\documentclass[reqno]{amsart} \usepackage{graphicx} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2003(2003), No. 38, pp.1--9. \newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu (login: ftp)} \thanks{\copyright 2003 Southwest Texas State University.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE--2003/38\hfil Multidimensional singular $\lambda$-lemma ] {Multidimensional singular $\lambda$-lemma} \author[Victoria Rayskin \hfil EJDE--2003/38\hfilneg] {Victoria Rayskin} \address{Victoria Rayskin \hfill\break Department of Mathematics\\ MS Bldg, 6363\\ University of California at Los Angeles, 155505\\ Los Angeles, CA 90095, USA} \email{vrayskin@math.ucla.edu} \date{} \thanks{Submitted November 4, 2002. Published April 11, 2003.} \subjclass[2000]{37B10, 37C05, 37C15, 37D10} \keywords{Homoclinic tangency, invariant manifolds, $\lambda$-Lemma, \hfill\break\indent order of contact, resonance} \begin{abstract} The well known $\lambda$-Lemma \cite{Pa} states the following: Let $f$ be a $C^1$-diffeomorphism of $\mathbb{R}^n$ with a hyperbolic fixed point at $0$ and $m$- and $p$-dimensional stable and unstable manifolds $W^S$ and $W^U$, respectively ($m+p=n$). Let $D$ be a $p$-disk in $W^U$ and $w$ be another $p$-disk in $W^U$ meeting $W^S$ at some point $A$ transversely. Then $\bigcup_{n\geq 0} f^n(w)$ contains $p$-disks arbitrarily $C^1$-close to $D$. In this paper we will show that the same assertion still holds outside of an arbitrarily small neighborhood of $0$, even in the case of non-transverse homoclinic intersections with finite order of contact, if we assume that $0$ is a low order non-resonant point. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{definition}[theorem]{Definition} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{claim}[theorem]{Claim} \section{Introduction} Let $M$ be a smooth manifold without boundary and $f:M\to M$ be a $C^1$ map that has a hyperbolic fixed point at the origin. The well known ${\lambda}$-Lemma \cite{Pa} gives an important description of chaotic dynamics. The basic assumption of this theorem is the presence of a transverse homoclinic point. \begin{theorem}[Palis]\label{Palis} Let $f$ be a $C^1$ diffeomorphism of $\bf R^n$ with a hyperbolic fixed point at $0$ and $m$- and $p$-dimensional stable and unstable manifolds $W^S$ and $W^U$ ($m+p=n$). Let $D$ be a $p$-disk in $W^U$, and $w$ be another $p$-disk in $W^U$ meeting $W^S$ at some point $A$ {\em transversely}. Then $\bigcup_{n\geq 0} f^n(w )$ contains $p$-disks arbitrarily $C^1$-close to $D$. \end{theorem} The assumption of transversality is not easy to verify for a concrete dynamical system. Obviously, the conclusion of the Theorem of Palis is not true for an arbitrary degenerate (non-transverse) crossing. Example by Newhouse illustrates this situation (See picture~\ref{newhousePic}). \begin{figure}[th] \begin{center} \includegraphics[width=0.7\textwidth]{fig1.eps} \end{center} \caption{Newhouse example. Branches of $W_U$ are not $C^1$-close near $0$} \label{newhousePic} \end{figure} In this paper we prove an analog of the ${\lambda}$-Lemma for the non-transverse case in arbitrary dimension. Suppose $W^S$ and $W^U$ are sufficiently smooth and cross non-transversally at an isolated homoclinic point, i.e. they have a {\it singular homoclinic crossing}. In Section~\ref{2} we define the order of contact for this crossing (Definition~\ref{def-ooc-manifold}) and show that it is preserved under a diffeomorphic transformation (Lemma~\ref{multidim-preserv-ooc-lemma}). We prove Singular ${\lambda}$-Lemma for the case of singular finite order homoclinic crossing of manifolds which have a graph portion (see Definition~\ref{def-p-p}), under non-resonance restriction. See Lemma~\ref{lemma} in Section~\ref{3}. \section{Definitions and Lemmas}\label{2} In this section we are considering two immersed $C^r$ manifolds in $\mathbb{R}^n$, $r>1$. Suppose they meet at an isolated point $A$. We will discuss the structure of these manifolds in the neighborhood of the point $A$. First, assume that each manifold is a curve. Hirsch in his work \cite{Hi} describes the order of contact for two curves and formulates the following definition: \begin{definition}\label{def-ooc-Hirsch} \rm Let ${\Lambda}_i$ ($i=1,2$) denote two immersed $C^r$ curves in $\mathbb{R}^2$ , $r>1$. Suppose the two curves meet at point $A$. Let $t\mapsto u_i(t)$ be a $C^r$ parameterization of ${\Lambda}_i$, both defined for $t$ in some interval $I$, with non-vanishing tangent vectors $u_i'(t)$. Suppose $0\in I$ and $A = u_i(0)$. The order of contact of the two curves at $A$ is the unique real number $l$ in the range $1\leq l \leq r$, if it exists, such that $u_1 - u_2$ has a root of order $l$ at $0$. \end{definition} For our higher-dimensional proof we can reformulate this definition for two curves in $\mathbb{R}^n$: \begin{definition}\label{def-ooc-curve} \rm Let ${\Lambda}_i$ ($i=1,2$) denote two immersed $C^r$ curves in $\mathbb{R}^n$, $r>1$. Suppose the two curves meet at point $A$. Let $t\mapsto u_i(t)$ be a $C^r$ parameterization of ${\Lambda}_i$, both defined for $t$ in some interval $I$, with non-vanishing tangent vectors $u_i'(t)$. Suppose $0\in I$ and $A = u_i(0)$. The order of contact of the two curves at $A$ is the unique real number $l$ in the range $1\leq l \leq r$, if it exists, such that $| u_1 - u_2|$ has a root of order $l$ at $0$. \end{definition} Now we can define the order of contact for two manifolds of arbitrary dimensions. \begin{definition}\label{def-ooc-manifold} \rm Let $W^S$ and $W^U$ denote two immersed $C^r$ manifolds in $\mathbb{R}^n$, $r>1$. Suppose the two manifolds meet at an isolated point $A$. The order of contact $\alpha$ at $A$ is the unique real number $\alpha$ in the range $1\leq \alpha \leq r$, if it exists, such that \begin{align*} \alpha = \sup \big\{& l| C^r\mbox{-curve } {\gamma}_1 \in W^S \mbox{ has order of contact } l \mbox{ with another }\\ &C^r\mbox{-curve } {\gamma}_2 \in W^U\ \mbox{and } A\in {\gamma}_1 \cap {\gamma}_2\big\} \end{align*} \end{definition} The order of contact is preserved under a diffeomorphism. This result is first proven for curves (Lemma~\ref{preserv-ooc-lemma}). \begin{lemma}\label{preserv-ooc-lemma} Consider a $C^\infty $ surface without boundary and a $C^r$ diffeomorphism $\phi$ that maps a neighborhood $N'$ of this surface onto some neighborhood $N \subset \bf R^2$. Assume that $u(t), v(t)$ are $C^r$ curves, such that $u(0)=v(0)$. Then, $\phi$ preserves the order of contact of these curves. \end{lemma} \begin{proof} Without lost of generality, we assume that $u(0)=v(0)=0$. We have curves \[ \phi \circ u(t),\quad \phi \circ v(t), \] transformed by the diffeomorphism $\phi$. There are positive constants $m$ and $M$ such that \[ m\leq \frac {| u(t)-v(t)| }{| t|^{l}} \leq M, \quad\mbox{as }t\to 0. \] By the $ C^1 $ Mean Value Theorem, \[ \phi (x)- \phi (y) =\Big[ \int_{0}^{1}(D \phi )_{\sigma (s)} ds \Big] (x-y), \] where $\sigma (s) =(1-s)x +sy$. Then \[ (\phi\circ u )(t)-(\phi\circ v )(t)=\Big[ \int_{0}^{1}(D \phi )_{\sigma (s)} d s \Big] (u(t)-v(t)), \] where $\sigma (s) =(1-s)u(t)+sv(t)$. Therefore, \[ \frac{(\phi\circ u )(t)-(\phi\circ v )(t)}{t^l}=\Big[ \int_{0}^{1}(D \phi )_{ \sigma (s)} ds \Big] ( \frac{u(t)-v(t)}{t^l} ) \] As $t \to 0$, $\sigma (s) \to u(0)$ and the matrix $ \int_{0}^{1}(D \phi )_{\sigma (s)} ds $ tends to the invertible matrix $ (D \phi )_{u(0)} $. The ratio $ \frac{u(t)-v(t)}{t^l} $ is a vector whose norm is bounded by $M$ and $m$, $00$, for an arbitrarily small $\epsilon$-neighborhood $\mathcal{U} \subset \mathbb{R}^n$ of the origin and for the graph portion ${\Lambda}$, $(\bigcup_{n\geq 0} f^n({\Lambda} ))\setminus \mathcal{U}$ contains disks $\rho$-$C^1$ close to $\mathcal{V} \setminus \mathcal{U}$. \end{lemma} \begin{remark} \label{rmk3} \rm There is no loss of generality to assume that $p\leq m$, because we can always replace $f$ with $f^{-1}$. \end{remark} \begin{figure}[th] \begin{center} \setlength{\unitlength}{1mm} \begin{picture}(70,54)(-8,-6) \put(-5,0){\line(1,0){65}} \put(59,-.85){$\to$} \put(0,-5){\line(0,1){50}} \put(-.85,45){$\downarrow$} \put(-.85,-5){$\uparrow$} \qbezier(0,10)(0,15)(60,15) \qbezier(0,20)(2,28)(40,28) \qbezier(0,30)(4,38)(20,40) \put(2,45){Y} \put(-4,29){$A$} \put(-8,19){$f(A)$} \put(-13,9){$f(f(A))$} \put(-8,1.5){$W_U$} \put(1,-6){$W_S$} \put(17,42){$\Lambda(x)$} \put(30,30){$f(x,\Lambda(x))$} \put(45,17){$f(f(x,\Lambda(x)))$} \put(60,-4){$X$} \put(-7,-.85){$\leftarrow$} \end{picture} \end{center} \caption{Iterations of the graph portion $\Lambda$ with the diffeomorphism $f$} \end{figure} \begin{proof}[Proof of Lemma \ref{lemma}] Let $\alpha = 1/l$ ($0 < \alpha <1$). Since ${\Lambda}$ is a graph portion that has finite order of contact with $W^S$, we can assume that locally ${\Lambda}$ is represented by the graph of the following form: \[ {\Lambda}(x) = A+r(x):\mathbb{R}^p \to \mathbb{R}^m, \quad r(0)=0, \] and for any sufficiently small $\sigma >0$ \[ | r(x)| \leq \mathop{\rm const} \cdot | x|^\alpha\quad \mbox{and}\quad | \frac{\partial}{\partial x_i}r(x)| \leq \mathop{\rm const} \cdot | x|^{\alpha -1}\ \] for all $| x| < \sigma$, $ i=1,\dots,p$. Let $x=(x_1,\dots,x_p)\in \bf R^p$, $y=(y_1,\dots,y_m)\in \bf R^m$ ($p+m=n$) and $f(x,y): \bf R^n \to \bf R^n $ has the linear part $$ ((\mathcal{A}x )_1,\dots,(\mathcal{A}x )_p,(\mathcal{B}y )_1, \dots,( \mathcal{B}y )_m). $$ Assume that $\| \mathcal{A}^{-1}\|$, $\| \mathcal{B} \| < {\lambda} <1$. Choose an arbitrarily small $\Delta$. If there is a cross terms $\mathop{\rm const}\cdot x_i y_j$ in the power expansion of this map around $0$, then we assume one-and-a-half-order non-resonance condition. Then, by Flattening Theorem (See \cite{ABZ}) there exists smooth change of coordinates, such that locally $f$ can be written in the form $f(x,y)=(S_1(x,y), S_2(x,y))$, where \begin{align*} S_1(x,y)= &\Big( \big( (\mathcal{A}x)_1 + {\phi_1}(x) + \sum_{i=1,\dots,p;j=1,\dots,m}x_i y_j U_{ij}^1 (x,y)\big),\dots ,\\ &\big( (\mathcal{A}x)_p + {\phi_p}(x) + \sum_{i=1,\dots,p;j=1,\dots,m}x_i y_jU_{ij}^p (x,y)\big) \Big) \end{align*} and \begin{align*} S_2(x,y)= &\Big( \big( (\mathcal{B}y)_1 + {\psi_1}(y) + \sum_{i=1,\dots,p;j=1,\dots,m}x_i y_j V_{ij}^1 (x,y)\big),\dots ,\\ &\big( (\mathcal{B}y)_m + {\psi_m}(y) + \sum_{i=1,\dots,p;j=1,\dots,m}x_i y_jV_{ij}^m (x,y)\big)\Big) . \end{align*} Here $U(0) = V(0) = 0$, $\|{\phi} \|_{C^1},\; \| {\psi} \|_{C^1},\; \| U\|_{C^0},\; \| V \|_{C^0}\leq \Delta$, and $\| U\|_{C^1},\ \| V \|_{C^1}$ are bounded. Consider $f(x, {\Lambda}(x))=(T_1^{\Lambda}(x),T_2^{\Lambda}(x))$. We will work with $(x,T_2^{\Lambda} \circ (T_1^{\Lambda})^{-1}(x))$ and deduce that $f^n(x, {\Lambda}(x))$ is $C^1$-small for $n$ big enough and $\sigma >0$ sufficiently small. First we will show that in $C^1$-topology $(T_1^{\Lambda})^{-1}$ is $\Delta$-close to $\mathcal{A}^{-1}$. For simplicity we will denote $T_1^{\Lambda}$ by $T_1$ and $T_2^{\Lambda}$ by $T_2$. \begin{align*} T_1(x)= &\Big( (\mathcal{A}x)_1 + {\phi_1}(x) + \sum_{i=1,\dots,p;j=1,\dots,m}x_i {\Lambda}_j(x) U_{ij}^1 (x,{\Lambda}(x)),\dots, \\ &(\mathcal{A}x)_p + {\phi_p}(x) + \sum_{i=1,\dots,p;j=1,\dots,m}x_i {\Lambda}_j(x)U_{ij}^p (x,{\Lambda}(x)) \Big). \end{align*} \begin{claim} \[ \| \sum_{i=1,\dots,p;j=1,\dots,m}x_i {\Lambda}_j(x) U_{ij}^t (x,{\Lambda}(x))\|_{C^1} < K\cdot\Delta \] for $| x | < \sigma$ ($\sigma >0 $ sufficiently small, $K >0$). \end{claim} \begin{proof} Fix some $l\in \{1,\dots,p\}$. Recall that $\Lambda (x) = A + r(x)$. \begin{align*} \big|\frac{\partial}{\partial x_l}x_i {\Lambda}_j(x)\big| &\leq \delta_{il}| {\Lambda}(x)| + | x_i| \cdot| \frac{\partial}{\partial x_l}{\Lambda}_j(x)|\\ &\leq \delta_{il}(| A|+ | x|^\alpha )+| x| \cdot O(1) | x|^{\alpha-1}\\ &\leq | A|\delta_{il}+(\delta_{il}+O(1))| x|^\alpha = O(1) \end{align*} Here \[ \delta_{il}= \begin{cases} 1 & \mbox{if } i=l,\\ 0 & \mbox{if } i\neq l. \end{cases} \] Through the proof of this Theorem, $O(1)$ will be the set \begin{align*} O(1) =& \big\{ \gamma (\zeta ) : \mathbb{R} \mapsto \mathbb{R} \mbox{ such that there exists a positive constant $c$ with }\\ &| \gamma (\zeta )| \leq c \mbox{ for all sufficiently small }\zeta \big\} \end{align*} Also, \[ \big| \frac{\partial}{\partial x_l}U_{ij}^t(x,{\Lambda}(x))\big| = \big| \frac{\partial}{\partial x_l}U_{ij}^t(x,y) + \sum_{k=1}^{m} \frac{\partial}{\partial y_k}U_{ij}^t(x,y) \cdot \frac{\partial}{\partial x_l}{\Lambda}_k(x)\big| = O(1)\,. \] Therefore, \begin{align*} &\big\| \sum_{i=1,\dots,p;j=1,\dots,m}x_i {\Lambda}_j(x) U_{ij}^t (x,{\Lambda}(x))\big\|_{C^1} \\ &\leq \sum_{i=1,\dots,p;j=1,\dots,m}\big| \sum_{l=1}^{p}\frac{\partial}{\partial x_l}(x_i {\Lambda}_j(x) U_{ij}^t (x,{\Lambda}(x)))\big| \\ &\leq \sum_{i=1,\dots,p;j=1,\dots,m}\sum_{l=1}^{p}\big|\frac{\partial}{\partial x_l}(x_i {\Lambda}_j(x))\cdot U_{ij}^t (x,{\Lambda}(x)) + x_i {\Lambda}_j(x)\cdot \frac{\partial}{\partial x_l}U_{ij}^t (x,{\Lambda}(x))\big|\\ &\leq \Delta \cdot O(1), \end{align*} if $\sigma$ is sufficiently small and $| x| < \sigma$ (Arbitrarily small $\Delta$ was chosen above). The estimate proves the claim. \end{proof} Now, we continue the proof of Lemma~\ref{lemma}. As it was noted earlier in the proof, $\|{\phi} \|_{C^1}\leq \Delta$, by Flattening Theorem. This estimate and the assertion of the Claim imply that $\| \mathcal{A} - T_1 \|_{C^1} = O(1) \cdot \Delta$. This obviously implies $\| \mathcal{A}^{-1} - T_1^{-1} \|_{C^1} = O(1) \cdot \Delta$. Now we can do the main estimate, -- the estimate for $\| T_2 \circ T_1^{-1} \|_{C^k} $ $(k=0,1)$. \begin{align*} T_2 \circ T_1^{-1} = &\Big(( \mathcal{B} {\Lambda}( T_1^{-1}))_1 + {\psi_1}({\Lambda}( T_1^{-1})) \\ &+\sum_{i=1,\dots,p;j=1,\dots,m}( T_1^{-1})_i ( {\Lambda} ( T_1^{-1}))_j V_{ij}^1 ( T_1^{-1},{\Lambda}( T_1^{-1})),\dots,\\ &\quad ( \mathcal{B} {\Lambda}( T_1^{-1}))_m + {\psi_m}({\Lambda}( T_1^{-1}))\\ &+ \sum_{i=1,\dots,p;j=1,\dots,m}( T_1^{-1})_i ( {\Lambda} ( T_1^{-1}))_j V_{ij}^m ( T_1^{-1}, {\Lambda}( T_1^{-1}))\Big) \end{align*} We will begin by estimating each term of this vector. \[ \mathcal{B} {\Lambda}( T_1^{-1}) = \mathcal{B} \cdot A + \mathcal{B} \cdot r( T_1^{-1}(x)). \] \[ | \mathcal{B} \cdot r( T_1^{-1}(x)) | = O(1) \cdot \| \mathcal{B}\| | T_1^{-1}(x)|^\alpha = O(1) \cdot \| \mathcal{B}\| ( \| \mathcal{A}^{-1} \| +\Delta )^ \alpha | x|^\alpha. \] By the chain rule, \begin{align*} &\big| \frac{\partial}{\partial x_l}\mathcal{B} \cdot r( T_1^{-1}(x)) \big|\\ &=O(1) \cdot\| \mathcal{B}\| \| T_1^{-1}\|_{C^1} | T_1^{-1}(x)|^{\alpha -1}\\ &=O(1) \cdot \| \mathcal{B}\| ( \| \mathcal{A}^{-1}\| + \Delta ) ( \| \mathcal{A}^{-1}\| + \Delta )^{\alpha -1} | x |^{\alpha -1}\\ &=O(1) \cdot \| \mathcal{B}\| ( \| \mathcal{A}^{-1}\| + \Delta )^\alpha | x |^{\alpha -1}\\ &= O(1) \cdot{\lambda} | x |^{\alpha -1} \end{align*} with $\lambda <1$. Moreover, \[ | \frac{\partial}{\partial x_l}\mathcal{B}^n \cdot r( T_1^{-n}(x)) | =O(1) \cdot\| \mathcal{B}\|^n ( \| \mathcal{A}^{-1}\|^n + \Delta )^\alpha | x |^{\alpha -1} = O(1) \cdot {\lambda}^n | x |^{\alpha -1} \] This term can be made small if we perform enough iterations by the map $f$. I.e., $( \mathcal{B}^n {\Lambda}T_1^{-n})_m$ is $C^1$-small outside of a fixed neighborhood of $0$, if $n$ is big enough. For the estimates of the next term one can use the following expansion: \[ {\psi_1}({\Lambda}( T_1^{-1}(x))) = {\psi_1} ( A + r ( T_1^{-1}(x)) ) = {\psi_1} ( A ) + D{\psi_1} ( A ) \cdot r( T_1^{-1}(x)) + R ( T_1^{-1}(x) ), \] where $R ( T_1^{-1}(x) ) = o(|( T_1^{-1}(x))^\alpha)|$. Here the set $o(1)$ is the following set of functions: \begin{align*} o(1) =& \big\{ \gamma (\zeta ) : \mathbb{R} \mapsto \mathbb{R} \mbox{ such that for any positive constant $c$}\\ &\mbox{and for all sufficiently small }\zeta <\sigma , | \gamma (\zeta )| < c \big\} \end{align*} Similar to the previous calculations ${\psi_1}({\Lambda}( T_1^{-1}(x)))$ can be made small in $C^1$-norm if we perform enough iterations with the map $f$. Finally, we will note that the last term \[ \sum_{i=1,\dots,p;j=1,\dots,m}( T_1^{-1})_i ( {\Lambda} ( T_1^{-1}))_j V_{ij}^t ( T_1^{-1},{\Lambda}( T_1^{-1})) \] can be written as a composition $ \Sigma^t \circ T_1^{-1}(x)$, where \[ \Sigma^t (x) = \sum_{i=1,\dots,p;j=1,\dots,m}x_i {\Lambda}_j(x) V_{ij}^t (x, {\Lambda}(x)) . \] Consider $\frac{\partial}{\partial x_l}\Sigma^t \circ T_1^{-1}(x)$. \[ \frac{\partial}{\partial x_l}\Sigma^t \circ T_1^{-1}(x) = \sum_{i=1}^{p} \frac{\partial}{\partial x_i}\Sigma^t \circ T_1^{-1}(x) \cdot \frac{\partial}{\partial x_l}(T_1^{-1}(x))_i. \] We have already shown that \[ \big\| \sum_{i=1,\dots,p;j=1,\dots,m}x_i {\Lambda}_j(x) U_{ij}^t (x,{\Lambda}(x))\big\|_{C^1}= O(1) \cdot \Delta . \] Similar, one can show that \[ \| \Sigma^t \|_{C^1}= \big\| \sum_{i=1,\dots,p;j=1,\dots,m}x_i {\Lambda}_j(x) V_{ij}^t (x,{\Lambda}(x))\big\|_{C^1}= O(1) \cdot \Delta . \] Also, \[ \| T_1^{-1} \|_{C^1} \leq \| \mathcal{A}^{-1} \|_{C^1} + \| T_1^{-1}-\mathcal{A}^{-1} \|_{C^1} \leq \| \mathcal{A}^{-1} \|_{C^1} + O(1) \cdot \Delta . \] The estimates on $\| \Sigma^t \|_{C^1}$ and $\| T_1^{-1} \|_{C^1}$, together with the fact that $T(0)= 0$, imply that \[ \| \Sigma^t \circ T_1^{-1} \|_{C^1} = O(1) \cdot \Delta . \] Thus, for any small positive number $\rho$ and for any small (but bigger than a fixed $\epsilon$) $| x |$ one can find $n$ such that $(x,(T_2^{\Lambda})^{n} \circ (T_1^{\Lambda})^{-n}(x))$ is $\rho$-$C^1$-close to $\mathcal{V}$. This implies that for any $\rho >0$ and for an arbitrarily small $\epsilon$-neighborhood $\mathcal{U} \subset \mathbb{R}^n$ of the origin, $(\bigcup_{n\geq 0} f^n({\Lambda} ))\setminus \mathcal{U}$ contains $p$-disks $\rho$-$C^1$-close to $\mathcal{V} \setminus \mathcal{U}$. \end{proof} \begin{thebibliography}{0} \bibitem{Ha} P. Hartman, {\em On local homeomorphisms of Euclidean spaces}, Boletin Sociedad Matematica Mexicana (2), 5, 220-241 (1960). \bibitem{Hi} M. Hirsch, {\em Degenerate homoclinic crossings in surface diffeomorphisms}, (preprint) (1993). \bibitem{Pa} J. Palis, {\em On Morse-Smale dynamical systems}, Topology, {\bf 8}, 385-405 (1969). \bibitem{ABZ} S. Aranson, G. Belitsky, E. Zhuzhoma, Introduction to the qualitative theory of dynamical systems on surfaces, Translations of Mathematical Monographs, Volume 153, 1996. \end{thebibliography} \end{document}