Discordant sets and ergodic Ramsey theory
Abstract.
We explore the properties of non-piecewise syndetic sets with positive upper density, which we call discordant, in countably infinite amenable (semi)groups. Sets of this kind are involved in many questions of Ramsey theory and manifest the difference in complexity between the classical van der Waerden’s theorem and Szemerédi’s theorem. We generalize and unify old constructions and obtain new results about these historically interesting sets. Along the way, we draw from various corners of mathematics, including classical Ramsey theory, ergodic theory, number theory, and topological and symbolic dynamics.
2020 Mathematics Subject Classification:
05D10, 37A441. Introduction
Let be a subset of the set of positive integers. We begin by defining some notions of largeness that can be imposed on .
-
•
is an IP set if it contains a finite sums set, i.e., a set of the form
for some infinite sequence of positive integers.
-
•
is thick if it contains arbitrarily long intervals: for each , there exists such that .
-
•
is syndetic if its gaps are bounded: this means that, if we write , the successive difference is bounded. Equivalently, is syndetic if there exist such that , where .
-
•
is piecewise syndetic if it can be written as the intersection of a thick set with a syndetic set. Equivalently, is piecewise syndetic if there exist for which is thick.
-
•
The upper density of is
When the limit in question exists, this quantity is called the density of , denoted by . A set should be thought of as large in terms of (upper) density if this quantity is positive.
Although we have called each of the above properties a “notion of largeness”, they are not on equal footing. For all these definitions, is large, and if has a large subset, then is itself large. However, only some of these notions of largeness are partition regular, meaning that if is large then, for any finite partition into disjoint sets, at least one is large. This third property underlies much of classical Ramsey theory. The partition regularity of IP sets is a famous theorem of Hindman [28], and the partition regularity of piecewise syndeticity is a classical fact, often rediscovered (see, e.g., [29, Lemma 2.4] and [22, Theorem 1.24]).111A related fact, that if then at least one is piecewise syndetic, appears to have been first observed by Brown [15, Lemma 1]. The partition regularity of positive upper density is an easy exercise for the reader, as is checking that thickness and syndeticity fail to be partition regular.
One of the earliest partition results of Ramsey theory is van der Waerden’s theorem, which says that given any finite partition , at least one is AP-rich, i.e., contains arithmetic progressions of any finite cardinality [35]. A useful equivalent form of this theorem states that the family of AP-rich sets is partition regular. Van der Waerden’s theorem has a corresponding density version, Szemerédi’s theorem [34], one of whose many equivalent formulations guarantees that any set having positive upper density is AP-rich.
In seeking a density version of Hindman’s theorem, Erdős asked whether is enough to imply that contains a shift of a finite sums set. (Of course we cannot ask that contain a finite sums set; consider, e.g., the set of odd numbers.) This question was answered negatively in unpublished work of E. Straus, who showed that, for any , one can find with such that is not an IP set for any . In [19, p. 105], Erdős writes the following:
I first hoped that Hindman’s theorem on all subsums can be generalised for sets of positive density, but the following example of E. Straus seems to give a counterexample to all such attempts: Let be a set of primes tending to infinity fast, and let . Consider the set of integers so that where . The density of this sequence is , and it is not difficult to see that this sequence furnishes the required counterexample.
See [4, Theorem 2.20] for more discussion and a detailed presentation of Straus’ construction.
Recall that if is piecewise syndetic, then a union of finitely many shifts of is thick. Since thick sets are obviously AP-rich, it follows from van der Waerden’s theorem that one of the shifts of (and hence itself) is AP-rich. This tempts us to ask, in analogy with Erdős’ question, whether implies that is piecewise syndetic: a positive answer to this question would mean that we have just proven Szemerédi’s theorem. However, Straus’ construction again provides a counterexample; it has positive density but is not piecewise syndetic. Indeed, it is an easy exercise to show that any thick set is IP, so any piecewise syndetic set contains a shift of an IP set by Hindman’s theorem. Thus Straus’ example cannot be piecewise syndetic.
It is counterexamples of the above kind we are interested in—those sets which are non-piecewise syndetic but have positive (upper) density. We will call such sets discordant. Another classical example of a discordant set is the set of squarefree numbers. However, this set has shifts which are IP [12, Theorem 2.8], and so has a different nature than Straus’ counterexample. More constructions of discordant sets can be found in [13] along with applications to Ramsey theory. A fruitful analogy can be drawn between discordant sets and nowhere dense sets of positive measure, both being small in a “topological” sense but large in an “analytic” sense. This analogy will be developed and more closely investigated in the following sections.
We will now more closely examine the relationship between van der Waerden’s theorem, Szemerédi’s theorem, and discordant sets. We observed above, using van der Waerden’s theorem, that any piecewise syndetic set is AP-rich. On the other hand, given that every piecewise syndetic set is AP-rich, the partition regularity of piecewise syndeticity tells us that, for any finite partition , there is some for which is AP-rich. Thus van der Waerden’s theorem is equivalent, in this informal sense, to the fact that every piecewise syndetic set is AP-rich. Because of this, the existence of discordant sets can be interpreted as a sign that Szemerédi’s infamously difficult theorem is deeper than van der Waerden’s.222Another equivalent form of Szemerédi’s theorem says that any subset of having positive upper Banach density is AP-rich. In contrast to density, for which one averages along the intervals , the upper Banach density of a set is defined to be the supremum of all densities of averaged along any sequence of intervals of increasing length. Any piecewise syndetic set has positive upper Banach density, a fact which clarifies the relationship between van der Waerden’s and Szemerédi’s theorems and strengthens the perspective we take here. The striking abundance of contexts in which discordant sets arise, some of which will be provided in the ensuing sections, can be taken as evidence for the versatility of Szemerédi’s theorem and strengthens our interest in discordant sets.
The distinction between van der Waerden’s theorem and Szemerédi’s theorem can also be profitably analyzed from the point of view of dynamics and ergodic theory. Indeed, as initiated by Furstenberg?s ergodic-theoretic proof of Szemerédi?s theorem [21], one can elucidate the properties of a subset of by examining the dynamical and ergodic properties of the orbit closure of its indicator function. To be more explicit, given , consider its indicator function , where we view as having the product topology (making homeomorphic to the Cantor set). Then, letting denote the left-shift map (i.e., , define the orbit closure of as
The resulting topological dynamical system ( can be made into a measure-preserving system. In this setting, the difference in difficulty between van der Waerden’s and Szemerédi’s theorems is indicated by the existence or nonexistence of nontrivial minimal subsystems of . In the presence of a nontrivial minimal subsystem, a purely topological-dynamical recurrence theorem, the topological analogue to van der Waerden’s theorem, can be used to obtain a streamlined proof of the classical van der Waerden theorem [23, Theorem 0.3]. On the other hand, when does not possess nontrivial minimal subsystems, a considerably more difficult proof via measurable dynamics is necessary to prove Szemerédi’s theorem. In particular, those sets for which has a nontrivial minimal subsystem are exactly the piecewise syndetic sets (Theorem 45). This dynamical and ergodic approach leads not only to a proof of these classical theorems but also to far-reaching generalizations which still elude purely combinatorial proofs, such as versions of the polynomial Szemerédi theorem (e.g. [10, Theorem 1.1] and [11, Theorem 7.3]).
Finally, because the existence of discordant sets is a nontrivial byproduct of the interplay between the notions of density and piecewise syndeticity, both very basic definitions of Ramsey theory, these special sets are particularly nice vehicles for advancing Ramsey theory in different settings and showing the bountiful connections between Ramsey theory and other areas of mathematics. Therefore our paper is not only about discordant sets but also is meant as an accessible (but admittedly very non-exhaustive) demonstration of the diversity of Ramsey theory.
Outline of the paper
This paper concerns itself with the study of subsets of countably infinite semigroups, so some definitions must be introduced to allow us to work in this generality. This is done in Section 2, where we also provide the necessary prerequisites from Ramsey theory, ergodic theory, and topological dynamics.
In Section 3 we give an ergodic-theoretic construction of discordant sets that can be done in many semigroups, generalizing a construction based on irrational rotation of the torus . In particular, we prove that the sets of return times to certain nowhere dense sets of positive measure form discordant sets. This section lays the foundation for the analogy between nowhere dense sets of positive measure and discordant sets, which we revisit and reexamine several times.
In Section 4, we prove in two ways that the squarefree numbers are discordant. The first proof is purely algebraic while the second uses the dynamical results of the previous section. We also give several wide-ranging generalizations of this fact.
Our view narrows in Section 5, where we take a brief excursion to . Although the classical notions of density fail to exist in this group since is not amenable, we define a natural analogue of density which allows us to construct discordant sets using the results of the preceding section.
In Section 6, we study a special kind of discordant set which is particularly bad for recurrence in measure-preserving systems. We establish the existence of such discordant sets with density for each for any countably infinite abelian group and also prove the existence of discordant sets with density for each in any countably infinite commutative cancellative semigroup.
We shift focus in Section 7. First, we generalize the notion of a disjunctive -sequence, one in which every finite -word occurs, to an arbitrary countably infinite semigroup, and show that such sequences are “topologically typical” in the sense that the family of such sequences is comeager. This allows us to state that, in cancellative semigroups, non-piecewise syndetic sets (and hence discordant sets) are topologically atypical. We then consider special types of disjunctive sequences by measuring the limiting frequency with which every finite word occurs. In contrast to how one defines a normal sequence by saying that these frequencies approach , where is the length of the word, we define an extremely non-averageable, or ENA, sequence by declaring that these frequencies should fail to converge as badly as possible. We prove that ENA sequences are topologically typical in arbitrary countably infinite amenable cancellative semigroups, strengthening the previous result for such semigroups.
Finally, in Section 8 we give a topological-dynamical characterization of non-piecewise syndetic sets in arbitrary countably infinite semigroups by looking at their orbit closures. We conclude the paper by giving combinatorial applications of these dynamical results.
Acknowledgements
We would like to thank Dona Strauss for helpful conversations. Her input was instrumental in guiding us toward the proof of Theorem 38. We are also indebted to the anonymous referee for their numerous helpful comments and corrections. The latter two authors were partially supported by the Ohio State Arts and Sciences Undergraduate Research Scholarship.
2. Preliminaries
Throughout this paper, will stand for a countably infinite semigroup.
When we say that is a left cancellative semigroup, we mean that for any , if then ; is right cancellative if instead implies ; and is cancellative if it is both left and right cancellative.
Given , define and ; the sets and are defined analogously. Also, given , define
For any set , we will write to denote the set of subsets of . We write for the complement of , when the superset is understood. In particular, for , always denotes the complement relative to .
2.1. Thickness, syndeticity, and piecewise syndeticity
Let .
-
•
is right thick if for any finite there exists such that .
-
•
is right syndetic if there exists some finite such that .
-
•
is right piecewise syndetic if there is a finite for which is right thick.
Since is not assumed to be commutative, there are also “left” versions of these definitions obtained by replacing with and with . We warn the reader that the left and right versions do not necessarily coincide in noncommutative (semi)groups.333Here are some examples to show that the left- and right-handed notions really are different. In , the free group on and , it is easily seen that is right thick but not left thick, and left syndetic but not right syndetic. For a less noncommutative example, consider the following subset of the Heisenberg group : It can be shown that is left thick but not right piecewise syndetic (and thus not right thick). The group is “almost abelian”, having nilpotency class , so we see from this example that not much noncommutativity is needed to separate the right- and left-handed notions. See [8, Section 2] for many similar examples. Henceforth we will deal only with the right-handed definitions and omit the word “right”.
Note that, if is an increasing sequence of finite subsets of satisfying , then is thick if and only if for some sequence in .
We now outline the relationship between thick, syndetic, and piecewise syndetic sets. First, the family of syndetic sets and the family of thick sets are dual, in the following sense.
Lemma 1.
Let . Then is syndetic if and only if is not thick.
Proof.
Suppose is syndetic, so for some finite . Let . Since , there exists with . Hence , i.e., . This proves that is not thick.
Conversely, suppose is not syndetic, and let be finite. Since , find such that . Then , so is thick. ∎
From this duality we obtain a useful equivalent description of piecewise syndetic sets, which we described in introduction for subsets of .
Lemma 2.
Let . Then is piecewise syndetic if and only if there exist a syndetic set and a thick set such that .
Proof.
First suppose is piecewise syndetic, and find a finite such that is thick. Let , and let , so that . Pick any ; we will show that . Indeed, let . If , then , so .
Now suppose is syndetic and is thick. Fix a finite satisfying . Let be finite and pick satisfying . By the syndeticity of , for any we can find such that . We also know that , so . This proves that , and hence that is thick. ∎
2.2. Densities in semigroups
In this subsection, is assumed to be a countably infinite amenable cancellative semigroup. Given these assumptions, amenability is equivalent to the existence of a Følner sequence.444Amenability is commonly defined as follows. For each , define by . Then is called (left) amenable if there exists a positive normalized linear functional on the space of bounded functions which satisfies for each ; is called an invariant mean for . For countably infinite left cancellative semigroups, the existence of Følner sequences and amenability are equivalent [1]. See [32] for a comprehensive study of the notion of amenability. A (left) Følner sequence in is a sequence of nonempty finite subsets of satisfying
for any . The upper density of with respect to is defined as
When the corresponding limit exists this quantity is the density of with respect to , denoted by , and we also define the lower density to be the corresponding . Note that . Observe also that it follows quickly from the pigeonhole principle and the cancellativity of that Følner sequences satisfy , guaranteeing a degree of non-degeneracy for these definitions.
A subset will be called discordant if it is not piecewise syndetic and there is some Følner sequence in such that . We would like to also guarantee, when possible, the stronger condition , or better, .
In the introduction, we defined density in with respect to the Følner sequence . We will reserve , leaving the Følner sequence unspecified, to refer to this density (and similarly with and ). A set for which exists is sometimes said to have natural density. Closely related are densities defined with respect to the Følner sequences in and in ; we will also use to refer to these without danger of ambiguity.
Upper density has some useful properties, which we briefly describe here. The most primitive is that it is invariant under shifts. In other words, for any and , one has . We record the following similar result as we will use it later on. (The proof of shift-invariance can be seen in the proof of Lemma 3.)
Lemma 3.
Suppose and satisfy . Then .
Proof.
It is clear that is subadditive. For the opposite inequality, let and find such that
Then
Now since by cancellativity, we get
Finally, let us remark that density is related to thickness and syndeticity in the following way: is thick if and only if for some , and is syndetic if and only if for all ; see [8, Theorem 2.6]. More discussion of thickness, syndeticity, piecewise syndeticity, the various notions of density, and other definitions of Ramsey theory can be found in, among other places, [8, Sections 1 and 2].
2.3. Partition regularity and Straus’ construction
To begin this subsection we will prove two lemmas about piecewise syndeticity and partition regularity. The following lemma states that the family of piecewise syndetic subsets of any countably infinite semigroup is partition regular. A different proof of this fact appears in [27, Lemma 7.2]. Since, as we will see, non-piecewise syndetic sets can be seen as discrete analogues of nowhere dense sets, Theorem 4 is a sort of discrete analogue to the Baire category theorem on unions of nowhere dense sets. We will not use Theorem 4 later on, but include it for completeness and to support the discussion given in the introduction and the analogy we are hoping to emphasize. A different proof, using topological dynamics, will be presented in the final section (Theorem 48).
Theorem 4.
Suppose is piecewise syndetic. Then for any finite partition into disjoint sets, there exists for which is piecewise syndetic.
Proof.
We will prove this for , and the claim will follow by induction. By Lemma 2, we may write for a syndetic set and thick . Let , so we get (since and are disjoint) and . Since either is syndetic or is thick by Lemma 1, one of and must be piecewise syndetic.555Note that this proof does not use the definition of piecewise syndeticity, but only that the family of piecewise syndetic sets is the family of intersections of two dual families (“dual” in the sense of Lemma 1). In other words, we have shown that if are any two dual families of subsets of that are upward closed, then the family is guaranteed to be partition regular. ∎
The following easy but important lemma demonstrates another connection between piecewise syndeticity and partition regularity. We record it as the general phenomenon behind why Straus’ construction produces discordant sets. We will apply this lemma by starting with a family and extending it to a -invariant family .
Lemma 5.
Suppose satisfies the following:
-
(a)
contains all thick subsets of .
-
(b)
is partition regular.
-
(c)
if and , then .
-
(d)
if and then .
Then contains all piecewise syndetic subsets of .
Proof.
Let be piecewise syndetic and find a finite set such that is thick. Then , so by (a) and (b), for some . Thus , so finally since . ∎
With this lemma, we give a simplified version of Straus’ construction to obtain a set with positive lower density, no shift of which is IP.
Theorem 6.
Let be a sequence in satisfying . Then the following set is discordant:
Proof.
We leave it as an exercise to verify that every thick subset of is IP; thus, by Hindman’s theorem and Lemma 5, it suffices to show that is not a shift of an IP set (that is, a set of the form for some IP set and ). This follows immediately from the construction of and the observation that if is IP, then for all .
To see that , we compute that
for each , since there are no more than multiples of in . ∎
For conciseness, we excuse ourselves from proving that exists, where is the set defined in Theorem 6. The reader is encouraged to look toward the sources we have mentioned or, alternatively, Theorem 23 to see how this could be done. One may also alter the set described in Theorem 6 to obtain a discordant set without invoking Hindman’s theorem as follows. Pick satisfying and set . One shows similarly that , and a totally self-contained proof that is not piecewise syndetic is not hard.
Note that Theorem 6 gives us discordant sets with lower densities arbitrarily close to . (A discordant set cannot have upper density equal to , since any set of upper density is thick.) By generalizing the notion of an IP set, a similar construction can be done in many different groups. Discussion of this construction and a complete characterization of such groups is given in [18, Chapter 4] and [7].
2.4. Preliminaries from ergodic theory and dynamics
In this section we review some basic notions and theorems of ergodic theory and dynamics. This section is utilitarian, and hence we do not venture past what we will need. For this reason we omit proofs or provide a rudimentary outline and refer the reader elsewhere.
Let be a probability space. A measurable map is measure preserving if for each . If is a countably infinite semigroup, a measure-preserving action of on is an action of on such that is measure-preserving for each . In this case, we say that is -invariant. We refer to and as measure-preserving systems.
Given a measure-preserving system, one has the following version of the von Neumann ergodic theorem.
Theorem 7.
Suppose is a countably infinite amenable cancellative semigroup with a Følner sequence and let be a measure-preserving system. Let , and let be the orthogonal projection of onto the subspace of consisting of -invariant functions. Then
Proof.
With additional constraints on the action of , more can be said. The system is called ergodic if there are no nontrivial -invariant subsets, that is, if satisfies for all , then . In this case, the orthogonal projection of onto the subspace of -invariant functions is , so Theorem 7 states that
If is a compact topological space on which acts by continuous maps, the pair is a (topological) dynamical system. If and , we define . When is or , can be thought of as the set of visiting times of to . The system is minimal if there is no proper nonempty closed subset invariant under the action of (in the sense that ). Equivalently, the action of on is minimal if, for each , its orbit is dense.
When is a compact metric space with a Borel -algebra , we say that is uniquely ergodic if there is exactly one -invariant Borel probability measure on , i.e., there exists exactly one normalized measure on which makes into a measure-preserving system. Under the condition of unique ergodicity, one has the following convenient pointwise ergodic theorem.
Theorem 8.
Let be a uniquely ergodic topological dynamical system where is a compact metric space and is a countably infinite amenable cancellative semigroup. Assume is the unique -invariant Borel probability measure and let be a Følner sequence in . Then, for each measurable whose points of discontinuity are contained in a set of measure zero and each ,
Proof.
For continuous , this can be proved in complete analogy with the well-known proof of the case ; see [36, Theorem 6.19]. For more general , the statement follows by a standard approximation argument. ∎
A fact we will find useful is that a minimal action by isometries on a compact metric space is automatically uniquely ergodic.
2.5. Cantor spaces
The following construction will be very useful. Given any sequence of finite sets having the discrete topology, their topological product is a compact space homeomorphic to the classical Cantor set. We will use this construction frequently, and we will call such a space a Cantor space. To make the notation less cluttered and easier to read, we will use boldface letters for elements of a Cantor space, e.g., . (This is also consistent with our use of for the indicator function.)
Given and for , we define the set
The collection of all sets having this form constitute an open base for the topology on .
It will be convenient to also give Cantor spaces a metric space structure. Giving each the discrete metric , the space can be made into a compact metric space via the product metric given by
(1) |
Assuming , let be the smallest integer satisfying and (take if ). Then is equivalent to the more common and conceptually simpler metric defined by . Observe that, given a collection of bijections for each , the induced map is an isometry of .
There is one more construction that we describe here. Analogously to the natural topological and metric structures on , we can endow each with the normalized counting measure. The resulting (uniform) product measure is a probability measure on the Borel -algebra of satisfying . Clearly, is invariant under the isometries described in the previous paragraph.
3. Discordant sets from visiting times
As foreshadowed by the introduction, discordant sets naturally arise from dynamical systems, one of the most familiar being irrational rotation of the one-dimensional torus . Our search for discordant sets is motivated by the fact that if is irrational and has nonempty interior, then is syndetic. The punchline of this direction of reasoning is that if is nowhere dense and has positive measure, then for almost every , the set of visiting times is discordant. Here is non-piecewise syndetic for any by Lemma 9 below, whereas the positive upper density of is guaranteed only for almost every by an application of the von Neumann ergodic theorem.666In fact, one can easily see that has positive natural density for almost every by using the pointwise ergodic theorem. Theorem 10 is a generalization of this construction which yields many discordant sets in many settings, and is the core of our analogy between discordant sets and nowhere dense sets of positive measure.
Let us note that the restriction “almost every ” is a significant and unavoidable constraint. One can construct a closed nowhere dense set of positive measure for which is empty for infinitely many .
The importance of the irrationality of is that is minimal if and only if is irrational. In our more general setting we replace this system with a minimal action of a semigroup on a compact space .
In this setting, we wish to connect properties of as a subset of with the properties of as a subset of . Lemma 9 connects the piecewise syndeticity of with the topological properties of and Theorem 10 connects the density of with the measure of , allowing us to construct discordant sets.
In analogy with our notation for subsets of , we write for and .
Lemma 9.
Let be a minimal topological dynamical system, where is a countably infinite cancellative semigroup. Then, for any nowhere dense and , the set is not piecewise syndetic.
Proof.
It is a classical fact that is minimal if and only if is syndetic for each nonempty open and ; see, e.g., [22, Theorem 1.15]. We reproduce the argument here for the convenience of the reader. The “if” direction is evident, so suppose is open and nonempty and . Let ; we claim that . Indeed, if , then for all , so cannot have dense orbit, contradicting the minimality of . By compactness, there is a finite set such that . Now fix and find such that . Then , so , proving that is syndetic.
Now let be nowhere dense and let be finite. Then , and is nowhere dense. Let be an open set lying in . Then is disjoint from the syndetic set , and so cannot be thick. Since was arbitrary, cannot be piecewise syndetic. ∎
Suppose is a topological dynamical system. If is amenable, there always exists a -invariant measure on by a version of the Bogolyubov–Krylov theorem (see [36, Theorem 6.9] for a proof of the case ; the theorem for general is proved analogously). Moreover, can taken to be ergodic by invoking the so-called ergodic decomposition.
Theorem 10.
Let be a minimal topological dynamical system, where is a countably infinite amenable cancellative semigroup with a Følner sequence . Let be a -invariant probability measure on the Borel -algebra of . If is nowhere dense with positive measure, then there exists such that is discordant. If, in addition, is ergodic, for -almost every .
Proof.
Let . By Theorem 7,
where is the orthogonal projection of onto the space of -invariant functions in , and the convergence is in . Extract a subsequence such that almost everywhere. For each ,
Hence, by the dominated convergence theorem, . Thus, there is a set of positive measure on which . Choose such that and
Then , and, since is minimal, is not piecewise syndetic by Lemma 9.
In the presence of ergodicity, is the constant function , so we may take to be any element for which . As we noted, such form a subset of of full measure. ∎
It is worth noting that, in the setting of Theorem 10, possesses a large supply of nowhere dense sets of positive measure if it is an infinite metric space.
Theorem 11.
Proof.
We first claim that is a non-atomic measure. Let . Then for any . Since is minimal, is dense in , and is in particular infinite. It follows that since is a probability measure.
Now let be a countable dense subset of . Since is non-atomic, for all . Hence by the continuity of from above, we can find an open neighborhood of with . Set , so that is a nowhere dense subset of with . The claim now follows from a theorem of Sierpiński [Si] which states that in any non-atomic measure space with , the measure takes every value in the interval . ∎
4. Discordant sets and generalizations of squarefree numbers
It is known that the set of squarefree numbers in is discordant, and in this section we describe a wide-ranging generalization of this fact and examine in detail some special cases.
4.1. A characterization and a construction
We will prove non-piecewise syndeticity in two ways; the first uses a characterization (Lemma 12) of non-piecewise syndetic sets, and the second uses Lemma 9. Ramsey theory aficionados will notice that Lemma 12 can be rephrased as a characterization of sets satisfying the dual property, in the sense of Lemma 1, to piecewise syndeticity, often denoted by PS∗.
Lemma 12.
Let be a countably infinite semigroup and let . Then is not piecewise syndetic if and only if, for any finite , there exists a syndetic such that .
Proof.
Let be finite and consider the set . Observe that , since if and only if for some . Thus, by Lemma 1, is syndetic if and only if is not thick. In particular, is syndetic for all finite if and only if is not piecewise syndetic. ∎
Before constructing non-piecewise syndetic sets using Lemma 12, we first recall a general group-theoretic form of the Chinese remainder theorem.888A generalization of this form of the Chinese remainder theorem is described in the StackExchange discussion SE2954401, although we will not need this generality.
Lemma 13 (Chinese remainder theorem).
Let be a group with normal subgroups , where and are pairwise coprime. Define and by . Then is an isomorphism.
Proof.
It is immediate from the first isomorphism theorem that is a well-defined injective homomorphism. So, it suffices to show that . Indeed, for each , divides since . Thus the inequality holds since are pairwise coprime. ∎
We now combine the Chinese remainder theorem with Lemmas 12 and 9 to obtain an abundance of non-piecewise syndetic sets. Although Lemma 12 shows that we can remove “thicker and thicker” syndetic sets to produce non-piecewise syndetic sets, Theorem 14 shows that in special situations (when the syndetic subsets are suitably “arithmetically/algebraically independent”, as the Chinese remainder theorem will guarantee) this is not necessary.
Theorem 14.
Let be a countably infinite group and be a sequence of normal finite-index subgroups of such that for any , the subgroups satisfy the hypotheses of Lemma 13. Then, for any sequence in , the set is not piecewise syndetic.
In principle, the set may be empty. Below we will see many examples where this is not the case. Notice that we can equivalently state the conclusion of Theorem 14 as “every piecewise syndetic subset of contains an element (in fact, infinitely many elements) of ”. We will give two proofs of this fact.
Proof 1.
Fix . Let , and let be the isomorphism guaranteed by the Chinese remainder theorem. Then, setting , we have for each , so . Since the subgroups have finite index, is syndetic, so Lemma 12 completes the proof. ∎
Our second proof of Theorem 14, though longer, uses Lemma 9 and strengthens our analogy between discordant sets and nowhere dense sets of positive measure. In the remainder of this subsection we will use the definitions made in Subsections 2.4 and 2.5.
Proof 2.
Consider the Cantor space . Given , define
(2) |
Sets having the form given in (2) constitute an open base for the topology on .
Let act by componentwise multiplication on , i.e., . The Chinese remainder theorem guarantees that this action is minimal. To see this, let and be given. By the Chinese remainder theorem, find so that . Then since . This proves that has dense orbit.
As one might expect from this second proof, we can obtain a density result for Theorem 14 by applying Theorem 10.
Theorem 15.
Let , , and be as in Proof 2 of Theorem 14, and set for each . Suppose is amenable, possessing a Følner sequence . Let be the probability measure on the Borel -algebra of satisfying . Finally, given , set . Then, for -almost every ,
(4) |
Proof.
4.2. Some number-theoretic examples of discordant sets
In this subsection we apply Theorem 14 to construct some discordant sets. In the following subsection we provide proofs of positive density for all but one of the constructions. The most primitive example is the following.
Example 16.
The set of squarefree integers is not piecewise syndetic:
It is classical that , so that is discordant.
Example 16 can be generalized a great deal. We provide five generalizations below.
Example 17.
Let be a sequence of pairwise coprime positive integers and let denote the set of -free integers
consisting of integers not divisible by any element of . Then is not piecewise syndetic, and if , then [25, Theorem 9, Section V.5]. We provide a proof of this fact below (Theorem 23). A well-known but notable special case is that the set of -free integers, those divisible by no perfect th power, has density .
Example 18.
Let be a finite extension of , let , and let denote the -free integers :
where denotes the ring of integers in . Using an appropriate ring-theoretic version of the Chinese remainder theorem, we may apply Theorem 14 to show that is not piecewise syndetic. Moreover, as is a finite extension, as additive groups, where is the degree of . Define to be a sequence of balls (pulled back from to under this isomorphism), with respect to the norm, centered at and having increasing radius. Then as shown in [17, Corollary 4.6], where denotes the Dedekind zeta function of . Equivalently, , where denotes the absolute norm of , i.e., the cardinality of the residue field . Hence is discordant.
Example 19.
Given and a prime , write to denote the exponent on in the prime factorization of . In [4, Theorem 3.9], it is shown in a somewhat cumbersome way that, for any infinite set of primes and any corresponding collection of nonempty subsets of , the set
is not piecewise syndetic. This strengthens the non-piecewise syndeticity result of Example 16 and, in certain cases, Example 17, since it is clear that is a superset of the -free numbers. Below, we state this result in better generality and deduce it easily from Theorem 14.
Example 20.
The set consisting of pairs of coprime integers can be expressed as
making clear that is not piecewise syndetic by Theorem 14. It is well known that , which will follow from the more general Theorem 26 (in other words, the probability that two “randomly chosen” integers are coprime is ; see, e.g., [26, Theorem 332]). Many possible generalizations in this direction can be imagined.
Example 21.
Our final example will take place in a noncommutative setting. Consider the Heisenberg group , the matrix group
Associating these matrices with triples , we view as with the group operation . For each , the kernel of the entrywise reduction map is a normal subgroup of of index . Hence, for any sequence of pairwise coprime positive integers, the set is not piecewise syndetic in by Theorem 14. We will see in Theorem 27 that this set has density with respect to an appropriate Følner sequence in .
4.3. Computing density
Before computing the density of some of the non-piecewise syndetic sets we constructed in Subsection 4.2, we make some definitions and prove a general lemma.
Suppose we have a countably infinite amenable cancellative semigroup with a Følner sequence . Given a function , we define
and likewise and in the obvious way. We will be most interested in the case when exists. Assuming this limit exists, it has some obvious but useful properties. First, given , we have . Second, if satisfy , then . Finally, is finitely additive: .
In order to prove that a set of the form has density, we will impose a number of nice properties on the family . We adopt the notation for . In addition, we use to refer to the set of -element subsets of .
Specifically, we say that is -inclusion-exclusion good, or -I.E. good if
-
(a)
exists for each , and ;
-
(b)
for every infinite ;
-
(c)
for every finite , ; and
-
(d)
for each , .
For -I.E. good families, the following lemma states that an asymptotic version of the inclusion-exclusion principle holds.
Lemma 22.
Suppose and are as above and is a -I.E. good family of subsets of . Write . Then .
Proof.
By property (c) of -I.E. goodness and the standard inclusion-exclusion principle, , since for each . Now let . Using the identity
which holds as long as , properties (a) and (c) of -I.E. goodness yield
Thus will follow if we can demonstrate that
To do this, it suffices, by the finite additivity of and property (d) of -I.E. goodness, to show
(5) |
Let denote the function on the right side of (5) and let . If , then clearly . Otherwise, due to property (b) of -I.E. goodness, we may suppose is contained in exactly elements of . Then
so indeed . This verifies that . ∎
Using this lemma, we compute the density of sets constructed in Examples 17 and 19. The following two theorems use the standard Følner sequence in , which (as with the notation ) we omit reference to.
Theorem 23.
Let be a sequence of pairwise coprime positive integers satisfying . Then .
Proof.
We prove this statement for for convenience, but it is of no consequence, since is symmetric about . We must show that is I.E. good (with respect to the standard Følner sequence as mentioned above). Only property (d) is not immediate. Notice that it suffices to prove the case , since any intersection of sets of the form again has this form.
Write . Since , clearly . To prove , the key point is that, for each ,
which is just a statement of the fact that there are no more than multiples of in . It follows that, for each ,
whence . Thus is I.E. good. ∎
Theorem 24.
Let be a sequence of pairwise coprime positive integers and let be a sequence of positive integers for which . Then
Proof.
As in Theorem 23 we confine ourselves to . For each define , so that . We aim to show that is an I.E. good family. (The density is easy to compute but is also derived below.) Clearly property (b) of I.E. goodness is satisfied.
Given , define also , , and . Observe that . Define to be the set of positive integers not divisible by for any . Observe that is -periodic, i.e., for each , one has if and only if . In particular, . Now consider the isomorphism afforded by the classical Chinese remainder theorem. An element is in if and only if is nonzero in every coordinate, from which it follows that . From this we get
proving that property (c) above holds for . In addition, the formula we have obtained for and the hypotheses on and verify property (a).
Now fix and set . As in Theorem 23 we have . To prove that , fix . Select such that
Now let and note that, since where is -periodic, we have, for each and ,
Find for which and set . Then, for each ,
(we have since ). This verifies that is indeed I.E. good and completes the proof. ∎
We next compute the density of the sets constructed in Examples 20 and 21. To do so, we will need a technical lemma.
Lemma 25.
Let and be countably infinite amenable cancellative semigroups with Følner sequences and , respectively. Let be a -I.E. good family of subsets of and be a -I.E. good family of subsets of . Then is a -I.E. good family of , where .
Proof.
The fact that is a Følner sequence for is easy and left to the reader. We outline the proof that satisfies property (d) of -I.E. goodness; the others are straightforward. We have, for each ,
But since
and likewise with , proving property (d) reduces to verifying the following claim: given doubly-indexed sequences and of nonnegative quantities satisfying
-
•
and for each , and
-
•
and ,
one has . ∎
Theorem 26.
Let and let be sequences such that, for each , consists of pairwise coprime positive integers. Then
To apply Lemma 25 to the Heisenberg group , viewed as with the group operation described in Example 21, a laborious but straightforward computation shows that
(6) |
is a Følner sequence in .
Theorem 27.
Let be a sequence of pairwise coprime positive integers. Then, letting be as in (6), we have
5. Discordant sets in
Let denote the free semigroup generated by and with identity . It is well known that is not amenable. To see why, suppose is a Følner sequence for , and form the partition . One of these parts must have positive upper density with respect to , and since does not, suppose without loss of generality that . Then are disjoint subsets of each with upper density . But Lemma 3 guarantees that this is impossible, which is our desired contradiction. This argument can be modified without too much difficulty to show that nonabelian free groups are not amenable, and moreover any group possessing a nonabelian free subgroup is not amenable.
We now turn our attention to the group . Since contains nonabelian free subgroups,999For example, it is well-known that the matrices and generate a free subgroup of (see [33]). it is not amenable. However, we will define a natural density-like notion for subsets of and construct sets which are large with respect to this notion but not piecewise syndetic. Although this density does not enjoy shift-invariance (which in the amenable case is assured by Følner sequences), we consider this result an interesting aside which has independent value. In particular, it makes more robust the analogy with nowhere dense sets of positive measure, showing that it extends beyond the amenable case.
We first leverage Theorem 14 to produce an ample supply of non-piecewise syndetic subsets of . Recall the congruence subgroups of , which are subgroups of the form
The subgroup is normal in for each since it is the kernel of the homomorphism defined by reducing each matrix entrywise modulo .
Theorem 28.
Let be a sequence of pairwise coprime positive integers, and define . Then is not piecewise syndetic.
Proof.
Fix any coprime . Since , to apply Theorem 14 it suffices to show that . We claim that we may always find, for any , some such that and . Granted this, fix , and find such that , , , and . Then , so there exists such that . Hence , as desired.
To prove the claim, let , treating the entries as elements of . Find such that and set
so and . Note that . It follows that . Letting denote the product of primes dividing but not , we see that and are coprime. Therefore, there exist such that . It also follows from that there exists such that
Then
is an element of satisfying and . ∎
We will now introduce a notion of density for subsets of . Define
We then define the upper density of to be
and define and analogously. We first make some crude estimates on the growth of the sets and .
Lemma 29.
There is a positive integer such that for all .
Proof.
We will produce, for sufficiently large , at least matrices such that and . As we saw in Example 20,
Thus there is a positive integer such that for all , there are at least pairs such that and . Given such a pair , using the Euclidean algorithm, we can find with and such that . Then , as desired. ∎
Lemma 30.
For each and , we have .
Proof.
Define
Given , either or the matrix given by swapping the rows and columns of is in . Thus .
There are at most pairs with such that and . In particular, there are at most such pairs with . Given such a pair, one can find, using the Euclidean algorithm, such that , , . Then
so there are at most three pairs such that and , and in particular there are at most three such that .
Combining all of this, we have . ∎
With these estimates, we can show that under certain conditions, the sets in Theorem 28 have positive lower density.
Theorem 31.
Let be a sequence of pairwise coprime positive integers such that . Then has positive lower density.
6. Recurrence and anti-recurrence
Let be a countably infinite group. A set is a set of (measurable) recurrence if for all measure preserving systems and for all with , there is a not equal to the identity such that . Recall that in a measure-preserving system, we assume , so sets of recurrence exist (for example, take ).
Call an anti-recurrence set, or an AR set, if for all , is not a set of recurrence. We will begin by showing that any piecewise syndetic set is a shift of a set of recurrence, so any AR set with positive upper density is therefore discordant.
Lemma 32.
Let be thick. Then is a set of recurrence.
Proof.
Let be thick. First, we will show that there is an infinite sequence of elements of such that for all pairs with , we have . Indeed, such a set can be constructed recursively: given , we can find such that .
Let be a measure preserving system, and let have . There exist with such that . Then, , as desired. ∎
Lemma 33.
Let be a set of recurrence partitioned as . Then at least one is a set of recurrence.
Proof.
Assume that no is a set of recurrence. Then, for each , there is a measure preserving system and with such that for all we have . The actions of on each induce an action on , which is measure preserving with respect to the product measure of the measures . Let . We have , but for all , since for some , we have , and thus . Therefore, is not a set of recurrence. This establishes the contrapositive. ∎
Theorem 34.
If is AR, then is not piecewise syndetic.
Proof.
Apply Lemma 5 with , noting the previous two lemmas. ∎
In [5], it is noted that Straus’ example of a discordant set is also an AR set. We will follow their terminology and refer to AR sets with positive upper density as Straus sets. In what follows, we investigate the question of whether a given discordant set is a Straus set. We begin by showing that Example 17 provides large class of non-examples.
Theorem 35.
Let be a sequence of pairwise coprime positive integers such that . Define . Then is a discordant set which is not AR.
Proof.
Not all countably infinite amenable groups admit Straus sets. For example, in the group of finitely supported even permutations of , every set with positive upper Banach density is a set of recurrence (in fact, such a set must be IP). In [5, Theorem 1.23], it is shown that Straus sets exist in a locally compact, second countable, amenable group if and only if the von Neumann kernel of that group is not cocompact. In such groups, for all , there is a Straus set with . Thus, it is natural to ask whether such a group has a Straus set with for each . We prove that this is the case for all countably infinite abelian groups in Theorem 38. First, we need the following lemma, a special case of a lemma appearing in [5].
Lemma 36 (cf. [5, Lemma 5.1]).
Let be a countably infinite amenable group and let be a Følner sequence in . Let be a sequence of subsets of such that exists for every , , and exists for every . Then there are cofinite subsets such that for , .
Before proving Theorem 38, we give a proof of the case which does not rely on tools from ergodic theory. This also serves as a model of the proof of Theorem 38 for arbitrary discrete countably infinite abelian groups.
For the proof of Theorem 37, the notion of well-distribution for a sequence in will be used. For a more detailed exposition of this theory, see [31]. A sequence in is called well-distributed if for any measurable with ,
where is the Lebesgue measure and is the boundary of , i.e., the closure of minus the interior of . Equivalently, is well-distributed if for any Følner sequence in and any measurable with ,
Theorem 37.
Let and let be a Følner sequence in . Then, there is a Straus set such that .
Proof.
Let , and let be irrational. For a set , let denote . For , let , let
and let . Since each of and have boundaries with measure zero, and since is well-distributed in , we have and for all . Since , by Lemma 36, there are cofinite subsets such that
Let , and let . For all ,
so the convergence is uniform. Since the boundary of each interval has measure zero, the functions are continuous, so is continuous. Moreover, (since ) and , so we can choose so that . Let . Then, . It remains to show that is AR. Let . Then
is finite, so cannot be a set of recurrence. ∎
In fact, a version of Theorem 37 holds true for all countably infinite abelian groups.
Theorem 38.
Let be a (discrete) countably infinite abelian group. Let , and let be a Følner sequence in . Then, there is a Straus set such that .
Proof.
We will construct a compact abelian group and a homomorphism with dense image. Depending on , this construction can occur three different ways.
First, assume that has an element of infinite order, i.e., is not a torsion group. We define a homomorphism from to as follows. By [20, Theorems 23.1 and 24.1], can be embedded in , where is a torsion group. Let be the embedding from into this direct sum. Let be an element of with infinite order. Then, there is a projection from onto such that is nonzero. Choose an irrational . Then the map is a homomorphism from to with dense image, so in this case, we will let and let be such that .
Next, assume that is a torsion group. For a prime and , let be the Prüfer -group . (By definition, for , .) By [20, Theorems 23.1 and 24.1], can be embedded into . Let be the corresponding inclusion map, and let be the projection from the direct sum onto . For convenience, we abbreviate . There are two cases.
First, assume there is some pair such that is infinite. Note that is isomorphic to the subgroup of ; call this isomorphism . Let . Then is a homomorphism with dense image, so we set .
On the other hand, assume that for all pairs , is finite. Then is embedded in . Since is a subgroup of a direct sum of finite cyclic groups, by [20, Theorem 18.1], there is an infinite sequence of integers greater than such that . We will assume without loss of generality that for all . Then there is a dense embedding from to the group . For each , define a metric on by , where and and are the equivalence classes of and in . Observe that for all . We can give the product metric . The Haar measure on is the product measure of the normalized counting measures on each . We will show that for any , the set has measure zero. Since is invariant under the group operation of , this will imply has measure zero for all .
Let ; we first claim that there is at most one sequence with for all and
(7) |
Assume there are two distinct such sequences . Let be the smallest positive integer with . Then
However,
so we cannot have
Now assume satisfies (7). We will show
Let be the normalized counting measure on , so for any . If , then , and if , then . Thus for all ,
so
and thus . If there is no sequence with satisfying (7), we also have .
Here, the casework ends, and for , the following is true. We have a homomorphism with dense image , where is a compact abelian group. Moreover, is endowed with an invariant metric such that for and , we have , where is the normalized Haar measure on . Additionally, for all .
We define a natural -action on by isometries via . Since is dense, the action is uniquely ergodic, with the normalized Haar measure as the unique invariant measure. Thus, by Theorem 8, for any for which the set of discontinuities has measure zero and ,
(8) |
Suppose is a Borel set with , and write . Taking and , equation (8) becomes
since the set of discontinuities of is exactly .
The remainder of the argument proceeds along the same lines as in Theorem 37. Enumerate , let , and for , let . Let
and let . Using Lemma 36, for each , find a cofinite such that
Let . Note that since the construction of the sets depends continuously on , the functions are continuous. Let . The convergence of to is uniform, so is continuous. Note that (since ) and , so we can choose so that . Then, if , we have . Let . Then is finite, so cannot be a set of recurrence. ∎
Theorem 38 can be extended to countably infinite commutative cancellative semigroups, allowing us to conclude that discordant subsets with any density in exist in, e.g., and as well.
Corollary 39.
Let be a countably infinite commutative cancellative semigroup, let be a Følner sequence in , and let . Then there is a discordant set with .
Proof.
We can embed in a countably infinite abelian group by a process similar to that of embedding an integral domain in a field (see [30, Exercise 1.1.8] for more details about this construction). The elements of can be thought formally as “fractions” of elements of , so for any , there is an such that . We will show that is a Følner sequence in (see [6, Theorem 2.12]). Let , and let be such that . For each ,
so
and thus is a Følner sequence in . By Theorem 38, there is a discordant set with . Set . We claim that is not piecewise syndetic in . Suppose, toward the contrapositive, that there is a finite such that is thick. Let be finite and select such that . Then there is an such that . But then is piecewise syndetic in , which establishes the claim. Thus is a non-piecewise syndetic subset of with . ∎
7. Large subsets of
Let be a countably infinite semigroup. In this section we will identify a subset with its indicator function . We will say that is thick, syndetic, etc. whenever possesses these properties. Thus, families of subsets of are identified with subsets of . To study the topological properties of subsets of , we give the product topology, making it a Cantor space. In this space, a base for the topology is given by the cylinder sets, which have the form
for finite, disjoint . Let us write
7.1. Background and motivation
Recall that if is a topological space homeomorphic to the Cantor set, it is a Baire space, meaning that it cannot be written as a countable union of nowhere dense sets. A subset of which is a countable union of nowhere dense sets is called meager, and a subset of whose complement is meager is comeager. The fact that is a Baire space means that no subset can be both meager and comeager, so one can think of the elements of a comeager set as being “topologically generic”. We wish to stress in this section the analogy between topology and measure theory in which “meager” corresponds to “measure 0” and “comeager” corresponds to “measure 1”.
We will describe comeager subsets of , whose elements can thus be thought of as “topologically generic” subsets of . Let us first motivate our results by examining them in a familiar setting. Consider the case , where is the space of infinite binary sequences. An infinite binary sequence is called disjunctive if every finite binary word occurs within it. (Formally, is disjunctive if, for each and , there exists such that for each .) The subset of consisting of disjunctive sequences is large both topologically and measure-theoretically: it is comeager and of measure 1.
However, the topological and measure-theoretical worlds diverge once we consider the limiting frequencies with which finite subwords occur. Given a sequence and a finite word , one says that occurs in with a limiting frequency
(9) |
It is a classical result of Émile Borel [14] that the set of normal sequences, those for which every word of length occurs as a subword with limiting frequency , has measure 1. Moreover, as we will see, it is a meager set.
On the other hand, one can consider sequences for which limits analogous to (9) (but modified slightly to avoid counting overlapping sequences) fail to exist as badly as possible. For these sequences, the of the frequency of any finite word is 1 and the corresponding is 0. We will call such sequences extremely non-averageable (or ENA). It turns out that the set of ENA sequences, in contrast to the set of normal sequences, is comeager [16, Theorem 1].
We will make the following generalizations of the facts we have just stated: assuming only cancellativity, we prove an analogue in a countably infinite semigroup to the statement that the set of disjunctive sequences is comeager (Theorem 40); assuming also amenability, we prove an analogue to the statement that the set of ENA sequences is comeager (Theorem 42). A generalization of measure-theoretical normality to countably infinite amenable semigroups and the corresponding results can be found in [6].
Since the set of normal sequences and the set of ENA sequences are disjoint, we see also that the set normal sequences is meager, and that the set of ENA sequences has measure 0 (see [6, Proposition 4.7] regarding the former statement in arbitrary amenable cancellative semigroups). This allows us to give another motivation for the results of this section. One reason we gave for studying discordant sets is that they illustrate how topological and analytic notions of largeness for subsets of diverge; in this section, we construct subsets of on which basic topological and analytic notions of largeness disagree.
Finally, we remark that, although we prove the results of this section in the space , their generalizations to (which can be thought of as the space of -part partitions, or -colorings, of ) are apparent and can be achieved with the same arguments.
7.2. Disjunctive elements of
In this subsection we will assume that is cancellative. Call disjunctive if for any there is some for which .101010In the notation of Section 8, we can equivalently say that is disjunctive if and only if, for every , there exists for which . This definition has the benefit of making sense for non-cancellative semigroups. Cancellativity here guarantees that for all . Note that, for , this definition specializes to the definition of disjunctive sequences we gave above.
Theorem 40.
The set of disjunctive elements of is comeager.
Proof.
For any , let , so that the set of disjunctive elements of is
Now let . Since is cancellative and infinite, we can find some for which . Then we have so that is dense, since it has nontrivial intersection with each element of the open base for . Since is a union of open sets, we have written as a countable intersection of open dense sets, which proves that is comeager. ∎
We deduce as a relevant consequence of this theorem that the set of non-piecewise syndetic elements of is meager, since any disjunctive element of is thick. If is non-cancellative, the set of piecewise syndetic elements of may fail to be comeager. For example, if there exists such that for all , one has , then is piecewise syndetic if and only if , so the family of piecewise syndetic sets cannot be dense.
7.3. Extremely non-averageable elements of
In this subsection we will assume that is both cancellative and amenable. Fix a Følner sequence in . We will begin by defining -normality of elements of , as defined in [6], for comparison with the definition we will give for ENA. We will also utilize the notion of -normality in Theorem 43.
Given , one says that is -normal if, for each finite and ,
The following classical lemma regarding Følner sequences, which states that they have a property analogous to thickness, assures us that the notion of -normality is meaningful.
Lemma 41.
Let be a Følner sequence in a countably infinite cancellative semigroup . Let be finite. Then there exists such that, for each , for some .
Proof.
Find large enough that, for each and ,
Then , so let be an element of this set. Then . ∎
Now we define extreme non-averageability in . The following definitions are admittedly unwieldy because we wish to measure the frequency of occurrences without overlap of our finite sequence. This is because we are interested in the maximal frequency with which the finite sequence can occur, and when allowing for overlaps, this frequency is difficult to predict. We would be pleased if a simpler equivalent definition of ENA in an arbitrary countably infinite amenable cancellative semigroup could be found.
Fix a finite . Given a finite set , define
Now let , and , and define
We observe that is well-defined since for only finitely many by Lemma 41. Define to be the corresponding . Hence, and give the (upper and lower) limiting frequencies, measured along , with which the finite binary word occurs without overlaps in . For convenience, we write
We say that is -ENA if and for each .
With these definitions in hand, we proceed to the main result of this subsection.
Theorem 42.
The set of -ENA elements of is comeager.
Proof.
For each , let
so that the set of -ENA elements of is . Fix ; it suffices to show that is comeager.
Find a subsequence of which satisfies
Define and, for , ; let denote the Følner sequence . Fix . Now for each choose and define
The definition of ensures that the sets are nonempty for all large enough , and they are clearly open since they are cylinder sets. In particular, , and either or , depending on whether is or . The set is chosen this way to ensure that for all , there is no such that for all . In particular, for all . Finally, define for all .
Let . Since the sets are disjoint and and are finite, cancellativity gives that for all large enough , and hence . It follows that is open and dense. Define , so that is a comeager subset of . Moreover, if , then
(10) |
by construction. Since , (10) must hold with respect to in addition to , and thus also with respect to . Hence is a comeager subset of , as we had hoped. ∎
It is important to note that the property of -ENA, as well as -normality, is dependent on . The following theorem shows that this is true in a strong sense.
Theorem 43.
Let be disjunctive. Then there exist Følner sequences and in such that is simultaneously -normal and -ENA.
Proof.
Let be a Følner sequence of , and suppose are -normal and -ENA, respectively. For each and , define and and find such that . Then define the Følner sequences by . These Følner sequences have the desired properties. ∎
8. Topological dynamics and piecewise syndeticity
Let denote any countably infinite (not necessarily cancellative) semigroup. In this section we explore piecewise syndeticity of subsets of in terms of dynamical properties of elements of and their orbit closures. Some of the results presented below are well-known for subsets of . Using these results we revisit and examine in greater detail the comparison between van der Waerden’s theorem and Szemerédi’s theorem made in the introduction.
We adopt the notation of the previous section. For each , let be the map . We will consider the dynamical system , where acts on by . In the familiar settings and , this is just the left-shift action on infinite one- or two-sided sequences. In light of the natural correspondence between and , we may view this dynamical system instead as . In this case, acts on by . This action is continuous, since for any , we have (for this identity to hold in the non-cancellative case, we adopt the convention that if , in view of the fact that if and only if ) .
Given , we define the orbit closure of to be
Now fix an increasing sequence of finite subsets of such that . Let us note that for any , we have if and only if, for all , there exists such that for each . This simple but crucial property of orbit closures in the system follows immediately from the definition of the product topology on .
We illustrate what this means in the familiar case . In this setting, if and only if, for any , can be shifted to agree with on . Put another way, means that one can find arbitrarily long initial segments of as subwords of .
Before stating the results, we give convenient reformulations of the notions of syndeticity, piecewise syndeticity, and thickness in terms of indicator functions. An element is syndetic if and only if there exists a finite set such that, for each , we have for some . Similarly, is piecewise syndetic if and only if there exists a finite and an infinite sequence in such that, for each and , we have for some . Finally, is thick if and only if , so by Lemma 1, is syndetic if and only if .
Lemma 44.
An element is piecewise syndetic if and only if the orbit closure contains a syndetic element.
Proof.
Suppose is piecewise syndetic and find a finite set and an infinite sequence in which witnesses ’s piecewise syndeticity, in the sense of the previous paragraph. We claim that there is a decreasing sequence of infinite subsets of with the property that for each and . Indeed, for any , there are finitely many functions , so such a sequence can easily be constructed inductively by applying the pigeonhole principle. Hence we can define by for any and . By construction, . To show that is syndetic, let , find such that , and find with . Since , we can find such that . Then , so is syndetic.
Now suppose and is syndetic. Let be a finite set which witnesses ’s syndeticity. Fix and find such that . Find such that for each . Let and find for which . Then , so . This proves that is piecewise syndetic. ∎
Recall that subsystem of a dynamical system is a nonempty closed -invariant subset of . A minimal subsystem is a subsystem which is itself a minimal dynamical system. By Zorn’s lemma, any dynamical system has a minimal subsystem (see, e.g., [24, Theorem 2.22]). We will call the subsystem of consisting of the fixed point (the zero function) the trivial subsystem.
The following theorem is a restatement of Lemma 44 in terms of the minimal subsystems of orbit closures.
Theorem 45.
An element is piecewise syndetic if and only if possesses a nontrivial minimal subsystem.
In particular, if is discordant, does not have a nontrivial minimal subsystem. Moreover, sets of positive upper density having this property are exactly the discordant sets.
Proof.
Suppose is piecewise syndetic. By Lemma 44, there exists which is syndetic. Then and , so has a nontrivial minimal subsystem. It follows that has a minimal subsystem as well.
Conversely, suppose possesses a nontrivial minimal subsystem . Then , so every element of is syndetic. Thus, by Lemma 44, is piecewise syndetic. ∎
The existence or nonexistence of nontrivial minimal subsystems in has significant combinatorial consequences, which we presently examine.
Let us specialize to the case . As stated in the introduction, there is an equivalent formulation of van der Waerden’s theorem as a topological recurrence theorem:
Theorem 46 (Topological van der Waerden’s theorem).
Let be a minimal111111In the sources, one will not see the condition of minimality explicitly imposed in the formulations of Theorem 46. We of course do not lose generality by making this assumption as we may always pass to a minimal subsystem. Moreover, the topological proofs of this theorem of which the authors are aware, with the notable exception of [9, Proposition L], use minimality in a crucial way. topological dynamical system where is a compact metric space. Then for any and there exists and such that the points are within of each other.
Proof.
From here the classical van der Waerden’s theorem can be derived along the following lines. Suppose is piecewise syndetic. Then possesses a nontrivial minimal subsystem by Theorem 45. Since , every element of is syndetic. Giving the product metric described in Subsection 2.5 and applying Theorem 46 inside shows that, for any , the system possesses an element with a length- arithmetic progression. Since , for any , there is an element of with a length- arithmetic progression. It follows that is AP-rich. (As we saw in the introduction, this is equivalent, via the partition regularity of piecewise syndetic sets, to the formulation of van der Waerden’s theorem in terms of partitions of .)
This discussion supports our position taken in the introduction on the distinction between van der Waerden’s and Szemerédi’s theorems. Specifically, if the orbit closure of any element of having positive upper density were to possess a nontrivial minimal subsystem, then Szemerédi’s theorem would follow immediately from the topological version of van der Waerden’s theorem using the argument we gave in the preceding paragraph.
As our next combinatorial application of Theorem 45, we will prove the partition regularity of piecewise syndetic sets from a dynamical perspective (cf. Theorem 4). We first address the following question: For which is itself minimal? The answer is that every finite subword occurring in must occur syndetically in , as the following lemma states. The result is well known in the case when is or ; see, e.g., [22, Proposition 1.22].
Lemma 47.
Let . Then is minimal if and only if, for all , the set is either empty or syndetic.
Proof.
This is a consequence of the following general phenomenon: Let be a topological dynamical system where is a metric space, and let . Then is minimal if and only if, for each open , is either empty or syndetic. Clearly, if is minimal, then this condition holds (see the proof of Lemma 9).
Conversely, suppose . We will show that , which will prove that is minimal. To this end, fix and find a finite set such that . Find a sufficiently small so that for each . Next find for which and find satisfying . Then , i.e., . This proves that , as claimed. ∎
Before continuing, we remark that, as in the previous section, the preceding can be easily generalized to statements about . In this setting, we say that is (piecewise) syndetic if is (piecewise) syndetic in the usual sense. We continue to refer to as the trivial subsystem. We leave the proofs of the generalizations to the interested reader. Using this, we give the promised result on the partition regularity of piecewise syndetic sets. The following proof in the case can be found in [22, Theorem 1.24].
Theorem 48.
Let be piecewise syndetic and suppose it is partitioned as . Then is piecewise syndetic for some .
Proof.
It is worth mentioning that Theorem 48 is easily proved with the help of topological algebra in , the Stone–Čech compactification of , that is, the space of ultrafilters on . It is well known that is piecewise syndetic if and only if some shift lies in a minimal idempotent of [30, Theorem 4.43]. From here partition regularity follows from the definition of an ultrafilter. This approach to Ramsey theory—learning about subsets of by studying —has proved very fruitful, leading to, for example, a slick proof of Hindman’s theorem (see [30, Corollary 5.9] or [3, Theorem 1.2]). We do not pursue this further but direct the curious reader towards the references we have just mentioned.
References
- [1] L. Argabright and C. Wilde “Semigroups satisfying a strong Følner condition” In Proceedings of the American Mathematical Society 18.4, 1967, pp. 587–591
- [2] V. Bergelson “Ergodic theory and diophantine problems” In Topics in Symbolic Dynamics and Applications, London Mathematical Society Lecture Note Series Cambridge University Press, 2000, pp. 167–206
- [3] V. Bergelson “Minimal idempotents and ergodic Ramsey theory” In Topics in Dynamics and Ergodic Theory, London Mathematical Society Lecture Note Series Cambridge University Press, 2003, pp. 8–39
- [4] V. Bergelson, M. Beiglböck, N. Hindman and D. Strauss “Multiplicative structures in additively large sets” In Journal of Combinatorial Theory, Series A 113.7, 2006, pp. 1219–1242
- [5] V. Bergelson, J.. Christopherson, D. Robertson and P. Zorin-Kranich “Finite product sets and minimally almost periodic groups” In Journal of Functional Analysis 270.6, 2016, pp. 2126–2167
- [6] V. Bergelson, T. Downarowicz and M. Misiurewicz “A fresh look at the notion of normality” In Annali della Scuola Normale Superiore di Pisa, Classe di Scienze, To appear
- [7] V. Bergelson and H. Furstenberg “WM groups and Ramsey theory” In Topology and its Applications 156.16, 2009, pp. 2572–2580
- [8] V. Bergelson, N. Hindman and R. McCutcheon “Notions of size and combinatorial properties of quotient sets in semigroups” In Topology Proceedings 23, 1998, pp. 23–60
- [9] V. Bergelson and A. Leibman “Set-polynomials and polynomial extension of the Hales-Jewett theorem” In Annals of Mathematics 150, 1999, pp. 33–75
- [10] V. Bergelson, A. Leibman and E. Lesigne “Intersective polynomials and polynomial Szemerédi theorem” In Advances in Mathematics 219, 2008, pp. 369–388
- [11] V. Bergelson and R. McCutcheon “An ergodic IP polynomial Szemerédi theorem” In Memoirs of the AMS 146.695, 2000, pp. 1–106
- [12] V. Bergelson and I. Rusza “Squarefree numbers, IP sets, and ergodic theory” In Paul Erdős and his Mathematics 4, 5 Janos Bolyai Mathematical Society, 2002, pp. 147–160
- [13] A. Bernardino, R. Pacheco and M. Silva “The gap structure of a family of integer subsets” #P1.47 In The Electronic Journal of Combinatorics 21.1, 2014
- [14] É. Borel “Les probabilités dénombrables et leurs applications arithmétiques” In Rendiconti del Circolo Matematico di Palermo 27, 1909, pp. 247–271
- [15] T.. Brown “An interesting combinatorial method in the theory of locally finite semigroups” In Pacific Journal of Mathematics 36, 1971, pp. 285–289
- [16] C. Calude and T. Zamfirescu “Most numbers obey no probability laws” In Publicationes Mathematicae Debrecen 54 (Supplement), 1999, pp. 619–623
- [17] F. Cellarosi and I. Vinogradov “Ergodic properties of -free integers in number fields” In Journal of Modern Dynamics 7.3, 2013, pp. 461–488
- [18] J.. Christopherson “Closed ideals in the Stone-Čech compactification of a countable semigroup and some applications to ergodic theory and topological dynamics”, 2014
- [19] P. Erdős “A survey of problems in combinatorial number theory” In Annals of Discrete Mathematics 6, 1980, pp. 89–115
- [20] L. Fuchs “Infinite Abelian Groups” Elsevier Science, 1970
- [21] H. Furstenberg “Ergodic behavior of diagonal measures and a theorem of Szemerédi on arithmetic progressions” In Journal d’Analyse Mathématique 31, 1977, pp. 204–256
- [22] H. Furstenberg “Recurrence in Ergodic Theory and Combinatorial Number Theory” Princeton University Press, 1981
- [23] H. Furstenberg and B. Weiss “Topological dynamics and combinatorial number theory” In Journal d’Analyse Mathématique 34, 1978, pp. 61–85
- [24] W.. Gottschalk and G.. Hedlund “Topological Dynamics” American Mathematical Society, 1955
- [25] H. Halberstam and K. Roth “Sequences” Oxford University Press, 1966
- [26] G.. Hardy and E.. Wright “An Introduction to the Theory of Numbers” Oxford University Press, 2008
- [27] N. Hindman “Algebra in and its applications to Ramsey theory” In Mathematica Japonica 44, 1996, pp. 581–625
- [28] N. Hindman “Finite sums from sequences within cells of a partition of ” In Journal of Combinatorial Theory, Series A 17.1, 1974, pp. 1–11
- [29] N. Hindman “Preimages of points under the natural map from to ” In Proceedings of the American Mathematical Society 37.2, 1973, pp. 603–608
- [30] N. Hindman and D. Strauss “Algebra in the Stone-Čech Compactificaton” De Gruyter, 2012
- [31] L. Kuipers and H. Neiderreiter “Uniform Distribution of Sequences” Wiley, 1974
- [32] A… Paterson “Amenability” American Mathematical Society, 1988
- [33] I.. Sanov “A property of a representation of a free group” In Doklady Akademii Nauk SSSR 57, 1947, pp. 657–659
- [34] E. Szemerédi “On sets of integers containing no elements in arithmetic progression” In Acta Arithmetica 27, 1975, pp. 199–243
- [35] B.. Waerden “Beweis einer Baudetschen Vermutung” In Nieuw. Arch. Wisk. 15, 1927, pp. 212–216
- [36] P. Walters “An Introduction to Ergodic Theory” Springer-Verlag, 1982
Si