This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\newverbcommand\CMRverb\newverbcommand\GRverb\newverbcommand\STverb\newverbcommand\Pverb\newverbcommand\Averb

Slow Entropy and Variational Dynamical Systems

Minhua Cheng DEPARTMENT OF MATHEMATICS, UNIVERSITY OF UTAH, SALT LAKE CITY, UT 84112 [email protected] Carlos Ospina DEPARTMENT OF MATHEMATICS, UNIVERSITY OF UTAH, SALT LAKE CITY, UT 84112 [email protected] Kurt Vinhage DEPARTMENT OF MATHEMATICS, UNIVERSITY OF UTAH, SALT LAKE CITY, UT 84112 [email protected]  and  Yibo Zhai DEPARTMENT OF MATHEMATICS, UNIVERSITY OF UTAH, SALT LAKE CITY, UT 84112 [email protected]
Abstract.

We define variational properties for dynamical systems with subexponential complexity, and study these properties in certain specific examples. By computing the value of slow entropy directly, we show that some subshifts are not variational, while a class of interval exchange transformations are variational.

1. Introduction

The metric and topological entropies for measure-preserving and topological dynamical systems are often the first and most important invariants to study. These notions of entropy are numbers assigned to a dynamical system which assigns a complexity based on the exponential growth rate of the number of distinguishable orbit segments.

We will investigate foundational properties of the slow entropy-type invariants introduced by Katok and Thouvenot [KT97], with an emphasis on establishing some results accepted as folklore, as well as some features of the usual entropy that do not pass to the invariants at subexponential rates.

We recall the usual definitions of entropy in dynamical contexts in Section 3.1. Entropy as a dynamical invariant stems from its formulation in information theory by Shannon. In the smooth setting, entropy is connected with the study of Lyapunov exponents due to the Pesin and Ledrappier-Young entropy formulas. These formulas and perspectives have played crucial roles in seemingly unrelated areas such as thermodynamical formalism, progress on the Furstenberg (×2,×3)(\times 2,\times 3)-conjecture and its generalizations, and the superrigidity phenomena for higher-rank Lie groups (the Zimmer program).

We refer the reader to [KH95, Sections 3.1, 4.3] for a more detailed introduction to the classical entropy theory, and [Kat23] for a review of the history of the standard entropy theory and more context on the history hinted at here.

1.1. The variational principle

The metric and topological entropies for continuous transformations are linked via the variational principle:

(1) htop(T)=supμThμ(T),h_{\operatorname{top}}(T)=\sup_{\mu\in\mathcal{M}^{T}}h_{\mu}(T),

where T\mathcal{M}^{T} is the set of TT-invariant Borel probability measures. If this supremum is achieved, a measure for which hμ(T)=htop(T)h_{\mu}(T)=h_{\operatorname{top}}(T) is called a measure of maximal entropy (or MME).

The variational principle is the fundamental connection between the two entropy theories, and features of a measure of maximal entropy can reveal many properties of the underlying dynamical system. For instance, for geodesic flows in negative curvature, it is conjectured that the measure of maximal entropy is the Liouville measure if and only if the underlying manifold is locally symmetric. This is known for surfaces [Kat82], but it remains open as the Katok entropy conjecture in higher dimensions.

For general flows and diffeomorphisms on surfaces, a measure of maximal entropy has fractional dimension, and the closer it is to the dimension of the manifold, the more “equidistributed” the divergence is in the space.

1.2. Slow entropy invariants

The slow entropy of a dynamical system is a class of invariants which can describe subexponential growth rates such as polynomial or logarithmic, and indeed can be given specific numerical values which are invariants of uniformly continuous or measure-preserving conjugacy, depending on the category. The terminology slow entropy was introduced by Katok and Thouvenot [KT97], but others have studied it under various other names, including measure-theoretic complexity [Fer97]. In the setting of shift spaces using language complexity (or simply the complexity, see Section 3.2). These correspond to the metric and topological slow entropy, respectively.

We discuss the definitions and basic properties in Section 2, but introduce some notation here, where the notation pχ(n)=nχp_{\chi}(n)=n^{\chi} represents the family of polynomial scales:

  • if hμ,pχ=dh_{\mu,p_{\chi}}=d, then the number of orbit types of length nn which can be distinguished by μ\mu is approximately ndn^{d}, and

  • if htop,pχ=dh_{\operatorname{top},p_{\chi}}=d, then the number of orbit types of length nn which are topologically distinguished is approximately ndn^{d}.

By analogy, the classical entropy can be seen as the slow entropy with respect to the family of exponential scales eχ(n)=eχne_{\chi}(n)=e^{\chi n}, and one may consider in general entropies at a more general family of scales aχa_{\chi}.

Slow entropy or measure-theoretic complexity has many applications in classification questions. One important characterization is those systems with minimal complexity. This was first proved by Ferenczi in [Fer97], where the result was phrased using measure-theoretic complexity. We present a proof of this theorem in Section 5.

1.3. Variational systems

It is natural to ask whether Equation 1 holds for slow entropy. We will see that the answer in general is no (1.2), motivating the following definition. We refer the reader to Section 2 for any undefined terms.

Definition 1.1.

Let (X,d)(X,d) be a metric space and T:XXT:X\to X be a transformation. We say that TT is variational at the family of scales {aχ}\set{a_{\chi}} if

htop,aχ=supμThμ,aχ.h_{\operatorname{top},a_{\chi}}=\sup_{\mu\in\mathcal{M}^{T}}h_{\mu,a_{\chi}}.

We say that TT is strongly variational at the family of scales {aχ}\set{a_{\chi}} if there exists a unique measure μ0\mu_{0} for which

htop,aχ=hμ0,aχ.h_{\operatorname{top},a_{\chi}}=h_{\mu_{0},a_{\chi}}.

Variational properties of several known examples can be deduced from from existing work:

  • Transitive translations on compact abelian groups are strongly variational at all scales (Ferenczi’s Theorem [Fer97], 5.2).

  • All continuous transformations of compact spaces are variational at exponential scale (This is the classical variational principle, see eg, [KH95, Theorem 4.5.3]).

  • Uniformly hyperbolic dynamical systems are strongly variational at exponential scale (Existence and uniqueness of MMEs for uniformly hyperbolic follows from classical works of Bowen and Margulis).

  • Transitive unipotent flows are strongly variational at polynomial scale ([KVW19]).

  • Some smooth systems obtained from combinatorial constructions are not variational ([BKW23a],[BKW23b]).

In this paper, we establish the following theorems towards understanding which systems are variational:

Theorem 1.2.

Sturmian subshifts and Denjoy circle transformations are not variational at polynomial scale.

Theorem 1.3.

There exists a full Hausdorff dimension subset of 3-IETs which are strongly variational at polynomial scale.

1.2 is proved in Section 6.1 and 1.3 is proved in Section 7.5. We also provide a precise description of a full Hausdorff dimension set of 3-IETs which are variational at polynomial scale.

1.4. Future directions

The proof of 1.3 requires strong Diophantine conditions to compute the slow entropy of the 3-IETs. While it is likely that these conditions can be relaxed, the proof suggests that intermediate behavior for slow entropy is possible. In particular, we believe that some 3-IETs are not variational.

This is perhaps less shocking after noting that interval exchanges have discontinuities. However, by adding roof functions with controlled singularities, such transformations are realized as first return maps for surface flows with stationary points. It is therefore natural to ask whether the surface flows are variational. The slow entropy of some surface flows was computed in [Kan18], so a computation of their topological slow entropy would determine their variational properties.

In Section 6.4, we observe that slow entropy does not behave like exponential entropy with respect to ergodic decompositions, and there is little hope to obtain a universal formula. In fact, one may have a system with positive slow entropy for which every ergodic component is Kronecker. It is natural to ask how large the gap between the entropy of the ergodic components and the entropy of the integrated system can become. By the usual formula for entropy given an ergodic decomposition, we know that the gap cannot be exponential.

For strongly variational systems, one can try to discern what information about the underlying system can be learned from properties of the entropy maximizing measure. Particularly, there may be analogs of the Katok entropy conjecture for systems with positive and finite entropy at polynomial scale. Correspondingly, for nonvariational systems, one should be able to identify some erratic divergence of orbits which is seen a the topological level but not detected by measures.

Finally, we note that for non-variational systems, the gap between
supμ(X)hμ,aχ\sup_{\mu\in\mathcal{M}(X)}h_{\mu,a_{\chi}} and htop,aχh_{\operatorname{top},a_{\chi}} can be very large (6.5 and 6.6). The examples we describe here are uniquely ergodic, and the metric slow entropy is 0 at all scales, but the topological slow entropy is very large. It would be interesting to find non-variational systems with many invariant measures, and variational systems which are not strongly variational at a subexponential scale.

1.5. Organization of the paper

In Section 2, we review three types of invariants for systems with zero entropy at an exponential scale. The first is the topological slow entropy which we denote htop,aχh_{\operatorname{top},a_{\chi}}. This is an invariant of dynamical systems under uniformly continuous conjugations. The second invariant is the metric slow entropy hμ,aχh_{\mu,a_{\chi}}, which is an invariant of measure-preserving systems. The third and last invariant of measure-preserving systems on metric spaces is the semi-topological slow entropy hsemi,μ,aχh_{\operatorname{semi},\mu,a_{\chi}}. We compare these definitions with the classical topological and metric entropies. In Section 3, we show that slow entropy coincides with other entropy and complexity invariants. In particular, 3.3 shows that the slow entropies are the classical entropies when aχ(n)=eχna_{\chi}(n)=e^{\chi n}, and 3.4 shows that the topological slow entropy of a subshift captures the usual complexity function studied in those settings.

In Section 4, we present two results, although not new and generally considered “folklore,” they are adaptations of standard arguments in the classical entropy setting to arbitrary scales. Particularly, 4.1 was proved for first time in [Goo69] to show that the topological entropy is larger or equal to the metric entropy. We prove that this result is still true without requiring continuity of the system, and also, to include in the inequality the semi-topological entropy. To summarize, we are able to prove, for any scale aχa_{\chi}, the inequality

hμ,aχhsemi,μ,aχhtop,aχ.h_{\mu,a_{\chi}}\leq h_{\operatorname{semi},\mu,a_{\chi}}\leq h_{\operatorname{top},a_{\chi}}.

Section 5 is mainly expository. We have parsed [Fer97, Proposition 3] in 5.2, to rewrite it to our notation and for referencing in later results of this paper. We remark that is crucial for proving 1.2. In Sections 6.1 and 6.3, we prove 1.2. We compute the semi-topological slow entropy of Sturmian systems, the main result is summarized in 6.4, and we add a discussion of Denjoy circle transformations to show that these systems are not variational. Finally, we discuss the interaction between slow entropy and the ergodic decomposition, see Section 6.4. In particular, in 6.11 we show that that the metric entropy of the geodesic flow on a flat torus is equal to 1 at polynmial scale, but 0 with respect to any ergodic measure and for any arbitrary scale.

In Section 7, we introduce the background of interval exchange transformations and prove 1.3.

Acknowledgements. The authors would like to thank Przemyslaw Berk, Jon Chaika, Adam Kanigowski, Scott Schmieding, and Daren Wei for discussions and recommendations on the direction of this paper. The authors also acknowledge the NSF award #1840190, the research training group Algebra, Geometry and Topology, which provided the space and opportunity for this work to be done.

2. Definitions

Let f:XXf\colon X\to X be a measurable transformation on a locally compact metric space. We denote (X,d)(X,d) the metric structure, and if μ\mu is a Radon measure, let (X,μ,)(X,\mu,\mathcal{B}) denote the triple determining a measure space and σ\sigma-algebra of Borel sets \mathcal{B}. If μ\mu is ff-invariant, i.e. B\forall B\in\mathcal{B} μ(f1(B))=μ(B)\mu(f^{-1}(B))=\mu(B), then we say that (X,μ,f,)(X,\mu,f,\mathcal{B}) is a measure preserving system. We do not assume μ\mu is ergodic, instead, we mention it whenever it is necessary. However, unless otherwise noted, we will assume that μ\mu is a probability measure, i.e. μ(X)=1\mu(X)=1.

2.1. Topological Slow Entropy

For xXx\in X, and ϵ>0\epsilon>0, let B(x,ϵ)={yX:d(x,y)<ϵ}B(x,\epsilon)=\{y\in X:d(x,y)<\epsilon\} denote an open ball. For nn a non-negative integer and ϵ\epsilon a positive real number, it is not hard to check that the map

dfn(x,y)=max0in1d(fi(x),fi(y))d^{n}_{f}(x,y)=\max_{0\leq i\leq n-1}d(f^{i}(x),f^{i}(y))

defines a new metric on XX, called the nn-Bowen metric, or simply Bowen metric. When ff is uniformly continuous, the metric dfnd_{f}^{n} is equivalent to dd. The (ϵ,n)(\epsilon,n)-Bowen ball is the ball in dfnd_{f}^{n} centered at xx of radius ϵ\epsilon, or equivalently the following set:

Bfn(x,ϵ)=i=0n1fi(B(fi(x),ϵ)).B^{n}_{f}(x,\epsilon)=\cap_{i=0}^{n-1}f^{-i}\left(B(f^{i}(x),\epsilon)\right).

Let Nf,K(ϵ,n)N_{f,K}(\epsilon,n) denote the minimal number of (ϵ,n)(\epsilon,n)-Bowen balls that cover a compact set KXK\subset X. Note that Nf,K(ϵ,n)<N_{f,K}(\epsilon,n)<\infty when ff is uniformly continuous.

Let Sf,K(ϵ,n)S_{f,K}(\epsilon,n) be the maximal number of disjoint (ϵ,n)(\epsilon,n)-Bowen balls that can be arranged with centers in KK, where all the centers of such a collection of Bowen balls form a maximal (ϵ,n)(\epsilon,n)-(Bowen) separating set.

The following inequalities are important in establishing that a well-defined invariant exists. They are used even in the usual definition of entropy (see, e.g. [KH95, Section 3.1.b])

(2) Sf,K(ϵ,n)\displaystyle S_{f,K}(\epsilon,n) \displaystyle\leq Nf,K(ϵ,n)\displaystyle N_{f,K}(\epsilon,n)
(3) Nf,K(2ϵ,n)\displaystyle N_{f,K}(2\epsilon,n) \displaystyle\leq Sf,K(ϵ,n).\displaystyle S_{f,K}(\epsilon,n).

Note that inequalities (2) and (3) still hold when ff is not continuous, since dfnd_{f}^{n} is still a metric in this case. Uniform continuity is usually used to guarantee the dfnd_{f}^{n} is equivalent to dd.

Definition 2.1.

A scale {aχ}\{a_{\chi}\} is a family of increasing functions,

aχ:>0>0a_{\chi}\colon\mathbb{Z}_{>0}\to\mathbb{R}_{>0}

indexed by χ0\chi\in\mathbb{R}_{\geq 0}, such that if χ<χ\chi<\chi^{\prime} then aχ=o(aχ)a_{\chi}=o(a_{\chi^{\prime}}).

Notation.

In subscripts, aχa_{\chi} is used to indicate the scale chosen beforehand, and χ\chi is used to indicate that the quantity depends on the value of the parameter χ\chi at the given scale {aχ}\{a_{\chi}\}.

We think of a scale as a family of functions indexed by χ\chi that describe the orbit growth. The corresponding slow entropy hh means that the orbits grow in time as the function aha_{h}. If the slow entropy with respect to a given scale is zero (resp. infinity) then, in time nn the orbits grow slower (resp. faster) than the sequence {ah(n)}n>0\{a_{h}(n)\}_{n\in\mathbb{Z}_{>0}} for all h+h\in\mathbb{R}_{+}.

Example 2.2.
  1. (1)

    At exponential scale aχ(n)=eχna_{\chi}(n)=e^{\chi n}, we will show, in 3.3, that the topological slow entropy (similarly, for metric slow entropy and semitopological slow entropy) and the classical topological entropy are equal.

  2. (2)

    At polynomial scale aχ(n)=nχa_{\chi}(n)=n^{\chi}, the slow topological entropy is often called polynomial topological entropy. Similar for other scales.

  3. (3)

    Another example is the logarithmic scale aχ(n)=n(logn)χa_{\chi}(n)=n(\log n)^{\chi}.

These scales have been used, for example, in [KVW19, Theorem 1.7], where it was shown that quasi-unipotent flows have positive polynomial entropy. Previously, [Kan18, Theorems 1.1 and 1.2] showed that the Korchegin flow has positive polynomial entropy, and the Arnol’d flow has positive logarithmic entropy.

We define

δf,K,χN(ϵ)=lim supnNf,K(ϵ,n)aχ(n),\delta_{f,K,\chi}^{N}(\epsilon)=\limsup_{n\to\infty}\frac{N_{f,K}(\epsilon,n)}{a_{\chi}(n)},

and

δf,K,χS(ϵ)=lim supnSf,K(ϵ,n)aχ(n).\delta_{f,K,\chi}^{S}(\epsilon)=\limsup_{n\to\infty}\frac{S_{f,K}(\epsilon,n)}{a_{\chi}(n)}.
Definition 2.3.

The slow topological entropy of ff for the scale aχa_{\chi} is

htop,aχ(f)=supKlimϵ0(sup{χ:δf,K,χN(ϵ)>0})\displaystyle h_{\operatorname{top},a_{\chi}}(f)=\sup_{K}\lim_{\epsilon\to 0}\left(\sup\left\{\chi:\delta_{f,K,\chi}^{N}(\epsilon)>0\right\}\right)
=supKlimϵ0(sup{χ:δf,K,χS(ϵ)>0}).\displaystyle=\sup_{K}\lim_{\epsilon\to 0}\left(\sup\left\{\chi:\delta_{f,K,\chi}^{S}(\epsilon)>0\right\}\right).

Here, the supremum is taken over all compact sets KK.

Note that this is well-defined by inequalities (2) and (3). Further, one may show that the innermost supremum is decreasing in ϵ\epsilon, so the limit exists as ϵ0\epsilon\to 0.

Remark 2.4.

We index the family of scales (aχ)(a_{\chi}) by the non-negative reals. When taking the supremum over some property PP of χ\chi, it can happen that the set {χ:P(χ) holds}\{\chi:P(\chi)\text{ holds}\} is empty. In such a case, we use the convention that the sup\sup is zero. The set {χ:δf,K,χN(ϵ)>0}\left\{\chi:\delta_{f,K,\chi}^{N}(\epsilon)>0\right\} is an interval that starts at zero. However, the point χ^=sup{χ:δf,K,χN(ϵ)>0}\hat{\chi}=\sup\left\{\chi:\delta_{f,K,\chi}^{N}(\epsilon)>0\right\} may or may not belong to the interval. All this follows because if χ>χ^\chi>\hat{\chi}, then by definition we have χ{χ:δf,K,χN(ϵ)>0};\chi\not\in\left\{\chi:\delta_{f,K,\chi}^{N}(\epsilon)>0\right\}; if χ<χ^,\chi<\hat{\chi}, again by definition we have γ(χ,χ^)\gamma\in(\chi,\hat{\chi}) such that δf,K,γN>0,\delta^{N}_{f,K,\gamma}>0, and thus

lim supnNf,K(ϵ,n)aχ(n)=lim supnNf,K(ϵ,n)aγ(n)aγ(n)aχ(n)lim supnNf,K(ϵ,n)aγ(n)>0\begin{split}\limsup_{n\to\infty}\frac{N_{f,K}(\epsilon,n)}{a_{\chi}(n)}&=\limsup_{n\to\infty}\frac{N_{f,K}(\epsilon,n)}{a_{\gamma}(n)}\cdot\frac{a_{\gamma}(n)}{a_{\chi}(n)}\\ &\geq\limsup_{n\to\infty}\frac{N_{f,K}(\epsilon,n)}{a_{\gamma}(n)}>0\end{split}

since aχ(n)=o(aγ(n))a_{\chi}(n)=o(a_{\gamma}(n)).

In conclusion, {χ:δf,K,χN(ϵ)>0}\left\{\chi:\delta_{f,K,\chi}^{N}(\epsilon)>0\right\} is of the form [0,χ^)[0,\hat{\chi}) or [0,χ^][0,\hat{\chi}].

Remark 2.5.

While many systems on compact metric spaces are continuous, there are certain natural systems which have discontinuities appearing. In this paper we treat the case of 3-IETs, whose discontinuities appear naturally when considering first return maps for Poincaré sections of flows. Crucially, it is important to note that it still makes sense to consider topological entropy for such systems, but the topological entropy may now depend on the choice of metric on XX.

2.2. Metric Slow Entropy

For a probability measure-preserving system (X,μ,f,)(X,\mu,f,\mathcal{B}), consider a finite measurable partition 𝒫={P1,,Pk}\mathcal{P}=\{P_{1},\dots,P_{k}\}. We call each set PiP_{i} an atom of 𝒫\mathcal{P}. Note that every xXx\in X defines a coding sequence (xs):=(xs)s0(x_{s})\mathrel{\mathop{:}}=(x_{s})_{s\in\mathbb{Z}_{\geq 0}}, where xs=jx_{s}=j if fs(x)Pjf^{s}(x)\in P_{j}. For any x,yXx,y\in X, the Hamming distance with respect to the partition 𝒫\mathcal{P} is the quantity

d¯f,𝒫n(x,y)=1|{0sn1:xs=ys}|n,\overline{d}_{f,\mathcal{P}}^{n}(x,y)=1-\frac{|\{0\leq s\leq n-1:x_{s}=y_{s}\}|}{n},

where |||\cdot| is the counting measure. The number d¯f,𝒫n(x,y)\overline{d}_{f,\mathcal{P}}^{n}(x,y) is the proportion of times for which the orbits of xx and yy lie in different atoms of the partition 𝒫\mathcal{P} up to time nn.

For n0n\geq 0 and ϵ>0\epsilon>0, the (ϵ,n)(\epsilon,n)-Hamming ball centered at xXx\in X is the set

Bf,𝒫n(x,ϵ)={yX:d¯f,𝒫n(x,y)<ϵ}.B^{n}_{f,\mathcal{P}}(x,\epsilon)=\{y\in X:\overline{d}_{f,\mathcal{P}}^{n}(x,y)<\epsilon\}.

Next, FF represents a finite subset of XX, and define the number

Sf,𝒫(ϵ,n)=min{card(F):μ(xFBf,𝒫n(x,ϵ))>1ϵ}.S_{f,\mathcal{P}}(\epsilon,n)=\min\left\{\operatorname{card}(F):\mu\left(\bigcup_{x\in F}B^{n}_{f,\mathcal{P}}(x,\epsilon)\right)>1-\epsilon\right\}.

For a given scale aχa_{\chi}, we define

δf,𝒫,χS(ϵ)=lim supnSf,𝒫(ϵ,n)aχ(n),\delta_{f,\mathcal{P},\chi}^{S}(\epsilon)=\limsup_{n\to\infty}\frac{S_{f,\mathcal{P}}(\epsilon,n)}{a_{\chi}(n)},

and the slow metric entropy for the partition 𝒫\mathcal{P} is

hμ,aχ,𝒫(f)=limϵ0(sup{χ:δf,𝒫,χS(ϵ)>0}).h_{\mu,a_{\chi},\mathcal{P}}(f)=\lim_{\epsilon\to 0}\left(\sup\left\{\chi:\delta_{f,\mathcal{P},\chi}^{S}(\epsilon)>0\right\}\right).
Definition 2.6.

The slow metric entropy of ff with respect to the scale aχa_{\chi} is defined as

hμ,aχ(f)=sup𝒫hμ,aχ,𝒫(f).h_{\mu,a_{\chi}}(f)=\sup_{\mathcal{P}}h_{\mu,a_{\chi},\mathcal{P}}(f).

Here the sup\sup is taken over all finite measurable partitions of XX.

2.3. Semi-topological slow entropy

A notion between topological and metric entropy can be obtained when a natural metric and measure are linked. Let

Ssemi,f(ϵ,n)=min{card(F):μ(xFBfn(x,ϵ))>1ϵ}.S_{\operatorname{semi},f}(\epsilon,n)=\min\left\{{\operatorname{card}(F):\mu\left(\bigcup_{x\in F}B_{f}^{n}(x,\epsilon)\right)>1-\epsilon}\right\}.

We similarly let δsemi,f,χS(ϵ)=lim supnSsemi,f(ϵ,n)aχ(n)\delta_{\operatorname{semi},f,\chi}^{S}(\epsilon)=\displaystyle\limsup_{n\to\infty}\dfrac{S_{\operatorname{semi},f}(\epsilon,n)}{a_{\chi}(n)}.

Definition 2.7.

The semi-topological slow entropy of ff with respect to μ\mu is

hsemi,μ,aχ(f)=lim supϵ0(supχ{χ:δsemi,f,χS(ϵ)>0}).h_{\operatorname{semi},\mu,a_{\chi}}(f)=\limsup_{\epsilon\to 0}\left(\sup_{\chi}\left\{\chi:\delta_{\operatorname{semi},f,\chi}^{S}(\epsilon)>0\right\}\right).

3. Slow Entropy as other growth invariants

The slow entropy invariants we have defined are in fact generalizations of the standard classification tools. We describe the connections here.

3.1. Classical entropy as slow entropy

Throughout, we assume that (X,d)(X,d) is a metric space, μ\mu is a probability measure on (X,)(X,\mathcal{B}), where \mathcal{B} is the Borel σ\sigma-algebra of (X,d)(X,d), and the f:XXf:X\to X is a μ\mu-preserving transformation.

Definition 3.1.

The (classical) topological entropy htop(f)h_{\operatorname{top}}(f) is

(4) htop(f)=supKlimϵ0lim supnlog(Nf,K(ϵ,n))nh_{\operatorname{top}}(f)=\sup_{K}\lim_{\epsilon\to 0}\limsup_{n\to\infty}\frac{\log(N_{f,K}(\epsilon,n))}{n}

We refer to the reader to [KH95, Section 3.1.b] or [Bow71] for an alternative definition using Sf,K(ϵ,n)S_{f,K}(\epsilon,n), and a discussion on how the classical topological entropy does not depend on the choice of metric dd determining the topology when XX is compact and ff is continuous.

Definition 3.2.

The (classical) metric entropy of hμ(f)h_{\mu}(f) is defined as

(5) hμ(f)=sup𝒫hμ,𝒫(f),h_{\mu}(f)=\sup_{\mathcal{P}}h_{\mu,\mathcal{P}}(f),

where the supremum is taken over all finite measurable partitions 𝒫\mathcal{P} and

(6) hμ,𝒫(f)=limn1nHμ(i=0n1fi(𝒫)).h_{\mu,\mathcal{P}}(f)=\lim_{n\to\infty}\frac{1}{n}\text{H}_{\mu}(\vee_{i=0}^{n-1}f^{-i}(\mathcal{P})).

In Equation 6, for any measurable partition 𝒫\mathcal{P},

Hμ(𝒫):=P𝒫μ(P)log(μ(P)).\text{H}_{\mu}(\mathcal{P})\mathrel{\mathop{:}}=\sum_{P\in\mathcal{P}}-\mu(P)\log(\mu(P)).

We have the following well-known theorem in the literature that very few authors proved.

Theorem 3.3 (Exponential scales in slow entropy).

Let aχ(n)=eχna_{\chi}(n)=e^{\chi n} and μ\mu be an ergodic probability measure. Then

hμ,aχ(f)=hsemi,μ,aχ=hμ(f)andhtop,aχ(f)=htop(f).h_{\mu,a_{\chi}}(f)=h_{\operatorname{semi},\mu,a_{\chi}}=h_{\mu}(f)\quad\text{and}\quad h_{\operatorname{top},a_{\chi}}(f)=h_{\operatorname{top}}(f).
Proof.

To prove that htop,aχ=htoph_{\operatorname{top},a_{\chi}}=h_{\operatorname{top}}, we claim that for any compact set KK and ϵ^>0\hat{\epsilon}>0,

(7) sup{χ:δf,K,χN(ϵ^)>0}=lim supnlog(Nf,K(ϵ^,n))n.\sup\{\chi:\delta^{N}_{f,K,\chi}(\hat{\epsilon})>0\}=\limsup_{n\to\infty}\frac{\log(N_{f,K}(\hat{\epsilon},n))}{n}.

First of all, lim suplog(Nf,K(ϵ^,n))n=\limsup\frac{\log(N_{f,K}(\hat{\epsilon},n))}{n}=\infty, is equivalent to: there exists nin_{i}\to\infty such that for all M,γ>0M,\gamma>0, there exists n,n_{*}, such that if ni>nn_{i}>n_{*}, then log(Nf,K(ϵ^,ni))ni>M+log(1+γ)\frac{\log(N_{f,K}(\hat{\epsilon},n_{i}))}{n_{i}}>M+\log(1+\gamma). This is equivalent to Nf,K(ϵ^,ni)eMni>(1+γ)ni\frac{N_{f,K}(\hat{\epsilon},n_{i})}{e^{Mn_{i}}}>(1+\gamma)^{n_{i}} for all ni>n.n_{i}>n_{*}. Remember that

(8) δf,K,χN(ϵ^)=lim supnNf,K(ϵ^,n)eχn.\delta^{N}_{f,K,\chi}(\hat{\epsilon})=\limsup_{n\to\infty}\frac{N_{f,K}(\hat{\epsilon},n)}{e^{\chi n}}.

So, δf,K,MN(ϵ^)=.\delta_{f,K,M}^{N}(\hat{\epsilon})=\infty. Since M>0M>0 is arbitrary; we conclude that it is equivalent that the left side in Equation 7 is equal to .\infty.

Now, assume that χ^=lim supnlog(Nf,K(ϵ^,n))n<\hat{\chi}=\limsup_{n\to\infty}\frac{\log(N_{f,K}(\hat{\epsilon},n))}{n}<\infty.

If χ>χ^\chi>\hat{\chi}, then the expression in Equation 8 is zero, because

(9) lim supnNf,K(ϵ^,n)eχn=lim supnNf,K(ϵ^,n)eχ^nlim supne(χ^χ)n=0.\limsup_{n\to\infty}\frac{N_{f,K}(\hat{\epsilon},n)}{e^{\chi n}}=\limsup_{n\to\infty}\frac{N_{f,K}(\hat{\epsilon},n)}{e^{\hat{\chi}n}}\limsup_{n\to\infty}e^{(\hat{\chi}-\chi)n}=0.

This implies that

(10) χ^sup{χ:δf,K,χN(ϵ^)>0}.\hat{\chi}\geq\sup\{\chi:\delta^{N}_{f,K,\chi}(\hat{\epsilon})>0\}.

If χ^=0\hat{\chi}=0, Equation 10 is an equality, and it proves Equation 7. If χ^>0\hat{\chi}>0, let χ<χ^\chi<\hat{\chi}, and nin_{i}\to\infty any sequence of positive integers such that Nf,K(ϵ^,ni)eχ^nic>0,\frac{N_{f,K}(\hat{\epsilon},n_{i})}{e^{\hat{\chi}n_{i}}}\geq c>0, for a positive constant cc. Then, we have that

(11) Nf,K(ϵ^,ni)eχnie(χχ^)nic.\frac{N_{f,K}(\hat{\epsilon},n_{i})}{e^{\chi n_{i}}}e^{(\chi-\hat{\chi})n_{i}}\geq c.

Since e(χχ^)ni0e^{(\chi-\hat{\chi})n_{i}}\to 0, then Nf,K(ϵ^,ni)eχni.\frac{N_{f,K}(\hat{\epsilon},n_{i})}{e^{\chi n_{i}}}\to\infty. And then Equation 8 with χ<χ^\chi<\hat{\chi} is equal to \infty. This implies that

(12) χ^=sup{χ:δf,K,χN(ϵ^)>0}.\hat{\chi}=\sup\{\chi:\delta^{N}_{f,K,\chi}(\hat{\epsilon})>0\}.

Finally, to prove that htop,aχ=htoph_{\operatorname{top},a_{\chi}}=h_{\operatorname{top}}, assume that htop,aχh_{\operatorname{top},a_{\chi}} and htoph_{\operatorname{top}} are finite. We focus only on the case when both entropies are finite and leave the case of infinite entropy to the reader.

It is enough to prove that for arbitrary ϵ>0\epsilon>0, then |htop,aχhtop|<ϵ.\left|h_{\operatorname{top},a_{\chi}}-h_{\operatorname{top}}\right|<\epsilon.

By definition of the entropies, and by triangle inequality, there exists KXK\subset X compact and ϵ^>0\hat{\epsilon}>0 sufficiently small, such that

|htop,aχhtop||htop,aχsup{χ:δf,K,χN(ϵ^)>0}|+|htoplim supnlog(Nf,K(ϵ^,n))n|+|sup{χ:δf,K,χN(ϵ^)>0}lim supnlog(Nf,K(ϵ^,n))n|<23ϵ+|sup{χ:δf,K,χN(ϵ^)>0}lim supnlog(Nf,K(ϵ^,n))n|.\begin{split}\left|h_{\operatorname{top},a_{\chi}}-h_{\operatorname{top}}\right|&\leq\left|h_{\operatorname{top},a_{\chi}}-\sup\{\chi:\delta^{N}_{f,K,\chi}(\hat{\epsilon})>0\}\right|\\ +&\left|h_{\operatorname{top}}-\limsup_{n\to\infty}\frac{\log(N_{f,K}(\hat{\epsilon},n))}{n}\right|\\ +&\left|\sup\{\chi:\delta^{N}_{f,K,\chi}(\hat{\epsilon})>0\}-\limsup_{n\to\infty}\frac{\log(N_{f,K}(\hat{\epsilon},n))}{n}\right|\\ <&\frac{2}{3}\epsilon+\left|\sup\{\chi:\delta^{N}_{f,K,\chi}(\hat{\epsilon})>0\}-\limsup_{n\to\infty}\frac{\log(N_{f,K}(\hat{\epsilon},n))}{n}\right|.\end{split}
|htop,aχhtop|<23ϵ.\left|h_{\operatorname{top},a_{\chi}}-h_{\operatorname{top}}\right|<\frac{2}{3}\epsilon.

To prove that hμ,aχ(f)=hμ(f)h_{\mu,a_{\chi}}(f)=h_{\mu}(f). First, assume that hμ(f)h_{\mu}(f) is finite. Let 𝒫={P1,,Pk}\mathcal{P}=\{P_{1},\dots,P_{k}\} be a finite measurable partition with hμ(f,𝒫)=hh_{\mu}(f,\mathcal{P})=h. By Shannon-McMillan-Breiman Theorem [VO16, Theorem 9.3.1], for the partition 𝒫\mathcal{P} we have that

1nlogμ([x]0n1)hμ-a.e.-\frac{1}{n}\log\mu([x]_{0}^{n-1})\rightarrow{h}\quad\mu{\text{-}a.e.}

We use [x]0n1[x]_{0}^{n-1} to denote the atom in i=0n1Ti𝒫\displaystyle\vee_{i=0}^{n-1}T^{-i}\mathcal{P} that contains xx. Let ϵ>0\epsilon>0, for nn be large enough, we get

μ(x:|1nlogμ([x]0n1)h|<ϵ)>1ϵ.\mu\Big{(}x:\left|-\frac{1}{n}\log\mu([x]_{0}^{n-1})-h\right|<\epsilon\Big{)}>1-\epsilon.

This implies:

(13) μ(x:en(h+ϵ)<μ([x]0n1)<en(hϵ))>1ϵ.\mu\Big{(}x:e^{-n(h+\epsilon)}<\mu([x]_{0}^{n-1})<e^{-n(h-\epsilon)}\Big{)}>1-\epsilon.

Note for any y[x]0n1y\in[x]_{0}^{n-1}, yj=xjy_{j}=x_{j} for any 0jn10\leq j\leq n-1, therefore Sf,𝒫(ϵ,n)en(h+ϵ)S_{f,\mathcal{P}}(\epsilon,n)\leq e^{n(h+\epsilon)}. Hence, when aχ(n)=eχna_{\chi}(n)=e^{\chi n}, hμ,aχ,𝒫(f)hh_{\mu,a_{\chi},\mathcal{P}}(f)\leq h. Moreover, Bf,𝒫n(x,ϵ)B_{f,\mathcal{P}}^{n}(x,\epsilon) is covered by at most (nnϵ)(nϵk)\dbinom{n}{\lfloor n\epsilon\rfloor}(\lfloor n\epsilon\rfloor^{k}) atoms in i=0n1Ti𝒫\displaystyle\bigvee_{i=0}^{n-1}T^{-i}\mathcal{P}. This is because for any yBf,𝒫n(x,ϵ)y\in B^{n}_{f,\mathcal{P}}(x,\epsilon), {xi}i=0n1\{x_{i}\}_{i=0}^{n-1} and {yi}i=0n1\{y_{i}\}_{i=0}^{n-1} differ in at most nϵ\lfloor n\epsilon\rfloor positions, each one has at most kk choices. From Equation 13, each atom is of measure at least en(h+ϵ)e^{-n(h+\epsilon)}, and they cover space of measure at least 1ϵ1-\epsilon. Hence, Sf,𝒫(ϵ,n)(1ϵ)en(h𝒪(ϵ))S_{f,\mathcal{P}}(\epsilon,n)\geq(1-\epsilon)e^{n(h-\mathcal{O}(\epsilon))} since kk is a fixed constant and

(nnϵ)en(ϵln(ϵ)+(1ϵ)log(1ϵ))/2πnϵ(1ϵ)\dbinom{n}{n\epsilon}\approx e^{n(\epsilon\ln(\epsilon)+(1-\epsilon)\log(1-\epsilon))}/\sqrt{2\pi n\epsilon(1-\epsilon)}

when nn is large enough. Therefore, hμ,aχ,𝒫(f)hh_{\mu,a_{\chi},\mathcal{P}}(f)\geq h, proving the equality hμ,𝒫(f)=hμ,aχ,𝒫(f)h_{\mu,\mathcal{P}}(f)=h_{\mu,a_{\chi},\mathcal{P}}(f). By definition, sup𝒫hμ,𝒫(f)=hμ(f)\displaystyle\sup_{\mathcal{P}}h_{\mu,\mathcal{P}}(f)=h_{\mu}(f) and hμ,aχ(f)=sup𝒫hμ,aχ,𝒫(f)\displaystyle h_{\mu,a_{\chi}}(f)=\sup_{\mathcal{P}}h_{\mu,a_{\chi},\mathcal{P}}(f), these give hμ(f)=hμ,aχ(f)h_{\mu}(f)=h_{\mu,a_{\chi}}(f). If hμ(f)=h_{\mu}(f)=\infty, then for any finite partition 𝒫\mathcal{P} with finite entropy, we still have hμ,𝒫(f)=hμ,aχ,𝒫(f)h_{\mu,\mathcal{P}}(f)=h_{\mu,a_{\chi},\mathcal{P}}(f), and the result follows by taking the sup\sup over a sequence of partitions with finite entropy that go to infinity. Therefore, hμ(f)=hμ,aχ(f)=h_{\mu}(f)=h_{\mu,a_{\chi}}(f)=\infty.

We omit the proof of hsemi,μ,aχ=hμh_{\operatorname{semi},\mu,a_{\chi}}=h_{\mu} and refer the reader to [Kat80, Theorem (I.I)]. ∎

3.2. Shift complexity as slow entropy

Consider a finite alphabet 𝒜={1,,n}\mathcal{A}=\set{1,\dots,n} and the space

Ω=𝒜={ω=(,ω2,ω1,ω0,ω1,ω2,):ωi𝒜 for all i}.\Omega=\mathcal{A}^{\mathbb{Z}}=\set{\omega=(\dots,\omega_{-2},\omega_{-1},\omega_{0},\omega_{1},\omega_{2},\dots):\omega_{i}\in\mathcal{A}\mbox{ for all }i\in\mathbb{Z}}.

The set Ω\Omega is called the shift space on nn symbols and has a canonical dynamical system attached, the shift map σ:ΩΩ\sigma:\Omega\to\Omega defined by

σ(ω)n:=ωn+1.\sigma(\omega)_{n}:=\omega_{n+1}.

In other words, the sequence σ(ω)\sigma(\omega) is the same as ω\omega, except that the 0 position of the sequence is shifted to the right by one index. A subshift is a closed σ\sigma-invariant set XX, and the language of XX is the set

={(α0,,αm):there exists ωX with ωi=αi for all i=0,,m}.\mathcal{L}=\set{(\alpha_{0},\dots,\alpha_{m}):\mbox{there exists }\omega\in X\mbox{ with }\omega_{i}=\alpha_{i}\mbox{ for all }i=0,\dots,m}.

That is, \mathcal{L} contains all of the finite words in XX. To clarify the dynamical system, we let σX\sigma_{X} denote the restriction of σ\sigma to XX. One may consider the (language) complexity of XX, which counts the growth rate of \mathcal{L}. That is, if we let m\mathcal{L}_{m} denote the words in the langauge of length mm, we consider the function

pn(X)=#m.p_{n}(X)=\#\mathcal{L}_{m}.

The language complexity has been studied carefully for a variety of subshifts, and we will not provide an exhaustive survey here, but some recent works on the complexity of subshifts and their applications include [CK20, CJKS22, DOP22, CP23, PS23a].

Fix the following metric on Ω\Omega (and correspondingly the induced metric on XX) as

d(ω,η)=2k, where k=inf{i0:ωiηi or ωiηi}.d(\omega,\eta)=2^{-k},\mbox{ where }k=\inf\set{i\geq 0:\omega_{i}\not=\eta_{i}\mbox{ or }\omega_{-i}\not=\eta_{-i}}.

We can use the complexity to compute the topological slow entropy. Recall the definition of Nf,Ω(ϵ,n)N_{f,\Omega}(\epsilon,n) as given at the start of Section 2.1

Proposition 3.4.

If σX:XX\sigma_{X}:X\to X is a subshift, then

Nf,Ω(2(k1),n)=p2k+1+n(X).N_{f,\Omega}(2^{-(k-1)},n)=p_{2k+1+n}(X).
Proof.

Observe that by definition of the metric, d(ω,η)<2(k1)d(\omega,\eta)<2^{-(k-1)} if and only if they agree on the indices ranging from k-k up to kk. Hence d(σ(ω),σ(η))<2(k1)d(\sigma^{\ell}(\omega),\sigma^{\ell}(\eta))<2^{-(k-1)} if and only if ω\omega and η\eta agree on the indices ranging from k\ell-k to +k\ell+k. Therefore, σ\sigma and η\eta are 2(k1)2^{-(k-1)}-close in the metric dσnd^{n}_{\sigma} if and only if they agree on the indices ranging from k-k to n+kn+k. Since there are 2k+1+n2k+1+n such indices, the follows that we for each finite word of length 2k+1+n2k+1+n, we must choose a representative to cover the corresponsing Bowen ball: every such word has an element which must belong to a Bowen cover, and each Bowen ball must be centered at some point and hence can only cover one such word. The result follows. ∎

This yields the immediate corollary, which shows that in shift spaces, the topological slow entropy captures the growth rate of the complexity function.

Corollary 3.5.

If aχa_{\chi} is a scale, σX:XX\sigma_{X}:X\to X is a subshift, and

lim supnpn(X)aχ(n)={,χ<χ00,χ>χ0\limsup_{n\to\infty}\dfrac{p_{n}(X)}{a_{\chi}(n)}=\left\{\begin{array}[]{ll}\infty,&\chi<\chi_{0}\\ 0,&\chi>\chi_{0}\end{array}\right.

then htop,aχ(σX)=χ0h_{\operatorname{top},a_{\chi}}(\sigma_{X})=\chi_{0}.

4. Structural Theorems

4.1. Slow Goodwyn’s theorem

The following result states a relation among the different entropies that we defined in Section 2. In the setting of classical entropy theory, this is part of the variational principle, due to Goodwyn [Goo69]. It states that under general circumstances, the metric entropy is bounded above by the semi-topological entropy and that the topological entropy is the largest of the previous two.

Theorem 4.1 (Goodwyn).

Let f:XXf\colon X\to X be a measurable transformation of a compact metric space, and f(X)\mathcal{M}^{f}(X) denote the space of ff-invariant Borel probability measures on XX. Then for any measure μf(X)\mu\in\mathcal{M}^{f}(X)

(14) hμ,aχ(f)hsemi,μ,aχ(f)htop,aχ(f).h_{\mu,a_{\chi}}(f)\leq h_{\operatorname{semi},\mu,a\chi}(f)\leq h_{\operatorname{top},a_{\chi}}(f).
Remark 4.2.

As in 2.5, we note that continuity is not required for this theorem.

To prove this theorem, we need an auxiliary result on the existence of certain partitions. In this result, the remarkable part is that the atoms of the partitions are small and their boundary is of measure zero.

Lemma 4.3.

Let (X,d)(X,d) be a compact metric space with a probability measure μ\mu. Then for every δ>0\delta>0, there exists a partition 𝒫={P1,,Pk}\mathcal{P}=\{P_{1},\dots,P_{k}\} such that diam(Pi)<δ\operatorname{diam}(P_{i})<\delta for every i=1,,ki=1,\dots,k and if Xϵ={xX:d(x,𝒫)<ϵ}X_{\epsilon}=\{x\in X:d(x,\partial\mathcal{P})<\epsilon\}, limϵ0μ(Xϵ)=0\lim_{\epsilon\to 0}\mu(X_{\epsilon})=0, where

d(x,𝒫):=maxP𝒫d(x,P).\displaystyle d(x,\partial\mathcal{P})\mathrel{\mathop{:}}=\max_{P\in\mathcal{P}}d(x,\partial P).

Furthermore, if x,yXx,y\in X belong to different atoms of the partition and x,yXϵx,y\not\in X_{\epsilon}, then d(x,y)ϵd(x,y)\geq\epsilon.

Proof of 4.1.

We will start by proving that hμ,aχhsemi,μ,aχh_{\mu,a_{\chi}}\leq h_{\operatorname{semi},\mu,a_{\chi}}. Let ϵ>0\epsilon>0. Let 𝒫ϵ\mathcal{P}_{\epsilon} be the partition in 4.3 for δ=ϵ\delta=\epsilon, so each set in 𝒫ϵ\mathcal{P}_{\epsilon} has diameter at most ϵ\epsilon. From 4.3, we conclude that the open set Xϵ^X_{\hat{\epsilon}} has μ\mu measure at most ϵ\epsilon for some ϵ^<ϵ\hat{\epsilon}<\epsilon, and if x,yXϵ^x,y\in X_{\hat{\epsilon}}, d(x,y)<ϵ^d(x,y)<\hat{\epsilon}, then [x]𝒫ϵ=[y]𝒫ϵ[x]_{\mathcal{P}_{\epsilon}}=[y]_{\mathcal{P}_{\epsilon}}. We have that the compact set Kϵ^:=X\Xϵ^K_{\hat{\epsilon}}\mathrel{\mathop{:}}=X\backslash X_{\hat{\epsilon}} has μ\mu measure at least 1ϵ1-\epsilon, and satisfies that for any xKϵ^x\in K_{\hat{\epsilon}}, we must have

(15) Bf,𝒫ϵn(x,ϵ)Bfn(x,ϵ^).B^{n}_{f,\mathcal{P}_{\epsilon}}(x,\epsilon)\supset B^{n}_{f}(x,\hat{\epsilon}).

By definition of Sf,𝒫ϵ(ϵ,n)S_{f,\mathcal{P}_{\epsilon}}(\epsilon,n), for every nn, we may choose FnKϵF_{n}\subset K_{\epsilon}, such that card(Fn)=Sf,𝒫ϵ(ϵ,n)\operatorname{card}(F_{n})=S_{f,\mathcal{P}_{\epsilon}}(\epsilon,n) and

μ(xFnBf,𝒫ϵn(x,ϵ))>1ϵ.\mu\left(\cup_{x\in F_{n}}B_{f,\mathcal{P}_{\epsilon}}^{n}(x,\epsilon)\right)>1-\epsilon.

Using Equation 15, we see that

Sf,𝒫ϵ(ϵ,n)Ssemi,f(ϵ^,n)=min{card(H):μ(xHBfn(x,ϵ^))>1ϵ^},S_{f,\mathcal{P}_{\epsilon}}(\epsilon,n)\leq S_{\operatorname{semi},f}(\hat{\epsilon},n)=\min\{\operatorname{card}(H):\mu\left(\cup_{x\in H}B^{n}_{f}(x,\hat{\epsilon})\right)>1-\hat{\epsilon}\},

and

δf,𝒫,χS(ϵ)δsemi,f,χS(ϵ^).\delta_{f,\mathcal{P},\chi}^{S}(\epsilon)\leq\delta_{\operatorname{semi},f,\chi}^{S}(\hat{\epsilon}).

Since ϵ0\epsilon\to 0 implies that ϵ^0\hat{\epsilon}\to 0, it follows that

(16) limϵ0(supχ{δf,𝒫,χS(ϵ)>0})limϵ^0(supχ{δsemi,f,χS(ϵ^)>0}).\lim_{\epsilon\to 0}\left(\sup_{\chi}\left\{\delta_{f,\mathcal{P},\chi}^{S}(\epsilon)>0\right\}\right)\leq\lim_{\hat{\epsilon}\to 0}\left(\sup_{\chi}\left\{\delta_{\operatorname{semi},f,\chi}^{S}(\hat{\epsilon})>0\right\}\right).

What we have accomplished with the LABEL:{eq:AlmostMeasureSmallerThanSemi} is that hμ,aχ,𝒫δhsemi,μ,aχ,h_{\mu,a_{\chi},\mathcal{P}_{\delta}}\leq h_{\operatorname{semi},\mu,a_{\chi}}, for a generating partition 𝒫δ\mathcal{P}_{\delta} with atoms of diameter at most δ>0.\delta>0. The result follows from [KT97, Proposition 1] where the authors proved that the sup\sup in hμ,aχ=sup𝒫hμ,aχ,𝒫h_{\mu,a_{\chi}}=\sup_{\mathcal{P}}h_{\mu,a_{\chi},\mathcal{P}} can be replaced by a limit limmhμ,aχ,𝒫m\lim_{m\to\infty}h_{\mu,a_{\chi},\mathcal{P}_{m}} over a sequence of generating partitions {𝒫m}\{\mathcal{P}_{m}\}.

Now we prove the inequality hsemi,μ,aχhtop,aχh_{\operatorname{semi},\mu,a_{\chi}}\leq h_{\operatorname{top},a_{\chi}}, this part does not require continuity and follows directly from definitions. Notice that in 2.3, we can drop the sup\sup over compact subsets of XX and substitute K=XK=X. Observe that for all ϵ>0\epsilon>0, and all nn sufficiently large:

Ssemi,f(ϵ,n)Nf,χ(ϵ,n).S_{\operatorname{semi},f}(\epsilon,n)\leq N_{f,\chi}(\epsilon,n).

This completes the proof of the inequalities in Equation 14. ∎

Proof of 4.3.

Since μ\mu is finite, the set XX^{\prime} of atoms of μ\mu is at most countable. For δ>0\delta>0, let 0<δ<δ/20<\delta^{\prime}<\delta/2 be such that if BB is a ball with center in an atom and radius δ\delta^{\prime}, then μ(B)=0\mu(\partial B)=0. Similarly, compact set X\xXB(x,δ)X\backslash\cup_{x\in X^{\prime}}B(x,\delta^{\prime}) has a finite covering of balls of radius at most δ/2\delta/2, and the boundary of these balls has measure zero. Hence, there exists a finite covering of balls B1,,BkB_{1},\dots,B_{k} with radius at most δ/2\delta/2 and μ(Bj)=0\mu(\partial B_{j})=0, for 1jk1\leq j\leq k. The partition 𝒫\mathcal{P} is constructed by recursion: We put P1=B1¯P_{1}=\overline{B_{1}}, and Pj=Bj¯\i=1j1Bi¯P_{j}=\overline{B_{j}}\backslash\cup_{i=1}^{j-1}\overline{B_{i}} for 2jk2\leq j\leq k.

It follows that limϵ0μ(Xϵ)=μ(i=1kPi)=0\lim_{\epsilon\to 0}\mu(X_{\epsilon})=\mu(\cup_{i=1}^{k}\partial P_{i})=0, because the probability measure is outer regular and i=1kPii=1kBi.\cup_{i=1}^{k}\partial P_{i}\subset\cup_{i=1}^{k}\partial B_{i}.

To prove the final observation, let x,yXϵx,y\not\in X_{\epsilon}, i.e. d(x,i=1kPi)>ϵd(x,\cup_{i=1}^{k}\partial P_{i})>\epsilon and d(y,i=1kPi)>ϵ.d(y,\cup_{i=1}^{k}\partial P_{i})>\epsilon. If there are 1i<jk1\leq i<j\leq k, with xPix\in P_{i}, and yPjy\in P_{j}, then d(x,y)>d(x,Bi)+d(y,Bi)2ϵ.d(x,y)>d(x,\partial B_{i})+d(y,\partial B_{i})\geq 2\epsilon.

4.2. Homogeneous Measures and Slow Entropy

The following result is of a classification type. It states that for a class of systems, both the semi-topological and topological entropies are the same. In other words, the inequality on the right in Equation 14 is in fact, an equality.

Definition 4.4.

Let f:XXf\colon X\to X be a measurable transformation. A measure μ\mu is called homogeneous with respect to ff, if for every ϵ>0\epsilon>0 there exists c>0c>0 such that for any x,yXx,y\in X, and every nn0n\geq n_{0}, for some n0n_{0} sufficiently large,

1cμ(Bfn(x,ϵ))μ(Bfn(y,ϵ))<c.\frac{1}{c}\leq\frac{\mu(B_{f}^{n}(x,\epsilon))}{\mu(B_{f}^{n}(y,\epsilon))}<c.
Theorem 4.5.

Suppose that f:XXf\colon X\to X is a measurable transformation, XX is compact, and μ\mu is homogeneous. Then

htop,aχ(f)=hsemi,μ,aχ(f)=limϵ0(sup{χ:Cf,χ(ϵ)>0})h_{\operatorname{top},a_{\chi}}(f)=h_{\operatorname{semi},\mu,a_{\chi}}(f)=\lim_{\epsilon\to 0}\left(\sup\{\chi:C_{f,\chi}(\epsilon)>0\}\right)

where

Cf,χ(ϵ)=lim supn1aχ(n)μ(Bfn(x,ϵ)).C_{f,\chi}(\epsilon)=\limsup_{n\to\infty}\dfrac{1}{a_{\chi}(n)\cdot\mu(B_{f}^{n}(x,\epsilon))}.
Proof.

Let kμ,aχ(f)\displaystyle k_{\mu,a_{\chi}}(f) be the quantity limϵ0(sup{χ:Cf,χ(ϵ)>0}).\displaystyle\lim_{\epsilon\to 0}\left(\sup\{\chi:C_{f,\chi}(\epsilon)>0\}\right).

First we show that htop,aχ(f)=kμ,aχ(f).h_{\operatorname{top},a_{\chi}}(f)=k_{\mu,a_{\chi}}(f). Given a compact set KK, by locally compactness of XX, we can choose a sufficiently small ϵ\epsilon and an open set UU such that B(K,ϵ)U.B(K,\epsilon)\subset U. Denote χ^:=sup{χ:δf,K,χS(ϵ)>0}\hat{\chi}:=\sup\left\{\chi:\delta^{S}_{f,K,\chi}(\epsilon)>0\right\}.

On the one hand, fix a set EE such that for all distinct x,yEx,y\in E we have Bfn(x,ϵ)Bfn(y,ϵ)=,B^{n}_{f}(x,\epsilon)\cap B^{n}_{f}(y,\epsilon)=\varnothing, and thus yEBfn(y,ϵ)U\cup_{y\in E}B^{n}_{f}(y,\epsilon)\subset U is a disjoint union. Since μ\mu is homogeneous, we can choose c>0c>0 such that for any x,yX,x,y\in X,

μ(Bfn(x,ϵ))cμ(Bfn(y,ϵ)).\displaystyle\mu\left(B^{n}_{f}(x,\epsilon)\right)\leq c\mu\left(B^{n}_{f}(y,\epsilon)\right).

Summing over yEy\in E we obtain

μ(Bfn(x,ϵ))card(E)cyEμ(Bfn(y,ϵ))cμ(U)c,\displaystyle\mu\left(B^{n}_{f}(x,\epsilon)\right)\operatorname{card}(E)\leq c\sum_{y\in E}\mu\left(B^{n}_{f}(y,\epsilon)\right)\leq c\mu(U)\leq c,

which implies that for any xXx\in X we have

μ(Bfn(x,ϵ))Sf,K(ϵ,n)c,\displaystyle\mu\left(B^{n}_{f}(x,\epsilon)\right)\cdot S_{f,K}(\epsilon,n)\leq c,

therefore

0<1clim supn1μ(Bfn(x,ϵ))Sf,K(ϵ,n).0<\frac{1}{c}\leq\limsup_{n\to\infty}\dfrac{1}{\mu(B_{f}^{n}(x,\epsilon))\cdot S_{f,K}(\epsilon,n)}.

We will use the following, although we do not prove it. This follows from the definitions of lim\lim and lim sup\limsup

Claim.

For two sequences {an}\{a_{n}\} and {bn}\{b_{n}\}, if limnan\lim_{n\to\infty}a_{n} exists and it is positive and lim supnbn\limsup_{n\to\infty}b_{n} is positive, then

lim supnanbn=limnanlim supnbn.\limsup_{n\to\infty}a_{n}\cdot b_{n}=\lim_{n\to\infty}a_{n}\limsup_{n\to\infty}b_{n}.

For χ<χ^,\chi<\hat{\chi}, take χ<χ<χ^\chi<\chi^{\prime}<\hat{\chi}, using the claim, we have that

Cf,χ(ϵ)\displaystyle C_{f,\chi}(\epsilon) =lim supn1aχ(n)μ(Bfn(x,ϵ))\displaystyle=\limsup_{n\to\infty}\dfrac{1}{a_{\chi}(n)\cdot\mu(B_{f}^{n}(x,\epsilon))}
=lim supnSf,K(ϵ,n)aχ(n)aχ(n)aχ(n)1μ(Bfn(x,ϵ))Sf,K(ϵ,n)\displaystyle=\limsup_{n\to\infty}\dfrac{S_{f,K}(\epsilon,n)}{a_{\chi^{\prime}}(n)}\cdot\dfrac{a_{\chi^{\prime}}(n)}{a_{\chi}(n)}\cdot\dfrac{1}{\mu(B_{f}^{n}(x,\epsilon))\cdot S_{f,K}(\epsilon,n)}
=.\displaystyle=\infty.

Denote kμ,aχ,ϵ(f):=sup{χ:Cf,χ(ϵ)>0}k_{\mu,a_{\chi},\epsilon}(f):=\sup\{\chi:C_{f,\chi}(\epsilon)>0\} and

htop,aχ,K,ϵ(f):=sup{χ:δf,K,χS(ϵ)>0}.h_{\operatorname{top},a_{\chi},K,\epsilon}(f):=\sup\left\{\chi:\delta_{f,K,\chi}^{S}(\epsilon)>0\right\}.

It follows that for any χ<χ^\chi<\hat{\chi} we have

kμ,aχ,ϵ(f)χ.\displaystyle k_{\mu,a_{\chi},\epsilon}(f)\geq\chi.

By taking supremum over χ<χ^\chi<\hat{\chi} on both sides we obtain

kμ,aχ,ϵ(f)χ^=htop,aχ,K,ϵ(f).\displaystyle k_{\mu,a_{\chi},\epsilon}(f)\geq\hat{\chi}=h_{\operatorname{top},a_{\chi},K,\epsilon}(f).

Since both KK and ϵ\epsilon are arbitrary, it is clear that

kμ,aχ(f)htop,aχ(f).\displaystyle k_{\mu,a_{\chi}}(f)\geq h_{\operatorname{top},a_{\chi}}(f).

On the other hand, assume μ(K)>0\mu(K)>0. Since μ\mu is homogeneous, we have

(17) μ(Bfn(x,ϵ))1cμ(Bfn(y,ϵ)).\displaystyle\mu\left(B^{n}_{f}(x,\epsilon)\right)\geq\frac{1}{c}\mu\left(B^{n}_{f}(y,\epsilon)\right).

Fix a finite covering {Bfn(x,ϵ):xF}\left\{B^{n}_{f}(x,\epsilon):x\in F\right\} of KK consisting of (n,ϵ)(n,\epsilon)-Bowen balls so that yFBfn(y,ϵ)K\cup_{y\in F}B^{n}_{f}(y,\epsilon)\supset K. By summing Equation 17 over yFy\in F, we obtain

μ(Bfn(x,ϵ))card(F)1cyFμ(Bfn(y,ϵ))1cμ(K),\displaystyle\mu\left(B^{n}_{f}(x,\epsilon)\right)\operatorname{card}(F)\geq\frac{1}{c}\cdot\sum_{y\in F}\mu\left(B^{n}_{f}(y,\epsilon)\right)\geq\frac{1}{c}\mu(K),

which implies that for any xXx\in X we have

μ(Bfn(x,ϵ))Nf,K(ϵ,n)1cμ(K).\displaystyle\mu\left(B^{n}_{f}(x,\epsilon)\right)N_{f,K}(\epsilon,n)\geq\frac{1}{c}\mu(K).

For χ>χ^\chi>\hat{\chi} we have

Cf,χ(ϵ)\displaystyle C_{f,\chi}(\epsilon) =lim supn1aχ(n)μ(Bfn(x,ϵ))\displaystyle=\limsup_{n\to\infty}\dfrac{1}{a_{\chi}(n)\cdot\mu(B_{f}^{n}(x,\epsilon))}
=lim supnNf,K(ϵ,n)aχ^(n)aχ^(n)aχ(n)1μ(Bfn(x,ϵ))Nf,K(ϵ,n)\displaystyle=\limsup_{n\to\infty}\dfrac{N_{f,K}(\epsilon,n)}{a_{\hat{\chi}(n)}}\cdot\dfrac{a_{\hat{\chi}}(n)}{a_{\chi}(n)}\cdot\dfrac{1}{\mu(B_{f}^{n}(x,\epsilon))N_{f,K}(\epsilon,n)}
=0,\displaystyle=0,

and thus kμ,aχ,ϵ(f)χk_{\mu,a_{\chi},\epsilon}(f)\leq\chi for any χ>χ^.\chi>\hat{\chi}. It follows that kμ,aχ,ϵ(f)χ^=htop,aχ,K,ϵ(f)k_{\mu,a_{\chi},\epsilon}(f)\leq\hat{\chi}=h_{\operatorname{top},a_{\chi},K,\epsilon}(f) and thus kμ,aχ(f)htop,aχ(f).k_{\mu,a_{\chi}}(f)\leq h_{\operatorname{top},a_{\chi}}(f). Now we can conclude that htop,aχ(f)=kμ,aχ(f).h_{\operatorname{top},a_{\chi}}(f)=k_{\mu,a_{\chi}}(f).

Secondly, we show that hsemi,aχ(f)kμ,aχ(f).h_{\operatorname{semi},a_{\chi}}(f)\geq k_{\mu,a_{\chi}}(f). Given ϵ>0,\epsilon>0, define

χ~\displaystyle\tilde{\chi} :=sup{χ:δsemi,f,χS(ϵ)>0},\displaystyle\mathrel{\mathop{:}}=\sup\left\{\chi:\delta_{\operatorname{semi},f,\chi}^{S}(\epsilon)>0\right\},

in which case we have δsemi,f,χ~S(ϵ)=lim supnSsemi,f(ϵ,n)aχ~(n),\displaystyle\delta_{\operatorname{semi},f,\tilde{\chi}}^{S}(\epsilon)=\limsup_{n\to\infty}\frac{S_{\operatorname{semi},f}(\epsilon,n)}{a_{\tilde{\chi}}(n)}, where

Ssemi,f(ϵ,n)=min{card(F):μ(xFBfn(x,ϵ))>1ϵ}.\displaystyle S_{\operatorname{semi},f}(\epsilon,n)=\min\left\{\operatorname{card}(F):\mu\left(\cup_{x\in F}B^{n}_{f}(x,\epsilon)\right)>1-\epsilon\right\}.

Consider a finite measurable set FF such that μ(xFBfn(x,ϵ))>1ϵ.\mu\left(\cup_{x\in F}B^{n}_{f}(x,\epsilon)\right)>1-\epsilon. Since μ\mu is homogeneous, then there exists c>0c>0 such that μ(Bfn(x,ϵ))1cμ(Bfn(y,ϵ)).\mu\left(B^{n}_{f}(x,\epsilon)\right)\geq\frac{1}{c}\mu\left(B^{n}_{f}(y,\epsilon)\right). Summing over yFy\in F we obtain

μ(Bfn(x,ϵ))card(F)yF1cμ(Bfn(y,ϵ))1cμ(yFBfn(y,ϵ))>1ϵc.\displaystyle\mu\left(B^{n}_{f}(x,\epsilon)\right)\operatorname{card}(F)\geq\sum_{y\in F}\frac{1}{c}\mu\left(B^{n}_{f}(y,\epsilon)\right)\geq\frac{1}{c}\mu\left(\cup_{y\in F}B^{n}_{f}(y,\epsilon)\right)>\frac{1-\epsilon}{c}.

Since FF is arbitrary, we have μ(Bfn(x,ϵ))Ssemi,f(ϵ,n)1ϵc.\mu\left(B^{n}_{f}(x,\epsilon)\right)S_{\operatorname{semi},f}(\epsilon,n)\geq\frac{1-\epsilon}{c}. Hence for χ>χ~\chi>\tilde{\chi}

Cf,χ(ϵ)\displaystyle C_{f,\chi}(\epsilon) =lim supn1aχ(n)μ(Bfn(x,ϵ))\displaystyle=\limsup_{n\to\infty}\frac{1}{a_{\chi}(n)\cdot\mu\left(B^{n}_{f}(x,\epsilon)\right)}
=lim supnaχ~(n)aχ(n)Ssemi,f(ϵ,n)aχ~(n)1μ(Bfn(x,ϵ))Ssemi,f(ϵ,n)\displaystyle=\limsup_{n\to\infty}\frac{a_{\tilde{\chi}}(n)}{a_{\chi}(n)}\cdot\frac{S_{\operatorname{semi},f}(\epsilon,n)}{a_{\tilde{\chi}}(n)}\cdot\frac{1}{\mu\left(B^{n}_{f}(x,\epsilon)\right)\cdot S_{\operatorname{semi},f}(\epsilon,n)}
c1ϵlimnaχ~(n)aχ(n)lim supnSsemi,f(ϵ,n)aχ~(n)=0.\displaystyle\leq\frac{c}{1-\epsilon}\lim_{n\to\infty}\frac{a_{\tilde{\chi}}(n)}{a_{\chi}(n)}\limsup_{n\to\infty}\frac{S_{\operatorname{semi},f}(\epsilon,n)}{a_{\tilde{\chi}}(n)}=0.

Letting ϵ0\epsilon\to 0, as in the first part, we obtain that kμ,aχ(f)hsemi,aχ(f).k_{\mu,a_{\chi}}(f)\leq h_{\operatorname{semi},a_{\chi}}(f).

Finally, by 4.1 we have hsemi,aχ(f)htop,aχh_{\operatorname{semi},a_{\chi}}(f)\leq h_{\operatorname{top},a_{\chi}}, we proved that htop,aχ(f)=kμ,aχ(f)h_{\operatorname{top},a_{\chi}}(f)=k_{\mu,a_{\chi}}(f) in the first part. Thus the 4.5 follows. ∎

5. Ferenczi’s Theorem

Here we will present the proof of [Fer97, Proposition 3], which provides a characterization of Kronecker systems via slow entropy. This section is purely expository, we include it to provide a more complete account of the current state of the theory.

Definition 5.1.

A topological dynamical system is called a Kronecker system if it is isomorphic to a group rotation on a compact abelian metrizable group.

Theorem 5.2 ([Fer97]).

Let X¯=(X,μ,f,)\overline{X}=(X,\mu,f,\mathcal{M}) be a probability measure preserving system. Then, X¯\overline{X} is isomorphic to the Kronecker system if and only if hμ,aχ(f)=0h_{\mu,a_{\chi}}(f)=0 for all scales aχ.a_{\chi}.

For the second part of the proof of 5.2, we need the following result.

Lemma 5.3.

If X¯=(X,μ,f,)\overline{X}=(X,\mu,f,\mathcal{M}) is not isomorphic to a Kronecker system, then we can find a partition 𝒫={P1,P2,Pl}\mathcal{P}=\{P_{1},P_{2}...,P_{l}\} and ϵ0>0\epsilon_{0}>0 such that, d(fn(𝒫),𝒫)>ϵ0d(f^{n}(\mathcal{P}),\mathcal{P})>\epsilon_{0} for every nn in a density 1 subset DD\subseteq\mathbb{N}.

For two partitions 𝒫={P1,,Pn}\mathcal{P}=\{P_{1},\dots,P_{n}\} and 𝒬={Q1,,Qn}\mathcal{Q}=\{Q_{1},\dots,Q_{n}\}, we denote the partition distance, between 𝒫\mathcal{P} and 𝒬\mathcal{Q} by

d(𝒫,𝒬):=i=1nμ(PiΔQi).\displaystyle d(\mathcal{P},\mathcal{Q})\mathrel{\mathop{:}}=\displaystyle\sum_{i=1}^{n}\mu(P_{i}\Delta Q_{i}).
Proof of 5.2.

If X¯\overline{X} is isomorphic to the Kronecker system, we can assume ff is an isometry and XX is a compact metric space with metric dd. Let δ>0\delta>0 be a fixed constant, and let 𝒫:=𝒫δ\mathcal{P}\mathrel{\mathop{:}}=\mathcal{P}_{\delta} be a measurable partition of XX described in 4.3. Given ϵ>0\epsilon>0, we want to show Sf,𝒫(ϵ,n)CϵS_{f,\mathcal{P}}(\epsilon,n)\leq C_{\epsilon} when nn is large enough where CϵC_{\epsilon} is a constant independent of nn. Let δ(ϵ)\delta(\epsilon) be a constant depending only on ϵ\epsilon, define Xϵ={xX:d(x,P)<δ(ϵ) for some P𝒫}X_{\epsilon}=\{x\in X:d(x,\partial{P})<\delta(\epsilon)\text{\;for some \;}P\in\mathcal{P}\}. By taking ϵ\epsilon small enough, μ(Xϵ)<ϵ\mu(X_{\epsilon})<\epsilon. Let CϵC_{\epsilon} be the minimal number of balls of radius δ(ϵ)/2\delta(\epsilon)/2 covering XX. Now, for every nn large enough

(18) μ({x:|1ni=0n1𝟙Xϵ(fi(x))|<ϵ})>1ϵ.\mu\left(\left\{x:\left|\frac{1}{n}\sum_{i=0}^{n-1}\mathds{1}_{X_{\epsilon}}(f^{i}(x))\right|<\epsilon\right\}\right)>1-\epsilon.

Therefore, by Equation 18, the set of indices

Exn={0j<n:fj(x)Xϵc}E_{x}^{n}=\left\{0\leq j<n:f^{j}(x)\in X_{\epsilon}^{c}\right\}

satisfies that |Exn|>n(1ϵ)|E_{x}^{n}|>n(1-\epsilon) for xx in a set with measure at least 1ϵ1-\epsilon. For such xx, suppose yy is in the same ball of radius δ(ϵ)/2\delta(\epsilon)/2 containing xx. Then, when jExnj\in E_{x}^{n}, fjxf^{j}x and fjyf^{j}y are in the same atom of partition 𝒫\mathcal{P} because ff is an isometry. Therefore, Sf,𝒫(ϵ,n)CϵS_{f,\mathcal{P}}(\epsilon,n)\leq C_{\epsilon} because each Hamming ball contains a Bowen ball of radius δ(ϵ)/2\delta(\epsilon)/2. Since the Bowen balls are exactly the balls in the metric dd, Sf,𝒫(ϵ,n)S_{f,\mathcal{P}}(\epsilon,n) is bounded by a constant depending only on 𝒫\mathcal{P} and ϵ\epsilon. Consequently, any Kronecker system has zero entropy at all scales.

To prove the other direction, we need 5.3. Assume hμ,aχ(f)=0h_{\mu,a_{\chi}}(f)=0 for all scales aχa_{\chi}, we claim that

lim infnSf,𝒫(ϵ,n)Cϵ\displaystyle\liminf_{n\rightarrow\infty}S_{f,\mathcal{P}}(\epsilon,n)\leq C_{\epsilon}

for any fixed measurable partition 𝒫\mathcal{P} and given ϵ\epsilon. Suppose not, then there exists a measurable partition 𝒫0\mathcal{P}_{0} and ϵ0\epsilon_{0} s.t.s.t. for any kk there exists nkn_{k} satisfying Sf,𝒫0(ϵ,n)>kS_{f,\mathcal{P}_{0}}(\epsilon,n)>k for every ϵ<ϵ0\epsilon<\epsilon_{0} and nk<nn_{k}<n. Choose aχ(n)a_{\chi}(n) to be some value between kk and k+1k+1 when nkn<nk+1n_{k}\leq n<n_{k+1}. This implies hμ,aχ,𝒫0(f)>0h_{\mu,a_{\chi},\mathcal{P}_{0}}(f)>0, which is a contradiction.

If the system X¯\overline{X} is not isomorphic to a Kronecker system, by applying 5.3 we can find a measurable partition 𝒫={P1,,Pl}\mathcal{P}=\{P_{1},...,P_{l}\}. Fixed an arbitrarily large NN, there exists a constant CC s.t.s.t. Sf,𝒫(ϵ2,N)CS_{f,\mathcal{P}}(\epsilon^{2},N)\leq C. By using MCM\leq C Hamming balls Bi=Bf,𝒫N(xi,ϵ2)B_{i}=B_{f,\mathcal{P}}^{N}(x^{i},\epsilon^{2}) where 1iM1\leq i\leq M, one can cover at least 1ϵ21-\epsilon^{2} space of XX. Consider space (X×{0,1,,N1},μ×ν)(X\times\{0,1,\dots,N-1\},\mu\times\nu), where ν(E)=1N×cardinality of E\nu(E)=\frac{1}{N}\times\text{cardinality of }E. Define a measurable function on X×{0,1,,N1}X\times\{0,1,\dots,N-1\} by

g(x,n)={1if there exists i s.t. xBi and fn(x) and fn(xi) are in the same atom0othersg(x,n)=\left\{\begin{aligned} 1&\;\text{if there exists $i$ $s.t.$ $x\in B_{i}$ and $f^{n}(x)$ and $f^{n}(x^{i})$ are }\\ &\;\text{in the same atom}\\ 0&\;\text{others}\\ \end{aligned}\right.

By definition of Hamming balls,

g𝑑μ×ν>(1ϵ2)(1ϵ2).\int g\,d\mu\times\nu>(1-\epsilon^{2})(1-\epsilon^{2}).

Now, there exists a subset EE of {0,1,,N1}\{0,1,\dots,N-1\} with cardinality larger than N(14ϵ)N(1-4\epsilon) satisfying for every nEn\in E, g𝑑μ>1ϵ/2\int g\,d\mu>1-\epsilon/2. This is because we can get g𝑑ν<12ϵ2\int gd\nu<1-2\epsilon^{2} otherwise.

We claim that we can find a positive density subset Λ\Lambda of \mathbb{N} s.t.s.t. for every nΛn\in\Lambda, d(fn(𝒫),𝒫)<ϵd(f^{n}(\mathcal{P}),\mathcal{P})<\epsilon, which is a contradiction to 5.3. Note that

d(fn(𝒫),fj(𝒫))=Xmin{|xnxj|,1}𝑑μ(x)d(f^{n}(\mathcal{P}),f^{j}(\mathcal{P}))=\int_{X}\min\{|x_{n}-x_{j}|,1\}d\mu(x)

for every nn and jj. Moreover, for each nn, (xn1,,xnM)(x_{n}^{1},\dots,x_{n}^{M}) has at most lMl^{M} choices. Therefore, for every nEn\in E, there exists an integer m(n)m(n), s.t.s.t. xni=xm(n)ix_{n}^{i}=x_{m(n)}^{i} for every 1iM1\leq i\leq M. Hence, d(fn(𝒫),fm(n)(𝒫))<ϵd(f^{n}(\mathcal{P}),f^{m(n)}(\mathcal{P}))<\epsilon because the set satisfying there exists ii s.t.s.t. xBix\in B_{i} and xni=xnx_{n}^{i}=x_{n} has measure larger than 1ϵ1-\epsilon for every nEn\in E. From the pigeonhole principle, there exists a subset ΛN\Lambda_{N} of {0,1,,N1}\{0,1,\dots,N-1\} with cardinality at least N(14ϵ)/lMN(1-4\epsilon)/l^{M} such that for every nΛNn\in\Lambda_{N}, d(fn(𝒫),fm(𝒫))<ϵd(f^{n}(\mathcal{P}),f^{m}(\mathcal{P}))<\epsilon for some m{0,1,,N1}m\in\{0,1,\dots,N-1\}. Since the metric is ff-invariant, we can conclude the result. ∎

Definition 5.4.

If X¯=(X,μ,f,)\overline{X}=(X,\mu,f,\mathcal{M}) is a measure preserving system, we say a function gL2(X,μ)g\in L^{2}(X,\mu) is almost periodic if {Ufng:n}\{U_{f}^{n}g:n\in\mathbb{Z}\} is a precompact set in L2(X,μ)L^{2}(X,\mu).

Proof of 5.3.

We give a sketch of proof here. If the system is isomorphic to a Kronecker system, then gg is almost periodic for any gL2(X)g\in L^{2}(X). Suppose X¯\overline{X} is not isomorphic to the Kronecker system, then there exists a non-zero measurable function gg with mean 0 such that gg is orthogonal to every almost periodic function. Then, there is a density 1 subset DD of \mathbb{N} s.t.s.t.

limn,nDUfn(g),g=0.\displaystyle\lim_{n\rightarrow\infty,n\in D}\langle U_{f}^{n}(g),g\rangle=0.

Assume w.l.o.g. that g=1|\!|g|\!|=1, given ϵ=1100,\epsilon=\frac{1}{100}, there is a simple function h=i=1lci𝟙Aih=\sum\limits_{i=1}^{l}c_{i}\mathds{1}_{A_{i}} such that ghϵ/4.|\!|g-h|\!|\leq\epsilon/4. Therefore, when nn is large enough,

|Ufn(h),h|ϵ.|\langle U_{f}^{n}(h),h\rangle|\leq\epsilon.

Choose partition 𝒫={A1,,Al}\mathcal{P}=\{A_{1},...,A_{l}\}, then there exists ϵ0<min{μ(A1),,μ(Al)}100\epsilon_{0}<\frac{\min\{\mu(A_{1}),...,\mu(A_{l})\}}{100}, such that d(fn(𝒫),𝒫)>ϵ0d(f^{n}(\mathcal{P}),\mathcal{P})>\epsilon_{0} for nD.n\in D.

6. Differences between Slow and Exponential Entropy

6.1. Sturmian subshifts

For more information about Sturmian systems, we refer the reader to [Fog02, Chapter 6].

Definition 6.1.

A sequence u{0,1}u\in\{0,1\}^{\mathbb{N}} ({0,1})(\in\{0,1\}^{\mathbb{Z}}) is a (bi-infinite) Sturmian sequence if for every n1n\geq 1, the number of words of length nn that appear in uu is equal to

pn(u)=n+1,p_{n}(u)=n+1,

and it is not eventually periodic.

Consider f:{0,1}{0,1}f\colon\{0,1\}^{\mathbb{Z}}\to\{0,1\}^{\mathbb{Z}}, the shift map defined as w=f(v)w=f(v), wi=vi+1w_{i}=v_{i+1}.

Definition 6.2.

A Sturmian system (Xu,f)(X_{u},f) for a bi-infinite Sturmian sequence uu is the shift map ff on Xu={fk(u):k}¯X_{u}=\overline{\{f^{k}(u):k\in\mathbb{Z}\}}.

Theorem 6.3.

Any Sturmian system (Xu,f)(X_{u},f) is measurably equivalent to an irrational rotation (/,Rθ)(\mathbb{R}/\mathbb{Z},R_{\theta}).

Sketch of the proof with ideas in [Fog02, Chapter 6] .

Any initial point β/\beta\in\mathbb{R}/\mathbb{Z}, generates a Sturmian sequence v=(vi)v=(v_{i}), by taking vi=jv_{i}=j if Rθi(β)IjR_{\theta}^{i}(\beta)\in I_{j}, I0=[0,1θ)I_{0}=[0,1-\theta) and I1=[1θ,1)I_{1}=[1-\theta,1).

The converse is very elaborate see [Fog02, Sections 6.3 and 6.4]. The idea is that every vXuv\in X_{u} has a coding (an,bn)(a_{n},b_{n}). The coefficients (an)(a_{n}) are the partial quotients of the continued fraction expansion of θ\theta. The initial point β\beta is determined by bnb_{n}; these are the coefficients of its Ostrowski expansion. Things are very subtle; for instance, see [Fog02, Exercise 6.2.13 item 5]. ∎

The 6.3 combined with 5.2 implies that Sturmian systems have 0 metric entropy at all scales. The following immediately implies 1.2:

Theorem 6.4.

For any Sturmian system (Xu,f)(X_{u},f), we have that

0=hμ,aχ<hsemi,μ,aχ=htop,aχ=1,0=h_{\mu,a_{\chi}}<h_{\operatorname{semi},\mu,a_{\chi}}=h_{\operatorname{top},a_{\chi}}=1,

for the polynomial scale aχ(N)=Nχa_{\chi}(N)=N^{\chi}, and the unique ff-invariant measure μ\mu.

Proof.

Proof of htop,aχ=1h_{\operatorname{top},a_{\chi}}=1.

This follows from 3.4 and 3.5. For Sturmian shifts, we have that for a fixed integer k>0k>0,

Nf,Xu(2(k1),n)=p2k+n+1(Xu)=2k+n+2.N_{f,X_{u}}(2^{-(k-1)},n)=p_{2k+n+1}(X_{u})=2k+n+2.

Hence,

δf,Xu,χ(2(k1))=lim supnNf,Xu(2(k1),n)nχ=limn2k+n+2nχ={0,χ>1,1,χ=1,,χ<1.\begin{split}\delta_{f,X_{u},\chi}(2^{-(k-1)})&=\limsup_{n\to\infty}\frac{N_{f,X_{u}}(2^{-(k-1)},n)}{n^{\chi}}\\ &=\lim_{n\to\infty}\frac{2k+n+2}{n^{\chi}}=\left\{\begin{array}[]{cc}0,&\chi>1,\\ 1,&\chi=1,\\ \infty,&\chi<1.\end{array}\right.\end{split}

For the linear scale aχ(N)=Nχa_{\chi}(N)=N^{\chi},

htop,aχ=limϵ(sup{χ:δf,Xu,χ(ϵ)>0})=1.h_{\operatorname{top},a_{\chi}}=\lim_{\epsilon}(\sup\{\chi:\delta_{f,X_{u},\chi}(\epsilon)>0\})=1.

Proof of hsemi,μ,aχ=1h_{\operatorname{semi},\mu,a_{\chi}}=1.

Let μ\mu be the invariant probability measure for (Xu,f)(X_{u},f). If the sequence uu was coded by the partition 𝒫={P0=[0,1θ),P1=[1θ,1)}\mathcal{P}=\{P_{0}=[0,1-\theta),P_{1}=[1-\theta,1)\}, then for every cylinder set ωXu\omega\subset X_{u} defined by a word b0b1bn1Lnb_{0}b_{1}\dots b_{n-1}\in L_{n}, we must have that μ(ω)=Leb(P)\mu(\omega)=\operatorname{Leb}(P) for some P𝒫n:=i=0n1Rθi𝒫.P\in\mathcal{P}^{n}\mathrel{\mathop{:}}=\vee_{i=0}^{n-1}R^{-i}_{\theta}\mathcal{P}. The atoms in the partition 𝒫n\mathcal{P}^{n} have endpoints in

{0,R1θ(0),,R1θn(0)}.\{0,R_{1-\theta}(0),\dots,R^{n}_{1-\theta}(0)\}.

Let 1θ=[0;a1,a2,a3,]1-\theta=[0;a_{1},a_{2},a_{3},\dots] be the continued fraction expansion, and {pkqk}\{\frac{p_{k}}{q_{k}}\} the sequence of best approximants. Writing n=mqk+qk1+rn=mq_{k}+q_{k-1}+r with 1mak+11\leq m\leq a_{k+1}, and 0r<qk0\leq r<q_{k}. By the Three Gap Theorem, see [AB98, Section 3], the measure of the cylinder ω\omega is

  1. (1)

    ηk:=(1)k(qk(1θ)pk)\eta_{k}\mathrel{\mathop{:}}=(-1)^{k}(q_{k}(1-\theta)-p_{k}), in which case there are n+1qkn+1-q_{k} cylinders of this measure. These are the smallest gaps.

  2. (2)

    ηk1mηk\eta_{k-1}-m\eta_{k}. There are r+1r+1 cylinders of this measure.

  3. (3)

    ηk1(m1)ηk\eta_{k-1}-(m-1)\eta_{k}. There are qk(r+1)q_{k}-(r+1) cylinders of this measure. These are the biggest gaps, and their length is the sum of the lengths of the previous types.

When n=(m+1)qk+qk11n=(m+1)q_{k}+q_{k-1}-1, for 1mak+11\leq m\leq a_{k+1} (r=qk1r=q_{k}-1), then there are no cylinders with the length in item 3.

If

C(n):=min{card(F):ωFμ(ω)>1ϵ}C(n)\mathrel{\mathop{:}}=\min\left\{\operatorname{card}(F):\sum_{\omega\in F}\mu(\omega)>1-\epsilon\right\}

where FF is a family of cylinders sets of size nn, then for some 0l,s,κ10\leq l,s,\kappa\leq 1 depending on nn,

(19) C(n)=n+1[κ(n+1qk)+s(r+1)+l(qk(r+1))]C(n)=n+1-\left[\kappa(n+1-q_{k})+s(r+1)+l(q_{k}-(r+1))\right]

where

(20) κ(n+1qk)ηk+s(r+1)(ηk1mηk)+l(qk(r+1))(ηk1(m1)ηk)<ϵ.\kappa(n+1-q_{k})\eta_{k}+s(r+1)(\eta_{k-1}-m\eta_{k})+l(q_{k}-(r+1))(\eta_{k-1}-(m-1)\eta_{k})<\epsilon.

Specializing to the case where there are no cylinders of type item 3

n=nk=(ak+1+1)qk+qk11=qk+1+qk1,n=n_{k}=(a_{k+1}+1)q_{k}+q_{k-1}-1=q_{k+1}+q_{k}-1,

we can substitute l=0l=0 and r=qk1r=q_{k}-1 into Equation 20 to obtain:

(21) κ>s(r+1)(ηk1ak+1ηk)ϵ(n+1qk)ηk=sqkηk+1ϵqk+1.\begin{split}-\kappa&>\frac{s(r+1)(\eta_{k-1}-a_{k+1}\eta_{k})-\epsilon}{(n+1-q_{k})\eta_{k}}\\ &=\frac{sq_{k}\eta_{k+1}-\epsilon}{q_{k+1}}.\end{split}

In this special case, combining Equation 19 and Equation 21, we obtain that

(22) C(nk)nk=1+1nkκqk+1nksqknk>1+1nk+(sqkηk+1ϵ)qk+1ηkqk+1nksqknk=1+1nk+sqknk(ηk+1ηk1)ϵnkηk.\begin{split}\frac{C(n_{k})}{n_{k}}&=1+\frac{1}{n_{k}}-\kappa\frac{q_{k+1}}{n_{k}}-s\frac{q_{k}}{n_{k}}\\ &>1+\frac{1}{n_{k}}+\frac{(sq_{k}\eta_{k+1}-\epsilon)}{q_{k+1}\eta_{k}}\frac{q_{k+1}}{n_{k}}-s\frac{q_{k}}{n_{k}}\\ &=1+\frac{1}{n_{k}}+s\frac{q_{k}}{n_{k}}\left(\frac{\eta_{k+1}}{\eta_{k}}-1\right)-\frac{\epsilon}{n_{k}\eta_{k}}.\end{split}

Using Khinchin’s inequality, see for instance [Khi97, Theorems 9 and 13]:

1qk+1+qk<ηk<1qk+1\frac{1}{q_{k+1}+q_{k}}<\eta_{k}<\frac{1}{q_{k+1}}

and that nk=qk+1+qk1n_{k}=q_{k+1}+q_{k}-1, we obtain

(23) 1=1limk1qk+1+qklim supknkηk1+lim supkqkqk+1<.1=1-\lim_{k\to\infty}\frac{1}{q_{k+1}+q_{k}}\leq\limsup_{k\to\infty}n_{k}\eta_{k}\leq 1+\limsup_{k\to\infty}\frac{q_{k}}{q_{k+1}}<\infty.

By Equation 22 and Equation 23, we have that

δsemi,f,1S(ϵ)=lim supnC(n)nlim supkC(nk)nk1o(ϵ).\delta_{\operatorname{semi},f,1}^{S}(\epsilon)=\limsup_{n\to\infty}\frac{C(n)}{n}\geq\limsup_{k\to\infty}\frac{C(n_{k})}{n_{k}}\geq 1-o(\epsilon).

We conclude that hsemi,μ,aχ=1h_{\operatorname{semi},\mu,a_{\chi}}=1 by Equation 14 in 4.1

6.2. Large gaps

We thank Scott Schmieding for pointing out the constructions in this section. Fix the polynomial scale pχ(N)=Nχp_{\chi}(N)=N^{\chi}. The following shows that we can achieve gaps of polynomial size for the variational principle:

Theorem 6.5.

For every mm\in\mathbb{N}, there exists a uniquely ergodic homeomorphism of a compact metric space f:XXf:X\to X preserving μ\mu such that hμ,pχ(f)=0h_{\mu,p_{\chi}}(f)=0, but htop,pχ(f)=mh_{\operatorname{top},p_{\chi}}(f)=m.

Proof.

Consider finitely many irrational numbers β1,,βm(0,1)\beta_{1},\dots,\beta_{m}\in(0,1) which are rationally independent. Then the translation on 𝕋m\mathbb{T}^{m} by the vector v=(β1,,βm)v=(\beta_{1},\dots,\beta_{m}) is uniquely ergodic, and is isomorphic to the product of the rotations Rβ1××RβmR_{\beta_{1}}\times\dots\times R_{\beta_{m}}.

For each i=1,,mi=1,\dots,m, consider the Sturmian subshift σi:XiXi\sigma_{i}:X_{i}\to X_{i} isomorphic to the rotation RβiR_{\beta_{i}}, and let X=X1××XmX=X_{1}\times\dots\times X_{m} and f:XXf:X\to X be the product f=σ1××σmf=\sigma_{1}\times\dots\times\sigma_{m}. Note that an element xXx\in X is an mm-tuple of infinite words in the symbols 0 and 11. Equivalently, one may consider it as a single infinite word whose entries are mm-tuples of 0’s and 11’s. Therefore, ff may be considered a subshift of a shift on 2m2^{m} symbols. Since a word is admissible if and only if each of its components are admissible, and each component may be chosen independently from each corresponding Sturmian language, there are (n+1)m(n+1)^{m} words of length nn in the language of ff. Thus, the topological slow entropy at polynomial scale is mm by 3.5.

On the other hand, we claim that ff is uniquely ergodic (in which case ff is measurably isomorphic to Rβ1××RβmR_{\beta_{1}}\times\dots\times R_{\beta_{m}}). Indeed, given an ff-invariant measure μ\mu, it must project to a σi\sigma_{i}-invariant measure on each XiX_{i}. Since Sturmian subshifts are uniquely ergodic, it follows that μ\mu is a joining of the circle rotations RβiR_{\beta_{i}}. Hence μ\mu must correspond to the Haar measure. ∎

Fix the scale aχ(N)=eNχa_{\chi}(N)=e^{N^{\chi}}, which we call the stretched exponential scales. Note that aχa_{\chi} is faster than polynomial scales for all χ>0\chi>0, but for χ<1\chi<1, the rate aχa_{\chi} is slower than exponential.

Theorem 6.6.

There exists a uniquely ergodic subshift σ:XX\sigma:X\to X preserving a measure μ\mu such that hμ,bχ(σ)=0h_{\mu,b_{\chi}}(\sigma)=0 for every family of scales bχb_{\chi}, but htop,aχ(σ)=1h_{\operatorname{top},a_{\chi}}(\sigma)=1.

6.6 shows that there are systems that have very large gap between the metric and topological slow entropies, achieving stretched exponential rates arbitrarily close to 1.

Remark 6.7.

6.6 heavily relies on our use of lim sup\limsup rather than lim inf\liminf when defining our slow entropies. When using the lim inf\liminf definition, the topological slow entropy grows linearly. That is, the growth rate of Nσ,X(ϵ,n)N_{\sigma,X}(\epsilon,n) (which is linked to the language complexity by 3.4) oscillates between linear and stretched exponential rates.

Proof of 6.6.

In [PS23b, Theorem 5.15, Proposition 5.23], it is shown that among transitive subshifts σ:XX\sigma:X\to X, the following properties (among others) are generic:

  • σ\sigma is a regular Toeplitz subshift

  • For every γ(0,1)\gamma\in(0,1), pn:=pn(X)p_{n}:=p_{n}(X) has subsequences satisfying

    limnloglogpnklognk=γ\lim_{n\to\infty}\dfrac{\log\log p_{n_{k}}}{\log n_{k}}=\gamma

Since such a σ\sigma is a regular Toeplitz subshift, it is uniquely ergodic and measurably isomorphic to translation on a compact abelian group [JK69]. It follows that the metric slow entropy is 0 at all scales. On the other hand, if 0<γ<γ<γ′′0<\gamma<\gamma^{\prime}<\gamma^{\prime\prime}, identify a subsequence such that limkloglogpnklognk=γ′′\lim_{k\to\infty}\dfrac{\log\log p_{n_{k}}}{\log n_{k}}=\gamma^{\prime\prime}. Then for sufficiently large kk,

loglogpnklognk\displaystyle\dfrac{\log\log p_{n_{k}}}{\log n_{k}} >\displaystyle> γ\displaystyle\gamma^{\prime}
loglogpnk\displaystyle\log\log p_{n_{k}} >\displaystyle> γlognk\displaystyle\gamma^{\prime}\log n_{k}
logpnk\displaystyle\log p_{n_{k}} >\displaystyle> nkγ\displaystyle n_{k}^{\gamma^{\prime}}
pnk\displaystyle p_{n_{k}} >\displaystyle> enkγ.\displaystyle e^{n_{k}^{\gamma^{\prime}}}.

Since γ\gamma was aribtrary, it follows that

lim supnpnenγ=\limsup_{n\to\infty}\dfrac{p_{n}}{e^{n^{\gamma}}}=\infty

whenever γ(0,1)\gamma\in(0,1). On the other hand, when γ=1\gamma=1, the lim sup\limsup must be 0 since the system has 0 topological entropy (since the variational principle holds at exponential scale, and the system is uniquely ergodic with 0 exponential metric entropy). It follows that the topological slow entropy at stretched exponential scale is 1 by 3.5. ∎

6.3. Denjoy circle transformations

The Sturmian systems considered above can be realized as invariant sets for transformations of the circle. Indeed, one may build C1,αC^{1,\alpha} circle diffeomorphisms by starting with an irrational circle rotation and “blowing up” an orbit by inserting an interval at each point of the orbit. Such examples were first studied by Denjoy and their construction can be found in [KH95, Section 12.2]. We characterize them here:

Definition 6.8.

We say that a circle homeomorphism f:S1S1f:S^{1}\to S^{1} is Denjoy if the rotation number θ\theta of ff is irrational, and there is a semiconjugacy h:S1S1h:S^{1}\to S^{1} and a point x0S1x_{0}\in S^{1} such that

  • hf=Rθhh\circ f=R_{\theta}\circ h,

  • h1(fn(x0))h^{-1}(f^{n}(x_{0})) is a nontrivial closed interval for all nn\in\mathbb{Z},

  • h1(x)h^{-1}(x) is a single point for all xx outside of the orbit of x0x_{0}.

Lemma 6.9.

If ff is a Denjoy circle transformation, then f|NW(f)f|_{NW(f)} is topologically conjugated to a Sturmian subshift.

Sketch of proof.

Recall that Sturmian sequences can be obtained by looking at codes appearing of the rotation RθR_{\theta} using the intervals [0,1θ)[0,1-\theta) and [1θ,1)[1-\theta,1). In the case of a circle rotation the map which sends the code to the point is not one-to-one. However, in the case of a Denjoy transformation, the coding intervals can be taken to cover only the nonwandering set, and are therefore disjoint. This yields a conjugacy instead of a semiconjugacy. ∎

Corollary 6.10.

Denjoy circle transformations are not variational.

Proof.

By Poincaré recurrence, any invariant measure for a Denjoy transformation must be supported on its nonwandering set. Since restricted to this set, the system is topologically conjugated to a Sturmian shift, we conclude that it is uniquely ergodic and that the unique invariant measure is Kronecker. Thus, the metric entropy has 0 entropy at all scales. However, since there is a compact invariant set topologically conjugated to a Sturmian subshift, the semi-topological and topological entropies are both linear. ∎

6.4. Geodesic flow on 𝕋2\mathbb{T}^{2}

Another unexpected feature of slow entropy is the failure of additivity over ergodic decompositions. Let μ=ν𝑑μ^(ν)\mu=\int_{\mathcal{E}}\nu d\hat{\mu}(\nu) be the ergodic decomposition of μ\mu, where \mathcal{E} is the space of ergodic invariant measures and μ^\hat{\mu} is a probability measure on \mathcal{E}. For the classical entropy at exponential scale [VO16, Theorem 9.6.2], we have that

(24) hμ(f)=hν(f)𝑑μ^(ν).h_{\mu}(f)=\int_{\mathcal{E}}h_{\nu}(f)d\hat{\mu}(\nu).

In this section, we explain that such a formula cannot hold for slow entropy, even when restricting so a fixed scale such as the polynomial scale.

It is well-known that the geodesic flow on T1𝕋2T^{1}\mathbb{T}^{2}, the unit tangent bundle to 𝕋2\mathbb{T}^{2}, is not ergodic and has a smooth ergodic decomposition. Each ergodic component is diffeomorphic to 𝕋2\mathbb{T}^{2} and corresponds to the unit speed linear flow in an irrational direction (the rational directions have measure 0, so we may omit them from the ergodic decomposition). Hence, at any scale, the ergodic components of the Haar measure on T1𝕋2T^{1}\mathbb{T}^{2} all have 0 entropy at all scales. The following Lemma shows that we can obtain a positive slow entropy by “gluing” several copies of Kronecker systems together in an interesting way.

Lemma 6.11.

If φt:T1𝕋2T1𝕋2\varphi_{t}:T^{1}\mathbb{T}^{2}\to T^{1}\mathbb{T}^{2} is the geodesic flow on 𝕋2\mathbb{T}^{2}, then the topological and Haar slow entropy of φt\varphi_{t} is 1 at polynomial scales nχ(t)=tχn_{\chi}(t)=t^{\chi}.

Proof.

Observe that Isom(2)S12\operatorname{Isom}(\mathbb{R}^{2})\cong S^{1}\ltimes\mathbb{R}^{2} acts simply transitively on T12T^{1}\mathbb{R}^{2}, and that T1𝕋2T^{1}\mathbb{T}^{2} is the quotient of T12T^{1}\mathbb{R}^{2} by 2\mathbb{Z}^{2}. Furthermore, since the isometry group takes orbits of the geodesic flow to orbits of the geodesic flow, it follows that the geodesic flow is smoothly conjugated to homogeneous flow on Isom(2)/2\operatorname{Isom}(\mathbb{R}^{2})/\mathbb{Z}^{2} by a one-parameter subgroup of 2𝕋12\mathbb{R}^{2}\subset\mathbb{T}^{1}\mathbb{R}^{2}. If ΘLie(Isom(2))\Theta\in\operatorname{Lie}(\operatorname{Isom}(\mathbb{R}^{2})) represents the generator of the subgroup S1S^{1}, and XX and YY represent orthonormal generators of 2\mathbb{R}^{2}, then (up to choice of orientation), we have structure relations

[Θ,X]=Y[Θ,Y]=X[X,Y]=0.[\Theta,X]=Y\qquad[\Theta,Y]=-X\qquad[X,Y]=0.

From this, one easily checks that in this basis ad(X)\operatorname{ad}(X) is

(010000000),\begin{pmatrix}0&1&0\\ 0&0&0\\ 0&0&0\end{pmatrix},

so by [KVW19], it follows that the polynomial slow entropy of φt\varphi_{t} is 1. ∎

We remark that this geodesic flow is also conjugated to the suspension of the affine map (x,y)(x,y+x)(x,y)\mapsto(x,y+x), so we have the phenomenon for transformations as well.

7. The slow entropy of some interval exchanges

Interval exchange transformations or IETs are piecewise isometries, with a finite number of discontinuities. Moreover, IETs preserve the orientation. These maps can be regarded as generalizations of rotations. In this section, we will compute the metric slow entropy of 3-IETs. The computations for the metric entropy will occupy most of Section 7. We will prove that for a large class of 3-IETs the metric slow entropy is 1 for the polynomial scale aχ(n)=nχa_{\chi}(n)=n^{\chi}, see 7.1.

Theorem 7.1.

Let Δ3\Delta\subset\mathbb{R}^{3} be the set {(x,y,z):x+y+z=1,x,y,z>0}\{(x,y,z):x+y+z=1,\,x,y,z>0\}. There exists a set AΔA\subset\Delta of Hausdorff dimension 2 such that if gg is a 3-IET determined by λA\lambda\in A, then hLeb,aχ(g)=1.h_{\operatorname{Leb},a_{\chi}}(g)=1.

We will easily see that the topological slow entropy of the corresponding symbolic system is at most 1 with respect to the same scale, see 7.5. This combined with 7.1 prove 1.3.

Remark 7.2.

We have used the convention that the limits appearing in the definitions of the functions δf,,χS\delta_{f,\cdot,\chi}^{S} slow entropy are lim sup\limsup’s. In general these are not actual limits, and we rely on this choice several times in the proof. It would be interesting to make similar computations for the lim inf\liminf definitions. It is already known that a gap may exist, and special attention is paid to this subtlety in [BKW23b].

7.1. Preliminaries of IET’s

We refer the reader to [Yoc10, Via06] for more details about interval exchange transformations (IET).

Let 𝒜\mathcal{A} be a collection of dd symbols and λ>0𝒜\lambda\in\mathbb{R}^{\mathcal{A}}_{>0} be a vector of positive entries. Given two bijective functions πt:𝒜{1,,d}\pi_{t}:\mathcal{A}\to\{1,\dots,d\}, πb:𝒜{1,,d}\pi_{b}:\mathcal{A}\to\{1,\dots,d\}, we obtain a permutation of the symbols in 𝒜\mathcal{A} defined by

π=(πt1(1)πt1(d)πb1(1)πb1(d)).\pi=\begin{pmatrix}\pi_{t}^{-1}(1)&\dots&\pi_{t}^{-1}(d)\\ \pi_{b}^{-1}(1)&\dots&\pi_{b}^{-1}(d)\\ \end{pmatrix}.

Let II\subset\mathbb{R} be a bounded interval, closed on the left and open on the right, and denote the length of II by |I||I|. From now on, we will assume that the left endpoint of II is 0. The vector λ\lambda and the permutation π\pi determine a partition {Ia}a𝒜\{I_{a}\}_{a\in\mathcal{A}} where

Ia=[{b:πt1(b)<πt1(a)}λb,{b:πt1(b)πt1(a)}λb),I_{a}=\left[\sum_{\{b:\pi_{t}^{-1}(b)<\pi_{t}^{-1}(a)\}}\lambda_{b},\sum_{\{b:\pi_{t}^{-1}(b)\leq\pi_{t}^{-1}(a)\}}\lambda_{b}\right),

and the interval I=[0,|I|)I=\left[0,|I|\right), where |I|=A𝒜λa.|I|=\sum_{A\in\mathcal{A}}\lambda_{a}.

Definition 7.3.

An interval exchange transformation on dd intervals (dd-IET) g:=gλ,π:IIg\mathrel{\mathop{:}}=g_{\lambda,\pi}\colon I\to I determined by a length vector λ\lambda and permutation π\pi is the bijective map defined by

(25) g(x)=x+{b:πb1(b)<πb1(a)}λb{b:πt1(b)<πt1(a)}λb,if xIa.g(x)=x+\sum_{\{b:\pi_{b}^{-1}(b)<\pi^{-1}_{b}(a)\}}\lambda_{b}-\sum_{\{b:\pi_{t}^{-1}(b)<\pi_{t}^{-1}(a)\}}\lambda_{b},\quad\text{if }x\in I_{a}.

Any IET is a measure preserving transformation with respect to the Lebesgue measure Leb\operatorname{Leb} on the interval I.I. If for some k<dk<d, the set {1,,k}\{1,\dots,k\} is π\pi invariant, the IET is a concatenation of IETs with fewer intervals. If {1,,k}\{1,\dots,k\} is not invariant for every k<dk<d, we say that the permutation (and the IET) is irreducible.

7.2. Topological and semi-topological slow entropy of a class of dd-IETs

Let g=gλ,πg=g_{\lambda,\pi} be a dd-IET, and denote D={β1,,βd1}D=\{\beta_{1},\dots,\beta_{d-1}\} the set of discontinuities of gg. Here, βi=j=1iλj\beta_{i}=\sum_{j=1}^{i}\lambda_{j}. For convenience, denote β0=0\beta_{0}=0 and βd=|I|\beta_{d}=|I|. The map gg has the idoc property (infinite distinct orbit condition) if for all n>1n>1, Dgn(D)=.D\cap g^{-n}(D)=\varnothing. Keane in [Kea75] proved that an idoc IET is minimal.

From now on, we will assume that π\pi is an irreducible permutation. The set I\DI\backslash D has the dd continuity intervals of g.g. By including the left endpoint, let 𝒫:={Ia}a𝒜\mathcal{P}\mathrel{\mathop{:}}=\{I_{a}\}_{a\in\mathcal{A}} be the natural partition of gg.

We have the following results regarding the number of Bowen balls.

Lemma 7.4.

Let ϵ>0\epsilon>0 be sufficiently small. Suppose that gg has the idoc property. There exists n0n_{0} such that Bgn+1(x,ϵ)𝒫n(x)B_{g}^{n+1}(x,\epsilon)\subset\mathcal{P}^{n}(x) and 𝒫n(x)Bgn(x,ϵ)\mathcal{P}^{n}(x)\subset B_{g}^{n}(x,\epsilon) for all nn0n\geq n_{0} and xI.x\in I.

Proof.

Assume that ϵ\epsilon is small enough so that it has the following property: if d(x,y)<ϵd(x,y)<\epsilon but y𝒫(x)y\not\in\mathcal{P}(x), then d(g(x),g(y))ϵd(g(x),g(y))\geq\epsilon. Equivalently, if d(x,y)<ϵd(x,y)<\epsilon and d(g(x),g(y))<ϵd(g(x),g(y))<\epsilon, then y𝒫(x)y\in\mathcal{P}(x). If no such ϵ\epsilon existed, then at least one of the points βi\beta_{i} would be a removable discontinuity. Since gg has the idoc property, then ||𝒫n||:=maxP𝒫n|P|0||\mathcal{P}^{n}||\mathrel{\mathop{:}}=\max_{P\in\mathcal{P}^{n}}|P|\downarrow 0. Let n0n_{0} be such that 𝒫n0<ϵd|I|||\mathcal{P}^{n_{0}}||<\frac{\epsilon}{d|I|}.

Since 𝒫n=i=0n1gi(P)\mathcal{P}^{n}=\bigwedge_{i=0}^{n-1}g^{-i}(P), then for all nn0n\geq n_{0} and y𝒫n(x)y\in\mathcal{P}^{n}(x), we have that d(gi(x),gi(y))<ϵd(g^{i}(x),g^{i}(y))<\epsilon for 0in10\leq i\leq n-1. This proves that 𝒫n(x)Bgn(x,ϵ).\mathcal{P}^{n}(x)\subset B^{n}_{g}(x,\epsilon).

To see the other containment, suppose that d(gi(x),gi(y))<ϵd(g^{i}(x),g^{i}(y))<\epsilon for i=0,,ni=0,\dots,n. By the property establishing the smallness of ϵ\epsilon, it follows that gi(x)𝒫(gi(y))g^{i}(x)\in\mathcal{P}(g^{i}(y)) for i=0,,n1i=0,\dots,n-1. ∎

With the above, we can conclude the following about the topological entropy.

Proposition 7.5.

Let gλ,πg_{\lambda,\pi} be dd-IET. Suppose that it has the idoc property. Then |𝒫n|=dn|\mathcal{P}^{n}|=dn and htop,aχ=1h_{top,a_{\chi}}=1 with the polynomial scale aχ(n)=nχa_{\chi}(n)=n^{\chi}.

Proof.

The idoc property implies that 𝒫n=i=0n1gi(𝒫)\mathcal{P}^{n}=\bigwedge_{i=0}^{n-1}g^{-i}(\mathcal{P}) has exactly dndn atoms. Applying 7.4,

δg,I,χ(ϵ)=lim supndnχ1.\delta_{g,I,\chi}(\epsilon)=\limsup_{n\to\infty}\frac{d}{n^{\chi-1}}.

Then

htop,aχ(g)=limϵ0(sup{χ:δg,I,χ(ϵ)>0})=1.h_{top,a_{\chi}}(g)=\lim_{\epsilon\to 0}(\sup\{\chi:\delta_{g,I,\chi}(\epsilon)>0\})=1.

Define ϵn:=minP𝒫n|P|\epsilon_{n}\mathrel{\mathop{:}}=\min_{P\in\mathcal{P}^{n}}|P|, the length of the smallest atom in the partition 𝒫n.\mathcal{P}^{n}. An IET is linearly recurrent if there exists a constant C>0C>0 such that for every n1n\geq 1, then nϵnC.n\epsilon_{n}\geq C.

We have the following characterization of a big class of dd-IETs.

Proposition 7.6.

Suppose that gg is an idoc dd-IET. The following are equivalent:

  1. (1)

    The Lebesgue measure is homogeneous, see 4.4.

  2. (2)

    gg is linearly recurrent.

Proof.

If we assume (1), fix ϵ>0\epsilon>0 and c>0c>0 as in the definition of homogeneous measure. Then

1cLeb(Bgn(x,ϵ))Leb(Bgn(y,ϵ))c\frac{1}{c}\leq\frac{\operatorname{Leb}(B^{n}_{g}(x,\epsilon))}{\operatorname{Leb}(B^{n}_{g}(y,\epsilon))}\leq c

for all nn sufficiently large. In particular, it follows that for every P𝒫nP\in\mathcal{P}^{n},

|P(y)|c=Leb(Bgn(y,ϵ))cϵn,\frac{|P(y)|}{c}=\frac{\operatorname{Leb}(B^{n}_{g}(y,\epsilon))}{c}\leq\epsilon_{n},

adding both sides of the inequality over P𝒫nP\in\mathcal{P}^{n} we obtain C:=1dcnϵnC\mathrel{\mathop{:}}=\frac{1}{dc}\leq n\epsilon_{n} for all nn sufficiently large. This proves (2).

Assume (2), and suppose by contradiction that (1) does not happen. So, for every 1>k>01>k>0 there exists nn very large such that

Leb(Bgn(x,ϵ))Leb(Bgn(y,ϵ))<k.\frac{\operatorname{Leb}(B^{n}_{g}(x,\epsilon))}{\operatorname{Leb}(B^{n}_{g}(y,\epsilon))}<k.

Without loss of generality, we can assume that ϵn=Leb(Bgn(x,ϵ))\epsilon_{n}=\operatorname{Leb}(B^{n}_{g}(x,\epsilon)) and Leb(Bgn(y,ϵ))<1n\operatorname{Leb}(B^{n}_{g}(y,\epsilon))<\frac{1}{n}. Then we have that nϵn=nLeb(Bgn(x,ϵ))<kn\epsilon_{n}=n\operatorname{Leb}(B^{n}_{g}(x,\epsilon))<k. In particular, if k=Ck=C, which contradicts (2). ∎

We have the main result of this section:

Corollary 7.7.

Suppose that gg is an idoc, linearly recurrent dd-IET. Then, for aχ=nχa_{\chi}=n^{\chi}, we have that

(26) hsemi,Leb,aχ(g)=htop,aχ=1h_{semi,Leb,a_{\chi}}(g)=h_{top,a_{\chi}}=1
Proof.

4.5 and 7.6 imply that hsemi,Leb,aχ(g)=htop,aχ(g)h_{semi,\operatorname{Leb},a_{\chi}}(g)=h_{top,a_{\chi}}(g). 7.5 implies that the equality is 1. ∎

We note that this does not say anything about the metric slow entropy of IETs. The class of linearly recurrent IET is uniquely ergodic by [Vee87, Theorem 1.2] then, computing the metric slow entropy of a linearly recurrent IET with respect to the Lebesgue measure will say whether such IET is variational. We also want to remark that the conditions in 7.7 are satisfied by a set of parameters λ\lambda of Hausdorff dimension dd. This was noted by D. Robertson following his proof of [Rob19, Proposition 4] and the proof of [CCM13, Theorem 1.4].

7.3. On 3-IETs

Any vector of positive entries λ=(λA,λB,λC)\lambda=(\lambda_{A},\lambda_{B},\lambda_{C}) and the symmetric permutation π=(ABCCBA)\pi=\begin{pmatrix}A&B&C\\ C&B&A\end{pmatrix} determines a 3-IET given by Equation 25, which in a simpler form is

g(x)={x+λB+λC if x[0,λA),x+λCλA if x[λA,λA+λB),xλAλB if x[λA+λB,|I|).g(x)=\left\{\begin{array}[]{ll}x+\lambda_{B}+\lambda_{C}&\text{ if }x\in[0,\lambda_{A}),\\ x+\lambda_{C}-\lambda_{A}&\text{ if }x\in[\lambda_{A},\lambda_{A}+\lambda_{B}),\\ x-\lambda_{A}-\lambda_{B}&\text{ if }x\in[\lambda_{A}+\lambda_{B},|I|).\end{array}\right.

It is common to think of 3-IETs over I=[0,1)I=[0,1), but we will consider 3-IETs with I=[0,1+ξ)I=[0,1+\xi), 𝒜={A,B,C}\mathcal{A}=\{A,B,C\}, and the symmetric permutation π=(ABCCBA)\pi=\begin{pmatrix}A&B&C\\ C&B&A\end{pmatrix}.

Let α(0,1)\alpha\in(0,1) be an irrational number, and let {pmqm}m1\left\{\frac{p_{m}}{q_{m}}\right\}_{m\geq 1} be the sequence of best approximations. Let x|\!|x|\!| be equal to

||x||:=minn|xn|.|\!|x|\!|\mathrel{\mathop{:}}=\min_{n\in\mathbb{Z}}|x-n|.

The number α\alpha is badly approximable if there exists Cα>0C_{\alpha}>0 depending on α\alpha, satisfying

qm+1Cαqm,q_{m+1}\leq C_{\alpha}q_{m},

for every m.m\in\mathbb{N}. Equivalently, α\alpha is badly approximable, if there exists Dα>0,D_{\alpha}>0, such that

mαDαm|\!|m\alpha|\!|\geq\frac{D_{\alpha}}{m}

for every m.m\in\mathbb{N}.

For every real number α\alpha, we denote SαS_{\alpha} the set

Sα={ξ:ξjαCξqn for every qn<j<qn}.S_{\alpha}=\{\xi\in\mathbb{R}:|\!|\xi-j\alpha|\!|\geq\frac{C_{\xi}}{q_{n}}\textit{\;for every $-q_{n}<j<q_{n}$}\}.

We have the following property regarding the Hausdorff dimension of the numbers.

Lemma 7.8.

The set of badly approximable real numbers has Hausdorff dimension one. For arbitrary α\alpha, the set SαS_{\alpha} has Hausdorff dimension one.

Proof.

For badly approximable numbers, this was proved in [Sch1, Theorem 3]. Lastly, [Tse09, Theorem 1] proved that dimSα=1\dim S_{\alpha}=1 for every real number α\alpha. ∎

Lemma 7.9.

Define

X={(α,ξ)(0,1)×(0,1):α is badly approximable and ξSα}.X=\{(\alpha,\xi)\in(0,1)\times(0,1):\alpha\text{ is badly approximable and }\xi\in S_{\alpha}\}.

Then Hausdorff dimension of XX, dimX=2.\dim X=2.

Proof.

We apply the following theorem about the Hausdorff dimension of products [Fal85, Theorem 5.8 and Excercise 5.2]: Let A,BA,B be Borel subsets of a Euclidean spaces and let YA×BY\subset A\times B be such that for all aAa\in A, dim{bB:(a,b)Y}d.\dim\{b\in B:(a,b)\in Y\}\geq d. Then

dimYdimA+d.\dim Y\geq\dim A+d.

Let AA be the set of badly approximable numbers, B=(0,1)B=(0,1), and replace XX into YY. The set {b:(a,b)X}\{b:(a,b)\in X\} is the set Sa.S_{a}. From 7.8, dimSa=1.\dim S_{a}=1.

Therefore dimXdimA+1.\dim X\geq\dim A+1. Thus dimX=2\dim X=2, because dimA=1\dim A=1 as stated in 7.8. ∎

The proof of 7.1 follows from 7.10 and 7.11. 7.10 is the core of the argument and its proof will be postponed until Section 7.5.

Proposition 7.10.

Let ξ(0,1)\xi\in(0,1) and let λ=(ξ,λB,λC)+3\lambda=(\xi,\lambda_{B},\lambda_{C})\in\mathbb{R}^{3}_{+} such that λB+λC=1\lambda_{B}+\lambda_{C}=1. Define

α={λCξ if λC>ξ,1+λCξ if λC<ξ.\alpha=\left\{\begin{array}[]{cc}\lambda_{C}-\xi&\text{ if }\lambda_{C}>\xi,\\ 1+\lambda_{C}-\xi&\text{ if }\lambda_{C}<\xi.\end{array}\right.

If α\alpha is badly approximable, and ξSα\xi\in S_{\alpha}, then the 3-IET f:=fλ,πf\mathrel{\mathop{:}}=f_{\lambda,\pi} determined in this way satisfies hLeb,aχ(f)=1h_{\operatorname{Leb},a_{\chi}}(f)=1 with polynomial scale aχ(n)=nχ.a_{\chi}(n)=n^{\chi}.

Lemma 7.11.

Let λ>03\lambda\in\mathbb{R}^{3}_{>0}, c>0c>0 and π\pi be a symmetric permutation. The 3-IETs g:=gλ,πg\mathrel{\mathop{:}}=g_{\lambda,\pi} and h:=hcλ,πh\mathrel{\mathop{:}}=h_{c\lambda,\pi} are smoothly conjugated by the map xcxx\mapsto cx. Moreover, any 3-IET g:=g(λA,λB,λC),πg\mathrel{\mathop{:}}=g_{(\lambda_{A},\lambda_{B},\lambda_{C}),\pi} is measurably conjugated to its inverse g1=g(λC,λB,λA),π1g^{-1}=g^{-1}_{(\lambda_{C},\lambda_{B},\lambda_{A}),\pi}, by the hyperelliptic involution map ι(x)=|I|x.\iota(x)=|I|-x.

Proof.

For c>0c>0, let f:[0,αλα)[0,cαλα)f:[0,\sum_{\alpha}\lambda_{\alpha})\to[0,c\sum_{\alpha}\lambda_{\alpha}) be the map xcx.x\mapsto cx. Let gg be the 3-IET determined by the length vector λ\lambda and the permutation π\pi. Let hh be the 3-IET determined by the length vector cλc\lambda and the permutation π.\pi. The map ff is a diffeomorphism that sends the Lebesgue measure on [0,αλα)[0,\sum_{\alpha}\lambda_{\alpha}) to the Lebesgue measure on [0,cαλα)[0,c\sum_{\alpha}\lambda_{\alpha}). The IET gg maps by translation the segment [πt(α)<iλα,πt(α)iλα)[\sum_{\pi_{t}(\alpha)<i}\lambda_{\alpha},\sum_{\pi_{t}(\alpha)\leq i}\lambda_{\alpha}) to the segment [πb(α)<4iλα,πb(α)4iλα)[\sum_{\pi_{b}(\alpha)<4-i}\lambda_{\alpha},\sum_{\pi_{b}(\alpha)\leq 4-i}\lambda_{\alpha}). Thus the composition fgf\circ g maps the segment [πt(α)<iλα,πt(α)iλα)[\sum_{\pi_{t}(\alpha)<i}\lambda_{\alpha},\sum_{\pi_{t}(\alpha)\leq i}\lambda_{\alpha}) by stretching and translating to the segment [cπb(α)<4iλα,cπb(α)4iλα)[c\sum_{\pi_{b}(\alpha)<4-i}\lambda_{\alpha},c\sum_{\pi_{b}(\alpha)\leq 4-i}\lambda_{\alpha}).

Similarly, hh maps by translation the interval [cπt(α)<iλα,cπt(α)iλα)[c\sum_{\pi_{t}(\alpha)<i}\lambda_{\alpha},c\sum_{\pi_{t}(\alpha)\leq i}\lambda_{\alpha}) to the interval [cπb(α)<4iλα,cπb(α)4iλα).[c\sum_{\pi_{b}(\alpha)<4-i}\lambda_{\alpha},c\sum_{\pi_{b}(\alpha)\leq 4-i}\lambda_{\alpha}). In conclusion, for all x[0,αλα)x\in[0,\sum_{\alpha}\lambda_{\alpha}), we have fg(x)=hf(x).f\circ g(x)=h\circ f(x). Proving that gg and hh are smoothly conjugated.

Denote I=[0,λA+λB+λC)I=[0,\lambda_{A}+\lambda_{B}+\lambda_{C}), and fix the permutation π=(ABCCBA)\pi=\begin{pmatrix}A&B&C\\ C&B&A\end{pmatrix}. Let gg be the 3-IET determined by the length vector λ=(λA,λB,λC)\lambda=(\lambda_{A},\lambda_{B},\lambda_{C}) and the permutation π.\pi. Let g1g^{-1} be the inverse map of gg. We let the reader verify that this is a 3-IET determined by the length vector λ=(λA,λB,λC)(λC,λB,λA)\lambda^{\prime}=(\lambda^{\prime}_{A},\lambda^{\prime}_{B},\lambda^{\prime}_{C})\equiv(\lambda_{C},\lambda_{B},\lambda_{A}) and the permutation π.\pi. Let ι:II\iota:I\to I be the map x|I|x.x\mapsto|I|-x. The map ι\iota sends the open interval (π(α)<iλα,π(α)iλα)(\sum_{\pi_{*}(\alpha)<i}\lambda_{\alpha},\sum_{\pi_{*}(\alpha)\leq i}\lambda_{\alpha}) to the open interval (π(α)<4iλα,π(α)4iλα)(\sum_{\pi_{*}(\alpha)<4-i}\lambda^{\prime}_{\alpha},\sum_{\pi_{*}(\alpha)\leq 4-i}\lambda^{\prime}_{\alpha}) by translation and order reversing, where {t,b}*\in\{t,b\}. The composition ιg\iota\circ g maps the interval (πt(α)<iλα,πt(α)iλα)(\sum_{\pi_{t}(\alpha)<i}\lambda_{\alpha},\sum_{\pi_{t}(\alpha)\leq i}\lambda_{\alpha}) to the interval (πb(α)<iλα,πb(α)iλα)(\sum_{\pi_{b}(\alpha)<i}\lambda^{\prime}_{\alpha},\sum_{\pi_{b}(\alpha)\leq i}\lambda^{\prime}_{\alpha}) by a translation and order reversing. Also, the composition g1ιg^{-1}\circ\iota maps the open interval (πt(α)<iλα,πt(α)iλα)(\sum_{\pi_{t}(\alpha)<i}\lambda_{\alpha},\sum_{\pi_{t}(\alpha)\leq i}\lambda_{\alpha}) to the interval (πb(α)<iλα,πb(α)iλα)(\sum_{\pi_{b}(\alpha)<i}\lambda^{\prime}_{\alpha},\sum_{\pi_{b}(\alpha)\leq i}\lambda^{\prime}_{\alpha}) by a translation and order reversing. Thus, we have the equality ιg=g1ι\iota\circ g=g^{-1}\circ\iota on the interior of the intervals IiI_{i} for i{1,2,3}i\in\{1,2,3\}. The equality does not occur at the left endpoints of the intervals IαI_{\alpha}, for example ι(g(0))=|I|g(0)=|I|λAλB=λC\iota(g(0))=|I|-g(0)=|I|-\lambda_{A}-\lambda_{B}=\lambda_{C}, but g1(ι(0))g^{-1}(\iota(0)) is not defined because ι(0)=|I|\iota(0)=|I| and g1g^{-1} is not defined at |I|.|I|. Thus the maps gg and g1g^{-1} are measurable conjugated, since the map ι\iota is a diffeomorphism that preserves the Lebesgue measure, and ι(g(x))=g(ι(x))\iota(g(x))=g(\iota(x)) except at finitely many xI.x\in I.

7.4. Suspensions of a 3-IET and a rotation

Let (T,X,μ)(T,X,\mu) be measure preserving transformation and f:X>0f\colon X\to\mathbb{R}_{>0} an L1(μ)L^{1}(\mu) function. Let XfX^{f} be the quotient

Xf={(x,s):xX and 0sf(x)}/,X^{f}=\left\{(x,s):x\in X\text{ and }0\leq s\leq f(x)\right\}/\sim,

with (x,f(x))(T(x),0).(x,f(x))\sim(T(x),0). The suspension flow determined by TT and ff is the flow Ttf:=t:XfXfT^{f}_{t}\mathrel{\mathop{:}}=\mathcal{F}_{t}\colon X^{f}\to X^{f} defined by t(x,s)=(Tn(x),s+ti=0n1f(Tix))\mathcal{F}_{t}(x,s)=(T^{n}(x),s+t-\sum_{i=0}^{n-1}f(T^{i}x)), where n0n\geq 0 is such that f(n)(x)s+t<f(n+1)(x),f^{(n)}(x)\leq s+t<f^{(n+1)}(x), where f(n)(x)f^{(n)}(x) is the Birkhoff sum

f(n)(x):=i=0n1f(Tix).\displaystyle f^{(n)}(x)\mathrel{\mathop{:}}=\sum_{i=0}^{n-1}f(T^{i}x).

7.4.1. A specific suspension for a 3-IET

Consider a 3-IET with |I|=1+ξ|I|=1+\xi, and the conditions λA=ξ<λB+λC=1\lambda_{A}=\xi<\lambda_{B}+\lambda_{C}=1. The suspension with the constant function 1 of this 3-IET is presented in Figure 1A. This construction starts with a rectangle of length 1+ξ1+\xi and height 1, by convenience assume that the left bottom corner is placed at the origin. The sides of the rectangle are identified by translation as follows

(27) [0,λC)×{0}[λA+λB,1+ξ)×{1}[λC,λB+λC)×{0}[λA,λA+λB)×{1}[1,1+ξ)×{0}[0,λA)×{1}{0}×[0,1){1+ξ}×[0,1).\begin{split}[0,\lambda_{C})\times\{0\}&\sim[\lambda_{A}+\lambda_{B},1+\xi)\times\{1\}\\ [\lambda_{C},\lambda_{B}+\lambda_{C})\times\{0\}&\sim[\lambda_{A},\lambda_{A}+\lambda_{B})\times\{1\}\\ [1,1+\xi)\times\{0\}&\sim[0,\lambda_{A})\times\{1\}\\ \{0\}\times[0,1)&\sim\{1+\xi\}\times[0,1).\end{split}
AAAABBBBCCCC1111
A
BBBBC1C_{1}C2C_{2}AAC1C_{1}C2C_{2}
B
Figure 1. This is case λC>ξ\lambda_{C}>\xi in Equation 28. In Figure 1A, the sides of a rectangle [0,1+ξ)×[0,1][0,1+\xi)\times[0,1] are identified according to Equation 27. The vertical flow’s first return time to [0,1+ξ)[0,1+\xi) is a 3-IET. In Figure 1B, suspension of a rotation on [0,1][0,1] with f=2𝟙[0,ξ)+𝟙[ξ,1)f=2\mathds{1}_{[0,\xi)}+\mathds{1}_{[\xi,1)}. This can also be obtained by cutting over the dotted line in Figure 1A and gluing the small piece over the segment [1,1+ξ)×{1}.[1,1+\xi)\times\{1\}.
AAAABBBBCCCC11
A
CCB2B_{2}B1B_{1}CCAAB1B_{1}B2B_{2}
B
Figure 2. This is case λC<ξ\lambda_{C}<\xi in Equation 28. In Figure 2A, the sides of a rectangle [0,1+ξ)×[0,1][0,1+\xi)\times[0,1] are identified according to Equation 27. The vertical flow’s first return time to [0,1+ξ)[0,1+\xi) is a 3-IET. In Figure 2B, suspension of a rotation on [0,1][0,1] with f=2𝟙[0,ξ)+𝟙[ξ,1)f=2\mathds{1}_{[0,\xi)}+\mathds{1}_{[\xi,1)}. This can also be obtained by cutting over the dotted line in Figure 2A and gluing the small piece over the segment [1,1+ξ)×{1}.[1,1+\xi)\times\{1\}.

7.4.2. Suspension flow of an irrational rotation and proof of 7.1

Towards the computation of the slow entropy of the 3-IET mentioned in Section 7.4.1, we will compute the slow entropy of the vertical flow in the construction involving the suspension of the 3-IET in Figure 1B. An equivalent construction is the suspension of a rotation T:[0,1)[0,1),xx+αmod1T:[0,1)\to[0,1),x\mapsto x+\alpha\mod 1. The suspension function is f=2𝟙[0,ξ)+𝟙[ξ,1).f=2\mathds{1}_{[0,\xi)}+\mathds{1}_{[\xi,1)}. Computations for the first return time of the vertical flow to the segment [0,1)×{0}[0,1)\times\{0\} show that

(28) α={λCξif λC>ξ,1+λCξif λC<ξ.\alpha=\begin{dcases}\lambda_{C}-\xi&\text{if }\lambda_{C}>\xi,\\ 1+\lambda_{C}-\xi&\text{if }\lambda_{C}<\xi.\end{dcases}

Although we are suspending a rotation, the 3-IET is still present in the first return map of the vertical flow to the segment I=[0,1)×{0}[0,ξ){1}[0,1)f.I=[0,1)\times\{0\}\cup[0,\xi)\cup\{1\}\subset[0,1)^{f}. Figure 1A and Figure 1B are proof by picture of the case λC>λA=ξ\lambda_{C}>\lambda_{A}=\xi. Because the rotation angle α\alpha in Figure 1B is equal to the length of the segment labeled C1C_{1}, then α\alpha is equal to λCλA=λCξ\lambda_{C}-\lambda_{A}=\lambda_{C}-\xi. Figure 2A and Figure 2B are proof by picture of the case λC<λA=ξ\lambda_{C}<\lambda_{A}=\xi. The rotation angle α\alpha of the suspension in Figure 2B is equal to the sum of the lengths of the segments labeled CC and B1B_{1}. The length of the latter is equal to 1λA=1ξ1-\lambda_{A}=1-\xi, then α=1+λCξ.\alpha=1+\lambda_{C}-\xi.

Starting with α,ξ[0,1]\alpha,\xi\in[0,1], the corresponding 3-IET is given by the length vector:

(29) F0(α,ξ):=(λA,λB,λC)={(ξ,1αξ,α+ξ)if α+ξ<1,(ξ,2αξ,α+ξ1)if α+ξ>1.F_{0}(\alpha,\xi)\mathrel{\mathop{:}}=(\lambda_{A},\lambda_{B},\lambda_{C})=\begin{dcases}(\xi,1-\alpha-\xi,\alpha+\xi)&\text{if }\alpha+\xi<1,\\ (\xi,2-\alpha-\xi,\alpha+\xi-1)&\text{if }\alpha+\xi>1.\end{dcases}

The proof of Equation 29 follows from similar computations and pictures as a verification of Equation 28.

Proof of 7.1.

Let Δ\Delta be the set {(x,y,z):x+y+z=1,x,y,z>0}.\{(x,y,z):x+y+z=1,\,x,y,z>0\}. The subset XX in 7.9 is of Hausdorff dimension 2. Let F:[0,1]×[0,1]ΔF:[0,1]\times[0,1]\to\Delta be the function defined by F(α,ξ)=11+ξF0(α,ξ).F(\alpha,\xi)=\frac{1}{1+\xi}F_{0}(\alpha,\xi). Since the function F0F_{0} is linear and of rank 2 in connected components, the Hausdorff dimension of the set F0(X)F_{0}(X) is 2. Additionally, the factor of 1/(1+ξ)1/(1+\xi) preserves the Hausdorff dimension since it is the normalization factor, the l1l_{1}-norm of any point in the image of F0F_{0} is 1/(1+ξ).1/(1+\xi). This proves that the set A:=F(X)A\mathrel{\mathop{:}}=F(X) is of Hausdorff dimension 2.

Note that by 7.10, the metric slow entropy of the 3-IET determined by the vector F0(α,ξ)F_{0}(\alpha,\xi) is 1, by 7.11, the 3-IET determined by F0(α,ξ)F_{0}(\alpha,\xi) and the 3-IET determined by F(α,ξ)=11+ξF0(α,ξ)F(\alpha,\xi)=\frac{1}{1+\xi}F_{0}(\alpha,\xi) are measurably conjugated. Then, every 3-IET determined by a vector in F(X)F(X) must have metric slow entropy 1 with aχ=nχ.a_{\chi}=n^{\chi}.

7.5. Computation of slow entropy and proof of 7.10

We aim to compute the growth rate of the number of Hamming balls of time RR of the suspension flow TtfT^{f}_{t}, because as we mentioned in the previous section, this suspension is equivalent to the suspension of a 3-IET with constant roof function 1. Therefore, these Hamming balls’ growth rate is the same as the Hamming balls’ of the suspended 3-IET.

Let α\alpha and ξ\xi as before and f=d1𝟙[0,ξ)+d2𝟙[ξ,1)f=d_{1}\mathds{1}_{[0,\xi)}+d_{2}\mathds{1}_{[\xi,1)}. The conditions on α\alpha and ξ\xi will be given later. Given a subset A𝕋A\subset\mathbb{T}, we use AfA^{f} to denote the set {(y,t)𝕋f:yA}\{(y,t)\in\mathbb{T}^{f}:y\in A\}. We also let λf\lambda^{f} denote the normalized Lebesgue measure restricted to 𝕋f\mathbb{T}^{f}. Let M=max{d1,d2,1d1,1d2}M=\max\{d_{1},d_{2},\frac{1}{d_{1}},\frac{1}{d_{2}}\} and e=|d1d2|.e=|d_{1}-d_{2}|. The special situation in which we prove 7.10 is by setting d1=2d_{1}=2 and d2=1.d_{2}=1. Given R>0R>0 very large and ϵ>0\epsilon>0, fix a generating partition 𝒫\mathcal{P} such that all atoms in the partition are squares with length between 1/2k1/2k to 1/k1/k for some large value kk to be specified later. We want to show that STf,𝒫(ϵ,R)CRS_{T^{f},\mathcal{P}}(\epsilon,R)\geq CR holds for some constant CC, hence

lim supRSTf,𝒫(ϵ,R)Rχ=\displaystyle\limsup_{R\rightarrow\infty}\frac{S_{T^{f},\mathcal{P}}(\epsilon,R)}{R^{\chi}}=\infty

for any χ<1\chi<1. This gives us a lower bound for the metric slow entropy, i.e. hμ,aχ(Tf)1.h_{\mu,a_{\chi}}(T^{f})\geq 1.

From now on, we will assume that α\alpha is badly approximable; otherwise, we will specify.

To prove 7.10, we combine several lemmas which describe recurrence properties for the base circle rotation RαR_{\alpha}. In fact, this will allow us to compute the slow entropy of some other special flows with piecewise constant roof functions (see 7.16). We therefore state a few lemmas (7.12 - 7.15), which we use to prove 7.10, delaying their proof until later in this section.

The following proposition uses similar ideas in [FLL07, Lemma 4], and is a Ratner-type property for the special flows we consider.

Proposition 7.12.

Suppose α(0,1)\alpha\in(0,1) is badly approximable, and ξSα\xi\in S_{\alpha}. There exists κ,c>0\kappa,c>0 and a finite set V{0}V\subset{\mathbb{R}\setminus\{0\}} such that for all large enough mm, and x,y𝕋x,y\in\mathbb{T} with

Cξ2qm+1xy<Cξ2qm,\frac{C_{\xi}}{2q_{m+1}}\leq|\!|x-y|\!|<\frac{C_{\xi}}{2q_{m}},

for some time NN where

qmNcqm+1+2qm,q_{m}\leq N\leq cq_{m+1}+2q_{m},

the following set

Λ={n[0,N]:f(n)(x)f(n)(y)V}\Lambda=\{n\in[0,N]\cap\mathbb{Z}:f^{(n)}(x)-f^{(n)}(y)\in V\}

has cardinality larger than κN\kappa N.

7.12 shows that in some definite proportion, the Birkhoff sum f(n)f^{(n)} of nearby points will differ. Moreover, the next lemma shows that the splitting phenomenon occurring in the Birkhoff sums implies the splitting time of two orbits in special flow space by a definite proportion.

Lemma 7.13.

Let x~=(x,s)\tilde{x}=(x,s), y~=(y,s)\tilde{y}=(y,s^{\prime}) be two elements in the same atom of partition. If

Cξ2qm+1xy<Cξ2qm\frac{C_{\xi}}{2q_{m+1}}\leq|\!|x-y|\!|<\frac{C_{\xi}}{2q_{m}}

for some mm, there exists some time TT and constant DD where TDqmT\leq Dq_{m} and κ>0\kappa^{\prime}>0 such that the following set

Λ~={t<T:Ttfx~ and Ttfy~ are not in the same atom }\tilde{\Lambda}=\{t<T:T_{t}^{f}{\tilde{x}}\text{\;and\;}T_{t}^{f}{\tilde{y}}\text{\;are not in the same atom }\}

has measure |Λ~||\tilde{\Lambda}| larger than κT\kappa^{\prime}T.

Let x~=(x,s)\tilde{x}=(x,s) be an element in 𝕋f\mathbb{T}^{f}. Consider a Hamming ball BTf,𝒫R(x~,ϵ)BR(x~,ϵ)B_{T^{f},\mathcal{P}}^{R}(\tilde{x},\epsilon)\eqqcolon B^{R}(\tilde{x},\epsilon) centered at x~\tilde{x}, then for any y~=(y,s)BR(x~,ϵ)\tilde{y}=(y,s^{\prime})\in B^{R}(\tilde{x},\epsilon) we want to show there exists n1n_{1} and n2n_{2} such that x+n1α(y+n2α)<H/R|\!|x+n_{1}\alpha-(y+n_{2}\alpha)|\!|<H/R for some constant HH.

Lemma 7.14.

There exists a constant H,H, such that for any y~=(y,s)BR(x~,ϵ)\tilde{y}=(y,s^{\prime})\in B^{R}(\tilde{x},\epsilon), there are t0,s0,s0(0,R)t_{0},s_{0},s_{0}^{\prime}\in(0,R), n1=n1(t0,x,s)n_{1}=n_{1}(t_{0},x,s)\in\mathbb{Z} and n2=n2(t0,y,s)n_{2}=n_{2}(t_{0},y,s^{\prime})\in\mathbb{Z} satisfying x+n1α(y+n2α)<H/R|\!|x+n_{1}\alpha-(y+n_{2}\alpha)|\!|<H/R, where Tt0fx~=(x+n1α,s0)T_{t_{0}}^{f}\tilde{x}=(x+n_{1}\alpha,s_{0}) and Tt0fy~=(y+n2α,s0)T_{t_{0}}^{f}\tilde{y}=(y+n_{2}\alpha,s_{0}^{\prime}) are in the same atom of 𝒫\mathcal{P}.

From 7.14, for every y~=(y,s)BR(x~,ϵ)\tilde{y}=(y,s^{\prime})\in B^{R}(\tilde{x},\epsilon), we obtain xy(n2n1)αH/R.|\!|x-y-(n_{2}-n_{1})\alpha|\!|\leq H/R. Therefore, y~\tilde{y} is in

j=|n2n1||n2n1|[x+jαHR,x+jα+HR]f.\displaystyle\bigcup_{j=-|n_{2}-n_{1}|}^{|n_{2}-n_{1}|}\left[x+j\alpha-\frac{H}{R},x+j\alpha+\frac{H}{R}\right]^{f}.

To get a lower bound of STf,𝒫(ϵ,R)S_{T^{f},\mathcal{P}}(\epsilon,R), we will compute a uniform upper bound for the Lebesgue measure of BR(x~,ϵ)B^{R}(\tilde{x},\epsilon), it suffices to find an upper bound for |n2n1||n_{2}-n_{1}| for all x~\tilde{x} and y~BR(x~,ϵ)\tilde{y}\in B^{R}(\tilde{x},\epsilon). 7.15 below, summarizes this idea.

Lemma 7.15.

There exists a constant G>0G>0 depending only on α\alpha and ξ\xi such that |n2n1|G|n_{2}-n_{1}|\leq G, where n1n_{1} and n2n_{2} are the integers in 7.14.

Let 𝒟[0,1]\mathcal{D}\subset[0,1] be the set consisting of irrational numbers α\alpha such that there exists C1>0C_{1}>0 depending on α\alpha, satisfying

qm+1C1qmlog2qmq_{m+1}\leq C_{1}q_{m}\log^{2}q_{m}

for every m.m. In particular 𝒟\mathcal{D} contains badly approximable irrational numbers.

Proposition 7.16.

If α𝒟\alpha\in\mathcal{D}, ξ𝕋\xi\in\mathbb{T} and we consider the roof function f=d1𝟙[0,ξ)+d2𝟙[ξ,1)f=d_{1}\mathds{1}_{[0,\xi)}+d_{2}\mathds{1}_{[\xi,1)}, the metric slow entropy of the special flow system is at most 1 for scale aχ(t)=tχa_{\chi}(t)=t^{\chi}.

Proof of 7.10.

From 7.14, for RR large enough and y~BR(x,ϵ)\tilde{y}\in B^{R}(x,\epsilon), there are n1,n2n_{1},n_{2} natural numbers, such that

y~j=|n2n1||n2n1|[x+jαHR,x+jα+HR]f.\tilde{y}\in\displaystyle\bigcup_{j=-|n_{2}-n_{1}|}^{|n_{2}-n_{1}|}\left[x+j\alpha-\frac{H}{R},x+j\alpha+\frac{H}{R}\right]^{f}.

From 7.15, there is an upper bound GG for |n2n1||n_{2}-n_{1}| independent of y~.\tilde{y}. Thus Leb(BR(x,ϵ))4GHR.\operatorname{Leb}(B^{R}(x,\epsilon))\leq\frac{4GH}{R}. In other words, for any large R>0R>0, the measure of each Hamming ball is bounded above by C/RC^{\prime}/R for some constant CC^{\prime}. Thus, there exists a constant C>0C>0 independent of RR and ϵ\epsilon such that STf,𝒫(ϵ,R)CRS_{T^{f},\mathcal{P}}(\epsilon,R)\geq CR. This shows the slow entropy χ\chi for this special flow is at least 1 using time scale aχ(t)=tχa_{\chi}(t)=t^{\chi}.

Since α𝒟\alpha\in\mathcal{D} (α\alpha is badly approximable), applying 7.16 the upper bound of the metric slow entropy is 1. ∎

The remaining of this paper is dedicated to prove 7.12,
7.13, 7.14, 7.15, and 7.16.

Proof of 7.12.

By the definition of f(n)f^{(n)}, and assuming that 0<x<y<10<x<y<1, it follows:

f(n)(x)f(n)(y)=ej=0n1𝟙(x,y](ξjα)ej=0n1𝟙(x,y](jα).f^{(n)}(x)-f^{(n)}(y)=e\sum_{j=0}^{n-1}\mathds{1}_{(x,y]}(\xi-j\alpha)-e\sum_{j=0}^{n-1}\mathds{1}_{(x,y]}(-j\alpha).

By assumption on ξ\xi, the interval of (x,y](x,y] can be crossed by the orbit of ξ\xi (and the orbit of 0) for at most cqm+1+2qmqm+1[cCα]+4U\frac{cq_{m+1}+2q_{m}}{q_{m}}+1\leq[cC_{\alpha}]+4\eqqcolon U times. Hence, choose

V={ne:UnU and n0}.V=\{ne:-U\leq n\leq U\text{ and }n\neq 0\}.

Taking κ=12cCα+4\kappa=\frac{1}{2cC_{\alpha}+4}, assume for all 0nqm,0\leq n\leq q_{m}, f(n)(x)f(n)(y)Vf^{(n)}(x)-f^{(n)}(y)\in V, then we are done. Suppose n0n_{0} is the minimum number such that

f(n0)(x)f(n0)(y)=0.f^{(n_{0})}(x)-f^{(n_{0})}(y)=0.

There exists I0I_{0}, where I0I_{0} is the minimum time such that j=0I0Rαj(x,y]\bigcup\limits_{j=0}^{I_{0}}R_{\alpha}^{j}(x,y] covers 0 or ξ\xi. Because α\alpha is badly approximable, let cc be the constant such that I0cqm+1.I_{0}\leq cq_{m+1}. Here cc is a constant depending only on α\alpha and ξ.\xi. Assume without loss of generality, ξ(x+I0α,y+I0α].\xi\in(x+I_{0}\alpha,y+I_{0}\alpha]. Assume at time J0,J_{0}, 0(x+(I0+J0)α,y+(I0+J0)α].0\in(x+(I_{0}+J_{0})\alpha,y+(I_{0}+J_{0})\alpha]. This means

ξ(x+I0α)Cξ2qm|\!|\xi-(x+I_{0}\alpha)|\!|\leq\frac{C_{\xi}}{2q_{m}}

and

x+(I0+J0)αCξ2qm.|\!|x+(I_{0}+J_{0})\alpha|\!|\leq\frac{C_{\xi}}{2q_{m}}.

By triangle inequality, we obtain ξ+J0αCξqm.|\!|\xi+J_{0}\alpha|\!|\leq\frac{C_{\xi}}{q_{m}}. By definition of ξ\xi, this means J0qm.J_{0}\geq q_{m}. Therefore, when n[I0+1,I0+J0]n\in[I_{0}+1,I_{0}+J_{0}],

f(n)(x)f(n)(y)V.f^{(n)}(x)-f^{(n)}(y)\in V.

Hence, choose NN to be the time n0+I0+qmn_{0}+I_{0}+q_{m}, we complete the proof. ∎

Proof of 7.13.

From 7.12, we can find such constants κ\kappa, and NN, and the interval Λ=[a,b]\Lambda=[a,b]\cap\mathbb{N}. For any time t0[f(a)(x)s,f(b)(x)s]t_{0}\in[f^{(a)}(x)-s,f^{(b)}(x)-s], Tt0fx~=(x+n1α,s0)T_{t_{0}}^{f}{\tilde{x}}=(x+n_{1}\alpha,s_{0}), Tt0fy~=(y+n2α,s0)T_{t_{0}}^{f}{\tilde{y}}=(y+n_{2}\alpha,s_{0}^{\prime}). Then

t0=f(n2)(y)+s0s=f(n1)(x)+s0s.t_{0}=f^{(n_{2})}(y)+s_{0}^{\prime}-s^{\prime}=f^{(n_{1})}(x)+s_{0}-s.

If Tt0f(x~)T_{t_{0}}^{f}(\tilde{x}) and Tt0f(y~)T_{t_{0}}^{f}(\tilde{y}) are in the same atom, we can obtain |s0s0|<1/k|s_{0}-s_{0}^{\prime}|<1/k. Since x~\tilde{x} and y~\tilde{y} are also in a same atom, we have

|f(n2)(y)f(n1)(x)|2/k.|f^{(n_{2})}(y)-f^{(n_{1})}(x)|\leq 2/k.

But from 7.12, there exists pVp\in V such that

f(n1)(x)=f(n1)(y)+p.f^{(n_{1})}(x)=f^{(n_{1})}(y)+p.

Thus, by combining the above two equations, we deduce:

|f(n2)(y)f(n1)(y)p|2/k.|f^{(n_{2})}(y)-f^{(n_{1})}(y)-p|\leq 2/k.

Since VV is a fixed finite set, we can assume 2/kmin{|q|:qV}2/k\ll\min{\{|q|:q\in V}\}, so n2n1n_{2}\neq n_{1}. Because Tt0f(x~)T_{t_{0}}^{f}(\tilde{x}) and Tt0f(y~)T_{t_{0}}^{f}(\tilde{y}) are in the same atom and f1Mf\geq\frac{1}{M}, we have the following relations:

(30) Dα|n2n1|(n2n1)α<2k,\frac{D_{\alpha}}{|n_{2}-n_{1}|}\leq|\!|(n_{2}-n_{1})\alpha|\!|<\frac{2}{k},

and

(31) |n2n1|(max{|q|:qV}+1)M.|n_{2}-n_{1}|\leq(\max{\{|q|:q\in V\}}+1)M.

The above two equations are contradictory since we can choose kk to be arbitrarily large. Therefore, for any such time t0[f(a)(x)s,f(b)(x)s]t_{0}\in[f^{(a)}(x)-s,f^{(b)}(x)-s], Tt0f(x~)T_{t_{0}}^{f}(\tilde{x}) and Tt0f(y~)T_{t_{0}}^{f}(\tilde{y}) are not in the same atom. Choosing T=f(b)(x)sT=f^{(b)}(x)-s, and κ=κ/100M\kappa^{\prime}=\kappa/100M. We know that

T(N+1)M(cCα+3)qmM.T\leq(N+1)M\leq(cC_{\alpha}+3)q_{m}M.

Hence, choose D=(cCα+3)MD=(cC_{\alpha}+3)M, we complete the proof. ∎

Proof of 7.14.

By our choice of generating partitions, the set of times for which Ttfx~T_{t}^{f}\tilde{x} and Ttfy~T_{t}^{f}\tilde{y} belong to the (closure of) the same partition element is a union of closed intervals. We may therefore write (0,R)(0,R) as a disjoint union of intervals j=1JIj\displaystyle\bigcup_{j=1}^{J}I_{j}, where for each IjI_{j}, Ttf(x~)T_{t}^{f}(\tilde{x}) and Ttf(y~)T_{t}^{f}(\tilde{y}) are either in the same atom of 𝒫\mathcal{P}, or they are in different atoms of 𝒫\mathcal{P} and the intervals IiI_{i} are maximal among such choices. In particular, either for all odd values or all even values of ii, the interior of IiI_{i} consists of matching times for x~\tilde{x} and y~\tilde{y}.
Case 1. If |J|=1|J|=1 or 2, Ttf(x~)T_{t}^{f}(\tilde{x}) and Ttf(y~)T_{t}^{f}(\tilde{y}) stays in the same atom all time in an interval of length at least (1ϵ)R(1-\epsilon)R. Let xx and yy be the corresponding first coordinates when they stay in the same atom for the first time. We know that

(32) f(n)(x)f(n)(y)=ej=0n1𝟙(x,y](ξjα)ej=0n1𝟙(x,y](jα).f^{(n)}(x)-f^{(n)}(y)=e\sum_{j=0}^{n-1}\mathds{1}_{(x,y]}(\xi-j\alpha)-e\sum_{j=0}^{n-1}\mathds{1}_{(x,y]}(-j\alpha).

Since ff is bounded between d1d_{1} and d2d_{2}, R/2M<N<2MRR/2M<N<2MR where NN is an upper bound of the number of terms in Equation 32. Let d=xyd=|\!|x-y|\!|, as in the proof of 7.12, there exists a constant N(d)N(d) where N(d)KdN(d)\leq\frac{K}{d} for some constant KK, s.t. 𝕋j=1N(d)(x+jα,y+jα]\mathbb{T}\subset\displaystyle\bigcup_{j=1}^{N(d)}(x+j\alpha,y+j\alpha]. By assumption, (x,y](x,y] cannot cross discontinuity points before N/2N/2 times of rotation when the orbit of x~\tilde{x} and y~\tilde{y} stay in the same atom. Hence, N(d)R/4MN(d)\geq R/4M. Therefore, there exists a constant H1H_{1} s.t. dH1/Rd\leq H_{1}/R.
Case 2. If |J|>2|J|>2, we regroup those intervals in the following way.

Let t0t_{0} be the first time when Tt0(x~)T_{t_{0}}(\tilde{x}) and Tt0(y~)T_{t_{0}}(\tilde{y}) stay in the same atom. Denote I~0={t:t<t0}\tilde{I}_{0}=\{t:t<t_{0}\}, note I~0\tilde{I}_{0} could be an empty set. Then let Tt0f(x~)=(x0,s0)T_{t_{0}}^{f}(\tilde{x})=(x_{0},s_{0}), Tt0f(y~)=(y0,s0)T_{t_{0}}^{f}(\tilde{y})=(y_{0},s_{0}^{\prime}). Since they are in the same atom, there exists an integer mm s.t.

Cξ2qm+1x0y0<Cξ2qm1/k.\frac{C_{\xi}}{2q_{m+1}}\leq|\!|x_{0}-y_{0}|\!|<\frac{C_{\xi}}{2q_{m}}\leq 1/k.

Applying 7.13, we can find the corresponding time TT. By definition, t0Ij0t_{0}\in I_{j_{0}} for some j0j_{0}. Let j1>j0j_{1}>j_{0} be the minimal natural number (if exists) such that the following interval l=j0j11Il\displaystyle\bigcup_{l=j_{0}}^{j_{1}-1}I_{l}, has a length larger than TT, and Ij1I_{j_{1}} is the interval when Tt(x~)T_{t}(\tilde{x}) and Tt(y~)T_{t}(\tilde{y}) are in the same atom. If such j1j_{1} exists, denote

I~1:=l=j0j11Il.\tilde{I}_{1}:=\displaystyle\bigcup_{l=j_{0}}^{j_{1}-1}I_{l}.

Otherwise, denote I~1=[t0,R)\tilde{I}_{1}=[t_{0},R). When j=01I~j(0,R)\displaystyle\cup_{j=0}^{1}\tilde{I}_{j}\neq(0,R), we know at the starting time t1t_{1} of Ij1I_{j_{1}}, Tt1(x~)T_{t_{1}}(\tilde{x}) and Tt1(y~)T_{t_{1}}(\tilde{y}) are in the same atom, we can then repeat the above procedure to find I~2\tilde{I}_{2}. Inductively, we can get

(0,R)=j=0LI~j(0,R)=\displaystyle\bigcup_{j=0}^{L}\tilde{I}_{j}

where L+1L+1 is the number of such intervals. By assumption, L1.L\geq 1.

If L=1L=1, for the first time when the orbit of x~\tilde{x} and y~\tilde{y} are in the same atom, let mm be the number s.t. the distance of the first coordinate is between Cξ2qm+1\frac{C_{\xi}}{2q_{m+1}} and Cξ2qm\frac{C_{\xi}}{2q_{m}}, then we can obtain R<TR<T. Otherwise, from 7.13, there will be a constant κ\kappa^{\prime} which is independent of RR and ϵ\epsilon, s.t. the total time when those two orbits are not in the same atom is larger than κR\kappa^{\prime}R. Since we can take ϵ<κ\epsilon<\kappa^{\prime} because κ\kappa^{\prime} is a fixed constant, this is a contradiction. Hence,

1Dqm1T1R.\frac{1}{Dq_{m}}\leq\frac{1}{T}\leq\frac{1}{R}.

Therefore, the distance of first coordinate if less than H2R\frac{H_{2}}{R} for some constant H2.H_{2}.

If L>1L>1, the sum of the length of I~j\tilde{I}_{j} where j=1,,L1j=1,\dots,L-1 is less than ϵR/κ\epsilon R/\kappa^{\prime}. Therefore, the last interval is of length larger than (1ϵ/κ)R(1-\epsilon/\kappa^{\prime})R. Using a similar argument, there exists a constant H3H_{3} satisfying the distance of the first coordinate is less than H3/RH_{3}/R. Finally, taking H=max{H1,H2,H3}H=\max\{H_{1},H_{2},H_{3}\}, we finish the proof. ∎

Proof of 7.15.

From the definition of n1n_{1} and n2n_{2}, we know

|f(n1)(x)f(n2)(y)||f(n1)(x)f(n2)(y)+s0s0|+|s0s0|G1|f^{(n_{1})}(x)-f^{(n_{2})}(y)|\leq|f^{(n_{1})}(x)-f^{(n_{2})}(y)+s_{0}-s_{0}^{\prime}|+|s_{0}-s_{0}^{\prime}|\leq G_{1}

for some constant G1.G_{1}. Since n12MRn_{1}\leq 2MR, we know there exists constant G2G_{2} s.t.

|f(n1)(x+n1α)f(n2)(y+n2α)|G2,|f^{(-n_{1})}(x+n_{1}\alpha)-f^{(-n_{2})}(y+n_{2}\alpha)|\leq G_{2},

this is because the interval [x+n1α,y+n2α)[x+n_{1}\alpha,y+n_{2}\alpha) only cross discontinuity points finitely many times. Using triangle inequality, cocycle identity, and f1Mf\geq\frac{1}{M}, we get

(33) G1\displaystyle G_{1} |f(n1)(x)f(n2)(y)|\displaystyle\geq|f^{(n_{1})}(x)-f^{(n_{2})}(y)|
=|f(n1)(x+n1α)f(n2)(y+n2α)+f(n2n1)(y+n2α)|\displaystyle=|f^{(-n_{1})}(x+n_{1}\alpha)-f^{(-n_{2})}(y+n_{2}\alpha)+f^{(n_{2}-n_{1})}(y+n_{2}\alpha)|
|f(n2n1)(y+n2α)|G2\displaystyle\geq|f^{(n_{2}-n_{1})}(y+n_{2}\alpha)|-G_{2}
|n2n1|MG2\displaystyle\geq\frac{|n_{2}-n_{1}|}{M}-G_{2}

Thus we can obtain |n2n1|(G2+G1)MG|n_{2}-n_{1}|\leq(G_{2}+G_{1})M\eqqcolon G. ∎

Proof of 7.16.

Choose δ=η2\delta=\frac{\eta}{2}. Let ZnZ_{n} be the subset

j=qnqn{x𝕋:x+jα(2qn1δ,2qn1δ)(2qn1δ+ξ,ξ+2qn1δ)}.\displaystyle\bigcup_{j=-q_{n}}^{q_{n}}\left\{x\in\mathbb{T}:x+j\alpha\in(-2q_{n}^{-1-\delta},2q_{n}^{-1-\delta})\cup(-2q_{n}^{-1-\delta}+\xi,\xi+2q_{n}^{-1-\delta})\right\}.

Since n=11qnδ<,\sum_{n=1}^{\infty}\frac{1}{q_{n}^{\delta}}<\infty, when NN is large enough, λ(n=NZn)<ϵ/4.\lambda\left(\displaystyle\cup_{n=N}^{\infty}Z_{n}\right)<\epsilon/4.

Denote X~=(n=NZn)c,\tilde{X}=\left(\displaystyle\cup_{n=N}^{\infty}Z_{n}\right)^{c}, then λf(X~f)>1ϵ/2\lambda^{f}\left(\tilde{X}^{f}\right)>1-\epsilon/2. Define

Ωκ={y~𝕋f:d(y~,𝒫)<κ}.\Omega_{\kappa}=\{\tilde{y}\in\mathbb{T}^{f}:d(\tilde{y},\partial\mathcal{P})<\kappa\}.

There exists 0<ϵ~<ϵ100Mk0<\tilde{\epsilon}<\frac{\epsilon}{100Mk}, such that λf(Ωe~)<ϵ/4.\lambda^{f}(\Omega_{\tilde{e}})<\epsilon/4. There exists some integer ll such that qlR<ql+1,q_{l}\leq R<q_{l+1}, choose RR to be a large value such that l2Nl\geq 2N and ql1<e~/100M.q_{l}^{-1}<\tilde{e}/100M. Take A=X~Ωe~.A=\tilde{X}\cap\Omega_{\tilde{e}}. Note that for x~=(x,s)A\tilde{x}=(x,s)\in A the Hamming ball BR(x~,ϵ)B^{R}(\tilde{x},\epsilon) centered at x~\tilde{x} containing

{(y,t)𝕋f:y(1ql+11+δ+x,x+1ql+11+δ),t(se~,s+e~)}.\left\{(y,t)\in\mathbb{T}^{f}:y\in\left(-\frac{1}{q_{l+1}^{1+\delta}}+x,x+\frac{1}{q_{l+1}^{1+\delta}}\right),t\in(s-\tilde{e},s+\tilde{e})\right\}.

Hence, STf,𝒫(ϵ,R)CR1+δlog2+2δRS_{T^{f},\mathcal{P}}(\epsilon,R)\leq CR^{1+\delta}\log^{2+2\delta}R for some constant CC independent of RR. It follows that δTf,𝒫,χS(ϵ)=0\delta_{T^{f},\mathcal{P},\chi}^{S}(\epsilon)=0 when χ=1+η.\chi=1+\eta.

References

  • [AB98] Pascal Alessandri and Valérie Berthé. Three distance theorems and combinatorics on words. L’Enseignement Mathématique, 44:103–132, 1998.
  • [BKW23a] Shilpak Banerjee, Philipp Kunde, and Daren Wei. Slow entropy for some Anosov-Katok diffeomorphisms. Discrete and Continuos Dynamical Systems, 43(1):209–255, 2023.
  • [BKW23b] Shilpak Banerjee, Philipp Kunde, and Daren Wei. Slow entropy of some combinatorial constructions. Nonlinearity, 36(6):2923–2974, 2023.
  • [Bow71] Rufus Bowen. Entropy for group endomorphisms and homogeneous spaces. Transactions of the American Mathematical Society, 153:401–414, 1971.
  • [CCM13] Jonathan Chaika, Yitwah Cheung, and Howard Masur. Winning games for bounded geodesics in moduli spaces of quadratic differentials. Journal of Modern Dynamics, 7(3):395–427, 2013.
  • [CJKS22] Van Cyr, Aimee Johnson, Bryna Kra, and Ayse Sahin. The complexity threshold for the emergence of Kakutani inequivalence. Israel J. Math., 251(1):271–300, 2022.
  • [CK20] Van Cyr and Bryna Kra. The automorphism group of a shift of slow growth is amenable. Ergodic Theory Dynam. Systems, 40(7):1788–1804, 2020.
  • [CP23] Darren Creutz and Ronnie Pavlov. Low complexity subshifts have discrete spectrum. Forum of Mathematics, Sigma, 11:e96, 2023.
  • [DOP22] Andrew Dykstra, Nicholas Ormes, and Ronnie Pavlov. Subsystems of transitive subshifts with linear complexity. Ergodic Theory and Dynamical Systems, 42(6):1967–1993, 2022.
  • [Fal85] Kenneth John Falconer. The geometry of fractal sets. Cambridge University Press, USA, 1985.
  • [Fer97] Sébastien Ferenczi. Measure-theoretic complexity of ergodic systems. Israel Journal of Mathematics, 100(1):189–207, 1997.
  • [FLL07] Krzysztof Fraczek, Mariusz Lemańczyk, and Emmanuel Lesigne. Mild mixing property for special flows under piecewise constant functions, 2007.
  • [Fog02] N. Pytheas Fogg. Substitutions in Dynamics, Arithmetics and Combinatorics. Lecture Notes in Mathematics. Springer Berlin Heidelberg, 2002.
  • [Goo69] L Wayne Goodwyn. Topological entropy bounds measure-theoretic entropy. Proceedings of the American Mathematical Society, 23(3):679–688, 1969.
  • [JK69] Konrad Jacobs and Michael Keane. 010-1-sequences of Toeplitz type. Z. Wahrscheinlichkeitstheorie und Verw. Gebiete, 13:123–131, 1969.
  • [Kan18] Adam Kanigowski. Slow entropy for some smooth flows on surfaces. Israel Journal of Mathematics, 226:535–577, 2018.
  • [Kat80] Anatole Katok. Lyapunov exponents, entropy and periodic orbits for diffeomorphisms. Publications Mathématiques de l’IHÉS, 51:137–173, 1980.
  • [Kat82] Anatole Katok. Entropy and closed geodesies. Ergodic Theory and Dynamical Systems, 2(3–4):339–365, 1982.
  • [Kat23] Anatole Katok. Fifty years of entropy in dynamics: 1958–2007. In Frontiers in Entropy Across the Disciplines: Panorama of Entropy: Theory, Computation, and Applications, pages 353–404. World Scientific, 2023.
  • [Kea75] Michael Keane. Interval exchange transformations. Mathematische Zeitschrift, 141:25–32, 1975.
  • [KH95] Anatole Katok and Boris Hasselblatt. Introduction to the modern theory of dynamical systems. Number 54 in Encyclopedia of Mathematics and its Applications. Cambridge university press, 1995.
  • [Khi97] A. Ya. Khinchin. Continued Fractions. Dover books on mathematics. Dover Publications, 1997.
  • [KT97] Anatole Katok and Jean-Paul Thouvenot. Slow entropy type invariants and smooth realization of commuting measure-preserving transformations. Annales de l’I.H.P. Probabilités et statistiques, 33(3):323–338, 1997.
  • [KVW19] Adam Kanigowski, Kurt Vinhage, and Daren Wei. Slow entropy of some parabolic flows. Communications in Mathematical Physics, 370:449–474, 2019.
  • [PS23a] Ronnie Pavlov and Scott Schmieding. Local finiteness and automorphism groups of low complexity subshifts. Ergodic Theory Dynam. Systems, 43(6):1980–2001, 2023.
  • [PS23b] Ronnie Pavlov and Scott Schmieding. On the structure of generic subshifts. Nonlinearity, 36(9):4904, aug 2023.
  • [Rob19] Donald Robertson. Mild mixing of certain interval-exchange transformations. Ergodic Theory and Dynamical Systems, 39(1):248–256, 2019.
  • [Sch1] Wolfgang M. Schmidt. On badly approximable numbers and certain games. Transactions of the American Mathematical Society, 123(1966):178–199, 1.
  • [Tse09] Jimmy Tseng. Badly approximable affine forms and schmidt games. Journal of Number Theory, 129(12):3020–3025, 2009.
  • [Vee87] William A. Veech. Boshernitzan’s criterion for unique ergodicity of an interval exchange transformation. Ergodic Theory and Dynamical Systems, 7(1):149–153, 1987.
  • [Via06] Marcelo Viana. Ergodic theory of interval exchange maps. Revista Matemática Complutense, 19(1):7–100, 2006.
  • [VO16] Marcelo Viana and Krerley Oliveira. Foundations of Ergodic Theory. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2016.
  • [Yoc10] Jean-Christophe Yoccoz. Interval exchange maps and translation surfaces. In Homogeneous Flows, Moduli Spaces, and Arithmetic, volume 10 of Clay mathematics proceedings, pages 1–69. American Mathematical Society, 2010.