First order rigidity of homeomorphism groups of manifolds
Abstract.
For every compact, connected manifold , we prove the existence of a sentence in the language of groups such that the homeomorphism group of another compact manifold satisfies if and only if is homeomorphic to . 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 equivalence2020 Mathematics Subject Classification:
Primary: 20A15, 57S05, ; Secondary: 03C07, 57S25, 57M601. 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 (possibly equipped with a probability measure ), we let and denote the homeomorphism group of and its –preserving subgroup, respectively. We denote by and the identity components of and , respectively. For general topological spaces and , we write if and are homeomorphic.
We denote by the class of all pairs , where is a compact, connected manifold and is a group satisfying
We also let denote the class of all where is further assumed to be equipped with some Oxtoby–Ulam measure (that is, a nonatomic Borel probability measure having full support and assigning measure zero to the boundary), and is a group satisfying
Note that in this case, we have
cf. [17].
Remark 1.1.
In statements that apply to both of the classes and , we will often use the notation ; 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 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 in some category, it is natural to ask the degree to which the group of automorphisms determines the object . This question is not completely precise, since the terms “degree” and “determine” do not have a mathematical meaning here. In our context, the object 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 , the extent to which the first order theory of can be used to reconstruct the homeomorphism type of .
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 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 and be compact manifolds, and suppose
is an isomorphism of groups. Then there exists a homeomorphism
such that for all , we have .
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 and are smooth and closed, and if the diffeomorphism groups and are isomorphic as groups, then and each isomorphism between the groups is induced by some –diffeomorphism between and .
In the continuous category, a different generalization was given by Rubin. We say that a topological action of a group on a topological space is locally dense if for each pair of a point and a neighborhood of , the orbit of by the action of the group
is somewhere dense; that is, the closure of has nonempty interior. Rubin’s Theorem can be stated as followsL
Theorem 1.3 ([41]).
Let and be perfect, locally compact, Hausdorff topological spaces, and let be locally dense subgroups for . If there exists an isomorphism if groups
then there exists a homeomorphism
such that for all , we have .
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 and are elementarily equivalent, written , if a first order sentence in the language of groups holds in if and only if it holds in ; this is sometimes expressed as saying that the theories of and agree, i.e.
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 we prove the existence of a group theoretic sentence that asserts “I am homeomorphic to ”:
Theorem 1.4.
For each compact, connected manifold , there exists a sentence in the language of groups such that when , we have that
In other words, the theories of homeomorphism groups of manifolds are quasi-finitely axiomatizable within the class , a property that is stronger than first order rigidity.
In Theorem 1.4, we emphasize that and 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 and 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 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 be a compact, connected manifold. Under what conditions is there a finitely generated (or countable) group such that whenever is a compact manifold with on which acts faithfully with a dense orbit, we have ?
Related results for actions of the full homeomorphism group of are given by Chen–Mann [10]. They show that if the identity component of acts transitively on a connected manifold or CW–complex , then is homeomorphic to a cover of a configuration space of points of . In our context, we have the following immediate consequence of the downward Löwenheim–Skolem Theorem:
Corollary 1.6.
To each compact connected manifold one can associate a countable group which is elementarily equivalent to , such that for two compact, connected manifolds and we have
In particular if and only if .
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 (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 . 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 remains unclear, and perhaps the goal was to promote first order equivalence to second order equivalence, using the assumption 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 . The universe of an structure corresponding to will contain the group , the regular open sets of , the real numbers , the set of continuous maps for
and the discrete subsets of . Since the expansion is specified by first order definitions, we deduce the following, which roughly means that each sentence in the theory of can be interpreted (in a way that is uniform in ) as a sentence in the theory of the group ; see Section 2 for a precise definition of uniform interpretation:
Theorem 1.8.
For , the structure is uniformly interpretable in the group structure .
The second part of the proof consists in showing that the language can express the sentence that “I am homeomorphic to ”:
Theorem 1.9.
For each , there exists an –sentence such that for all , we have
By Theorem 1.8, we can interpret –sentences
as group theoretic sentences and respectively, which distinguish from all the other non-homeomorphic manifolds ; 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 ; 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 is expressive enough to interpret the full second order theory of countable subsets of , 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 . In this structure, we interpret the regular open sets in , 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 and are Oxtoby–Ulam measures on a compact connected manifold , then there exists a homeomorphism of isotopic to the identity and fixing such that .
Thus, for Oxtoby–Ulam measures, the groups of measure-preserving homeomorphisms of are all conjugate to each other. In particular, each 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, means the compact ball of radius with the center at the origin in .
Lemma 2.2 ([7, 17]; cf. [11, Chapter 17]).
For each compact connected –manifold , there exists a continuous surjection
such that the following hold:
-
(i)
the restriction of on is an embedding onto an open dense subset of ;
-
(ii)
we have , and in particular, ;
If is equipped with an Oxtoby–Ulam measure , we can further require the following:
-
(iii)
we have ;
-
(iv)
the measure is the pushforward of Lebesgue measure by .
The conditions (i) and (ii) already imply that an Oxtoby–Ulam measure on exists. For instance, one can pull back the Lebesgue measure on a ball using the surjection
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 –manifold has nonempty boundary and if is equipped with a nonatomic, fully supported Radon measure that assigns zero measure to , then there exists an open embedding
such that the following hold:
-
(i)
and ;
-
(ii)
is closed and of measure zero.
We call a topologically embedded image of in a manifold a ball. The same goes for an open ball in . If there exists an embedding
then the image is called a collared ball [11, Chapter 17]. The same goes for a collared open ball. In the case when is equipped with an Oxtoby–Ulam measure , we say a collared ball is –good (or, simply good) if has measure zero. There exists an arbitrarily small covering of by –good balls [17]. For brevity of exposition, by a good ball, we mean both a collared ball in the context of and a –good ball in the context of . The same goes for a good open ball. Note that a good ball is always contained in the interior of .
Recall the topological action of a group on is path–transitive if for all paths
and for all neighborhoods of there exists such that . We say the action of on is –transitive if it induces a transitive action on the configuration space of distinct points in . A path-transitive action on is always –transitive whenever ; 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 with , we have the following.
-
(1)
The action of on is path–transitive and –transitive for all .
-
(2)
If and are good balls of the same measure in an open connected set , then there exists such that .
Proof.
The path–transitivity of part (1) is well-known; see [3, Section 7.7] for , and [17, p. 85] for . The –transitivity follows immediately.
The case when 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 by a sequence of compact bounded manifolds so that some contains and in its interior; this can be seen from [40], as explained in [37]. We can further require that has measure zero by countable additivity. Applying Le Roux’s argument for , we obtain the desired transitivity.∎
Lemma 2.5.
Let be a compact, connected –manifold with , equipped with an Oxtoby–Ulam measure . If is an open connected subset, then for each positive real number , there exists a good ball of measure precisely inside . Moreover, we may require that is connected.
Proof.
Note the general fact that for a connected open subset of and for a collared ball in , the set 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 such that the connected –manifold
has measure larger than and has nonempty boundary. Applying Lemma 2.3 to , we have an open embedding
such that
Since is a countable increasing union of collared balls, we can find a collared ball in having measure larger than ; moreover, we can further require that is good by countable additivity of . Applying Theorem 2.1 to , we see that the restriction of is conjugate to a Lebesgue measure on a cube. It is then trivial to find a good ball with measure precisely . ∎
2.2. Regular open sets and homeomorphism groups
Let be a topological space. If is a subset then we write and for its closure and interior, respectively, and
for the frontier of .
A set is regular open if . For instance, a good ball is always regular open. The set of regular open subsets of forms a Boolean algebra, denoted as . In this Boolean structure, the minimal and maximal elements are the empty set and respectively. The meet is the intersection, and the join of two regular open sets and is given by
We write
when is the disjoint union of two sets and .
The complement coincides with the exterior:
Consequently, the Boolean partial order coincides with the inclusion for . For each subcollection of regular open sets we can define its supremum as
In particular, 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 .
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 is an open cover of a normal space, then there exists a regular open cover such that for each .
Proof.
Under the given hypothesis, one can find an open cover satisfying for each ; see [12, Corollary 1.6.4]. It then suffices for us to take , which is clearly a regular open set. ∎
Let . We denote its fixed point set by , and define its (open) support as . We then define its extended support as
Let . We define the rigid stabilizer (group) of as
If is regular open in , we note that
Recall from the introduction that the group is locally dense if for each nonempty open set and for each we have hat
More weakly, we say is locally moving if the rigid stabilizer of each nonempty open set is nontrivial.
If is a locally moving group of homeomorphisms of then has no atoms, and the set of extended supports
is dense in the complete Boolean algebra , i.e. for all there exists such that ; 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 with , or that with . Then each regular open set of is the extended support of some element of .
Proof.
Pick a countable dense subset of . Set , and pick a good ball containing such that and such that . Suppose we have constructed a sequence
and a disjoint collection of good balls such that and such that
for each ; furthermore, we require that
If
then we terminate the procedure; otherwise, we let be the minimal index such that
Pick a good ball containing such that
Thus, we build an infinite disjoint collection of good balls in such that
We claim that there exists for each such that . 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 is part of the data, we first pick a homeomorphism in whose fixed point set has empty interior; here, the condition that is used. Let us also pick a homeomorphism
We see from Theorem 2.1 that the pullback measure of on under the map is conjugate to (a rescaling of) the Lebesgue measure by a homeomorphism. Hence, by conjugation and extension by the identity, we obtain a homeomorphism satisfying
for some closed set with empty interior. This proves the claim.
Since we have
for all , we see from the uniform convergence theorem that the infinite product converges in , and is isotopic to the identity. By definition,
Hence, this map 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 , there exist a group theoretic formula such that when , we have that
if and only if and are homeomorphic.
Proof.
Since , the group theoretic sentence stating that there are at most two elements in the group is satisfied by a pair if and only if . Since contains the abelian group as the index–two subgroup, a pair of the form satisfies the sentence
Finally, if with , then is not virtually abelian and hence does not satisfy the above formulae. ∎
2.3. First order logic
Proposition 2.8 establishes the measure-preserving case of the main theorem with . 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 is specified by logical symbols and a signature. Logical symbols include quantifiers (, ), logical connectives (, , , ), 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) –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 is a relation symbol with the arity value for some sort symbols and , and if and are variables with sort values and , respectively, then is a formula. We write instead of for the ease of reading. The language 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 –structure (or, an –model) , which is specified by a set called the universe, a sort function from to the set of sort symbols, and an assignment that is a correspondence from each relation symbol to an actual relation among tuples of the elements in the universe. For each sort symbol , we call the domain of in . It is required that the relation respects the arity value of . For instance, if is as in the previous paragraph, then will be a subset of . A function symbol is assigned the graph of some function, and often written as a function notation such as . A constant symbol is fixed as an element in the universe by an assignment. An assignment (for relations) naturally extends to an assignment for each formula . We sometimes omit from when the meaning is clear.
For an –formula with a tuple of free variables , and for a tuple of elements in , we write if holds after has been substituted for . We define as the set of all –sentences such that .
Let , and let be a –tuple of elements of . A subset of is definable (by ) with parameters if for some formula with free variables, the set coincides with the set
If we simply say is definable, in which case we denote the above set as . We now formalize the concept of “interpreting” a new language.
Definition 2.9.
Let and be languages. Suppose we have a class of ordered pairs in the form with being an –structure. We say is interpretable in uniformly for in if there exist some –formulae and , and there also exists a map from the set of –formulae to the set of –formulae such that the following hold.
for each , we have a surjection
with its fiber uniformly defined by in the sense that
Furthermore, it is required for each –formula that
The bijection
along with the map is called a uniform interpretation of in .
Remark 2.10.
-
(1)
In the above, if is –ary (as a relation) and is –ary, then is –ary. In practice, we only need to consider relation symbols (in a broad sense, including function and constant symbols) rather than all possible –formulae.
-
(2)
In various instances of this paper, it will be the case that and that the interpretation restricts to the identity on . As a consequence of such interpretability, we will have that is a conservative extension of for each . Also, we will often add a function symbol in corresponding to the surjection , which is clearly justified.
Lemma 2.11.
Suppose and are as in Definition 2.9 so that is interpretable in uniformly for . Let and be in . Then for each sentence belonging to , the interpretation belongs to . In particular, if , then .
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 and the structure
The ultimate language we will work in will be called , which stands for “Action of a Group, Analysis, Points and Exponentiation”. This language is denoted as and contains the following different sort symbols for :
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 structure assigned to each pair in the class or . In this structure, we give the “intended” choice of the domain of each sort symbol.
The group sort. The domain of the sort symbol will be the group , under our standing assumption that . The signatures only relevant for this sort are
which are respectively assigned with the natural meanings in the group theory. These symbols, along with variables, form the language of groups . The group is regarded an -structure . We will usually not write the symbol.
The sort of regular open sets. The domain of the sort symbol is the set of the regular open sets in . The newly introduced signatures for this sort are
The symbol means the manifold in the structure. By the natural assignment as before, we have Boolean symbols
for the Boolean algebra . We let the function symbol mean the map defined as
We have an assignment for so that
with and . The symbols introduce so far (along with countably many variables for each sort) form the language of a group action on a Boolean algebra . The –structure described above on the universe is denoted as .
The sorts from the analysis We then introduce new sort symbols, which are and for . The signatures introduced here are
Standard second order arithmetic
is given the sort symbols and , as well as with relevant non-logical symbols. We note the ambiguity of our notation that the sort symbols and will be assigned with the set of the natural numbers and its power set , respectively. The symbol is interpreted so that
means has connected components. See Section 4 for details. The ordered ring of the real numbers
is assigned with the sort symbol and the signatures above. Note that, as is usual, is considered as a subsort of , by identifying each integer as a real number.
The domain of the sort symbol will be the set of continuous functions. We also have a formula when the sort value of is , and when and are tuples of variables assigned with the sort symbol . We have the –norm
which will be also a part of the language. Combining these symbols with , we obtain the language . An –structure is assigned to each having the universe
The point and the discrete subset sorts and . The domain of the sort symbol will be the set of the points in a manifold. We also introduce the sort symbol to mean a subset of every point of which is isolated in . By abuse of notation, the symbols and introduced above will have multiple meanings (depending on the context), so that they have the arity values , and .
We also have a cardinality function
meaning that the cardinality of is , assuming that every point in is isolated.
The interpretation of points of the manifold will allow us to include symbols such as and , 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 from the interior ones, and hence justified to use the notations
for point sort variables and . The function symbol has a natural additional meaning as below:
In all contexts, we abbreviate by when the sort of is either or and when the sort of is (tuples of) , or .
The omnibus language, combining all of the previous sorts and relevant symbols, is denoted by
or simply as . We have so far described the –structure corresponding to .
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 , and let be a formula in . Suppose for each that a function or relation is defined by in . Then the collection
is said to be uniformly defined over .
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 | , , , | , |
Regular open sets | , , , , , | , , |
Natural numbers | , , , | |
Sets of natural numbers | , | |
Real numbers | , , , | , |
Sets of points | , , , | , , |
Functions | , , , |
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 to denote indices. The symbols and 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 , and uniformly for , the –structure is interpretable in the –structure .
The proof of this proposition will require the construction of –formulae and , and a surjection
for all 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 or the space .
Theorem 3.4 (Rubin’s Expressibility Theorem, cf. [42]).
There exist first order formulae
in the language of groups such that if be a locally moving group of homeomorphisms of a Hausdorff topological space , then the following hold for all .
-
(1)
.
-
(2)
.
-
(3)
.
-
(4)
.
-
(5)
.
Proof.
Let . By Proposition 2.7, we have a surjection
defined as . Since is locally moving on , Rubin’s expressibility theorem implies that the fiber
of is definable, and that the Boolean symbols and the function symbols and 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 .
Corollary 3.5.
Uniformly for , the –structure is interpretable in the group structure .
Corollary 3.5 can be summarized as saying that interprets the group action structure of on the algebra of regular open sets, in a way that preserves the meaning of . This interpretation is uniform in the underlying pair , and any formula in the language of and can be expressed entirely in , since the formulae in Theorem 3.4 are independent of . Henceforth, we will assume that we work in the expanded language .
3.3. First order descriptions of basic topological properties
Recall that whenever the expression is used it is assumed that and are disjoint.
We now produce first order expressions for some standard point-set–topological properties.
Lemma 3.6.
The following hold for .
-
(1)
For , we have that if and only if .
-
(2)
For each , we have that
-
(3)
An open subset is path-connected if and only if it is connected.
-
(4)
An arbitrary union of connected components of a regular open set is necessarily regular open. More specifically, if a regular open set can be written as for some disjoint pair of open sets and , then and are regular open and . Moreover, we have .
-
(5)
For disjoint pair of regular open sets, we have (i)(ii)(iii).
-
(i)
is connected, and ;
-
(ii)
Every satisfies either or ;
-
(iii)
-
(i)
-
(6)
Let and are regular open sets such that is connected and such that . Then is a connected component of if and only if for some regular open that is disjoint from , and every satisfies either or .
-
(7)
The following are all equivalent for a regular open set .
-
(i)
is disconnected;
-
(ii)
for some disjoint pair of nonempty regular open sets and such that is connected;
-
(iii)
for some disjoint pair of nonempty regular open sets and , and every satisfies either or ;
-
(iv)
for some disjoint pair of nonempty regular open sets and .
-
(i)
-
(8)
For two regular open subsets and satisfying , we have that if and only if each connected component of is contained either in or in .
Proof.
(1) If , then there exists some satisfying ; see [24, Lemma 3.2.3] for instance. In particular, we have
This proves the nontrivial part of the given implication. We remark that the same statement holds without the assumption that and are regular open, under the extra hypothesis that . Part (2) is similar.
(3) This part is clear from the fact that every manifold is locally path-connected.
(4) Whenever two open sets and are disjoint we have that and are also disjoint; see [24, Lemma 3.6.4 (4)], for instance. From
we see that and are actually regular open and . It is clear that , which implies .
(5) The implication (i)(ii) is clear from that every setwise stabilizer of permutes connected components of .
For the implication (ii)(iii), assume we have a point
Take a sufficiently small open ball around so that
Note also that because
it follows that . Similarly, . This implies that we can choose distinct points
and . Since is –transitive on for all , we can find a supported in satisfying and ; see also Lemma 2.4. Then is neither nor disjoint from .
(8) The forward direction comes from the observation that is a disconnection of . The backward direction is trivial since the hypothesis implies that . ∎
Let us note the following consequences of Lemma 3.6.
Corollary 3.7.
There exist first order formulae in the language as follows:
-
(1)
A formula , also abbreviated as such that
-
(2)
A formula such that
-
(3)
A formula , also abbreviated as such that
-
(4)
A formula such that
-
(5)
For all , a formula such that
-
(6)
A formula such that
-
(7)
A formula such that
-
(8)
A formula such that for all regular open sets and having finitely many connected components, we have
if and only if and have the same number of connected components.
Proof.
The existence of the formula is trivial since and belong to the signature of . The formulae and exist by parts (6) and (7) of Lemma 3.6. We can then set
The construction of the formulae and follows from the same lemma, which also implies that the formula
has the meaning required in part (7). Finally, we set
From the transitivity on good balls (of equal measure, in the measure preserving case) as in Lemma 2.4, we see has the intended meaning. ∎
Using the above formula, we can distinguish the case that among all compact connected manifolds.
Corollary 3.8.
For each compact connected one–manifold , there exist –formulae such that when , we have that
if and only if and are homeomorphic.
Proof.
We let be the –formula expressing that for all pairwise disjoint, proper, nonempty regular open sets and the exterior of is disconnected for some . This formula holds for since at least one of does not intersect , and hence separates the two endpoints of . It is clear that is never satisfied by other compact connected manifolds.
We now suppose that is a compact connected manifold not homeomorphic to . Then for all disjoint, proper, non-empty regular open sets and satisfying , the set is disconnected. From Corollary 3.7, we obtain the formula expressing that and are homeomorphic. ∎
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 is contained in a collared ball inside of another regular open set . This is not particularly difficult to state and prove in the class , but is substantially harder in . For the rest of this section, we will make a standing assumption that , and that the underlying structure is .
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 is a connected, compact –manifold with , equipped with an Oxtoby–Ulam measure .
Lemma 3.9.
There exists a formula in the language such that for all with an Oxtoby–Ulam measure on , and for all any triple with and connected and , we have the following:
-
(1)
If then holds.
-
(2)
If holds then .
Proof.
Suppose first that , and let be arbitrary. By Lemma 2.5, we can find a good ball such that
and such that connected. Lemma 2.4 furnishes such that , but clearly there is no -preserving such that
We have just established that holds with , where
Let us now suppose for a contradiction that holds but that
Let be the interior of a good ball in with measure . It suffices to show that there is no witness as required by .
The foregoing discussion allows us to characterize when a regular open set is contained in a collared ball inside a regular open set . 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 such that for each , we have that
if and only if is relatively compact in some good ball contained in .
Recall our convention that this lemma actually claims to produce two formulae, namely and .
Proof of Lemma 3.10.
Let us consider the formula , which expresses that there exists some component of satisfying the following two conditions:
-
•
contains ;
-
•
for each nonempty, regular open set contained in , there exists some element that moves into .
We first claim that this formula satisfies the conclusion for . Indeed, if is relatively compact in a collared ball , then there exists a unique containing , and hence . For each nonempty regular open , we see from Lemma 2.4 that some satisfies
as desired. Conversely, suppose holds and let be the connected component of containing . Let us fix a collared ball in and set . By assumption, we can find such that . Then is relatively compact in the collared ball in .
For the case when , we set
In order to prove the forward direction, assume that holds for some nonempty . Let and be witnesses for the existentially quantified variables and . Since , the Boolean subtraction is nonempty. We now see that
So, Lemma 2.5 furnishes a good ball satisfying
By Lemma 3.9, we have that , and that some satisfies that . It follows that
as desired.
For the backward direction, we pick a good ball satisfying for a suitable and set . Consider an arbitrary connected regular open set satisfying . From Lemma 3.9 again, we see that
We may therefore find some such that . This shows that holds. ∎
When using Lemma 3.10, we will write both in the case of the full homeomorphism group and the measure-preserving homeomorphism group, suppressing the symbol from the notation.
Many of the formulae below will actually have different meanings for and for , though sometimes coincide in their implications; we record the fact that implies that .
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 connected component in the theory of for each fixed . It is not obvious a priori how to express the infinitude of the connected components of , as such an infinitude would be equivalent to the infinite conjunction
However, one can express such an infinitude in a single formula.
Definition 3.11.
Let us set
We say a regular open set is dispersed if holds.
Note that implies that
for each connected component of . 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 –formula such that
for and if and only if the following conditions hold for a unique :
-
(i)
the set is dispersed;
-
(ii)
we have that ;
-
(iii)
for all , the set is nonempty and connected;
-
(iv)
for all , the set is nonempty and connected;
-
(v)
we have that ;
-
(vi)
if a union of connected components of satisfies that and that , then .
In a situation as in Lemma 3.12, we can enumerate the components of as
so that for each . Furthermore, we have an injection
sending to for each . We also note that for each there exists a uniformly definable function such that
We can now establish the main result of this subsection.
Lemma 3.13.
There exists a formula such that
Proof.
Let us define
In order to prove the forward direction, suppose we have for some nonempty , such that each connected component of satisfies . In particular, we have . The injection above certifies that is an infinite set. Hence, is infinite as well.
For the backward direction, suppose that has infinitely many components. We will establish only in the case of , since the case is strictly easier. We use an idea similar to the proof of Proposition 2.7. We first find distinct components of such that some sequence satisfying converges to some point . We consider a sufficiently small compact chart neighborhood of , which still intersects infinitely many components of . Let . By the Oxtoby–Ulam theorem, we can simply identify with or equipped with the Lebesgue measure. The point is then identified with the origin . By shrinking each to and passing to a subsequence, we can further require the following for all .
-
•
The open set is an open Euclidean ball, converging to ;
-
•
We have .
We set
and . We can find a disjoint collection of compact topological balls such that intersects both and , and no other ’s. Using the path–transitivity as in Lemma 2.4, we can inductively find a
sending some good ball onto another good ball inside . We will set
By the uniform convergence theorem, the sequence
converges to a homeomorphism , which witnesses the properties that the formula requires. ∎
It follows immediately that we may also test whether a regular open set has finitely many components, and write
3.4.3. Touching and containing the boundary
By a collar (embedding) of the boundary in a manifold , we mean an embedding
that extends the identity map
we sometimes allow to be an embedding of . 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 –formulae as follows:
-
(1)
A formula such that
-
(2)
A formula such that
Proof.
(1) Let us define the formula
It is clear from the formulation that
if and only if meets in finitely many components on . We now set
Suppose that . Choose a sequence of points in converging to a point in , and choose small open balls in with pairwise disjoint closures and with radii tending to zero. Let be the union of these balls. Now, if fails to satisfy , then must meet infinitely many of the balls ; thus . In particular, .
Conversely, suppose that , and let have infinitely many components . As in Lemma 3.13, by shrinking components of and passing to a subsequence, we may assume that each is an open ball, that the sequence has shrinking radii, and converges monotonically to the origin in an open chart in . Moreover, the origin in this chart lies in the interior of , by assumption.
We may take to be a neighborhood of the origin in this chart, which then satisfies and meets infinitely many components of . Thus, fails to witness , and so does not hold.
We claim that holds for if and only if setwise stabilizes each component of
For the forward direction, suppose we have . By the aforementioned result of Brown, we can pick a closure–disjoint collection of collar neighborhoods of the components of . Defining
we see from the hypothesis that for each , which trivially implies . The backward direction is clear after observing that the hypothesis of simply says that contains at least one boundary component.∎
4. Interpretation of second-order arithmetic
The goal of this section is to prove that the group interprets second order arithmetic and analysis uniformly for , establishing the case of 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
in the structure . For this, we consider the surjection
sending each to , namely the cardinality of . The domain of this surjection is clearly definable, and so is the fiber by the formula in Corollary 3.7. To complete an interpretation of , it suffices to establish the following:
Lemma 4.1.
There exist –formulae and such that the following hold for all having finitely many connected components.
-
(1)
We have .
-
(2)
We have .
Proof.
Recall the meaning of the formula from Corollary 3.7. Let us make the following definitions.
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
which has two sorts, namely and . In particular, we will have to be able to quantify over subsets of .
In order to achieve this, we will consider more restricted class of regular open sets , the components of which admit a linear order as described by the formula ; see Section 3.4.2. In this linear order of , the –th component will interpret the integer , and a union of the connected components 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 such that if
then for the unique satisfying , we have that
Proof.
It is routine to check that the following has the intended meaning:
Let us consider the set
which is definable in uniformly for . We have a surjection
This surjection satisfies
The fiber of is
and hence uniformly definable. It is trivial to check that produces a uniform interpretation of to . For instance, we have
if and only if
After this interpretation of , the symbol has an intended meaning as a function from to . We have uniformly defined functions and satisfying
Similarly, we consider another uniformly definable set
We have a surjection
defined by the condition
Since the fiber of is uniformly definable, so is that of . We will introduce the function symbol in interpreted as .
Finally, we have
if and only if
for some . Hence, the pair of surjections produces the desired interpretation of the two–sorted structure . We note that the order relation symbol , the successor symbol , and the inclusion symbol are naturally interpreted as a consequence.
4.3. Analysis
The interpretation of is now standard. From , we interpret , together with addition, multiplication, and order, by imposing a suitable definable equivalence relation on a suitable definable subset of . We similarly interpret by imposing a suitable definable equivalence relation on a definable subset of .
We define together with addition, multiplication, and order via Dedekind cuts of ; all this is interpretable because of our access to . Finally, we have canonical identifications of
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 without further comment. We further may assume to have in the universe of the structure for all .
In order to justify the introduction of the sort symbol in the structure, let us first note that each function in is uniquely determined by its restriction on . Since
we have an interpretation of by , and hence, that of
This latter set is the domain of , and the function symbols
are interpreted accordingly. In practice, we write
for the above formulae. The expanded language containing structure, second order arithmetic, and analysis will be written . This establishes the uniform interpretability of to , namely Proposition 3.3 for the case .
5. Interpretation of points
We now wish to be able to talk about points of more directly, and prove Proposition 3.3 for the case . 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 with , unless stated otherwise.
5.1. Encoding points of a manifold
Using the –formulae introduced in the preceding sections, we define the following new formulae:
Note that when , we can find some whose connected components can be written as
with the property that each is contained in some relatively compact ball inside ; moreover, no two components of belong to the same component of , and similarly for and .
We consider the definable set
which we call as the limit stabilizer of in . Intuitively, each element of this set fixes some open set that is arbitrarily close to a certain limit point of the components of . We will write for the formula corresponding to .
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 and interpret different points and if and only if the group
acts sufficiently transitively, in the sense that for some , every satisfying is an element of the set
In our approach, we will utilize the sufficient transitivity of the limit stabilizer characterized in terms of the formula .
Consider the set , defined by the following formula:
The following lemma furnishes an interpretation of the points.
Lemma 5.2.
For each and for an arbitrary sequence satisfying
for all , the limit
exists in , and is independent of the choice of . Moreover, the following conclusions hold:
-
(1)
The map is surjective.
-
(2)
We have
if and only if some satisfies
-
(3)
We have
if and only if
-
(4)
We have if and only if some satisfies
-
(5)
We have if and only if there exists some such that and such that .
Proof.
Let , and let
be a sequence. In particular, we have . Suppose two subsequences
converge to two distinct points and . For and , we let be the union of sufficiently small good open balls centered at . In particular, we may assume that , and that
in the Hausdorff sense. By hypothesis, we have some such that
Since fixes points arbitrarily close to , we have . It follows that
This proves the existence of the claimed limit. The same argument also implies the independence of the limit from the choice of , and also the backward direction of part (2). The surjectivity of in part (1) is clear, after choosing 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 and such that
As in the proof of Lemma 3.13, we can find a disjoint collection of good balls of decreasing sizes such that each contains and , after passing to a subsequence if necessary. By the uniform convergence theorem, we have some such that for all , and such that pointwise fixes some nonempty open set inside
In particular, we have that and that , as claimed. The remaining parts of the lemma are straightforward to check. ∎
In part (2) of the lemma, we see that the relation
is first order expressible; hence, we deduce that the functional relation and the membership relation in parts (3) and (4) are interpretable for , and . 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 by
Clearly, for regular open sets if and only if . Henceforth, we will include the usual set-theoretic union symbol in the language such as and . We are also able now to talk directly about the closure of a regular open set , both in and in for arbitrary ; for this, it suffices to note that if and only if .
5.2. Encoding discrete sets of points in a manifold
We now interpret the set
In particular, every finite subset of belongs to .
We recall from Lemma 3.14 the formula . We first let be the set of quadruples defined by the following formula:
For such a quadruple, we set
It is routine to check that this map defines a surjection
with a definable fiber. Namely, we have
if and only if there exists some regular open sets satisfying that
and that
for some .
We interpret the membership between a point and a set; namely, we have
if and only if there exists some satisfying and
We also interpret the group action
as
Finally, the set
has finite cardinality if and only if has finitely many connected components. In this case, the cardinality function for is clearly definable by
We omit the details, which are very similar to those in Section 5.1. We denote by the sort symbol for sets belong to .
5.3. Interpreting exponentiation
We now interpret the map
so that the exponentiation map
is definable. Note that holds with if and only if we can write for some integers and such that we have a period– orbit
and a sequence of distinct points
Let us now define formulae and , which will express the existences of a periodic orbit and of a sequence without repetitions, respectively. More precisely, we set
We see that with if and only if the tuple satisfies the formula
It is then trivial to extend the definition for the case , establishing the definability of the exponentiation function.
5.4. The structure
We now define our ultimate structure
as the extension of by including the points in and adding the relations
for , and . We are then justified to use expressions such as
for points , regular open sets , group elements and integer within .
6. Balls with definable parametrizations
From this point on, we work in the language , containing second order arithmetic and points. The underlying structure will be ; recall our further standing assumption that . We will use the notation and
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 , there exists a formula such that if and only if is an –manifold.
Theorem 6.2.
For each , there exist formulae
such that the following hold for all with .
-
(1)
Let , and . If
then there exists a unique homeomorphism
the graph of which satisfies
and also .
-
(2)
Let and be good open balls inside such that ; if , we further assume that is sufficiently small compared to some positive number determined by . Then we have
In Section 8, we will modify the definition of so that the domain is , instead of . We emphasize again that the above formulae for and may differ; for instance, the abbreviated sentence could be more precisely denoted by and 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 , the order of a finite open cover is defined as the number
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 is at most , and write , if every finite open cover of is refined by an open cover with order at most . The topological dimension is defined to be , if holds but does not. A topological –manifold has the topological dimension .
A collection of open sets is said to shrink to another collection if holds for each in the index set . Let us note the following well-known facts.
Lemma 6.3.
-
(1)
(Lebesgue’s Covering Theorem [21, Theorem IV.2]) If is a finite open cover of such that no element of intersects an opposite pair of codimension one faces, then cannot be refined by an open cover of order at most .
-
(2)
(Čech [8]) If is a metrizable space and if , then .
-
(3)
(Ostrand’s Theorem [33, Theorem 3]) If is a locally finite open cover of a normal space satisfying , then for each , the cover shrinks to some pairwise disjoint collection of open sets such that the collection is a cover.
We can now give a characterization of manifold dimension.
Lemma 6.4.
For each positive integer and for each compact manifold , the following two conditions are equivalent.
-
(A)
The dimension of is at most ;
-
(B)
Let be a regular open set in . If
is a regular open cover of , then there exists a pairwise disjoint collection
of regular open sets for each such that shrinks to each , and such that is a cover of .
Proof.
Suppose we have , and assume the hypothesis of part (B). We see from Lemma 6.3 (2) that . Part (3) of the same lemma implies that shrinks to a pairwise disjoint collection of (not necessarily regular) open sets
for each with the property that is a cover of the normal space . By Lemma 2.6, there exists a regular open cover
of satisfying
for all and . This implies the conclusion of (B).
Conversely, suppose we have condition (B) and assume for contradiction that . We first note the following:
Claim.
The unit –cube admits a finite regular open cover of cardinality that cannot be refined by another open cover with order at most .
Let denote the unit cube in , which is embedded in as the subset with the last coordinates being zero. For each vertex , let us consider the translated open cube
We then have a regular open cover
of with cardinality . Note that each open cube does not intersect an opposite pair of codimension one faces of . By taking the Cartesian product of each with , we obtain a finite regular open cover
of . If is refined by another finite open cover of with order at most , then the intersection of the elements in with is a finite open cover of with order at most . This violates Lebesgue’s Covering Theorem (Lemma 6.3), and the claim is thus proved.
Let us now consider a good ball in , which comes with an embedding
satisfying . By applying the above claim, we obtain a finite regular open cover of that cannot be refined by a finite open cover with order at most . This contradicts condition (B), which we have assumed. ∎
Note that the cardinalities of covers and in condition (B) of the above lemma are explicitly bounded above by and , respectively. Note also that conditions such as
are expressible in the language. It is therefore clear that condition (B) is expressible in this language, for each fixed positive integer . As a consequence, we obtain Theorem 6.1.
6.2. Parametrizing balls in in dimension two and higher
For the proof of Theorem 6.2, let us consider the quotient map
defined by
The image of is dense in the circle , equipped with the natural cyclic order. The expression will be regarded as a (definable) constant symbol in . 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 satifying
if and only if the (unsigned) angular metric between and is .
Let us also define an formula
We also use the formula
We will equip with a compatible metric , and denote by the induced uniform metric on the homeomorphism group. We have the following characterization of uniform convergence:
Lemma 6.5.
Let be a regular open set in such that , and let
be a sequence of subsets of such that each setwise stabilizes . Then the following two conditions are equivalent.
-
(A)
We have
-
(B)
Suppose we have two tuples of regular open sets
such that
Then there exist some such that whenever a pair belongs to
each satisfies
Proof.
Let us assume part (A), and also the hypotheses of (B). We set
which is positive since is finite. Choosing so that
for all , we obtain the conclusion.
Conversely, we assume the condition (B) and pick an arbitrary . Let be a finite cover of by regular open sets with radius less than . Applying Lemma 6.4 (after replacing the number in the lemma by the size of ), we obtain a tuple of regular open sets
such that every connected component of each has diameter at most , and such that holds. By Lemma 2.6 and by compactness of , we obtain
such that
Pick as given by the condition (B), and let and be arbitrary. Since there exists some such that , we see that
This implies that and that condition (A) holds. ∎
We now interpret non-integral powers of group elements, in the following sense:
Lemma 6.6.
There exist formulae
such that the following hold for each .
-
(1)
For group elements , a regular open set , and a real number satisfying and
we have
if and only if
-
(2)
For and satisfying and , we have
if and only if there exists a unique topological flow
such that, with the notation , we have the conditions below:
-
•
for each , we have ;
-
•
the map is a topological embedding of into the group
-
•
for each , we have .
In this case, for and , the map
is definable.
-
•
-
(3)
If , then for and , we have
Proof.
It is straightforward to check
satisfies the desired conditions in (2). In particular, the uniquenss is a consequence of the fact that the formula uniquely determines the restriction of on , as an approximation of the form
satisfying
in . The definability of the flow in (2) and the independence on the choice of in part (3) also follow by the same reason, completing the proof. ∎
In the situation of Lemma 6.6, we will say that defines a circular flow on the open set . When we have , the element is viewed as an irrational rotation through a specified angle, and is the rotation of the –multiple of this angle. By the definability of for , we are justified to use an expression such as
in an formula with the hypothesis that . When the meaning is clear, we also use the more succinct notation
We are now ready to complete the proof of Theorem 6.2:
Proof of Theorem 6.2.
By Lemma 6.6, we have an formula that expresses the following:
-
•
there exists some such that
for each , and such that
-
•
there exists a continuous bijection defined by
-
•
For all and for all permutation of , we have
Here, it is implicitly required that
for all , so that
is well-defined. The formula is simply obtained from the map
This proves part (1).
For part (2), we may identify and for some sufficiently large . We can then choose independent circular flows such that each flow rotates in some compact solid torus with the rotation number , and such that on the outside of the restrictions of the flows are the identity; see Figure 1 (a), where a suitable homeomorphism is applied to for illustrative purposes. Such choices of flows will yield the desired conclusion. ∎
We remark that in the measure preserving case, if is not sufficiently small, then there may not be enough room for a solid torus inside such that occupies –fraction of the torus. For instance, one may consider an annulus that is homeomorphic to , 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 . We now describe a definable parametrization of collar neighborhoods of the boundary of a compact –manifold. More specifically, we will establish the following.
Theorem 7.1.
Then there exist formulae
for some tuple of variables in the language such that each pair with satisfies the following:
-
(1)
We have that .
-
(2)
Let be a tuple of elements in satisfying
Then there exists a unique collar embedding
of such that for all points and , and for all we have
7.1. Decomposition of a unit cube
Let us fix . We will use a certain partition of a cube to parametrize a collar neighborhood of . We set
For convention, we also let
By abuse of notation, we move or remove parantheses rather freely and often write
when the vector is used to index certain objects . For each
with , we let be the dyadic cube of side length that contains the following two points as opposite vertices:
For instance, we have
and so on. We have partitions (with disjoint interiors):
We have a unique parametrization
of the regular cube obtained by a positive homothety and translation.
7.2. The condition for a collar neighborhood
Let us first consider the case that . For a tuple
in the universe of , we consider the collection of conditions with appropriate notation as itemized in (a) through (i) below; see Figure 2 for an illustration when .
Condition
-
(a)
We have regular open sets and such that
and such that every regular open neighborhood of contains for some ; moreover, each has finitely many components, and the closures of distinct components are disjoint.
-
(b)
We have dispersed (see Definition 3.11) regular open sets
for each and ; moreover, we have for each that
-
(c)
For each , we have and
such that is a nonempty, finite, minimal –invariant set; moreover, the map
is a bijection
-
(d)
For each and , there exists a unique satisfying
For each , there also exists a unique such that
We further have closure–disjoint unions
-
(e)
For each , we have . Setting , we also have
For all and we have that
-
(f)
For each in the index set
there exists a unique
the closure of which contains .
-
(g)
For each and , we have
We further have that
-
(h)
For each , there exists a homeomorphism
such that for each we have
and such that
-
(i)
If , then some and satisfy
Moreover, in this case we require that for each , we have
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 satisfies these conditions with a suitable choice of .
The first point is trivial to check from the preceding results, possibly except for the continuity condition in (h) at the level– subset of . At such a point , we then can simply require the convergence of the values of the form
whenever gets arbitrarily close to ; we also require the bijectivity of the resulting map onto . We can now let be the formula expressing the condition .
Regarding the second point, we note the following:
Claim.
Under the hypothesis , we have a collar embedding
which is unambiguously defined by
for all
In particular, the image of the level– set under the map coincides with .
Proof.
From the above claim and from the definability of , we obtain the desired formula expressing the map . We complete the proof of part (2) in Theorem 7.1 by simply reparametrizing so that the level– 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 being contained in a collar neighborhood. Hence, we may simply start with a homeomorphism
that satisfies
Using Ostrand’s theorem (Lemma 6.3 (3)), we can write
for some , each of whose components is homeomorphic to . We have a natural homeomorphism
Denote by the image of under this homeomorphism. We can find a homeomorphism that permutes the components of as in condition (d). We let and . We further define
and set
The regular open sets are similar and straightforward to define. The homeomorphism is clearly defined, so that . After decomposing modeled on , we find for the current setup using the uniform convergence theorem. Here, it is crucial that the diameters of the cubes converge to zero as they approach the boundary. This completes the proof of the case .
Slightly more care is needed in the measure preserving case . 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 to be sufficiently far from each other. More precisely, we will pick a sufficiently large depending on , and replace condition (g) by the following two conditions; we also change the definition of the tuple , which is now required to contain the group tuple variables as below.
-
(f)’
For each , and , we have
We further have that
-
(f)”
For each and , after setting we have that
Part (2) of Theorem 7.1 is still proved in the same way, even independently of the choice of . For part (1), we choose sufficiently large, under a fixed metric and a measure on some chart neighborhood of . We will require that for each fixed , each open set in the collection
is contained in some closure–disjoint collection of open balls
with the additional requirement that 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 , 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 and .
For positive integers and , we denote by the set of all tuples of embeddings
from to such that the following conditions hold:
-
(i)
The following set is a compact connected –manifold:
-
(ii)
There exists a collar such that for all .
-
(iii)
We have that
-
(iv)
For each , the restriction
is an embedding of into such that
where here and , and such that
Every compact –manifold is homeomorphic to for some tuple
as above, which we call as a parametrized cover of . The space inherits the uniform separable metric from the space
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 and for each , there exists such that every that is at most –far from admits a homeomorphism
that is at most –far from the identity map.
We choose a sufficiently small for which the conclusion of Lemma 8.1 holds, and call it as a Cheeger–Kister number of ; for our purposes, we will further require to be rational. Our strategy for proving Theorem 1.4 is providing a sentence in which is modeled by an input manifold , such that the sentence holds for the structure if and only if admits an embedding into Euclidean space that is within the Cheeger–Kister number of a fixed parametrized cover of .
In order to execute this strategy, let us fix a pair with . We will slightly modify the definition in Theorem 6.2 by affine transformations, so that is a map from into , sending to .
We let and be positive integers, and consider a tuple
of functions in . Let us denote by the collection of all the conditions below from (a) through (e); see also Figure 3:
-
(a)
each restricts to an embedding of into ;
-
(b)
for all indices as above, we have some
satisfying , corresponding to the homeomorphism
-
(c)
there exists a collar
such that , and such that
-
(d)
for each , the restriction
is an embedding of into such that
where here and , and such that
-
(e)
whenever for some , we have
The condition implies that
defines an embedding
and that the tuple is a parametrized cover of the image.
Recall the domain of the sort symbol is . By the preceding results, there exists a formula
expressing in . We emphasize that although the maps do not belong to the universe of , 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 satisfy . For positive integers and , there exists a formula with a –tuple of variables
in the language such that
if and only if the condition is satisfied.
We can now establish the main result of this paper.
Proof of Theorem 1.4.
We may assume that and that . Consider a parametrized cover
of . We have a corresponding Cheeger–Kister rational number
Let us pick such that
We can find a partition of having diameters less than such that each is the intersection of with a cube with rational corners. Each is definable in , since so is every rational number. We arbitrarily pick in , and choose such that
Let us now consider the following conditions for an arbitrary , which are first order expressible in by preceding results:
-
•
;
-
•
some tuple satisfies that
and also
The above conditions are obviously met in the case when . We also note that for each that
By Lemma 8.1, we see that is homeomorphic to .∎
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 –analogue of Theorem 1.4 holds.
Question 9.1.
Let be a compact, connected, smooth manifold, and let be an arbitrary smooth manifold. Is there a sentence in the language of groups such that if satisfies then is diffeomorphic to ?
Relatedly, leaving the framework of first order rigidity, we have the following.
Question 9.2.
Let be a compact, connected, smooth manifold. Is there a finitely generated (or countable) group such that acts faithfully by diffeomorphisms of a compact, connected, smooth manifold of the same dimension as if and only is diffeomorphic to ?
The discussion in the present article depended heavily on the compactness of the comparison manifold.
Question 9.3.
Let be an arbitrary manifold. Under what conditions is there a sentence in the language of groups such that if is an arbitrary manifold then satisfies if and only if is homeomorphic to ? More generally, under what conditions does imply ?
We conclude by asking what the weakest hypotheses on can be.
Question 9.4.
For what classes of subgroups of do the conclusions of Theorem 1.4 hold?
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 and , 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 and , 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