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

Multiplicity of topological systems

David Burguet Sorbonne Universite, LPSM, 75005 Paris, France david.burguet@upmc.fr  and  Ruxi Shi Sorbonne Universite, LPSM, 75005 Paris, France ruxi.shi@upmc.fr
Abstract.

We define the topological multiplicity of an invertible topological system (X,T)(X,T) as the minimal number kk of real continuous functions f1,,fkf_{1},\cdots,f_{k} such that the functions fiTnf_{i}\circ T^{n}, nn\in\mathbb{Z}, 1ik,1\leq i\leq k, span a dense linear vector space in the space of real continuous functions on XX endowed with the supremum norm. We study some properties of topological systems with finite multiplicity. After giving some examples, we investigate the multiplicity of subshifts with linear growth complexity.

2020 Mathematics Subject Classification:

1. Introduction

The multiplicity of an invertible bounded operator U:EU:E\circlearrowleft on a normed vector space EE is the minimal cardinality of subsets FEF\subset E, whose cyclic space (i.e. the vector space spanned by UkxU^{k}x, kk\in\mathbb{Z}, xFx\in F) is dense in EE.

For an ergodic measure preserving system (X,f,,μ)(X,f,\mathcal{B},\mu), the multiplicity Mult(μ)\text{\rm Mult}(\mu) of the Koopman operator, which is the operator of composition by ff on the Hilbert space L2(μ)L^{2}(\mu) is a dynamical invariant, which has been investigated in many works (see e.g. [Dan13] and the references therein).

Cyclicity, which corresponds to simple multiplicity (i.e. there is an element whose cyclic space is dense in the whole vector space), has been also established for operators of composition on the Hardy space H2(D)H^{2}(D) [BS97]. In this context a pioneering work of Birkhoff [Bir29] states that there is an entire function ϕ\phi in the complex plane such that the set {ϕ(+n),n}\{\phi(\cdot+n),\ n\in\mathbb{N}\} is dense itself in the set of entire functions endowed with the uniform topology on compact subsets, i.e. the operator of translation by 11 is hypercyclic.

Quite surprisingly the corresponding topological invariant has not been studied in full generality. More precisely we consider here topological dynamical systems (X,T)(X,T), where XX is a compact metrizable space and T:XT:X\circlearrowleft is a homeomorphism and we study the operator of composition by TT on the Banach space C(X)C(X) of real continuous functions endowed with the uniform topology. We call topological multiplicity of (X,T)(X,T) the associated multiplicity and we denote it by Mult(T)\text{\rm Mult}(T). We remark that our definitions and results can be extended to the noninvertible continuous map T:XT:X\circlearrowleft. But for sake of simplicity, we focus on homeomorphisms T:XT:X\circlearrowleft.

In this paper, we mostly focus on topological systems with finite multiplicity. We first show the following properties for such systems.

Theorem.

Let (X,T)(X,T) be a topological system with finite multiplicity. Then the following properties are satisfied.

  1. (1)

    (X,T)(X,T) has zero topological entropy.

  2. (2)

    (X,T)(X,T) has finitely many ergodic measures.

These properties are the main contents of Section 2. The property (1) is proven in Proposition 3.5 in two ways: one uses the variational principal of topological entropy; the other is purely topological. The property (2) is proven in Lemma 2.6 and Corollary 2.8. In fact we show more precisely that the number of ergodic measures is equal to the multiplicity of the operator induced on the quotient of C(X)C(X) by the closure of coboundaries.

In Section 3, we relate the topological multiplicity with the dimension of cubical shifts, in which the action T:(X)T_{*}:\mathcal{M}(X)\circlearrowleft induced by TT on the set (X)\mathcal{M}(X) of Borel probability measures on XX may be affinely embedded. In Theorem 3.3, we show a necessary and sufficient condition for the existence of affinely embedding of ((X),T)(\mathcal{M}(X),T_{*}) to the shift on ([0,1]d)([0,1]^{d})^{\mathbb{Z}}. Furthermore, we compare our result to Lindenstrauss-Tsukamoto conjecture for dynamical embedding (Corollary 3.4).

In Section 4, we state a generalized Banach version of a lemma due to Baxter [Bax71] which is a classical criterion of simplicity for ergodic transformations. The generalized Baxter’s Lemma (Lemma 4.1) will play an important role on estimating the topological multiplicity in next sections.

For minimal Cantor systems a topological analogue of the rank of a measure preserving system has been defined and studied (see [DP22]). In Section 5, under this setting, we compare the topological multiplicity with the topological rank (Theorem 5.1).

In Section 6, we study some examples and estimate their topological multiplicity: minimal rotations on compact groups, Sturmian and substitution subshifts, homeomorphisms of the interval, etc.

In Section 7, we estimate the topological multiplicity of subshifts with linear growth complexity, i.e. subshifts XX such that the cardinality pX(n)p_{X}(n) of nn-words in XX satisfies lim infnpX(n)n<+\liminf_{n\to\infty}\frac{p_{X}(n)}{n}<+\infty. Such subshifts aroused a great deal of interest, specially recently [Bos92, CK19, CP23, DDMP21]. In [Bos92] it is proved that an aperiodic subshift XX has at most kk ergodic measures if lim infnpX(n)nk\liminf_{n}\frac{p_{X}(n)}{n}\leq k\in\mathbb{N}. Our main related result states as follows (Theorem 7.1 and Theorem 7.6):

Theorem.

Let XX be an aperiodic subshift with lim infnpX(n)nk\liminf_{n\to\infty}\frac{p_{X}(n)}{n}\leq k\in\mathbb{N}. Then

Mult(T)2k and μergodicMult(μ)2k.\text{\rm Mult}(T)\leq 2k\text{ and }\sum_{\mu\ ergodic}\text{\rm Mult}(\mu)\leq 2k.

Except the results on multiplicity that we investigate, we propose several questions in the current paper.

2. Topological multiplicity, definition and first properties

2.1. Multiplicity of a linear operator

Let (E,)(E,\|\cdot\|) be a normed vector space over \mathbb{R}. We consider a linear invertible bounded operator U:EU:E\circlearrowleft. A subset FF of EE is called a generating family of UU when the vector space spanned by UkxU^{k}x, kk\in\mathbb{Z}, xFx\in F, is dense in EE. In the following we denote by span(G)\text{span}(G) (resp. span¯(G)\overline{\text{span}}(G)) the vector space spanned by a subset GG of EE (resp. its closure) and we then let VFU:=span¯{Ukx:k,xF}V^{U}_{F}:=\overline{\text{span}}\{U^{k}x:k\in\mathbb{Z},x\in F\}. Sometimes, we write VFV_{F} instead of VFUV^{U}_{F} whenever the operator is fixed. The multiplicity Mult(U){}\text{\rm Mult}(U)\in\mathbb{N}\cup\{\infty\} of UU is then the smallest cardinality of generating families of UU. By convention we let Mult(U)=0\text{\rm Mult}(U)=0 when EE is reduced to {0}\{0\}. A linear operator with multiplicity one is called cyclic.

We first study the equivariant map between two normed vector spaces with linear invertible bounded operators.

Lemma 2.1.

Let Ui:EiU_{i}:E_{i}\circlearrowleft, i=1,2i=1,2 be two linear invertible bounded operators. Assume that there is a linear bounded operator W:E1E2W:E_{1}\rightarrow E_{2} satisfying WU1=U2WW\circ U_{1}=U_{2}\circ W then

Mult(U2|Im(W)¯)Mult(U1).\text{\rm Mult}(U_{2}|_{\overline{\mathrm{Im}(W)}})\leq\text{\rm Mult}(U_{1}).
Proof.

One checks easily that if FF is a generating family for U1U_{1} then W(F)W(F) is a generating family for the restriction of U2U_{2} to the closure of the image of WW. Therefore Mult(U2|Im(W)¯)Mult(U1)\text{\rm Mult}(U_{2}|_{\overline{\mathrm{Im}(W)}})\leq\text{\rm Mult}(U_{1}). ∎

A direct consequence of Lemma 2.1 is that the multiplicity is a spectral invariant : if UiU_{i} are linear invertible operators on EiE_{i} ,i=1,2,i=1,2, satisfying WU1=U2WW\circ U_{1}=U_{2}\circ W for some invertible bounded linear operator W:E1E2W:E_{1}\rightarrow E_{2}, then U1U_{1} and U2U_{2} have the same multiplicities.

When EE^{\prime} is a closed subspace of EE we endow the quotient E/EE/E^{\prime} space with the norm u¯=inf{u+v,vE}\|\overline{u}\|^{\prime}=\inf\{\|u+v\|,\ v\in E^{\prime}\}. If EE^{\prime} is invariant by UU we let UE/EU^{E/E^{\prime}} be the action induced by UU on the quotient normed space E/EE/E^{\prime}. In this context, by applying Lemma 2.1 with W:EE/EW:E\rightarrow E/E^{\prime} being the natural projection, we get

(2\cdot1) Mult(UE/E)Mult(U).\displaystyle\text{\rm Mult}(U^{E/E^{\prime}})\leq\text{\rm Mult}(U).

2.2. Operator of composition : Topological and ergodic multiplicities.

Ergodic theory focuses on the study of invertible measure preserving systems (X,f,,μ)(X,f,\mathcal{B},\mu). In particular the spectral properties of the unitary operator Uf:L2(μ)U_{f}:L^{2}(\mu)\circlearrowleft, ϕϕf\phi\mapsto\phi\circ f, are investigated. We let f2:=(X|f(x)|2𝑑μ)1/2\|f\|_{2}:=(\int_{X}|f(x)|^{2}d\mu)^{1/2} be the L2L^{2}-norm of fL2(μ)f\in L^{2}(\mu).

Definition 2.2.

The ergodic multiplicity Mult(μ)\text{\rm Mult}(\mu) of an ergodic system (X,f,,μ)(X,f,\mathcal{B},\mu) is the multiplicity of the restriction of UfU_{f} to the Hilbert space L02(μ):={fL2(μ),f𝑑μ=0}L_{0}^{2}(\mu):=\{f\in L^{2}(\mu),\ \int f\,d\mu=0\}, that is to say, Mult(μ)=Mult(Uf)\text{\rm Mult}(\mu)=\text{\rm Mult}(U_{f}).

This quantity has been intensely studied in ergodic theory (see Danilenko’s survey [Dan13]).

Next we consider here an invertible topological dynamical system (X,T)(X,T), i.e. T:XT:X\circlearrowleft is a homeomorphism of a compact metric space XX. We denote by C(X)C(X) the Banach space of real continuous functions endowed with the topology of uniform convergence. We let f:=supxX|f(x)|\|f\|_{\infty}:=\sup_{x\in X}|f(x)| be the supremum norm of fC(X)f\in C(X).

Definition 2.3.

The topological multiplicity Mult(T)\text{\rm Mult}(T) of (X,T)(X,T) is the multiplicity of the operator of composition UT:C(X)U_{T}:C(X)\circlearrowleft, ϕϕT\phi\mapsto\phi\circ T.

Quite surprisingly this last notion seems to be new (note however that cyclicity of UTU_{T} has already been investigated in some cases). Let us first observe that the topological multiplicity bounds from above the ergodic multiplicity of ergodic TT-invariant measures.

Lemma 2.4.

Let (X,T)(X,T) be an invertible topological dynamical system. For any ergodic TT-invariant measure μ\mu, we have

Mult(μ)Mult(T).\text{\rm Mult}(\mu)\leq\text{\rm Mult}(T).
Proof.

Let FF be a generating family with minimal cardinality of UT:C(X)U_{T}:C(X)\circlearrowleft. Then the vector space spanned by FF is dense in (C(X),)(C(X),\|\cdot\|_{\infty}), therefore in (L2(μ),2)(L^{2}(\mu),\|\cdot\|_{2}). As p:L2(μ)L02(μ)p:L^{2}(\mu)\rightarrow L^{2}_{0}(\mu), fff𝑑μf\mapsto f-\int f\ d\mu is continuous and pUT=UTpp\circ U_{T}=U_{T}\circ p, the vector space spanned by p(F)p(F) is dense in L02(μ)L^{2}_{0}(\mu).

Let (X)\mathcal{M}(X) be the set of Borel probability measures endowed with the weak-* topology. It is standard that (X)\mathcal{M}(X) is a compact metrizable space. The compact subset (X,T)(X)\mathcal{M}(X,T)\subset\mathcal{M}(X) of Borel TT-invariant probability measures of (X,T)(X,T) is a simplex, whose extreme set is given by the subset e(X,T)\mathcal{M}_{e}(X,T) of ergodic measures. A topological system with a unique (ergodic) invariant measure is said to be uniquely ergodic. Jewett-Krieger theorem states that every ergodic system has a uniquely ergodic model. Several proofs have been given of this theorem, e.g. see [DGS06, Section 29]. One may wonder if the multiplicity may be preserved:

Question 2.5.

Given an ergodic system with measure μ\mu, is there a uniquely ergodic model (X,T)(X,T) of it such that Mult(T)=Mult(μ)\text{\rm Mult}(T)=\text{\rm Mult}(\mu)?

2.3. The number of ergodic measures as a multiplicity.

Let (X,T)(X,T) be an invertible topological dynamical system. A function ψC(X)\psi\in C(X) is a called a continuous TT-coboundary, if ψ\psi is equal to ϕTϕ\phi\circ T-\phi for some ϕC(X)\phi\in C(X). In other terms the set BT(X)B_{T}(X) of continuous TT-coboundaries is the image of UTIdU_{T}-\text{\rm Id}, in particular it is a vector space. Observe that UT(BT(X))=BT(X)U_{T}(B_{T}(X))=B_{T}(X). To simplify the notations we write U~T\tilde{U}_{T} for the action induced by UTU_{T} on the quotient Banach space C(X)/BT(X)¯C(X)/\overline{B_{T}(X)} and U¯T\underline{U}_{T} for the restriction of UTU_{T} to the closure BT(X)¯\overline{B_{T}(X)} of continuous coboundaries. By a standard application of Hahn-Banach theorem (see e.g. Proposition 2.13 in [Kat01]), a function ψ\psi belongs to BT(X)¯\overline{B_{T}(X)} if and only if ψ𝑑μ=0\int\psi\,d\mu=0 for any μ(X,T)\mu\in\mathcal{M}(X,T) (resp. μe(X,T)\mu\in\mathcal{M}_{e}(X,T)). It is well-known that unique ergodicity is equivalent to the decomposition C(X)=𝟙BT(X)¯C(X)=\mathbb{R}\mathbb{1}\oplus\overline{B_{T}(X)} (see e.g. Lemma 1 in [LV97]), where 𝟙\mathbb{1} denotes the constant function equal to 11. In particular in the case of unique ergodicity, we have C(X)/BT(X)¯𝟙C(X)/\overline{B_{T}(X)}\simeq\mathbb{R}\mathbb{1} and therefore Mult(U~T)=1\text{\rm Mult}(\tilde{U}_{T})=1. It may be generalized as follows.

Lemma 2.6.

Let (X,T)(X,T) be an invertible topological dynamical system. We have

Mult(U~T)=e(X,T).\text{\rm Mult}(\tilde{U}_{T})=\sharp\mathcal{M}_{e}(X,T).
Proof.

We first show that Mult(U~T)e(X,T)\text{\rm Mult}(\tilde{U}_{T})\geq\sharp\mathcal{M}_{e}(X,T). Assume that :

  • ν1,,νp\nu_{1},\cdots,\nu_{p} are distinct ergodic measures,

  • F¯={f1¯,,fq¯}C(X)/BT(X)¯\overline{F}=\{\overline{f_{1}},\cdots,\overline{f_{q}}\}\in C(X)/\overline{B_{T}(X)} is a generating family of U~T\tilde{U}_{T}.

For 1lq1\leq l\leq q, let flC(X)f_{l}\in C(X) be a function (a priori not unique) such that fl¯=flmodBT(X)¯\overline{f_{l}}=f_{l}\mod\overline{B_{T}(X)}. If q<pq<p then the pp vectors

Xi=(fl𝑑νi)l=1,2,,q,i=1,2,,pX_{i}=\left(\int f_{l}\,d\nu_{i}\right)_{l=1,2,\cdots,q},i=1,2,\cdots,p

are linearly dependent in q\mathbb{R}^{q}, i.e. there is (ci)1ipp(0,0,,0)(c_{i})_{1\leq i\leq p}\in\mathbb{R}^{p}\setminus(0,0,\dots,0) such that

(2\cdot2) 1ipciXi\displaystyle\sum_{1\leq i\leq p}c_{i}X_{i} =0.\displaystyle=0.

Let ν\nu be the signed measure ν=1ipciνi\nu=\sum_{1\leq i\leq p}c_{i}\nu_{i}. Then Equality (2\cdot2) may be rewritten as follows:

1lq,fl𝑑ν=0.\forall 1\leq l\leq q,\ \int f_{l}\,d\nu=0.

The measures νi\nu_{i} being invariant for 1ip{1\leq i\leq p}, so is ν\nu. Therefore we get

(2\cdot3) 1lq,k,flTk𝑑ν=0.\forall 1\leq l\leq q,\forall k\in\mathbb{Z},\ \int f_{l}\circ T^{k}\,d\nu=0.

But VF¯U~T=C(X)/BT(X)¯V^{\tilde{U}_{T}}_{\overline{F}}=C(X)/\overline{B_{T}(X)}, so that for any ϵ>0\epsilon>0 and for any gC(X)g\in C(X), we may find hspan(flTk, 1lq,k)h\in\mathrm{span}(f_{l}\circ T^{k},\ 1\leq l\leq q,k\in\mathbb{Z}) and uBT(X)u\in B_{T}(X) with g(h+u)<ϵ\|g-(h+u)\|_{\infty}<\epsilon. By 2\cdot3 we have h𝑑ν=0\int h\,d\nu=0. As uu is a coboundary, we have also u𝑑ν=0\int u\,d\nu=0. Therefore

|g𝑑ν|\displaystyle\left|\int g\,d\nu\right| |(h+u)𝑑ν|+g(h+u)<ϵ.\displaystyle\leq\left|\int(h+u)\,d\nu\right|+\|g-(h+u)\|_{\infty}<\epsilon.

Since ϵ>0\epsilon>0 and gC(X)g\in C(X) are chosen arbitrarily, we obtain g𝑑ν=0\int g\,d\nu=0, for any gC(X)g\in C(X), therefore ν=0\nu=0. This contradicts the ergodicity of the measures νi\nu_{i} for 1ip1\leq i\leq p. Consequently we have qpq\geq p and therefore Mult(U~T)e(X,T)\text{\rm Mult}(\tilde{U}_{T})\geq\sharp\mathcal{M}_{e}(X,T).

Let us show now the converse inequality. Without loss of generality we may assume that p=e(X,T)<Mult(U~T)=q<p=\sharp\mathcal{M}_{e}(X,T)<\text{\rm Mult}(\tilde{U}_{T})=q<\infty. We let again:

  • e(X,T)={ν1,,νp}\mathcal{M}_{e}(X,T)=\{\nu_{1},\cdots,\nu_{p}\},

  • F¯=f1¯,,fq¯C(X)/BT(X)¯\overline{F}={\overline{f_{1}},\cdots,\overline{f_{q}}}\in C(X)/\overline{B_{T}(X)} a generating family of U~T\tilde{U}_{T} with minimal cardinality.

Then the qq vectors

Yl=(fl𝑑νi)i=1,,p,l=1,,q,Y_{l}=\left(\int f_{l}\,d\nu_{i}\right)_{i=1,\cdots,p},\ l=1,\cdots,q,

are linearly dependent in p\mathbb{R}^{p}, i.e. there is (cl)1lqq(0,0,0)(c_{l})_{1\leq l\leq q}\in\mathbb{R}^{q}\setminus(0,0\dots,0) such that

1lqclYl=0.\sum_{1\leq l\leq q}c_{l}Y_{l}=0.

Let gg be the function g=1lqclflg=\sum_{1\leq l\leq q}c_{l}f_{l}. Then we have

g𝑑νi=0,1ip,\int g\,d\nu_{i}=0,\ \forall 1\leq i\leq p,

A previously mentioned, it implies that gg lies in BT(X)¯\overline{B_{T}(X)}. This contradicts the minimality of the generating family F¯\overline{F}. ∎

Remark 2.7.

It follows from the proof of Lemma 2.6 that if e(X,T)={ν1,,νp}\mathcal{M}_{e}(X,T)=\{\nu_{1},\cdots,\nu_{p}\}, then f1¯,,fp¯{\overline{f_{1}},\cdots,\overline{f_{p}}} is a generating family of U~T\tilde{U}_{T} if and only if the matrix A=(fj𝑑νi)1i,jpMp()A=(\int f_{j}\,d\nu_{i})_{1\leq i,j\leq p}\in M_{p}(\mathbb{R}) is invertible.

By inequality (2\cdot1) and Lemma 2.6 we get:

Corollary 2.8.
e(X,T)Mult(T).\sharp\mathcal{M}_{e}(X,T)\leq\text{\rm Mult}(T).

2.4. Relating Mult(T)\text{\rm Mult}(T) and Mult(U¯T)\text{\rm Mult}(\underline{U}_{T})

It follows from definition of BT(X)¯\overline{B_{T}(X)} that the map

W:C(X)BT(X)¯,ffTfW:C(X)\rightarrow\overline{B_{T}(X)},\ f\mapsto f\circ T-f

has dense image and commutes with UTU_{T}. By applying Lemma 2.1 with U1=U2=UU_{1}=U_{2}=U and E1=C(X)E_{1}=C(X), E2=BT(X)¯E_{2}=\overline{B_{T}(X)}, we obtain Mult(U¯T)Mult(T)\text{\rm Mult}(\underline{U}_{T})\leq\text{\rm Mult}(T).

We show then in this subsection the following inequality.

Proposition 2.9.
Mult(T)Mult(UT~)+Mult(U¯T)1.\text{\rm Mult}(T)\leq\text{\rm Mult}(\tilde{U_{T}})+\text{\rm Mult}(\underline{U}_{T})-1.

In particular if (X,T)(X,T) is uniquely ergodic, Mult(T)=Mult(U¯T)\text{\rm Mult}(T)=\text{\rm Mult}(\underline{U}_{T}) by Lemma 2.6. Let us now prove Proposition 2.9. For a family FF of C(X)C(X), we write F¯\overline{F} the subset of C(X)/BT(X)¯C(X)/\overline{B_{T}(X)} consisting of f¯=fmodBT(X)¯\overline{f}=f\mod\overline{B_{T}(X)} for fFf\in F. We start with a technical lemma.

Lemma 2.10.

Let (X,T)(X,T) be an invertible dynamical system with e(X,T)<\sharp\mathcal{M}_{e}(X,T)<\infty. If FF is a family of C(X)C(X) such that F¯\overline{F} is generating for U~T\tilde{U}_{T}, then the constant function 𝟙\mathbb{1} belongs to VFV_{F}.

Proof.

Let e(X,T)={ν1,,νp}\mathcal{M}_{e}(X,T)=\{\nu_{1},\cdots,\nu_{p}\}. By Remark 2.7 the matrix (fi𝑑νj)1i,jp(\int f_{i}\,d\nu_{j})_{1\leq i,j\leq p} is invertible. Then by replacing F={f1,,fp}F=\{f_{1},\cdots,f_{p}\} by some invertible linear combinations we can assume

1i,jp,fi𝑑νj=δi,j,\forall 1\leq i,j\leq p,\ \int f_{i}\,d\nu_{j}=\delta_{i,j},

where δi,j\delta_{i,j} is equal to 11 if i=ji=j and 0 otherwise. Let f=i=1pfif=\sum_{i=1}^{p}f_{i}. We have

1jp,\displaystyle\forall 1\leq j\leq p, f𝑑νj=iδi,j=1,\displaystyle\int f\,d\nu_{j}=\sum_{i}\delta_{i,j}=1,
(2\cdot4) therefore, ν(X,T),\displaystyle\text{therefore, }\forall\nu\in\mathcal{M}(X,T), f𝑑ν=1.\displaystyle\int f\,d\nu=1.

We claim that 1Nn=0N1fTn\frac{1}{N}\sum_{n=0}^{N-1}f\circ T^{n} is converging uniformly to 𝟙\mathbb{1} as NN goes to infinity. If not, there would exist a positive number ϵ\epsilon, a sequence (xk)k1(x_{k})_{k\geq 1} and an increasing sequence (Nk)k1(N_{k})_{k\geq 1} of positive integers such that

(2\cdot5) |1Nkn=0Nk1f(Tn(xk))1|>ϵ,k1.\left|\frac{1}{N_{k}}\sum_{n=0}^{N_{k}-1}f(T^{n}(x_{k}))-1\right|>\epsilon,\ \forall k\geq 1.

After passing to a subsequence of (Nk)k1(N_{k})_{k\geq 1}, we might assume that 1Nkn=0Nk1δTn(xk)\frac{1}{N_{k}}\sum_{n=0}^{N_{k}-1}\delta_{T^{n}(x_{k})} is converging to a TT-invariant measure μ\mu in the weak-* topology. It follows from (2\cdot5) that

|f𝑑μ1|>ϵ>0.\left|\int fd\mu-1\right|>\epsilon>0.

It is a contradiction to (2.4). Therefore, 1Nn=0N1fTn\frac{1}{N}\sum_{n=0}^{N-1}f\circ T^{n} is converging uniformly to 𝟙\mathbb{1} as NN goes to infinity, in particular 𝟙VF\mathbb{1}\in V_{F}. ∎

Proof of Proposition 2.9.

Let ¯={f1¯,,fp¯}\overline{\mathcal{F}}=\{\overline{f_{1}},\cdots,\overline{f_{p}}\} and 𝒢={g1,,gq}\mathcal{G}=\{g_{1},\cdots,g_{q}\} be generating families of U~T\tilde{U}_{T} and U¯T\underline{U}_{T} with p=Mult(U~T)p=\text{\rm Mult}(\tilde{U}_{T}) and q=Mult(U¯T)q=\text{\rm Mult}(\underline{U}_{T}). For 1lp1\leq l\leq p, take flC(X)f_{l}\in C(X) be a function such that fl¯=flmodBT(X)¯\overline{f_{l}}=f_{l}\mod\overline{B_{T}(X)}, then let ={f1,,fl}\mathcal{F}=\{f_{1},\cdots,f_{l}\}. One easily checks that 𝒢\mathcal{F}\cup\mathcal{G} is a generating family of UTU_{T}. By Lemma 2.6 we may write e(X,T)={ν1,,νp}\mathcal{M}_{e}(X,T)=\{\nu_{1},\cdots,\nu_{p}\}. As in the proof of Lemma 2.10 we may assume without loss of generality fi𝑑νj=δi,j\int f_{i}\,d\nu_{j}=\delta_{i,j} for any 1i,jp1\leq i,j\leq p. Let g1=g1+𝟙g^{\prime}_{1}=g_{1}+\mathbb{1}, hence

(g1𝑑νi)1ip=(1,,1).\left(\int g^{\prime}_{1}\,d\nu_{i}\right)_{1\leq i\leq p}=(1,\cdots,1).

By Remark 2.7 the family {g1¯,fj¯: 1<jp}\{\overline{g^{\prime}_{1}},\overline{f_{j}}:\ 1<j\leq p\} is generating for U~T\tilde{U}_{T}. By Lemma 2.10 the constant functions, therefore also g1g_{1}, belongs to V{g1,fj: 1<jp}V_{\{g^{\prime}_{1},f_{j}:\ 1<j\leq p\}}. Then V{g1,fj,gi: 1<jp, 1<iq}=V{g1,fj,gi: 1<jp, 1iq}V_{\{g^{\prime}_{1},f_{j},g_{i}\ :\ 1<j\leq p,\,1<i\leq q\}}=V_{\{g^{\prime}_{1},f_{j},g_{i}\ :\ 1<j\leq p,\,1\leq i\leq q\}} and f1V{g1,fj,gi: 1<jp, 1iq}f_{1}\in V_{\{g^{\prime}_{1},f_{j},g_{i}\ :\ 1<j\leq p,\,1\leq i\leq q\}}. Consequently we get

V{g1,fj,gi: 1<jp, 1<iq}\displaystyle V_{\{g^{\prime}_{1},f_{j},g_{i}\ :\ 1<j\leq p,\,1<i\leq q\}} V𝒢=C(X).\displaystyle\supset V_{\mathcal{F}\cup\mathcal{G}}=C(X).

We conclude that Mult(T)p+q1=Mult(UT~)+Mult(U¯T)1.\text{\rm Mult}(T)\leq p+q-1=\text{\rm Mult}(\tilde{U_{T}})+\text{\rm Mult}(\underline{U}_{T})-1.

3. Affine embedding of ((X),T)(\mathcal{M}(X),T_{*}) in cubical shifts

For a topological system (X,T)(X,T) we denote by TT_{*} the action induced by TT on the compact set (X)\mathcal{M}(X), i.e. Tμ()=μ(T1)T_{*}\mu(\cdot)=\mu(T^{-1}\cdot) for all μ(X)\mu\in\mathcal{M}(X). Then ((X),T)(\mathcal{M}(X),T_{*}) is also a topological system, which is called the induced system of (X,T)(X,T).

For dd\in\mathbb{N} we let σd\sigma_{d} be the shift on the simplex ([0,1]d)([0,1]^{d})^{\mathbb{Z}}. An embedding of (X,T)(X,T) in ([0,1]d)([0,1]^{d})^{\mathbb{Z}} is a continuous injective map ϕ:X([0,1]d)\phi:X\rightarrow([0,1]^{d})^{\mathbb{Z}} satisfying ϕT=σdϕ\phi\circ T=\sigma_{d}\circ\phi. Existence of such embedding is related to the mean dimension theory (we refer to [Coo15] for an introduction). Such an embedding implies that the mean dimension of (X,T)(X,T) is less than or equal to dd. Moreover the topological dimension (i.e. Lebesgue covering dimension) dnTd_{n}^{T} of the set of nn-periodic points then also satisfy dnTnd\frac{d_{n}^{T}}{n}\leq d. Conversely it has been shown that minimal systems with mean dimension less than d/2d/2 can be embedded in the cubical shift σd\sigma_{d} [Lin99, GT20].

In this section we consider affine embedding of the induced system ((X),T)(\mathcal{M}(X),T_{*}) in cubical shift σd\sigma_{d}, i.e. the embedding ϕ:(X)([0,1]d)\phi:\mathcal{M}(X)\rightarrow([0,1]^{d})^{\mathbb{Z}} is affine. In particular we will relate the embedding dimension dd with the multiplicity of (X,T)(X,T).

3.1. Case of finite sets

We first deal with the case of a finite set XX. Then TT is just a permutation of XX and (X)\mathcal{M}(X) is a finite dimensional simplex. We classify the possible affine embedding of ((X),T)(\mathcal{M}(X),T_{*}) in the following proposition.

Proposition 3.1.

Suppose XX is a finite set and TT is a a permutation of XX. Let τ1τk\tau_{1}\cdots\tau_{k} be the decomposition of TT into disjoint cycles τi\tau_{i} of length rir_{i} for 1ik1\leq i\leq k.

  • (1)

    If there is a nontrivial common factor of rir_{i} for 1ik1\leq i\leq k, then there is an affine embedding of ((X),T)(\mathcal{M}(X),T_{*}) in (([0,1]k),σk)\left(([0,1]^{k})^{\mathbb{Z}},\sigma_{k}\right). Such kk is sharp.

  • (2)

    If there is no nontrivial common factor of rir_{i} for 1ik1\leq i\leq k, then there is an affine embedding of ((X),T)(\mathcal{M}(X),T_{*}) in (([0,1]k1),σk1)\left(([0,1]^{k-1})^{\mathbb{Z}},\sigma_{k-1}\right). Such k1k-1 is sharp.

Proof.

For each 1ik1\leq i\leq k we fix a point eiXe_{i}\in X in each cycle τi\tau_{i}, i.e. {Tjei:0jri}=X\{T^{j}e_{i}:0\leq j\leq r_{i}\}=X. Notice that there are continuous maps ae:(X)[0,1]a_{e}:\mathcal{M}(X)\rightarrow[0,1] , eXe\in X, with eXae=𝟙\sum_{e\in X}a_{e}=\mathbb{1} satisfying μ=eXae(μ)δe\mu=\sum_{e\in X}a_{e}(\mu)\delta_{e} for all μ(X)\mu\in\mathcal{M}(X).
(1) Assume there is a nontrivial common factor pp of rir_{i} for 1ik1\leq i\leq k. Then

dim(Fix(Tp))=kp1,\mathrm{dim}(\mathrm{Fix}(T_{*}^{p}))=kp-1,

where Fix(Tp)={μ(X):Tpμ=μ}\mathrm{Fix}(T_{*}^{p})=\{\mu\in\mathcal{M}(X):\ T_{*}^{p}\mu=\mu\}. Since p>1p>1 and dim(Fix(σk1p))=kpp\mathrm{dim}(\mathrm{Fix}(\sigma_{k-1}^{p}))=kp-p, the dynamical system ((X),T)(\mathcal{M}(X),T_{*}) can not embed in (([0,1]k1),σk1)\left(([0,1]^{k-1})^{\mathbb{Z}},\sigma_{k-1}\right).

Now we construct the embedding of ((X),T)(\mathcal{M}(X),T_{*}) in (([0,1]k),σk)\left(([0,1]^{k})^{\mathbb{Z}},\sigma_{k}\right). We define firstly a dynamical embedding Ψ\Psi of the set of extreme points in (X)\mathcal{M}(X), which is identified with XX through the map xδxx\mapsto\delta_{x}, into ([0,1]k)([0,1]^{k})^{\mathbb{Z}} by letting

i=1,,k,l,(Ψ(Tlei))i=σl((10ri1));\forall i=1,\cdots,k,\ \forall l\in\mathbb{Z},\ \ (\Psi(T^{l}e_{i}))_{i}=\sigma^{l}\left((10^{r_{i}-1})^{\infty}\right);

the other components (Ψ(Tlei))j(\Psi(T^{l}e_{i}))_{j}, jij\neq i, being chosen to be equal to the 00^{\infty} sequence. Then we may extend Ψ\Psi affinely from the set of extreme points on (X)\mathcal{M}(X) by letting

Ψ(μ)=eXae(μ)Ψ(δe).\Psi(\mu)=\sum_{e\in X}a_{e}(\mu)\Psi(\delta_{e}).

It is easy to check that Ψ\Psi is injective which deduces a dynamical embedding ((X),T)(\mathcal{M}(X),T_{*}) in (([0,1]k),σk)\left(([0,1]^{k})^{\mathbb{Z}},\sigma_{k}\right).

(2) Assume there is no nontrivial common factor of rir_{i} for 1ik1\leq i\leq k. We have that k1k-1 numbers qi:=(rk,ri),q_{i}:=(r_{k},r_{i}), 1ik11\leq i\leq k-1 are co-prime where (a,b)(a,b) are the highest common factor of aa and bb. We define firstly a continuous map Ψ\Psi of the set of extreme points in (X)\mathcal{M}(X) into ([0,1]k1)([0,1]^{k-1})^{\mathbb{Z}} by letting

l,(Ψ(Tlek))j=σl((10ri1)),1jk1,\ \forall l\in\mathbb{Z},\ \ (\Psi(T^{l}e_{k}))_{j}=\sigma^{l}\left((10^{r_{i}-1})^{\infty}\right),\forall 1\leq j\leq k-1,

and

1ik1,l,(Ψ(Tlei))i=σl((10ri1));ji,l,(Ψ(Tlei))j=0.\forall 1\leq i\leq k-1,\ \forall l\in\mathbb{Z},\ \ (\Psi(T^{l}e_{i}))_{i}=\sigma^{l}\left((10^{r_{i}-1})^{\infty}\right);\ \forall j\not=i,\forall l\in\mathbb{Z},(\Psi(T^{l}e_{i}))_{j}=0^{\infty}.

Then we may extend Ψ\Psi affinely from the set of extreme points on (X)\mathcal{M}(X) by letting

Ψ(μ)=eXae(μ)Ψ(δe).\Psi(\mu)=\sum_{e\in X}a_{e}(\mu)\Psi(\delta_{e}).

It remains to show that Ψ\Psi is injective. Let μ=eXbeδe\mu=\sum_{e\in X}b_{e}\delta_{e} and μ=eXbeδe\mu^{\prime}=\sum_{e\in X}b_{e}^{\prime}\delta_{e}. Suppose

Ψ(μ)=Ψ(μ)=eXaeΨ(δe)=(ci,j)1ik1,j.\Psi(\mu)=\Psi(\mu^{\prime})=\sum_{e\in X}a_{e}\Psi(\delta_{e})=(c_{i,j})_{1\leq i\leq k-1,j\in\mathbb{Z}}.

Since qi:=(rk,ri)q_{i}:=(r_{k},r_{i}) then there are integers sis_{i} and tit_{i} such that siritirk=qi.s_{i}r_{i}-t_{i}r_{k}=q_{i}. Let

ui,l=siri+l=tirk+qi+l.u_{i,l}=s_{i}r_{i}+l=t_{i}r_{k}+q_{i}+l.

It implies that

bTlekbTl+qiek=bTlekbTl+qiek=ci,lci,ui,l.b_{T^{l}e_{k}}^{\prime}-b_{T^{l+q_{i}}e_{k}}^{\prime}=b_{T^{l}e_{k}}-b_{T^{l+q_{i}}e_{k}}=c_{i,l}-c_{i,u_{i,l}}.

Since qi,1ik1q_{i},1\leq i\leq k-1 are co-prime, there are integers wi,1ik1w_{i},1\leq i\leq k-1 such that i=1k1wiqi=1\sum_{i=1}^{k-1}w_{i}q_{i}=1. Since

(bekbTw1q1ek)+\displaystyle\left(b_{e_{k}}-b_{T^{w_{1}q_{1}}e_{k}}\right)+ (bTw1q1ekbTw1q1+w2q2ek)\displaystyle\left(b_{T^{w_{1}q_{1}}e_{k}}-b_{T^{w_{1}q_{1}+w_{2}q_{2}}e_{k}}\right)
++(bTw1q1+w2q2++wk2qk2ekbTw1q1+w2q2++wk1qk1ek)=bekbTek,\displaystyle+\dots+\left(b_{T^{w_{1}q_{1}+w_{2}q_{2}+\dots+w_{k-2}q_{k-2}}e_{k}}-b_{T^{w_{1}q_{1}+w_{2}q_{2}+\dots+w_{k-1}q_{k-1}}e_{k}}\right)=b_{e_{k}}-b_{Te_{k}},

we have

bekbTlek=bekbTlek,l.b_{e_{k}}-b_{T^{l}e_{k}}=b_{e_{k}}^{\prime}-b_{T^{l}e_{k}}^{\prime},\ \forall l\in\mathbb{Z}.

Since eXbe=eXbe=1\sum_{e\in X}b_{e}=\sum_{e\in X}b_{e}^{\prime}=1, we conclude that bek=bekb_{e_{k}}=b_{e_{k}^{\prime}} and consequently bei=beib_{e_{i}}=b_{e_{i}^{\prime}} by bek+bei=bek+bei=ci,0b_{e_{k}}+b_{e_{i}}=b_{e_{k}}^{\prime}+b_{e_{i}}^{\prime}=c_{i,0} for 1ik11\leq i\leq k-1. It means that μ=μ\mu=\mu^{\prime} and Ψ\Psi is injective.

Remark 3.2.

For such a permutation TT, we have Mult(T)=e(X,T)=k\text{\rm Mult}(T)=\sharp\mathcal{M}_{e}(X,T)=k, with kk being the number of cycles in the decomposition of TT.

3.1.1. General case

We consider now a general topological system and relates the dimension of the cubical shift in an affine embedding with the multiplicity of (X,T)(X,T).

Theorem 3.3.

Let (X,T)(X,T) be a topological system. If Mult(T)\text{\rm Mult}(T) is equal to dd, then there is an affine embedding of ((X),T)(\mathcal{M}(X),T_{*}) in (([0,1]d),σd)(([0,1]^{d})^{\mathbb{Z}},\sigma_{d}). Conversely if ((X),T)(\mathcal{M}(X),T_{*}) embeds into (([0,1]d),σd)(([0,1]^{d})^{\mathbb{Z}},\sigma_{d}) then

  • either e(X,T)d\sharp\mathcal{M}_{e}(X,T)\leq d and Mult(T)d\text{\rm Mult}(T)\leq d,

  • or e(X,T)=d+1\sharp\mathcal{M}_{e}(X,T)=d+1 and Mult(T)=d+1\text{\rm Mult}(T)=d+1.

Proof.

Firstly, notice that any affine equivariant map Ψ:((X),T)(([0,1]d),σd)\Psi:(\mathcal{M}(X),T_{*})\rightarrow(([0,1]^{d})^{\mathbb{Z}},\sigma_{d}) is of the form

Ψf:μ(fTk𝑑μ)k,\Psi_{f}:\mu\mapsto\left(\int f\circ T^{k}\,d\mu\right)_{k\in\mathbb{Z}},

for some continuous function f=(f1,,fd):X[0,1]df=(f_{1},\cdots,f_{d}):X\rightarrow[0,1]^{d}.

Assume the topological multiplicity Mult(X,T)\text{\rm Mult}(X,T) is equal to dd, i.e. there is a family F={f1,,fd}F=\{f_{1},\cdots,f_{d}\} of continuous functions such that VF=C(X)V_{F}=C(X). Let us show the associated map Ψf\Psi_{f} is injective. Let μ1,μ2(X)\mu_{1},\mu_{2}\in\mathcal{M}(X) with Ψf(μ1)=Ψf(μ1)\Psi_{f}(\mu_{1})=\Psi_{f}(\mu_{1}) i.e. fiTk𝑑μ1=fiTk𝑑μ1\int f_{i}\circ T^{k}\,d\mu_{1}=\int f_{i}\circ T^{k}\,d\mu_{1} for any i=1,,di=1,\cdots,d and any kk\in\mathbb{Z}. Then by density of span(fiTk,i,k)\mathrm{span}(f_{i}\circ T^{k},\ i,\ k) in C(X)C(X) we have

g𝑑μ1=g𝑑μ1 for all gC(X),\int g\,d\mu_{1}=\int g\,d\mu_{1}\text{ for all }g\in C(X),

which implies μ1=μ2\mu_{1}=\mu_{2}. Therefore we get the injectivity of Ψf\Psi_{f}.

Conversely, assume Ψf\Psi_{f} is injective for f=(f1,,fd):X[0,1]df=(f_{1},\cdots,f_{d}):X\rightarrow[0,1]^{d}. Let F={𝟙,f1,,fd}F=\{\mathbb{1},f_{1},\cdots,f_{d}\}. We claim that VF=C(X)V_{F}=C(X). Then if e(X,T)d\sharp\mathcal{M}_{e}(X,T)\leq d, we get by injectivity of Ψf\Psi_{f} that there exists A{1,2,,d}A\subset\{1,2,\dots,d\} with A=e(X,T)\sharp A=\sharp\mathcal{M}_{e}(X,T) such that the matrix (fi𝑑ν)iA,νe(X,T)\left(\int f_{i}d\nu\right)_{i\in A,\nu\in\mathcal{M}_{e}(X,T)} is invertible. Then by Remark 2.7 the family F{𝟙}¯\overline{F\setminus\{\mathbb{1}\}} is generating for U~T\tilde{U}_{T} and consequently VF=VF{𝟙}V_{F}=V_{F\setminus\{\mathbb{1}\}} by Lemma 2.10. If e(X,T)=d+1\sharp\mathcal{M}_{e}(X,T)=d+1 then we only get Mult(T)=d+1\text{\rm Mult}(T)=d+1.

It remains to show our claim. Assume to the contrary that VFC(X)V_{F}\not=C(X). Then by Riesz Theorem there is a signed finite measure μ\mu vanishing on each function in FF. Let μ=μ+μ\mu=\mu^{+}-\mu^{-} be the Jordan decomposition of μ\mu (i.e. the measures μ+\mu^{+} and μ\mu^{-} are two finite positive measures which are mutually singular). Evaluating on the constant function 𝟙\mathbb{1}, we get μ+(X)=μ(X)\mu^{+}(X)=\mu^{-}(X). Then by rescaling, we may assume both μ\mu^{-} and μ+\mu^{+} belong to (X)\mathcal{M}(X). Finally we get Ψf(μ+)=Ψf(μ)\Psi_{f}(\mu^{+})=\Psi_{f}(\mu^{-}), and therefore μ+=μ\mu^{+}=\mu^{-} by injectivity of Ψf\Psi_{f} contradicting therefore the mutual singularity of μ+\mu^{+} and μ\mu^{-}.

3.2. Affine embeddings and Lindenstrauss-Tsukamoto conjecture

Lindenstrauss and Tsukamoto [LT14] have conjectured that any topological system with mean dimension mdim(X,T)\text{\rm mdim}(X,T) less than d/2d/2 and such that the dimension dnTd_{n}^{T} of the set of nn-periodic points satisfies dnTn<d/2\frac{d_{n}^{T}}{n}<d/2 for any nn\in\mathbb{N} may be embedded in the shift over ([0,1]d)([0,1]^{d})^{\mathbb{Z}}. As mentioned above it is known for minimal systems. We consider here affine systems, i.e. affine maps of a simplex. Such maps are never minimal, as they always admits at least one fixed point.

The example below shows that Lindenstrauss-Tsukamoto conjecture does not hold true in the affine category. Recall that an ergodic system (X,f,,μ)(X,f,\mathcal{B},\mu) has a countable Lebesgue spectrum, when there is a countable family (ψn)n(\psi_{n})_{n\in\mathbb{N}} in L02(μ)L^{2}_{0}(\mu) such that ψnfk\psi_{n}\circ f^{k}, kk\in\mathbb{Z}, nn\in\mathbb{N} form a Hilbert basis of L02(μ)L^{2}_{0}(\mu).

Corollary 3.4.

There is an affine system with a unique periodic (fixed) point and zero-topological entropy (in particular mdim(T)=0mdim(T)=0 and dnT=0d_{n}^{T}=0 for all n1n\geq 1) which does not embed affinely in (([0,1]k),σ)(([0,1]^{k})^{\mathbb{Z}},\sigma) for any k1k\geq 1.

Proof.

There exists an ergodic measure preserving system (Y,𝒜,f,μ)(Y,\mathcal{A},f,\mu) with zero entropy and countable Lebesgue spectrum [NP66, Par53] (in particular totally ergodic, i.e. fnf^{n} is ergodic for any nn\in\mathbb{Z}). Then by Jewett-Krieger theorem there is a uniquely ergodic topological system (X,T)(X,T) with measure ν\nu realizing such a measure preserving system. All powers of TT are uniquely ergodic as μ\mu was chosen totally ergodic. Moreover the topological entropy of TT, thus that of TT_{*} is zero by Glasner-Weiss [GW95]. As the unique invariant measure ν\nu has countable Lebesgue spectrum, the topological multiplicity of (X,T)(X,T) is infinite by Lemma 2.4. We conclude with Theorem 3.3. ∎

3.3. Application: zero topological entropy

A classical result in ergodic theory states that any ergodic system (X,f,,μ)(X,f,\mathcal{B},\mu) with positive entropy has a countable Lebesgue spectrum. In particular [h(μ)>0][Mult(μ)=][h(\mu)>0]\Rightarrow[\text{\rm Mult}(\mu)=\infty]. Then it follows from the variational principle for the topological entropy :

Proposition 3.5.

Any topological system (X,T)(X,T) with Mult(T)<\text{\rm Mult}(T)<\infty has zero topological entropy.

We may also give a purely topological proof of Proposition 3.5 based on mean dimension theory. More precisely we use the main result of [BS22], which states as follows:

Theorem 3.6.

[BS22] For any topological system (X,T)(X,T) with positive topological entropy, the induced system ((X),T)(\mathcal{M}(X),T_{*}) has infinite topological mean dimension. Therefore,

htop(T)>0mdim(T)>0mdim(T)=.h_{top}(T)>0\Leftrightarrow\text{\rm mdim}(T_{*})>0\Leftrightarrow\text{\rm mdim}(T_{*})=\infty.
Topological proof of Proposition 3.5.

Assume Mult(T)=d\text{\rm Mult}(T)=d is finite. Then by Theorem 3.3 the induced system ((X),T)(\mathcal{M}(X),T_{*}) embeds in the cubical shift (([0,1]d),σ)\left(([0,1]^{d})^{\mathbb{Z}},\sigma\right). In particular the mean dimension of TT_{*} is less than or equal to the mean dimension of the shift (([0,1]d),σ)\left(([0,1]^{d})^{\mathbb{Z}},\sigma\right), which is equal to dd. By Theorem 3.6, it implies that TT has zero topological entropy. ∎

4. Baxter’s Lemma in Banach spaces

In [Bax71], Baxter gave a useful criterion to show simple spectrum of ergodic transformations. It may be extended more generally to bound the multiplicity of the spectrum (e.g. see Proposition 2.12 in [Que10]). We generalize this criterion for operators defined on a Banach space. It will be used in the next section to estimate the topological multiplicity in some examples.

Lemma 4.1.

Let BB be a separable Banach space and (B)\mathcal{L}(B) be the set of bounded linear operator on BB. We consider an invertible isometry U(B)U\in\mathcal{L}(B). If (n)n(\mathcal{F}_{n})_{n} is a sequence of finite subsets in HH satisfying for all fBf\in B

(4\cdot1) infFnVnFnfn0\displaystyle\inf_{F_{n}\in V_{\mathcal{F}_{n}}}\|F_{n}-f\|\xrightarrow{n\rightarrow\infty}0

then there exists a family B\mathcal{F}\subset B with supnn\sharp\mathcal{F}\leq\sup_{n}\sharp\mathcal{F}_{n} and B=V.B=V_{\mathcal{F}}.

Classical proofs of Baxter’s lemma strongly used the Hilbert structure. Here we use a Baire argument as in Lemma 5.2.10 [Fog02]. Note also that we do not require the sequence of vectors spaces (Vn)n(V_{\mathcal{F}_{n}})_{n} to be nondecreasing. Observe finally that it is enough to assume (4\cdot1) for ff in SS, where SS spans a dense subset of BB.

Proof.

Let m=supnnm=\sup_{n}\sharp\mathcal{F}_{n}. If m=m=\infty, it is trivial. Assume m<m<\infty. By passing to a subsequence, we assume n=m\sharp\mathcal{F}_{n}=m for all nn. Let B(m)B^{(m)} be the space of finite subsets of HH whose cardinality is smaller than or equal to mm. When endowed with the Hausdorff distance dHaud_{Hau}, the space B(m)B^{(m)} is a metric space, which is complete and separable. We assume the following claim, which we prove later on.

Claim 4.2.

For any ϵ>0\epsilon>0 and B(m)\mathcal{F}\in B^{(m)}, the set

O(,ϵ)={𝒢B(m),finfGV𝒢Gf<ϵ}O(\mathcal{F},\epsilon)=\left\{\mathcal{G}\in B^{(m)},\ \forall f\in\mathcal{F}\ \inf_{G\in V_{\mathcal{G}}}\|G-f\|<\epsilon\right\}

is open and dense.

Let (gq)q\left(g_{q}\right)_{q\in\mathbb{N}} be a countable dense family in BB. Let 𝒢q\mathcal{G}_{q} be the finite family {g1,,gq}\{g_{1},\cdots,g_{q}\}. For any pp\in\mathbb{N}^{*} and any qq\in\mathbb{N} we consider the open and dense set

Op,q:=O(𝒢q,1/p)={𝒢B(m),g𝒢qinfGV𝒢Gg<1/p}.O_{p,q}:=O(\mathcal{G}_{q},1/p)=\left\{\mathcal{G}\in B^{(m)},\ \forall g\in\mathcal{G}_{q}\ \inf_{G\in V_{\mathcal{G}}}\|G-g\|<1/p\right\}.

According to Baire’s theorem, the intersection p,qOp,q\bigcap_{p,q}O_{p,q} is not empty. Clearly any family \mathcal{F} in the intersection satisfies V=BV_{\mathcal{F}}=B. It remains to prove Claim 4.2.

Proof of Claim 4.2.

The set O(,ϵ)O(\mathcal{F},\epsilon) is open. We focus on the denseness property. Pick arbitrary δ>0\delta>0 and B(m)\mathcal{H}\in B^{(m)}. We will show that there is O(,ϵ)\mathcal{H}^{\prime}\in O(\mathcal{F},\epsilon) with dHau(,)<δd_{Hau}(\mathcal{H},\mathcal{H}^{\prime})<\delta. As the elements of B(m)B^{(m)} with cardinality mm are dense in B(m)B^{(m)} we can assume without loss of generality that =m\sharp\mathcal{H}=m. By assumptions on the sequence n\mathcal{F}_{n}, there exists nn such that

(4\cdot2) finfFnVnFnf<ϵ,\displaystyle\forall f\in\mathcal{F}\ \inf_{F_{n}\in V_{\mathcal{F}_{n}}}\|F_{n}-f\|<\epsilon,
(4\cdot3) hinfFnVnFnh<δ.\displaystyle\forall h\in\mathcal{H}\ \inf_{F_{n}\in V_{\mathcal{F}_{n}}}\|F_{n}-h\|<\delta.

We write ={h1,h2,,hm}\mathcal{H}=\{h_{1},h_{2},\dots,h_{m}\} and n={f1,f2,,fm}\mathcal{F}_{n}=\{f_{1},f_{2},\dots,f_{m}\}. By (4\cdot3) there are polynomials (Pi,j)1i,jm(P_{i,j})_{1\leq i,j\leq m} in [X]\mathbb{R}[X] and a nonnegative integer pp such that

(4\cdot4) i=1,,m,hij=1mUpPi,j(U)fj<δ\displaystyle\forall i=1,\cdots,m,\ \left\|h_{i}-\sum_{j=1}^{m}U^{-p}P_{i,j}(U)f_{j}\right\|<\delta

Let Q[X]Q\in\mathbb{R}[X] be the polynomial given by the determinant of the matrix M=(Pi,j)1i,jmMm([X])M=(P_{i,j})_{1\leq i,j\leq m}\in M_{m}(\mathbb{R}[X]). The spectrum Sp(U)\mathrm{Sp}(U) of UU is contained in the unit circle. In particular for arbitrarily small λ\lambda\in\mathbb{R}, the polynomial Q(+λ)Q(\cdot+\lambda) does not vanish on Sp(U)\mathrm{Sp}(U). Hence by replacing Pi,jP_{i,j} by Pi,j(+λ)P_{i,j}(\cdot+\lambda) we may assume that QQ does not vanish on the spectrum of UU. Then Q(U)=λ,Q(λ)=0(UλId)Q(U)=\prod_{\lambda,\,Q(\lambda)=0}(U-\lambda\mathrm{Id}) is invertible and its inverse may be approximated by polynomials in UU and U1U^{-1}, because for λ\lambda with Q(λ)=0Q(\lambda)=0 we have (UλId)1=kUkλk+1(U-\lambda\mathrm{Id})^{-1}=-\sum_{k\in\mathbb{N}}\frac{U^{k}}{\lambda^{k+1}} for |λ|>1|\lambda|>1 and (UλId)1=kU(k+1)λk(U-\lambda\mathrm{Id})^{-1}=-\sum_{k\in\mathbb{N}}\frac{U^{-(k+1)}}{\lambda^{k}} for |λ|<1|\lambda|<1 (these sequences are normally convergent in (B)\mathcal{L}(B) as we assume U=U1=1\|U\|=\|U^{-1}\|=1). Let ={h1,h2,,hm}\mathcal{H}^{\prime}=\{h^{\prime}_{1},h^{\prime}_{2},\dots,h^{\prime}_{m}\} with hi=j=1mUpPi,j(U)fjh^{\prime}_{i}=\sum_{j=1}^{m}U^{-p}P_{i,j}(U)f_{j}. We have

n\displaystyle\mathcal{F}_{n} =M(U)1Up,\displaystyle=M(U)^{-1}U^{p}\mathcal{H}^{\prime},
=comtM(U)UpQ(U)1.\displaystyle=\mathrm{{}^{t}com}\,M(U)U^{p}Q(U)^{-1}\mathcal{H}^{\prime}.

Then from the above observations we get

nV.\mathcal{F}_{n}\subset V_{\mathcal{H}^{\prime}}.

In particular VnVV_{\mathcal{F}_{n}}\subset V_{\mathcal{H}^{\prime}} and it follows finally from (4\cdot2) that O(,ϵ)\mathcal{H}^{\prime}\in O(\mathcal{F},\epsilon), i.e.

finfHVHf<ϵ.\displaystyle\forall f\in\mathcal{F}\ \inf_{H^{\prime}\in V_{\mathcal{H}^{\prime}}}\|H^{\prime}-f\|<\epsilon.

This completes the proof as we have dHau(,)<δd_{Hau}(\mathcal{H},\mathcal{H}^{\prime})<\delta by (4\cdot4), where \mathcal{H} and δ\delta have been chosen arbitrarily. ∎

5. Cantor systems with finite topological rank

Roughly speaking an ergodic measure preserving system is of finite rank rr, when it may be obtained by cutting and staking with rr Kakutani-Rohlin towers. For ergodic systems, Baxter’s lemma implies that the ergodic multiplicity is less than or equal to the rank. Topological rank as been defined and studied for minimal Cantor systems (see e.g. [DDMP21] and the references therein). For such systems we show now with Lemma 4.1 that the same inequality holds for the topological quantities : the topological multiplicity is less than or equal to the topological rank.

Firstly we recall the definition of topological rank. Let (X,T)(X,T) be a minimal Cantor system. A Kakutani-Rohlin partition of XX is given by

𝒯={TjB(k):1kd,0j<h(k)},\mathcal{T}=\{T^{-j}B(k):1\leq k\leq d,0\leq j<h(k)\},

where d,h(k),1kdd,h(k),1\leq k\leq d are positive integers and B(k),1kdB(k),1\leq k\leq d are clopen subsets of XX such that

k=1dTh(k)B(k)=k=1dB(k).\cup_{k=1}^{d}T^{-h(k)}B(k)=\cup_{k=1}^{d}B(k).

The base of 𝒯\mathcal{T} is the set B(𝒯)=k=1dB(k)B(\mathcal{T})=\cup_{k=1}^{d}B(k). A sequence of Kakutani-Rohlin partitions

𝒯n={TjBn(k):1kdn,0j<hn(k)},n1,\mathcal{T}_{n}=\{T^{-j}B_{n}(k):1\leq k\leq d_{n},0\leq j<h_{n}(k)\},n\geq 1,

is nested if

  1. (1)

    𝒯0\mathcal{T}_{0} is the trivial partition, i.e. d0=1,h0=1d_{0}=1,h_{0}=1 and B0(1)=XB_{0}(1)=X.

  2. (2)

    B(𝒯n+1)B(𝒯n)B(\mathcal{T}_{n+1})\subset B(\mathcal{T}_{n}).

  3. (3)

    𝒯n+1𝒯n\mathcal{T}_{n+1}\succ\mathcal{T}_{n}.

  4. (4)

    (n0B(𝒯n))=1\sharp(\cap_{n\geq 0}B(\mathcal{T}_{n}))=1.

  5. (5)

    n1𝒯n\cup_{n\geq 1}\mathcal{T}_{n} spans the topology of XX.

Moreover, it is primitive if for all n1n\geq 1 there exists N>nN>n such that for all 1kdN1\leq k\leq d_{N} and for each xT(hN(k)1)BN(k)x\in T^{-(hN(k)-1)}B_{N}(k),

{Ti(x):0ihN(k)1}Bn(j),1jdn.\{T^{i}(x):0\leq i\leq h_{N}(k)-1\}\cap B_{n}(j)\not=\emptyset,\forall 1\leq j\leq d_{n}.

Following [DDMP21], a minimal Cantor system is of topological rank dd if it admits a primitive sequence of nested Kakutani-Rohlin partitions with dndd_{n}\leq d for all nn\in\mathbb{N}.

Theorem 5.1.

Let (X,T)(X,T) be a minimal Cantor system with topological rank dd. Then Mult(X,T)d\text{\rm Mult}(X,T)\leq d.

Proof.

Let (𝒯n)n(\mathcal{T}_{n})_{n\in\mathbb{N}} be the primitive sequence of nested Kakutani-Rohlin partitions with dndd_{n}\leq d for all nn\in\mathbb{N}. Let

n={χBn(k):1kdn}.\mathcal{F}_{n}=\{\chi_{B_{n}(k)}:1\leq k\leq d_{n}\}.

Since n1𝒯n\cup_{n\geq 1}\mathcal{T}_{n} spans the topology of XX, we have

fC(X),infFnVnFnfn0.\forall f\in C(X),\ \inf_{F_{n}\in V_{\mathcal{F}_{n}}}\|F_{n}-f\|\xrightarrow{n\rightarrow\infty}0.

It follows from Lemma 4.1 and dndd_{n}\leq d for all nn\in\mathbb{N} that Mult(X,T)d\text{\rm Mult}(X,T)\leq d. ∎

Remark 5.2.

It was shown in [DDMP21] that the Thue-Morse subshift has topological rank 33. By Theorem 5.1, Thue-Morse subshift has therefore topological multiplicity at most 33. We give another direct proof of this fact in Subsection 6.4.

Examples of ergodic systems with rank rr and multiplicity mm have been built for any 1mr1\leq m\leq r in [KL97]. We then propose the following question.

Question 5.3.

Can one build for any 1mr1\leq m\leq r a minimal Cantor system with topological multiplicity mm and topological rank rr?

6. Examples of finite topological multiplicity

An invertible dynamical system is called topological simple or have simple topological spectrum if Mult(T)=1\text{\rm Mult}(T)=1.

6.1. Minimal rotation on compact groups

Let GG be a compact abelian group. Denote by G^\hat{G} the dual group of GG and by λ\lambda the Haar measure on GG. For fC(G)f\in C(G), we write f^\hat{f} the Fourier transformation of ff.

Proposition 6.1.

Any minimal translation τ\tau on a compact abelian group GG is topologically simple.

Proof.

We claim that any fC(G)f\in C(G) with f^(χ)0\hat{f}(\chi)\neq 0 for all χG^\chi\in\hat{G} is cyclic, i.e. the vector space spanned by fτkf\circ\tau^{k}, kk\in\mathbb{N} is dense in C(X)C(X). As characters of a compact abelian group separates points it is enough to show by Stone-Weierstrass theorem that any character belongs to the complex vector space spanned by fτkf\circ\tau^{k}, kk\in\mathbb{N}. But for all χG^\chi\in\hat{G} we have

f^(χ)χ=χf=f(y)χ(y)dλ(y).\hat{f}(\chi)\chi=\chi*f=\int f(\cdot-y)\chi(y)\,d\lambda(y).

Then the function ff being uniformly continuous, there are functions of the form kf(yk)χ(yk)\sum_{k}f(\cdot-y_{k})\chi(y_{k}) arbitrarily close to f^(χ)χ\hat{f}(\chi)\chi for the supremum norm. By minimality of τ\tau, there are integers lkl_{k}\in\mathbb{N} such that fτlkf\circ\tau^{l_{k}} and f(yk)f(\cdot-y_{k}) are arbitrarily closed. It concludes the proof. ∎

6.2. Sturmian subshift

A word u{0,1}u\in\{0,1\}^{\mathbb{Z}} is called Sturmian if it is recurrent under the shift σ\sigma, and the number of nn-words in uu equals n+1n+1 for each n1n\geq 1. Take the shift-orbit closure Xu=Oσ(u)¯X_{u}=\overline{O_{\sigma}(u)}. The corresponding subshift (Xu,σ)(X_{u},\sigma) is called a Sturmian subshift. Sturmian sequences are symbolic representation of circle irrational rotations.

We first recall some standard notations in symbolic dynamics. For a subset YY of 𝒜\mathcal{A}^{\mathbb{Z}} with 𝒜\mathcal{A} being a finite alphabet we let n(Y)\mathcal{L}_{n}(Y) be the number of nn-words appearing in the sequences of YY. Then for wn(Y)w\in\mathcal{L}_{n}(Y) we let [w][w] be the associated cylinder defined as [w]:={(xn)nY:x0xn1=w}.[w]:=\{(x_{n})_{n\in\mathbb{Z}}\in Y:\ x_{0}\cdots x_{n-1}=w\}. The indicator function of a subset EE of XX will be denoted by χE\chi_{E}.

Proposition 6.2.

Any Sturmian subshift has simple topological spectrum.

Proof.

Let uu be a Sturmian sequence. Let

Fn=span{χ[w]:wn(u)}.F_{n}=\text{span}\{\chi_{[w]}:w\in\mathcal{L}_{n}(u)\}.

It follows that

C(Xu)=VnFn.C(X_{u})=V_{\cup_{n}F_{n}}.

Notice dim(Fn)=n(u)=n+1\dim(F_{n})=\sharp\mathcal{L}_{n}(u)=n+1. We let f:Xuf:X_{u}\rightarrow\mathbb{R} be the continuous function defined as f:x=(xn)n(1)x0f:x=(x_{n})_{n}\mapsto(-1)^{x_{0}}. Let

Gn=span{𝟙,fσk:0kn1}.G_{n}=\text{span}\{\mathbb{1},f\circ\sigma^{k}:0\leq k\leq n-1\}.

Clearly, GnFnG_{n}\subset F_{n}. To prove that (Xu,σ)(X_{u},\sigma) has simple topological spectrum, it is sufficient to show dim(Gn)=n+1\dim(G_{n})=n+1. Thus it is enough to show the functions {𝟙,fσk:0kn1}\{\mathbb{R}\mathbb{1},f\circ\sigma^{k}:0\leq k\leq n-1\} are linearly independent. If not, for some nn there exists a nonzero vector (a0,a1,,an)(a_{0},a_{1},\dots,a_{n}) such that

a0(1)x0+a1(1)x1++an1(1)xn1+an=0,xXu.a_{0}(-1)^{x_{0}}+a_{1}(-1)^{x_{1}}+\dots+a_{n-1}(-1)^{x_{n-1}}+a_{n}=0,\forall x\in X_{u}.

Since n1(u)>n2(u)\sharp\mathcal{L}_{n-1}(u)>\sharp\mathcal{L}_{n-2}(u), we can find distinct x,xXux,x^{\prime}\in X_{u} such that x|0n2=x|0n2x|_{0}^{n-2}=x^{\prime}|_{0}^{n-2} but xn1xn1x_{n-1}\not=x^{\prime}_{n-1}. It follows that an1=0a_{n-1}=0. Since for each 0kn20\leq k\leq n-2 we can always find y,yy,y^{\prime} such that y|0k1=y|0k1y|_{0}^{k-1}=y^{\prime}|_{0}^{k-1} but ykyky_{k}\not=y^{\prime}_{k}, we obtain that an2=an3==a0=0a_{n-2}=a_{n-3}=\dots=a_{0}=0. Finally, we get an=0a_{n}=0. This is a contradiction. Therefore we conclude that dim(Gn)=n+1\dim(G_{n})=n+1, then Fn=GnF_{n}=G_{n}. Then by Lemma 4.1 and Lemma 2.10, (Xu,σ)(X_{u},\sigma) has simple topological spectrum. ∎

6.3. Homeomorphism of the interval

Estimating the multiplicity of non-zero dimensional systems is difficult in general. Below we focus on homeomorphisms of the interval (see [Jav19] for related results on the circle). We use the following result due to Atzmon and Olevskii [AO96]. We denote by C0()C_{0}(\mathbb{R}) the set of continuous map on \mathbb{R} with zero limits in ±\pm\infty. For fC0()f\in C_{0}(\mathbb{R}) and nn\in\mathbb{Z} we let fn=f(+n)f_{n}=f(\cdot+n) be the translation of ff by nn.

Theorem 6.3.

[AO96] There exists gC0()g\in C_{0}(\mathbb{R}) such that the vector space spanned by gng_{n}, nn\in\mathbb{N} is dense in C0()C_{0}(\mathbb{R}).

In particular the operator V:C0()V:C_{0}(\mathbb{R})\circlearrowleft, ff(+1)f\mapsto f(\cdot+1), is cyclic. A Borel set SS of \mathbb{R} is called a set of uniqueness if the sets Sn:=(S+2πn)[π,π]S_{n}:=(S+2\pi n)\cap[-\pi,\pi], nn\in\mathbb{Z} satisfy the following properties:

  1. (1)

    SnS_{n}, nn\in\mathbb{Z}, are pairwise disjoint,

  2. (2)

    Leb(SnU)>0\mathrm{Leb}(S_{n}\cap U)>0 for any nn\in\mathbb{Z} and any open set UU of [π,π][-\pi,\pi],

  3. (3)

    Leb(S)<\mathrm{Leb}(S)<\infty,

where Leb\mathrm{Leb} denotes the Lebesgue measure on \mathbb{R}.

Atzmon and Olevskii proved for any set of uniqueness SS (such sets exist!) the conclusion of Theorem 6.3 holds true with gg being the the Fourier transform of the indicator function of SS. Let us just remark that if SS is a set of uniqueness then

Sl=n(Snk+l+2πk),1lk,S^{l}=\bigcup_{n\in\mathbb{Z}}(S_{nk+l}+2\pi k),1\leq l\leq k,

are kk disjoints sets of uniqueness. Let C0(;)C_{0}(\mathbb{R};\mathbb{C}) be the set of continuous map on \mathbb{C} with zero limits in infinity.

Lemma 6.4.

The operator V:C0(;)kV:C_{0}(\mathbb{R};\mathbb{C})^{k}\circlearrowleft, (fi)1ik(fi(+1))1ik(f_{i})_{1\leq i\leq k}\mapsto\left(f_{i}(\cdot+1)\right)_{1\leq i\leq k} is cyclic. In particular, the operator U:C0()kU:C_{0}(\mathbb{R})^{k}\circlearrowleft, (fi)1ik(fi(+1))1ik(f_{i})_{1\leq i\leq k}\mapsto\left(f_{i}(\cdot+1)\right)_{1\leq i\leq k} is cyclic.

Proof.

Let SS, SlS^{l}, 1lk1\leq l\leq k, be sets of uniqueness as above. By following [AO96] we show that the vector space generated by the translates of g:=(χSl^)1lkg:=(\widehat{\chi_{S^{l}}})_{1\leq l\leq k} is dense in C0(;)kC_{0}(\mathbb{R};\mathbb{C})^{k} with χSl^\widehat{\chi_{S^{l}}} be the Fourier transform of the indicator function χSl\chi_{S^{l}} of SlS^{l}. It follows that the translates of Re(g)Re(g) is dense in C0()kC_{0}(\mathbb{R})^{k}. Let μ=(μl)1lk\mu=(\mu_{l})_{1\leq l\leq k} be a complex bounded measure with

Vn(g),μ=1lk(χSl^)n𝑑μl=0,\langle V^{n}(g),\mu\rangle=\sum_{1\leq l\leq k}\int(\widehat{\chi_{S^{l}}})_{n}d\mu_{l}=0,

for all nn\in\mathbb{Z}. It is enough to prove μl=0\mu_{l}=0 for all 1lk1\leq l\leq k. By Plancherel-Parseval formula we have

(χSl^)n𝑑μl=(χSl^)n^(t)μl^(t)𝑑t=χSl(t)eintμl^(t)𝑑t.\int(\widehat{\chi_{S^{l}}})_{n}d\mu_{l}=\int\widehat{(\widehat{\chi_{S^{l}}})_{n}}(t)\hat{\mu_{l}}(-t)\,dt=-\int\chi_{S^{l}}(t)e^{-int}\hat{\mu_{l}}(t)\,dt.

Therefore we have

1lkχSl(t)μl^(t)eint𝑑t=0,\sum_{1\leq l\leq k}\int\chi_{S^{l}}(t)\hat{\mu_{l}}(t)e^{-int}\,dt=0,

for all nn. But this term is just the nthn^{th} coefficient of the function of L1([π,π])L^{1}([-\pi,\pi]) given by m(1lkχSlμ^l)(+2πm)\sum_{m\in\mathbb{Z}}(\sum_{1\leq l\leq k}\chi_{S^{l}}\hat{\mu}_{l})(\cdot+2\pi m), which should therefore be 0. As the sets (Sl+2πm)[π,π],m,(S^{l}+2\pi m)\cap[-\pi,\pi],m\in\mathbb{Z}, are pairwise disjoint, each term of the previous sum should be zero ; that is (χSlμ^l)(x+2πk)=0(\chi_{S^{l}}\hat{\mu}_{l})(x+2\pi k)=0 for all m,lm,l and for Lebesgue almost every x[π,π]x\in[-\pi,\pi]. By Property (2) in the definition of a set of uniqueness, we conclude μl^=0\widehat{\mu_{l}}=0. Therefore μl=0\mu_{l}=0 for each 1lk1\leq l\leq k and consequently the translates of gg is dense in C0(;)kC_{0}(\mathbb{R};\mathbb{C})^{k}. ∎

Proposition 6.5.

Let f:[0,1]f:[0,1]\circlearrowleft be a homeomorphism of the interval. Then

Mult(f)=e([0,1],f).\text{\rm Mult}(f)=\sharp\mathcal{M}_{e}([0,1],f).
Proof.

We first deal with the case of an increasing homeomorphism. The ergodic measures of ff are the Dirac measures at these fixed points. Notice that ff has at least two fixed points, 0 and 11. If it has infinitely many fixed points, then Mult(Uf)e([0,1],f)=\text{\rm Mult}(U_{f})\geq\sharp\mathcal{M}_{e}([0,1],f)=\infty. Now assume it has finitely many fixed points. Let 2k+1<+2\leq k+1<+\infty be the number of fixed points. Since φ(x)φf(x)=0\varphi(x)-\varphi\circ f(x)=0 for any continuous function φC(X)\varphi\in C(X) and any fixed point xx, the space Bf([0,1])¯\overline{B_{f}([0,1])} is the set of real continuous maps on the interval which vanishes at the fixed points. It follows that the operator U¯f\underline{U}_{f} is spectrally conjugate to V:C0()kV:C_{0}(\mathbb{R})^{k}\circlearrowleft, (fi)1ik(fi(+1))1ik(f_{i})_{1\leq i\leq k}\mapsto\left(f_{i}(\cdot+1)\right)_{1\leq i\leq k}. By Lemma 6.4 we have Mult(Uf¯)=1\text{\rm Mult}(\underline{U_{f}})=1. It follows then from Proposition 2.9 and Proposition 2.6 that

(6\cdot1) e([0,1],f)Mult(Uf)e([0,1],f)+Mult(Uf¯)1=e([0,1],f).\sharp\mathcal{M}_{e}([0,1],f)\leq\text{\rm Mult}(U_{f})\leq\sharp\mathcal{M}_{e}([0,1],f)+\text{\rm Mult}(\underline{U_{f}})-1=\sharp\mathcal{M}_{e}([0,1],f).

It remains to consider the case of a decreasing homeomorphism ff. Let 0<a<10<a<1 be the unique fixed point of ff. Then f2:[0,a]f^{2}:[0,a]\circlearrowleft is an increasing homeomorphism. Let 0=x1<x2<<xk=a0=x_{1}<x_{2}<\cdots<x_{k}=a be the fixed points of f2|[0,a]f^{2}|_{[0,a]}. Then the ergodic measures of ff are the atomic periodic measures δa\delta_{a} and 12(δxi+δf(xi))\frac{1}{2}\left(\delta_{x_{i}}+\delta_{f(x_{i})}\right) for i=1,,k1i=1,\cdots,k-1. In particular we have k=e([0,1],f)k=\sharp\mathcal{M}_{e}([0,1],f). From the previous case there is a generating family 𝒢={g1,,gk}\mathcal{G}=\{g_{1},\cdots,g_{k}\} for f2:[0,a]f^{2}:[0,a]\circlearrowleft. Let hC([0,1])h\in C([0,1]). For any ϵ>0\epsilon>0, there are NN\in\mathbb{N}, al,na_{l,n} and bl,nb_{l,n}, for l=1,,kl=1,\cdots,k and |n|N|n|\leq N, (depending on ϵ\epsilon), such that

(6\cdot2) hl,nal,nglf2n[0,a],\displaystyle\|h-\sum_{l,n}a_{l,n}g_{l}\circ f^{2n}\|_{{[0,a]},\infty} <ϵ,\displaystyle<\epsilon,

and

(6\cdot3) hf1l,nbl,nglf2nh(a)[0,a],\displaystyle\|h\circ f^{-1}-\sum_{l,n}b_{l,n}g_{l}\circ f^{2n}-h(a)\|_{{[0,a]},\infty} <ϵ,\displaystyle<\epsilon,

where g[0,a],=supx[0,a]|g(x)|\|g\|_{{[0,a]},\infty}=\sup_{x\in[0,a]}|g(x)|. We consider the extension gl~\tilde{g_{l}} of glg_{l} to [0,1][0,1] with gl~=gl(a)\tilde{g_{l}}=g_{l}(a) on [a,1][a,1]. We check now that 𝒢~={g~1,,g~k}\tilde{\mathcal{G}}=\{\tilde{g}_{1},\cdots,\tilde{g}_{k}\} is generating for ff. It follows from (6\cdot2) and (6\cdot3) at x=ax=a that

(6\cdot4) |h(a)l,nal,ngl(a)|<ϵ and |l,nbl,ngl(a)|<ϵ.\left|h(a)-\sum_{l,n}a_{l,n}g_{l}(a)\right|<\epsilon\text{ and }\left|\sum_{l,n}b_{l,n}g_{l}(a)\right|<\epsilon.

Observe that

h\displaystyle h =h|[0,a]+(hf1|[0,a])f|[a,1].\displaystyle=h|_{[0,a]}+(h\circ f^{-1}|_{[0,a]})\circ f|_{[a,1]}.

Combining (6\cdot4) with (6\cdot2), we obtain that

hl,nal,ngl~f2nl,nbl,ngl~f2n+1[0,a],\displaystyle\|h-\sum_{l,n}a_{l,n}\tilde{g_{l}}\circ f^{2n}-\sum_{l,n}b_{l,n}\tilde{g_{l}}\circ f^{2n+1}\|_{[0,a],\infty}
=\displaystyle= hl,nal,nglf2nl,nbl,ngl(a)[0,a]<2ϵ.\displaystyle\|h-\sum_{l,n}a_{l,n}{g_{l}}\circ f^{2n}-\sum_{l,n}b_{l,n}{g_{l}}(a)\|_{[0,a]\infty}<2\epsilon.

Similarly, combining (6\cdot4) with (6\cdot3) we get

hl,nal,ngl~f2nl,nbl,ngl~f2n+1[a,1],\displaystyle\|h-\sum_{l,n}a_{l,n}\tilde{g_{l}}\circ f^{2n}-\sum_{l,n}b_{l,n}\tilde{g_{l}}\circ f^{2n+1}\|_{[a,1],\infty}
=\displaystyle= hl,nal,ngl(a)l,nbl,nglf2n+1[a,1]\displaystyle\|h-\sum_{l,n}a_{l,n}{g_{l}}(a)-\sum_{l,n}b_{l,n}{g_{l}}\circ f^{2n+1}\|_{[a,1]\infty}
\displaystyle\leq |h(a)l,nal,ngl(a)|+hl,nbl,nglf2n+1h(a)[a,1]<2ϵ.\displaystyle|h(a)-\sum_{l,n}a_{l,n}{g_{l}}(a)|+\|h-\sum_{l,n}b_{l,n}{g_{l}}\circ f^{2n+1}-h(a)\|_{[a,1]\infty}<2\epsilon.

Therefore we have

hl,nal,ngl~f2nl,nbl,ngl~f2n+1<2ϵ.\|h-\sum_{l,n}a_{l,n}\tilde{g_{l}}\circ f^{2n}-\sum_{l,n}b_{l,n}\tilde{g_{l}}\circ f^{2n+1}\|_{\infty}<2\epsilon.

We conclude that 𝒢~\tilde{\mathcal{G}} is generating for ff as ϵ>0\epsilon>0 and hC([0,1])h\in C([0,1]) have been chosen arbitrarily.

Question 6.6.

What is the topological multiplicity of a Morse-Smale diffeomorphism?

6.4. Substitution Subshift

For a finite alphabet 𝒜\mathcal{A} with 𝒜2\sharp\mathcal{A}\geq 2 we let 𝒜+\mathcal{A}^{+} be the set of associated finite words, i.e. finite sequences of letters in 𝒜\mathcal{A}. For a word ww in 𝒜+\mathcal{A}^{+} we let |w||w| be the number of letters of ww. A substitution is a map ζ:𝒜𝒜+\zeta:\mathcal{A}\rightarrow\mathcal{A}^{+}. We say ζ\zeta has the constant length if |ζ(a)||\zeta(a)| does not depend on a𝒜a\in\mathcal{A}. The substitution may be extended to finite and infinite (one-sided) words by concatenation: ζ(u0u1)=ζ(u0)ζ(u1)\zeta(u_{0}u_{1}\cdots)=\zeta(u_{0})\zeta(u_{1})\cdots for ui𝒜u_{i}\in\mathcal{A} and ii\in\mathbb{N}. In this way ζ\zeta can be iterated.

The langage ζ\mathcal{L}_{\zeta} of ζ\zeta is the set of all subwords of ζn(a)\zeta^{n}(a) over a𝒜a\in\mathcal{A} and nn\in\mathbb{N}, i.e. the set of finite sequence of consecutive letters appearing in the words ζn(a)\zeta^{n}(a). The associated subshift Xζ𝒜X_{\zeta}\subset\mathcal{A}^{\mathbb{Z}} is the set of sequences (xn)n(x_{n})_{n\in\mathbb{Z}} such that xpxp+1xqx_{p}x_{p+1}\cdots x_{q} belongs to ζ\mathcal{L}_{\zeta} for any pqp\leq q\in\mathbb{Z}. The map ζ\zeta may be extended to map from XζX_{\zeta} to XζX_{\zeta} by letting

ζ(x)=ζ(x1)ζ(x1)|ζ(x0)ζ(x1)\zeta(x)=\cdots\zeta(x_{-1})\zeta(x_{-1})\,|\,\zeta(x_{0})\cdots\zeta(x_{1})\cdots

for any x=(xn)nXζx=(x_{n})_{n\in\mathbb{Z}}\in X_{\zeta}, where the non negative coordinates of ζ(x)\zeta(x) lie after the symbol “||”. For w𝒜w\in\mathcal{A} we recall that the associated cylinder [w][w] in XζX_{\zeta} is given by the sequences in XζX_{\zeta} whose |w||w| first coordinates coincide with ww.

A substitution ζ\zeta is said to be primitive when there exists nn\in\mathbb{N}^{*} such that for any letters a,ba,b the word ζn(a)\zeta^{n}(a) contains the letter bb. It is well-known that the shift σ\sigma on XζX_{\zeta} with primitive ζ\zeta is then minimal and uniquely ergodic (see [Que10]). In the followings, we only consider aperiodic primitive substitutions, i.e. primitive substitutions such that XζX_{\zeta} is not reduced to a periodic orbit.

We will make use of the standard recognazibility property of primitive substitutions:

Theorem 6.7.

[Mos96] Let ζ\zeta be an aperiodic primitive substitution. Then for any positive integer nn, the family

Pn:={σk(ζn[a]),a𝒜 and 0k<|ζn(a)|}P_{n}:=\{\sigma^{k}\left(\zeta^{n}[a]\right),\ a\in\mathcal{A}\ \text{ and }0\leq k<|\zeta^{n}(a)|\}

defines a partition of XζX_{\zeta} into clopen sets.

Moreover ζn:Xζζn(Xζ)\zeta^{n}:X_{\zeta}\rightarrow\zeta^{n}(X_{\zeta}) is a topological conjugacy between (Xζ,σ)(X_{\zeta},\sigma) and the induced system (ζn(Xζ),σζn(Xζ))\left(\zeta^{n}(X_{\zeta}),\sigma_{\zeta^{n}(X_{\zeta})}\right) defined as σζn(Xζ)(x)=σ|ζn(x0)|(x)\sigma_{\zeta^{n}(X_{\zeta})}(x)=\sigma^{|\zeta^{n}(x_{0})|}(x) for any x=(xk)kζn(Xζ)x=(x_{k})_{k\in\mathbb{Z}}\in\zeta^{n}(X_{\zeta}).

We define two equivalence relations on 𝒜\mathcal{A} as follows:

adb if and only if  ζ(a) and ζ(b) begin with the same letter;a\sim^{d}b\text{ if and only if }\text{ $\zeta(a)$ and $\zeta(b)$ begin with the same letter};
afb if and only if  ζ(a) and ζ(b) end with the same letter.a\sim^{f}b\text{ if and only if }\text{ $\zeta(a)$ and $\zeta(b)$ end with the same letter}.

We then let 𝒜d\mathcal{A}_{d} and 𝒜f\mathcal{A}_{f} be respectively the equivalence classes of d\sim^{d} and f\sim^{f}. We let NdN_{d} and NfN_{f} be the number of equivalence classes for d\sim^{d} and f\sim^{f}.

Theorem 6.8.

Let ζ\zeta be a primitive aperiodic substitution (resp. with constant length). Then the multiplicity (Xζ,σ)(X_{\zeta},\sigma) is less than or equal to (resp. strictly less than)

Kζ=𝒜(Nd+Nf)+min(Nd,Nf)(𝒜+NdNf),K_{\zeta}=\sharp\mathcal{A}(N_{d}+N_{f})+\min(N_{d},N_{f})-(\sharp\mathcal{A}+N_{d}N_{f}),

which is less than or equal to 𝒜2\sharp\mathcal{A}^{2}.

For a proper substitution (i.e. when Nd=Nf=1N_{d}=N_{f}=1) we get Kζ=𝒜K_{\zeta}=\sharp\mathcal{A}. The partitions (Pn)n(P_{n})_{n\in\mathbb{N}} in Theorem 6.7 are not refining, i.e. the diameter of PnP_{n} does not go to zero with nn in general. To prove Theorem 6.8 we use the following finer clopen partitions (Qn)n(Q_{n})_{n\in\mathbb{N}}. For a subset J𝒜J\subset\mathcal{A} and a𝒜a\in\mathcal{A} we let [a|J]Xζ[a|J]\subset X_{\zeta} (resp. [J|a][J|a]) be the union of σ([aj])\sigma([aj]) (resp. σ([ja])\sigma([ja])) over jJj\in J).

Lemma 6.9.

Let ζ\zeta be an aperiodic primitive substitution and let MnM_{n} be the integer part of mina𝒜|ζn1(a)|/2\min_{a\in\mathcal{A}}|\zeta^{n-1}(a)|/2. Then for any positive integer nn, the family

Qn:=\displaystyle Q_{n}:= {σl(ζn[Jf|a]):a𝒜,Jf𝒜f and 0l<Mn}\displaystyle\{\sigma^{l}\left(\zeta^{n}[J_{f}|a]\right):\ a\in\mathcal{A},\ J_{f}\in\mathcal{A}_{f}\text{ and }0\leq l<M_{n}\}\bigcup
{σl(ζn[a|Jd]):a𝒜,Jd𝒜d and 0l<|ζn(a)|Mn}\displaystyle\{\sigma^{-l}\left(\zeta^{n}[a|J_{d}]\right):\ a\in\mathcal{A},\ J_{d}\in\mathcal{A}_{d}\text{ and }0\leq l<|\zeta^{n}(a)|-M_{n}\}

defines a partition of XζX_{\zeta} into clopen sets whose diameter is going to zero when nn goes to infinity.

Proof.

Clearly, for 0k<Mn0\leq k<M_{n}, the clopen sets σk(ζn[Jf|a])\sigma^{k}\left(\zeta^{n}[J_{f}|a]\right) for Jf𝒜fJ_{f}\in\mathcal{A}_{f} defines a partition of σk(ζn[a])\sigma^{k}(\zeta^{n}[a]). On the other hand, for 0l<|ζn(a)|Mn0\leq l<|\zeta^{n}(a)|-M_{n}, the clopen sets σl(ζn[a|Jd])\sigma^{-l}\left(\zeta^{n}[a|J_{d}]\right) for Jd𝒜dJ_{d}\in\mathcal{A}_{d} defines a partition of σ|ζn(a)|l(ζn[a])\sigma^{|\zeta^{n}(a)|-l}(\zeta^{n}[a]). Therefore we deduce then from Theorem 6.7 that QnQ_{n} is a partition of XζX_{\zeta}. It remains to show that their diameter is going to zero when nn goes to infinity. Fix a𝒜a\in\mathcal{A} and Jf𝒜fJ_{f}\in\mathcal{A}_{f} and let b𝒜b\in\mathcal{A} such that ζ(j)\zeta(j) ends with the letter bb for any jJfj\in J_{f}. Then any sequence in ζn[Jf|a]\zeta^{n}[J_{f}|a] lies in σ|ζn1(b)|([ζn1(b)ζn(a)])\sigma^{|\zeta^{n-1}(b)|}\left([\zeta^{n-1}(b)\zeta^{n}(a)]\right). In particular, for any x,yσl(ζn[Jf|a])x,y\in\sigma^{l}\left(\zeta^{n}[J_{f}|a]\right) with 0l<Mn0\leq l<M_{n}, we have xMn+1xMn1=yMn+1yMn1x_{-M_{n}+1}\cdots x_{M_{n}-1}=y_{-M_{n}+1}\cdots y_{M_{n}-1} since we have Mnmax(|ζn(a)|/2,|ζn1(b)|)M_{n}\leq\max(|\zeta^{n}(a)|/2,|\zeta^{n-1}(b)|). Arguing similarly with σl(ζn[a|Jd])\sigma^{-l}\left(\zeta^{n}[a|J_{d}]\right) we conclude that the diameter of QnQ_{n} goes to zero when nn goes to infinity, because MnM_{n} goes to infinity. ∎

For a finite collection 𝒞\mathcal{C} of clopen sets and for nn\in\mathbb{N}^{*} we let n𝒞\mathcal{F}_{n}^{\mathcal{C}} be the set of indicator functions χζn(C)\chi_{\zeta^{n}(C)} with C𝒞C\in\mathcal{C}. We also write (Qn)\mathcal{F}(Q_{n}) for the set of indicator functions χA\chi_{A} with AQnA\in Q_{n}.

Lemma 6.10.

There is a finite collection 𝒞\mathcal{C} of clopen sets with 𝒞Kζ\sharp\mathcal{C}\leq K_{\zeta}, such that

n,V(Qn)=Vn𝒞.\forall n\in\mathbb{N}^{*},\ V_{\mathcal{F}(Q_{n})}=V_{\mathcal{F}_{n}^{\mathcal{C}}}.
Proof.

Without loss of generality we may assume NdNfN_{d}\leq N_{f}. Fix Jf¯𝒜f\overline{J_{f}}\in\mathcal{A}_{f} and id¯Jd\overline{i_{d}}\in J_{d} for any Jd𝒜dJ_{d}\in\mathcal{A}_{d}. We consider the collection

𝒞:={[a|Jd],[Jf|b]:a𝒜,Jd𝒜d,Jf𝒜f{Jf¯},bJd{id¯}}.\mathcal{C}:=\left\{[a|J_{d}],\ [J_{f}|b]\ :\ a\in\mathcal{A},\ J_{d}\in\mathcal{A}_{d},\ J_{f}\in\mathcal{A}_{f}\setminus\{\overline{J_{f}}\},\ b\in J_{d}\setminus\{\overline{i_{d}}\}\ \right\}.

Clearly V(Qn)Vn𝒞V_{\mathcal{F}(Q_{n})}\supset V_{\mathcal{F}_{n}^{\mathcal{C}}} and

𝒞Nd𝒜+(Nf1)(𝒜Nd)=Kζ.\sharp\mathcal{C}\leq N_{d}\sharp\mathcal{A}+(N_{f}-1)(\sharp\mathcal{A}-N_{d})=K_{\zeta}.

We show now the other inclusion V(Qn)Vn𝒞V_{\mathcal{F}(Q_{n})}\subset V_{\mathcal{F}_{n}^{\mathcal{C}}}. Let 𝒞~n\tilde{\mathcal{C}}_{n} be the collection of clopen sets EE satisfying χζn(E)Vn𝒞\chi_{\zeta^{n}(E)}\in V_{\mathcal{F}_{n}^{\mathcal{C}}}. It is enough to check [a|Jd][a|J_{d}] and [Jf|b][J_{f}|b] lie in 𝒞~n\tilde{\mathcal{C}}_{n} for any a,b𝒜a,b\in\mathcal{A}, Jd𝒜dJ_{d}\in\mathcal{A}_{d} and Jf𝒜fJ_{f}\in\mathcal{A}_{f}. For E,E𝒞~nE,E^{\prime}\in\tilde{\mathcal{C}}_{n}, the union EEE\cup E^{\prime} (resp. EEE\setminus E^{\prime}) also belongs to 𝒞~n\tilde{\mathcal{C}}_{n} if EE=E\cap E^{\prime}=\emptyset (resp. EEE^{\prime}\subset E), because χζn(EE)=χζn(E)+χζn(E)\chi_{\zeta^{n}(E\cup E^{\prime})}=\chi_{\zeta^{n}(E^{\prime})}+\chi_{\zeta^{n}(E)} (resp. χζn(EE)=χζn(E)χζn(E)\chi_{\zeta^{n}(E\setminus E^{\prime})}=\chi_{\zeta^{n}(E)}-\chi_{\zeta^{n}(E^{\prime})} ) as ζn\zeta^{n} is one-to-one.

Then for any Jd𝒜dJ_{d}\in\mathcal{A}_{d} and any Jf𝒜fJ_{f}\in\mathcal{A}_{f}, the set [Jf|Jd]=aJf[a|Jd][J_{f}|J_{d}]=\coprod_{a\in J_{f}}[a|J_{d}] lie in 𝒞~n\tilde{\mathcal{C}}_{n}, because they are disjoint unions of elements of 𝒞𝒞~n\mathcal{C}\subset\tilde{\mathcal{C}}_{n}. We have also [Jf|id¯]=[Jf|Jd](Jdbid¯[Jf|b])𝒞~n[J_{f}|\overline{i_{d}}]=[J_{f}|J_{d}]\setminus\left(\coprod_{J_{d}\ni b\neq\overline{i_{d}}}[J_{f}|b]\right)\in\tilde{\mathcal{C}}_{n} for any JfJf¯J_{f}\neq\overline{J_{f}}. Finally for any b𝒜b\in\mathcal{A} we get [Jf¯|b]=[b](JfJf¯[Jf|b])[\overline{J_{f}}|b]=[b]\setminus\left(\coprod_{J_{f}\neq\overline{J_{f}}}[J_{f}|b]\right). But [b]=σ1([b|])[b]=\sigma^{-1}([b|]), then ζn([b])=σ|ζn(b)|(ζn([b|]))\zeta^{n}([b])=\sigma^{-|\zeta^{n}(b)|}\left(\zeta^{n}([b|])\right) and [b|]=Jd[b|Jd]𝒞~n[b|]=\coprod_{J_{d}}[b|J_{d}]\in\tilde{\mathcal{C}}_{n}. Therefore χζn([b])=χζn([b|])σ|ζn(b)|Vn𝒞\chi_{\zeta^{n}([b])}=\chi_{\zeta^{n}([b|])}\circ\sigma^{|\zeta^{n}(b)|}\in V_{\mathcal{F}_{n}^{\mathcal{C}}}, thus [b]𝒞~n[b]\in\tilde{\mathcal{C}}_{n}. We conclude that [Jf¯|b][\overline{J_{f}}|b] also belongs to 𝒞~n\tilde{\mathcal{C}}_{n}.

Proof of Theorem 6.9.

Let 𝒞\mathcal{C} be the collection given by Lemma 6.10. Then for any AQnA\in Q_{n}, the function χA\chi_{A} belongs to Vn𝒞V_{\mathcal{F}_{n}^{\mathcal{C}}}. As the diameter of QnQ_{n} goes to zero with nn, the hypothesis of Lemma 4.1 are satisfied, so that Mult(Xζ,σ)𝒞Kζ\text{\rm Mult}(X_{\zeta},\sigma)\leq\sharp\mathcal{C}\leq K_{\zeta}.

When moreover the substitution has constant lenght LL, then the induced system on ζn(Xζ)\zeta^{n}(X_{\zeta}) is just the power σLn\sigma^{L^{n}}. By Theorem 6.7 the induced system is topologically conjugated to the substitution system, thus it is also uniquely ergodic. Let νn\nu_{n} be the corresponding unique invariant measure. In particular for any clopen subset CC of ζn(X)\zeta^{n}(X), the sequence of continuous functions

1K0kKχCσLn\frac{1}{K}\sum_{0\leq k\leq K}\chi_{C}\circ\sigma^{L^{n}}

is converging uniformly to νn(C)χζn(X)\nu_{n}(C)\chi_{\zeta^{n}(X)} when KK goes to infinity. When defining 𝒞\mathcal{C} in the previous lemma, we let all [a|Jd][a|J_{d}] in 𝒞\mathcal{C} for a𝒜a\in\mathcal{A} and Jd𝒜dJ_{d}\in\mathcal{A}_{d}. The elements of this form defines a partition and we may choose one such element [a¯|Jd¯][\overline{a}|\overline{J_{d}}] such that ζn([a¯|Jd¯])\zeta^{n}([\overline{a}|\overline{J_{d}}]) has not full measure for the induced system. We can remove this element from 𝒞\mathcal{C}. We let 𝒞\mathcal{C}^{\prime} be this new collection. If C=[a|Jd]C^{\prime}=[a|J_{d}] is an element of 𝒞\mathcal{C}^{\prime} with νn(ζn(C))>0\nu_{n}(\zeta^{n}(C^{\prime}))>0, then χζn(X)\chi_{\zeta^{n}(X)} belongs to Vχζn(C)V_{\chi_{\zeta^{n}(C^{\prime})}} therefore χζn([a¯|Jd¯])\chi_{\zeta^{n}([\overline{a}|\overline{J_{d}}])} lies in Vn𝒞V_{\mathcal{F}_{n}^{\mathcal{C}^{\prime}}}, i.e. Vn𝒞=Vn𝒞V_{\mathcal{F}_{n}^{\mathcal{C}}}=V_{\mathcal{F}_{n}^{\mathcal{C}^{\prime}}}. Therefore Mult(Xζ,σ)𝒞=𝒞1<Kζ\text{\rm Mult}(X_{\zeta},\sigma)\leq\sharp\mathcal{C}^{\prime}=\sharp\mathcal{C}-1<K_{\zeta}.

For the Thue-Morse substitution: 0010\mapsto 01 and 1101\mapsto 10, we get Nd=Nf=2N_{d}=N_{f}=2, thus Mult(Xζ,σ)<Kζ=4\text{\rm Mult}(X_{\zeta},\sigma)<K_{\zeta}=4.

Question 6.11.

What is the topological multiplicity of the Thue-Morse substitution?

7. Subshifts with linear growth complexity

We consider a subshift X𝒜X\subset\mathcal{A}^{\mathbb{Z}} with letters in a finite alphabet 𝒜\mathcal{A}. For x𝒜x\in\mathcal{A}^{\mathbb{Z}} we denote by x=(xn)nx=(x_{n})_{n\in\mathbb{Z}} for xn𝒜x_{n}\in\mathcal{A}. Let n(X)𝒜n\mathcal{L}_{n}(X)\subset\mathcal{A}^{n} be the finite words of XX of length nn, i.e. n(X)={xkxk+1xk+n1:xX,k}\mathcal{L}_{n}(X)=\{x_{k}x_{k+1}\dots x_{k+n-1}:x\in X,k\in\mathbb{Z}\}. The word complexity of XX is given by

n,pX(n)=n(X).\forall n\in\mathbb{N},\ \ p_{X}(n)=\sharp\mathcal{L}_{n}(X).

We suppose that XX is aperiodic and has linear growth, that is, for some kk\in\mathbb{N}^{*}

(7\cdot1) lim infnpX(n)nk.\liminf_{n}\frac{p_{X}(n)}{n}\leq k.

Boshernitzan [Bos92] showed that such a subshift admits at most kk ergodic measures. By Theorem 5.5 in [DDMP21] such subshifts, when assumed to be moreover minimal, have topological rank less than or equal to (1+k𝒜2)2(k+2)(1+k\sharp\mathcal{A}^{2})^{2(k+2)}. We show in this section the following upper bound on the topological multiplicity.

Theorem 7.1.

Any aperiodic subshift XX with lim infnpX(n)nk\liminf_{n\to\infty}\frac{p_{X}(n)}{n}\leq k has topological multiplicity less than or equal to 2k2k.

One may wonder if the upper bound in Theorem 7.1 is sharp.

Question 7.2.

Is an aperiodic subshift XX with lim infnpX(n)n=1\liminf_{n\to\infty}\frac{p_{X}(n)}{n}=1 topologically simple?

In order to prove Theorem 7.1, we define some notations. Let QnQ_{n} be the subset of n(X)\mathcal{L}_{n}(X) given by words ww such that there are several letters a𝒜a\in\mathcal{A} with wan+1(X)wa\in\mathcal{L}_{n+1}(X). We also let Qn+1Q^{\prime}_{n+1} be the (n+1)(n+1)-words wawa as above. Clearly, we have

(7\cdot2) QnpX(n+1)pX(n) and Qn+1=Qn+pX(n+1)pX(n).\sharp Q_{n}\leq p_{X}(n+1)-p_{X}(n)\text{ and }\sharp Q^{\prime}_{n+1}=\sharp Q_{n}+p_{X}(n+1)-p_{X}(n).

Through this section, we always assume the subshift is aperiodic and satisfies the linear growth (7\cdot1).

Lemma 7.3.

The subset of integers

𝒩={n:pX(n+1)<(k+1)(n+1) and pX(n+1)pX(n)k}.\mathcal{N}=\{n\in\mathbb{N}:p_{X}(n+1)<(k+1)(n+1)\text{ and }p_{X}(n+1)-p_{X}(n)\leq k\}.

is infinite.

For the sake of completeness, we reproduce the proof, which is contained in Theorem 2.2 of [Bos84].

Proof.

By (7\cdot1), we have

lim infn(pX(n)(k+1)n)=.\liminf_{n}(p_{X}(n)-(k+1)n)=-\infty.

It follows that

={n:pX(n+1)(k+1)(n+1)min{0,min1mn{pX(m)(k+1)m}}1}\mathcal{M}=\left\{n\in\mathbb{N}:p_{X}(n+1)-(k+1)(n+1)\leq\min\{0,\min_{1\leq m\leq n}\{p_{X}(m)-(k+1)m\}\}-1\right\}

is an infinite set. For any nn\in\mathcal{M}, we have

pX(n+1)pX(n)(k+1)(n+1)(k+1)n1=k.p_{X}(n+1)-p_{X}(n)\leq(k+1)(n+1)-(k+1)n-1=k.

On the other hand, for any nn\in\mathcal{M}, we get

pX(n+1)(k+1)(n+1)1.p_{X}(n+1)\leq(k+1)(n+1)-1.

This implies that 𝒩\mathcal{M}\subset\mathcal{N}. Therefore, the set 𝒩\mathcal{N} is infinite. ∎

Lemma 7.4 ([Bos84], Lemma 4.1).

For any n𝒩n\in\mathcal{N} and m(k+2)(n+1)m\geq(k+2)(n+1), any word wmw\in\mathcal{L}_{m} contains a subword in QnQ_{n}.

For the sake of completeness we provide a proof here.

Proof.

We prove it by contradiction. Assume to the contrary that all (mn+1)(m-n+1) nn-subwords of ww do not belong to QnQ_{n}. That means that each of these nn-blocks determines uniquely the next letter. Since mn+12(k+1)n>pX(n)m-n+1\geq 2(k+1)n>p_{X}(n), at least one nn-word appears more than one time as a subword of ww. Therefore XX contains a periodic point. This contradicts our assumption. ∎

Now we show that any cylinder of length less than nn can be decomposed as the cylinders of elements in Qn+1Q^{\prime}_{n+1} after translations.

Lemma 7.5.

Let n𝒩n\in\mathcal{N}. Any cylinder [w][w] with length of ww less than nn may be written uniquely as a finite disjoint union of sets of the form σp[qn+1]\sigma^{p}[q^{\prime}_{n+1}] with pp\in\mathbb{N}, qn+1Qn+1q^{\prime}_{n+1}\in Q^{\prime}_{n+1}, such that σt[qn+1][qn]=\sigma^{t}[q^{\prime}_{n+1}]\cap[q_{n}]=\emptyset for any 0<t<p0<t<p and any qnQnq_{n}\in Q_{n}.

Remark that by Lemma 7.4 the integers pp belongs to [0,(k+2)(n+1)][0,(k+2)(n+1)].

Proof.

Let [w][w] be a cylinder associated to a word wl(X)w\in\mathcal{L}_{l}(X) with l<nl<n. For x[w]x\in[w], we let KxK_{x}\in\mathbb{Z} be the largest integer jj less than ll such that xjn+1xjx_{j-n+1}\cdots x_{j} belongs to QnQ_{n}. Then the word wn+1x=xKxn+1xKx+1w_{n+1}^{x}=x_{K_{x}-n+1}\cdots x_{K_{x}+1} belongs to Qn+1Q^{\prime}_{n+1}. Observe also that by Lemma 7.4 we have l1Kx(k+2)(n+1)l-1-K_{x}\leq(k+2)(n+1), thus n1Kx(k+3)(n+1)n-1-K_{x}\leq(k+3)(n+1). Let Wn+1W_{n+1} be the collection of these words wn+1xw_{n+1}^{x} over x[w]x\in[w]. By definition of KxK_{x} and QnQ_{n} the word wn+1xw_{n+1}^{x} completely determines the l1Kxl-1-K_{x} next letters, that is to say,

[wn+1x]=[xKxn+1xl1].[w_{n+1}^{x}]=[x_{K_{x}-n+1}\cdots x_{l-1}].

For x,y[w]x,y\in[w] the sets σn1Kx[wn+1x]\sigma^{n-1-K_{x}}[w_{n+1}^{x}] and σn1Ky[wn+1y]\sigma^{n-1-K_{y}}[w_{n+1}^{y}] are either disjoint or equal. As xx belongs to [w][w] we have xσn1Kx[wn+1x][w]x\in\sigma^{n-1-K_{x}}[w_{n+1}^{x}]\subset[w] and finally

[w]=wn+1xWn+1σn1Kxwn+1x.[w]=\coprod_{w_{n+1}^{x}\in W_{n+1}}\sigma^{n-1-K_{x}}w^{x}_{n+1}.

We complete the proof. ∎

Proof of Theorem 7.1.

By (7\cdot2) and the definition of 𝒩\mathcal{N} we have for n𝒩n\in\mathcal{N} :

Qn+1\displaystyle\sharp Q^{\prime}_{n+1} =Qn+pX(n+1)pX(n),\displaystyle=\sharp Q_{n}+p_{X}(n+1)-p_{X}(n),
2(pX(n+1)pX(n)),\displaystyle\leq 2(p_{X}(n+1)-p_{X}(n)),
2k.\displaystyle\leq 2k.

For n𝒩n\in\mathcal{N} we let Fn={χ[qn+1],qn+1Qn+1}F_{n}=\{\chi_{[q^{\prime}_{n+1}]},\ q^{\prime}_{n+1}\in Q^{\prime}_{n+1}\}. By Lemma 7.5, any cylinder [w][w] with length less than nn is a finite disjoint union of σp[qn+1]\sigma^{p}[q^{\prime}_{n+1}]. In particular χ[w]\chi_{[w]} lies in VFnV_{F_{n}}. We may therefore apply Lemma 4.1 to (Fn)n𝒩(F_{n})_{n\in\mathcal{N}} and we get

Mult(X,σ)supn𝒩Qn+12k.\text{\rm Mult}(X,\sigma)\leq\sup_{n\in\mathcal{N}}\sharp Q^{\prime}_{n+1}\leq 2k.

7.1. Multiplicity of invariant measures

It follows from Theorem 7.1 and Lemma 2.4 that any ergodic measure has (ergodic) multiplicity bounded by 2k2k. In fact we may refine this result as follows:

Theorem 7.6.

Let XX be an aperiodic subshift with lim infnpX(n)nk\liminf_{n\to\infty}\frac{p_{X}(n)}{n}\leq k. Then

μe(X,σ)Mult(μ)2k.\sum_{\mu\in\mathcal{M}_{e}(X,\sigma)}\text{\rm Mult}(\mu)\leq 2k.

In order to prove Theorem 7.6, we first recall some notations and then show two lemmas for general aperiodic subshifts. We have learned from the referee that some parts of our proofs overlap with results in [Cre23][Esp23]. Let (Y,σ)(Y,\sigma) be an aperiodic subshift. For two finite words ww and vv, we denote by N(w|v)N(w|v) the number of times that ww appears as a subword of vv. Also, We define d(w|v)=N(w|v)/|v|d(w|v)=N(w|v)/|v|. For a generic point xx of a measure μ\mu, we have

limnd(w|x1n)=μ([w]),\lim_{n\to\infty}d(w|x_{1}^{n})=\mu([w]),

where x1n=x1x2xnx_{1}^{n}=x_{1}x_{2}\dots x_{n}. For a finite word vv, we denote by vm=vvvm timesv^{\otimes m}=\underbrace{vv\dots v}_{m\text{ times}}. For a finite word ww, we denote by

νw=1|w|k=0|w|1δσk(w¯),\nu_{w}=\frac{1}{|w|}\sum_{k=0}^{|w|-1}\delta_{\sigma^{k}(\bar{w})},

where |w||w| is the length of ww and w¯𝒜\bar{w}\in\mathcal{A}^{\mathbb{Z}} is the periodization of ww, i.e. ww^{\otimes\infty}.

Let wnw_{n} be a word of length nn. Foy any nn, we put n=(wn):=min{1<n:[wn]σ([wn])}\ell_{n}=\ell(w_{n}):=\min\{1\leq\ell<n:[w_{n}]\cap\sigma^{\ell}([w_{n}])\not=\emptyset\} and Ln:=1+{1<n:[wn]σ([wn])}L_{n}:=1+\sharp\{1\leq\ell<n:[w_{n}]\cap\sigma^{\ell}([w_{n}])\not=\emptyset\}, with the convention min=n\min\emptyset=n. Let vn=v(wn)v_{n}=v(w_{n}) be the first n\ell_{n}-subword of wnw_{n}. It follows that wn=vnKnv^nw_{n}=v_{n}^{\otimes K_{n}}\hat{v}_{n} with v^nvn\hat{v}_{n}\neq v_{n} being a prefix of vnv_{n}. Then Kn=n/nLnK_{n}=\lfloor n/\ell_{n}\rfloor\geq L_{n}. Observe that for any xXx\in X, pnp\geq\ell_{n} and any word uu of length less than n\ell_{n} we have

(7\cdot3) N(u|x1p)\displaystyle N(u|x_{1}^{p}) N(u|wn)N(wn|x1p)Ln,\displaystyle\geq N(u|w_{n})\frac{N(w_{n}|x_{1}^{p})}{L_{n}},
N(u|vn)KnN(wn|x1p)Ln,\displaystyle\geq N(u|v_{n})K_{n}\frac{N(w_{n}|x_{1}^{p})}{L_{n}},
N(u|vn)N(wn|x1p).\displaystyle\geq N(u|v_{n})N(w_{n}|x_{1}^{p}).

In the next two lemmas we assume that

  • the subshift (Y,σ)(Y,\sigma) is aperiodic;

  • wnn(Y)w_{n}\in\mathcal{L}_{n}(Y) for nn\in\mathbb{N};

  • νwnn𝒩n+νe(Y,σ)\nu_{w_{n}}\xrightarrow[n\in\mathcal{N}]{n\to+\infty}\nu\in\mathcal{M}_{e}(Y,\sigma), i.e. νwn\nu_{w_{n}} is weakly converging to an ergodic measure ν\nu when nn goes to infinity along a subsequence 𝒩\mathcal{N}.

Lemma 7.7.

Under the above assumption, we have

n=|vn|n𝒩n++\ell_{n}=|v_{n}|\xrightarrow[n\in\mathcal{N}]{n\to+\infty}+\infty

and

νvnn𝒩n+ν.\nu_{v_{n}}\xrightarrow[n\in\mathcal{N}]{n\to+\infty}\nu.
Proof.

We argue by contradiction. Assume (n)n(\ell_{n})_{n\in\mathbb{N}} has a bounded infinite subsequence 𝒩\mathcal{N}^{\prime} of 𝒩\mathcal{N}. Then there are finite words vv and v^n\hat{v}_{n} with |v^n|<|v||\hat{v}_{n}|<|v| such that wn=vKnv^nw_{n}=v^{\otimes K_{n}}\hat{v}_{n} for n𝒩′′n\in\mathcal{N}^{\prime\prime} where 𝒩′′\mathcal{N}^{\prime\prime} is some infinite subsequence of 𝒩\mathcal{N}^{\prime}. Observe firstly that the length of wnw_{n} goes to infinity. As a consequence, KnK_{n} goes also to infinity as nn goes to infinite along 𝒩′′\mathcal{N}^{\prime\prime}. But then XX should contain the periodic point v¯\overline{v} associated to vv which is a contradiction to the aperiodicity of (Y,σ)(Y,\sigma). Therefore nn𝒩n++\ell_{n}\xrightarrow[n\in\mathcal{N}]{n\to+\infty}+\infty.

Let us check now that νvnn𝒩n+ν\nu_{v_{n}}\xrightarrow[n\in\mathcal{N}]{n\to+\infty}\nu. Let ν=limkνvnk\nu^{\prime}=\lim_{k\to\infty}\nu_{v_{n_{k}}} be a weak limit of (νvn)n𝒩(\nu_{v_{n}})_{n\in\mathcal{N}} with a subsequence (nk)k(n_{k})_{k\in\mathbb{N}} of 𝒩\mathcal{N}. For any word uu with |u|<n|u|<\ell_{n}, by (7\cdot3) we have

N(u|wn)N(u|vn)Kn,N(u|w_{n})\geq N(u|v_{n})K_{n},

and consequently

d(u|wn)\displaystyle d(u|w_{n})\geq d(u|vn)Kn|vn|(Kn+1)|vn|,\displaystyle d(u|v_{n})\frac{K_{n}|v_{n}|}{(K_{n}+1)|v_{n}|},
\displaystyle\geq 12d(u|vn).\displaystyle\frac{1}{2}d(u|v_{n}).

For any cylinder [u][u], By letting nn got infinity we get that

ν([u])12ν([u]).\nu([u])\geq\frac{1}{2}\nu^{\prime}([u]).

It implies that ν12ν\nu-\frac{1}{2}\nu^{\prime} is an σ\sigma-invariant measure. It follows from the ergodicity of ν\nu that ν=ν\nu=\nu^{\prime}. ∎

Lemma 7.8.

For any ergodic measure μν\mu\not=\nu, we have

limn𝒩,n|vn|μ([wn])=0.\lim_{n\in\mathcal{N},n\to\infty}|v_{n}|\mu([w_{n}])=0.
Proof.

Assume lim supn𝒩,n|vn|μ([wn])>0.\limsup_{n\in\mathcal{N},n\to\infty}|v_{n}|\mu([w_{n}])>0. By passing to an infinite subsequence 𝒩\mathcal{N}^{\prime} of 𝒩\mathcal{N} we have limn𝒩,n|vn|μ([wn])=b>0\lim_{n\in\mathcal{N}^{\prime},n\to\infty}|v_{n}|\mu([w_{n}])=b>0. By Lemma 7.7 the sequence νvn\nu_{v_{n}}, n𝒩n\in\mathcal{N}^{\prime}, is converging to the measure ν\nu.

Let xx be a generic point of μ\mu. Then we have for any nn

(7\cdot4) limp1p=0p1χ[wn](σ(x))=limpN(wn|x1p)p=μ([wn]).\lim_{p\to\infty}\frac{1}{p}\sum_{\ell=0}^{p-1}\chi_{[w_{n}]}(\sigma^{\ell}(x))=\lim_{p\to\infty}\frac{N(w_{n}|x_{1}^{p})}{p}=\mu([w_{n}]).

In particular for any nn we can choose PnP_{n}\in\mathbb{N} such that for pPnp\geq P_{n}

(7\cdot5) N(wn|x1p)pμ([wn])2.\frac{N(w_{n}|x_{1}^{p})}{p}\geq\frac{\mu([w_{n}])}{2}.

Pick an arbitrary cylinder [u][u]. by Lemma 7.7 there exists an integer NN such that for n>Nn>N we have

(7\cdot6) N(u|vn)12ν([u])|vn|.N(u|v_{n})\geq\frac{1}{2}\nu([u])|v_{n}|.

It follows from (7\cdot3) that

d(u|x1p)14ν([u])|vn|μ(wn).d(u|x^{p}_{1})\geq\frac{1}{4}\nu([u])|v_{n}|\mu(w_{n}).

By letting pp, then n𝒩n\in\mathcal{N}^{\prime} go to infinity, we get for any cylinder [u][u]

μ([u])b4ν([u]).\mu([u])\geq\frac{b}{4}\nu([u]).

It implies that μb4ν\mu-\frac{b}{4}\nu is an σ\sigma-invariant measure which is a contradiction to the ergodicity of μ\mu. ∎

We recall now briefly the proof of Boshernitzan that an aperiodic subshift of linear growth has finite many ergodic measures. Let 𝒩\mathcal{N} be the infinite set as in Lemma 7.3. For any n𝒩n\in\mathcal{N}, one can choose (not uniquely) an ordered kk-tuple of nn-words Kn:={qn,1,,qn,k}K_{n}:=\{q_{n,1},\dots,q_{n,k}\} which coincides with QnQ_{n}. By passing to a subsequence 𝒩\mathcal{N}^{\prime} of 𝒩\mathcal{N}, we can make each of the sequences of νqn,i\nu_{q_{n,i}} weakly convergences to some measures μi(X,T)\mu_{i}\in\mathcal{M}(X,T). Boshernitzan showed that

e(X,T){μ1,μ2,,μk}.\mathcal{M}_{e}(X,T)\subset\{\mu_{1},\mu_{2},\dots,\mu_{k}\}.

Since μi\mu_{i} may coincide with the other μj\mu_{j} for jij\not=i, we define Ii={1jk:μj=μi}I_{i}=\{1\leq j\leq k:\mu_{j}=\mu_{i}\}.

We will use the following complement of Lemma 7.5.

Lemma 7.9.

In the decomposition of a cylinder [w][w] given by Lemma 7.5, for any term σp[qn+1]\sigma^{p}[q^{\prime}_{n+1}] with |v(qn+1)|<n+1|v(q^{\prime}_{n+1})|<n+1 we have p|v(qn+1)|p\leq|v(q^{\prime}_{n+1})|.

Proof.

We argue by contradiction. To simplify the notations we write vn=v(qn+1)v_{n}=v(q^{\prime}_{n+1}). Assume |vn|<n+1|v_{n}|<n+1 and p>|vn|p>|v_{n}|. By definition of vnv_{n} we have

σp[qn+1]σp|vn|[qn+1].\emptyset\not=\sigma^{p}[q^{\prime}_{n+1}]\cap\sigma^{p-|v_{n}|}[q^{\prime}_{n+1}].

But it follows from Lemma 7.5 that σp[qn+1]\sigma^{p}[q^{\prime}_{n+1}] does not intersect σl(qnQn[qn])\sigma^{l}\left(\bigcup_{q_{n}\in Q_{n}}[q_{n}]\right) for 0<l<p0<l<p, therefore with qnQnq_{n}\in Q_{n} being the prefix of qn+1q^{\prime}_{n+1} we get the contradiction

σp[qn+1]σp|vn|[qn+1]σp[qn+1]σp|vn|[qn]=.\sigma^{p}[q^{\prime}_{n+1}]\cap\sigma^{p-|v_{n}|}[q^{\prime}_{n+1}]\subset\sigma^{p}[q^{\prime}_{n+1}]\cap\sigma^{p-|v_{n}|}[q_{n}]=\emptyset.

Thus we have p|v(qn+1)|p\leq|v(q^{\prime}_{n+1})|. ∎

For a given ii we let (qn,il)l𝒬n,i(q^{l}_{n,i})_{l\in\mathcal{Q}_{n,i}} be the elements of Qn+1Q^{\prime}_{n+1} with prefix qn,iq_{n,i}, where 𝒬n,i\mathcal{Q}_{n,i} is a subset of 𝒜\mathcal{A} for each n𝒩,1ikn\in\mathcal{N}^{\prime},1\leq i\leq k. Note that νqn,il\nu_{q^{l}_{n,i}} is also converging to μi\mu_{i} for any ll when nn goes to infinity along 𝒩\mathcal{N}^{\prime}. Finally we let vn,il=vn(qn,il)v_{n,i}^{l}=v_{n}(q^{l}_{n,i}) for each n𝒩,1ikn\in\mathcal{N}^{\prime},1\leq i\leq k and l𝒬n,il\in\mathcal{Q}_{n,i}.

Proof of Theorem 7.6.

Pick an arbitrary cylinder [w][w]. Let

[w]=j,lPn,jlσp[qn,jl],[w]=\coprod_{j,l}\coprod_{\in P_{n,j}^{l}}\sigma^{p}[q_{n,j}^{l}],

be the decomposition of [w][w] given by Lemma 7.5. Recall that Pj,lP_{j,l} is a subset of [0,(k+2)(n+1)][0,(k+2)(n+1)] for any ll and by Lemma 7.9 we have also Pj,l[0,|vn,jl|1]P_{j,l}\subset[0,|v^{l}_{n,j}|-1] if |vn,jl|<n+1|v^{l}_{n,j}|<n+1. For each n𝒩n\in\mathcal{N}^{\prime}, we decompose {(j,l):1jk,l𝒬n,j}\{(j,l):1\leq j\leq k,l\in\mathcal{Q}_{n,j}\} into three set Jn,iJ_{n,i}, Jn,iJ_{n,i}^{\prime} and Jn,i′′J_{n,i}^{\prime\prime}, where Jn,i:={(j,l):jIi}J_{n,i}:=\{(j,l):j\in I_{i}\} , Jn,i:={(j,l):jIi,|vn,jl|=n+1}J_{n,i}^{\prime}:=\{(j,l):j\notin I_{i},|v^{l}_{n,j}|=n+1\} and Jn,i′′J_{n,i}^{\prime\prime} is the rest. Then for (j,l)Jn,i(j,l)\in J_{n,i}^{\prime} we have

(7\cdot7) μi(pPn,jlσp[qn,jl])(k+2)(n+1)μi([qn,jl])=(k+2)|vn,jl|μi([qn,jl]).\mu_{i}\left(\coprod_{p\in P_{n,j}^{l}}\sigma^{p}[q_{n,j}^{l}]\right)\leq(k+2)(n+1)\mu_{i}([q_{n,j}^{l}])=(k+2)|v^{l}_{n,j}|\mu_{i}([q_{n,j}^{l}]).

On the other hand, for (j,l)Jn,i′′(j,l)\in J_{n,i}^{\prime\prime}, we have

(7\cdot8) μi(pPn,jlσp[qn,jl])|vn,jl|μi([qn,jl]).\mu_{i}\left(\coprod_{p\in P_{n,j}^{l}}\sigma^{p}[q_{n,j}^{l}]\right)\leq|v^{l}_{n,j}|\mu_{i}([q_{n,j}^{l}]).

By summing up (7\cdot7) and (7\cdot8), we have

(7\cdot9) μi((j,l)Jn,iJn,i′′pPn,jlσp[qn,jl])2k(k+2)(j,l)Jn,iJn,i′′|vn,jl|μi([qn,jl]).\mu_{i}\left(\coprod_{(j,l)\in J_{n,i}^{\prime}\cup J_{n,i}^{\prime\prime}}\coprod_{p\in P_{n,j}^{l}}\sigma^{p}[q_{n,j}^{l}]\right)\leq 2k(k+2)\sum_{(j,l)\in J_{n,i}^{\prime}\cup J_{n,i}^{\prime\prime}}|v^{l}_{n,j}|\mu_{i}([q_{n,j}^{l}]).

Combining this with Lemma 7.8, we obtain

limnμi((j,l)Jn,iJn,i′′pPn,jlσp[qn,jl])=0.\lim_{n\to\infty}\mu_{i}\left(\coprod_{(j,l)\in J_{n,i}^{\prime}\cup J_{n,i}^{\prime\prime}}\coprod_{p\in P_{n,j}^{l}}\sigma^{p}[q_{n,j}^{l}]\right)=0.

Therefore we have

χ[w](n,j)Jn,i,pPn,jlχ[qn,jl]σpL2(μi)2\displaystyle\left\|\chi_{[w]}-\sum_{(n,j)\in J_{n,i},\ p\in P_{n,j}^{l}}\chi_{[q_{n,j}^{l}]}\circ\sigma^{-p}\right\|^{2}_{L^{2}(\mu_{i})} =χ[w]χ(n,j)Jn,i,pPn,jlσp[qn,jl]L2(μi)2,\displaystyle=\left\|\chi_{[w]}-\chi_{\coprod_{(n,j)\in J_{n,i},\ p\in P_{n,j}^{l}}\sigma^{p}[q_{n,j}^{l}]}\right\|^{2}_{L^{2}(\mu_{i})},
=μi((j,l)Jn,iJn,i′′pPn,jlσp[qn,jl])n𝒩n+0.\displaystyle=\mu_{i}\left(\coprod_{(j,l)\in J_{n,i}^{\prime}\cup J_{n,i}^{\prime\prime}}\coprod_{p\in P_{n,j}^{l}}\sigma^{p}[q_{n,j}^{l}]\right)\xrightarrow[n\in\mathcal{N}^{\prime}]{n\to+\infty}0.

Thus we can apply Lemma 4.1 in L02(μi)L^{2}_{0}(\mu_{i}) with Fn={χ[qn,jl]:(j,l)Jn,i}F_{n}=\{\chi_{[q_{n,j}^{l}]}\ :\ (j,l)\in J_{n,i}\} to get

Mult(μi)lim infn𝒩,nJn,i.\text{\rm Mult}(\mu_{i})\leq\liminf_{n\in\mathcal{N}^{\prime},n\to\infty}\sharp J_{n,i}.

By summing it up, we conclude that

μe(X,σ)Mult(μ)\displaystyle\sum_{\mu\in\mathcal{M}_{e}(X,\sigma)}\text{\rm Mult}(\mu)\leq ilim infn𝒩,nJn,i,\displaystyle\sum_{i}\liminf_{n\in\mathcal{N}^{\prime},n\to\infty}\sharp J_{n,i},
\displaystyle\leq lim infn𝒩,nQn+1,\displaystyle\liminf_{n\in\mathcal{N}^{\prime},n\to\infty}\sharp Q^{\prime}_{n+1},
\displaystyle\leq 2k.\displaystyle 2k.

References

  • [AO96] A. Atzmon and A. Olevskiĭ. Completeness of integer translates in function spaces on 𝐑{\bf R}. J. Approx. Theory, 87(3):291–327, 1996.
  • [Bax71] J. R. Baxter. A class of ergodic transformations having simple spectrum. Proc. Amer. Math. Soc., 27:275–279, 1971.
  • [Bir29] G. D. Birkhoff. Demonstration d’un theoreme elementaire sur les fonctions entieres. C. R. Acad. Sci. Paris, 189:473–475, 1929.
  • [Bos84] Michael Boshernitzan. A unique ergodicity of minimal symbolic flows with linear block growth. Journal d’Analyse Mathématique, 44(1):77–96, 1984.
  • [Bos92] Michael D Boshernitzan. A condition for unique ergodicity of minimal symbolic flows. Ergodic Theory and Dynamical Systems, 12(3):425–428, 1992.
  • [BS97] Paul S. Bourdon and Joel H. Shapiro. Cyclic phenomena for composition operators. Mem. Amer. Math. Soc., 125(596):x+105, 1997.
  • [BS22] David Burguet and Ruxi Shi. Topological mean dimension of induced systems. 2022.
  • [CK19] Van Cyr and Bryna Kra. Counting generic measures for a subshift of linear growth. J. Eur. Math. Soc. (JEMS), 21(2):355–380, 2019.
  • [Coo15] Michel Coornaert. Topological dimension and dynamical systems. Springer, 2015.
  • [CP23] Darren Creutz and Ronnie Pavlov. Low complexity subshifts have discrete spectrum. arXiv preprint arXiv:2302.10336, 2023.
  • [Cre23] Darren Creutz. Measure-theoretically mixing subshifts of minimal word complexity, 2023.
  • [Dan13] Alexandre I. Danilenko. A survey on spectral multiplicities of ergodic actions. Ergodic Theory Dynam. Systems, 33(1):81–117, 2013.
  • [DDMP21] Sebastián Donoso, Fabien Durand, Alejandro Maass, and Samuel Petite. Interplay between finite topological rank minimal cantor systems, s-adic subshifts and their complexity. Transactions of the American Mathematical Society, 374(5):3453–3489, 2021.
  • [DGS06] Manfred Denker, Christian Grillenberger, and Karl Sigmund. Ergodic theory on compact spaces, volume 527. Springer, 2006.
  • [DP22] Fabien Durand and Dominique 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.
  • [Esp23] Bastián Espinoza. The structure of low complexity subshifts, 2023.
  • [Fog02] N. Pytheas Fogg. Substitutions in dynamics, arithmetics and combinatorics, volume 1794 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2002. Edited by V. Berthé, S. Ferenczi, C. Mauduit and A. Siegel.
  • [GT20] Yonatan Gutman and Masaki Tsukamoto. Embedding minimal dynamical systems into hilbert cubes. Inventiones mathematicae, 221(1):113–166, 2020.
  • [GW95] Eli Glasner and Benjamin Weiss. Quasi-factors of zero-entropy systems. Journal of the American Mathematical Society, 8(3):665–686, 1995.
  • [Jav19] Mohammad Javaheri. Cyclic composition operators on separable locally compact metric spaces. Topology Appl., 258:126–141, 2019.
  • [Kat01] Anatole Katok. Cocycles, cohomology and combinatorial constructions in ergodic theory. In Smooth ergodic theory and its applications (Seattle, WA, 1999), volume 69 of Proc. Sympos. Pure Math., pages 107–173. Amer. Math. Soc., Providence, RI, 2001. In collaboration with E. A. Robinson, Jr.
  • [KL97] J. Kwiatkowski and Y. Lacroix. Multiplicity, rank pairs. J. Anal. Math., 71:205–235, 1997.
  • [Lin99] Elon Lindenstrauss. Mean dimension, small entropy factors and an embedding theorem. Inst. Hautes Études Sci. Publ. Math., 89(1):227–262, 1999.
  • [LT14] Elon Lindenstrauss and Masaki Tsukamoto. Mean dimension and an embedding problem: an example. Israel Journal of Mathematics, 199(2):573–584, 2014.
  • [LV97] Pierre Liardet and Dalibor Volnỳ. Sums of continuous and differentiable functions in dynamical systems. Israel Journal of Mathematics, 98:29–60, 1997.
  • [Mos96] Brigitte Mossé. Recognizability of substitutions and complexity of automatic sequences. Bull. Soc. Math. Fr., 124(2):329–346, 1996.
  • [NP66] D. Newton and W. Parry. On a factor automorphism of a normal dynamical system. Ann. Math. Statist., 37:1528–1533, 1966.
  • [Par53] O. S. Parasyuk. Flows of horocycles on surfaces of constant negative curvature. Uspehi Matem. Nauk (N.S.), 8(3(55)):125–126, 1953.
  • [Que10] Martine Queffélec. Substitution dynamical systems-spectral analysis, volume 1294. Springer, 2010.