Instituto Nacional de Matemática Pura e Aplicada

Doctoral Thesis

KNEADING SEQUENCES FOR TOY MODELS OF HÉNON MAPS

Ermerson Rocha Araujo

Rio de Janeiro July 30, 2018 Instituto Nacional de Matemática Pura e Aplicada

Ermerson Rocha Araujo

KNEADING SEQUENCES FOR TOY MODELS OF HÉNON MAPS

Thesis presented to the Post-graduate Program in Math- ematics at Instituto Nacional de Matemática Pura e Aplicada as partial fulfillment of the requirements for the degree of Doctor in Philosophy in .

Advisor: Enrique Ramiro Pujals

Rio de Janeiro 2018 Amar e mudar as coisas me interessa mais. Alucinação - Belchior To Sheila Cristina and José Ribamar Acknowledgments

I would like to especially express my gratitude to my advisor Enrique Pujals. I thank him for constant encouragement and for many insightful conversa- tions during the preparation of the thesis. Our meetings were almost always held in the IMPA’s restaurant. Which was very good for me, because this took away the pressure that the elaboration of a thesis puts on the students. At each meeting with him many new ideas arose for the problems we were working on. He made extremely complicated arguments seem simple. Besides being an extraordinary mathematician, he is a great person, generous and very attentive. So I can say that it was an honor and an immense adventure to have been a Pujals’ student. I thank IMPA for these four years in which I had the opportunity to grow as a mathematician. In addition, I would like to express my gratitude to the Graduate Center-CUNY in New York City, USA, where I have been twice. Here I thank Pujals again for his help in making possible my visit at the Graduate Center. I thank especially to Adriana Sánchez and Yaya Tall for many discussions during our journey at IMPA as well as by your friendships, particularly dur- ing our preparation for the Qualification Exam. We are the Trinity of the Dynamics-2014. I would like to thank Alex Zamudio, Cayo Dória, Fernando Lenarduzzi, Pedro Gaspar and Sandoel Vieira for many inspiring discussions and in par- ticular spending many hours with me during the final part of the preparation of the thesis. Especially, I thank Alex Zamudio again for his detailed read-

v Ermerson Rocha Araujo Kneading Sequences ing and suggestions for early versions of the thesis that helped me fix small mistakes. I thank my committee members, Alex Zamudio, Carlos Gustavo Moreira (Gugu), , Patrícia Cirilo, Pierre Arnoux and Yuri Lima, for their comments and suggestions. In Rio de Janeiro, Brazil, I had the company of excellent people. To all, my thanks. Finally, I wish to thank my parents Sheila Cristina and José Ribamar and my brother Herbert Araujo for the support, encouragement and love. Without them the walk would certainly be more difficult. I would like to thank CNPq for giving me the financial support.

IMPA vi July, 2018 Abstract

We present a study on how a certain type of combinatorial equivalence implies topological conjugacy. In this work, we present the concept of kneading se- quences for a more general setting than one-dimensional dynamics. For this, we consider a two-dimensional family introduced by Benedicks and Carleson [BC91] as a toy model for Hénon maps and define the notion of kneading se- quences of the critical line for a toy model. We show that these sequences are a complete invariant of the conjugacy class of the toy model. Furthermore, we show a version of Singer’s Theorem for toy models and a combinatorial equivalence result for nonautonomous discrete dynamical systems.

Keywords: Toy models, kneading sequences, negative Schwarzian derivative, nonautonomous discrete dynamical systems

vii Resumo

Nós estudamos condições para que uma equivalêcia combinatória implique uma conjugação topológica. Neste trabalho, apresentamos a noção de se- quências kneading para um contexto mais geral em relação a dinâmicas unidi- mensionais. Para este fim, consideramos uma família bidimensional definida por Benedicks e Carleson [BC91] como um “toy model” para aplicações de Hénon e definimos sequências kneading da linha crítica. Mostramos que tais sequências constituem um invariante completo para as classes de conjugação dos “toy models”. Além disso, mostramos um resultado análogo ao Teorema de Singer para nossa família bem como uma equivalência combinatória para sistemas dinâmicos não autônomos.

Palavras-chave: Toy models, sequências kneading, derivada Schwarziana negativa, sistemas dinâmicos não autônomos

viii Contents

Acknowledgments ...... v Abstract ...... vii Resumo ...... viii

1 Introduction 1

2 The one-dimensional case revisited 6 2.1 The continuous case ...... 6 2.2 The discontinuous case ...... 11 2.2.1 One-side topology ...... 12 2.2.2 Kneading sequence for discontinuous unimodal maps . 12

3 Toy models and kneading sequences 14 3.1 Notations and preliminaries ...... 15 3.2 Kneading sequences and main results ...... 20

4 Proofs of the main results 24 4.1 Proof of Theorem 2 ...... 24 4.2 Proof of Theorem A ...... 28 4.3 Proof of Theorem C ...... 32

5 On Singer’s Theorem for toy models 37

6 Open questions and future work 41 6.1 Toy model and further results ...... 41

ix 6.2 Looking for combinatorial structure for two dimensional dissi- pative diffeomorphisms ...... 42

A Generalized toy models 44

B Combinatorial equivalence for nonautonomous discrete dy- namical systems 47 B.1 Combinatorial equivalence for NDS ...... 48 CHAPTER 1

Introduction

One of the main questions in Dynamical Systems is whether two systems are ‘the same’, where by ‘the same’ we mean some type of equivalence between two systems. In this sense, one of the most simple ways to say that two systems are equivalent is by obtaining an orbit equivalence map between them. Roughly speaking, we say that two topological dynamical systems are topologically orbit equivalent if there exists a homeomorphism between their phase spaces that preserves their structures and induces a one-to-one correspondence be- tween their orbits. Formally, let X and Y be two topological spaces, and let f : X → X and g : Y → Y be functions. We say that f and g are orbit equivalent whenever there is a homeomorphism h : X → Y sending orbits to orbits, that is, h(Of (x)) = Og(h(x)) for every x ∈ X. When f and g are homeomorphisms, the definition means that there exist func- tions α, β : X → Z such that for all x ∈ X, h ◦ f(x) = gα(x) ◦ h(x) and h ◦ f β(x)(x) = g ◦ h(x). So, given f : X → X a system we want to charac- terize the orbit equivalence class of f, that is, we want to determine the set [f] := {g : X → X; g is orbit equivalent to f}. The notion of orbit equivalence was firstly studied in the context of prob- ability measure preserving group actions, where the homeomorphism h is replaced by a measurable isomorphism. It follows from works of Dye [Dye59] and [Dye63], Ornstein and Weiss [OW80], and Connes, Feldman and Weiss

1 Ermerson Rocha Araujo Kneading Sequences

[CFW81] one of the most remarkable results: Any probability measure pre- serving action of an amenable group is orbit equivalent to a probability mea- sure preserving action of Z. This implies that in the measurable setting there is only one orbit equivalence class, at least when the action is made by an amenable group. One notion stronger than orbit equivalence and which is the main problem addressed in the thesis is topological conjugacy. Let X and Y be topological spaces, and let f : X → X and g : Y → Y be continuous functions. We say that f : X → X and g : Y → Y are topologically conjugate (or conjugate for simple) if there exists a homeomorphism h : X → Y satisfying the conjugacy equation h ◦ f = g ◦ h. Note that if f is topologically conjugate to g, then f is orbit equivalent to g. The space where the characterization of the topological conjugacy class was firstly constructed is the circle. In the late XIX century, Poincaré intro- duced the notion of rotation number for orientation preserving homeomor- phisms of circle. He showed that if f : S1 → S1 is an orientation preserving homeomorphism with irrational rotation number, then there exists a rigid rotation R with the same rotation number such that f is topologically semi- conjugate to R. Furthermore, if f contains a point x whose orbit is dense in S1, then f is topologically conjugate to R. The next step is to consider an interval instead of the circle. However, if f : I → I is a homeomorphism, then the dynamic of f is trivial, because for all points x ∈ I we get that either x is a periodic point or f n(x) converges to a periodic orbit. Therefore, for the interval the class of endomorphisms is more interesting to work with. In this direction, Milnor and Thurston, in their famous paper [MT88], gave origin to the Kneading Theory for endomorphisms of the interval, which is an analogue to the Poincaré theory for homeomorphisms of the circle. More specifically, they considered maps f : I → I, where I is an interval, with a finite number of turning points (a point of change of monotonicity) and defined the notion of kneading sequences (the itinerary of the turning points). They proved the following theorem: Theorem 1.1. Suppose that f, g : I → I are two l-modal maps with turning f f g g points c1 < ··· < cl and c1 < ··· < cl . Assume that f and g have no wandering intervals, no intervals of periodic points and no attracting periodic points. If f and g have the same kneading sequences, then f and g are topologically conjugate.

IMPA 2 July, 2018 Ermerson Rocha Araujo Kneading Sequences

As our main result uses the same ideas, we will give the proof of Theorem 1.1 in the next section. The conjugacy is constructed by matching the orbits. We match “the inverse orbits” of the turning points. Besides this, the knead- ing theory plays an important role in one-dimensional dynamics, such as the continuity of the topological entropy and the monotonicity of the kneading sequence for the quadratic family. See de Melo and van Strien [dMvS93]. Over the last years, much research has been done attempting to construct a similar theory in dimension two. However, not much progress has been made. The great difficulty is the lack of critical points in the usual sense and the fact that the plane does not have a natural order like in dimension one. In [PRH07], Pujals and Hertz with the goal to characterize the dynamical phenomenon that obstructs hyperbolicity, defined a notion of critical point for dissipative surface diffeomorphisms. Based on the numerical results of [CGP88], Cvitanović introduced in [Cvi91] the concept of pruning fronts to homeomorphisms of the plane like Hénon family and Lozi family. The definition of pruning fronts is somewhat technical so we will not give it here. Cvitanović conjectured that every map

Ha,b in the Hénon family can be understood as a pruned horseshoe. That is, if F : R2 → R2 is the Smale’s horseshoe map, then after pruning (destroying) ˜ some orbits of F we have a map F equivalent, in some sense, to Ha,b. This is known as the Pruning Front Conjecture. In this direction, Carvalho showed in [dC99] the following:

Theorem 1.2. Let f : R2 → R2 be a homeomorphism of the plane and P ⊂ R2 a pruning front of f. Then there exists an isotopy H : R2 × [0, 1] → R2 with supp(H) ⊂ S f k(P ), such that H(·, 0) = f(·) and H(·, 1) = f (·) is k∈Z P a homeomorphism under which every point of P is wandering.

In other words, given a pruning front P of a homeomorphism f of the plane, up to isotopy, we can destroy all orbits of f which enter P (that is, transform these orbits into dynamically irrelevant orbits), while the dynamic outside P does not change. Although Carvalho did not prove the Pruning Front Conjecture, he constructed a family containing the Hénon and Lozi families of two-dimensional homeomorphisms going from trivial dynamics to a complex dynamics (that is, a horseshoe) like the full logistic family. Mendonza proved in [Men13] that the Pruning Front Conjecture holds in an open set of parameter space. More specifically,

Theorem 1.3. There exists an open set A in the real parameter plane, such

IMPA 3 July, 2018 Ermerson Rocha Araujo Kneading Sequences

that if (a, b) ∈ A then Ha,b is topologically conjugate to a pruning homeomor- phism of the horseshoe.

In [Ish97], Ishii gave a solution of the Pruning Front Conjecture for the Lozi family. For more detail on pruning and its relationship with kneading theory, see [dCH02] and [dCH03]. With an approach different from the pruning techniques, Mendes and Sousa Ramos, in [MR04], developed a kneading theory for two-dimensional triangular maps.

Recently in [MŠ16], Misiurewicz and Štimac studied the Lozi family La,b and their strange attractors Λa,b. They introduced a countable set of knead- ing sequences for the Lozi map and proved that all the dynamics in Λa,b is characterized by this set. In this work we will introduce the concept of Kneading Sequences for the two-dimensional family studied by Benedicks and Carleson, in [BC91], as a toy model for the Hénon maps. Namely, we will consider maps of the form F (x, y) = (f(x, y),K(x, y)) (toy model ) acting on a two-dimensional rectangle, where f is a family of unimodal maps and K is an inverse branch of a Cantor map. In [MMP13], Matheus et al proved that Smale’s Axiom A property is C1-dense among the systems in this family and, on the C2- topology, there exists an open subset where we have a type of Newhouse phenomenon. This indicates that this family may have other interesting properties. Now we can state the main result of this thesis.

Theorem A. Let F and G be two toy models. Assume that F and G have no wandering intervals, no interval of periodic points and no weakly attracting periodic points. If F and G have the same kneading sequences, then F and G are topologically conjugate.

Following this approach we obtain an analogue theorem for more general toy models. Besides that, the same ideas allow us to prove a combinatorial equivalence for nonautonomous discrete dynamical systems. We also prove a Singer’s Theorem for toy models:

F Theorem B. Let F (x, y) = (f(y)(x),Ksign(x)(y)) be a toy model. Suppose that f(y):[−1, 1] → [−1, 1] has negative Schwarzian derivative for all y ∈ [0, 1]. Then the closure of the immediate basin of any strong attracting periodic orbit contains either a point of the critical line or a point of Λ := {−1, 1} × [0, 1].

IMPA 4 July, 2018 Ermerson Rocha Araujo Kneading Sequences

This thesis is organized as follows. In the second chapter we recall the notions of kneading sequences for the one-dimensional case and prove The- orem 1.1 in a particular case. In the third chapter, we deal with the notion of kneading sequence for toy models and present our main results. In the fourth chapter, we prove our main results. In the fifth chapter, we present and proof Singer’s Theorem for toy models. Finally, in the last chapter, we discuss some open problems related to the thesis.

IMPA 5 July, 2018 CHAPTER 2

The one-dimensional case revisited

The purpose of this chapter is to revisit the notion of kneading sequence for unimodal maps. Besides that, we extend the results for discontinuous unimodal maps.

2.1 The continuous case

Let I = [a, b] be a compact interval and f : I → I be a continuous map. We say that f is a unimodal map if f(∂I) ⊂ ∂I and there exists a unique point c ∈ I\∂I, the turning point, such that f is increasing to the left and decreasing to the right of c. N Consider the space Σf = {L, c, R} , where L = [a, c) and R = (c, b].

For each x ∈ I define the itinerary if (x) = (i0(x), i1(x), . . . , in(x),...) ∈ Σf , n n where in(x) = L if f (x) ∈ L, in(x) = R if f (x) ∈ R and in(x) = c if n f (x) = c. It is known that we can define an order structure 4f on Σf such that the map if is order preserving. The sequence

if (c) = (i0(c), i1(c), . . . , in(c),...) is called the kneading invariant of f.

The importance of the sequence if (c) is that it contains all information about the conjugacy class of f. Consider the set of pre-critical points

C (f) = {x ∈ [a, b]; f n(x) = c for some n ≥ 0}.

6 Ermerson Rocha Araujo Kneading Sequences

The theorem below is well known and a particular case of Theorem 1.1, but, for the sake of completeness, we will proof it. This will give us intuition about how we will generalize it in the Theorem A. A similar proof of Theorem 2.1 can be found in [Ran78, Thm1].

Theorem 2.1. Suppose that f, g : I → I are two unimodal maps with turn- ing points cf and cg. Assume that if (cf ) = ig(cg). Then there exists a strictly increasing bijection h : C (f) → C (g) such that h ◦ f = g ◦ h on C (f)\{cf }. Furthermore, if g has no wandering intervals, no intervals of periodic points and no attracting periodic points, then f and g are topologically semiconju- gate.

i Proof. Let Cn(f) = {x ∈ I; f (x) = cf for some 0 ≤ i ≤ n − 1}, n ∈ N. Note that [ C (f) = Cn(f). n≥1 −n Since Cn+1(f) = Cn(f) ∪ f (cf ), we have that Cn(f) ⊂ Cn+1(f). Consider f f f Pn(f) = {Ii ⊂ I ; ∂Ii ⊂ Cn(f) ∪ {a, b} and 1 ≤ i ≤ kn}, where the f increasing order of the index of Ii is the same order as the intervals are f f placed in I and kn := #Pn(f). By definition, Ii is a maximal monotonicity n f closed interval of f for all 1 ≤ i ≤ kn.

Denote f| and f| by f−, f+ respectively. Furthermore, given a [a,cf ] [cf ,b] m sequence j = j1j2 . . . jm ∈ {−, +} we define fj := fjm ◦ · · · ◦ fj2 ◦ fj1 . Take k x ∈ Cn(f)\{cf }. Consider j(x) = j1j2 . . . jk ∈ {−, +} , with 1 ≤ k ≤ n − 1, the minimal sequence such that fj(x)(x) = cf . We will denote x by xj(x). f i Note that for each Ii ∈ Pn(f) there exists a unique sequence j = i i i n f j1j2 . . . jn ∈ {−, +} such that fji is strictly monotone on Ii . Thus we get

f ∂I ∈ {a, b, cf , x i i i }, i j1j2...jk for some 1 ≤ k ≤ n − 1. For each n ∈ N we set

k An(f) = {j1j2 . . . jk ∈ {−, +} ; ∃ x ∈ Cn(f)\{cf } such that x = xj1j2...jk

for some 1 ≤ k ≤ n − 1}.

In the same way we can define Cn(g), Pn(g) and An(g). Suppose that f(cf ) ≤ cf , then C (f) = Cn(f) = {cf } for all n ∈ N. Since if (cf ) = ig(cg) the same happen with C (g) and the proof is complete. Note that this case does not happen if g has no wandering intervals, no intervals of periodic points and no periodic attractors.

IMPA 7 July, 2018 Ermerson Rocha Araujo Kneading Sequences

From now on we make the assumption that f(cf ) > cf . Since if (cf ) = ig(cg) we have that g(cg) > cg. So C1(f) = {cf }, C1(g) = {cg}, C2(f) = f f g g f g {x−, cf , x+} and C2(g) = {x−, cg, x+}, where fj(xj ) = cf and gj(xj ) = cg, f f g g with j = −, +. By construction, x− < cf < x+ and x− < cg < x+. See Figure

2.1. This implies that there exist hi : Ci(f) → Ci(g), i = 1, 2, such that hi is a strictly increasing bijection and h2| = h1. Namely h1(cf ) := cg, C1(f) f g h2(cf ) := cg and h2(xj ) := xj , with j = −, +.

f g h2 cf cg

f cf f g cg g x− x+ x− x+

f f g Figure 2.1: Construction of x− and x+. Similarly we can construct x− and g x+.

We now proceed by induction on n. Let us suppose that there exists hn : Cn(f) → Cn(g) strictly increasing bijection such that hn ◦ f = g ◦ hn on Cn(f)\{cf } and hn| = hn−1. This implies that #Pn(f) = #Pn(g) Cn−1(f) and An(f) = An(g). f n Let Ii ∈ Pn(f) and j1j2 . . . jn ∈ {−, +} such that

If = [xf , xf ], i j1j2...jk j1j2...jl with 1 ≤ k 6= l ≤ n−1. By the manner that we order the intervals on Pn(f) and Pn(g) we have

h (xf ) = xg and h (xf ) = xg , n j1j2...jk j1j2...jk n j1j2...jl j1j2...jl where Ig = [xg , xg ]. i j1j2...jk j1j2...jl Furthermore,

f n−k n−l fjn ◦· · ·◦fj1 (Ii ) = [fjn ◦· · ·◦fjk+1 (cf ), fjn ◦· · ·◦fjl+1 (cf )] = [f (cf ), f (cf )]

IMPA 8 July, 2018 Ermerson Rocha Araujo Kneading Sequences and

g n−k n−l gjn ◦· · ·◦gj1 (Ii ) = [gjn ◦· · ·◦gjk+1 (cg), gjn ◦· · ·◦gjl+1 (cg)] = [g (cg), g (cg)].

Since if (cf ) = ig(cg), we get

n−k n−l n−k n−l [f (cf ), f (cf )] ∩ {cf }= 6 ∅ if and only if [g (cg), g (cg)] ∩ {cg}= 6 ∅.

n−k n−l f If [f (cf ), f (cf )] ∩ {cf } = ∅, then Ii ∈ Pn+1(f) and, consequently, g n−k n−l Ii ∈ Pn+1(g) too. If {f (cf ), f (cf )} ∩ {cf } 6= ∅, then either cf = n−k n−l f g f (cf ) or cf = f (cf ). Thus Ii ∈ Pn+1(f). Hence, we also have Ii ∈ n−k n−l Pn+1(g). On the other hand, if (f (cf ), f (cf )) ∩ {cf }= 6 ∅ then there exists a unique

xf := f −1 ◦ f −1 ◦ · · · ◦ f −1(c ) ∈ int(If ) ∩ f −n(c ). j1j2...jn j1 j2 jn f i f Hence, there also exists a unique

xg := g−1 ◦ g−1 ◦ · · · ◦ g−1(c ) ∈ int(Ig) ∩ g−n(c ). j1j2...jn j1 j2 jn g i g So we can define h (xf ) := xg , where xf ∈ (f)\ (f) n+1 j1j2...jn j1j2...jn j1j2...jn Cn+1 Cn and xg ∈ (g)\ (g). Moreover, if x ∈ (f), then we put h (x) = j1j2...jn Cn+1 Cn Cn n+1 hn(x). f f f The cases where Ii = [a, xj1j2...jk ], Ii = [xj1j2...jk , b], or Ii = [xj1j2...jk , cf ] are similar. Thus we have that hn+1 : Cn+1(f) → Cn+1(g) is a strictly increasing bijection such that hn+1 ◦ f = hn+1 ◦ g on Cn+1(f)\{cf } and hn+1 = hn. |Cn(f) Therefore, we define

h : C (f) −→ C (g)

y 7−→ h(y) = hn(y), where n is such that y ∈ Cn(f). Note that h ◦ f = g ◦ h on C (f)\{cf } Now, we will assume that g has no wandering intervals, no intervals of periodic points and no attracting periodic points. We claim that we can extend continuously h to C (f). In fact, take y ∈ C (f)\C (f). Suppose that j 1 2 there exist yn ∈ C (f), j = 1, 2, such that yn ↑ y and yn ↓ y. Without loss of j generality, we can assume that yn ∈ Cn(f) for all n ≥ 1, with j = 1, 2. From 0 1 2 1 1 the definition of the hns, it follows that hn(yn) < hn(yn), hn(yn) < hn+1(yn+1) 2 2 and hn(yn) > hn+1(yn+1) for all n ≥ 1. Since hn is strictly increasing there are unique 1 1 2 2 h (y) := lim hn(yn) and h (y) := lim hn(yn). n n

IMPA 9 July, 2018 Ermerson Rocha Araujo Kneading Sequences

1 2 j j Note that h (y) ≤ h (y). Besides that, h (y) do not depend of the sequence yn which converges to y. Now, it is enough to show that h1(y) = h2(y). Assume that h1(y) < h2(y). Let J := [h1(y), h2(y)]. Observe that J ∩ C (g) = ∅. Indeed, suppose that there exists zg ∈ J ∩ C (g). Consequently

1 g 2 hn(yn) < z < hn(yn) for all n ≥ 1.

f f g Let n0 ≥ 1 and z ∈ C (f) such that hn(z ) = z for all n ≥ n0. Hence 1 f 2 f yn < z < yn for all n ≥ n0. Thus we get z = y, which gives an absurd. Since g has no wandering intervals, there exist 1 ≤ l1 < l2 such that gl1 (J) ∩ gl2 (J) 6= ∅. As J ∩ C (g) = ∅, we may suppose, without loss of generality, that l1 = 0, l2 = l and l ≥ 1 is the smallest integer with such property. Put [ L = gml(J). m≥0 l It follows that L is a non-empty interval that contains no {cg} and g (L) ⊂ L. So gl is strictly monotone in L. Thus either L contains an interval of periodic points for gl, or some open interval in L converges to a single periodic point, which is a contradiction with the hypotheses on g. Therefore h1(y) = h2(y).

The cases where we have only yn ↑ y or yn ↓ y, with yn ∈ C (f) are similar.

Hence, if y ∈ C (f)\C (f) and yn → y, with yn ∈ C (f), then

h(y) := lim hn(yn) n defines a continuous extension of h to C (f). By a similar argument, if g has no wandering intervals, no intervals of periodic points and no attracting periodic points, then we also have that C (g) = I. Whence h : C (f) → I is a strictly increasing surjective map. From this, if W = (α, β) is a connected component of I\C (f), then we can extend h as h(z) := h(α) for all z ∈ W , once h(α) = h(β).

Claim: h ◦ f = g ◦ h.

It is clear that h ◦ f = g ◦ h on C (f). Note that for all W ⊂ I\C (f) 0 n 0 and n ≥ 0 there is Wn ⊂ I\C (f) such that f (W ) ⊂ Wn. As f is injective n−1 n 0 on f (W ), we get f (W ) = Wn. Take now z ∈ (α, β) ⊂ I\C (f). Thus h(f(z)) = h(f(α)) = g(h(α)) = g(h(z)). This proves the claim.

IMPA 10 July, 2018 Ermerson Rocha Araujo Kneading Sequences

Therefore, h is a semiconjugacy on I between f and g. 

The following corollary is immediate.

Corollary 2.1. Suppose that f, g : I → I are two unimodal maps with turning points cf and cg. Assume that f and g have no wandering intervals, no intervals of periodic points and no attracting periodic points. If if (cf ) = ig(cg), then f and g are topologically conjugate.

2.2 The discontinuous case

In this section we prove a theorem analogue to Theorem 2.1 for discontinuous unimodal maps. In particular, we extend these maps to a continuous map. This argument will be useful to our main results. Let I = [a, b] be a compact interval and c ∈ (a, b). Let f : I\{c} → I be a map such that f|[a,c) is strictly increasing, f|(c,b] is strictly decreasing and f({a, b}) = a. We call a map like this discontinuous unimodal map. See Figure 2.2. We extend f to c by adding two points c− and c+ and taking f(c−) := lim f(x) and f(c+) := lim f(x). x%c x&c Let f− and f+ be two maps defined by

− − + f− :[a, c ] −→ [−1, c ] ∪ (c , b] x 7−→ f(x) and + − + f+ :[c , b] −→ [−1, c ) ∪ [c , b] x 7−→ f(x).

f− f− f− f+

f+

f+

Figure 2.2: Some examples of discontinuous unimodal maps: (a), (b) and (c).

IMPA 11 July, 2018 Ermerson Rocha Araujo Kneading Sequences

From now on f will be the map with c± glued into its domain. We say that f is continuous if the functions f− and f+ are continuous. For this, we need define a topology into [a, c−] ∪ [c+, b].

2.2.1 One-side topology

First, let us define a usual neighborhood of a point x ∈ [a, c−] ∪ [c+, b]. If x ∈ [a, c−) ∪ (c+, b], then a neighborhood of x is the standard Euclidean neighborhood. On the other hand, if x = c± then a neighborhood of x is obtained from the intersection of [a, c] or [c, b] with a standard Euclidean neighborhood of x. This we call Euclidean Topology on [a, c−] ∪ [c+, b]. Now we will refine that neighborhood basis in each point. Set

C (f) = {x ∈ [a, c−] ∪ [c+, b]; f l(x) ∈ {c−, c+} for some l ≥ 0}.

For x∈ / C (f) we do not add any open neighborhood. Take now x ∈ C (f)\{c−, c+} and l > 0 such that f l(x) ∈ {c−, c+}. Suppose that f l(x) = c− (The other case is similar). Let ε > 0. Then there is δ > 0 such that ei- ther (x − δ, x] satisfies f l((x − δ, x]) ⊂ (c− − ε, c−] or [x, x + δ) satisfies f l([x, x+δ)) ⊂ (c− −ε, c−]. On the first case, we add the set {(x−δ, x]; δ > 0} to the neighborhood base of x and on the second case we add the set {[x, x + δ); δ > 0}.

Let Bf be the family of sets constructed above. Note that Bf is a basis − + for a topology on [a, c ] ∪ [c , b]. We call One-side topology the topology Tf generated by Bf and the euclidean topology. So f is continuous with this topology.

Remark 1. In some cases, we get ask that a point should be itself an open set. That happen, for example, in the case (b) on Figure 2.2 where the point c+ is an open set on the one-side topology.

2.2.2 Kneading sequence for discontinuous unimodal maps

Let f be a (dis)continuous unimodal map under the topological space ([a, c−]∪ + [c , b], Tf ). − + N − Consider the space Σf = {L, c , c ,R} , where L = [a, c ) and R = (c+, b]. So for each x ∈ I,its itinerary is defined by

if (x) = (i0(x), i1(x), . . . , in(x),...) ∈ Σf ,

IMPA 12 July, 2018 Ermerson Rocha Araujo Kneading Sequences

n n ± where in(x) = L if f (x) ∈ L, in(x) = R if f (x) ∈ R and in(x) = c if f n(x) = c±. Now, the kneading invariants of f are the sequences

± ± ± ± if (c ) = (i0(c ), i1(c ), . . . , in(c ),...).

The proof of the theorem below is similar to Theorem 2.1. So, it will be omitted.

Theorem 1. Let f and g be two discontinuous unimodal maps. Assume that g has no wandering intervals, no intervals of periodic points and no ± ± attracting periodic points. If if (cf ) = ig(cg ), then f and g are topologically semiconjugate.

IMPA 13 July, 2018 CHAPTER 3

Toy models and kneading sequences

Once the conjugacy class of one-dimensional maps is well understood, it is natural to look for approaches analogous for higher dimensional maps. In this chapter we define the class of maps with which we shall to work and its kneading sequences. Following [MMP13], we study a two-dimensional class of maps defined as follows. Consider a one parameter family

f(y):[−1, 1] → [−1, 1], with y ∈ [0, 1], depending continuously on y, such that f(y) is a unimodal map verifying that 0 is the turning point and f(y)(−1) = f(y)(1) = −1 for all y ∈ [0, 1]. Let k : [0, a] ∪ [b, 1] → [0, 1] be a differentiable function such that k(0) = k(1) = 0, k(a) = k(b) = 1 and |k0| > γ > 1. We put ( K (y), if x > 0 K(x, y) = + K−(y), if x < 0,

−1 −1 where K+ = (k|[0,a] ) and K− = (k|[b,1] ) .

We will study the map F , called Toy Model, defined by

F : ([−1, 1]\{0}) × [0, 1] −→ [−1, 1] × [0, 1]

(x, y) 7−→ (f(y)(x),Ksign(x)(y)).

14 Ermerson Rocha Araujo Kneading Sequences

1 1 F b a 0 0 −1 0 1 −1 0 1

Figure 3.1: Dynamics of F .

Given a toy model F and its Cantor map k : [0, a] ∪ [b, 1] → [0, 1] we can consider the Cantor set, denoted by K F , induced by k, that is, \ K F = k−n([0, a] ∪ [b, 1]). n≥0

In order to make the domain of F compact we introduce the points (0±, y) ± and extend F to them via the formula F (0 , y) = (f(y)(0),K±(y)). These points will be called the turning points of the Toy Model F and the set

± Lc(F ) = {(0 , y); y ∈ [0, 1]} is called the critical line of F . Let us define the topology on

Dom(F ) := (([−1, 1]\{0}) × [0, 1]) ∪ Lc(F ).

If (x, y) ∈ Dom(F )\Lc(F ), then a neighborhood of (x, y) is the standard Eu- clidean neighborhood. On the other hand, if (x, y) ∈ Lc(F ) then a neighbor- hood of (x, y) is obtained from the intersection of [−1, 0]×[0, 1] or [0, 1]×[0, 1] with a standard Euclidean neighborhood of (x, y).

Remark 2. Notice that F is now defined on a compact set, but we have to pay a price. The map F is not continuous.

3.1 Notations and preliminaries

In this section we establish some notations and define some notions that will be used throughout the Thesis. For each y ∈ [0, 1] we will use the notation f (y) and f (y) for f(y) , − + |[−1,0−] f(y) respectively. See Figure 3.2. |[0+,1]

IMPA 15 July, 2018 Ermerson Rocha Araujo Kneading Sequences

1

f−(y) f+(y) 0

−1 −1 0 1

Figure 3.2: Dynamics of f(y).

Consider the maps F− and F+ defined as follow

− − + F− :[−1, 0 ] × [0, 1] −→ ([−1, 0 ] ∪ (0 , 1]) × [0, 1]

(x, y) 7−→ (f−(y)(x),K−(y)).

F−

Figure 3.3: Dynamics of F−.

and

+ − + F+ : [0 , 1] × [0, 1] −→ ([−1, 0 ) ∪ [0 , 1]) × [0, 1]

(x, y) 7−→ (f+(y)(x),K+(y)).

F+

Figure 3.4: Dynamics of F+.

This allows us to see the orbit of a point (x, y) ∈ Dom(F ) by compositions of the functions F±. More specifically, for each (x, y) ∈ Dom(F ) there is a N sequence j(x, y) = (j1j2 ··· jm ··· ) ∈ {−, +} such that

m+1 F (x, y) = Fjm+1 (xm, ym)

IMPA 16 July, 2018 Ermerson Rocha Araujo Kneading Sequences

for all m ≥ 0, where xm = fjm (yjm−1···j1 ) ◦ · · · ◦ fj2 (yj1 ) ◦ fj1 (y)(x) and ym =

Kjm ◦ · · · ◦ Kj1 (y) := yjm···j1 . We denote the set ([−1, 0−] ∪ [0+, 1]) × {y} by I(y). In addition, a subset − + J ⊂ I(y) is called an interval if there are x1, x2 ∈ [−1, 0 ](or [0 , 1]) such that − + J = {(x, y) ∈ [−1, 0 ] ∪ [0 , 1] × {y} ; x1 < x < x2}.

Sometimes we will denote an interval J by [(x1, x2), y]. If

− + J = {(x, y) ∈ [−1, 0 ] ∪ [0 , 1] × {y} ; x1 ≤ x ≤ x2}, then we denote it by [[x1, x2], y]. Let (p, q) be a periodic point of F and set

B(p, q) = {(x, y); F l(x, y) → O(p, q) as l → ∞}.

We say that (p, q) is an weakly attracting periodic point if B(p, q) contains an interval. If B(p, q) contains an open subset, then we say that (p, q) is a strong attracting periodic point. When we have OF (p, q)∩Lc(F ) = ∅, then the immediate basin B0(p, q) of O(p, q) is the union of the connected components of B(p, q) which contain points from {(p, q),F (p, q),...,F n−1(p, q)} so that

B0(p, q) ∩ Lc(F ) = ∅, where n is the period of (p, q). In this case, there is a n sequence j1 ··· jn ∈ {−, +} such that for each (x, y) ∈ B0(p, q) we get  fjm (yjm−1···j1 ) ◦ · · · ◦ fj2 (yj1 ) ◦ fj1 (y)(x), yjm···j1 ∈ B0(p, q), for all 0 ≤ m ≤ n − 1.

Furthermore, if B(p, q) ⊂ B0(p, q) is the component containing (p, q) then  fjn (yjn−1···j1 ) ◦ · · · ◦ fj2 (yj1 ) ◦ fj1 (y)(x), yjn···j1 ∈ U(p, q), for all (x, y) ∈ B(p, q). Note that if (p, q) is a strong attracting periodic point, then (p, q) is a weakly attracting periodic point. Given y ∈ [0, 1], an interval J ⊂ I(y) is called interval of periodic points if all (x, y) ∈ J are periodic points for F .

Definition 1. Let F be a toy model and let J ∈ Dom(F ) be an interval. We say that J is wandering if

1. F n(J) ∩ F m(J) = ∅ for all n 6= m;

n 2. For all n ≥ 0 we get F (J) ∩ Lc(F ) = ∅.

IMPA 17 July, 2018 Ermerson Rocha Araujo Kneading Sequences

It is well known that if f : I → I is an unimodal map and f has no intervals of periodic points and no attracting periodic points, then the set C (f) is dense on the non-wandering set of f. The same holds for toy models. For this, let Ω(F ) be the non-wandering set of F and set

 l C (F ) = (x, y) ∈ Dom(F ); ∃ l ≥ 0 such that F (x, y) ∈ Lc(F ) .

We get

Proposition 1. Let F be a toy model and let (x, y) ∈ Ω(F ). Suppose that F has no weakly attracting periodic point and no interval of periodic points. Then (x, y) ∈ C (F ).

Proof. Take (x, y) ∈ Ω(F ) and suppose that (x, y) ∈/ C (F ). Hence, there n exists V neighborhood of (x, y) such that F (V ) ∩ Lc(F ) = ∅ for all n ≥ 0. We will assume that V is of the form I × J, where x ∈ I and y ∈ J and I,J are open intervals. Since (x, y) ∈ Ω(F ), there exists m ≥ 1 such that F m(V ) ∩ V 6= ∅. Put [ Lb = F n(V ). n≥0 S jm+r From choice of m, it follows that j≥0 F (V ) are connected sets for all r = 0, . . . , m−1. This forces that Lb has finitely many connected components. Set Lb = L1 t · · · t Ls, where L1,...,Ls are connected. Since F (Lb) ⊂ Lb, we get for all i there is j such that F (Li) ⊂ Lj. Let L be the connected component which contains (x, y). Thus we have F m(L) ∩ L 6= ∅ and conse- quently F m(L) ⊂ L. Consider l = min{d ≥ 1; F d(L) ⊂ L}. This implies l i that L is a domain and F (L) ⊂ L. In addition, F (L) ∩ Lc(F ) = ∅ and F i(L) ∩ F j(L) = ∅ for all i, j ∈ {0, . . . l − 1} with i 6= j. Thus, there exists l a sequence j0j1 ··· jl−1 ∈ {−, +} such that, for any (a, b) ∈ L and n ≥ 0 we nl n have F (a, b) = (anl, b(jl−1···j1j0) ), where

n n b(jl−1···j1j0) = Kjl−1 ◦ · · · ◦ Kj0 (b) and

n−1 anl = fjl−1 (bjl−2···j0(jl−1···j0) ) ◦ · · · ◦ fjl−1 (bjl−2···j0 ) ◦ · · · ◦ fj1 (bj0 ) ◦ fj0 (b)(a).

Let π2 be the projection on the second coordinate. Since Kjl−1 ◦ · · · ◦ Kj0 is a contraction, there exists w ∈ π2(L) such that Kjl−1 ◦ · · · ◦ Kj0 (w) = w. We claim that w = y. In fact, for each n ≥ 1 take Vn := In × Jn & I × J

IMPA 18 July, 2018 Ermerson Rocha Araujo Kneading Sequences

neighborhood of (x, y) such that In+1 × Jn+1 & In × Jn with |In| → 0 and |Jn| → 0 as n → ∞. As before, for each n ≥ 1 there are (xn, yn) ∈ Vn and kn kn kn ≥ 1 such that F (xn, yn) ∈ Vn and F (xn, yn) → (x, y) as n → ∞. We may suppose that kn → ∞ as n → ∞. Furthermore, by construction of L, we have kn = rnl for some rn ≥ 1. As K := Kjl−1 ◦ · · · ◦ Kj0 is a contraction, there exists λ ≤ 1 such that |K(x) − K(y)| ≤ λ|x − y|, for all x, y ∈ [0, 1]. Note that

rn rn rn rn |K (yn) − w| ≤ |K (yn) − K (y)| + |K (y) − w|

rn rn ≤ λ |yn − y| + |K (y) − w| → 0. Therefore kn rn π2(F (xn, yn)) = K (yn) → w.

kn As F (xn, yn) → (x, y), we have that w = y. This proves the claim. For every (v, y) ∈ L ∩ I(y) and n ≥ 1 we get

ml m  F (v, y) = f(jl−1···j0)(y) (v), y ∈ L ∩ I(y), where f(jl−1···j0)(y) := fjl−1 (yjl−2···j0 ) ◦ · · · ◦ fj0 (y) is a strictly monotone map.

Without loss of generality, assume that f(jl−1···j0)(y) is strictly increasing. Let

{Ii}i be the connected components of L ∩ I(y).

l Claim: There exists i0 such that F (Ii0 ) ⊂ Ii0 .

Assume that the claim does not hold. Since f(jl−1···j0)(y) is strictly increas- l ing, for every i there exists σ(i) such that F (Ii) ⊂ Iσ(i) and sup Ii < inf Iσ(i). ln Fixed any k, this implies that the sequence {F (Ik)}n≥0 is formed by dis- ln l(n+1) ln joint intervals satisfying sup F (Ik) < inf F (Ik) and |F (Ik)| → 0 as l n goes to infinity. Thus, there is (v0, y) such that F (v0, y) = (v0, y) and ln F (Ik) → (v0, y), contradicting the nonexistence of weakly attracting peri- odic point. This proves the claim. l Let i0 given by claim. So, F : Ii0 → Ii0 is a strictly increasing interval l map. Therefore, either Ii0 contains an interval of periodic points for F , or some open interval in Ii0 converges to a single periodic point, which is again a contradiction with the hypotheses on F . The proof of the proposition is finished. 

IMPA 19 July, 2018 Ermerson Rocha Araujo Kneading Sequences

3.2 Kneading sequences and main results

In this section we define sequences of symbols and show that these sequences completely determine the combinatorial type of the toy models.

Let π1 be the projection on first coordinate and let F be a toy model. − + Consider the alphabet A = {L, 0 , 0 ,R} the address tF (x, y) of a point (x, y) ∈ Dom(F ) is defined by   L if π1(x, y) < 0  − −  0 if π1(x, y) = 0 tF (x, y) = + +  0 if π1(x, y) = 0   R if π1(x, y) > 0 Applying this map to an orbit of a given point (x, y) ∈ Dom(F ), we associate to that orbit one sequence of symbols.

Definition 2. Consider the sequence of symbols in A

n TF (x, y) = (tF (x, y), tF (F (x, y)), . . . , tF (F (x, y)),...).

This infinite sequence is called itinerary of (x, y) ∈ Dom(F ).

This definition allows us to define the map

TF : Dom(F ) −→ Σ

(x, y) 7−→ TF (x, y), where Σ := {L, 0−, 0+,R}N. Let Φ:Σ → Σ be the one-side shift, we get

TF ◦ F = Φ ◦ TF .

We call Kneading Sequences of F , denoted by KF (Lc), the set of all sequences associated to elements of Lc(F ). So, it is reasonable to ask whether the Theorem 2.1 still holds for Toy Models family. Indeed, the Theorem A answers positively this question. The following proposition is clear, but we will present a proof for com- pleteness.

F Proposition 2. Let F (x, y) = (f(y)(x),Ksign(x)(y)) and G(x, y) = (g(y)(x), G Ksign(x)(y)) be two toy models. There exists a strictly increasing continuous map ψ : [0, 1] → [0, 1] such that

F G ψ ◦ K± = K± ◦ ψ.

Moreover, ψ|K F is independent of choice of ψ.

IMPA 20 July, 2018 Ermerson Rocha Araujo Kneading Sequences

Proof. Let kF : [0, aF ] ∪ [bF , 1] → [0, 1] and kG : [0, aG] ∪ [bG, 1] → [0, 1] the Cantor maps of F and G. If K F and K G are the Cantor sets generated by kF , kG respectively, then there is ψ : K F → K G strictly increasing map such that ψ◦kF = kG ◦ψ. Now we will extend ψ continuously to gaps of K F . Take any strictly increasing homeomorphism h :(aF , bF ) → (aG, bG). Denote F,G F,G F,G F,G k and k by k+ , k− respectively. Observe that given a gap |[0,aF,G] |[bF,G,1] F n J ⊂ [0, 1] of K there exist unique n ≥ 1 and a sequence j1 ··· jn ∈ {−, +} F F F F such that kjn ◦ · · · ◦ kj1 (J) = (a , b ). Thus, for each x ∈ J we define

G −1 G −1 F F ψ(x) := (kj1 ) ◦ · · · ◦ (kjn ) ◦ h ◦ kjn ◦ · · · ◦ kj1 (x). It is clear that ψ : [0, 1] → [0, 1] is a strictly increasing continuous map F G which ψ|K F is independent of choice of ψ and ψ ◦ K± = K± ◦ ψ. 

The following definition gives a notion of equality between two kneading sequences.

F Definition 3. Let F (x, y) = (f(y)(x),Ksign(x)(y)) and G(x, y) = (g(y)(x), G Ksign(x)(y)) be two toy models and let ψ : [0, 1] → [0, 1] be the map con- structed on the Proposition 2. We say that KF (Lc) = KG(Lc) if

± ± TF (0 , y) = TG(0 , ψ(y)) for all y ∈ [0, 1].

We are now ready to state the main step on the proof of Theorem A.

Theorem 2. Let F and G be two toy models with kneading sequences KF (Lc) and KG(Lc). Assume that KF (Lc) = KG(Lc). Then there exists a bijective map H : C (F ) → C (G) such that H ◦ F = G ◦ H on C (F )\Lc(F ) and

H(C (F ) ∩ I(y)) ⊂ C (G) ∩ I(ψ(y)) for each y ∈ [0, 1]. Moreover, H|C (F )∩I(y) is strictly increasing for every y ∈ [0, 1].

The proof of this theorem will be given in the next chapter. In order to classify toy models, we need a notion of equivalence between them.

F Definition 4. Let F (x, y) = (f(y)(x),Ksign(x)(y)) and G(x, y) = (g(y)(x), G Ksign(x)(y)) be two toy models. We say that F and G are topologically conjugate if there is a homeomorphism which is strictly increasing on inter- vals H : Dom(F ) → Dom(G),

IMPA 21 July, 2018 Ermerson Rocha Araujo Kneading Sequences such that H ◦ F = G ◦ H and H(I(y)) ⊂ I(ψ(y)), where ψ : [0, 1] → [0, 1] is the map constructed on the Proposition 2.

Note that if H conjugates F and G, then H(Lc(F )) = Lc(G), that is, H(0±, y) = (0±, ψ(y)). The proposition below is an immediate consequence of the Definition 4 and implies that Kneading Sequences are a topological invariant.

Proposition 3. Let F and G be two toy models. Suppose that F and G are topologically conjugate. Then KF (Lc) = KG(Lc).

± ± Proof. We will show that for all y ∈ [0, 1] we have TF (0 , y) = TG(0 , ψ(y)). ± ± Fix y ∈ [0, 1]. It is clear that tF (0 , y) = tG(0 , ψ(y)). Now, as F and G are conjugate we get H(F n(0±, y)) = Gn(0±, ψ(y)) for all n ≥ 1. Since H is strictly increasing on intervals, we have

n ± n ± tF (F (0 , y)) = tG(G (0 , ψ(y)))

± ± for all n ≥ 1. Therefore, TF (0 , y) = TG(0 , ψ(y)). 

Remark 3. Notice that the Proposition above and the Theorem A imply that the Kneading Sequences are a ‘complete’ topological invariant.

Definition 5. Let F be a toy model and let V ⊂ Dom(F ) be a domain. We say that V is wandering if

1. F n(V ) ∩ F m(V ) = ∅ for all n 6= m;

n 2. For all n ≥ 0 we have F (V ) ∩ Lc(F ) = ∅.

It is not clear that if we change the hypothesis of nonexistence of wander- ing intervals by nonexistence of wandering domains the Theorem A would still be true. On the other hand, if we add a technical hypothesis we can give another version of Theorem A. Let F be a toy model. For each y ∈ [0, 1] and n ≥ 1 we define

F  l Cn (y) := (x, y) ∈ Dom(F ) ∩ I(y); F (x, y) ∈ Lc(F ) for some 0 ≤ l ≤ n − 1 .

F Dom(F ) F F Consider the map φn : [0, 1] → 2 defined by φn (y) := Cn (y). We can state another formulation of the Theorem A.

IMPA 22 July, 2018 Ermerson Rocha Araujo Kneading Sequences

Theorem C. Let F and G be two toy models. Assume that F and G G have no wandering domain and KF (Lc) = KG(Lc). If the family {φn } is equicontinuous, then F and G are topologically semiconjugate.

Remark 4. Here, the equicontinuity means that for all ε > 0 there is δ > 0 G G such that |y1 − y2| < δ implies dH(Cn (y1), Cn (y2)) < ε for all n ≥ 1, where dH is the Hausdorff distance.

G Remark 5. Note that if G has no wandering domain and {φn } is equicon- tinuous, then G has no wandering intervals. Whence the Theorem C follows from Theorem A. However, we will give an independent proof.

The proof of Theorem C will be given in the next chapter.

Example 1. Let f :[−1, 1] → [−1, 1] be the tend map, that is, f is defined by f(x) = 1 + 2x if x ≤ 0 and f(x) = 1 − 2x if x ≥ 0. Let k : [0, 1] → [0, 1] be the Cantor map defined by k(y) = 3y if y ∈ [0, 1/3] and k(y) = 3 − 3y if y ∈ [2/3, 1]. Take the toy model F (x, y) := (f(x),K(x, y)). Note that F satisfies both hypotheses of Theorem A and Theorem C.

Example 2. Let f :[−1, 1] → [−1, 1] be a unimodal map containing a wandering interval I0 and 0 as turning point. Let k : [0, 1] → [0, 1] be the Cantor map defined by k(y) = 3y if y ∈ [0, 1/3] and k(y) = 3 − 3y if y ∈ [2/3, 1]. Take the toy model F (x, y) := (f(x),K(x, y)). Note that F has a wandering interval, namely I0 × {y}.

Remark 6. It is clear that the construction of the toy model using unimodal maps has nothing special. So, in the Appendix A we build more general toy models.

IMPA 23 July, 2018 CHAPTER 4

Proofs of the main results

In this chapter, we present the proofs of our main results.

4.1 Proof of Theorem 2

As before

 l C (F ) = (x, y) ∈ Dom(F ); ∃ l ≥ 0 such that F (x, y) ∈ Lc(F ) and

F  l Cn (y) := (x, y) ∈ Dom(F ) ∩ I(y); F (x, y) ∈ Lc(F ) for some 0 ≤ l ≤ n − 1 , for each y ∈ [0, 1] and n ≥ 1. Set F [ F C (y) := Cn (y). n≥1

Hence

a C (F ) = C F (y). y∈[0,1]

24 Ermerson Rocha Araujo Kneading Sequences

G G Interchanging F and G we get C (G), Cn (y) and C (y) for each y ∈ [0, 1] and n ≥ 1. Suppose that for all y ∈ [0, 1] and n ≥ 1 there exists F G an strictly increasing bijection map Hn(y): Cn (y) → Cn (ψ(y)) such that H (y) ◦ F = G ◦ H (y) on F (y)\{(0±, y)} and H (y) = H (y). As n n Cn n | F n−1 Cn−1(y) ψ : [0, 1] → [0, 1] is a bijective map, it follows that a C (G) = C G(ψ(y)). y∈[0,1]

Therefore, we can define

H : C (F ) −→ C (G)

(x, y) 7−→ Hn(y)(x, y),

F where n is such that (x, y) ∈ Cn (y). Then H ◦ F = G ◦ H on C (F )\Lc(F ). Hence to prove theorem 2 it is enough construct the maps Hn(y). The proof follows by the same method as in the one-dimensional case. F F There exists a partition of I(y) induced by Cn (y), denoted by Pn (y). More specifically,

F  F F F F Pn (y) = Ii (y) ⊂ I(y); ∂Ii (y) ⊂ Cn (y) ∪ {(−1, y), (1, y)}, 1 ≤ i ≤ kn (y) ,

F where the increasing order of the index of Ii (y) is the same order as the F F intervals are placed in I(y) and kn (y) := #Pn (y). Note that the partition F + Pn (y) includes two intervals, one with left endpoint 0 and one with right endpoint 0−. m Given a sequence j = j1j2 . . . jm ∈ {−, +} we will use the following notation

fj(y) := fjm (yjm−1···j1 ) ◦ · · · ◦ fj3 (yj2j1 ) ◦ fj2 (yj1 ) ◦ fj1 (y),

where Kjl ◦ · · · ◦ Kj1 (y) := yjl···j1 , for all 1 ≤ l ≤ m. F ± Let (x, y) ∈ Cn (y)\{(0 , y)} and we will identify the pair (x, y) with F F x(y)j(x,y). Let us remark that we will abuse the notation of x(y)j(x,y) to denote when we either apply F , seeing (x, y) as a vector, or applying the unimodal maps to it, seeing (x, y) as a real number. Here, j(x, y) = j1j2 . . . jk ∈ {−, +}k is the minimal sequence such that

k F  F  jk  F x(y)j(x,y) = fj(x,y)(y)(x(y)j(x,y)), yjk···j1 = 0 , yjk···j1 , for some 1 ≤ k ≤ n − 1.

IMPA 25 July, 2018 Ermerson Rocha Araujo Kneading Sequences

F F More precisely, for each Ii (y) ∈ Pn (y) there is a unique sequence F i i i n j(Ii (y)) = j1j2 . . . jn ∈ {−, +} such that the function

F    (z, y) ∈ int I (y) 7→ f F (y)(z), y i i i j(Ii (y)) jn···j1 is strictly monotone. Thus we get

F  n ± F o ∂ Ii (y) ⊂ (−1, y), (1, y), (0 , y), x(y) i i i , j1j2...jk for some 1 ≤ k ≤ n − 1. So for each n ≥ 1 set

F k F ± An (y) = {j1j2 . . . jk ∈ {−, +} ; ∃ (x, y) ∈ Cn (y)\{(0 , y)} such that

(x, y) = x(y)F for some 1 ≤ k ≤ n − 1}. j1j2...jk G G It is clear that we can define Pn (ψ(y)) and An (ψ(y)) similarly. Now, we are able to finish the proof of the theorem. Our hypotheses imply that we F G ± ± can define H1(y): C1 (y) → C1 (ψ(y)) as (0 , y) 7→ (0 , ψ(y)). F G Suppose by induction that there exists Hn(y): Cn (y) → Cn (ψ(y)) a F ± strictly increasing map such that Hn(y)◦F = G◦Hn(y) on Cn (y)\{(0 , y)}, F G F for every y ∈ [0, 1]. This implies that #Pn (y) = #Pn (ψ(y)) and An (y) = G An (ψ(y)). F F n Let Ii (y) ∈ Pn (y). There is a unique sequence j = j1j2 ··· jn ∈ {−, +} such that ∂ IF (y) ⊂ (±1, y), (0±, y), x(y)F , i j1···jk for some 1 ≤ k ≤ n − 1. First suppose that

IF (y) = x(y)F , x(y)F  , y , i j1···jk j1···jl with 1 ≤ k 6= l ≤ n − 1. By the induction hypothesis we have

IG(ψ(y)) = x(ψ(y))G , x(ψ(y))G  , y . i j1···jk j1···jl

Moreover

H (y)(x(y)F ) = x(ψ(y))G and H (y)(x(y)F ) = x(ψ(y))G . n j1···jk j1···jk n j1···jl j1···jl

From the definition of F and G it follows that

n F  jk+1 F (Ii (y)) = fjn (yjn−1···j1 ) ◦ · · · ◦ fjk+1 (yjk···j1 )(0 ),

jl+1   F fjn (yjn−1···j1 ) ◦ · · · ◦ fjl+1 (yjl···j1 )(0 ) , yjn···j1 ∪ Λ(Ii (y))

IMPA 26 July, 2018 Ermerson Rocha Araujo Kneading Sequences

and

n G  jk+1 G (Ii (ψ(y))) = gjn (ψ(y)jn−1···j1 ) ◦ · · · ◦ gjk+1 (ψ(y)jk···j1 )(0 ),

jl+1   gjn (ψ(y)jn−1···j1 ) ◦ · · · ◦ gjl+1 (ψ(y)jl···j1 )(0 ) , yjn···j1 G ∪ Λ(Ii (ψ(y))),

F n−k jk n−l jl G where Λ(Ii (y)) = {F (0 , yjk···j1 ),F (0 , yjl···j1 )} and Λ(Ii (ψ(y))) = n−k jk n−l jl {G (0 , ψ(y)jk···j1 ),G (0 , ψ(y)jl···j1 )}. Since KF (Lc) = KG(Lc), we get

n F F (Ii (y)) ∩ Lc(F ) 6= ∅ if, and only if n G G (Ii (ψ(y))) ∩ Lc(G) 6= ∅.

n F F F If F (Ii (y)) ∩ Lc(F ) = ∅, then Ii (y) ∈ Pn+1(y). On the other hand, if F F F Λ(Ii (y)) ∩ Lc(F ) 6= ∅ we get Ii (y) ∈ Pn+1(y), since F is a one-to-one map. G G In any case, we also have Ii (ψ(y)) ∈ Pn+1(ψ(y)). n F F  Now, if F (Ii (y))\Λ(Ii (y)) ∩ Lc(F ) 6= ∅ then there exists a unique

x(y)F := f −1(y) ◦ · · · ◦ f −1(y )(0jn ), y ∈ int IF (y) . j1j2···jn j1 jn jn−1···j1 i

n G G  Hence, G (Ii (ψ(y)))\Λ(Ii (ψ(y))) ∩ Lc(G) 6= ∅ and also there exists a unique x(ψ(y))G := g−1(ψ(y)) ◦ · · · ◦ g−1(ψ(y) )(0jn ), ψ(y) ∈ int IG(ψ(y)) . j1j2···jn j1 jn jn−1···j1 i

F G So we can define Hn+1(y)(x(y)j1j2···jn ) := x(ψ(y))j1j2···jn . Moreover, if F F (x, y) ∈ Cn (y), then we put Hn+1(x, y) = hn(x, y). The cases where Ii = [[−1, x(y)F ], y], IF = [[x(y)F , 1], y], or IF = [[x(y)F , 0±], y] are j1j2...jk i j1j2...jk i j1j2...jk similar. Thus we have

F G Hn+1(y): Cn+1(y) → Cn+1(ψ(y)) a strictly increasing map such that

Hn+1(y) ◦ F = G ◦ Hn+1(y)

F ± on Cn+1(y)\{(0 , y)}. Moreover, Hn+1(y)| F = Hn(y). This concludes the Cn (y) proof of theorem 2.

IMPA 27 July, 2018 Ermerson Rocha Araujo Kneading Sequences

4.2 Proof of Theorem A

We will first prove the following Theorem 3. Let F and G be two toy models. Assume that G has no wander- ing intervals, no interval of periodic points and no weakly attracting periodic points. If F and G have the same kneading sequences, then there is a con- tinuous surjective map H : Dom(F ) → Dom(G) such that H ◦ F = G ◦ H. We need the following lemma. Lemma 1. Let y ∈ [0, 1] 7→ f(y):[−1, 1] → [−1, 1] be the family on the definition of a toy model. For all y ∈ [0, 1] and ε > 0 there is δ > 0 satisfying: If y0 ∈ [0, 1], x, x0 ∈ [−1, 1] are such that d(y, y0) < δ, d(x, x0) < δ −1 0 0 −1 0 0 and there exist x(y) = fj (y)(x) and x (y ) = fj (y )(x ), where j = − or +, then d(x(y), x0(y0)) < ε. Proof. Suppose that there is ε > 0 such that for all n ≥ 1 there are

y yn y yn yn ∈ [0, 1] and xn, xn ∈ [−1, 1] with d(y, yn) < 1/n and d(xn, xn ) < −1 y −1 yn 1/n such that xn(y) = fj (y)(xn) and xn(yn) = fj (yn)(xn ) exist and y d(xn(y), xn(yn)) ≥ ε. By compactness, we may suppose that lim xn = yn lim xn = x0, lim xn(y) = x1 and lim xn(yn) = x2. Note that x1 6= x2 and x1, x2 ∈ [−1, 0] or x1, x2 ∈ [0, 1]. So f(y)(x1) 6= f(y)(x2). On the other hand, as f(yn) converges uniformly to f(y) we get f(y)(x1) = lim f(y)(xn(y)) = y yn lim xn = lim xn = lim f(yn)(xn(yn)) = f(y)(x2). This contradiction finishes the proof of the lemma. 

− + − − + Let ξ− : ([−1, 0 ]∪(0 , 1])×[0, 1] → [−1, 0 ] and ξ+ : ([−1, 0 )∪[0 , 1])× [0, 1] → [0+, 1] be two maps defined by ( f −1(y)(x) , if x ∈ Im(f (y)) ξ (x, y) = j j j 0j , otherwise. As a consequence of Lemma 1 we get the following.

Corollary 1. The maps ξj defined above are continuous.

j Proof. Let (x, y) ∈ Dom(ξj) and ε > 0. Assume that x ∈ [−1, fj(y)(0 )).

Since the family {f(y)}y∈[0,1] depends continuously on y, there exists δ > 0 such that for every x0 and y0 satisfying d(x, x0) < δ and d(y, y0) < δ, we have 0 0 j x ∈ [−1, fj(y )(0 )). From Lemma 1 it follows that

0 0 −1 −1 0 0 d(ξj(x, y), ξj(x , y )) = d(fj (y)(x), fj (y )(x )) < ε

IMPA 28 July, 2018 Ermerson Rocha Araujo Kneading Sequences if δ > 0 is small enough. j Suppose now that x > fj(y)(0 ). Again by continuity, there exists δ > 0 such that for all x0 and y0 satisfying d(x, x0) < δ and d(y, y0) < δ we have 0 0 j j 0 0 x > fj(y )(0 ). Therefore, ξj(x, y) = 0 = ξj(x , y ) and so

0 0 d(ξj(x, y), ξj(x , y )) = 0.

j The case where x = fj(y)(0 ) is analogous. 

The following lemma is the fundamental step in the proof of Theorem 3.

Lemma 2. For every (x, y) ∈ C (F ) there exists a continuous curve γF : F F F [0, 1] → C (F ) of the form γ (w) = (γe (w), w) such that it satisfies γ (y) = (x, y).

Proof. Given (x, y) ∈ C (F ) there exist n ≥ 1 and a sequence (j1 ··· jn) ∈ {−, +}n such that (x, y) = (f −1(y) ◦ · · · ◦ f −1(y )(0jn ), y). Consider the j1 jn jn−1···j1 following sequence of curves defined inductively:

F F j1 • γj1 : [0, 1] → Dom(F ) is defined by γj1 (w) = (ξj1 (0 , w), w);

F F F • γj1j2 : [0, 1] → Dom(F ) is defined by γj1j2 (w) = (ξj1 (π1(γj2 (wj1 )), w), w); . .

• γF : [0, 1] → Dom(F ) is defined by γF (w) = (γF (w), w), j1···jn j1···jn ej1···jn where γF (w) := ξ (π (γF (w )), w). ej1···jn j1 1 j2···jn j1 F It follows from Corollary 1 that γj1···jn : [0, 1] → Dom(F ) is a continuous F curve. Furthermore, by construction we have that γj1···jn (y) = (x, y). Now it F is quite easy to see that γj1···jn ([0, 1]) ⊂ C (F ). For each w ∈ [0, 1] put l (w) := max{1 ≤ k ≤ n; f −1(w) ◦ · · · ◦ f −1(w )(0jk ) exists}. j1···jn j1 jk jk−1···j1

Hence, if lj1···jn (w) exists, then  j  F −1 −1 lj ···j (w) γ (w) = f (w) ◦ · · · ◦ f (wj ···j )(0 1 n ), w . j1···jn j1 jl (w) lj ···j (w)−1 1 j1···jn 1 n

On the other hand, if lj1···jn (w) does not exist, then

F j1 γj1···jn (w) = (0 , w).

F So in any case we get γj1···jn ([0, 1]) ⊂ C (F ). 

IMPA 29 July, 2018 Ermerson Rocha Araujo Kneading Sequences

Remark 7. Suppose that (x, y), (z, w) ∈ C (F ) are such that there exist n ≥ 1 and a sequence (j ··· j ) ∈ {−, +}n so that (x, y) = (f −1(y) ◦ · · · ◦ 1 n j1 f −1(y )(0jn ), y) and (z, w) = (f −1(w)◦· · ·◦f −1(w )(0jn ), w). The jn jn−1···j1 j1 jn jn−1···j1 Lemma 2 implies that there exists a continuous curve γF : [0, 1] → C (F ) satisfying γF (y) = (x, y) and γF (w) = (z, w).

Now we are able to prove Theorem 3.

Proof of Theorem 3. From the proof of Theorem 2, for every y ∈ [0, 1], let F G H(y): C (y) → C (ψ(y)) be the map defined by H(y)(x, y) := Hn(y)(x, y), F where n is such that (x, y) ∈ Cn (y). Like in the one-dimensional case, we will extend H(y) to I(y). Take (z, y) ∈ C F (y)\C F (y) and suppose j F 1 that there exist (zn, y) ∈ C (y), j = 1, 2, such that (zn, y) ↑ (z, y) 2 and (zn, y) ↓ (z, y). The cases where we have only (zn, y) ↑ (z, y) or F (zn, y) ↓ (z, y), with (zn, y) ∈ C (y) are similar. Since G has no wandering intervals, no intervals of periodic points and no weakly attracting periodic 1 1 2 2 point, the limits lim H((zn, y) ) := H (z, y) and lim H((zn, y) ) := H (z, y) exist and H1(z, y) = H2(z, y). Hence, if (z, y) ∈ C F (y)\C F (y) we define

H(z, y) := lim H(zn, y),

F where (zn, y) ∈ C (y) is any sequence such that (zn, y) → (z, y). As C G(ψ(y)) is dense on I(ψ(y)), for all W = [(α, β), y] connected component of I(y)\C F (y) and (z, y) ∈ W we can extend H(y) as H(y)(z, y) := H(y)(α, y), once H(y)(α, y) = H(y)(β, y). Let H : Dom(F ) → Dom(G) be the map defined by H(x, y) := H(y)(x, y). Note that H is non-decreasing on each fiber I(y).

Claim: H : Dom(F ) → Dom(G) is continuous.

Take (x, y) ∈ Dom(F )\{(±1, y), (0±, y); y ∈ [0, 1]}. If we get (x, y) ∈ ± {(±1, y), (0 , y); y ∈ [0, 1]} the argument is similar. Let Bε(H(x, y)) be a neighborhood of H(x, y) for some ε > 0. Since C G(ψ(y)) = I(ψ(y)), there are z1 < x < z2 such that (z1, y), (z2, y) ∈ C (F ) and H(zi, y) ∈

Bε(H(x, y)) ∩ I(ψ(y)). Note that π1(H(z1, y)) < π1(H(x, y)) < π1(H(z2, y)). From Lemma 2, it follows that there exist continuous curves γG and j1···jl γG such that γG (ψ(y)) = H(z , y) and γG (ψ(y)) = H(z , y). By e1···ek j1···jl 1 e1···ek 2 continuity, there are y1 < y < y2 so that the domain D(H(x, y)) lim- ited by curves γG ([ψ(y ), ψ(y )]), γG ([ψ(y ), ψ(y )]) and the intervals j1···jl 1 2 e1···ek 1 2

IMPA 30 July, 2018 Ermerson Rocha Araujo Kneading Sequences

[(γG (ψ(y )), γG (ψ(y ))), ψ(y )], [(γG (ψ(y )), γG (ψ(y ))), ψ(y )] con- j1···jl 1 e1···ek 1 1 j1···jl 2 e1···ek 2 2 tains H(x, y) and satisfies D(H(x, y)) ⊂ Bε(H(x, y)). Again by Lemma 2, m F for every sequence t1 ··· tm ∈ {−, +} and w ∈ [0, 1] we get H(γt1···tm (w)) = γG (ψ(w)). Let D(x, y) be the subset limited by curves γF ([y , y ]), t1···tm j1···jl 1 2 γF ([y , y ]) and the intervals [(γF (y ), γF (y )), y ], [(γF (y ), γF e1···ek 1 2 j1···jl 1 e1···ek 1 1 j1···jl 2 e1···ek (y2)), y2]. Note that D(x, y) contains a neighborhood of (x, y). As H is mono- tone on the fibers we have H(D(x, y)) = D(H(x, y)). See Figure 4.1. This proves the claim.

B (H(x, y)) D(x, y) D(H(x, y)) ε H (x, y) H(x, y)

Figure 4.1: Action of H on D(x, y).

The Claim implies that H is surjective. Now, take (z, y) ∈ C F (y)\C F (y) F ± and let (zn, y) ∈ C (y)\{0 , y} such that lim(zn, y) = (z, y). As F is contin- uous on Dom(F )\C (F ) and H(z, y) ∈ C G(ψ(y))\C G(ψ(y)) it follows that

(H ◦ F )(z, y) = lim(H ◦ F )(zn, yn)

= lim(G ◦ H)(zn, yn) = (G ◦ H)(z, y).

On the other hand, for all connected component W ⊂ I(y)\C F (y) and 0 n F n n ≥ 0 there is a connected component Wn ⊂ I(π2(F (y)))\C (π2(F (y))) n 0 n 0 such that F (W ) ⊂ Wn. As F is injective, we get F (W ) = Wn. Let [(α, β), y] be a connected component of I(y)\C F (y). Since (α, y) ∈ C F (y) we have H(F (α, y)) = G(H(α, y)). Thus

H(F (z, y)) = H(F (α, y)) = G(H(α, y)) = G(H(z, y)), for all (z, y) ∈ [(α, β), y]. This finishes the proof of Theorem. 

Theorem A is now an immediate consequence of Theorem 3. Proof of Theorem A. As, for all y ∈ [0, 1] we have C F (y) = I(y) and C G(y) = I(y), Theorem A follows from Theorem 3. 

IMPA 31 July, 2018 Ermerson Rocha Araujo Kneading Sequences

4.3 Proof of Theorem C

Let’s start with a lemma which proof is similar to the proof of Proposition 1.

Lemma 3. Let F be a toy model. Suppose that F has no wandering domain. Then C (F ) = Dom(F ).

m Proof. Take V ⊂ Dom(F ) open set such that F (V ) ∩ Lc(F ) = ∅ for all m ≥ 0. We will assume that V is of the form I × J, where I and J are open intervals. Since F has no wandering domain, there exist 1 ≤ m1 < m2 such that F m1 (V ) ∩ F m2 (V ) 6= ∅. By the choice of V , we may suppose, without loss of generality, that m1 = 0, m2 = m for some m ≥ 1 and m is the small- est integer with such property F m(V ) ∩ V 6= ∅. We now construct, as in the proof of Proposition 1, a connected set L containing V such that F l(L) ⊂ L i i j for some 0 < l ≤ m. In addition, F (L) ∩ Lc(F ) = ∅ and F (L) ∩ F (L) = ∅ for all i, j ∈ {0, . . . l − 1} with i 6= j. Therefore, there exists a sequence l j0j1 ··· jl−1 ∈ {−, +} such that, for any (a, b) ∈ L and n ≥ 0 we have nl n n n F (a, b) = (anl, b(jl−1···j1j0) ), where b(jl−1···j1j0) := Kjl−1 ◦ · · · ◦ Kj0 (b)

n−1 n−1 and anl = fjl−1 (bjl−2···j0(jl−1···j0) )◦· · ·◦fj0 (b(jl−1···j0) )◦· · ·◦fjl−1 (bjl−2···j0 )◦

· · · ◦ fj1 (bj0 ) ◦ fj0 (b)(a). Since Kjl−1 ◦ · · · ◦ Kj0 is a contraction, there is w ∈ π2(L) such that Kjl−1 ◦ · · · ◦ Kj0 (w) = w. For each (x, y) ∈ V , we can prove that y = w. Therefore V ⊂ I(w), which contradicts the fact that V is an open set. 

For all n ≥ 1 consider the sets

F  l Cn = (x, y) ∈ Dom(F ); ∃ 0 ≤ l ≤ n − 1 such that F (x, y) ∈ Lc(F ) .

F G For each y ∈ [0, 1], let Hn(y): Cn (y) → Cn (ψ(y)) be the map con- structed in the proof of Theorem 2. Let Hfn(y): I(y) → I(ψ(y)) be the piecewise linear homeomorphism such that

Hn(y)(±1, y) = (±1, ψ(y)), Hfn(y)| F = Hn(y) Cn (y)

F and Hfn(y) is linear in each connected component of I(y)\Cn (y). Now we define Hfn : Dom(F ) → Dom(G) as Hfn(x, y) := Hfn(y)(x, y). See Figure 4.2.

IMPA 32 July, 2018 Ermerson Rocha Araujo Kneading Sequences

Dom(F ) Dom(G) 1 1

Hfn I(y) I(ψ(y))

0 0 −1 0 1 −1 0 1

F G Figure 4.2: Dynamics of Hfn and some curves of Cn and Cn .

Lemma 4. For each n ≥ 1, Hfn : Dom(F ) → Dom(G) is a continuous function.

Proof. Let us start by invoking the curves defined in the proof of Lemma l 2. For each sequence j1j2 ··· jl ∈ {−, +} , with 0 ≤ l ≤ n − 1, let

γF : [0, 1] → F j1j2···jl Cn

F be the continuous curve that appears in the construction of Cn . Note that F Cn is a finite union of curves of this kind, that is, [ F = γF ([0, 1]). Cn j1j2···jl l j1j2···jl ∈ {−,+} 0 ≤ l ≤ n−1

F F Take (x, y) ∈ Dom(F )\Cn . Since we have finite curves in Cn there exists F δ > 0 such that Bδ(x, y)∩Cn = ∅. Let x1 < x < x2 such that (x1, y), (x2, y) ∈ F F 1 Cn and does not exist point of Cn (y) on the interval [(x1, x2), y]. In addition, there are curves γF and γF such that γF (y) = (x , y) j1j2···jl e1e2···ek j1j2···jl 1 and γF (y) = (x , y) for some 1 ≤ l, k ≤ n − 1. For any other curve e1e2···ek 2 γF , with 1 ≤ i ≤ n − 1, we have γF (y) ∩ [(x , x ), y] = ∅. See Figure t1t2···ti t1t2···ti 1 2 4.3.

γF t1t2···ti γF j1j2···jl γF e1e2···ek

(x1, y) (x2, y) (x, y)

Figure 4.3: Construction of the domain D(x, y).

1 It is clear that (x1, y) or (x2, y) may not exist. In this case, either x1 = −1 or x2 = 1.

IMPA 33 July, 2018 Ermerson Rocha Araujo Kneading Sequences

By finiteness, there are y1 < y < y2 and continuous curves αi :[y1, y2] → F Cn , with i = 1, 2, such that αi(y) = (xi, y) and, if D(x, y) is the domain lim- ited by the curves α1([y1, y2]), α2([y1, y2]) and the intervals [(α1(y1), α2(y1)), y1], F [(α1(y2), α2(y2)), y2], then D(x, y) ∩ Cn = α1([y1, y2]) ∪ α2([y1, y2]). Besides that Hfn(D(x, y)) is a domain D(Hfn(x, y)) such that

G D(Hfn(x, y)) ∩ Cn = αf1[ψ(y1), ψ(y2)] ∪ αf2[ψ(y1), ψ(y2)], where αei = Hfn(αi). See Figure 4.4.

D(x, y) D(Hfn(x, y))

Hfn (x, y) Hfn(x, y)

Figure 4.4: Action of Hfn on neighborhood of (x, y).

Now, by definition of Hfn, for all (z, w) ∈ D(x, y) we get Hfn(z, w) =

(ξx,y,w(z), ψ(w)) where

π1(αf2(ψ(w))) − π1(αf1(ψ(w))) ξx,y,w(z) = π1(αf1(ψ(w))) + [z − π1(α1(w))] . π1(α2(w)) − π1(α1(w))

This implies that Hfn is continuous in (x, y). In fact, Hfn is continuous in D(x, y) ∪ ∂D(x, y). F l Suppose now that (x, y) ∈ Cn . Let s ∈ {−, +} such that (x, y) = F (γs (y), y) for some 0 ≤ l ≤ n − 1. Consider (xj, yj) → (x, y). Note that there exist curves γF and γF such that (x , y ) belongs to the inter- aj bj j j val [[γF (y ), γF (y )], y ] and does not exist point of F (y) on the intervals aj j bj j j Cn [(γF (y ), γF (y )), y ], for each j ≥ 1. Therefore, in order to prove that aj j bj j j

Hfn(xj, yj) → Hfn(x, y) it is enough to prove that any subsequence (xjk , yjk ) contains a subsequence (xj , yj ) so that Hn(xj , yj ) → Hn(x, y). Let kl kl f kl kl f

(xjk , yjk ) be any subsequence of (xj, yj). There is a subsequence such that F F (xj , yj ) ∈ [[γ (yj ), γ (yj )], yj ] for some sequences a and b fix. Since kl kl a kl b kl kl F F F F F γ (yj ) ≤ xj ≤ γ (yj ), we get that γ (y) ≤ γ (y) ≤ γ (y) by taking the a kl kl b kl a s b F F limit when l goes to infinity. If γ (y) = γ (y), then Hn(xj , yj ) → Hn(x, y), a b f kl kl f since F F Hn(γ (yj ), yj ) ≤ Hn(xj , yj ) ≤ Hn(γ (yj ), yj ) f a kl kl f kl kl f b kl kl

IMPA 34 July, 2018 Ermerson Rocha Araujo Kneading Sequences and

F F lim Hfn(γa (yjk ), yjk ) = Hfn(γa (y), y) l→∞ l l F = Hfn(γs (y), y) F = Hfn(γb (y), y) F = lim Hfn(γb (yjk ), yjk ). l→∞ l l

F F F F F Assume now that γa (y) 6= γb (y). Then either γs (y) = γa (y) or γs (y) = F γ (y). Thus, we can build a domain D(x, y) such that (xj , yj ) ∈ D(x, y) b kl kl for all l ≥ 1. Hence Hn(xj , yj ) → Hn(x, y). It follows that Hn is continu- f kl kl f f ous. 

The goal now is to prove that the sequence {Hfn}n≥1 is a Cauchy sequence.

Lemma 5. The sequence {Hfn}n≥1 form a Cauchy sequence.

G Proof. Let ε > 0. By equicontinuity of the family {φn }, there is δ > 0 such G G that |y1 −y2| < δ implies dH(Cn (y1), Cn (y2)) < ε for all n ≥ 1. From Lemma 3 it follows that C (G) is dense in Dom(G), so there exists a number N > 0 G such that {Bδ(x, y); (x, y) ∈ CN } covers Dom(G). Let (x, y) ∈ Dom(F ) and n, m ≥ N. As Hfn is a surjective map, we can take x1 < x < x2 such that |Hfn(x1, y) − Hfn(x, y)| = |Hfn(x2, y) − Hfn(x, y)| = 2ε. Thus there is 0 0 00 00 F 0 0 (x1, y ), (x2, y ) ∈ CN such that Hfn(x1, y) ∈ Bδ(HgN (x1, y )) and Hfn(x2, y) ∈ 00 00 F Bδ(HgN (x1, y )). Again by equicontinuity there exist (xe1, y), (xe2, y) ∈ CN such that Hfn(x, y) belongs to the interval [[HgN (xe1, y), HgN (xe2, y)], ψ(y)] and |HN (x1, y) − HN (x2, y)| < 6ε. Since Hl| F = HN | F for all l ≥ N we get g e g e f CN g CN Hfn(x, y), Hgm(x, y) ∈ [[HgN (xe1, y), HgN (xe2, y)], ψ(y)]. Therefore, |Hfn(x, y) − Hgm(x, y)| < 6ε. Hence,

kHn − Hmk = sup |Hfn(x, y) − Hgm(x, y)| < 6ε. (x,y)∈Dom(F )

Thus, it follows that {Hfn}n≥1 is a Cauchy sequence. 

From Lemma 5 there exist a continuous map He : Dom(F ) → Dom(G) such that lim Hfn = He. The proof is completed by showing that He is a semiconjugacy between F and G.

IMPA 35 July, 2018 Ermerson Rocha Araujo Kneading Sequences

F Note that the mapping He agrees with Hfn on Cn whence He ◦ F = G ◦ He on C (F ). Take (x, y) ∈ C (F )\C (F ) and (xn, yn) ∈ C (F ) such that lim(xn, yn) = (x, y). As F is continuous on Dom(F )\C (F ) and He(x, y) ∈ C (G)\C (G) it follows that

(He ◦ F )(x, y) = lim(He ◦ F )(xn, yn)

= lim(G ◦ He)(xn, yn) = (G ◦ He)(x, y).

So He ◦ F and G ◦ He agree on C (F ) = Dom(F ). Thus He is the required semiconjugacy and Theorem C is proved.

F Remark 8. Notice that if the family {φn } is equicontinuous, then He is a conjugacy.

IMPA 36 July, 2018 CHAPTER 5

On Singer’s Theorem for toy models

The goal of this short chapter is to prove a version of Singer’s theorem for toy models. We will recall the definition of the Schwarzian derivative. Let f : I → I be a C3 interval map and let x ∈ I such that Df(x) 6= 0. The Schwarzian derivative Sf(x) of f at x is defined by D3f(x) 3 D2f(x)2 Sf(x) := − . Df(x) 2 Df(x) We say that f has negative Schwarzian derivative if Sf(x) < 0 for all points x ∈ I except, possibly, the turning points. In these points we define Sf(x) = −∞. From the definition, if f, g : I → I are C3 interval maps and x ∈ I is such that Df(x) 6= 0 and Dg(f(x)) 6= 0 then S(g ◦ f)(x) = Sg(f(x)) · (Df(x))2 + Sf(x). This implies that if f and g have negative Schwarzian derivative, then g ◦ f also has negative Schwarzian derivative. The lemma below gives an intuition about the types of graphs that can not occur for functions f with negative Schwarzian derivative. Lemma 5.1. (Minimum Principle) Let I be a closed interval with end- points a, b and f : I → I a map with negative Schwarzian derivative. If Df(x) 6= 0 for all x ∈ I then |Df(x)| > min{|Df(a)|, |Df(b)|}, for all x ∈ (a, b).

37 Ermerson Rocha Araujo Kneading Sequences

We say that c is a critical point of a C1 map f if Df(c) = 0. As conse- quence of the Minimum Principle we have

Theorem 5.1. (Singer) If f : I → I is a C3 map with negative Schwarzian derivative then the immediate basin of any attracting periodic orbit contains either a critical point of f or a boundary point of the interval I.

Now we can state the main result of this chapter.

F Theorem B. Let F (x, y) = (f(y)(x),Ksign(x)(y)) be a toy model. Sup- pose that f(y):[−1, 1] → [−1, 1] has negative Schwarzian derivative for all y ∈ [0, 1]. Then the closure of the immediate basin of any strong attract- ing periodic orbit contains either a point of the critical line or a point of Λ := {−1, 1} × [0, 1].

The proof of this Theorem is similar to one dimensional case. Before proving Theorem B, let us establish some notation. We write the derivative F of the toy model F (x, y) = (f(y)(x),Ksign(x)(y)) as ! f f DF = x y . 0 Ky

We know that given (x, y) ∈ Dom(F ), there is a sequence j(x,y) = (j1 ··· jm ··· ) ∈ {−, +}N such that

m  F (x, y) = fjm (yjm−1···j1 ) ◦ · · · ◦ fj2 (yj1 ) ◦ fj1 (y)(x), yjm···j1 , for all m ≥ 0. So the derivative of F m at (x, y) is given by ! Am(x, y) Bm(x, y) DF m(x, y) = , 0 Dm(x, y) where

m−1 m−1 m Y l m Y l A (x, y) = fx(F (x, y)),D (x, y) = Ky(F (x, y)), l=0 l=0

m−1 j−1 m−1 ! m X j Y l Y l B (x, y) = fy(F (x, y)) Ky(F (x, y)) fx(F (x, y)) . j=0 l=0 l=j+1

IMPA 38 July, 2018 Ermerson Rocha Araujo Kneading Sequences

Proof of Theorem B. Let (p, q) ∈ Dom(F ) be a strong attracting periodic point of period m and B be the closure of its immediate basin. We assume that B does not contain points of Λ neither points of the critical line. Let B0 m be the interior of the connected component containing (p, q). Then F (B0) ⊂

B0. Let [(a, b), q] be the connected component of B0 ∩ I(q) containing (p, q). See Figure 5.1.

y B0

(p, q) I(q)

x a b

Figure 5.1: Construction of the interval [(a, b), q].

Since F m(p, q) = (p, q) this implies

F m([(a, b), q]) ⊂ ([(a, b), q]) .

m By our assumptions, there exists a sequence j = (j1 ··· jm) ∈ {−, +} such that m F (x, q) = (fj(q)(x), q) and Am(x, q) 6= 0

for all x ∈ (a, b), where fj(q) = fjm (qjm−1···j1 ) ◦ · · · ◦ fj2 (qj1 ) ◦ fj1 (q).

Claim: fj(q)({a, b}) ⊂ {a, b}.

Suppose that fj(q)(a) ∈ (a, b) (the case where fj(q)(b) ∈ (a, b) is similar). As (a, q) is a continuity point of F m there is a neighborhood V of (a, q) m such that F (V ) ⊂ B0. This is a contradiction since B0 is the connected component of the immediate basin containing (p, q). Therefore the claim is proved. m We can assume that A (x, q) > 0 for all x ∈ (a, b) and fj(q)(α) = α for α ∈ {a, b} (otherwise, consider A2m(x, q) instead of Am(x, q)). As (a, q) and (b, q) can not be attractor periodic points we get Am(α, q) ≥ 1 for α ∈ {a, b}.

IMPA 39 July, 2018 Ermerson Rocha Araujo Kneading Sequences

Note that fj(q) has negative Schwarzian derivative. From the Minimum Prin- ciple, it follows that Am(x, q) > 1 for all (x, q) ∈ [(a, b), q], contradicting that m F ([(a, b), q]) = [(a, b), q]. 

IMPA 40 July, 2018 CHAPTER 6

Open questions and future work

The present work has at least two possible directions of research. One is related to increase the knowledge on the toy model and, the other is to try to extend the results of the toy model to the two dimensional dissipative diffeomorphisms of the disc.

6.1 Toy model and further results

It is known that in the context of one-dimensional dynamics the set of peri- odic points is dense in the set of recurrent points for all f : I → I continuous. Inspired by this result we would like to prove the following:

Question 1: Let µ a F -invariant ergodic measure. Is supp(µ) contained in Per(F )?

Once Question 1 is answered, it is natural to wonder if the periodic points of F are dense in its recurrent set (or, more generally, non-wandering set). For this it is enough to answer the next question:

Question 2: C (F |K F ) ⊂ Per(F )?

41 Ermerson Rocha Araujo Kneading Sequences

On the other hand, it is well-known that for unimodal maps f with neg- ative Schwarzian derivative we get absence of wandering intervals. More precisely,

Theorem 6.1. (Guckenheimer) Let f : I → I be a C3 unimodal map with negative Schwarzian derivative and such that f 00(c) 6= 0 at the unique critical point c of f. Then f has no wandering intervals.

As we have seen in the proof of Theorem A, the hypothesis of nonexis- tence of wandering intervals is fundamental to construct our conjugacy. So it is natural to ask:

F Question 3: Let F (x, y) = (f(y)(x),Ksign(x)(y)) be a toy model. If f(y):[−1, 1] → [−1, 1] has negative Schwarzian derivative for all y ∈ [0, 1] does F have no wandering intervals?

6.2 Looking for combinatorial structure for two dimensional dissipative diffeomorphisms

As we saw on Chapter 2, the classical results in one-dimensional dynamics follow from the fact that the line is totally ordered: the orbit equivalence (provided by the preimages of the turning points) can be extended to the closure as a semi-conjugacy, since it is monotone. In our two dimensional counterpart, the one-dimensional order structure is preserved along “unstable leaves” providing a partial order allowing us to extend the orbit equivalence to the closure. We wonder if a similar type of approach can be used for two dimensional diffeomorphisms. In [CP17], it was introduced an open class of diffeomor- phisms of the disc, namely strongly dissipative: In short that means that stable manifolds of regular points of a non-trivial ergodic measure separate the disc.

Definition 6. Let S be a boundaryless surface. A Cr-diffeomorphism f : S → f(S) ⊂ S is strongly dissipative if

• f(S) is contained in a compact subset of S,

• f is dissipative, that is, |det(Df(x))| < 1 for any x ∈ S,

IMPA 42 July, 2018 Ermerson Rocha Araujo Kneading Sequences

• for any ergodic measure µ which is not supported on a hyperbolic sink, and for µ-almost every point x, each of the two connected components s of WS(x)\{x} meets S\f(S).

In [CP17], for each f strongly dissipative diffeomorphism of the disc a one-dimensional dynamics is associated: there exists a real tree (that is, a path connected metric space such that for any two points x, y there exists a unique subset homeomorphic to [0, 1] whose endpoints are x and y) and a continuous map defined on the tree which is measure equivalent to f. More precisely, they proved the following:

Theorem 6.2. Let f be a Cr, r > 1, strongly dissipative diffeomorphism of the disc. Then there exists a semi-conjugacy π :(D, f) → (X, h) to a continuous map h on a compact real tree which induces an injective map on the set of non-atomic ergodic measures µ of f. Moreover the entropies of µ and π∗(µ) are the same.

Furthermore, conjugate dynamics on the disc have their induced dynam- ics on the tree conjugated. The tree, by definition, is partially ordered, so it is natural to associate to it a kneading sequence involving the turning points and regions of monotonicity. The goal is to answer the following question.

Question 4: Let f and g be two strongly dissipative diffeomorphisms of the disc such that the kneading sequences of their associated dynamics over their trees are equal. Does there exist a (semi)conjugacy between f and g?

IMPA 43 July, 2018 APPENDIX A

Generalized toy models

In this appendix, we build toy models with families of l-modal maps. Consider a one-parameter family f(y) : [0, 1] → [0, 1], with y ∈ [0, 1], depending continuously on y, such that f(y) is a l-modal map verifying that

0 < c1 < ··· < cl are the turning points and f(y)({0, 1}) ⊂ {0, 1} for all y ∈ [0, 1]. This means that f(y) has local extrema at 0 < c1 < ··· < cl and that f(y) is strictly monotone in each of the l + 1 intervals I1 = [c0, c1),

I2 = (c1, c2),..., Il+1 = (cl, cl+1], where c0 = 0 and cl+1 = 1. For each y ∈ [0, 1] we use the notation f (y) for f(y) . See Figure A.1. i |Ii

f1(y) fi(y)

··· ···

fl+1(y)

1 i i l+1 c1 ci−1 ci cl

Figure A.1: Dynamics of f(y).

44 Ermerson Rocha Araujo Kneading Sequences

Let 0 = a1 < b1 < a2 < b2 < ··· < al+1 < bl+1 = 1 and

k :[a1, b1] ∪ · · · ∪ [al+1, bl+1] → [0, 1]

0 be a differentiable function such that k({ai, bi}) ⊂ {0, 1} and |k | > γ > 1. We put

K(x, y) = Ki(y), if x ∈ Ii,

−1 where Ki = (k| ) for all i = 1, . . . , l + 1. [ai,bi] We call Generalized Toy Model the map defined as

F : ([0, 1]\{c1, . . . , cl}) × [0, 1] −→ [0, 1] × [0, 1] (x, y) 7−→ (f(y)(x),K(x, y)).

As before we will make the domain of F compact introducing the points i i+1 i ci and ci with i = 1, . . . , l and extend F to them via the formula F (ci, y) = i+1 (fi(y)(ci),Ki(y)) and F (ci , y) = (fi+1(y)(ci),Ki+1(y)). See Figure A.2.

fi(y) bi i Ki ci ai

i ci−1

i i ci−1 ci Fi

i i Figure A.2: Action of Fi on [ci−1, ci] × [0, 1].

Let Ii, i = 1, . . . , l + 1, be an interval on the construction of the family y 7→ f(y). We make the following convention. If f(y) is strictly increasing at Ii, then we put

i i 1 2 2 l+1 Fi :[ci−1, ci] × [0, 1] −→ ([0, c1] ∪ (c1, c2] ∪ · · · ∪ (cl , 1]) × [0, 1] (x, y) 7−→ (fi(y)(x),Ki(y)).

IMPA 45 July, 2018 Ermerson Rocha Araujo Kneading Sequences

If f(y) is strictly decreasing at Ii, then we put

i i 1 2 2 l+1 Fi :[ci−1, ci] × [0, 1] −→ ([0, c1) ∪ [c1, c2) ∪ · · · ∪ [cl , 1]) × [0, 1] (x, y) 7−→ (fi(y)(x),Ki(y)).

So, for each (x, y) ∈ Dom(F ) there is a sequence j(x, y) = (j1j2 ··· jm ··· ) ∈ {1, . . . , l + 1}N such that

m+1 F (x, y) = Fjm+1 (xm, ym)

for all m ≥ 0, where xm = fjm (yjm−1···j1 ) ◦ · · · ◦ fj2 (yj1 ) ◦ fj1 (y)(x) and ym =

Kjm ◦ · · · ◦ Kj1 (y) := yjm···j1 . 1 2 l l+1 Similarly, if we consider the alphabet A = {I1, c1, c1,I2, . . . , cl, cl ,Il+1}, the address tF (x, y) of a point (x, y) ∈ Dom(F ) is defined by  I if π (x, y) ∈ I  i 1 i i i tF (x, y) = ci if π1(x, y) = ci  i+1 i+1  ci if π1(x, y) = ci

Besides that, the itinerary of (x, y) is the sequence

n TF (x, y) = (tF (x, y), tF (F (x, y)), . . . , tF (F (x, y)),...).

The critical line of F is the set

i i+1 Lc(F ) = {(ci, y), (ci , y); y ∈ [0, 1] and i = 1, . . . , l} and the Kneading Sequences of F , denoted by KF (Lc), are the set of all itineraries of points of Lc(F ). Now we can state a Theorem analogue to Theorem A.

Theorem A’. Let F and G be two generalized toy models. Assume that G has no wandering intervals, no interval of periodic points and no weakly attracting periodic points. If F and G have the same kneading sequences, then F and G are topologically semiconjugate.

The proof of Theorem A’ follows the same ideas of the proof of Theorem 3.

IMPA 46 July, 2018 APPENDIX B

Combinatorial equivalence for nonautonomous discrete dynamical systems

A nonautonomous discrete (short NDS) is a pair (X , F), where X = (Xn)n≥1 is a sequence of metric spaces and F = (fn)n≥1 is a sequence of continuous maps fn : Xn → Xn+1. Orbits of the system are l described by the maps fn : Xn → Xn+l, defined by

l fn(x) := (fn+(l−1) ◦ · · · ◦ fn)(x) for each n, l ∈ N and x ∈ Xn,

0 fn := idXn for each n ∈ N.

The classical autonomous setting is obtained by letting fn = f and Xn = X, −l l −1 for every n ≥ 1. Furthemore, we define fn := (fn) , which is only applied to sets. (We do not assume that the maps fn are invertible.) Nonautonomous discrete dynamical systems were introduced in [KS96]. In this paper, S. Kolyada and L. Snoha introduced and studied the notion of topological entropy for the nonautonomous discrete dynamical systems given by a sequence of endomorphisms (fn)n≥1 of a compact topological space X. They were motivated by the desire to understand better the topological entropy of skew products. In the past twenty years, a large number of papers have been devoted to dynamical properties in nonautonomous discrete systems. Kolyada et al [KMS99] generalized the definitions of [KS96]. Huang et al [HWZ08] intro- duced and studied the notion of topological pressure for the nonautonomous

47 Ermerson Rocha Araujo Kneading Sequences discrete dynamical systems. Metric entropy of NDS has been studied in [Kaw14] and [KL16]. The notion of chaos was extended to NDS setting by many researchers (see, e.g., [ZSS16],[WZ13],[Shi12], [TC06]). So, although recognizably distinct from that of classical autonomous dy- namic systems, the theory of the nonautonomuos discrete dynamical systems has developed into a highly active field of research. In particular, on [KS96, Section 5] Kolyada and Snoha studied the invariance of the topological en- tropy by conjugacy between two NDS’s.

Definition 7. Let (X , F) and (Y, G) be two nonautonomous discrete dynam- ical systems. We say that (X , F) and (Y, G) are topologically semiconjugate if there exist a sequence hn : Xn → Yn of continuous surjective maps such that hn+1 ◦ fn = gn ◦ hn for every n ≥ 1. If hn is a homeomorphism for each n ≥ 1, then we say that (X , F) and (Y, G) are topologically conjugate.

So it is natural to ask:

Question 1: Let (X , F) and (Y, G) be two nonautonomous discrete dy- namical systems. Under which conditions (X , F) and (Y, G) are topologically (semi)conjugate?

The main aim of this Appendix is to give a notion weaker than topological conjugacy to NDS’s on the particular case where Xn are intervals and fn are unimodal maps for every n ≥ 1, namely combinatorial equivalence.

B.1 Combinatorial equivalence for NDS

Let (X , F) be the NDS defined as follows: For each n ≥ 1 let Jn = [an, bn] be an interval and fn : Jn → Jn+1 be a continuous map satisfying fn(an) = fn(bn) = an+1. Besides that, there is cn ∈ (an, bn) such that fn|[an,cn] is strictly increasing and fn|[cn,bn] is strictly decreasing. The points {cn; n ≥ 1} are the turning points of the NDS. We call the NDS defined above unimodal nonautonomous discrete dynamical system (short UNDS) For each n ≥ 1 set

 F l Cn(F) = x ∈ Jn ; ∃ l ≥ 0 such that fn(x) = cn+l .

Definition 8. We say that two UNDS’s (X , F) and (Y, G) with turning F G points {cn ; n ≥ 1} respectively {cn; n ≥ 1} are combinatorially equiva-

IMPA 48 July, 2018 Ermerson Rocha Araujo Kneading Sequences

lent if there exist a family of order preserving bijections hn : Cn(F) → Cn(G) such that hn+1 ◦ fn = gn ◦ hn on Cn(F)\{cn} for all n ≥ 1.

Let’s proceed, as in previous chapters, by constructing the symbolic space for UNDS. Consider the alphabet A = {L, cn,R; n ≥ 1}. Let x ∈ Jn for some n ≥ 1. The address iF (x) is defined by  L if x ∈ [a , c )  n n iF (x) = cn if x = cn   R if x ∈ (cn, bn]

Applying this map to an orbit of a given point x ∈ Jn, we associate to that orbit one infinite symbolic sequence.

Definition 9. Consider the sequence of symbols in A

1 l N IF (x) = (iF (x), iF (fn(x)), . . . , iF (fn(x)),...) ∈ A .

This infinite sequence is called itinerary of x ∈ Jn.

The kneading invariant of (X , F) is the sequence V(F) = {Vn}n≥1, where Vn := IF (cn). Now we can state the main result of this appendix.

Theorem D. Let (X , F) and (Y, G) be two UNDS’s with kneading invariants V(F) and V(G). If V(F) = V(G), then (X , F) and (Y, G) are combinatorially G equivalent. Furthermore, if Cn(G) is dense in Jn for all n ≥ 1, then (X , F) and (Y, G) are topologically semiconjugate.

Again, the proof of the theorem above is similar to Theorem 2.1. Note that many concepts used in classical dynamics such as periodicity, recurrence and wandering domains do not seem to make sense for nonau- tonomous systems, while other concepts such as entropy can be generalized. Thus, if we want analogous hypotheses to Theorem 2.1 to ensure that G be G dense in Jn , firstly we have define compatible hypotheses. So we can ask:

Question 2: Let (X , F) be a unimodal nonautonomous discrete dynam- F ical systems. Under which conditions we get Cn(F) = Jn for all n ≥ 1?

IMPA 49 July, 2018 Bibliography

[BC91] Michael Benedicks and Lennart Carleson. The dynamics of the Hénon map. Ann. of Math. (2), 133(1):73–169, 1991.

[CFW81] A. Connes, J. Feldman, and B. Weiss. An amenable equivalence relation is generated by a single transformation. Dynamical Systems, 1(4):431–450 (1982), 1981.

[CGP88] Predrag Cvitanović, Gemunu H. Gunaratne, and Itamar Procac- cia. Topological and metric properties of Hénon-type strange at- tractors. Phys. Rev. A (3), 38(3):1503–1520, 1988.

[CP17] S. Crovisier and E. Pujals. Strongly dissipative surface diffeomor- phisms. arXiv:1608.05999, 2017.

[Cvi91] Predrag Cvitanović. Periodic orbits as the skeleton of classical and quantum chaos. Phys. D, 51(1-3):138–151, 1991. Nonlinear science: the next decade (Los Alamos, NM, 1990).

[dC99] André de Carvalho. Pruning fronts and the formation of horse- shoes. Ergodic Theory Dynam. Systems, 19(4):851–894, 1999.

[dCH02] André de Carvalho and Toby Hall. How to prune a horseshoe. Nonlinearity, 15(3):R19–R68, 2002.

[dCH03] André de Carvalho and Toby Hall. Symbolic dynamics and topo- logical models in dimensions 1 and 2. In Topics in dynamics and

50 Ermerson Rocha Araujo Kneading Sequences

ergodic theory, volume 310 of London Math. Soc. Lecture Note Ser., pages 40–59. Cambridge Univ. Press, Cambridge, 2003.

[dMvS93] Welington de Melo and Sebastian van Strien. One-dimensional dynamics, volume 25 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1993.

[Dye59] H. A. Dye. On groups of measure preserving transformation. I. Amer. J. Math., 81:119–159, 1959.

[Dye63] H. A. Dye. On groups of measure preserving transformations. II. Amer. J. Math., 85:551–576, 1963.

[HWZ08] Xianjiu Huang, Xi Wen, and Fanping Zeng. Topological pressure of nonautonomous dynamical systems. Nonlinear Dyn. Syst. Theory, 8(1):43–48, 2008.

[Ish97] Yutaka Ishii. Towards a kneading theory for Lozi mappings. I. A solution of the pruning front conjecture and the first tangency problem. Nonlinearity, 10(3):731–747, 1997.

[Kaw14] Christoph Kawan. Metric entropy of nonautonomous dynamical systems. Nonauton. Dyn. Syst., 1:26–52, 2014.

[KL16] Christoph Kawan and Yuri Latushkin. Some results on the en- tropy of nonautonomous dynamical systems. Dynamical Systems, 31(3):251–279, 2016.

[KMS99] Sergi˘ı Kolyada, Michal Misiurewicz, and L’ubomír Snoha. Topo- logical entropy of nonautonomous piecewise monotone dynamical systems on the interval. Fund. Math., 160(2):161–181, 1999.

[KS96] Sergi˘ı Kolyada and L’ubomír Snoha. Topological entropy of nonautonomous dynamical systems. Random Comput. Dynam., 4(2-3):205–233, 1996.

[Men13] Valentín Mendoza. Proof of the pruning front conjecture for cer- tain Hénon parameters. Nonlinearity, 26(3):679–690, 2013.

[MMP13] Carlos Matheus, Carlos G. Moreira, and Enrique R. Pujals. Ax- iom A versus Newhouse phenomena for Benedicks-Carleson toy models. Ann. Sci. Éc. Norm. Supér. (4), 46(6):857–878, 2013.

IMPA 51 July, 2018 Ermerson Rocha Araujo Kneading Sequences

[MR04] Diana Aldea Mendes and José Sousa Ramos. Kneading theory for triangular maps. Int. J. Pure Appl. Math., 10(4):421–450, 2004.

[MŠ16] M. Misiurewicz and S. Štimac. Symbolic dynamics for Lozi maps. Nonlinearity, 29(10):3031–3046, 2016.

[MT88] John Milnor and William Thurston. On iterated maps of the inter- val. In Dynamical systems (College Park, MD, 1986–87), volume 1342 of Lecture Notes in Math., pages 465–563. Springer, Berlin, 1988.

[OW80] Donald S. Ornstein and Benjamin Weiss. Ergodic theory of amenable group actions. I. The Rohlin lemma. Bull. Amer. Math. Soc. (N.S.), 2(1):161–164, 1980.

[PRH07] Enrique R. Pujals and Federico Rodriguez Hertz. Critical points for surface diffeomorphisms. J. Mod. Dyn., 1(4):615–648, 2007.

[Ran78] David Rand. The topological classification of Lorenz attractors. Math. Proc. Cambridge Philos. Soc., 83(3):451–460, 1978.

[Shi12] Yuming Shi. Chaos in nonautonomous discrete dynamical systems approached by their induced systems. Internat. J. Bifur. Chaos Appl. Sci. Engrg., 22(11):1250284, 12, 2012.

[TC06] Chuanjun Tian and Guanrong Chen. Chaos of a sequence of maps in a metric space. Chaos Solitons Fractals, 28(4):1067–1075, 2006.

[WZ13] Xinxing Wu and Peiyong Zhu. Chaos in a class of nonautonomous discrete systems. Appl. Math. Lett., 26(4):431–436, 2013.

[ZSS16] Hao Zhu, Yuming Shi, and Hua Shao. Devaney chaos in nonau- tonomous discrete systems. International Journal of Bifurcation and Chaos, 26(11):1650190, 2016.

IMPA 52 July, 2018