1 Introduction

An important task in ergodic theory is to describe the topology of the space of invariant and/or ergodic measures which are supported on a given invariant set. Here in many cases the weak\(*\) topology is considered, though one also studies convergence in the weak\(*\) topology and entropy. Recently there happened a certain revival of this type of problems in the context of nonhyperbolic dynamical systems [2, 12, 16, 17], most of them revisiting the pioneering work of Sigmund on topological dynamical systems satisfying the specification property [27, 28].

For a general continuous map F on a metric space \(\Lambda \), consider the set of F-invariant Borel probability measures \(\mathcal {M}(\Lambda )\) and denote by \(\mathcal {M}_{\mathrm{erg}}(\Lambda )\) the subset of ergodic ones. If \(\Lambda \) is compact then \(\mathcal {M}=\mathcal {M}(\Lambda )\) is a Choquet simplex whose extremal elements are the ergodic measures. Density of ergodic measures in \(\mathcal {M}\) implies that either \(\mathcal {M}\) is a singleton (when F is uniquely ergodic) or a nontrivial simplex whose extreme points are dense. In the latter case, it is the so-called Poulsen simplex and by [23] it has immediately a number of further strong properties such as arcwise connectedness. Sigmund [27, 28] addressed first the questions on the density of ergodic measures and also the properties of generic invariant measures. He showed that for a map F satisfying the so-called periodic specification property the periodic measures (and thus the ergodic ones) are dense in \(\mathcal {M}\). Here a measure is periodic if it is the invariant probability measure supported on a periodic orbit. Moreover, the sets of ergodic measures and of measures with entropy zero are both residual in \(\mathcal {M}\). For an updated discussion and more references, see [16].

Observe that Sigmund’s results [27, 28] immediately apply to any basic set of a smooth Axiom A diffeomorphism. In a (more) general context, to address the general question if the space \(\mathcal {M}\) has dense extreme points or at least is connected, some natural requirements are to be satisfied. An important one is certainly topological transitivity, which is however far from being sufficient as for example there exist minimal systems with exactly two ergodic measures.

Nowadays arguments which provide the connectedness of \(\mathcal {M}\) are largely based on the approximation of invariant measures by periodic measures or Markov ergodic measures supported on horseshoes (a specific type of basic set). This demands that the periodic orbits involved are hyperbolic and somehow dynamically related among themselves. A natural relation introduced by Newhouse [24], and used in this context, is the homoclinic relation, that is, the un-/stable invariant sets of these orbits intersect cyclically and transversally.

A natural strategy is to study the components of the space of measures which each are candidate to correspond to one of the “elementary” undecomposable pieces of the dynamics. One of the possibilities to define properly what is meant by elementary is the homoclinic class, that is, the closure of the hyperbolic periodic orbits which are homoclinically related to the orbit of a hyperbolic periodic point P and denoted by H(P). Note that one of the fundamental properties is that the dynamics on each class is topologically transitive. Basic sets of the hyperbolic theory mentioned above are the simplest examples of homoclinic classes.

Notice that, when defining a homoclinic class, taking the closure can incorporate other orbits which are dynamically related but which are of different type of hyperbolicity. In this way, homoclinic classes may fail to be hyperbolic, contain saddles of different types of hyperbolicity (different \(\mathrm{u}\)-index, that is, dimension of unstable manifold), exhibit internal cycles, and support nonhyperbolic measures (also with positive entropy). Homoclinic classes of periodic points of different indices may even coincide. Furthermore, there are examples where a homoclinic class H(P) of a periodic point P properly contains another class \(H(P')\) of a periodic point \(P'\) of the same index as P. Note that this precisely occurs if \(P'\in H(P)\) was not homoclinically related to P. One sometimes refers to \(H(P')\) as an exposed piece of H(P) [10]. This type of phenomenon is a key ingredient in this paper. This gives only a rough idea what complicated structure these classes may have, see also [3, Chapter 10.4] for a more complete discussion.

To be more precise for the following, we say that an ergodic measure \(\mu \) is hyperbolic if its Lyapunov exponents are nonzero. Moreover, almost all points have the same number \(u=u(\mu )\) of positive Lyapunov exponents and we call this number u the \(\mathrm{u}\)-index of\(\mu \) (analogously to hyperbolic periodic measures above). Given u, we denote by denote by \(\mathcal {M}_{\mathrm{erg},u}\) the set of ergodic measures of \(\mathrm{u}\)-index u. Note that in general one may have \(\mathcal {M}_{\mathrm{erg},u}(H(P))\ne \varnothing \) for several values of u.

For the following let us study the topological structure of \(\mathcal {M}_{\mathrm{erg},u}(H(P))\) for u being the index of P. Assuming that H(P) is locally maximal and that all the saddles of index u are homoclinically related, in [17] it is shown that \(\mathcal {M}_{\mathrm{erg},u}(H(P))\) is path connected with periodic measures being dense and that its closure is a Poulsen simplex. Note that \(\mathcal {M}_{\mathrm{erg},u}(H(P))\) may only capture some part of \(\mathcal {M}_{\mathrm{erg}}(H(P))\). Indeed this occurs when H(P) contains saddles of different indices. Still in this context, assume now that there coexists a saddle Q of index \(v\ne u\) and having the property that \(H(Q)\subset H(P)\) (in an extreme case, these classes can even coincide as sets) and assume that all the saddles of index v in H(Q) are homoclinically related with Q and consider \(\mathcal {M}_{\mathrm{erg},v}(H(Q))\). Though the interrelation between \(\mathcal {M}_{\mathrm{erg},u}(H(P))\) and \(\mathcal {M}_{\mathrm{erg},v}(H(Q))\) is not addressed in [17], note that, by the very definition, they are disjoint. Nevertheless, their closures may intersect or may not. Indeed, the space \(\mathcal {M}_{\mathrm{erg}}(H(P))\) may be connected or may not. To address this point is precisely the goal of this paper.

We introduce a class of examples of saddles P and Q of different indices whose homoclinic classes coincide \(H(P)=H(Q)=\Lambda \) such that \(\Lambda \) is the disjoint union of two invariant sets \(\Lambda _{\mathrm {ex}}\) (a compact set that is a topological horseshoe) and \(\Lambda _{\mathrm {core}}\). Moreover, these sets satisfy the following properties: (i) \(P, Q\in \Lambda _{\mathrm {core}}\) and the closure of \(\Lambda _{\mathrm {core}}\) is the whole homoclinic class, (ii) every pair of saddles of the same index in \(\Lambda _{\mathrm {core}}\) (respectively, \(\Lambda _{\mathrm {ex}}\)) are homoclinically related, and (iii) no saddle in \(\Lambda _{\mathrm{core}}\) is homoclinically related to any one in \(\Lambda _{\mathrm{ex}}\). We refer to \(\Lambda _{\mathrm {ex}}\) as the exposed piece of \(\Lambda =H(P)=H(Q)\) and to \(\Lambda _{\mathrm{core}}\) as its core. We study the space \(\mathcal {M}_{\mathrm{erg}}(\Lambda )\) and show that it has an interesting topological structure: the set \(\mathcal {M}_{\mathrm{erg}}(\Lambda )\) has three pairwise disjoint parts \(\mathcal {M}_{\mathrm{erg},u}(\Lambda )\), \(\mathcal {M}_{\mathrm{erg},v}(\Lambda )\), \(v=u+1\) and uv are the indices of P and Q, and \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm {ex}})\), such that

$$\begin{aligned} \mathcal {M}_{\mathrm{erg}}(\Lambda )= \mathcal {M}_{\mathrm{erg},u}(\Lambda ) \cup \mathcal {M}_{\mathrm{erg},u}(\Lambda ) \cup \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm {ex}}) \cup \mathcal {M}_{\mathrm{erg, nhyp}}(\Lambda ) \end{aligned}$$

where \(\mathcal {M}_{\mathrm{erg, nhyp}}(\Lambda )\) is the set of of nonhyperbolic ergodic measures of \(\Lambda \). Note that \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm {ex}})\) and \(\mathcal {M}_{\mathrm{erg, nhyp}}(\Lambda )\) may intersect. Moreover, the sets \(\mathrm {closure}(\mathcal {M}_{\mathrm{erg},u}(\Lambda ))\), \(\mathrm {closure}(\mathcal {M}_{\mathrm{erg},v}(\Lambda ))\), and \(\mathrm {closure}(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm {ex}}))\), are Poulsen simplices whose intersection is contained in \(\mathcal {M}_{\mathrm{erg, nhyp}}(\Lambda )\), see Theorem 2.5. Figure 1 below illustrates the interrelation between the measure space components.

Fig. 1
figure 1

The space \(\mathcal {M}_{\mathrm{erg}}(\Lambda )\)

Let us say a few additional words about the topological structure of the set \(\Lambda =H(P)=H(Q)\). There are two exposed saddles \(P_{\mathrm {ex}}, Q_{\mathrm {ex}} \in \Lambda _{\mathrm {ex}}\) of the same indices such as P and Q, respectively, which are involved in a heterodimensional cycle (i.e., the invariant sets of these saddles meet cyclically), Indeed, the intersections of these invariant sets give rise to the exposed piece of dynamics that satisfy \(\Lambda _{\mathrm {ex}}=H(P_{\mathrm {ex}})=H(Q_{\mathrm {ex}})\subsetneq \Lambda \). We are aware that on one hand this is a quite specific dynamical configuration, on the other hand it provides paradigmatic examples. We also observe that this dynamical configuration resembles in some aspects the so-called Bowen eye (a two dimensional vector field having two saddle singularities involved in a double saddle connection) in [15, 30] and the examples due to Kan of intermingled basins of attractions (where an important property is that the boundary of an annulus is preserved) [19]. Finally, if we considered systems satisfying some boundary conditions or preserving a boundary, the conditions considered are quite general.

A particular emphasize is given to the measures of maximal entropy. In some cases, these measures can be nonhyperbolic. We give a (non-transitive) example where the unique measure of maximal entropy is nonhyperbolic.

Finally, we state our results for step skew products (these examples have differentiable realizations as partially hyperbolic sets with one dimensional central direction) and throughout the paper we do not aim generality, on the contrary our goal is to make the construction in the simplest setting emphasizing the key ingredients behind the constructions.

This paper is organized as follows. In Sect. 2 we state precisely our setting and our examples and state our main results. In Sect. 3 we study the “symmetries” between certain measures and investigate entropy. In Sect. 4 we study the approximation of “boundary measures”. In Sect. 5 we study the measures supported in \(\Lambda _{\mathrm {core}}\). In “Appendix A” we provide details on transitivity and homoclinic relations in our examples and we analyze examples with nonhyperbolic measures of maximal entropy.

2 Setting and statement of results

We now define precisely the dynamics that we will study. Consider \(C^1\) diffeomorphisms \(f_0,f_1:[0,1]\rightarrow [0,1]\) satisfying the following properties:

  1. (H1)

    The map \(f_0\) has (exactly) two fixed points \(f_0(0)=0\) and \(f_0(1)=1\), satisfies \(f_0'(0)=\beta >1\) and \(f_0'(1)=\lambda \in (0,1)\).

  2. (H2)

    The map \(f_1\) has negative derivative and satisfies \(f_1(0)=1\) and \(f_1(1)=0\).

The simplest (and also paradigmatic) example occurs when \(f_1(x)=1-x\).

Let \(\sigma :\Sigma _2 \rightarrow \Sigma _2\) be the standard shift map on the shift space \(\Sigma _2=\{0,1\}^{{\mathbb {Z}}}\) of two-sided sequences, endowed with the usual metric. Consider the one step-skew product map F associated to \(\sigma \) and the maps \(f_0\) and \(f_1\) defined by

$$\begin{aligned} F:\Sigma _2\times [0,1]\rightarrow \Sigma _2 \times [0,1],\quad ( \xi , x) \mapsto \big ( \sigma (\xi ) , f_{\xi _0}(x) \big ). \end{aligned}$$
(2.1)

We consider the following F-invariant subsets of \(\Lambda {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\Sigma _2\times [0,1]\)

$$\begin{aligned} \Lambda _{\mathrm{{ex}}}{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\Sigma _2\times \{0,1\}, \quad \Lambda _{\mathrm{{core}}} {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}( \Sigma _2\times [0,1]){\setminus }\Lambda _{\mathrm{{exc}}} =\Sigma _2\times (0,1). \end{aligned}$$
(2.2)

We say that \(\Lambda _{\mathrm{{ex}}}\) is the exposed piece of \(\Sigma _2\times [0,1]\) and that \(\Lambda _{\mathrm{{core}}}\) is the core of \(\Sigma _2\times [0,1]\) (these denominations are justified below). Note that \(\Lambda _{\mathrm{{ex}}}\) is closed while \(\Lambda _{\mathrm{{core}}}\) is not. Moreover, \(F|_{\Lambda _{\mathrm{ex}}}\) is topologically transitive. In fact, \(F|_{\Lambda _{\mathrm{ex}}}\) is conjugate to a subshift of finite type, one may think this dynamical system as a horseshoe in a “plane”, in that plane any pair of saddles are “homoclinically related”.

Remark 2.1

(Topological dynamics on \(\Sigma _2\times [0,1]\)) Note again that \(F|_{\Lambda _{\mathrm{ex}}}\) is conjugate to a subshift of finite type. While the dynamics in \(\Lambda _{\mathrm{ex}}\) is completely characterized, in our quite general setting very few can be said about the dynamics of F in \(\Lambda _{\mathrm{core}}\). The most interesting case certainly occurs when \(F|_{\Lambda _{\mathrm{core}}}\) is topologically transitive. Below we will see more specific examples where this transitivity indeed holds and, moreover, hyperbolic periodic orbits of positive and negative Lyapunov exponent are both dense in \(\Sigma _2\times [0,1]\) and homoclinically related. We will see that nevertheless the measure space \(\mathcal {M}(\Lambda _{\mathrm{ex}})\) is “semi-detached” from \(\mathcal {M}(\Lambda _{\mathrm{core}})\).

Consider now more specific hypotheses on the \(C^1\) diffeomorphisms interval maps \(f_0,f_1:[0,1]\rightarrow [0,1]\):

  1. (H2’)

    \(f_1(x)=1-x\).

  2. (H3)

    The derivative \(f_0^\prime \) is decreasing. Considering the point \(c\in (0,1)\) defined by the condition \(f_0^\prime (c)=1\), it holds \( f_1 \circ f_0^2(c)> f_0^2(c). \)

  3. (H4)

    The numbers \(\lambda \) and \(\beta \) given in (H1) satisfy

    $$\begin{aligned} \varkappa {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\frac{\lambda ^2\, (1-\lambda )}{\beta \, (\beta -1)}>1. \end{aligned}$$

Observe that for fixed \(\lambda \), the inequality in (H4) holds whenever \(\beta \) is close enough to 1.

Fig. 2
figure 2

Fiber maps of the example in Remark 2.4

Proposition 2.2

Assume that F defined in (2.1) satisfies the hypotheses (H1), (H2’), (H3), and (H4). Then F is topologically transitive. Moreover, every pair of fiber expanding hyperbolic periodic orbits and every pair of fiber contracting hyperbolic periodic orbits in \(\Lambda _{\mathrm{core}}\) are homoclinically related, respectively.

Remark 2.3

(Discussion of hypotheses) Homoclinic relations for skew products are recalled in “Appendix A”, where also the above proposition is proved. Condition (H4) will provide so-called expanding itineraries which in turn imply the homoclinic relations and their density for expanding points, while condition (H3) takes care of so-called contracting itineraries and the corresponding homoclinic relations. Thus, we conclude transitivity. The proof follows largely blender-like standard arguments used in [8]. Condition (H2’) is only used for simplicity and also to follow more closely the model in [14]. The key facts remain true assuming only (H2), in particular we never use the fact that for (H2’) the map \(f_1\) is an involution.

We observe that (H3) and (H4) demand a certain “asymmetry” of the fiber map \(f_0\). In Sect. A.4 we will provide a “symmetric” example which satisfies (H1) and for which the associated skew product fails to be transitive and its only measure of maximal entropy is nonhyperbolic and supported on \(\Lambda _{\mathrm{ex}}\).

Remark 2.4

(Examples in \(\Sigma _3\times \mathbb {S}^1\) and \(\Sigma _2\times \mathbb {S}^1\)) We can produce a transitive example in \(\Sigma _3\times \mathbb {S}^1\) with properties analogous to the one in Proposition 2.2 as follows. Obtain \(\mathbb {S}^1\) by identifying the boundary points of [0, 2]. Define \(g_0,g_1,g_2:\mathbb {S}^1\rightarrow \mathbb {S}^1\) as follows

  • \(g_0(x)=f_0(x)\) if \(x\in [0,1]\) and \(g_0(x)=2-f_0 (2-x)\) if \(x\in (1,2)\).

  • \(g_1(x)=f_1(x)=1-x\) if \(x\in [0,1]\) and \(g_1(x)=3-x\) if \(x\in (1,2)\),

  • \(g_2(x)=2-x \mod 2\) (or any appropriate map preserving \(\{0,1\}\) and interchanging the intervals (0, 1) and (1, 2)).

These maps are depicted in Fig. 2. In this case, \(\Lambda _{\mathrm {ex}}=\Sigma _3\times \{0,1\}\) and \(\Lambda _{\mathrm {core}}=\Sigma _3\times ((0,1)\cup (1,2))\). We observe that the IFS \(\{g_0,g_1,g_2\}\) does not satisfy the axioms stated in [11] which would prevent the existence of exposed pieces of dynamics. Although the Axioms Transitivity and CEC (controlled expanding forward/backward covering) can be verified, the Axiom Accessibility is not satisfied (the points \(\{0,1,2\}\) cannot “be reached from outside”).

Note that the skew product on \(\Sigma _2\times {\mathbb {S}}^1\) generated by the fiber maps \(\{g_0,g_1\}\) as above is not transitive and has two open “transitive” components \(\Lambda _{\mathrm{core}}^-\) and \(\Lambda _{\mathrm{core}}^+\) contained in \(\Sigma _2\times (0,1)\) and \(\Sigma _2\times (1,2)\), respectively, which are glued at the “exposed” piece \(\Sigma _2\times \{0,1\}\). The additional map \(g_2\) in the previous example just mixes the two components \(\Lambda _{\mathrm{core}}^\pm \) while preserving the exposed piece.

Let \(\mathcal {M}\) be the space of all F-invariant measures and equip it with the weak\(*\) topology. It is well known that it is a compact metrizable topological space [31, Chapter 6.1]. Denote by \(\mathcal {M}_{\mathrm{erg}}=\mathcal {M}_{\mathrm{erg}}(\Sigma _2\times [0,1])\) the subset of ergodic measures. We denote by \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) the ergodic measures supported on \(\Lambda _{\mathrm{ex}}\) and by \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\) the ergodic measures supported on \(\Lambda _{\mathrm{core}}\). Observe that

$$\begin{aligned} \mathcal {M}_{\mathrm{erg}} =\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\cup \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}}). \end{aligned}$$

We will study this system by separately looking at measures supported on these two sets. A crucial point for us is how these two components “glue”.

Given \(X=(\xi ,x)\in \Sigma _k\times [0,1]\), we consider the (fiber) Lyapunov exponent of the map F at X which is defined by

$$\begin{aligned} \chi (X) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\lim _{n\rightarrow \pm \infty }\frac{1}{ n}\log \,|(f^n_\xi )'(x)|, \quad \text { where }\quad f_\xi ^n{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}f_{\xi _{n-1}}\circ \cdots \circ f_{\xi _0}, \end{aligned}$$

where we assume that both limits exist and are equal. Note that it is nothing but the Birkhoff average of a continuous function. Given an invariant measure \(\mu \in \mathcal {M}(\Lambda )\), define its Lyapunov exponent by

$$\begin{aligned} \chi (\mu ) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\int \log \,|(f_\xi ^n)'(x)|\,d\mu (\xi ,x). \end{aligned}$$

For every F-ergodic Borel probability measure \(\mu \) the Lyapunov exponent is almost everywhere well defined and constant equal to \(\chi (\mu )\). An ergodic measure \(\mu \) is nonhyperbolic if \(\chi (\mu )=0\) and hyperbolic otherwise.Footnote 1

Accordingly, we split the set of all ergodic measures in \(\Lambda _{\mathrm{core}}\) and consider the decomposition

$$\begin{aligned} \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}}) =\mathcal {M}_{\mathrm{erg,<0}}(\Lambda _{\mathrm{core}})\cup \mathcal {M}_{\mathrm{erg,0}}(\Lambda _{\mathrm{core}}) \cup \mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}}) \end{aligned}$$

into measures with negative, zero, and positive fiber Lyapunov exponent, respectively. Analogously, we consider

$$\begin{aligned} \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}}) =\mathcal {M}_{\mathrm{erg,<0}}(\Lambda _{\mathrm{ex}})\cup \mathcal {M}_{\mathrm{erg,0}}(\Lambda _{\mathrm{ex}}) \cup \mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{ex}}). \end{aligned}$$

Properties of the space of measures are summarized in the next theorem. Given \(\mathcal {N}\subset \mathcal {M}\), its closed convex hull is the smallest closed convex set containing \(\mathcal {N}\).

Theorem 2.5

Assume that F defined in (2.1) satisfies the hypotheses (H1) and (H2). Then the space \(\mathcal {M}(\Sigma _2\times [0,1])\) has the following properties:

  1. 1.

    Periodic orbit measures are dense in the closed convex hull of \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\).

  2. 2.

    Every measure \(\mathcal {M}(\Lambda _{\mathrm{ex}})\) with \(\chi (\mu )\ne 0\) has positive weak\(*\) distance from \(\mathcal {M}(\Lambda _{\mathrm{core}})\).

  3. 3.

    Every measure \(\mu \in \mathcal {M}(\Lambda _{\mathrm{ex}})\) with \(\chi (\mu )=0\) can be weak\(*\) approximated by periodic measures in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\).

  4. 4.

    Each of the components \(\mathcal {M}_{\mathrm{erg,\star }}(\Lambda _{\mathrm{ex}})\), \(\star \in \{<0,0,>0\}\) is nonempty.

Moreover, if hypotheses (H2’), (H3), and (H4) additionally hold, then

  1. 5.

    The set \(\mathcal {M}_{\mathrm{erg,\star }}(\Lambda _{\mathrm{core}})\), \(\star \in \{<0,>0\}\), is nonempty.

  2. 6.

    The set \(\mathcal {M}_{\mathrm{erg,<0}}(\Lambda _{\mathrm{core}})\) and the set \(\mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}})\) are arcwise connected, respectively.

Fig. 3
figure 3

Porcupine-like horseshoes

The fact that there are ergodic measures with zero Lyapunov exponent and positive entropy in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\) can be shown using methods in [1], we refrain from discussing this here. We also refrain from studying how such measures are approached by hyperbolic ergodic measures in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\) as this is much more elaborate and will be part of an ongoing project (see [11] for techniques in a slightly different but technically simpler context).

Fig. 4
figure 4

Fiber maps in (2.1)

Remark 2.6

(Porcupine vs. totally spiny porcupine) Let us compare the porcupine-like horseshoes corresponding to the interval maps in Fig. 3 with the “totally spiny porcupine” discussed here (corresponding to Fig. 4). Porcupine-like horseshoes were introduced in [13] as model for internal heterodimensional cycles in horseshoes. Later these horseshoes were generalized and studied in a series of papers from various points of view: topological [8,9,10], thermodynamical [10, 22, 25, 26] and fractal [14].Footnote 2 This line of research is also closely related to the study of so-called bony attractors and sets (see [18] for a survey and references). One important motivation to study those models is that they serve as a prototype of partially hyperbolic dynamics.

Let us consider the map \(F_t\) defined as in (2.1) but with the maps \(f_0,f_{1,t}\) as in Fig. 3 in the place of \(f_0,f_1\) in Fig. 4. Let \(\Gamma ^t\) be the maximal invariant set of \(F_t\). In the above cited porcupine-like horseshoes, one also splits the maximal invariant set \(\Gamma ^t\) (which is nonhyperbolic and transitive) into two parts \(\Gamma ^t_{\mathrm{ex}}\) and \(\Gamma ^t_{\mathrm{core}}\) in the same spirit as in (2.2) (and with analogous properties as in Proposition 2.2). In that case \(\Gamma ^t_{\mathrm{ex}}\) consists only of one fiber expanding point \(Q=(0^{\mathbb {Z}},0)\) and \(\Gamma ^t_{\mathrm{core}}\) is its complement that contains the fiber contracting point \(P=(0^{\mathbb {Z}},1)\). The space of ergodic measures of \(\Gamma ^t\) splits into two components, each of them connected but at positive distance from each other, which are \(\{\delta _Q\}\) and \(\mathcal {M}_{\mathrm{erg}}(\Gamma ^t_{\mathrm{core}})\) (see, in particular, [22]). In the transition from a porcupine to a totally spiny porcupine (which occurs at \(t=1\)), the space of ergodic measures becomes connected (stated in Theorem 2.5) and this happens as follows. The measures \(\delta _Q\) and \(\delta _P\) form part of the space of ergodic measures of an abstract horseshoe \(\Lambda _{\mathrm{ex}}\). At the same time, the measure \(\delta _P\) detaches from \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\) which is a consequence of the fact that the saddle P is not homoclinically related to any saddle in \(\Lambda _{\mathrm{core}}\), similarly for Q. The components \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) and \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\) become glued through nonhyperbolic measures.

Theorem 2.7

Assume that F defined in (2.1) satisfies the hypotheses (H1) and (H2). Then there is a unique measure \(\mu _{\mathrm{max}}^{\mathrm{ex}}\) of maximal entropy \(\log 2\) in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) and its Lyapunov exponent is given by

$$\begin{aligned} \chi (\mu _{\mathrm{max}}^{\mathrm{ex}}) = \frac{1}{4}\left( \log \, f_0'(0)+\log \, f_0'(1) +\log \, |f_1'(0)|+\log \, |f_1'(1)|\right) . \end{aligned}$$

Moreover, if the measure \(\mu _{\mathrm{max}}^{\mathrm{ex}}\) is hyperbolic then there exists at least one measure of maximal entropy in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\). More precisely, if the measure \(\mu _{\mathrm{max}}^{\mathrm{ex}}\) has positive (negative) Lyapunov exponent then there exists a measure of maximal entropy with nonpositive (nonnegative) exponent in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\).

Note that the topological structure of \(\mathcal {M}(\Lambda _{\mathrm{ex}})\) (items 1. and 4. in Theorem 2.7) are immediate consequences of the fact that the dynamics of F on \(\Lambda _{\mathrm{ex}}\) is conjugate to a subshift of finite type (see Sect. 3 for details).

Note that the under the hypotheses of the above theorem, we do not know if the measure of maximal entropy in \(\Lambda _{\mathrm{core}}\) is hyperbolic or not.

Remark 2.8

In view of Theorem 2.7, choosing the derivatives of the fiber maps at 0 and 1 appropriately, one obtains one measure of maximal entropy \(\mu _{\mathrm{max}}^{\mathrm{ex}}\) which is nonhyperbolic. Note that condition (H4) is incompatible with such a choice, and hence it is unclear if the system is transitive (compare Proposition 2.2).

Similar arguments apply to the examples discussed in Remark 2.4. It is interesting to compare to the results in [29] where ergodic measures with “sufficiently high entropy” are always hyperbolic, though there a key ingredient is accessibility which is missing here.

In “Appendix A.3”, we provide examples where the system is transitive and exhibits a nonhyperbolic measure of maximal entropy in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\), proving the following theorem.

Theorem 2.9

There are maps \({\tilde{F}}\) defined as in (2.1) whose fiber maps \({\tilde{f}}_0,{\tilde{f}}_1\) satisfy

  1. 1.

    \({\tilde{f}}_0\) has (exactly) two fixed points \({\tilde{f}}_0(0)=0\) and \({\tilde{f}}_0(1)=1\), with \({\tilde{f}}_0'(0)=1={\tilde{f}}_0'(1)\),

  2. 2.

    \({\tilde{f}}_1(x)=1-x\),

such that \({\tilde{F}}\) is topologically transitive and that every pair of fiber expanding hyperbolic periodic orbits and every pair of fiber contracting hyperbolic periodic orbits in \(\Lambda _{\mathrm{core}}\) are homoclinically related, respectively. In particular, the unique measure of maximal entropy in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) is nonhyperbolic.

Note that in the above theorem this measure is also a measure of maximal entropy in \(\mathcal {M}_{\mathrm{erg}}(\Lambda )\), however we do not know if there is some hyperbolic measure of maximal entropy in \(\mathcal {M}_{\mathrm{erg}}(\Lambda )\).

Mutatis mutandi, we can perform a version of the map \({\tilde{F}}\) in \(\Sigma _3\times \mathbb {S}^1\) as in Remark 2.4.

Finally, in “Appendix A.4” we present an example with a unique measure of maximal entropy which is nonhyperbolic and supported on \(\Lambda _{\mathrm{ex}}\). However this example fails to be transitive.

Theorem 2.10

There are maps F defined as in (2.1) whose fiber maps \( f_0, f_1\) satisfy

  1. 1.

    \( f_0\) has (exactly) two fixed points \( f_0(0)=0\) and \( f_0(1)=1\), with \( f_0'(0)=1=f_0'(1)\),

  2. 2.

    \(f_1(x)=1-x\).

such that F is not topologically transitive and has a unique (hence ergodic) measure of maximal entropy supported on \(\Lambda _{\mathrm{ex}}\), which is nonhyperbolic.

One of the key properties of the class of examples in the above theorem is that \(f_0\) is conjugate to its inverse \(f_0^{-1}\) by \(f_1\). The proof of the result is based on an analysis of random walks on \({\mathbb {R}}\) and of somewhat different flavor.

3 Symmetric, mirror, and twin measures

Recalling well-known facts about shift spaces, we will see that there is a unique measure of maximal entropy for \(F|_{\Lambda _{\mathrm{ex}}}\) and we will deduce that, in the case this measure is hyperbolic, there is (at least) one “twin” measure in \(\Lambda _{\mathrm{core}}\) with the same (maximal) entropy. The latter is either hyperbolic with opposite sign of its exponent or nonhyperbolic.

Recall that on the full shift \(\sigma :\Sigma _2\rightarrow \Sigma _2\) there is a unique measure \({\widehat{\nu }}_{\mathrm{max}}\) of maximal entropy \(\log 2\) which is the \((\frac{1}{2},\frac{1}{2})\)-Bernoulli measure.

To study the structure of the invariant set \(\Lambda _{\mathrm{ex}}\), consider the “first level” rectangles \({\mathbb {C}}_k{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\{\xi \in \Sigma _2:\xi _0=k\}\) and the subsets

$$\begin{aligned} {\widehat{{\mathbb {C}}}}_{0_L}{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{\mathbb {C}}_0\times \{0\} ,\quad {\widehat{{\mathbb {C}}}}_{1_L}{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{\mathbb {C}}_1\times \{0\} ,\quad {\widehat{{\mathbb {C}}}}_{0_R}{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{\mathbb {C}}_0\times \{1\} ,\quad {\widehat{{\mathbb {C}}}}_{1_R}{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{\mathbb {C}}_1\times \{1\} , \end{aligned}$$

of \(\Sigma _2 \times [0,1]\). Consider the transition matrix A given by

$$\begin{aligned} A{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\left( \begin{matrix} 1&{}\quad 1&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 1&{}\quad 1\\ 0&{}\quad 0&{}\quad 1&{}\quad 1\\ 1&{}\quad 1&{}\quad 0&{}\quad 0 \end{matrix}\right) . \end{aligned}$$

This matrix codes the transitions between the symbols \(\{0_L,1_L,0_R,1_R\}\) modelling the transitions between the sets \({\widehat{{\mathbb {C}}}}_{0_L}, {\widehat{{\mathbb {C}}}}_{1_L}, {\widehat{{\mathbb {C}}}}_{0_R}, {\widehat{{\mathbb {C}}}}_{1_R}\) by the map F. More precisely, note that the restriction of \(F|_{\Lambda _{\mathrm{ex}}}\) is topologically conjugate to the subshift of finite type \(\sigma _A:\Sigma _A\rightarrow \Sigma _A\) by means of a map \(\varpi :\Lambda _{\mathrm {ex}}\rightarrow \Sigma _A\). Note that there is a unique measure of maximal entropy \(\nu ^{\mathrm{ex}}_{\mathrm{max}}\) for \(\sigma _A:\Sigma _A\rightarrow \Sigma _A\).Footnote 3 Note that \(h_{\nu ^{\mathrm{ex}}_{\mathrm{max}}}(\sigma )=\log 2\). Hence, by conjugation, the measure \(\mu ^{\mathrm{ex}}_{\mathrm{max}}=(\varpi ^{-1})_*\nu ^{\mathrm{ex}}_{\mathrm{max}}\) is the unique measure of maximal entropy \(\log 2\) for \(F:\Lambda _{\mathrm{ex}}\rightarrow \Lambda _{\mathrm{ex}}\).

We define the following projection

$$\begin{aligned} \begin{aligned} \Pi :\Sigma _A\rightarrow \Sigma _2, \quad \Pi (\ldots i_{-1}.i_0\ldots )&{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}(\ldots \xi _{-1}.\xi _0\ldots ),\\ \xi _k&{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{\left\{ \begin{array}{ll} 0&{}\quad \text {if }i_k\in \{0_L,0_R\}\\ 1&{}\quad \text {if }i_k\in \{1_L,1_R\}. \end{array}\right. } \end{aligned} \end{aligned}$$

It is immediate to check that

$$\begin{aligned} {\widehat{\nu }}_{\mathrm{max}} = \Pi _*\nu ^{\mathrm{ex}}_{\mathrm{max}}. \end{aligned}$$

We say that the symbols \(i_R\), \(j_L\) are the mirrors of \(i_L\) and \(j_R\), respectively, for i and j in \(\{0,1\}\) and denote \(i_R=\bar{i_L}\) and \(j_L=\bar{j_R}\). Given a sequence \(\xi =(\ldots \xi _{-1}.\xi _0\ldots )\in \Sigma _A\), we define by \(\bar{\xi }=(\ldots \bar{\xi _{-1}}.\bar{\xi _0}\ldots )\) the mirrored sequence of \(\xi \). Note that \({\bar{\xi }}\in \Sigma _A\). Given a subset \(B\subset \Sigma _A\), we denote by \({\bar{B}}{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\{{\bar{\xi }}:\xi \in B\}\) its mirrored set.

Now we are ready to define symmetric sets and measures.

Definition 3.1

(Symmetric sets and measures) A measurable set \(B\subset \Sigma _A\) is symmetric if \(B={\bar{B}}\). We say that B is symmetric\(\nu \)-almost surely if \(\nu (B\Delta {\bar{B}})=0\). A measure \(\nu \in \mathcal {M}(\Sigma _A)\) is symmetric if \(\nu (\bar{B})=\nu (B)\) for every \(B\subset \Sigma _A\). If a measure is not symmetric then we call it asymmetric. A measure \(\mu \in \mathcal {M}(\Lambda _{\mathrm {ex}})\) is symmetric if \(\varpi _*\mu \) is symmetric, otherwise we call it asymmetric.

We denote by \(\mathcal {M}_{\mathrm{erg}}^{\mathrm{sym}}(\Lambda _{\mathrm{ex}})\) and \(\mathcal {M}_{\mathrm{erg}}^{\mathrm{asym}}(\Lambda _{\mathrm{ex}})\) the sets of symmetric and asymmetric ergodic measures in \(\Lambda _{\mathrm{ex}}\), respectively.

We will use the following lemma.

Lemma 3.2

Let \(\nu \in \mathcal {M}(\Sigma _A)\) be a symmetric measure. Then any set \(B\subset \Sigma _A\) which is \(\nu \)-almost symmetric satisfies

$$\begin{aligned} \nu ((\Pi ^{-1}\circ \Pi )(B) ) = \nu (B). \end{aligned}$$

Proof

Indeed, by \(\nu \)-almost symmetry of B, setting \(C{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}B\cap {\bar{B}}\), \(D{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{\bar{B}}{\setminus }C\), and \(E{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}B{\setminus }C\), we have \(B=E\cup C\), \(\nu (C)=\nu ({\bar{C}})=\nu (B)=\nu ({\bar{B}})\), and \(\nu (D)=\nu (E)=0\). Hence \(\nu ({\bar{D}})=\nu (D)=0\). Observing that

$$\begin{aligned} (\Pi ^{-1}\circ \Pi )(B) =B\cup {\bar{B}} = E\cup C\cup D \end{aligned}$$

the claim follows. \(\square \)

Lemma 3.3

For every \(\nu \in \mathcal {M}_{\mathrm{erg}}(\Sigma _A)\), there exist at most one measure \({\bar{\nu }}\in \mathcal {M}_{\mathrm{erg}}(\Sigma _A)\), \({\bar{\nu }}\ne \nu \), such that \(\Pi _*{\bar{\nu }}=\Pi _*\nu \). There is no such measure if, and only if, \(\nu \) is symmetric.

Proof

It suffices to observe that the product \(\sigma \)-algebra of Borel measurable sets of \(\Sigma _A\) is generated by the semi-algebra generated by the family of all finite cylinder sets \(\{[i_k\ldots i_\ell ]\}\). Note also that the mirror \({\bar{C}}\) of a cylinder C in \(\Sigma _A\) is again a cylinder in \(\Sigma _A\). Now given \(\nu \in \mathcal {M}_{\mathrm{erg}}(\Sigma _A)\), define a measure \({\bar{\nu }}\) by setting \({\bar{\nu }}(C){\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\nu ({\bar{C}})\) for every cylinder C and extend it to the generated \(\sigma \)-algebra.

By definition, we immediately obtain that \(\Pi _*{\bar{\nu }}=\Pi _*\nu \) and that \(h_{{\bar{\nu }}}(\sigma _A)=h_\nu (\sigma _A)\).

To prove that \(\nu \) and \({\bar{\nu }}\) are the only ergodic measures satisfying \(\Pi _*{\bar{\nu }}=\Pi _*\nu \), by contradiction assume that there exists \({\widehat{\nu }}\in \mathcal {M}_{\mathrm{erg}}(\Sigma _A)\), \({\bar{\nu }}\ne {\widehat{\nu }}\ne \nu \) satisfying \(\Pi _*{\widehat{\nu }}=\Pi _*\nu \). Consider the measure \({\widetilde{\nu }}{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\frac{1}{2}(\nu +{\bar{\nu }})\). Note that \({\widetilde{\nu }}\) is symmetric. Also note that \(\Pi _*{\widetilde{\nu }}=\Pi _*{\widehat{\nu }}\). Finally note that \({\widetilde{\nu }}\) is singular with respect to \({\widehat{\nu }}\) and hence there is a set \(B\subset \Sigma _A\) satisfying \({\widetilde{\nu }}(B)=0={\widehat{\nu }}(B^c)\). Since \({\widetilde{\nu }}\) is symmetric, we have \({\widetilde{\nu }}({\bar{B}})={\widetilde{\nu }}(B)\). So we obtain \({\widetilde{\nu }}({\bar{B}}\triangle B)=0\) and hence B is \({\widetilde{\nu }}\)-almost symmetric. Hence, we have

$$\begin{aligned} \begin{aligned} 0&= {\widetilde{\nu }}(B) \\ (\text {by Lemma}~3.2)&= {\widetilde{\nu }}((\Pi ^{-1}\circ \Pi )(B)) = \Pi _*{\widetilde{\nu }}(\Pi (B)) \\ (\text {since }\Pi _*{\widetilde{\nu }}=\Pi _*{\widehat{\nu }}\,)&= \Pi _*{\widehat{\nu }}(\Pi (B)) = {\widehat{\nu }}((\Pi ^{-1}\circ \Pi )(B)) \ge {\widehat{\nu }}(B) =1, \end{aligned} \end{aligned}$$

a contradiction. This proves that \({\bar{\nu }}\) is uniquely defined.

By definition, \(\nu \) is symmetric if, and only if, \({\bar{\nu }}=\nu \). \(\square \)

Definition 3.4

(Mirror measure) We call the measure \({\bar{\nu }}\) provided by Lemma 3.3 the mirror measure of \(\nu \). We call the measure \(\varpi ^{-1}_*{\bar{\nu }}\in \mathcal {M}(\Lambda _{\mathrm {ex}})\) the mirror measure of \(\mu =\varpi ^{-1}_*\nu \) and denote it by \({\bar{\mu }}\).

The following is an immediate consequence of Lemma 3.3 and the uniqueness of the measure of maximal entropy.

Corollary 3.5

The measure of maximal entropy \(\mu ^{\mathrm{ex}}_{\mathrm{max}}\) is symmetric.

Lemma 3.6

If \({\bar{\mu }}\) is a mirror measure of \(\mu \in \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) then \(h_{{\bar{\mu }}}(F)=h_\mu (F)\). Moreover, we have

$$\begin{aligned} \chi ({\bar{\mu }})+\chi (\mu ) =N(0)\log ( f_0'(0)\cdot f_0'(1)) +N(1)\log ( f_1'(0)\cdot f_1'(1)), \end{aligned}$$

where \(N(0){\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\mu (\Sigma _2\times \{0\})\) and \(N(1){\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\mu (\Sigma _2\times \{1\})\).

Proof

Let \(\nu =\varpi _*\mu \). It suffices to observe that a sequence \(\xi \) is \(\nu \)-generic if, and only if, \({\bar{\xi }}\) is \({\bar{\nu }}\)-generic and to do the straightforward calculation. \(\square \)

Definition 3.7

Given an ergodic measure \(\mu \in \mathcal {M}_{\mathrm{erg}}(\Sigma _2\times [0,1])\), an ergodic measure \({\widetilde{\mu }}\in \mathcal {M}_{\mathrm{erg}}(\Sigma _2\times [0,1])\), \({\widetilde{\mu }}\ne \mu \), is called a twin measure of \(\mu \) if \(\pi _*{\widetilde{\mu }}=\pi _*\mu \).

Note that the above immediately implies that if \(\mu \in \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) is symmetric then all its twin measures are in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\).

Lemma 3.8

(Existence of twin measures) For every measure \(\lambda \in \mathcal {M}(\Sigma _2)\) there exist a measure \(\mu _1\in \mathcal {M}(\Sigma _2\times [0,1])\) satisfying \(\pi _*\mu _1=\lambda \) and \(\chi (\mu _1)\ge 0\) and a measure \(\mu _2\in \mathcal {M}(\Sigma _2\times [0,1])\) satisfying \(\pi _*\mu _2=\lambda \) and \(\chi (\mu _2)\le 0\).

Moreover, if \(\lambda \) was ergodic then \(\mu _1\) and \(\mu _2\) can be chosen ergodic.

Note that the measures \(\mu _1\) and \(\mu _2\) in the above lemma may coincide.

Proof

First observe that \(\lambda \in \mathcal {M}(\Sigma _2)\) is weak\(*\) approximated by measures \(\lambda _\ell \in \mathcal {M}(\Sigma _2)\) supported on periodic sequences.

For each such measure \(\lambda _\ell \) there exists a measure \(\mu _\ell \in \mathcal {M}(\Sigma _2\times [0,1])\) which is supported on a F-periodic orbit in \(\Sigma _2\times [0,1]\) and satisfies \(\pi _*\mu _\ell =\lambda _\ell \) and \(\chi (\mu _\ell )\ge 0\). Indeed, assume that \(\lambda _\ell \) is supported on the orbit of a periodic sequence \(\xi \in \Sigma _2\) of period n. Recall that the fiber maps \(f_0\) and \(f_1\) and hence the map \(f_\xi ^n\) preserve the boundary \(\{0,1\}\). Hence, this map \(f_\xi ^n\) has a fixed point \(x\in [0,1]\) satisfying \(|(f_\xi ^n)'(x)|\ge 1\). Now observe that the orbit of \((\xi ,x)\) is F-periodic of period n and taking the measure \(\mu _\ell \) supported on it we have \(\frac{1}{n}\log \,|(f_\xi ^n)'(x)|=\chi (\mu _\ell )\).

Now take any weak\(*\) accumulation point \(\mu \) of the sequence \((\mu _\ell )_\ell \). Note that by continuity of \(\pi _*\) we have \(\pi _*\mu =\lambda \).

If \(\lambda \) was ergodic, \(\mu \) might not be ergodic. However, any ergodic measure in the ergodic decomposition of \(\mu \) also projects to \(\lambda \) and hence there must exist one measure \(\mu '\) in this decomposition satisfying \(\chi (\mu ')\ge 0\).

The same arguments work for the case \(\chi (\cdot )\le 0\). \(\square \)

Corollary 3.9

For every hyperbolic symmetric ergodic measure \(\mu \in \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) there exists an ergodic twin measure \({\widetilde{\mu }}\in \mathcal {M}_{\mathrm{erg}}(\Sigma _2\times (0,1))\), \({\widetilde{\mu }}\ne \mu \), satisfying \(h_{{\widetilde{\mu }}}(F)=h_\mu (F)\).

Proof

Assume that \(\chi (\mu )>0\), the other case \(\chi (\mu )<0\) is analogous. Let \(\lambda {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\pi _*\mu \). By Lemma 3.8, there exists a twin measure \({\widetilde{\mu }}\in \mathcal {M}_{\mathrm{erg}}(\Sigma _2\times [0,1])\) of \(\mu \) satisfying \(\chi ({\widetilde{\mu }})\le 0\). Note that \(h_{\pi _*{\widetilde{\mu }}}(\sigma )\le h_{{\widetilde{\mu }}}(F)\) and \(h_{\pi _*\mu }(\sigma )\le h_\mu (F)\). On the other hand, by [21]

$$\begin{aligned} \max \{h_\mu (F),h_{{\widetilde{\mu }}}(F)\}\le & {} \sup _{m:\pi _*m=\pi _*\mu }h_m(F) = h_{\pi _*\mu }(\sigma )\\&+\int h_{\mathrm{top}}(F,\pi ^{-1}(\xi ))\,d\pi _*\mu (\xi ). \end{aligned}$$

Since \(\pi \) is 2–1, we have \(h_{\mathrm{top}}(F,\pi ^{-1}(\xi ))=0\) for every \(\xi \). Thus, we conclude \(h_\mu (F)=h_{{\widetilde{\mu }}}(F)=h_{\pi _*\mu }(\sigma )\).

By conjugation \(\varpi \) between \(F|_{\Lambda _{\mathrm{ex}}}\) and \(\sigma _A|_{\Sigma _A}\), there can be at most one other ergodic measure in \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) which project to the same measure on \(\Sigma _2\), namely \(\varpi ^{-1}_*{\bar{\nu }}\), where \(\nu =\varpi ^{-1}_*\mu \) and \({\bar{\nu }}\) is the mirror measure of \(\nu \). For symmetric \(\mu \), no such mirror exists. Hence, we must have \({\widetilde{\mu }}\in \mathcal {M}_{\mathrm{erg}}(\Sigma _2\times (0,1))\). \(\square \)

By Corollary 3.5, the above applies in particular to \(\mu _{\mathrm{max}}^{\mathrm{ex}}\).

Corollary 3.10

If the measure of maximal entropy \(\mu _{\mathrm{max}}^{\mathrm{ex}}\in \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) is hyperbolic then there exists an ergodic twin measure of maximal entropy \({\widetilde{\mu }}\in \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\) such that \(\chi ({\widetilde{\mu }})\chi (\mu _{\mathrm{max}}^{\mathrm{ex}})\le 0\).

Proof of Theorem 2.7

As recalled already, there is a unique measure of maximal entropy for \(F|_{\Lambda _{\mathrm{ex}}}\) and its Lyapunov exponents can be easily calculated. The fact that there may exist another measure of maximal entropy for \(F|_{\Lambda _{\mathrm{core}}}\) follows immediately from Corollary 3.10. \(\square \)

4 Approximations of boundary measures

This section discusses the approximation of measures in \(\mathcal {M}(\Lambda _{\mathrm{ex}})\) by (ergodic) measures in \(\mathcal {M}(\Lambda _{\mathrm{core}})\). In particular, we will complete the proof of Theorem 2.5. We will always work with the system satisfying hypotheses (H1) and (H2).

Remark 4.1

Recall again that \(\mathcal {M}\) equipped with the weak\(*\) topology it is a compact metrizable topological space [31, Chapter 6.1]. Recall that \(X\in \Sigma _2\times [0,1]\) is a generic point of a measure \(\mu \in \mathcal {M}(\Sigma _2\times [0,1])\) if the sequence \(\frac{1}{n}(\delta _X+\delta _{F(X)}+\cdots +\delta _{F^{n-1}(X)})\) converges to \(\mu \) in the weak\(*\) topology, where \(\delta _Y\) denotes the Dirac measure supported on Y. Recall that for every ergodic measure there exists a set of generic points with full measure. Note that \(F|_{\Lambda _{\mathrm{ex}}}\) is conjugate to a subshift of finite type (see Remark 2.1). Hence, by [27, Corollary to Theorem 4], for every invariant measure \(\mu \in \mathcal {M}(\Lambda _{\mathrm{ex}})\) there exists a \(\mu \)-generic point.

Given \(\delta \in (0,1/2)\), we consider the local distortion map

$$\begin{aligned} \Delta (\delta ) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\max _{i=0,1} \left\{ \max _{z\in [0,\delta ]}\left|\log \frac{|f_i'(z)|}{|f_i'(0)|}\right|, \max _{z\in [1-\delta ,1]}\left|\log \frac{|f_i'(z)|}{|f_i'(1)|} \right|\right\} . \end{aligned}$$
(4.1)

Note that \(\Delta (\delta )\rightarrow 0\) as \(\delta \rightarrow 0\). We state the following simple facts without proof.

Lemma 4.2

For every \(\delta \in (0,1/2)\) and every \(x\in [0,\delta ]\) we have

$$\begin{aligned} e^{-\Delta (\delta )} \le \frac{ f_i'(x)}{ f_i'(0)}, \frac{ f_i'(1-x)}{ f_i'(1)} \le e^{\Delta (\delta )} \end{aligned}$$

and

$$\begin{aligned} e^{-\Delta (\delta )} \le \frac{|f_i(x)-f_i(0)|}{|x||f_i'(0)|}, \frac{|f_i(1-x)-f_i(1)|}{|1-(1-x)||f_i'(1)|} \le e^{\Delta (\delta )} . \end{aligned}$$

Proposition 4.3

For every \(\mu \in \mathcal {M}(\Lambda _{\mathrm{ex}})\) satisfying \(\chi (\mu )=0\) there exists a sequence \((\mu _k)_k\subset \mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\) of measures supported on periodic orbits which converge to \(\mu \) in the weak\(*\) topology.

Proof

Let \(\mu \) be an invariant measure supported in \(\Lambda _{\mathrm{ex}}\) and satisfying the hypothesis \(\chi (\mu )=0\) and let \(X=(\xi ,x)\in \Lambda _{\mathrm{ex}}\) be a \(\mu \)-generic point. By Remark 4.1 such point indeed exists. Hence \(\chi (X)=0\). Note that \(\xi \) hence has infinitely many symbols 1 by our hypothesis \(f_0'(0)\ne 1\ne f_0'(1)\). Hence, without loss of generality, we can assume that \(x=1\).

Given the sequence \(\xi =(\ldots \xi _{-1}.\xi _0\xi _1\ldots )\), for \(n\ge 1\) define

$$\begin{aligned} \begin{aligned} p_n&{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{{\,\mathrm{card}\,}}\big \{i\in \{0,\ldots ,n-1\}:\xi _i=0, {{\,\mathrm{card}\,}}\{j<i:\xi _j=1\}\text { even}\big \},\\ q_n&{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{{\,\mathrm{card}\,}}\big \{i\in \{0,\ldots ,n-1\}:\xi _i=0, {{\,\mathrm{card}\,}}\{j<i:\xi _j=1\}\text { odd}\big \},\\ r_n&{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{{\,\mathrm{card}\,}}\big \{i\in \{0,\ldots ,n-1\}:\xi _i=1, {{\,\mathrm{card}\,}}\{j<i:\xi _j=1\}\text { even}\big \},\\ s_n&{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{{\,\mathrm{card}\,}}\big \{i\in \{0,\ldots ,n-1\}:\xi _i=1, {{\,\mathrm{card}\,}}\{j<i:\xi _j=1\}\text { odd}\big \}. \end{aligned} \end{aligned}$$

Note that \(n=p_n+q_n+r_n+s_n\). Let

$$\begin{aligned} \phi (n) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}p_n\log \,|f_0'(1)|+q_n\log \,|f_0'(0)|+r_n\log \,|f_1'(1)|+ s_n\log \,|f_1'(0)|. \end{aligned}$$
(4.2)

Observe that \(\phi (n)=\log \,|(f_\xi ^n)'(1)|\) and hence \(\chi (X)=0\) implies

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{\phi (n)}{n}=0. \end{aligned}$$

Let

$$\begin{aligned} \psi (n) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\max _{i=1,\ldots ,n}|\phi (i)|\end{aligned}$$

and note that

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{ \psi (n)}{n}=0. \end{aligned}$$
(4.3)

Let \(B=\{0,1\}\). Given a fiber point \(x_0\in (0,1)\) let us use the following notation of its orbit under the fiber dynamics determined by the sequence \(\xi \):

$$\begin{aligned} x_i {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}f_\xi ^i(x_0). \end{aligned}$$
(4.4)

Partially affine case To sketch the idea of the proof, assume for a moment that \(f_0|_{I_\delta }\) and \(f_1|_{I_\delta }\) are affine, where \(I_\delta =[0,\delta ]\cup [1-\delta ,1]\) for some small \(\delta >0\). Note that for given n and a point \(x_0 \in [1-\delta , 1)\) satisfying

$$\begin{aligned} x_i\in I_\delta \quad \text { for all }\quad i\in \{0,\ldots ,n-1\} \end{aligned}$$
(4.5)

we have

$$\begin{aligned} \frac{{{\,\mathrm{Dist}\,}}(x_{i+1},B)}{{{\,\mathrm{Dist}\,}}(x_i,B)} = e^{\phi (i+1)-\phi (i)}, \end{aligned}$$
(4.6)

where \({{\,\mathrm{Dist}\,}}(x,B)\) denotes the distance of x from a set B. Hence

$$\begin{aligned} \frac{{{\,\mathrm{Dist}\,}}(x_n,B)}{{{\,\mathrm{Dist}\,}}(x_0,B)} = e^{\phi (n)}. \end{aligned}$$

This implies

$$\begin{aligned} e^{-\psi (n)} \le \frac{{{\,\mathrm{Dist}\,}}(x_i,B)}{{{\,\mathrm{Dist}\,}}(x_0,B)} \le e^{\psi (n)} \quad \text { for all }\quad i\in \{0,\ldots ,n\}. \end{aligned}$$
(4.7)

Note that (4.5) is satisfied provided \(x_0\) was chosen to satisfy \({{\,\mathrm{Dist}\,}}(x_0,B)<\delta e^{-\psi (n)}\).

Note that \(e^{-\psi (n)}\) may not converge to 0. For this reason, let us choose

$$\begin{aligned} \delta (n) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\delta e^{-2\max \{\psi (n),\sqrt{n}\}}. \end{aligned}$$
(4.8)

Note that

$$\begin{aligned} \lim _{n\rightarrow \infty }\delta (n)e^{\psi (n)}=0. \end{aligned}$$
(4.9)

Let now n be a sufficiently large integer such that \({{\,\mathrm{card}\,}}\{j\le n-1:\xi _j=1\}\) is odd. Note that this implies \(f_\xi ^n\) is orientation reversing. Let N(n) be the smallest positive integer such that \(x_0{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}f_0^{N(n)}(1/2)\in [1-\delta (n),1)\). Note that

$$\begin{aligned} N(n) \sim |\log \delta (n)|, \end{aligned}$$
(4.10)

where the approximation is up to some universal multiplicative factor, independent on n. Note that \(f_0^{N(n)}\) is orientation preserving. We now apply the above arguments to the chosen point \(x_0\). We consider the sequence \((x_i)_{i=0}^n\) as defined in (4.4). First, note that \({{\,\mathrm{Dist}\,}}(x_0,B)\sim \delta (n)\) and with (4.7) we have

$$\begin{aligned} \delta (n) e^{-\psi (n)} \le x_i \le {{\,\mathrm{Dist}\,}}(x_0,B) e^{\psi (n)} \le \delta (n) e^{\psi (n)} \quad \text { for all }\quad i\in \{0,\ldots ,n\}. \end{aligned}$$

Further note that \(x_n\) by our choice of n is close to 0. Then let M(n) be the smallest positive integer such that \(f_0^{M(n)}(x_n)\ge 1/2\). Note that \(f_0^{M(n)}\) is orientation preserving. Note that

$$\begin{aligned} M(n) \sim |\log \delta (n)+\psi (n)|. \end{aligned}$$
(4.11)

Now consider the map \(g{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}f_0^{M(n)}\circ f_\xi ^n\circ f_0^{N(n)}\) and note that it reverses orientation. Hence, there exists a point y in the fundamental domain \([1/2,f_0(1/2))\) such that \(g(y)=y\). Note that by the estimates of N(n) and M(n) in (4.10) and (4.11) and our choice of \(\delta (n)\) in (4.8) and by (4.3) we have

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{N(n)+M(n)}{n} =0. \end{aligned}$$
(4.12)

We now consider the (invariant) measure \(\mu _{n,\delta }\) supported on the periodic orbit of the point \(Y=(\eta ,y)\), where \(\eta =(0^{N(n)}\xi _0\ldots \xi _{n-1}0^{M(n)})^{\mathbb {Z}}\). It remains to show that this measure is close to \(\mu \) in the weak\(*\) topology provided that n was big. Note that we can write \(\mu _{n,\delta }\) as

$$\begin{aligned} \begin{aligned} \mu _{n,\delta }=&\,\frac{N(n)}{N(n)+n+M(n)}\mu _1+\frac{n}{N(n)+n+M(n)}\frac{1}{n}\sum _{k=0}^{n-1}\delta _{F^{N(n)+k}(Y)} \\&+ \frac{M(n)}{N(n)+n+M(n)}\mu _2, \end{aligned} \end{aligned}$$

where \(\mu _1\) and \(\mu _2\) are some probability measures. Note that by (4.12) the first and the last term converges to 0 as n tends to \(\infty \). The second term is close to \(\mu \) because X was a \(\mu \)-generic point and the orbit piece \(\{F^{N(n)}(Y),F^{N(n)+1}(Y),\ldots ,F^{N(n)+n}(Y)\}\) is \(\delta (n)e^{\psi (n)}\)-close to the orbit piece \(\{X,F(X),\ldots ,F^{n}(X)\}\). Recalling (4.9), this completes the proof in the affine case.

General case In the nonaffine case the proof goes similarly. Applying Lemma 4.2, we choose the number \(\delta (n)\) in an appropriate way. First note that instead of (4.6) by this lemma we have

$$\begin{aligned} \frac{{{\,\mathrm{Dist}\,}}(x_{i+1},B)}{{{\,\mathrm{Dist}\,}}(x_i,B)} \le e^{\phi (i+1)-\phi (i)} e^{\Delta (\delta )}. \end{aligned}$$

Arguing as above, let now

$$\begin{aligned} \delta (n) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\delta e^{-2\max \{\psi (n),\sqrt{n}\}} e^{-n\Delta (\delta e^{-\sqrt{n}})}. \end{aligned}$$

Observe that with this choice, for every \(x_0\in [1-\delta (n),1)\) for every \(i\in \{1,\ldots , n\}\) we have

$$\begin{aligned} {{\,\mathrm{Dist}\,}}(x_i,B) \le \delta (n)e^{\phi (i)}e^{i\Delta (\delta e^{-\sqrt{n}})} \le \delta e^{-\sqrt{n}} \end{aligned}$$

provided that \({{\,\mathrm{Dist}\,}}(x_j,B)\le \delta e^{-\sqrt{n}}\) for all \(j\in \{0,\ldots ,i-1\}\). By induction, we will get that for all \(i\in \{0,\ldots , n\}\) we have

$$\begin{aligned} {{\,\mathrm{Dist}\,}}(x_i,B) \le \delta e^{-\sqrt{n}}. \end{aligned}$$

Note that with the above definition of \(\delta (n)\) the estimates of N(n) and M(n) in (4.10) and (4.11) remain without changes. And the rest of the proof is analogous to the partially affine case. \(\square \)

We now prove the converse to Proposition 4.3.

Proposition 4.4

For every \(\mu \in \mathcal {M}(\Lambda _{\mathrm{ex}})\) for which there exists a sequence \((\nu _k)_k\subset \mathcal {M}(\Lambda _{\mathrm{core}})\) of measures which converge to \(\mu \) in the weak\(*\) topology we have \(\chi (\mu )=0\).

The proof of the above proposition will be an immediate consequence of the following lemma. Recall the definition of \(\Delta (\cdot )\) in (4.1).

Lemma 4.5

There exist constants \(K_1,K_2>0\) such that for every \(\delta \in (0,1/2)\) and every measure \(\nu \in \mathcal {M}(\Lambda _{\mathrm{core}})\) we have

$$\begin{aligned} |\chi (\nu )|\le K_1\nu (\Sigma _2\times [\delta ,1-\delta ])+K_2\Delta (\delta ). \end{aligned}$$

Proof

Note that it is enough to prove the claim for \(\nu \in \mathcal {M}(\Lambda _{\mathrm{core}})\) being ergodic. Indeed, for a general invariant measure \(\nu \in \mathcal {M}(\Lambda _{\mathrm{core}})\) with ergodic decomposition \(\nu =\int \nu _\theta \,d\lambda (\nu _\theta )\), applying the above claim to any (ergodic) \(\nu _\theta \) in this decomposition we have

$$\begin{aligned} \chi (\nu ) = \int \chi (\nu )\,d\lambda (\nu _\theta ) \le K_1\nu (\Sigma _2\times [\delta ,1-\delta ]) + K_2\Delta (\delta ) \end{aligned}$$

with the analogous lower bound.

Let us hence assume that \(\nu \in \mathcal {M}(\Lambda _{\mathrm{core}})\) is ergodic. Since \(\nu \) is not supported on \(\Lambda _{\mathrm{ex}}\), there exists \(\delta '\in (0,\delta )\) such that \(\nu (\Sigma _2\times [\delta ',1-\delta '])>0\). Let \(X=(\xi ,x)\) be a generic point for \(\nu \) satisfying \(x\in [\delta ',1-\delta ']\) and consider the sequence of points \(x_i{\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}f_\xi ^i(x)\) for \(i\ge 0\). Since \(\nu \) is ergodic, there are infinitely many \(n\ge 0\) such that \(x_n\in [\delta ',1-\delta ']\) and hence

$$\begin{aligned} 2\delta ' \le \frac{{{\,\mathrm{Dist}\,}}(x_n,B)}{{{\,\mathrm{Dist}\,}}(x_0,B)} \le \frac{1}{2\delta '}, \end{aligned}$$
(4.13)

where \(B=\{0,1\}\). Because X is a generic point, given any \(\varepsilon \), for n large enough we have

$$\begin{aligned} \Big |\frac{1}{n}\log \,|(f_\xi ^n)'(x)|- \chi (\nu )\Big |\le \varepsilon \end{aligned}$$

and also

$$\begin{aligned} \Big |\frac{1}{n}{{\,\mathrm{card}\,}}\{i\in \{0,\ldots ,n-1\}:x_i\in [\delta ,1-\delta ]\} -\nu (\Sigma _2\times [\delta ,1-\delta ])\Big |\le \varepsilon .\nonumber \\ \end{aligned}$$
(4.14)

Applying Lemma 4.2, we have

$$\begin{aligned}&e^{\omega (i+1)}e^{-\Delta (\delta )} \le \frac{{{\,\mathrm{Dist}\,}}(x_{i+1},B)}{{{\,\mathrm{Dist}\,}}(x_i,B)} \le e^{\omega (i+1)}e^{\Delta (\delta )} \quad \text { if }x_i\in (0,\delta ]\cup [1-\delta ,1),\\&e^{\omega (i+1)}K^{-1} \le \frac{{{\,\mathrm{Dist}\,}}(x_{i+1},B)}{{{\,\mathrm{Dist}\,}}(x_i,B)} \le e^{\omega (i+1)}K \quad \text { if }x_i\not \in (0,\delta ]\cup [1-\delta ,1), \end{aligned}$$

where \(K>1\) is some universal constant and

$$\begin{aligned} \omega (i) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}{\left\{ \begin{array}{ll} \log \,|f'_{\xi _i}(0)|&{}\quad \text {if }x_{i-1}\in (0,\delta ],\\ \log \,|f'_{\xi _i}(1)|&{}\quad \text {if }x_{i-1}\in [1-\delta ,1),\\ 0 &{}\text { otherwise}. \end{array}\right. } \end{aligned}$$

By a telescoping sum, we have

$$\begin{aligned} \begin{aligned} \frac{{{\,\mathrm{Dist}\,}}(x_{n},B)}{{{\,\mathrm{Dist}\,}}(x_0,B)} = \prod _{i=0}^{n-1}\frac{{{\,\mathrm{Dist}\,}}(x_{i+1},B)}{{{\,\mathrm{Dist}\,}}(x_i,B)}. \end{aligned} \end{aligned}$$

We split the index set \(\{0,\ldots ,n-1\}=I_1\cup I_2\) according to the rule that \(x_i\in [\delta ,1-\delta ]\) for all \(i\in I_1\) and \(x_i\not \in [\delta ,1-\delta ]\) for all \(i\in I_2\). Let

$$\begin{aligned} \phi (n) {\mathop {=}\limits ^{\scriptscriptstyle \mathrm{def}}}\sum _{i=1}^n\omega (i) \end{aligned}$$

and note that this function was also used in the previous proof, see (4.2). By the above estimates, we hence have

$$\begin{aligned} e^{\phi (n)}K^{-{{\,\mathrm{card}\,}}I_1} \cdot e^{-\Delta (\delta ){{\,\mathrm{card}\,}}I_2} \le \frac{{{\,\mathrm{Dist}\,}}(x_{n},B)}{{{\,\mathrm{Dist}\,}}(x_0,B)} \le e^{\phi (n)}K^{{{\,\mathrm{card}\,}}I_1} \cdot e^{\Delta (\delta ){{\,\mathrm{card}\,}}I_2}. \end{aligned}$$

This implies

$$\begin{aligned} e^{\phi (n)} \le \frac{{{\,\mathrm{Dist}\,}}(x_{n},B)}{{{\,\mathrm{Dist}\,}}(x_0,B)} K^{{{\,\mathrm{card}\,}}I_1} \cdot e^{\Delta (\delta ){{\,\mathrm{card}\,}}I_2}. \end{aligned}$$
(4.15)

Again applying Lemma 4.2, if \({{\,\mathrm{Dist}\,}}(x_i,B)<\delta \) then we have

$$\begin{aligned} |(f_{\xi _{i+1}})'(x_i)|\le e^{\omega (i+1)} e^{\Delta (\delta )} \end{aligned}$$

and if \({{\,\mathrm{Dist}\,}}(x_i,B)\ge \delta \) then we have

$$\begin{aligned} |(f_{\xi _{i+1}})'(x_i)|\le e^{\omega (i+1)} L \end{aligned}$$

for some universal \(L>1\). Hence, decomposing the orbit piece \((x_i)_{i\ge 0}^{n-1}\) as above into index sets \(I_1\) and \(I_2\), we obtain

$$\begin{aligned} |(f_\xi ^n)'(x)|\le e^{\phi (n)}L^{{{\,\mathrm{card}\,}}I_1}\cdot e^{\Delta (\delta ){{\,\mathrm{card}\,}}I_2} \end{aligned}$$

with the analogous lower bound.

By (4.14), we have \({{\,\mathrm{card}\,}}I_1\le n(\nu (\Sigma _2\times [\delta ,1-\delta ])+\varepsilon )\).

Substituting the estimate for \(e^{\phi (n)}\) in (4.15) we obtain

$$\begin{aligned} |(f_\xi ^n)'(x)|\le \frac{{{\,\mathrm{Dist}\,}}(x_{n},B)}{{{\,\mathrm{Dist}\,}}(x_0,B)} (KL)^{{{\,\mathrm{card}\,}}I_1} \cdot e^{2\Delta (\delta ){{\,\mathrm{card}\,}}I_2} \le \frac{1}{2\delta '} (KL)^{{{\,\mathrm{card}\,}}I_1} \cdot e^{2\Delta (\delta ){{\,\mathrm{card}\,}}I_2}, \end{aligned}$$

where we also used (4.13). Hence

$$\begin{aligned} \frac{1}{n}\log \,|(f_\xi ^n)'(x)|\le \frac{1}{n}|\log (2\delta ')|+\log (KL)\big (\nu (\Sigma _2\times [\delta ,1-\delta ])+\varepsilon \big ) +2\Delta (\delta ), \end{aligned}$$

with the analogous lower bound. Since \(|\log \,|(f_\xi ^n)'(x)|/n-\chi (\nu )|\le \varepsilon \), passing \(n\rightarrow \infty \) and then \(\varepsilon \rightarrow 0\) this ends the proof of the lemma. \(\square \)

Proof of Theorem 2.5

Item 1 is a well-known fact, see for example [27, Proposition 2 item (a)]. This fact implies that \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) is a Poulsen simplex (see [23] or in the particular case of the shift space [28]). Item 4 is then an immediate consequence from the facts that \(\mu \mapsto \chi (\mu )\) is continuous and that the Dirac measure on \((0^{\mathbb {Z}},0)\) has Lyapunov exponent \(\log \,f_0'(0)>0\) and the Dirac measure on \((0^{\mathbb {Z}},1)\) has Lyapunov exponent \(\log \,f_0'(1)<0\) together with the fact that \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{ex}})\) is path-connected.

Item 2 follows from Proposition 4.4.

Item 3 follows from Proposition 4.3.

Let us now assume (H1), (H2’), (H3), and (H4). By Lemmas A.4 and A.6 there exist hyperbolic periodic points in \(\Lambda _{\mathrm{core}}\) with positive and negative exponent, respectively. This proves item 5. By Proposition 2.2 we can apply 5.1. This implies item 6. \(\square \)

5 The core measures

In this section we will investigate a bit further the topological structure of \(\mathcal {M}_{\mathrm{erg}}(\Lambda _{\mathrm{core}})\). The overall hypotheses are again (H1) and (H2), and we will discuss further additional conditions under which we are able to say more than in the previous sections.

Proposition 5.1

Assume that every pair of fiber expanding hyperbolic periodic orbits in \(\Lambda _{\mathrm{core}}\) are homoclinically related. Then the set \(\mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}})\) is arcwise-connected. The analogous result holds true for fiber contracting hyperbolic periodic orbits in \(\Lambda _{\mathrm{core}}\) and the set \(\mathcal {M}_{\mathrm{erg,<0}}(\Lambda _{\mathrm{core}})\).

A map F whose fiber maps \(f_0,f_1\) satisfy the hypotheses (H1), (H2’), (H3), and (H4) will satisfy the hypotheses of the above proposition.

We will several times refer to a slightly strengthened version of [6, Proposition 1.4] which, in fact, is contained in its proof in [6] and which can be seen as an ersatz of Katok’s horseshoe construction (see [20, Supplement S.5]) in the \(C^1\) dominated setting. We formulate it in our setting. Note that to guarantee that the approximating periodic orbits are indeed contained in \(\Sigma _2\times (0,1)\) it suffices to observe that in the approximation arguments one can consider any sufficiently large (in measure \(\mu \)) set and hence restrict to points which are uniformly away from the “boundary” \(\Sigma _2\times \{0,1\}\). Indeed, the projection to [0, 1] of the support of \(\mu \) can be the whole interval [0, 1] but it does not “concentrate” in \(\{0,1\}\).

Lemma 5.2

Let \(\star \in \{<0,>0\}\) and \(\mu \in \mathcal {M}_{\mathrm{erg,\star }}(\Lambda _{\mathrm{core}})\).

Then for every \(\rho \in (0,1)\) there exist \(\alpha >0\) and a set \(\Gamma _\rho \subset \Sigma _2\times (2\alpha ,1-2\alpha )\) and a number \(\delta =\delta (\rho ,\mu )>0\) such that \(\mu (\Gamma _\rho )>1-\rho \) and for every point \(X\in \Gamma _\rho \) there is a sequence \((p_n)_n\subset \Sigma _2\times (\alpha ,1-\alpha )\) of hyperbolic periodic points such that:

  • \(p_n\) converges to X as \(n\rightarrow \infty \);

  • the invariant measures \(\mu _n\) supported on the orbit of \(p_n\) are contained in \(\mathcal {M}_{\mathrm{erg,\star }}(\Lambda _{\mathrm{core}})\) and converge to \(\mu \) in the weak\(*\) topology;

Proof of Proposition 5.1

Similar results were shown before, though in slightly different contexts (see [17] and [12, Theorem 3.2]). For completeness, we sketch the proof.

Assume that \(\mu ^0,\mu ^1\in \mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}})\). By Lemma 5.2, \(\mu ^i\) is accumulated by a sequence of hyperbolic periodic measures \(\nu _n^i\in \mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}})\) supported on the orbits of fiber expanding hyperbolic periodic points \(P_n^i\in \Lambda _{\mathrm{core}}\), \(i=0,1\). Since, by hypothesis, \(P_1^0\) and \(P_1^1\) are homoclinically related, there exists a horseshoe \(\Gamma _1^{0,1}\subset \Lambda _{\mathrm{core}}\) containing these two points. Hence, since \(\mathcal {M}(\Gamma _1^{0,1})\) is a Poulsen simplex [23, 28], there is a continuous arc \(\mu _0:[1/3,2/3]\rightarrow \mathcal {M}_{\mathrm{erg,>0}}(\Gamma _1^{0,1})\subset \mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}})\) joining the measures \(\nu _1^0\) and \(\nu _1^1\). For any pair of measures \(\nu _n^0,\nu _{n+1}^0\), the same arguments apply and, in particular, there exists a continuous arc \(\mu _n^0:[1/3^{n+1},1/3^n]\rightarrow \mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}})\) joining the measure \(\nu _n^0\) with \(\nu _{n+1}^0\). Using those arcs and concatenating their domains (or appropriate parts of), we can construct an arc \({\bar{\mu }}_n^0:[1/3^{n+1},1/3]\) joining \(\nu _{n+1}^0\) and \(\nu _{1}^0\). The same applies to the measures \(\nu _n^1\), defining arcs \({\bar{\mu }}_n^1 :[1-1/3^n,2/3]\rightarrow \mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}})\) joining \(\nu _{n+1}^1\) and \(\nu _{1}^1\). Defining \(\mu _\infty |_{(0,1)}:(0,1)\rightarrow \mathcal {M}_{\mathrm{erg,>0}}(\Lambda _{\mathrm{core}})\) by concatenating (appropriate parts of) the domains of those arcs, we complete the definition of the arc \(\mu _\infty \) by letting \(\mu _\infty (0)=\lim _{n\rightarrow \infty }{\bar{\mu }}_n^0(1/3^n)\) and \(\mu _\infty (1)=\lim _{n\rightarrow \infty }{\bar{\mu }}_n^1(1-1/3^n)\), joining \(\mu ^0\) and \(\mu ^1\). Note that in the last step we assume that \(\mu ^1,\mu ^2\) do not belong to the image of \(\mu _\infty \), if one of these measures belongs it is enough to cut the domain of definition of \(\mu _\infty \) appropriately. \(\square \)