Fluctuation bounds for ergodic averages of amenable groups
Abstract.
We study fluctuations of ergodic averages generated by actions of amenable groups. In the setting of an abstract ergodic theorem for locally compact second countable amenable groups acting on uniformly convex Banach spaces, we deduce a highly uniform bound on the number of fluctuations of the ergodic average for a class of Følner sequences satisfying an analogue of Lindenstrauss’s temperedness condition. Equivalently, we deduce a uniform bound on the number of fluctuations over long distances for arbitrary Følner sequences. As a corollary, these results imply associated bounds for a continuous action of an amenable group on a -finite space with .
In this article, we consider a problem at the interface of effective ergodic theory and the ergodic theory of group actions.
By effective ergodic theory, we mean the following programme: insofar as ergodic theory provides theorems which tell us about the long-term behaviour of dynamical systems, we may ask for more quantitative, or computationally explicit, analogues of those theorems. For instance, the mean ergodic theorem of von Neumann asserts that, whenever is a -finite measure space, is a measure-preserving transformation, and , the sequence of ergodic averages converges in norm; but how fast does this convergence occur?
Effective ergodic theory is complicated by some rather general negative results. Indeed, it is known [19] that in the aforementioned setting of the mean ergodic theorem, there is no uniform rate of convergence of the averages in norm if one considers arbitrary functions in . Consequently, if we are to find a quantitative analogue of the mean ergodic theorem, viz. one which is more explicit about the nature of the convergence of , it is necessary to look for more subtle convergence data than a uniform rate of convergence. One such form of convergence data is the number of -fluctuations of a sequence for each , or possibly -fluctuations satisfying some side condition, such as the fluctuations being “over long distances”. This is exactly the type of convergence data that we investigate here; a precise explanation of these terms is given in Definition 1.
On the other hand, the ergodic theory of group actions seeks to modify the machinery of classical ergodic theory, by replacing the action of a single measure-preserving transformation with a measure-preserving action by a group. In fact, since the action by a single invertible measure-preserving transformation can be identified with a measure-preserving action of , many theorems in the ergodic theory of group actions contain results from classical ergodic theory as special cases. In particular, this is typically the case where one is interested in the ergodic theory of a class of group actions where the groups involved belong to a class of groups containing , such as abelian groups, nilpotent groups, or, in the case of the present article, amenable groups (see Definition 12 below).
Consider, therefore, the following version of the mean ergodic theorem for actions of amenable groups:
Theorem.
Let be such that either is -finite and or and , and let . Let be a locally compact second countable amenable group with Haar measure , let act continuously on by measure-preserving transformations, and let be a Følner sequence of compact subsets of . Then converges in .
This result is originally due to Greenleaf [11], whose proof goes by way of an abstract Banach space analogue of the mean ergodic theorem which is simultaneously general enough to deduce the mean ergodic theorem for an amenable group acting on any reflexive Banach space or any with a finite measure. Central to Greenleaf’s proof is a fixed point argument which, in particular, does not give any effective convergence information about the averages . (If one specialises to the case where , a simpler proof is available [7, Thm. 8.13], but this proof also does not give any effective convergence information.)
Here our aim is to give an effective analogue of Greenleaf’s theorem. At the cost of some generality — here, we only consider actions of amenable groups on uniformly convex Banach spaces — we obtain an explicit uniform fluctuation bound for . In other words, we deduce a result of the following form:
Theorem.
(“Main theorem”) Let be such that is -finite and , and let . Let be a locally compact second countable amenable group with Haar measure , let act continuously on by measure-preserving transformations, and let be a Følner sequence of compact subsets of . Then converges in with a uniform bound on the number of -fluctuations over long distances for each ; where the uniform bound (both the number of fluctuations and the “long distances”) depends exclusively, and explicitly, on: the choice of , the quantity , and the Følner sequence .
In fact, the dependence of the number of -fluctuations over long distances of on the choice of Følner sequence is only via a specific type of data from the Følner sequence, which expresses the fact that is a Følner sequence in a slightly more quantitatively explicit fashion. We call this data the Følner convergence modulus; see Definition 15. It also turns out to be possible to delete the “over long distances” clause from the main theorem if, instead, one adds a side condition on the Følner sequence used, namely the condition that the Følner sequence be “fast”, as detailed in Definition 17.
We deduce this theorem as a special case of an “explicit fluctuation bound” analogue of an abstract mean ergodic theorem for a certain class of amenable group actions on uniformly convex Banach spaces, which we also call the “main theorem” of the article; this result is stated as Theorem 21 below, which also gives the explicit form of the uniform bound. The modified version of this theorem where the fluctuations are not “over long distances” but the Følner sequence is assumed to be “fast” is given as Corollary 22.
The plan of the article is as follows. In Section 1, we establish a number of background facts from functional analysis and the theory of amenable groups which are needed to state and prove the main theorem. In Section 2, we first provide a specialised proof of an abstract mean ergodic theorem for lcsc amenable groups acting on uniformly convex Banach spaces, and then modify this proof so as to be sufficiently quantitatively explicit that we are able to deduce the main theorem. Lastly, in Section 3, a discussion of some related literature and open problems is provided.
1. Preliminaries
Definition 1.
Fix an . Given a sequence in some metric space, we say that has at most -fluctuations if for every finite sequence such that for each , , it always holds that . A weaker notion is “-fluctuations at distance ”: given some function with for every , we say that has at most -fluctuations at distance if for every finite sequence with the property that for every such that for each , , it always holds that .
Remark.
It holds that a sequence is Cauchy (viz. that for every there exists an such that for , ) iff for every there exists some such that has at most -fluctuations, iff for any and with , there exists some such that has at most -fluctuations at distance . (More precisely: if is Cauchy then for any with , has only finitely many -fluctuations at distance for each ; whereas conversely, if there is any such so that has only finitely many -fluctuations at distance for each , then it follows that is Cauchy.) Likewise, it is obvious that if for a specific sequence we happen to know an explicit witnessing the Cauchy property, then this also serves as an explicit upper bound for , and likewise any explicit serves as an explicit upper bound on (for any ). However the converses are all false in a strong sense: there exist examples of sequences where is a computable function of but is not computable [1], and likewise with (for ) and respectively [18].
These phenomena are certainly present in ergodic theory. As mentioned in the introduction, it has long been known that a single measure-preserving transformation acting on , then when , there exist functions for which the convergence indicated by the mean (and pointwise) ergodic theorem occurs arbitrarily slowly [19]. In other words does not exhibit a uniform rate of convergence. However, it was shown by Avigad and Rute [1] that when in this setting — in fact, more generally, if the acted-upon space is replaced with any uniformly convex Banach space with modulus of uniform convexity ) (see definition below) — then there exists a uniform bound on the number of -fluctuations in the sequence which depends only on and .
Definition 2.
A normed vector space is said to be uniformly convex if there exists a nondecreasing function such that for all with and , it follows that . Such a function is then referred to as a modulus of uniform convexity for . We say that is -uniformly convex if is a modulus of uniform convexity for , where is some constant.
Remark.
There are a number of equivalent ways to define uniform convexity. We have chosen the preceding definition because it is the most convenient for our argument, but it is worth mentioning another characterisation which is perhaps more standard: a space is uniformly convex provided there is a nondecreasing function (also called a modulus of uniform convexity) such that for all with and , it follows that . It is not hard to show (cf. [17, Lemma 3.2]) that a function is a modulus of convexity in this sense iff is a modulus of uniform convexity in the sense of our definition. (This indicates the origin of the off-by-one issue in our definition of -uniform convexity.)
It is well known [6] that the and spaces are uniformly convex when . Hanner showed [12] that for , the sharp modulus for has an especially nice form, namely . In particular, this implies that is a modulus of uniform convexity, in our sense, for with . (The same work shows that, even though is also uniformly convex for , the sharp modulus does not have as nice of an explicit form.)
We recall some basic notions from the theory of vector-valued integration. We shall closely follow the recent textbook by Hytönen et al. [13]; for the convenience of the reader, we will sometimes refer directly to specific definitions, theorems, etc. therein.
Consider some measure space with some function , with a Banach space. We say that is a simple function with respect to the -algebra and space , if it is of the form , with an indicator function for , and . We then say that a function is strongly measurable if it is a pointwise limit of simple functions, i.e. if there exists a sequence of simple functions such that for every , [13, Def. 1.1.4]. By contrast, a subtly different notion (but more standard in the vector integration literature) is that of -strong measurability, which only asserts this limit for -almost all , but requires that the sets in the definition of simple function have finite -measure (see [13, Def. 1.1.13 and Def. 1.1.14]). We do not directly use -strong measurability in the main theorem of this paper — in particular, it is too weak of a form of measurability for Propositions 8 and 9 below. The relationship between strong measurability and -strong measurability is the following (quoting from [13, Prop. 1.1.16]):
Fact 3.
Consider a measure space , a Banach space , and a function .
-
(1)
If is strongly -measurable, then is -almost everywhere equal to a strongly measurable function.
-
(2)
If is -finite and is -almost everywhere equal to a strongly measurable function, then is strongly -measurable.
As a particular case of Fact 3, we see that when is -finite, a strongly measurable function is also -strongly measurable.
For -strongly measurable functions, one can define a form of integration, namely the Bochner integral, in direct analogy with the Lebesgue integral. Bochner integration is denoted by . A function is Bochner -integrable iff it is both -strongly measurable and , that is, is integrable in the Lebesgue sense [13, Prop. 1.2.2].
We record some other basic facts about the Bochner integral. (Each of these will be ultimately used in the proof of Lemma 19.)
Fact 4.
Let be a measure space and be -strongly measurable.
(1) . [13, Prop. 1.2.2]
(2) If , then . [13, Eqn. 1.2]
(3) Fubini’s theorem holds for the Bochner integral. [13, Prop. 1.2.7] In particular, if is another measure space, and and are -finite, and is Bochner integrable, then
A more general notion than strong measurability is weak measurability: we say that is weakly measurable if for every , the function is measurable (in the ordinary sense as a function from to with the Borel -algebra). The following classical result indicates when weak measurability implies strong measurability.
Proposition 5.
(Pettis measurability criterion [13, Thm. 1.1.6]) Let be a measure space and a Banach space. For a function the following are equivalent:
-
(1)
is strongly measurable.
-
(2)
f is weakly measurable, and is separable in .
An easy consequence of the Pettis measurability criterion is the following.
Proposition 6.
If the measure space is also a separable topological space and every Borel set in is -measurable, and is continuous, then is strongly measurable.
Proof.
Observe that for any , is a composition of continuous functions, and is therefore continuous. Hence is weakly measurable. Moreover, it holds that the continuous image of a separable space is separable. ∎
Before the next proposition, it will be convenient to introduce the following terminology.
Definition 7.
If is a measure space which is also a topological space, and extends the Borel -algebra on , then we say that is weakly Borel if is Borel, for all . Likewise we say that is strongly Borel provided it is weakly Borel and that is separable in .
Of course, in the case where is precisely the Borel -algebra on , then the notions of weakly and strongly Borel functions coincide with weak and strong measurability of .
Proposition 8.
If and are both topological spaces where every Borel set is measurable, is Borel, and is strongly Borel, then is strongly Borel (in particular strongly measurable).
Proof.
By hypothesis, is also weakly Borel, so for any , we have that is Borel. Therefore is Borel, and thus is weakly Borel. Now, let be the image of in . Note that . Since is separable in , it follows that is also separable in since it is contained in . ∎
We now fix some notation and terminology regarding group actions on Banach spaces and measure spaces.
A locally compact group will always come equipped with a Haar measure. In the countable discrete case this coincides with the counting measure. Regardless of whether the group is discrete or continuous, we will use the notations and interchangeably to refer to the Haar measure. Conventions differ on the issue of whether the measure space is understood to come equipped with the Borel -algebra or its completion; in the latter case, the assertion that a function is strongly Borel is stronger than the assertion that is strongly measurable. Therefore, in what follows, we assume only that the -algebra on contains the Borel -algebra, and carefully note when a function is assumed to be strongly Borel rather than strongly measurable. We use the abbreviation lcsc for topological groups which are locally compact and second countable; importantly, for lcsc groups it always holds that the Haar measure is -finite.
In general, we say that a group acts on a Banach space if there is a function that returns an operator on for every , is the identity operator, and for all , . Together these imply that . We say that acts linearly on provided that in addition, maps from to the space of linear operators on . Writing to indicate the set of all linear operators from to with supremum norm , another way to say that acts both linearly and with unit norm on is to say that acts on via .111We remark that any group that acts via a representation such that every is nonexpansive actually does so via , by the fact that and the general fact about linear operators that . Nonexpansivity is required for the proof of our main result. Likewise, we say that a topological group acts continuously on provided that for every , if then . In other words is continuous from to . In the case where also acts linearly (resp. and with unit norm) on , this is equivalent to requiring that (resp. ) is continuous when is equipped with the strong operator topology.
Relatedly, in the case where a group acts on a measure space , we say that acts by measure-preserving transformations if for every and . Likewise, we say that acts continuously on provided that for every , if then . This notion of continuous group action is related to our previous notion of continuous action on a Banach space by Proposition 10 below (which may be taken as a justification for the terminology).
Furthermore, we say that if is understood as a measure space, then acts strongly on a Banach space provided that for every , is strongly measurable from to .222By analogy with the previous situation where acts continuously on : in the case where also acts linearly (resp. and with unit norm) on , this is equivalent to requiring that (resp. ) satisfies a “strongly measurable with respect to the strong operator topology” like condition: for a precise statement, see [13, Cor. 1.4.7]. However we do not use this alternative, strong operator topology based, characterisation in our argument. Likewise, we say that acts Borel strongly on provided that for every , is strongly Borel from to . Our argument will simultaneously require that acts Borel strongly on , and linearly and with unit norm. To refer to this last condition, we will say that acts Borel strongly on via the representation .
In the situation where acts Borel strongly on via the representation , we can immediately deduce the following facts, which will be used in the proof of Lemma 19.
Proposition 9.
Suppose that acts Borel strongly on via the representation . Then,
-
(1)
for each with , we have that is Bochner integrable for each , i.e. is -strongly measurable and . Moreover,
-
(2)
for each with and for each , we have that is Bochner integrable.
Proof.
(1) Since is a strongly Borel function from to , and is a continuous and therefore Borel function from to , it follows from Proposition 8 that is strongly measurable (in fact strongly Borel). Since the Haar measure on is -finite, this implies that is -strongly measurable, by Fact 3. For the second condition of Bochner integrability, it suffices to observe that since for every ,
(2) Since is a topological group, it holds that group multiplication is continuous. Therefore is strongly measurable (in fact strongly Borel) thanks to Proposition 8, since it is a composition of the continuous (and therefore Borel) multiplication function and the strongly Borel function . Consequently, Fact 3 implies that is also -strongly measurable. We also have Bochner integrability because again,
Let us tie all this discussion of Bochner integration and group actions on Banach spaces back to our original setting of interest. In the “concrete version” of Greenleaf’s mean ergodic theorem, stated in the introduction, acts continuously and by measure-preserving transformations on a -finite measure space , and with . In the next proposition and corollary, we briefly check that in fact, this particular situation is actually covered by our abstract “group action on Banach space” framework.
Proposition 10.
Suppose that is a topological group which acts continuously and by measure-preserving transformations on a -finite measure space . Denote this action by
Fix .
-
(1)
This action of on induces an action of on , given by
where for a fixed , the function is defined in the following manner: . In other words, there is a representation of into the space of operators on , with .
-
(2)
Moreover, acts linearly and isometrically on , in the sense that for all ; so in particular, .
-
(3)
In addition, acts continuously on , that is, for each , is a continuous function from to .
Remark.
This proposition is basically a modification of the rather classical Koopman operator formalism; except here, instead of a single measure-preserving transformation, we have a topological group of measure-preserving transformations. In the former case, the argument is quite standard; its generalisation to a group action is straightforward, but we explicitly go through the proof here for completeness. (The argument is nearly given in [7, Ch. 8], for instance, albeit using a more restrictive definition of continuous -action on a measure space.)
Additionally, we note that the statement of the proposition includes the case , which is excluded from our main theorem; nor does the proposition require that be lcsc or amenable.
Proof.
(1) Here, we need only to check explicitly that is indeed a group action of on . To wit, we must show that , and , for every .
The first is obvious, since
and for every , since is a group action on . Likewise, the second condition also follows directly from the fact that is a group action on :
(2) To see that is a linear operator on for each , simply note that
The isometry property follows immediately from the change of variable formula and the fact that the action of on is measure preserving. Explicitly: fix , and compute that
(3) We need to show that, for arbitrary , if then . We first prove the claim for an indicator function; then, for simple functions; then pass to general functions in by a density argument.
Fix a measurable set with . By the assumption that the action of on is continuous, it holds that: for every , there exists a such that for all , .
Let denote the indicator function for . Let be a neighbourhood of the origin such that for every , . Let . Observe that . Since , we have that . But observe that
Since our choice of was arbitrary, we conclude that is continuous when is an indicator function of a set of finite measure.
Now let be a simple function of the form . Then, . In this case, it suffices to pick a neighbourhood which is small enough that for each , we have that
Indeed, from the triangle inequality, and the previous result for indicator functions, we observe that
Lastly, take to be arbitrary. Suppose that is a simple function such that . By the previous case, we can pick a neighbourhood such that for all . Now, observe that
We have already seen that the first two terms are each . For the last term, observe that
by part (2) of the proposition. Consequently, as well, so that as desired. ∎
Corollary 11.
Let be a lcsc group acting continuously by measure-preserving transformations on a -finite measure space . Fix . Then the induced action of on is linear, strongly Borel, and has unit norm.
Proof.
That the induced action of on is linear with unit norm is immediate from Proposition 10 (2). From Proposition 10 (3), we know that is continuous from to for each ; since is separable, Proposition 6 indicates that is strongly measurable. Thus, we need only to check that is strongly Borel, as per Definition 7.
The continuity of and the separability of imply that the image of under this mapping is separable in . Likewise, since is continuous, post-composing the mapping with an element of the dual space of results in a continuous map from to , hence automatically Borel also. Therefore is both weakly Borel and has separable image, for any choice of ; so we’ve shown that acts Borel strongly on . ∎
We now turn our attention to amenable groups. The following serves as our preferred characterisation of amenability.
Definition 12.
(1) Let be a countable discrete group. A sequence of finite subsets of is said to be a Følner sequence if for every and finite , there exists an such that for all and for all , .
(2) Let be a locally compact second countable (lcsc) group with Haar measure . A sequence of compact subsets of is said to be a Følner sequence if for every and compact , there exists an such that for all , there exists a subset of with such that for all , .
Remark.
It has been observed, for instance, by Ornstein and Weiss [23] that (2) is one of several equivalent “correct” generalisations of (1) to the lcsc setting. Note however, that we do not assume is nested ( for all ) or exhausts (), nor do we assume, in the lcsc case, that is unimodular. (Each of these is a common additional technical assumption when working with amenable groups.) Conversely, some authors use a version of (2) where the sets in are merely assumed to have finite volume, rather than compact; thanks to the regularity of the Haar measure, our definition results in no loss of generality.
Definition 13.
If is either a countable discrete or lcsc amenable group, and has some distinguished Følner sequence , and acts Borel strongly on via a representation , then we define the th ergodic average operator as follows: . Here denotes the normalised integral, that is, .
Proposition 14.
With the notation above, .
Proof.
Evidently is a linear operator from to . From Proposition 9 we know that is Bochner integrable, and in particular
A key piece of quantitative information for us will be how large has to be if is chosen to be an element of , in order for to be small. This information is encoded by the following modulus:
Definition 15.
Let be an amenable group, either countable discrete or lcsc, with Følner sequence . A Følner convergence modulus for returns an integer such that:
-
(1)
If is countable discrete,
-
(2)
If is lcsc,
We remark that if is an increasing Følner sequence (that is, for all ) then it follows trivially that is a nondecreasing function for any fixed . However, in what follows we do not always assume that is increasing. In some instances it is technically convenient to assume that is nondecreasing; in this case, we can upper bound using an “envelope” of the form . Hence, in any case we are free to assume that is nondecreasing in if necessary.
It should be clear that if we have a “sufficiently explicit” amenable group with a “sufficiently explicit” Følner sequence , then we can explicitly write down a Følner convergence modulus for . What is meant by this? Simply, the following: if we know “explicitly” that is a Følner sequence, this amounts to knowing “explicitly”, for each , that (resp. the analogous statement for when is lcsc), which in turn amounts to being able to write down how exactly this convergence occurs in an explicit way, i.e. that one can explicitly write down a rate of convergence for . But a Følner convergence modulus is just a function which dominates for every — in particular we can take .
Likewise (but less heuristically), it is easy to see that we can select a Følner sequence in such a way that can be chosen to be a computable function (for an appropriate restriction on the domain of the second variable). The following argument has essentially already been observed by previous authors [4, 5, 22] working with slightly different objects, but we include it for completeness. (In the interest of simplicity, we restrict our attention to countable discrete groups; one can say something similar for lcsc groups with computable topology, but we do not pursue this here.)
Proposition 16.
Let be a countable discrete finitely generated amenable group with the solvable word property. Fix . Then has a Følner sequence such that is a Følner convergence modulus for . Moreover can be chosen in a computable fashion.
Proof.
Fix a computable enumeration of the elements of , as well as a computable enumeration of the finite subsets of . The solvable word property ensures that we can do this, and also that the cardinality of can always be computed for any and finite set . So, take to be an arbitrary finite subset of containing , the first element of . Given , take to be the least (with respect to the enumeration) finite subset of containing , such that for all , . Such an exists since is amenable. Then, put where is the th element of . This is indeed a Følner sequence: for a fixed , we see that for all greater than the first such that , because
Hence . Moreover, we see that if , then
and so is indeed a Følner convergence modulus for . ∎
Remark.
The previous proposition is not sharp. It has been shown that there are groups without the solvable word property which nonetheless have computable Følner sequences with computable convergence behaviour [5]. (The cited paper uses a different explicit modulus of convergence for Følner sequences than the present paper, although the argument carries over to our setting without modification.)
Finally, it will be convenient to introduce the notion of a Følner sequence where the gap between and is controlled.
Definition 17.
Let be a countable discrete or lcsc amenable group and a Følner sequence. Let and . We say that is a -fast Følner sequence if
-
(1)
For countable and discrete, it holds that for all that for all , for all , .
-
(2)
For lcsc, it holds that for all that for all , there exists a set such that , so that for all , .
In other words, a Følner sequence is -fast provided that is a Følner convergence modulus for .
It is clear that any Følner sequence can be refined into a -fast Følner sequence.
Proposition 18.
Given and , any Følner sequence can be refined into a -fast Følner sequence.
Proof.
It suffices to produce a -fast refinement. For simplicity, we only state the argument for the case where is countable and discrete.
Suppose we have already selected the first Følner sets in our refinement . Then, take to be the next element of the sequence after such that, for all , . Such a term exists since is a Følner sequence. ∎
Less obvious is the relationship between a Følner sequence being fast and the property of being tempered which is used in Lindenstrauss’s pointwise ergodic theorem [20], although they are somewhat similar in spirit. Nonetheless, we quickly observe that the previous proposition indicates that any tempered Følner sequence can be refined into a Følner sequence which is both tempered and -fast, simply by the fact that any subsequence of a tempered Følner sequence is again tempered.
2. The Main Theorem
Frequently in ergodic theory, one argues that if , then . The following lemma makes this precise in terms of the modulus .
Lemma 19.
Let be a normed vector space. Let be a lcsc amenable group with Følner sequence , and let act Borel strongly on via the representation . Fix and . Let be the Følner convergence modulus and suppose . Then for any , . (If is countable discrete, the “strongly Borel” part is trivially satisfied, and we have the sharper estimate .)
Proof.
From the definition of Følner convergence modulus, we know that there exists an such that and such that for all , . Now perform the following computation (justification for each step addressed below):
If is countable discrete, we instead assume that for all (rather than ), . Therefore, the penultimate line reduces to , and the last line reduces to .
Finally let’s discuss which properties of the Bochner integral we had to use. Thanks to Proposition 9, we have that is Bochner integrable. Given that, for each , is a bounded linear operator, then indeed it follows that . From Fubini’s theorem and the fact (also from Proposition 9) that is Bochner integrable, we see that . Lastly, we repeatedly invoked the fact that . It’s worth noting that in the case where is countable discrete, only the first fact (that acts by bounded linear operators with unit norm) is needed as an assumption, as the latter two properties hold trivially for finite averages. ∎
Remark.
It is possible to generalise this argument to the case where the action of is “power bounded” in the sense that there is some uniform constant such that for (-almost) all , . However the argument for our main theorem necessitates setting .
The following argument is a generalisation of a proof of Garrett Birkhoff [2] to the amenable setting. The statement of the theorem is weaker than results which are already in Greenleaf’s article [11], but we include the argument for several reasons. One is that it is very short; another is that we will ultimately derive a bound on -fluctuations via a modification of this proof; and finally, the proof indicates additional information about the limiting behaviour of the norm of , namely that .
Theorem 20.
Let be a lcsc amenable group with compact Følner sequence , and let be a uniformly convex Banach space such that acts Borel strongly on via the representation . Then for every , the sequence of averages converges in norm .
Proof.
Without loss of generality, we assume (otherwise, simply replace with ). Define . Fix an arbitrary , and let be some index such that . Let denote the modulus of uniform convexity. Fix a Følner convergence modulus for , and let be arbitrary. Suppose is an index such that . (If no such exists then this means that after , the sequence has converged to within .) Then, from uniform convexity, we know that
Additionally, it follows from Lemma 19 that , and thus
But , so this implies
In turn, we know that , so
In fact, it follows that also, for any index , since . Now, choosing , we have that both and . Thus,
Since can be chosen to be arbitrarily small (this merely implies that and are very large), we see that . But since our choice of was arbitrary, it follows that in fact .
Moreover this implies that converges in norm. For if this were not the case, then we could find some such that infinitely often; in fact, there must be some such that infinitely often, with the further restriction333Note, here, that we are invoking the the fact that a sequence in a metric space converges iff, for any with , the sequence contains only finitely many -fluctuations at distance , for each ; cf. discussion of this equivalence in the remark immediately following Definition 1. that . Now, pick and small enough that , and pick both and such that: and are sufficiently large that (which we can do since ); and, moreover, pick and in such a way that and . The computation above shows that for larger than and we have that ; but since we chose , this implies that , which contradicts the definition of . ∎
We now proceed to deriving a quantitative analogue of this result. We first do so “at distance ”, and then recover a global bound in the case where the Følner sequence is fast (in the sense of Definition 17). The only really non-explicit part of the proof of the preceding theorem was the step where we used the fact that an infimum of a real sequence exists. Therefore the principal innovation in what follows is the avoidance of the direct invocation of this fact.
Fix a nondecreasing Følner convergence modulus for . Initially let us consider the case where . Suppose that . Moreover, we pick some , and suppose that . Then the computation from the previous proof shows that
More generally, if with , it follows that
Now, choosing , we have that both and . Thus,
Therefore let equal the least index greater than (equivalently, greater than , since is nondecreasing in ) such that . The previous calculation shows that . More generally, we have that
So simply from the fact that , these expressions derive a contradiction on the least such that
since this would imply that . That is, the contradiction implies that the th epsilon fluctuation could not have occurred. We have no a priori information on the values of and , except that both are at most . Therefore, we have the following uniform bound:
where tracks the indices of the subsequence along which -fluctuations occur. This is actually one more than the number of -fluctuations, so instead we have that the number of -fluctuations is bounded by .
If we happen to have any lower bound on the infimum of , we can sharpen the previous calculation. Instead of using the fact that for all , , we use the fact that for some . To wit, if is large enough that
then this would imply that , a contradiction. Therefore we have the bound
and so the number of -fluctuations is bounded by .
In the case where , a small modification must be made. We can make the substitution , so that , and conclude (provided that that has at most -fluctuations at distance , or rather at most -fluctuations at distance since . But since iff , this actually tells us that has at most -fluctuations at distance . Likewise, if we happen to have a lower bound on the infimum of , since iff we have that has at most -fluctuations at distance .
To summarise, we have shown that:
Theorem 21.
(“Main theorem”) Let be a uniformly convex Banach space with modulus . Fix and . Pick some such that if , or if . Then if is a lcsc amenable group that acts Borel strongly on via the representation , with Følner sequence , the sequence has at most -fluctuations at distance if , or at most -fluctuations at distance if . If we know that , then we can sharpen the bound to in the case where , or in the case where .
Corollary 22.
In the above setting, suppose that is -fast if , or -fast if respectively. Then the sequence has at most -fluctuations (resp. -fluctuations).
Proof.
We know from the theorem that, if , there are at most -fluctuations at distance . This leaves the possibility that there are some -fluctuations in the many gaps of width , and also that there are some -fluctuations in between the last possible index given by the previous theorem, and the index at which contradiction is achieved. This last interval at the end is at most wide as well. The reasoning for the case is identical. ∎
Example 23.
We mention two special cases of interest in the above setting. For brevity we restrict ourselves to the case where and , but the reader can make the obvious substitutions.
(1) In the case where is -uniformly convex with constant , then in this setting has at most -fluctuations.
(2) In the case where for , then in this setting has at most -fluctuations.
3. Discussion
The prospect of obtaining quantitative convergence information for ergodic averages via a modification of Garrett Birkhoff’s argument [2] has been previously explored, in the -action setting, by Kohlenbach and Leuştean [17] and subsequently by Avigad and Rute [1].
Our proof was carried out in the setting where the acted upon space was assumed to be uniformly convex, and indeed our bound on the number of fluctuations explicitly depends on the modulus of uniform convexity. Nonetheless, it is natural to ask whether an analogous result might be obtained for a more general class of acted upon spaces.
However, it has already been observed, in the case where , that there exists a separable, reflexive, and strictly convex Banach space such that for every and , there exists an such that has at least -fluctuations [1]. This counterexample applies equally to bounds on the rate of metastability (for more on metastability and its relationship with fluctuation bounds, see for instance [1, Section 5]). However, this counterexample does not directly eliminate the possibility of a fluctuation bound for , so the question of a “quantitative mean ergodic theorem for amenable groups” remains unresolved.
What about the choice of acting group? Our assumptions on (amenable and countable discrete or lcsc) where selected because this is the most general class of groups which have Følner sequences. Our argument depends essentially on Følner sequences; indeed, proofs of ergodic theorems for actions of non-amenable groups have a qualitatively different structure. Remarkably, there are certain classes of non-amenable groups whose associated ergodic theorems have much stronger convergence behaviour than the classical () setting; for recent progress on quantitative ergodic theorems in the non-amenable setting, we refer the reader to the book and survey article of Gorodnik and Nevo [9, 10].
We should also mention quantitative bounds for pointwise ergodic theorems. For such results go as far back as Bishop’s upcrossing inequality [3]. Inequalities of this type have also been found for by Kalikow and Weiss (for ) [16]; more recently Moriakov has modified the Kalikow and Weiss argument to give an upcrossing inequality for symmetric ball averages in groups of polynomial growth [21], and (simultaneously with the preparation of this article) Gabor [8] has extended this strategy to give an upcrossing inequality for countable discrete amenable groups with Følner sequences satisfying a strengthening of Lindenstrauss’s temperedness condition.
For both norm and pointwise convergence of ergodic averages, it is sometimes possible to deduce convergence behaviour which is stronger than -fluctuations and/or upcrossings but weaker than an explicit rate of convergence, namely that there exists an appropriate variational inequality. Jones et al. have succeeded in proving numerous variational inequalities, both for norm and pointwise convergence, for a large class of Følner sequences in and [14, 15]. (In particular, the result proved in the present article is known not to be sharp when specialised to the case where and , due to these existing results.) However, their methods, which rely on a martingale comparison estimate and a Calderón-Zygmund decomposition argument, exploit numerous incidental geometric properties of which do not hold for many other groups. It would be interesting to determine which other groups enjoy similar variational inequalities.
Acknowledgements
The author thanks his advisor Jeremy Avigad, under whom this work was completed, for his guidance and support. The author also thanks Clinton Conley, Yves Cornulier, and Henry Towsner for helpful discussions. In addition, the author thanks the anonymous referee, whose comments led to a number of improvements to the exposition of the article.
References
- [1] Jeremy Avigad and Jason Rute, Oscillation and the mean ergodic theorem for uniformly convex Banach spaces, Ergodic Theory and Dynamical Systems 35 (2015), no. 4, 1009–1027.
- [2] Garrett Birkhoff, The mean ergodic theorem, Duke Mathematical Journal 5 (1939), no. 1, 19,20.
- [3] Errett Bishop, A constructive ergodic theorem, Journal of Mathematics and Mechanics 17 (1968), no. 7, 631–639.
- [4] Matteo Cavaleri, Computability of Følner sets, International Journal of Algebra and Computation 27 (2017), no. 07, 819–830.
- [5] by same author, Følner functions and the generic word problem for finitely generated amenable groups, Journal of Algebra 511 (2018), 388–404.
- [6] James A. Clarkson, Uniformly convex spaces, Transactions of the American Mathematical Society 40 (1936), no. 3, 396–414.
- [7] Manfred Einsiedler and Thomas Ward, Ergodic theory: with a view towards number theory, vol. 259, Springer, 2010.
- [8] Uri Gabor, Fluctuations of ergodic averages for amenable group actions, arXiv preprint arXiv:1902.07912 (2019).
- [9] Alexander Gorodnik and Amos Nevo, The ergodic theory of lattice subgroups, Princeton University Press, 2009.
- [10] by same author, Quantitative ergodic theorems and their number-theoretic applications, Bulletin of the American Mathematical Society 52 (2015), no. 1, 65–113.
- [11] Frederick P. Greenleaf, Ergodic theorems and the construction of summing sequences in amenable locally compact groups, Communications on Pure and Applied Mathematics 26 (1973), no. 1, 29–46.
- [12] Olof Hanner, On the uniform convexity of and , Arkiv för Matematik 3 (1956), no. 3, 239–244.
- [13] Tuomas Hytönen, Jan van Neerven, Mark Veraar, and Lutz Weis, Analysis in Banach spaces: Volume I: Martingales and Littlewood-Paley theory, vol. 63, Springer, 2016.
- [14] Roger L. Jones, Robert Kaufman, Joseph M. Rosenblatt, and Máté Wierdl, Oscillation in ergodic theory, Ergodic Theory and Dynamical Systems 18 (1998), no. 4, 889–935.
- [15] Roger L. Jones, Joseph M. Rosenblatt, and Máté Wierdl, Oscillation in ergodic theory: higher dimensional results, Israel Journal of Mathematics 135 (2003), no. 1, 1–27.
- [16] Steven Kalikow and Benjamin Weiss, Fluctuations of ergodic averages, Illinois Journal of Mathematics 43 (1999), no. 3, 480–488.
- [17] Ulrich Kohlenbach and Laurenţiu Leuştean, A quantitative mean ergodic theorem for uniformly convex Banach spaces, Ergodic Theory and Dynamical Systems 29 (2009), no. 6, 1907–1915.
- [18] Ulrich Kohlenbach and Pavol Safarik, Fluctuations, effective learnability and metastability in analysis, Annals of Pure and Applied Logic 165 (2014), no. 1, 266–304.
- [19] Ulrich Krengel, On the speed of convergence in the ergodic theorem, Monatshefte für Mathematik 86 (1978), no. 1, 3–6.
- [20] Elon Lindenstrauss, Pointwise theorems for amenable groups, Inventiones Mathematicae 146 (2001), no. 2, 259–295.
- [21] Nikita Moriakov, Fluctuations of ergodic averages for actions of groups of polynomial growth, Studia Mathematica 240 (2018), no. 3, 255–273.
- [22] Nikita Moriakov, On effective Birkhoff’s ergodic theorem for computable actions of amenable groups, Theory of Computing Systems 62 (2018), no. 5, 1269–1287.
- [23] Donald S. Ornstein and Benjamin Weiss, Entropy and isomorphism theorems for actions of amenable groups, Journal d’Analyse Mathématique 48 (1987), no. 1, 1–141.