofmove}(h,U_{0},U_{1}), as claimed. The remaining parts of the lemma are straightforward to check. ∎

In part (2) of the lemma, we see that the relation

ρ2(U,V,g)=ρ2(U,V,g)\rho_{2}(U,V,g)=\rho_{2}(U^{\prime},V^{\prime},g^{\prime})

is first order expressible; hence, we deduce that the functional relation g(p)=qg(p)=q and the membership relation pWp\in W in parts (3) and (4) are interpretable for p,qMp,q\in M, gMg\in M and WRO(M)W\in\operatorname{RO}(M). Part (5) of the lemma separates out the interior points.

Direct access to points allows us to make direct reference to set theoretic operations. For instance, we can define union(u,v,w)\mathrm{union}(u,v,w) by

union(u,v,w)(π)[(πuπv)πw].\mathrm{union}(u,v,w)\equiv(\forall\pi)[(\pi\in u\vee\pi\in v)\leftrightarrow\pi\in w].

Clearly, union(U,V,W)\mathrm{union}(U,V,W) for regular open sets {U,V,W}\{U,V,W\} if and only if W=UVW=U\cup V. Henceforth, we will include the usual set-theoretic union symbol in the language such as ,\cup,\cap and \setminus. We are also able now to talk directly about the closure clU\operatorname{cl}U of a regular open set UU, both in MM and in VV for arbitrary UVU\subseteq V; for this, it suffices to note that pclUp\in\operatorname{cl}U if and only if pUp\not\in U^{\perp}.

5.2. Encoding discrete sets of points in a manifold

We now interpret the set

𝒫disc(intM):={AintMA is discrete}.\mathscr{P}^{\operatorname{disc}}(\operatorname{int}M):=\{A\subseteq\operatorname{int}M\mid A\text{ is discrete}\}.

In particular, every finite subset of intM\operatorname{int}M belongs to 𝒫disc(intM)\mathscr{P}^{\operatorname{disc}}(\operatorname{int}M).

We recall from Lemma 3.14 the formula finint(u,w)\operatorname{finint}(u,w). We first let X2X_{2}^{\prime} be the set of quadruples (U,V,g,W)(U,V,g,W) defined by the following formula:

ψdom2(u,v,γ,w)\displaystyle\psi_{\operatorname{dom}}^{2}(u,v,\gamma,w)\equiv dispersed(w)(usuppeγw)\displaystyle\operatorname{dispersed}(w)\wedge(u\oplus\operatorname{{supp}^{\mathrm{e}}}\gamma\subseteq w)\wedge
w^π0(w)[ϕdom2(uw^,vw^,γ)RCB((usuppeγ)w^,w^)].\displaystyle\forall\hat{w}\in\pi_{0}(w)[\phi^{2}_{\operatorname{dom}}(u\cap\hat{w},v\cap\hat{w},\gamma)\wedge\operatorname{RCB}((u\oplus\operatorname{{supp}^{\mathrm{e}}}\gamma)\cap\hat{w},\hat{w})].

For such a quadruple, we set

ρ2(U,V,g,W):={ρ2(UW^,VW^,g)W^π0(W)}.\rho_{2}^{\prime}(U,V,g,W):=\{\rho_{2}(U\cap\hat{W},V\cap\hat{W},g)\mid\hat{W}\in\pi_{0}(W)\}.

It is routine to check that this map defines a surjection

ρ2:X2𝒫disc(intM)\rho_{2}^{\prime}\colon X_{2}^{\prime}\longrightarrow\mathscr{P}^{\operatorname{disc}}(\operatorname{int}M)

with a definable fiber. Namely, we have

ρ2(U0,V0,g0,W0)ρ2(U1,V1,g1,W1)\rho_{2}^{\prime}(U_{0},V_{0},g_{0},W_{0})\neq\rho_{2}^{\prime}(U_{1},V_{1},g_{1},W_{1})

if and only if there exists some regular open sets W,W′′W^{\prime},W^{\prime\prime} satisfying that

RCB(W,W′′)\operatorname{RCB}(W^{\prime},W^{\prime\prime})

and that

¬finint(Ui,W)finint(U1i,W′′)\neg\operatorname{finint}(U_{i},W^{\prime})\wedge\operatorname{finint}(U_{1-i},W^{\prime\prime})

for some i{0,1}i\in\{0,1\}.

We interpret the membership between a point and a set; namely, we have

ρ2(U,V,g)ρ2(U,V,g,W)\rho_{2}(U,V,g)\in\rho_{2}^{\prime}(U^{\prime},V^{\prime},g^{\prime},W^{\prime})

if and only if there exists some W′′W^{\prime\prime} satisfying RCB(W′′,W)\operatorname{RCB}(W^{\prime\prime},W^{\prime}) and

cofcontain(W′′,U).\operatorname{cofcontain}(W^{\prime\prime},U).

We also interpret the group action

h(ρ2(U,V,g,W))=ρ2(U,V,g,W))h(\rho_{2}^{\prime}(U,V,g,W))=\rho_{2}^{\prime}(U^{\prime},V^{\prime},g^{\prime},W^{\prime}))

as

ρ2(h(U),h(V),hgh1,h(W)))=ρ2(U,V,g,W))\rho_{2}^{\prime}(h(U),h(V),hgh^{-1},h(W)))=\rho_{2}^{\prime}(U^{\prime},V^{\prime},g^{\prime},W^{\prime}))

Finally, the set

ρ2(U,V,g,W)𝒫disc(intM)\rho_{2}^{\prime}(U,V,g,W)\in\mathscr{P}^{\operatorname{disc}}(\operatorname{int}M)

has finite cardinality if and only if WW has finitely many connected components. In this case, the cardinality function #\# for T𝒫disc(intM)T\in\mathscr{P}^{\operatorname{disc}}(\operatorname{int}M) is clearly definable by

#(ρ2(U,V,g,W))=#π0(W).\#(\rho_{2}^{\prime}(U,V,g,W))=\#\pi_{0}(W).

We omit the details, which are very similar to those in Section 5.1. We denote by 𝐌discint\mathbf{M}^{\operatorname{disc-int}} the sort symbol for sets belong to 𝒫disc(intM)\mathscr{P}^{\operatorname{disc}}(\operatorname{int}M).

5.3. Interpreting exponentiation

We now interpret the map

G××MM,(g,k,p)gkp,G\times\mathbb{Z}\times M\longrightarrow M,\quad(g,k,p)\mapsto g^{k}\cdot p,

so that the exponentiation map

exp:G×G\exp:G\times\mathbb{Z}\longrightarrow G

is definable. Note that gk(p)=pg^{k}(p)=p^{\prime} holds with kωk\in\omega if and only if we can write k=mq+rk=mq+r for some integers 0r<m0\leq r<m and qq such that we have a period–mm orbit

p,g(p),,gm(p)=p,p,g(p),\ldots,g^{m}(p)=p,

and a sequence of distinct points

p,g(p),,gr(p)=p.p,g(p),\ldots,g^{r}(p)=p^{\prime}.

Let us now define formulae expcyc\operatorname{exp}_{\operatorname{cyc}} and explin\operatorname{exp}_{\operatorname{lin}}, which will express the existences of a periodic orbit and of a sequence without repetitions, respectively. More precisely, we set

expcyc(γ,α,π)\displaystyle\exp_{\operatorname{cyc}}(\gamma,\alpha,\pi)\equiv (α=0)(τ𝐌discint)[#τ=α(πτ=γ(τ))\displaystyle(\alpha=0)\vee(\exists\tau\in\mathbf{M}^{\operatorname{disc-int}})[\#\tau=\alpha\wedge(\pi\in\tau=\gamma(\tau))
¬(τ𝐌discint)[ττγτ=τ]],\displaystyle\wedge\neg(\exists\tau^{\prime}\in\mathbf{M}^{\operatorname{disc-int}})[\varnothing\neq\tau^{\prime}\subsetneq\tau\wedge\gamma\cdot\tau^{\prime}=\tau^{\prime}]],
explin(γ,α,π,π)\displaystyle\exp_{\operatorname{lin}}(\gamma,\alpha,\pi,\pi^{\prime})\equiv (τ𝐌discint)[#τ=α+1{π,π}τγ(τ{π})=τ{π}\displaystyle(\exists\tau\in\mathbf{M}^{\operatorname{disc-int}})[\#\tau=\alpha+1\wedge\{\pi,\pi^{\prime}\}\subseteq\tau\wedge\gamma(\tau\setminus\{\pi^{\prime}\})=\tau\setminus\{\pi\}
¬(τ𝐌discint)[ττγτ=τ]].\displaystyle\wedge\neg(\exists\tau^{\prime}\in\mathbf{M}^{\operatorname{disc-int}})[\varnothing\neq\tau^{\prime}\subsetneq\tau\wedge\gamma\cdot\tau^{\prime}=\tau^{\prime}]].

We see that exp(g,k)p=p\exp(g,k)\cdot p=p^{\prime} with k0k\geq 0 if and only if the tuple (g,k,p,p)(g,k,p,p^{\prime}) satisfies the formula

exp(γ,α,π,π)(α,β1,β2)[α=β2α+β1expcyc(γ,β2,π)explin(γ,β1,π,π)].\exp(\gamma,\alpha,\pi,\pi^{\prime})\equiv(\exists\alpha^{\prime},\beta_{1},\beta_{2})[\alpha=\beta_{2}\alpha^{\prime}+\beta_{1}\wedge\exp_{\operatorname{cyc}}(\gamma,\beta_{2},\pi)\wedge\exp_{\operatorname{lin}}(\gamma,\beta_{1},\pi,\pi^{\prime})].

It is then trivial to extend the definition for the case k<0k<0, establishing the definability of the exponentiation function.

5.4. The AGAPE\operatorname{AGAPE} structure

We now define our ultimate structure

Act3(M,G)=Act𝐆,RO,,𝐌(M,G)=AGAPE(M,G)\operatorname{{Act}}^{3}(M,G)=\operatorname{{Act}}_{\mathbf{G},\operatorname{RO},\mathbb{R},\mathbf{M}}(M,G)=\operatorname{AGAPE}(M,G)

as the extension of Act2(M,G)=Act𝐆,RO,(M,G)\operatorname{{Act}}^{2}(M,G)=\operatorname{{Act}}_{\mathbf{G},\operatorname{RO},\mathbb{R}}(M,G) by including the points in MM and adding the relations

g(p)=q,pWg(p)=q,\quad p\in W

for gGg\in G, p,qMp,q\in M and WRO(M)W\in\operatorname{RO}(M). We are then justified to use expressions such as

pintM,pM,pclU,gn=h,fixg=clU,UV=Wp\in\operatorname{int}M,\quad p\in\partial M,\quad p\in\operatorname{cl}U,g^{n}=h,\quad\operatorname{fix}g=\operatorname{cl}U,\quad U\cup V=W

for points pp, regular open sets U,V,WU,V,W, group elements g,hGg,h\in G and integer nn\in\mathbb{Z} within AGAPE(M,G)\operatorname{AGAPE}(M,G).

6. Balls with definable parametrizations

From this point on, we work in the AGAPE\operatorname{AGAPE} language LAGAPE=L𝐆,RO,,𝐌L_{\operatorname{AGAPE}}=L_{\mathbf{G},\operatorname{RO},\mathbb{R},\mathbf{M}}, containing second order arithmetic and points. The underlying structure will be AGAPE(M,G)\operatorname{AGAPE}(M,G); recall our further standing assumption that dimM>1\dim M>1. We will use the notation In=[0,1]nI^{n}=[0,1]^{n} and

Qn(r):=[r,r]n.Q^{n}(r):=[-r,r]^{n}.

The main objective of this section is to interpret the dimension and collared balls inside of a manifold, as described in the following two theorems:

Theorem 6.1.

For each n2n\geq 2, there exists a formula dimn\dim_{n} such that dimn\models\dim_{n} if and only if MM is an nn–manifold.

Theorem 6.2.

For each n2n\geq 2, there exist formulae

flowsn(u,γ,π),Paramn(u,γ,π,ρ,π)\operatorname{flows}_{n}(u,\gamma,\pi),\operatorname{Param}_{n}(u,\gamma,\pi,\rho,\pi^{\prime})

such that the following hold for all (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} with n=dimMn=\dim M.

  1. (1)

    Let URO(M)U\in\operatorname{RO}(M), g¯Gn\underline{g}\in G^{n} and pMp\in M. If

    flowsn(U,g¯,p)\models\operatorname{flows}_{n}(U,\underline{g},p)

    then there exists a unique homeomorphism

    Ψ=Ψ[U,g¯,p]:InclU\Psi=\Psi[U,\underline{g},p]\colon I^{n}\longrightarrow\operatorname{cl}U

    the graph Γ\Gamma of which satisfies

    Γ={(r,q)In×M:AGAPE(M,G)Paramn(U,g¯,p,r,q)},\Gamma=\{(r,q)\in I^{n}\times M\colon\operatorname{AGAPE}(M,G)\models\operatorname{{Param}}_{n}(U,\underline{g},p,r,q)\},

    and also (0,p)Γ(0,p)\in\Gamma.

  2. (2)

    Let UU and VV be good open balls inside intM\operatorname{int}M such that clUV\operatorname{cl}U\subseteq V; if (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}}, we further assume that vol(U)/vol(V)\operatorname{vol}(U)/\operatorname{vol}(V) is sufficiently small compared to some positive number determined by nn. Then we have

    (γ¯π)flowsn(U,γ¯,π).\models(\exists\underline{\gamma}\exists\pi)\,\operatorname{flows}_{n}(U,\underline{\gamma},\pi).

In Section 8, we will modify the definition of Ψ[U,g¯,p]\Psi[U,\underline{g},p] so that the domain is Qn(2)Q^{n}(2), instead of InI^{n}. We emphasize again that the above formulae for \mathscr{M} and vol\mathscr{M}_{\operatorname{vol}} may differ; for instance, the abbreviated sentence dimn\dim_{n} could be more precisely denoted by dimn\dim_{n} and dimnvol\dim_{n}^{\operatorname{vol}} separately depending on the context.

6.1. Detecting the dimension of a manifold

We prove Theorem 6.1 by interpreting a sufficient amount of dimension theory. For a topological space XX, the order of a finite open cover 𝒰\mathscr{U} is defined as the number

supxX|{U𝒰xU}|.\sup_{x\in X}\;\left\lvert\{U\in\mathscr{U}\mid x\in U\}\right\rvert.

Though in classical literature one considers general open covers, it is sufficient (especially in our situation) to consider finite covers only; cf. [12, 15].

We say the topological dimension of XX is at most nn, and write dimXn\dim X\leq n, if every finite open cover of XX is refined by an open cover with order at most n+1n+1. The topological dimension dimX\dim X is defined to be nn, if dimXn\dim X\leq n holds but dimXn1\dim X\leq n-1 does not. A topological nn–manifold has the topological dimension nn.

A collection of open sets 𝒱={Vi}i\mathscr{V}=\{V_{i}\}_{i\in\mathscr{I}} is said to shrink to another collection 𝒲={Wi}i\mathscr{W}=\{W_{i}\}_{i\in\mathscr{I}} if WiViW_{i}\subseteq V_{i} holds for each ii in the index set \mathscr{I}. Let us note the following well-known facts.

Lemma 6.3.
  1. (1)

    (Lebesgue’s Covering Theorem [21, Theorem IV.2]) If 𝒰\mathscr{U} is a finite open cover of InI^{n} such that no element of 𝒰\mathscr{U} intersects an opposite pair of codimension one faces, then 𝒰\mathscr{U} cannot be refined by an open cover of order at most nn.

  2. (2)

    (Čech [8]) If XX is a metrizable space and if YXY\subseteq X, then dimYdimX\dim Y\leq\dim X.

  3. (3)

    (Ostrand’s Theorem [33, Theorem 3]) If 𝒰={Ui}i\mathscr{U}=\{U_{i}\}_{i\in\mathscr{I}} is a locally finite open cover of a normal space XX satisfying dimXn\dim X\leq n, then for each j=0,,nj=0,\ldots,n, the cover 𝒰\mathscr{U} shrinks to some pairwise disjoint collection 𝒱j={Vij}i\mathscr{V}^{j}=\{V_{i}^{j}\}_{i\in\mathscr{I}} of open sets such that the collection j𝒱j\bigcup_{j}\mathscr{V}^{j} is a cover.

We can now give a characterization of manifold dimension.

Lemma 6.4.

For each positive integer nn and for each compact manifold MM, the following two conditions are equivalent.

  1. (A)

    The dimension of MM is at most nn;

  2. (B)

    Let WW be a regular open set in MM. If

    𝒰={Ui:i=1,2,,2n+1}\mathscr{U}=\{U_{i}\colon i=1,2,\ldots,2^{n+1}\}

    is a regular open cover of clW\operatorname{cl}W, then there exists a pairwise disjoint collection

    𝒱j={Vij:i=1,2,,2n+1}\mathscr{V}^{j}=\{V^{j}_{i}\colon i=1,2,\ldots,2^{n+1}\}

    of regular open sets for each j{0,1,,n}j\in\{0,1,\ldots,n\} such that 𝒰\mathscr{U} shrinks to each 𝒱j\mathscr{V}^{j}, and such that j𝒱j\bigcup_{j}\mathscr{V}^{j} is a cover of clW\operatorname{cl}W.

Proof.

Suppose we have dimMn\dim M\leq n, and assume the hypothesis of part (B). We see from Lemma 6.3 (2) that dimclWn\dim\operatorname{cl}W\leq n. Part (3) of the same lemma implies that 𝒰\mathscr{U} shrinks to a pairwise disjoint collection of (not necessarily regular) open sets

𝒲j={Wij}i=1,,2n+1\mathscr{W}^{j}=\{W_{i}^{j}\}_{i=1,\ldots,2^{n+1}}

for each j{0,1,,n}j\in\{0,1,\ldots,n\} with the property that j𝒲j\bigcup_{j}\mathscr{W}^{j} is a cover of the normal space clW\operatorname{cl}W. By Lemma 2.6, there exists a regular open cover

𝒱:={Vij}i,j\mathscr{V}:=\{V^{j}_{i}\}_{i,j}

of clW\operatorname{cl}W satisfying

clVijWijUi\operatorname{cl}V^{j}_{i}\subseteq W^{j}_{i}\subseteq U_{i}

for all ii and jj. This implies the conclusion of (B).

Conversely, suppose we have condition (B) and assume for contradiction that m:=dimM>nm:=\dim M>n. We first note the following:

Claim.

The unit mm–cube [0,1]m[0,1]^{m} admits a finite regular open cover of cardinality 2n+12^{n+1} that cannot be refined by another open cover with order at most n+1n+1.

Let CC denote the unit cube [0,1]n+1[0,1]^{n+1} in n+1\mathbb{R}^{n+1}, which is embedded in m\mathbb{R}^{m} as the subset with the last mn1m-n-1 coordinates being zero. For each vertex vC(0)v\in C^{(0)}, let us consider the translated open cube

Uv:=v+(1,1)n+1n+1.U_{v}:=v+(-1,1)^{n+1}\subseteq\mathbb{R}^{n+1}.

We then have a regular open cover

𝒰:={Uv:vC(0)}\mathscr{U}:=\{U_{v}\colon v\in C^{(0)}\}

of CC with cardinality 2n+12^{n+1}. Note that each open cube UvU_{v} does not intersect an opposite pair of codimension one faces of CC. By taking the Cartesian product UvU^{\prime}_{v} of each UvU_{v} with (1,2)mn1(-1,2)^{m-n-1}, we obtain a finite regular open cover

𝒰={UvvC(0)}\mathscr{U}^{\prime}=\{U^{\prime}_{v}\mid v\in C^{(0)}\}

of [0,1]m[0,1]^{m}. If 𝒰\mathscr{U}^{\prime} is refined by another finite open cover 𝒱\mathscr{V} of [0,1]m[0,1]^{m} with order at most n+1n+1, then the intersection of the elements in 𝒱\mathscr{V} with n+1m\mathbb{R}^{n+1}\subseteq\mathbb{R}^{m} is a finite open cover of C=[0,1]n+1C=[0,1]^{n+1} with order at most n+1n+1. This violates Lebesgue’s Covering Theorem (Lemma 6.3), and the claim is thus proved.

Let us now consider a good ball QQ in MM, which comes with an embedding

ϕ:mM\phi\colon\mathbb{R}^{m}\longrightarrow M

satisfying ϕ[0,1]m=clQ\phi[0,1]^{m}=\operatorname{cl}Q. By applying the above claim, we obtain a finite regular open cover of clQ\operatorname{cl}Q that cannot be refined by a finite open cover with order at most n+1n+1. This contradicts condition (B), which we have assumed. ∎

Note that the cardinalities of covers 𝒰\mathscr{U} and j𝒱j\bigcup_{j}\mathscr{V}^{j} in condition (B) of the above lemma are explicitly bounded above by 2n+12^{n+1} and (n+1)2n+1(n+1)2^{n+1}, respectively. Note also that conditions such as

clWU1U2n+1\operatorname{cl}W\subseteq U_{1}\cup\cdots\cup U_{2^{n+1}}

are expressible in the AGAPE\operatorname{AGAPE} language. It is therefore clear that condition (B) is expressible in this language, for each fixed positive integer nn. As a consequence, we obtain Theorem 6.1.

6.2. Parametrizing balls in MM in dimension two and higher

For the proof of Theorem 6.2, let us consider the quotient map

pr:/2\mathrm{pr}:\mathbb{R}\longrightarrow\mathbb{R}/\sqrt{2}\mathbb{Z}

defined by

x[x]:=x+2.x\mapsto[x]:=x+\sqrt{2}\mathbb{Z}.

The image of \mathbb{Z} is dense in the circle /2\mathbb{R}/\sqrt{2}\mathbb{Z}, equipped with the natural cyclic order. The expression 2\sqrt{2} will be regarded as a (definable) constant symbol in LAGAPEL_{\operatorname{AGAPE}}. We have chosen this value for concreteness, but for our purpose we could use an arbitrary irrational number that is definable without parameters in arithmetic. There exists a definable function ang(ρ1,ρ2)\operatorname{ang}(\rho_{1},\rho_{2}) satifying

r=ang(r1,r2)r=\operatorname{ang}(r_{1},r_{2})

if and only if the (unsigned) angular metric between [r1][r_{1}] and [r2][r_{2}] is r[0,2)r\in[0,\sqrt{2}).

Let us also define an LAGAPEL_{\operatorname{AGAPE}} formula

fcov(u,v0,,vn)(clu)i=0nvii=0nfincomp(vi).\operatorname{fcov}(u,v_{0},\ldots,v_{n})\equiv(\operatorname{cl}u)\subseteq\bigcup_{i=0}^{n}v_{i}\wedge\bigwedge_{i=0}^{n}\operatorname{fincomp}(v_{i}).

We also use the formula

clshrink(v0,,vn,v0,,vn)i=0nclvivi.\operatorname{clshrink}(v_{0},\ldots,v_{n},v_{0}^{\prime},\ldots,v_{n}^{\prime})\equiv\bigwedge_{i=0}^{n}\operatorname{cl}v_{i}^{\prime}\subseteq v_{i}.

We will equip MM with a compatible metric dd, and denote by dd_{\infty} the induced uniform metric on the homeomorphism group. We have the following characterization of uniform convergence:

Lemma 6.5.

Let UU be a regular open set in MM such that clUintM\operatorname{cl}U\subseteq\operatorname{int}M, and let

F1F2F_{1}\supseteq F_{2}\supseteq\cdots

be a sequence of subsets of Homeo(M)\operatorname{Homeo}(M) such that each fF1f\in F_{1} setwise stabilizes UU. Then the following two conditions are equivalent.

  1. (A)

    We have

    limisup{d(fU,IdU)fFi}=0.\lim_{i\to\infty}\sup\{d_{\infty}(f\restriction_{U},\operatorname{Id}\restriction_{U})\mid f\in F_{i}\}=0.
  2. (B)

    Suppose we have two tuples of regular open sets

    V¯=(V0,,Vn),V¯=(V0,,Vn)\underline{V}=(V_{0},\ldots,V_{n}),\quad\underline{V}^{\prime}=(V_{0}^{\prime},\ldots,V_{n}^{\prime})

    such that

    fcov(U,V¯)fcov(U,V¯)clshrink(V¯,V¯).\operatorname{fcov}(U,\underline{V})\wedge\operatorname{fcov}(U,\underline{V}^{\prime})\wedge\operatorname{clshrink}(\underline{V},\underline{V}^{\prime}).

    Then there exist some iωi\in\omega such that whenever a pair (V^,V^)(\hat{V}^{\prime},\hat{V}) belongs to

    A:={(V^,V^)j=0n(π0(Vj)×π0(Vj))|V^V^},A:=\left\{(\hat{V}^{\prime},\hat{V})\in\bigcup_{j=0}^{n}\left(\pi_{0}(V_{j}^{\prime})\times\pi_{0}(V_{j})\right)\middle|\;\hat{V}^{\prime}\subseteq\hat{V}\right\},

    each fFif\in F_{i} satisfies

    f(V^clU)V^.f(\hat{V}^{\prime}\cap\operatorname{cl}U)\subseteq\hat{V}.
Proof.

Let us assume part (A), and also the hypotheses of (B). We set

ϵ0:=inf{d(clV^,MV^)(V^,V^)A},\epsilon_{0}:=\inf\left\{d(\operatorname{cl}\hat{V}^{\prime},M\setminus\hat{V})\mid(\hat{V}^{\prime},\hat{V})\in A\right\},

which is positive since AA is finite. Choosing ii so that

d(fU,IdU)<ϵ0d_{\infty}(f\restriction_{U},\operatorname{Id}\restriction_{U})<\epsilon_{0}

for all fFif\in F_{i}, we obtain the conclusion.

Conversely, we assume the condition (B) and pick an arbitrary ϵ>0\epsilon>0. Let 𝒰\mathscr{U} be a finite cover of clU\operatorname{cl}U by regular open sets with radius less than ϵ\epsilon. Applying Lemma 6.4 (after replacing the number 2n+12^{n+1} in the lemma by the size of 𝒰\mathscr{U}), we obtain a tuple of regular open sets

V¯=(V0,,Vn)\underline{V}=(V_{0},\ldots,V_{n})

such that every connected component of each VjV_{j} has diameter at most 2ϵ2\epsilon, and such that fcov(U,V¯)\operatorname{fcov}(U,\underline{V}) holds. By Lemma 2.6 and by compactness of clU\operatorname{cl}U, we obtain

V¯=(V0,,Vn)\underline{V}^{\prime}=(V_{0}^{\prime},\ldots,V_{n}^{\prime})

such that

fcov(U,V¯)clshrink(V¯,V¯).\operatorname{fcov}(U,\underline{V}^{\prime})\wedge\operatorname{clshrink}(\underline{V},\underline{V}^{\prime}).

Pick iωi\in\omega as given by the condition (B), and let fFif\in F_{i} and xclUx\in\operatorname{cl}U be arbitrary. Since there exists some (V^,V^)A(\hat{V}^{\prime},\hat{V})\in A such that xV^x\in\hat{V}^{\prime}, we see that

d(x,f(x))diamV^2ϵ.d(x,f(x))\leq\operatorname{diam}\hat{V}\leq 2\epsilon.

This implies that d(fU,IdU)2ϵd_{\infty}(f\restriction_{U},\operatorname{Id}\restriction_{U})\leq 2\epsilon and that condition (A) holds. ∎

We now interpret non-integral powers of group elements, in the following sense:

Lemma 6.6.

There exist formulae

conv(u,γ,ρ,δ),flow(u,γ)\mathrm{conv}(u,\gamma,\rho,\delta),\quad\operatorname{flow}(u,\gamma)

such that the following hold for each (M,G)(vol)(M,G)\in\mathscr{M}_{(vol)}.

  1. (1)

    For group elements {g,h}G\{g,h\}\subseteq G, a regular open set URO(M)U\in\operatorname{RO}(M), and a real number rr\in\mathbb{R} satisfying clUintM\operatorname{cl}U\subseteq\operatorname{int}M and

    g(U)=U=h(U),g(U)=U=h(U),

    we have

    conv(U,g,r,h)\models\mathrm{conv}(U,g,r,h)

    if and only if

    limδ+0sup{d(gsU,hU)s and ang(s,r)<δ}=0.\lim_{\delta\to+0}\sup\left\{d_{\infty}(g^{s}\restriction_{U},h\restriction_{U})\mid s\in\mathbb{Z}\text{ and }\operatorname{ang}(s,r)<\delta\right\}=0.
  2. (2)

    For gGg\in G and URO(M)U\in\operatorname{RO}(M) satisfying clUintM\operatorname{cl}U\subseteq\operatorname{int}M and g(U)=Ug(U)=U, we have

    flow(U,g)\models\operatorname{flow}(U,g)

    if and only if there exists a unique topological flow

    Φ=ΦU,g:/2×UU\Phi=\Phi_{U,g}\colon\mathbb{R}/\sqrt{2}\mathbb{Z}\times U\longrightarrow U

    such that, with the notation Φ([t],p)=Φt(p)\Phi([t],p)=\Phi^{t}(p), we have the conditions below:

    • for each mm\in\mathbb{Z}, we have Φm=gmU\Phi^{m}=g^{m}\restriction_{U};

    • the map [t]Φt[t]\mapsto\Phi^{t} is a topological embedding of /2\mathbb{R}/\sqrt{2}\mathbb{Z} into the group

      GU:={hUhG and h(U)=U}Homeo(U);G\restriction_{U}:=\{h\restriction_{U}\mid h\in G\text{ and }h(U)=U\}\leq\operatorname{Homeo}(U);
    • for each [t][0][t]\neq[0], we have fixΦtU=\operatorname{fix}\Phi^{t}\cap U=\varnothing.

    In this case, for rr\in\mathbb{R} and pUp\in U, the map

    (U,g,r,p)ΦU,gr(p)(U,g,r,p)\mapsto\Phi^{r}_{U,g}(p)

    is definable.

  3. (3)

    If flow(U,g)flow(V,g)\models\operatorname{flow}(U,g)\wedge\operatorname{flow}(V,g), then for pUVp\in U\cap V and rr\in\mathbb{R}, we have

    ΦU,gr(p)=ΦV,gr(p).\Phi^{r}_{U,g}(p)=\Phi^{r}_{V,g}(p).
Proof.

Applying Lemma 6.5 for the definable set

Fi:={h1gss and ang(s,r)<1/i}G,F_{i}:=\{h^{-1}g^{s}\mid s\in\mathbb{N}\text{ and }\operatorname{ang}(s,r)<1/i\}\subseteq G,

we immediately obtain a desired formula conv(γ,δ,u,ρ)\operatorname{conv}(\gamma,\delta,u,\rho).

It is straightforward to check

flow(u,γ)(ρδ)[conv(u,γ,ρ,δ)(ρ2fixδu=)]\operatorname{flow}(u,\gamma)\equiv(\forall\rho\exists\delta)[\mathrm{conv}(u,\gamma,\rho,\delta)\wedge(\rho\in\sqrt{2}\mathbb{Z}\vee\operatorname{fix}\delta\cap u=\varnothing)]

satisfies the desired conditions in (2). In particular, the uniquenss is a consequence of the fact that the formula conv(U,g,r,h)\operatorname{conv}(U,g,r,h) uniquely determines the restriction of hh on UU, as an approximation of the form

{gknU}\{g^{k_{n}}\restriction_{U}\}

satisfying

knrk_{n}\longrightarrow r

in /2\mathbb{R}/\sqrt{2}\mathbb{Z}. The definability of the flow in (2) and the independence on the choice of UU in part (3) also follow by the same reason, completing the proof. ∎

In the situation of Lemma 6.6, we will say that gg defines a circular flow on the open set UU. When we have conv(U,g,r,h)\mathrm{conv}(U,g,r,h), the element gg is viewed as an irrational rotation through a specified angle, and hh is the rotation of the rr–multiple of this angle. By the definability of ΦU,gr(p)\Phi_{U,g}^{r}(p) for pUp\in U, we are justified to use an expression such as

Φu,γρ(π)=π\Phi_{u,\gamma}^{\rho}(\pi)=\pi^{\prime}

in an LAGAPEL_{\operatorname{AGAPE}} formula with the hypothesis that πu\pi\in u. When the meaning is clear, we also use the more succinct notation

gr:=ΦU,gr.g^{r}:=\Phi_{U,g}^{r}.

We are now ready to complete the proof of Theorem 6.2:

Proof of Theorem 6.2.

By Lemma 6.6, we have an LAGAPEL_{\operatorname{AGAPE}} formula flowsn(U,g¯,p)\operatorname{flows}_{n}(U,\underline{g},p) that expresses the following:

  • there exists some V¯={Vi}\underline{V}=\{V_{i}\} such that

    clViintMgi(Vi)=Viflow(Vi,gi)\models\operatorname{cl}V_{i}\subseteq\operatorname{int}M\wedge g_{i}(V_{i})=V_{i}\wedge\operatorname{flow}(V_{i},g_{i})

    for each ii, and such that

    pUclUiVi;p\in U\in\operatorname{cl}U\subseteq\cap_{i}V_{i};
  • there exists a continuous bijection [0,1]nclU[0,1]^{n}\longrightarrow\operatorname{cl}U defined by

    (r1,,rn)igiri(p);(r_{1},\ldots,r_{n})\mapsto\prod_{i}g_{i}^{r_{i}}(p);
  • For all ri[0,1]r_{i}\in[0,1] and for all permutation σ\sigma of {1,,n}\{1,\ldots,n\}, we have

    igiri(p)=igσ(i)rσ(i)(p).\prod_{i}g_{i}^{r_{i}}(p)=\prod_{i}g_{\sigma(i)}^{r_{\sigma(i)}}(p).

Here, it is implicitly required that

i=1j1giri(p)clU\prod_{i=1}^{j-1}g_{i}^{r_{i}}(p)\in\operatorname{cl}U

for all jnj\leq n, so that

i=1jgiri(p)=gjrji=1j1giri(p)\prod_{i=1}^{j}g_{i}^{r_{i}}(p)=g_{j}^{r_{j}}\circ\prod_{i=1}^{j-1}g_{i}^{r_{i}}(p)

is well-defined. The formula Paramn\operatorname{Param}_{n} is simply obtained from the map

(U,gi,ri,p)giri(p).(U,g_{i},r_{i},p)\mapsto g_{i}^{r_{i}}(p).

This proves part (1).

For part (2), we may identify clU=Qn(1)\operatorname{cl}U=Q^{n}(1) and V=intQn(R)V=\operatorname{int}Q^{n}(R) for some sufficiently large RR. We can then choose nn independent circular flows such that each flow rotates UU in some compact solid torus Bn1(1)×S1B^{n-1}(1)\times S^{1} with the rotation number 1/21/\sqrt{2}, and such that on the outside of VV the restrictions of the flows are the identity; see Figure 1 (a), where a suitable homeomorphism is applied to UU for illustrative purposes. Such choices of flows will yield the desired conclusion. ∎

2π2\frac{2\pi}{\sqrt{2}}UUVV
a The ball UU is exactly 1/21/\sqrt{2} fraction of the domain of a flow.
UUVV
b There may not be enough room for a desired measure–preserving flow.
Figure 1. The proof of Theorem 6.2 (2) and a potential issue when UU is not “spaciously collared”.

We remark that in the measure preserving case, if vol(U)/vol(V)\operatorname{vol}(U)/\operatorname{vol}(V) is not sufficiently small, then there may not be enough room for a solid torus inside VV such that clU\operatorname{cl}U occupies (1/2)(1/\sqrt{2})–fraction of the torus. For instance, one may consider an annulus that is homeomorphic to S1×IS^{1}\times I, but which is equipped with a measure that is not the product of the Lebesgue measures on the two factors. Thus, the annulus may be “throttled” in some interval as in Figure 1 (b), and thus there may be no measure preserving flow that globally rotates the annulus.

7. Parametrization of collar neighborhoods

Let us fix an integer n>1n>1. We now describe a definable parametrization of collar neighborhoods of the boundary of a compact nn–manifold. More specifically, we will establish the following.

Theorem 7.1.

Then there exist formulae

collar(κ¯),collarembed(κ¯,π,ρ,π)\operatorname{collar}(\underline{\kappa}),\quad\operatorname{collar-embed}(\underline{\kappa},\pi,\rho,\pi^{\prime})

for some tuple κ¯\underline{\kappa} of variables in the AGAPE\operatorname{AGAPE} language such that each pair (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} with dimM=n\dim M=n satisfies the following:

  1. (1)

    We have that (κ¯)[collar(κ¯)]\models(\exists{\underline{\kappa}})[\operatorname{{collar}}(\underline{\kappa})].

  2. (2)

    Let K¯\underline{K} be a tuple of elements in AGAPE(M,G)\operatorname{AGAPE}(M,G) satisfying

    collar(K¯).\operatorname{{collar}}(\underline{K}).

    Then there exists a unique collar embedding

    u=u[K¯]:M×[0,1)Mu=u[\underline{K}]\colon\partial M\times[0,1)\longrightarrow M

    of M\partial M such that for all points pMp\in\partial M and qMq\in M, and for all r[0,1)r\in[0,1) we have

    u(p,r)=q(AGAPE(M,G)collarembed(K¯,p,r,q)).u(p,r)=q\Longleftrightarrow\left(\operatorname{AGAPE}(M,G)\models\operatorname{collar-embed}(\underline{K},p,r,q)\right).

7.1. Decomposition of a unit cube

Let us fix n>1n>1. We will use a certain partition of a cube to parametrize a collar neighborhood of M\partial M. We set

Λ:={0,1}n1In1,\displaystyle\Lambda:=\{0,1\}^{n-1}\subseteq I^{n-1},
𝟎:=(0,,0),𝟏:=(1,,1)Λ,\displaystyle\mathbf{0}:=(0,\ldots,0),\quad\mathbf{1}:=(1,\ldots,1)\in\Lambda,
𝟎k,𝟏kΛk(In1)k\displaystyle\mathbf{0}^{k},\mathbf{1}^{k}\in\Lambda^{k}\subseteq(I^{n-1})^{k} for k>0,\displaystyle\text{ for }k>0,
len(w):=k\displaystyle\operatorname{len}(w):=k for wΛk,\displaystyle\text{ for }w\in\Lambda^{k},
par(m):=m2m/2\displaystyle\operatorname{par}(m):=m-2\lfloor m/2\rfloor for mω.\displaystyle\text{ for }m\in\omega.

For convention, we also let

Λ0={𝟎0}={}.\Lambda^{0}=\{\mathbf{0}^{0}\}=\{\varnothing\}.

By abuse of notation, we move or remove parantheses rather freely and often write

X(v1,,vk)=X(v1,,vk1),vk=Xv1,,vkX^{(v_{1},\ldots,v_{k})}=X^{(v_{1},\ldots,v_{k-1}),v_{k}}=X^{v_{1},\ldots,v_{k}}

when the vector (v1,,vk)(v_{1},\ldots,v_{k}) is used to index certain objects XX^{*}. For each

w=(v1,,vk)Λkw=(v_{1},\ldots,v_{k})\in\Lambda^{k}

with kωk\in\omega, we let S¯w\bar{S}^{w} be the dyadic cube of side length 1/2k1/2^{k} that contains the following two points as opposite vertices:

i=1kvi/2i,i=1kvi/2i+𝟏/2k.\sum_{i=1}^{k}{v_{i}}/{2^{i}},\quad\sum_{i=1}^{k}{v_{i}}/{2^{i}}+{\mathbf{1}}/{2^{k}}.

For instance, we have

S¯=In1,S¯𝟎=[0,1/2]n1,S¯(𝟎,𝟏)=[1/4,1/2]n1,\bar{S}^{\varnothing}=I^{n-1},\quad\bar{S}^{\mathbf{0}}=[0,1/2]^{n-1},\quad\bar{S}^{(\mathbf{0},\mathbf{1})}=[1/4,1/2]^{n-1},

and so on. We have partitions (with disjoint interiors):

In1\displaystyle I^{n-1} ={S¯w|wΛk} for each kω,\displaystyle=\bigcup\left\{\bar{S}^{w}\middle|w\in\Lambda^{k}\right\}\qquad\qquad\qquad\text{ for each }k\in\omega,
In1×[0,2)\displaystyle I^{n-1}\times[0,2) ={Sw:=S¯w×[212len(w)1,212len(w)]|wkωΛk}.\displaystyle=\bigcup\left\{S^{w}:=\bar{S}^{w}\times\left[2-\frac{1}{2^{\operatorname{len}(w)-1}},2-\frac{1}{2^{\operatorname{len}(w)}}\right]\ \middle|\ w\in\bigcup_{k\in\omega}\Lambda^{k}\right\}.

We have a unique parametrization

σw:InSw\sigma^{w}\colon I^{n}\longrightarrow S^{w}

of the regular cube SwS^{w} obtained by a positive homothety and translation.

7.2. The condition for a collar neighborhood

Let us first consider the case that (M,G)(M,G)\in\mathscr{M}. For a tuple

K¯=(Ui,Ui0,Ui1,Ui0,v,Ui1,v,Ti0,Ti1,pi,0,hori,verti,si0,v,si1,v1in and vΛ)\underline{K}=(U_{i},U_{i}^{0},U_{i}^{1},U_{i}^{0,v},U_{i}^{1,v},T_{i}^{0},T_{i}^{1},p_{i,0}^{\varnothing},\operatorname{hor}_{i},\operatorname{vert}_{i},s_{i}^{0,v},s_{i}^{1,v}\mid 1\leq i\leq n\text{ and }v\in\Lambda)

in the universe of AGAPE(M,G)\operatorname{AGAPE}(M,G), we consider the collection of conditions with appropriate notation as itemized in (a) through (i) below; see Figure 2 for an illustration when n=2n=2.

M\partial M22012\frac{1}{2}34\frac{3}{4}78\frac{7}{8}Vi,jV_{i,j}^{\varnothing}Vi,j𝟏V_{i,j}^{\mathbf{1}}pi,jp_{i,j}^{\varnothing}pi,j𝟎p_{i,j}^{\mathbf{0}}pi,j(𝟎,𝟎)p_{i,j}^{(\mathbf{0},\mathbf{0})}Vi,jV_{i^{\prime},j^{\prime}}^{\varnothing}pi,jp_{i^{\prime},j^{\prime}}^{\varnothing}pi,j𝟎p_{i^{\prime},j^{\prime}}^{\mathbf{0}}pi,j(𝟎,𝟎)p_{i^{\prime},j^{\prime}}^{(\mathbf{0},\mathbf{0})}si0,𝟎¯\underline{s_{i}^{0,\mathbf{0}}}si1,𝟏¯\underline{s_{i}^{1,\mathbf{1}}}si1,𝟎¯\underline{s_{i}^{1,\mathbf{0}}}si0,𝟏¯\underline{s_{i}^{0,\mathbf{1}}}si0,𝟎¯\underline{s_{i}^{0,\mathbf{0}}}si0,𝟏¯\underline{s_{i}^{0,\mathbf{1}}}si0,𝟎¯\underline{s_{i}^{0,\mathbf{0}}}Ui0U_{i}^{0}Ui1U_{i}^{1}Ui0U_{i}^{0}
Figure 2. The condition COL(M,G;K¯)\operatorname{COL}(M,G;\underline{K}).
Condition COL(M,G;K¯)\operatorname{COL}(M,G;\underline{K})
  1. (a)

    We have regular open sets UU^{*} and U1,,UnU_{1},\ldots,U_{n} such that

    MU=1inUi,\partial M\subseteq U^{*}=\bigcup_{1\leq i\leq n}U_{i},

    and such that every regular open neighborhood of M\partial M contains g(U)g(U^{*}) for some gGg\in G; moreover, each UiU_{i} has finitely many components, and the closures of distinct components are disjoint.

  2. (b)

    We have dispersed (see Definition 3.11) regular open sets

    Ui0,Ui1,Ui0,v,Ui1,vU_{i}^{0},U_{i}^{1},U_{i}^{0,v},U_{i}^{1,v}

    for each ini\leq n and vΛv\in\Lambda; moreover, we have for each ϵ{0,1}\epsilon\in\{0,1\} that

    Ui=Ui0Ui1,Uiϵ=vΛUiϵ,v.U_{i}=U_{i}^{0}\oplus U_{i}^{1},\quad U_{i}^{\epsilon}=\oplus_{v\in\Lambda}U_{i}^{\epsilon,v}.
  3. (c)

    For each ini\leq n, we have hori,vertiG\operatorname{hor}_{i},\operatorname{vert}_{i}\in G and

    pi,0Ti0Ti1𝒫disc(intM)p_{i,0}^{\varnothing}\in T_{i}^{0}\subseteq T_{i}^{1}\in\mathscr{P}^{\operatorname{disc}}(\operatorname{int}M)

    such that Ti0T_{i}^{0} is a nonempty, finite, minimal hori\operatorname{hor}_{i}–invariant set; moreover, the map

    (j,k)pi,j𝟎k:=vertikhorij(pi,0)(j,k)\mapsto p_{i,j}^{\mathbf{0}^{k}}:=\operatorname{vert}_{i}^{k}\circ\operatorname{hor}_{i}^{j}\left(p_{i,0}^{\varnothing}\right)

    is a bijection

    {0,,#Ti01}×ωTi1.\{0,\ldots,\#T_{i}^{0}-1\}\times\omega\longrightarrow T_{i}^{1}.
  4. (d)

    For each ini\leq n and j<#Ti0j<\#T_{i}^{0}, there exists a unique Ui,jπ0UiU_{i,j}\in\pi_{0}U_{i} satisfying

    pi,jfrUi,j.p_{i,j}^{\varnothing}\in\operatorname{fr}U_{i,j}.

    For each kωk\in\omega, there also exists a unique Ui,jkπ0Uipar(k)U_{i,j}^{k}\in\pi_{0}U_{i}^{\operatorname{par}(k)} such that

    pi,j𝟎kfrUi,jk.p_{i,j}^{\mathbf{0}^{k}}\in\operatorname{fr}U_{i,j}^{k}.

    We further have closure–disjoint unions

    Ui=jUi,j,Ui0=j,kUi,j2k,Ui1=j,kUi,j2k+1.U_{i}=\bigsqcup_{j}U_{i,j},\qquad U_{i}^{0}=\bigsqcup_{j,k}U_{i,j}^{2k},\qquad U_{i}^{1}=\bigsqcup_{j,k}U_{i,j}^{2k+1}.
  5. (e)

    For each ini\leq n, we have s¯iGn\underline{s}_{i}^{\varnothing}\in G^{n}. Setting Vi,j:=Ui,j0V_{i,j}^{\varnothing}:=U_{i,j}^{0}, we also have

    flowsn(Vi,j,s¯i,pi,j),\displaystyle\models\operatorname{flows}_{n}\left(V_{i,j}^{\varnothing},\underline{s}_{i}^{\varnothing},p_{i,j}^{\varnothing}\right),
    Ψi,j:=Ψ[Vi,j,s¯i,pi,j]:InclVi,j.\displaystyle\Psi_{i,j}^{\varnothing}:=\Psi\left[V_{i,j}^{\varnothing},\underline{s}_{i}^{\varnothing},p_{i,j}^{\varnothing}\right]\colon I^{n}\longrightarrow\operatorname{cl}V_{i,j}^{\varnothing}.

    For all k>0k>0 and (v1,,vk)Λk(v_{1},\ldots,v_{k})\in\Lambda^{k} we have that

    pi,j(v1,,vk):=vertikΨi,j(i=1kvi2i,0)frUi,jk1frUi,jk.p_{i,j}^{(v_{1},\ldots,v_{k})}:=\operatorname{vert}_{i}^{k}\circ\Psi_{i,j}^{\varnothing}\left(\sum_{i=1}^{k}\frac{v_{i}}{2^{i}},0\right)\in\operatorname{fr}U_{i,j}^{k-1}\cap\operatorname{fr}U_{i,j}^{k}.
  6. (f)

    For each (i,j,w=(v1,,vk))(i,j,w=(v_{1},\ldots,v_{k})) in the index set

    :={(i,j,w)|1in,0j<#Ti0,wkωΛk},\mathscr{I}:=\left\{(i,j,w)\middle|1\leq i\leq n,0\leq j<\#T_{i}^{0},w\in\bigcup_{k\in\omega}\Lambda^{k}\right\},

    there exists a unique

    Vi,jwπ0Uipar(k),vk,V_{i,j}^{w}\in\pi_{0}U_{i}^{\operatorname{par}(k),v_{k}},

    the closure of which contains pi,jwp_{i,j}^{w}.

  7. (g)

    For each ini\leq n and vΛv\in\Lambda, we have

    s¯i0,v,s¯i1,vGn.\underline{s}_{i}^{0,v},\underline{s}_{i}^{1,v}\in G^{n}.

    We further have that

    flowsn(Vi,jw,s¯ipar(k),vk,pi,jw),\displaystyle\models\operatorname{flows}_{n}\left(V_{i,j}^{w},\underline{s}_{i}^{\operatorname{par}(k),v_{k}},p_{i,j}^{w}\right),
    Ψi,jw:=Ψ[Vi,jw,s¯ipar(k),vk,pi,jw]:InclVi,jw.\displaystyle\Psi_{i,j}^{w}:=\Psi\left[V_{i,j}^{w},\underline{s}_{i}^{\operatorname{par}(k),v_{k}},p_{i,j}^{w}\right]\colon I^{n}\longrightarrow\operatorname{cl}V_{i,j}^{w}.
  8. (h)

    For each (i,j,)(i,j,\varnothing)\in\mathscr{I}, there exists a homeomorphism

    Ψi,j:In1×[0,2]clUi,j\Psi_{i,j}\colon I^{n-1}\times[0,2]\longrightarrow\operatorname{cl}U_{i,j}

    such that for each wΛkw\in\Lambda^{k} we have

    Ψi,jSw=Ψi,jw(σw)1,\Psi_{i,j}\restriction_{S^{w}}=\Psi_{i,j}^{w}\circ(\sigma^{w})^{-1},

    and such that

    Ψi,j(In1×{2})M.\Psi_{i,j}(I^{n-1}\times\{2\})\subseteq\partial M.
  9. (i)

    If xclUi,jclUi,jx\in\operatorname{cl}U_{i,j}\cap\operatorname{cl}U_{i^{\prime},j^{\prime}}, then some v,vIn1v,v^{\prime}\in I^{n-1} and t[0,2]t\in[0,2] satisfy

    x=Ψi,j(v,t)=Ψi,j(v,t).x=\Psi_{i,j}(v,t)=\Psi_{i^{\prime},j^{\prime}}(v^{\prime},t).

    Moreover, in this case we require that for each t[0,2]t^{\prime}\in[0,2], we have

    Ψi,j(v,t)=Ψi,j(v,t).\Psi_{i,j}(v,t^{\prime})=\Psi_{i^{\prime},j^{\prime}}(v^{\prime},t^{\prime}).

We now make three claims. First, these conditions are first order expressible. Second, these conditions produce a definable collar embedding; for this, we will actualy need only the conditions (h) and (i). Third, every pair (M,G)(M,G)\in\mathscr{M} satisfies these conditions with a suitable choice of K¯\underline{K}.

The first point is trivial to check from the preceding results, possibly except for the continuity condition in (h) at the level–22 subset of In1×[0,2]I^{n-1}\times[0,2]. At such a point x0x_{0}, we then can simply require the convergence of the values of the form

Ψi,jw(σw)1(x)\Psi_{i,j}^{w}\circ(\sigma^{w})^{-1}(x)

whenever xSwx\in S^{w} gets arbitrarily close to x0x_{0}; we also require the bijectivity of the resulting map onto clUi,j\operatorname{cl}U_{i,j}. We can now let collar(κ¯)\operatorname{collar}(\underline{\kappa}) be the formula expressing the condition COL(M,G;K¯)\operatorname{COL}(M,G;\underline{K}).

Regarding the second point, we note the following:

Claim.

Under the hypothesis COL(M,G;K¯)\operatorname{COL}(M,G;\underline{K}), we have a collar embedding

u=u[K¯]:M×[0,2]Mu=u\left[\underline{K}\right]\colon\partial M\times[0,2]\longrightarrow M

which is unambiguously defined by

u(Ψi,j(v,2),r)=Ψi,j(v,r)u(\Psi_{i,j}(v,2),r)=\Psi_{i,j}(v,r)

for all

(i,j,),vIn1,r[0,2].(i,j,\varnothing)\in\mathscr{I},\quad v\in I^{n-1},\quad r\in[0,2].

In particular, the image of the level–22 set under the map uu coincides with M\partial M.

Proof.

The well-definedness and the injectivity follow from the condition (i) above. This map uu is continuous because Ψi,j\Psi_{i,j} is for all ii and jj. The condition (h) further implies that this map uu is a collar embedding of the boundary. ∎

From the above claim and from the definability of Ψi,jw\Psi_{i,j}^{w}, we obtain the desired formula collaremb(κ¯,π,ρ,π)\operatorname{collar-emb}(\underline{\kappa},\pi,\rho,\pi^{\prime}) expressing the map uu. We complete the proof of part (2) in Theorem 7.1 by simply reparametrizing uu so that the level–0 set corresponds to the boundary.

For the third claim, and hence part (1) of the theorem, we note that the condition (a) is equivalent to clU\operatorname{cl}U^{*} being contained in a collar neighborhood. Hence, we may simply start with a homeomorphism

u:M×[0,2]clUu\colon\partial M\times[0,2]\longrightarrow\operatorname{cl}U^{*}

that satisfies

u(x,2)=xM.u(x,2)=x\in\partial M.

Using Ostrand’s theorem (Lemma 6.3 (3)), we can write

M=1inclW¯i=1inW¯i\partial M=\bigcup_{1\leq i\leq n}\operatorname{cl}\bar{W}_{i}=\bigcup_{1\leq i\leq n}\bar{W}_{i}

for some clW¯iM\operatorname{cl}\bar{W}_{i}\subseteq\partial M, each of whose components clWi,j\operatorname{cl}W_{i,j} is homeomorphic to In1I^{n-1}. We have a natural homeomorphism

ui,j:In1×[0,2]Ui,j:=u(Wi,j×[0,2]).u_{i,j}\colon I^{n-1}\times[0,2]\longrightarrow U_{i,j}:=u(W_{i,j}\times[0,2]).

Denote by pi,j𝟎kp_{i,j}^{\mathbf{0}^{k}} the image of (0,0,,0,21/2k1)(0,0,\ldots,0,2-1/2^{k-1}) under this homeomorphism. We can find a homeomorphism hori\operatorname{hor}_{i} that permutes the components clUi,j\operatorname{cl}U_{i,j} of clUi\operatorname{cl}U_{i} as in condition (d). We let Ti0:={pi,j}jT_{i}^{0}:=\{p_{i,j}^{\varnothing}\}_{j} and Ti1:={pi,j𝟎k}j,kT_{i}^{1}:=\{p_{i,j}^{\mathbf{0}^{k}}\}_{j,k}. We further define

Vi,jw:=ui,j(Sw),V_{i,j}^{w}:=u_{i,j}(S^{w}),

and set

Ui,jk:=wΛkVi,jw,Ui0:=j,kUi,j2k.U_{i,j}^{k}:=\oplus_{w\in\Lambda^{k}}V_{i,j}^{w},\quad U_{i}^{0}:=\bigsqcup_{j,k}U_{i,j}^{2k}.

The regular open sets Ui1,Ui0,v,Ui1,vU_{i}^{1},U_{i}^{0,v},U_{i}^{1,v} are similar and straightforward to define. The homeomorphism verti\operatorname{vert}_{i} is clearly defined, so that verti(pi,jw)=pi,jw,𝟎\operatorname{vert}_{i}(p_{i,j}^{w})=p_{i,j}^{w,\mathbf{0}}. After decomposing Ui,jkU_{i,j}^{k} modeled on {Sw}\{S^{w}\}, we find s¯ipar(k),v\underline{s}_{i}^{\operatorname{par}(k),v} for the current setup using the uniform convergence theorem. Here, it is crucial that the diameters of the cubes Vi,jwV_{i,j}^{w} converge to zero as they approach the boundary. This completes the proof of the case (M,G)(M,G)\in\mathscr{M}.

Slightly more care is needed in the measure preserving case (M,G)vol(M,G)\in\mathscr{M}_{\operatorname{vol}}. To guarantee the existence of a measure preserving flow avoiding issues as described in Figure 1, we need that the components of the supports of flow-generating homeomorphsms s¯iv\underline{s}_{i}^{v} to be sufficiently far from each other. More precisely, we will pick a sufficiently large n0>0n_{0}>0 depending on MM, and replace condition (g) by the following two conditions; we also change the definition of the tuple κ¯\underline{\kappa}, which is now required to contain the group tuple variables s¯i,jw\underline{s}_{i,j}^{w} as below.

  1. (f)’

    For each k{1,,n0}k\in\{1,\ldots,n_{0}\}, ϵ{0,1}\epsilon\in\{0,1\} and w=(v1,,vk)Λkw=(v_{1},\ldots,v_{k})\in\Lambda^{k}, we have

    s¯w,s¯ϵ,wGn.\underline{s}^{w},\underline{s}^{\epsilon,w}\in G^{n}.

    We further have that

    flowsn(Vi,jw,s¯iw,pi,jw),\displaystyle\models\operatorname{flows}_{n}\left(V_{i,j}^{w},\underline{s}_{i}^{w},p_{i,j}^{w}\right),
    Ψi,jw:=Ψ[Vi,jw,s¯iw,pi,jw]:InclVi,jw.\displaystyle\Psi_{i,j}^{w}:=\Psi\left[V_{i,j}^{w},\underline{s}_{i}^{w},p_{i,j}^{w}\right]\colon I^{n}\longrightarrow\operatorname{cl}V_{i,j}^{w}.
  2. (f)”

    For each k>n0k>n_{0} and w=(v1,,vk)w=(v_{1},\ldots,v_{k}), after setting w:=(vkn0+1,,vk)w^{\prime}:=(v_{k-n_{0}+1},\ldots,v_{k}) we have that

    flowsn(Vi,jw,s¯ipar(k),w,pi,jw),\displaystyle\models\operatorname{flows}_{n}\left(V_{i,j}^{w},\underline{s}_{i}^{\operatorname{par}(k),w^{\prime}},p_{i,j}^{w}\right),
    Ψi,jw:=Ψ[Vi,jw,s¯ipar(k),w,pi,jw]:InclVi,jw.\displaystyle\Psi_{i,j}^{w}:=\Psi\left[V_{i,j}^{w},\underline{s}_{i}^{\operatorname{par}(k),w^{\prime}},p_{i,j}^{w}\right]\colon I^{n}\longrightarrow\operatorname{cl}V_{i,j}^{w}.

Part (2) of Theorem 7.1 is still proved in the same way, even independently of the choice of n01n_{0}\geq 1. For part (1), we choose n0n_{0} sufficiently large, under a fixed metric and a measure on some chart neighborhood of MM. We will require that for each fixed w:=(v1,,vn0)Λn0w^{\prime}:=(v_{1},\ldots,v_{n_{0}})\in\Lambda^{n_{0}}, each open set in the collection

{Vi,jw|w=(,v1,,vn0)k>n0Λk}\left\{V_{i,j}^{w}\middle|w=(\ldots,v_{1},\ldots,v_{n_{0}})\in\bigcup_{k>{n_{0}}}\Lambda^{k}\right\}

is contained in some closure–disjoint collection of open balls

{Wi,jw|w=(,v1,,vn0)k>n0Λk}\left\{W_{i,j}^{w}\middle|w=(\ldots,v_{1},\ldots,v_{n_{0}})\in\bigcup_{k>{n_{0}}}\Lambda^{k}\right\}

with the additional requirement that vol(Vi,jw)/vol(Wi,jw)\operatorname{vol}(V_{i,j}^{w})/\operatorname{vol}(W_{i,j}^{w}) is sufficiently small, in the sense of Theorem 6.2. This guarantees the existence and the convergence of each measure preserving homeomorphism of the required form s¯ϵ,w\underline{s}^{\epsilon,w^{\prime}}, thus completing the proof.

8. Completing the proof

Cheeger and Kister [9] proved that there exist only countably many homeomorphism types of compact manifolds. A key step in their proof is that the topological type of a manifold is invariant under “small” perturbations, in some quantitatively precise sense. As is more concretely described below, this step will be crucial for the construction of the sentences ϕM\phi_{M} and ϕMvol\phi_{M}^{\operatorname{vol}}.

For positive integers n,kn,k and \ell, we denote by (n,k,)\mathscr{E}(n,k,\ell) the set of all tuples of embeddings

f¯=(f1,1,,f1,k,f2,1,,f2,)\underline{f}=(f_{1,1},\ldots,f_{1,k},f_{2,1},\ldots,f_{2,\ell})

from Qn(2)Q^{n}(2) to 2n+1\mathbb{R}^{2n+1} such that the following conditions hold:

  1. (i)

    The following set is a compact connected nn–manifold:

    M=C(f¯):=i,jimfi,j2n+1.M=C(\underline{f}):=\bigcup_{i,j}\operatorname{im}f_{i,j}\subseteq\mathbb{R}^{2n+1}.
  2. (ii)

    There exists a collar u:M×[2,2]Mu\colon\partial M\times[-2,2]\longrightarrow M such that u(x,2)=xu(x,-2)=x for all xx.

  3. (iii)

    We have that

    Mu(M×[2,0))if1,j(intQn(1))if1,j(Qn(2))Mu(M×[2,1]).M\setminus u(\partial M\times[-2,0))\subseteq\bigcup_{i}f_{1,j}(\operatorname{int}Q^{n}(1))\subseteq\bigcup_{i}f_{1,j}(Q^{n}(2))\subseteq M\setminus u(\partial M\times[-2,-1]).
  4. (iv)

    For each i=1,,i=1,\ldots,\ell, the restriction

    f2,jQn1(2)×{2}f_{2,j}\restriction_{Q^{n-1}(2)\times\{-2\}}

    is an embedding of Qn1(2)Q^{n-1}(2) into M\partial M such that

    f2,j(x,t)=u(f2,j(x,2),t),f_{2,j}(x,t)=u(f_{2,j}(x,-2),t),

    where here xQn1(2)x\in Q^{n-1}(2) and t[2,2]t\in[-2,2], and such that

    M=if2,j(intQn1(1)×{2}).\partial M=\bigcup_{i}f_{2,j}(\operatorname{int}Q^{n-1}(1)\times\{-2\}).

Every compact nn–manifold MM is homeomorphic to C(f¯)C(\underline{f}) for some tuple

f¯=(fi,j)(n,k,)\underline{f}=(f_{i,j})\in\mathscr{E}(n,k,\ell)

as above, which we call as a parametrized cover of C(f¯)C(\underline{f}). The space (n,k,)\mathscr{E}(n,k,\ell) inherits the uniform separable metric from the space

C0(Qn(2),(2n+1)(k+)).C^{0}(Q^{n}(2),\mathbb{R}^{(2n+1)(k+\ell)}).

The proof of Cheeger and Kister essentially boils down to the following rigidity result, along with a deep result of Edwards and Kirby on deformation of embeddings in manifolds [16].

Lemma 8.1.

[9] For each f¯(n,k,)\underline{f}\in\mathscr{E}(n,k,\ell) and for each ϵ>0\epsilon>0, there exists δ>0\delta>0 such that every g¯(n,k,)\underline{g}\in\mathscr{E}(n,k,\ell) that is at most δ\delta–far from f¯\underline{f} admits a homeomorphism

C(f¯)C(g¯)C(\underline{f})\longrightarrow C(\underline{g})

that is at most ϵ\epsilon–far from the identity map.

We choose a sufficiently small δ>0\delta>0 for which the conclusion of Lemma 8.1 holds, and call it as a Cheeger–Kister number of f¯(n,k)\underline{f}\in\mathscr{E}(n,k); for our purposes, we will further require δ\delta to be rational. Our strategy for proving Theorem 1.4 is providing a sentence in LAGAPEL_{\operatorname{AGAPE}} which is modeled by an input manifold M2n+1M\subseteq\mathbb{R}^{2n+1}, such that the sentence holds for the structure AGAPE(N,H)\operatorname{AGAPE}(N,H) if and only if NN admits an embedding into Euclidean space that is within the Cheeger–Kister number of a fixed parametrized cover of MM.

In order to execute this strategy, let us fix a pair (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} with dimM=n\dim M=n. We will slightly modify the definition in Theorem 6.2 by affine transformations, so that Ψ[U,g¯,p]\Psi[U,\underline{g},p] is a map from Qn(2)Q^{n}(2) into MM, sending (2,,2)(-2,\ldots,-2) to pp.

We let kk and \ell be positive integers, and consider a tuple

f¯=(f1,1,,f1,k,f2,1,,f2,)\underline{f}=(f_{1,1},\ldots,f_{1,k},f_{2,1},\ldots,f_{2,\ell})

of functions in C0(n,2n+1)C^{0}(\mathbb{R}^{n},\mathbb{R}^{2n+1}). Let us denote by EMB(M,G;f¯)\operatorname{EMB}(M,G;\underline{f}) the collection of all the conditions below from (a) through (e); see also Figure 3:

  1. (a)

    each fi,jf_{i,j} restricts to an embedding of Qn(2)Q^{n}(2) into 2n+1\mathbb{R}^{2n+1};

  2. (b)

    for all indices as above, we have some

    Ui,jRO(M),pi,jintM,g¯i,jGnU_{i,j}\in\operatorname{RO}(M),\quad p_{i,j}\in\operatorname{int}M,\quad\underline{g}_{i,j}\in G^{n}

    satisfying flowsn(Ui,j,g¯i,j,pi,j)\operatorname{flows}_{n}(U_{i,j},\underline{g}_{i,j},p_{i,j}), corresponding to the homeomorphism

    hi,j:=Ψ[Ui,j,g¯i,j,pi,j]:Qn(2)clUi,jintM;h_{i,j}:=\Psi\left[U_{i,j},\underline{g}_{i,j},p_{i,j}\right]\colon Q^{n}(2)\longrightarrow\operatorname{cl}U_{i,j}\subseteq\operatorname{int}M;
  3. (c)

    there exists a collar

    u:M×[3,2]Mu\colon\partial M\times[-3,2]\longrightarrow M

    such that uM×{3}=1Mu\restriction_{\partial M\times\{-3\}}=1\restriction_{\partial M}, and such that

    Mu(M×[3,0))jh1,j(intQn(1))jh1,j(Qn(2))Mu(M×[3,1]);M\setminus u(\partial M\times[-3,0))\subseteq\bigcup_{j}h_{1,j}(\operatorname{int}Q^{n}(1))\subseteq\bigcup_{j}h_{1,j}(Q^{n}(2))\subseteq M\setminus u(\partial M\times[-3,-1]);
  4. (d)

    for each j=1,,j=1,\ldots,\ell, the restriction

    h2,jQn1(2)×{2}h_{2,j}\restriction_{Q^{n-1}(2)\times\{-2\}}

    is an embedding of Qn1(2)Q^{n-1}(2) into u(M×{2})u(\partial M\times\{-2\}) such that

    h2,j(x,t)=u(h2,j(x,2),t),h_{2,j}(x,t)=u(h_{2,j}(x,-2),t),

    where here xQn1(2)x\in Q^{n-1}(2) and t[2,2]t\in[-2,2], and such that

    u(M×{2})=jh2,j(intQn1(1)×{2}).u(\partial M\times\{-2\})=\bigcup_{j}h_{2,j}(\operatorname{int}Q^{n-1}(1)\times\{-2\}).
  5. (e)

    whenever xclUa,bclUc,dx\in\operatorname{cl}U_{a,b}\cap\operatorname{cl}U_{c,d} for some a,b,c,da,b,c,d, we have

    fa,bha,b1(x)=fc,dhc,d1(x).f_{a,b}\circ h_{a,b}^{-1}(x)=f_{c,d}\circ h_{c,d}^{-1}(x).

The condition EMB(M,G;f¯)\operatorname{EMB}(M,G;\underline{f}) implies that

f¯h¯1:=(fi,jhi,j1)\underline{f}\circ\underline{h}^{-1}:=(f_{i,j}\circ h_{i,j}^{-1})

defines an embedding

M:=Mu(M×[3,2))2n+1,M^{\prime}:=M\setminus u(\partial M\times[-3,-2))\hookrightarrow\mathbb{R}^{2n+1},

and that the tuple f¯\underline{f} is a parametrized cover of the image.

M\partial M3-3MM^{\prime}2-22\phantom{-}21-10\phantom{-}0imh1,i\operatorname{im}h_{1,i}imh2,j\operatorname{im}h_{2,j}
Figure 3. Parts (c) and (d) of the condition EMB(M,G;f)\operatorname{EMB}(M,G;f).

Recall the domain of the sort symbol Contn,2n+1\operatorname{Cont}_{n,2n+1} is C0(n,2n+1)C^{0}(\mathbb{R}^{n},\mathbb{R}^{2n+1}). By the preceding results, there exists a formula

Embedn,k,(χ¯),\operatorname{Embed}_{n,k,\ell}(\underline{\chi}),

expressing EMB(M,G;f¯)\operatorname{EMB}(M,G;\underline{f}) in AGAPE(M,G)\operatorname{AGAPE}(M,G). We emphasize that although the maps hi,jh_{i,j} do not belong to the universe of AGAPE(M,G)\operatorname{AGAPE}(M,G), Theorem 7.1 together with our access to the real numbers enables us to use such expressions. Let us record this fact:

Lemma 8.2.

Let (M,G)(vol)(M,G)\in\mathscr{M}_{(\operatorname{vol})} satisfy dimM=n\dim M=n. For positive integers n,kn,k and \ell, there exists a formula Embedn,k,(χ¯)\operatorname{{Embed}}_{n,k,\ell}(\underline{\chi}) with a (k+)(k+\ell)–tuple of Contn,2n+1\operatorname{Cont}_{n,2n+1} variables

χ¯=(χ1,1,,χ1,k,χ2,1,,χ2,)\underline{\chi}=(\chi_{1,1},\ldots,\chi_{1,k},\chi_{2,1},\ldots,\chi_{2,\ell})

in the AGAPE\operatorname{AGAPE} language such that

Embedn,k,(f¯)\models\operatorname{{Embed}}_{n,k,\ell}(\underline{f})

if and only if the condition EMB(M,G;f¯)\operatorname{EMB}(M,G;\underline{f}) is satisfied.

We can now establish the main result of this paper.

Proof of Theorem 1.4.

We may assume that n:=dimM>1n:=\dim M>1 and that M2n+1M\subseteq\mathbb{R}^{2n+1}. Consider a parametrized cover

f¯(n,k,)C0(Qn(2),(2n+1)(k+))\underline{f}\in\mathscr{E}(n,k,\ell)\subseteq C^{0}(Q^{n}(2),\mathbb{R}^{(2n+1)(k+\ell)})

of M=C(f¯)M=C(\underline{f}). We have a corresponding Cheeger–Kister rational number

δ=δ(M,f¯)>0.\delta=\delta(M,\underline{f})>0.

Let us pick δ0>0\delta_{0}>0 such that

supxyδ0f¯(x)f¯(y)<δ/3.\sup_{\|x-y\|\leq\delta_{0}}\|\underline{f}(x)-\underline{f}(y)\|<\delta/3.

We can find a partition {C1,,Cs}\{C_{1},\ldots,C_{s}\} of Qn(2)Q^{n}(2) having diameters less than δ0\delta_{0} such that each CiC_{i} is the intersection of Qn(2)Q^{n}(2) with a cube with rational corners. Each CiC_{i} is definable in LAGAPEL_{\operatorname{AGAPE}}, since so is every rational number. We arbitrarily pick xix_{i} in CiC_{i}, and choose qi(2n+1)(k+)q_{i}\in\mathbb{Q}^{(2n+1)(k+\ell)} such that

f¯(xi)qi<δ/3.\|\underline{f}(x_{i})-q_{i}\|<\delta/3.

Let us now consider the following conditions for an arbitrary (N,H)(vol)(N,H)\in\mathscr{M}_{(\operatorname{vol})}, which are first order expressible in LAGAPEL_{\operatorname{AGAPE}} by preceding results:

  • dimN=n\dim N=n;

  • some tuple g¯C0(n,(2n+1)(k+))\underline{g}\in C^{0}(\mathbb{R}^{n},\mathbb{R}^{(2n+1)(k+\ell)}) satisfies that

    AGAPE(N,H)Embedn(g¯),\operatorname{AGAPE}(N,H)\models\operatorname{{Embed}}_{n}(\underline{g}),

    and also

    supxCig¯(x)qi<δ/3.\sup_{x\in C_{i}}\|\underline{g}(x)-q_{i}\|<\delta/3.

The above conditions are obviously met in the case when (N,H)=(M,G)(N,H)=(M,G). We also note that for each xCix\in C_{i} that

g¯(x)f¯(x)g¯(x)qi+qif¯(xi)+f¯(xi)f¯(x)<δ.\|\underline{g}(x)-\underline{f}(x)\|\leq\|\underline{g}(x)-q_{i}\|+\|q_{i}-\underline{f}(x_{i})\|+\|\underline{f}(x_{i})-\underline{f}(x)\|<\delta.

By Lemma 8.1, we see that NN is homeomorphic to MM.∎

9. Further questions

A large number of interesting open questions remain. We already mentioned Question 1.5. Part of the motivation for this question is the theory of critical regularity of groups, which seeks to distinguish between diffeomorphism groups of various regularities of a given manifold by the isomorphism types of finitely generated subgroups; cf. [23, 30]. Along this line of question, one may ask whether or not the CkC^{k}–analogue of Theorem 1.4 holds.

Question 9.1.

Let MM be a compact, connected, smooth manifold, and let NN be an arbitrary smooth manifold. Is there a sentence ϕk,M\phi_{k,M} in the language of groups such that if Diffk(N)\operatorname{Diff}^{k}(N) satisfies ϕk,M\phi_{k,M} then NN is diffeomorphic to MM?

Relatedly, leaving the framework of first order rigidity, we have the following.

Question 9.2.

Let MM be a compact, connected, smooth manifold. Is there a finitely generated (or countable) group GMG_{M} such that GMG_{M} acts faithfully by CkC^{k} diffeomorphisms of a compact, connected, smooth manifold NN of the same dimension as MM if and only NN is CkC^{k} diffeomorphic to MM?

The discussion in the present article depended heavily on the compactness of the comparison manifold.

Question 9.3.

Let MM be an arbitrary manifold. Under what conditions is there a sentence ϕM\phi_{M} in the language of groups such that if NN is an arbitrary manifold then Homeo(N)\operatorname{Homeo}(N) satisfies ϕM\phi_{M} if and only if NN is homeomorphic to MM? More generally, under what conditions does Homeo(M)Homeo(N)\operatorname{Homeo}(M)\equiv\operatorname{Homeo}(N) imply MNM\cong N?

We conclude by asking what the weakest hypotheses on GG can be.

Question 9.4.

For what classes of subgroups of Homeo(M)\operatorname{Homeo}(M) do the conclusions of Theorem 1.4 hold?

A partial answer to Question 9.4 is given in [27], as noted in Remark 1.10.

Acknowledgements

The first and the third author are supported by Mid-Career Researcher Program (RS-2023-00278510) through the National Research Foundation funded by the government of Korea. The first and the third authors are also supported by KIAS Individual Grants (MG073601 and MG084001, respectively) at Korea Institute for Advanced Study and by Samsung Science and Technology Foundation under Project Number SSTF-BA1301-51. The second author is partially supported by NSF Grants DMS-2002596 and DMS-2349814, Simons Foundation International Grant SFI-MPS-SFM-00005890. The authors thank M. Brin, J. Hanson and O. Kharlampovich, and for helpful discussions. The authors are deeply grateful to C. Rosendal for introducing the result of Cheeger–Kister to them.

References

  • [1] M. Ajtai, Isomorphism and higher order equivalence, Ann. Math. Logic 16 (1979), no. 3, 181–203. MR548473
  • [2] Nir Avni, Alexander Lubotzky, and Chen Meiri, First order rigidity of non-uniform higher rank arithmetic groups, Invent. Math. 217 (2019), no. 1, 219–240. MR3958794
  • [3] Augustin Banyaga, The structure of classical diffeomorphism groups, Mathematics and its Applications, vol. 400, Kluwer Academic Publishers Group, Dordrecht, 1997. MR1445290 (98h:22024)
  • [4] Thomas William Barrett and Hans Halvorson, Quine’s conjecture on many-sorted logic, Synthese 194 (2017), no. 9, 3563–3582.
  • [5] Salomon Bochner and Deane Montgomery, Groups of differentiable and real or complex analytic transformations, Ann. of Math. (2) 46 (1945), 685–694. MR14102
  • [6] Morton Brown, Locally flat imbeddings of topological manifolds, Ann. of Math. (2) 75 (1962), 331–341. MR133812
  • [7] by same author, A mapping theorem for untriangulated manifolds, Topology of 3-manifolds and related topics (Proc. The Univ. of Georgia Institute, 1961), Prentice-Hall, Englewood Cliffs, N.J., 1962, pp. 92–94. MR0158374
  • [8] Eduard Čech, Příspěvek k theorii dimense [contribution to the theory of dimension], Časopis Pěst. Mat. 62 (1933), 277–291.
  • [9] J. Cheeger and J. M. Kister, Counting topological manifolds, Topology 9 (1970), 149–151. MR256399
  • [10] Lei Chen and Kathryn Mann, Structure theorems for actions of homeomorphism groups, Duke Mathematical Journal 172 (2023), no. 5, 915 – 962.
  • [11] Charles O. Christenson and William L. Voxman, Aspects of topology, Pure and Applied Mathematics, Vol. 39, Marcel Dekker, Inc., New York-Basel, 1977. MR0487938
  • [12] Michel Coornaert, Topological dimension and dynamical systems, Universitext, Springer, Cham, 2015, Translated and revised from the 2005 French original. MR3242807
  • [13] Jean-Louis Duret, Sur la théorie élémentaire des corps de fonctions, J. Symbolic Logic 51 (1986), no. 4, 948–956. MR865921
  • [14] by same author, Équivalence élémentaire et isomorphisme des corps de courbe sur un corps algébriquement clos, J. Symbolic Logic 57 (1992), no. 3, 808–823. MR1187449
  • [15] Gerald Edgar, Measure, topology, and fractal geometry, second ed., Undergraduate Texts in Mathematics, Springer, New York, 2008. MR2356043
  • [16] Robert D. Edwards and Robion C. Kirby, Deformations of spaces of imbeddings, Ann. of Math. (2) 93 (1971), 63–88. MR283802
  • [17] A. Fathi, Structure of the group of homeomorphisms preserving a good measure on a compact manifold, Ann. Sci. École Norm. Sup. (4) 13 (1980), no. 1, 45–93. MR584082
  • [18] R. P. Filipkiewicz, Isomorphisms between diffeomorphism groups, Ergodic Theory Dynam. Systems 2 (1982), no. 2, 159–171 (1983). MR693972
  • [19] Casper Goffman and George Pedrick, A proof of the homeomorphism of Lebesgue-Stieltjes measure with Lebesgue measure, Proc. Amer. Math. Soc. 52 (1975), 196–198. MR376995
  • [20] Be’eri Greenfeld, First-order rigidity of rings satisfying polynomial identities, Ann. Pure Appl. Logic 173 (2022), no. 6, Paper No. 103109, 8. MR4393207
  • [21] Witold Hurewicz and Henry Wallman, Dimension Theory, Princeton Mathematical Series, vol. 4, Princeton University Press, Princeton, N. J., 1941. MR0006493
  • [22] Olga Kharlampovich and Alexei Myasnikov, Elementary theory of free non-abelian groups, J. Algebra 302 (2006), no. 2, 451–552. MR2293770
  • [23] S.-h. Kim and T. Koberda, Diffeomorphism groups of critical regularity, Invent. Math. 221 (2020), no. 2, 421–501. MR4121156
  • [24] Sang-hyun Kim and Thomas Koberda, Structure and regularity of group actions on one-manifolds, Springer Monographs in Mathematics, Springer, Cham, [2021] ©2021. MR4381312
  • [25] Robion C. Kirby, Stable homeomorphisms and the annulus conjecture, Ann. of Math. (2) 89 (1969), 575–582. MR242165
  • [26] Thomas Koberda and J. de la Nuez González, Uniform first order interpretation of the second order theory of countable groups of homeomorphisms, 2023.
  • [27] by same author, Locally approximating groups of homeomorphisms of manifolds, 2024.
  • [28] Frédéric Le Roux, On closed subgroups of the group of homeomorphisms of a manifold, J. Éc. polytech. Math. 1 (2014), 147–159. MR3322786
  • [29] Gérard Leloup, Équivalence élémentaire entre anneaux de monoïdes, Séminaire Général de Logique 1983–1984 (Paris, 1983–1984), Publ. Math. Univ. Paris VII, vol. 27, Univ. Paris VII, Paris, 1986, pp. 149–157. MR938759
  • [30] Kathryn Mann and Maxime Wolff, Reconstructing maps out of groups, arXiv preprint (2019), 1–16, Prepint, arXiv:1907.03024.
  • [31] David Marker, Model theory, Graduate Texts in Mathematics, vol. 217, Springer-Verlag, New York, 2002, An introduction. MR1924282
  • [32] Francis Oger, Équivalence élémentaire entre groupes finis-par-abéliens de type fini, Comment. Math. Helv. 57 (1982), no. 3, 469–480. MR689074
  • [33] Phillip A. Ostrand, Covering dimension in general spaces, General Topology and Appl. 1 (1971), no. 3, 209–221. MR288741
  • [34] J. C. Oxtoby and S. M. Ulam, Measure-preserving homeomorphisms and metrical transitivity, Ann. of Math. (2) 42 (1941), 874–920. MR5803
  • [35] Eugene Plotkin, On first order rigidity for linear groups, Toyama Math. J. 41 (2020), 45–60. MR4353090
  • [36] Florian Pop, Elementary equivalence versus isomorphism, Invent. Math. 150 (2002), no. 2, 385–408. MR1933588
  • [37] MathOverflow Post, Does a *topological* manifold have an exhaustion by compact submanifolds with boundary?, https://mathoverflow.net/questions/129321.
  • [38] by same author, On a result by rubin on elementary equivalence of homeomorphism groups and homeomorphisms of the underlying spaces, https://mathoverflow.net/questions/400602.
  • [39] Frank Quinn, Ends of maps. III. Dimensions 44 and 55, J. Differential Geometry 17 (1982), no. 3, 503–521. MR679069
  • [40] by same author, Topological transversality holds in all dimensions, Bull. Amer. Math. Soc. (N.S.) 18 (1988), no. 2, 145–148. MR929089
  • [41] Matatyahu Rubin, On the reconstruction of topological spaces from their groups of homeomorphisms, Trans. Amer. Math. Soc. 312 (1989), no. 2, 487–538. MR988881
  • [42] by same author, Locally moving groups and reconstruction problems, Ordered groups and infinite permutation groups, Math. Appl., vol. 354, Kluwer Acad. Publ., Dordrecht, 1996, pp. 121–157. MR1486199
  • [43] Dan Segal and Katrin Tent, Defining RR and G(R)G(R), J. Eur. Math. Soc. (JEMS) 25 (2023), no. 8, 3325–3358. MR4612113
  • [44] Zlil Sela, Diophantine geometry over groups x: The elementary theory of free products of groups, (2010).
  • [45] Floris Takens, Characterization of a differentiable structure by its group of diffeomorphisms, Bol. Soc. Brasil. Mat. 10 (1979), no. 1, 17–25. MR552032
  • [46] Katrin Tent and Martin Ziegler, A course in model theory, Lecture Notes in Logic, vol. 40, Association for Symbolic Logic, La Jolla, CA; Cambridge University Press, Cambridge, 2012. MR2908005
  • [47] Xavier Vidaux, Équivalence élémentaire de corps elliptiques, C. R. Acad. Sci. Paris Sér. I Math. 330 (2000), no. 1, 1–4. MR1741158
  • [48] John von Neumann, Collected works. Vol. II: Operators, ergodic theory and almost periodic functions in a group, Pergamon Press, New York-Oxford-London-Paris, 1961, General editor: A. H. Taub. MR0157872
  • [49] James V. Whittaker, On isomorphic groups and homeomorphic spaces, Ann. of Math. (2) 78 (1963), 74–91. MR150750