Zero entropy actions of amenable groups are not dominant
Abstract.
A probability measure preserving action of a discrete amenable group is said to be dominant if it is isomorphic to a generic extension of itself. In [AGTW21], it was shown that for , an action is dominant if and only if it has positive entropy and that for any , positive entropy implies dominance. In this paper we show that the converse also holds for any , i.e. that zero entropy implies non-dominance.
1. Introduction
1.1. Definitions and results
Let be a standard Lebesgue space and let be a free, ergodic, -preserving action of a discrete amenable group on . It is natural to ask what properties of are preserved by a generic extension (a precise definition of “generic extension” is discussed in section 3). For example, it was shown in [GTW21] that if is a nontrivial Bernoulli shift then a generic is also Bernoulli and that a generic has the same entropy as . A system is said to be dominant if it is isomorphic to a generic extension . So, for example, the aforementioned results from [GTW21] together with Ornstein’s famous isomorphism theorem [Orn70] imply that all nontrivial Bernoulli shifts are dominant. More generally, it has been shown in [AGTW21] that
-
(1)
if , then is dominant if and only if it has positive Kolmogorov-Sinai entropy, and
-
(2)
for any , if has positive entropy, then it is dominant.
In this paper, we complete the picture by proving the following result.
Theorem 1.1.
Let be any discrete amenable group, and let be any free ergodic action with zero entropy. Then is not dominant.
1.2. Outline
1.3. Acknowledgements
I am grateful to Tim Austin for originally suggesting this project and for endless guidance. I also thank Benjy Weiss for pointing out an incorrect reference in an earlier version.
This project was partially supported by NSF grant DMS-1855694.
2. Slow entropy
Fix a Følner sequence for . For , write for the action of on the point , and for a subset , write . If is a partition of , then for denote by the index of the cell of containing . Sometimes we will use the same notation to mean the cell itself; which meaning is intended will be clear from the context. Given a finite subset , the -name of for the action is the tuple . Similarly, we also define the partition , and in some contexts we will use the same notation to refer to the cell of containing .
For a finite subset and any finite alphabet , the symbolic space is equipped with the normalized Hamming distance .
Definition 2.1.
Given a partition , a finite set , and , define
We refer to this set as the “-Hamming ball of radius centered at ”. Formally, it is the preimage under the map of the ball of radius centered at in the metric space .
Definition 2.2.
The Hamming covering number of is defined to be the minimum number of -Hamming balls of radius required to cover a subset of of -measure at least , and is denoted by .
Lemma 2.3.
Let be an isomorphism. Also let be a finite partition of , a finite subset of , and . Then .
Proof.
It is immediate from the definition of isomorphism that for -a.e.
Therefore it follows that for -a.e. . So any collection of -Hamming balls in covering a set of -measure is directly mapped by to a collection of -Hamming balls in covering a set of -measure . Therefore . The reverse inequality holds by doing the same argument with in place of . ∎
The goal of the rest of this section is to show that for a given action , the sequence of covering numbers grows at a rate that is bounded uniformly for any choice of partition . A key ingredient is an analogue of the classical Shannon-McMillan theorem for actions of amenable groups [MO85, Theorem 4.4.2].
Theorem 2.4.
Let be a countable amenable group, and let be any Følner sequence for . Let be an ergodic action of , and let be any finite partition of . Then
where denotes the entropy. In particular, for any fixed ,
Lemma 2.5.
For any partition , any Følner sequence , and any , let be the minimum number of -cells required to cover a subset of of measure . Then
Proof.
Let . Let . By Theorem 2.4, for sufficiently large depending on , we have
Let denote the set in the above equation. Let be the family of cells of the partition that meet . Then clearly and . Therefore
and this holds for arbitrary , so we are done. ∎
At this point, fix for all time . We can also now omit from all of the notations defined above, because it will never change. Also assume from now on that the system has zero entropy.
Lemma 2.6.
If is a Følner sequence for and is any finite subset of , then is also a Følner sequence for .
Proof.
First, because is finite and is Følner we have
Now fix any and observe that
which shows that is a Følner sequence. ∎
Lemma 2.7.
Let be real numbers satisfying
-
•
for each fixed , and
-
•
for all .
Then there exists a sequence such that and for each fixed , for sufficiently large (depending on ).
Proof.
For each , let be such that for all . Without loss of generality, we may assume that . Then we define the sequence by for and for . We have because for all . Finally, the fact that implies that for every fixed , as soon as . ∎
Proposition 2.8.
There is a sequence such that
-
(1)
, and
-
(2)
for any finite partition , there exists an such that for all .
Proof.
Because has zero entropy, there exists a finite generating partition for (see for example [Sew19, Corollary 1.2] or [Ros88, Theorem 2′]). Fix such a partition and let be any given partition. Because is generating, there is an integer and another partition such that is refined by and
By the mean ergodic theorem, we can write
so in particular, for sufficiently large, we have
Let denote the set .
Recall that is refined by , so is refined by . Let be the minimum number of cells required to cover a set of -measure at least , and let be such a collection of cells satisfying . If any of the do not meet the set , then drop them from the list. Because we can still assume after dropping that . Choose a set of representatives with each .
Now we claim that . To see this, let . Then there is one index such that and are in the same cell of . We can then estimate
The bounds for the first and third terms come from the fact that . The second term is because is refined by and was chosen so that and are in the same -cell. Therefore, . So, the proof is complete once we find a fixed sequence that is subexponential in and eventually dominates for each fixed .
3. Cocycles and extensions
Let be the unit interval , and let be Lebesgue measure on . Denote by the group of invertible -preserving transformations of . A cocycle on is a family of measurable maps indexed by that satisfies the cocycle condition: for every and -a.e. , . A cocycle can equivalently be thought of as a measurable map , where is the orbit equivalence relation induced by (i.e. if and only if for some ). With this perspective, the cocycle condition takes the form . A cocycle induces the skew product action of on the larger space defined by
This action preserves the measure and factors onto the original action .
By a classical theorem of Rokhlin (see for example [Gla03, Theorem 3.18]), any infinite-to-one extension of is isomorphic to a skew product action for some cocycle. Therefore, by topologizing the space of all cocycles on we can capture the notion of a “generic” extension – a property is said to hold for a generic extension if it holds for a dense set of cocycles. Denote the space of all cocycles on by . Topologizing is done in a few stages.
-
(1)
Let be the Borel sets in , and let be a sequence in that is dense in the metric. For example, could be an enumeration of the family of all finite unions of intervals with rational endpoints.
-
(2)
is completely metrizable via the metric
Notice that with this metric, has diameter at most . See for example [Kec10, Section 1.1]
-
(3)
If are maps , then define .
-
(4)
The metric defined in the previous step induces a topology on . Therefore, because is just a certain (closed) subset of , it just inherits the product topology.
To summarize, if is a cocycle, then a basic open neighborhood is specified by two parameters: a finite subset and . The -neighborhood of is . In practice, we will always arrange things so that for all on a set of of measure , which is sufficient to guarantee that is in the -neighborhood of .
Let be the partition of . We derive Theorem 1.1 from the following result about covering numbers of extensions, which is the main technical result of the paper.
Theorem 3.1.
For any sequence satisfying , there is a dense set such that for any , for infinitely many .
Proof that Theorem 3.1 implies Theorem 1.1.
Choose a sequence as in Proposition 2.8 such that for any partition , for sufficiently large . Let be the dense set of cocycles associated to as guaranteed by Theorem 3.1, and let , so we know that for infinitely many . Now if were an isomorphism, then by Lemma 2.3, would be a partition of satisfying for infinitely many , contradicting the conclusion of Proposition 2.8. Therefore, we have produced a dense set of cocycles such that , which implies Theorem 1.1. ∎
To prove Theorem 3.1, we need to show roughly that is both open and dense. We will address the open part here and leave the density part until the next section. Let be the partition of .
Lemma 3.2.
If is a sequence of cocycles converging to , then for any finite , we have
Proof.
For the names and to be the same means that for every ,
which is equivalent to
(1) |
The idea is the following. For fixed and , if and are close in , then (1) fails for only a small measure set of . And if is very close to in the cocycle topology, then and are close for all and most . Then, by Fubini’s theorem, we will get that the measure of the set of failing (1) is small.
Here are the details. Fix ; we will show that the measure of the desired set is at least for sufficiently large. First, let be so small that for any ,
This is possible because
Then, from the definition of the cocycle topology, we have
Let be large enough so that the above is larger than . Then, by Fubini’s theorem, we have
We have arranged things so that the integrand above is on a set of of -measure , so the integral is at least as desired. ∎
Lemma 3.3.
For any finite and any , is open in .
Proof.
Suppose is a sequence of cocycles converging to and satisfying for all . We will show that as well. The covering number is a quantity which really depends only on the measure , which we now call for short. The assumption that for all says that for each , there is a collection of words such that the Hamming balls of radius centered at these words cover a set of -measure at least . Since is a finite set, there are only finitely many possibilities for the collection . Therefore by passing to a subsequence and relabeling we may assume that there is a fixed collection of words with the property that if we let be the Hamming ball of radius centered at , then for every .
Now, by Lemma 3.2, the map agrees with on a set of measure converging to as . This implies that the measures converge in the total variation norm on to . Since for every , we conclude that also, which implies that as desired. ∎
Define . By Lemma 3.3, each is a union of open sets and therefore open. Also, is exactly the set of satisfying infinitely often. Therefore, by the Baire category theorem, in order to prove Theorem 3.1 it suffices to prove
Proposition 3.4.
For each , is dense in .
The proof of this proposition is the content of the next section.
4. Proof of Proposition 3.4
4.1. Setup
Let be fixed and let be an arbitrary cocycle. Consider an arbitrary neighborhood of determined by a finite set and . We can assume without loss of generality that . We will produce a new cocycle such that there is a set of measure on which for all . The construction of such an is based on the fact that the orbit equivalence relation is hyperfinite.
Theorem 4.1 ([CFW81, Theorem 10]).
There is an increasing sequence of equivalence relations such that
-
•
each is measurable as a subset of ,
-
•
every cell of every is finite, and
-
•
agrees -a.e. with .
Fix such a sequence and for , write to denote the cell of that contains .
Lemma 4.2.
There exists an such that .
Proof.
Almost every satisfies , so in particular, for -a.e. , there is an such that for all . Letting , we see that the sets are increasing and exhaust almost all of . Therefore we can pick so that . ∎
Now we drop from the sequence and assume that .
Lemma 4.3.
There exists a such that .
Proof.
Every -cell is finite, so if we define , then the are increasing and exhaust all of . So we pick so that . ∎
Continue to use the notation .
Lemma 4.4.
For all sufficiently large, .
Proof.
We have . By the mean ergodic theorem, we get
Therefore, in particular, as , so this measure is for all sufficiently large. ∎
From now on, let be a fixed number that is large enough so that the above lemma holds, , and . This is possible because is assumed to be subexponential in . The relevance of the final condition will appear at the end.
Lemma 4.5.
There is an such that .
Proof.
Same proof as Lemma 4.2. ∎
Again, drop from the sequence of partitions and assume .
4.2. Construction of the perturbed cocycle
Let be the relabeled sequence of equivalence relations from the previous section. The following measure theoretic fact is well known. Recall that two partitions and of are said to be independent with respect to if for any , .
Lemma 4.6.
Let and be two finite partitions of . Then there exists a such that and are independent with respect to .
Proposition 4.7.
For any , there is an such that
-
(1)
whenever , and
-
(2)
for -a.e. , the following holds. If is an -cell contained in , consider the map as a random variable on the underlying space . Then as ranges over all such -cells, the random variables are independent.
Proof.
We give here only an intuitive sketch of the proof and leave the full details to Appendix A. It is more convenient to adopt the perspective of a cocycle as a map satisfying the condition .
Step 1. For , let .
Step 2. Fix an -cell . Enumerate by all of the -cells contained in and choose from each a representative .
Step 3. Let denote the partition of . Define to be an element of such that
are independent. These expressions are well defined because has already been defined on and we use Lemma 4.6 to guarantee that such an element of exists.
Step 4. There is now a unique way to extend the definition of to that is consistent with the cocycle conditition. For arbitrary , define
The middle term in the first equation was defined in the previous step and the outer two terms were defined in step 1.
Step 5. Extend the definition of to the rest of the inductively, making each cell independent of all the previous ones. Suppose has been defined on . Using Lemma 4.6 again, define to be an element of such that
are independent. Then, just as in step 4, there is a unique way to extend the definition of to all of . At the end of this process, has been defined on all of . This was done for an arbitrary -cell , so now is defined on .
Step 6. For each , extend the definition of from to with the same procedure, but there is no need to set up any independence. Instead, every time there is a choice for how to define between two of the cell representatives, just take it to be the identity. This defines on , which is equal mod to the full orbit equivalence relation, so is a well defined cocycle.
Now we verify the two claimed properties of . Property (1) is immediate from step 1 of the construction. To check property (2), fix and let be any of the -cells contained in . Note that the name records the data for all such that , which, by switching to the other notation is the same data as for . So, the set of for which is equal to a particular word is given by a corresponding particular cell of the partition . The construction of was defined exactly so that the partitions are all independent and the names are determined by these independent partitions pulled back by the fixed -preserving map , so they are also independent.
The reason this is only a sketch is because it is not clear that the construction described here can be done in a way so that the resulting is a measurable function. To do it properly requires a slightly different approach; see Appendix A for full details. ∎
Letting , this construction guarantees that for all . By Lemma 4.2, , so this shows that is in the neighborhood of .
4.3. Estimating the size of Hamming balls
Let be the cocycle constructed in the previous section. We will estimate the -measure of -Hamming balls in order to get a lower bound for the covering number. The following formulation of Hoeffding’s inequality will be quite useful [Ver18, Theorem 2.2.6].
Theorem 4.8.
Let be independent random variables such that each almost surely. Let . Then for any ,
Proposition 4.9.
For any ,
(2) |
Proof.
Let be the collection of -cells that meet and satisfy . For each , let . Define
and for each , define
Then we have
so to get an upper bound for , it is sufficient to control .
View each as a random variable on the underlying probability space . Our construction of the cocycle guarantees that the collection of names as ranges over all of the -cells contained in is an independent collection. Therefore, in particular, the for are independent (the assumption that guarantees that all are contained in ).
We also have that each and the expectation of the sum is
where the final inequality is true because . So, we can apply Theorem 4.8 with to conclude
The final inequality holds because and is small enough so that . ∎
Corollary 4.10.
Let . If is any -Hamming ball of radius , then .
Proof.
If does not meet the fiber above , then obviously . So assume for some . Then applying the triangle inequality in the space shows that . Now apply Proposition 4.9 with in place of . The proof goes through exactly the same and we get the same constant in the final estimate because and are small enough so that is still . ∎
Corollary 4.11.
We have .
Proof.
Let be a collection of -Hamming balls of radius such that
Then
implying that . ∎
Our choice of at the beginning now guarantees that , showing that as desired. This completes the proof of Proposition 3.4.
Appendix A Measurability of the perturbed cocycle
In this section, we give a more careful proof of Proposition 4.7 that addresses the issue of measurability. We will need to use at some point the following measurable selector theorem [Fre06, Proposition 433F].
Theorem A.1.
Let and be standard Borel spaces. Let be a probability measure on and suppose that is measurable and surjective. Then there exists a measurable selector which is defined -a.e. (meaning for -a.e. ).
Given , there is a natural bijection between and because is a free action. We can also identify subsets – if , then we will write . Note that this set depends on the “base point” . If and are two points in the same -orbit, then the set based at is a translate of the same set based at . It will always be clear from context what the intended base point is.
Definition A.2.
A pattern in is a pair , where is a finite subset of and is a partition of .
Definition A.3.
For , define to be the pattern , where and is the partition of into the sets where ranges over all of the -cells contained in .
Lemma A.4.
is a measurable function of .
Proof.
Because there are only countably many possible patterns, it is enough to fix a pattern and show that is measurable. Enumerate . Saying that is the same as saying that and each is a cell of . We can express the set of satisfying this as
Because each is a measurable set and each is a measurable map, this whole thing is measurable. ∎
For each pattern , let . We will define our cocycle inductively on the equivalence relations . For each , the sets partition into countably many measurable sets, so it will be enough to define measurably on each . At this point, fix a pattern , fix , and write instead of . Define
where is said to be -independent if for any fixed , the partitions
as ranges over are independent with respect to .
Proposition A.5.
For every , there is some that extends .
Proof.
The idea is exactly the same as the construction described in steps 3-5 in the sketched proof of Proposition 4.7, but we write it out here also for completeness.
Enumerate and for each fix an element . First, obviously we will define on each . Next, define to be an element of such that
are independent. Then, define on all of by setting
for any . Continue this definition inductively, making each new step independent of all the steps that came before it. If has been defined on , then define to be an element of such that
are independent. Then extend the definition of to all of in the exact same way.
At the end of this process, has been defined on , and it satisfies the cocycle condition by construction. To verify that it also satisfies the independence condition, notice that the construction has guaranteed that
are independent partitions as ranges over . To get the same conclusion for an arbitrary base point , pull everything back by the fixed map . Because this map is measure preserving, pulling back all of the partitions by it preserves their independence. ∎
Now we would like to take this information about cocycles defined on patterns and use it to produce cocycles defined on the actual space . Define the map by . Note that this is a measurable map because is a measurable cocycle.
By Theorem A.1 applied to the measure , we get a measurable map defined -a.e. such that extends . Denote the composition by and write the image of under this map as . To summarize, for every pattern , there is a measurable map defined -a.e. with the property that extends .
It is now natural to define our desired cocycle on the equivalence relation by the formula . It is then immediate to verify the two properties of claimed in the statement of Proposition 4.7. The fact that agrees with on follows from the fact that extends and the claimed independence property of translates directly from the independence property that the were constructed to have (see also the discussion after step 6 in the sketched proof of Proposition 4.7). Also, is measurable because for each fixed , the map is simply a composition of other maps already determined to be measurable. The only problem is that , when defined in this way, need not satisfy the cocycle condition. To see why, observe that the cocycle condition is equivalent to the condition
(3) |
But in defining the maps , we have simply applied Theorem A.1 arbitrarily to each pattern separately, so and have nothing to do with each other. However, we can fix this problem with a little extra work, and once we do, we will have defined with all of the desired properties.
Start by declaring two patterns equivalent if they are translates of each other, and fix a choice of one pattern from each equivalence class. Since there are only countably many patterns in total, there is no need to worry about how to make this choice. For each representative pattern , apply Theorem A.1 arbitrarily to get a map . This does not cause any problems because two patterns that are not translates of each other can not appear in the same orbit (this follows from the easy fact that ), so it doesn’t matter that their maps are not coordinated with each other. For convenience, let us denote the representative of the equivalence class of by . Now for every , let be the unique element of with the property that . Notice that the maps and are both constant on each subset and are therefore measurable.
Now for an arbitrary pattern and , we define the map by
Notice that this is still just a composition of measurable functions, so is measurable. All that remains is to verify that this definition satisfies (3). The right hand side of (3) is
which is by definition equal to the left hand side of (3) as desired.
This, together with the discussion surrounding (3), shows that if we construct the maps in this way, then making the definition gives us a true measurable cocycle with all of the desired properties. Finally, to extend the definition of to with , repeat the exact same process, except it is even easier because there is no need to force any independence. The maps only need to be measurable selections into the space , and then everything else proceeds in exactly the same way.
References
- [AGTW21] Tim Austin, Eli Glasner, Jean-Paul Thouvenot, and Benjamin Weiss. An ergodic system is dominant exactly when it has positive entropy. arXiv:2112.03800, 2021.
- [CFW81] Alain Connes, Jacob Feldman, and Benjamin Weiss. An amenable equivalence relation is generated by a single transformation. Ergodic Theory Dynam. Systems, 1(4):431–450, 1981.
- [Fre06] D. H. Fremlin. Measure theory. Vol. 4. Torres Fremlin, Colchester, 2006. Topological measure spaces. Part I, II, Corrected second printing of the 2003 original.
- [Gla03] Eli Glasner. Ergodic theory via joinings, volume 101 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2003.
- [GTW21] E. Glasner, J.-P. Thouvenot, and B. Weiss. On some generic classes of ergodic measure preserving transformations. Trans. Moscow Math. Soc., 82:15–36, 2021.
- [Kec10] Alexander S. Kechris. Global aspects of ergodic group actions, volume 160 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2010.
- [KT97] Anatole Katok and Jean-Paul Thouvenot. Slow entropy type invariants and smooth realization of commuting measure-preserving transformations. Ann. Inst. H. Poincaré Probab. Statist., 33(3):323–338, 1997.
- [MO85] Jean Moulin Ollagnier. Ergodic theory and statistical mechanics, volume 1115 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1985.
- [Orn70] Donald Ornstein. Bernoulli shifts with the same entropy are isomorphic. Advances in Math., 4:337–352, 1970.
- [Ros88] A. Rosenthal. Finite uniform generators for ergodic, finite entropy, free actions of amenable groups. Probab. Theory Related Fields, 77(2):147–166, 1988.
- [Sew19] Brandon Seward. Krieger’s finite generator theorem for actions of countable groups I. Invent. Math., 215(1):265–310, 2019.
- [Ver18] Roman Vershynin. High-dimensional probability, volume 47 of Cambridge Series in Statistical and Probabilistic Mathematics. Cambridge University Press, Cambridge, 2018. An introduction with applications in data science, With a foreword by Sara van de Geer.