Research article

Phase transition for piecewise linear fibonacci bimodal map

  • Received: 22 September 2022 Revised: 17 January 2023 Accepted: 19 January 2023 Published: 03 February 2023
  • MSC : 37E05, 37A10

  • In this paper we concern with the Fibonacci bimodal maps. We first study the topological properties of the Fibonacci bimodal maps in the context of kneading map and give an equivalent description of Fibonacci combinatorics. Then we construct a one-parameter family fλ of countably piecewise linear Fibonacci bimodal maps depending on the parameter λ which are all odd functions. By a random walk argument on its induced Markov map, we will show that a phase transition occurs from Lebesgue conservative to Lebesgue dissipative behaviors.

    Citation: Haoyang Ji, Wenxiu Ma. Phase transition for piecewise linear fibonacci bimodal map[J]. AIMS Mathematics, 2023, 8(4): 8403-8430. doi: 10.3934/math.2023424

    Related Papers:

    [1] Binbin Miao, Chongbin Xu, Caidi Zhao . Statistical solution and piecewise Liouville theorem for the impulsive discrete Zakharov equations. AIMS Mathematics, 2022, 7(5): 9089-9116. doi: 10.3934/math.2022505
    [2] Najmeddine Attia . Some remarks on recursive sequence of fibonacci type. AIMS Mathematics, 2024, 9(9): 25834-25848. doi: 10.3934/math.20241262
    [3] Can Kızılateş, Halit Öztürk . On parametric types of Apostol Bernoulli-Fibonacci, Apostol Euler-Fibonacci, and Apostol Genocchi-Fibonacci polynomials via Golden calculus. AIMS Mathematics, 2023, 8(4): 8386-8402. doi: 10.3934/math.2023423
    [4] Ning Cui, Junhong Li . A new 4D hyperchaotic system and its control. AIMS Mathematics, 2023, 8(1): 905-923. doi: 10.3934/math.2023044
    [5] Yanshou Dong, Junfang Zhao, Xu Miao, Ming Kang . Piecewise pseudo almost periodic solutions of interval general BAM neural networks with mixed time-varying delays and impulsive perturbations. AIMS Mathematics, 2023, 8(9): 21828-21855. doi: 10.3934/math.20231113
    [6] Abdon Atangana, Seda İğret Araz . Piecewise differential equations: theory, methods and applications. AIMS Mathematics, 2023, 8(7): 15352-15382. doi: 10.3934/math.2023785
    [7] Rajiniganth Pandurangan, Sabri T. M. Thabet, Imed Kedim, Miguel Vivas-Cortez . On the Generalized ¯θ(t)-Fibonacci sequences and its bifurcation analysis. AIMS Mathematics, 2025, 10(1): 972-987. doi: 10.3934/math.2025046
    [8] Jingyuan Zhang, Ruikun Zhang, Xue Lin . A stabilized multiple time step method for coupled Stokes-Darcy flows and transport model. AIMS Mathematics, 2023, 8(9): 21406-21438. doi: 10.3934/math.20231091
    [9] Tingting Du, Li Wang . On the power sums problem of bi-periodic Fibonacci and Lucas polynomials. AIMS Mathematics, 2024, 9(4): 7810-7818. doi: 10.3934/math.2024379
    [10] F. Z. Geng . Piecewise reproducing kernel-based symmetric collocation approach for linear stationary singularly perturbed problems. AIMS Mathematics, 2020, 5(6): 6020-6029. doi: 10.3934/math.2020385
  • In this paper we concern with the Fibonacci bimodal maps. We first study the topological properties of the Fibonacci bimodal maps in the context of kneading map and give an equivalent description of Fibonacci combinatorics. Then we construct a one-parameter family fλ of countably piecewise linear Fibonacci bimodal maps depending on the parameter λ which are all odd functions. By a random walk argument on its induced Markov map, we will show that a phase transition occurs from Lebesgue conservative to Lebesgue dissipative behaviors.



    The dynamical properties of unimodal interval maps have been extensively studied recently, for backgrounds and history, see for example [14,17] and the references therein. Major breakthroughs were the complete solution of Milnor's attractor problem for smooth unimodal maps with small critical order and the proof of Palis' conjecture for real analytic unimodal maps.

    From a modern perspective, it is essential (for both of the two problems) to study the local geometry around the critical point of a non-renormalizable map. A useful tool is the so-called principal nest. Let f:[0,1][0,1] be a (smooth) unimodal map. Let I0=(ˆq,q) where q(0,1) is the orientation reversing fixed point of f and f(ˆq)=f(q). Let I0I1I2 be the principal nest of f. The scaling factor of f is defined as μn:=|In+1|/|In|. One of the main studies in interval dynamics is the classification of the asymptotic property of λn when n tends to . It is well-known, see [12], that decay of geometry property, which means μni decreases to 0 exponentially fast for a subsequence {ni}, excludes the existence of wild attractor. It was proved in [6] that if μn is sufficiently small for all n large enough, then f has an acip. On the other hand, a wild attractor occurs for a unimodal map f with very large critical order and specific combinatorics including Fibonacci combinatorics. We remark here that a wild attractor is a Cantor attractor of a non-renormalizable map with full Lebesgue measure attraction of basin but is of the first Baire category. It is a metric attractor, but fails to be a topological one in the sense of Milnor.

    Fibonacci unimodal maps can be treated as the prototype of non-renormalizable unimodal maps. A unimodal map f has Fibonacci combinatorics if the first return time sn of the critical point c to the principal nest In coincides with that Fibonacci numbers. Fibonacci combinatorics were first introduced by Branner and Hubbard [1] for cubic polynomials with one critical point escaping to infinity, and by Hofbauer and Keller [7] for unimodal maps with slow recurrence. Existence of wild attractor for Fibonacci unimodal map with sufficiently high critical order was verified in [3]. It is quite interesting that the geometry and metic properties of a Fibonacci unimodal map will change as the critical order grows. To be precise, let f be the unique map from the family xa(1|2x1|) with Fibonacci combinatorics. It is now well-known that: if (1,2], then f possess decay of geometry and admits an absolutely continuous (respect to Lebesgue measure) invariant probability measure (acip for short) [10,11]; if (2,), then f possess bounded geometry in the sense that μn is uniformly bounded from above and below by constants depending only on [8,12]; if is sufficient large, then f has a wild attractor and a dissipative σ-finite acim, [3] (Here an acim means an absolutely continuous (respect to Lebesgue measure) invariant measure).

    But what about multimodal maps? Even for the simplest case, the cubic bimodal maps, is rarely known. A conceptual dichotomy was due to Shen [15] where he showed that a polynomial possesses either 'decay of geometry' or 'essentially bounded geometry'. A concrete example of a non-renormalizable bimodal cubic polynomial with bounded geometry was given by Świtiek and Vargas in [18]. Principal nest is a useful tool when studying geometric properties of interval maps, but seems not convenient to treat about metric problems in the multimodal case. Unlike in the unimodal case, the scaling factors fail to give distortion control of the first return map of the principal nest, because the critical branches may fold many times. However, in [19] Vargas constructed the Fibonacci bimodal map by use of the natural symmetry of bimodal maps using twin principal nest (see subsection 2.1). It was proved by Vargas (unpublished) and the authors [9] that the cubic Fibonacci polynomial possesses decay of geometry. Actually, we proved such a property for a general class of maps.

    A natural question arises now: What are the metric properties of the Fibonacci bimodal map when the critical order grows? It is quite a difficult question, for we do not even have 'a priori bounds' (or real bounds) in this settings. We cannot adapt the proofs in [9] to do so because the complex tools used there rely heavily on the fact that the two critical points are locally quasi-quadratic. Despite difficulties we study Fibonacci bimodal maps under a restrictive condition in this paper. We aim to give a somewhat hypothetical picture of smooth Fibonacci bimodal maps, in particular when the critical order changes. We first study the combinatorics of the Fibonacci bimodal map in the context of kneading map. We also give an equivalent description of Fibonacci combinatorics. And by doing this we can construct an induced Markov map over such a map naturally. The metric property of the original map depends on the conservative and dissipative behaviors of the induced Markov map. To avoid the difficulty of distortion control, we construct a countably piecewise linear bimodal map f with Fibonacci combinatorics which is also an odd function. By carefully choosing the slope, we show that the induced Markov map F is also piecewise linear on each branches in forms of [zj1,zj] or [yj,yj1], where zj and yj are closest precritical points. We let zj tends to the critical points in a geometric manner (|zjc|=O(λj)) so that f depends solely on the single parameter λ(0,1). The change of λ will be reflected on the change of the critical order. Then we will show that the one-parameter family fλ has a phase transition from Lebesgue conservative to dissipative behaviors.

    The Main Theorem is stated as follows. Note that in this non-differentiable setting, the critical order is defined by the property that C1|xc||f(x)f(c)|C|xc| for some C>0 and all x[1,0]. We use a single symbol here since the critical order of the two critical points are equal. Our theorem yields the precise values of critical orders =(λ), where each of the different behaviors occurs.

    Main Theorem. The piecewise linear bimodal map fλ (i.e., with a geometric manner defined in Sect. 3) satisfies the following properties:

    (1) The critical order =3+2log(1λ)logλ.

    (2) If λ(12,1), i.e., >5, then fλ has a wild attractor.

    (3) If λ[23+5,12], i.e., 45, then fλ has no wild attractor, but an infinite σ-finite acim.

    (4) If λ(0,23+5), i.e., 3<<4, then fλ has an acip.

    This paper is organized as follows. In Section 2 we study the topological properties of the Fibonacci bimodal map in the context of kneading map. We also give an equivalent description of Fibonacci combinatorics and construct the induced Markov map. In Section 3 we construct the piecewise linear bimodal map f which is an odd function and has the Fibonacci combinatorics. In Section 4 we let the closest precritical points tend to the critical points in a geometric manner and turn the original system into a one-parameter family. In Section 5 we use a random walk argument to prove the Main Theorem.

    Denote I=[1,1]. A continuous map f:II is called bimodal if f({1,1})={1,1} and f has exactly one local maximum and one local minimum in (1,1). The two extreme points specified by c<d are called turning points and f is strictly monotone on subintervals determined by these points. If the points {1,1} are fixed then we say that the bimodal map f is positive and in the case that these points are permuted we say that f is negative.

    Definition 2.1. A bimodal map f is called combinatorially symmetric if there exists an orientation-reversing homeomorphism h:II such that hf=fh.

    Let B denote the collection of bimodal maps f:II which have no wandering intervals and no attracting periodic cycles. Let B+ and B denote, respectively, the subset of positive and negative bimodal maps from class B. Let B denote the collection of maps from B which are combinatorially symmetric. If a bimodal map fB, then there is a fixed point p between c and d with three preimages {p,p1,p2} specifying by p1<p<p2. Define I0=(p1,p), J0=(p,p2).

    Assume that both c and d are recurrent and define

    I0I1I2{c} and J0J1J2{d}

    inductively such that, for k1, the intervals Ik and Jk are components of the domain of the first return map ϕk to Ik1Jk1, which are called critical domains of ϕk. The critical return times sk are defined by ϕk(c)=fsk(c) and ϕk(d)=fsk(d). The first return map ϕk is called central return if ϕk(c)IkJk or ϕk(d)IkJk; otherwise ϕk is called non-central return.

    Definition 2.2. A bimodal map fB has Fibonacci combinatorics if and only if sk is well-defined for all k1 and coincide with the Fibonacci sequence 2,3,5,.

    According to [19], Fibonacci bimodal map is combinatorially symmetric and has no central return. Then for each k1, let CkIk1 and DkJk1 be the return domains intersecting {ϕk(c),ϕk(d)}, which are called post critical domains. Note that the post critical branches ϕk|Ck and ϕk|Dk are monotone and onto. The Fibonacci combinatorics implies some constraints on the position of the post-critical domains and their images. This leads us to consider the three types of first return map ϕk below:

    ● Type A: if ϕk(Ck)=Jk1, ϕk(Ik)Jk1 and ϕk(Dk)=Ik1, ϕk(Jk)Ik1;

    ● Type B: if ϕk(Ck)=Jk1, ϕk(Ik)Ik1 and ϕk(Dk)=Ik1, ϕk(Jk)Jk1;

    ● Type C: if ϕk(Ck)=Ik1, ϕk(Ik)Jk1 and ϕk(Dk)=Jk1, ϕk(Jk)Ik1.

    Figure 1 illustrates a possible position but still without the orientation. Furthermore, the Fibonacci combinatorics shows that the sequence of first return maps ϕ1,ϕ2,ϕ3, exhibits a specific sequence of types as in Fact 2.1 below. This together with an analysis of the orientation and the precise position of the branches of their first return maps will determine the topological properties of a Fibonacci bimodal map. So for each type A, B and C, subdivide in subtypes Aij, Bij and Cij with i,j{+,}. Where i=+ or i= if the post critical branches of ϕk are orientation-preserving or orientation-reversing; and j=+ or j= if ϕk is local maximal or minimal at c.

    Figure 1.  Examples of types A B C.

    Fact 2.1. [9,19] If fB has Fibonacci combinatorics, then for all k1 the following holds:

    (1) ϕk(IkJk)IkJk.

    (2) Denote s0=1, s1=2 and sn+1=sn+sn1 for n1. Then ϕk|Ik and ϕk|Jk equal to fsk, while ϕk|Ck and ϕk|Dk equal to fsk1. Furthermore, ϕk|IkJk=ϕ2k1 while ϕk|CkDk=ϕk1.

    (3) The sequence ϕ1,ϕ2,ϕ3, of the first return maps exhibits the sequence of types

    A++B+CA+B+C+A+BC+AB++C++A++

    or

    C+AB++C++A++B+CA+B+C+A+BC+

    depending respectively on fB+ or fB. See Figure 2.

    Figure 2.  Types A++, B+, C, A+, B+, C+, A+, B, C+, A, B++, C++.

    The definition of Fibonacci combinatorics for bimodal maps was stated in the sense of generalized renormalization. In the unimodal case, the Fibonacci unimodal map can be understood in the context of both generalized renormalization and kneading theory (including kneading invariants and kneading map). Kneading map (which was introduced by Hofbauer and Keller) and the associated Hofbauer tower construction are very useful tools to study metric properties for unimodal maps. But these tools seem to be of no use in multimodal case. For example, it is rather difficult to give a proper definition. However, since we only consider maps which are combinatorially symmetric, it is still possible to describe the kneading map for such maps, especially for specific combinatorics. For the Fibonacci bimodal map, we will give an equivalent description.

    Let I0 and J0 be defined as above. For xI0 or xJ0, the involution of x, denoted ˆx, is defined as the point in I0J0 and such that f(x)=f(ˆx).

    We may assume that f(c)>d and f(d)<c when fB+ and f(c)<c and f(d)>d when fB. For otherwise, one can show that fj(c),fj(d)(fj1(c),fj1(d)) for all j1, hence the orbits of c and d both converge to the fixed point p. Then either c or d has two preimages inside I0=(p1,p). Denote Crit:={c,d}. Define S0:=1, define

    z0:=f1(Crit)(p1,c) and y0:=f1(Crit)(d,p2).

    Define inductively,

    Sk+1:=min{n>Sk|fn(Crit)(zk,c)}

    and

    zk+1:=fSk+1(Crit)(zk,c) and yk+1:=fSk+1(Crit)(d,yk).

    Sk are called cutting times while zk and yk are called the closest precritical points. If Sk<nSk+1, then (zk,c), (c,ˆzk), (ˆyk,d) and (d,yk) are maximal intervals on which fn is monotone. Let Ak:=(zk1,zk) and Bk:=(yk,yk1) for k1. Let ˆAk:=(ˆzk,ˆzk1) and ˆBk:=(ˆyk1,ˆyk).

    By construction, fSk1(zk) has the form fm(Crit) and is contained in (zr,ˆzr) or (ˆyr,yr) for some r0, see Figure 3. Hence SkSk1 is still a cutting time. The kneading map of a bimodal map fB is defined as

    Q:NN so that Sk=Sk1+SQ(k).
    Figure 3.  One possible position for zk under fSk1.

    It follows that

    fSk1(zk){zQ(k),ˆzQ(k),yQ(k),ˆyQ(k)}.

    The kneading map determines the combinatorics of the map. By Figure 3 and the construction of the closest preimages,

    {fSk1(c),fSk1(d)}AQ(k)ˆAQ(k)BQ(k)ˆBQ(k). (1)

    This is true for all k1. If Q(k)=0, then fSk1(c) and fSk1(d) are outside (z0,ˆz0)(ˆy0,y0). Notice also that fSk maps (zk1,c)(b,fSk(c)) is monotone and onto, where b{fSQ(k)(c),fSQ(k)(d)}, and its image contains one turning point c or d. See Figure 4 for one possible case.

    Figure 4.  One possible position for zk1 under fSQ(k).

    Lemma 2.1. If f has no periodic attractor, then Q(k)<k for all k1.

    Proof. Consider fSk1 on (zk1,c). We may assume that fSk1:(zk1,c)(d,fSk1(c)) is monotone and increasing. Then fSk1(c)(yQ(k),yQ(k)1). If Q(k)>k, then fSk1(c)(d,yk1). Hence fSk1fSk1 maps (zk1,c) into itself, yields a periodic attractor. Contradiction.

    Lemma 2.1 also implies the fact that if f has no periodic attractor, then the cutting times are well-defined for all k1.

    In this subsection we will state some topological properties of the Fibonacci bimodal map in terms of kneading map. For simplicity, we only focus on positive Fibonacci bimodal maps. The case of negative maps is analogous.

    Suppose fB+ has Fibonacci combinatorics. Let {In}n0 and {Jn}n0 be its twin principal nest. Denote In:=(un,ˆun) with un<c and Jn:=(ˆvn,vn) with d<vn. Note that ˆu0=ˆv0=p, u0=p1 and v0=p2.

    Proposition 1. The cutting times Sk and the critical return times sk are coincident, the kneading map is clarified as Q(k)=max{k2,0}. Moreover, for k0,

    (1) zk(uk,uk+1) and yk(vk+1,vk).

    (2) If k0,1mod3, then zkfSk(d) and ykfSk(c).

    (3) If k2mod3, then zkfSk(c) and ykfSk(d).

    Proof. We prove this lemma by induction. For k=0, statement (2) holds by the chosen of z0 and y0. By Fact 2.1 ϕ1 is of type A++, then fs1=f2:u1p. In particular, C1=(u0,u1) and fs0=f:(u0,u1)J0 is monotone and onto. Hence statement (1) holds.

    We now assume that the lemma holds for k and prove that it holds for k+1. Without loss of generality, we may assume that ϕk is of type C++ and ϕk+1 is of type A++. Since sk=Sk and ϕk+1|Ck+1=fsk with Ck+1 is on the left side of c, we have zkCk+1. Since ϕk+1(Ik+1)=fsk+1(Ik+1)Jk+1d, sk+1 is a cutting time. Now it suffices to show that there is no cutting time between sk and sk+1. Let sk<n<sk+1, consider fn|(zk,c). Firstly note that fsk:(zk,c)(d,fsk(c)) where fsk(c)DkJk1. Since ϕk1|Jk1=fsk1 is the first return map, fj(Jk1)(Ik1Jk1)= for 1jsk1. This implies fn((zk,c)){c,d}= for all sk<n<sk+1. Hence sk+1=Sk+1. Since ϕk+1(Ik+2)Dk+1, zk+1Ik+1Ik+2. This proves statement (1). Statements (2) and (3) follow from Fact 2.1 immediately. This finishes the proof.

    Combining this with Fact 2.1, we can prove the following lemma by induction.

    Lemma 2.2. For k2, we have

    (1) If k0,1mod3, then fSk1(zk){yk2,ˆyk2} and fSk1(yk){zk2,ˆzk2};

    (2) If k2mod3, then fSk1(zk){zk2,ˆzk2} and fSk1(yk){yk2,ˆyk2}.

    Moreover, for k1, we have

    (3) If k0,1,2,4,9,11mod12, then fSk is increasing on (zk,c) and (d,yk);

    (4) If k3,5,6,7,8,10mod12, then fSk is decreasing on (zk,c) and (d,yk).

    Actually, by Eq (1) and Proposition 1 we can state the position of fSk(c) and fSk(d) in more details.

    Since f:(1,u0)(1,ˆu0)=(1,p) is increasing and onto, c has a preimage z1(1,u0). Similarly, d has a preimage y1(v0,1).

    Figure 5 illustrates one possible possition for points and their images under fSk11 for k2mod12.

    Figure 5.  Points and their images under fSk11 for k2mod12. This corresponds to the case when ϕk1 is of type A++.

    Lemma 2.3. Suppose that fB+ has Fibonacci combinatorics. Then f satisfies the following starting condition:

    (1) f(c)(y1,1) and f(d)(1,z1);

    (2) f2(c),f2(d)I0J0;

    (3) f4(c),f4(d)I0J0.

    Proof. By Fact 2.1, the first return map ϕ1 is of type A++. Then ϕ1:I1J0 is local maximal at c with ϕ1|I1=fS1=f2. Statement 2 follows immediately. Since ϕ1(I1)J1, f2(c)>d, hence f(c)(y1,1). Similarly we have f(d)(1,z1). To prove statement 3, note that f3(c)=fS2(c)I1. Recall that I1=(u1,ˆu1) with f(u1)=v0. Then f:(u0,u1)J0 is monotone and onto. Since f is increasing on (1,c), f(I1)(v0,1), hence f4(c)I0J0.

    For any x[1,1], let xf:=f(x) and xi:=fi(x). Furthermore, for k1, let αk:={fSk(c),fSk(d)}(1,p) and ak:={fSk(c),fSk(d)}(p,1). Similarly, let βk:={fSk+Sk+2(c),fSk+Sk+2(d)}(1,p) and bk:={fSk+Sk+2(c),fSk+Sk+2(d)}(p,1).

    Lemma 2.4. The points ufk1,αfk,zfk1,βfk,ufk are ordered in the following way (we state the ordering near cf rather than near c so that we do not need to be careful about on which side of c these points lie):

    ufk1<αfk<zfk1<βfk<ufk<<cf.

    Proof. Firstly note that αkCk by the chosen of Ck. By Proposition 1, zk1(uk1,uk), hence ufk1<zfk1<ufk<cf. In particular, since fSk1|Ck is monotone and onto Ik1 or Jk1, zk1 or ˆzk1Ck. We claim that zk1(αk,ˆαk).

    Indeed, without loss of generality, we may assume that ϕk is of type A++. Then Ck is on the left side of Ik and ϕk:CkJk1 is increasing. Moreover, αk=fSk(d) and fSk1:αkfSk+1(d)Dk+1. Since ϕk+1 is of type B+, Dk+1 is on the left side of Jk+1, then βk+1<d and hence αk<zk1. This proves the claim.

    Since ϕk+2 is of type C, fSk+2(c)Dk+2Jk+1. Recall that fSk:JkIk1 and hence fSk(Jk+1)Ck. Therefore fSk+Sk+2(c)Ck. Furthermore, fSk1fSk+Sk+2(c)=fSk+3(c)Dk+3 which is on the right side of d. Hence fSk+Sk+2(c)(zk1,ˆzk1). This finishes the proof.

    The following proposition gives an equivalent description of Fibonacci combinatorics for bimodal maps.

    Proposition 2. Suppose that fB+ has kneading map Q(k)=max{k2,0} and satisfies the starting condition stated in Lemma 2.3, then f has Fibonacci combinatorics.

    Proof. We prove this by induction. Firstly note that the kneading map Q(k)=max{k2,0} shows that S0=1, S1=2, S2=3 and {Sk}k1 is the Fibonacci sequence.

    From the starting condition, f(d)(1,z1). Since f is positive and f(z1)=c, f2(d) is on the left side of c. Hence f2(d)I0=(u0,ˆu0) (recall that I0 and J0 are always well-defined). By equation (1), f2(d) is outside (z0,ˆz0), hence f2(d)(u0,z0). Since the first return time of c and d to I0J0 is 2, I1 and J1 are well-defined. To be precise, since S1=2, by the chosen of z1 and y1 we have f(y1)=z1 and f(z1)=y1. Then f:(z0,z1)(d,y1) which contains v0. Let u1f1(v0)(z0,z1). Then f2:u1p=ˆv0. Indeed, I1=(u1,ˆu1). Furthermore, f2=fS1:I1J0 is local maximal at c. Since f:(u0,u1)J0 is monotone and onto, fS2(d)=f3(d)J0. By equation (1), f3(d)(ˆy0,ˆy1)(y1,y0). We claim that f3(d)(ˆv1,ˆy1)(y1,v1). For otherwise, since f:(v1,y0)(u0,c) is monotone and onto, we have f4(d)I0 and which contradicts the starting condition. Moreover, since f|(u0,z0) is increasing, we have f3(d)<d and hence f3(d)(ˆv1,ˆy1). Since fS1:J1I0, fS3(d)=fS1+S2(d)I0. Since S3 is a cutting time, (fS1(d),fS3(d)) contains c, then fS3(d) is on the right side of c. This implies f4(d)=fS0+S2(d)(z1,u0).

    For the induction step, assume that the first return times of c and d to Ik1Jk1 equal to Sk and the following induction hypothesis holds:

    zfk2<βfk1<ufk1<αfk<zfk1<cf.

    Then Ik and Jk are well-defined such that the first return times to Ik1Jk1 are Sk. Since fSk1:zk1c or d, the return time of zk1 to Ik1Jk1 is not larger that Sk1. Hence zk1Ik. For simplicity, we may assume that fSk:IkIk1 is local maximal at c. Since Sk is a cutting time, fSk|(zk1,c) is monotone and increasing. Since fSk((zk1,c)) contains the critical point c, fSk(c)>c and hence fSk(Ik)c. Then zkIk. This shows that zfk1<ufk<zfk. Moreover, since αfk<zfk1, fSk(c)Ik and then fSk(Ik)Ik. This implies that zk is contained in a return domain to Ik with return time Sk.

    For simplicity, we may assume that fSk1:Ik1Jk2 is local maximal at c. (Keep in mind that this is the case when ϕk1 is of type A++ and ϕk is of type B+). By the induction hypothesis, fSk1(c) is on the right side of d. Then fSk+1(c) is on the left side of d. By equation (1), fSk+1(c)(ˆyk1,ˆyk).

    Now let Rk1 be the first return domain to Ik2Jk2 which contains yk2. Then the return time on Rk1 is Sk2. Since we assume that fSk1:Ik1Jk2 is local maximal at c, it is clear that by the induction hypothesis, d<fSk+1+Sk1(c)<yk2<fSk1(c). Since fSk2fSk1(c)=fSk(c)Ik2Jk2, fSk1(c)Rk1. Also since fSk2fSk+1+Sk1(c)=fSk+2(c)Ik2Jk2, fSk+1+Sk1(c)Rk1. Then fSk1(Ik)Rk1. In particular, fSk1:IkRk1. This fact implies that fSk+1(c)Ik and hence zfk1<αfk+1<zfk.

    Finally, by Eq (1) αk+2(zk,zk+1)(ˆzk+1,ˆzk). Consider fSk|(zk,zk+1) which is monotone and increasing. Since Sk+1 is a cutting time, then fSk:(zk,zk+1)(c,zk1). This shows βk(c,zk1). Since fSk1(βk)Ik1Jk1, βkIk. This is because the return times are different. This shows zfk1<βfk<ufk and proves the induction hypothesis. The proof is now complete.

    Inducing schemes is a standard tool for interval maps that relates the dynamics to Markov chains. In order to prove, among other things, the existence of invariant measures, mixing rates and stochastic laws. We will first construct a countable interval partition, and then the induced Markov map for the Fibonacci bimodal map f.

    The partition of I=[1,1] is defined as the following. Denote

    {W3=(1,z1)V3=(y1,1)W2=(z1,u0)V2=(v0,y1)W1=(u0,z0)V1=(y0,v0)W0=(z0,u1)V0=(v1,y0)W1=(u1,z1)V1=(y1,v1)Wi=(zi1,zi)Vi=(yi,yi1) for i2.

    For j1, define ˆWj and ˆVj so that

    f(ˆWj)=f(Wj) and f(ˆVj)=f(Vj).

    The induced Markov map F is defined as the following.

    {F|WjVj=f for j=3,2F|WjˆWjVjˆVj=f for j=1,0F|W1ˆW1V1ˆV1=f2F|WjˆWjVjˆVj=fSj1 for j2.

    Clearly, F is monotone on each state Wk and Vk. Moreover, F preserves the partition, hence is a Markov map. To be precise, for j=3,2,1,0,1 we have

    {F(W3)=j3Wj=(1,c)F(W2)=j1ˆWj=(c,p)F(W1)=j1ˆVj=(p,d)=F(ˆW1)F(W0)=j1Vj=(d,v0)=F(ˆW0)F(W1)=j1ˆVj=(p,d)=F(ˆW1)

    and due to the symmetry

    {F(V3)=j3Vj=(d,1)F(V2)=j1ˆVj=(p,d)F(V1)=j1ˆWj=(c,p)=F(ˆV1)F(V0)=j1Wj=(u0,c)=F(ˆV0)F(V1)=j1ˆWj=(c,p)=F(ˆV1).

    For j2, by Proposition 1 and Lemma 2.2 we have the following.

    Fact 2.2.

    (1) For j=2, F(W2)=F(ˆW2)=j0Vj and F(V2)=F(ˆV2)=j0Wj;

    (2) For j3, we have

    (2.1) if j1,2,5,10mod12, then F(Wj)=F(ˆWj)=kj1Vk and F(Vj)=F(ˆVj)=kj1Wk;

    (2.2) if j4,7,8,11mod12, then F(Wj)=F(ˆWj)=kj1ˆVk and F(Vj)=F(ˆVj)=kj1ˆWk;

    (2.3) if j0,3mod12, then F(Wj)=F(ˆWj)=kj1ˆWk and F(Vj)=F(ˆVj)=kj1ˆVk;

    (2.4) if j6,9mod12, then F(Wj)=F(ˆWj)=kj1Wk and F(Vj)=F(ˆVj)=kj1Vk.

    In this section we will construct a piecewise linear bimodal map which is an odd function and has Fibonacci combinatorics. The idea comes from the unimodal model studied in [5]. The induced Markov map F over f is countably piecewise linear with definite slope on each domain. The construction proceeds along the following steps:

    I: Fix the Fibonacci sequence Sj such that S0=1,S1=2 and Sj+1=Sj+Sj1 for j1.

    II: Choose 0<q<p<1. For j1, choose a strictly increasing sequences of points zjc=p/2 satisfying z1<p<z0<q<z1. For j0, set ˆzj=pzjp/2. Finally, set yj=zj for j1 and ˆyj=ˆzj for j0. Clearly, yjp/2 satisfying y1<q<y0<p<y1. These points {zj}j1 and {yj}j1 will play the role of the closest precritical points. The points p and q should be treated as u0 and u1. Then we have a partition of the interval [1,1] as in subsection 2.4. Set

    ϵj:=|Wj|=|ˆWj|=|Vj|=|ˆVj|>0.

    Note that ˆW3,ˆW2,ˆV3 and ˆV2 have no sense. Therefore we have

    ϵ3+ϵ2+2j=1ϵj=1. (1)

    III: Define

    {s3=1ϵ3i=3ϵi=|(1,c)|ϵ3,sj=1ϵji=1ϵi=|(p,c)|ϵj for j=2,1,0,1,s2=1ϵ2i=0ϵi=|(z0,c)|ϵ2,sj=1ϵji=j1ϵi=|(zj2,c)||(zj1,zj)| for j3. (2)

    These numbers will turn out to be the absolute values of the slopes of F|Wj for the induced Markov map.

    IV: For j3, define kj>0 that will denote the slope of f|Wj. Let

    {kj:=sj for j=3,2,1,0,k1:=s1s2,k2:=s2s3,k3:=s3s1k2s2,k4:=s4s1k3s3,k5:=s5s0s2k4s4,kj:=sjsj4sj3kj1sj1 forj6. (3)

    V: Let f be the unique continuous bimodal map such that

    {f(1)=1,f(0)=0,f(1)=1,Df|W3=Df|V3=k3,Df|W2=Df|V2=k2,Df|Wj=Df|Vj=Df|ˆWj=Df|ˆVj=kj for j1,

    so that |f(Wj)|=kjϵj,j3 and each interval f(Wj) is adjacent to f(Wj+1).

    VI: Comparing Lemma 2.3, Lemma 2.4 and Proposition 2, we need some additional conditions on the sequence {ϵj} to ensure that f has Fibonacci combinatorics:

    {s2k2|cfzf2|=s2k2i=3kiϵiϵ1,sjkj|cfzfj|=sjkji=j+1kiϵiϵj2 for j3 (4)

    and

    {s4k4|cfzf4|=s4k4i=5kiϵiϵ0s2,sjkj|cfzfj|=sjkji=j+1kiϵiϵj3sj2 for j5. (5)

    Actually, in the next section, we will let ϵj tends to 0 in a geometric manner so that f depends solely on a parameter λ(0,1), and we will check these conditions are satisfied under the geometric manner.

    VII: The induced Markov map F:[1,1][1,1] is defined the same as in subsection 2.4.

    Proposition 3. Let f be the map constructed above. Then f is well-defined and has kneading map Q(k)=max{k2,0}. The induced map F is linear on each set Wj,ˆWj,ˆVj and Vj, having slope ±sj.

    Proof.

    (I) For j=3,2,1,0. We check directly that

    {f(z1)=f(1)+k3ϵ3=1+|(1,c)|=cf(p)=f(z1)+k2ϵ2=c+|(p,c)|=c+|(c,0)|=0f(z0)=f(p)+k1ϵ1=0+|(p,c)|=0+|(0,d)|=df(q)=f(z0)+k0ϵ0=d+|(p,c)|=p.

    From the definition of the induced Markov map F, it is clear that F is linear on Wj and has slope sj for j=3,2,1,0.

    (II) For j=1. Since f(q)=p,k1=s1s2, we have

    f(z1)=f(q)+k1ϵ1=p+s1s2ϵ1=p+ϵ2ϵ1ϵ1=p+ϵ2=y1.

    It follows that f(W1)=f((q,z1))=(p,y1)=V2. Therefore

    F|W1=f2|W1=f|f(W1)=f|V2

    is linear with slope k2k1=s1. Moreover, cf(y1,1)=V3.

    (III) For j=2. Since (zf1,cf)V3, fS11|(zf1,cf)=f|(zf1,cf) is linear (and also increasing) with slope k3=s3. Hence F|W2=fS1|W2 is increasing and linear with slope k2s3=s2. Furthermore,

    fS1(z1)=f2(z1)=f(y1)=d

    and

    fS1(z2)=fS1(z1)+s2ϵ2=d+|(p,c)|=d+|(d,y0)|=y0.

    Since zf2>zf1,(zf2,cf)V3. By Eq (4), we have

    fS1(c)=fS11(zf2)+s3|cfzf2|=y0+s2k2|cfzf2|y0+ϵ1=p.

    It follows that fS1(c)V1=(y0,p).

    (IV) For j=3. Since (zf2,cf)V3, f|(zf2,cf) is increasing and linear with slope k3=s3. See from above that f=fS11:(zf2,cf)(y0,fS1(c))V1. Therefore fS21|(zf2,cf)=ffS11|(zf2,cf) is increasing and linear with slope k1k3=s1s3=s3k3. Hence F|W3=fS2|W3 is increasing and linear with slope s3. Furthermore, we have

    fS2(z2)=f(fS1(z2))=f(y0)=c

    and

    fS2(z3)=fS2(z2)+s3ϵ3=c+|(z1,c)|=c+|(c,ˆz1)|=ˆz1.

    By Eq (4) again,

    fS2(c)=fS21(zf3)+s3k3|cfzf3|ˆz1+ϵ1ˆq.

    It follows that fS2(c)ˆW1. Recall that f:ˆW1V2 is monotone and onto. Then fS0+S2(c)V2.

    (V) For j4, we argue by induction, using the induction hypothesis:

    {fSj11|(zfj1,cf) is linear with slope ±sjkj.fSj1(zj1)=c or d.(Clearly,fSj1(yj1)=d or c).fSj1(c)Wj2 or ˆWj2 or Vj2 or ˆVj2. (IHj)

    From the first statement, it immediately follows F|Wj=fSj1|Wj is linear with slope ±sj. Then by the second statement, for j4,

    fSj1(zj)=fSj1(zj1)±sjϵj=c±i=j1ϵi  or  d±i=j1ϵi=zj2 or ˆzj2 or yj2 or ˆyj2.

    (VI) For j=4. Since (zf3,cf)(zf2,cf), fS21|(zf3,cf) is increasing and linear with slope s3k3. Since fS2(z3)=ˆz1 and fS2(c)ˆW1, fS21((wf3,cf))ˆW1. Note that fS1|ˆW1 is decreasing and has slope s1. Then fS31|(zf3,cf)=fS1fS21|(zf3,cf) is decreasing with slope s1s3k3=s4k4. Therefore F|W4=fS3|W4 is decreasing and linear with slope s4. Furthermore,

    fS3(z3)=fS1(fS2(z3))=fS1(ˆz1)=fS1(z1)=d

    and

    fS3(z4)=fS3(z3)s4ϵ4=ˆy2.

    By Eq (4),

    fS3(c)=fS31(cf)=fS31(zf4)s4k4|cfzf4|=ˆy2s4k4|cfzf4|ˆy2ϵ2.

    This shows fS3(c)ˆV2.

    (VII) For j=5. We have

    fS41|(zf4,cf)=ffS1fS31|(zf4,cf).

    Since (zf4,cf)(zf3,cf), fS31|(zf4,cf) is decreasing and linear with slope s4k4. Since fS3(c)ˆV2, fS31:(zf4,cf)(fS3(c),ˆy2)ˆV2. Since fS1|ˆV2 has slope s2. We have

    fS1fS31:(zf4,cf)(z0,fS1+S3(c))

    is increasing and linear with slope s2s4k4. By Eq (5),

    s2s4k4|zf4cf|ϵ0.

    Therefore fS1+S3(c)z0ϵ0, hence (z0,fS1+S3(c))W0. Since f|W0 is increasing and linear with slope k0, we have fS41|(zf4,cf) is linear with slope k0s2s4k4=s5k5. Hence F|W5=fS4|W5 is increasing and linear with slope s5. Furthermore,

    fS4(z4)=fS2(fS3(z4))=fS2(ˆy2)=d,

    and

    fS4(z5)=fS4(z4)+s5ϵ5=y3.

    By Eq (4),

    fS4(c)=fS41(zf5)+s5k5|cfzf5|=ω3+s5k5|cfwf5|ω3+ϵ3.

    This shows fS4(c)V3.

    (VIII) For j6. Now assume that the induction hypothesis (IHi) holds for 5i<j, we will prove that (IHj) holds.

    Consider

    fSj11|(zfj1,cf)=fSj5fSj4fSj21|(zfj1,cf).

    Since (zfj1,cf)(zfj2,cf), using (IHj1), we have fSj21|(zfj1,cf) is linear with slope ±sj1kj1. By Eq (4),

    |fSj2(c)fSj2(zj1)|=sj1kj1|zfj1cf|ϵj3.

    Then fSj2(c)Wj3 or ˆWj3 or Vj3 or ˆVj3. It follows that

    fSj4fSj21|(zfj1,cf)

    is linear with slope ±sj3sj1kj1. Using (IHj1) and (IHj3),

    fSj4fSj21(zfj1)=zj5 or ˆzj5 or yj5 or ˆyj5.

    By Eq (5), the length of the interval

    (fSj4fSj21(zfj1),fSj4fSj21(cf))

    is sj3sj1kj1|cfzfj1|sj3ϵj4sj3=ϵj4. Then

    (fSj4fSj21(zfj1),fSj4fSj21(cf))Wj4 or ˆWj4 or Vj4 or ˆVj4.

    Using (IHj4), F=fSj5 is linear on these sets with slope ±sj4. It follows that fSj11|(zfj1,cf) is linear with slope ±sj4sj3sj1kj1=±sjkj.

    For simplicity, we may assume that fSj1(zj1)=c and fSj11|(zfj1,cf) is increasing and linear with slope sjkj. Then

    fSj1(zj)=fSj11(zfj1)+sjkj|zfjzfj1|=c+sjϵj=ˆzj2.

    By Eq (4),

    fSj1(c)=fSj1(zj1)+sjkj|cfzfj|ˆzj2+ϵj2.

    Hence fSj1(c)ˆWj2. This concludes the induction.

    The following proposition shows that the map f constructed as above has Fibonacci combinatorics.

    Proposition 4. Let f be the map constructed as above. Then f has Fibonacci combinatorics.

    Proof. It follows from the proof of Proposition 3 that f satisfies

    (1) f(c)V3 and f(d)W3;

    (2) f2(c)=fS1(c)V1J0 and f2(d)=fS1(d)W1I0;

    (3) f4(c)=fS0+S2(c)V2 and f4(d)=fS0+S2(d)W2, hence f4(c),f4(d)I0J0.

    Then f satisfies the starting condition. Now this proposition follows from the fact that f has kneading map Q(k)=max{k2,0} and Proposition 2.

    In smooth interval dynamics, the uniform expansion of a map was mainly described in two ways: the expansion of derivatives along critical orbits and the local geometry around the critical points. In the latter case, the local geometry was often clarified by the ratio between principal nest (in our setting we use twin principal nest instead). In other words, the uniform expansion of a map depends on the asymptotic properties of its scaling factor, i.e. the sequences |unc||un1c| and |vnc||vn1c|.

    In this section, we let ϵj tends to 0 in a geometric manner. To be precise, set

    {ϵ3=38, ϵ2=18,ϵj=(1λ)λj+14 for j1, (6)

    where λ(0,1). Then f depends solely on a single parameter λ and we can calculate sj and kj under this geometric manner. Since we use the preimages of the critical points instead of the preimages of the fixed point p to construct the induced Markov map, the ratio |zjc||zj1c| (which equals λ for j1) plays the same role as the scaling factor. Figures 6 and 7 show the partial graphs of f and F under this geometric manner.

    Figure 6.  Graph of f on the 4 outmost branches.
    Figure 7.  Graph of the induced Markov map F on [12,12].

    By formula (2),

    {s3=s2=2,s1=11λs0=1λ(1λ)s1=s2=1λ2(1λ)sj=1λ(1λ) for j3. (7)

    Using formula (3) we obtain

    {k3=k2=2,k1=11λ,k0=1λ(1λ),k1=k2=12λ2(1λ),k3=12λ,k4=λ(1λ)2,k5=λ4(1λ)32,kj=λ2j5(1λ)2j72 for j6. (8)

    Let us check that formulas (4) and (5) in Step are true for all λ(0,1).

    Lemma 4.1.

    {s2k2|cfzf2|=s2k2i=3kiϵiϵ1,sjkj|cfzfj|=sjkji=j+1kiϵiϵj2forj3 (4.1)

    and

    {s4k4|cfzf4|=s4k4i=5kiϵiϵ0s2,sjkj|cfzfj|=sjkji=j+1kiϵiϵj3sj2forj5. (4.2)

    are true for all λ(0,1) under the geometric manner.

    Proof. For simplicity, we write ϵj=C1λj, j1, kj=C2ωj, j6, where C1=λ(1λ)4,C2=12λ5(1λ)7,ω=λ2(1λ)2, then C2i=6(λω)i=λ13(1λ)52(1λ3(1λ)2). Let us first check formula (4.1).

    ● For j=2,

    s2k2|cfzf2|ϵ1s2k2(k3ϵ3+k4ϵ4+k5ϵ5+i=6kiϵi)ϵ12(12λC1λ3+λ(1λ)2C1λ4+λ4(1λ)32C1λ5+C1C2i=6(λω)i)C1λλ3+λ6(1λ)+λ10(1λ)3+λ14(1λ)51λ3(1λ)210.

    ● For j=3,

    s3k3|cfzf3|ϵ1s3k3(k4ϵ4+k5ϵ5+i=6kiϵi)ϵ121λ(λ(1λ)2C1λ4+λ4(1λ)32C1λ5+C1C2i=6(λω)i)C1λλ4+λ8(1λ)2+λ12(1λ)41λ3(1λ)210.

    ● For j=4,

    s4k4|cfzf4|ϵ2s4k4(k5ϵ5+i=6kiϵi)ϵ22λ2(1λ)2(λ4(1λ)32C1λ5+C1C2i=6(λω)i)C1λ2λ5(1λ)+λ9(1λ)31λ3(1λ)210.

    Figure 8 shows the graphs of functions x3+x6(1x)+x10(1x)3+x14(1x)51x3(1x)21, x4+x8(1x)2+x12(1x)41x3(1x)21 and x5(1x)+x9(1x)31x3(1x)21.

    Figure 8.  Graphs of functions x3+x6(1x)+x10(1x)3+x14(1x)51x3(1x)21, x4+x8(1x)2+x12(1x)41x3(1x)21 and x5(1x)+x9(1x)31x3(1x)21.

    ● For j=5,

    s5k5|cfzf5|ϵ3s5k5i=6kiϵiϵ32λ5(1λ)4C1C2i=6(λω)iC1λ3λ5(1λ)1λ3(1λ)210.

    ● For j6,

    sjkj|cfzfj|ϵj21λ(1λ)1C2ωji=j+1C1C2(λω)iC1λj2λ4(1λ)1λ3(1λ)210.

    Figure 9 shows the graphs of functions x5(1x)1x3(1x)21 and x4(1x)1x3(1x)21.

    Figure 9.  Graphs of functions x5(1x)1x3(1x)21 and x4(1x)1x3(1x)21.

    These are true for all λ(0,1). Now we check formula (4.2).

    ● For j=4,

    s4k4|cfzf4|ϵ0s2s4k4(k5ϵ5+i=6kiϵi)ϵ0s22λ2(1λ)2(λ4(1λ)32C1λ5+C1C2i=6(λω)i)C1λ2(1λ)λ5+λ9(1λ)21λ3(1λ)210.

    ● For j=5,

    s5k5|cfzf5|ϵ2s3s5k5i=6kiϵiϵ2s32λ5(1λ)4C1C2i=6(λω)iC1λ3(1λ)λ51λ3(1λ)210.

    ● For j6,

    sjkj|cfzfj|ϵj3sj21λ(1λ)1C2ωji=j+1C1C2(λω)iC1λj2(1λ)λ41λ3(1λ)210.

    Figure 10 shows the graphs of functions x5+x9(1x)21x3(1x)21, x51x3(1x)21 and x41x3(1x)21.

    Figure 10.  Graphs of functions x5+x9(1x)21x3(1x)21, x51x3(1x)21 and x41x3(1x)21.

    Again, these are true for all λ(0,1).

    In this section, we prove the Main Theorem. We first compute the critical order of f in terms of λ. This is statement (1) of the Main Theorem.

    Proof of statement (1) of the Main Theorem. Note that by formula (8) we have |Df(x)|=kj=O(λ2j(1λ)2j). Then by formula (6) and the fact that ij+1ϵi|xc|ijϵi, |xc|=O(λj) if xWj for j5. On the other hand, by definition, |Df(x)|=O(|xc|1). Hence

    1=2logλ(1λ)logλ=2+2log(1λ)logλ.

    This finishes the proof.

    Now we will compute the values of λ for which there is an Fλ-invariant probability measure which is an acip. For simplicity, we will consider the following system instead.

    Given any λ(0,1), let Vn=(λn,λn1] for n1. Then Vn form a Markov partition of the interval (0,1]. Define the countably piecewise linear interval map Tλ:(0,1](0,1] as:

    For the graph of Tλ, see Figure 11.

    Figure 11.  Graph of Tλ.

    Note that the system Fλ and Tλ are not semi-conjugate. This is because there are infinitely branches F|Wn and T|Vn with different orientations. However, one can show that Fλ admits an acip whenever Tλ does. Moreover, if Tλ admits an acip μ, then Fλ admits an acip ν satisfying ν(Wn)=14μ(Vn) for all n1.

    To continue, we will consider a random walk governed by Tλ:(0,1](0,1]. When there is no confusion, we will omit the parameter λ and use F and T for instead. To describe the random walk, let αk(x)=n if Tk(x)Vn. Then the sequence of random variables α0,α1,α2, can be considered as a Markov chain with the following transition probability:

    P(αk+1=i|αk=j):=Leb(αk=j and αk+1=i)Leb(αk=n),

    where Leb() denotes the Lebesgue measure of the set.

    Let (Ai,j)i,j be the transition matrix corresponding to Tλ, one can compute that in matrix form

    (Ai,j)i,j=(1λ)(1λλ2λ3λ4λ51λλ2λ3λ4λ51λλ2λ3λ4λ501λλ2λ3λ4λ501λλ2λ3λ401λλ2λ301λλ2).

    Now write yn,k=Leb({αk=n})=Leb(Tk(Vn)). Let vk=(y1,k,y2,k,y3,k,,yn,k,) be a row vector.

    Lemma 5.1. For k1, we have

    vk=vk1(Ai,j)i,j.

    Proof. Choose n4. Note that T1(Vn) has n+1 connected components. Let JiVi, 1in+1, denote these components. Now let J be any components of Tk1 such that Tk1(J)=Vi. Since T is piecewise linear, we have

    Leb(xJ;Tk1(x)Ji)|J|=|Ji||Vi|=1si|Vn||Vi|.

    Summing over all J and 1in+1, we have

    yn,k=|Vn|n+1i=11si|Vi|yi,k1.

    The cases for n=1,2,3 are similar. Now this lemma follows from formulas (6) and (7).

    For λ(0,12), take

    {v1=2λ3λ22λ+1λ32λ2λ+1v2=λ2λ2λ32λ2λ+1v3=λ22λ3λ32λ2λ+1v4=λ3(2λ1)(λ1)(λ32λ2λ+1)k=(2λ1)(λ33λ2+3λ1)λ(λ32λ2λ+1).

    The functions vi(λ) and k(λ) remain positive. Figure 12 shows the graphs of vi(λ) and k(λ).

    Figure 12.  Graphs of v1,v2,v3,v4 and k.

    For i5, we set vi=k(λ1λ)i. One can check that v=(v1,v2,,vi,) is the left eigenvector of the matrix (Ai,j)i,j with eigenvalue 1. In fact, k and vi, 1i4 are the solutions of the following equations:

    {v1=(1λ)(v1+v2+v3)v2=λv1+(1λ)v4v3=λv2v4=λv3+(1λ)v5Σivi=1.

    Lemma 5.2. If λ(0,12), then the system ((0,1],Tλ) admits an acip μ. Furthermore, for i5

    μ(Vi)(λ1λ)i.

    Proof. We will use the result by Straube [16] claiming that T has an acip if (and only if) for any η(0,1), there exists δ>0 such that for every measurable set A of measure Leb(A)<δ holds Leb(Tk(A))<η for all k1.

    To check the Straube's condition, we will prove by induction that yn,kCvn(λ) for all n,k0. For fixed λ, take C=C(λ) large enough such that 0<yn,0Cvn for all n. This is possible because

    yn,0=|Vn|=λn1(1λ)=(λ1λ)n(1λ)n+1λ2kk(λ1λ)n,

    for all n5. Since v is the left eigenvector, now this claim follows from Lemma 5.1.

    Now take η(0,1). Since yn,k is exponentially small in terms of n, pick n1 large enough such that nn1yn,k<η/2. Assume that AVi for some 1i<n1. Since T is piecewise linear, we have

    Leb(Tk(A))=Leb(A)yi,kCLeb(A)η2n1

    provided Leb(A)<δ is sufficiently small. It follows that if Ai<n1Vi has sufficiently small measure, then Leb(Tk(A))<n1η/(2n1)=η/2. This concludes the verification of Straube's condition.

    Since T has an acip μ, it can be written as μ(A)=limn1nn1i=0Leb(Ti(A)). Therefore,

    μ(Vn)Cvn.

    Indeed, one can show that μ=(μ(V1),μ(V2),,μ(Vn),) is also an eigenvector of (Ai,j)i,j with eigenvalue 1 and satisfying nμ(Vn)=1. Argue as [4][Theorem 1] shows that μ(Vn)(λ1λ)n for all n5.

    Proof of statement (4) of the Main Theorem. By a standard pullback argument, the original map f has acip if and only if

    jSj1μ(Wj)<.

    And when it fails, then f has an absolutely continuous σ-finite measure. This follows because f is conservative and ω(c)=ω(d) is a Cantor set with 0 Lebesgue measure. See [13] [Lemma 3.1] and [7] [Theorem 1].

    Note that {Sj} is Fibonacci so Sj1γj1, where γ=1+52 is the golden mean. Since μ(Wi)(λ1λ)i for i5, then

    jSj1μ(Wj)<1+52λ1λ<1λ<23+5.

    This corresponds to the critical order 3<<4. If 23+5λ<12, the series divergence, f has an absolutely continuous σ-finite measure. This corresponds to the critical order 4<5.

    When λ=12, the proof is quite different and is somewhat out of the context of this paper. However, one can argue as [4] [Lemma 3] to show that Tλ (equivalently Fλ) is null recurrent w.r.t. Lebesgue measure and hence is conservative with a σ-finite measure.

    In this subsection we prove that fλ has a wild attractor when λ(12,1), which is equivalent to show that Fλ (and also Tλ) is dissipative.

    The main tool in the proof of the existence of wild Cantor attractor is a random walk. In this subsection we will use a random walk governed by the system Fλ which is somewhat similar with the argument used in the previous subsection. To describe the random walk, write φn(x)=k if Fn(x)WkˆWkˆVkVk. In principle, if a point x escapes to infinity under iteration of F, then x tends to ω(c) under f. And if a positive Lebesgue set points escape to infinity, then ω(c) is a wild attractor. We will have to calculate the probabilities from one state to another.

    The asymptotic behavior of the random variable φn can be computed from the expectation E(φnφn1), taken with respect to normalized Lebesgue measure on [12,12]. If E(φnφn1)ϵ>0, then one can expect that limnφn=. To prove this we will use the conditional expectation. We also need the boundedness of the variance. The drift of the random walk is defined as the conditional expectation

    Dr(f)=Dr(λ)=E(φnk|φn1=k)=ik1(ik)P(φn=i,φn1=k)P(φn1=k)=ik1(ik)ϵiik1ϵi=ik1(ik)(1λ)λi+14ik1(1λ)λi+14=λ1λ1=2λ11λ.

    Such a concept was first used in the study of Fibonacci unimodal maps in [3]. The second moment Var(φnk|φn1=k) is clarified as

    ik1(ik)2ϵiik1ϵi=ik1(ik)2(1λ)λi+14ik1(1λ)λi+14=λ2(1λ)22λ1λ+1.

    Lemma 5.3. Let Fλ be the induced Markov map over fλ under the geometric manner. If there exists kN and δ>0 such that

    (1) E(φnk|φn1=k)δ;

    (2) Var(φnk|φn1=k) is uniformly bounded from above.

    Then limnφn(x)= for Lebesgue a.e. x.

    Proof. The proof is just a modification of [2] [Theorem 5.2] since in our case the system has strong symmetry.

    Proof of statement (2) of Main Theorem. Now the statement follows from the remark before and Lemma 5.3.

    Remark 5.1. For a general smooth bimodal map f with Fibonacci combinatorics, even in the case that the two critical points are non-flat with the same critical order , it is quite difficult to show that f has a wild attractor when is sufficiently large. A crucial step in the proof is to estimate the upper and lower bounds of the derivatives along the orbit of the critical points DfSn(f(c)). Suppose that n0,1mod3, fSn(c)Jn1. Then by the standard argument,

    |DfSn(f(c))|=|Df(fSn(c))||DfSn1(f(c))||Jn|1K|fSn(In)||f(In)|K(|Jn||In|)|Jn1||Jn|.

    But in fact we do not have any estimate of |Jn|/|In|. However, if the Fibonacci bimodal map is also an odd function, one can show that its metric properties are essentially the same as those of the Fibonacci unimodal map. Since in this case we always have |In|=|Jn| and Dfn(f(c))=Dfn(f(d)) for all n1. One can use the topological properties stated in Lemma 2.4 and the formulas in [10] to do so.

    Remark 5.2. One possible way to find a Fibonacci bimodal map which is an odd function is by topology. If we can find a one-parameter family of odd polynomials with two critical points and show that such a family is a 'full family' (see [13]) for any combinatorially symmetric combinatorics (note that it is never a full family for any combinatorics since non symmetric combinatorics cannot appear in this family), then the family must contain the Fibonacci combinatorics. Another possible way is by renormalization. Note that the first return map gn restricted on the four return domains is a special class of box mapping. Rescaling gn to the same scale and obtaining a new map, such a step is called 'generalized renormalization'. The Fibonacci bimodal map is clearly infinitely renormalizable in this sense. Denote the n-th renormalization by Rnf. If the renormalization sequence Rnf converges to a periodic cycle (this is reasonable by Fact 2.1), then any limit map is an odd function and has Fibonacci combinatorics. We emphasize here that the 'bounded geometry' property plays essential way in the renormalization theory, see [8].

    We study Fibonacci bimodal maps under a restrictive condition and show that the one-parameter family fλ has a phase transition from Lebesgue conservative to dissipative behaviors in this paper.

    In Section 2, we study the topological properties of the Fibonacci bimodal maps in the context of kneading map and give an equivalent description of Fibonacci combinatorics. In Proposition 1, we show that the cutting times Sk and the critical return times sk are coincident, the kneading map is clarified as Q(k)=max{k2,0}. And, in Proposition 2, we prove that if fB+ has kneading map Q(k)=max{k2,0} and satisfies the starting condition stated in Lemma 2.3, then f has Fibonacci combinatorics. Therefore we can study the Fibonacci combinatorics for bimodal maps by two equivalent descriptions.

    In Section 3, we construct a piecewise linear bimodal map f which is an odd function and has Fibonacci combinatorics. After construction, we prove in Proposition 3 and Proposition 4 that f is well-defined and has kneading map Q(k)=max{k2,0}. The induced map F is linear on each set Wj,ˆWj,ˆVj and Vj, having slope ±sj.

    In Section 4, we let ϵj tends to 0 in a geometric manner, where ϵj only depends on λ. Then f depends solely on a single parameter λ. We denote f by fλ for emphasizing λ and calculate sj and kj under this geometric manner. In Lemma 4.1 we show that the additional conditions VI in Section 3 are true for all λ(0,1).

    In Section 5, we first compute the critical order of f in terms of λ. In Subsection 5.1, we compute the values of λ for which there is an Fλ-invariant probability measure which is an acip, using a random walk govered by Tλ:(0,1](0,1]. Finally, in Subsection 5.2, we prove that fλ has a wild attractor when λ(12,1), which is equivalent to show that Fλ is dissipative.

    All authors declare no conflicts of interest in this paper.



    [1] B. Branner, J. H. Hubbard, The iteration of cubic polynomials I, Acta Math., 160 (1988), 143–206.
    [2] H. Bruin, Topological conditions for the existence of absorbing Cantor sets, Trans. Amer. Math. Soc., 350 (1998), 2229–2263.
    [3] H. Bruin, G. Keller, T. Nowicki, S. van Strien, Wild Cantor attractors exist, Ann. Math., 143 (1996), 97–130. https://doi.org/10.2307/2118654 doi: 10.2307/2118654
    [4] H. Bruin, M. Todd, Transience and thermodynamic formalism for infinitely branched interval maps, J. London Math. Soc., 86 (2012), 171–194. https://doi.org/10.1112/jlms/jdr081 doi: 10.1112/jlms/jdr081
    [5] H. Bruin, M. Todd, Wild attractors and thermodynamic formalism, Monatsh. Math. 178 (2015), 39–83. https://doi.org/10.1007/s00605-015-0747-2 doi: 10.1007/s00605-015-0747-2
    [6] H. Bruin, W. Shen, S. van Strien, Invariant measure exists without a growth condition, Commun. Math. Phys., 241 (2003), 287–306. https://doi.org/10.1007/s00220-003-0928-z doi: 10.1007/s00220-003-0928-z
    [7] F. Hofbauer, G. Keller, Some remarks on recent results about S-unimodal maps, Ann. Inst. Henri Poincaré, 53 (1990), 413–425.
    [8] H. Ji, S. Li, The attractor of Fibonacci-like renormalization operator, Acta Math. Sin., English Ser., 36 (2020), 1256–1278. https://doi.org/10.1007/s10114-020-9185-8 doi: 10.1007/s10114-020-9185-8
    [9] H. Ji, W. Ma, Decay of geometry for a class of cubic polynomials, Preprint.
    [10] G. Keller, T. Nowicki, Fibonacci maps re(al)-visited, Ergod. Theor. Dyn. Sys., 15 (1995), 99–120. https://doi.org/10.1017/S0143385700008269 doi: 10.1017/S0143385700008269
    [11] M. Lyubich, J. Milnor, The Fibonacci unimodal map, J. Amer. Math. Soc., 6 (1993), 425–457.
    [12] M. Lyubich, Combinatorics, geometry and attractors of quasi-quadratic maps, Ann. Math., 140 (1994), 347–404. https://doi.org/10.2307/2118604 doi: 10.2307/2118604
    [13] W. de Melo, S. van Strien, One-dimensional dynamics, Springer-Verlag, Berlin, 1993.
    [14] M. L. Nascimento, R. J. S. Mota, The importance of Fibonacci combinatorics for dynamic systems in the interval and in the circle and applications, Eur. Int. J. Sci. Technol., 10 (2021), 82–96.
    [15] W. Shen, On the metric properties of multimodal interval maps and C2 density of Axiom A, Invent. Math., 156 (2004), 301–403. https://doi.org/10.1007/s00222-003-0343-2 doi: 10.1007/s00222-003-0343-2
    [16] E. Straube, On the existence of invariant absolutely continuous measures, Commun. Math. Phys., 81 (1981), 27–30. https://doi.org/10.1007/BF01941798 doi: 10.1007/BF01941798
    [17] A. K. Supriatna, E. Carnia, M. Z. Ndii, Fibonacci numbers: a population dynamics perspective, Heliyon, 5 (2019), e01130. https://doi.org/10.1016/j.heliyon.2019.e01130 doi: 10.1016/j.heliyon.2019.e01130
    [18] G. Światek, E. Vargas, Decay of geometry in the cubic family, Ergod. Theor. Dyn. Sys., 18 (1998), 1311–1329. https://doi.org/10.1017/S0143385798117558 doi: 10.1017/S0143385798117558
    [19] E. Vargas, Fibonacci bimodal maps, Discrete Contin. Dyn. Sys., 22 (2008), 807–815.
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1462) PDF downloads(44) Cited by(0)

Figures and Tables

Figures(12)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog