This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Odd period cycles and ergodic properties in price dynamics for an exchange economy

Tomohiro Uchiyama
Faculty of International Liberal Arts, Soka University,
1-236 Tangi-machi, Hachioji-shi, Tokyo 192-8577, Japan
Email:[email protected]
Abstract

In the first part of this paper (Sections 1-4), we study a standard exchange economy model with Cobb-Douglas type consumers and give a necessary and sufficient condition for the existence of an odd period cycle in the Walras-Samuelson (tatonnement) price adjustment process. We also give a sufficient condition for a price to be eventually attracted to a chaotic region. In the second part (Sections 5 and 6), we investigate ergodic properties of the price dynamics showing that the existence of chaos is not necessarily bad. (The future is still predictable on average.) Moreover, supported by a celebrated work of Avila et al. (Invent. Math., 2003), we conduct a sensitivity analysis to investigate a relationship between the ergodic sum (of prices) and the speed of price adjustment. We believe that our methods in this paper can be used to analyse many other chaotic economic models.

Keywords: chaos, odd period cycle, exchange economy, price dynamics, ergodic theory, sensitivity analysis
JEL classification: D11, D41, D51

1 Introduction

In this paper, we study a standard exchange economy model with two consumers of Cobb-Douglas type and two goods. Denote the two consumers by i=1,2i=1,2 and the two goods by xx and yy. Let α,β(0,1)\alpha,\beta\in(0,1). We assume that the consumer 11 has the utility function u1(x,y)=xαy1αu^{1}(x,y)=x^{\alpha}y^{1-\alpha} and the consumer 22 has the utility function u2(x,y)=xβy1βu^{2}(x,y)=x^{\beta}y^{1-\beta} in the consumption space +2\mathbb{R}^{2}_{+}. We also assume that the consumer 11 has the initial endowment w1=(x¯,0)w^{1}=(\bar{x},0) and the consumer 22 has the initial endowment w2=(0,y¯)w^{2}=(0,\bar{y}) where x¯,y¯>0\bar{x},\bar{y}>0. Fix the price of good yy as 11 and that of good xx as p>0p>0. Then by the standard optimisation result under the budget constraints, we obtain the excess demand function z(p)z(p) for good xx, that is given by z(p)=y¯β/px¯(1α)z(p)=\bar{y}\beta/p-\bar{x}(1-\alpha). Now we define the Walras-Samuelson (tatonnement) price adjustment process by

pt+1=f(pt)=pt+λz(pt)=pt+λ[y¯β/ptx¯(1α)] where λ++.p_{t+1}=f(p_{t})=p_{t}+\lambda z(p_{t})=p_{t}+\lambda[\bar{y}\beta/p_{t}-\bar{x}(1-\alpha)]\textup{ where }\lambda\in\mathbb{R}_{++}. (1.1)

Note that λ\lambda denotes the speed of adjustment and pjp_{j} denotes the price of good xx at time jj.

Let E=(0,1)×(0,1)E=(0,1)\times(0,1). Then the following (that is a slight extension of [Bhattacharya and Majumdar, 2007, Prop. 9.10]) is not difficult to show (see Section 4 for a proof):

Proposition 1.1.

There exist open sets AEA\subset E, B++2B\subset\mathbb{R}^{2}_{++}, and C++C\subset\mathbb{R}_{++} such that if (α,β)A(\alpha,\beta)\in A, (x¯,y¯)B(\bar{x},\bar{y})\in B, and λC\lambda\in C then the process (1.1) has a period three cycle (hence exhibits a Li-Yorke chaos).

It is well-known that the existence of a period three cycle implies that of a Li-Yorke chaos (by the famous Li-Yorke theorem [Li and Yorke, 1975, Thm. 1]), and this argument has been used a lot in economic literature, see [Benhabib and Day, 1980][Benhabib and Day, 1982][Day and Shafer, 1985][Nishimura and Yano, 1996] for example. However, this is a bit overkill: by [Block and Coppel, 1992, Chap. \@slowromancapii@], we know that the existence of a cycle of any odd length (not necessarily of period three) implies that of a Li-Yorke chaos. In the first part of this paper, extending Proposition 1.1, we obtain:

Theorem 1.2.

Let y¯βx¯2(1α)2<λ<4y¯βx¯2(1α)2\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}. Then the map ff in Equation (1.1) has the following properties:

  1. 1.

    f|E𝔊{\left.\kern-1.2ptf\right|_{E}}\in\mathfrak{G}.

  2. 2.

    f|E{\left.\kern-1.2ptf\right|_{E}} has an odd period cycle if and only if 25y¯β9x¯2(1α)2<λ<4y¯βx¯2(1α)2\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}.

  3. 3.

    The second iterate (f|E)2({\left.\kern-1.2ptf\right|_{E}})^{2} is turbulent if and only if 25y¯β9x¯2(1α)2λ<4y¯βx¯2(1α)2\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}}\leq\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}.

  4. 4.

    If 25y¯β9x¯2(1α)2λ<4y¯βx¯2(1α)2\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}}\leq\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}, then the closed interval EE is attracting for ff, that is, fn(p)Ef^{n}(p)\in E for some nn\in\mathbb{N}.

Our first main result of this paper (Theorem 1.2) extends Proposition 1.1 in the following way: 1. Replace a period three cycle with an odd period cycle (or a turbulence for the second iterate), 2. Give a specific (algebraic) form of the sets AA, BB, and CC, 3. Give a necessary and sufficient condition for the existence of an odd period cycle (or a turbulence for the second iterate) rather than giving a sufficient condition only (that is usually done in many economic literature such as [Benhabib and Day, 1980][Benhabib and Day, 1982][Day and Shafer, 1985][Nishimura and Yano, 1996]).

We defer detailed explanations, (mathematical/economic) interpretations, and all the necessary definitions (such as EE, 𝔊\mathfrak{G}, and ”turbulent”) for Theorem 1.2 to Section 2. Here, we just note two things: (1) EE is some compact interval in \mathbb{R} and 𝔊\mathfrak{G} is a ”nice” class of continuous unimodal maps. (2) The upper bound for λ\lambda directly follows from the requirement p>0p>0. So, the upper bound in Theorem 1.2 is not really interesting. The real meat is in the lower bounds.

In the second part of this paper (Section 5 onwards), we investigate the price dynamics defined by Equation (1.1) using a probabilistic method. The upshot of Theorem 1.2 is that if λ\lambda (the speed of price adjustment) is large enough (with some bound to keep p>0p>0), the price dynamics show chaotic behaviours for (uncountable) many initial pp (see the definition of a Li-Yorke chaos in Section 2). This means that it is hard to predict the future (fn(p)f^{n}(p) for large nn) for these pp. This sounds pretty bad, but the main results in the second part of the paper (Theorem 1.4) show that this not necessarily so. Roughly speaking, we can still ”predict” the future on average.

Our (numerical/theoretical) argument in the second part of the paper use ergodic theory (a quick overview of ergodic theory is given in Section 5 to make the paper self-contained). What we need now is a bare minimum: we just need one definition.

Definition 1.3.

limn1nk=0n1fk(p)\lim_{n\rightarrow\infty}\frac{1}{n}\sum_{k=0}^{n-1}f^{k}(p) is called the ergodic sum of ff with respect to pp. (If the limit exists, the ergodic sum of ff is same for almost all pp, so we omit ”with respect to pp”.)

We are ready to state our main result in the second part of the paper. Here, to simplify the exposition and to obtain sharp numerical results, we fix (α,β)=(0.75,0.5)(\alpha,\beta)=(0.75,0.5), (x¯,y¯)=(4,2)(\bar{x},\bar{y})=(4,2). A similar analysis (as follows) can be done for any α,β,x¯,y¯\alpha,\beta,\bar{x},\bar{y} (our choice of the parameters is completely ad hoc and our method is quite generic). We obtain

Theorem 1.4.

For 1<λ<41<\lambda<4, the ergodic sums of ff are as in Figure 1 (possibly except some λ\lambda values whose total Lebesgue measure is 0).

Refer to caption
Figure 1: Ergodic sums of ff

We defer the detailed explanations/comments on Theorem 1.4 to Sections 5 and 6, but to be honest, we found it very surprising that the ergodic sums of ff change so smooth and stable (except a few bumps) as λ\lambda increases considering the fact (as can be seen from the bifurcation diagram of ff, Figure 13) that as λ\lambda increases fn(p)f^{n}(p) go through a quite a bit of qualitative changes from a stable fixed point, attracting periodic orbit of different periods, and finally chaotic behaviours.

Here is the structure of the rest of the paper. In Section 2, we give necessary definitions involving chaos and explain how we got our first main theorem (Theorem 1.2) together with its (mathematical/economic) interpretations. Next, in Section 3, we give the proofs of Theorem 1.2 and Proposition 1.1. In Section 4, after giving a quick overview of ergodic theory, we explain a generic method to obtain our second main result (Theorem 1.4) with its important stepping stone (Theorem 5.11). Our argument here is a delicate combinations of results from ergodic theory and numerical computations. In Section 5, we give some (mathematical) interpretations of Theorem 1.4 including its connection to a deep result in ergodic theory (Proposition 6.3) of Avila et.al.

All programming files used for numerical calculations and for generating plots in this paper (Jupyter notebooks) are available upon request.

2 Odd period cycles

2.1 Definitions involving a chaos

First, we clarify what we mean by a Li-Yorke chaos, a turbulence, and a topological chaos. (There are several definitions of a chaos in the literature.) The following definitions are taken from [Ruette, 2017, Def. 5.1] and [Block and Coppel, 1992, Chap. \@slowromancapii@]. Let gg be a continuous map of a closed interval II into itself:

Definition 2.1.

We say that gg exhibits a Li-Yorke chaos if there exists an uncountable scrambled set SIS\subset I, that is, for any x,ySx,y\in S we have

lim supngn(x)gn(y)>0 and lim infngn(x)gn(y)=0,\limsup_{n\rightarrow\infty}\mid g^{n}(x)-g^{n}(y)\mid>0\textup{ and }\liminf_{n\rightarrow\infty}\mid g^{n}(x)-g^{n}(y)\mid=0,

and for xSx\in S and yy being a periodic point of gg,

lim supngn(x)gn(y)>0.\limsup_{n\rightarrow\infty}\mid g^{n}(x)-g^{n}(y)\mid>0.
Definition 2.2.

We call gg turbulent if there exist three points, x1x_{1}, x2x_{2}, and x3x_{3} in II such that g(x2)=g(x1)=x1g(x_{2})=g(x_{1})=x_{1} and g(x3)=x2g(x_{3})=x_{2} with either x1<x3<x2x_{1}<x_{3}<x_{2} or x2<x3<x1x_{2}<x_{3}<x_{1}. Moreover, we call gg (topologically) chaotic if some iterate of gg is turbulent.

It is known that a map gg is topologically chaotic if and only if gg has a periodic point whose period is not a power of 22, see [Block and Coppel, 1992, Chap. \@slowromancapii@]. This implies that a map gg is topologically chaotic if and only if the topological entropy of gg is positive, see [Block and Coppel, 1992, Chap. \@slowromancapviii@]. In the first part of this paper, we focus on a topological chaos (or a positive topological entropy) in the context of price dynamics. See [Ruette, 2017] for more characterisations and (subtle) mutual relations of various kinds of chaos.

2.2 Explanations and Interpretations of Theorem 1.2

Here, we explain a key result to obtain Theorem 1.2. Along the way, we give definitions for all the unexplained notation (such as 𝔊\mathfrak{G} and EE) in Theorem 1.2. First, we recall the following mathematical result characterising the existence of a topological chaos for a unimodal interval map [Deng et al., 2022, Cor. 3]. Theorem 1.2 is a (highly non-trivial as seen in Section 3) consequence (or a special case) of [Deng et al., 2022, Cor. 3]. Let 𝔊\mathfrak{G} be the set of continuous maps from a closed interval [a,b][a,b] to itself so that an arbitrary element g𝔊g\in\mathfrak{G} satisfies the following two properties:

  1. 1.

    there exists m(a,b)m\in(a,b) with the map gg strictly decreasing on [a,m][a,m] and strictly increasing on [m,b][m,b].

  2. 2.

    g(a)>ag(a)>a, g(b)bg(b)\leq b, and g(x)<xg(x)<x for all x[m,b)x\in[m,b).

For g𝔊g\in\mathfrak{G}, let Π:={x[a,m]g(x)[a,m] and g2(x)=x}\Pi:=\{x\in[a,m]\mid g(x)\in[a,m]\textup{ and }g^{2}(x)=x\}. Now we are ready to state [Deng et al., 2022, Cor. 3]:

Proposition 2.3.

Let g𝔊g\in\mathfrak{G}. The map gg has an odd-period cycle if and only if g2(m)>mg^{2}(m)>m and g3(m)>max{xΠ}g^{3}(m)>\max\{x\in\Pi\} and the second iterate g2g^{2} is turbulent if and only if g2(m)>mg^{2}(m)>m and g3(m)min{xΠ}g^{3}(m)\geq\min\{x\in\Pi\}.

We keep the same notation ff, x¯\bar{x}, y¯\bar{y}, α\alpha, β\beta, and λ\lambda from Equation (1.1). We write f|E{\left.\kern-1.2ptf\right|_{E}} for the restriction of ff to the closed interval E:=[f(y¯λβ),f2(y¯λβ)+y¯λβ]E:=[f(\sqrt{\bar{y}\lambda\beta}),f^{2}(\sqrt{\bar{y}\lambda\beta})+\sqrt{\bar{y}\lambda\beta}]. (We will explain the significance of the number y¯λβ\sqrt{\bar{y}\lambda\beta} and the reason for the choice of EE in the next section.) Note that in the next section we will show that f(y¯λβ)<f2(y¯λβ)+y¯λβ)f(\sqrt{\bar{y}\lambda\beta})<f^{2}(\sqrt{\bar{y}\lambda\beta})+\sqrt{\bar{y}\lambda\beta}) (so, EE is a non-degenerate closed interval) and that f|E{\left.\kern-1.2ptf\right|_{E}} is a map to EE.

Now, we give several comments/interpretations on Theorem 1.2. First, in the next section, we will show that the condition y¯βx¯2(1α)2<λ<4y¯βx¯2(1α)2\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}} in Theorem 1.2 is the weakest condition for ff to be economically meaningful (to keep p>0p>0) and also for f|E{\left.\kern-1.2ptf\right|_{E}} to be in 𝔊\mathfrak{G}. As stated in Introduction, the upper bound for λ\lambda is not really interesting and the real meat is in the lower bounds in parts 2,3, and 4.

Second, it is well-know that the vertical stretch the graph of ff controls the existence of a chaos: if we stretch the graph further, we are likely to obtain a chaos. Now, looking at Equation (1.1), we see that if we make α\alpha, β\beta, y¯\bar{y} small, or x¯\bar{x} large, the ”valley” of the graph of ff goes deep down. Also, it is clear that a small λ\lambda makes the graph of ff ”flat”. Our lower bound L:=25y¯β9x¯2(1α)2L:=\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}} agrees with these observations: LL is an increasing function of α\alpha, β\beta, y¯\bar{y} and a decreasing function of x¯\bar{x}. In other words, it is easy to generate a chaos (small λ\lambda gives a chaos) if α\alpha, β\beta, or y¯\bar{y} is small or x¯\bar{x} is large. We give an economic explanation for this. By looking at the form of the excess demand function z(p)=y¯β/px¯(1α)z(p)=\bar{y}\beta/p-\bar{x}(1-\alpha) (or the demand function of each consumer for good xx, that is, x=αx¯x=\alpha\bar{x} for consumer 1 and x=βy¯px=\frac{\beta\bar{y}}{p} for consumer 2 respectively), we see that pp goes up very sharply (almost irrespective of the parameter values) after pp gets close to 0. To generate a chaos, pp needs to drop sharply afterwords, that is possible (or at least easy) if α\alpha, β\beta, or y¯\bar{y} is small or x¯\bar{x} is large for the following reasons: 1. if α\alpha or β\beta is small, the demand for good 11 is weak (thus pp drops sharply), 2. if y¯\bar{y} is small, then the demand of consumer 2 for good xx (that is excessively strong when pp is close to 0) drops sharply (since the budget for consumer 2 is tight), 3. if x¯\bar{x} is large, when pp is very high, a large excess supply happens.

Third, part 4 of Theorem 1.2 shows that if λ\lambda is sufficiently large, for any initial p>0p>0, our price dynamics eventually trap pp inside a (Li-Yorke) chaotic region EE. This is interesting since parts 1, 2, and 3 say nothing about what is going on outside of EE. This sort of analysis is not done in [Deng et al., 2022].

2.3 Odd period (but no period three) cycles

To end this section, we give an application of Theorem 1.2. By the Sharkovsky order in [Sharkovsky, 1964], we know that if the map ff has a cycle of period three, then it also has a cycle of any odd order. Thus, it is natural to guess that if λ\lambda is close to the lower bound for λ\lambda in Theorem 1.2 (but still above the lower bound), the map ff has an odd period cycle but no period three cycle. We give one example where this is actually the case. Using our concrete characterisation of the existence of an odd period cycle, we obtain:

Proposition 2.4.

Let x¯=4,y¯=2,α=0.75,β=0.5\bar{x}=4,\bar{y}=2,\alpha=0.75,\beta=0.5. If 2.77<λ3.002.77<\lambda\leq 3.00, then the map ff in Equation (1.1) has an odd period cycle but no period three cycle.

Note that in this case, by Theorem 1.2 a necessary and sufficient condition for the existence of an odd period cycle is 25y¯β9x¯2(1α)2=2.77<λ<4=4y¯βx¯2(1α)2\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}}=2.77<\lambda<4=\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}. (We use this condition in Section 6.) Our numerical computation (see Section 4 for details) shows that Proposition 2.4 is almost an if and only if statement: for λ3.01\lambda\geq 3.01, we get a period three cycle. Although many examples of this sort can be obtained by the same method, a complete characterisation for the existence of a period three cycle for a unimodal map is not known. (Thus it is not possible to obtain the precise λ\lambda value where a bifurcation happens.) We leave it for a future work.

3 Proof of Theorem 1.2

In the following proof, most results follow from direct (but a fairly complicated) algebraic calculations. We give some sketches of our manipulations while pointing some important steps out rather than writing all the details. All calculations can be checked by a computer algebra system, say, Magma [Bosma et al., 1997], Python [Rossum and Drake, 2009], etc.

First, for the function ff in Equation (1.1), we have f(p)=1y¯λβp2f^{\prime}(p)=1-\frac{\bar{y}\lambda\beta}{p^{2}} and f′′(p)=2y¯λβp3>0f^{\prime\prime}(p)=\frac{2\bar{y}\lambda\beta}{p^{3}}>0 for any p>0p>0. So, ff is strictly convex (unimodal) and takes its minimum at p=y¯λβp=\sqrt{\bar{y}\lambda\beta}. We sometimes write ss for y¯λβ\sqrt{\bar{y}\lambda\beta} to ease the notation. First of all, since we assume that p>0p>0, we must have f(p)>0f(p)>0 for any p>0p>0, hence f(s)>0f(s)>0. This gives that λ<4y¯βx¯2(1α)2\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}.

Next, we want (a restriction of) ff to be in 𝔊\mathfrak{G}. Note that any function ff in 𝔊\mathfrak{G} must have some mm in its interior of the domain, say [a,b][a,b], with ff strictly decreasing on [a,m][a,m] and strictly increasing on [m,b][m,b]. Since our ff is unimodal, this forces mm to be ss. Moreover ff needs to satisfy f(p)<pf(p)<p for all p[s,b)p\in[s,b). So in particular, we must have f(s)<sf(s)<s. This implies y¯βx¯2(1α)2<λ\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda. Note that s<f2(s)+ss<f^{2}(s)+s since f2(s)>0f^{2}(s)>0 (under the condition λ<4y¯βx¯2(1α)2\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}), so we have f(s)<s<f2(s)+sf(s)<s<f^{2}(s)+s. Now we set a:=f(s)a:=f(s) and b:=f2(s)+sb:=f^{2}(s)+s. (Now it is clear that [a,b][a,b] is non-degenerate.)

Lemma 3.1.

If y¯βx¯2(1α)2<λ<4y¯βx¯2(1α)2\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}, then f|E𝔊{\left.\kern-1.2ptf\right|_{E}}\in\mathfrak{G}.

Proof.

We need to show three things: (1) f(a)>af(a)>a and f(b)bf(b)\leq b. (2) f(p)<pf(p)<p for all p[s,b)p\in[s,b). (3) f|E{\left.\kern-1.2ptf\right|_{E}} is a map from EE to itself. We begin with (2). Let p[s,b)p\in[s,b). Then we have pf(p)=λ[y¯βpx¯(1α)]p-f(p)=-\lambda[\frac{\bar{y}\beta}{p}-\bar{x}(1-\alpha)]. Since λ>0\lambda>0 and y¯βpx¯(1α)y¯βsx¯(1α)\frac{\bar{y}\beta}{p}-\bar{x}(1-\alpha)\leq\frac{\bar{y}\beta}{s}-\bar{x}(1-\alpha), it is enough to show that y¯βsx¯(1α)<0\frac{\bar{y}\beta}{s}-\bar{x}(1-\alpha)<0. Now y¯βsx¯(1α)=y¯βy¯λβx¯(1α)\frac{\bar{y}\beta}{s}-\bar{x}(1-\alpha)=\frac{\bar{y}\beta}{\sqrt{\bar{y}\lambda\beta}}-\bar{x}(1-\alpha), so y¯βsx¯(1α)<0\frac{\bar{y}\beta}{s}-\bar{x}(1-\alpha)<0 is equivalent to λ>y¯βx¯2(1α)2\lambda>\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}, that is certainly true (since this is our assumption). Now we show that the same argument gives f(b)<bf(b)<b in (1). (This strict inequality is stronger than we need here, but we need it to prove part 4 of Theorem 1.2, see the proof of Lemma 3.6 below.) We have bf(b)=λ[y¯βf2(s)+sx¯(1α)]b-f(b)=-\lambda[\frac{\bar{y}\beta}{f^{2}(s)+s}-\bar{x}(1-\alpha)]. Since λ>0\lambda>0 and y¯βf2(s)+sx¯(1α)<y¯βsx¯(1α)\frac{\bar{y}\beta}{f^{2}(s)+s}-\bar{x}(1-\alpha)<\frac{\bar{y}\beta}{s}-\bar{x}(1-\alpha), it is enough to show that y¯βsx¯(1α)<0\frac{\bar{y}\beta}{s}-\bar{x}(1-\alpha)<0. We have already shown that this is true. Next we show that f(a)>af(a)>a. By a direct calculation, we have that f(a)a=f(f(s))f(s)=λ[y¯βf(s)x¯(1α)]>0f(a)-a=f(f(s))-f(s)=\lambda[\frac{\bar{y}\beta}{f(s)}-\bar{x}(1-\alpha)]>0 is equivalent to (λy¯βx¯(1α))2>0\left(\sqrt{\lambda}-\frac{\sqrt{\bar{y}\beta}}{\bar{x}(1-\alpha)}\right)^{2}>0. Now the assumption y¯βx¯2(1α)2<λ\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda forces f(a)a>0f(a)-a>0.

To prove (3), we need to show that the maximum value of f|E{\left.\kern-1.2ptf\right|_{E}} does not exceed b=f2(s)+sb=f^{2}(s)+s (that the minimum value of f|E{\left.\kern-1.2ptf\right|_{E}}, that is f(s)f(s), is in EE is clear). Since ff is unimodal, the maximum of ff on EE is taken either at aa or at bb. For the first case, we have f(a)=f(f(s))=f2(s)<f2(s)+m=bf(a)=f(f(s))=f^{2}(s)<f^{2}(s)+m=b. For the second case, we need f(b)bf(b)\leq b, but this is true by part (1) above.

We have proved part 1 of Theorem 1.2. We assume y¯βx¯2(1α)2<λ<4y¯βx¯2(1α)2\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}} for the rest of this section (actually for the rest of the paper). Now we are ready to prove parts 2 and 3 of the theorem. In view of Proposition 2.3 and Lemma 3.1, we only need to translate two conditions f2(m)>mf^{2}(m)>m and f3(m)>max{xΠ}f^{3}(m)>\max\{x\in\Pi\} (or f3(m)min{xΠ}f^{3}(m)\geq\min\{x\in\Pi\}) in terms of x¯\bar{x}, y¯\bar{y}, α\alpha, β\beta, and λ\lambda. First we show that

Lemma 3.2.

f2(s)>sf^{2}(s)>s if and only if 9y¯β4x¯2(1α)2<λ\frac{9\bar{y}\beta}{4{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda.

Proof.

Under the condition y¯βx¯2(1α)2<λ\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda, a direct computation shows that f2(s)>sf^{2}(s)>s is equivalent to 2x¯2(1α)2λ5x¯(1α)y¯βλ+3βy¯>02{\bar{x}}^{2}(1-\alpha)^{2}\lambda-5\bar{x}(1-\alpha)\sqrt{\bar{y}\beta\lambda}+3\beta\bar{y}>0. Now the statement follows. ∎

Next we show that

Lemma 3.3.

If 9y¯β4x¯2(1α)2<λ\frac{9\bar{y}\beta}{4{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda, then the set Π={x[a,s]f(x)[a,s] and f2(x)=x}\Pi=\{x\in[a,s]\mid f(x)\in[a,s]\textup{ and }f^{2}(x)=x\} is a singleton, namely Π={y¯βx¯(1α)}\Pi=\{\frac{\bar{y}\beta}{\bar{x}(1-\alpha)}\} (the unique fixed point for f|E{\left.\kern-1.2ptf\right|_{E}}).

Proof.

First, we compute the fixed points of f|E{\left.\kern-1.2ptf\right|_{E}}. Solving f(p)=pf(p)=p for p>0p>0, we obtain p=y¯βx¯(1α)p=\frac{\bar{y}\beta}{\bar{x}(1-\alpha)}. So, ff has the unique fixed point, which we name zz. It is clear that z<sz<s (this follows from y¯βx¯2(1α)2<λ\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda) and aza\leq z (since a=f(s)a=f(s) is the minimum of ff). So, we have zΠz\in\Pi. Next, we compute the period 22 points for ff on [a,s][a,s]. Solving f2(p)=pf^{2}(p)=p for pp (with Python), we get p=y¯βx¯(1α),x¯αλ2+x¯λ212x¯2α2λ22x¯2αλ22y¯βλ+x¯2λ2,x¯αλ2+x¯λ2+12x¯2α2λ22x¯2αλ22y¯βλ+x¯2λ2p=\frac{\bar{y}\beta}{\bar{x}(1-\alpha)},-\frac{\bar{x}\alpha\lambda}{2}+\frac{\bar{x}\lambda}{2}-\frac{1}{2}\sqrt{{\bar{x}}^{2}\alpha^{2}\lambda^{2}-2{\bar{x}}^{2}\alpha\lambda^{2}-2\bar{y}\beta\lambda+{\bar{x}}^{2}\lambda^{2}},-\frac{\bar{x}\alpha\lambda}{2}+\frac{\bar{x}\lambda}{2}+\frac{1}{2}\sqrt{{\bar{x}}^{2}\alpha^{2}\lambda^{2}-2{\bar{x}}^{2}\alpha\lambda^{2}-2\bar{y}\beta\lambda+{\bar{x}}^{2}\lambda^{2}}. The first pp is zz (the fixed point), and the other two points are the period 22 points. We name the last two points as w1w_{1} and w2w_{2} respectively (w1w2w_{1}\leq w_{2}). If we show that s<w2s<w_{2}, we are done. A direct calculation (with Python) shows that s<w2s<w_{2} is equivalent to 9y¯β4x¯2(1α)2<λ\frac{9\bar{y}\beta}{4{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda (this is our assumption). ∎

Now we assume 9y¯β4x¯2(1α)2<λ\frac{9\bar{y}\beta}{4{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda. (So Π\Pi is a singleton.) Finally we show

Lemma 3.4.

f3(s)>max{xΠ}f^{3}(s)>\max\{x\in\Pi\} if and only if 25y¯β9x¯2(1α)<λ\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)}<\lambda.

Proof.

This calculation is a bit involved, so we give some details. Let zz be the fixed point of ff. Since we know that max{xΠ}={z}\max\{x\in\Pi\}=\{z\}, we have

f3(s)max{xΠ}\displaystyle f^{3}(s)-\max\{x\in\Pi\} =f(f2(s))z\displaystyle=f(f^{2}(s))-z
=f(f2(s))f(z)\displaystyle=f(f^{2}(s))-f(z)
=(f2(s)+λ[y¯βf2(s)x¯(1α)])(z+λ[y¯βzx¯(1α)])\displaystyle=\left(f^{2}(s)+\lambda\left[\frac{\bar{y}\beta}{f^{2}(s)}-\bar{x}(1-\alpha)\right]\right)-\left(z+\lambda\left[\frac{\bar{y}\beta}{z}-\bar{x}(1-\alpha)\right]\right)
=f2(s)z+λ[y¯βf2(s)y¯βz]\displaystyle=f^{2}(s)-z+\lambda\left[\frac{\bar{y}\beta}{f^{2}(s)}-\frac{\bar{y}\beta}{z}\right]
=f2(s)zy¯βλ[f2(s)zzf2(s)]\displaystyle=f^{2}(s)-z-\bar{y}\beta\lambda\left[\frac{f^{2}(s)-z}{zf^{2}(s)}\right]
=(f2(s)z)(1y¯βλf2(s)z)\displaystyle=(f^{2}(s)-z)\left(1-\frac{\bar{y}\beta\lambda}{f^{2}(s)z}\right)

We consider the first term of the last expression. A direct calculation shows that s>zs>z if and only if λ>y¯βx¯2(1α)2\lambda>\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}} (which we already assumed). So, we have f2(s)z>f2(s)s>0f^{2}(s)-z>f^{2}(s)-s>0. (The last inequality followed from Lemma 3.2.) Next, a direct calculation shows that the second term of the last expression is positive if and only if 3x¯2(1α)2λ8x¯(1α)y¯βλ+5y¯β>03{\bar{x}}^{2}(1-\alpha)^{2}\lambda-8\bar{x}(1-\alpha)\sqrt{\bar{y}\beta\lambda}+5\bar{y}\beta>0. Now we see that under the condition y¯βx¯2(1α)2<λ\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda, this is equivalent to 25y¯β9x¯2(1α)2<λ\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda. ∎

It is clear that min{xΠ}={z}\min\{x\in\Pi\}=\{z\}. So by the same argument, we obtain

Lemma 3.5.

f3(s)min{xΠ}f^{3}(s)\geq\min\{x\in\Pi\} if and only if 25y¯β9x¯2(1α)λ\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)}\leq\lambda.

Note that 9y¯β4x¯2(1α)22.25y¯βx¯2(1α)2<2.78y¯βx¯2(1α)225y¯β9x¯2(1α)2\frac{9\bar{y}\beta}{4{\bar{x}}^{2}(1-\alpha)^{2}}\approx\frac{2.25\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\frac{2.78\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}\approx\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}}. By Proposition 2.3 and Lemmas 3.1, 3.2, 3.3, 3.4, 3.5, we have proved parts 2 and 3 of Theorem 1.2. Finally, we are left to show (part 4 of Theorem 1.2):

Lemma 3.6.

If 25y¯β9x¯2(1α)2λ<4y¯βx¯2(1α)2\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}}\leq\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}, then the closed interval EE is attracting for ff, that is, fn(p)Ef^{n}(p)\in E for some n>0n>0.

Proof.

Since E=[a,b]E=[a,b], we need to prove: (1) if 0<p<a0<p<a, then fn1(p)Ef^{n_{1}}(p)\in E for some n1n_{1}\in\mathbb{N}, (2) if b<pb<p, then fn2(p)Ef^{n_{2}}(p)\in E for some n2n_{2}\in\mathbb{N}. Note that if 0<p<a0<p<a (the first case), then f(p)f(s)=af(p)\geq f(s)=a since f(s)f(s) is a global minimum for ff, thus we just need to consider the second case. Let p>bp>b. Then we have f(p)p=λ[y¯βpx¯(1α)]λ[y¯βbx¯(1α)]=f(b)b<0f(p)-p=\lambda[\frac{\bar{y}\beta}{p}-\bar{x}(1-\alpha)]\leq\lambda[\frac{\bar{y}\beta}{b}-\bar{x}(1-\alpha)]=f(b)-b<0. (The last strict inequality follows from y¯βx¯2(1α)2<λ\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda, see the proof of Lemma 3.1) This shows that if p>bp>b, in each iteration of ff, the value of pp drops by bf(b)>0b-f(b)>0 at least. So, to prove that pp is attracted to EE, it is enough to show that the size of the drop is not too big (thus pp does not jump over EE). Therefore, it is sufficient to have pf(p)bap-f(p)\leq b-a. Since this is equivalent to λ[x¯(1α)]f2(s)+sf(s)\lambda[\bar{x}(1-\alpha)]\leq f^{2}(s)+s-f(s) and sf(s)>0s-f(s)>0 under our assumption on λ\lambda, it suffices to show that λ[x¯(1α)]f2(s)\lambda[\bar{x}(1-\alpha)]\leq f^{2}(s). After some computations, we obtain

f2(s)λ[x¯(1α)]=λ[3x¯2(1α)2λ8x¯y¯(1α)βλ+5y¯β]x¯(α1)λ+2y¯βλ.f^{2}(s)-\lambda[\bar{x}(1-\alpha)]=\frac{\lambda\left[3{\bar{x}}^{2}(1-\alpha)^{2}\lambda-8\bar{x}\sqrt{\bar{y}}(1-\alpha)\sqrt{\beta\lambda}+5\bar{y}\beta\right]}{\bar{x}(\alpha-1)\lambda+2\sqrt{\bar{y}\beta\lambda}}.

We check that in the last expression, the denominator is strictly positive if 0<λ<4y¯βx¯2(1α)20<\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}} (this follows from our assumption). Also, we see that the numerator is positive if 25y¯β9x¯2(1α)2λ\frac{25\bar{y}\beta}{9{\bar{x}}^{2}(1-\alpha)^{2}}\leq\lambda.

4 Proofs of Propositions 1.1 and 2.4

Proof of Proposition 1.1.

The first part (this paragraph) of the following argument is a replicate of [Bhattacharya and Majumdar, 2007, Proof of Prop. 9.10]. We include this to make the paper self-contained. Let (α,β)=(0.75,0.5)(\alpha,\beta)=(0.75,0.5), (x¯,y¯)=(4,2)(\bar{x},\bar{y})=(4,2), and λ=3.61\lambda=3.61. Then we have f(3.75)=1.9f(3.75)=1.9, f(1.9)=0.19f(1.9)=0.19, f(0.19)=15.58>3.75f(0.19)=15.58>3.75. So, by the Li-Yorke theorem, there exists a period three cycle. Also, by [Bhattacharya and Majumdar, 2007, Prop. 9.10 and its proof], we know that the choices of (α,β)(\alpha,\beta) and λ\lambda are robust (for this fixed (x¯,y¯)(\bar{x},\bar{y})), so it is clear that there exist open sets AA and CC as in Proposition 1.1. Thus, the only thing we need to show is that the choice of (x¯,y¯)(\bar{x},\bar{y}) is also robust. This follows from the following numerical/graphical argument.

In Figure 2, we plot the graphs of pt+1=f(pt)p_{t+1}=f(p_{t}) (dashed curve) and pt+3=f3(pt)p_{t+3}=f^{3}(p_{t}) (solid curve) together with the 4545^{\circ} line. We see that the solid curve crosses the 4545^{\circ} line at pp^{*} (the big dot) between p=2p=2 and p=3p=3. A numerical computation gives p2.66p^{*}\cong 2.66, and it is clear that pp^{*} is a point of period three. (From the picture we see that pp^{*} is not a fixed point.) Since the solid curve crosses (but not touching) the 4545^{\circ} line at pp^{*} and f3f^{3} is continuous in x¯\bar{x} and y¯\bar{y}, a small perturbation of x¯\bar{x} and y¯\bar{y} does not affect the existence of a period three cycle. So, there exists an open set BB as required. (Actually, this graphical argument gives the existence of open sets AA and CC as well.) ∎

Proof of Proposition 2.4.

Let (x¯,y¯)=(4,2)(\bar{x},\bar{y})=(4,2), (α,β)=(0.75,0.5)(\alpha,\beta)=(0.75,0.5), and 2.77<λ3.002.77<\lambda\leq 3.00. Then by Theorem 1.2 it is clear that the map ff has an odd period cycle. Only thing we need to show now is that ff does not have a period three cycle. We use a graphical/numerical argument. In Figures 34, and 5 we plot the graphs of ff (dashed curve) and f3f^{3} (solid curve) with the 4545^{\circ} line. A general pattern is that if we increase λ\lambda, the graph of ff gets slightly deeper down, and f3f^{3} becomes more ”wavy”. (Compare Figure 2 (λ=3.61\lambda=3.61) to Figure 3 (λ=2.78\lambda=2.78) for example.) For Figure 3 (λ=2.78\lambda=2.78), using a numerical computation, we have checked that the solid curve does not touch/cross the 4545^{\circ} line except at p=1p=1 (the fixed point of ff). Thus, there is no period three cycle in this case. Likewise, for Figure 4 (λ=3\lambda=3), there is no period three cycle (although the solid curve seems touching the 4545^{\circ} line but a numerical computation shows that it is not touching). However, if λ=3.01\lambda=3.01, a numerical computation shows that there exists a period three cycle. In Figure 5, we plotted the λ=3.05\lambda=3.05 case since the λ=3.01\lambda=3.01 case is indistinguishable from the λ=3\lambda=3 case in picture. It is clear that the solid curve crosses the 4545^{\circ} line near p=4p=4 (that is not a fixed point of ff).

Refer to caption
Figure 2: λ=3.61\lambda=3.61
Refer to caption
Figure 3: λ=2.78\lambda=2.78
Refer to caption
Figure 4: λ=3.0\lambda=3.0
Refer to caption
Figure 5: λ=3.05\lambda=3.05

5 Ergodic properties: an experimental approach

Our (numerical/theoretical) argument in this and the next sections use ergodic theory. Here, we give a quick review of ergodic theory. If the reader is familiar with ergodic theory, skip Subsection 5.1. Our basic references for ergodic theory are classical [Collet and Eckmann, 1980][Day, 1998], and [W. de Melo, 1993]. Note that our strategy (philosophy) in this and the next sections stems from [Lyubich, 2012] and [Shen and van Strien, 2014] (these are quite readable expository articles on recent developments of unimodal dynamics). We stress that a deep result by Avila et.al. (Proposition 6.3) theoretically supports our argument.

5.1 Background from ergodic theory

Let II be a compact interval on \mathbb{R}. Let \mathcal{B} denote the Borel σ\sigma-algebra of II. Let ζ:[0,]\zeta:\mathcal{B}\rightarrow[0,\infty] be a measure on II. A measurable map g:IIg:I\rightarrow I is called ergodic with respect to ζ\zeta if whenever g1(A)=Ag^{-1}(A)=A for AA\in\mathcal{B}, then either ζ(A)=0\zeta(A)=0 or ζ(I\A)=0\zeta(I\backslash A)=0. We say that a measure ζ\zeta is g-invariant if ζ(g1(A))=ζ(A)\zeta(g^{-1}(A))=\zeta(A) for any AA\in\mathcal{B}. We write μ\mu for the Lebesgue measure on II. A measure ζ\zeta is called absolutely continuous with respect to μ\mu if whenever μ(A)=0\mu(A)=0 for AA\in\mathcal{B} then ζ(A)=0\zeta(A)=0. We write an ”acim” for a measure that is absolutely continuous with respect to μ\mu and gg-invariant (if gg is clear from the context, we just say ”invariant”). Remember that if ζ\zeta is an acim then there exists a (μ\mu-integrable) density function (Radon-Nikodym derivative) ξ:I\xi:I\rightarrow\mathbb{R} with dζ=ξdμd\zeta=\xi d\mu, see [Collet and Eckmann, 1980, \@slowromancap[email protected]]. It is well-known that if gg is ergodic with respect to an acim ζ\zeta, then we can give an estimate of ξ\xi (hence an estimate of ζ\zeta) using some iterates of gg, see [Day, 1998, 8.5.2 and 8.5.3] (we use this argument below in Theorem 5.11).

Here, we recall (a special case of) the famous Birkhoff’s ergodic theorem [Day, 1998, Thm. 8.2], which states that under certain conditions the time average of gg is equal to its space average:

Proposition 5.1.

If gg has an absolutely continuous invariant measure ζ\zeta, gg is ergodic with respect to ζ\zeta, and gg is ζ\zeta-integrable, then

limn1nk=0n1gk(x)=Ig𝑑ζ for ζ-almost all x.\lim_{n\rightarrow\infty}\frac{1}{n}\sum_{k=0}^{n-1}g^{k}(x)=\int_{I}g\;d\zeta\text{ for $\zeta$-almost all $x$.} (5.1)

In particular, Proposition 5.1 says that (if the conditions are met) the time average of gg converges to a constant (for almost all xx), in other words, we can ”predict” the future on average. We call the integral on the right-hand side (or sometimes the sum inside the limit on the left-hand side) of Equation (5.1), that is Ig𝑑ζ\int_{I}g\;d\zeta (or sometimes 1nk=0n1gk(x)\frac{1}{n}\sum_{k=0}^{n-1}g^{k}(x)), the ergodic sum of gg (which we are referring to would be clear from the context). In the following, we try to apply Proposition 5.1 to our price dynamics defined by f|E{\left.\kern-1.2ptf\right|_{E}} (with the condition y¯βx¯2(1α)2<λ<4y¯βx¯2(1α)2\frac{\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}}<\lambda<\frac{4\bar{y}\beta}{{\bar{x}}^{2}(1-\alpha)^{2}} to force that p>0p>0 and f|E𝔊{\left.\kern-1.2ptf\right|_{E}}\in\mathfrak{G}). In the following, we write ff for f|E{\left.\kern-1.2ptf\right|_{E}} to ease the notation. So we need to show that ff has an absolutely continuous invariant measure ζ\zeta, ff is ergodic with respect to ζ\zeta, and ff is ζ\zeta-integrable. (The last condition is clear since ff is continuous and bounded on the compact interval EE.)

5.2 SS-unimodal maps

In general, it is pretty difficult to prove the existence of an acim for a measurable transformation gg except for some special cases such as when gg is ”expansive” or ”iteratively expansive” as studied in a classical paper of Lasota and Yorke [Lasota and Yorke, 1973]. Recall that gg is called expansive if gg is piecewise C2C^{2} and |g(x)|>1|g^{\prime}(x)|>1 for μ\mu-almost all xx. A typical example of an expansive map is a well-studied ”tent map” [Day, 1998, 8.5.4 and 8.5.5]. Further recall that a slightly more general ”iteratively expansive” map, that is a piecewise C2C^{2} map with |g(x)|n>1|g^{\prime}(x)|^{n}>1 for some positive integer n>1n>1 for μ\mu-almost all xx. From [Lasota and Yorke, 1973], we know that a density function ξ\xi with dζ=ξdμd\zeta=\xi d\mu is a fixed point of the Perron-Frobenius operator PP from the set of measurable function on II to itself, and that a uniform expansion of a map gg helps a lot to make PP a ”nice” operator. As a result, it is not too hard to prove the existence of an acim in these (iteratively) expansive cases. See [W. de Melo, 1993, Chap. 5][Shen and van Strien, 2014, Sec. 4], and [Lyubich, 2012] for an overview of this problem and also see [Sato and Yano, 2012] and [Sato and Yano, 2013] for applications of an acim for an iteratively expansive map in economics.

In this paper, our function ff is not (even iteratively) expansive since it has a critical point ss. (So this is a hard case.) Thus, to establish the existence of an acim, we need some deep analytical results to investigate the counter play between the contraction of ff near the critical point ss and the expansion of ff at f(s)f(s) (that is far from ss). Now, we restrict the class of functions we consider to so-called SS-unimodal maps (due to Singer [Singer, 1978]), whose ergodic properties are well studied, see [Collet and Eckmann, 1980, Part \@slowromancapii@][W. de Melo, 1993, Chap. 5][Avila and Moreira, 2005][Avila et al., 2003] for example. (Our function ff is actually SS-unimodal as we will show below.) Let gg be a measurable transformation defined on a compact interval I=[a,b]I=[a,b] of \mathbb{R}.

Definition 5.2.

A function gg is called SS-unimodal if the following conditions are satisfied:

  1. 1.

    gg is C3C^{3}.

  2. 2.

    gg is unimodal with the unique critical point x=cx=c in (a,b)(a,b) and g(x)0g^{\prime}(x)\neq 0 except when x=cx=c.

  3. 3.

    The Schwarzian derivative of gg, that is, Sg(x)=g′′′(x)g(x)32(g′′(x)g(x))2Sg(x)=\frac{g^{\prime\prime\prime}(x)}{g^{\prime}(x)}-\frac{3}{2}\left(\frac{g^{\prime\prime}(x)}{g^{\prime}(x)}\right)^{2} is negative except at x=cx=c.

Moreover the critical point x=cx=c is called non-flat and of order ll if there are positive constants O1O_{1}, O2O_{2} with

O1|xc|l1|g(x)|O2|xc|l1.O_{1}|x-c|^{l-1}\leq|g^{\prime}(x)|\leq O_{2}|x-c|^{l-1}.

Note that a well-known ”logistic map” (g(x)=rx(1x)g(x)=rx(1-x) for r(0,4]r\in(0,4]) is SS-unimodal. The first key result in this section is

Proposition 5.3.

If gg is SS-unimodal with a non-flat critical point and without an attracting periodic orbit, then gg is ergodic with respect to any absolutely continuous measure ζ\zeta.

Proof.

Let gg be SS-unimodal with a non-flat critical point and without an attracting periodic orbit. Suppose that g1(A)=Ag^{-1}(A)=A for some AA\in\mathcal{B}. Then we have μ(g1(A))=μ(A)\mu(g^{-1}(A))=\mu(A). From [W. de Melo, 1993, Thm. 1.2] we know that gg is ergodic with respect to μ\mu, so we obtain that μ(A)=0\mu(A)=0 or μ(I\A)=0\mu(I\backslash A)=0. This yields ζ(A)=0\zeta(A)=0 or ζ(I\A)=0\zeta(I\backslash A)=0 since ζ\zeta is absolutely continuous. ∎

To end this subsection, we prove that

Lemma 5.4.

The function ff (restricted to E=[f(y¯λβ),f2(y¯λβ)+y¯λβ]=[a,b]E=[f(\sqrt{\bar{y}\lambda\beta}),f^{2}(\sqrt{\bar{y}\lambda\beta})+\sqrt{\bar{y}\lambda\beta}]=[a,b]) is SS-unimodal and the unique critical point cc of ff is non-flat and of order 22.

Proof.

First, we show that f(p)=p+λ[y¯β/px¯(1α)]f(p)=p+\lambda[\bar{y}\beta/p-\bar{x}(1-\alpha)] is C3C^{3}. We have f(p)=1λy¯β/p2f^{\prime}(p)=1-\lambda\bar{y}\beta/{p^{2}}, f′′(p)=2λy¯β/p3f^{\prime\prime}(p)=2\lambda\bar{y}\beta/{p^{3}}, and f′′′(p)=6λy¯β/p4f^{\prime\prime\prime}(p)=-6\lambda\bar{y}\beta/{p^{4}}. Since pp is positive (so it is not zero), it is clear that ff is C3C^{3}. Second, from Section 2, we know that ff is unimodal and has a unique critical point at p=λy¯β=sp=\sqrt{\lambda\bar{y}\beta}=s in (a,b)(a,b). It is easy to see that f(p)0f^{\prime}(p)\neq 0 except when p=sp=s. Third, we have Sf(p)=6λy¯β/(p2λy¯β)2<0Sf(p)=-6\lambda\bar{y}\beta/(p^{2}-\lambda\bar{y}\beta)^{2}<0 except when p=λy¯β=sp=\sqrt{\lambda\bar{y}\beta}=s. Finally, we show that the critical point p=sp=s is non-flat of order l=2l=2. It is clear that |f(p)||f^{\prime}(p)| is strictly concave and monotone increasing on the compact interval [s,b][s,b]. Also note that the righthand derivative of |f(p)||f^{\prime}(p)| at ss is 2/λy¯β2/\sqrt{\lambda\bar{y}\beta}. So f(p)|[s,b]{\left.\kern-1.2ptf^{\prime}(p)\right|_{[s,b]}} is bounded below by f(b)/(bs)|ps|f^{\prime}(b)/(b-s)|p-s| and is bounded above by 2/λy¯β|ps|2/\sqrt{\lambda\bar{y}\beta}|p-s|. Likewise, we have that |f(p)||f^{\prime}(p)| is strictly convex and monotone decreasing on [a,s][a,s]. Also, the lefthand derivative of |f(p)||f^{\prime}(p)| is 2/λy¯β-2/\sqrt{\lambda\bar{y}\beta}. Therefore, f(p)|[a,s]{\left.\kern-1.2ptf^{\prime}(p)\right|_{[a,s]}} is bounded below by 2/λy¯β|ps|2/\sqrt{\lambda\bar{y}\beta}|p-s| and is bounded above by f(a)(sa)|ps|f^{\prime}(a)(s-a)|p-s|. Thus we see that f(p)f^{\prime}(p) (on EE) is bounded below by min{f(b)/(bs),2/λy¯β}|ps|\min\{f^{\prime}(b)/(b-s),2/\sqrt{\lambda\bar{y}\beta}\}|p-s| and is bounded above by max{2/λy¯β,f(a)(sa)}|ps|\max\{2/\sqrt{\lambda\bar{y}\beta},f^{\prime}(a)(s-a)\}|p-s|. ∎

5.3 Our strategy and the existence of an acim

The second key result in this section is

Proposition 5.5.

[Collet and Eckmann, 1980, Thm. \@slowromancap[email protected]] If gg is SS-unimodal, then every stable periodic orbit attracts at least one of aa, bb, or cc (i.e. the endpoints of II or the critical point of gg).

Proposition 5.5 means that all ”visible” orbits (in numerical experiments) are orbits containing aa, bb, or cc only (in the long run). We consider that only these visible orbits are meaningful in economics (or in real life) since it is widely believed that every economic modelling is some sort of an approximation of real economic activities and contains inevitable errors. We know that there are totally different point of view for economic modellings, but we do not argue here. Here is the third key result for this section:

Proposition 5.6.

[Collet and Eckmann, 1980, Cor. \@slowromancap[email protected]] If gg is SS-unimodal, then gg has at most one stable periodic orbit, plus possibly a stable fixed point. If the critical point cc is not attracted to a stable periodic orbit, then gg has no stable periodic orbit.

In this paper, we interpret our numerical calculations based on Propositions 5.5 and 5.6. In particular, we look at the orbit starting from the critical point ss, that is {s,f(s),f2(s),}\{s,f(s),f^{2}(s),\cdots\} (we call this orbit the ”critical orbit”). If the critical orbit seems to eventually converge to a periodic orbit, we conclude that we can see the future: the average price in the long run will be the average price in this attracting periodic orbit. Note that in this case, ff is neither ergodic nor has an acim since most fn(s)f^{n}(s) accumulate around this attracting periodic orbit, but we do not care (since we can still predict the future). Otherwise, we compute (or give an estimate for) the following Lyapunov exponent at the critical point p=sp=s since the existence of a positive Lyapunov exponent at the critical point implies that the critical orbit is repelling and also is a strong indication for the existence of a chaos (hence the existence of an acim, see (CE1) in Proposition 5.8 below):

Definition 5.7.

limn1ni=1nln|Dgn(g(c))|\lim_{n\rightarrow\infty}\frac{1}{n}\sum_{i=1}^{n}\ln{|Dg^{n}(g(c))|} is called the Lyapunov exponent of gg at cc (if the limit exists).

If the Lyapunov exponent (at cc) is positive, we test one of the following well-known sufficient conditions (within some numerical bound) to confirm the existence of an acim.

Proposition 5.8.

Suppose that gg is SS-unimodal, gg has no attracting periodic orbit, and the critical point cc is non-flat. Then gg has a unique acim ζ\zeta and gg is ergodic with respect to ζ\zeta if one of the following conditions is satisfied:

  1. 1.

    Collet-Eckmann conditions (together with some regularity conditions) [Collet and Eckmann, 1983, Sec. 1]:

    (CE1)lim infn1nln|Dgn(g(c))|>0,\displaystyle\textup{(CE1)}\;\;\liminf_{n\rightarrow\infty}\frac{1}{n}\ln\left|Dg^{n}(g(c))\right|>0,
    (CE2)lim infn1ninfpgn(c)ln|Dgn(p))|>0.\displaystyle\textup{(CE2)}\;\;\liminf_{n\rightarrow\infty}\frac{1}{n}\inf_{p\in g^{-n}(c)}\ln\left|Dg^{n}(p))\right|>0.
  2. 2.

    Misiurewicz condition [Misiurewicz, 1981, Thms. 6.2 and 6.3]: The ω\omega-limit set of cc does not contain cc, that is,

    (MC)cn0{gi(c):in}¯.(MC)\;\;c\notin\bigcap_{n\geq 0}\overline{\{g^{i}(c):i\geq n\}}.
  3. 3.

    Nowicki-Van Strien summation condition [Nowicki and van Strien, 1991, Main Thm.]: if ll is the order of cc, then

    (SC)n=1|Dgn(g(c))|1/l<.\displaystyle(SC)\;\;\sum_{n=1}^{\infty}|Dg^{n}(g(c))|^{-1/l}<\infty.

If one of the conditions in Proposition 5.8 is satisfied (within some numerical limitation), we conclude that we can predict the future by Proposition 5.1. We must admit that our argument in this section is not rigorous (we hope to make it rigorous in the future), but we believe that we have provided enough (numerical/theoretical) evidence to support it. We stress that it is very hard to prove the existence of an acim for any non-expansive function gg (even for an SS-unimodal gg) by a rigorous analytic argument. There are only a few known examples of such, see a famous g(x)=4x(1x)g(x)=4x(1-x) example due to Ulam and Neumann [Ulam and von Neumann, 1947], also see [Misiurewicz, 1981, Sec. 7 Examples] for more examples.

In this paper, we test Condition 3 (SC) in Proposition 5.8 since it is easy to compute (numerically) and covers the most general class of functions, see [Shen and van Strien, 2014, 4.2] for a comparison of these three sufficient conditions for the existence of an acim, also see [W. de Melo, 1993, Chap. \@slowromancapv@, Sec. 4] for more on (SC). We found that (CE2) was hard to compute since the set gn(c)g^{-n}(c) can be very large for a large nn. Also, the ω\omega-limit set of cc was difficult to compute for us although (MC) is theoretically beautiful. (For a numerical computation, it is not clear where to set the numerical bound to estimate the ω\omega-limit set.)

Remark 5.9.

Roughly speaking, all three condtions (CE1), (MC), and (SC) are basically testing the same thing: they (more or less) guarantee that the critical orbit does not accumulate around the critical point cc. (So, on the critical orbit, the expansion far from the critical point wins against the contraction near the critical point.)

5.4 λ=3.61\lambda=3.61 case

For the rest of the paper, to simplify the exposition and to obtain sharp numerical results, we fix (α,β)=(0.75,0.5)(\alpha,\beta)=(0.75,0.5), (x¯,y¯)=(4,2)(\bar{x},\bar{y})=(4,2). Also, in this subsection, we fix λ=3.61\lambda=3.61 (in the next subsection, we let λ\lambda vary). Then we have E=[0.19,17.48]E=[0.19,17.48]. In this case, the dynamics exhibit a Li-Yorke chaos as shown in the previous sections. It is easy to see that a similar analysis (as follows) can be done for any α,β,x¯,y¯,λ\alpha,\beta,\bar{x},\bar{y},\lambda.

Here is our main result in this section. (We can predict the future even if ff is chaotic.)

Theorem 5.10.

There exists an acim ζ\zeta (whose estimate is as in Figure 9) for ff on EE. Moreover, we have limn1nk=0n1fk(p)4.5\lim_{n\rightarrow\infty}\frac{1}{n}\sum_{k=0}^{n-1}f^{k}(p)\approx 4.5 for ζ\zeta-almost all pEp\in E.

A few comments are in order. Although we give an estimate of ζ\zeta in Figure 9, it is hard to give an explicit formula for ζ\zeta (thus it is hard to express ζ\zeta in a concrete way). Also, looking at Figure 9 (and its numerical data), we see that the density of pp is positive in [0.19,15.58]=[f(s),f2(s)][0.19,15.58]=[f(s),f^{2}(s)] and is zero in [15.58,17.48]=[f2(s),f2(s)+s][15.58,17.48]=[f^{2}(s),f^{2}(s)+s]. Thus, without a loss, if one wishes, Theorem 5.10 can be stated (in a more readable form) as follows:

Theorem 5.11.

limn1nk=0n1fk(p)4.5\lim_{n\rightarrow\infty}\frac{1}{n}\sum_{k=0}^{n-1}f^{k}(p)\approx 4.5 for μ\mu-almost all pE~:=[f(s),f2(s)]=[0.19,15.58]p\in\tilde{E}:=[f(s),f^{2}(s)]=[0.19,15.58].

Now we start looking at the model closely. First, following our strategy as in the last subsection, we consider the critical orbit of ff. Note that the critical point of ff is s=y¯λβ=1.9s=\sqrt{\bar{y}\lambda\beta}=1.9. Using the first 100100 iterates of fn(s)f^{n}(s), we obtain Figure 6 that shows a chaotic behaviour of the iterates of ff.

Refer to caption
Figure 6: The first 100100 iterates of fn(s)f^{n}(s) look chaotic

In particular, it seems that ff has no attracting periodic orbit. To convince the reader that this really the case, we give an estimate for the Lyapunov exponent (using the first 10000 terms of fn(s)f^{n}(s)). Figure 7 shows that the first 1000 terms are enough to estimate the Lyapunov exponent (but we used the first 10000 terms to be safe). We obtain

Lemma 5.12.

limn1nln|Dfn(f(s))|110000ln|Df10000(f(s))|0.5882>0\lim_{n\rightarrow\infty}\frac{1}{n}\ln{|Df^{n}(f(s))|}\approx\frac{1}{10000}\ln{|Df^{10000}(f(s))|}\approx 0.5882>0.

Remark 5.13.

Let c0=s,c1=f(c0),c2=f(c1),,cn=f(cn1)c_{0}=s,c_{1}=f(c_{0}),c_{2}=f(c_{1}),\cdots,c_{n}=f(c_{n-1}). Note that to compute the Lyapunov exponent we have used the chain rule, that is, Dfn(f(s))=f(c1)×f(c2)×f(c3)××f(cn)Df^{n}(f(s))=f^{\prime}(c_{1})\times f^{\prime}(c_{2})\times f^{\prime}(c_{3})\times\cdots\times f^{\prime}(c_{n}) (so it is not hard to compute).

Refer to caption
Figure 7: Convergence of the Lyapunov exponent

Now we check the Nowicki-Van Strien summation condition (SC). Figure 8 shows that the sum in (SC) stabilises if we use the first 100 terms. To be safe, we use 1000 terms to estimate the infinite sum in (SC). We obtain

Refer to caption
Figure 8: Convergence of the sum in SC
Lemma 5.14.

n=1|Dfn(f(s))|1/ln=11000|Dfn(f(s))|1/l0.935065399839560\sum_{n=1}^{\infty}|Df^{n}(f(s))|^{-1/l}\approx\sum_{n=1}^{1000}|Df^{n}(f(s))|^{-1/l}\approx 0.935065399839560.

Remark 5.15.

Since Lemma 5.14 is crucial for our argument, we have double-checked (SC) with 100000 terms obtaining n=1100000|Dfn(f(s))|1/l0.935065399839560\sum_{n=1}^{100000}|Df^{n}(f(s))|^{-1/l}\approx 0.935065399839560. (This is the same number as in Lemma 5.14!.)

Now we conclude that ff has a unique acim ζ\zeta and ff is ergodic with respect to ζ\zeta by Proposition 5.8. Next, in Figure 9 using fn(c)f^{n}(c) with nn from 10001000 to 1000010000 (after removing the effect of a transient period) we obatin an estimate of the density function (Radon-Nikodym derivative) ξ:E\xi:E\rightarrow\mathbb{R} with dζ=ξdμd\zeta=\xi d\mu.

Refer to caption
Figure 9: Estimate of the density function

Next, we directly compute the ergodic sum of fn(p)f^{n}(p) using the initial p=sp=s and the first 100000100000 terms of fn(p)f^{n}(p). We get

Lemma 5.16.

limn1nk=0n1fk(s)1100000k=099999fk(s)4.483627795147089\lim_{n\rightarrow\infty}\frac{1}{n}\sum_{k=0}^{n-1}f^{k}(s)\approx\frac{1}{100000}\sum_{k=0}^{99999}f^{k}(s)\approx 4.483627795147089.

Note that Figure 10 shows that the ergodic sum converges if we use more than 20002000 terms to estimate it (we use 100000100000 terms to be safe).

Refer to caption
Figure 10: Convergence of the ergodic sum of fn(c)f^{n}(c)

To end this section, we compute the ergodic sums using various initial values. Figure 11 shows the result where the initial pp is taken from 0.19,0.20,0.21,,17.480.19,0.20,0.21,\cdots,17.48 (excluding p=1p=1 (the unique fixed point of ff). By Figure 11 (where each ergodic sum is estimated using the first 1000010000 terms of fn(p)f^{n}(p)), we conclude that Theorems 5.10 and 5.11 hold and the ergodic sums of fn(p)f^{n}(p) converge to somewhere around 4.54.5 for ζ\zeta (or μ\mu) almost all pEp\in E (or pE~p\in\tilde{E}).

Refer to caption
Figure 11: Ergodic sums of fn(p)f^{n}(p) using various initial pp

6 Ergodic properties: a sensitivity analysis

6.1 Chaos is not too bad

In this final section, we still keep (α,β)=(0.75,0.5)(\alpha,\beta)=(0.75,0.5), (x¯,y¯)=(4,2)(\bar{x},\bar{y})=(4,2), but we let λ\lambda vary. Recall that, from Subsection 2.3, we know that there exists an odd period (hence a Li-Yorke chaos) for 2.77<λ<42.77<\lambda<4. We conduct a sensitivity analysis, that is, how the ergodic sums vary when λ\lambda changes with 1<λ<41<\lambda<4. We know that the unique fixed point of ff is p=y¯βx¯(1α)=1p=\frac{\bar{y}\beta}{\bar{x}(1-\alpha)}=1 (independent of λ\lambda), and the (unique) critical point of ff is s=y¯λβ=λs=\sqrt{\bar{y}\lambda\beta}=\sqrt{\lambda} (dependent of λ\lambda).

Now, using the same strategy as in the last section, we investigate how the critical orbit {s,f(s),f2(s),}\{s,f(s),f^{2}(s),\cdots\} behaves for each λ\lambda. Figure 12 is the bifurcation diagram for ff. We (roughly) see that: (1) for λ<2\lambda<2, fn(s)f^{n}(s) converges to the unique attracting fixed point (that is p=1p=1), (2) for 2<λ<2.52<\lambda<2.5, fn(s)f^{n}(s) converges to a period-22 orbit, (3) for 2.5<λ<2.72.5<\lambda<2.7 (roughly), period doubling bifurcations occur and fn(s)f^{n}(s) converges to a period-44 (8, 16, and so on) orbit, (4) for 2.7<λ<32.7<\lambda<3 (possibly starting at around 2.772.77 based on results in previous sections), we see a chaos except a few ”windows”, (5) at λ=3\lambda=3 (or 3.013.01), we see a period three orbit, (6) for 3<λ<3.13<\lambda<3.1 (roughly), again, we see period doubling bifurcations and fn(s)f^{n}(s) converges to an orbit of period 32,322,3233\cdot 2,3\cdot 2^{2},3\cdot 2^{3} and so on, (7) for 3.1<λ3.1<\lambda, we see a chaos again except a few windows.

Remark 6.1.

Actually, we have obtained the bifurcation diagram of ff for 1<λ<41<\lambda<4, but In Figure 12, we have chopped the diagram at λ=3.5\lambda=3.5. The reason is that for 3.5<λ<43.5<\lambda<4, the maximum value of fn(p)f^{n}(p) grows exponentially, so the diagram gets very skewed (and the part of the diagram shown in Figure 12 becomes almost invisible). For 3.5<λ<43.5<\lambda<4 we have obtained a chaotic region (with some windows).

Refer to caption
Figure 12: The bifurcation diagram of ff

Next, we compute Lyapunov exponents of ff at ss for various λ\lambda and obtain Figure 13. We see that roughly speaking the Lyapunov exponent is negative for 1<λ<2.71<\lambda<2.7 (thick dots), and positive for 2.7<λ<42.7<\lambda<4 (thin dots) except a few ”windows” corresponding to the windows in the bifurcation diagram (Figure 12).

Refer to caption
Figure 13: Lyapunov exponents of ff

Now, we test (SC) for 2.7<λ<42.7<\lambda<4 (we do not care for λ<2.7\lambda<2.7 since it is a non-chaotic region and easy to predict the future). We obtain Figure 14.

Refer to caption
Figure 14: Infinite sums in SC

Our argument is based on a numerical computation using 10001000 terms to estimate the infinite sum in (SC), so we need to decide when we conclude that the infinite sum is finite. We draw the (ad hoc) line at 1010 (that is the horizontal line in Figure 14). Basically, we only avoid λ\lambda where (the estimate of) the infinite sum grows exponentially. We hope to rigorously prove that the infinite sums here are finite in the future work.

Here are main results in this section.

Theorem 6.2.

For 2.75<λ<42.75<\lambda<4 except λ\lambda values corresponding to the few windows in Figure 12 (and possibly except some λ\lambda values whose total Lebesgue measure is 0, see Proposition 6.3 below), there exists a unique acim for ff. Moreover for these λ\lambda values, the ergodic sums of ff are as in Figure 15.

For λ\lambda values as in Theorem 6.2 (satisfying the SC), we obtain a pretty smooth relation between λ\lambda and the ergodic sums of ff as in Figure 15 (using 50005000 terms to estimate the ergodic sums). Extending Theorem 6.2 (and Figure 15) using the naive estimates of the ergodic sum (that is k=09999fk(s)\sum_{k=0}^{9999}f^{k}(s)) for λ\lambda that does not satisfy (SC), we obtain Theorem 1.4 (and Figure 16).

Refer to caption
Figure 15: Ergodic sums of ff

6.2 Comments and interpretations on Theorem 1.4

Here, we add a few comments and interpretations on Theorem 1.4: (1) We found it surprising that the overall behaviour of the ergodic sums of ff is quite smooth and stable considering the fact that as λ\lambda increases fn(x)f^{n}(x) go through a quite a bit of qualitative changes from a stable fixed point, attracting periodic orbit of different periods, and finally chaotic behaviours. (2) The gradual increase of the ergodic sum of ff as λ\lambda increases was unexpected (for us). We knew that as λ\lambda increases, ff takes more extreme values (very high/very low), but knew nothing about their distributions. To give some (mathematical) explanation for the behaviour (2), we obtained the estimates of distributions of fn(p)f^{n}(p) for various λ\lambda using 1000010000 terms in Figures 1617, and 18 (see Figure 9 also).

Our computation shows that: (1) The ”shapes” of the distributions of fn(s)f^{n}(s) for various λ\lambda look similar: the density is high for a low range and low for a high range. (2) ff takes more extreme values as λ\lambda becomes large, however, the extension of the upper bound is much greater than that of the lower bound (meaning that for E=[f(s),f2(s)+s]E=[f(s),f^{2}(s)+s], as λ\lambda becomes large, f(s)f(s) gets smaller a little bit, but f2(s)+sf^{2}(s)+s gets larger quite a bit), as shown in Figures 16,17,18, and 9, (3) The distribution gets smoother as λ\lambda becomes large.

We conclude that the gradual increase of the ergodic sum of ff happens as λ\lambda increases because: (1) ff takes more extreme values with a great extension of the upper bound (and with a little extension of the lower bound) as λ\lambda increases, (2) The shapes of the distributions do not change much as λ\lambda increases (although it gets smoother), (3) Thus, the average of fn(s)f^{n}(s) goes up as λ\lambda increases. We do not know why the overall behaviour of ergodic sums of ff is so smooth and stable. Also, we do not know why the density curve gets smoother as λ\lambda increases. We leave these issues for a future work.

Refer to caption
Figure 16: Density of fn(s)f^{n}(s) with λ=2.8\lambda=2.8
Refer to caption
Figure 17: Density of fn(s)f^{n}(s) with λ=3.2\lambda=3.2
Refer to caption
Figure 18: Density of fn(s)f^{n}(s) with λ=3.99\lambda=3.99

Theorem 1.4 (and the whole results in this paper) says that a naive estimate of the ergodic sums of ff (estimate of the future) using a reasonably large number of terms (say 5000100005000~{}10000 terms) is not too bad. To end the paper, we quote a deep result of Avila (2014 fields medalist) and others [Avila et al., 2003, Sec. 3.1, Theorem B] that supports Theorem 1.4.

Proposition 6.3.

In any non-trivial real analytic family of quasiquadratic maps (that contains SS-unimodal maps), (Lebesgue) almost any map is either regular (i.e., it has an attracting cycle) or stochastic (i.e., it has an acim).

Remark 6.4.

A technical note: ”non-triviality” is guaranteed for our set of maps ff (parametrised by λ\lambda) since there exist two maps in this set that are not topologically conjugate. For example, take ff with λ=1.5\lambda=1.5 (non-chaotic) and ff with λ=3.61\lambda=3.61 (chaotic). See [Avila et al., 2003, Sec. 2.8, Sec. 2.13, Sec. 3.1] for the precise definitions of ”non-trivial” and ”quasiquadratic” (those are a bit too technical to state here). Also, see [Avila and Moreira, 2005], and [Lyubich, 2012] for more on this.

Remark 6.5.

We need ”Lebesgue almost” (or ”except a set of measure zero”) in Theorems 1.46.2, and Proposition 6.3 since the following (anomalous) examples are known, see [Hofbauer and Keller, 1990] and [Johnson, 1987]: for a quadratic map Tλ(x)=λx(1x)T_{\lambda}(x)=\lambda x(1-x) (parametrised by λ\lambda), there exists λ\lambda such that TλT_{\lambda} does not have an attracting periodic orbit and shows a chaotic behaviour, but does not have an acim. We expect that for our ff, we obtain examples of the same properties (although we have not checked yet). The point is that we do not care such anomalous cases since the λ\lambda values corresponding to such examples are of Lebesgue measure zero and our approach in this (and the last) section is probabilistic.

Acknowledgements

This research was supported by a JSPS grant-in-aid for early-career scientists (22K13904) and an Alexander von Humboldt Japan-Germany joint research fellowship.

References

  • [Avila et al., 2003] Avila, A., Lyubich, M., and de Melo, W. (2003). Regular or stochastic dynamics in real analytic families of unimodal maps. Invent. Math., 154:451–550.
  • [Avila and Moreira, 2005] Avila, A. and Moreira, C. (2005). Statistical properties of unimodal maps: the quadratic family. Ann. Math., 161:831–881.
  • [Benhabib and Day, 1980] Benhabib, J. and Day, R. (1980). Erratic accumulation. Econ. Lett., 6:113–117.
  • [Benhabib and Day, 1982] Benhabib, J. and Day, R. (1982). A characterization of erratic dynamics in the overlapping generations model. J. Econ. Dyn. Control, 4:37–55.
  • [Bhattacharya and Majumdar, 2007] Bhattacharya, R. and Majumdar, M. (2007). Random Dynamical Systems. Cambridge University Press.
  • [Block and Coppel, 1992] Block, L. and Coppel, W. (1992). Dynamics in One Dimension. Springer, Berlin.
  • [Bosma et al., 1997] Bosma, W., Cannon, J., and Playoust, C. (1997). The Magma algebra system \@slowromancapi@: The user language. J. Symb. Comp, 24:235–265.
  • [Collet and Eckmann, 1980] Collet, P. and Eckmann, J.-P. (1980). Iterated Maps on the Interval as Dynamical Systems. Birkhäuser, Progress in Mathematics.
  • [Collet and Eckmann, 1983] Collet, P. and Eckmann, J.-P. (1983). Positive liapunov exponents and absolute continuity for maps of the interval. Ergodic Theory Dyn. Syst., 3:13–46.
  • [Day, 1998] Day, R. (1998). Complex Economic Dynamics, Volume 1: An Introduction to Dynamical Systems and Market Mechanisms. The MIT Press.
  • [Day and Shafer, 1985] Day, R. and Shafer, W. (1985). Keynesian chaos. J. Macroecon., 7:277–295.
  • [Deng et al., 2022] Deng, L., Khan, M., and Mitra, T. (2022). Continuous unimodal maps in economic dynamics: On easily verifiable conditions for topological chaos. J. Econ. Theory, 201. Article 105446.
  • [Hofbauer and Keller, 1990] Hofbauer, F. and Keller, G. (1990). Quadratic maps without asymptotic measure. Comm. Math. Phis, 127:319–337.
  • [Johnson, 1987] Johnson, S. (1987). Singular measures without restrictive intervals. Comm. Math. Phys, 110:185–190.
  • [Lasota and Yorke, 1973] Lasota, A. and Yorke, J. (1973). On the existence of invariant measures for piecewise monotonic transformations. Trans. Amer. Math. Soc., 186:481–488.
  • [Li and Yorke, 1975] Li, T. and Yorke, J. (1975). Period three implies chaos. Amer. Math. Monthly, 82:985–992.
  • [Lyubich, 2012] Lyubich, M. (2012). Forty years of unimodal dynamics: on the occasion of artur avila winning the brin prize. J. Mod. Dyn., 6:183–203.
  • [Misiurewicz, 1981] Misiurewicz, M. (1981). Absolutely continuous measures for certain maps of an interval. IHÉS Publ. Math., 53:17–51.
  • [Nishimura and Yano, 1996] Nishimura, K. and Yano, M. (1996). Chaotic solutions in dynamic linear programming. Chaos Solitons Fractals, 7:1941–1953.
  • [Nowicki and van Strien, 1991] Nowicki, T. and van Strien, S. (1991). Invariant measures exist under a summability condition for unimodal maps. Invent. Math., 105:123–136.
  • [Rossum and Drake, 2009] Rossum, G. and Drake, F. (2009). Python 3 Reference Manual. CreateSpace, Scotts Valley, CA.
  • [Ruette, 2017] Ruette, S. (2017). Chaos on the Interval. American Mathematical Society, Providence.
  • [Sato and Yano, 2012] Sato, K. and Yano, M. (2012). A simple condition for uniqueness of the absolutely continuous ergodic measure and its application to economic models. AIP Conf. Proc., 1479:737–740.
  • [Sato and Yano, 2013] Sato, K. and Yano, M. (2013). Ergodic chaos for non-expansive economic models. Int. J. Econ. Theory, 9:57–67.
  • [Sharkovsky, 1964] Sharkovsky, A. (1964). Co-existence of cycles of a continuous mapping of the line into itself. Ukr. Math. J., 16:61–71.
  • [Shen and van Strien, 2014] Shen, W. and van Strien, S. (2014). Recent developments in interval dynamics. Proc. Int. Cong. Math., 3:699–719.
  • [Singer, 1978] Singer, D. (1978). Stable orbits and bifurcations of maps of the interval. SIAM J. Appl. Math, 35:260–267.
  • [Ulam and von Neumann, 1947] Ulam, S. and von Neumann, J. (1947). On combinations of stochastic and deterministic processes. Bull. Amer. Math. Soc., 53:1120.
  • [W. de Melo, 1993] W. de Melo, S. v. S. (1993). One-Dimensional Dynamics. Springer.