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

Unfolding globally resonant homoclinic tangencies.

Sishu Shankar Muni, Robert I. McLachlan, David J.W. Simpson

School of Fundamental Sciences
Massey University
Palmerston North
New Zealand
Abstract

Global resonance is a mechanism by which a homoclinic tangency of a smooth map can have infinitely many asymptotically stable, single-round periodic solutions. To understand the bifurcation structure one would expect to see near such a tangency, in this paper we study one-parameter perturbations of typical globally resonant homoclinic tangencies. We assume the tangencies are formed by the stable and unstable manifolds of saddle fixed points of two-dimensional maps. We show the perturbations display two infinite sequences of bifurcations, one saddle-node the other period-doubling, between which single-round periodic solutions are asymptotically stable. Generically these scale like |λ|2k|\lambda|^{2k}, as kk\to\infty, where 1<λ<1-1<\lambda<1 is the stable eigenvalue associated with the fixed point. If the perturbation is taken tangent to the surface of codimension-one homoclinic tangencies, they instead scale like |λ|kk\frac{|\lambda|^{k}}{k}. We also show slower scaling laws are possible if the perturbation admits further degeneracies.

1 Introduction

Homoclinic tangencies are perhaps the simplest mechanism in nonlinear dynamical systems for the loss of hyperbolicity and the creation of chaotic dynamics [12]. They occur most simply for saddle fixed points of two-dimensional maps. A tangential intersection between the stable and unstable manifolds of a fixed point is a homoclinic tangency, see Fig 1. This intersection is one point of an orbit that is homoclinic to the fixed point. This is a codimension-one phenomenon, meaning it can be equated to a single scalar condition. At a homoclinic tangency the map typically has infinitely many periodic solutions. Some of these are single-round, roughly meaning that they shadow the homoclinic orbit once before repeating, Fig. 1. Newhouse in [11] showed that infinitely many stable multi-round periodic solutions can coexist at a homoclinic tangency. At a generic homoclinic tangency all single-round periodic solutions of sufficiently large period are unstable [3, 4].

Refer to caption
Figure 1: A homoclinic tangency for a saddle fixed point of a two-dimensional map. In this illustration the eigenvalues associated with the fixed points are positive, i.e. 0<λ<10<\lambda<1 and σ>1\sigma>1. A coordinate change has been applied so that in the region 𝒩\mathcal{N} (shaded) the coordinate axes coincide with the stable and unstable manifolds. The homoclinic orbit ΓHC\Gamma_{\rm HC} is shown with black dots. A typical single-round periodic solution is shown with blue triangles.

In our previous work [10] we determined what degeneracies are needed for the infinitely many single-round periodic solutions to be asymptotically stable for smooth maps on 2\mathbb{R}^{2}. We found that the eigenvalues associated with the saddle fixed point, call them λ\lambda and σ\sigma satisfying 0<|λ|<1<|σ|0<|\lambda|<1<|\sigma|, need to multiply to 11 or 1-1. If λσ=1\lambda\sigma=-1 (so the map is orientation-reversing, at least locally), then the phenomenon is codimension-three. It involves a ‘global resonance’ condition on the reinjection mechanism for taking iterates back to a neighbourhood of the fixed point.

If a family of maps fμf_{\mu}, with μn\mu\in\mathbb{R}^{n}, has this phenomenon at some point μ\mu^{*} in parameter space, then for any positive jj\in\mathbb{Z} there exists an open set containing μ\mu^{*} in which jj asymptotically stable, single-round periodic solutions coexist. Due to the high codimension, a precise description of the shape of these sets (for large jj) is beyond the scope of this paper. The approach we take here is to consider one-parameter families that perturb from a globally resonant homoclinic tangency. Some information about the size and shapes of the sets can then be inferred from our results. Globally resonant homoclinic tangencies are hubs for extreme multi-stability. They should occur generically in some families of maps with three or more parameters, such as the generalised Hénon map [8, 9], but to our knowledge they have yet to have been identified.

We find that as the value of μ\mu is varied from μ\mu^{*}, there occurs infinite sequence of either saddle-node or period-doubling bifurcations that destroy the periodic solutions or make them unstable. Generically these sequences converge exponentially to μ\mu^{*} with the distance (in parameter space) to the bifurcation asymptotically proportional to |λ|2k|\lambda|^{2k}, where the periodic solutions have period k+mk+m, for some fixed m1m\geq 1. If we move away from μ\mu^{*} without a linear change to the codimension-one condition of a homoclinic tangency, the bifurcation values instead generically scale like |λ|kk\frac{|\lambda|^{k}}{k}. If the perturbation suffers further degeneracies, the scaling can be slower. Specifically we observe |λ|k|\lambda|^{k} and 1k\frac{1}{k} for an abstract example that we believe is representative of how the bifurcations scale in general.

Similar results have been obtained for more restrictive classes of maps. For area-preserving families the phenomenon is codimension-two and there exist infinitely many elliptic, single-round periodic solutions [7]. As shown in [1, 5] the periodic solutions are destroyed or lose stability in bifurcations that scale like |α|2k|\alpha|^{2k}, matching our result. For piecewise-linear families the phenomenon is codimension-three [15, 16]. In this setting the bifurcation values instead scale like |α|k|\alpha|^{k} [14], see also [2].

The phenomenon is also reminscient of a celebrated result of Newhouse [11]. Newhouse showed that for a generic (codimension-one) homoclinic tangency with |λσ|<1|\lambda\sigma|<1, infinitely many asymptotically stable periodic solutions coexist for a dense set of parameter values near that of the tangency. However, in this scenario the periodic solutions are multi-round (shadow the homoclinic orbit several times before repeating).

The remainder of the paper is organised as follows. In §2 we introduce two maps, T0T_{0} and T1T_{1}, to describe the dynamics near the homoclinic orbit. The map T0T_{0} is a local map describing the saddle fixed point, while T1T_{1} represents the reinjection mechanism. The single-round periodic solutions correspond to fixed points of T0kT1T_{0}^{k}\circ T_{1}. In §3 we summarise the necessary and sufficient conditions derived in [10] for the coexistence of infinitely many asymptotically stable, single-round periodic solutions. Then in §4–6 we introduce parameter-dependence and state our main result (Theorem 6.1) for the scaling properties of the saddle-node and period-doubling bifurcations. A proof of Theorem 6.1 is provided in §7. Next in §8 we illustrate Theorem 6.1 with a four-parameter family of C1C^{1} maps (an abstract minimal example). We observe numerically computed bifurcation values match those predicted by Theorem 6.1, to leading order, and observe slower scaling laws in special cases. Finally §9 provides a discussion and outlook for further studies.

2 A quantitative description of the dynamics near a homoclinic connection

Let ff be a CC^{\infty} map on 2\mathbb{R}^{2}. Suppose the origin (x,y)=(0,0)(x,y)=(0,0) is a saddle fixed point of ff. That is, Df(0,0){\rm D}f(0,0) has eigenvalues λ,σ\lambda,\sigma\in\mathbb{R} satisfying

0<|λ|<1<|σ|.0<|\lambda|<1<|\sigma|. (2.1)

By the results of Sternberg [13, 17] there exists a CC^{\infty} coordinate change that transforms ff to

T0(x,y)=[λx(1+𝒪(xy))σy(1+𝒪(xy))].T_{0}(x,y)=\begin{bmatrix}\lambda x\mathopen{}\mathclose{{}\left(1+\mathcal{O}\mathopen{}\mathclose{{}\left(xy}\right)}\right)\\ \sigma y\mathopen{}\mathclose{{}\left(1+\mathcal{O}\mathopen{}\mathclose{{}\left(xy}\right)}\right)\end{bmatrix}. (2.2)

In these new coordinates let 𝒩\mathcal{N} be a convex neighbourhood of the origin for which

f(x,y)=T0(x,y),for all (x,y)𝒩.f(x,y)=T_{0}(x,y),\qquad\text{for all $(x,y)\in\mathcal{N}$}. (2.3)

See Fig. 1. If λpσq1\lambda^{p}\sigma^{q}\neq 1 for all integers p,q1p,q\geq 1, then the eigenvalues are said to be non-resonant and the coordinate change can be chosen so that T0T_{0} is linear. If not, then T0T_{0} must contain resonant terms that cannot be eliminated by the coordinate change. As explained in §4, if λσ=1\lambda\sigma=1 we can reach the form

T0(x,y)=[λx(1+a1xy+𝒪(x2y2))σy(1+b1xy+𝒪(x2y2))],T_{0}(x,y)=\begin{bmatrix}\lambda x\mathopen{}\mathclose{{}\left(1+a_{1}xy+\mathcal{O}\mathopen{}\mathclose{{}\left(x^{2}y^{2}}\right)}\right)\\ \sigma y\mathopen{}\mathclose{{}\left(1+b_{1}xy+\mathcal{O}\mathopen{}\mathclose{{}\left(x^{2}y^{2}}\right)}\right)\end{bmatrix}, (2.4)

where a1,b1a_{1},b_{1}\in\mathbb{R}. If λσ=1\lambda\sigma=-1 we can obtain (2.4) with a1=b1=0a_{1}=b_{1}=0.

Now suppose there exists an orbit homoclinic to the origin, ΓHC\Gamma_{\rm HC}. By scaling xx and yy we may assume that (1,0)(1,0) and (0,1)(0,1) are points on ΓHC\Gamma_{\rm HC} and

(1,0),(λ,0),(0,1σ),(0,1σ2)𝒩,(1,0),(\lambda,0),\mathopen{}\mathclose{{}\left(0,\tfrac{1}{\sigma}}\right),\mathopen{}\mathclose{{}\left(0,\tfrac{1}{\sigma^{2}}}\right)\in\mathcal{N}, (2.5)

By assumption there exists m1m\geq 1 such that fm(0,1)=(1,0)f^{m}(0,1)=(1,0). We let T1=fmT_{1}=f^{m} and expand T1T_{1} in a Taylor series centred at (x,y)=(0,1)(x,y)=(0,1):

T1(x,y)=[c0+c1x+c2(y1)+𝒪((x,y1)2)d0+d1x+d2(y1)+d3x2+d4x(y1)+d5(y1)2+𝒪((x,y1)3)],T_{1}(x,y)=\begin{bmatrix}c_{0}+c_{1}x+c_{2}(y-1)+\mathcal{O}\mathopen{}\mathclose{{}\left((x,y-1)^{2}}\right)\\ d_{0}+d_{1}x+d_{2}(y-1)+d_{3}x^{2}+d_{4}x(y-1)+d_{5}(y-1)^{2}+\mathcal{O}\mathopen{}\mathclose{{}\left((x,y-1)^{3}}\right)\end{bmatrix}, (2.6)

where c0=1c_{0}=1 and d0=0d_{0}=0. In (2.6) we have written explicitly the terms that will be important below.

3 Conditions for infinitely many stable single-round solutions

In this section we state the main results of our previous work [10]. First Theorem 3.4 gives necessary conditions for the existence of infinitely many stable, single-round periodic solutions. Then Theorem 3.2 gives sufficient conditions for these to exist and be asymptotically stable.

Theorem 3.1.

Suppose ff satisfies (2.3) with (2.2) and (2.5). Suppose in (2.6) we have c0=1c_{0}=1, d0=0d_{0}=0, d50d_{5}\neq 0, and c1d2c2d10c_{1}d_{2}-c_{2}d_{1}\neq 0. Suppose ff has an infinite sequence of stable, single-round, periodic solutions accumulating on ΓHC\Gamma_{\rm HC}. Then

d2\displaystyle d_{2} =0,\displaystyle=0, (3.1)
|λσ|\displaystyle|\lambda\sigma| =1,\displaystyle=1, (3.2)
|d1|\displaystyle|d_{1}| =1,\displaystyle=1, (3.3)

with d1=1d_{1}=1 in the case λσ=1\lambda\sigma=1. Moreover, if λσ=1\lambda\sigma=1 and T0T_{0} has the form (2.4) then

a1+b1=0.a_{1}+b_{1}=0. (3.4)

The equation d2=0d_{2}=0 corresponds to a homoclinic tangency, as shown in Fig. 1. With |λσ|=1|\lambda\sigma|=1 we have |det(Df)|=1|\det({\rm D}f)|=1 at the origin. The condition d1=1d_{1}=1 is a global condition, see [10] for a geometric interpretation, with which the tangency is termed globally resonant. Finally if λσ=1\lambda\sigma=1 and T0T_{0} has the form (2.4), then det(Df)=1+2(a1+b1)xy+𝒪(x2y2)\det({\rm D}f)=1+2(a_{1}+b_{1})xy+\mathcal{O}\mathopen{}\mathclose{{}\left(x^{2}y^{2}}\right). Thus the condition a1+b1=0a_{1}+b_{1}=0 implies that as (x,y)(x,y) is varied from (0,0)(0,0), the value of det(Df)\det({\rm D}f) varies quadratically instead of linearly (as is generically the case). Now suppose |λσ|=1|\lambda\sigma|=1. Given kmink_{\rm min}\in\mathbb{Z}, let

K(kmin)={k|kkmin,(λσ)k=d1}.K(k_{\rm min})=\mathopen{}\mathclose{{}\left\{k\in\mathbb{Z}\,\big{|}\,k\geq k_{\rm min},\mathopen{}\mathclose{{}\left(\lambda\sigma}\right)^{k}=d_{1}}\right\}. (3.5)

If λσ=d1=1\lambda\sigma=d_{1}=1, then KK is the set of all integers greater than or equal to kmink_{\rm min}. If λσ=1\lambda\sigma=-1 and d1=1d_{1}=1 [resp. d1=1d_{1}=-1], then KK is the set of all even [resp. odd] integers greater than or equal to kmink_{\rm min}. Also let

Δ=(1c2d1d4)24d5(d3+c1d1).\Delta=\mathopen{}\mathclose{{}\left(1-c_{2}-d_{1}d_{4}}\right)^{2}-4d_{5}(d_{3}+c_{1}d_{1}). (3.6)
Theorem 3.2.

Suppose ff satisfies (2.3) with (2.4) and (2.5). Suppose c0=1c_{0}=1, d0=0d_{0}=0, and d50d_{5}\neq 0. Suppose (3.1)–(3.4) are satisfied, Δ>0\Delta>0, and

1<c2<1Δ2.-1<c_{2}<1-\frac{\sqrt{\Delta}}{2}. (3.7)

Then there exists kmink_{\rm min}\in\mathbb{Z} such that ff has an asymptotically stable period-(k+m)(k+m) solution for all kK(kmin)k\in K(k_{\rm min}).

Theorems 3.4 and 3.2 imply that the phenomenon of infinitely many asymptotically stable, single-round periodic solutions is codimension-three in the case λσ=1\lambda\sigma=-1. Specifically the three independent codimension-one conditions (3.1)–(3.3) need to hold, and the phenomenon indeed occurs if Δ>0\Delta>0 and (3.7) holds. In the case λσ=1\lambda\sigma=1 the phenomenon is codimension-four as we also require (3.4). This condition is absent when λσ=1\lambda\sigma=-1 because in this case the cubic terms in (2.4) are removable.

4 Smooth parameter dependence

Now suppose fμf_{\mu} is a CC^{\infty} map on 2\mathbb{R}^{2} with a CC^{\infty} dependence on a parameter μn\mu\in\mathbb{R}^{n}. Let 𝟎n{\bf 0}\in\mathbb{R}^{n} denote the origin in parameter space. Suppose that for all μ\mu in some region containing 𝟎{\bf 0}, the origin in phase space (x,y)=(0,0)(x,y)=(0,0) is a fixed point of fμf_{\mu}. Let λ=λ(μ)\lambda=\lambda(\mu) and σ=σ(μ)\sigma=\sigma(\mu) be its associated eigenvalues (these are CC^{\infty} functions of μ\mu) and suppose

λ(𝟎)\displaystyle\lambda({\bf 0}) =α,\displaystyle=\alpha, (4.1)
σ(𝟎)\displaystyle\sigma({\bf 0}) =χeigα,\displaystyle=\frac{\chi_{\rm eig}}{\alpha}, (4.2)

with 0<α<10<\alpha<1 and χeig{1,1}\chi_{\rm eig}\in\{-1,1\}. With μ=𝟎\mu={\bf 0} we have |λσ|=1|\lambda\sigma|=1, so as described above T0T_{0} can be assumed to have the form (2.4). We now show we can assume T0T_{0} has this form when the value of μ\mu is sufficiently small.

Lemma 4.1.

There exists a neighbourhood 𝒩paramn\mathcal{N}_{\rm param}\subset\mathbb{R}^{n} of 𝟎{\bf 0} and a CC^{\infty} coordinate change that puts fμf_{\mu} in the form (2.4) for all μ𝒩param\mu\in\mathcal{N}_{\rm param}.

Proof.

Via a linear transformation fμf_{\mu} can be transformed to

[xy][λx+i0,j0,i+j2aijxiyjσy+i0,j0,i+j2bijxiyj],\begin{bmatrix}x\\ y\end{bmatrix}\mapsto\begin{bmatrix}\lambda x+\sum\limits_{i\geq 0,\,j\geq 0,\,i+j\geq 2}a_{ij}x^{i}y^{j}\\ \sigma y+\sum\limits_{i\geq 0,\,j\geq 0,\,i+j\geq 2}b_{ij}x^{i}y^{j}\end{bmatrix},

for some aij,bija_{ij},b_{ij}\in\mathbb{R}. It is a standard asymptotic matching exercise to show that via an additional CC^{\infty} coordinate change we can achieve aij=0a_{ij}=0 if λi1σj1\lambda^{i-1}\sigma^{j}\neq 1, and bij=0b_{ij}=0 if λiσj11\lambda^{i}\sigma^{j-1}\neq 1. The remainder of the proof is based on this fact.

Assume the value of μ\mu is small enough that (2.1) is satisfied. Then λpσq=1\lambda^{p}\sigma^{q}=1 is only possible with p,q1p,q\geq 1, so a CC^{\infty} coordinate change can be performed to reduce the map to

[xy][λx+i2,j1aijxiyjσy+i1,j2bijxiyj].\begin{bmatrix}x\\ y\end{bmatrix}\mapsto\begin{bmatrix}\lambda x+\sum\limits_{i\geq 2,\,j\geq 1}a_{ij}x^{i}y^{j}\\ \sigma y+\sum\limits_{i\geq 1,\,j\geq 2}b_{ij}x^{i}y^{j}\end{bmatrix}.

Since |λσ|=1|\lambda\sigma|=1 when μ=𝟎\mu={\bf 0} we can assume μ\mu is small enough that λp1σ1\lambda^{p-1}\sigma\neq 1 for all p3p\geq 3 and λσq11\lambda\sigma^{q-1}\neq 1 for all q3q\geq 3. Consequently the map can further be reduced to

[xy][λx+a21x2y+i3,j2aijxiyjσy+b12xy2+i2,j3bijxiyj],\begin{bmatrix}x\\ y\end{bmatrix}\mapsto\begin{bmatrix}\lambda x+a_{21}x^{2}y+\sum\limits_{i\geq 3,\,j\geq 2}a_{ij}x^{i}y^{j}\\ \sigma y+b_{12}xy^{2}+\sum\limits_{i\geq 2,\,j\geq 3}b_{ij}x^{i}y^{j}\end{bmatrix},

which can be rewritten as (2.4). ∎

The product of the eigenvalues is

λ(μ)σ(μ)=χeig+𝐧eig𝖳μ+𝒪(μ2),\lambda(\mu)\sigma(\mu)=\chi_{\rm eig}+{\bf n}_{\rm eig}^{\sf T}\mu+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right), (4.3)

where 𝐧eig{\bf n}_{\rm eig} is the gradient of λσ\lambda\sigma evaluated at μ=𝟎\mu={\bf 0}. The following result is an elementary application of the implicit function theorem.

Lemma 4.2.

Suppose 𝐧eig𝟎{\bf n}_{\rm eig}\neq{\bf 0}. Then |λ(μ)σ(μ)|=1|\lambda(\mu)\sigma(\mu)|=1 on a CC^{\infty} codimension-one surface intersecting μ=𝟎\mu={\bf 0} and with normal vector 𝐧eig{\bf n}_{\rm eig} at μ=𝟎\mu={\bf 0} (as illustrated in Fig. 2).

Refer to caption
Figure 2: A sketch of codimension-one surfaces of homoclinic tangencies (green) and where λ(μ)σ(μ)=1\lambda(\mu)\sigma(\mu)=1 (purple). The vectors 𝐧tang{\bf n}_{\rm tang} and 𝐧eig{\bf n}_{\rm eig}, respectively, are normal to these surfaces at the origin μ=𝟎\mu={\bf 0}.

5 The codimension-one surface of homoclinic tangencies

In this section we describe the codimension-one surface of homoclinic tangencies that intersects μ=𝟎\mu={\bf 0} where we will be assuming that a globally resonant homoclinic tangency occurs.

Suppose (2.5) is satisfied when μ=𝟎\mu={\bf 0}. Write fμmf_{\mu}^{m} as (2.6) and suppose

c0(𝟎)\displaystyle c_{0}({\bf 0}) =1,\displaystyle=1, (5.1)
d0(𝟎)\displaystyle d_{0}({\bf 0}) =0,\displaystyle=0, (5.2)

so that f𝟎f_{\bf 0} has an orbit homoclinic to the origin through (x,y)=(1,0)(x,y)=(1,0) and (0,1)(0,1). Also suppose

d2(𝟎)\displaystyle d_{2}({\bf 0}) =0,\displaystyle=0, (5.3)
d5(𝟎)\displaystyle d_{5}({\bf 0}) =d5,00,\displaystyle=d_{5,0}\neq 0, (5.4)

for a quadratic tangency. Also write

d0(μ)=𝐧tang𝖳μ+𝒪(μ2).d_{0}(\mu)={\bf n}_{\rm tang}^{\sf T}\mu+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right). (5.5)
Lemma 5.1.

Suppose (2.5) is satisfied when μ=𝟎\mu={\bf 0}, (5.1)–(5.4) are satisfied, and 𝐧tang𝟎{\bf n}_{\rm tang}\neq{\bf 0}. Then fμf_{\mu} has a quadratic homoclinic tangency to (x,y)=(0,0)(x,y)=(0,0) on a CC^{\infty} codimension-one surface intersecting μ=𝟎\mu={\bf 0} and with normal vector 𝐧tang{\bf n}_{\rm tang} at μ=𝟎\mu={\bf 0}. Moreover, this tangency occurs at (x,y)=(0,Y(μ))(x,y)=(0,Y(\mu)) where YY is a CC^{\infty} function with Y(𝟎)=1Y({\bf 0})=1.

Proof.

Let T1,2T_{1,2} denote the second component of T1T_{1} (2.6). The function

h(u;μ)=T1,2y(0,1+u)=d2(μ)+2d5(μ)u+𝒪(2),h(u;\mu)=\frac{\partial T_{1,2}}{\partial y}(0,1+u)=d_{2}(\mu)+2d_{5}(\mu)u+\mathcal{O}(2),

is a CC^{\infty} function of uu\in\mathbb{R} and μ\mu. Since h(0;𝟎)=0h(0;\mathbf{0})=0 by (5.3) and hu(0;𝟎)0\frac{\partial h}{\partial u}(0;\mathbf{0})\neq 0 by (5.4), the implicit function theorem implies there exists a CC^{\infty} function utang(μ)u_{\rm tang}(\mu) such that h(utang(μ);μ)=0h(u_{\rm tang}(\mu);\mu)=0 locally.

By construction, the unstable manifold of (x,y)=(0,0)(x,y)=(0,0) is tangent to the xx-axis at T1(0,1+utang)T_{1}(0,1+u_{\rm tang}). Moreover this tangency is quadratic by (5.4). Thus a homoclinic tangency occurs if T1,2(0,1+utang)=d0+d2utang+d5utang2+𝒪(3)=0T_{1,2}(0,1+u_{\rm tang})=d_{0}+d_{2}u_{\rm tang}+d_{5}u_{\rm tang}^{2}+\mathcal{O}(3)=0. This function is CC^{\infty} and

T1,2(0,1+utang)=𝐧tang𝖳μ+𝒪(μ2).T_{1,2}(0,1+u_{\rm tang})={\bf n}_{\rm tang}^{\sf T}\mu+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right).

Since 𝐧tang𝟎{\bf n}_{\rm tang}\neq{\bf 0} the result follows from another application of the implicit function theorem. ∎

6 Sequences of saddle-node and period-doubling bifurcations

Suppose

d1(𝟎)=χ,d_{1}({\bf 0})=\chi, (6.1)

where χ{1,1}\chi\in\{-1,1\}. Write

a1(𝟎)\displaystyle a_{1}({\bf 0}) =a1,0,\displaystyle=a_{1,0}\,, c1(𝟎)\displaystyle\qquad c_{1}({\bf 0}) =c1,0,\displaystyle=c_{1,0}\,, d3(𝟎)\displaystyle\qquad d_{3}({\bf 0}) =d3,0,\displaystyle=d_{3,0}\,, (6.2)
b1(𝟎)\displaystyle b_{1}({\bf 0}) =b1,0,\displaystyle=b_{1,0}\,, c2(𝟎)\displaystyle\qquad c_{2}({\bf 0}) =c2,0,\displaystyle=c_{2,0}\,, d4(𝟎)\displaystyle\qquad d_{4}({\bf 0}) =d4,0,\displaystyle=d_{4,0}\,,

and, recalling (3.6), let

Δ0=(1c2,0χd4,0)24d5,0(d3,0+χc1,0).\Delta_{0}=\mathopen{}\mathclose{{}\left(1-c_{2,0}-\chi d_{4,0}}\right)^{2}-4d_{5,0}(d_{3,0}+\chi c_{1,0}). (6.3)
Theorem 6.1.

Suppose fμf_{\mu} satisfies (4.1), (4.2), (5.1)–(5.4), a1,0+b1,0=0a_{1,0}+b_{1,0}=0, Δ0>0\Delta_{0}>0, and 1<c2,0<1Δ02-1<c_{2,0}<1-\frac{\sqrt{\Delta_{0}}}{2}. Let 𝐯n{\bf v}\in\mathbb{R}^{n}. If 𝐧tang𝖳𝐯0{\bf n}_{\rm tang}^{\sf T}{\bf v}\neq 0 then there exists kmink_{\rm min}\in\mathbb{Z} such that for all kK(kmin)k\in K(k_{\rm min}) there exist εk<0\varepsilon_{k}^{-}<0 and εk+>0\varepsilon_{k}^{+}>0 with εk±=𝒪(α2k)\varepsilon_{k}^{\pm}=\mathcal{O}\mathopen{}\mathclose{{}\left(\alpha^{2k}}\right) such that ff has an asymptotically stable period-(k+m)(k+m) solution for all μ=ε𝐯\mu=\varepsilon{\bf v} with εk<ε<εk+\varepsilon_{k}^{-}<\varepsilon<\varepsilon_{k}^{+}. Moreover, one sequence, {εk}\{\varepsilon_{k}^{-}\} or {εk+}\{\varepsilon_{k}^{+}\}, corresponds to saddle-node bifurcations of the periodic solutions, the other to period-doubling bifurcations. If instead 𝐧tang𝖳𝐯=0{\bf n}_{\rm tang}^{\sf T}{\bf v}=0 and 𝐧eig𝖳𝐯0{\bf n}_{\rm eig}^{\sf T}{\bf v}\neq 0 the same results hold now εk±=𝒪(αkk)\varepsilon_{k}^{\pm}=\mathcal{O}\mathopen{}\mathclose{{}\left(\frac{\alpha^{k}}{k}}\right).

7 Proof of the main result

To prove Theorem 6.1 we use the following lemma which is proved in Appendix A by carefully estimating the error terms in (7.1).. If χeig=1\chi_{\rm eig}=-1 then (7.1) is true if a1=b1=0a_{1}=b_{1}=0 (and this can be proved in the same fashion).

Lemma 7.1.

Suppose χeig=1\chi_{\rm eig}=1 and μ=𝒪(αk)\mu=\mathcal{O}\mathopen{}\mathclose{{}\left(\alpha^{k}}\right). If |x1||x-1| and |αky1|\mathopen{}\mathclose{{}\left|\alpha^{-k}y-1}\right| are sufficiently small for all sufficiently large values of kk, then

T0k(x,y)=[λkx(1+ka1xy+𝒪(k2α2k))σky(1+kb1xy+𝒪(k2α2k))].T_{0}^{k}(x,y)=\begin{bmatrix}\lambda^{k}x\mathopen{}\mathclose{{}\left(1+ka_{1}xy+\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}\alpha^{2k}}\right)}\right)\\ \sigma^{k}y\mathopen{}\mathclose{{}\left(1+kb_{1}xy+\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}\alpha^{2k}}\right)}\right)\end{bmatrix}. (7.1)
Proof of Theorem 6.1.

Step 1 — Coordinate changes to distinguish the surface of homoclinic tangencies.
First we perform two coordinate changes on parameter space. There exists an n×nn\times n orthogonal matrix AA such that after μ\mu is replaced with AμA\mu, and μ\mu is scaled, we have 𝐧tang𝖳=[1,0,,0]{\bf n}_{\rm tang}^{\sf T}=[1,0,\ldots,0] — the first coordinate vector of n\mathbb{R}^{n}. Then d0(μ)=μ1+𝒪(μ2)d_{0}(\mu)=\mu_{1}+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right). Second, for convenience, we apply a near-identity transformation to remove the higher order terms, resulting in

d0(μ)=μ1.d_{0}(\mu)=\mu_{1}\,. (7.2)

These coordinate changes do not alter the signs of the dot products 𝐧tang𝖳𝐯{\bf n}_{\rm tang}^{\sf T}{\bf v} and 𝐧eig𝖳𝐯{\bf n}_{\rm eig}^{\sf T}{\bf v}. Now write

c0(μ)\displaystyle c_{0}(\mu) =1+i=1npiμi+𝒪(μ2),\displaystyle=1+\sum_{i=1}^{n}p_{i}\mu_{i}+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right), (7.3)
d1(μ)\displaystyle d_{1}(\mu) =χ+i=1nqiμi+𝒪(μ2),\displaystyle=\chi+\sum_{i=1}^{n}q_{i}\mu_{i}+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right), (7.4)
d2(μ)\displaystyle d_{2}(\mu) =i=1nriμi+𝒪(μ2),\displaystyle=\sum_{i=1}^{n}r_{i}\mu_{i}+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right), (7.5)
λ(μ)\displaystyle\lambda(\mu) =α+i=1nsiμi+𝒪(μ2),\displaystyle=\alpha+\sum_{i=1}^{n}s_{i}\mu_{i}+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right), (7.6)
σ(μ)\displaystyle\sigma(\mu) =χeigα+i=1ntiμi+𝒪(μ2),\displaystyle=\frac{\chi_{\rm eig}}{\alpha}+\sum_{i=1}^{n}t_{i}\mu_{i}+\mathcal{O}\mathopen{}\mathclose{{}\left(\|\mu\|^{2}}\right), (7.7)

where pi,,tip_{i},\ldots,t_{i}\in\mathbb{R} are constants. Step 2 — Apply a kk-dependent scaling to μ\mu.
In view of the coordinate changes applied in the previous step, the surface of homoclinic surfaces of Lemma 5.1 is tangent to the μ1=0\mu_{1}=0 coordinate hyperplane. In order to show that bifurcation values scale like |α|2k|\alpha|^{2k} if we adjust the value of μ\mu in a direction transverse to this surface, and, generically, scale like |α|kk\frac{|\alpha|^{k}}{k} otherwise, we scale the components of μ\mu as follows:

μi={α2kμ~i,i=1,αkμ~i,i1.\mu_{i}=\begin{cases}\alpha^{2k}\tilde{\mu}_{i}\,,&i=1,\\ \alpha^{k}\tilde{\mu}_{i}\,,&i\neq 1.\end{cases} (7.8)

Below we will see that the resulting asymptotic expansions are consistent and this will justify (7.8). Notice that μ~1\tilde{\mu}_{1}-terms are higher order than μ~i\tilde{\mu}_{i}-terms, for i1i\neq 1. For example (7.6) now becomes λ=α+i=2nsiμ~iαk+𝒪(|α|2k)\lambda=\alpha+\sum_{i=2}^{n}s_{i}\tilde{\mu}_{i}\alpha^{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(|\alpha|^{2k}}\right). Further, let kk be such that λ(𝟎)kσ(𝟎)k=d1(𝟎)\lambda({\bf 0})^{k}\sigma({\bf 0})^{k}=d_{1}({\bf 0}), that is χeigk=χ\chi_{\rm eig}^{k}=\chi. Then from (7.6)–(7.8) we obtain

λk\displaystyle\lambda^{k} =αk(1+kαi=2nsiμ~iαk+𝒪(k2|α|2k)),\displaystyle=\alpha^{k}\mathopen{}\mathclose{{}\left(1+\frac{k}{\alpha}\sum_{i=2}^{n}s_{i}\tilde{\mu}_{i}\alpha^{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}|\alpha|^{2k}}\right)}\right), (7.9)
σk\displaystyle\sigma^{k} =χαk(1+kαχeigi=2ntiμ~iαk+𝒪(k2|α|2k)).\displaystyle=\frac{\chi}{\alpha^{k}}\mathopen{}\mathclose{{}\left(1+k\alpha\chi_{\rm eig}\sum_{i=2}^{n}t_{i}\tilde{\mu}_{i}\alpha^{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}|\alpha|^{2k}}\right)}\right). (7.10)

Step 3 — Calculate one point of each periodic solution.
One point of a single-round periodic solution is a fixed point of T0kT1T_{0}^{k}\circ T_{1}. We look for fixed points (x(k),y(k))\mathopen{}\mathclose{{}\left(x^{(k)},y^{(k)}}\right) of T0kT1T_{0}^{k}\circ T_{1} of the form

x(k)=αk(1+ϕkαk+𝒪(k2|α|2k)),y(k)=1+ψkαk+𝒪(k2|α|2k).\begin{split}x^{(k)}&=\alpha^{k}\mathopen{}\mathclose{{}\left(1+\phi_{k}\alpha^{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}|\alpha|^{2k}}\right)}\right),\\ y^{(k)}&=1+\psi_{k}\alpha^{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}|\alpha|^{2k}}\right).\end{split} (7.12)

By substituting (7.12) into (2.6) and the above various asymptotic expressions for the coefficients in T1T_{1}, we obtain

T1(x(k),y(k))=[1+(c1,0+c2,0ψk+i=2npiμ~i)αk+𝒪(k2|α|2k)χαk+(μ~1+d3,0+d4,0ψk+d5,0ψk2+χϕk+i=2n(qi+riψk)μ~i)α2k+𝒪(k2|α|3k)].T_{1}\mathopen{}\mathclose{{}\left(x^{(k)},y^{(k)}}\right)=\begin{bmatrix}1+\mathopen{}\mathclose{{}\left(c_{1,0}+c_{2,0}\psi_{k}+\sum_{i=2}^{n}p_{i}\tilde{\mu}_{i}}\right)\alpha^{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}|\alpha|^{2k}}\right)\\ \begin{aligned} \chi\alpha^{k}+\mathopen{}\mathclose{{}\left(\tilde{\mu}_{1}+d_{3,0}+d_{4,0}\psi_{k}+d_{5,0}\psi_{k}^{2}+\chi\phi_{k}+\sum_{i=2}^{n}\mathopen{}\mathclose{{}\left(q_{i}+r_{i}\psi_{k}}\right)\tilde{\mu}_{i}}\right)\alpha^{2k}&\\[5.69054pt] +\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}|\alpha|^{3k}}\right)&\end{aligned}\end{bmatrix}.

Then by (7.1),

(T0kT1)(x(k),y(k))=[αk+(a1,0χ+1αi=2nsiμ~i)kα2k+(c1,0+c2,0ψk+i=2npiμ~i)α2k+𝒪(k2|α|3k)1+(b1,0χ+αχeigi=2ntiμ~i)kαk+(χμ~1+χd3,0+χd4,0ψk+χd5,0ψk2+ϕk+χi=2n(qi+riψk)μ~i)αk+𝒪(k2|α|2k)].\mathopen{}\mathclose{{}\left(T_{0}^{k}\circ T_{1}}\right)\mathopen{}\mathclose{{}\left(x^{(k)},y^{(k)}}\right)=\begin{bmatrix}\begin{aligned} \alpha^{k}+\bigg{(}a_{1,0}\chi+\frac{1}{\alpha}\sum_{i=2}^{n}s_{i}\tilde{\mu}_{i}\bigg{)}k\alpha^{2k}+\bigg{(}c_{1,0}+c_{2,0}\psi_{k}+\sum_{i=2}^{n}p_{i}\tilde{\mu}_{i}\bigg{)}\alpha^{2k}&\\[5.69054pt] +\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}|\alpha|^{3k}}\right)&\end{aligned}\\ \begin{aligned} 1+\bigg{(}b_{1,0}\chi+\alpha\chi_{\rm eig}\sum_{i=2}^{n}t_{i}\tilde{\mu}_{i}\bigg{)}k\alpha^{k}+\bigg{(}\chi\tilde{\mu}_{1}+\chi d_{3,0}+\chi d_{4,0}\psi_{k}&\\ +\chi d_{5,0}\psi_{k}^{2}+\phi_{k}+\chi\sum_{i=2}^{n}\mathopen{}\mathclose{{}\left(q_{i}+r_{i}\psi_{k}}\right)\tilde{\mu}_{i}\bigg{)}\alpha^{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(k^{2}|\alpha|^{2k}}\right)&\end{aligned}\end{bmatrix}. (7.13)

By matching (7.12) and (7.13) and eliminating ϕk\phi_{k} we obtain the following expression that determines the possible values of ψk\psi_{k}:

χd5,0ψk2Pψk+Q=0,\chi d_{5,0}\psi_{k}^{2}-P\psi_{k}+Q=0, (7.14)

where

P\displaystyle P =1c2,0χd4,0χi=2nriμ~i,\displaystyle=1-c_{2,0}-\chi d_{4,0}-\chi\sum_{i=2}^{n}r_{i}\tilde{\mu}_{i}\,, (7.15)
Q\displaystyle Q =χμ~1+c1,0+χd3,0+i=2n(pi+χqi)μ~i+i=2n(siα+αχeigti)μ~ik,\displaystyle=\chi\tilde{\mu}_{1}+c_{1,0}+\chi d_{3,0}+\sum_{i=2}^{n}\mathopen{}\mathclose{{}\left(p_{i}+\chi q_{i}}\right)\tilde{\mu}_{i}+\sum_{i=2}^{n}\mathopen{}\mathclose{{}\left(\frac{s_{i}}{\alpha}+\alpha\chi_{\rm eig}t_{i}}\right)\tilde{\mu}_{i}k, (7.16)

and we have also used a1,0+b1,0=0a_{1,0}+b_{1,0}=0. Of the two solutions to (7.14), the one that yields an asymptotically stable solution when μ=𝟎\mu={\bf 0} is

ψk=12χd5,0(PP24χd5,0Q).\psi_{k}=\frac{1}{2\chi d_{5,0}}\mathopen{}\mathclose{{}\left(P-\sqrt{P^{2}-4\chi d_{5,0}Q}}\right). (7.17)

Note that this solution exists and is real-valued for sufficiently small values of μ\mu because when μ=𝟎\mu={\bf 0} the discriminant is P24χd5,0Q=Δ0>0P^{2}-4\chi d_{5,0}Q=\Delta_{0}>0.

Step 4 — Stability of the periodic solution.
By using (2.6), (7.1), (7.9), (7.10), and (7.12),

D(T0kT1)(x(k),y(k))=[𝒪(|α|k)c2,0αk+𝒪(k|α|2k)1αk(1+𝒪(k|α|k))χd4,0+χi=2nriμ~i+2χd5,0ψk+𝒪(k|α|k)].{\rm D}\mathopen{}\mathclose{{}\left(T_{0}^{k}\circ T_{1}}\right)\mathopen{}\mathclose{{}\left(x^{(k)},y^{(k)}}\right)=\begin{bmatrix}\mathcal{O}\mathopen{}\mathclose{{}\left(|\alpha|^{k}}\right)&c_{2,0}\alpha^{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(k|\alpha|^{2k}}\right)\\ \frac{1}{\alpha^{k}}\mathopen{}\mathclose{{}\left(1+\mathcal{O}\mathopen{}\mathclose{{}\left(k|\alpha|^{k}}\right)}\right)&\chi d_{4,0}+\chi\sum_{i=2}^{n}r_{i}\tilde{\mu}_{i}+2\chi d_{5,0}\psi_{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(k|\alpha|^{k}}\right)\end{bmatrix}. (7.18)

Let τk\tau_{k} and δk\delta_{k} denote the trace and determinant of this matrix, respectively. By (7.15), (7.17), and (7.18) we obtain

τk\displaystyle\tau_{k} =1c2,0P24χd5,0Q+𝒪(k|α|k),\displaystyle=1-c_{2,0}-\sqrt{P^{2}-4\chi d_{5,0}Q}+\mathcal{O}\mathopen{}\mathclose{{}\left(k|\alpha|^{k}}\right), (7.19)
δk\displaystyle\delta_{k} =c2,0+𝒪(k|α|k).\displaystyle=-c_{2,0}+\mathcal{O}\mathopen{}\mathclose{{}\left(k|\alpha|^{k}}\right). (7.20)

By substituting μ=𝟎\mu={\bf 0} into (7.19) we obtain τk=1c2,0Δ0\tau_{k}=1-c_{2,0}-\sqrt{\Delta_{0}}. It immediately follows from the assumption 1<c2,0<1Δ02-1<c_{2,0}<1-\frac{\sqrt{\Delta_{0}}}{2} that the periodic solution is asymptotically stable for sufficiently large values of kk.

Step 5 — The generic case 𝐧tang𝖳𝐯0{\bf n}_{\rm tang}^{\sf T}{\bf v}\neq 0.
Now suppose 𝐧tang𝖳𝐯0{\bf n}_{\rm tang}^{\sf T}{\bf v}\neq 0, that is, v10v_{1}\neq 0 (in view of the earlier coordinate change). Write μ=ε𝐯\mu=\varepsilon{\bf v} and ε=ε~α2k\varepsilon=\tilde{\varepsilon}\alpha^{2k}. By (7.8), μ~1=ε~v1\tilde{\mu}_{1}=\tilde{\varepsilon}v_{1} and μ~i=ε~viαk\tilde{\mu}_{i}=\tilde{\varepsilon}v_{i}\alpha^{k} for i1i\neq 1. Then by (7.15) and (7.16), P=1c2,0χd4,0+𝒪(|α|k)P=1-c_{2,0}-\chi d_{4,0}+\mathcal{O}\mathopen{}\mathclose{{}\left(|\alpha|^{k}}\right) and Q=c1,0+χd3,0+χε~v1+𝒪(|α|k)Q=c_{1,0}+\chi d_{3,0}+\chi\tilde{\varepsilon}v_{1}+\mathcal{O}\mathopen{}\mathclose{{}\left(|\alpha|^{k}}\right), so

P24χd5,0Q=(1c2,0χd4,0)24χd5,0(c1,0+χd3,0+χε~v1)+𝒪(|α|k).P^{2}-4\chi d_{5,0}Q=\mathopen{}\mathclose{{}\left(1-c_{2,0}-\chi d_{4,0}}\right)^{2}-4\chi d_{5,0}\mathopen{}\mathclose{{}\left(c_{1,0}+\chi d_{3,0}+\chi\tilde{\varepsilon}v_{1}}\right)+\mathcal{O}\mathopen{}\mathclose{{}\left(|\alpha|^{k}}\right). (7.21)

Since v10v_{1}\neq 0 and d5,00d_{5,0}\neq 0 we can solve δkτk+1=0\delta_{k}-\tau_{k}+1=0 for ε~\tilde{\varepsilon} (formally this is achieved via the implicit function theorem) and the solution is

ε~SN=Δ04d5,0v1+𝒪(k|α|k).\tilde{\varepsilon}_{\rm SN}=\frac{\Delta_{0}}{4d_{5,0}v_{1}}+\mathcal{O}\mathopen{}\mathclose{{}\left(k|\alpha|^{k}}\right). (7.22)

Also we can use (7.21) to solve δk+τk+1=0\delta_{k}+\tau_{k}+1=0 for ε~\tilde{\varepsilon}:

ε~PD=Δ04(1c2,0)24d5,0v1+𝒪(k|α|k).\tilde{\varepsilon}_{\rm PD}=\frac{\Delta_{0}-4\mathopen{}\mathclose{{}\left(1-c_{2,0}}\right)^{2}}{4d_{5,0}v_{1}}+\mathcal{O}\mathopen{}\mathclose{{}\left(k|\alpha|^{k}}\right). (7.23)

Since ε~SN\tilde{\varepsilon}_{\rm SN} and ε~PD\tilde{\varepsilon}_{\rm PD} evidently have opposite signs, this completes the proof in the first case.

Step 6 — The degenerate case 𝐧tang𝖳𝐯=0{\bf n}_{\rm tang}^{\sf T}{\bf v}=0.
Now suppose v1=0v_{1}=0 and 𝐧eig𝖳𝐯0{\bf n}_{\rm eig}^{\sf T}{\bf v}\neq 0. Again write μ=ε𝐯\mu=\varepsilon{\bf v} but now write ε=ε~αkk\varepsilon=\frac{\tilde{\varepsilon}\alpha^{k}}{k}. By (7.8), μ~1=0\tilde{\mu}_{1}=0 and μ~i=ε~vik\tilde{\mu}_{i}=\frac{\tilde{\varepsilon}v_{i}}{k} for i1i\neq 1.

We first evaluate 𝐧eig𝖳𝐯{\bf n}_{\rm eig}^{\sf T}{\bf v}. By multiplying (7.6) and (7.7) and comparing the result to (4.3) we see that the ithi^{\rm th} element of 𝐧eig{\bf n}_{\rm eig} is χeigsiα+αti\frac{\chi_{\rm eig}s_{i}}{\alpha}+\alpha t_{i}. Then since v1=0v_{1}=0 we have 𝐧eig𝖳𝐯=i=2n(χeigsiα+αti)vi{\bf n}_{\rm eig}^{\sf T}{\bf v}=\sum_{i=2}^{n}\mathopen{}\mathclose{{}\left(\frac{\chi_{\rm eig}s_{i}}{\alpha}+\alpha t_{i}}\right)v_{i}. Further, μ=ε𝐯\mu=\varepsilon{\bf v} and ε=ε~αkk\varepsilon=\frac{\tilde{\varepsilon}\alpha^{k}}{k},

χeigε~𝐧eig𝖳𝐯=i=2n(siα+αχeigti)μ~ik,\chi_{\rm eig}\tilde{\varepsilon}{\bf n}_{\rm eig}^{\sf T}{\bf v}=\sum_{i=2}^{n}\mathopen{}\mathclose{{}\left(\frac{s_{i}}{\alpha}+\alpha\chi_{\rm eig}t_{i}}\right)\tilde{\mu}_{i}k,

which is a term appearing in (7.16). So by (7.15) and (7.16), P=1c2,0χd4,0+𝒪(1k)P=1-c_{2,0}-\chi d_{4,0}+\mathcal{O}\mathopen{}\mathclose{{}\left(\frac{1}{k}}\right) and Q=c1,0+χd3,0+ε~χeig𝐧eig𝖳𝐯k+𝒪(1k)Q=c_{1,0}+\chi d_{3,0}+\frac{\tilde{\varepsilon}\chi_{\rm eig}{\bf n}_{\rm eig}^{\sf T}{\bf v}}{k}+\mathcal{O}\mathopen{}\mathclose{{}\left(\frac{1}{k}}\right). By solving δkτk+1=0\delta_{k}-\tau_{k}+1=0,

ε~SN=Δ04d5,0χχeig𝐧eig𝖳𝐯+𝒪(1k),\tilde{\varepsilon}_{\rm SN}=\frac{\Delta_{0}}{4d_{5,0}\chi\chi_{\rm eig}{\bf n}_{\rm eig}^{\sf T}{\bf v}}+\mathcal{O}\mathopen{}\mathclose{{}\left(\frac{1}{k}}\right), (7.24)

and by solving δkτk+1=0\delta_{k}-\tau_{k}+1=0,

ε~PD=Δ04(1c2,0)24d5,0χχeig𝐧eig𝖳𝐯+𝒪(1k).\tilde{\varepsilon}_{\rm PD}=\frac{\Delta_{0}-4\mathopen{}\mathclose{{}\left(1-c_{2,0}}\right)^{2}}{4d_{5,0}\chi\chi_{\rm eig}{\bf n}_{\rm eig}^{\sf T}{\bf v}}+\mathcal{O}\mathopen{}\mathclose{{}\left(\frac{1}{k}}\right). (7.25)

As in the previous case, ε~SN\tilde{\varepsilon}_{\rm SN} and ε~PD\tilde{\varepsilon}_{\rm PD} have opposite signs, and this completes the proof in the second case. Notice how the assumptions we have made ensure the denominators of (7.24) and (7.25) are nonzero.

8 A comparison to numerically computed bifurcation values.

Here we extend the example given in [10] and illustrate the results with the following C1C^{1} family of maps

f(x,y)={U0(x,y),y2α+13,(1r(y))U0(x,y)+r(y)U1(x,y),2α+13yα+23,U1(x,y),yα+23,f(x,y)=\begin{cases}U_{0}(x,y),&y\leq\frac{2\alpha+1}{3},\\ (1-r(y))U_{0}(x,y)+r(y)U_{1}(x,y),&\frac{2\alpha+1}{3}\leq y\leq\frac{\alpha+2}{3},\\ U_{1}(x,y),&y\geq\frac{\alpha+2}{3},\end{cases} (8.1)

where

U0(x,y)\displaystyle U_{0}(x,y) =[(α+μ2)x(1+(a1,0+μ4)xy)1αy(1a1,0xy)],\displaystyle=\begin{bmatrix}(\alpha+\mu_{2})x\big{(}1+(a_{1,0}+\mu_{4})xy\big{)}\\ \frac{1}{\alpha}y(1-a_{1,0}xy)\end{bmatrix}, (8.2)
U1(x,y)\displaystyle U_{1}(x,y) =[1+c2,0(y1)μ1+(1+μ3)x+d5,0(y1)2],\displaystyle=\begin{bmatrix}1+c_{2,0}(y-1)\\ \mu_{1}+(1+\mu_{3})x+d_{5,0}(y-1)^{2}\end{bmatrix}, (8.3)
r(y)\displaystyle r(y) =s(yh0h1h0),\displaystyle=s\mathopen{}\mathclose{{}\left(\frac{y-h_{0}}{h_{1}-h_{0}}}\right), (8.4)

where

s(z)=3z22z3,s(z)=3z^{2}-2z^{3},\\ (8.5)

Below we fix

α=0.8,a1,0=0.2,c2,0=0.5,d5,0=1,\begin{split}\alpha&=0.8,\\ a_{1,0}&=0.2,\\ c_{2,0}&=-0.5,\\ d_{5,0}&=1,\end{split} (8.6)

and vary μ=(μ1,μ2,μ3,μ4)4\mu=(\mu_{1},\mu_{2},\mu_{3},\mu_{4})\in\mathbb{R}^{4}.

Refer to caption
Refer to caption
Figure 3: A phase portrait of (8.1) with (8.6) and μ=𝟎\mu={\bf 0}. The shaded horizontal strip is where the middle component of (8.1) applies. We show parts of the stable and unstable manifolds of (x,y)=(0,0)(x,y)=(0,0). Note the unstable manifold has very high curvature at (x,y)(0,1.1)(x,y)\approx(0,1.1) because (8.1) is highly nonlinear in the horizontal strip. For the given parameter values (8.1) has an asymptotically stable, single-round periodic solutions of period k+1k+1 for all k1k\geq 1. These are shown for k=1,2,,15k=1,2,\ldots,15; different colours correspond to different values of kk. The map also has an asymptotically stable fixed point at (x,y)=(1,1)(x,y)=(1,1).

With μ=𝟎\mu={\bf 0}, (8.1) satisfies the conditions of Theorem 3.2. In particular Δ0=2.25\Delta_{0}=2.25 so Δ0>0\Delta_{0}>0 and 1<c2,0<1Δ02-1<c_{2,0}<1-\frac{\Delta_{0}}{2}. Therefore (8.1) has an asymptotically stable single-round periodic solution for all sufficiently large values of kk. In fact these exist for all k1k\geq 1, see Fig. 3, plus there exists an asymptotically stable fixed point at (x,y)=(1,1)(x,y)=(1,1) that can be interpreted as corresponding to k=0k=0.

Refer to captionRefer to captiona)b)
Figure 4: Panel (a) is a numerically computed bifurcation diagram of (8.1) with (8.6) and μ2=μ3=μ4=0\mu_{2}=\mu_{3}=\mu_{4}=0. The triangles [circles] are saddle-node [period-doubling] bifurcations of single-round periodic solutions of period k+1k+1. Panel (b) shows the same points but with the horizontal axis scaled in such a way that the asymptotic approximations to these bifurcations, given by the leading-order terms in (7.22) and (7.23), appear as vertical lines.

In this remainder of this section we study how the infinite coexistence is destroyed by varying each the components of μ\mu from zero in turn. We identify saddle-node and period-doubling bifurcations numerically and compare these to our above asymptotic results. First, by varying the value of μ1\mu_{1} from zero we destroy the homoclinic tangency. Indeed in (5.5) we have 𝐧tang𝖳=[1,0,0,0]{\bf n}_{\rm tang}^{\sf T}=[1,0,0,0]. Thus if we fix μ2=μ3=μ4=0\mu_{2}=\mu_{3}=\mu_{4}=0 and vary the value of μ1\mu_{1}, by Theorem 6.1 there must exist sequences of saddle-node and period-doubling bifurcations occurring at values of μ1\mu_{1} that are asymptotically proportional to α2k\alpha^{2k}. Fig. 4a shows the bifurcation values (obtained numerically) for six different values of kk. We have designed (8.1) so that it satisfies (7.2). Consequently the formulas (7.22) and (7.23) for the bifurcation values can be applied directly. In panel (b) we observe that the numerically computed bifurcation values indeed converge to their leading-order approximations.

Refer to captionRefer to captiona)b)
Figure 5: Panel (a) is a numerically computed bifurcation diagram of (8.1) with (8.6) and μ1=μ3=μ4=0\mu_{1}=\mu_{3}=\mu_{4}=0. Panel (b) shows convergence to the leading-order terms of (7.24) and (7.25).

We now fix μ1=μ3=μ4=0\mu_{1}=\mu_{3}=\mu_{4}=0 and vary the value of μ2\mu_{2}. This parameter variation alters the product of the eigenvalues associated with the origin. Specifically 𝐧eig𝖳=[0,1α,0,0]{\bf n}_{\rm eig}^{\sf T}=\mathopen{}\mathclose{{}\left[0,\frac{1}{\alpha},0,0}\right] in (4.3) so by Theorem 6.1 the bifurcation values are asymptotically proportional to αkk\frac{\alpha^{k}}{k}. In Fig. 5 we see the numerically computed bifurcation values converging to their leading order approximations (7.24) and (7.25).

Refer to captionRefer to captiona)b)
Figure 6: Panel (a) is a numerically computed bifurcation diagram of (8.1) with (8.6) and μ1=μ2=μ4=0\mu_{1}=\mu_{2}=\mu_{4}=0. Panel (b) shows convergence to the leading-order terms of (8.7) and (8.8).

Next we fix μ1=μ2=μ4=0\mu_{1}=\mu_{2}=\mu_{4}=0 and vary the value of μ3\mu_{3} which breaks the global resonance condition. Here 𝐧tang𝖳𝐯=0{\bf n}_{\rm tang}^{\sf T}{\bf v}=0 and 𝐧eig𝖳𝐯=0{\bf n}_{\rm eig}^{\sf T}{\bf v}=0 so Theorem (6.1) does not apply. But by performing calculations analogous to those given above in the proof of Theorem 6.1 directly to the map (8.1), we obtain the following expressions for the saddle-node and period-doubling bifurcation values

ϵSN=(1c2,0)24d5,0αk+𝒪(|α|2k),\epsilon_{\rm SN}=\frac{(1-c_{2,0})^{2}}{4d_{5,0}}\alpha^{k}+\mathcal{O}(|\alpha|^{2k}), (8.7)
ϵPD=3(1c2,0)24d5,0αk+𝒪(|α|2k).\epsilon_{\rm PD}=-\frac{3(1-c_{2,0})^{2}}{4d_{5,0}}\alpha^{k}+\mathcal{O}(|\alpha|^{2k}). (8.8)

Fig. 6 shows that the numerically computed bifurcations do indeed appear to be converging to the leading-order components of (8.7) and (8.8). Notice the bifurcation values are asymptotically proportional to αk\alpha^{k} (a slightly slower rate than that in Fig. 5).

Refer to captionRefer to captiona)b)
Figure 7: Panel (a) is a numerically computed bifurcation diagram of (8.1) with (8.6) and μ1=μ2=μ3=0\mu_{1}=\mu_{2}=\mu_{3}=0. Panel (b) shows convergence to the leading-order terms of (8.9) and (8.10).

Finally we fix μ1=μ2=μ3=0\mu_{1}=\mu_{2}=\mu_{3}=0 and vary the value of μ4\mu_{4} which breaks the condition on the resonance terms in T0T_{0}. As in the previous case Theorem 6.1 does not apply. By again calculating the bifurcations as above we obtain

ϵSN=(1c2,0)24d5,0k+𝒪(1k2),\epsilon_{{\rm SN}}=\frac{(1-c_{2,0})^{2}}{4d_{5,0}k}+\mathcal{O}\mathopen{}\mathclose{{}\left(\frac{1}{k^{2}}}\right), (8.9)
ϵPD=3(1c2,0)24d5,0k+𝒪(1k2),\epsilon_{{\rm PD}}=-\frac{3(1-c_{2,0})^{2}}{4d_{5,0}k}+\mathcal{O}\mathopen{}\mathclose{{}\left(\frac{1}{k^{2}}}\right), (8.10)

and these agree with the numerically computed bifurcation values as shown in Fig. 7. The bifurcation values are asymptotically proportional to 1k\frac{1}{k} which is substantially slower than in the previous three cases.

9 Discussion

In this paper we have considered globally resonant homoclinic tangencies in smooth two-dimensional maps and determined scaling laws for the size of parameter intervals in which single-round periodic solutions are asymptotically stable. We have illustrated the results with an abstract four-parameter family. It remains to identify globally resonant homoclinic tangencies in prototypical maps and maps derived from physical applications.

About a parameter point satisfying the conditions of Theorem 3.2, for any positive jj\in\mathbb{Z} there exists an open region of parameter space in which the family of maps has jj coexisting asymptotically stable periodic solutions. It follows from Theorem 6.1 that the largest ball (sphere) that the region contains has a diameter asymptotically proportional to |α|2j|\alpha|^{2j}. But, as we have shown, different directions of perturbation yield different scaling laws. Consequently we expect such regions to have an elongated shape for large values of jj. Indeed preliminary investigations reveal that such regions may have a particularly complicated shape, bounded by many of the saddle-node and periodic-doubling bifurcations identified above.

We believe the primary |α|2k|\alpha|^{2k} scaling law holds true for higher-dimensional maps. Certainly similar aspects of homoclinic tangencies have been shown to be independent of dimension, [6]. Also in the piecewise-linear setting, globally resonant homoclinic tangencies were analysed in [16].

Appendix A Proof of Lemma 7.1

Write

T0(x,y)=[λx(1+a1xy+x2y2F~(x,y))σy(1+b1xy+x2y2G~(x,y))].T_{0}(x,y)=\begin{bmatrix}\lambda x\mathopen{}\mathclose{{}\left(1+a_{1}xy+x^{2}y^{2}\tilde{F}(x,y)}\right)\\ \sigma y\mathopen{}\mathclose{{}\left(1+b_{1}xy+x^{2}y^{2}\tilde{G}(x,y)}\right)\end{bmatrix}. (A.1)

Let R>0R>0 be such that

|a1|,|b1|,|F~(x,y)|,|G~(x,y)|R|a_{1}|,|b_{1}|,|\tilde{F}(x,y)|,|\tilde{G}(x,y)|\leq R (A.2)

for all (x,y)𝒩(x,y)\in\mathcal{N} and all sufficiently small values of μ\mu. For simplicity we assume α>0\alpha>0; if instead α<0\alpha<0 the proof can be completed in the same fashion.

We have λ=α+𝒪(|α|k)\lambda=\alpha+\mathcal{O}\mathopen{}\mathclose{{}\left(|\alpha|^{k}}\right) and σ=1α+𝒪(|α|k)\sigma=\frac{1}{\alpha}+\mathcal{O}\mathopen{}\mathclose{{}\left(|\alpha|^{k}}\right). Thus there exists M2RM\geq 2R such that λα(1+Mαk)\lambda\leq\alpha\mathopen{}\mathclose{{}\left(1+M\alpha^{k}}\right) and σ1α(1+Mαk)\sigma\leq\frac{1}{\alpha}\mathopen{}\mathclose{{}\left(1+M\alpha^{k}}\right) for sufficiently large values of kk. It follows (by induction on jj) that

σjαj(1+2Mjαk),\sigma^{j}\leq\alpha^{-j}\mathopen{}\mathclose{{}\left(1+2Mj\alpha^{k}}\right), (A.3)

and

(λσ)j1+4Mjαk,\mathopen{}\mathclose{{}\left(\lambda\sigma}\right)^{j}\leq 1+4Mj\alpha^{k}, (A.4)

for all j=1,2,,kj=1,2,\ldots,k again assuming kk is sufficiently large.

Let ε>0\varepsilon>0. Assume

|x1|\displaystyle|x-1| ε,\displaystyle\leq\varepsilon, |αky1|\displaystyle\mathopen{}\mathclose{{}\left|\alpha^{-k}y-1}\right| ε,\displaystyle\leq\varepsilon, (A.5)

for sufficiently large values of kk. We can assume ε<112\varepsilon<1-\frac{1}{\sqrt{2}} so then

αk2|xy|2αk.\frac{\alpha^{k}}{2}\leq|xy|\leq 2\alpha^{k}. (A.6)

Write

T0j(x,y)=[λjx(1+ja1xy+x2y2F~j(x,y))σjy(1+jb1xy+x2y2G~j(x,y))].T_{0}^{j}(x,y)=\begin{bmatrix}\lambda^{j}x\mathopen{}\mathclose{{}\left(1+ja_{1}xy+x^{2}y^{2}\tilde{F}_{j}(x,y)}\right)\\ \sigma^{j}y\mathopen{}\mathclose{{}\left(1+jb_{1}xy+x^{2}y^{2}\tilde{G}_{j}(x,y)}\right)\end{bmatrix}. (A.7)

Below we will use induction on jj to show that

|F~j(x,y)|,|G~j(x,y)|12Mj2,\mathopen{}\mathclose{{}\left|\tilde{F}_{j}(x,y)}\right|,\mathopen{}\mathclose{{}\left|\tilde{G}_{j}(x,y)}\right|\leq 12Mj^{2}, (A.8)

for all j=1,2,,kj=1,2,\ldots,k, assuming kk is sufficiently large. This will complete the proof because with j=kj=k, (A.8) implies (7.1).

Clearly (A.8) is true for j=1j=1: |F~1(x,y)|=|F~(x,y)|RM2<12M\mathopen{}\mathclose{{}\left|\tilde{F}_{1}(x,y)}\right|=\mathopen{}\mathclose{{}\left|\tilde{F}(x,y)}\right|\leq R\leq\frac{M}{2}<12M and similarly for G~1\tilde{G}_{1}. Suppose (A.8) is true for some j<kj<k. It remains for us to verify (A.8) for j+1j+1. First observe that by using |a1|R|a_{1}|\leq R, (A.6), and the induction hypothesis,

|1+ja1xy+x2y2F~j(x,y)|1+2Rjαk+48Mj2α2k.\mathopen{}\mathclose{{}\left|1+ja_{1}xy+x^{2}y^{2}\tilde{F}_{j}(x,y)}\right|\leq 1+2Rj\alpha^{k}+48Mj^{2}\alpha{2k}.

For sufficiently large kk this implies

|1+ja1xy+x2y2F~j(x,y)|1+2Mjαk,\mathopen{}\mathclose{{}\left|1+ja_{1}xy+x^{2}y^{2}\tilde{F}_{j}(x,y)}\right|\leq 1+2Mj\alpha^{k}, (A.9)

where we have also used M2RM\geq 2R. Similarly

|1+jb1xy+x2y2G~j(x,y)|1+2Mjαk.\mathopen{}\mathclose{{}\left|1+jb_{1}xy+x^{2}y^{2}\tilde{G}_{j}(x,y)}\right|\leq 1+2Mj\alpha^{k}. (A.10)

Write T0j(x,y)=(xj,yj)T_{0}^{j}(x,y)=(x_{j},y_{j}). By using (A.3), (A.5), and (A.10) we obtain

|yj|αkj(1+ε)(1+2Mjαk)2.|y_{j}|\leq\alpha^{k-j}(1+\varepsilon)\mathopen{}\mathclose{{}\left(1+2Mj\alpha^{k}}\right)^{2}.

Thus |yj|αkj(1+2ε)|y_{j}|\leq\alpha^{k-j}(1+2\varepsilon), say, for sufficiently large values of kk. Also |xjyj||x_{j}y_{j}| is clearly small, so we can conclude that (xj,yj)𝒩(x_{j},y_{j})\in\mathcal{N} (in particular we have shown that (xk1,yk1)(x_{k-1},y_{k-1}) can be made as close to (0,1α)\mathopen{}\mathclose{{}\left(0,\frac{1}{\alpha}}\right) as we like).

By matching the first components of T0j+1(x,y)=(T0T0j)(x,y)T_{0}^{j+1}(x,y)=\mathopen{}\mathclose{{}\left(T_{0}\circ T_{0}^{j}}\right)(x,y) we obtain

λj+1x(1+(j+1)a1xy+x2y2F~j+1(x,y))\displaystyle\lambda^{j+1}x\mathopen{}\mathclose{{}\left(1+(j+1)a_{1}xy+x^{2}y^{2}\tilde{F}_{j+1}(x,y)}\right) =λj+1x(1+ja1xy+x2y2F~j(x,y))\displaystyle=\lambda^{j+1}x\mathopen{}\mathclose{{}\left(1+ja_{1}xy+x^{2}y^{2}\tilde{F}_{j}(x,y)}\right)
+λj+1a1x2y(1+P)+λj+1x3y2Q,\displaystyle\quad+\lambda^{j+1}a_{1}x^{2}y(1+P)+\lambda^{j+1}x^{3}y^{2}Q, (A.11)

where

1+P\displaystyle 1+P =(λσ)j(1+ja1xy+x2y2F~j(x,y))2(1+jb1xy+x2y2G~j(x,y)),\displaystyle=\mathopen{}\mathclose{{}\left(\lambda\sigma}\right)^{j}\mathopen{}\mathclose{{}\left(1+ja_{1}xy+x^{2}y^{2}\tilde{F}_{j}(x,y)}\right)^{2}\mathopen{}\mathclose{{}\left(1+jb_{1}xy+x^{2}y^{2}\tilde{G}_{j}(x,y)}\right), (A.12)
Q\displaystyle Q =(λσ)2j(1+ja1xy+x2y2F~j(x,y))3(1+jb1xy+x2y2G~j(x,y))2F~(xj,yj).\displaystyle=\mathopen{}\mathclose{{}\left(\lambda\sigma}\right)^{2j}\mathopen{}\mathclose{{}\left(1+ja_{1}xy+x^{2}y^{2}\tilde{F}_{j}(x,y)}\right)^{3}\mathopen{}\mathclose{{}\left(1+jb_{1}xy+x^{2}y^{2}\tilde{G}_{j}(x,y)}\right)^{2}\tilde{F}(x_{j},y_{j}). (A.13)

By (A.4), (A.9), and (A.10), we obtain

1+P\displaystyle 1+P (1+4Mjαk)(1+2Mjαk)3\displaystyle\leq\mathopen{}\mathclose{{}\left(1+4Mj\alpha^{k}}\right)\mathopen{}\mathclose{{}\left(1+2Mj\alpha^{k}}\right)^{3}
1+11Mjαk,\displaystyle\leq 1+11Mj\alpha^{k},
Q\displaystyle Q (1+4Mjαk)2(1+2Mjαk)5R\displaystyle\leq\mathopen{}\mathclose{{}\left(1+4Mj\alpha^{k}}\right)^{2}\mathopen{}\mathclose{{}\left(1+2Mj\alpha^{k}}\right)^{5}R
2R,\displaystyle\leq 2R,

assuming kk is sufficiently large and where we have also used |F~(xj,yj)|R|\tilde{F}(x_{j},y_{j})|\leq R (valid because (xj,yj)𝒩(x_{j},y_{j})\in\mathcal{N}). From (A.11),

F~j+1(x,y)=F~j(x,y)+Pxy+Q.\tilde{F}_{j+1}(x,y)=\tilde{F}_{j}(x,y)+\frac{P}{xy}+Q.

Then by using the induction hypothesis, the lower bound on |xy||xy| (A.6), and the above bounds on PP and QQ, we arrive at

|F~j+1(x,y)|\displaystyle\mathopen{}\mathclose{{}\left|\tilde{F}_{j+1}(x,y)}\right| 12Mj2+22Mj+2R\displaystyle\leq 12Mj^{2}+22Mj+2R
12Mj2+24Mj\displaystyle\leq 12Mj^{2}+24Mj
<12M(j+1)2.\displaystyle<12M(j+1)^{2}.

In a similar fashion by matching the second components of T0j+1(x,y)=(T0T0j)(x,y)T_{0}^{j+1}(x,y)=\mathopen{}\mathclose{{}\left(T_{0}\circ T_{0}^{j}}\right)(x,y) we obtain |G~j+1(x,y)|<12M(j+1)2\mathopen{}\mathclose{{}\left|\tilde{G}_{j+1}(x,y)}\right|<12M(j+1)^{2}. This verifies (A.8) for j+1j+1 and so completes the proof. \Box

References

  • [1] A. Delshams, M. Gonchenko, and S. Gonchenko. On dynamics and bifurcations of area preserving maps with homoclinic tangencies. Nonlinearity, 28:3027–3071, 2015.
  • [2] Y. Do and Y. C. Lai. Multistability and arithmetically period-adding bifurcations in piecewise smooth dynamical systems. Chaos, 18(4):043107, 2008.
  • [3] N.K. Gavrilov and L.P. Silnikov. On three dimensional dynamical systems close to systems with structurally unstable homoclinic curve. I. Math. USSR Sbornik, 17:467–485, 1972.
  • [4] N.K. Gavrilov and L.P. Silnikov. On three-dimensional dynamical systems close to systems with a structurally unstable homoclinic curve. II. Math. USSR Sbornik, 19:139–156, 1973.
  • [5] M.S. Gonchenko and S.V. Gonchenko. On cascades of elliptic periodic points in two-dimensional symplectic maps with homoclinic tangencies. Regul. Chaotic Dyns., 14:116–136, 2009.
  • [6] S. V. Gonchenko, L. P. Shil’nikov, and D. V. Turaev. Quasiattractors and homoclinic tangencies. Comput. Math. Appl., 34(2–4):195–227, 1997.
  • [7] S.V. Gonchenko and L.P. Shilnikov. On two-dimensional area-preserving maps with homoclinic tangencies that have infinitely many generic elliptic periodic points. J. Math. Sci. (N. Y.), 128:2767–2773, 2005.
  • [8] V.S. Gonchenko, Yu.A. Kuznetsov, and H.G.E. Meijer. Generalized Hénon map and bifurcations of homoclinic tangencies. SIAM J. Appl. Dyn. Syst., 2005.
  • [9] Y. A. Kuznetsov and H. G. E. Meijer. Numerical Bifurcation Analysis of Maps: From Theory to Software. Cambridge Monographs on Applied and Computational Mathematics. Cambridge University press, 2019.
  • [10] S.S. Muni, R.I. McLachlan, and D.J.W. Simpson. Homoclinic tangencies with infinitely many asymptotically stable single-round periodic solutions. Discrete Contin Dyn Syst Ser A, 2021.
  • [11] S.E. Newhouse. Diffeomorphisms with infinitely many sinks. Topology, 13:9–18, 1974.
  • [12] J. Palis and F. Takens. Hyperbolicity and sensitive chaotic dynamics at homoclinic bifurcations. Cambridge University Press, New York, 1993.
  • [13] L. P. Shilnikov, A. L. Shilnikov, D. V. Turaev, and L. O. Chua. Methods of Qualitative Theory in Nonlinear Dynamics. World Scientific, 1998.
  • [14] D.J.W. Simpson. Scaling laws for large numbers of coexisting attracting periodic solutions in the border-collision normal form. Int. J. Bifurcation Chaos, 24:1450118, 2014.
  • [15] D.J.W. Simpson. Sequences of periodic solutions and infinitely many coexisting attractors in the border-collision normal form. Int. J. Bifurcation Chaos, 24:1430018, 2014.
  • [16] D.J.W. Simpson and C.P. Tuffley. Subsumed homoclinic connections and infinitely many coexisting attractors in piecewise-linear continuous maps. Int. J. Bifurcation Chaos, 27, 2017.
  • [17] S. Sternberg. On the structure of local homeomorphisms of Euclidean nn-space, II. Amer. J. Math., 1958.