The homeomorphism group of the universal Knaster continuum
Abstract.
This paper has two aims: the first is to define a projective Fraissé family whose limit approximates the universal Knaster continuum. The family is such that the group of automorphisms of the Fraissé limit is a dense subgroup of the group, , of homeomorphisms of the universal Knaster continuum.
The second aim is to compute the universal minimal flows of and . We prove that both have universal minimal flow homeomorphic to the universal minimal flow of the free abelian group on countably many generators. The computation involves proving that both groups contain an open, normal subgroup which is extremely amenable.
2010 Mathematics Subject Classification: 05D10, 03C98, 37B05, 54F15, 03E15
Keywords: universal minimal flows, projective Fraissé limits, Ramsey theorem, continua
1. Introduction
Knaster continua are a well-studied class of indecomposable continua (precise definitions are in Section 2). We focus here on the universal Knaster continuum–a Knaster continuum which continuously and openly surjects onto all other Knaster continua and particularly on the dynamics of the Polish group of homeomorphisms of the universal Knaster continuum. Some recent results on the dynamics of continua homeomorphism groups of the sort we are interested in here can be found in [3] and [16]. In each case, an appropriate extremely amenable subgroup of the homeomorphism group is found and studied. A topological group is extremely amenable when every continuous action of it on a compact Hausdorff space has a fixed point (see [15] for more on the topic).
The main result of the current paper is a structural theorem about the group :
Theorem 1.1.
The group is isomorphic as a topological group to where is an extremely amenable Polish group and is the free abelian group on countably many generators with the discrete topology.
Recall that the topology on the semidirect product of two topological groups is the product topology on the underlying product space. Theorem 1.1 is proven in Section 7 (as Theorem 7.1). We also show (Proposition 7.10) that is not isomorphic to the direct product .
A few definitions before we say more about the motivation for Theorem 1.1; if is a topological group, we call a compact Hausdorff space equipped with a continuous -action, a -flow. It is minimal if the orbit of each point is dense in . By abstract topological dynamics, for each topological group we can find a unique (up to isomorphism) flow that is minimal and has the property that for any minimal -flow , there is a -equivariant, continuous surjection (see [17] for a short proof of these facts). The flow is called the universal minimal flow of . Observe that is extremely amenable if and only if is a singleton. After extreme amenability the next dividing line in complexity is whether or not is a metrizable space. By results of Ben Yaacov, Melleray, and Tsankov in [4], when is metrizable, must contain a large (closed, co-precompact) extremely amenable subgroup and the universal minimal flow of is the translation action on the completion of . Our case is firmly outside the realm of [4]; it is immediate from Theorem 1.1 that is non-metrizable and with a little more work that is homeomorphic to (Corollary 7.8). Even so it is interesting to compare the situation here with the theory of metrizable universal minimal flows developed in [4]. The group also has a “large”–in that it is open, normal (it is of course, not pre-compact)– extremely amenable subgroup whose quotient determines the behavior of its universal minimal flow.
The main tools used in the paper are projective Fraissé families and the Kechris-Pestov-Todorcevic correspondence. Irwin and Solecki in [9] introduced projective Fraissé families of finite structures, a dual version of a classical construction from model theory. The Fraissé limits in the dual setting carry zero-dimensional compact topologies (rather than being countable spaces as in the classical setting). Under mild assumptions, the projective Fraissé family will have a canonical quotient that is a compact and often connected topological space (precise definitions are in Section 3). Projective Fraissé families are by now a well-used tool for studying continua, they are used for example in the work of Bartošovà and Kwiatkowska in computing the universal minimal flow of the homeomorphism group of the Lelek fan (see [2], [3]). The Kechris-Pestov-Todorcevic correspondence is a general result that connects extreme amenability of automorphism groups of Fraissé limits with the Ramsey property for classes of finite structures (see [11]).
Now we outline the structure of the paper. Section 2 contains some background on Knaster continua. In Section 3, we define a family of finite structures and prove that it is a projective Fraissé family (Theorem 3.2). In Section 4, we prove that the quotient of the Fraissé limit, , is homeomorphic to the universal Knaster continuum, . In Section 5, we show that the group embeds densely into the group via an approximate projective homogeneity property for (Theorem 5.1). We note that Wickman in her Ph.D. thesis, [19], independently discovered a projective Fraissé class whose limit approximates the universal Knaster continuum. The automorphism group of the Fraissé limit of the class from [19] however does not embed densely into the group which is the main property we are concerned with in this paper.
In Section 6, we collect some results of Dębski that we need for the computation of universal minimal flows. The central notion needed is Dębski’s definition of degree for homeomorphisms of the Knaster continuum. In Section 7, we prove the main theorem (Theorem 7.1).
Acknowledgements
I would like to thank Sławek Solecki and Lauren Wickman for helpful conversations about this project. I would also like to thank Sławek Solecki for suggesting this question to work on.
2. Background on Knaster continua
Let be the closed unit interval. A Knaster continuum is a continuum of the form where each is continuous, open, surjection that maps to . In the case that all but finitely many of the maps are monotone, the resulting continuum is homeomorphic to an arc. Except for this trivial case, all other Knaster continua are indecomposable–they cannot be written as the union of two non-trivial proper subcontinua. The simplest example of a Knaster continuum is one in which each bonding map is the tent-map given by
and is known as Brouwer’s “buckethandle" continuum. Indeed, it can be embedded into in such a way that it resembles a thick buckethandle whose sections are Cantor sets (several pictures of Knaster continua can be found in [18]). We collect below a few basic facts and definitions about Knaster continua that we will need.
First, an observation about continuous open maps (it can be found as [8], Theorem 2.1 for reference).
Lemma 2.1.
Let be open and continuous, with . Then, there exists so that is a homeomorphism of onto and for all .
Given as in Lemma 2.1, the number is unique and is called the degree of . We denote it by . An obvious but somehow vital observation is that if and are open continuous maps , then .
Let be a Knaster continuum. The topology on is the product topology and it is induced by the metric:
for all . We always consider the group of all homeomorphisms of , which we denote , with the topology of uniform convergence. This is the topology induced by the metric:
for all . With this topology, is a Polish group (a topological group whose topology is completely metrizable and separable).
3. A Fraisse category of finite linear graphs
We define in this section a projective Fraissé category of finite graphs to approximate the universal Knaster continuum. A finite linear graph, , is a finite set of vertices together with an edge relation satisfying:
-
(1)
for each , and
-
(2)
for each , where
-
(3)
Note that our graphs are different than usual combinatorial graphs in that we allow (in fact, require) loops at vertices.
These two special vertices that satisfy condition (3) will be called end vertices of V. A subset of a finite linear graph is connected if it is connected with respect to the edge relation. Also we have the usual graph metric, , on where is the length of the shortest -path in connecting to .
A pointed finite linear graph is a finite linear graph with one special designated end vertex called the zero-vertex. As suggested by the name, a convenient notation for working with finite linear graphs is the following: we label the vertices of a graph with vertices by so that corresponds to the zero-vertex and vertices and are connected by an edge if and only if . By , we denote the pointed finite linear graph with vertices labeled as just described. An epimorphism between pointed finite linear graphs is a function that is a graph homomorphism (i.e., takes edges to edges) which is surjective on vertices and edges and which preserves the zero-vertex.
To put things into the model-theoretic context in which Fraissé theory is usually developed: we are working with a language consisting of one binary relation symbol and one constant symbol respectively and with the class of finite -structures in which is interpreted as an edge relation satisfying (1)-(3) above and is interpreted as an endpoint. Epimorphisms as defined above correspond exactly to the usual model theory notion of epimorphisms between structures.
We will consider a Fraissé category in which objects are pointed finite linear graphs and morphisms are restricted to a proper subset of all epimorphisms.
Definition 3.1.
Let be the category of pointed finite linear graphs where morphisms are all epimorphisms, , such that: there exist such that
-
(1)
is either non-increasing or non-decreasing for
-
(2)
for odd
-
(3)
for even
The morphisms are essentially discrete versions of the open, continuous surjections . For a morphism as in Definition 3.1, define the degree of , denoted , to be .
Following [14] and [9], we say that a class of finite -structures with a fixed family of morphisms is a projective Fraissé family if:
-
(1)
is a category–morphisms are closed under composition and the identity map on each structure is a morphism
-
(2)
contains countably many structures (up to isomorphism)
-
(3)
for any , there exist and morphisms and
-
(4)
for any and morphisms and , there exists and morphisms and so that
Properties (3) and (4) above are called the joint-projection property and the projective amalgamation property respectively.
Theorem 3.2.
is a projective Fraissé family.
Points (1)-(3) in the definition are easy to check: we will show that the projective amalgamation property holds. The proof proceeds by amalgamating “slope-by-slope” via a sequence of lemmas. The following notation is convenient. For a function, we let be the string . For a finite string , we denote by the string . For and , by we mean the string . The symbol denotes the concatenation of string with string .
Let be the set of non-decreasing morphisms in . Let be the set of non-increasing functions between objects of which send edges to edges and are surjective on both vertices and edges. Note that the members of are not technically epimorphisms since they do not preserve the zero-vertex.
Lemma 3.3.
Let be a morphism. For all , , if we let
then there is , with .
Proof.
Let
One can check that and clearly is non-decreasing. ∎
Applying Lemma 3.3 times and composing yields the following:
Lemma 3.4.
Let be a morphism, and let
Then there is so that .
We can now amalgamate along one slope:
Lemma 3.5.
Let and both be in or in . Then there exist and morphisms and so that . Further and are both in .
Proof.
Suppose that . As and are non-decreasing, they have the following form:
We let for and let
Then set and apply Lemma 3.4 to find and so that and .
A nearly identical argument proves the claim for . ∎
An interval in is just a subset of the form for some and we denote such an interval by . For an interval , let be the greatest integer such that and let be the least integer such that . We write for intervals if .
Proof of Theorem.
3.2 Let and be morphisms. We want to find with common domain such that .
We claim that it is enough to prove the Theorem for morphisms and satisfying the condition below:
A morphism has condition if there exist , disjoint intervals of so that and for each , is either a non-decreasing surjection or non-increasing surjection. Suppose that . The fact that implies that there is as in Definition 3.1. Then, consider the string
and by Lemma 3.4, find so that . It is easy to see that has condition . We do the same with to produce with the property above. Then, notice that amalgamating over and produces an amalgamation over and as desired.
So from now on we assume and have . Divide into intervals with so that for each , is a non-decreasing surjection or a non-increasing surjection. Similarly subdivide into intervals for .
Let be the function given by and let be given by . Define as and by . For any , . When are such that and are both in or both in , we let and let and be non-decreasing morphisms such that
The existence of , and is by Lemma 3.5.
We define two strings and as follows. Let and and let and . We refer to and as blocks. Then
where there are blocks in the concatenation above and depending on the parity of . Also
where there are blocks in the concatenation above. Note that .
Further for all , and are both in or both in . This can be seen via two observations: (1) and are both in and (2) for all , and and . It is easy to see that (2) holds so long as and are in the same block. If is in a different block from , then note that and so (2) follows.
We set
and note that is naturally formed as the disjoint union of intervals of the form:
where we set .
We now define interval-by-interval as follows:
For , we let . Suppose is defined on for all . If , then for , we let . If , then for , we let .
And similarly, we define interval-by-interval: For , we let . Suppose is defined on for all . If , then for , we let . If , then for , we let .
This choice of and work; one can verify that and . ∎
4. The Fraissé limit of
4.1. Enlarging category
We first must define an enlargement of the category to include possible limit objects. This type of enlargement is used in [14] and we follow the development from that paper. A topological pointed graph, , consists of a compact, metrizable, zero-dimensional domain space, ; a reflexsive, and symmetric relation which is a closed subset of ; and a designated point . Epimorphisms between topological pointed graphs must take edges to edges, preserve the designated point, be surjective on vertives and edges, and moreover must be continuous. Of course, finite pointed linear graphs (with the discrete topology) and epimorphisms as defined earlier for finite pointed linear graphs are examples of topological pointed graphs and epimorphisms respectively.
We will be interested only in a proper subset of topological pointed graphs and of morphisms between them. Loosely, we want those structures which are “approximated" by the category and therefore are possible limit objects for and subcategories of . Define the category as follows. If is a sequence consisting of finite pointed linear graphs, , and where each is in , then there is a natural way in which to view as a topological pointed graph. The topology on is the compact, zero-dimensional product topology that arises by taking each as a finite discrete space. The graph relation, , defined by
is easily seen to be reflexsive and symmetric and the designated point is given by . Now, contains all topological pointed graphs of this kind: those which arise as inverse limits of sequences of finite pointed linear graphs with bonding maps from . In particular, is contained in . Morphisms in are defined as follows: given and in , is a morphism if there is an increasing sequence of natural numbers and maps with each and so that for each , where is the projection of onto its th coordinate and is defined analogously for . It can be easily seen that morphisms in are a proper subset of all epimorphisms between as topological pointed graphs. Notice that if and , then morphisms are all maps of the form for some and some in . For example, the canonical projections are morphisms.
For , we call a pre-continuum if is transitive (i.e., an equivalence relation). Each pre-continuum has a quotient, that is, the space with the quotient topology. This space is a compact metrizable space since the relation is always a closed relation and the topology on is always compact, metrizable. The terminology of pre-continuum is justified by the following:
Theorem 4.1.
A topological space is a Knaster continuum if and only if is homeomorphic to the quotient of some pre-continuum in .
For the proof, recall that the mesh of an open cover of a metric space is given by . We will also need the following folklore/ easy to check fact. If is a topological pointed graph formed as the inverse limit of finite pointed linear graphs such that each bonding map is a monotone epimorphism (monotone meaning that pre-images of connected sets are connected) and is an equivalence relation, then is homeomorphic to an arc. Further, if is the quotient map, then is an endpoint of the arc . An endpoint of a continuum is a point such that if is any arc in with , then is an endpoint of .
In the proof below we will at one point mention the property of Kelley (a property of continua) but we will never actually use the definition so we do not define it here. For a definition, see [5], Definition 6.2 on p.23. Following Charatonik and Roe in [5], we say that an epimorphism between finite linear graphs, is confluent if for any connected subset , every connected component, , of is such that (see [5], Definition 4.1 on p.15).
Proof.
Let be a pre-continuum in , let be the projection map on the th coordinate, and let . We want to show that is a Knaster continuum. As per a characterization due to Krupski (Theorem 3 of [12]), is a Knaster continuum if:
-
(1)
is a chainable continuum
-
(2)
has the property of Kelley
-
(3)
every proper subcontinua of is an arc
-
(4)
has one or two endpoints
Point (1) above follows from applying Lemma 4.3 of [9] since any open cover of is refined by some , which is an epimorphism onto a finite linear graph.
The fact that has the property of Kelley is a consequence of Theorem 6.5 of [5] along with the observation that every morphism in is confluent. Let be a proper subcontinua of . Then, we have that for each , is an -connected subset of . A moment of thought about the definition of the maps in checks that must be an inverse limit of connected linear graphs with monotone bonding maps; so must be an arc. This proves (2) and (3).
To see (4), first we show that is an end point. If is an arc in containing , then for each , is an -connected subset of containing , i.e., is a finite pointed linear graph. Since each is connected, for all but (possibly) finitely many , the restriction of to is a monotone map from onto that sends to . Now it follows that is an endpoint of . If every map has odd degree, then a similar argument shows that the point where is the endpoint of not equal to is also an endpoint. If is any other point in , then we can show that is not an endpoint as follows: let for some . There is some so that is not an endpoint of . Let be an -connected subset of so that and each -neighbor of . Let and denote the two distinct -neighbors of other than itself. For , let be the connected component of containing . Then: is an inverse limit of finite linear graphs with monotone bonding maps; so is an arc contained in . Because is transitive,
and it follows that is an interior point of . By the theorem of Krupski, is a Knaster continuum.
Now, let be a Knaster continuum. By Lemma 4 of [6], up to homeomorphism, we may assume that each is a standard tent-map of the form:
where is the degree of . Denote by the projection of onto it’s th coordinate. For each , we construct a chain on so that
-
(1)
-
(2)
-
(3)
if and are such that and
, then -
(4)
for each , there exists so that
-
(5)
and so that there exists so that
-
(6)
if
-
(7)
for each , there is so that
-
(8)
if with , then there is some so that
A chain satisfying properties (6)-(8) above is called an -fine chain. Given any , it is easy to construct a chain on that is -fine for some and so that the diameter of each element of the chain is less than .
We now show how to construct a sequence of chains satisfying (1)-(6). Let be a -fine chain on , for some , with elements of the chain labelled so that conditions (1) and (2) are satisfied. Now given chain and satisfying (1)-(6), we construct chain as follows: let and be such that there is an -fine chain, , on so that each element of has diameter less than
Label the elements of so that conditions (1) and (2) are satisfied. Clearly, (5) is satisfied. Notice that for each :
(we are using, above, that is a standard tent-map and therefore can expand the length of an interval by at most a factor of ). So, by (7) for , and the fact that is surjective, we must have for each , some with –i.e, condition (4) is satisfied for . Condition (6) for implies that (3) holds for . This concludes the construction.
Let be the number of elements of chain . For , let be a pointed linear graph with vertices. We think of as having its vertices labelled by so that is vertex zero. Define by
Such a exists since and by condition (8) for chain . Clearly, . Condition (4) implies that is surjective on vertices and condition (5) implies that is surjective on edges. The fact that and conditions (2) and (7) imply that preserves edges. So, is an epimorphism. Lemma 2.1 for gives that .
Let . First, we will show that condition (3) implies that for any there is at most one element so that and . Assume for contradiction that are distinct so that and . Then, there is some so that , and are pairwise disjoint. Since is a linear graph, it follows that . So, by condition (3), which implies that either or , a contradiction. Since any element of is -connected to at most one element (other than itself) of , we have that is transitive.
We claim that is homeomorphic to . Define as follows. Given , let
Note that the cover has mesh at most .
The sequence is a sequence of nested compact sets whose diameters converge to zero, so there is a unique point in the intersection above. We claim that:
-
(1)
-
(2)
is continuous
-
(3)
is onto
and from points (1)-(3) and the fact that is compact and is Hausdorff it follows that descends to a homeomorphism .
To see (1), assume first that and are such that . It follows that for each , and intersect. Notice that the sequence
is a sequence of nested compact sets whose diameters converge to zero and that and are both in the intersection of this sequence of compact sets by the definition of . It follows that . If , then there is some so that and so by (3). Then, and and so . Point (2) follows essentially from the definition of the map .
To see (3), let . Consider the tree of all sequences of the form
where for all : , , and . The partial order on the tree is given if extends . Then, it is easy to see that this tree is finitely branching and countably infinite; so Koenig’s lemma implies that there is an infinite branch of the form . Note that defined by is an element of and for each which implies that .
∎
4.2. Universal Knaster continuum
The idea of classical Fraissé theory is that given a suitably rich category of finite structures and embeddings (where suitably rich means satisfies an injective amalgamation property), there is a unique ultrahomogeneous limit object: a countable structure so that every member of the category has an arrow into the limit object and every arrow between finite subobjects of the limit extends to an automorphism of the full limit. In the dual setting, the limit satisfies analogous properties with the arrow directions “swapped." We will follow [9] and [14] for the development of projective Fraissé limits.
Theorem 4.2 (Theorem 3.1 of [14]).
If is a projective Fraissé family of finite (pointed) graphs, then there exists a unique up to isomorphism
(pointed) topological graph so that
-
(1)
for each , there is a morphism in from
-
(2)
for and morphisms and ,
, there is a morphism so that
First a remark about point (2) of Theorem 4.2 above: this property is sometimes called the projective expansion property. Because of the definition of morphisms in , point (2) has the following consequence which we refer to as projective ultrahomogeneity: for any morphisms , there is some so that . The proof of this fact is a simple diagram chase.
The structure in Theorem 4.2 is referred to as the projective Fraisse limit of the family . It will be helpful in Section 7 to understand a bit about how this limit is constructed, so we will say a few words about it here. Given a Fraissé family , we will say that a sequence of objects in with morphisms from is a generic sequence if the following two conditions hold:
-
(1)
for any , there is and in so that
-
(2)
for any and and , there is some and so that
The inverse limit of any generic sequence is the Fraissé limit of (this will be the only fact that we need in Section 7). The construction of a generic sequence in a Fraissé category uses the amalgamation property repeatedly to build a sequence that is “fully saturated" with the respect to the morphism types of (of which there are only countably many). The paper [9] develops the construction in detail. In the case that the graph relation on the Fraissé limit, , is an equivalence relation then we call the canonical quotient of .
From now on, we denote by the projective Fraissé limit of the category defined in Section 3.
Proposition 4.3.
The canonical quotient of is the universal Knaster continuum.
Proof.
Write for a generic sequence. First we need to show that has a canonical quotient, i.e., that is transitive. An inspection of the proof of Theorem 4.1 reveals that is transitive (in fact, has equivalence classes of size at most 2) so long as the following condition holds: for any and any with there is some so that any and have the property that . Given , let be an increasing morphism such that every and every are so that (this is easy to arrange simply by ensuring that is an interval of length 2 for some that is between and in ). Apply property (2) of a generic sequence to find some and with and note that this has the property desired.
Let be any Knaster continuum; by Theorem 4.1, there is
so that is homeomorphic to . By property (1) in Theorem 4.2 of projective Fraissé limits, there is a morphism . By property (2), there is a morphism so that . Continuing on in this way, we produce for each , so that . Then, define by for each and notice that is a morphism. As each sends edges to edges, sends edges in to edges in . So descends to a continuous map . ∎
With some extra care, one can show that the map constructed in Proposition 4.3 is open as well.
5. Density of in
Proposition 4.3 pushes the universality of the projective Fraissé limit (point (1) in Theorem 4.2) to the quotient of the limit. Using approximation results of Dębski, we can push projective ultrahomogeneity to the quotient of the limit. This gives an approximate ultrahomogeneity property for the universal Knaster continuum and more important– establishes that is dense in .
From now on, let denote the quotient map. We define a map . Any automorphism gives rise to a homeomorphism, , of defined by
and further, is a continuous embedding of into . When is a compact space, we always take with the topology induced by the supremum metric. We take with the topology induced by considering it as a subgroup of the group of homeomorphisms of the topological space .
Theorem 5.1.
Let be continuous open surjections onto a Knaster continuum and let . Then, there exists so that
Proof.
Let be the projection onto the th coordinate. Similarly, is the projection of onto its th coordinate. Let , where each , and denote by the projection of onto its th coordinate. Let be a sequence of open covers of as in the proof of Theorem 4.1 (satisfying conditions (1)-(8)) for . Fix be large enough so that refines the cover of by balls of radius . For ease of notation we let be labelled as in Theorem 4.1. Let be so that the chain is -fine. Consider ; this is a continuous, open map that sends to 0. By a theorem of Dębski (Corollary on p. 207 of [6]), there is a sequence of continuous open maps mapping 0 to 0 so that converges uniformly (in the supremum metric) to . Let and and be continuous open maps such that:
and
Let be a chain on of mesh less than where is chosen, by the uniform continuity of , so that
and where the chain is labeled so that if and only if and . Now let so that the cover
refines the open cover
Define by
Claim 5.2.
The map is in .
Proof of Claim 5.2.
Notice that
and that
So, . To see that sends edges to edges, suppose that and ; then there exists and so that . So which implies that . So
which implies in turn that
and so . The fact that follows from Lemma 2.1 applied to the continuous open map . ∎
Analogously, we may find and morphism so that
Projective ultrahomogeneity of (see the remarks following Theorem 4.2) implies that there exists such that . Now it is easy to check that . ∎
The following corollary is most important to us as we are interested in studying the extent to which the dynamics of controls the dynamics of .
Corollary 5.3.
The image of is dense in .
6. The notion of degree
In this section, we discuss an important definition, due to Dębski, of degree of homeomorphisms of Knaster continua. This will be necessary later on when we compute the universal minimal flow of . As is a Knaster continuum, we can write where each and each is a standard tent-map. Let be a homeomorphism. Dębski proves ([6], Corollary on p.207) that there is a sequence of continuous, open maps so that
where the convergence above is with respect to the supremum metric on the space of continuous maps . Further, the sequence
is eventually constant (Lemma 7 of [6]) and in fact independent of the sequence chosen to approximate . We let
and note that is a positive rational number.
Lemmas 11 and 12 of [6] tell us that for , and that . In particular,
is a group homomorphism, where is the group of positive rationals with multiplication. Note that is a free abelian group with countably many generators (the generators are the primes). Lemma 9 of the same paper implies that in fact is continuous when is taken with the topology induced by the supremum metric and with the discrete topology.
We will at this point define an analogous definition of degree for automorphisms of . Suppose that is an automorphism. By definition of morphisms in , there exists and with such that . Then we define:
Recall that is the map induced by the quotient.
Lemma 6.1.
For any , .
Proof.
Let and the projection of onto its th coordinate. For each , let be the number of vertices in . Now, for , we define a map as follows. First, we define a map . Let . For , let
Then, we let be the piece-wise linear map with and with breakpoints at the points
for each . One can check that is a continuous, open surjection with . Given and some , we define
where
and
We have that the map
that sends a point to where
descends to a homeomorphism
There some details to check here, but the proofs are very similar to the proof of Theorem 4.1, so we omit the details. The point of this representation of is that it is clear that if such that for some and , then is such that:
for some continuous open of degree equal to the degree of . So we have that
∎
Lemma 6.2.
The degree map is surjective.
Proof.
Let . Suppose that where and are positive natural numbers. Let be a finite pointed linear graph with enough vertices such that there exists morphisms of degree and morphisms of degree from to . Fix any morphism with . By the projective extension property, there is some and a morphism so that . So:
(6.1) |
Now let be a morphism with . Projective ultrahomogeneity implies that there is some so that and now by 6.1:
∎

The map is continuous as the composition of two continuous maps (see Figure 1 for a diagram of the situation we have thus far). It is a general fact that if and are topological groups and there is a continuous surjective group homomorphism , then induces a surjective, continuous, -equivariant map . This is simply because
defines a continuous action of on and the action is minimal because is surjective. This fact gives us the existence of the two maps on the bottom row of Figure 1 and that they are both surjections. Notice that since is non-metrizable (this is true of any countable discrete group by a theorem of Veech, see [11], Appendix), this immediately implies that the universal minimal flows of and are non-metrizable.
In the remainder of the paper, we will show that the bottom two arrows are actually injective as well; and hence that and are both homeomorphic to .
7. The main theorem
In this section we prove the main theorem:
Theorem 7.1.
The group where is Polish and extremely amenable and is free abelian on countably many generators.
Let:
be the subgroup of all homeomorphisms with degree 1; that is the kernel of the degree map. By the considerations of the previous section, is an open (and so Polish), normal subgroup of . This will end up being the group in Theorem 7.1. We defer the proof of Theorem 7.1 for a bit as it requires some set-up and construction. We first will focus on showing that is extremely amenable using a projective Fraissé construction and the Kechris-Pestov-Todorcevic correspondence.
7.1. Approximating the group of degree one homeomorphisms
We define a Fraissé family, , that is a slight modification of as follows. The objects are all pairs where is a finite pointed linear graph and . For and , is a morphism if is in and . So, between any two fixed objects of , all morphisms have the same degree. In the case that , there are no morphisms between and . Checking that is a Fraissé family is easy. For example to see that the projective amalgamation property holds, given and , let be the finite pointed linear graph and and be the morphisms constructed in the proof of Theorem 3.2. Then, an inspection of the proof of Theorem 3.2 gives that is a morphism and is a morphism ; so witness the amalgamation property for . From now on, will be used to denote the projective Fraissé limit of . There is an obvious forgetful functor from to which simply takes an object to and a morphism just to the morphism of the underlying finite linear graph . Abusing notation slightly, if is in , we will also use to denote the morphism in of the underlying finite graphs .
Proposition 7.2.
Any generic sequence for is a generic sequence for .
Proof.
Let and be a generic sequence in . Now, let . Since , we have a morphism for some and of course, is in . Suppose are finite pointed linear graphs and are morphisms in with and . Note that so and also . Further, and are morphisms in . As is generic for , there is and so that , and of course, . So, the sequence is generic for . ∎
So, let for a generic sequence in . By Proposition 7.2, , the projective Fraissé limit of . Notice that any automorphism of is an automorphism of . We have the following situation:
Proposition 7.3.
The image, , is dense in .
Proof of Proposition 7.3.
We show that:
and from this fact and Corollary 5.3, the proposition follows. First if and is in so that , then note . But, is also a morphism in and so . It follows that .
Conversely, if is a degree one automorphism, then let and so that . We know that
and so . So is in . Now for any , the fact that implies
and so is in . It follows that . ∎
7.2. Extreme amenability
We prove in this section:
Theorem 7.4.
-
(i)
The group is extremely amenable.
-
(ii)
The group is extremely amenable.
Theorem 7.4 will be proved via the dual of the Kechris-Pestov-Todorcecvic correspondence. First we need some set-up and definitions. By a -coloring of a set , we simply mean a function . Having fixed a coloring, , a subset of is monochromatic if there is some so that . Fix a category of finite structures with surjective morphisms. For objects in , let be the collection of all morphisms in . Now we say that an object has the Ramsey property if for any and any , there exists so that for any coloring , there is some so that
is monochromatic. The category is a Ramsey category if every object in has the Ramsey property. Here is the dual of the Kechris-Pestov-Todorcecvic correspondence; it is due to Bartošová and Kwiatkowska ([1]). An object is rigid if there are no nontrivial morphisms .
Theorem 7.5 (Theorem 2.2. of [1]).
Let be a projective Fraissé class
with projective Fraissé limit . The following are equivalent:
-
(1)
the group is extremely amenable
-
(2)
is a Ramsey category and every object in is rigid
The Ramsey statement involved in the category , when reformulated appropriately, ends up being a direct application of the classical Ramsey theorem. For natural numbers, we denote by the set of all increasing injections . If is an increasing injection, by we mean the collection of all functions of the form where is an increasing injection; of course every such is in .
Theorem 7.6 (Finite Ramsey Theorem).
Let and be natural numbers. Then, there exists , so that for any coloring , there exists so that is monochromatic.
We denote by the least such satisfying Theorem 7.6 for as above.
Proof of Theorem 7.4.
Since by Proposition 7.3 is a dense subgroup of , point (i) of the theorem implies point (ii).
By Theorem 7.5, to show that is extremely amenable, we need to prove that is a Ramsey class. So let and let . The statement of the Ramsey property for and is vacuously true if is empty so we may assume that and thus . Now by Theorem 7.6, let and we claim that witness the Ramsey property for and in ; i.e, for any , there is some so that is monochromatic.
Given we associate to the increasing injection defined as follows:
The key observation is that for a morphism of degree , there are exactly distinct vertices, in so that and further the values of these distinct vertices fully determine the morphism . This implies that the set displayed above is nonempty for each and thus that is defined for each . It is clear from the definition of that for each and so . The observation at the beginning of the paragraph implies that map is injective.
Let be a coloring. Define
by (and is arbitrary for elements of not of the form ). Let be so that
is -monochromatic.
Let be the unique degree one morphism defined by the condition:
Now we are done so long as for each ,
(7.1) |
Take , notice that checking 7.1 for is equivalent to showing that for each . To see this, compute that for any :
where we know that exists so that because . ∎
7.3. The proof of Theorem 7.1
We can now prove the main theorem.
Proof of Theorem 7.1.
From the diagram in Figure 1, we have a short exact sequence of Polish groups (every arrow below is continuous):
By [10] (p.9), to show that as a topological group, we need only show that the short exact sequence above splits continuously, i.e., that there is a continuous group homomorphism which is a right inverse of the degree map. It suffices to show that for each prime there exists a homeomorphism of degree such that for any , and commute. Given this, we can extend the mapping to a group homomorphism by first mapping for each reciprocal of prime and then extending using the fact that elements of can be uniquely factored into primes and reciprocals of primes.
Let be the universal Knaster continuum and for each , the projection map onto the th coordinate. We use maps from what is called in [8] the semigroup of standard induced maps (see second paragraph on p. 129 of [8]). For , let be the standard degree- tent-map given by
It is easy to check that for the maps and commute. For each prime , choose to be the homeomorphism of such that
There is a unique such homeomorphism by Lemma 3.3 and Theorem 3.10 of [8]. Now let be primes. Observe that
and similarly
By uniqueness (Lemma 3.3 of [8]), . So we get that . As noted before is free abelian and generated by the set of primes. By Theorem 7.4, is extremely amenable. ∎
7.4. Dynamical consequences
We return to the diagram in Figure 1. The kernel of the map
is extremely amenable by Theorem 7.4. The following Proposition is a general and, surely, folklore fact which applies to our situation.
Proposition 7.7.
Let be a continuous, surjective group homomorphism between Polish groups and suppose that is extremely amenable. Then, the continuous surjective map induced by is a -flow isomorphism.
Proof.
Consider the action of on ; there is some fixed by . We claim there is a well-defined action of on given by:
(7.2) |
This follows from the fact that
(7.3) |
for all and . Equation 7.3 holds for in the orbit of since for any , if is such that then:
where the second equality uses that is normal. Then since the -orbit of is dense in , Equation 7.3 holds everywhere by continuity of the action. The action defined in 7.2 is easily seen to be continuous and -equivariant and it is minimal by 7.3. Now uniqueness of the universal minimal flow implies that the action of on is the universal minimal flow of . As the map induced by is an -flow morphism, it is in fact an -flow isomorphism (see [17], Proposition 3.3). ∎
It follows from Proposition 7.7, that the bottom two arrows on the diagram in Figure 1 are -flow isomorphisms. In particular, we have that:
Corollary 7.8.
The flows and are homeomorphic to .
7.5. is not a product
We show now that is not a direct product of and .
Lemma 7.9.
Every non-trivial conjugacy class in is uncountable.
Proof.
Let be in . Let where each . Let and with and . Let such that . We will assume that (the other case is analogous). For each let be such that and . Each as a map from to induces a unique open continuous map (the other coordinate maps are fully determined once is set, see [8], Lemma 3.3) which we denote by . Each is a homeomorphism (see [8] top of p.129, where such maps are called vertically induced homeomorphisms). One may check that and that there exists with and . So:
In particular for , since the equation above implies that and thus the conjugacy class of is uncountable. ∎
Proposition 7.10.
The group is not isomorphic to the direct product .
Proof.
If , then would contain a countably infinite normal subgroup. This is clearly impossibly by Lemma 7.9. ∎
8. Remarks on Ramsey degree
An object in a projective Fraissé class has infinite Ramsey degree if for any , there exists so that for any with we have that: there exists a coloring such that for each , for each . Recall from Section 6 that the group has non-metrizable universal minimal flow. It is a theorem of Zucker that if is a Fraissé class of rigid structures, then has non-metrizable universal minimal flow if and only if contains an object of infinite Ramsey degree (see [20], and for a proof of the dual of Zucker’s theorem see [13]). So we know abstractly that must contain an object of infinite Ramsey degree. It is perhaps worth mentioning that one can prove this fact directly:
Proposition 8.1.
There is an object in with infinite Ramsey degree.
Proof.
Let be a pointed linear graph with two vertices. Fix . Let be a pointed linear graph with vertices. We claim that for any with there is a -coloring of so that for any , the set has morphisms colored by every one of the colors.
For a natural number , denote by the largest natural number such that . Let be as above and then color as follows:
Since , it is easy to check that for any ,
∎
We note that just the fact that has non-metrizable universal minimal flow and is dense in is not enough to conclude that has non-metrizable universal minimal flow. For example, any countable dense subgroup of (for instance, the group of all piece-wise linear homeomorphisms with finitely many pieces, breakpoints at rationals, and taking only rational values at breakpoints) has non-metrizable universal minimal flow whereas is extremely amenable (Pestov, [16], Theorem 6.2).
References
- [1] D. Bartošová and A. Kwiatkowska, Gowers’ Ramsey theorem with multiple operations and dynamics of the homeomorphism group of the Lelek fan. J. Combin. Theor Ser. A 150 (2017), 108–136.
- [2] D. Bartošová and A. Kwiatkowska, Lelek fan from a projective Fraissé limit. Fund. Math. 231 (2015), no.1, 57–79.
- [3] D. Bartošová and A. Kwiatkowska, The universal minimal flow of the homeomorphism group of the Lelek fan. Trans. Amer. Math Soc. 371 (2019), no. 10, 6995–7027.
- [4] I. Ben Yaacov, J. Melleray, and T. Tsankov. Metrizable universal minimal flows of Polish groups have a comeagre orbit. Geom. Funct. Anal. 27 (2017), no. 1, 67–77.
- [5] W.J. Charatonik and R.P. Roe, Projective Fraissé limits of trees. https://web.mst.edu/ rroe/Fraisse.pdf.
- [6] W. Dębski, On topological types of the simplest indecomposable continuaa. Colloq. Math. 49 (1985), no. 2, 203-211.
- [7] C. Eberhart, J.B. Fugate, S. Schumann, The lattice of Knaster continua. Proceedings of the First International Meeting on Continuum Theory (Puebla City, 2000) Topology Appl. 126 (2002), no.3, 343-349.
- [8] C. Eberhart, J. B. Fugate, and S. Schumann, Open maps between Knaster continua. Fund. Math. 162 (1999), no.2, 119–148.
- [9] T. Irwin and S. Solecki, Projective Fraissé limits and the pseudo-arc. Trans. Amer. Math. Soc. 358 (2006), no.7, 3077–3096.
- [10] C. Jahel and A. Zucker, Topological dynamics of Polish group extensions. arXiv:1902.04901.
- [11] A. S. Kechris, V. G. Pestov, and S. Todorcevic, Fraissé limits, Ramsey theory, and topological dynamics of automorphism groups. Geom. Funct. Anal. 15 (2005), no. 1, 106–189.
- [12] P. Krupski, Solenoids and inverse limits of sequences of arcs with open bonding maps. Fund. Math. 120 (1984), no.1, 41–52.
- [13] D. Mašulović, The Kechris-Pestov-Todorcevic correspondence from the point of view of category theory. Appl. Categ. Structures 29 (2021), no.1, 141–169.
- [14] A. Panagiotopoulos and S. Solecki, A combinatorial model for the Menger curve. J. Topol. Anal. 14 (2022), no.1, 203–229.
- [15] V. Pestov, Dynamics of infinite-dimensional groups: the Ramsey-Dvoretzky-Milman phenomenon, AMS University Lecture Series 40, American Mathematical Society, Providence, RI (2006).
- [16] V. Pestov, On free actions, minimal flows, and a problem by Ellis. Trans. Amer. Math. Soc. 350 (1998), no. 10, 4149–4165.
- [17] V. Uspenskij, On universal minimal compact -spaces. Proeceedings of the 2000 Topology and Dynamics Conference, Topology Proc. 25 (2000), 301–308.
- [18] W.T. Watkins, Homeomorphic classification of certain inverse limit spaces with open bonding maps. Pacific J. Math. 103 (1982), no. 2, 589–601.
- [19] L. Wickman, Projective Fraissé theory and Knaster continua, Ph.D. Thesis, University of Florida, 2022.
- [20] A. Zucker, Topological dynamics of automorphism groups, ultrafilter combinatorics, and the generic point problem. Trans. Amer. Math. Soc 368 (2016), no. 9, 6715-6740.