Skip to main content
Log in

Bohr chaoticity of topological dynamical systems

  • Published:
Mathematische Zeitschrift Aims and scope Submit manuscript

Abstract

We introduce the notion of Bohr chaoticity, which is a topological invariant for topological dynamical systems, and which means that the system is not orthogonal to any non-trivial weights. We prove the Bohr chaoticity for all systems which have a horseshoe and for all toral affine dynamical systems of positive entropy, some of which don’t have a horseshoe.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Béguin, F., Crovisier, S., Le Roux, F.: Construction of curious minimal uniquely ergodic homeomorphisms on manifolds: the Denjoy-Rees technique. Ann. Sci. École Norm. Sup. (4) 40(2), 251–308 (2007)

    Article  MathSciNet  Google Scholar 

  2. Béguin, F., Crovisier, S., Le Roux, F.: Realisation of measured dynamics as uniquely ergodic minimal homeomorphisms on manifolds. Math. Z. 170(1–2), 59–102 (2012)

    Article  MathSciNet  Google Scholar 

  3. Blum, J.R., Friedman, N.: On commuting transformations and roots. Proc. Am. Math. Soc. 17, 1370–1374 (1966)

    Article  MathSciNet  Google Scholar 

  4. Downarowicz, T., Serafin, J.: Almost full entropy subshifts uncorrelated to the Möbius function. Int. Math. Res. Not. IMRN 11, 3459–3472 (2019)

    Article  Google Scholar 

  5. El Abdalaoui, H.E., Kulaga-Przymus, J., Lemańczyk, M., de la Rue, T.: The Chowla and the Sarnak conjectures from ergodic theory point of view. Discrete Contin. Dyn. Syst. 37(6), 2899–2944 (2017)

    Article  MathSciNet  Google Scholar 

  6. Fan, A.: Sur les dimensions de mesures. Stud. Math. 111(1), 1–17 (1994)

    Article  MathSciNet  Google Scholar 

  7. Fan, A.: Weighted Birkhoff ergodic theorem with oscillating weights. Ergodic Theory Dynam. Syst. 39(5), 1275–1289 (2019)

    Article  MathSciNet  Google Scholar 

  8. Fan, A.: Lacunarité à la Hadamard et équirépartition. Colloq. Math. 66(1), 151–163 (1993)

    Article  MathSciNet  Google Scholar 

  9. Fan, A.: Quelques propriétés des produits de Riesz. Bull. Sci. Math. 117(4), 421–439 (1993)

    MathSciNet  MATH  Google Scholar 

  10. Fan, A.: Fully oscillating sequences and weighted multiple ergodic limit. C. R. Math. Acad. Sci. Paris 355(8), 866–870 (2017)

    Article  MathSciNet  Google Scholar 

  11. Fan, A.: Topological Wiener-Wintner ergodic theorem with polynomial weights. Chaos, Solitons Fractals 117, 105–116 (2018)

    Article  MathSciNet  Google Scholar 

  12. Fan, A.: Multifractal analysis of weighted ergodic averages. Adv. Math. 377, 107488 (2021)

    Article  MathSciNet  Google Scholar 

  13. Fan, A., Jiang, Y.: Oscillating sequences, MMA and MMLS flows and Sarnak’s conjecture. Ergod. Th. Dynam. Sys. 38, 1709–1744 (2018)

    Article  MathSciNet  Google Scholar 

  14. Fan, A., Klaus, S., Evgeny, V.: Bohr chaoticity of principle algebraic \(\mathbb{Z}^d\)-actions and Riesz products. arXiv:2103.04767

  15. Frantzikinakis, N.: Uniformity in the polynomial wiener-wintner theorem. Ergod. Th. Dynam. Sys. 26, 1061–71 (2006)

    Article  MathSciNet  Google Scholar 

  16. Hedlund, G.A.: Endomorphisms and automorphisms of the shift dynamical system. Math. Syst. Theory 3, 320–375 (1969)

    Article  MathSciNet  Google Scholar 

  17. Herman, M.-R.: Construction d’un difféomorphisme minimal d’entropie topologique non nulle. Ergodic Theory Dyn. Syst. 1(1), 65–76 (1981)

    Article  Google Scholar 

  18. Hewitt, E., Zuckerman, H.S.: Singular measures with absolutely continuous convolution squares. Proc. Cambridge Philos. Soc. 62, 399–420 (1966)

    Article  MathSciNet  Google Scholar 

  19. Huang, W., Li, J., Leiye, X., Ye, X.: The existence of semi-horseshoes for partially hyperbolic diffeomorphisms. Adv. Math. 381, 107616 (2021). https://doi.org/10.1016/j.aim.2021.107616

    Article  MathSciNet  MATH  Google Scholar 

  20. Huang, W., Ye, X.: A local variational relation and applications. Israel J. Math. 151, 237–279 (2006)

    Article  MathSciNet  Google Scholar 

  21. Karagulyan, D.: On Möbius orthogonality for subshifts of finite type with positive topological entropy. Studi. Math. 237(3), 277–282 (2017)

    Article  MathSciNet  Google Scholar 

  22. Katok, A.: Lyapunov exponents, entropy and periodic orbits for diffeomorphisms. Inst. Hautes Études Sci. Publ. Math. 51, 137–173 (1980)

    Article  MathSciNet  Google Scholar 

  23. Krieger, W.: On unique ergodicity. In Proceedings of the Sixth Berkeley Symposium on Mathematical Statistics and Probability (Univ. California, Berkeley, Calif., 1970/1971), Vol. II: Probability theory, pages 327–346, (1972)

  24. Kwietniak, D., Oprocha, P., Rams, M.: On entropy of dynamical systems with almost specification. Israel J. Math. 213(1), 475–503 (2016)

    Article  MathSciNet  Google Scholar 

  25. Lesigne, E.: Spectre quasi-discret et théorème ergodique de Wiener-Wintner pour les polymômes. Ergod. Th. Dynam. Syst. 13, 676–684 (1993)

    MATH  Google Scholar 

  26. Lian, Z., Young, L.-S.: Lyapunov exponents, periodic orbits and horseshoes for mappings of Hilbert spaces. Ann. Henri Poincaré 12(6), 1081–1108 (2011)

    Article  MathSciNet  Google Scholar 

  27. Lind, D., Schmidt, K.: Homoclinic points of algebraic \({ Z}^d\)-actions. J. Am. Math. Soc. 12(4), 953–980 (1999)

    Article  Google Scholar 

  28. Robinson, E.A.: On uniform convergence in the Wiener-Wintner theorem. J. Lond. Math. Soc. 49(3), 493–501 (1994)

    Article  MathSciNet  Google Scholar 

  29. Sarnak, P.: Three lectures on the möbius function, randomness and dynamics. IAS Lecture Notes, 1980

  30. Shi, R.: Etude de la conjecture de fuglede et les suites oscillantes. Ph. D Thesis, University of Picardie, France, 2018

  31. Smale, S.: Diffeomorphisms with many periodic points. In Differential and Combinatorial Topology (A Symposium in Honor of Marston Morse), pages 63–80. Princeton Univ. Press, Princeton, N.J., 1965

  32. Tal, M.: Some Remarks on the Notion of Bohr Chaos and Invariant Measures. arXiv:2109.04531v2

  33. Walters, P.: An introduction to ergodic theory. Graduate Texts in Mathematics, vol. 79. Springer-Verlag, New York-Berlin (1982)

  34. Young, L.S.: On the prevalence of horseshoes. Trans. Am. Math. Soc. 263(1), 75–88 (1981)

    Article  MathSciNet  Google Scholar 

  35. Weiss, B.: Countable Generators in Dynamics–Universal Minimal Models. Contemporary.Mathematics, vol. 94 (Measure and Measurable Dynamics), 1989, pages 321-326

  36. Wiener, N., Wintner, A.: Harmonic analysis and ergodic theory. Am. J. Math. 63, 415–426 (1941)

    Article  MathSciNet  Google Scholar 

  37. Zhang, G.: Personal communication

  38. Zygmund, A.: Trigonometric series. Vol. I, II. Cambridge Mathematical Library. Cambridge University Press, Cambridge, third edition, 2002. With a foreword by Robert A. Fefferman

Download references

Acknowledgements

The authors would like to thank Klaus Schmidt, Evgeny Verbitskiy, Meng Wu and Weisheng Wu for helpful discussions and valuable comments. Thanks also go to Xiangdong Ye for providing the preprint [19] before it is diffused, to B. Weiss for valuable informations. Special thanks go to Evgeny Verbitskiy for proposing to use an argument from Blum and Friedman [3] at the end of the proof of Proposition 6.1. We are grateful to the Shanghai Center for Mathematical Sciences for their hospitality, where part of this paper was written.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shilei Fan.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

A. H. FAN was supported by NSF of China (Grant No. 11971192); S. L. FAN was supported by NSF of China (Grant No. 11971190) and Fok Ying-Tong Education Foundation, China (Grant No.171001); W. X. SHEN was supported by NSF of China (Grant No 11731003)

Appendices

7 Appendix A: Extended horseshoes have no roots

The extended horseshoe has no root. A self-contained proof of this fact will be given below for the case of one-sided horseshoe. A different proof is given for the case of two-sided horseshoe and it replies on the no existence of root for the two-sided full shift, which is known ( [16]). It seems that the first proof can not be adapted to the case of two-sided horseshoe.

1.1 7.1 No root of \(T: \Lambda ^*\rightarrow \Lambda ^*\) (one-sided case)

Here we give a proof of Proposition 5.2.

The proof is based on the investigation of the maximal entropy measure of the dynamical system \((\Lambda ^*,T)\). Given a topological dynamical system (XT) and an integer \(s\ge 2\). A T-invariant measure is \(T^s\)-invariant. If such a measure is \(T^s\)-ergodic, it is T-ergodic. But in general, a T-ergodic measure is not necessarily \(T^s\)-ergodic. That is the case for the invariant measure \(\frac{1}{2}(\delta _{\frac{1}{3}} + \delta _{\frac{2}{3}})\) of the doubling dynamics \(x \rightarrow 2 x \mod 1\). The proof of Proposition 5.2 is based on the following lemma, which shows that the existence of a T-invariant and \(T^s\)-ergodic measure is an obstruction for \(T^{ns}\) to be both 2-to-1 and surjective.

Lemma 7.1

Let (XT) be a topological dynamical system and \(s\ge 2\) an integer. Suppose that there exists a T-invariant measure which is \(T^s\)-ergodic. Then \(T^{ns}\) can not be both exactly 2-to-1 and surjective for any integer \(n\ge 1\).

Proof

Suppose that \(T^{ns}:X\rightarrow X\) is 2-to-1 and surjective for some integer \(n\ge 1\). Then the map \(T:X\rightarrow X\) must be surjective and each point \(x\in X\) has at most two pre-images by T, i.e. \(1\le \#T^{-1}(x) \le 2\). This allows us to decompose X into two disjoint sets

$$\begin{aligned}A_0=\{x\in X: \#T^{-1}(x)=1\}, \qquad B_0=\{x\in X: \#T^{-1}(x)=2\}.\end{aligned}$$

We claim that these two sets are measurable. For any \(\epsilon >0\), let

$$\begin{aligned}B_0^{\epsilon }=\{x\in X: \#T^{-1}(x)=2 \text { and } |y_1-y_2| \ge \epsilon \text { for distinct } y_1, y_2 \in T^{-1}(x)\}. \end{aligned}$$

Then \(B_0^\epsilon \) is closed. Since \(B_0=\cup _{n\ge 1} B_0^{1/n}\), it is an \(F_\sigma \) set. Hence, \(B_0\) is measurable which implies \(A_0\) is also measurable.

For \(0\le k\le ns\), let \(B_{k}=T^{-k}(B_0)\). First notice that we have

$$\begin{aligned} T^{-1}(B_k) = B_{k+1}, \quad T(B_{k+1}) = B_k \qquad (0\le k\le ns -1) \end{aligned}$$

where the second equality is because of the surjectivity of T. We claim that the maps \(T:B_{k+1}\rightarrow B_k\) are injective and then bijective for \(1\le k\le ns-1\).

Indeed, otherwise, for some k and some \(v\in B_k\) there are two distinct points \(u', u''\in B_{k+1}\) such that \(T(u') = T(u'')=v\). Let \(w=T^{k-1}(v)\), which belongs to \(B_1=T^{-1}(B_0)\). Let \(z=T(w)\), which belongs to \(B_0\). By the definition of \(B_0\), there exists a point \(w^*\in B_1\) different from w such that \(T(w^*)=z\) and then there exists a point \(u^*\in B_{k+1}\) different from \(u', u''\) such that \(T^k(u^*) = w^*\). Therefore, z has at least three \(T^{k+1}\)-preimages \(u',u'', u^*\) and then at least three \(T^{ns}\)-preimages. This contradicts the fact that \(T^{ns}\) is 2-to-1.

The above claim implies that \(B_k\subset A_0\) for \(1\le k \le ns-1\). Since \(A_0\) and \(B_0\) form a partition of X, we have \(B_{k}\cap B_0=\emptyset \) for \(1\le k \le ns-1\). Consequently, all \(B_j\)’s for \(0\le j \le ns-1\) are disjoint. We claim that all \(B_j\)’s for \(0\le j \le ns-1\) form a partition of X. To that end, it suffices to prove

$$\begin{aligned} A_0=\bigsqcup _{k=1}^{ns-1}B_{k}. \end{aligned}$$
(7.1)

Suppose that (7.1) is not true, which means there exists a point \(x\in A_0\setminus \bigsqcup _{k=1}^{ns-1}B_{k}\). Hence \(T^{k}(x)\in A_0\) for \(0\le k\le ns-1 \), by the definition of \(B_k\) and the fact that \(\{A_0, B_0\}\) is a partition of X. Therefore, the point \(T^{ns-1}(x)\in A_0\) has a unique \(T^{ns}\)-preimage \(T^{-1}(x)\), which contradicts to the assumption that \(T^{ns}\) is 2-to-1.

Since \(X= A_0\sqcup B_0\), we get the decomposition

$$\begin{aligned}X=\bigsqcup _{k=0}^{ns-1}B_{k} = \bigsqcup _{j=0}^{s-1} T^{-j} (X') \ \ \ \mathrm{with }\ \ X^{\prime }=\bigsqcup _{k=0}^{n-1}B_{ks}. \end{aligned}$$

By the hypothesis, there exists a T-invariant measure \(\mu \) which is \(T^s\)-ergodic. The T-invariance of \(\mu \) implies that \(\mu (X^{\prime })=1/s\).

Let \(A_1=T^{-1}(A_0)\). Recall that \(B_1 =T^{-1}(B_0)\). Since \(\{A_0, B_0\}\) is a partition of X, so is \(\{A_1, B_1\}\). As \( B_1\subset A_0\) which is proved above, we have \(B_0\subset A_1\) and actually \(B_0=A_1\setminus A_0\). By (7.1) and the definition of \(A_1\), we get \(A_1= \bigsqcup _{k=2}^{ns}B_{k}\). Then from \(A_0\sqcup B_0 = A_1\sqcup B_1\) we get \(B_0=B_{ns}\), in other words,

$$\begin{aligned} T^{-ns}(B_0)=B_0. \end{aligned}$$

Consequently, \(T^{-s}(X^{\prime })=X^{\prime }\). Then, by the \(T^s\)-ergodicity of \(\mu \), we have \(\mu (X^{\prime })=0 \text { or } 1\), which contradicts to \(\mu (X') =\frac{1}{s}\) with \(s\ge 2\). \(\square \)

Now we shall prove Proposition 5.2 by contradiction. Basic properties of entropy function will be used. We refer to Walters’ book [33] (cf. Theorem 4.13, Theorem 7.5, Theorem 7.10, Theorem 8.1).

Proof

(Proof of Proposition 5.2)

Assume that \(S:\Lambda ^*\rightarrow \Lambda ^*\) is a continuous map such that \(S^n=T\) for some integer \(n\ge 2\).

For \(1\le j\le N-1\), we consider the dynamical system \((T^{j}(\Lambda ), T^{N})\). By the hypothesis, the map \(T^j:\Lambda \rightarrow T^{j}(\Lambda )\) is a homeomorphism. So, the system \((T^{j}(\Lambda ), T^{N})\) is conjugate to \((\Lambda , T^{N})\) with the conjugation \(T^{j}\). Hence, the topological entropy of the system \((T^{j}(\Lambda ), T^{N})\) is equal to \( \log 2\). It follows that \( h_\mathrm{top}(\Lambda ^*,T^{N})=\log 2\). Therefore,

$$\begin{aligned} h_\mathrm{top}(\Lambda ^*,T)=\frac{1}{N}\log 2. \end{aligned}$$

It follows that the root dynamical system \((\Lambda ^*,S)\) admits its topological entropy

$$\begin{aligned}h_\mathrm{top}(\Lambda ^*,S)=\frac{1}{nN}\log 2.\end{aligned}$$

Let \(\mu \) be a maximal entropy measure of the system \((\Lambda ^*,S)\). It is also a maximal entropy measure of dynamical systems \((\Lambda ^*, T)\) and \((\Lambda ^*,T^{N})\).

Let \(\mu _j = \mu |_{T^{j}(\Lambda )}\) be the restriction for \(0\le j\le N-1\). By the T-invariance of \(\mu \) and the disjointness of \(T^{j}(\Lambda )\)’s (\(0\le j \le N-1\)), it is easy to get

$$\begin{aligned} \mu _{j+1} = \mu _j \circ T^{-1} \quad (0\le j \le N-1) \end{aligned}$$

with the convention \(\mu _N =\mu _0\). It follows that \(\mu _j\) are all \(T^{N}\)-invariant. Let \(\nu _j= N\cdot \mu _j\), which is a probability measure concentrated on \(T^{j}(\Lambda )\). Since the systems \((T^{j}(\Lambda ), \nu _j, T^{N})\) are all conjugate, the measure-theoretic entropies \(h_{\nu _j}(T^{j}(\Lambda ), T^{N})\) are equal and the common value is \(h_{\nu _j}(\Lambda ^*, T^{N})\). From this, the relation \(\mu = \frac{1}{N} \sum _{j=0}^{N-1}\nu _j\) and the affinity of entropy, we get

$$\begin{aligned} \log 2=h_\mu (\Lambda ^*, T^{N})= \frac{1}{N}\sum _{j=0}^{N-1} h_{\nu _j}(\Lambda ^*, T^{N}) = h_{\nu _0}(\Lambda ^*, T^{N}) = h_{\nu _0}(\Lambda , T^{N}). \end{aligned}$$

So, the measure \(\nu _0\) is the unique maximal entropy measure of the horseshoe \((\Lambda , T^{N})\), i.e. the Bernoulli \((\frac{1}{2},\frac{1}{2})\)-measure on \(\{0,1\}^{\mathbb {N}}\). Since \(\nu _0\) is \(T^{N}\)-ergodic, so are \(\nu _j\)’s. Now we claim that \(\mu \) is T-ergodic. In fact, assume that \(A\subset \Lambda ^*\) is a T-invariant set. Then \(A\cap T^{j}(\Lambda )\) is \(T^{N}\)-invariant for every \(0\le j\le N-1\). By the \(T^{N}\)-ergodicity of \(\nu _j\), \(\nu _j(A\cap T^{j}(\Lambda )) =0 \ \mathrm{or} \ 1\) for each j. But \(\nu _j(A\cap T^{j}(\Lambda ))\) are equal for different j’s. Then we get \(\mu (A) =0 \ \mathrm{or}\ 1\), because

$$\begin{aligned} \mu (A) = \frac{1}{N}\sum _{j=0}^{N-1} \nu _j(A\cap T^{j}(\Lambda )). \end{aligned}$$

The existence of S-invariant measure \(\mu \) which is T-ergodic measure, contradicts the fact that \(T^N: \Lambda ^*\rightarrow \Lambda ^*\) is 2-to-1, by Lemma 7.1. \(\square \)

1.2 7.2 No root of \(T: \Lambda ^*\rightarrow \Lambda ^*\) (two-sided case)

Here we prove Proposition 6.2. The proof is based on the fact that the shift map \(\sigma : \{0,1\}^\mathbb {Z}\rightarrow \{0,1\}^\mathbb {Z}\) has no root, which is well known (cf. [16], Corollary 18.2, p.371).

Given a two-sided horseshoe \((\Lambda ,T^{N})\) with disjoint steps, it is convenient to identify the subsystem \((\Lambda ^*,T)\) with the following system \((\{0,1\}^{\mathbb {Z}}\times \mathbb {Z}/N\mathbb {Z}, \sigma _N)\), a tower of height N, where \(\sigma _N\) is defined by

$$\begin{aligned} \sigma _N(\omega ,k)=\left\{ \begin{array}{ll} (\sigma (\omega ),0), &{} \text{ if } k=N-1; \\ (\omega ,k+1), &{} \text{ otherwise. } \end{array} \right. \end{aligned}$$
(7.2)

The tower \(\{0,1\}^{\mathbb {Z}}\times \mathbb {Z}/N\mathbb {Z}\) has N floors \(F_i = \{0,1\}^{\mathbb {Z}}\times \{i\}\) for \(0\le i <N\). We extend \(F_i\) for all integers \(i\ge 0\) by defining \(F_i = F_{i \mod N}\). Especially \(F_N = F_0\). In the following we denote \(\sigma _N\) by T.

Suppose that T has a root S, i.e. \(S^p=T\) for some \(p\ge 2\). It is clear that S is bijective and commutes with \(\sigma _N\). We claim that S permute floors, that is to say, for any i there exists a j such that \(S (F_i) =F_j\). We shall prove this claim by the commutativity of S with T. Assume this claim for the moment. Then each floor \(F_i\) (\(0\le i <N\)) is mapped back to \(F_i\) by \(S^N\), that is to say,

$$\begin{aligned} \forall \omega , \quad S^N(\omega , i) = (R_i\omega , i), \end{aligned}$$
(7.3)

where \(R_i: \{0, 1\}^\mathbb {Z}\rightarrow \{0, 1\}^\mathbb {Z}\) is some map, which depends on i. It is easy to see that \(R_i\) is continuous. Thus, on one hand, we have

$$\begin{aligned} \forall \omega , \quad S^{Np}(\omega , i) =(R_i^p\omega , i); \end{aligned}$$

on the other hand, we have

$$\begin{aligned} \forall \omega , \quad S^{Np}(\omega , i) = T^N( \omega , i)= ( \sigma \omega , i). \end{aligned}$$

It follows that \(\sigma =R_i^p\), which is impossible.

Now let us prove the claim. Let \(\mu \) the ergodic probability measure of the maximal entropy of T (its restriction on each floor is the symmetric Bernoulli measure). We have \(\mu (F_i) =\frac{1}{N}\). Let \(C_i = S(F_0)\cap F_i\), the portion of \(SF_0\) contained in \(F_i\), for \(0\le i <N\). Suppose \(0<\mu (C_i)<\mu (F_i)\) for some i. There would be a contradiction. Indeed, we have the invariance

$$\begin{aligned} T \left( \bigcup _{j=0}^{N-1} T^j (C_i)\right) = \bigcup _{j=0}^{N-1} T^j (C_i). \end{aligned}$$

This is because

$$\begin{aligned} T^j(C_i)= T^j(S(F_0)\cap F_i)= ST^j (F_0) \cap T^jF_{i} = S(F_j) \cap F_{i+j} \end{aligned}$$

and

$$\begin{aligned} \bigcup _{j=1}^{N} S(F_j) \cap F_{i+j} = \bigcup _{j=0}^{N-1} S(F_j) \cap F_{i+j}. \end{aligned}$$

The invariant set \(\bigcup _{j=0}^{N-1} T^j (C_i)\) has its measure between 0 and 1 because of \(0<\mu (C_i)<\frac{1}{N}\). This contradicts the ergodicity of \(\mu \). We have thus proved that for every i, the measure \(\mu (S (F_0) \cap F_i)\) is equal to 0 or \(\mu (F_i)\). There exists one i such that \(\mu (S (F_0) \cap F_i)= \mu (F_i)\), otherwise \(\mu (S(F_0))=0\) so that

$$\begin{aligned} \mu (S(F_j))= \mu (ST^j (F_0))= \mu (T^j S (F_0)) = \mu (S(F_0))=0, \end{aligned}$$

which implies that \(\mu \) is the null measure. There is at most one i such that \(\mu (S (F_0) \cap F_i)= \mu (F_i)\), otherwise \(\mu (S(F_0))\ge \frac{2}{N}\), which implies that \(\mu \) has a total measure equal to at least 2. So, \(S(F_0)\) and \(F_i\) are equal almost everywhere. If we take into account the continuity of S, we get that the two compact sets \(S(F_0)\) and \(F_i\) must be equal. In place of \(F_0\), we can consider any \(F_j\). The same argument shows that \(S(F_j)\) must be equal to some \(F_k\). In this way, S defines a permutation on floors. Otherwise, under S we have a cycle

$$\begin{aligned} F_{i_0} \rightarrow F_{i_1}\rightarrow \cdots \rightarrow F_{i_{\ell -1}}\rightarrow F_{i_0} \end{aligned}$$

with \(\ell <N\). Then the union U of these \(\ell \) floors is S-invariant and it is also \(S^p\)-invariant, i.e. T-invariant, an obvious contradiction.

8 Appendix B: Any cylinder contains a horseshoe

1.1 8.1 Horseshoe with disjoint steps in any cylinder: one-sided case

Consider the one-sided full shift system \((\{0,1\}^{\mathbb {N}}, \sigma )\). For any non-empty open set \(U\subset \{0,1\}^{\mathbb {N}}\), we shall show that there exists a horseshoe \((\Lambda ,\sigma ^N)\) with disjoint steps for some sufficiently large integer \(N\ge 1\) such that \( \Lambda \subset U\) and the map \(\sigma ^{N-1}: \Lambda \rightarrow \sigma ^{N-1}(\Lambda )\) is bijective. In other word, this horseshoe \((\Lambda ,\sigma ^N)\) satisfies the conditions (i) and (ii) required by Proposition 5.2.

For a word \(a_0\cdots a_{n-1}\) of length n, denote by \([a_0\cdots a_{n-1}]\) the cylinder of rank n

$$\begin{aligned}=\{y \in \{0,1\}^{\mathbb {N}}: y_i=a_i \text { for } 0\le i \le n-1\}.\end{aligned}$$

Denote the n-prefix of \(x=(x_j)_{j\ge 0}\) by \(x|_n\), i.e. \(x|_n=x_0x_1\cdots x_{n-1}\). Hence \([x|_{n}]\) denote a cylinder of rank n. Let uv denote the concatenation \(u_0u_1\cdots u_{n-1}v_0\cdots v_{m-1}\) of two words \(u=u_0u_1\cdots u_{n-1}\) and \(v= v_0\cdots v_{m-1}\). So, \(u^{r}\) denotes \(u\cdots u\) (r times). In particular, \(1^r\) means \(1\cdots 1\) (r times). The following lemma is a preparation for proving the above announced existence of horseshoe in a given cylinder.

Lemma 8.1

For any cylinder C in \(\{0,1\}^\mathbb {N}\) of rank \(M\ge 1\), there exists a sub-cylinder \(C'\subset C\) of rank N with \(N\ge M\) such that

  1. (i)

    \(\sigma ^N (C') = \{0,1\}^\mathbb {N}; \)

  2. (ii)

    \(C'\cap \sigma ^{n}(C')=\emptyset \ \text {for } 1\le n \le N-1.\)

Proof

Assume \(C=[a_0a_1\cdots a_{M-1}]\). We assume \(a_0=0\), without loss of generality. Note that \(\sigma ^{M}(C)=\{0,1\}^{\mathbb {N}}\). Let \(n_0\ge 1 \) be the minimal positive integer such that \(C \subset \sigma ^{n_0}(C)\), so that

$$\begin{aligned} C\cap \sigma ^{n}(C)=\emptyset \quad \ \text {for\ all }\ 1\le n\le n_0-1. \end{aligned}$$
(8.1)

We restate these facts as follow

$$\begin{aligned}{}[0a_1\cdots a_{M-1}]\cap [a_1\cdots a_{M-1}]=\emptyset ,&\cdots , [0a_1\cdots a_{M-1}]\cap [a_{n_0-1}a_{n_0}\cdots a_{M-1}]=\emptyset ; \end{aligned}$$
(8.2)
$$\begin{aligned}{}[0a_1\cdots a_{M-1}]\subset [a_{n_0}\cdots a_{M-1}]. \end{aligned}$$
(8.3)

We have \(n_0\le M\). If \(n_0=M\), we are done and we can take \(C'=C\). In the following, we assume \(n_0<M\).

The inclusion (8.3) means

$$\begin{aligned} a_{n_0+j}=a_j \quad \text {for } 0\le j \le M-n_0-1. \end{aligned}$$
(8.4)

Let \(x=\overline{a_0a_1\cdots a_{n_0-1}}\), which is \(n_0\)-periodic. By (8.4), \(a_0a_1\cdots a_{M-1}\) is a prefix of x, so x is in C. By (8.2), \(n_0\) is the minimal period of x. Define the sub-cylinder \(C^{\prime }=[0a_1a_2\cdots a_{M-1}1^{n_0}]\), or more precisely

$$\begin{aligned} C^{\prime } =[(0a_1\cdots a_{n_0-1})^{q}0a_1\cdots a_{j-1}1^{n_0}] \end{aligned}$$

where \(q\ge 0\) and \(0\le j \le n_0-1\) are determined by \(M=qn_0+j\). Now we shall check that the sub-cylinder \(C^{\prime }\)(of C) of rank \(N=M+n_0\) has property (ii). We distinguish three cases.

Case I. \(1\le n\le n_{0}-1\). Since \(C^{\prime }\subset C\), we proved \(C^{\prime }\cap \sigma ^{n}(C^{\prime })=\emptyset \), by (8.1).

Case II. \(n_0\le n \le M-1\). Suppose \(C^{\prime }\cap \sigma ^{n}(C^{\prime })\ne \emptyset \). Then \(C\cap \sigma ^{n}(C^{\prime })\ne \emptyset \). Since \(|\sigma ^{n}(C^{\prime })|\le M=|C|\), we get \(C\subset \sigma ^{n}(C^{\prime })\). Hence, the word \(a_n\cdots a_{M-1}1^{n_0}\) defining the cylinder \(\sigma ^{n}(C^{\prime })\) has \(1^{n_0}\) as suffix, which is a word contained in \(a_0a_1\cdots a_{M-1}\) defining the cylinder C. But, on the other hand, any word of length \(n_0\) contained in \(a_0a_1\cdots a_{M-1}\) contains 0, a contradiction.

Case III. \(M\le n \le N-1\). This case is evident because \(\sigma ^{n}(C^{\prime })\subset [1]\), but \(C'\subset [0]\). \(\square \)

Let us look at the cylinders C of rank \(M=4\) and the cylinders \(C'\) constructed in Lemma 8.1:

$$\begin{aligned} C=[0000],&C' = [00001^1]\\ C=[0001],&C' = C\ \ \ \ \ \ \ \ \\ C=[0010],&C' = [00101^3]\\ C=[0011],&C' = C\ \ \ \ \ \ \ \ \\ C=[0100],&C' = [01001^3]\\ C=[0101],&C' = [01011^2]\\ C=[0110],&C' = [01101^3]\\ C=[0111],&C' = C\ \ \ \ \ \ \ \ \end{aligned}$$

where the exponent represents \(n_0\).

We are now ready to prove the existence of horseshoe contained in a given cylinder.

Proposition 8.2

Let \(C\subset \{0,1\}^{\mathbb {N}}\) be an arbitrary cylinder. There exists a \(\sigma ^{N}\)-invariant closed subset \(\Lambda \subset C\) for some integer \(N\ge |C|\), such that

  1. (i)

    The system \((\Lambda , \sigma ^N)\) is topologically conjugate to the full shift \((\{0,1\}^{\mathbb {N}}, \sigma )\);

  2. (ii)

    The maps \(\sigma ^{n}: \Lambda \rightarrow \sigma ^{n}(\Lambda ) \ (1\le n\le N-1)\) are bijections;

  3. (iii)

    The sets \(\Lambda , \sigma (\Lambda ), \ldots , \sigma ^{N-1}(\Lambda )\) are disjoint.

Proof

Assume \(C=[y_0y_1\cdots y_{m-1}]\) be a cylinder of rank m. We can assume \(y_0=0\) without loss of generality. By Lemma 8.1, there exists an integer \(n_* \ge m\) and a sub-cylinder \(C^{\prime }\subset C\) of rank \(n_*\) such that

$$\begin{aligned} C^{\prime }\cap \sigma ^{n}(C^{\prime })=\emptyset \ \ \text { for } 1\le n \le n_*-1. \end{aligned}$$
(8.5)

It is obvious that \(\sigma ^{n_*}(C^{\prime })=\{0,1\}^\mathbb {N}\) and \(\sigma ^{n_*}: C' \rightarrow \{0,1\}^\mathbb {N}\) is bijective. Assume that \(C' =[y_0y_1\cdots y_{m-1} y_m \cdots y_{n_*-1}]\). Notice that \(y_{n_*-1}=1\) because \(C'\cap \sigma ^{n_*-1}(C')=\emptyset \). Let \(y= \overline{y_0y_1\cdots y_{n_*-1}}\). This periodic point y is the unique periodic point contained in \(C^{\prime }\) of exact period \(n_*\), by (8.5). Let \(n_1=\min \{0\le i\le n_*-1: y_i=1\}\). Since \(y_0=0\) and \(y_{n_*-1}=1\), we have \(1\le n_1\le n_*-1\).

Define two sub-cylinders of \(C'\) of rank \(n_*+n_1+2\):

$$\begin{aligned} C_1= & {} [y_0y_1\cdots y_{n_*-1}10 0^{n_1}] =[0^{n_1}y_{n_1}\cdots y_{n_*-1}10 0^{n_1}]=[0^{n_1}a0^{n_1}], \\ C_2= & {} [y_0y_1\cdots y_{n_*-1}11 0^{n_1}] = [0^{n_1}y_{n_1}\cdots y_{n_*-1}11 0^{n_1}] =[0^{n_1}b0^{n_1}] \end{aligned}$$

where \(a = y_{n_1}\cdots y_{n_*-1}10\) and \(b = y_{n_1}\cdots y_{n_*-1}11\).

Also observe that

(iv) for \(1\le n\le n_*-1\), \(\sigma ^{n}(C_1\cup C_2)\cap C^{\prime }=\emptyset \) and \(\sigma ^{n}\) is injective on \(C_1\cup C_2\).

This follows from the relation \(C_{1}\cup C_2 \subset C^{\prime }\), the disjointness (8.5) and the injectivity of \(\sigma ^{n_*}: C^{\prime } \rightarrow \{0,1\}^{\mathbb {N}}\). By the definition of \(n_1\), we have \(C_{1}\cup C_{2}\subset [0^{n_1}1]\). Note that

$$\begin{aligned}\sigma ^{n_*}(C_1)=[100^{n_1}], \quad \sigma ^{n_*+1}(C_1)=[0^{n_1+1}]; \quad \sigma ^{n_*}(C_2)=[110^{n_1}], \quad \sigma ^{n_*+1}(C_2)=[10^{n_1}]\end{aligned}$$

which are all disjoint from \([0^{n_1}1]\). Hence,

(v) for \( n=n_*\) or \(n_*+1\), \(\sigma ^{n}(C_1\cup C_2)\cap (C_{1}\cup C_{2})=\emptyset \) and \(\sigma ^n\) is injective on \(C_1\cup C_2\).

Now take \(N=n_*+2\) and define

$$\begin{aligned} \Lambda =\{x\in \{0,1\}^{\mathbb {N}}: \forall k\ge 0, \sigma ^{kN}(x) \in C_{1}\cup C_{2}\}. \end{aligned}$$

Observe that both the words defining \(C_1\) and \(C_2\) have \(0^{n_1}\) as their prefix and as well as their suffix. Therefore \(\Lambda \) can be identified with the symbolic space \(\{0^{n_1}a, 0^{n_1} b\}^\mathbb {N}\) and \((\Lambda , \sigma ^N)\) is conjugate to the shift map on \(\{0^{n_1}a, 0^{n_1} b\}^\mathbb {N}\). This is the property (i) of \((\Lambda , \sigma ^N)\). The required properties (ii) and (iii) follow from (iv) and (v).

\(\square \)

1.2 8.2 Horseshoe with disjoint steps in any cylinder: two-sided case

Consider the full shift system \((\{0,1\}^{\mathbb {Z}}, \sigma )\). For any non-empty cylinder \(C\subset \{0,1\}^{\mathbb {Z}}\), we shall show that there exists a horseshoe \((\Lambda ,\sigma ^N)\) with disjoint steps for some sufficiently large integer \(N\ge 1\), where \(\Lambda \subset C\). We start with the following preparative lemma, the counterpart of Lemma 8.1.

Lemma 8.3

For any cylinder C in \(\{0,1\}^\mathbb {Z}\) of rank \(M\ge 1\), there exists a sub-cylinder \(C'\subset C\) of rank N with \(N\ge M\) such that

  1. (i)

    \(\sigma ^N (C') \cap C'\ne \emptyset ; \)

  2. (ii)

    \(C'\cap \sigma ^{n}(C')=\emptyset \ \text {for } 1\le n \le N-1.\)

Proof

The proof is almost the same as that of Lemma 8.1. We assume \(C=[a_0a_1\cdots a_{M-1}]\), and \(a_0=0\) without loss of generality. Note that \(C\cap \sigma ^{M}(C)\ne \emptyset \). Let \(n_0\ge 1 \) be the minimal positive integer such that \(C\cap \sigma ^{n_0}(C)\ne \emptyset \), so that

$$\begin{aligned} C\cap \sigma ^{n}(C)=\emptyset \quad \ \text {for\ all }\ 1\le n\le n_0-1. \end{aligned}$$
(8.6)

We restate these facts as follow

$$\begin{aligned}{}[0a_1\cdots a_{M-1}]_0\cap [a_1\cdots a_{M-1}]_0=\emptyset ,&\cdots , [0a_1\cdots a_{M-1}]_0\cap [a_{n_0-1}a_{n_0}\cdots a_{M-1}]_0=\emptyset ; \end{aligned}$$
(8.7)
$$\begin{aligned}{}[0a_1\cdots a_{M-1}]_0\subset [a_{n_0}\cdots a_{M-1}]_0. \end{aligned}$$
(8.8)

The rest of the proof is the same if we replace cylinders of the for \([*]\) by \([*]_0\) where \(*\) represents a word.

\(\square \)

We are now ready to prove the existence of horseshoe contained in a given cylinder.

Proposition 8.4

Let \(C\subset \{0,1\}^{\mathbb {N}}\) be an arbitrary cylinder. There exists a \(\sigma ^{N}\)-invariant closed subset \(\Lambda \subset C\) for some integer \(N\ge |C|\), such that

  1. (i)

    The system \((\Lambda , \sigma ^N)\) is topologically conjugate to the full shift \((\{0,1\}^{\mathbb {N}}, \sigma )\);

  2. (ii)

    The sets \(\Lambda , \sigma (\Lambda ), \ldots , \sigma ^{N-1}(\Lambda )\) are disjoint.

Proof

It is exactly the same proof as that of Proposition 8.2, if we replace cylinders of the form \([*]\) by \([*]_0\).

\(\square \)

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fan, A., Fan, S., Ryzhikov, V.V. et al. Bohr chaoticity of topological dynamical systems. Math. Z. 302, 1127–1154 (2022). https://doi.org/10.1007/s00209-022-03093-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00209-022-03093-6

Navigation