1,167
Views
4
CrossRef citations to date
0
Altmetric
Articles

Topological properties of Lorenz maps derived from unimodal maps

, &
Pages 1174-1191 | Received 17 Oct 2019, Accepted 17 Apr 2020, Published online: 29 May 2020

Abstract

A symmetric Lorenz map is obtained by ‘flipping’ one of the two branches of a symmetric unimodal map. We use this to derive a Sharkovsky-like theorem for symmetric Lorenz maps, and also to find cases where the unimodal map restricted to the critical omega-limit set is conjugate to a Sturmian shift. This has connections with properties of unimodal inverse limit spaces embedded as attractors of some planar homeomorphisms.

MATHEMATICS SUBJECT CLASSIFICATIONS 2010:

This article is part of the following collections:
Journal of Difference Equations and Applications Best Paper Award

1. Introduction

Topological properties of a continuous dynamical system are, in general, easier to understand than those of discontinuous systems. For example, for continuous functions of the real line there is the celebrated Sharkovsky Theorem [Citation29], which says that if the map has a periodic point of prime period n, it also has a periodic point of prime period m for every mn in the Sharkovsky order 1248474543272523753. However, in general there is no analogue of the Sharkovsky theorem for discontinuous functions of the reals.

In this paper we study Lorenz maps, which are piecewise monotone interval maps with a single discontinuity point. Such Lorenz maps appear as Poincaré maps of geometric models of Lorenz attractors described independently by Guckenheimer [Citation20], Williams [Citation30] and Afraimovich, Bykov and Shil'nikov [Citation1]. For the class of ‘old’ maps (discontinuous degree one interval maps) which also include Lorenz maps, a characterization of periodic orbit forcing was given by Alsedá, Llibre, Misurewicz and Tresser in [Citation3]. Hofbauer in [Citation22] obtained a result similar as in [Citation3] using an oriented graph with infinitely many vertices whose closed paths represent the periodic orbits of the map except that he did not characterize completely the set of periodic points. The increasing Lorenz maps ϕ were studied earlier by Rand in [Citation28], where he noticed, based on observations from [Citation23], that periods follow Sharkovsky's order. In particular, [Citation28, Theorem 4] is an analogue of our Theorem 3.1. In [Citation2] the connection between β-expansions and Sharkovsky's orderi is given. Recently, Cosper [Citation18] proved a direct analogue of Sharkovsky's Theorem in special families of piecewise monotone maps (a truncated tent map family).

In the present paper we combine some old and more recent results on the relation between unimodal and Lorenz maps, including a version of Sharkovsky's Theorem. The basic idea is to explore the relation between a unimodal map f and symmetric Lorenz maps ϕ and ψ obtained by ‘flipping the right branch’ and ‘flipping the left branch’ of the graph of f respectively, see Figure .

Figure 1. An increasing and decreasing symmetric Lorenz map ϕ and ψ obtained from a unimodal map f.

Figure 1. An increasing and decreasing symmetric Lorenz map ϕ and ψ obtained from a unimodal map f.

For increasing Lorenz maps ϕ we prove in Theorem 3.1 that Sharkovsky's Theorem holds with the exception of the fixed points, and for decreasing Lorenz maps ψ we prove in Theorem 3.2 that Sharkovsky's Theorem holds possibly except for periods 2r,r1.

We can turn ϕ into a proper circle endomorphism (with unique rotation number independent of xS1) by setting (see also Figure ): φ¯(x)={φ(1)=f(1)~,x[0,a];wherea<cissuchthatφ(a)=φ(1),φ(x),otherwise. In Proposition 5.1 we calculate the rotation number of the family of such maps, and prove that in the irrational rotation number case the restriction to omega limit set is a minimal homeomorphism. We use techniques developed primarily for unimodal interval maps.

Figure 2. The points ζSQ(k)<cSk1<ζSQ(k)1 and their images under fSQ(k).

Figure 2. The points ζSQ(k)<cSk−1<ζSQ(k)−1 and their images under fSQ(k).

Figure 3. A stunted symmetric Lorenz map φ¯ as a circle endomorphism.

Figure 3. A stunted symmetric Lorenz map φ¯ as a circle endomorphism.

Next, we also give an implementation of Sturmian shifts in interval maps. For every Sturmian shift we assign a unimodal map (basically a kneading sequence) so that the unimodal map restricted to its omega limit set is conjugate to that Sturmian shift.

Maps φ¯, besides being interesting on their own, prove also to be very useful in surface dynamics. Namely, knowledge of their dynamics can be related to special orientation preserving planar embeddings of inverse limit spaces with bonding maps being f. In the last section of the paper we connect the map φ¯ to the study of unimodal inverse limit spaces represented as attractors of some planar homeomorphisms (this was initially done in [Citation9] using a map conjugated to φ¯). In Theorem 6.1 we give a compete characterization of accessible points of tent inverse limit spaces embedded in such a way. Then Corollary 6.1 gives a partial answer to Problem 1 in [Citation4] by giving an example of tent inverse limit space which has uncountably many inhomogeneities with only countably infinitely many of them not being endpoints.

2. Preliminaries

Let I:=[0,1] be the unit interval, and f:II a symmetric unimodal map, i.e. given the involution x~=1x, we assume that f(x~)=f(x) for every x. This means that the critical point c=12, and by an appropriate scaling, we can assume that f(c)=1. For example, fa(x)=1a(x12)2 with a(0,4] is the logistic family in this scaling.

We can turn f into an (increasing) symmetric Lorenz map φ:II by flipping the right half of the graph vertically around c=12, see Figure , giving the following result: φ(x)={f(x)ifx[0,c],f(x)~ifx(c,1]. The choice φ(c)=f(c)=1 is arbitrary, only made to be definite.

Then, ϕ is semi-conjugate to f: fφ=ff. In fact (1) φn(x)={fn(x)iffnisincreasingatx;fn(x)~iffnisdecreasingatx.(1) We can also flip the left branch of f and obtain ψ:=φ~ which is called a decreasing symmetric Lorenz map. Then ψ(x)~=ψ(x~) for all x, and by induction ψn(x)={φ~n(x)=φn(x)~ifnisodd,φn(x)ifniseven. Suppose then ψn is continuous at x. Then (Equation1) implies that ψn(x)=fn(x)ifandonlyif{fnisdecreasingatxandnisodd,fnisincreasingatxandniseven, and ψn(x)=fn(x)~ otherwise.

3. Sharkovsky's Theorem for Lorenz maps

We can describe the dynamics of f using the standard symbolic dynamics with the alphabet {0,,1}, where the symbols stand for the sets [0,c),{c} and (c,1] respectively. It is also enough to restrict the study to the dynamical core [f(0),1], since points from [0,f(0)) will be mapped to the core under f. The kneading invariant ν{0,,1}N is the itinerary of the point 1=f(c). Since the itinerary map xi(x) is monotone in the parity-lexicographical order on {0,,1}N, the kneading invariant is maximal admissible sequence, i.e. i(x)plν for all xI. Also, i(x)plσ(ν) for all x[f(0),1]. It can be shown that every itinerary for which every shift is in parity-lexicographical ordering between σ(ν) and ν can be realized by a point in the dynamical core (see e.g. [Citation25]).

Also, if an m-periodic point y is closest to c from all the points in its orbit, and i(f(y))=e1em¯, then σn(e1em¯)ple1em¯ for all n1. As a corollary, e1em1em¯ (if admissible) is periodic of period k = m or k=m/2, which we prove in the rest of this paragraph. To prove that, assume that there is k3 such that for j=m/k we can write e1em1em=(e1ej)k. Then,since e1em¯ is maximal among its shifts, e1ej<ple1ej, and thus #1(e1ej) is odd. But then e1eje1ej>pl(e1ej)2, so σ(k2)j(e1em1em¯)>ple1em1em¯, violating the parity-lexicographical shift-maximality of e1em¯.

Lemma 3.1

Let f be a unimodal map with a periodic point x of period n. Then for every mn,m>1, there are periodic points y and y of f such that

  1. y has prime period m and fm is decreasing at y, and

  2. y has prime period m and fm is increasing at y or y has prime period m/2 and fm/2 is decreasing at y.

If fn is decreasing at x, then the statement holds for m = n as well.

Proof.

By Sharkovsky's Theorem, f has at least one periodic orbit of period m. Take the m-periodic point y closest to c, so the itinerary e:=i(f(y)) is maximal (w.r.t. the party-lexicographical order pl) among all admissible m-periodic itineraries. Find e by setting ei=1ei if ei and i=km,kN. Otherwise we set ei=ei. Let us first show that e is admissible. Let j1 be the smallest integer such that ejνj. If j<m, then both e,e<plν. If mj and #{1im:ei=1} is odd, then e<ple<plν. The remaining case is #{1im:ei=1} is even and mj.

Assume that m = j. Thus e=ν1νm¯. To show that e=ν1νm¯ is admissible, assume that e>plν. Since #{1im:νi=1} is odd, we have ν<plσm(e)=e<plσm(ν), which contradicts shift-maximality of ν. Thus, e is admissible in this case. Also, e cannot be periodic of period m/2 since e1em1em has an odd number of ones. It follows that e<ple, which contradicts the assumption that e is the closest to ν among m-periodic itineraries, so this case is not possible.

Assume that m<j. Then σm(e)plσm(ν) but since the first symbol at which σm(e)=e and σm(ν) differ is jm, the parity argument and m-periodicity of e imply that σm(ν)>plν, which contradicts the shift-maximality of ν. So this case cannot occur either.

We conclude that e<plν, and since it is shift-maximal, σn(e)<plν for every n0. We still have to argue that σn(e)>plσ(ν) for every n0. Assume there is n0 such that σn(e)<plσ(ν) and take the smallest such n. Since m>1,e starts with 1, and thus n>0. Also, since n is the smallest such integer, σn1(e)=1σn(e). Then σn(e)<plσ(ν) implies that σn1(e)=1σn(e)>plν, which is a contradiction. We conclude that e is admissible, i.e. realized by a point y in [f(0),1].

Moreover, we also conclude that fm is decreasing in y and increasing in y. From the discussion preceding the statement of the lemma, we conclude that the prime period of y is m or m/2.

If y has prime period m/2=k, then e=e1eke1ek1ek¯ and e=e1ek¯. Since σk(e)<ple, we conclude that #{1ik:ei=1} is odd, from which is follows that fk is decreasing in y.

For discontinuous interval maps, there are previous results regarding the forcing relation between periods, see e.g. [Citation3] which however do not give the following result.

Theorem 3.1

Symmetric increasing Lorenz maps ϕ satisfy Sharkovsky's Theorem, except for the fixed points.

Proof.

We start the proof for the symmetric Lorenz map ϕ with two claims.

  1. We first show that if ϕ has a periodic point of prime period n1, then f also has a periodic point of prime period n, unless, possibly, n is a power of 2, and then f has a periodic point of prime period n or n2.

    Let φn(x)=x and assume φk(x)x for all k<n. Then the same holds for x~. At exactly one of x and x~, say at x,fn is increasing, so fn(x)=x. Assume k<n is such that fk(x)=x and take the smallest such (so that x has prime period k). If k is not a power of two, then, since k divides n, Sharkovsky's Theorem gives the existence of a periodic point of prime period n as well. So we only have to consider the case that k=2r.

    If fk is increasing at x, then φk(x)=fk(x)=x, a contradiction. Thus fk must be decreasing at x. In that case f2k is increasing at x, and thus φ2k(x)=f2k(x)=x, from which we conclude that n=2k=2r+1 and x is a periodic point of f of prime period k=n2.

  2. Next we show that if m>1 is such that f has an m-periodic point, then there exists an m-periodic point of ϕ. Assume fm(x)=x and fk(x)x for all k<m. If fm is increasing at x, then φm(x)=x. Assume that there is k<m such that φk(x)=x. Then fk must be decreasing at x, and we get f2k(x)=φ2k(x)=x, thus m = 2k. Now fk(x~)=fk(x)=φk(x)~=x~, so f2k(x)=fk(x~)=x~, but on the other hand f2k(x)=fm(x)=x, which gives a contradiction.

    The remaining case is when fm is decreasing at x. By Lemma 3.1 and its proof, we find a point x such that fm(x)=x and fm is increasing at x. If m is indeed the prime period of x, then we can use the above argument to conclude that x is m-periodic point of ϕ. Otherwise, the prime period of x is m/2 and fm/2 is decreasing in x. But then φm/2(x)=x~ and φm(x)=x, so x is periodic for ϕ with prime period m.

    However, if m = 1, then e=1¯,e=0¯ and x lies in general outside the core (and in fact outside I), so it is lost in the construction of ϕ. Indeed, ϕ has a fixed point only if it comes from a ‘full’ unimodal map f (i.e. a unimodal map that exhibits all possible itineraries of points, such as e.g. the quadratic Chebyshev polynomial f(x)=4x(1x)).

To finish the proof, assume that ϕ has an n-periodic point. By the first part of the proof, there exists an n-periodic point for f (or possibly an n/2-periodic point if n is a power of 2). Sharkovsky's Theorem implies that f has an m-periodic point for every mn. The second part of the proof implies that there exists an m-periodic point of ϕ provided m1.

There are maps f with periodic points of period 2r and no other periods. If r is maximal with this property, we say that f is of type 2r. If f has periodic points of all periods of the form 2r we say that f is of type 2. The union of these two is called type 2. If x is a 2r-periodic point of a unimodal map f of type 2, then we say that x has the pattern from the first period doubling cascade; itinerary of such point is the (shift of the) 2r-periodic continuation of the Feigenbaum itinerary νF=ν1Fν2Fν3F which equals (2) 1.0.11.1010.10111011.1011101010111010.1011101010111011101110101(2) where the dots indicate the powers of 2.

Studying the decreasing symmetric Lorenz maps ψ we can obtain a theorem similar to Theorem 3.1.

Theorem 3.2

Decreasing symmetric Lorenz maps ψ satisfy Sharkovsky's Theorem, possibly except for periods 2r,r1.

Proof.

The proof for a decreasing symmetric Lorenz map ψ is similar as for increasing Lorenz maps. We only need to repeat the two claims.

  1. Let ψn(x)=x and assume ψk(x)x for all k<n. Then the same holds for x~. For even n the proof is the same as for ϕ in Theorem 3.1, so assume that n is odd. At exactly one of x and x~, say at x,fn is decreasing, so fn(x)=x. Assume k is a divisor of n is such that fk(x)=x. Then k and n/k are odd and fk is decreasing as well, so ψk(x)=x, which is a contradiction.

  2. Assume that f has a n-periodic point and take mn. (We note that the claim does not hold for m = n. Indeed, if n = m = 3 and the 3-periodic point is emerging in a saddle node bifurcation, then ψ does not yet have a 3-periodic point.) By Lemma 3.1, f has periodic points x of prime period m, and if m is not a power of two, then we can take x orientation preserving as well as orientation reversing. Assume that fk(x)x for all proper divisors k of m.

    • If m is odd, we take x orientation reversing, so that ψm(x)=x. Suppose that j is a proper divisor of m such that ψj(x)=x. Then fj(x)=x~ because fj(x)x by assumption. Also, x=ψmj(ψj(x))=ψmj(x) so we also conclude that fmj(x)=x~. But then x=fm(x)=fm(x~)=fmj(fj(x~))=fmj(x~)=x~, a contradiction. Therefore m is the prime period of x for ψ.

    • If m is even, we take x orientation preserving, so that ψm(x)=x. Analogously as above we prove that m is the prime period of x for ψ.

This shows that ψ satisfies Sharkovsky's Theorem with the potential exception of periodic points in the first period doubling cascade. For instance, if fa(x)=1a(x12)2 with 4>a>aFeig (where aFeig is the Feigenbaum parameter, then ψ does not have a point of prime period 2, despite the fact that it has periods n2. More generally, if fa is r−1 renormalizable of period 2 (so in contrast with Theorem 3.1 the final renormalization has period 2r1), then ψ has no periodic point of period 2r. The map ψ always has a fixed point, so we don't need to make exceptions for fixed points.

4. Cutting times

We recall some notation from Hofbauer towers and kneading maps that we use later in the paper; for more information on these topics, see e.g. [Citation11, Chapter 6].

Recall that c denotes the critical point 1/2. For nN denote by cn:=fn(c). We assume that c2<c (otherwise the dynamics of f is trivial).

Define inductively D1:=[c,c1], and Dn+1:={[cn+1,c1]ifcDn;f(Dn)ifcDn. We say that n is a cutting time if cDn. The cutting times are denoted by S0,S1,S2, (where S0=1 and S1=2). They were introduced in the late 1970s by Hofbauer [Citation21]. The difference between consecutive cutting times is again a cutting time (see e.g. Subsection 6.1 in [Citation11]), so we can define the kneading map Q:NN{0} as SQ(k):=SkSk1. We call f long-branched if lim infn|Dn|>0, which is equivalent to lim infk|DSk|>0 and also to lim supkQ(k)<.

A purely symbolic way of obtaining the cutting times is the following. Recall that we use the itinerary map i for f (and also for ϕ) with codes 0 for [0,c) and 1 for (c,1]. We will use the modified kneading sequence ν=limxci(x)=10{0,1}N, where we traditionally omit the zero-th symbol. Note that if c is not periodic, ν=i(c1) and the modification is only made so that the itineraries do not contain symbol (we take the smaller of the two sequences in parity-lexicographical ordering).

We can split any sequence e{0,1}N into maximal pieces (up to the last symbol) that coincide with a prefix of ν. To this end, define (3) ρ:NN,ρ(n)=max{k>n:en+1en+2en+k1isprefixofν}.(3) That is, the function ρ depends on e and ν, but we will suppress this dependence. When we apply this for e=ν, we obtain S0=1,Sk+1=ρ(Sk), or in other words Sk=ρk(1) for e=ν and k0.

Define the closest precritical points ζI as any point such that fn(ζ)=c for some n1 and fk(x)c for all kn and x(ζ,c). By symmetry, if ζ is a closest precritical point, ζ~=1ζ is also a closest precritical point. If ζ(ζ,c) is a closest precritical point of the lowest n>n, then the itineraries of f(c) and f(x),x(ζ~,ζ~) coincide for exactly n2 entries, and differ at entry n1. Hence n is a cutting time, say n=Sk for some k1. We use the notation ζ=ζk if n=Sk. That is (4) <ζk<ζk+1<<c<<ζ~k+1<ζ~k<fSk(ζk)=fSk(ζ~k)=c.(4) and (5) xΥk:=(ζk1,ζk][ζ~k,ζ~k1)i(f(x))=ν1νSk1νSk(5)

Applying this to x=fm(c), we obtain that ρ(m)m is a cutting time.

In particular, (6) fSk1(c)ΥQ(k)=(ζQ(k)1,ζQ(k)][ζ~Q(k),ζ~Q(k)1),(6) see Figure , and the larger Q(k), the closer fSk1(c) is to c.

Let κ=min{j>1:νj=1}. Then we can define the co-cutting times as S^0=κ,S^k+1=ρ(S^k), The cutting and co-cutting times are always disjoint sequences (see [Citation13, Lemma 2]), and {S^k}= if f is the full unimodal map (because then ν=10000 and κ is not defined). Furthermore, there is a co-kneading map Q^:NN{0} such that S^k=S^k1+SQ^(k).

Proposition 4.1

Let f be a unimodal map with the kneading map Q. If Q(k), then Q^(k) and ω(c) is a minimal Cantor set.

Proof.

In [Citation12, Lemma 3.6 and Proposition 3.2] and [Citation13, Lemma 4 and Proposition 2] it was shown that Q(k) implies Q^(k) and that c is persistently recurrent. This property was introduced by Blokh and implies minimality of ω(c), see [Citation7] and also [Citation14, Section 3].

In fact, lim supkQ(k)= implies that lim supkQ^(k)=, but not vice versa. If both lim supkQ(k)< and lim supkQ^(k)<, then c is non-recurrent, but as we will see in Section 6, there are maps where lim supkQ(k)<lim supkQ^(k)=.

5. Sturmian shifts

There are multiple ways of defining Sturmian shifts and we take the one using the symbolic dynamics of circle rotations.

Definition 5.1

Let Rα:S1S1,xx+αmod1, be the rotation over an irrational angle α. Let βS1 and build the itinerary u=(un)n0 by (7) un={1ifRαn(x)[0,α),0ifRαn(x)[0,α).(7) Then u is called a rotational sequence. The minimal (and uniquely ergodic) shift space obtained as Xα={σn(u):nN}¯ is the Sturmian shift of frequency α, and each xXα is called a Sturmian sequence.

The purpose of this section is to describe cases when unimodal maps restricted to their critical omega-limit sets ω(c) are conjugate to Sturmian shift. There are in fact multiple ways of choosing the kneading map Q so that (ω(c),f) is Sturmian. The simplest way is by means of the Ostrowski numeration, see [Citation26]. Indeed, let αI be some irrational number and let pn/qn be the convergent of its continued fraction expansion. Thus q1=0,q0=1 and qn=anqn1+qn2. Take kn=j=0naj and then cutting times as follows: {Sk=k+1for0ka1,Skn=qnforn1,Skn+a=aqn+qn1for1aan,n1. It is clear that Q(k) in this case, and the {Sk} interpolate between the numbers qn, see also [Citation16]. However, f:ω(c)ω(c) is in general not invertible, since c itself and/or other points in the backward orbit of c have two preimages in ω(c), see also [Citation15]. As such (ω(c),f) is conjugate to the one-sided Sturmian shift.

However, also when Q(k) is bounded (in fact also when Q(k)1) there are examples where (ω(c),f) is Sturmian, see [Citation12, Chapter III, 3.6]. Let φ:II be an increasing symmetric Lorenz map as in previous sections. In addition to i, another way of coding orbits of unimodal maps (used by Milnor & Thurston [Citation25], Collet & Eckmann [Citation17] and Derrida et al. [Citation19]) is as follows: set ϑ0(x)=+1 and for n1, (8) ϑn(x)=j=0n1(1)ij(x)={+1iffnisincreasingatx;1iffnisdecreasingatx.(8) It follows that ϑ(f(x))=σ(ϑ(x)) if i0(x)=0 and ϑ(f(x))=σ(ϑ(x)) if i0(x)=1. For the itinerary iφ of xIj=0nφj(c) under the function ϕ this means that inφ(x)=0{in(x)=0andϑn(x)=+1,orin(x)=1andϑn(x)=1.ϑn+1(x)=+1, and inφ(x)=1{in(x)=1andϑn(x)=+1,orin(x)=0andϑn(x)=1.ϑn+1(x)=1, In other words, inφ=(1ϑn+1(x))/2. This gives iφφ(x)=σiφ(x). Also, if νφ=limxciφ(x) with the first symbol neglected, and defined ρφ(n)=min{k>n:νkφ=νknφ}, then we recover the cutting times as S0=1,Sk+1=ρφ(Sk). (The co-cutting times can be recovered as S^0=κ=min{k1:νkφ=0} and S^i+1=min{k>S^i:νkφνkS^iφ}.) See the example in the proof of Proposition 5.1.

To each xI we can assign a rotation number by first assigning a lift Φ:RR to the Lorenz map ϕ: Φ(x)={φ(x)ifx[0,c],φ(c)=1;φ(x)+1ifx(c,1);Φ(xn)+nifx[n,n+1). Then Φ(x)mod1=φ(xmod1) and the rotation number is defined as (9) α(x)=lim supnΦn(x)xn,(9) Since Φ(x)=x if and only if xmod1[0,c) and Φ(x)=x+1 otherwise, we obtain (10) α(x)=lim supn1n#{0k<n:ikφ(x)=1}=lim supn1n#{1kn:ϑk(x)=1}.(10) Next we turn ϕ into a proper circle endomorphism (with unique rotation number independent of xS1) by setting: φ¯(x)={φ(1)=f(1)~,x[0,a];wherea<cissuchthatφ(a)=φ(1),φ(x),otherwise. Also let b>c be such that φ(b)=a, see Figure .

The circle endomorphism φ¯ obtained from ϕ was already studied in the last section of [Citation12].

Proposition 5.1

Assume that f is a unimodal map with cutting times {Sj}j0. Let b>c be such that φ¯(b)=a, see Figure . Then the rotation number of the corresponding φ¯ equals α={kSk[12,1]QifkisminimalsuchthatfSk(c)(b^,b),limkkSk[12,1]ifnosuchkexists. In the latter case, the kneading map Q(j)1 for all jN, and if αQ, then f:ω(c)ω(c) is a minimal homeomorphism.

Proof.

Recall that f(c)=1 and assume that there is a minimal integer n1 such that φn(1)(c,b]. Then φ¯n+1(1)(0,a] and φ¯n+2(1)=φ¯(1) is periodic with period n + 1.

Recall that b>c is such that φ¯(b)=a, so f(b)=a~>c, and f2(b)=f(a)=f2(c)~>c. Therefore b(ζ~2,ζ~1) for closest precritical points ζ~1>ζ~2>c, see (Equation4), and b~(ζ1,ζ2). There are two possibilities:

  • φn(1)=fn(1). In this case fn is increasing at 1 and thus n+1=Sk is a cutting time.

  • φn(1)=fn(1)~. In this case fn is decreasing at 1 and again n+1=Sk is a cutting time.

By minimality of k,fSj(c)[b~,b]{c} for all j<k, and hence the kneading sequence ν of f consists of blocks 0 or 11. For example: ν=1.0.0.11.0.11.0.11.101ϑ=+1111+111+111+11+1+11νφ=1.1.1.01.1.01.1.01.001 where dots indicate cutting times and the bold symbol the position Sk. Since n + 1 is the period of φ¯(1), this shows that #{1jSk:ϑj=1}=k, and in view of (Equation10) we have α=k/Sk.

If there is no such minimal n, i.e. φn(1)(b~,b) for all n1, then fn(1)(b~,b) for all n1 (and in particular Q(j)1) for all j1. A counting argument similar to the above shows that α=lim supkk/Sk=limkk/Sk. It is possible that α is rational, e.g. for the logistic map fa(x)=1a(x12)2 with a = 3.5097. In this case, ν=(101) and φ¯i(1) converges to an attracting orbit of period 3. Also for the tent map T(x)=1λ|x12| with λ=12(1+5), the critical orbit {12,1,34145} has period three and avoids [0,a].

If αQ, then ωφ¯(c) is the Cantor set, disjoint from [0,a] and minimal w.r.t. the action of φ¯. Under the semi-conjugacy f between f and ϕ (indeed ff=fφ), this projects to a minimal map f:ωf(c)ωf(c). We will show that f:ωφ(c)ωf(c) is in fact a homeomorphism, from which it follows that f:ωf(c)ωf(c) is also a homeomorphism. Assume by contradiction that x<c<x~ are points in ωφ(c) such that f(x)=f(x~)=yωf(c). Then, since f is the semi-conjugacy between ϕ and f, we must have f(φn(x~))=f(φn(x))=fn(y) for every nN. Note that φn(x~)=φn(x)~ for every nN, and thus φn(x~)φn(x), unless φn(x)=c, and thus fn(x)=c. Since c is not periodic, there exists NN such that φn(x~)φn(x) for all nN, and thus fn(y) has two f-preimages in ωφ(c). Since f:ωf(c)ωf(c) is minimal, for every ε>0 there exists infinitely many yorbf(y) which are ϵ-close to f2(c)=f(1). For sufficiently small ϵ, an f-preimage of a point ϵ-close to f(1) will be contained in [0,a]. Since every point in orbf(y) eventually has both f-preimages in ωφ(c), we conclude that ωφ¯(c)[0,a]=ωφ(c)[0,a], which is a contradiction.

We argued so far that there exist stunted Lorenz maps for which orbφ¯(c)¯ is a Cantor set with dynamics similar to circle rotations (in fact to Denjoy circle maps) with irrational rotation number, and that there are also unimodal maps with kneading map bounded by 1, such that f|ω(c) is semi-conjugate to a circle rotation, and in fact, the rotation number is α=limkk/Sk. Therefore (ω(c),f) represents a Sturmian shift.

In fact, every irrational rotation number (hence every Sturmian shift) can be realized this way, as we can prove by studying this rotation number closer. Indeed, let α=[0;a1,a2,a3,] be the continued fraction expansion of ρ, with convergents piqi. For the irrational rotation Rα, the denominators qi are the times of closest returns of any point xS1 to itself, and these returns occur alternatingly on the left and on the right. If we assume that Rαqi(x) is to the right of x, and set Aqi=[x,Rαqi(x)], then the first iterate k such that Rαk(Aqi)x is k=qi+1 and Rαqi+1(x) is to the left of x.

For the map φ¯, the closest returns on the left indeed accumulate on c, but the right neighbourhood [c,b) is the preimage of the plateau [0,a) and no further iterates of c enter that region. Instead, returns on the left accumulate on b.

Translating this back to the unimodal map f with kneading sequence ν=ν1ν2ν3, the closest returns on the left correspond to closest returns at co-cutting times (recall that there are no cutting times Sj so that fSj(c)(b~,b)). If qi is such a co-cutting time, then (recalling the function ρ from (Equation3) and using the above argument), the Farey convergents ρa(qi)=qi+aqi+1 are also the next co-cutting times for 1aai+1, and in particular, ρai+1(qi)=qi+2.

The closest returns on the right correspond to cutting times, but this time fqi(c) accumulate on b, and because f3(b)=f3(c), the itinerary of b is (11) i(b)=b1b2b3b4b5=11ν3ν4ν5(11) Therefore we need to consider the analogous function ρb(m)=min{n>m:bnbnm}, and find that ρba(qi)=qi+aqi+1 for 1aai+1, and in particular, ρbai+1(qi)=qi+2.

For example, if ai2, so the qis are the Pell numbers 2,5,12,29,70,189,, then we obtain ν=10.11.11.0.11.11.0.11.11.11.0.11.11.0.11.11.11.0 where dots indicate cutting times and primes co-cutting times. The bold symbols indicate the positions qi. In fact, for each i νqi+1qi+1νqi+11νqi+1=ν1νqi1νqiorν1νqi1νqiforeacheveniN, and therefore c has two limit itineraries limxci(x)=0ν and limxci(x)=1ν, but c has only one preimage in ω(c).

6. Outside maps and unimodal inverse limit spaces

Boyland, de Carvalho and Hall in [Citation9, Section 3] present a different way of creating a circle endomorphism from a unimodal map. They call this the outside map B, and use it to study the inverse limit space of the unimodal map as attractors of sphere homeomorphisms. Starting from a unimodal map f:II such that the second branch is surjective (i.e. f([c,1])=I), they double the interval to a circle R/2Z=[0,2]/02, and let B map the second branch onto [1,2] by flipping this branch, and then extend the definition of f on [1,2d] for the unique point d(c,1] for which f(d)=f(0) to cover the interval [0,f(0)]. The remaining interval [2d,2] is then mapped to the constant f(0). That is B(x)={f(x)ifx[0,c);2f(x)ifx[c,1);f(2x)ifx[1,2d);f(0)ifx[2d,2), see Figure . Let us carry this out for the family of cores of tent maps Tλ:II, Tλ={λx+2λ,x[0,λ1λ],λx+λ,x[λ1λ,1], for all λ(1,2].

Figure 4. Constructing the outside and stunted Lorenz map for a tent map Tλ.

Figure 4. Constructing the outside and stunted Lorenz map for a tent map Tλ.

Then the map (12) φ¯(x)={λ2if0xa=λ1λ;λ(x12)mod1ifa=λ1λx<1,onR/Z(12) and the outside map B(x)={λ(x1)+2mod2if0x<2λ;2λif2λx<2,onR/2Z are conjugate with conjugacy G:R/ZR/2Z,G(x)=2(1x)mod2, i.e. Gφ¯=BG. But the conjugacy reverses orientation, so the rotation numbers are each others opposite, α for φ¯ versus 1α for B.

Outside map B was used in [Citation9] to give a complete description of the prime end and accessible points structure in unimodal inverse limits embedded in the plane as attractors of an orientation-preserving homeomorphism of the plane (or the two-dimensional sphere S2).

In the rest of the section we restate some results from [Citation9] and relate them to the established conjugacy between maps B and φ¯.

Recall that I denotes the unit interval [0,1]. The inverse limit space with the bonding map g:II is a subspace of the Hilbert cube IN0 defined by lim(I,g):={(x0,x1,x2,)IN0:g(xi+1)=xi,iN0}. Equipped with the product topology, the space lim(I,g) is a continuum, i.e. compact and connected metric space. Define the shift homeomorphism g^:lim(I,g)lim(I,g), g^((x0,x1,)):=(g(x0),x0,x1,)for(x0,x1,)lim(I,g). There is a natural way to make lim(I,g) an attractor of an orientation preserving sphere homeomorphism. Such embeddings are called Brown-Barge-Martin embeddings (abbreviated BBM embeddings), see [Citation6] for the original construction, [Citation8] for generalization of the construction to parametrized families and [Citation9] for the construction applied to unimodal inverse limits. As the outcome of the BBM embedding of lim(I,g), one obtains an orientation-preserving homeomorphism H:S2S2 so that H|lim(I,g) is topologically conjugate to g^ and for every xS2{point}, and ω(x,H) is contained in lim(I,g).

In [Citation9] the authors study in detail BBM embeddings of inverse limits of unimodal maps satisfying certain (mild) conditions, which are in particular satisfied for the tent map family Tλ,λ(2,2]. For simplicity we state the following results for tent maps only, noting that they can be generalized to a much wider class of unimodal maps.

Fix the slope λ(2,2] for the tent map Tλ. Let B be the corresponding outside map. Denote by S^=lim{S1,B}. Theorem 4.28 from [Citation9] shows that there is a natural homeomorphism h between S^ and the circle P of prime ends of lim(I,Tλ). Then h conjugates the shift homeomorphism B^ of the outside map to the action of H on P, so that the prime end rotation number of lim(I,Tλ) is equal to the rotation number (as defined in (Equation9)) of B^, see [Citation9, Lemma 4.30]. Finally, Corollary 4.36 in [Citation9] gives that the prime end rotation number of B^ is equal to the rotation number of B. Since φ¯ and B are conjugate, the results above follow analogously and by Proposition 5.1 we obtain that the prime end rotation number of B^ equals 1α.

Proposition 6.1

Let Tλ be a tent map with slope λ(2,2] and let φ¯ be corresponding stunted Lorenz map with rotation number α. Let lim([0,1],Tλ) be embedded in S2 by a BBM construction. Then the prime end rotation number of T^λ on lim([0,1],Tλ) equals 1α.

Remark 6.1

In [Citation9], the prime end rotation number is expressed in terms of the height q(ν) of the kneading sequence of a unimodal map f (see the definition of height in e.g. [Citation9, Section 2.6]). Proposition 5.1 thus gives an algorithm to compute the height of the kneading sequence in the following way: find the smallest nN such that cn(1b,b), and n is a cutting time n=Sk. Then the height equals 1k/Sk. If no such n exists, then the height equals 1limkk/Sk. Recall that b>c is such that f2(b)=f2(c)^, so the itinerary of b is i(b)=11ν3ν4ν5, see (Equation11), where ν=i(f(c))=10ν3ν4ν5 is the kneading sequence. Hence, the previous condition can be expressed with symbols as 01ν3ν4ν5νnνn+1νn+211ν3ν4ν5, where ≺ denotes the parity-lexicographic ordering on symbolic sequences.

Furthermore, [Citation9] gives the complete characterization of accessible points the BBM embeddings of lim(I,Tλ) using the outside map. We emphasize it here and connect it to the stunted Lorenz map φ¯.

Let S:[0,2]R2 be a circle parametrisation defined by S(t)=(12+12cos(π+tπ),12sin(π+tπ)) for t[0,2]. Let τ:S([0,2])I be the vertical projection, i.e. τ((x,y))=x for (x,y)S([0,2]). Furthermore, let γ˚=(S(2/λ),S(2)). As before, let S^=lim{S([0,2]),B}.

Proposition 6.2

Theorem 4.28(d), Remark 4.15, Definition 4.12, Corollary 4.14 in [Citation9]

Let lim(I,Tλ) be embedded in R2 by an orientation-preserving BBM embedding. Then (x0,x1,x2,)lim(I,Tλ) is accessible if and only if there exists N0 and y=(y0,y1,y2,)S^ such that yiγ˚ for all i>N and such that xN+j=τ(yN+j) for all j0.

Using the conjugacy of B and φ¯, we can state the previous theorem in terms of φ¯ directly. We parametrize the circle above as T:IR2 as T(t)=(12+12cos(π+2tπ),12sin(π+2tπ)) and let δ˚=(T(0),T(λ1λ)). The vertical projection onto a horizontal diameter is denoted by τ as above (that is actually τG, where G is the conjugacy between φ¯ and B and it is equal to the vertical projection). In particular, τφ¯(x)=Tλτ(x) for all xδ˚.

Theorem 6.1

Let Tλ be a tent map with slope λ(2,2] and let φ¯ be corresponding stunted Lorenz map with rotation number α. Let lim(I,Tλ) be embedded in S2 by an orientation-preserving BBM embedding. Let S^=lim(S1,φ¯). A point (x0,x1,x2,)lim(I,Tλ) is accessible if and only if there exists N0 and y=(y0,y1,y2,)S^ such that yiδ˚ for all i>N and such that xN+j=τ(yN+j) for all j0.

We say that a point x=(x0,x1,)lim(I,Tλ) is a folding point if xnωTλ(c) for every n0. In the context of inverse limits on intervals this is equivalent to saying that x has no neighbourhood homeomorphic to the Cantor set times an open interval (see [Citation27]). In the case when rotation number of φ¯ is irrational, Proposition 5.1 and its proof imply that f:ωφ(c)ωTλ(c) is a homeomorphism (recall that orbits of c under ϕ and φ¯ are the same when the corresponding height of the tent map is irrational). From that and Theorem 6.1 we have the following:

Corollary 6.1

If λ(2,2] and the rotation number of φ¯ is irrational (i.e. the height of the kneading sequence of Tλ is irrational), then every folding point of lim(I,Tλ) embedded in R2 by the orientation-preserving BBM embedding is accessible.

Proof.

We first note that τ:ωφ¯(c)ωTλ(c) is well defined and bijective. The first part follows since τ(limiφ¯ni(c))=limiτφ¯ni(c)=limiTλniτ(c)=limiTλni(1)ωTλ(c), when limiφ¯ni(c) exists. Similar argument also shows that τ is surjective. For the proof of injectivity, it is enough to note that τ(x)=τ(y) implies that y=x~ or y = x and apply the fact that Tλ:ωφ¯(c)ωTλ(c) is a homeomorphism.

Now let (x0,x1,x2,)lim(I,Tλ) be such that xiωTλ(c) for every i0. Then φ¯(τ1(xi))=τ1(Tλ(xi))=τ1(xi1) for every i>0, so (τ1(x0),τ1(x1),τ1(x2),)lim(S1,φ¯). We apply Theorem 6.1 to conclude that (x0,x1,x2,) is accessible.

Remark 6.2

Given a continuum X, we say that point xX is an endpoint if for every subcontinua A,BX such that xAB, we have AB or BA. In [Citation5] it was shown that if lim(I,Tλ) is embedded in the plane by an orientation-preserving BBM embedding, and if φ¯ has irrational rotation number (i.e. the height of the kneading sequence of Tλ is irrational), then all endpoints are accessible. Moreover, it was shown that there also exist countably many accessible non-end folding points. Corollary 6.1Footnote1 in particular implies that there are uncountably many endpoints and only countably many non-end folding point in lim(I,Tλ). This partially answers Problem 1 in [Citation4].

Let us discuss the irrational rotation number case in more details. If 1α (and hence α) is irrational, then Proposition 5.1 gives that Q(j)1 for all j, and Tλ:ω(c)ω(c) is a Cantor minimal homeomorphism conjugate to a Sturmian shift. This implies that H^ induces Denjoy-like dynamics on the corresponding circle of prime ends P. In [Citation5], a detailed characterization of accessible points for BBM embeddings of tent inverse limit spaces is given (there, also accessible endpoints and non-end folding points are distinguished), based solely on symbolic dynamics techniques from kneading theory. It follows (see [Citation5, Theorem 11.20]) that there is a Cantor set CP corresponding to accessible folding points in lim(I,Tλ) uncountably many of which are endpoints and countably many are non-end folding points. Furthermore, all endpoints are accessible. The remaining countably infinitely many open arcs PC correspond to countably infinitely many open arcs in different arc-components of lim(I,Tλ) (unions of all arcs containing some point from lim(I,Tλ)) that are accessible at more than one point. Thus this planar continua are interesting also from a topological perspective. A theorem of Mazurkiewicz [Citation24] shows that for every indecomposable planar continuum there are at most countably infinitely many arc-components accessible at more than one point. Our examples confirm that it is possible to find planar continua indeed having countably infinitely many arc-components accessible at more than one point. Furthermore, Tλ is long-branched since supjQ(j)=1. Therefore, all proper subcontinua of lim(I,Tλ) are arcs (see e.g. [Citation10, Proposition 3]).

Thus, from the discussion in this section we have a complete understanding of topology of lim(I,Tλ) as well as their orientation-preserving BBM planar embedding in the case when the rotation number of φ¯ is irrational (that is, the height of the kneading sequence of Tλ is irrational).

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

AA was supported by grant 2018/17585-5, São Paulo Research Foundation (FAPESP). HB gratefully acknowledges the support of the FWF stand-alone grant number P31950-N45. JČ was supported by the FWF Schrödinger Fellowship stand-alone project J 4276-N35 and University of Ostrava grant IRP201824 ‘Complex topological structures’; Austrian Science Fund (J-4276,P31950-N45).

Notes

1 Its statement was suggested to us by Boyland, de Carvalho, Hall through personal communication.

References

  • V.S. Afraimovich, V.V. Bykov, and L.P. Shil'nikov, On structurally unstable attracting limit sets of the Lorenz attractor type, Trudy Moskov. Mat. Obsch. 44 (1982), pp. 150–212.
  • J.-P. Allouche, M. Clarke, and N. Sidorov, Periodic unique beta-expansions: The Sharkovski ordering, Ergodic Theory Dynam. Syst. 29 (2009), pp. 1055–1074. doi: 10.1017/S0143385708000746
  • L. Alsedà, J. Llibre, M. Misiurewicz, and C. Tresser, Periods and entropy for Lorenz maps, Ann. Inst. Fourier 39 (1989), pp. 929–952. doi: 10.5802/aif.1195
  • L. Alvin, A. Anušić, H. Bruin, J. Činč, Folding points of unimodal inverse limit spaces. Nonlinearity 33(1) (2020), pp. 224–248. doi: 10.1088/1361-6544/ab4e31
  • A. Anušić, J. Činč, Accessible points of planar embeddings of tent inverse limit spaces. Diss. Math. 541 (2019), pp. 1–57.
  • M. Barge and J. Martin, Construction of global attractors, Proc. Am. Math. Soc. 110 (1990), pp. 523–525. doi: 10.1090/S0002-9939-1990-1023342-1
  • A. Blokh and L. Lyubich, Measurable dynamics of S-unimodal maps of the interval, Ann. Sci. Ec. Norm. Sup. 24 (1991), pp. 545–573. doi: 10.24033/asens.1636
  • P. Boyland, A. de Carvalho, and T. Hall, Inverse limits as attractors in parametrized families, Bull. Lond. Math. Soc. 45(5) (2013), pp. 1075–1085. doi: 10.1112/blms/bdt032
  • P. Boyland, A. de Carvalho, and T. Hall, Natural extensions of unimodal maps: Virtual sphere homeomorphisms and prime ends of basin boundaries, Geom. Topol. (2017). preprint arXiv:1704.06624, to appear in.
  • K. Brucks and H. Bruin, Subcontinua of inverse limit spaces of unimodal maps, Fund. Math. 160 (1999), pp. 219–246.
  • K. Brucks and H. Bruin, Topics from One-Dimensional Dynamics, London Mathematical Society Student Texts 62, Cambridge University Press, 2004.
  • H. Bruin, Invariant measures of interval maps, PhD-Thesis, University of Delft, 1994.
  • H. Bruin, Combinatorics of the kneading map, Internat. J. Bifur. Chaos Appl. Sci. Engrg. 5 (1995), pp. 1339–1349. doi: 10.1142/S0218127495001010
  • H. Bruin, Topological conditions for the existence of Cantor attractors, Trans. Amer. Math. Soc. 350 (1998), pp. 2229–2263. doi: 10.1090/S0002-9947-98-02109-6
  • H. Bruin, Homeomorphic restrictions of unimodal maps, Contemp. Math. 246 (1999), pp. 47–56. doi: 10.1090/conm/246/03773
  • H. Bruin, G. Keller, and M. St. Pierre, Adding machines and wild attractors, Ergod. Th. Dyn. Sys.17 (1997), pp. 1267–1287. doi: 10.1017/S0143385797086392
  • P. Collet and J.-P. Eckmann, Iterated Maps of the Interval As Dynamical Systems, Birkhäuser, Boston, 1980.
  • D. Cosper, Periodic orbits of piecewise monotone maps, PhD.-thesis, Purdue University, May 2018.
  • B. Derrida, A. Gervois, and Y. Pomeau, Iteration of endomorphisms on the real axis and representation of numbers, Ann. Inst. H. Poincaré Sect. A (N.S.) 29 (1978), pp. 305–356.
  • J. Guckenheimer, The strange, strange attractor, in The Hopf Bifurcation and its Applications, J. E. Marsden and M. McCracken, eds., (Springer Lecture Notes in Applied Mathematics), New York, 1976, pp. 368–381.
  • F. Hofbauer, The topological entropy of a transformation x↦ax(1−x), Monatsh. Math. 90 (1980), pp. 117–141. doi: 10.1007/BF01303262
  • F. Hofbauer, Periodic points for piecewise monotonic transformation, Ergod. Th. Dyn. Sys. 5 (1985), pp. 237-–256. doi: 10.1017/S014338570000287X
  • L. Jonker, Periodic orbits and kneading invariants, Proc. Lon. Math. Soc. 39(3) (1979), pp. 428–450. doi: 10.1112/plms/s3-39.3.428
  • S. Mazurkiewicz, Un théorème sur l'accessibilité des continus indécomposables, Fund. Math. 14 (1929), pp. 271–276. doi: 10.4064/fm-14-1-271-276
  • J. Milnor and W. Thurston, On iterated maps of the interval: I, II, Preprint 1977. Published in Lect. Notes in Math. 1342, Springer, Berlin New York, 1988, pp. 465–563.
  • A. Ostrowski, Bemerkungen zur Theorie der diophantischen Approximationen, Hamb. Abh. 1 (1921), pp. 77–98. doi: 10.1007/BF02940581
  • B. Raines, Inhomogeneities in non-hyperbolic one-dimensional invariant sets, Fund. Math. 182 (2004), pp. 241–268. doi: 10.4064/fm182-3-4
  • D. Rand, The topological classification of Lorenz attractors, Math. Proc. Camb. Phil. Soc. 83 (1978), pp. 451–460. doi: 10.1017/S0305004100054736
  • A. Sharkovsky, Coexistence of cycles of a continuous map of the line into itself, (Russian) Ukrain. Math. Zh. 16 (1964), pp. 61–71.
  • R.F. Williams, The structure of Lorenz attractors, Publ. Math. Inst. Hautes Études Sci. 50 (1979), pp. 73–99. doi: 10.1007/BF02684770