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

thanks: I am thankful to C. Pugh and M. Hirsch for helpful and instructive discussions

Homoclinic tangencies in n{\mathbb{R}}^{n}

Victoria Rayskin Victoria Rayskin Department of Mathematics
520 Portola Plaza
Box 951555
University of California
Los Angeles, CA 90095-1555, USA
vrayskin@math.ucla.edu
Abstract.

Let f:MMf:M\rightarrow M denote a diffeomorphism of a smooth manifold MM. Let pMp\in M be its hyperbolic fixed point with stable and unstable manifolds WSW_{S} and WUW_{U} respectively. Assume that WSW_{S} is a curve. Suppose that WUW_{U} and WSW_{S} have a degenerate homoclinic crossing at a point BpB\neq p, i.e., they cross at BB tangentially with a finite order of contact.

It is shown that, subject to C1C^{1}-linearizability and certain conditions on the invariant manifolds, a transverse homoclinic crossing will arise arbitrarily close to BB. This proves the existence of a horseshoe structure arbitrarily close to BB, and extends a similar planar result of Homburg and Weiss [10].

Key words and phrases:
Homoclinic tangency, invariant manifolds, λ\lambda-Lemma, order of contact, horseshoe structure
1991 Mathematics Subject Classification:
37B10, 37C05, 37C15, 37D10

1. Introduction

The celebrated Theorem of Birkhoff-Smale (see, for instance, [17]) provides a method for rigorously concluding the existence of a horseshoe structure. The basic assumption of this theorem is the presence of a transverse homoclinic point.

Theorem 1.1 (Birkhoff-Smale).

Let f:nnf:{\mathbb{R}}^{n}\to{\mathbb{R}}^{n} be a diffeomorphism with a hyperbolic fixed point pp. If qpq\not=p is a transverse homoclinic point, then there exist a hyperbolic invariant set Λ\Lambda, a set of bi-infinite sequences Σ{\Sigma} and a shift map σ\sigma such that for some mm the following diagram commutes:

ΛfmΛhhΣσΣ\begin{array}[]{ccc}\Lambda&\stackrel{{\scriptstyle f^{m}}}{{\rightarrow}}&\Lambda\\ h\downarrow&&\ \ \downarrow h\\ \Sigma&\stackrel{{\scriptstyle\sigma}}{{\rightarrow}}&\Sigma\end{array}

Here Σ={A2A1.A0A1A2}\Sigma=\{...A_{-2}A_{-1}.A_{0}A_{1}A_{2}...\}, Ai{0,,N}A_{i}\in\{0,...,N\},
σ{A2A1.A0A1A2}={A2A1A0.A1A2}\sigma\{...A_{-2}A_{-1}.A_{0}A_{1}A_{2}...\}=\{...A_{-2}A_{-1}A_{0}.A_{1}A_{2}...\} and hh is a homeomorphism mapping Λ\Lambda onto Σ\Sigma.

The assumption of transversality is not easy to verify for a concrete dynamical system. There were many attempts to remove this assumption. (See K. Burns and H. Weiss [3], A. J. Homburg and H. Weiss [10], M. Hirsch [9], R. Churchill and D. Rod [4], Gavrilov and Shilnikov [5][6], Gonchenko and Shilnikov [7]). These papers address the question of existence of horseshoes arbitrarily close to a homoclinic point in planar dynamics.

The object of the papers of Hirsch [9] and Churchill and Rod [4] was to show that, in the planar case, the manifolds WSW_{S} and WUW_{U} cross transversely at some other point, arbitrarily close to the original crossing, if the original crossing is two-sided (Figure 5, case IV) and of finite contact.

Churchill and Rod [4] imposed the additional assumptions that ff is analytic and area-preserving. In that case, by Moser’s Theorem [13], this diffeomorphism can be assumed to be in the Birkhoff normal form, and any branch of the stable (unstable) manifold can be re-parameterized so that fn(WU)WSf^{n}(W_{U})\cap W_{S} contains a positive orbit q,fq,f2q,,fkqq^{\prime},fq^{\prime},f^{2}q^{\prime},...,f^{k}q^{\prime} that lies on a hyperbola. Then, direct calculations imply that qq^{\prime} is a transverse homoclinic point. But area preservation is a strong restriction. Furthermore, analyticity allows one to consider the case of a finite order of contact only.

Hirsch [9], replaced the above assumptions with the smooth linearizability assumption (near pp). If we substitute LL, the linear part of ff, for ff, the order of contact will not be changed, and we can calculate the “slope” between two manifolds at the point qq^{\prime}, where Ln(WU)L^{n}(W_{U}) crosses WSW_{S}. If the crossing at qq^{\prime} is transverse, then fn(WU)=ΦLnΦ1(WU)f^{n}(W_{U})=\Phi{L}^{n}{\Phi}^{-1}(W_{U}) also crosses WSW_{S} non-degenerately, where Φ\Phi is the C1C^{1} linearizing map.

Homburg and Weiss [10] considered a surface diffeomorphism and its homoclinic one-sided (Figures 5, 4) tangency of arbitrary high (possibly infinite) order of contact. They proved that some power of this diffeomorphism has the full shift on two symbols as a topological factor. Then, Katok’s theorem [11], [12] implies that the map possesses a horseshoe arbitrarily close to homoclinic tangency.

Important cases of homoclinic tangencies were also described in works of Shilnikov, Gonchenko and Gavrilov (see [7], [5], [6]). They considered one-sided planar homoclinic tangencies.

Gavrilov and Shilnikov [5], [6] established a topological conjugacy (on a closed invariant set) between a surface diffeomorphism having a dissipative hyperbolic periodic point with certain types of quadratic (one-sided) tangencies (Figures 5, 4) and the full shift on two symbols.

Gonchenko and Shilnikov considered higher orders of contact. Homburg and Weiss write: “Gonchenko claims that a proof of this result for finite order tangencies appeared in his unpublished (Russian) thesis in 1984. The result for finite order tangencies was announced by Gonchenko and Shilnikov [7] in 1986, but we are unable to find any proof in the literature.” We were unable to find such a proof in the existing literature either.

Homburg and Weiss, in their paper [10], state an Open Problem, asking whether their results could be extended to dimensions higher than two. In the present paper, we consider a diffeomorphism ff of a smooth manifold MM (of an arbitrary dimension) with a hyperbolic fixed point. We assume that one of the invariant manifolds of ff is a curve, that this curve has finite tangential contact with the other invariant manifold, and that the contact is not of Newhouse type (see Definition 3.1 and Figure 4). We prove (Theorem 3.2) that, subject to C1C^{1}-linearizability of ff and the existence of a graph portion of the invariant manifold, a transverse homoclinic crossing will arise arbitrarily close to the tangential crossing. This implies the existence of a horseshoe arbitrarily close to the tangential crossing.

To prove this result, we consider (Section 3) tangent subspaces of stable and unstable manifolds at a certain sequence of their intersection points. We show that, at some intersection points, a pair of the tangent subspaces spans the entire n{\mathbb{R}}^{n}. In Section 2 we present basic definitions and lemmas that are used in Section 3. In particular, we prove tangential λ\lambda-Lemma 2.14, essential for the proof of the main result. Also, in Section 2, we present a new definition of the order of contact of two manifolds. This definition does not require high differentiability and plays a key role in our approach.

2. Definitions and Preliminaries

In this section we give basic definitions and state several preliminary lemmas to be used in the proof of the main result. The references for this Section are V. I. Arnold, S. M. Guseĭn-Zade [1] and A. N. Varchenko, K. Burns and H. Weiss [3], M. Hirsch [9], A. J. Homburg and H. Weiss [10], S. Newhouse and J. Palis [14], J. Palis [15], V. Rayskin [16], R. Uribe-Vargas [18].

Throughout this section, we will be considering two submanifolds in a certain ambient smooth manifold (of arbitrary dimension). Suppose two such submanifolds meet at an isolated point AA. We will discuss the contact properties of these submanifolds at point AA. Since these properties are local, we may assume without loss of generality that the ambient manifold is just n{\mathbb{R}}^{n}.

First, assume that each manifold is a curve. The order of contact of two smooth curves has been defined and studied in works of Arnold, Guseĭn-Zade [1] and A. N. Varchenko, Burns and Weiss [3], Hirsch [9], Homburg and Weiss [10], Uribe-Vargas [18]. Their definitions of the ll-th order of contact require high differentiability; i.e. ll is the maximal integer, such that the first l1l-1 derivatives of the two curves coincide at the point of their contact. Our definition below does not require high differentiability and thus, allows us to apply C1C^{1}-linearization in our main Theorem 3.2.

Definition 2.1.

Let γi{\gamma}_{i} (i=1,2i=1,2) denote two immersed C1C^{1}-curves in n{\mathbb{R}}^{n} . Suppose the two curves meet at an isolated point AA. Then we will say that a number ll is the order of contact of the curve γ1\gamma_{1} with the curve γ2\gamma_{2} at the point AA if there exist positive real numbers mm and MM such that for all points xx on γ1\gamma_{1} sufficiently close to AA

md(x,γ2)xAlM.m\leq\frac{d(x,\gamma_{2})}{\|x-A\|^{l}}\leq M\;.
Proposition 2.2.

If the curves are C1C^{1}-smooth, then the definition of the order of contact is symmetric with respect to γ1\gamma_{1} and γ2\gamma_{2}.

Proof.

Since the curves are C1C^{1}-smooth, we can assume that γ2\gamma_{2} is the XX-axis (in some coordinate system), and γ1\gamma_{1} is a graph of a C1C^{1} function ff in this system. Let the order of contact of γ1\gamma_{1} with γ2\gamma_{2} be ll. By Definition 2.1 it means that

mf(x)(x2+f(x)2)l/2M,m\leq\frac{\|f(x)\|}{\left(x^{2}+\|f(x)\|^{2}\right)^{l/2}}\leq M\;,

so that there exists a constant CC such that for all sufficiently small xx

mf(x)|x|lC,m\leq\frac{\|f(x)\|}{|x|^{l}}\leq C\;,

which means that γ1\gamma_{1} is sandwiched between two order ll paraboloids 𝒫m{\mathcal{P}}_{m} and 𝒫C{\mathcal{P}}_{C}, respectively.

Let us now look at the order of contact of γ2\gamma_{2} with γ1\gamma_{1}. By the above, for any point xXx\in X

d(x,𝒫m)d(x,γ1)d(x,𝒫C).d(x,{\mathcal{P}}_{m})\leq d(x,\gamma_{1})\leq d(x,{\mathcal{P}}_{C})\;.

Since 𝒫m{\mathcal{P}}_{m} and 𝒫C{\mathcal{P}}_{C} are order ll paraboloids, the distances to them are uniformly equivalent to |x|l|x|^{l}, whence the claim ∎

Remark 2.3.

Our notion of the order of contact is more general than the classical one. Obviously, if the order of contact of two curves is ll in the classical sense, then it is also ll in our sense. The converse, however, need not be true.

Remark 2.4.

Without the C1C^{1} assumption Proposition 2.2 may fail. For instance, let us consider the graphs of the functions γ1(x)=0\gamma_{1}(x)=0 and γ2(x)=x2+(1+sin(1/x))x\gamma_{2}(x)=x^{2}+(1+\sin(1/x))x on the plane. γ2\gamma_{2} oscillates between x2x^{2} and xx with the distance between consecutive crests of order x2x^{2}. Then, the order of contact of γ1\gamma_{1} with γ2\gamma_{2} in the sense of our Definition 2.1 is 2, whereas the order of contact of γ2\gamma_{2} with γ1\gamma_{1} does not exists.

Naturally, the ll-th order of contact (in the sense of [1], [3], [9], [10], [18]) between two ClC^{l}-curves is preserved under a C1C^{1}-diffeomorphism. Lemma 2.5 below shows that the order of contact in the sense of our Definition 2.1 is preserved by C1C^{1}-diffeomorphisms. In particular, our order of contact from Definition 2.1 does not depend on the choice of a coordinate chart in the ambient manifold, so that it is well-defined in the general manifold setup as well.

Lemma 2.5.

Given two C1C^{1}-curves in n{\mathbb{R}}^{n} intersecting at an isolated point, any diffeomorphism of a neighborhood of this point preserves the order of contact of these curves.

Proof.

Let curves γ1\gamma_{1} and γ2\gamma_{2} have order of contact ll at some point AA. Then there are positive constants mm and MM such that for all points xγ1x\in\gamma_{1} sufficiently close to AA

mxyxAlM,m\leq\frac{\|x-y\|}{\|x-A\|^{l}}\leq M,

where y=y(x)y=y(x) denotes a point on γ2\gamma_{2} minimizing the distance from xx to γ2\gamma_{2}. By the C1C^{1} Mean Value Theorem applied to the diffeomorphism ϕ\phi,

ϕ(x)ϕ(y)=[01(Dϕ)σ(s)𝑑s](xy),\phi(x)-\phi(y)=\Bigg{[}\int_{0}^{1}(D\phi)_{\sigma(s)}ds\Bigg{]}(x-y),

where σ:[0,1]n\sigma:[0,1]\to{\mathbb{R}}^{n} is a path connecting the points x=σ(0)x=\sigma(0) and y=σ(1)y=\sigma(1). As xAx\to A, y(x)Ay(x)\to A. Then, σ(s)A\sigma(s)\to A, and the matrix 01(Dϕ)σ(s)𝑑s\int_{0}^{1}(D\phi)_{\sigma(s)}ds tends to the invertible matrix (Dϕ)A(D\phi)_{A}, hence the claim. ∎

Let us now pass to a discussion of the order of contact for higher dimensional intersecting manifolds.

Definition 2.6.

Let S1S_{1} and S2S_{2} denote two immersed C1C^{1}-manifolds in n{\mathbb{R}}^{n}. Suppose the two manifolds meet at an isolated point AA. Their order of contact ll at AA is the supremum of the orders of contact of curves γ1S1\gamma_{1}\subset S_{1} and γ2S2\gamma_{2}\subset S_{2} passing through point AA.

As it follows from Lemma 2.5, the order of contact is preserved under C1C^{1}-diffeomorphisms.


The well known λ{\lambda}-Lemma of Palis [15] gives an important description of chaotic dynamics. The basic assumption of this theorem is the presence of a transverse homoclinic point, which means that the manifolds have contact of order 1.

Theorem 2.7 (λ\lambda-Lemma, Palis).

Let ff be a diffeomorphism of n{\mathbb{R}}^{n} with a hyperbolic fixed point at 0 and mm- and pp-dimensional stable and unstable manifolds WSW_{S} and WUW_{U} (m+p=nm+p=n). Let DD be a pp-disk in WUW_{U}, and ww be another pp-disk in WUW_{U} meeting WSW_{S} at some point AA transversely. Then n0fn(w)\bigcup_{n\geq 0}f^{n}(w) contains pp-disks arbitrarily C1C^{1}-close to DD.

Later in this section we will prove an analog of this lemma (Lemma 2.14) for non-transversal homoclinic intersections.

For the estimates in the proofs of the Singular λ\lambda-Lemma (i.e., in the case when the order of contact is greater than 1) and the Main Theorem we need the following definitions. The first one, the definition of a graph portion, deals with the shape of a global invariant manifold in the vicinity of a homoclinic point. This homoclinic point must be chosen close to a hyperbolic reference point. Then, the local properties of the invariant manifold are checked. The second definition describes a group of diffeomorphisms that can be locally represented (after a C1C^{1} change of coordinates) in some special normal form.

Definition 2.8.

Let ff be a diffeomorphism of n{\mathbb{R}}^{n} with a hyperbolic fixed point at the origin and UU\in n{\mathbb{R}}^{n} be some small neighborhood of 0. Denote by WSW_{S} (resp., WUW_{U}) the associated stable (resp., unstable) manifold, and by mm (resp., pp) its dimension (m+p=nm+p=n). Let AA be a homoclinic point of WSW_{S} and WUW_{U}. Denote by 𝒱\mathcal{V} a small pp-neighborhood in WUW_{U} around the origin. Define a local coordinate system E1E_{1} at 0, which spans 𝒱\mathcal{V}. Similarly, define a local coordinate system E2E_{2} which spans WSW_{S} near 0. Let E=E1+E2E=E_{1}+E_{2}, EUE\subset U. A pp-neighborhood ΛWU\Lambda\subset W_{U} is a graph portion in UU, associated with the homoclinic point AA, if for some n>0n>0 fn(A)Λf^{n}(A)\in\Lambda, Λ(UWU)\Lambda\subset(U\cap W_{U}), and Λ\Lambda is a graph in EE-coordinates of some C1C^{1}-function defined on 𝒱\mathcal{V}.

\psfrag{L}{$\Lambda$}\psfrag{A}{$A$}\psfrag{0}{$0$}\psfrag{V}{$\mathcal{V}$}\psfrag{WS}{$W_{S}$}\psfrag{WU}{$W_{U}$}\epsfbox{fig_graph.eps}
Figure 1. In this figure, Λ\Lambda is not a graph portion of the manifold WUW_{U}, because, for any nn, an iteration of Λ\Lambda with diffeomorphism fnf^{n} is never a graph in the given coordinates.
Definition 2.9.

We will say that diffeomorphism ff with a hyperbolic fixed point 0 satisfies the normal form condition, if there exists a local C1C^{1} change of coordinates in a neighbourhood of 0, under which ff takes the following normal form: f(x,y)=(S1(x,y),S2(x,y))f(x,y)=(S_{1}(x,y),S_{2}(x,y)), where S1S_{1} and S2S_{2} belong to expanding and contracting subspaces respectively, and S2S_{2} is linear, i.e S2(x,y)=yS_{2}(x,y)={\mathcal{B}}y, with <1\|{\mathcal{B}}\|<1.

We are interested in the class of diffeomorphisms satisfying the normal form condition described above. The next lemma (Lemma 2.12) allows us to ascertain belonging to this class by verifying certain conditions (called resonances) on the eigenvalues of the linear part of the map. The following two definitions are used in the lemma.

Definition 2.10.

Let ff be a diffeomorphism of n{\mathbb{R}}^{n} with the linear part

((𝒜x)1,,(𝒜x)p,(y)1,,(y)m)(({\mathcal{A}}x)_{1},\dots,({\mathcal{A}}x)_{p},({\mathcal{B}}y)_{1},\dots,({\mathcal{B}}y)_{m})

at a hyperbolic fixed point 0, where mm and pp are the dimensions of its stable and unstable manifolds, respectively. Then, ff satisfies the mixed second order resonance condition if there exist aspec𝒜a\in\mathop{\rm spec}{\mathcal{A}} and bspecb\in\mathop{\rm spec}{\mathcal{B}} with ab(spec𝒜spec)ab\in(\mathop{\rm spec}{\mathcal{A}}\cup\mathop{\rm spec}{\mathcal{B}}).

Definition 2.11.

We will say that ff has mixed second order resonance in its contracting coordinates if there exist aspec𝒜a\in\mathop{\rm spec}{\mathcal{A}}, bspecb\in\mathop{\rm spec}{\mathcal{B}} with abspecab\in\mathop{\rm spec}{\mathcal{B}}.

Lemma 2.12.

If a CC^{\infty} diffeomorphism ff with a hyperbolic fixed point 0 has no mixed second order resonances in the contracting coordinates, then it satisfies the normal form condition.

Remark 2.13.

The CC^{\infty} assumption in the above Lemma 2.12 can be replaced with a CkC^{k} assumption, where kk depends on the spectrum of the linear part, like in Bronstein and Kopanskii [2] (Theorem 11.9). As this dependence is a bit complicated we assume CC^{\infty}.

Proof.

Let x=(x1,,xp)px=(x_{1},\dots,x_{p})\in{\mathbb{R}}^{p}, y=(y1,,ym)my=(y_{1},\dots,y_{m})\in{\mathbb{R}}^{m} (p+m=np+m=n) and f(x,y):nnf(x,y):{\mathbb{R}}^{n}\to{\mathbb{R}}^{n} have the linear part

((𝒜x)1,,(𝒜x)p,(y)1,,(y)m),(({\mathcal{A}}x)_{1},\dots,({\mathcal{A}}x)_{p},({\mathcal{B}}y)_{1},\dots,({\mathcal{B}}y)_{m}),

with 𝒜1\|{\mathcal{A}}^{-1}\|, <λ<1\|{\mathcal{B}}\|<{\lambda}<1.

First, we will establish conjugation between ff and the following normal form:

𝒜x+(i=1,,p;j=1,,maij1xiyj,,i=1,,p;j=1,,maijpxiyj),{\mathcal{A}}x+\left(\sum_{{i=1,\dots,p;}\atop{j=1,\dots,m}}a_{ij}^{1}x_{i}y_{j},\dots,\sum_{{i=1,\dots,p;}\atop{j=1,\dots,m}}a_{ij}^{p}x_{i}y_{j}\right),
y+(i=1,,p;j=1,,mbij1xiyj,,i=1,,p;j=1,,mbijmxiyj).{\mathcal{B}}y+\left(\sum_{{i=1,\dots,p;}\atop{j=1,\dots,m}}b_{ij}^{1}x_{i}y_{j},\dots,\sum_{{i=1,\dots,p;}\atop{j=1,\dots,m}}b_{ij}^{m}x_{i}y_{j}\right).

Conjugation of ff with this quadratic polynomial can be constructed C1C^{1}-smooth, because any term of degree higher than 2 in the polynomial expansion, of a CC^{\infty} diffeomorphism with a hyperbolic fixed point satisfies the S(1)S(1) condition defined by Bronstein and Kopanskii [2] (Definition 7.4, page 110).

Suppose that the polynomial expansion of ff contains a term cxkylcx^{k}y^{l} and at least one of the two inequalities holds: |k|>1|k|>1, |l|>1|l|>1. For definiteness let us assume that |k|>1|k|>1 (where |k||k| denotes the length of the multiindex kk). Also, we can assume that the eigenvalues are ordered: 𝒜1𝒜p{\mathcal{A}}_{1}\leq\dots\leq{\mathcal{A}}_{p}. Let k=(k1,,kp)k=(k_{1},\dots,k_{p}) and j=max{i:ki0}j=\max\{i:k_{i}\neq 0\}. Then,

𝒜1k1𝒜jkj>𝒜j,{\mathcal{A}}_{1}^{k_{1}}\cdots{\mathcal{A}}_{j}^{k_{j}}>{\mathcal{A}}_{j},

i.e. cxkylcx^{k}y^{l} satisfies the S(1)S(1) condition.

Remark 7.6 in Bronstein and Kopanskii [2] (page 111) asserts that S(1)𝒜(1)S(1)\implies{\mathcal{A}}(1). Then, Samovol’s theorem ([2], Theorem 10.1, page 179) implies C1C^{1}-conjugation of ff with the quadratic polynomial.

Moreover, since ff has no mixed second order resonances in its contracting coordinates, by Samovol’s theorem it can actually be C1C^{1}-conjugated to

𝒜x\displaystyle{\mathcal{A}}x +(i=1,,p;j=1,,maij1xiyj,,i=1,,p;j=1,,maijpxiyj),\displaystyle+\left(\sum_{{i=1,\dots,p;}\atop{j=1,\dots,m}}a_{ij}^{1}x_{i}y_{j},\dots,\sum_{{i=1,\dots,p;}\atop{j=1,\dots,m}}a_{ij}^{p}x_{i}y_{j}\right),
y.\displaystyle{\mathcal{B}}y.


Obviously, the conclusion of the λ\lambda-Lemma of Palis is not true for an arbitrary degenerate (non-transverse) crossing (see, for example, figure 4).

We will now prove an analog of the λ{\lambda}-Lemma for the non-transverse case in n{\mathbb{R}}^{n}.

Lemma 2.14 (Singular λ\lambda-Lemma).

Let ff be a C2C^{2} diffeomorphism of n{\mathbb{R}}^{n} with a hyperbolic fixed point at 0 and with mm- and pp-dimensional stable and unstable manifolds WSW_{S} and WUW_{U} (m+p=nm+p=n), respectively. Assume that ff satisfies the normal form condition and there exists ΛWU\Lambda\subset W_{U}, a graph portion in a small neighborhood of 0, associated with some degenerate homoclinic point AA (cf. Definition 2.8). Also assume that ll, the order of contact at AA, is finite (1<l<1<l<\infty).

Then, for any ρ>0\rho>0, for an arbitrarily small ϵ\epsilon-neighborhood 𝒰n{\mathcal{U}}\subset{\mathbb{R}}^{n} of the origin and for the graph portion Λ{\Lambda}, the set (n0fn(Λ))𝒰(\bigcup_{n\geq 0}f^{n}({\Lambda}))\setminus{\mathcal{U}} contains disks ρ\rho-C1C^{1} close to 𝒱𝒰{\mathcal{V}}\setminus{\mathcal{U}} (here 𝒱{\mathcal{V}} is a small neighborhood of 0 in the local unstable manifold).

\psfrag{L}{$\Lambda$}\psfrag{f}{$f(\Lambda)$}\psfrag{ff}{$f^{2}(\Lambda)$}\psfrag{A}{$A$}\psfrag{f(A)}{$f(A)$}\psfrag{f(f(A))}{$f^{2}(A)$}\psfrag{V}{$\mathcal{V}$}\psfrag{S}{$W_{S}$}\psfrag{U}{$W_{U}$}\psfrag{X}{$X$}\psfrag{Y}{$Y$}\psfrag{o}{$0$}\epsfbox{lemma.eps}
Figure 2. Iterations of the graph portion Λ\Lambda with the diffeomorphism ff.
Proof.

Let α=1/l\alpha=1/l (0<α<10<\alpha<1). Since Λ{\Lambda} is a graph portion that has order of contact ll with WSW_{S} at point AA, we can assume that locally Λ{\Lambda} is the graph of a function

Λ(x)=A+r(x):pn,r(0)=0,{\Lambda}(x)=A+r(x):{\mathbb{R}}^{p}\to{\mathbb{R}}^{n},\quad r(0)=0,

and for any sufficiently small σ>0\sigma>0

|r(x)|const|x|αand|xir(x)|const|x|α1|r(x)|\leq\mathop{\rm const}\cdot|x|^{\alpha}\quad\mbox{and}\quad\left|\frac{\partial}{\partial x_{i}}r(x)\right|\leq\mathop{\rm const}\cdot|x|^{\alpha-1}\

for all |x|<σ|x|<\sigma, i=1,,pi=1,\dots,p. Let x=(x1,,xp)px=(x_{1},\dots,x_{p})\in{\mathbb{R}}^{p}, y=(y1,,ym)my=(y_{1},\dots,y_{m})\in{\mathbb{R}}^{m} (p+m=np+m=n) and f(x,y):nnf(x,y):{\mathbb{R}}^{n}\to{\mathbb{R}}^{n} have the linear part

((𝒜x)1,,(𝒜x)p,(y)1,,(y)m).(({\mathcal{A}}x)_{1},\dots,({\mathcal{A}}x)_{p},({\mathcal{B}}y)_{1},\dots,({\mathcal{B}}y)_{m}).

Assume that 𝒜1\|{\mathcal{A}}^{-1}\|, <λ<1\|{\mathcal{B}}\|<{\lambda}<1. Choose an arbitrarily small Δ\Delta. We know that locally ff can be written in the form f(x,y)=(S1(x,y),S2(x,y))f(x,y)=(S_{1}(x,y),S_{2}(x,y)), where S2(x,y)S_{2}(x,y) is linear, S2(x,y)=yS_{2}(x,y)={\mathcal{B}}y. Since ff is C2C^{2}, the existence of invariant manifolds implies that, in appropriate coordinates and in a sufficiently small neighborhood of the origin, ff can be written as

S1(x,y)=𝒜x\displaystyle S_{1}(x,y)={\mathcal{A}}x +(i=1,,pxiUi1(x,y),,i=1,,pxiUip(x,y)),\displaystyle+\left(\sum_{i=1,\dots,p}x_{i}U_{i}^{1}(x,y),\dots,\sum_{i=1,\dots,p}x_{i}U_{i}^{p}(x,y)\right),
S2(x,y)=y,\displaystyle S_{2}(x,y)={\mathcal{B}}y,

with U(0)=0U(0)=0, UC0Δ\|U\|_{C^{0}}\leq\Delta (for an arbitrary small Δ\Delta), and UC1\|U\|_{C^{1}} bounded.

Consider f(x,Λ(x))=(T1Λ(x),T2Λ(x))f(x,{\Lambda}(x))=(T_{1}^{\Lambda}(x),T_{2}^{\Lambda}(x)). We will work with (x,T2Λ(T1Λ)1(x))(x,T_{2}^{\Lambda}\circ(T_{1}^{\Lambda})^{-1}(x)) and deduce that fn(x,Λ(x))f^{n}(x,{\Lambda}(x)) is C1C^{1}-small for nn big enough, |x|(ϵ,σ)|x|\in(\epsilon,\sigma), and σ>0\sigma>0 sufficiently small. First we will show that, in C1C^{1}-topology, (T1Λ)1(T_{1}^{\Lambda})^{-1} is Δ\Delta-close to 𝒜1{\mathcal{A}}^{-1}. For simplicity, we will denote T1ΛT_{1}^{\Lambda} by T1T_{1} and T2ΛT_{2}^{\Lambda} by T2T_{2}. Then

T1(x)=𝒜x+(i=1,,pxiUi1(x,Λ(x)),,i=1,,pxiUip(x,Λ(x))).T_{1}(x)={\mathcal{A}}x+\left(\sum_{i=1,\dots,p}x_{i}U_{i}^{1}(x,{\Lambda}(x)),\dots,\sum_{i=1,\dots,p}x_{i}U_{i}^{p}(x,{\Lambda}(x))\right).
Claim 2.15.
T1nC1(A1C1+KΔ)n\bigl{\|}T_{1}^{-n}\bigr{\|}_{C^{1}}\leq\left(\bigl{\|}A^{-1}\bigr{\|}_{C^{1}}+K\cdot\Delta\right)^{n}

for |x|<σ|x|<\sigma (σ>0\sigma>0 sufficiently small, K>0K>0).

Proof.

Fix some l{1,,p}l\in\{1,\dots,p\}. Recall that Λ(x)=A+r(x)\Lambda(x)=A+r(x).

|xixlΛj(x)|\displaystyle\bigg{|}x_{i}\frac{\partial}{\partial x_{l}}{\Lambda}_{j}(x)\bigg{|} |xi||xlΛj(x)|\displaystyle\leq|x_{i}|\cdot\bigg{|}\frac{\partial}{\partial x_{l}}{\Lambda}_{j}(x)\bigg{|}
|x|O(1)|x|α1\displaystyle\leq|x|\cdot O(1)|x|^{\alpha-1}
O(1)|x|α\displaystyle\leq O(1)|x|^{\alpha}

Through the proof of this Theorem, O(1)O(1) will denote the set of functions

O(1)=\displaystyle O(1)= {γ:, such that there exists a positive constant c with\displaystyle\big{\{}\gamma:{\mathbb{R}}\mapsto{\mathbb{R}}\mbox{, such that there exists a positive constant $c$ with }
|γ(ζ)|c for all sufficiently small ζ}\displaystyle|\gamma(\zeta)|\leq c\mbox{ for all sufficiently small }\zeta\big{\}}

Also,

|x||xlUit(x,Λ(x))|\displaystyle|x|\cdot\bigg{|}\frac{\partial}{\partial x_{l}}U_{i}^{t}(x,{\Lambda}(x))\bigg{|} =|x||xlUit(x,y)+k=1mykUit(x,y)xlΛk(x)|\displaystyle=|x|\cdot\bigg{|}\frac{\partial}{\partial x_{l}}U_{i}^{t}(x,y)+\sum_{k=1}^{m}\frac{\partial}{\partial y_{k}}U_{i}^{t}(x,y)\cdot\frac{\partial}{\partial x_{l}}{\Lambda}_{k}(x)\bigg{|}
=O(1)|x|α,\displaystyle=O(1)|x|^{\alpha},

and

U(x,Λ(x))C0Δ.\big{\|}U(x,{\Lambda}(x))\big{\|}_{C^{0}}\leq\Delta.

Therefore,

(i=1,,pxiUi1(x,Λ(x)),,i=1,,pxiUip(x,Λ(x)))C1ΔO(1),\displaystyle\Bigg{\|}\Bigg{(}\sum_{i=1,\dots,p}x_{i}U_{i}^{1}(x,{\Lambda}(x)),\dots,\sum_{i=1,\dots,p}x_{i}U_{i}^{p}(x,{\Lambda}(x))\Bigg{)}\Bigg{\|}_{C^{1}}\leq\Delta\cdot O(1),

if σ\sigma is sufficiently small and |x|<σ|x|<\sigma. (An arbitrarily small Δ\Delta was chosen earlier). Then,

T1nC1(A1C1+KΔ)n\bigl{\|}T_{1}^{-n}\bigr{\|}_{C^{1}}\leq\left(\bigl{\|}A^{-1}\bigr{\|}_{C^{1}}+K\cdot\Delta\right)^{n}

Now, we continue the proof of Lemma 2.14 and show that T2nT1nC1\|T_{2}^{n}\circ T_{1}^{-n}\|_{C^{1}} is ρ\rho-small.

T2nT1nC1\displaystyle\|T_{2}^{n}\circ T_{1}^{-n}\|_{C^{1}} =nΛ(T1n)C1\displaystyle=\|{\mathcal{B}}^{n}\cdot\Lambda(T_{1}^{-n})\|_{C^{1}}
nΛ(T1n)C1\displaystyle\leq\|{\mathcal{B}}\|^{n}\cdot\|\Lambda(T_{1}^{-n})\|_{C^{1}}
=O(1)(T1n)α1C0T1nC1\displaystyle=O(1)\cdot\|(T_{1}^{-n})^{\alpha-1}\|_{C^{0}}\cdot\|T_{1}^{-n}\|_{C^{1}}
=O(1)T1nC0α1T1nC1\displaystyle=O(1)\cdot\|T_{1}^{-n}\|_{C^{0}}^{\alpha-1}\cdot\|T_{1}^{-n}\|_{C^{1}}
=O(1)(T1nC1x)α1T1nC1\displaystyle=O(1)\cdot\big{(}\|T_{1}^{-n}\|_{C^{1}}\|x\|\big{)}^{\alpha-1}\cdot\|T_{1}^{-n}\|_{C^{1}}
=O(1)T1nC1αxα1.\displaystyle=O(1)\cdot\|T_{1}^{-n}\|_{C^{1}}^{\alpha}\cdot\|x\|^{\alpha-1}.

Applying Claim 2.15 to the last expression, we obtain

T2nT1nC1=O(1)(A1+ΔK)αnxα1.\|T_{2}^{n}\circ T_{1}^{-n}\|_{C^{1}}=O(1)\cdot\Big{(}\big{\|}A^{-1}\big{\|}+\Delta K\Big{)}^{\alpha n}\big{\|}x\big{\|}^{\alpha-1}.

The last expression is small if x>ϵ\big{\|}x\big{\|}>\epsilon and nn is big enough. ∎

\psfrag{q}{$A$}\psfrag{f(f(q))}{$f^{2}(A)$}\psfrag{f(q)}{$f(A)$}\psfrag{p}{$0$}\epsfbox{lemma_counter_ex.eps}
Figure 3. The iterated manifold is not a graph portion. The iterations of this manifold do not come C1C^{1}-close to the neighborhood of 0 in WUW_{U}.
Remark 2.16.

Clearly, if WUW_{U} does not contain any graph portion Λ\Lambda, the conclusion of the λ\lambda-Lemma 2.14 is not true. Figure 3 illustrates this situation.

3. The Main Theorem

In this section, we extend the results of Hirsch [9], Homburg and Weiss [10], and Gavrilov and Shilnikov [5], [6] to certain types of non-planar homoclinic tangencies with finite orders of contact. Our goal is to establish a transversal crossing arbitrarily close to the non-transversal one. Then, it would follow that there exist horseshoes, located arbitrarily close to the degenerate crossing.

Consider a diffeomorphism ff on n{\mathbb{R}}^{n} which has 11-dimensional stable and (n1)(n-1)-dimensional unstable manifolds, and a homoclinic point BB. Since the stable manifold is 11-dimensional, we can obviously define one- and two-sided homoclinic intersections.

Definition 3.1.

We will say that ff has a homoclinic tangency of Newhouse type, if ff possesses only one-sided homoclinic tangencies.

If ff possesses only one-sided homoclinic tangencies, then transversal intersection need not exist in the neighborhood of BB even in the planar case. The Newhouse example (Figure 4) illustrates this situation. In the planar case, this type of tangency (Newhouse type) was studied by Homburg and Weiss [10]. Unfortunately, we can not extend their results to n{\mathbb{R}}^{n}.

\psfrag{p}{$p$}\psfrag{q}{$B$}\epsfbox{newhouse.eps}
Figure 4. Newhouse example.

The other three cases of one-sided intersections in 2{\mathbb{R}}^{2} were studied by Gavrilov and Shilnikov, and Homburg and Weiss (Figure 5, cases I, II, III). The case of two-sided intersection in 2{\mathbb{R}}^{2} ((Figure 5, case IV) was investigated by Hirsch and Chirchill and Rod.

We will consider, in higher dimensions, the situation similar to the planar cases I, II, III and IV, although we clearly have a different number of cases in n{\mathbb{R}}^{n}. For example, cases I and II (with 1-dimensional stable, and 2-dimensional unstable manifolds) are the same in 3{\mathbb{R}}^{3}. But this is not important for our proof.

If a one-sided intersection is preceded or followed by a two-sided intersection (see Figure 5, cases I, II, III), then we can assume that a two-sided intersection is arbitrarily close to a one-sided touching. Consequently, it is enough to find a transversal crossing arbitrarily close to a two-sided crossing.

\psfrag{1}{case I}\psfrag{2}{case II}\psfrag{3}{case III}\psfrag{4}{case IV}\psfrag{B}{$B$}\epsfbox{twoSided.eps}
Figure 5. Two-sided intersections of invariant manifolds (non-Newhouse types).
Theorem 3.2.

Let MM be a smooth manifold, and let f:MMf:M\to M be a diffeomorphism with a hyperbolic fixed point pMp\in M and with stable and unstable invariant manifolds WSW_{S} and WUW_{U} at pp. Assume:

  • (i)

    ff is locally C1C^{1}-linearizable in a neighborhood of pp;

  • (ii)

    WUW_{U} and WSW_{S} have complementary dimensions n1n-1 and 11;

  • (iii)

    WUW_{U} and WSW_{S} have a finite tangential contact at an isolated homoclinic point BMB\in M (BpB\neq p), which is not of Newhouse type;

  • (iv)

    WUW_{U} contains a graph portion Λ\Lambda associated with BB in a small neighborhood, contained in the neighborhood of linearization.

Then the invariant manifolds have a point of transverse intersection arbitrarily close to the point BB. Therefore, ff has a horseshoe (arbitrarily close to BB) according to the Birkhoff-Smale Theorem.

Proof.

Consider a C1C^{1} linearization ϕ:NN\phi:N^{\prime}\mapsto N from a neighborhood NMN^{\prime}\subset M of pp onto a neighborhood NnN\subset{\mathbb{R}}^{n} of the origin. We choose a linearization such that ff has a normal Jordan form in the new coordinates. We may assume that ϕ\phi takes the local stable manifold at pp to a neighborhood in the YY-axis, and the local unstable manifold to a neighborhood in the XX-hyperplane.

There are images of ϕ(B)\phi(B) in NN, under forward and backward iterates of ff, on both the local stable and local unstable manifolds of 0. There is a compact curve, ΓN\Gamma\subset N (see Figure 6), with the following properties:

  • Γ\Gamma is an image under ff of the stable manifold;

  • Γ\Gamma has finite contact with the XX-plane at some isolated point
    A=(A1,,An1,0)A=(A_{1},...,A_{n-1},0);

  • Γ\Gamma has two-sided intersection with the XX-plane at AA;

  • AA is arbitrarily close to the origin.

Any C1C^{1}-curve Γ\Gamma immersed in n{\mathbb{R}}^{n}, which has finite order of contact with the XX-hyperplane at point A=(A1,,An1,0)A=(A_{1},...,A_{n-1},0), locally can be parameterized as:

Γ={u1=a1(τ)+A1.........un1=an1(τ)+An1v=τl,\Gamma=\left\{\begin{array}[]{lll}u^{1}&=&a_{1}(\tau)+A_{1}\\ .&.&.\\ .&.&.\\ .&.&.\\ u^{n-1}&=&a_{n-1}(\tau)+A_{n-1}\\ v&=&{\tau}^{l},\end{array}\right.

for some constants l>1l>1, AjA_{j}, and C1C^{1} functions aja_{j}, such that aj(0)=0a_{j}(0)=0 and aj(τ)ra_{j}^{\prime}(\tau)\leq r in a small neighborhood of the origin, maxj|Aj|>0\max_{j}|A_{j}|>0 (j=1,,n1j=1,...,n-1).

Since Λ\Lambda is a graph portion, it can be parameterized as:

Λ={x1=t1.........xn1=tn1y=ϕ(t1,,tn1)+B,\Lambda=\left\{\begin{array}[]{lll}x^{1}&=&t_{1}\\ .&.&.\\ .&.&.\\ .&.&.\\ x^{n-1}&=&t_{n-1}\\ y&=&\phi(t_{1},...,t_{n-1})+B,\end{array}\right.

for some constant B0B\neq 0 and C1C^{1} function ϕ\phi, ϕ(0)=0\phi(0)=0 .

Now it suffices to refer to the following lemma, in order to have the proof of this theorem completed.

Lemma 3.3.

Suppose

f:(x,y)(β1x1+σ2x2,β2x2+σ3x3,,βn2xn2+σn1xn1,βn1xn1,αy),f:(x,y)\to(\beta_{1}x^{1}+\sigma_{2}x^{2},\beta_{2}x^{2}+\sigma_{3}x^{3},...,\beta_{n-2}x^{n-2}+\sigma_{n-1}x^{n-1},\beta_{n-1}x^{n-1},\alpha y),
0<α<1<βj(j=1,,n1)0<\alpha<1<\beta_{j}\quad(j=1,...,n-1)

and σk\sigma_{k} are 0 or 11 – according to the Jordan normal form of ff.

Then there exists a natural number kk_{*} such that fkΛf^{k}\Lambda crosses Γ\Gamma transversely if k>k.k>k_{*}.

\psfrag{xyn}{$(x_{n},y_{n})$}\psfrag{xy1}{$(x_{1},y_{1})$}\psfrag{f(xyn)}{$f^{n}(x_{n},y_{n})$}\psfrag{f(xy1)}{$f^{n}(x_{1},y_{1})$}\psfrag{T}{$f$}\psfrag{f(l)}{$f(\Lambda)$}\psfrag{f(f(l))}{$f^{2}(\Lambda)$}\psfrag{fn(l)}{$f^{n}(\Lambda)$}\psfrag{G}{$\Gamma$}\psfrag{o}{$0$}\psfrag{A}{$B$}\psfrag{aA}{$\alpha B$}\psfrag{a2A}{$\alpha^{2}B$}\psfrag{anA}{$\alpha^{n}B$}\psfrag{B}{$A$}\psfrag{l}{$\Lambda$}\psfrag{X}{$X$}\psfrag{Y}{$Y$}\epsfbox{par.eps}
Figure 6. Degenerate homoclinic crossing near the fixed point 0. Illustration of the proof.
Proof.

To simplify notations, first assume that ff has a diagonal Jordan normal form, i.e., all σk\sigma_{k} are 0. Consider the iterations of Λ\Lambda by ff:

fkiΛ={u1=β1kit1.........un1=βn1kitn1v=αki(ϕ(t1,,tn1)+B)f^{k_{i}}\Lambda=\left\{\begin{array}[]{lll}u^{1}&=&\beta_{1}^{k_{i}}t_{1}\\ .&.&.\\ .&.&.\\ .&.&.\\ u^{n-1}&=&\beta_{n-1}^{k_{i}}t_{n-1}\\ v&=&\alpha^{k_{i}}\left(\phi(t_{1},...,t_{n-1})+B\right)\\ \end{array}\right.

We can find sequences (xi,yi)Λ(x_{i},y_{i})\in\Lambda and kik_{i}\rightarrow\infty such that (xi,yi)(0,B)(x_{i},y_{i})\rightarrow(0,B) and fki(xi,yi)(A,0)f^{k_{i}}(x_{i},y_{i})\rightarrow(A,0). This implies that fkiΛf^{k_{i}}\Lambda and Γ\Gamma will cross at some point other than (0,B)(0,B) and (A,0)(A,0) (see Figure 6). Thus, we only need to show that this intersection is transversal.

Consider the matrix, constructed in the following way:

  • The first n1n-1 rows represent the linear space of the tangent hyperplane to the (n1)(n-1)-manifold fkiΛf^{k_{i}}\Lambda.

  • The last row represents the linear space of the tangent line to the curve Γ\Gamma.

In order to prove that fkiΛf^{k_{i}}\Lambda crosses Γ\Gamma transversely at some points of the form (u1(ti),,un1(ti),v(ti))(u^{1}(t_{i}),...,u^{n-1}(t_{i}),v(t_{i})), we show that the determinant of the matrix below is never 0 for ki{k_{i}} sufficiently large. We consider the following determinant DiD_{i}, which corresponds to the kik_{i}-th iteration of ff:

Di=|β1ki0αkit1ϕ(ti)0βn1kiαkitn1ϕ(ti)a1(τi)an1(τi)lτil1|D_{i}=\left|\begin{array}[]{ccccc}\beta_{1}^{k_{i}}&&0&&\alpha^{k_{i}}\frac{\partial}{\partial t_{1}}\phi(t_{i})\\ &&&&\\ &&&&\\ &\ddots&&&\vdots\\ &&&&\\ 0&&\beta_{n-1}^{k_{i}}&&\alpha^{k_{i}}\frac{\partial}{\partial t_{n-1}}\phi(t_{i})\\ &&&&\\ &&&&\\ a_{1}^{\prime}(\tau_{i})&\ldots&a_{n-1}^{\prime}(\tau_{i})&&l\tau_{i}^{l-1}\\ \end{array}\right|

Since Λ\Lambda intersects Γ\Gamma,

aj(τi)+Aj=βjkiti,j=1,,n1a_{j}(\tau_{i})+A_{j}=\beta_{j}^{k_{i}}t_{i},\ \ \ \ \ \ \ \ j=1,...,n-1

and

τil=αki(ϕ(ti)+B).\tau_{i}^{l}=\alpha^{k_{i}}(\phi(t_{i})+B).

Then, it follows that

τl1=αkil1l(ϕ(t(τ))+B)l1l\tau^{l-1}=\alpha^{k_{i}\frac{l-1}{l}}\left(\phi(t(\tau))+B\right)^{\frac{l-1}{l}}

for any small value of the parameter τ\tau, and the corresponding small value of the parameter t(τ)t(\tau).

We substitute the above expression in the matrix DiD_{i} and calculate its determinant:

Di=β1kiβn1ki[(1)n1lαkil1l(ϕ(t(τ))+B)l1lD_{i}=\beta_{1}^{k_{i}}...\beta_{n-1}^{k_{i}}\Bigg{[}(-1)^{n-1}l\alpha^{k_{i}\frac{l-1}{l}}(\phi(t(\tau))+B)^{\frac{l-1}{l}}
+j=1n1αkitjϕ(t(τ))(1)jaj(τ)βjki]\left.+\sum_{j=1}^{n-1}\alpha^{k_{i}}\frac{\partial}{\partial t_{j}}\phi(t(\tau))(-1)^{j}a_{j}^{\prime}(\tau)\beta_{j}^{-k_{i}}\right]

We can assume that ϕ(t(τ))+B>B/2\phi(t(\tau))+B>B/2. Also, aj(τ)a_{j}^{\prime}(\tau) in our parameterization of Γ\Gamma are bounded from above by rr.

By Lemma 2.14 for any ϵ>0\epsilon>0 there exists kk_{*} such that for any k>kk>k_{*}

αktjϕ(t(τ))<ϵ,\alpha^{k}\frac{\partial}{\partial t_{j}}\phi(t(\tau))<\epsilon,

for all j=1,n1j=1,...n-1. Then

|Di|>β1kiβn1ki[lαkil1l(B/2)l1lαkiϵ(n1)rβjki]|D_{i}|>\beta_{1}^{k_{i}}...\beta_{n-1}^{k_{i}}\left[l\alpha^{k_{i}\frac{l-1}{l}}(B/2)^{\frac{l-1}{l}}-\alpha^{k_{i}}\epsilon(n-1)r\beta_{j}^{-k_{i}}\right]
=β1kiβn1kiαkil1l[l(B/2)l1lαki/lϵ(n1)rβjki]=\beta_{1}^{k_{i}}...\beta_{n-1}^{k_{i}}\alpha^{k_{i}\frac{l-1}{l}}\left[l(B/2)^{\frac{l-1}{l}}-\alpha^{k_{i}/l}\epsilon(n-1)r\beta_{j}^{-k_{i}}\right]

Take

ϵ<l(B/2)l1lr(n1)\epsilon<\frac{l(B/2)^{\frac{l-1}{l}}}{r(n-1)}

Then,

[l(B/2)l1lαki/lϵ(n1)rβjki]\left[l(B/2)^{\frac{l-1}{l}}-\alpha^{k_{i}/l}\epsilon(n-1)r\beta_{j}^{-k_{i}}\right]

is positive for any sufficiently big iteration kik_{i}.


If ff is non-diagonalizable, using the Jordan normal form, one can see that

Di=β1kiβn1ki[(1)n1lαkil1l(ϕ(t(τ))+B)l1lD_{i}=\beta_{1}^{k_{i}}...\beta_{n-1}^{k_{i}}\left[(-1)^{n-1}l\alpha^{k_{i}\frac{l-1}{l}}(\phi(t(\tau))+B)^{\frac{l-1}{l}}\right.
+j=1n1αkitjϕ(t(τ))(1)jaj(τ)βjki\left.+\sum_{j=1}^{n-1}\alpha^{k_{i}}\frac{\partial}{\partial t_{j}}\phi(t(\tau))(-1)^{j}a_{j}^{\prime}(\tau)\beta_{j}^{-k_{i}}\right.
j=1n1αkitjϕ(t(τ))Pki1(βj)βjki],\left.-\sum_{j=1}^{n-1}\alpha^{k_{i}}\frac{\partial}{\partial t_{j}}\phi(t(\tau))P_{k_{i}-1}(\beta_{j})\beta_{j}^{-k_{i}}\right],

where Pki1(βj)P_{k_{i}-1}(\beta_{j}) are polynomial functions of degree ki1k_{i}-1.
Then,

|Di|>β1kiβn1kiαkil1l[l(B/2)l1l|D_{i}|>\beta_{1}^{k_{i}}...\beta_{n-1}^{k_{i}}\alpha^{k_{i}\frac{l-1}{l}}\left[l(B/2)^{\frac{l-1}{l}}-\right.
(αki/lϵ(n1)rβjki+αki/lϵ(n1)|Pki1(βj)|βjki)]\left.\left(\alpha^{k_{i}/l}\epsilon(n-1)r\beta_{j}^{-k_{i}}+\alpha^{k_{i}/l}\epsilon(n-1)\left|P_{k_{i}-1}(\beta_{j})\right|\beta_{j}^{-k_{i}}\right)\right]

and

[l(B/2)l1l(αki/lϵ(n1)rβjki+αki/lϵ(n1)|Pki1(βj)|βjki)]\left[l(B/2)^{\frac{l-1}{l}}-\left(\alpha^{k_{i}/l}\epsilon(n-1)r\beta_{j}^{-k_{i}}+\alpha^{k_{i}/l}\epsilon(n-1)\left|P_{k_{i}-1}(\beta_{j})\right|\beta_{j}^{-k_{i}}\right)\right]

is also non-zero, because αki/l\alpha^{k_{i}/l} tends to 0 exponentially, as kik_{i}\to\infty.

This implies that for some kk_{*}, the iterations fkif^{k_{i}} (ki>kk_{i}>k_{*}) applied to the manifold Λ\Lambda near the origin will produce a transverse crossing with Γ\Gamma.

Applying Lemma 3.3, we complete the proof of Theorem 3.2. ∎

\psfrag{p}{$0$}\psfrag{B}{$A$}\psfrag{U}{$W_{U}$}\psfrag{S}{$W_{S}$}\psfrag{L}{$\Lambda$}\psfrag{fL}{$f(\Lambda)$}\psfrag{ffL}{$f^{2}(\Lambda)$}\psfrag{G}{$\Gamma$}\epsfbox{theorem_counter_ex.eps}
Figure 7. Λ\Lambda is not a graph portion. Intersections of Γ\Gamma with fn(Λ)f^{n}(\Lambda) are indicated by a thicker line. Γ\Gamma and fn(Λ)f^{n}(\Lambda) have no transversal intersections.
Remark 3.4.

It is clear from the proof of Theorem 3.2 that, if WUW_{U} does not contain any graph portion, then WUW_{U} possibly has no transversal crossings with Γ\Gamma. Figure 7 shows that Γfn(Λ)\Gamma\cap f^{n}(\Lambda) can be a 1-dimensional manifold.

References

  • [1] V. I. Arnold, S. M. Guseĭn-Zade, A. N. Varchenko, Singularities of differentiable maps. Vol. I. The classification of critical points, caustics and wave fronts, Birkhäuser, Boston, 1985.
  • [2] I. U. Bronstein, A. Ya. Kopanskii, Smooth Invariant Manifolds and Normal Forms, World Scientific, Singapore, 1994.
  • [3] K. Burns, H. Weiss, A geometric criterion for positive topological entropy, Comm. Math. Phys., 172 (1995), 95-118.
  • [4] R. Churchill and D. Rod, Pathology in dynamical systems. III. Analytic Hamiltonians, J. Diff. Equations, 37 (1980), 23-38.
  • [5] N. Gavrilov, L. Silnikov On 3-dimensional dynamical systems close to systems with a structurally unstable homoclinic curve. I, Math. USSR Sb., 88 (1972), 467-485.
  • [6] N. Gavrilov, L. Silnikov On 3-dimensional dynamical systems close to systems with a structurally unstable homoclinic curve, Math. USSR Sb., 90 (1973), 139-156.
  • [7] S. Gonchenko, L. Shilnikov Dynamical systems with structurally unstable homoclinic curves, Soviet Math. Dokl., 33 (1986), 234-238.
  • [8] P. Hartman, On local homeomorphisms of Euclidean spaces, Bol. Soc. Mat. Mexicana (2), 5 (1960), 220-241.
  • [9] M. Hirsch, Degenerate homoclinic crossings in surface diffeomorphisms, preprint (1993).
  • [10] A. J. Homburg, H. Weiss, A geometric criterion for positive topological entropy. II: Homoclinic tangencies, Comm. Math. Phys., 208 (1999), 267-273.
  • [11] A. Katok, Lyapunov exponents, entropy and periodic orbits for diffeomorphisms, Publ. Math. IHES, 51 (1980), 137-173.
  • [12] A. Katok, Nonuniform hyperbolicity and structure of smooth dynamical systems, Proc. International Congress of Mathematicians, Warszawa, 2 (1983), 1245-1254.
  • [13] J. Moser, The analytic invariants of an area-preserving mapping near a hyperbolic fixed point, Comm. Pure Appl. Math., 9 (1956), 673-692.
  • [14] S. Newhouse and J. Palis, Cycles and bifurcation theory, Asterisque, 31, Soc. Math. France, Paris, 43-140 (1976).
  • [15] J. Palis, On Morse-Smale dynamical systems, Topology, 8 (1969), 385-405.
  • [16] V. Rayskin, Multidimensional singular λ\lambda-Lemma, Electron. J. Diff. Eqns., 38 (2003), 1-9.
  • [17] S. Smale, Diffeomorphisms with many periodic points, in:Differential and Combinatorial topology (edited by S. S. Cairnes), Princeton University Press, 63-80, (1965).
  • [18] R. Uribe-Vargas, Four-vertex theorems in higher dimensional spaces for a larger class of curves than the convex ones, C.R. Acad. Sci. Paris, Ser. I, 330 (2000), 1085-1090.

Received October 2003; revised September 2004.