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

\AtBeginEnvironment

proof

Multiple partial rigidity rates in low complexity subshifts

Tristán Radić Department of mathematics, Northwestern University, 2033 Sheridan Rd, Evanston, IL, United States of America [email protected]
(Date: January 28, 2025)
Abstract.

Partial rigidity is a quantitative notion of recurrence and provides a global obstruction which prevents the system from being strongly mixing. A dynamical system (X,𝒳,μ,T)(X,{\mathcal{X}},\mu,T) is partially rigid if there is a constant δ>0\delta>0 and sequence (nk)k(n_{k})_{k\in{\mathbb{N}}} such that lim infkμ(ATnkA)δμ(A)\displaystyle\liminf_{k\to\infty}\mu(A\cap T^{n_{k}}A)\geq\delta\mu(A) for every A𝒳A\in{\mathcal{X}}, and the partial rigidity rate is the largest δ\delta achieved over all sequences. For every integer d1d\geq 1, via an explicit construction, we prove the existence of a minimal subshift (X,S)(X,S) with dd ergodic measures having distinct partial rigidity rates. The systems built are 𝒮{\mathcal{S}}-adic subshifts of finite alphabetic rank that have non-superlinear word complexity and, in particular, have zero entropy.

Key words and phrases:
partial rigidity, partial rigidity rate, S-adic subshifts
2020 Mathematics Subject Classification:
Primary: 37A05; Secondary: 37B10,37B02
Northwestern University

1. Introduction

For measure preserving systems, partial rigidity quantitatively captures recurrence along a particular trajectory. Roughly speaking, this measurement ensures that at least a proportion δ(0,1]\delta\in(0,1] of any measurable set AA returns to AA along some sequence of iterates. The notion was introduced by Friedman [18] and defined formally by King [24]. An important property of partially rigid systems is that, besides the trivial system, they are not strongly mixing. Although the converse does not hold, many common examples of non-mixing systems are partially rigid, see for example [11, 23, 8, 6, 10, 9, 22].

To be more precise, a measure-preserving systems (X,𝒳,μ,T)(X,{\mathcal{X}},\mu,T) is partially rigid if there exists δ>0\delta>0 and an increasing sequence (nk)k(n_{k})_{k\in{\mathbb{N}}} of integers such that

lim infkμ(ATnkA)δμ(A)\liminf_{k\to\infty}\mu(A\cap T^{-n_{k}}A)\geq\delta\mu(A) (1)

for every measurable set AA. A constant δ>0\delta>0 and a sequence (nk)k(n_{k})_{k\in{\mathbb{N}}} satisfying (1) are respectively called a constant of partial rigidity and a partial rigidity sequence.

Once we know that a system is partially rigid, computing the largest value of δ\delta provides valuable information on how strongly the system exhibits recurrent behavior. In particular, as was remarked by King in 1988 [24, Proposition 1.13], this constant is invariant under measurable isomorphisms and increases under factor maps. We call this constant the partial rigidity rate, we denote it δμ\delta_{\mu} and it is given by

δμ=sup{δ>0δ is a partial rigidity constant for some sequence (nk)k},\delta_{\mu}=\sup\{\delta>0\mid\delta\text{ is a partial rigidity constant for some sequence }(n_{k})_{k\in{\mathbb{N}}}\},

with the convention that δμ=0\delta_{\mu}=0 whenever the system is not partially rigid. There are only limited partially rigid systems for which that constant is known. One major case is rigid systems, that is when δμ=1\delta_{\mu}=1. Such systems have been well studied after Furstenberg and Weiss introduced them in [20], see for instance [3, 7, 14, 17, 21]. The only non-rigid examples for which the partial rigidity rates are calculated are some specific substitution subshifts studied in [13, Section 7].

Since minimal substitution subshifts are uniquely ergodic, it is natural to ask whether it is possible to construct a minimal, low-complexity system with more than one ergodic measure and distinct partial rigidity rates. Via an explicit construction, we fully resolve this question. More precisely, we show

Theorem 1.1.

For any natural number d2d\geq 2, there exists a minimal subshift with non-superlinear complexity that has dd distinct ergodic measures μ0,,μd1\mu_{0},\ldots,\mu_{d-1} for which the partial rigidity rates 0<δμ0<<δμd1<10<\delta_{\mu_{0}}<\ldots<\delta_{\mu_{d-1}}<1 are also distinct.

Moreover, the partial rigidity sequence (nk)k(n_{k})_{k\in{\mathbb{N}}} associated to each δμi\delta_{\mu_{i}} is the same for all i{0,,d1}i\in\{0,\ldots,d-1\}.

Constructing measures all of which share the same partial rigidity sequence is a key aspect because, in general, an invariant measure can be partially rigid for two different sequences (nk)k(n_{k})_{k\in{\mathbb{N}}} and (nk)k(n^{\prime}_{k})_{k\in{\mathbb{N}}} and have different partial rigidity constants δ\delta and δ\delta^{\prime} for each sequence. For instance, in [13, Theorem 7.1] it is proven that for the Thue-Morse substitution subshift equipped with its unique invariant measure ν\nu, δν=2/3\delta_{\nu}=2/3 and its associated partial rigidity sequence is (32n)n(3\cdot 2^{n})_{n\in{\mathbb{N}}}. Using a similar proof, the largest constant of partial rigidity for the sequence (2n)n(2^{n})_{n\in{\mathbb{N}}} is 1/31/3. In contrast, the discrepancy between the values in Theorem 1.1 is not due to quantifying along a different trajectory, but rather that for each measure the returning mass takes on a different value.

The system constructed to prove Theorem 1.1 is an 𝒮{\mathcal{S}}-adic subshift, that is a symbolic system formed as a limit of morphisms 𝝈=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A_{n+1}^{*}\to A_{n}^{*})_{n\in{\mathbb{N}}} (see Section 2 for the precise definitions). We introduce a novel technique that allows us to build minimal 𝒮{\mathcal{S}}-adic subshift with dd ergodic measures, where each ergodic measure “behaves like” a substitution subshift for which we already know its partial rigidity rate. The idea is that the measures of the cylinder sets “closely approximate” the values assigned by the unique invariant measure of the substitution subshift that is “imitating”. For the precise statement, see Theorem 3.2. This gluing technique is of interest on its own, as it gives a general way for controlling distinct ergodic measures in some specific 𝒮{\mathcal{S}}-adic subshift.

For each ergodic measure μi\mu_{i}, with i{0,,d1}i\in\{0,\ldots,d-1\}, the gluing technique gives us a lower bound for the partial rigidity rate (see Corollary 3.4). The lower bound corresponds to the partial rigidity rate associated to the uniquely ergodic system that the measure μi\mu_{i} is “imitating”. In Section 4, we restrict to a specific example in which that lower bound is achieved. In that section, we prove that the number of morphisms needed for building the 𝒮{\mathcal{S}}-adic subshift can be reduced to three. Combining results from Sections 3 and 4, we complete the proof of Theorem 1.1. An extended version of the theorem that includes the values of δμi\delta_{\mu_{i}} for i{0,,d1}i\in\{0,\ldots,d-1\} and the partial rigidity sequence is stated in Theorem 4.8.

Acknowledgments. The author thanks B. Kra for her careful reading and helpful suggestions on the earlier versions of this paper. He is also grateful to A. Maass and S. Donoso for their insights in the early stages of this project, and extends his thanks to F. Arbulu for providing valuable references. Special thanks to S. Petite, who, during the author’s first visit to the UPJV in Amiens, asked whether an example with multiple partial rigidity rates, such as the one described in this paper, could be constructed.

2. Preliminaries and notation

2.1. Topological and symbolic dynamical systems

In this paper, a topological dynamical system is a pair (X,T)(X,T), where XX is a compact metric space and T:XXT\colon X\to X is a homeomorphism. We say that (X,T)(X,T) is minimal if for every xXx\in X the orbit {Tnx:n}\{T^{n}x:n\in{\mathbb{Z}}\} is dense in XX. A continuous and onto map π:X1X2\pi\colon X_{1}\to X_{2} between two topological dynamical systems (X1,T1)(X_{1},T_{1}) and (X2,T2)(X_{2},T_{2}) is a factor map if for every xX1x\in X_{1}, T2π(x)=πT1(x)T_{2}\circ\pi(x)=\pi\circ T_{1}(x).

We focus on a special family of topological dynamical system, symbolic systems. To define them, let AA be a finite set that we call alphabet. The elements of AA are called letters. For \ell\in{\mathbb{N}}, the set of concatenations of \ell letters is denoted by AA^{\ell} and w=w1wAw=w_{1}\ldots w_{\ell}\in A^{\ell} is a word of length \ell. The length of a word ww is denoted by |w||w|. We set A=nAA^{*}=\bigcup_{n\in{\mathbb{N}}}A^{\ell} and by convention, A0={ε}A^{0}=\{\varepsilon\} where ε\varepsilon is the empty word.

For a word w=w1ww=w_{1}\ldots w_{\ell} and two integers 1i<j1\leq i<j\leq\ell, we write w[i,j+1)=w[i,j]=wiwjw_{[i,j+1)}=w_{[i,j]}=w_{i}\ldots w_{j}. We say that uu appears or occurs in ww if there is an index 1i|w|1\leq i\leq|w| such that u=w[i,i+|u|)u=w_{[i,i+|u|)} and we denote this by uwu\sqsubseteq w. The index ii is an occurrence of uu in ww and |w|u|w|_{u} denotes the number of (possibly overleaped) occurrences of uu in ww. We also write freq(u,w)=|w|u|w|{\rm freq}(u,w)=\frac{|w|_{u}}{|w|}, the frequency of uu in ww.

Let AA^{{\mathbb{Z}}} be the set of two-sided sequences (xn)n(x_{n})_{n\in{\mathbb{Z}}}, where xnAx_{n}\in A for all nn\in{\mathbb{Z}}. Like for finite words, for xAx\in A^{{\mathbb{Z}}} and <i<j<-\infty<i<j<\infty we write x[i,j]=x[i,j+1)x_{[i,j]}=x_{[i,j+1)} for the finite word given by xixi+1xjx_{i}x_{i+1}\ldots x_{j}. The set AA^{{\mathbb{Z}}} endowed with the product topology is a compact and metrizable space. The shift map S:AAS\colon A^{{\mathbb{Z}}}\to A^{{\mathbb{Z}}} is the homeomorphism defined by S((xn)n)=(xn+1)nS((x_{n})_{n\in{\mathbb{Z}}})=(x_{n+1})_{n\in{\mathbb{Z}}}. Notice that, the collection of cylinder sets {Sj[w]:wA,j}\{S^{j}[w]\colon w\in A^{*},j\in{\mathbb{Z}}\} where [w]={xA:x[0,|w|)=w}[w]=\{x\in A^{{\mathbb{Z}}}\colon x_{[0,|w|)}=w\}, is a basis of clopen subsets for the topology of AA^{{\mathbb{Z}}}.

A subshift is a topological dynamical system (X,S)(X,S), where XX is a closed and SS-invariant subset of AA^{{\mathbb{Z}}}. In this case the topology is also given by cylinder sets, denoted [w]X=[w]X[w]_{X}=[w]\cap X, but when there is no ambiguity we just write [w][w]. Given an element xXx\in X, the language (x){\mathcal{L}}(x) is the set of all words appearing in xx and (X)=xX(x){\mathcal{L}}(X)=\bigcup_{x\in X}{\mathcal{L}}(x). Notice that [w]X[w]_{X}\neq\emptyset if and only if w(X)w\in{\mathcal{L}}(X). Also, (X,S)(X,S) is minimal if and only if (X)=(x){\mathcal{L}}(X)={\mathcal{L}}(x) for all xXx\in X.

Let AA and BB be finite alphabets and σ:AB\sigma\colon A^{*}\to B^{*} be a morphism for the concatenation, that is σ(uw)=σ(u)σ(w)\sigma(uw)=\sigma(u)\sigma(w) for all u,wAu,w\in A^{*}. A morphism σ:AB\sigma\colon A^{*}\to B^{*} is completely determined by the values of σ(a)\sigma(a) for every letter aAa\in A. We only consider non-erasing morphisms, that is σ(a)ε\sigma(a)\neq\varepsilon for every aAa\in A, where ε\varepsilon is the empty word in BB^{*}. A morphism σ:AA\sigma\colon A^{*}\to A^{*} is called a substitution if for every aAa\in A, limn|σn(a)|=\displaystyle\lim_{n\to\infty}|\sigma^{n}(a)|=\infty.

A directive sequence 𝝈=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A^{*}_{n+1}\to A^{*}_{n})_{n\in{\mathbb{N}}} is a sequence of (non-erasing) morphisms. Given a directive sequence 𝝈\boldsymbol{\sigma} and nn\in{\mathbb{N}}, define (n)(𝝈){\mathcal{L}}^{(n)}(\boldsymbol{\sigma}), the language of level nn associated to 𝝈\boldsymbol{\sigma} by

(n)(𝝈)={wAn:wσ[n,N)(a) for some aAN and N>n}{\mathcal{L}}^{(n)}(\boldsymbol{\sigma})=\{w\in A_{n}^{*}:w\sqsubseteq\sigma_{[n,N)}(a)\text{ for some }a\in A_{N}\text{ and }N>n\}

where σ[n,N)=σnσn+1σN1\sigma_{[n,N)}=\sigma_{n}\circ\sigma_{n+1}\circ\ldots\circ\sigma_{N-1}. For nn\in{\mathbb{N}}, we define X𝝈(n)X_{\boldsymbol{\sigma}}^{(n)}, the nn-th level subshift generated by 𝝈\boldsymbol{\sigma}, as the set of elements xAnx\in A_{n}^{{\mathbb{Z}}} such that (x)(n)(𝝈){\mathcal{L}}(x)\subseteq{\mathcal{L}}^{(n)}(\boldsymbol{\sigma}). For the special case n=0n=0, we write X𝝈X_{\boldsymbol{\sigma}} instead of X𝝈(0)X_{\boldsymbol{\sigma}}^{(0)} and we call it the 𝒮{\mathcal{S}}-adic subshift generated by 𝝈\boldsymbol{\sigma}.

A morphism σ:AB\sigma\colon A^{*}\to B^{*} has a composition matrix M(σ)B×AM(\sigma)\in{\mathbb{N}}^{B\times A} given by M(σ)b,a=|σ(a)|bM(\sigma)_{b,a}=|\sigma(a)|_{b} for all bBb\in B and aAa\in A. If τ:BC\tau\colon B^{*}\to C^{*} is another morphism, then M(τσ)=M(τ)M(σ)M(\tau\circ\sigma)=M(\tau)M(\sigma). Therefore, for a substitution, σ:AA\sigma\colon A^{*}\to A^{*}, M(σ2)=M(σ)2M(\sigma^{2})=M(\sigma)^{2}. We say that 𝝈\boldsymbol{\sigma} is primitive if for every nn\in{\mathbb{N}} there exists k1k\geq 1 such that the matrix M(σ[n,n+k])=M(σn)M(σn+1)M(σn+k)M(\sigma_{[n,n+k]})=M(\sigma_{n})M(\sigma_{n+1})\cdots M(\sigma_{n+k}) has only positive entries. When 𝝈\boldsymbol{\sigma} is primitive, then for every nn\in{\mathbb{N}} (X𝝈(n),S)(X_{\boldsymbol{\sigma}}^{(n)},S) is minimal and (X𝝈(n))=(n)(𝝈){\mathcal{L}}(X^{(n)}_{\boldsymbol{\sigma}})={\mathcal{L}}^{(n)}(\boldsymbol{\sigma}).

When 𝝈\boldsymbol{\sigma} is the constant directive sequence σn=σ\sigma_{n}=\sigma for all nn\in{\mathbb{N}}, where σ:AA\sigma\colon A^{*}\to A^{*} is a substitution, then X𝝈X_{\boldsymbol{\sigma}} is denoted XσX_{\sigma} and it is called substitution subshift. Similarly (𝝈){\mathcal{L}}(\boldsymbol{\sigma}) is denoted (σ){\mathcal{L}}(\sigma). Also if in that context 𝝈\boldsymbol{\sigma} is primitive, we say that the substitution σ\sigma itself is primitive, which is equivalent to saying that the composition matrix M(σ)M(\sigma) is primitive. We also say that the substitution σ\sigma is positive if M(σ)M(\sigma) only have positive entries. By definition, every positive substitution is also primitive.

A morphism σ:AB\sigma\colon A^{*}\to B^{*} has constant length if there exists a number 1\ell\geq 1 such that |σ(a)|=|\sigma(a)|=\ell for all aAa\in A. In this case, we write |σ|=|\sigma|=\ell. More generally, a directive sequence 𝝈=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A^{*}_{n+1}\to A^{*}_{n})_{n\in{\mathbb{N}}} is of constant-length if each morphism σn\sigma_{n} is of constant length. Notice that we do not require that |σn|=|σm||\sigma_{n}|=|\sigma_{m}| for distinct n,mn,m\in{\mathbb{N}}.

We define the alphabet rank ARAR of 𝝈=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A^{*}_{n+1}\to A^{*}_{n})_{n\in{\mathbb{N}}} as AR(𝝈)=lim infn|An|\displaystyle AR(\boldsymbol{\sigma})=\liminf_{n\to\infty}|A_{n}|. Having finite alphabet rank has many consequences, for instance if AR(𝝈)<AR(\boldsymbol{\sigma})<\infty then X𝝈X_{\boldsymbol{\sigma}} has zero topological entropy.

For a general subshift (X,S)(X,S), let pX:p_{X}\colon{\mathbb{N}}\to{\mathbb{N}} denote the word complexity function of XX given by pX(n)=|n(X)|p_{X}(n)=|{\mathcal{L}}_{n}(X)| for all nn\in{\mathbb{N}}. Here n(X)={w(X):|w|=n}{\mathcal{L}}_{n}(X)=\{w\in{\mathcal{L}}(X)\colon|w|=n\}. If lim infnpX(n)n=\displaystyle\liminf_{n\to\infty}\frac{p_{X}(n)}{n}=\infty we say that XX has superlinear complexity. Otherwise we say XX has non-superlinear complexity.

We say that a primitive substitution τ:AA\tau\colon A^{*}\to A^{*} is right prolongable (resp. left prolongable) on uAu\in A^{*} if τ(u)\tau(u) starts (resp. ends) with uu. If, for every letter aAa\in A, τ:AA\tau\colon A^{*}\to A^{*} is left and right prolongable on aa, then τ:AA\tau\colon A^{*}\to A^{*} is said to be prolongable. A word w=w1w𝒜w=w_{1}\ldots w_{\ell}\in{\mathcal{A}}^{*} is complete if 2\ell\geq 2 and w1=ww_{1}=w_{\ell}. Notice that if a substitution τ:AA\tau\colon A^{*}\to A^{*} is primitive and prolongable, then τ(a)\tau(a) is a complete word for every aAa\in A. If WW is a set of words, then we denote

𝒞W={wW:|w|2,w1=w|w|}.{\mathcal{C}}W=\{w\in W\colon|w|\geq 2,w_{1}=w_{|w|}\}. (2)

the set of complete words in WW. In particular, for k2k\geq 2, 𝒞Ak{\mathcal{C}}A^{k} is the set of complete words of length kk with letters in AA, for example, 𝒞{a,b}3={aaa,aba,bab,bbb}{\mathcal{C}}\{a,b\}^{3}=\{aaa,aba,bab,bbb\}.

Finally, when the alphabet has two letters 𝒜={a,b}{\mathcal{A}}=\{a,b\}, the complement of a word w=w1w𝒜w=w_{1}\ldots w_{\ell}\in{\mathcal{A}}^{*} denoted w¯\overline{w} is given by w¯1w¯\overline{w}_{1}\ldots\overline{w}_{\ell} where a¯=b\overline{a}=b and b¯=a\overline{b}=a. A morphism τ:𝒜𝒜\tau\colon{\mathcal{A}}^{*}\to{\mathcal{A}}^{*} is said to be a mirror morphism if τ(w¯)=τ(w)¯\tau(\overline{w})=\overline{\tau(w)} (the name is taken from [25, Chapter 8.2] with a slight modification).

2.2. Invariant measures

A measure preserving system is a tuple (X,𝒳,μ,T)(X,\mathcal{X},\mu,T), where (X,𝒳,μ)(X,\mathcal{X},\mu) is a probability space and T:XXT\colon X\to X is a measurable and measure preserving transformation. That is, T1A𝒳T^{-1}A\in\mathcal{X} and μ(T1A)=μ(A)\mu(T^{-1}A)=\mu(A) for all A𝒳A\in{\mathcal{X}}, and we say that μ\mu is TT-invariant. An invariant measure μ\mu is said to be ergodic if whenever AXA\subseteq X is measurable and μ(AΔT1A)=0\mu(A\Delta T^{-1}A)=0, then μ(A)=0\mu(A)=0 or 11.

Given a topological dynamical system (X,T)(X,T), we denote (X,T){\mathcal{M}}(X,T) (resp. (X,T){\mathcal{E}}(X,T)) the set of Borel TT-invariant probability measures (resp. the set of ergodic probability measures). For any topological dynamical system, (X,T){\mathcal{E}}(X,T) is nonempty and when (X,T)={μ}{\mathcal{E}}(X,T)=\{\mu\} the system is said to be uniquely ergodic.

If (X,S)(X,S) is a subshift over an alphabet AA, then any invariant measure μ(X,S)\mu\in{\mathcal{M}}(X,S) is uniquely determined by the values of μ([w]X)\mu([w]_{X}) for w(X)w\in{\mathcal{L}}(X). Since XAX\subset A^{{\mathbb{Z}}}, μ(X,S)\mu\in{\mathcal{M}}(X,S) can be extended to AA^{{\mathbb{Z}}} by μ~(B)=μ(BX)\tilde{\mu}(B)=\mu(B\cap X) for all BAB\subset A^{{\mathbb{Z}}} measurable. In particular, μ~([w])=μ([w]X)\tilde{\mu}([w])=\mu([w]_{X}) for all wAw\in A^{*}. We use this extension many times, making a slight abuse of notation and not distinguishing between μ\mu and μ~\tilde{\mu}. Moreover, for wAw\in A^{*}, since there is no ambiguity with the value of the cylinder set we write μ(w)\mu(w) instead of μ([w])\mu([w]). This can also be done when we deal with two alphabets ABA\subset B, every invariant measure μ\mu in AA^{{\mathbb{Z}}} can be extended to an invariant measure in BB^{{\mathbb{Z}}}, where in particular, μ(b)=0\mu(b)=0 for all bB\Ab\in B\backslash A.

A sequence of non-empty subsets of the integers, 𝚽=(Φn)n\boldsymbol{\Phi}=(\Phi_{n})_{n\in{\mathbb{N}}} is a Følner sequence if for all tt\in{\mathbb{Z}}, limn|ΦnΔ(Φn+t)||Φn|=0\displaystyle\lim_{n\to\infty}\frac{|\Phi_{n}\Delta(\Phi_{n}+t)|}{|\Phi_{n}|}=0. Let (X,T)(X,T) be a topological system and let μ\mu be an invariant measur, an element xXx\in X is said to be generic along 𝚽\boldsymbol{\Phi} if for every continuous function fC(X)f\in C(X)

limn1|Φn|kΦnf(Tx)=Xf𝑑μ.\lim_{n\to\infty}\frac{1}{|\Phi_{n}|}\sum_{k\in\Phi_{n}}f(Tx)=\int_{X}fd\mu.

Every point in a minimal system is generic for some Følner sequence 𝚽\boldsymbol{\Phi}, more precisely

Proposition 2.1.

[19, Proposition 3.9] Let (X,T)(X,T) be a minimal system and μ\mu an ergodic measure. Then for every xXx\in X there exists sequences (mn)n,(mn)n(m_{n})_{n\in{\mathbb{N}}},(m^{\prime}_{n})_{n\in{\mathbb{N}}}\subset{\mathbb{N}} such that mn<mnm_{n}<m^{\prime}_{n} for every nn\in{\mathbb{N}} and limnmnmn=\displaystyle\lim_{n\to\infty}m^{\prime}_{n}-m_{n}=\infty such that xx is generic along 𝚽=({mn,,mn})n\boldsymbol{\Phi}=(\{m_{n},\ldots,m^{\prime}_{n}\})_{n\in{\mathbb{N}}}.

In particular, for an 𝒮{\mathcal{S}}-adic subshift with primitive directive sequence 𝝈=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A_{n+1}^{*}\to A_{n}^{*})_{n\in{\mathbb{N}}}, when the infinite word 𝒘=limnσ0σ1σn1(an)\boldsymbol{w}=\displaystyle\lim_{n\to\infty}\sigma_{0}\circ\sigma_{1}\circ\cdots\circ\sigma_{n-1}(a_{n}) is well-defined then every invariant measure μ(X𝝈,S)\mu\in{\mathcal{M}}(X_{\boldsymbol{\sigma}},S) is given by

μ(u)=limn|𝒘[mn,mn]|umnmn+1=limnfreq(u,𝒘[mn,mn])u(X𝝈),\mu(u)=\lim_{n\to\infty}\frac{|\boldsymbol{w}_{[m_{n},m^{\prime}_{n}]}|_{u}}{m^{\prime}_{n}-m_{n}+1}=\lim_{n\to\infty}{\rm freq}(u,\boldsymbol{w}_{[m_{n},m^{\prime}_{n}]})\quad\forall u\in{\mathcal{L}}(X_{\boldsymbol{\sigma}}), (3)

for some (mn)n,(mn)n(m_{n})_{n\in{\mathbb{N}}},(m^{\prime}_{n})_{n\in{\mathbb{N}}}\subset{\mathbb{N}} as before. Notice that such infinite word 𝒘\boldsymbol{w} is well-defined for example when An=AA_{n}=A, an=aa_{n}=a and σn:AA\sigma_{n}\colon A^{*}\to A^{*} is prolongable, for all nn\in{\mathbb{N}}, where AA and aAa\in A are a fixed alphabet and letter respectively. Those are the condition for the construction of the system announced in Theorem 1.1.

We remark that for a primitive substitution, σ:AA\sigma\colon A^{*}\to A^{*} the substitution subshift (Xσ,S)(X_{\sigma},S) is uniquely ergodic and the invariant measure is given by any limit of the form (3).

2.3. Partial rigidity rate for 𝒮{\mathcal{S}}-adic subshifts

Every 𝒮{\mathcal{S}}-adic subshift can be endowed with a natural sequence of Kakutani-Rokhlin partitions see for instance [4, Lemma 6.3], [16, Chapter 6] or [13, section 5]. To do this appropriately, one requires recognizability of the directive sequence 𝝈=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A_{n+1}^{*}\to A_{n}^{*})_{n\in{\mathbb{N}}}, where we are using the term recognizable as defined in [4]. We do not define it here, but if every morphism σn:An+1An\sigma_{n}\colon A_{n+1}^{*}\to A_{n}^{*} is left-permutative, that is the first letter of σn(a)\sigma_{n}(a) is distinct from the first letter of σn(a)\sigma_{n}(a^{\prime}) for all aaa\neq a^{\prime} in AnA_{n}, then the directive sequence is recognizable. In this case we say that the directive sequence 𝝈\boldsymbol{\sigma} itself is left-permutative. If τ:AA\tau\colon A^{*}\to A^{*} is prolongable, then it is left-permutative.

Once we use the Kakutani-Rokhlin partition structure, X𝝈(n)X^{(n)}_{\boldsymbol{\sigma}} can be identified as the induced system in the nn-th basis and for every invariant measure μ\mu^{\prime} in X𝝈(n)X^{(n)}_{\boldsymbol{\sigma}}, there is an invariant measure μ\mu in X𝝈X_{\boldsymbol{\sigma}} such that μ\mu^{\prime} is the induced measure of μ\mu in X𝝈(n)X^{(n)}_{\boldsymbol{\sigma}}. We write μ=μ(n)\mu^{\prime}=\mu^{(n)} and this correspondence is one-to-one. This is a crucial fact for computing the partial rigidity rate for an 𝒮{\mathcal{S}}-adic subshift, for instance, if 𝝈\boldsymbol{\sigma} is a directive sequence of constant-length, δμ=δμ(n)\delta_{\mu}=\delta_{\mu^{(n)}} for all μ(X𝝈,S)\mu\in{\mathcal{E}}(X_{\boldsymbol{\sigma}},S) and n1n\geq 1 (see Theorem 2.3). Since the aim of this paper is building a specific example, we give a way to characterize μ(n)\mu^{(n)} for a more restricted family of 𝒮{\mathcal{S}}-adic subshift that allows us to carry out computations.

In what follows, we restrict the analysis to less general directive sequences 𝝈\boldsymbol{\sigma}. To do so, from now on, 𝒜{\mathcal{A}} always denotes the two letters alphabet {a,b}\{a,b\}. Likewise, for d2d\geq 2, 𝒜i={ai,bi}{\mathcal{A}}_{i}=\{a_{i},b_{i}\} for i{0,,d1}i\in\{0,\ldots,d-1\} and Λd=i=0d1𝒜i\Lambda_{d}=\bigcup_{i=0}^{d-1}{\mathcal{A}}_{i}.

We cite a simplified version of [5, Theorem 4.9], the original proposition is stated for Bratelli-Vershik transformations, but under recognizability, it can be stated for 𝒮{\mathcal{S}}-adic subshifts, see [4, Theorem 6.5].

Lemma 2.2.

Let 𝛔=(σn:ΛdΛd)n1\boldsymbol{\sigma}=(\sigma_{n}\colon\Lambda_{d}^{*}\to\Lambda_{d}^{*})_{n\geq 1} be a recognizable constant-length and primitive directive sequence, such that for all i{0,,d1}i\in\{0,\ldots,d-1\},

limn1|σn|ji|σn(ai)|aj+|σn(ai)|bj+|σn(bi)|aj+|σn(bi)|bj=0\lim_{n\to\infty}\frac{1}{|\sigma_{n}|}\sum_{j\neq i}|\sigma_{n}(a_{i})|_{a_{j}}+|\sigma_{n}(a_{i})|_{b_{j}}+|\sigma_{n}(b_{i})|_{a_{j}}+|\sigma_{n}(b_{i})|_{b_{j}}=0 (4)
n1(1minc𝒜i1|σn|(|σn(c)|ai+|σn(c)|bi))<\sum_{n\geq 1}\left(1-\min_{c\in{\mathcal{A}}_{i}}\frac{1}{|\sigma_{n}|}\left(|\sigma_{n}(c)|_{a_{i}}+|\sigma_{n}(c)|_{b_{i}}\right)\right)<\infty (5)
and limn1|σn|maxc,c𝒜idΛd||σn(c)|d|σn(c)|d|=0.\text{and }\quad\lim_{n\to\infty}\frac{1}{|\sigma_{n}|}\max_{c,c^{\prime}\in{\mathcal{A}}_{i}}\sum_{d\in\Lambda_{d}}||\sigma_{n}(c)|_{d}-|\sigma_{n}(c^{\prime})|_{d}|=0. (6)

Then the system (X𝛔,S)(X_{\boldsymbol{\sigma}},S) has dd ergodic measures μ0,,μd1\mu_{0},\ldots,\mu_{d-1}.

Moreover, for NN\in{\mathbb{N}} sufficiently large, the measures μi(n)\mu^{(n)}_{i} are characterized by μi(n)(ai)+μi(n)(bi)=max{μ(ai)+μ(bi):ν(X𝛔(n),S)}\mu^{(n)}_{i}(a_{i})+\mu^{(n)}_{i}(b_{i})=\max\{\mu^{\prime}(a_{i})+\mu^{\prime}(b_{i})\colon\nu\in{\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n)},S)\} for all nNn\geq N. Also, for all jij\neq i,

limnμi(n)(aj)+μi(n)(bj)=0.\lim_{n\to\infty}\mu_{i}^{(n)}(a_{j})+\mu_{i}^{(n)}(b_{j})=0.

Whenever 𝝈=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A_{n+1}^{*}\to A_{n}^{*})_{n\in{\mathbb{N}}} is a constant-length directive sequence, we write h(n)=|σ[0,n)|h^{(n)}=|\sigma_{[0,n)}| where we recall that σ[0,n)=σ0σ1σn1\sigma_{[0,n)}=\sigma_{0}\circ\sigma_{1}\circ\cdots\circ\sigma_{n-1}.

Theorem 2.3.

[13, Theorem 7.1] Let 𝛔=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A_{n+1}^{*}\to A_{n}^{*})_{n\in{\mathbb{N}}} be a recognizable, constant-length and primitive directive sequence. Let μ\mu be an SS-invariant ergodic measure on X𝛔X_{\boldsymbol{\sigma}}. Then

δμ=limnsupk2{w𝒞Ankμ(n)(w)},\delta_{\mu}=\lim_{n\to\infty}\sup_{k\geq 2}\left\{\sum_{w\in{\mathcal{C}}A^{k}_{n}}\mu^{(n)}(w)\right\}, (7)

where 𝒞Ank{\mathcal{C}}A^{k}_{n} is defined in (2). Moreover, if (kn)n(k_{n})_{n\in{\mathbb{N}}} is a sequence of integers (posibly constant), with kn2k_{n}\geq 2 for all nn\in{\mathbb{N}}, such that

δμ=limn{w𝒞Anknμ(n)(w)},\delta_{\mu}=\lim_{n\to\infty}\left\{\sum_{w\in{\mathcal{C}}A_{n}^{k_{n}}}\mu^{(n)}(w)\right\}, (8)

then the partial rigidity sequence is ((kn1)h(n))n((k_{n}-1)h^{(n)})_{n\in{\mathbb{N}}}.

Another useful characterization of the invariant measures is given by explicit formulas between the invariant measures of X𝝈(n)X_{\boldsymbol{\sigma}}^{(n)} and X𝝈(n+1)X_{\boldsymbol{\sigma}}^{(n+1)}. To do so we combine [2, Proposition 1.1, Theorem 1.4] and [1, Proposition 1.4]. In the original statements one needs to normalize the measures to get a probability measure (see [1, Proposition 1.3]), but for constant length morphisms the normalization constant is precisely the length of the morphism. Before stating the lemma, for σ:AB\sigma\colon A^{*}\to B^{*}, wAw\in A^{*} and uBu\in B^{*}, we define σ(w)u\lfloor\sigma(w)\rfloor_{u}, the essential occurrence of uu on σ(w)\sigma(w), that is the number of times such that uu occurs on ww for which the first letter of uu occurs in the image of the first letter of ww under σ\sigma, and the last letter of uu occurs in the image of last letter of ww under σ\sigma.

Example.

Let σ:𝒜𝒜\sigma\colon{\mathcal{A}}^{*}\to{\mathcal{A}}^{*} given by σ(a)=abab\sigma(a)=abab and σ(b)=babb\sigma(b)=babb. Then σ(ab)=ababbabb\sigma(ab)=ababbabb and |σ(ab)|abb=2|\sigma(ab)|_{abb}=2 but σ(ab)abb=1\lfloor\sigma(ab)\rfloor_{abb}=1.

Lemma 2.4.

Let 𝛔=(σn:An+1An)n\boldsymbol{\sigma}=(\sigma_{n}\colon A_{n+1}^{*}\to A_{n}^{*})_{n\in{\mathbb{N}}} be a recognizable constant-length and primitive directive sequence and fix an arbitrary nn\in{\mathbb{N}}. Then there is a bijection between (X𝛔(n),S){\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n)},S) and (X𝛔(n+1),S){\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n+1)},S).

Moreover, for every invariant measure μ(X𝛔(n),S)\mu^{\prime}\in{\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n)},S), there is an invariant measure μ(X𝛔(n+1),S)\mu\in{\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n+1)},S) such that for all words uAnu\in A_{n}^{*},

μ(u)=1|σn|wW(u)σn(w)uμ(w),\mu^{\prime}(u)=\frac{1}{|\sigma_{n}|}\sum_{w\in W(u)}\lfloor\sigma_{n}(w)\rfloor_{u}\cdot\mu(w), (9)

where W(u)={w:|w||u|2|σn|+2}\displaystyle W(u)=\left\{w\colon|w|\leq\frac{|u|-2}{|\sigma_{n}|}+2\right\}. Finally, if μ\mu is ergodic, then μ\mu^{\prime} is also ergodic.

Corollary 2.5.

Let 𝛔=(σn:ΛdΛd)n\boldsymbol{\sigma}=(\sigma_{n}\colon\Lambda_{d}^{*}\to\Lambda_{d}^{*})_{n\in{\mathbb{N}}} be a recognizable constant-length and primitive directive sequence that fulfills (4),(5) and (6) from Lemma 2.2. Letting μ0,,μd1\mu_{0},\ldots,\mu_{d-1} denote the dd ergodic measures, then for nn\in{\mathbb{N}} sufficiently large

μi(n)(u)=1|σn|wW(u)σn(w)uμi(n+1)(w)uΛd.\mu^{(n)}_{i}(u)=\frac{1}{|\sigma_{n}|}\sum_{w\in W(u)}\lfloor\sigma_{n}(w)\rfloor_{u}\cdot\mu^{(n+1)}_{i}(w)\quad\forall u\in\Lambda_{d}^{*}. (10)
Proof.

By the characterization given by Lemma 2.2 and using (9)

μi(n)(ai)\displaystyle\mu^{(n)}_{i}(a_{i}) +μi(n)(bi)=max{ν(ai)+ν(bi):ν(X𝝈(n),S)}\displaystyle+\mu^{(n)}_{i}(b_{i})=\max\{\nu(a_{i})+\nu(b_{i})\colon\nu\in{\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n)},S)\}
=1|σn|max{cΛd(|σn(c)|ai+|σn(c)|bi)ν(c)ν(X𝝈(n+1),S)}.\displaystyle=\frac{1}{|\sigma_{n}|}\max\left\{\sum_{c\in\Lambda_{d}}(|\sigma_{n}(c)|_{a_{i}}+|\sigma_{n}(c)|_{b_{i}})\cdot\nu^{\prime}(c)\mid\nu^{\prime}\in{\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n+1)},S)\right\}.

Using (5), for big enough nn\in{\mathbb{N}}, the invariant measure ν\nu^{\prime} that maximizes this equation has to be the invariant measure that maximize ν(ai)+ν(bi)\nu^{\prime}(a_{i})+\nu^{\prime}(b_{i}) which is in fact μi(n+1)\mu^{(n+1)}_{i}.

Remark 2.6.

When ϕ:AB\phi\colon A^{*}\to B^{*} is a letter to letter morphism, that is |ϕ(c)|=1|\phi(c)|=1 for all cAc\in A, we have that ϕ\phi induces a continuous map from AA^{{\mathbb{Z}}} to BB^{{\mathbb{Z}}} and that if μ\mu is an invariant measure in BB^{{\mathbb{Z}}}, then μ(w)=uϕ1(w)μ(u)\mu^{\prime}(w)=\displaystyle\sum_{u\in\phi^{-1}(w)}\mu(u) corresponds to the pushforward measure ϕμ\phi_{*}\mu.

3. The gluing technique and lower bound for the partial rigidity rates

We recall that 𝒜i={ai,bi}{\mathcal{A}}_{i}=\{a_{i},b_{i}\} and Λd=i=0d1𝒜i\Lambda_{d}=\bigcup_{i=0}^{d-1}{\mathcal{A}}_{i}. Let κ:ΛdΛd\kappa\colon\Lambda^{*}_{d}\to\Lambda_{d}^{*} be the function that for every word of the form uaiua_{i} (resp. ubiub_{i}) with uΛdu\in\Lambda_{d}^{*}, κ(uai)=uai+1\kappa(ua_{i})=ua_{i+1} (resp. κ(ubi)=ubi+1\kappa(ub_{i})=ub_{i+1}) where the index i{0,,d1}i\in\{0,\ldots,d-1\} is taken modulo dd. For example, if d=2d=2, κ(a0a0)=a0a1\kappa(a_{0}a_{0})=a_{0}a_{1}, κ(a0b0)=a0b1\kappa(a_{0}b_{0})=a_{0}b_{1}, κ(a0a1)=a0a0\kappa(a_{0}a_{1})=a_{0}a_{0} and κ(a0b1)=a0b0\kappa(a_{0}b_{1})=a_{0}b_{0}. We highlight that the function κ:ΛdΛd\kappa\colon\Lambda^{*}_{d}\to\Lambda_{d}^{*} is not a morphism.

For a finite collection of substitutions {τi:𝒜i𝒜ii=0,,d1}\{\tau_{i}\colon{\mathcal{A}}_{i}^{*}\to{\mathcal{A}}_{i}^{*}\mid i=0,\ldots,d-1\} we call the morphism σ=Γ(τ0,,τd1):ΛdΛd\sigma=\Gamma(\tau_{0},\ldots,\tau_{d-1})\colon\Lambda_{d}^{*}\to\Lambda_{d}^{*} given by

σ(ai)\displaystyle\sigma(a_{i}) =κ(τi(ai))\displaystyle=\kappa(\tau_{i}(a_{i}))
σ(bi)\displaystyle\sigma(b_{i}) =κ(τi(bi))\displaystyle=\kappa(\tau_{i}(b_{i}))

for all i{0,,d1}i\in\{0,\ldots,d-1\}, the glued substitution . This family of substitutions is the main ingredient for our construction.

Example.

Let d=2d=2, τ0:𝒜0𝒜0\tau_{0}\colon{\mathcal{A}}_{0}^{*}\to{\mathcal{A}}_{0}^{*} and τ1:𝒜1𝒜1\tau_{1}\colon{\mathcal{A}}_{1}^{*}\to{\mathcal{A}}_{1}^{*} be the substitutions given by

τ0(a0)=a0b0b0a0τ0(b0)=b0a0a0b0,τ1(a1)=a1b1b1b1τ1(b1)=b1a1a1a1.\begin{array}[]{cccc}\tau_{0}(a_{0})&=a_{0}b_{0}b_{0}a_{0}&\tau_{0}(b_{0})&=b_{0}a_{0}a_{0}b_{0},\\ \tau_{1}(a_{1})&=a_{1}b_{1}b_{1}b_{1}&\tau_{1}(b_{1})&=b_{1}a_{1}a_{1}a_{1}.\end{array}

Then σ=Γ(τ0,τ1):Λ2Λ2\sigma=\Gamma(\tau_{0},\tau_{1})\colon\Lambda_{2}^{*}\to\Lambda_{2}^{*} is given by

σ(a0)=a0b0b0a1σ(b0)=b0a0a0b1,σ(a1)=a1b1b1b0σ(b1)=b1a1a1a0\begin{array}[]{cccc}\sigma(a_{0})&=a_{0}b_{0}b_{0}a_{1}&\sigma(b_{0})&=b_{0}a_{0}a_{0}b_{1},\\ \sigma(a_{1})&=a_{1}b_{1}b_{1}b_{0}&\sigma(b_{1})&=b_{1}a_{1}a_{1}a_{0}\end{array}
Lemma 3.1.

Let τi:𝒜i𝒜i\tau_{i}\colon{\mathcal{A}}_{i}^{*}\to{\mathcal{A}}_{i}^{*} for i=0,d1i=0,\ldots d-1 be a collection of positive and prolongable substitutions. Let 𝛔=(σn:ΛdΛd)n\boldsymbol{\sigma}=(\sigma_{n}\colon\Lambda_{d}\to\Lambda_{d})_{n\in{\mathbb{N}}} be the directive sequence for which σn=Γ(τ0n+1,,τd1n+1)\sigma_{n}=\Gamma(\tau^{n+1}_{0},\ldots,\tau^{n+1}_{d-1}), that is

σn(ai)\displaystyle\sigma_{n}(a_{i}) =κ(τin+1(ai))\displaystyle=\kappa(\tau_{i}^{n+1}(a_{i}))
σn(bi)\displaystyle\sigma_{n}(b_{i}) =κ(τin+1(bi))\displaystyle=\kappa(\tau_{i}^{n+1}(b_{i}))

for all i{0,,d1}i\in\{0,\ldots,d-1\}. Then 𝛔\boldsymbol{\sigma} is primitive and left-permutative.

Proof.

Firstly, τ0,,τd1\tau_{0},\ldots,\tau_{d-1} are prolongable, in particular they are left-permutative and min{|τi(ai)|,|τi(bi)|}2\min\{|\tau_{i}(a_{i})|,|\tau_{i}(b_{i})|\}\geq 2 for all i{0,,d1}i\in\{0,\ldots,d-1\}. Since the function κ:ΛdΛd\kappa\colon\Lambda^{*}_{d}\to\Lambda^{*}_{d} does not change the first letter and every τi\tau_{i} is defined over a different alphabet, the left permutativity is preserved.

Secondly, M(σn)c,d=M(τin+1)c,d𝟙c=dM(\sigma_{n})_{c,d}=M(\tau_{i}^{n+1})_{c,d}-\mathds{1}_{c=d} if c,dc,d are in the same alphabet 𝒜i{\mathcal{A}}_{i}, M(σn)ai+1,ai=M(σn)bi+1,bi=1M(\sigma_{n})_{a_{i+1},a_{i}}=M(\sigma_{n})_{b_{i+1},b_{i}}=1 and M(σn)c,d=0M(\sigma_{n})_{c,d}=0 otherwise. Notice that by positivity and prolongability, the sub-blocks (M(σn)c,d)c,d𝒜i(M(\sigma_{n})_{c,d})_{c,d\in{\mathcal{A}}_{i}} are positive and therefore, for every nn\in{\mathbb{N}}, M(σ[n,n+d))M(\sigma_{[n,n+d)}) only has positive entries. ∎

Theorem 3.2.

Let τi:𝒜i𝒜i\tau_{i}\colon{\mathcal{A}}_{i}^{*}\to{\mathcal{A}}_{i}^{*} for i=0,,d1i=0,\ldots,d-1 be a collection of positive and prolongable substitutions. Suppose that every substitution τi\tau_{i} has constant length for the same length. Let 𝛔=(σn:ΛdΛd)n\boldsymbol{\sigma}=(\sigma_{n}\colon\Lambda_{d}\to\Lambda_{d})_{n\in{\mathbb{N}}} be the directive sequence of glued substitutions σn=Γ(τ0n+1,,τd1n+1)\sigma_{n}=\Gamma(\tau^{n+1}_{0},\ldots,\tau^{n+1}_{d-1}). Then the 𝒮{\mathcal{S}}-adic subshift (X𝛔,S)(X_{\boldsymbol{\sigma}},S) is minimal and has dd ergodic measures μ0,,μd1\mu_{0},\ldots,\mu_{d-1} such that for every i{0,,d1}i\in\{0,\ldots,d-1\}

limnμi(n)(w)=νi(w) for all w𝒜i\displaystyle\lim_{n\to\infty}\mu^{(n)}_{i}(w)=\nu_{i}(w)\quad\text{ for all }w\in{\mathcal{A}}_{i}^{*} (11)

where νi\nu_{i} is the unique invariant measure of the substitution subshift given by τi\tau_{i}.

Remark.

From (11), we get that limnμi(n)(ai)+μi(n)(bi)=1\displaystyle\lim_{n\to\infty}\mu^{(n)}_{i}(a_{i})+\mu_{i}^{(n)}(b_{i})=1 and therefore
limnμi(n)(w)=0\displaystyle\lim_{n\to\infty}\mu^{(n)}_{i}(w)=0 for all w𝒜iw\not\in{\mathcal{A}}_{i}^{*}.

Before proving the theorem, we want to emphasize that this gluing technique can be easily generalized. Indeed, many of the hypothesis are not necessary but we include them to simplify notation and computations. For instance, restricting the analysis to substitutions defined over two letter alphabets is arbitrary. Also, the function κ:ΛdΛd\kappa\colon\Lambda^{*}_{d}\to\Lambda_{d}^{*} could change more than one letter at the end of words. Furthermore, with an appropriated control of the growth, the number of letters replaced could even increase with the levels.

One fact that seems critical for the conclusion of Theorem 3.2 is that 𝝈\boldsymbol{\sigma} is a constant-length directive sequence and that 1|σn|M(σn)c,d\frac{1}{|\sigma_{n}|}M(\sigma_{n})_{c,d} for two letters cc and dd in distinct alphabets 𝒜i{\mathcal{A}}_{i}, 𝒜j{\mathcal{A}}_{j} goes to zero when nn goes to infinity.

Proof.

By Lemma 3.1, (X𝝈,S)(X_{\boldsymbol{\sigma}},S) is minimal. Let |τi|=|\tau_{i}|=\ell, which is well defined because the substitutions τ0,,τd1\tau_{0},\ldots,\tau_{d-1} all have the same length. Then, for every nn\in{\mathbb{N}}, σn=Γ(τ0n+1,,τd1n+1)\sigma_{n}=\Gamma(\tau_{0}^{n+1},\ldots,\tau_{d-1}^{n+1}) has constant length n+1\ell^{n+1}.

We need to prove that (X𝝈,S)(X_{\boldsymbol{\sigma}},S) has dd ergodic measures, and so we check the hypotheses of Lemma 2.2,

limn1|σn|ji|σn(ai)|aj+|σn(ai)|bj+|σn(bi)|aj+|σn(bi)|bj\displaystyle\lim_{n\to\infty}\frac{1}{|\sigma_{n}|}\sum_{j\neq i}|\sigma_{n}(a_{i})|_{a_{j}}+|\sigma_{n}(a_{i})|_{b_{j}}+|\sigma_{n}(b_{i})|_{a_{j}}+|\sigma_{n}(b_{i})|_{b_{j}}
=limn1n+1(|σn(ai)|ai+1+|σn(bi)|bi+1)=limn2n+1=0.\displaystyle=\lim_{n\to\infty}\frac{1}{\ell^{n+1}}(|\sigma_{n}(a_{i})|_{a_{i+1}}+|\sigma_{n}(b_{i})|_{b_{i+1}})=\lim_{n\to\infty}\frac{2}{\ell^{n+1}}=0.

This verifies (4). Similarly for (5),

n1(11n+1(|σn(ai)|ai+|σn(ai)|bi))=n1(1n+11n+1)<.\sum_{n\geq 1}\left(1-\frac{1}{\ell^{n+1}}(|\sigma_{n}(a_{i})|_{a_{i}}+|\sigma_{n}(a_{i})|_{b_{i}})\right)=\sum_{n\geq 1}\left(1-\frac{\ell^{n+1}-1}{\ell^{n+1}}\right)<\infty.

For (6), notice that |σn(ai)|ai=|τin+1(ai)|ai1|\sigma_{n}(a_{i})|_{a_{i}}=|\tau_{i}^{n+1}(a_{i})|_{a_{i}}-1, therefore 1n+1|σn(ai)|ai=freq(ai,τn+1(ai))1n+1\frac{1}{\ell^{n+1}}|\sigma_{n}(a_{i})|_{a_{i}}={\rm freq}(a_{i},\tau^{n+1}(a_{i}))-\frac{1}{\ell^{n+1}}. Similarly for |σn(ai)|bi,|σn(bi)|ai|\sigma_{n}(a_{i})|_{b_{i}},|\sigma_{n}(b_{i})|_{a_{i}} and |σn(bi)|bi|\sigma_{n}(b_{i})|_{b_{i}}. Therefore

limn1n+1||σn(ai)|ai|σn(bi)|ai|\displaystyle\lim_{n\to\infty}\frac{1}{\ell^{n+1}}||\sigma_{n}(a_{i})|_{a_{i}}-|\sigma_{n}(b_{i})|_{a_{i}}|
=\displaystyle= limn|freq(ai,τin+1(ai))freq(ai,τin+1(bi))|=νi(ai)νi(ai)=0.\displaystyle\lim_{n\to\infty}|{\rm freq}(a_{i},\tau_{i}^{n+1}(a_{i}))-{\rm freq}(a_{i},\tau_{i}^{n+1}(b_{i}))|=\nu_{i}(a_{i})-\nu_{i}(a_{i})=0.

Likewise limn1n+1||σn(ai)|bi|σn(bi)|bi|=νi(bi)νi(bi)=0\displaystyle\lim_{n\to\infty}\frac{1}{\ell^{n+1}}||\sigma_{n}(a_{i})|_{b_{i}}-|\sigma_{n}(b_{i})|_{b_{i}}|=\nu_{i}(b_{i})-\nu_{i}(b_{i})=0.

Thus, by Lemma 2.2, there are dd ergodic measures, μ0,,μd1\mu_{0},\ldots,\mu_{d-1} which are characterize by

μi(n)(ai)+μi(n)(bi)=max{μ(ai)+μ(bi):μ(X𝝈(n),S)}\mu^{(n)}_{i}(a_{i})+\mu^{(n)}_{i}(b_{i})=\max\{\mu^{\prime}(a_{i})+\mu^{\prime}(b_{i})\colon\mu^{\prime}\in{\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n)},S)\} (12)

for sufficiently large nn\in{\mathbb{N}}. The invariant measure that reaches the maximum in (12) can be characterize as a limit like in (3). Indeed, fix nn\in{\mathbb{N}} sufficiently large, i{0,,d1}i\in\{0,\ldots,d-1\} and define the infinite one-sided word 𝒘(n)=limkσ[n,n+k](ai)=limk(σnσn+k)(ai)\displaystyle\boldsymbol{w}^{(n)}=\lim_{k\to\infty}\sigma_{[n,n+k]}(a_{i})=\lim_{k\to\infty}(\sigma_{n}\circ\cdots\circ\sigma_{n+k})(a_{i}) and the number Nk(n)=|σ[n,n+k](ai)|N_{k}^{(n)}=|\sigma_{[n,n+k]}(a_{i})| for every kk\in{\mathbb{N}}. Let μn(X𝝈,S)\mu_{n}\in{\mathcal{M}}(X_{\boldsymbol{\sigma}},S) be the measure given by

μn(u)=limk1Nk(n)|𝒘[1,Nk(n)](n)|u=limkfreq(u,σ[n,n+k](ai))\mu_{n}(u)=\lim_{k\to\infty}\frac{1}{N^{(n)}_{k}}\left|\boldsymbol{w}^{(n)}_{[1,N^{(n)}_{k}]}\right|_{u}=\lim_{k\to\infty}{\rm freq}(u,\sigma_{[n,n+k]}(a_{i}))

for all uΛdu\in\Lambda_{d}^{*}. Notice that for any other Følner sequence of the form ({mk,mk+1,,mk})k(\{m_{k},m_{k}+1,\ldots,m^{\prime}_{k}\})_{k\in{\mathbb{N}}}, limk1mkmk(|𝒘[mk,mk)(n)|ai+|𝒘[mk,mk)(n)|bi)μn(ai)+μn(bi)\displaystyle\lim_{k\to\infty}\frac{1}{m^{\prime}_{k}-m_{k}}\left(\left|\boldsymbol{w}^{(n)}_{[m_{k},m^{\prime}_{k})}\right|_{a_{i}}+\left|\boldsymbol{w}^{(n)}_{[m_{k},m^{\prime}_{k})}\right|_{b_{i}}\right)\leq\mu_{n}(a_{i})+\mu_{n}(b_{i}). Thus, if μ\mu^{\prime} is given by μ(u)=limk1mkmk|𝒘[mk,mk)(n)|u\displaystyle\mu^{\prime}(u)=\lim_{k\to\infty}\frac{1}{m^{\prime}_{k}-m_{k}}\left|\boldsymbol{w}^{(n)}_{[m_{k},m^{\prime}_{k})}\right|_{u} we get that μ(ai)+μ(bi)μn(ai)+μn(bi)\mu^{\prime}(a_{i})+\mu^{\prime}(b_{i})\leq\mu_{n}(a_{i})+\mu_{n}(b_{i}) and since every invariant measure μ(X𝝈(n),S)\mu^{\prime}\in{\mathcal{M}}(X_{\boldsymbol{\sigma}}^{(n)},S) has this form, μn=μi(n)\mu_{n}=\mu_{i}^{(n)} by (12).

To prove (11), fix w𝒜iw\in{\mathcal{A}}_{i}^{*} and nn\in{\mathbb{N}} large enough, then

μi(n)(w)\displaystyle\mu_{i}^{(n)}(w) =limk|σ[n,n+k](ai)|w|σ[n,n+k](ai)|=limk|σ[n,n+k)κ(τin+k+1(ai))|w|σ[n,n+k](ai)|\displaystyle=\lim_{k\to\infty}\frac{|\sigma_{[n,n+k]}(a_{i})|_{w}}{|\sigma_{[n,n+k]}(a_{i})|}=\lim_{k\to\infty}\frac{|\sigma_{[n,n+k)}\circ\kappa(\tau_{i}^{n+k+1}(a_{i}))|_{w}}{|\sigma_{[n,n+k]}(a_{i})|}
limk1|σ[n,n+k](ai)|(|σ[n,n+k)(τin+k+1(ai))|w1+|σ[n,n+k)(ai+1)|w)\displaystyle\geq\lim_{k\to\infty}\frac{1}{|\sigma_{[n,n+k]}(a_{i})|}\left(|\sigma_{[n,n+k)}(\tau_{i}^{n+k+1}(a_{i}))|_{w}-1+|\sigma_{[n,n+k)}(a_{i+1})|_{w}\right)
limk|σ[n,n+k)(τin+k+1(ai))|w|σ[n,n+k](ai)|,\displaystyle\geq\lim_{k\to\infty}\frac{|\sigma_{[n,n+k)}(\tau_{i}^{n+k+1}(a_{i}))|_{w}}{|\sigma_{[n,n+k]}(a_{i})|}, (13)

where in the last inequality we use that |σ[n,n+k]|=nn+1n+k+1|\sigma_{[n,n+k]}|=\ell^{n}\cdot\ell^{n+1}\cdots\ell^{n+k+1} and therefore |σ[n,n+k)||σ[n,n+k]|=1n+k+1k0\frac{|\sigma_{[n,n+k)}|}{|\sigma_{[n,n+k]}|}=\frac{1}{\ell^{n+k+1}}\xrightarrow{k\to\infty}0. Notice that

|σ[n,n+k)(τin+k+1(ai))|w\displaystyle|\sigma_{[n,n+k)}(\tau_{i}^{n+k+1}(a_{i}))|_{w} |σ[n,n+k)(ai)|w|τin+k+1(ai)|ai\displaystyle\geq|\sigma_{[n,n+k)}(a_{i})|_{w}|\tau_{i}^{n+k+1}(a_{i})|_{a_{i}}
+|σ[n,n+k)(bi)|w|τin+k+1(ai)|bi\displaystyle+|\sigma_{[n,n+k)}(b_{i})|_{w}|\tau_{i}^{n+k+1}(a_{i})|_{b_{i}}

and since |τin+k+1(ai)|ai+|τin+k+1(ai)|bi=n+k+1|\tau_{i}^{n+k+1}(a_{i})|_{a_{i}}+|\tau_{i}^{n+k+1}(a_{i})|_{b_{i}}=\ell^{n+k+1} there exists λ(0,1)\lambda\in(0,1) such that

|σ[n,n+k)(τin+k+1(ai))|wn+k+1(λ|σ[n,n+k)(ai)|w+(1λ)|σ[n,n+k)(bi)|w).|\sigma_{[n,n+k)}(\tau_{i}^{n+k+1}(a_{i}))|_{w}\geq\ell^{n+k+1}\left(\lambda|\sigma_{[n,n+k)}(a_{i})|_{w}+(1-\lambda)|\sigma_{[n,n+k)}(b_{i})|_{w}\right).

Combining the previous inequality with (13) and supposing, without lost of generality, that |σ[n,n+k)(ai)|w=min{|σ[n,n+k)(ai)|w,|σ[n,n+k)(bi)|w}\displaystyle|\sigma_{[n,n+k)}(a_{i})|_{w}=\min\{|\sigma_{[n,n+k)}(a_{i})|_{w},|\sigma_{[n,n+k)}(b_{i})|_{w}\}, we get that

μi(n)(w)limkn+k+1|σ[n,n+k](ai)||σ[n,n+k)(ai)|w.\mu_{i}^{(n)}(w)\geq\lim_{k\to\infty}\frac{\ell^{n+k+1}}{|\sigma_{[n,n+k]}(a_{i})|}|\sigma_{[n,n+k)}(a_{i})|_{w}.

Now inductively

μi(n)(w)\displaystyle\mu_{i}^{(n)}(w) limkn+2n+3n+k+1|σ[n,n+k](ai)||τin+1(ai)|w=|τin+1(ai)|wn+1,\displaystyle\geq\lim_{k\to\infty}\frac{\ell^{n+2}\ell^{n+3}\cdots\ell^{n+k+1}}{|\sigma_{[n,n+k]}(a_{i})|}|\tau_{i}^{n+1}(a_{i})|_{w}=\frac{|\tau_{i}^{n+1}(a_{i})|_{w}}{\ell^{n+1}},

where in the last equality we use again that |σ[n,n+k]|=nn+1n+k+1|\sigma_{[n,n+k]}|=\ell^{n}\cdot\ell^{n+1}\cdots\ell^{n+k+1}. We conclude that μi(n)(w)freq(w,τin+1(ai))\displaystyle\mu_{i}^{(n)}(w)\geq{\rm freq}(w,\tau_{i}^{n+1}(a_{i})), and then taking nn\to\infty,

limnμi(n)(w)limnfreq(w,τin(ai))=νi(w).\lim_{n\to\infty}\mu_{i}^{(n)}(w)\geq\lim_{n\to\infty}{\rm freq}(w,\tau_{i}^{n}(a_{i}))=\nu_{i}(w). (14)

Since w𝒜iw\in{\mathcal{A}}_{i}^{*} was arbitrary (14) holds for every word with letters in 𝒜i{\mathcal{A}}_{i}. In particular, for every k1k\geq 1, 1=u𝒜ikνi(u)limnu𝒜ikμi(n)(u)1\displaystyle 1=\sum_{u\in{\mathcal{A}}_{i}^{k}}\nu_{i}(u)\leq\lim_{n\to\infty}\sum_{u\in{\mathcal{A}}_{i}^{k}}\mu_{i}^{(n)}(u)\leq 1 which implies that the inequality in (14) is an equality for every word w𝒜iw\in{\mathcal{A}}_{i}^{*}. ∎

In what follows every system (X𝝈,S)(X_{\boldsymbol{\sigma}},S) and family of substitutions τi:𝒜i𝒜i\tau_{i}\colon{\mathcal{A}}^{*}_{i}\to{\mathcal{A}}^{*}_{i} for i=0,,d1i=0,\ldots,d-1 satisfy the assumption of Theorem 3.2.

Corollary 3.3.

(X𝝈,S)(X_{\boldsymbol{\sigma}},S) has non-superlinear complexity.

Proof.

This is direct from [12, Corollary 6.7] where 𝒮{\mathcal{S}}-adic subshifts with finite alphabet rank and constant-length primitive directive sequences have non-superlinear complexity. ∎

Corollary 3.4.

If μ0,,μd1\mu_{0},\ldots,\mu_{d-1} are the ergodic measures of (X𝛔,S)(X_{\boldsymbol{\sigma}},S), then

δνiδμi\delta_{\nu_{i}}\leq\delta_{\mu_{i}} (15)

for all i{0,,d1}i\in\{0,\ldots,d-1\}, where each νi\nu_{i} is the unique invariant measure of XτiX_{\tau_{i}}.

Proof.

By Theorem 2.3 equation (8), there exists a sequence of (kt)t(k_{t})_{t\in{\mathbb{N}}} such that

δνi=limtw𝒞𝒜iktνi(w)\delta_{\nu_{i}}=\lim_{t\to\infty}\sum_{w\in{\mathcal{C}}{\mathcal{A}}_{i}^{k_{t}}}\nu_{i}(w)

and by (11) for every tt\in{\mathbb{N}}, there exists ntn_{t} such that

w𝒞𝒜iktμi(n)(w)w𝒞𝒜iktνi(w)1t for all nnt.\sum_{w\in{\mathcal{C}}{\mathcal{A}}_{i}^{k_{t}}}\mu_{i}^{(n)}(w)\geq\sum_{w\in{\mathcal{C}}{\mathcal{A}}_{i}^{k_{t}}}\nu_{i}(w)-\frac{1}{t}\quad\text{ for all }n\geq n_{t}.

Taking limits we have,

δμilimt(w𝒞𝒜iktνi(w)1t)=δνi.\delta_{\mu_{i}}\geq\lim_{t\to\infty}\left(\sum_{w\in{\mathcal{C}}{\mathcal{A}}_{i}^{k_{t}}}\nu_{i}(w)-\frac{1}{t}\right)=\delta_{\nu_{i}}.\qed

We finish this section with a case where the lower bound in (15) is trivially achieved. For that, when we define a substitution τ:𝒜𝒜\tau\colon{\mathcal{A}}^{*}\to{\mathcal{A}}^{*} we abuse notation and write τ:𝒜i𝒜i\tau\colon{\mathcal{A}}_{i}^{*}\to{\mathcal{A}}_{i}^{*}, by replacing the letters aa and bb by aia_{i} and bib_{i} respectively. Using that abuse of notation for iji\neq j, we say that τ:𝒜i𝒜i\tau\colon{\mathcal{A}}_{i}^{*}\to{\mathcal{A}}_{i}^{*} and τ:𝒜j𝒜j\tau\colon{\mathcal{A}}_{j}^{*}\to{\mathcal{A}}_{j}^{*} are the same substitution even though they are defined over different alphabets. We write Γ(τ,d):ΛdΛd\Gamma(\tau,d)\colon\Lambda_{d}^{*}\to\Lambda_{d}^{*} when we are gluing dd times the same substitution. In the next corollary we prove that if we glue the same substitutions then we achieve the bound.

Corollary 3.5.

Let τ:𝒜𝒜\tau\colon{\mathcal{A}}^{*}\to{\mathcal{A}}^{*} be a positive, prolongable and constant length substitution. Let 𝛔=(σn:ΛdΛd)n\boldsymbol{\sigma}=(\sigma_{n}\colon\Lambda_{d}\to\Lambda_{d})_{n\in{\mathbb{N}}} be the directive sequence of glued substitutions σn=Γ(τn+1,d)\sigma_{n}=\Gamma(\tau^{n+1},d). Then (X𝛔,S)(X_{\boldsymbol{\sigma}},S) has dd ergodic measures with the same partial rigidity rate δν\delta_{\nu}, where ν\nu denotes the unique invariant measure of the substitution subshift (Xτ,S)(X_{\tau},S).

Proof.

The letter-to-letter morphism ϕ:Λd𝒜\phi\colon\Lambda_{d}^{*}\to{\mathcal{A}}^{*} given by aiaa_{i}\mapsto a and bibb_{i}\mapsto b for all i=0,,d1i=0,\ldots,d-1 induce a factor map from X𝝈X_{\boldsymbol{\sigma}} to XτX_{\tau} and therefore δμδν\delta_{\mu}\leq\delta_{\nu} for all μ(X𝝈,S)\mu\in{\mathcal{E}}(X_{\boldsymbol{\sigma}},S) (see [24, Proposition 1.13]). The opposite inequality is given by Corollary 3.4. ∎

4. Computation of the partial rigidity rates

4.1. Decomposition of the directive sequence

We maintain the notation, using 𝒜i={ai,bi}{\mathcal{A}}_{i}=\{a_{i},b_{i}\} and Λd=i=0d1𝒜i\Lambda_{d}=\bigcup_{i=0}^{d-1}{\mathcal{A}}_{i} and we also fix 𝒜i={ai,bi}{\mathcal{A}}_{i}^{\prime}=\{a_{i}^{\prime},b_{i}^{\prime}\}, Λd=i=0d1𝒜i𝒜i\Lambda_{d}^{\prime}=\bigcup_{i=0}^{d-1}{\mathcal{A}}_{i}\cup{\mathcal{A}}_{i}^{\prime}. In this section, τi:𝒜i𝒜i\tau_{i}\colon{\mathcal{A}}^{*}_{i}\to{\mathcal{A}}_{i}^{*} for i=0,,d1i=0,\ldots,d-1 is a collection of mirror substitutions satisfying the hypothesis of Theorem 3.2, =|τi|\ell=|\tau_{i}| and 𝝈=(Γ(τ0n+1,,τd1n+1))n\boldsymbol{\sigma}=(\Gamma(\tau_{0}^{n+1},\ldots,\tau_{d-1}^{n+1}))_{n\in{\mathbb{N}}}, that is

σn(ai)\displaystyle\sigma_{n}(a_{i}) =κ(τin+1(ai))\displaystyle=\kappa(\tau_{i}^{n+1}(a_{i}))
σn(bi)\displaystyle\sigma_{n}(b_{i}) =κ(τin+1(bi))\displaystyle=\kappa(\tau_{i}^{n+1}(b_{i}))

for all i{0,,d1}i\in\{0,\ldots,d-1\}. We also write {\mathcal{E}} instead of (X𝝈,S)={μ0,,μd1}{\mathcal{E}}(X_{\boldsymbol{\sigma}},S)=\{\mu_{0},\ldots,\mu_{d-1}\} for the set of ergodic measures.

Proposition 4.1.

The directive sequence 𝛔\boldsymbol{\sigma} can be decomposed using 33 morphisms in the following way: for every nn\in{\mathbb{N}}, σn=ϕρnψ\sigma_{n}=\phi\circ\rho^{n}\circ\psi where

ψ:Λd(Λd)\displaystyle\psi\colon\Lambda_{d}^{*}\to(\Lambda_{d}^{\prime})^{*} aiuiai+1\displaystyle\quad a_{i}\mapsto u_{i}a_{i+1}^{\prime}
bivibi+1\displaystyle\quad b_{i}\mapsto v_{i}b_{i+1}^{\prime}
ρ:(Λd)(Λd)\displaystyle\rho\colon(\Lambda_{d}^{\prime})^{*}\to(\Lambda_{d}^{\prime})^{*} aiτi(ai)aiui1ai\displaystyle\quad a_{i}\mapsto\tau_{i}(a_{i})\quad a_{i}^{\prime}\mapsto u_{i-1}a_{i}^{\prime}
biτi(bi)bivi1bi\displaystyle\quad b_{i}\mapsto\tau_{i}(b_{i})\quad b_{i}^{\prime}\mapsto v_{i-1}b_{i}^{\prime}
ϕ:(Λd)Λd\displaystyle\phi\colon(\Lambda_{d}^{\prime})^{*}\to\Lambda_{d}^{*} aiaiaiai\displaystyle\quad a_{i}\mapsto a_{i}\quad a_{i}^{\prime}\mapsto a_{i}
bibibibi.\displaystyle\quad b_{i}\mapsto b_{i}\quad b_{i}^{\prime}\mapsto b_{i}.

with ui=τi(ai)[1,)u_{i}=\tau_{i}(a_{i})_{[1,\ell)} and vi=τi(bi)[1,)v_{i}=\tau_{i}(b_{i})_{[1,\ell)} and the index ii is taken modulo dd.

Proof.

Fix i{0,,d1}i\in\{0,\ldots,d-1\}. Consider first that for every n1n\geq 1, ρn(ai+1)=ρn1(ui)ρn1(ai+1)=τin1(ui)ρn1(ai+1)\rho^{n}(a_{i+1}^{\prime})=\rho^{n-1}(u_{i})\rho^{n-1}(a_{i+1}^{\prime})=\tau_{i}^{n-1}(u_{i})\rho^{n-1}(a_{i+1}^{\prime}), therefore by induction

ρn(ai+1)=τin1(ui)τin2(ui)τi(ui)uiai+1.\rho^{n}(a_{i+1}^{\prime})=\tau_{i}^{n-1}(u_{i})\tau_{i}^{n-2}(u_{i})\cdots\tau_{i}(u_{i})u_{i}a_{i+1}^{\prime}.

Since, by assumption, the last letter of τi(ai)\tau_{i}(a_{i}) is aia_{i}, one gets that τin1(ui)τin2(ui)\tau_{i}^{n-1}(u_{i})\tau_{i}^{n-2}(u_{i}) τi(ui)ui=τn(ai)[1,n)\cdots\tau_{i}(u_{i})u_{i}=\tau^{n}(a_{i})_{[1,\ell^{n})} and then ρn(ai+1)=τn(ai)[1,n)ai+1\rho^{n}(a_{i+1}^{\prime})=\tau^{n}(a_{i})_{[1,\ell^{n})}a_{i+1}^{\prime}. Also, we notice that ψ(ai)=ρ(ai+1)\psi(a_{i})=\rho(a_{i+1}^{\prime}) and therefore ρnψ(ai)=ρn+1(ai+1)=τn+1(ai)[1,n+1)ai+1\rho^{n}\circ\psi(a_{i})=\rho^{n+1}(a_{i+1}^{\prime})=\tau^{n+1}(a_{i})_{[1,\ell^{n+1})}a_{i+1}^{\prime}.

Finally, ϕρnψ(ai)=ϕ(τn+1(ai)[1,n+1))ϕ(ai+1)=τn+1(ai)[1,n+1)ai+1=κ(τn+1(ai))=σn(ai).\displaystyle\phi\circ\rho^{n}\circ\psi(a_{i})=\phi(\tau^{n+1}(a_{i})_{[1,\ell^{n+1})})\phi(a_{i+1}^{\prime})=\tau^{n+1}(a_{i})_{[1,\ell^{n+1})}a_{i+1}=\kappa(\tau^{n+1}(a_{i}))=\sigma_{n}(a_{i}). We conclude noticing that the same proof works for bib_{i}. ∎

With this decomposition, we make an abuse of notation and define a directive sequence 𝝈\boldsymbol{\sigma}^{\prime} over an index QQ different from {\mathbb{N}}.

Set Q={0}n1{n+mn+2:m=0,,n+1}\displaystyle Q=\{0\}\cup\bigcup_{n\geq 1}\left\{n+\frac{m}{n+2}:m=0,\ldots,n+1\right\} we define the directive sequence 𝝈\boldsymbol{\sigma}^{\prime} indexed by QQ given by

σq={ϕ if q=nρ if q=n+m/(n+2) for m=1,,nψ if q=n+(n+1)/(n+2)\sigma^{\prime}_{q}=\begin{cases}\begin{array}[]{cc}\phi&\text{ if }q=n\\ \rho&\text{ if }q=n+m/(n+2)\text{ for }m=1,\ldots,n\\ \psi&\text{ if }q=n+(n+1)/(n+2)\end{array}\end{cases}

for all n1n\geq 1. We use this abuse of notation, in order to get X𝝈(n)=X𝝈(n)X^{(n)}_{\boldsymbol{\sigma}}=X^{(n)}_{\boldsymbol{\sigma}^{\prime}} for every positive integer nn, and therefore we maintain the notation for μi(n)\mu^{(n)}_{i}. The advantage of decomposing the directive sequence is that every morphism in 𝝈\boldsymbol{\sigma} has constant length, either \ell in the case of ψ\psi and ρ\rho or 11 in the case of ϕ\phi. This simplifies the study of the complete words at each level. Notice that, the morphisms ϕ\phi, ρ\rho and ψ\psi are not positive, otherwise the 𝒮{\mathcal{S}}-adic subshift would automatically be uniquely ergodic, see [15], which does not happen as we show in Theorem 3.2.

4.2. Recurrence formulas for complete words

The formulas in this section are analogous to those presented in [13, Lemma 7.7], and aside from technicalities, the proofs are not so different.

We define four sets of words that are useful in what follows,

Cki\displaystyle C_{k}^{i} ={wΛdk:w1,wk𝒜i𝒜i+1,w1=wk}\displaystyle=\{w\in\Lambda_{d}^{k}\colon w_{1},w_{k}\in{\mathcal{A}}_{i}\cup{\mathcal{A}}_{i+1}^{\prime},w_{1}=w_{k}\} (16)
Dki\displaystyle D_{k}^{i} ={w(Λd)k:w1,wk𝒜i𝒜i+1,η(w1)=η(wk)}\displaystyle=\{w\in(\Lambda_{d}^{\prime})^{k}\colon w_{1},w_{k}\in{\mathcal{A}}_{i}\cup{\mathcal{A}}_{i+1}^{\prime},\eta(w_{1})=\eta(w_{k})\} (17)
C¯ki\displaystyle\overline{C}_{k}^{i} ={wΛdk:w1,wk𝒜i𝒜i+1,w1=wk¯}\displaystyle=\{w\in\Lambda_{d}^{k}\colon w_{1},w_{k}\in{\mathcal{A}}_{i}\cup{\mathcal{A}}_{i+1}^{\prime},w_{1}=\overline{w_{k}}\} (18)
D¯ki\displaystyle\overline{D}_{k}^{i} ={w(Λd)k:w1,wk𝒜i𝒜i+1,η(w1)=η(wk)¯}\displaystyle=\{w\in(\Lambda_{d}^{\prime})^{k}\colon w_{1},w_{k}\in{\mathcal{A}}_{i}\cup{\mathcal{A}}_{i+1}^{\prime},\eta(w_{1})=\overline{\eta(w_{k})}\} (19)

where η:ΛdΛd\eta\colon\Lambda_{d}^{\prime}\to\Lambda_{d} is a letter-to-letter function for which aiaia_{i}\mapsto a_{i}, bibib_{i}\mapsto b_{i}, ai+1aia_{i+1}^{\prime}\mapsto a_{i} and bi+1bib_{i+1}^{\prime}\mapsto b_{i}. For instance if wDkiw\in D_{k}^{i} and w1=aiw_{1}=a_{i} then wk{ai,ai+1}w_{k}\in\{a_{i},a_{i+1}^{\prime}\}.

To simplify the notation, we enumerate the index set Q={qm:m}Q=\{q_{m}\colon m\in{\mathbb{N}}\} where qm<qm+1q_{m}<q_{m+1} for all mm\in{\mathbb{N}}. We continue using the abuse of notation μ(w)=μ([w])\mu(w)=\mu([w]) and for a set of words WW, μ(W)=μ(wW[w])\displaystyle\mu(W)=\mu\left(\bigcup_{w\in W}[w]\right).

For i{0,,d1}i\in\{0,\ldots,d-1\}, fix the word v=τi(ai)v=\tau_{i}(a_{i}) and we define δj,ji=𝟙vj=vj\delta_{j,j^{\prime}}^{i}=\mathds{1}_{v_{j}=v_{j^{\prime}}} for j,j={1,,}j,j^{\prime}=\{1,\ldots,\ell\} where =|v|\ell=|v|. Notice that if one defines δj,ji\delta_{j,j^{\prime}}^{i} with the word τi(bi)\tau_{i}(b_{i}) instead of τi(ai)\tau_{i}(a_{i}), by the mirror property, the value remains the same. Now, for j{1,,}j\in\{1,\ldots,\ell\}, we define

rji=j=1jδj+j,ji and r~ji=j=1jδj,j+ji.r_{j}^{i}=\sum^{j}_{j^{\prime}=1}\delta_{\ell-j+j^{\prime},j^{\prime}}^{i}\quad\text{ and }\quad\tilde{r}_{j}^{i}=\sum^{\ell-j}_{j^{\prime}=1}\delta_{j^{\prime},j+j^{\prime}}^{i}.
Lemma 4.2.

If 𝛔=(σq)qQ\boldsymbol{\sigma}^{\prime}=(\sigma^{\prime}_{q})_{q\in Q} and μ\mu\in{\mathcal{E}}, then for every nn\in{\mathbb{N}}, and every qm=n+mn+2q_{m}=n+\frac{m^{\prime}}{n+2} for m{1,,n}m^{\prime}\in\{1,\ldots,n\},

μ(qm)(Dk+ji)=\displaystyle\ell\cdot\mu^{(q_{m})}(D^{i}_{\ell k+j})= rjiμ(qm+1)(Dk+2i)+r~jiμ(qm+1)(Dk+1i)\displaystyle r^{i}_{j}\cdot\mu^{(q_{m+1})}(D^{i}_{k+2})+\tilde{r}^{i}_{j}\cdot\mu^{(q_{m+1})}(D^{i}_{k+1})
+(jrji)μ(qm+1)(D¯k+2i)+(jr~ji)μ(qm+1)(D¯k+1i)\displaystyle+(j-r^{i}_{j})\mu^{(q_{m+1})}(\overline{D}^{i}_{k+2})+(\ell-j-\tilde{r}^{i}_{j})\mu^{(q_{m+1})}(\overline{D}^{i}_{k+1})
μ(qm)(D¯k+ji)=\displaystyle\ell\cdot\mu^{(q_{m})}(\overline{D}^{i}_{\ell k+j})= (jrji)μ(qm+1)(Dk+2i)+(jr~ji)μ(qm+1)(Dk+1i)\displaystyle(j-r^{i}_{j})\mu^{(q_{m+1})}(D^{i}_{k+2})+(\ell-j-\tilde{r}^{i}_{j})\mu^{(q_{m+1})}(D^{i}_{k+1})
+rjiμ(qm+1)(D¯k+2i)+r~jiμ(qm+1)(D¯k+1i)\displaystyle+r^{i}_{j}\cdot\mu^{(q_{m+1})}(\overline{D}^{i}_{k+2})+\tilde{r}^{i}_{j}\cdot\mu^{(q_{m+1})}(\overline{D}^{i}_{k+1})

for j{1,,}j\in\{1,\ldots,\ell\}, where the set DkiD^{i}_{k} was defined in (17).

Proof.

Notice that in this case σq=ρ\sigma^{\prime}_{q}=\rho.

If w(X𝝈(qm))w\in{\mathcal{L}}(X^{(q_{m})}_{\boldsymbol{\sigma^{\prime}}}) for which w1𝒜i𝒜i+1w_{1}\in{\mathcal{A}}_{i}\cup{\mathcal{A}}_{i+1}^{\prime}, then wρ(u)w\sqsubseteq\rho(u), where u(X𝝈(qm+1))u\in{\mathcal{L}}(X^{(q_{m+1})}_{\boldsymbol{\sigma^{\prime}}}) and u1𝒜i𝒜i+1u_{1}\in{\mathcal{A}}_{i}\cup{\mathcal{A}}_{i+1}^{\prime}. This is equivalent to the condition η(u1)𝒜i\eta(u_{1})\in{\mathcal{A}}_{i} . Since η(ρ(ai))=η(ρ(ai+1))=τi(ai)\eta(\rho(a_{i}))=\eta(\rho(a_{i+1}^{\prime}))=\tau_{i}(a_{i}) and η(ρ(bi))=η(ρ(bi+1))=τi(bi)\eta(\rho(b_{i}))=\eta(\rho(b_{i+1}^{\prime}))=\tau_{i}(b_{i}), for u(X𝝈(qm+1))u\in{\mathcal{L}}(X^{(q_{m+1})}_{\boldsymbol{\sigma^{\prime}}}) satisfying η(u1)𝒜i\eta(u_{1})\in{\mathcal{A}}_{i}, we deduce that if |u|=k+2|u|=k+2 with η(u1)=η(uk)\eta(u_{1})=\eta(u_{k}), then

rji=j=1j𝟙η(ρ(u1)jj)=η(ρ(uk+2)j)r^{i}_{j}=\sum_{j^{\prime}=1}^{j}\mathds{1}_{\eta(\rho(u_{1})_{\ell-j-j^{\prime}})=\eta(\rho(u_{k+2})_{j^{\prime}})}

and when we consider η(u1)=η(uk+2)¯\eta(u_{1})=\overline{\eta(u_{k+2})}, jrji=j=1j𝟙η(ρ(u¯1)jj)=η(ρ(uk+2)j)\displaystyle j-r^{i}_{j}=\sum_{j^{\prime}=1}^{j}\mathds{1}_{\eta(\rho(\overline{u}_{1})_{\ell-j-j^{\prime}})=\eta(\rho(u_{k+2})_{j^{\prime}})}. If |u|=k+1|u|=k+1 with η(u1)=η(uk)\eta(u_{1})=\eta(u_{k})

r~ji=j=1j𝟙η(ρ(u1)j)=η(ρ(uk+1)j+j)\tilde{r}^{i}_{j}=\sum_{j^{\prime}=1}^{\ell-j}\mathds{1}_{\eta(\rho(u_{1})_{j^{\prime}})=\eta(\rho(u_{k+1})_{j+j^{\prime}})}

and when we consider η(u1)=η(uk+1)¯\eta(u_{1})=\overline{\eta(u_{k+1})}, jr~ji=j=1j𝟙η(ρ(u¯1)j)=η(ρ(uk+1)j+j)\displaystyle\ell-j-\tilde{r}^{i}_{j}=\sum_{j^{\prime}=1}^{\ell-j}\mathds{1}_{\eta(\rho(\overline{u}_{1})_{j^{\prime}})=\eta(\rho(u_{k+1})_{j+j^{\prime}})}.

Thus, the first equality of the lemma is a direct consequence of (10) and the second equality is completely analogous.

Lemma 4.3.

If 𝛔=(σq)qQ\boldsymbol{\sigma}^{\prime}=(\sigma^{\prime}_{q})_{q\in Q} and μ\mu\in{\mathcal{E}}, then for every nn\in{\mathbb{N}}, let q=n+n+1n+2q=n+\frac{n+1}{n+2}, we get

μ(qm)(Dk+ji)=\displaystyle\ell\cdot\mu^{(q_{m})}(D^{i}_{\ell k+j})= rjiμ(qm+1)(Ck+2i)+r~jiμ(qm+1)(Ck+1i)\displaystyle r^{i}_{j}\cdot\mu^{(q_{m+1})}(C^{i}_{k+2})+\tilde{r}^{i}_{j}\cdot\mu^{(q_{m+1})}(C^{i}_{k+1})
+(jrji)μ(qm+1)(C¯k+2i)+(jr~ji)μ(qm+1)(C¯k+1i)\displaystyle+(j-r^{i}_{j})\mu^{(q_{m+1})}(\overline{C}^{i}_{k+2})+(\ell-j-\tilde{r}^{i}_{j})\mu^{(q_{m+1})}(\overline{C}^{i}_{k+1})
μ(qm)(D¯k+ji)=\displaystyle\ell\cdot\mu^{(q_{m})}(\overline{D}^{i}_{\ell k+j})= (jrji)μ(qm+1)(Ck+2i)+(jr~ji)μ(qm+1)(Ck+1i)\displaystyle(j-r^{i}_{j})\mu^{(q_{m+1})}(C^{i}_{k+2})+(\ell-j-\tilde{r}^{i}_{j})\mu^{(q_{m+1})}(C^{i}_{k+1})
+rjiμ(qm+1)(C¯k+2i)+r~jiμ(qm+1)(C¯k+1i)\displaystyle+r^{i}_{j}\cdot\mu^{(q_{m+1})}(\overline{C}^{i}_{k+2})+\tilde{r}^{i}_{j}\cdot\mu^{(q_{m+1})}(\overline{C}^{i}_{k+1})

for j{1,,}j\in\{1,\ldots,\ell\}.

Proof.

Noting σqm=ψ\sigma^{\prime}_{q_{m}}=\psi and that ψ(ai)=ρ(ai+1)\psi(a_{i})=\rho(a_{i+1}^{\prime}) for all i{0,,d1}i\in\{0,\ldots,d-1\}, one can repeat the steps of Lemma 4.2 proof and deduce the formula. ∎

Lemma 4.4.

If 𝛔=(σq)qQ\boldsymbol{\sigma}^{\prime}=(\sigma^{\prime}_{q})_{q\in Q} and μ\mu\in{\mathcal{E}}, then for every qm=nq_{m}=n\in{\mathbb{N}},

μ(n)(Cki)\displaystyle\mu^{(n)}(C^{i}_{k}) μ(qm+1)(Dki)+2n+1\displaystyle\leq\mu^{(q_{m+1})}(D^{i}_{k})+\frac{2}{\ell^{n+1}} (20)
μ(n)(C¯ki)\displaystyle\mu^{(n)}(\overline{C}^{i}_{k}) μ(qm+1)(D¯ki)+2n+1\displaystyle\leq\mu^{(q_{m+1})}(\overline{D}^{i}_{k})+\frac{2}{\ell^{n+1}} (21)
Proof.

Notice that σn=ϕ\sigma^{\prime}_{n}=\phi is letter-to-letter so by Remark 2.6

μ(n)(w)=uϕ1(w)μ(qm+1)(u).\mu^{(n)}(w)=\sum_{u\in\phi^{-1}(w)}\mu^{(q_{m+1})}(u).

The set ϕ1(Cki)\phi^{-1}(C_{k}^{i}) is contained in UUU\cup U^{\prime} where UU is the set of complete words uu with length kk and first letter in 𝒜i{\mathcal{A}}_{i} and UU^{\prime} is the set of words uu with length kk and first or last letter in 𝒜i{\mathcal{A}}_{i}^{\prime}. With that,

μ(n)(Cki)\displaystyle\mu^{(n)}(C_{k}^{i})\leq μ(qm+1)(U)+μ(qm+1)(U)\displaystyle\mu^{(q_{m+1})}(U)+\mu^{(q_{m+1})}(U^{\prime})
\displaystyle\leq μ(qm+1)(Dki)+2(μ(qm+1)(ai)+μ(qm+1)(bi))μ(qm+1)(Dki)+2n+1.\displaystyle\mu^{(q_{m+1})}(D^{i}_{k})+2(\mu^{(q_{m+1})}(a_{i}^{\prime})+\mu^{(q_{m+1})}(b_{i}^{\prime}))\leq\mu^{(q_{m+1})}(D^{i}_{k})+\frac{2}{\ell^{n+1}}.

where the last inequality uses that, by induction, μ(qm+1)(ai)=1n+1μ(n+1)(ai1)12n+1\mu^{(q_{m+1})}(a_{i}^{\prime})=\frac{1}{\ell^{n+1}}\mu^{(n+1)}(a_{i-1})\leq\frac{1}{2\ell^{n+1}}. Likewise, μ(qm+1)(bi)12n+1\mu^{(q_{m+1})}(b_{i}^{\prime})\leq\frac{1}{2\ell^{n+1}}. Inequality (21) uses the same reasoning.

4.3. Upper bounds

Recall the definition of CkiC^{i}_{k}, DkiD^{i}_{k}, C¯ki\overline{C}^{i}_{k} and D¯ki\overline{D}^{i}_{k} given by the equations (16) to (19).

Lemma 4.5.

For every μ\mu\in{\mathcal{E}} nn\in{\mathbb{N}} and k2k\geq 2,

μ(n)(Cki)maxk=2,,qQ,qn{μ(q)(Dki),μ(q)(D¯ki)}+12n+1.\mu^{(n)}(C^{i}_{k})\leq\max_{\begin{subarray}{c}k^{\prime}=2,\ldots,\ell\\ q\in Q,q\geq n\end{subarray}}\{\mu^{(q)}(D^{i}_{k^{\prime}}),\mu^{(q)}(\overline{D}^{i}_{k^{\prime}})\}+\frac{\ell}{\ell-1}\frac{2}{\ell^{n+1}}. (22)
Remark.

Following what we discuss in Section 2.2 in the right hand side, if qq is an integer, μ(q)\mu^{(q)} is supported in Λd\Lambda_{d}^{{\mathbb{Z}}} and therefore it can be studied as a measure in (Λd)(\Lambda_{d}^{\prime})^{{\mathbb{Z}}}. In that context, μ(q)(Dki)=μ(q)(Cki)\mu^{(q)}(D^{i}_{k^{\prime}})=\mu^{(q)}(C^{i}_{k^{\prime}}) and μ(q)(D¯ki)=μ(q)(C¯ki)\mu^{(q)}(\overline{D}^{i}_{k^{\prime}})=\mu^{(q)}(\overline{C}^{i}_{k^{\prime}}), because μ(q)(w)=0\mu^{(q)}(w)=0 whenever ww contains a letter in Λd\Λd\Lambda_{d}^{\prime}\backslash\Lambda_{d}.

Proof.

Combining Lemmas 4.2 and 4.3 we deduce that for qmQ\q_{m}\in Q\backslash{\mathbb{N}}, μ(qm)(Dk+ji)\mu^{(q_{m})}(D^{i}_{\ell k+j}) and μ(qm)(D¯k+ji)\mu^{(q_{m})}(\overline{D}^{i}_{\ell k+j}) are convex combinations of μ(qm+1)(Dk+si)\mu^{(q_{m+1})}(D^{i}_{k+s}) and μ(qm+1)(D¯k+si)\mu^{(q_{m+1})}(\overline{D}^{i}_{k+s}) for s=1,2s=1,2. Therefore, if qmQ\q_{m}\in Q\backslash{\mathbb{N}}

μ(qm)(Dk+ji)maxs=1,2{μ(qm+1)(Dk+si),μ(qm+1)(D¯k+si)}\mu^{(q_{m})}(D^{i}_{\ell k+j})\leq\max_{s=1,2}\{\mu^{(q_{m+1})}(D^{i}_{k+s}),\mu^{(q_{m+1})}(\overline{D}^{i}_{k+s})\}

and the same bound holds for μ(qm)(D¯k+ji)\mu^{(q_{m})}(\overline{D}^{i}_{\ell k+j}). Likewise, using Lemma 4.4 for qmq_{m}\in{\mathbb{N}},

μ(qm)(Dki)\displaystyle\mu^{(q_{m})}(D^{i}_{k}) μ(qm+1)(Dki)+2n+1\displaystyle\leq\mu^{(q_{m+1})}(D^{i}_{k})+\frac{2}{\ell^{n+1}}
μ(qm)(D¯ki)\displaystyle\mu^{(q_{m})}(\overline{D}^{i}_{k}) μ(qm+1)(D¯ki)+2n+1\displaystyle\leq\mu^{(q_{m+1})}(\overline{D}^{i}_{k})+\frac{2}{\ell^{n+1}}

Notice that for 2k2\leq k\leq\ell, the proposition is trivial. Thus, fix k>k>\ell, there exists an integer k1k_{1}\in{\mathbb{N}} and m1{1,,}m_{1}\in\{1,\ldots,\ell\} such that k=k1+m1k=\ell\cdot k_{1}+m_{1}.

Now, take qm=nq_{m}=n\in{\mathbb{N}}, then by the previous inequalities

μ(n)(Cki)\displaystyle\mu^{(n)}(C^{i}_{k}) μ(qm+1)(Dki)+2n+1\displaystyle\leq\mu^{(q_{m+1})}(D^{i}_{k})+\frac{2}{\ell^{n+1}}
μ(qm+1)(Dki)\displaystyle\mu^{(q_{m+1})}(D^{i}_{k}) maxs=1,2{μ(qm+2)(Dk1+si),μ(qm+2)(D¯k1+si)}\displaystyle\leq\max_{s=1,2}\{\mu^{(q_{m+2})}(D^{i}_{k_{1}+s}),\mu^{(q_{m+2})}(\overline{D}^{i}_{k_{1}+s})\}

If k1{1,,2}k_{1}\in\{1,\ldots,\ell-2\} we are done. If k1=1k_{1}=\ell-1, we need to control the values indexed by k1+2=+1k_{1}+2=\ell+1, but for that we need to iterate the argument one more time. Otherwise, that is if k1k_{1}\geq\ell, we can find k21k_{2}\geq 1 and m2{1,,}m_{2}\in\{1,\ldots,\ell\} such that k1+1=k2+m2k_{1}+1=\ell k_{2}+m_{2} (similarly for k1+2=k2+m2+1k_{1}+2=\ell k_{2}+m_{2}+1 or, if m2=m_{2}=\ell, k1+2=(k2+1)+1k_{1}+2=\ell(k_{2}+1)+1). With that decomposition one can bound the right hand side of the second equality by maxs=1,2,3{μ(qm+3)(Dk2+si),μ(qm+3)(D¯k2+si)}\displaystyle\max_{s=1,2,3}\{\mu^{(q_{m+3})}(D^{i}_{k_{2}+s}),\mu^{(q_{m+3})}(\overline{D}^{i}_{k_{2}+s})\}.

Consider the sequence, (kt)t(k_{t})_{t\in{\mathbb{N}}} and (mt)t1(m_{t})_{t\geq 1} such that kt0k_{t}\geq 0 and mt{1,,}m_{t}\in\{1,\ldots,\ell\} and are defined as follow, k0=kk_{0}=k, k0=k1+m1k_{0}=\ell k_{1}+m_{1} and inductively kt=(kt+1+t)+mtk_{t}=\ell(k_{t+1}+t)+m_{t}. Then eventually kt=0k_{t}=0 for some tt\in{\mathbb{N}}. With that, one can iterate the previous argument a finite amount of time and be able to express everything with only values k{2,,}k^{\prime}\in\{2,\ldots,\ell\}. The only problem is when nn¯=qm+tn\leq\overline{n}=q_{m+t}\in{\mathbb{N}} in that case, we are force to add the term 2/n¯+12/\ell^{\overline{n}+1}. So we get

μ(n)(Cki)maxk=2,,qQ,nq<N{μ(q)(Dki),μ(q)(D¯ki)}+2n+1+2n+2++2N\mu^{(n)}(C^{i}_{k})\leq\max_{\begin{subarray}{c}k^{\prime}=2,\ldots,\ell\\ q\in Q,n\leq q<N\end{subarray}}\{\mu^{(q)}(D^{i}_{k^{\prime}}),\mu^{(q)}(\overline{D}^{i}_{k^{\prime}})\}+\frac{2}{\ell^{n+1}}+\frac{2}{\ell^{n+2}}+\cdots+\frac{2}{\ell^{N}}

for some NnN\geq n, but that value is bounded by

maxk=2,,qQ,qn{μ(q)(Dki),μ(q)(D¯ki)}+s12n+s,\max_{\begin{subarray}{c}k^{\prime}=2,\ldots,\ell\\ q\in Q,q\geq n\end{subarray}}\{\mu^{(q)}(D^{i}_{k^{\prime}}),\mu^{(q)}(\overline{D}^{i}_{k^{\prime}})\}+\sum_{s\geq 1}\frac{2}{\ell^{n+s}},

which finish the proof. ∎

Proposition 4.6.

For every i{0,,d1}i\in\{0,\ldots,d-1\},

δμimaxk=2,,{w𝒞𝒜ikνi(w),w𝒞¯𝒜ikνi(w)}\delta_{\mu_{i}}\leq\max_{k=2,\ldots,\ell}\left\{\sum_{w\in{\mathcal{C}}{\mathcal{A}}_{i}^{k}}\nu_{i}(w),\sum_{w\in\overline{{\mathcal{C}}}{\mathcal{A}}_{i}^{k}}\nu_{i}(w)\right\}

where the notation 𝒞𝒜ik{\mathcal{C}}{\mathcal{A}}_{i}^{k} is introduced in (2) and 𝒞¯𝒜ik\overline{{\mathcal{C}}}{\mathcal{A}}^{k}_{i} is the set of words w𝒜iw\in{\mathcal{A}}_{i}^{*} of length kk such that w1=w¯kw_{1}=\overline{w}_{k}

Proof.

First notice that, for every (kt)t(k_{t})_{t\in{\mathbb{N}}} a possibly constant sequence of integers greatest or equal than 22,

limtw𝒞Λdktμi(t)(w)\displaystyle\lim_{t\to\infty}\sum_{w\in{\mathcal{C}}\Lambda_{d}^{k_{t}}}\mu_{i}^{(t)}(w) =limtw𝒞Λdkt,w1𝒜iμi(t)(w)+limtw𝒞Λdkt,w1𝒜iμi(t)(w)\displaystyle=\lim_{t\to\infty}\sum_{w\in{\mathcal{C}}\Lambda_{d}^{k_{t}},w_{1}\in{\mathcal{A}}_{i}}\mu_{i}^{(t)}(w)+\lim_{t\to\infty}\sum_{w\in{\mathcal{C}}\Lambda_{d}^{k_{t}},w_{1}\not\in{\mathcal{A}}_{i}}\mu_{i}^{(t)}(w)
limtμi(t)(Ckti)+limtcΛd\𝒜iμi(t)(c)=limtμi(t)(Ckti)\displaystyle\leq\lim_{t\to\infty}\mu_{i}^{(t)}(C_{k_{t}}^{i})+\lim_{t\to\infty}\sum_{c\in\Lambda_{d}\backslash{\mathcal{A}}_{i}}\mu_{i}^{(t)}(c)=\lim_{t\to\infty}\mu_{i}^{(t)}(C_{k_{t}}^{i})

Therefore, by Theorem 2.3 we get that there exists (kt)t(k_{t})_{t\in{\mathbb{N}}} a possibly constant sequence of integers greatest or equal than 22 such that

δμi\displaystyle\delta_{\mu_{i}} =limtw𝒞Λdktμi(t)(w)limtμi(t)(Ckti)limtmaxk=2,,qQ,qt{μ(q)(Dki),μ(q)(D¯ki)}\displaystyle=\lim_{t\to\infty}\sum_{w\in{\mathcal{C}}\Lambda_{d}^{k_{t}}}\mu_{i}^{(t)}(w)\leq\lim_{t\to\infty}\mu_{i}^{(t)}(C_{k_{t}}^{i})\leq\lim_{t\to\infty}\max_{\begin{subarray}{c}k^{\prime}=2,\ldots,\ell\\ q\in Q,q\geq t\end{subarray}}\{\mu^{(q)}(D^{i}_{k^{\prime}}),\mu^{(q)}(\overline{D}^{i}_{k^{\prime}})\}

where the last inequality is a consequence of (22).

Thus, we only have to control the values of μ(q)(Dki)\mu^{(q)}(D^{i}_{k}) and μ(q)(D¯ki)\mu^{(q)}(\overline{D}^{i}_{k}) for k{2,,}k\in\{2,\ldots,\ell\} and big qQq\in Q. This is already controlled when qq is an integer because, Theorem 3.2 implies that for every ϵ>0\epsilon>0, there exists N1N\geq 1 such that for every nNn\geq N and every word w𝒜iw\in{\mathcal{A}}^{*}_{i}, with |w||w|\leq\ell, μi(n)(w)νi(w)+ε\mu_{i}^{(n)}(w)\leq\nu_{i}(w)+\varepsilon and w𝒜iw\not\in{\mathcal{A}}_{i}^{*}, μi(n)(w)ε2\mu_{i}^{(n)}(w)\leq\frac{\varepsilon}{2}.

Now, fix q=n1+mn1+2q=n_{1}+\frac{m^{\prime}}{n_{1}+2}\not\in{\mathbb{N}} and n1Nn_{1}\geq N , notice that for jij\neq i,

μi(q)(Dkj)c𝒜j𝒜j+1μi(q)(c)μi(n1+1)(aj)+μi(n1+1)(aj)ε.\mu^{(q)}_{i}(D^{j}_{k})\leq\sum_{c\in{\mathcal{A}}_{j}\cup{\mathcal{A}}_{j+1}^{\prime}}\mu^{(q)}_{i}(c)\leq\mu_{i}^{(n_{1}+1)}(a_{j})+\mu_{i}^{(n_{1}+1)}(a_{j})\leq\varepsilon.

If one repeats a proof similar to the one of Theorem 3.2 for the subshift η(X𝝈(q))\eta(X_{\boldsymbol{\sigma}^{\prime}}^{(q)}), we get that for every w𝒜iw\in{\mathcal{A}}^{*}_{i}, with |w||w|\leq\ell, ημi(q)(w)νi(w)+ε\eta_{*}\mu_{i}^{(q)}(w)\leq\nu_{i}(w)+\varepsilon. Noting that, for kk^{\prime}\leq\ell, if wDkiw\in D^{i}_{k^{\prime}} then η(w)𝒞𝒜ik\eta(w)\in{\mathcal{C}}{\mathcal{A}}_{i}^{k^{\prime}} we deduce

μi(q)(Dki)ημi(q)(𝒞𝒜ik)u𝒞𝒜ik(νi(u)+ε)2kε+νi(𝒞𝒜ik).\mu^{(q)}_{i}(D^{i}_{k^{\prime}})\leq\eta_{*}\mu^{(q)}_{i}({\mathcal{C}}{\mathcal{A}}_{i}^{k^{\prime}})\leq\sum_{u\in{\mathcal{C}}{\mathcal{A}}_{i}^{k^{\prime}}}(\nu_{i}(u)+\varepsilon)\leq 2^{k^{\prime}}\varepsilon+\nu_{i}({\mathcal{C}}{\mathcal{A}}_{i}^{k^{\prime}}).

Similarly μi(q)(D¯ki)2kε+νi(𝒞¯𝒜ik)\mu^{(q)}_{i}(\overline{D}^{i}_{k^{\prime}})\leq 2^{k^{\prime}}\varepsilon+\nu_{i}(\overline{{\mathcal{C}}}{\mathcal{A}}_{i}^{k^{\prime}}). Therefore for every ε>0\varepsilon>0 there exists NN, such that for every nNn\geq N

maxk=2,,qQ,qn{μ(q)(Cki),μ(q)(C¯ki)}2ε+maxk=2,,{νi(𝒞𝒜ik),νi(𝒞¯𝒜ik)}\max_{\begin{subarray}{c}k^{\prime}=2,\ldots,\ell\\ q\in Q,q\geq n\end{subarray}}\{\mu^{(q)}(C^{i}_{k^{\prime}}),\mu^{(q)}(\overline{C}^{i}_{k^{\prime}})\}\leq 2^{\ell}\varepsilon+\max_{k=2,\ldots,\ell}\left\{\nu_{i}({\mathcal{C}}{\mathcal{A}}_{i}^{k^{\prime}}),\nu_{i}(\overline{{\mathcal{C}}}{\mathcal{A}}_{i}^{k^{\prime}})\right\}

Thus taking limit nn\to\infty and ε0\varepsilon\to 0 and we conclude. ∎

4.4. System with multiple partial rigidity rates

We use the result of the last section of [13], for that fix L6L\geq 6 and let ζL:𝒜𝒜\zeta_{L}\colon{\mathcal{A}}^{*}\to{\mathcal{A}}^{*} given by

aaLb\displaystyle a\mapsto a^{L}b
bbLa.\displaystyle b\mapsto b^{L}a.

In particular ζL2\zeta_{L}^{2} is a prolongable and mirror morphism.

Proposition 4.7.

[13, Proposition 7.17] Fix L6L\geq 6 and let (XζL,,ν,S)(X_{\zeta_{L}},{\mathcal{B}},\nu,S) be the substitution subshift given by ζL:𝒜𝒜\zeta_{L}\colon{\mathcal{A}}^{*}\to{\mathcal{A}}^{*}, then

δν=ν(aa)+ν(bb)=maxk2{w𝒞𝒜kν(w),w𝒞¯𝒜kν(w)}=L1L+1\delta_{\nu}=\nu(aa)+\nu(bb)=\max_{k\geq 2}\left\{\sum_{w\in{\mathcal{C}}{\mathcal{A}}^{k}}\nu(w),\sum_{w\in\overline{{\mathcal{C}}}{\mathcal{A}}^{k}}\nu(w)\right\}=\frac{L-1}{L+1}

Now we can give a detailed version of Theorem 1.1 stated in the introduction. For that, as for Corollary 3.5, we write ζL:𝒜i𝒜i\zeta_{L}\colon{\mathcal{A}}_{i}^{*}\to{\mathcal{A}}_{i}^{*} even if it is originally define in the alphabet 𝒜{\mathcal{A}}.

Theorem 4.8.

For L6L\geq 6, let 𝛔\boldsymbol{\sigma} be the directive sequence of glued substitutions 𝛔=(Γ(ζL2i+1(n+1)2di:i=0,,d1))n\boldsymbol{\sigma}=(\Gamma(\zeta_{L^{2^{i+1}}}^{(n+1)2^{d-i}}\colon i=0,\ldots,d-1))_{n\in{\mathbb{N}}}. That is

σn(ai)=κ(ζL2i+1(n+1)2di(ai))σn(bi)=κ(ζL2i+1(n+1)2di(bi)) for i{0,,d1}.\begin{array}[]{cc}\sigma_{n}(a_{i})&=\kappa(\zeta_{L^{2^{i+1}}}^{(n+1)2^{d-i}}(a_{i}))\\ \sigma_{n}(b_{i})&=\kappa(\zeta_{L^{2^{i+1}}}^{(n+1)2^{d-i}}(b_{i}))\end{array}\quad\text{ for }i\in\{0,\ldots,d-1\}.

Then,

δμi=L2i+11L2i+1+1\delta_{\mu_{i}}=\frac{L^{2^{i+1}}-1}{L^{2^{i+1}}+1} (23)

and the rigidity sequence is (h(n))n(h^{(n)})_{n\in{\mathbb{N}}}.

Remark.

The directive sequence 𝝈\boldsymbol{\sigma} in the statement fullfils all the hypothesis of Theorem 3.2 with τi=ζL2i+12di\tau_{i}=\zeta_{L^{2^{i+1}}}^{2^{d-i}} for i=0,,d1i=0,\ldots,d-1. In particular, |ζL2i+12di(ai)|=(L2i+1)2di=L2i+12di=L2d+1|\zeta_{L^{2^{i+1}}}^{2^{d-i}}(a_{i})|=(L^{2^{i+1}})^{2^{d-i}}=L^{2^{i+1}\cdot 2^{d-i}}=L^{2^{d+1}}, therefore the morphism σn\sigma_{n} is indeed of constant length, for all nn\in{\mathbb{N}}. In this case h(n)=(L2d+1)n+1(L2d+1)nL2d+1h^{(n)}=(L^{2^{d+1}})^{n+1}\cdot(L^{2^{d+1}})^{n}\cdots L^{2^{d+1}}.

Proof.

By Proposition 4.7

maxk=2,,L2d+1{ν(𝒞𝒜ik),ν(𝒞¯𝒜ik)}=νi(aiai)+νi(bibi)=L2i+11L2i+1+1=δνi.\max_{k=2,\ldots,L^{2^{d+1}}}\left\{\nu({\mathcal{C}}{\mathcal{A}}_{i}^{k}),\nu(\overline{{\mathcal{C}}}{\mathcal{A}}_{i}^{k})\right\}=\nu_{i}(a_{i}a_{i})+\nu_{i}(b_{i}b_{i})=\frac{L^{2^{i+1}}-1}{L^{2^{i+1}}+1}=\delta_{\nu_{i}}.

Therefore, by Corollary 3.4 and Proposition 4.6, δμi=δνi\displaystyle\delta_{\mu_{i}}=\delta_{\nu_{i}}, concluding (23). Since

limnj=0d1μi(n)(ajaj)+μi(n)(bjbj)=limnμi(n)(aiai)+μi(n)(bibi)=δμi,\lim_{n\to\infty}\sum_{j=0}^{d-1}\mu_{i}^{(n)}(a_{j}a_{j})+\mu_{i}^{(n)}(b_{j}b_{j})=\lim_{n\to\infty}\mu_{i}^{(n)}(a_{i}a_{i})+\mu_{i}^{(n)}(b_{i}b_{i})=\delta_{\mu_{i}},

by Theorem 2.3, the partial rigidity sequence is given by (h(n))n(h^{(n)})_{n\in{\mathbb{N}}}. ∎

Remark.

The construction in Theorem 4.8 relies on the fact that the family of substitutions ζL\zeta_{L}, for L6L\geq 6, had been previously studied. A similar result could be achieved including ζ2\zeta_{2}, which corresponds to the Thue-Morse substitution, also studied in [13], but in that case the partial rigidity sequence for its corresponding ergodic measure would have been (3h(n))n(3\cdot h^{(n)})_{n\in{\mathbb{N}}} instead of (h(n))n(h^{(n)})_{n\in{\mathbb{N}}}. In general, studying the partial rigidity rates of more substitution subshifts should allow us to construct more examples like the one in Theorem 4.8. In particular, it would be interesting to construct systems with algebraically independent partial rigidity rates, but even in the uniquely ergodic case, there is no example in the literature of a system with an irrational partial rigidity rate.

We also notice that with the gluing technique introduced in Theorem 3.2 one can only build constant-length 𝒮{\mathcal{S}}-adic subshift, which have non-trivial rational eigenvalues, that is m2m\geq 2 and 1k<m1\leq k<m such that for some non-zero function fL2(X,μ)f\in L^{2}(X,\mu), fS=e2πik/mff\circ S=e^{2\pi ik/m}f. Thus, a refinement of Theorem 1.1 would be to construct a minimal system with distinct weak-mixing measures and distinct partial rigidity rates. To construct an explicit example following a similar approach to the one outlined in this paper, it would be necessary to use a non-constant-length directive sequence and then being forced to use the general formula for the partial rigidity rate from [13, Theorem B]. Additionally, the equation (9) no longer holds in the non-constant-length case.

References

  • [1] N. Bédaride, A. Hilion, and M. Lustig. The measure transfer for subshifts induced by a morphism of free monoids. arXiv preprint arXiv:2211.11234, 2022.
  • [2] N. Bédaride, A. Hilion, and M. Lustig. Measure transfer and s-adic developments for subshifts. Ergodic Theory and Dynamical Systems, pages 1–35, 2023.
  • [3] V. Bergelson, A. del Junco, M. Lemańczyk, and J. Rosenblatt. Rigidity and non-recurrence along sequences. Ergodic Theory Dynam. Systems, 34(5):1464–1502, 2014.
  • [4] V. Berthé, W. Steiner, J. M. Thuswaldner, and R. Yassawi. Recognizability for sequences of morphisms. Ergodic Theory Dynam. Systems, 39(11):2896–2931, 2019.
  • [5] S. Bezuglyi, O. Karpel, and J. Kwiatkowski. Exact number of ergodic invariant measures for bratteli diagrams. Journal of Mathematical Analysis and Applications, 480(2):123431, 2019.
  • [6] S. Bezuglyi, J. Kwiatkowski, K. Medynets, and B. Solomyak. Finite rank Bratteli diagrams: structure of invariant measures. Trans. Amer. Math. Soc., 365(5):2637–2679, 2013.
  • [7] A. Coronel, A. Maass, and S. Shao. Sequence entropy and rigid σ\sigma-algebras. Studia Math., 194(3):207–230, 2009.
  • [8] M. I. Cortez, F. Durand, B. Host, and A. Maass. Continuous and measurable eigenfunctions of linearly recurrent dynamical Cantor systems. J. London Math. Soc. (2), 67(3):790–804, 2003.
  • [9] D. Creutz. Measure-theoretically mixing subshifts of minimal word complexity. 2023.
  • [10] A. I. Danilenko. Actions of finite rank: weak rational ergodicity and partial rigidity. Ergodic Theory Dynam. Systems, 36(7):2138–2171, 2016.
  • [11] F. M. Dekking and M. Keane. Mixing properties of substitutions. Z. Wahrscheinlichkeitstheorie und Verw. Gebiete, 42(1):23–33, 1978.
  • [12] S. Donoso, F. Durand, A. Maass, and S. Petite. Interplay between finite topological rank minimal Cantor systems, 𝒮\mathcal{S}-adic subshifts and their complexity. Trans. Amer. Math. Soc., 374(5):3453–3489, 2021.
  • [13] S. Donoso, A. Maass, and T. Radić. On partial rigidity of 𝒮\mathcal{S}-adic subshifts. arXiv preprint arXiv:2312.12406, 2023.
  • [14] S. Donoso and S. Shao. Uniformly rigid models for rigid actions. Studia Math., 236(1):13–31, 2017.
  • [15] F. Durand. Linearly recurrent subshifts have a finite number of non-periodic subshift factors. Ergodic Theory Dynam. Systems, 20:1061–1078, 2000.
  • [16] F. Durand and D. Perrin. Dimension groups and dynamical systems—substitutions, Bratteli diagrams and Cantor systems, volume 196 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2022.
  • [17] B. Fayad and A. Kanigowski. Rigidity times for a weakly mixing dynamical system which are not rigidity times for any irrational rotation. Ergodic Theory Dynam. Systems, 35(8):2529–2534, 2015.
  • [18] N. A. Friedman. Partial mixing, partial rigidity, and factors. In Measure and measurable dynamics (Rochester, NY, 1987), volume 94 of Contemp. Math., pages 141–145. Amer. Math. Soc., Providence, RI, 1989.
  • [19] H. Furstenberg. Recurrence in ergodic theory and combinatorial number theory. Princeton University Press, Princeton, N.J., 1981. M. B. Porter Lectures.
  • [20] H. Furstenberg and B. Weiss. The finite multipliers of infinite ergodic transformations. In The structure of attractors in dynamical systems (Proc. Conf., North Dakota State Univ., Fargo, N.D., 1977), volume 668 of Lecture Notes in Math., pages 127–132. Springer, Berlin, 1978.
  • [21] S. Glasner and D. Maon. Rigidity in topological dynamics. Ergodic Theory Dynam. Systems, 9(2):309–320, 1989.
  • [22] G. R. Goodson and V. V. Ryzhikov. Conjugations, joinings, and direct products of locally rank one dynamical systems. J. Dynam. Control Systems, 3(3):321–341, 1997.
  • [23] A. Katok. Interval exchange transformations and some special flows are not mixing. Israel J. Math., 35(4):301–310, 1980.
  • [24] J. L. King. Joining-rank and the structure of finite rank mixing transformations. J. Analyse Math., 51:182–227, 1988.
  • [25] M. Queffélec. Substitution dynamical systems—spectral analysis, volume 1294 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1987.