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

The bifurcation structure within robust chaos for two-dimensional piecewise-linear maps.

I. Ghosh School of Mathematical and Computational Sciences
Massey University
Colombo Road, Palmerston North, 4410
New Zealand
R.I. McLachlan School of Mathematical and Computational Sciences
Massey University
Colombo Road, Palmerston North, 4410
New Zealand
D.J.W. Simpson School of Mathematical and Computational Sciences
Massey University
Colombo Road, Palmerston North, 4410
New Zealand
Abstract

We study two-dimensional, two-piece, piecewise-linear maps having two saddle fixed points. Such maps reduce to a four-parameter family and are well known to have a chaotic attractor throughout open regions of parameter space. The purpose of this paper is to determine where and how this attractor undergoes bifurcations. We explore the bifurcation structure numerically by using Eckstein’s greatest common divisor algorithm to estimate from sample orbits the number of connected components in the attractor. Where the map is orientation-preserving the numerical results agree with formal results obtained previously through renormalisation. Where the map is orientation-reversing or non-invertible the same renormalisation scheme appears to generate the bifurcation boundaries, but here we need to account for the possibility of some stable low-period solutions. Also the attractor can be destroyed in novel heteroclinic bifurcations (boundary crises) that do not correspond to simple algebraic constraints on the parameters. Overall the results reveal a broadly similar component-doubling bifurcation structure in the orientation-reversing and non-invertible settings, but with some additional complexities.

1 Introduction

Complex dynamics are caused by nonlinearity, and many physical systems involve an extreme form of nonlinearity: a switch. Examples include engineering systems with relay or on/off control which involve distinct modes of operation [5, 23, 43]. Mathematical models of systems with switches are naturally piecewise-smooth, where different pieces of the equations correspond to different positions of a switch.

Piecewise-smooth maps arise as discrete-time models of systems with switches [16, 33] and as Poincaré maps or stroboscopic maps of continuous-time models [7]. For multi-dimensional piecewise-smooth maps there are many abstract ergodic theory results stating that if an attractor satisfies certain properties (e.g. expansion) then it is chaotic in a certain sense (e.g. has an SRB measure) [6, 30, 35, 42, 45]. This paper in contrast provides explicit results for a physically-motivated family of maps. We study the four-parameter family

fξ(x,y)={[τLx+y+1δLx],x0,[τRx+y+1δRx],x0,\displaystyle f_{\xi}(x,y)=\begin{cases}\begin{bmatrix}\tau_{L}x+y+1\\ -\delta_{L}x\end{bmatrix},&x\leq 0,\\ \begin{bmatrix}\tau_{R}x+y+1\\ -\delta_{R}x\end{bmatrix},&x\geq 0,\end{cases} (1.1)

where ξ=(τL,δL,τR,δR)4\xi=(\tau_{L},\delta_{L},\tau_{R},\delta_{R})\in\mathbb{R}^{4}. This is known as the two-dimensional border-collision normal form (BCNF) [26]. This family provides leading-order approximations to border-collision bifurcations where a fixed point of a piecewise-smooth map collides with a switching manifold as parameters are varied [36]. As such (1.1) can be used to capture the change in dynamics as parameters are varied to pass through border-collision bifurcations. In this regards it has been applied to mechanical oscillators with stick-slip friction [41], DC/DC power converters [47], and various models in economics [33]. We stress that unlike Poincaré maps of smooth invertible flows, the piecewise-smooth maps arising from these applications are mostly non-invertible. We distinguish four generic cases based on the signs of the determinants δL\delta_{L} and δR\delta_{R}:

  • With δL>0\delta_{L}>0 and δR>0\delta_{R}>0 (1.1) is orientation-preserving. This is the classical scenario considered in the ‘robust chaos’ paper of Banerjee et al [4].

  • With δL<0\delta_{L}<0 and δR<0\delta_{R}<0 (1.1) is orientation-reversing. Misiurewicz [25] used this scenario in the subfamily of Lozi maps (τL=τR\tau_{L}=-\tau_{R}, δL=δR\delta_{L}=\delta_{R}) for his novel rigorous demonstration of robust chaos.

  • With δL>0\delta_{L}>0 and δR<0\delta_{R}<0 (1.1) is non-invertible mapping both the left half-plane (x0x\leq 0) and the right half-plane (x0x\geq 0) onto the upper half-plane (y0y\geq 0). This scenario applies to border-collision bifurcations in a wide range of power converters whereby a switch in a circuit creates non-invertible dynamics [3, 8, 46].

  • With δL<0\delta_{L}<0 and δR>0\delta_{R}>0 (1.1) is again non-invertible with applications to power electronics, but now maps to the lower half-plane (y0y\leq 0) and exhibits different dynamical features (notice (1.1) maps the origin to the right half-plane so does not have left/right symmetry).

Despite its apparent simplicity (1.1) has an incredibly complex bifurcation structure that remains to be fully understood [2, 10, 38, 40, 47]. The most significant bifurcations are those at which the attractor splits into pieces. In [12] we used renormalisation to uncover an infinite sequence of codimension-one bifurcations. However, this only accommodated the orientating-preserving setting of Banerjee et al [4]. In this paper we identify similar sequences of bifurcations in the orientation-reversing and non-invertible parameter regimes. Again we use renormalisation to find bifurcations, but now this approach misses some bifurcations. For this reason we combine the analytical framework with brute-force numerical simulations that compute the number of connected components from the behaviour of forward orbits.

The remainder of this paper is organised as follows. We start in §2 by reviewing the bifurcation structure of one-dimensional piecewise-linear maps (skew tent maps) and how this structure can be generated through renormalisation. This corresponds to the BCNF in the special case δL=δR=0\delta_{L}=\delta_{R}=0, and it is helpful to view the bifurcation structures described in later sections as extensions or perturbations from the structure that arises in one dimension. Then in §3 we review the parameter region Φ4\Phi\subset\mathbb{R}^{4} where the BCNF has two saddle fixed points. We describe special points on the stable and unstable manifolds of the fixed points whose relation to one another can be used to characterise the occurrence of homoclinic and heteroclinic bifurcations where the attractor is destroyed. In §4 we introduce the renormalisation scheme and describe some basic aspects of this scheme that hold for all values of δL\delta_{L} and δR\delta_{R}. Then in §5 we summarise the results of [12] for the orientation-preserving setting.

Next in §6 we describe an algorithm for numerically determining the number of connected components of attractor. The algorithm outputs the greatest common divisor of a set of iteration numbers required for an orbit to return close to its starting point. The algorithm is effective when the components of attractor are not too close to one another and the dynamics on the attractor is ergodic. Applied to the orientation-preserving setting it reproduces the bifurcation structure obtained by renormalisation.

In §7, §8, and §9 we study the orientation-reversing and non-invertible cases and explain why the same renormalisation scheme should be expected to work in these settings. Formal proofs are beyond the scope of this study because, as evident from [12], these require a detailed analysis of the four-dimensional nonlinear renormalisation operator, plus will involve additional complexities because now stable period-two and period-four solutions are possible. Also if δL<0\delta_{L}<0 and δR>0\delta_{R}>0 then the bifurcations that destroy the attractor are different and more difficult to characterise. In §10 we describe a special case unique to the non-invertible settings where the dynamics reduces to one dimension. Concluding remarks are provided in §11.

2 Skew tent maps

With δL=δR=0\delta_{L}=\delta_{R}=0 the yy-dynamics of (1.1) is trivial and the xx-dynamics reduces to

x{τLx+1,x0,τRx+1,x0.\displaystyle x\mapsto\begin{cases}\tau_{L}x+1,&x\leq 0,\\ \tau_{R}x+1,&x\geq 0.\end{cases} (2.1)

This is a two-parameter family of skew tent maps. For any values of the parameters τL\tau_{L} and τR\tau_{R} for which (2.1) has an attractor, this attractor is unique. Fig. 1 shows typical examples of the attractor using τL>1\tau_{L}>1 and τR<1\tau_{R}<-1 with which (2.1) is piecewise-expanding, so the attractor is chaotic in a strict sense by the results of Lasota and Yorke [21] and Li and Yorke [22]. Fig. 1 shows typical examples of the chaotic attractor. In panel (a) the attractor is the interval [τR+1,1][\tau_{R}+1,1] whose endpoints are the first and second iterates of the switching value x=0x=0. In panel (b) the attractor is a disjoint union of four intervals whose endpoints are the first through eighth iterates of x=0x=0.

Refer to caption Refer to caption
(a) (τL,τR)=(1.3,2)(\tau_{L},\tau_{R})=(1.3,-2). (b) (τL,τR)=(1.1,1.1)(\tau_{L},\tau_{R})=(1.1,-1.1).
Figure 1: Cobweb diagrams showing the attractor of the skew tent map (2.1) with two different combinations of the parameter values. In (a) the attractor is an interval; in (b) the attractor is the union of four disjoint intervals. The maps are also instances of the BCNF (1.1) with ξ=(1.3,0,2,0)\xi=(1.3,0,-2,0) in panel (a) and ξ=(1.1,0,1.1,0)\xi=(1.1,0,-1.1,0) in panel (b).

For all values of τL\tau_{L} and τR\tau_{R} the nature of the attractor is well understood [24, 27, 39]. Fig. 2 shows how the part of the (τL,τR)(\tau_{L},\tau_{R})-plane that is relevent to us divides into regions according to the number of intervals that comprise the attractor. In the top-left region labelled LRLR the map has a stable period-two solution. This solution has one point in the left half-plane and one point in the right half-plane, so is termed an LRLR-cycle. This region is bounded by the curve α0(τL,τR)=0\alpha_{0}(\tau_{L},\tau_{R})=0, where

α0(τL,τR)=τLτR+1,\displaystyle\alpha_{0}(\tau_{L},\tau_{R})=\tau_{L}\tau_{R}+1, (2.2)

which is where the LRLR-cycle loses stability by attaining a stability multiplier of 1-1. The boundary in the lower-right part of the figure is a homoclinic bifurcation where x=0x=0 maps in two iterations to the fixed point of (2.1) in x<0x<0. This occurs on the curve ϕ0(τL,τR)=0\phi_{0}(\tau_{L},\tau_{R})=0, where

ϕ0(τL,τR)=τLτR+τLτR.\displaystyle\phi_{0}(\tau_{L},\tau_{R})=\tau_{L}\tau_{R}+\tau_{L}-\tau_{R}. (2.3)
Refer to caption
Figure 2: A two-parameter bifurcation diagram of the skew tent map family (2.1). In 0\mathcal{R}_{0} the attractor is an interval. In each n\mathcal{R}_{n} with n1n\geq 1 the attractor is comprised of 2n2^{n} disjoint intervals. Below ϕ0(τL,τR)=0\phi_{0}(\tau_{L},\tau_{R})=0 the map has no attractor; above τR=1\tau_{R}=-1 it has a stable fixed point in x>0x>0; in the top-left region it has a stable LRLR-cycle (period-two solution). The triangles indicate the parameter values used in Fig. 1.

As shown by Ito et al [17, 18] the remaining boundaries can be identified through renormalisation as follows. Suppose there exists an interval neighbourhood II of x=0x=0 that maps into x>0x>0, then, on the second iteration, back into II. In this case the restriction of the second iterate of (2.1) to II is piecewise-linear with two pieces. The x<0x<0 piece corresponds to iterates with symbolic itinerary LRLR and has slope τLτR<1\tau_{L}\tau_{R}<-1, while the x>0x>0 piece corresponds to iterates with symbolic itinerary RRRR and has slope τR2>1\tau_{R}^{2}>1. Thus the second iterate of (2.1) on II is conjugate to (2.1) with τR2\tau_{R}^{2} in place of τL\tau_{L} and τLτR\tau_{L}\tau_{R} in place of τR\tau_{R}. This corresponds to the substitution rule (L,R)(RR,LR)(L,R)\mapsto(RR,LR), and induces the renormalisation operator

g(τL,τR)=(τR2,τLτR).\displaystyle g(\tau_{L},\tau_{R})=(\tau_{R}^{2},\tau_{L}\tau_{R}). (2.4)

By using the preimages of the homoclinic bifurcation boundary we define the sequence of regions

n={(τL,τR)|τR<1,ϕ0(gn(ξ))>0,ϕ0(gn+1(ξ))0,α0(τL,τR)<0},\displaystyle\mathcal{R}_{n}=\mathopen{}\mathclose{{}\left\{(\tau_{L},\tau_{R})\,\middle|\,\tau_{R}<-1,\phi_{0}\mathopen{}\mathclose{{}\left(g^{n}(\xi)}\right)>0,\phi_{0}\mathopen{}\mathclose{{}\left(g^{n+1}(\xi)}\right)\leq 0,\alpha_{0}(\tau_{L},\tau_{R})<0}\right\}, (2.5)

shown in Fig. 2. These are non-empty for all n0n\geq 0 and converge to (τL,τR)=(1,1)(\tau_{L},\tau_{R})=(1,-1) as nn\to\infty. In 0\mathcal{R}_{0} the attractor consists of one interval:

Theorem 2.1.

For any (τL,τR)0(\tau_{L},\tau_{R})\in\mathcal{R}_{0} with τL1\tau_{L}\geq 1, the interval [τR+1,1][\tau_{R}+1,1] is the unique attractor.

A simple proof of Theorem 2.1 can be found in Veitch and Glendinning [44]. The condition τL1\tau_{L}\geq 1 is needed because for some 0<τL<10<\tau_{L}<1 and τR<1\tau_{R}<-1 the attractor is periodic. This condition can be weakened but this is not needed for our purposes.

To characterise the attractor in the other regions n\mathcal{R}_{n} we first make the following observation that follows simply from the definitions of gg and n\mathcal{R}_{n}.

Proposition 2.2.

If (τL,τR)n(\tau_{L},\tau_{R})\in\mathcal{R}_{n} with n1n\geq 1, then g(τL,τR)n1g(\tau_{L},\tau_{R})\in\mathcal{R}_{n-1}.

Thus if (τL,τR)n(\tau_{L},\tau_{R})\in\mathcal{R}_{n} then gn(τL,τR)0g^{n}(\tau_{L},\tau_{R})\in\mathcal{R}_{0}. Theorem 2.1 can be applied at this parameter point because any point below τR=1\tau_{R}=-1 maps under gg to the right of τL=1\tau_{L}=1. Moreover, the above conjugacy can be shown to hold under all nn applications of gg, see Ito et al [17, 18], thus the 2n2^{n}-iterate of (2.1) has an interval attractor. The images of this interval under (2.1) give 2n2^{n} disjoint intervals and hence the following result.

Theorem 2.3.

If (τL,τR)n(\tau_{L},\tau_{R})\in\mathcal{R}_{n} with n1n\geq 1, then (2.1) has a unique attractor comprised of 2n2^{n} disjoint intervals.

3 Two-dimensional piecewise-linear maps with two saddle fixed points

We now return to the BCNF (1.1) with non-zero values of δL\delta_{L} and δR\delta_{R}. In this section we identify two saddle fixed points and compute some important points on their stable and unstable manifolds.

Motivated by Banerjee et al [4] we focus on the following subset of four-dimensional parameter space:

Φ={ξ4|τL>|δL+1|,τR<|δR+1|}.\displaystyle\Phi=\mathopen{}\mathclose{{}\left\{\xi\in\mathbb{R}^{4}\ \middle|\ \tau_{L}>|\delta_{L}+1|,\tau_{R}<-|\delta_{R}+1|}\right\}. (3.1)

It is a simple exercise to show that Φ\Phi is the set of all parameter values for which (1.1) has two saddle fixed points [11]. These fixed points are

X\displaystyle X =(1τRδR1,δRτRδR1),\displaystyle=\mathopen{}\mathclose{{}\left(\frac{-1}{\tau_{R}-\delta_{R}-1},\frac{\delta_{R}}{\tau_{R}-\delta_{R}-1}}\right), (3.2)
Y\displaystyle Y =(1τLδL1,δLτLδL1),\displaystyle=\mathopen{}\mathclose{{}\left(\frac{-1}{\tau_{L}-\delta_{L}-1},\frac{\delta_{L}}{\tau_{L}-\delta_{L}-1}}\right), (3.3)

where XX is in the open right half-plane x>0x>0, and YY is in the open left half-plane x<0x<0. Let

AL\displaystyle A_{L} =[τL1δL0],\displaystyle=\begin{bmatrix}\tau_{L}&1\\ -\delta_{L}&0\end{bmatrix}, AR\displaystyle A_{R} =[τR1δR0],\displaystyle=\begin{bmatrix}\tau_{R}&1\\ -\delta_{R}&0\end{bmatrix}, (3.4)

denote the Jacobian matrices associated with YY and XX respectively. With ξΦ\xi\in\Phi the matrix ALA_{L} has eigenvalues |λLs|<1|\lambda_{L}^{s}|<1 and λLu>1\lambda_{L}^{u}>1, while ARA_{R} has eigenvalues |λRs|<1|\lambda_{R}^{s}|<1 and λRu<1\lambda_{R}^{u}<-1.

Since XX and YY are saddles their stable and unstable manifolds are one-dimensional, and since the map is piecewise-linear these manifolds are piecewise-linear. Their stable manifolds Ws(X)W^{s}(X) and Ws(Y)W^{s}(Y) have kinks on the switching manifold x=0x=0 and at preimages of these points, while their unstable manifolds Wu(X)W^{u}(X) and Wu(Y)W^{u}(Y) have kinks on the image of the switching manifold y=0y=0 and at images of these points. For instance, as we grow Wu(X)W^{u}(X) outwards from XX, in one direction its first kink occurs at the point

T=(11λRs,0),\displaystyle T=\mathopen{}\mathclose{{}\left(\frac{1}{1-\lambda_{R}^{s}},0}\right), (3.5)

on y=0y=0, see Fig. 3. Similarly as we grow Wu(Y)W^{u}(Y) outwards from YY, in one direction its first kink occurs at

D=(11λLs,0).\displaystyle D=\mathopen{}\mathclose{{}\left(\frac{1}{1-\lambda_{L}^{s}},0}\right). (3.6)

A third point CC will be also central to our later calculations. This point is where Ws(Y)W^{s}(Y), when grown outwards from YY to the right, first intersects y=0y=0 at a point with x>0x>0 and is given by

C=(λLu2(δRτRλLu)(λLu1),0).\displaystyle C=\mathopen{}\mathclose{{}\left(\frac{{\lambda_{L}^{u}}^{2}}{(\delta_{R}-\tau_{R}\lambda_{L}^{u})(\lambda_{L}^{u}-1)},0}\right). (3.7)

Fig. 3 indicates the points TT, DD, and CC for four example phase portraits, one for each of the four cases for the signs of δL\delta_{L} and δR\delta_{R}. Notice if δL>0\delta_{L}>0 (upper plots) the left half-plane maps to the upper half-plane, while if δL<0\delta_{L}<0 (lower plots) the left half-plane maps to the lower half-plane. Similarly if δR>0\delta_{R}>0 (left plots) the right half-plane maps to the lower half-plane, while if δR<0\delta_{R}<0 (right plots) the right half-plane maps to the upper half-plane.

Refer to caption Refer to caption
(a) δL>0,δR>0\delta_{L}>0,\delta_{R}>0 (b) δL>0,δR<0\delta_{L}>0,\delta_{R}<0
Refer to caption Refer to caption
(c) δL<0,δR>0\delta_{L}<0,\delta_{R}>0 (d) δL<0,δR<0\delta_{L}<0,\delta_{R}<0
Figure 3: Phase portraits of the BCNF (1.1) for four different parameter combinations: (a) ξ=(2.1,0.06,1.7,0.18)\xi=(2.1,0.06,-1.7,0.18); (b) ξ=(2.1,0.4,1.7,0.55)\xi=(2.1,0.4,-1.7,-0.55); (c) ξ=(2.2,0.3,1.7,0.1)\xi=(2.2,-0.3,-1.7,0.1); (d) ξ=(1.8,0.75,1.6,0.4)\xi=(1.8,-0.75,-1.6,-0.4). Each plot shows the linear segments of the stable and unstable manifolds emanating from the fixed points XX and YY, as well as an adjoining segment of the stable manifold of YY. We also indicate some of their intersections with y=0y=0: TT, DD, and CC; formulas for these are given by (3.5), (3.6), and (3.7). To illustrate the chaotic attractor the black dots show 2000 iterates of a typical forward orbit after transient dynamics has decayed.

4 Renormalisation in two dimensions

In §2 we saw that renormalisation based on the substitution rule (L,R)(RR,LR)(L,R)\mapsto(RR,LR) explains the bifurcation structure of the one-dimensional skew tent map family shown in Fig. 2. For the BCNF (1.1) this rule defines the renormalisation operator

g(ξ)=(τR22δR,δR2,τLτRδLδR,δLδR).\displaystyle g(\xi)=\mathopen{}\mathclose{{}\left(\tau_{R}^{2}-2\delta_{R},\delta_{R}^{2},\tau_{L}\tau_{R}-\delta_{L}-\delta_{R},\delta_{L}\delta_{R}}\right). (4.1)

The four values on the right-hand side of (4.1) are the traces and determinants of AR2A_{R}^{2} and ARALA_{R}A_{L}. In this section we extend the basic aspects of this renormalisation operator beyond the orientation-preserving setting of our earlier work [12]. Renormalisation has also been applied to subfamilies of (1.1) by Pumariño et al [31, 32] and Ou [29].

We first identify the subset of Φ\Phi where the BCNF has a stable LRLR-cycle. Generalising (2.2) we have that the matrix ALARA_{L}A_{R} (whose eigenvalues are the stability multipliers of the LRLR-cycle) has an eigenvalue 1-1 when α(ξ)=0\alpha(\xi)=0, where

α(ξ)=τLτR+(δL1)(δR1).\displaystyle\alpha(\xi)=\tau_{L}\tau_{R}+(\delta_{L}-1)(\delta_{R}-1). (4.2)

If α(ξ)>0\alpha(\xi)>0 this eigenvalue is greater than 1-1. For the LRLR-cycle to be stable we also need det(ALAR)<1\det(A_{L}A_{R})<1, where det(ALAR)=δLδR\det(A_{L}A_{R})=\delta_{L}\delta_{R}, so we define

𝒫2={ξΦ|δLδR<1,α(ξ)>0}.\mathcal{P}_{2}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi\,\middle|\,\delta_{L}\delta_{R}<1,\,\alpha(\xi)>0}\right\}. (4.3)
Proposition 4.1.

If ξ𝒫2\xi\in\mathcal{P}_{2} then fξf_{\xi} has an asymptotically stable LRLR-cycle.

Proof.

Let τ~R=τLτRδLδR\tilde{\tau}_{R}=\tau_{L}\tau_{R}-\delta_{L}-\delta_{R} and δ~R=δLδR\tilde{\delta}_{R}=\delta_{L}\delta_{R} be the trace and determinant of ALARA_{L}A_{R}. All eigenvalues of ALARA_{L}A_{R} have modulus less than 11 if and only if δ~R<1\tilde{\delta}_{R}<1 and δ~R1<τ~R<δ~R+1-\tilde{\delta}_{R}-1<\tilde{\tau}_{R}<\tilde{\delta}_{R}+1. Certainly δ~R<1\tilde{\delta}_{R}<1 by assumption, and δ~R1<τ~R-\tilde{\delta}_{R}-1<\tilde{\tau}_{R} by the definition of α\alpha; also

τ~Rδ~R1=τLτR(δL+1)(δR+1)<|(δL+1)(δR+1)|(δL+1)(δR+1)0\tilde{\tau}_{R}-\tilde{\delta}_{R}-1=\tau_{L}\tau_{R}-(\delta_{L}+1)(\delta_{R}+1)<-|(\delta_{L}+1)(\delta_{R}+1)|-(\delta_{L}+1)(\delta_{R}+1)\leq 0

by the definition of Φ\Phi.

The composed map fRfLf_{R}\circ f_{L} is affine with unique fixed point

P=1τ~Rδ~R1(τR+δR+1,δR(τL+δL+1)).P=\frac{-1}{\tilde{\tau}_{R}-\tilde{\delta}_{R}-1}\big{(}\tau_{R}+\delta_{R}+1,-\delta_{R}(\tau_{L}+\delta_{L}+1)\big{)}.

Let

Q=fL(P)=1τ~Rδ~R1(τL+δL+1,δL(τR+δR+1)).Q=f_{L}(P)=\frac{-1}{\tilde{\tau}_{R}-\tilde{\delta}_{R}-1}\big{(}\tau_{L}+\delta_{L}+1,-\delta_{L}(\tau_{R}+\delta_{R}+1)\big{)}.

Notice the first component of PP is negative while the first component of QQ is positive, thus {P,Q}\{P,Q\} is a periodic solution of fξf_{\xi}. Above we showed all eigenvalues of ALARA_{L}A_{R} have modulus less than 11, hence {P,Q}\{P,Q\} is asymptotically stable. ∎

If α(ξ)<0\alpha(\xi)<0 the LRLR-cycle is unstable. This is always the case if δL+δR0\delta_{L}+\delta_{R}\geq 0:

Proposition 4.2.

If ξΦ\xi\in\Phi with δL+δR0\delta_{L}+\delta_{R}\geq 0 then α(ξ)<0\alpha(\xi)<0.

Proof.

Suppose ξΦ\xi\in\Phi with δL+δR0\delta_{L}+\delta_{R}\geq 0. If δL,δR1\delta_{L},\delta_{R}\geq-1 then

α(ξ)<(δL+1)(δR+1)+(δL1)(δR1)=2(δL+δR)0,\alpha(\xi)<-(\delta_{L}+1)(\delta_{R}+1)+(\delta_{L}-1)(\delta_{R}-1)=-2(\delta_{L}+\delta_{R})\leq 0,

using the definition of Φ\Phi in the first inequality. If instead δL<1\delta_{L}<-1 or δR<1\delta_{R}<-1 then δLδR<1\delta_{L}\delta_{R}<-1 because δL+δR>0\delta_{L}+\delta_{R}>0, so

α(ξ)<(δL+1)(δR+1)+(δL1)(δR1)=2(δLδR+1)<0.\alpha(\xi)<(\delta_{L}+1)(\delta_{R}+1)+(\delta_{L}-1)(\delta_{R}-1)=2(\delta_{L}\delta_{R}+1)<0.

Now we show that if α(ξ)<0\alpha(\xi)<0 then the renormalisation operator gg produces another instance of the BCNF in Φ\Phi.

Proposition 4.3.

If ξΦ\xi\in\Phi and α(ξ)<0\alpha(\xi)<0 then g(ξ)Φg(\xi)\in\Phi.

Proof.

Let τ~L=τR22δR\tilde{\tau}_{L}=\tau_{R}^{2}-2\delta_{R}, let δ~L=δR2\tilde{\delta}_{L}=\delta_{R}^{2}, and let τ~R\tilde{\tau}_{R} and δ~R\tilde{\delta}_{R} be as in the proof of Proposition 4.1; then g(ξ)=(τ~L,δ~L,τ~R,δ~R)g(\xi)=\big{(}\tilde{\tau}_{L},\tilde{\delta}_{L},\tilde{\tau}_{R},\tilde{\delta}_{R}\big{)}. Observe τ~Lδ~L1=τR2(δR+1)2>0\tilde{\tau}_{L}-\tilde{\delta}_{L}-1=\tau_{R}^{2}-(\delta_{R}+1)^{2}>0 because τR<(δR+1)\tau_{R}<-(\delta_{R}+1), and τ~L+δ~L+1=τR2+(δR+1)2>0\tilde{\tau}_{L}+\tilde{\delta}_{L}+1=\tau_{R}^{2}+(\delta_{R}+1)^{2}>0 because τR<0\tau_{R}<0, thus τ~L>|δ~L+1|\tilde{\tau}_{L}>|\tilde{\delta}_{L}+1|. Also τ~Rδ~R1=τLτR(δL+1)(δR+1)<0\tilde{\tau}_{R}-\tilde{\delta}_{R}-1=\tau_{L}\tau_{R}-(\delta_{L}+1)(\delta_{R}+1)<0 because τL>δL+1\tau_{L}>\delta_{L}+1 and τR<δR+1\tau_{R}<\delta_{R}+1, and τ~R+δ~R+1=α(ξ)<0\tilde{\tau}_{R}+\tilde{\delta}_{R}+1=\alpha(\xi)<0, by assumption, thus τ~R<|δ~R+1|\tilde{\tau}_{R}<-|\tilde{\delta}_{R}+1|. ∎

Now let

Πξ={fξ1(x,y)|x0},\Pi_{\xi}=\mathopen{}\mathclose{{}\left\{f_{\xi}^{-1}(x,y)\,\middle|\,x\geq 0}\right\}, (4.4)

be the set of all points that map to the right half-plane. On Πξ\Pi_{\xi} the second iterate fξ2f_{\xi}^{2} has only two pieces:

fξ2(x,y)={(fR,ξfL,ξ)(x,y),x0,fR,ξ2(x,y),x0.f_{\xi}^{2}(x,y)=\begin{cases}\mathopen{}\mathclose{{}\left(f_{R,\xi}\circ f_{L,\xi}}\right)(x,y),&x\leq 0,\\ f_{R,\xi}^{2}(x,y),&x\geq 0.\end{cases} (4.5)

For any ξΦ\xi\in\Phi (4.5) is affinely conjugate to fg(ξ)f_{g(\xi)}. Specifically fξ2=hξ1fg(ξ)hξf_{\xi}^{2}=h_{\xi}^{-1}\circ f_{g(\xi)}\circ h_{\xi} on Πξ\Pi_{\xi}, where

hξ(x,y)=1τR+δR+1[xδRx+τRyδR],h_{\xi}(x,y)=\frac{1}{\tau_{R}+\delta_{R}+1}\begin{bmatrix}x\\ \delta_{R}x+\tau_{R}y-\delta_{R}\end{bmatrix}, (4.6)

is the necessary change of coordinates. As an example Fig. 4-a shows Πξ\Pi_{\xi} (in the phase space of fξf_{\xi}) at the parameter point

ξex(1)=(1.5,0.4,1.5,0.4),\xi^{(1)}_{\rm ex}=(1.5,0.4,-1.5,0.4), (4.7)

while Fig. 4-b shows hξ(Πξ)h_{\xi}(\Pi_{\xi}) (in the phase space of fg(ξ)f_{g(\xi)}). The map fg(ξ)f_{g(\xi)} has an invariant set Λhξ(Πξ)\Lambda\subset h_{\xi}(\Pi_{\xi}), so, as shown below, hξ1(Λ)Πξh_{\xi}^{-1}(\Lambda)\subset\Pi_{\xi} is an invariant set of fξ2f_{\xi}^{2}. Moreover, the union of hξ1(Λ)h_{\xi}^{-1}(\Lambda) and its image under fξf_{\xi} is an invariant set of fξf_{\xi}. It is easy to infer the number of connected components in this set from the number of connected components of Λ\Lambda, and this forms the basis of the renormalisation scheme.

Refer to caption Refer to caption
(a) ξ=ξex(1)1(1)\xi=\xi^{(1)}_{\rm ex}\in\mathcal{R}^{(1)}_{1} (b) ξ=g(ξex(1))0(1)\xi=g\big{(}\xi^{(1)}_{\rm ex}\big{)}\in\mathcal{R}^{(1)}_{0}
Figure 4: Phase portraits of the BCNF (1.1). Panel (a) uses the example parameter point ξex(1)\xi^{(1)}_{\rm ex} given by (4.7) where the attracor Λ~\tilde{\Lambda} has two connected components, one of which lies entirely in Πξ\Pi_{\xi} (shaded). Panel (b) uses the parameter point g(ξex(1))g\big{(}\xi^{(1)}_{\rm ex}\big{)} where the attractor Λ\Lambda has one connected component in hξ(Πξ)h_{\xi}(\Pi_{\xi}) (shaded).
Proposition 4.4.

Suppose Λhξ(Πξ)\Lambda\subset h_{\xi}(\Pi_{\xi}) is an invariant set of fg(ξ)f_{g(\xi)}. Then Λ~=hξ1(Λ)fξ(hξ1(Λ))\tilde{\Lambda}=h_{\xi}^{-1}(\Lambda)\cup f_{\xi}\mathopen{}\mathclose{{}\left(h_{\xi}^{-1}(\Lambda)}\right) is an invariant set of fξf_{\xi}. Moreover, if hξ1(Λ)fξ(hξ1(Λ))=h_{\xi}^{-1}(\Lambda)\cap f_{\xi}\mathopen{}\mathclose{{}\left(h_{\xi}^{-1}(\Lambda)}\right)=\varnothing then the number of connected components of Λ~\tilde{\Lambda} is twice the number of connected components of Λ\Lambda.

Proof.

Let S=hξ1(Λ)S=h_{\xi}^{-1}(\Lambda). Since SΠξS\subset\Pi_{\xi},

fξ2(S)=hξ1(fg(ξ)(hξ(S))).f_{\xi}^{2}(S)=h_{\xi}^{-1}\mathopen{}\mathclose{{}\left(f_{g(\xi)}\mathopen{}\mathclose{{}\left(h_{\xi}(S)}\right)}\right). (4.8)

But hξ(S)=Λh_{\xi}(S)=\Lambda and fg(ξ)(Λ)=Λf_{g(\xi)}(\Lambda)=\Lambda, so the right-hand side of (4.8) is hξ1(Λ)=Sh_{\xi}^{-1}(\Lambda)=S. Thus SS is invariant under fξ2f_{\xi}^{2}, so Sfξ(S)S\cup f_{\xi}(S) is invariant under fξf_{\xi}. Since hξh_{\xi} is invertible the number of components of SS is the same as the number of components of Λ\Lambda. If Sfξ(S)=S\cap f_{\xi}(S)=\varnothing then fξ(S)f_{\xi}(S) also has this many components (for otherwise fξ2(S)=Sf_{\xi}^{2}(S)=S would not be possible because fξf_{\xi} is continuous). ∎

5 The orientation-preserving case

In this section we summarise the main results of our earlier work [12]. Let

Φ(1)={ξΦ|δL>0,δR>0},\displaystyle\Phi^{(1)}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi\ \middle|\ \delta_{L}>0,\delta_{R}>0}\right\}, (5.1)

be the subset of Φ\Phi for which the BCNF is orientation-preserving. In a large part of Φ(1)\Phi^{(1)} an attractor is destroyed when the points CC and DD (shown in Fig. 3-a) coincide. This is a homoclinic bifurcation, or homoclinic corner [37], where the stable and unstable manifolds of the fixed point YY attain non-trivial intersections. Straight-forward algebra gives

C1D1=ϕ+(ξ)(λLu1)(1λLs)(δRτRλLu),\displaystyle C_{1}-D_{1}=\frac{\phi^{+}(\xi)}{(\lambda_{L}^{u}-1)(1-\lambda_{L}^{s})(\delta_{R}-\tau_{R}\lambda_{L}^{u})}, (5.2)

where

ϕ+(ξ)=δR(τR+δL+δR(1+τR)λLu)λLu.\displaystyle\phi^{+}(\xi)=\delta_{R}-(\tau_{R}+\delta_{L}+\delta_{R}-(1+\tau_{R})\lambda_{L}^{u})\lambda_{L}^{u}. (5.3)

Fig. 5 shows an example. In panel (a) the closure of the unstable manifold of XX,

Λ=cl(Wu(X)),\displaystyle\Lambda={\rm cl}(W^{u}(X)), (5.4)

is a chaotic attractor. As the value of τL\tau_{L} is increased the attractor approaches the point DD and is destroyed at τL1.9083\tau_{L}\approx 1.9083 when ϕ+(ξ)=0\phi^{+}(\xi)=0, panel (b). At larger values of τL\tau_{L} almost all forward orbits diverge, panel (c).

Refer to caption
(a) τL=1.7\tau_{L}=1.7
Refer to caption
(b) τL1.9083\tau_{L}\approx 1.9083
Refer to caption
(c) τL=2.1\tau_{L}=2.1
Figure 5: Phase portraits of (1.1) with δL=0.1\delta_{L}=0.1, δR=0.1\delta_{R}=0.1, τR=2\tau_{R}=-2 and three different values of τL\tau_{L}. These parameter values correspond to the black triangles in Fig. 6-a. Panel (b) uses τL\tau_{L} such that ϕ+(ξ)=0\phi^{+}(\xi)=0 to ten decimal places. In (a) cl(Wu(X)){\rm cl}(W^{u}(X)) is a chaotic attractor; in (b) Ws(Y)W^{s}(Y) and Wu(Y)W^{u}(Y) form a homoclinic corner by intersecting at D=CD=C; in (c) there is no attractor.

In some parts of Φ(1)\Phi^{(1)} the attractor Λ\Lambda is not destroyed at ϕ+(ξ)=0\phi^{+}(\xi)=0. This occurs when it does not approach DD as CDC\to D; instead the attractor is destroyed in a subsequent heteroclinic bifurcation involving a period-three solution [13].

With δL=δR=0\delta_{L}=\delta_{R}=0 we have ϕ+(ξ)=τLϕ0(τL,τR)\phi^{+}(\xi)=\tau_{L}\phi_{0}(\tau_{L},\tau_{R}), so in this case the bifurcation surface ϕ+(ξ)=0\phi^{+}(\xi)=0 reduces to the homoclinic bifurcation ϕ0(τL,τR)=0\phi_{0}(\tau_{L},\tau_{R})=0 of the skew tent map family. This motivates the definition

n(1)={ξΦ(1)|ϕ+(gn(ξ))>0,ϕ+(gn+1(ξ))0},\displaystyle\mathcal{R}^{(1)}_{n}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi^{(1)}\,\middle|\,\phi^{+}(g^{n}(\xi))>0,\,\phi^{+}(g^{n+1}(\xi))\leq 0}\right\}, (5.5)

for all n0n\geq 0. The constraint α(ξ)<0\alpha(\xi)<0 is not needed in this definition because, by Proposition 4.2, if ξΦ(1)\xi\in\Phi^{(1)} then automatically α(ξ)<0\alpha(\xi)<0. As shown in [12] the regions n(1)\mathcal{R}^{(1)}_{n} are disjoint and cover the subset of Φ(1)\Phi^{(1)} for which ϕ+(ξ)>0\phi^{+}(\xi)>0. These regions are more difficult to visualise than those in §2 because they are four-dimensional. Fig. 6 shows two-dimensional slices of parameter space defined by fixing δL>0\delta_{L}>0 and δR>0\delta_{R}>0. In these slices only finitely many n(1)\mathcal{R}^{(1)}_{n} are visible because as nn\to\infty they converge to ξ=(1,0,1,0)\xi=(1,0,-1,0) (a fixed point of gg). The following result describes Λ\Lambda for parameter values in 0(1)\mathcal{R}^{(1)}_{0}. For a proof see [12].

Refer to caption Refer to caption
(a) δL=0.1,δR=0.1\delta_{L}=0.1,\delta_{R}=0.1. (b) δL=0.04,δR=0.04\delta_{L}=0.04,\delta_{R}=0.04.
Figure 6: Two-dimensional slices of the orientation-preserving parameter region Φ(1)\Phi^{(1)} showing the curves ϕ(gn(ξ))=0\phi(g^{n}(\xi))=0, for low values of nn, overlaid upon the numerical results of Eckstein’s greatest common divisor algorithm, explained in §6. Each point in a 200×200200\times 200 grid is coloured by the greatest common divisor of a set of close return iteration numbers JJ according to the colour bar on the right. In the algorithm we used ε=0.001\varepsilon=0.001 and M=106M=10^{6}; also points are coloured white if iterates appeared to diverge. The black triangles correspond to the parameter values of Fig. 5.
Theorem 5.1.

For any ξ0(1)\xi\in\mathcal{R}_{0}^{(1)},

  1. (i)

    Λ\Lambda is bounded, connected, and invariant under fξf_{\xi},

  2. (ii)

    has a positive Lyapunov exponent, and

  3. (iii)

    if δR<1\delta_{R}<1 there exists a forward invariant set Δ2\Delta\subset\mathbb{R}^{2} with non-empty interior such that

    n=0fξn(Δ)=Λ.\bigcap_{n=0}^{\infty}f_{\xi}^{n}(\Delta)=\Lambda. (5.6)

Item (ii) says Λ\Lambda is chaotic in the sense of having a positive Lyapunov exponent. We believe it satisfies Devaney’s definition of chaos throughout Φ(1)\Phi^{(1)}, but have only proved this in a subset of Φ(1)\Phi^{(1)} [13]. Item (iii) says Λ\Lambda is a Milnor attractor. We believe the condition δR>1\delta_{R}>1 is unnecessary and Λ\Lambda is in fact a topological attractor throughout 0(1)\mathcal{R}^{(1)}_{0}.

To understand the dynamics in all other regions n(1)\mathcal{R}_{n}^{(1)} we use the renormalisation operator gg. The following result follows simply from Proposition 4.3 and the definition of n(1)\mathcal{R}^{(1)}_{n}.

Proposition 5.2.

If ξn(1)\xi\in\mathcal{R}_{n}^{(1)} with n1n\geq 1, then g(ξ)n1(1)g(\xi)\in\mathcal{R}_{n-1}^{(1)}.

So if ξn(1)\xi\in\mathcal{R}_{n}^{(1)} then gn(ξ)0(1)g^{n}(\xi)\in\mathcal{R}_{0}^{(1)} where the attractor is connected. Then by nn applications of Proposition 4.4, fξf_{\xi} has an attractor with 2n2^{n} connected components as long as the assumptions of Proposition 4.4 are satisfied in each application. This is indeed the case, as shown in [12] through a series of careful calculations, and gives the following result.

Theorem 5.3.

For any ξn(1)\xi\in\mathcal{R}_{n}^{(1)} with n0n\geq 0, fξf_{\xi} has a chaotic Milnor attractor with exactly 2n2^{n} connected components.

For example Fig. 4-a uses ξ=ξex(1)\xi=\xi^{(1)}_{\rm ex} belonging to 1(1)\mathcal{R}^{(1)}_{1} so the attractor has two connected components, as shown. The renormalisation allows us to say more about this attractor. Specifically each piece of the attractor is an affine transformation of the one-component attractor of (1.1) with ξ=g(ξex(1))\xi=g\big{(}\xi^{(1)}_{\rm ex}\big{)} belonging to 0(1)\mathcal{R}^{(1)}_{0}, shown in Fig. 4-b.

6 A component counting algorithm.

To support an extension of the theoretical results of §5 to the orientation-reversing and non-invertible settings, we perform numerical simulations to count the number of connected components of the attractors. There are many methods for estimating the number of connected components of a set Ψ\Psi from a finite collection FF of points in Ψ\Psi. For example Robins et al [34] connect all pairs of points in FF with line segments to form a complete graph, then base their estimation from a minimal spanning tree. In our setting Ψ\Psi is generated by a map, so it is more effective to use a method that utilises the dynamics. For this reason we compute the number of components as the greatest common divisor of a certain set of computed values. This method originates with Eckstein [9] and is described by Avrutin et al [1]. As explained at the end of this section, the effectiveness of the method relies on the following result.

Lemma 6.1.

Suppose a compact invariant set Ψ\Psi of a continuous map ff has k2k\geq 2 connected components, and ff has an orbit that visits all components. Then the components can be labelled as Ψ1,Ψ2,,Ψk\Psi_{1},\Psi_{2},\ldots,\Psi_{k} such that f(Ψ1)=Ψ2,f(Ψ2)=Ψ3,,f(Ψk1)=Ψkf(\Psi_{1})=\Psi_{2},f(\Psi_{2})=\Psi_{3},\ldots,f(\Psi_{k-1})=\Psi_{k}, and f(Ψk)=Ψ1f(\Psi_{k})=\Psi_{1}.

Thus the components of Ψ\Psi are ordered cyclically, each has one ‘predecessor’ and one ‘successor’. The following proof is adapted from [1, 9]. It uses a small quantity ε\varepsilon to avoid reference to path-connectedness.

Proof of Lemma. 6.1.

By assumption there exists PΨP\in\Psi whose forward orbit enters all components. Let Ψ1\Psi_{1} be the component containing PP. Suppose for a contradiction f(Ψ1)f(\Psi_{1}) contains points in two different components, TT and UU. Since Ψ\Psi is compact there exists d>0d>0 such that any point in TT and any point in UU are at least a distance dd apart. Since Ψ1\Psi_{1} is connected for any ε>0\varepsilon>0 there exist Q,RΨ1Q,R\in\Psi_{1} with QR<ε\|Q-R\|<\varepsilon such that f(Q)Tf(Q)\in T and f(R)Uf(R)\in U. But f(Q)f(R)>d\|f(Q)-f(R)\|>d so this is not possible because ff is continuous, hence f(Ψ1)f(\Psi_{1}) is a subset of one component of Ψ\Psi. This component is not Ψ1\Psi_{1} (because then Ψ1\Psi_{1} would be forward invariant and the forward orbit of PP could not reach the other components), let us call it Ψ2\Psi_{2}. So f(Ψ1)Ψ2f(\Psi_{1})\subset\Psi_{2}.

By a similar argument f(Ψ2)f(\Psi_{2}) is a subset of one component, and if k>2k>2 this component is neither Ψ1\Psi_{1} nor Ψ2\Psi_{2}, call it Ψ3\Psi_{3}. Inductively we obtain Ψ1,Ψ2,,Ψk\Psi_{1},\Psi_{2},\ldots,\Psi_{k} with f(Ψ1)Ψ2,f(Ψ2)Ψ3,,f(Ψk1)Ψkf(\Psi_{1})\subset\Psi_{2},f(\Psi_{2})\subset\Psi_{3},\ldots,f(\Psi_{k-1})\subset\Psi_{k}. Also f(Ψk)f(\Psi_{k}) is a subset of one component.

But Ψ\Psi is invariant, meaning f(Ψ)=Ψf(\Psi)=\Psi. Thus f(Ψk)Ψ1f(\Psi_{k})\subset\Psi_{1}, and further f(Ψ1)=Ψ2,f(Ψ2)=Ψ3,,f(Ψk1)=Ψkf(\Psi_{1})=\Psi_{2},f(\Psi_{2})=\Psi_{3},\ldots,f(\Psi_{k-1})=\Psi_{k}, and f(Ψk)=Ψ1f(\Psi_{k})=\Psi_{1} as required. ∎

Now fix ξ\xi and suppose fξf_{\xi} has an attractor Λ\Lambda with k1k\geq 1 connected components. To compute the value kk, which is assumed to be unknown to us a priori, the algorithm proceeds as follows. Fix ε>0\varepsilon>0 (we used ε=0.001\varepsilon=0.001 or ε=0.0001\varepsilon=0.0001) and M>0M>0 (we used M=106M=10^{6} ) and let J=J=\varnothing. Choose some initial point assumed to be in the basin of attraction of Λ\Lambda and iterate it under fξf_{\xi} a reasonably large number of times (we used 10410^{4} iterations) to remove transient dynamics and obtain a point in Λ\Lambda, or extremely close to Λ\Lambda, call it (x0,y0)(x_{0},y_{0}). Iterate further, and for all i=1,2,,Mi=1,2,\ldots,M evaluate the distance (Euclidean norm in 2\mathbb{R}^{2}) between fξi(x0,y0)f_{\xi}^{i}(x_{0},y_{0}) and (x0,y0)(x_{0},y_{0}). If this distance is less than ε\varepsilon, append the number ii to the set JJ. Finally evaluate the greatest common divisor of the elements in JJ — this is our estimate for the value of kk.

For example using ξ=ξex(1)\xi=\xi_{\rm ex}^{(1)}, as in Fig. 4-a, the algorithm generated a set of 12491249 numbers

J={1292,3170,3778,,999930},J=\{1292,3170,3778,\ldots,999930\},

whose elements have greatest common divisor 22 (and indeed at this parameter point the attractor has two components).

Fig. 6 shows the output of this algorithm over a 200×200200\times 200 grid of parameter points. The results show excellent agreement to the theory described in §5. For example parameter points where the greatest common divisor of the numbers in JJ is two have boundaries indistinguishable from the true bifurcation boundaries ϕ(g(ξ))=0\phi(g(\xi))=0 and ϕ(g2(ξ))=0\phi(g^{2}(\xi))=0.

Two principles underlie the effectiveness of the algorithm. First, if the distance between any two components of Λ\Lambda is greater than ε\varepsilon, then fξi(x0,y0)(x0,y0)<ε\mathopen{}\mathclose{{}\left\|f_{\xi}^{i}(x_{0},y_{0})-(x_{0},y_{0})}\right\|<\varepsilon implies that fξi(x0,y0)f_{\xi}^{i}(x_{0},y_{0}) and (x0,y0)(x_{0},y_{0}) belong to the same component of Λ\Lambda (assuming (x0,y0)Λ(x_{0},y_{0})\in\Lambda). By Lemma 6.1 this is only possible if ii is a multiple of kk. Thus the elements of JJ are all multiples of kk.

Second, assuming fξf_{\xi} is ergodic on Λ\Lambda, the set of all i>0i>0 giving fξi(x0,y0)(x0,y0)<ε\mathopen{}\mathclose{{}\left\|f_{\xi}^{i}(x_{0},y_{0})-(x_{0},y_{0})}\right\|<\varepsilon will not be expected to share a multiple larger than kk. This is because ergodicity implies the dynamics of fξkf_{\xi}^{k} on any component are well mixed (see Haydn et al [15] and references within for theory on the probability distribution of the values of ii when ε\varepsilon is small). In our setting M=106M=10^{6} seems to generate large enough sets JJ to ensure the greatest common divisor of the numbers in JJ is kk, instead a multiple of kk. One can also modify the algorithm, as described in [1, 9], to search for a close return to several reference points, instead of the single point (x0,y0)(x_{0},y_{0}).

7 The orientation-reversing case

Let

Φ(2)={ξΦ|δL<0,δR<0},\displaystyle\Phi^{(2)}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi\ \middle|\ \delta_{L}<0,\delta_{R}<0}\right\}, (7.1)

be the subset of Φ\Phi for which the BCNF is orientation-reversing. As shown originally by Misiurewicz [25], the closure of the unstable manifold of the fixed point XX can be a chaotic attractor. This attractor contains the point TT, as shown in Fig. 3-d, so under parameter variation is destroyed when T=CT=C. This is a heteroclinic bifurcation beyond which the unstable manifold of XX is unbounded. From the formulas (3.5) and (3.7) we obtain

C1T1=ϕ(ξ)(λLu1)(1λRs)(δRτRλLu),\displaystyle C_{1}-T_{1}=\frac{\phi^{-}(\xi)}{(\lambda_{L}^{u}-1)(1-\lambda_{R}^{s})(\delta_{R}-\tau_{R}\lambda_{L}^{u})}, (7.2)

where

ϕ(ξ)=δR(δR+τR(1+λRu)λLu)λLu.\displaystyle\phi^{-}(\xi)=\delta_{R}-(\delta_{R}+\tau_{R}-(1+\lambda_{R}^{u})\lambda_{L}^{u})\lambda_{L}^{u}\,. (7.3)

An example illustrating the destruction of the attractor is shown in Fig. 7. As the value of τL\tau_{L} is increased the attractor is destroyed when T=CT=C at τL2.0104\tau_{L}\approx 2.0104.

Refer to caption
(a) τL=1.8\tau_{L}=1.8
Refer to caption
(b) τL2.0104\tau_{L}\approx 2.0104
Refer to caption
(c) τL=2.2\tau_{L}=2.2
Figure 7: Phase portraits of (1.1) with δL=0.2\delta_{L}=-0.2, δR=0.2\delta_{R}=-0.2, τR=1.8\tau_{R}=-1.8, and three different values of τL\tau_{L} corresponding to the black triangles in Fig. 8-b. Panel (b) uses τL\tau_{L} such that ϕ(ξ)=0\phi^{-}(\xi)=0 to ten decimal places. In (a) cl(Wu(X)){\rm cl}(W^{u}(X)) is a chaotic attractor; in (b) Wu(X)W^{u}(X) and Ws(Y)W^{s}(Y) have corner intersections; in (c) there is no attractor.

Analogous to the orientation-preserving case, if δL=δR=0\delta_{L}=\delta_{R}=0 the expression ϕ(ξ)\phi^{-}(\xi) reduces to ϕ(ξ)=τLϕ0(τL,τR)\phi^{-}(\xi)=\tau_{L}\phi_{0}(\tau_{L},\tau_{R}). Thus the heteroclinic bifurcation ϕ(ξ)=0\phi^{-}(\xi)=0 reduces to the familiar homoclinic bifurcation ϕ0(τL,τR)=0\phi_{0}(\tau_{L},\tau_{R})=0 of the skew tent map family in the limit (δL,δR)(0,0)(\delta_{L},\delta_{R})\to(0,0).

We are now ready to subdivide Φ(2)\Phi^{(2)} into regions based upon the renormalisation operator gg, (4.1). But ξΦ(2)\xi\in\Phi^{(2)} implies g(ξ)Φ(1)g(\xi)\in\Phi^{(1)}, so we again use the preimages of ϕ+(ξ)=0\phi^{+}(\xi)=0 under gg to define the region boundaries. Specifically we let

0(2)\displaystyle\mathcal{R}^{(2)}_{0} ={ξΦ(2)|ϕ(ξ)>0,ϕ+(g(ξ))0,α(ξ)<0},\displaystyle=\mathopen{}\mathclose{{}\left\{\xi\in\Phi^{(2)}\,\middle|\,\phi^{-}(\xi)>0,\phi^{+}(g(\xi))\leq 0,\alpha(\xi)<0}\right\}, (7.4)
n(2)\displaystyle\mathcal{R}^{(2)}_{n} ={ξΦ(2)|ϕ+(gn(ξ))>0,ϕ+(gn+1(ξ))0,α(ξ)<0},for all n1.\displaystyle=\mathopen{}\mathclose{{}\left\{\xi\in\Phi^{(2)}\,\middle|\,\phi^{+}\mathopen{}\mathclose{{}\left(g^{n}(\xi)}\right)>0,\phi^{+}\mathopen{}\mathclose{{}\left(g^{n+1}(\xi)}\right)\leq 0,\alpha(\xi)<0}\right\},\qquad\text{for all $n\geq 1$.} (7.5)

Fig. 8 shows these regions in two typical two-dimensional slices of parameter space. Unlike in the orientation-preserving setting we need to impose α(ξ)<0\alpha(\xi)<0 in these regions so that they don’t overlap 𝒫2\mathcal{P}_{2} where a stable LRLR-cycle exists.

Refer to caption Refer to caption
(a) δL=0.1,δR=0.2\delta_{L}=-0.1,\delta_{R}=-0.2. (b) δL=0.2,δR=0.2\delta_{L}=-0.2,\delta_{R}=-0.2.
Figure 8: Two-dimensional slices of the orientation-reversing parameter region Φ(2)\Phi^{(2)} showing the bifurcation curves overlaid upon numerical results obtained using ε=0.0001\varepsilon=0.0001 and M=106M=10^{6}. The black triangles correspond to the parameter values of Fig. 7.

We conjecture that throughout 0(2)\mathcal{R}^{(2)}_{0} the map has a unique chaotic attractor with one connected component equal to the closure of the unstable manifold of XX, as in Fig. 7-a. In earlier work [11] we proved this to be true on a subset of 0(2)\mathcal{R}^{(2)}_{0}. In the special case τL=τR\tau_{L}=-\tau_{R} and δL=δR\delta_{L}=\delta_{R} of the Lozi family of maps, the constraint ϕ(ξ)>0\phi^{-}(\xi)>0 reduces to equation (3) of Misiurewicz [25].

The following result is a simple consequence of Propositions 4.2 and 4.3 and (7.5).

Proposition 7.1.

If ξn(2)\xi\in\mathcal{R}_{n}^{(2)} with n1n\geq 1, then g(ξ)n1(1)g(\xi)\in\mathcal{R}_{n-1}^{(1)}.

This suggests that for any ξn(2)\xi\in\mathcal{R}_{n}^{(2)} the BCNF has an attractor with 2n2^{n} connected components. For example

ξex(2)=(2.5,0.1,1.1,0.2)\displaystyle\xi_{\rm ex}^{(2)}=\mathopen{}\mathclose{{}\left(2.5,-0.1,-1.1,-0.2}\right) (7.6)

belongs to 1(2)\mathcal{R}^{(2)}_{1} and indeed the attractor shown in Fig. 9-a appears to have two connected components. One component belongs to Πξ\Pi_{\xi} so Proposition 4.4 applies. Hence this component is an affine transformation of the attractor of (1.1) at the parameter point g(ξex(2))g\big{(}\xi_{\rm ex}^{(2)}\big{)} that belongs to 0(1)\mathcal{R}^{(1)}_{0}, and we know its attractor has one component by Theorem 5.1.

Fig. 8 shows that our conjecture is supported by the output of the greatest common divisor algorithm. The boundaries of the n(2)\mathcal{R}^{(2)}_{n} closely approximate the places with the value of the greatest common divisor changes. This value changes slightly above ϕ+(g(ξ))=0\phi^{+}(g(\xi))=0 because here the two components of the attractor are very close and ε=0.0001\varepsilon=0.0001 is insufficient to detect this difference. Also 0(2)\mathcal{R}^{(2)}_{0} has pixels erroneously corresponding to more than one component because here the attractor is relatively large and M=106M=10^{6} iterations are insufficient for the algorithm to consistently obtain a greatest common divisor of 11.

Refer to caption Refer to caption
(a) ξ=ξex(2)1(2)\xi=\xi_{\rm ex}^{(2)}\in\mathcal{R}_{1}^{(2)} (b) ξ=g(ξex(2))0(1)\xi=g\big{(}\xi_{\rm ex}^{(2)}\big{)}\in\mathcal{R}_{0}^{(1)}
Figure 9: Panel (a) is a phase portrait of (1.1) at the parameter point ξex(2)\xi_{\rm ex}^{(2)} (7.6) where the attractor has two connected components in the orientation-reversing case. Panel (b) uses instead g(ξex(2))g\big{(}\xi_{\rm ex}^{(2)}\big{)}.

8 The non-invertible case δL>0\delta_{L}>0, δR<0\delta_{R}<0.

Here we study the parameter region

Φ(3)={ξΦ|δL>0,δR<0},\displaystyle\Phi^{(3)}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi\ \middle|\ \delta_{L}>0,\delta_{R}<0}\right\}, (8.1)

where (1.1) is non-invertible. In this region an attractor of (1.1) can be destroyed by crossing the homoclinic bifurcation ϕ+(ξ)=0\phi^{+}(\xi)=0 or the heteroclinic bifurcation ϕ(ξ)=0\phi^{-}(\xi)=0. This is because near these boundaries the attractor contains the point TT and is close to the point DD so is destroyed when one of these points collides with CC. For example in Fig. 10 TT lies to the left of DD, so the attractor is destroyed when D=CD=C, i.e. when ϕ+(ξ)=0\phi^{+}(\xi)=0. In contrast in Fig. 11 TT lies to the right of DD, so the attractor is destroyed when T=CT=C, i.e. when ϕ(ξ)=0\phi^{-}(\xi)=0. For this reason we define

ϕmin(ξ)=min[ϕ+(ξ),ϕ(ξ)],\displaystyle\phi_{\rm min}(\xi)={\rm min}[\phi^{+}(\xi),\phi^{-}(\xi)], (8.2)

and

n(3)={ξΦ(3)|ϕmin(gn(ξ))>0,ϕmin(gn+1(ξ))0,α(ξ)<0},\displaystyle\mathcal{R}^{(3)}_{n}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi^{(3)}\,\middle|\,\phi_{\rm min}\mathopen{}\mathclose{{}\left(g^{n}(\xi)}\right)>0,\,\phi_{\rm min}\mathopen{}\mathclose{{}\left(g^{n+1}(\xi)}\right)\leq 0,\,\alpha(\xi)<0}\right\}, (8.3)

for all n0n\geq 0. Two-dimensional slices of these regions are shown in Fig. 12. We conjecture that throughout the first region 0(3)\mathcal{R}_{0}^{(3)} the BCNF has a unique chaotic attractor with one connected component, and this is supported by the numerical results shown in Fig. 12. To explain the dynamics in the remaining regions we use the following analogy to Propositions 5.2 and 7.1. Here, however, the result is not trivial, so we provide a proof.

Refer to caption
(a) τL=1.7\tau_{L}=1.7
Refer to caption
(b) τL1.7948\tau_{L}\approx 1.7948
Refer to caption
(c) τL=1.8\tau_{L}=1.8
Figure 10: Phase portraits of (1.1) with δL=0.3\delta_{L}=0.3, δR=0.4\delta_{R}=-0.4, τR=2.4\tau_{R}=-2.4 and three different values of τL\tau_{L} corresponding to the blue triangles in Fig. 12-b.
Refer to caption
(a) τR=1.1\tau_{R}=-1.1
Refer to caption
(b) τR1.2654\tau_{R}\approx-1.2654
Refer to caption
(c) τR=1.4\tau_{R}=-1.4
Figure 11: Phase portraits of (1.1) with δL=0.3\delta_{L}=0.3, δR=0.4\delta_{R}=-0.4, τL=3\tau_{L}=3, and three different values of τR\tau_{R} corresponding to the black triangles in Fig. 12-b.
Refer to caption Refer to caption
(a) δL=0.5,δR=0.4\delta_{L}=0.5,\delta_{R}=-0.4. (b) δL=0.3,δR=0.4\delta_{L}=0.3,\delta_{R}=-0.4.
Figure 12: Two-dimensional slices of the non-invertible parameter region Φ(3)\Phi^{(3)} showing the bifurcation curves overlaid upon numerical results obtained using ε=0.001\varepsilon=0.001 and M=106M=10^{6}. The blue [black] triangles correspond to the parameter values of Fig. 10 [Fig. 11].
Proposition 8.1.

If ξn(3)\xi\in\mathcal{R}^{(3)}_{n} with n1n\geq 1, then g(ξ)n1(3)g(\xi)\in\mathcal{R}^{(3)}_{n-1}.

Proof.

Choose any ξn(3)\xi\in\mathcal{R}^{(3)}_{n} with n1n\geq 1 and write g(ξ)=(τ~L,δ~L,τ~R,δ~R)g(\xi)=\big{(}\tilde{\tau}_{L},\tilde{\delta}_{L},\tilde{\tau}_{R},\tilde{\delta}_{R}\big{)} as in the proof of Proposition 4.3. By Proposition 4.3 we have g(ξ)Φg(\xi)\in\Phi. Further, g(ξ)Φ(3)g(\xi)\in\Phi^{(3)} because δ~L=δR2>0\tilde{\delta}_{L}=\delta_{R}^{2}>0 and δ~R=δLδR<0\tilde{\delta}_{R}=\delta_{L}\delta_{R}<0. Also ϕmin(gn1(g(ξ)))>0\phi_{\rm min}\big{(}g^{n-1}(g(\xi))\big{)}>0 and ϕmin(gn(g(ξ)))0\phi_{\rm min}\big{(}g^{n}(g(\xi))\big{)}\leq 0. Finally α(g(ξ))=τ~Lτ~R+(δ~L1)(δ~R1)\alpha(g(\xi))=\tilde{\tau}_{L}\tilde{\tau}_{R}+(\tilde{\delta}_{L}-1)(\tilde{\delta}_{R}-1) where τ~L=τR22δR\tilde{\tau}_{L}=\tau_{R}^{2}-2\delta_{R}, δ~L=δR2\tilde{\delta}_{L}=\delta_{R}^{2}, τ~R=τLτRδLδR\tilde{\tau}_{R}=\tau_{L}\tau_{R}-\delta_{L}-\delta_{R}, and δ~R=δLδR\tilde{\delta}_{R}=\delta_{L}\delta_{R}. By substituting τL>δL+1\tau_{L}>\delta_{L}+1 and τR<(δR+1)\tau_{R}<-(\delta_{R}+1) (true because ξΦ\xi\in\Phi) we obtain (after simplification)

α(g(ξ))<2(1+δR+δR2)(δL+δR).\alpha(g(\xi))<-2\mathopen{}\mathclose{{}\left(1+\delta_{R}+\delta_{R}^{2}}\right)(\delta_{L}+\delta_{R}).

So if δL+δR>0\delta_{L}+\delta_{R}>0 then α(g(ξ))<0\alpha(g(\xi))<0 (because 1+δR+δR234>01+\delta_{R}+\delta_{R}^{2}\geq\frac{3}{4}>0). If instead δL+δR0\delta_{L}+\delta_{R}\leq 0 then δ~L+δ~R=δR(δL+δR)0\tilde{\delta}_{L}+\tilde{\delta}_{R}=\delta_{R}(\delta_{L}+\delta_{R})\geq 0 because δR<0\delta_{R}<0, so again α(g(ξ))<0\alpha(g(\xi))<0 by Proposition 4.2 applied to the parameter point g(ξ)g(\xi). ∎

Proposition 8.1 suggests that throughout each n(3)\mathcal{R}^{(3)}_{n} with n1n\geq 1 the BCNF has an attractor with exactly 2n2^{n} connected components, and Fig. 12 supports this conjecture. Fig. 13-a provides an example using the parameter point

ξex(3)=(3,0.3,.9,0.4),\displaystyle\xi_{\rm ex}^{(3)}=\mathopen{}\mathclose{{}\left(3,0.3,-.9,-0.4}\right), (8.4)

which belongs to 1(3)\mathcal{R}_{1}^{(3)}. The attractor has two pieces, one of which belongs to Πξ\Pi_{\xi}, so as above this piece is an affine transformation of the attractor of g(ξex(3))0(3)g\big{(}\xi^{(3)}_{\rm ex}\big{)}\in\mathcal{R}^{(3)}_{0} shown in Fig. 13-b.

Refer to caption Refer to caption
(a) ξ=ξex(3)1(3)\xi=\xi_{\rm ex}^{(3)}\in\mathcal{R}_{1}^{(3)} (b) ξ=g(ξex(3))0(3)\xi=g\big{(}\xi_{\rm ex}^{(3)}\big{)}\in\mathcal{R}_{0}^{(3)}
Figure 13: Panel (a) is a phase portrait of (1.1) at the parameter point ξex(3)\xi_{\rm ex}^{(3)} (8.4) where the attractor has two connected components in the δL>0\delta_{L}>0 non-invertible case. Panel (b) uses instead g(ξex(3))g\big{(}\xi_{\rm ex}^{(3)}\big{)}.

9 The non-invertible case δL<0\delta_{L}<0, δR>0\delta_{R}>0

It remains for us to consider

Φ(4)={ξΦ|δL<0,δR>0},\displaystyle\Phi^{(4)}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi\ \middle|\ \delta_{L}<0,\delta_{R}>0}\right\}, (9.1)

where (1.1) is non-invertible. In this region the attractor is usually destroyed before the boundaries ϕ+(ξ)=0\phi^{+}(\xi)=0 and ϕ(ξ)=0\phi^{-}(\xi)=0 in a heteroclinic bifurcation that cannot be characterised by an explicit condition on the parameter values. This occurs when the attractor contains points on y=0y=0 that lie to the right of DD and TT and is destroyed when its right-most point collides with CC.

Fig. 14 shows an example. In panel (a) the attractor is the closure of the unstable manifold of XX. As we grow the unstable manifold outwards from XX it develops points on y=0y=0 that lie further and further to the right, but not beyond CC. This occurs as a consequence of the geometric configuration afforded by fRf_{R} being orientation-preserving and fLf_{L} being orientation-reversing. As the value of τL\tau_{L} is increased the attractor is destroyed when its right-most point collides with CC, panel (b). With a slightly value of τL\tau_{L}, panel (c), typical forward orbits diverge even though DD and TT still lie to the left of CC. Fig. 15 provides a second example. Here the parameter values are such that DD and TT are switched around but the attractor is destroyed in the same way.

Refer to caption
(a) τL=1.2\tau_{L}=1.2
Refer to caption
(b) τL1.3439\tau_{L}\approx 1.3439
Refer to caption
(c) τL=1.35\tau_{L}=1.35
Figure 14: Phase portraits of (1.1) with δL=0.4\delta_{L}=-0.4, δR=0.4\delta_{R}=0.4, τR=2.8\tau_{R}=-2.8, and three different values of τL\tau_{L} corresponding to the blue triangles in Fig. 16-a. The parameter value used in panel (b) is approximately where the attractor is destroyed. In all three panels DD lies to the left of TT which lies to the left of CC.
Refer to caption
(a) τL=2\tau_{L}=2
Refer to caption
(b) τL2.2285\tau_{L}\approx 2.2285
Refer to caption
(c) τL=2.3\tau_{L}=2.3
Figure 15: Phase portraits of (1.1) with δL=0.4\delta_{L}=-0.4, δR=0.4\delta_{R}=0.4, τR=1.8\tau_{R}=-1.8, and three different values of τL\tau_{L} corresponding to the black triangles in Fig. 16-a. The parameter value used in panel (b) is approximately where the attractor is destroyed. In all three panels TT lies to the left of DD which lies to the left of CC.
Refer to caption Refer to caption
(a) δL=0.4,δR=0.4\delta_{L}=-0.4,\delta_{R}=0.4. (b) δL=0.5,δR=0.4\delta_{L}=-0.5,\delta_{R}=0.4.
Figure 16: Two-dimensional slices of the non-invertible parameter region Φ(4)\Phi^{(4)} showing the bifurcation curves overlaid upon numerical results obtained using ε=0.001\varepsilon=0.001 and M=106M=10^{6}. The blue [black] triangles correspond to the parameter values of Fig. 14 [Fig. 15].

Fig. 16 shows the results of numerical simulations applied to two slices of Φ(4)\Phi^{(4)}. The white areas just to the left of ϕ+(ξ)=0\phi^{+}(\xi)=0 and ϕ(ξ)=0\phi^{-}(\xi)=0 correspond to the phenomenon that we have just described. These areas are bounded on the left by a curve of heteroclinic bifurcations. Since we cannot identify this curve with hand calculations, it is natural to attempt to compute the curve numerically with numerical continuation methods. However, we do not show the result of such a computation because we suspect this curve highly irregular, e.g. non-differentiable at infinitely many points, as explained by Osinga [28].

In each plot in Fig. 16 the attractor does in fact persist up to where the boundaries ϕ+(ξ)=0\phi^{+}(\xi)=0 and ϕ(ξ)=0\phi^{-}(\xi)=0 meet. This is because on the curve λLs=λRs\lambda_{L}^{s}=\lambda_{R}^{s} the right-most point of the attractor is DD and TT, which are equal. Also near the top of Fig. 16-a the attractor persists slightly beyond ϕ+(ξ)=0\phi^{+}(\xi)=0 because here the attractor fails to approach DD so is not destroyed when D=CD=C at ϕ+(ξ)=0\phi^{+}(\xi)=0.

Despite the extra complexities in Φ(4)\Phi^{(4)} it still appears that renormalisation is helpful for explaining the bifurcation structure. Let

0(4)={ξΦ(4)|ϕmin(ξ)>0,ϕmin(g(ξ))0,α(ξ)<0}.\displaystyle\mathcal{R}^{(4)}_{0}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi^{(4)}\middle|\ \phi_{\rm min}(\xi)>0,\,\phi_{\rm min}(g(\xi))\leq 0,\,\alpha(\xi)<0}\right\}. (9.2)

Based on Fig. 16 we conjecture that for any ξ0(4)\xi\in\mathcal{R}^{(4)}_{0}, if (1.1) has an attractor then this attractor is chaotic with one connected component. Now let

n(4)={ξΦ(4)|ϕmin(gn(ξ))>0,ϕmin(gn+1(ξ))0,α(ξ)<0,α(g(ξ))<0}.\displaystyle\mathcal{R}^{(4)}_{n}=\mathopen{}\mathclose{{}\left\{\xi\in\Phi^{(4)}\middle|\ \phi_{\rm min}(g^{n}(\xi))>0,\,\phi_{\rm min}(g^{n+1}(\xi))\leq 0,\,\alpha(\xi)<0,\alpha(g(\xi))<0}\right\}. (9.3)

Unlike in previous sections here we have included the extra constraint α(g(ξ))<0\alpha(g(\xi))<0 so that the n(4)\mathcal{R}^{(4)}_{n} do not include the region 𝒫4\mathcal{P}_{4}, defined to be where (1.1) has a stable period-four solution with symbolic itinerary LRRRLRRR. This region is visible in Fig. 16-b and shown more clearly in the magnification, Fig. 17. Note that here we do not show the result of the numerics because the component counting algorithm does not work efficiently close to the curve α(g(ξ))=0\alpha(g(\xi))=0. The following result is a trivial consequence of our definitions.

Proposition 9.1.

If ξn(4)\xi\in\mathcal{R}^{(4)}_{n} with n1n\geq 1, then g(ξ)n1(3)g(\xi)\in\mathcal{R}^{(3)}_{n-1}.

Refer to caption
Figure 17: A magnified version of Fig. 16-b showing the regions having a stable LRLR-cycle (period-two solution) and a stable LRRRLRRR-cycle (period-four solution).

Based on this we conjecture that for any ξn(4)\xi\in\mathcal{R}^{(4)}_{n} with n1n\geq 1 the map (1.1) has a chaotic attractor with exactly 2n2^{n} connected components, and this is supported by the numerics in Fig. 16. Fig. 18-a shows the attractor of (1.1) for a typical parameter point in 1(4)\mathcal{R}^{(4)}_{1}, specifically

ξex(4)=(2.2,0.4,1.5,0.4).\displaystyle\xi_{\rm ex}^{(4)}=\mathopen{}\mathclose{{}\left(2.2,-0.4,-1.5,0.4}\right). (9.4)

As expected it has two connected components, one of which is contained in Πξ\Pi_{\xi}, and both of which are affine transformations of the single-component attractor of (1.1) with g(ξex(4))0(3)g\big{(}\xi^{(4)}_{\rm ex}\big{)}\in\mathcal{R}^{(3)}_{0} shown in Fig. 18-b.

Refer to caption Refer to caption
(a) ξ=ξex(4)1(4)\xi=\xi_{\rm ex}^{(4)}\in\mathcal{R}_{1}^{(4)} (b) ξ=g(ξex(4))0(3)\xi=g\big{(}\xi_{\rm ex}^{(4)}\big{)}\in\mathcal{R}_{0}^{(3)}
Figure 18: Panel (a) is a phase portrait of (1.1) at the parameter point ξex(4)\xi_{\rm ex}^{(4)} (9.4) where the attractor has two connected components in the δL<0\delta_{L}<0 non-invertible case. Panel (b) uses instead g(ξex(4))g\big{(}\xi_{\rm ex}^{(4)}\big{)}.

10 Reduction to one dimension

Finally we address a novelty of the non-invertible settings. In these settings the stable eigenvalues λLs\lambda_{L}^{s} and λRs\lambda_{R}^{s} have the same sign (they are both positive if δL>0\delta_{L}>0 and δR<0\delta_{R}<0, and both negative if δL<0\delta_{L}<0 and δR>0\delta_{R}>0), so it is possible for them to be equal. From the formulas (3.5) and (3.6) for the points DD and TT we have

D1T1=λLsλRs(1λLs)(1λRs).\displaystyle D_{1}-T_{1}=\frac{\lambda_{L}^{s}-\lambda_{R}^{s}}{\mathopen{}\mathclose{{}\left(1-\lambda_{L}^{s}}\right)\mathopen{}\mathclose{{}\left(1-\lambda_{R}^{s}}\right)}. (10.1)

Thus DD and TT coincide when the stable eigenvalues are equal. Thus the boundaries ϕ+(ξ)=0\phi^{+}(\xi)=0 and ϕ(ξ)=0\phi^{-}(\xi)=0, where C=DC=D and C=TC=T respectively, intersect where λLs=λRs\lambda_{L}^{s}=\lambda_{R}^{s}, and this is evident in Figs. 12 and 16. Also for each n1n\geq 1 the boundaries ϕ+(gn(ξ))=0\phi^{+}(g^{n}(\xi))=0 and ϕ(gn(ξ))=0\phi^{-}(g^{n}(\xi))=0 intersect where λLs=λRs\lambda_{L}^{s}=\lambda_{R}^{s}.

We now show that if λLs=λRs\lambda_{L}^{s}=\lambda_{R}^{s} then the pertinent dynamics reduces to one dimension, as in Fig. 19.

Refer to caption Refer to caption
(a) ξ=(2.323,0.5,1.427,0.4)\xi=\mathopen{}\mathclose{{}\left(2.323,0.5,-1.427,-0.4}\right) (b) corresponding cobweb diagram
Figure 19: Panel (a) shows the attractor of (1.1) at the given parameter point which belongs to 0(3)\mathcal{R}^{(3)}_{0}. This attractor is the line segment from TT to fξ(T)f_{\xi}(T). The restriction of fξf_{\xi} to the attractor is the one-dimensional map (10.2) indicated in panel (b).
Proposition 10.1.

If ξΦ\xi\in\Phi with ϕmin(ξ)>0\phi_{\rm min}(\xi)>0 and λLs=λRs\lambda_{L}^{s}=\lambda_{R}^{s}, then fξf_{\xi} is forward invariant on the line segment from TT to fξ(T)f_{\xi}(T). Moreover, on this segment fξf_{\xi} is conjugate to the skew tent map

z{λLuz+1,z0,λRuz+1,z0,z\mapsto\begin{cases}\lambda_{L}^{u}z+1,&z\leq 0,\\ \lambda_{R}^{u}z+1,&z\geq 0,\end{cases} (10.2)

on [λRu+1,1]\mathopen{}\mathclose{{}\left[\lambda_{R}^{u}+1,1}\right].

Proof.

The line segment from TT to fξ(T)f_{\xi}(T) is

Γ={γ(z)|z[λRu+1,1]}, where γ(z)=(z1λRs,(1z)λRs1λRs),\Gamma=\mathopen{}\mathclose{{}\left\{\gamma(z)\,\middle|\,z\in\big{[}\lambda_{R}^{u}+1,1\big{]}}\right\},\text{~{}where~{}}\gamma(z)=\mathopen{}\mathclose{{}\left(\frac{z}{1-\lambda_{R}^{s}},\frac{(1-z)\lambda_{R}^{s}}{1-\lambda_{R}^{s}}}\right),

because putting z=1z=1 gives (x,y)=(11λRs,0)=T(x,y)=\mathopen{}\mathclose{{}\left(\frac{1}{1-\lambda_{R}^{s}},0}\right)=T by (3.5), while putting z=λRu+1z=\lambda_{R}^{u}+1 gives (x,y)=(1+λRu1λRs,λRsλRu1λRs)(x,y)=\mathopen{}\mathclose{{}\left(\frac{1+\lambda_{R}^{u}}{1-\lambda_{R}^{s}},\frac{-\lambda_{R}^{s}\lambda_{R}^{u}}{1-\lambda_{R}^{s}}}\right) which is identical to fξ(T)=(τRT1+1,δRT1)f_{\xi}(T)=(\tau_{R}T_{1}+1,-\delta_{R}T_{1}) (since τR=λRs+λRu\tau_{R}=\lambda_{R}^{s}+\lambda_{R}^{u} and δR=λRsλRu\delta_{R}=\lambda_{R}^{s}\lambda_{R}^{u}). If z[0,1]z\in[0,1] then

fξ(γ(z))=(τRx+y+1,δRx)=(1+λRuz1λRs,λRsλRuz1λRs)=γ(λRuz+1),f_{\xi}(\gamma(z))=\mathopen{}\mathclose{{}\left(\tau_{R}x+y+1,-\delta_{R}x}\right)=\mathopen{}\mathclose{{}\left(\frac{1+\lambda_{R}^{u}z}{1-\lambda_{R}^{s}},\frac{-\lambda_{R}^{s}\lambda_{R}^{u}z}{1-\lambda_{R}^{s}}}\right)=\gamma(\lambda_{R}^{u}z+1),

while if z[λRu+1,1]z\in[\lambda_{R}^{u}+1,1] then

fξ(γ(z))=(τLx+y+1,δLx)=(1+λLuz1λRs,λRsλLuz1λRs)=γ(λLuz+1),f_{\xi}(\gamma(z))=\mathopen{}\mathclose{{}\left(\tau_{L}x+y+1,-\delta_{L}x}\right)=\mathopen{}\mathclose{{}\left(\frac{1+\lambda_{L}^{u}z}{1-\lambda_{R}^{s}},\frac{-\lambda_{R}^{s}\lambda_{L}^{u}z}{1-\lambda_{R}^{s}}}\right)=\gamma(\lambda_{L}^{u}z+1),

and the result follows. ∎

Notice λLu>1\lambda_{L}^{u}>1 and λRu<1\lambda_{R}^{u}<-1, which corresponds in Fig. 2 to a point in some n\mathcal{R}_{n}, with n0n\geq 0. Thus, with this value of nn, the attractor of the skew tent map (10.2) is comprised of 2n2^{n} disjoint intervals. Consequently the attractor of fξf_{\xi} is comprised of 2n2^{n} disjoint line segments. Places where the value of nn changes can be computed by solving for where ϕ0gn=0\phi_{0}\circ g^{n}=0 in Fig. 2. We computed these points numerically and have plotted them as black dots in Figs. 12 and 16 from which we see that, as expected, these points are where ϕ+gn=0\phi^{+}\circ g^{n}=0 and ϕgn=0\phi^{-}\circ g^{n}=0. In this way the conjectures in §8 and §9 on the number of connected components of the attractors are confirmed in the special codimension-one scenario that the stable stability multipliers associated with XX and YY are equal.

11 Discussion

It has long been known that symmetric tent maps and skew tent maps readily admit attractors with 2n2^{n} connected components, and that for any given slopes for the two pieces of the map the value of nn can be determined through renormalisation. Here we have shown how this can be realised for two-dimensional maps and uncovered some novel complexities relating to the possibility of stable period-two and period-four solutions. We have also shown how the attractor can be destroyed at heteroclinic bifurcations (boundary crises) that cannot be characterised algebraically.

It remains to verify the conjectures in §7–§9 on the number of connected components of the attractor in the various parameter regions we have defined. We feel this should be possible by following the methodology used in [12] but suspect it will be a substantial undertaking. Chaos in higher-dimensional piecewise-linear maps has applications to cryptography [19], and it remains to see what aspects of the renormalisation can be described for the NN-dimensional border-collision normal form [36] with no restriction on NN. Also it remains to see if renormalisation schemes based on other symbolic substitution rules can be used to explain parameter regimes where (1.1) has attractors with other numbers of components, e.g. three components, as described in [14].

Finally we note that Proposition 10.1 provides a rare example of dimension reduction in a piecewise-smooth setting. Dimension reduction is core element of smooth bifurcation theory whereby centre manifolds allow us to explain the dynamics of high dimensional systems with low dimensional equations [20]. In general the bifurcation theory of piecewise-smooth systems is hampered by an inability to do this usually centre manifolds usually do not exist [7]. It remains to see how far Proposition 10.1 can be generalised to help us understand border-collision bifurcations involving two eigenvalues that are identical, or are nearly identical.

Acknowledgements

This work was supported by Marsden Fund contracts MAU1809 and MAU2209 managed by Royal Society Te Apārangi.

References

  • [1] V. Avrutin, B. Eckstein, and M. Schanz. On detection of multi-band chaotic attractors. Proc. R. Soc. A, 463(2081):1339–1358, 2007.
  • [2] S. Banerjee and C. Grebogi. Border collision bifurcations in two-dimensional piecewise smooth maps. Phys. Rev. E, 59(4):4052, 1999.
  • [3] S. Banerjee and G.C. Verghese, editors. Nonlinear Phenomena in Power Electronics. IEEE Press, New York, 2001.
  • [4] S. Banerjee, J.A. Yorke, and C. Grebogi. Robust chaos. Phys. Rev. Lett., 80(14):3049–3052, 1998.
  • [5] R.S. Burns. Advanced Control Engineering. Butterworth-Heinemann, Oxford, 2001.
  • [6] J. Buzzi. Absolutely continuous invariant measures for generic multi-dimensional piecewise affine expanding maps. Int. J. Bifurcation Chaos, 9(9):1743–1750, 1999.
  • [7] M. di Bernardo, C. Budd, A.R. Champneys, and P. Kowalczyk. Piecewise-smooth dynamical systems: theory and applications, volume 163. Springer-Verlag, New York, 2008.
  • [8] M. di Bernardo and C.K. Tse. Chaos in power electronics: An overview. Chaos in circuits and systems, pages 317–340, 2002.
  • [9] B. Eckstein. Bandcounter: counting bands of multiband chaotic attractors. Universitätsbibliothek der Universität Stuttgart, 2006.
  • [10] H.O. Fatoyinbo and D.J.W. Simpson. A synopsis of the non-invertible, two-dimensional, border-collision normal form with applications to power converters. Int. J. Bifurcation Chaos, 33(8):2330019, 2023.
  • [11] I. Ghosh, R.I. McLachlan, and D.J.W. Simpson. Robust chaos in orientation-reversing and non-invertible two-dimensional piecewise-linear maps. arXiv preprint arXiv:2307.05144, 2023.
  • [12] I. Ghosh and D. J. W. Simpson. Renormalization of the two-dimensional border-collision normal form. Int. J. Bifurcation Chaos, 32(12):2250181, 2022.
  • [13] I. Ghosh and D.J.W. Simpson. Robust Devaney chaos in the two-dimensional border-collision normal form. Chaos, 32(4):043120, 2022.
  • [14] P. A. Glendinning and D. J. W. Simpson. Robust chaos and the continuity of attractors. Trans. Math. Appl., 4(1):tnaa002, 2020.
  • [15] N. Haydn, Y. Lacroix, and S. Vaienti. Hitting and return times in ergodic dynamical systems. Ann. Probab., 33(5):2043–2050, 2005.
  • [16] B. Ibarz, J.M. Casado, and M.A.F. Sanjuán. Map-based models in neuronal dynamics. Phys. Rep., 501:1–74, 2011.
  • [17] S. Ito, S. Tanaka, and H. Nakada. On unimodal linear transformations and chaos. Proc. Japan Acad., 55, Ser. A(71):231–236, 1979.
  • [18] S. Ito, S. Tanaka, and H. Nakada. On unimodal linear transformations and chaos II. Tokyo J. Math., 2(2):241–259, 1979.
  • [19] L. Kocarev and S. Lian, editors. Chaos-Based Cryptography: Theory, Algorithms and Applications. Springer, New York, 2011.
  • [20] Y.A. Kuznetsov. Elements of applied bifurcation theory, volume 112. Springer, 2004.
  • [21] A. Lasota and J.A. Yorke. On the existence of invariant measures for piecewise monotonic transformations. Trans. Amer. Math. Soc., 186:481–488, 1973.
  • [22] T. Li and J.A. Yorke. Ergodic transformations from an interval into itself. Trans. Amer. Math. Soc., 235(1):183–192, 1978.
  • [23] D. Liberzon. Switching in Systems and Control. Birkhauser, Boston, 2003.
  • [24] Yu.L. Maistrenko, V.L. Maistrenko, and L.O. Chua. Cycles of chaotic intervals in a time-delayed Chua’s circuit. Int. J. Bifurcation Chaos., 3(6):1557–1572, 1993.
  • [25] M. Misiurewicz. Strange attractors for the Lozi mappings. Nonlinear dynamics, Annals of the New York Academy of Sciences, 357(1):348–358, 1980.
  • [26] H.E. Nusse and J.A. Yorke. Border-collision bifurcations including “period two to period three” for piecewise smooth systems. Phys. D: Nonlinear Phenomena, 57(1-2):39–57, 1992.
  • [27] H.E. Nusse and J.A. Yorke. Border-collision bifurcations for piecewise-smooth one-dimensional maps. Int. J. Bifurcation Chaos., 5(1):189–207, 1995.
  • [28] H. M. Osinga. Boundary crisis bifurcation in two parameters. J. Diff. Eq. Appl., 12(10):997–1008, 2006.
  • [29] D. Ou. Critical points in higher dimensions, I: Reverse order of periodic orbit creations in the Lozi family. arXiv preprint arXiv:2203.02326, 2022.
  • [30] Ya.B. Pesin. Dynamical systems with generalized hyperbolic attractors: hyperbolic, ergodic and topological properties. Ergod. Th. & Dynam. Sys., 12:123–151, 1992.
  • [31] A. Pumariño and E. Rodríguez, J. Á.and Vigil. Renormalization of two-dimensional piecewise linear maps: Abundance of 2-D strange attractors. Discrete Contin. Dyn. Syst., 38:941–966, 2018.
  • [32] A. Pumariño, J. Á. Rodríguez, and E. Vigil. Persistent two-dimensional strange attractors for a two-parameter family of expanding baker maps. Discrete Contin. Dyn. Syst. Ser. B, 24(2):657–670, 2019.
  • [33] T. Puu and I. Sushko. Business Cycle Dynamics: Models and Tools. Springer-Verlag, New York, 2006.
  • [34] V. Robins, J.D. Meiss, and E. Bradley. Computing connectedness: Disconnectedness and discreteness. Phys. D, 139(3–4):276–300, 2000.
  • [35] M. Rychlik. Invariant Measures and the Variational Principle for Lozi Mappings., pages 190–221. Springer, New York, 2004.
  • [36] D.J.W Simpson. Border-collision bifurcations in n\mathbb{R}^{n}. SIAM Rev., 58(2):177–226, 2016.
  • [37] D.J.W. Simpson. Unfolding homoclinic connections formed by corner intersections in piecewise-smooth maps. Chaos, 26(7):073105, 2016.
  • [38] D.J.W. Simpson and J.D. Meiss. Neimark-Sacker bifurcations in planar, piecewise-smooth, continuous maps. SIAM J. Appl. Dyn. Sys., 7(3):795–824, 2008.
  • [39] I. Sushko, V. Avrutin, and L. Gardini. Bifurcation structure in the skew tent map and its application as a border collision normal form. J. Diff. Eq. Appl., 22(8):1040–1087, 2016.
  • [40] I. Sushko and L. Gardini. Center bifurcation for two-dimensional border-collision normal form. Int. J. Bifurcation Chaos, 18(4):1029–1050, 2008.
  • [41] R. Szalai and H.M. Osinga. Arnol’d tongues arising from a grazing-sliding bifurcation. SIAM J. Appl. Dyn. Sys., 8(4):1434–1461, 2009.
  • [42] M. Tsujii. Absolutely continuous invariant measures for expanding piecewise linear maps. Invent. Math., 143:349–373, 2001.
  • [43] Ya.Z. Tsypkin. Relay Control Systems. Cambridge University Press, New York, 1984.
  • [44] D. Veitch and P. Glendinning. Explicit renormalisation in piecewise linear bimodal maps. Phys. D, 44(1-2):149–167, 1990.
  • [45] L.-S. Young. Bowen-Ruelle measures for certain piecewise hyperbolic maps. Trans. Amer. Math. Soc., 287(1):41–48, 1985.
  • [46] Z.T. Zhusubaliyev and E. Mosekilde. Bifurcations and Chaos in Piecewise-Smooth Dynamical Systems. World Scientific, Singapore, 2003.
  • [47] Z.T. Zhusubaliyev, E. Mosekilde, S. Maity, S. Mohanan, and S. Banerjee. Border collision route to quasiperiodicity: Numerical investigation and experimental confirmation. Chaos, 16(2):023122, 2006.