1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%
3: %% Time-stamp: <30-08-2007 16:25:13 /home/lbutler/research-laptop/research/so4/t4.tex>
4: %%
5: %% Title: Positive-Entropy Geodesic Flows on Nilmanifolds
6: %%
7: %% Author: Leo Butler, 6214 JCMB, School of Maths
8: %% University of Edinburgh
9: %% Edinburgh, UK, EH9 3JZ
10: %% l . butler at ed . ac . uk
11: %% www.maths.ed.ac.uk/~lbutler
12: %%
13: %% Vassili Gelfreich
14: %%
15: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
16:
17: \documentclass[a4paper]{amsart}
18: \usepackage{amssymb,graphicx,rotating,hyperref}
19: %\usepackage{showkeys}
20: \usepackage[all]{xy}
21: \xyoption{import}
22:
23:
24: \def\d{{\rm d}}
25: \def\sech{\,{\rm sech}}
26: \def\t{{\mathfrak t}}
27: \def\g{{\mathfrak g}}
28: \def\R{{\bf R}}
29: \def\Z{{\bf Z}}
30: \def\ad#1{{\rm ad}_{#1}}
31: \def\Ad#1{{\rm Ad}_{#1}}
32: \def\vf{{\mathcal X}}
33: \def\ds{\displaystyle}
34: \def\ie{{\it i.e.\ }}
35: \def\im{{\rm{Im}\, }}
36: \def\rmt#1{{\rmfamily\tiny #1}}
37: \def\H{{\sf H}}
38:
39: \newtheorem{theorem}{Theorem}[section]
40: \newtheorem{thm}{Theorem}[section]
41: \newtheorem{lemma}[thm]{Lemma}
42:
43: \begin{document}
44:
45: \title{Positive-Entropy Geodesic Flows on Nilmanifolds}
46: \author{Leo T. Butler, Vassili Gelfreich}
47: \date{\today}
48:
49:
50:
51:
52: \begin{abstract} Let $T_n$ be the nilpotent group of real $n
53: \times n$ upper-triangular matrices with $1$s on the diagonal. The
54: hamiltonian flow of a left-invariant hamiltonian on $T^*T_n$ naturally
55: reduces to the Euler flow on $\t_n^*$, the dual of $\t_n = {\rm
56: Lie}(T_n)$. This paper shows that the Euler flows of the standard
57: riemannian and subriemannian structures of $T_4$ have transverse
58: homoclinic points on all regular coadjoint orbits. As a corollary,
59: left-invariant riemannian metrics with positive topological entropy
60: are constructed on all quotients $D \backslash T_n$ where $D$ is a
61: discrete subgroup of $T_n$ and $n \geq 4$.
62: \end{abstract}
63:
64: \keywords{Geodesic flows, entropy, nilmanifolds, nonintegrability,
65: subriemmannian geometry}
66: %\classification{58F17, 53D25, 37D40}
67:
68:
69: \maketitle
70:
71:
72:
73:
74: \section{Introduction}
75: \label{int}
76:
77: Let $\Sigma$ be a nilmanifold, \ie homogeneous space of a connected
78: nilpotent Lie group $G$. Each homogeneous riemannian metric on $G$
79: induces a locally-homogeneous metric on $\Sigma$. These riemannian
80: geometries, which will be called {\it left-invariant}, are of interest
81: in both geometry and dynamics. A basis question is
82: %%
83: %% Let $G$ be a connected nilpotent Lie group, $D < G$ a discrete
84: %% subgroup and ${\bf g}$ a left-invariant metric on $G$. The manifold
85: %% $\Sigma = D \backslash G$ is a {\it nilmanifold}; the metric (and
86: %% geodesic flow) induced by ${\bf g}$ on $\Sigma$ is said to be
87: %% left-invariant. A basic question concerning the dynamics of a
88: %% left-invariant geodesic flow on $\Sigma$ is:
89: %%
90:
91: \medskip
92: \noindent
93: {\bf Question A}: {\it Which left-invariant geodesic flows on a
94: compact nilmanifold have zero topological entropy?}
95: \medskip
96:
97: A mistaken answer to question A appears in Theorem 3
98: of~\cite{Manning}. In~\cite{Butler:2003a}, the first author showed
99: that on $2$-step nilmanifolds, all left-invariant geodesic flows have
100: zero entropy. In~\cite{Butler:2003b}, metrics on compact quotients of
101: the $3$-step nilpotent Lie group $T_4 \oplus T_3$ are constructed
102: whose geodesic flows have positive topological entropy. The paper also
103: speculated that the standard geodesic flow on $T_4$ also had such
104: horseshoes. Montgomery, Shapiro and Stolin~\cite{Montgomeryetal}
105: investigated the standard {\em subriemannian} geodesic flow on
106: $T_4$~\cite{Butler:2003a}; they showed that it reduces to the
107: Yang-Mills hamiltonian flow which is known to be algebraically
108: non-integrable~\cite{Zig1,Zig2}.
109:
110: Let us state the first result of the present paper. The Lie algebra of
111: $T_4$, $\t_4$, has the standard basis consisting of those $4\times 4$
112: matrices $X_{ij}$ with a unit in the $i$-th row and $j$-th column, $i
113: < j$, and zeros everywhere else. A quadratic hamiltonian $h : \t_4^*
114: \to \R$ is {\em diagonal} if it is expressed as $h(p) = \sum_{i<j}
115: a_{ij} \langle p,X_{ij} \rangle^2$ for some constants $a_{ij}$. The
116: standard riemannian metric has $a_{ij}=1$ for all $i,j$; the standard
117: Carnot (subriemannian) metric has $a_{12}=a_{23}=a_{34}=1$ and all other
118: coefficients zero.
119:
120: \begin{theorem} \label{thm:main}
121: If $h : \t_4^* \to \R$ is a diagonal hamiltonian with
122: $a_{12}a_{13}a_{23}a_{34} \neq 0$ and $a_{13}a_{34}=a_{12}a_{24}$,
123: then for all but at most countably many regular coadjoint orbits in
124: $\t_4^*$, the Euler vector field of $h$ has a horseshoe. In
125: particular, the Euler vector field of the standard riemannian metric
126: (resp. sub-riemannian metric with $a_{13} \neq 0$) is analytically
127: non-integrable.
128: \end{theorem}
129:
130:
131: \medskip
132: The condition that $a_{13}a_{34}=a_{12}a_{24}$ is only a device to
133: simplify the proof: all nearby hamiltonians also have a horseshoe. We
134: also show, by means of a numerical computation of an integral (see
135: Table \ref{tab:1} in section \ref{sec:deg}), that when $a_{13}=0$,
136: the conclusions of Theorem \ref{thm:main} hold. This shows that the
137: subriemannian geodesic flow of Montgomery, Shapiro and Stolin is
138: real-analytically non-integrable. Related numerical computations
139: (figure \ref{fig:1} in section \ref{sec:deg}) suggest that the
140: hamiltonian $h$ has a horseshoe on {\em every} regular coadjoint orbit
141: provided only that $a_{12}a_{23}a_{34} \neq 0$ and
142: $a_{13}a_{34}=a_{12}a_{24}$. Let us formulate a corollary to Theorem
143: \ref{thm:main}: Let $D < T_n$ be a discrete subgroup of $T_n$, $\Sigma
144: = D \backslash T_n$ and $S\Sigma$ is the unit sphere bundle.
145:
146: \medskip
147: \begin{theorem} \label{thm:cor1}
148: If $n \geq 4$, then there is a left-invariant geodesic flow $\phi_t :
149: S\Sigma \to S\Sigma$ such that $h_{top}(\phi_1) > 0$.
150: \end{theorem}
151:
152: \medskip
153: It appears likely that {\em all} left-invariant geodesic flows on
154: $S\Sigma$ have positive topological entropy and are non-integrable
155: with smooth integrals.
156:
157: \medskip
158:
159: Theorem \ref{thm:cor1} is interesting from a riemannian point of
160: view. Let $(M,g)$ be a smooth ($C^{\infty}$) riemannian manifold, and
161: $\phi_t : SM \to SM$ the geodesic flow of $g$. For each $T > 0$ and
162: $p,q \in M$, let $n_T(p,q)$ denote the number of distinct geodesics of
163: length no more than $T$ that join $p$ to $q$. Ma\~{n}\'e~\cite{Man}
164: showed that if $M$ is compact then $h_{top}(\phi_1 | SM) = \lim_{T \to
165: \infty} T^{-1} \log \int_{M \times M}\ n_T(p,q)\, dpdq$. Thus, for the
166: geodesic flows constructed here, for generic points $p$ and $q$ on a
167: compact quotient of $T_n$, $n_T(p,q)$ grows exponentially fast. In
168: constrast, Karidi showed that the volume growth on the universal cover
169: of these manifolds is polynomial of degree
170: $\frac{1}{6}n(n^2-1)$~\cite{Kar}.
171:
172: Theorem \ref{thm:main} is proved by reducing the hamiltonian flow of
173: $\phi_t$ on $T^* \Sigma$ to a hamiltonian flow on the coadjoint orbits
174: of ${\rm Lie(}T_4{\rm )}^*$. In the appropriate coordinate system, the
175: reduced hamiltonian is a small perturbation of a hamiltonian on $\R^4$
176: that is the sum of an unforced Duffing hamiltonian and a forced linear
177: system whose solutions can be expressed in terms of Legendre
178: functions. The Poincar\'e-Melnikov technique developed
179: in~\cite{MarsdenHolmes:1983a,Robinson:1988a} for autonomous
180: Hamiltonian systems is adapted here to show that a suspended Smale
181: horseshoe appears in the perturbed hamiltonian flow.
182:
183:
184: \section{The construction on Lie($T_4$)${}^*$}
185:
186: In this section, we will first recall a number of key facts about
187: geodesic flows and left-invariant hamiltonian systems on the cotangent
188: bundle of a Lie group; for more details, see~\cite{GS}. We will
189: then reduce the equations of motion of a left-invariant geodesic flow
190: on $T^* T_4$ to the equations of motion of a hamiltonian system on
191: $T^*\R^2$.
192:
193: \subsection{Poisson geometry of left-invariant hamiltonians}
194:
195: A {\em Poisson manifold} is a smooth manifold $M$ such
196: that $C^{\infty}(M)$ is equipped with a skew-symmetric bracket $\{,\}$
197: that makes $( C^{\infty}(M), \{,\} )$ into a Lie algebra of
198: derivations of $C^{\infty}(M)$. The centre of $( C^{\infty}(M), \{,\}
199: )$ is traditionally called the set of {\em Casimirs}. If $f$ is a
200: Casimir then $X_f \equiv 0$ and $f$ is a first integral of all
201: hamiltonian vector fields. If the set of Casimirs of $( C^{\infty}(M),
202: \{,\} )$ are the constant functions, then we say that $(
203: C^{\infty}(M), \{,\} )$ is a {\em symplectic} manifold. In this case,
204: the Poisson bracket naturally induces a closed, non-degenerate skew
205: $2$-form on $M$ which is called a symplectic structure. We will say
206: that a smooth map $f : M \to N$ is a {\em Poisson map} if $f^* : (
207: C^{\infty}(N), \{,\}_N ) \to ( C^{\infty}(M), \{,\}_M )$ is a Lie
208: algebra homomorphism.
209:
210: The most basic example of a Poisson manifold that is also symplectic
211: is provided by $T^* \R = \{ (a,A)\ : a,A \in \R\}$ equipped with the
212: Poisson bracket satisfying $\{a,A\}_{T^* \R} = 1$.
213:
214: The dual space of a Lie algebra gives an example of a Poisson manifold
215: that is not (in general) a symplectic manifold. Let $\g$ be a
216: finite-dimensional real Lie algebra and let $\g^*$ be the dual vector
217: space of $\g$. $T^*_p \g^*$ is identified with $\g$ for all $p \in
218: \g^*$. The Poisson bracket on $\g^*$ is defined for all $f,h \in
219: C^{\infty}(\g^*)$ and $p \in \g^*$ by
220: \begin{equation} \label{eq:pb}
221: \{f,h\}(p) := - \langle p, [ df_p, dh_p ] \rangle,
222: \end{equation}
223: where $\langle \cdot,\cdot \rangle : \g^* \times \g \to \R$ is the
224: natural pairing. Recall that for $\xi \in \g$, $\ad{\xi}^* : \g^* \to
225: \g^*$ is the linear map defined by $\langle \ad{\xi}^*p, \eta \rangle
226: = -\langle p, [\xi,\eta] \rangle$. $\ad{}^* : \g \to gl(\g^*)$ is the
227: representation contragredient to the adjoint representation. For any
228: $h \in C^{\infty}(\g^*)$, the hamiltonian vector field $E_h = \{\cdot
229: , h\}$ equals $-\ad{dh_p}^* p$. The standard example of a hamiltonian
230: vector field is obtained from a positive-definite linear map $\phi :
231: \g^* \to \g$ by setting $h(p) = \frac{1}{2} \langle p, \phi(p)
232: \rangle$, in which case $E_h(p) = -\ad{\phi(p)}^* p$.
233:
234: Let $G$ be a connected Lie group whose Lie algebra is $\g$. The
235: adjoint representation of $G$ on $\g$, $\Ad{g}\xi =
236: \frac{d}{dt}|_{t=0}\, g \exp(t\xi) g^{-1}$, induces the coadjoint
237: representation $\langle \Ad{g}^* p, \xi \rangle = \langle p,
238: \Ad{g^{-1}} \xi \rangle$ for all $p \in \g^*$, $g \in G$ and $\xi \in
239: \g$. As each vector field $p \to \ad{\xi}^* p$ is hamiltonian on
240: $\g^*$, with linear hamiltonian $h_{\xi}(p) = - \langle p, \xi
241: \rangle$, the coadjoint action of $G$ on $\g^*$ preserves the Poisson
242: bracket. The orbits of the coadjoint action are called the coadjoint
243: orbits. Each coadjoint orbit is a homogeneous $G$-space, and {\em
244: every} hamiltonian vector field on $\g^*$ is tangent to each coadjoint
245: orbit. For this reason, the Poisson bracket $\{,\}_{\g^*}$ restricts
246: to each coadjoint orbit, and is non-degenerate on each coadjoint
247: orbit. Thus, the coadjoint orbits are naturally symplectic
248: manifolds. A Casimir is necessarily constant on each coadjoint orbit,
249: and in many cases (as in this paper) each coadjoint orbit is the
250: common level set of all Casimirs.
251:
252: The Poisson bracket on $\g^*$ also arises in a natural way from the
253: Poisson bracket on $T^* G$. The group $G$ acts from the left on $T^*
254: G$, and this action preserves the Poisson structure
255: $\{,\}_{T^*G}$. The set of smooth left-invariant functions
256: $C^{\infty}(T^* G)^G$ is therefore a Lie subalgebra of $C^{\infty}(T^*
257: G)$ with respect to $\{,\}_{T^*G}$. This subalgebra is naturally
258: identified with $( C^{\infty}(\g^*), \{,\}_{\g^*})$ as follows: the
259: left-trivialization of $T^* G = G \times \g^*$ induces the projection
260: map $r : T^* G \to \g^*$ onto the second factor; $r^* C^{\infty}(\g^*)
261: = C^{\infty}(T^* G)^G$ and $r^*$ is a Lie algebra monomorphism.
262:
263:
264: The hamiltonian flow of a left-invariant hamiltonian $H$ on $T^* G$
265: therefore has the equations of motion:
266: \begin{equation}
267: X_H(g,p) = \left\{
268: \begin{array}{ccc}
269: \dot{g} &=& T_e L_g dh(p),\\
270: \dot{p} &=& -\ad{dh(p)}^*\ p,
271: \end{array}
272: \right.
273: \end{equation}
274: Note that $\d r(X_H) = E_h$, where $h \in C^{\infty}(\g^*)$ satisfies
275: $r^* h = H$. The vector field $E_h$ is called the {\em Euler vector
276: field}. It is a natural reduction of $X_H$ by $G$. If $h(p) =
277: \frac{1}{2} \langle p, \phi(p) \rangle$ for a positive-definite linear
278: map $\phi : \g^* \to \g$ then $H$ is induced by a left-invariant
279: metric on $T^* G$ and $X_H$ is the geodesic vector field.
280:
281: Finally, if $D < G$ is a discrete subgroup, then $T^* \Sigma = \Sigma
282: \times \g^*$ where $\Sigma = D \backslash G$. The projection map $r :
283: T^* G \to \g^*$ is naturally left-invariant, so it factors through to
284: a map $r_o : T^* \Sigma \to \g^*$. If $H_o = r_o^*h$ for some $h \in
285: C^{\infty}(\g^*)$ then $Tr_o(X_{H_o}) = E_h$. Thus, the hamiltonian
286: flow of a left-invariant hamiltonian on $T^* \Sigma$ always projects
287: to a hamiltonian flow on $\g^*$.
288:
289:
290: \subsection{Poisson geometry of $T^*T_4$}
291:
292: Let $\t_4$ denote the Lie algebra of $T_4$, so
293: $$
294: \t_4 = \left\{ {
295: \left[ \begin{array}{cccc}
296: 0 & x & z & w\\
297: 0 & 0 & y & u\\
298: 0 & 0 & 0 & v\\
299: 0 & 0 & 0 & 0
300: \end{array} \right] }
301: \ :\ u,v,w,x,y,z \in \R \right\}.
302: $$ For each $a \in \{u,v,w,x,y,z\}$, let $A \in \t_4$ be the
303: element obtained by setting $a$ equal to one and all other
304: coefficients equal to zero. Then $\{U,V,W,X,Y,Z\}$ is a basis of
305: $\t_4$ whose commutation relations given by: $[X,Y]=Z$,$[Y,V]=U$,
306: $[X,U]=W$, $[Z,V]=W$, and all others are trivial or obtained by
307: skew-symmetry.
308:
309: Let $p_{a} : \g^* \to \R{}$ be the linear function given by $p_{a}(p)
310: = \langle p, A \rangle$ for $A \in \g$ and all $p \in \g^*$. From the
311: definition of the Poisson bracket on $\t_4^*$ ({\em
312: c.f.}~Eq.~\ref{eq:pb}), along with the commutation relations, we
313: conclude that: $$\{p_x,p_y\}=-p_z,\ \{p_y,p_v\}=-p_u,\
314: \{p_x,p_u\}=-p_w,\ \{p_z,p_v\}=-p_w.$$
315:
316: There are two independent Casimirs of $\t_4^*$ are $K_1(p) = p_w$, $K_2(p) =
317: p_w p_y - p_z p_u$. Let $K : \t_4^* \to \R^2$ be
318: defined by $K = (K_1,K_2)$. The level sets of $K$ are the
319: coadjoint orbits of $T_4$'s action on $\t_4^*$ and will be denoted by
320: ${\mathcal O}_{k}$, where $k=(k_1,k_2)$. We will say that
321: ${\mathcal O}_{k}$ is a {\em regular coadjoint orbit} if $k_1 k_2
322: \neq 0$.
323:
324: \begin{lemma} \label{lem:ok}
325: Let ${\mathcal O}_k$ be a regular coadjoint orbit. Then ${\mathcal
326: O}_{k}$ is symplectomorphic to $(T^* \R)^{2}$ equipped with its
327: canonical symplectic structure.
328: \end{lemma}
329:
330: \begin{proof}
331: The Poisson bracket on $\t_4^*$ restricts to ${\mathcal O}_k$. We will
332: denote the restricted bracket by $\{,\}_k$. Let $(T^* \R)^{2} = \{
333: (a,A,b,B)\ :\ a,A,b,B \in \R \}$. The canonical Poisson bracket, which
334: we denote by $[,]$, satisfies $[a,A]=1$, $[b,B]=1$ and all other
335: brackets are zero.
336:
337: Let $\lambda$ and $\mu$ be two non-zero parameters (the parameters are
338: included because we will further transform coordinate systems). The
339: map $f_{k}(p) = (a,A,b,B)$ defined by
340: $$a= -\lambda p_x,\ A=(k_1\lambda)^{-1} p_u,\ b= - \mu p_v,\ B=
341: (k_1\mu)^{-1} p_z
342: $$
343: is a diffeomorphism of ${\mathcal O}_{k}$ onto $T^*\R^2$. Indeed,
344: $f_k$ is clearly smooth. And $g_k(a,A,b,B) = (p_u, \ldots, p_z)$
345: defined by
346: $$p_v = -\mu^{-1}b,\ p_w = k_1,\ p_x = -\lambda^{-1}a,\ p_y =
347: (k_2+k_1^2 \lambda \mu AB)/k_1,\ {\rm and\ } p_z = k_1 \mu B,$$
348: satisfies $K \circ g_k = k$ and $f_k \circ g_k = id$, $g_k \circ f_k =
349: id$. Since $g_k$ is an algebraic map, it is smooth, and so we see
350: ${\mathcal O}_k$ is diffeomorphic to $T^* \R^2$.
351:
352: The commutation relations for the Poisson bracket on $\t_4^*$ allow us
353: to compute that $\{f_k^* a, f_k^* A\}_{k} = \{f_k^* b, f_k^* B\}_{k} =
354: 1$ and all other Poisson brackets are zero. It follows that $f_k^* : (
355: C^{\infty}( T^* \R^2) , [,] ) \to ( C^{\infty}({\mathcal O}_k),
356: \{,\}_{k})$ is a Lie algebra isomorphism. Hence $f_k : {\mathcal O}_k
357: \to T^* \R^2$ is a symplectomorphism.
358: \end{proof}
359:
360:
361: \subsection{The hamiltonians}
362:
363: Let $a_{ij} > 0$ be constants such that $a_{13}a_{34}=a_{12}a_{24}$ and let
364: \begin{equation} \label{eq:h}
365: 4H(p) = a_{12} p_x^2 + a_{23} p_y^2 + a_{13} p_z^2 + a_{24} p_u^2 +
366: a_{34} p_v^2 + a_{14} p_w^2
367: \end{equation}
368: Since the vector field $E_{H}$ is unaffected by the addition of a
369: Casimir, the term $a_{14} p_w^2$ can be ignored.
370:
371: Let us introduce a symplectic change of variables on $T^* \R^2$: $A =
372: \frac{1}{\sqrt{2}} (X-Y)$, $B = \frac{1}{\sqrt{2}} (X+Y)$, $a =
373: \frac{1}{\sqrt{2}} (x-y)$, $b = \frac{1}{\sqrt{2}} (x+y)$, $z = c$ and
374: $Z = C$. Because $a_{13}a_{34}=a_{12}a_{24}$, there exists unique
375: $\lambda,\mu > 0$ so that $0 = a_{34} \mu^{-2} - a_{12} \lambda^{-2} =
376: a_{13} \mu^2 - a_{24} \lambda^2$, and $a_{12} \lambda^{-2} + a_{34}
377: \mu^{-2} = 1$. Indeed, we can choose $\lambda^2 = 2a_{12}$ and $\mu^2
378: = 2a_{34}$. Then:
379: \begin{equation} \label{eq:h1}
380: 2{\bf H}_{k} = (x^2 - \xi X^2 + \nu X^4) + (y^2 + \omega Y^2 + \nu
381: Y^4 - 2 \nu X^2 Y^2),
382: \end{equation}
383: where $\xi = - ( a_{13} a_{34} k_1^2 + a_{23} k_2 \sqrt{a_{12} a_{34}})
384: $, $\omega = \xi + 2 a_{13} a_{34} k_1^2 = a_{13} a_{34} k_1^2 - a_{23} k_2
385: \sqrt{a_{12} a_{34}}$ and $\nu = a_{12} a_{23} a_{34} k_1^2$. Note that we
386: can write $\omega=\xi + 2c\nu$ where $c= \frac{a_{13}}{a_{12}a_{23}}$.
387:
388: \begin{lemma} \label{lem:h2}
389: The Euler vector field of $H$ on the regular coadjoint orbit
390: ${\mathcal O}_k$ (equation \ref{eq:h}) is a time change of the hamiltonian
391: vector field of
392: \begin{equation} \label{eq:h2}
393: 2{\bf H} = x^2 + \left(X^2-\frac{1}{2}\right)^2 + y^2 + \alpha^2 Y^2 +
394: Y^4 - 2X^2Y^2
395: \end{equation}
396: on $T^* \R^2$, where $\alpha^2 = 1 + 2c\nu^{\frac{1}{3}}$.
397: \end{lemma}
398:
399: \begin{proof}
400: Define the coordinate change $g_{\nu}(x,X,y,Y) = (ax, a^{-1}X, ay,
401: a^{-1}Y)$ where $a=\nu^{\frac{1}{6}}$. Then $g^*({\bf H}_k) = a^2 {\bf
402: H}$.
403: \end{proof}
404:
405:
406:
407: \begin{lemma} \label{lem:h3}
408: For all $\epsilon > 0$, the hamiltonian flow of ${\bf H}$ (equation
409: \ref{eq:h2}) is conjugate to the flow of the vector field
410: \begin{equation} \label{eq:vf}
411: \vf_{\epsilon} = \left\{
412: \begin{array}{lclclcl}
413: \dot{X} &=& x, & \hspace{2mm} & \dot{Y} &=& y,\\
414: \dot{x} &=& X - 2X^3 + 2\epsilon XY^2, & & \dot{y} &=& \left[ -\alpha^2
415: + 2 X^2 \right]\, Y + 2\epsilon Y^3.
416: \end{array}
417: \right.
418: \end{equation}
419: \end{lemma}
420:
421: \begin{proof}
422: Introduce the coordinate transformation $h_{\epsilon}(x,X,y,Y) =
423: (x,X,\sqrt{\epsilon}y,\sqrt{\epsilon}Y)$.
424: \end{proof}
425:
426: \medskip
427: \noindent{\bf Remark.} It is clear from (\ref{eq:vf}) that the vector
428: field $\vf_{\epsilon}$ depends on the parameter $\alpha$. Inspection
429: of the formula for $\alpha$ (lemma \ref{lem:h2} and immediately above)
430: shows that $\alpha$ is identically unity when the coefficient $a_{13}$
431: vanishes. This is the case for the standard subriemannian metric,
432: where $a_{12}=a_{23}=a_{34}=1$ and the other coefficients vanish. Rather than
433: specializing to $\alpha = 1$, we have elected to carry $\alpha$
434: through our analysis. The rationale for this will be apparent in
435: section \ref{ssec:mel}.
436: \medskip
437:
438:
439: \section{Analysis of $\vf_{\epsilon}$} \label{sec:anax}
440: For $\epsilon=0$, the map $h_{\epsilon}$ is singular. However, the
441: vector field $\vf_0$ is well-defined. We will show that $\vf_0$ has a
442: normally hyperbolic invariant manifold $S$ whose stable and unstable
443: manifolds coincide, and that this manifold $S$ persists for
444: $\epsilon>0$ but the stable and unstable manifolds
445: $W^{\pm}_{\epsilon}(S)$ no longer coincide. This implies that the
446: Euler vector field $E_H | {\mathcal O}_k$ has transverse homoclinic
447: points for all regular coadjoint orbits.
448:
449: \subsection{The normally hyperbolic manifold $S$}
450: Inspection of the vector field $\vf_{\epsilon}$ shows that the set
451: $$S = \{ (x,X,y,Y)\ :\ x=X=0 \},$$
452: is invariant for all $\epsilon$. One sees that for $\epsilon=0$, the
453: vector field is
454: $$\vf_0 = \left\{
455: \begin{array}{lclclcl}
456: \dot{X} &=& x, & \hspace{2mm} & \dot{Y} &=& y,\\
457: \dot{x} &=& X + O(X^3), & & \dot{y} &=& \left[ -\alpha^2 + 2X^2 \right]\, Y,
458: \end{array}
459: \right.
460: $$
461: which shows that $S$ is normally hyperbolic. Therefore $S$ is
462: normally hyperbolic for all $\epsilon$ sufficiently small. Since
463: $\vf_{\epsilon}$ is conjugate to the same vector field for all
464: non-zero $\epsilon$, one concludes that $S$ is a normally hyperbolic
465: manifold for {\em all} $\epsilon$.
466:
467: \subsection{The stable and unstable manifolds of $S$} \label{ssec:wpmS}
468: The function $h = x^2 + (X^2-\frac{1}{2})^2$ is a first integral of
469: $\vf_0$. The set $h^{-1}(\frac{1}{4})$ is the stable and unstable
470: manifold of $S$, which we denote by $W^{\pm}_0(S)$. On $W^{\pm}_0(S)
471: - S$, the flow of $\vf_0$ satisfies
472: \begin{align} \label{eq:separatrix}
473: \left\{
474: \begin{array}{lcllcl}
475: X &=& \pm \sech(t+t_0), & x &=& \mp
476: \tanh(t+t_0)\sech(t+t_0)^2,\\
477: Y &=&
478: c_0 Y_0(t+t_0) + c_1Y_1(t+t_0), & y &=& \dot{Y},
479: \end{array}
480: \right.
481: \end{align}
482: where $X(0)=\pm \sech(t_0), x(0)=\mp \tanh(t_0)\sech(t_0)^2$ and $Y_j$
483: solves the initial-value problem
484: \begin{align} \label{eq:deY}
485: \left\{
486: \begin{array}{rclrcl}
487: \ddot{Y} + \left[ \alpha^2 - 2\sech(t)^2 \right]\, Y &=& 0, \hspace{20mm}
488: (*)\\
489: & \\
490: Y(0) = 1-j,\quad \dot{Y}(0)=j
491: \end{array}
492: \right.
493: \end{align}
494: while $Y(0)=c_0Y_0(t_0) + c_1Y_1(t_0)$,
495: $y(0)=c_0\dot{Y}_0(t_0) + c_1\dot{Y}_1(t_0)$. The solutions $Y_j$ are
496: chosen so that they are even ($j=0$) and odd ($j=1$) functions of
497: time.
498:
499: \subsection{The Melnikov function} \label{ssec:mel}
500: To determine if the flow of $\vf_{\epsilon}$ has transverse homoclinic
501: points for non-zero $\epsilon$, we appeal to the following theorem.
502:
503: \begin{thm} \label{thm:melint}
504: Let $\varphi_{\epsilon} : M \times \R \to M$ be a complete, smooth
505: flow that depends smoothly on $\epsilon$. Assume that $\varphi_0$
506: possesses a normally hyperbolic, invariant manifold $S \subset M$, and
507: that there is a smooth function $h : M \to \R$ such that
508: \begin{enumerate}
509:
510: \item the stable and unstable manifolds of $S$ coincide and equal
511: $h^{-1}(0)$;
512: \item $\d h$ does not vanish on $ W^{\pm}_0(S) - S.$
513:
514: \end{enumerate}
515: Then, for all sufficiently small non-zero $\epsilon$,
516: $\varphi_{\epsilon}$ possesses a normally hyperbolic invariant
517: manifold $S_{\epsilon}$ and the local stable and unstable manifolds of
518: $S_{\epsilon}$ ($W^+_{\epsilon}(S)$ and $W^-_{\epsilon}(S)$,
519: respectively) can be written as the graph of a function
520: $s_{\epsilon}^{\pm} : W^{\pm}_0(S) \to W^{\pm}_{\epsilon}(S)$. The
521: splitting distance, defined for $p \in W^{\pm}_0(S)$ by
522: $s_{\epsilon}(p) = h \circ s^+_{\epsilon}(p) - h \circ
523: s^-_{\epsilon}(p)$, is a smooth function of $\epsilon$ and
524: $s_{\epsilon}(p) = \epsilon m(p) + O(\epsilon^2)$ where
525: \begin{equation} \label{eq:melint}
526: m(p) = \int_{t \in \R} \langle \d h, {\mathcal Y} \rangle \circ
527: \varphi_0^t(p)\, \, \d t,
528: \end{equation}
529: and ${\mathcal Y} = \displaystyle \left. \frac{\partial\ }{\partial
530: \epsilon} \frac{\partial\ }{\partial t} \right|_{t=\epsilon=0}\,
531: \varphi^t_{\epsilon} $.
532: \end{thm}
533:
534: \begin{proof}
535: The proof of this theorem is a standard application of invariant
536: manifold theory plus an adaptation of the proof of the Melnikov
537: formula~\cite{MarsdenHolmes:1983a,Robinson:1988a,Gruendler}.
538: \end{proof}
539:
540: \medskip
541: \noindent{\bf Remark.} If $m$ changes sign, then, for all $\epsilon
542: \neq 0$ sufficiently small, the perturbed stable and unstable
543: manifolds intersect but do not coincide. In our case, a result of
544: Burns and Weiss implies that the topological entropy is non-zero. If
545: $m$ has a non-degenerate zero, then the implicit function theorem
546: implies that, for all $\epsilon \neq 0$ sufficiently small,
547: $W^+_{\epsilon}(S)$ has a transverse intersection with
548: $W^-_{\epsilon}(S)$. Note that the intersection of the surface $S$
549: with a constant energy level is a periodic orbit. Therefore each
550: trajectory in the intersection is doubly asymptotic to a periodic
551: orbit in $S$.
552: \bigskip
553:
554: For the flow defined by $\vf_{\epsilon}$, we have that
555: \begin{equation} \label{eq:y}
556: {\mathcal Y} =
557: \left\{
558: \begin{array}{lclclcl}
559: \dot{X} &=& 0, & \hspace{2mm} & \dot{Y} &=& 0,\\
560: \dot{x} &=& 2 XY^2, & & \dot{y} &=& 2 Y^3.
561: \end{array}
562: \right.
563: \end{equation}
564: Whence
565: \begin{equation} \label{eq:dhy}
566: \langle \d h, {\mathcal Y} \rangle = 4 xXY^2,
567: \end{equation}
568: since $h = x^2 + (X-\frac{1}{2})^2$. The equations in
569: (\ref{eq:separatrix}) imply that the Melnikov function is
570: \begin{equation} \label{eq:melfn}
571: m(p) = c_0c_1 \times \int_{\tau \in \R} -4\, \tanh(\tau)\, \sech(\tau)^2\
572: Y_0(\tau)\, Y_1(\tau)\ \d\tau.
573: \end{equation}
574:
575: \medskip
576: \noindent{\bf Remarks.} (1) The formula for the Melnikov integral
577: (\ref{eq:melfn}) appears to be a function on $S$ not
578: $W^{\pm}_0(S)-S$. This does not contradict Theorem
579: (\ref{thm:melint}). Inspection of the integral (equation
580: \ref{eq:melint}) shows that $m(\varphi_0^s(p)) = m(p)$ for all $s$ and
581: $p$. The coordinate system on $W^{\pm}_0(S)-S$ determined by
582: (\ref{eq:separatrix}) uses time along the flow as one coordinate
583: ($t_0$), so only the other two coordinates, $c_0$ and $c_1$, ought to
584: appear in the Melnikov function. (2) If we write $m(p) = 2 c_0c_1 \times
585: I$, where
586: \begin{equation} \label{eq:I}
587: I=\int_{0}^{\infty} -4\, \tanh(\tau)\, \sech(\tau)^2\ Y_0(\tau)\,
588: Y_1(\tau)\ \d\tau,
589: \end{equation} then $m$ has
590: non-degenerate zeros along $\{ c_0=0,c_1\neq0\ {\rm or}\
591: c_1=0,c_0\neq0 \}$, provided that $I\neq 0$.
592: \medskip
593:
594: \subsection{The Legendre functions and $I$}
595: Substitution of $z=\tanh(t)$ transforms the differential equation
596: (\ref{eq:deY}*) into the Legendre differential equation
597: \begin{equation} \label{eq:legendre}
598: (1-z^2)Y'' -2zY' + \left( \nu(\nu+1) - \frac{\mu^2}{1-z^2}
599: \right) Y = 0,
600: \end{equation}
601: where $\mu=i\alpha$, $\nu=-\frac{1}{2}+\frac{\sqrt{-7}}{2}$ and
602: $'=\frac{\d\ }{\d z}$. The integral $I$ (equation \ref{eq:I}) is
603: transformed to
604: \begin{equation} \label{eq:Imod}
605: I = \int_{0}^1 z\, U_0(z)\, U_1(z)\, \d z,
606: \end{equation}
607: where $U_j(z) = Y_j(t)$.
608:
609: \subsection{The Melnikov function is non-zero: $I\neq0$}
610: \label{ssec:melfn0}
611: For the remainder of this note, $q~:~[0,\infty)~\to~\R$ is a
612: continuously differentiable function such that $\lim_{t\to\infty} q(t)
613: = 0$ and $\alpha > 0$ is a fixed positive number. The function
614: $z=z(t)$ is assumed to solve
615: \begin{equation} \label{eq:dez}
616: \ddot{z} + [\alpha^2 - q(t)]\, z = 0.
617: \end{equation}
618: In analogy with the integral $I$ in equation (\ref{eq:I}), define an
619: integral
620: \begin{equation} \label{eq:Iz}
621: I = \int_{0}^{\infty} \dot{q}(t)\, z_0(t)\, z_1(t)\, \d t
622: \end{equation}
623: $I$ is implicitly a function of $\alpha$; one wants to prove that
624: $I$ can vanish at most countably many times.
625:
626: Let us first prove the following.
627: \begin{lemma} \label{lem:bdd}
628: If $z$ solves equation (\ref{eq:dez}), then $z$ is bounded with
629: bounded derivative.
630: \end{lemma}
631: \begin{proof}
632: Define $H = \frac{1}{2} (\alpha^2 z^2 + \dot{z}^2)$. From (\ref{eq:dez}) one
633: computes that $\dot{H} = qz\dot{z}$. Integrating by parts yields $H =
634: C_0 + \frac{1}{2} q(t)z(t)^2 + \int_0^t - \frac{1}{2} \dot{q}(s)\, z(s)^2\
635: \d s$, where $C_0$ is a constant that depends only on $z(0)$ and
636: $\dot{z}(0)$. Thus
637: \begin{equation} \label{eq:zbd}
638: \frac{1}{2} \left[\, \alpha^2 - q(t) \right]\, z(t)^2 \leq C_0 +
639: \int_0^t - \frac{1}{2}\, \dot{q}(s)\, z(s)^2\ \d s.
640: \end{equation}
641: Since $q(t) \to 0$ as $t \to \infty$, there is a $T \geq 0$ such that
642: $\alpha^2 - q(t) \geq \frac{1}{2}\alpha^2$ for all $t \geq
643: T$. Therefore, equation (\ref{eq:zbd}) implies that there is a constant
644: $C_1$ such that for all $t \geq 0$
645: \begin{equation} \label{eq:zbd2}
646: z(t)^2 \leq C_1 + \frac{4}{\alpha^2} \times \int_0^t -\dot{q}(s)\, z(s)^2\
647: \d s
648: \end{equation}
649: This is a Gronwall inequality for $u=z^2$. Thus
650: $$z(t)^2 \leq C_1 \, \exp(-\frac{4}{\alpha^2} q(t) )$$
651: for all $t \geq 0$. Since $q$ is continuous and converges to $0$ at
652: infinity, it is bounded. Therefore, $z$ is bounded.
653:
654: To prove that $w=\dot{z}$ is bounded, define $H=\frac{1}{2} (\alpha^2
655: w^2 + \dot{w}^2)$. Since $\ddot{w}+[\alpha^2 - q(t)]w = f$, $f=\dot{q}z$,
656: one computes that $\dot{H} = qw\dot{w} + f\dot{w}$. One can bound
657: $\int_0^t f(s)\dot{w}(s)\d s$ using that $\ddot{z}=\dot{w}$ and that
658: $q$ and $z$ are already bounded. One then obtains a Gronwall
659: inequality like (\ref{eq:zbd2}) for $w(t)^2$.
660: \end{proof}
661:
662:
663: \begin{lemma} \label{lem:zero}
664: If $z_0,z_1$ are solutions to equation (\ref{eq:dez}), then the limit
665: \begin{equation} \label{eq:limit}
666: I = -\lim_{t\to \infty} \left[ \dot{z}_0(t)\, \dot{z}_1(t)\, + \alpha^2
667: z_0(t)\, z_1(t) \right]
668: \end{equation}
669: exists and equals $W \alpha\cot(B)$ where the angle $B$ is defined
670: below, and $W$ is the Wronskian of the solutions $z_0,z_1$.
671: \end{lemma}
672:
673: \begin{proof}
674: Let $z$ be any solution of (\ref{eq:dez}). Let $t_n$ be the sequence
675: of zeros of $z(t)$, indexed in increasing order. Let $\phi_n = \alpha
676: t_n \bmod 2\pi$ and $a_n = \dot{z}(t_n)$. There is a sequence $n_k$
677: such that $\phi_{n_k} \to \phi \bmod 2\pi$ and $a_{n_k} \to a>0$. The
678: former follows by compactness of $\R/2\pi\Z$ and the latter because
679: $\dot{z}$ is bounded.
680:
681: Since $q(t) \to 0$, the Sturm comparison theorem implies that the
682: sequence $\phi_n$ converges to $\phi$ and $t_{n+1}-t_n$ converges to
683: $\pi/\alpha$~\cite{Zettl}. One sees that positive and negative zeros
684: of $z$ must alternate for all $n$. Hence, without loss of generality,
685: one may assume that $\dot{z}(t_{2n}) \to a$ and $\dot{z}(t_{2n+1}) \to
686: -a$. The continuous dependence of solutions on initial data therefore
687: implies that $z^n(t) := z(t+\frac{2n\pi}{\alpha})$ converges in the
688: weak Whitney $C^1$ topology to $a \sin(\alpha t-\phi)$. In particular,
689: $z^n(t)$ converges uniformly to $a \sin(\alpha t-\phi)$ for $t \in
690: [0,4\pi/\alpha]$.
691:
692: To apply these observations to the limit (\ref{eq:limit}), let $N$ be
693: sufficiently large so that the $n$-th and $n+1$-th zeros of both $z_0$
694: and $z_1$ are at most $\frac{2\pi}{\alpha}$ apart for all $n\geq
695: N$. Let $s\in [\frac{2n\pi}{\alpha},\frac{2(n+1)\pi}{\alpha}]$ and write
696: $s=t+\frac{n\pi}{\alpha}$ so that $t\in [0,\frac{2\pi}{\alpha}]$. Then
697: \begin{align*}
698: &\ \ \ \ |\dot{z}_0(s) \dot{z}_1(s) + \alpha^2 z_0(s)z_1(s) - \alpha^2 a_0a_1
699: \cos(\phi_1-\phi_0)|\\
700: &\leq |\dot{z}^n_0(t) \dot{z}^n_1(t) - \alpha^2
701: a_0a_1 \cos(\alpha t - \phi_0) \cos(\alpha t - \phi_1)| +\\
702: &\ \ \ \ \ \alpha^2 | z^n_0(t)z^n_1(t) - a_0a_1 \sin(\alpha t - \phi_0)
703: \sin(\alpha t - \phi_1)|.
704: \end{align*}
705: If $s \to \infty$, then $n\to \infty$. The above-mentioned uniform
706: convergence for $t\in [0,2\pi/\alpha]$ shows that the limit
707: (\ref{eq:limit}) exists and equals $A\cos(B)$ where $A=\alpha^2
708: a_0a_1$ and $B=\phi_1-\phi_0 \bmod 2\pi$.
709:
710: On the other hand, the Wronskian $W$ of $z_0,z_1$ is
711: constant and
712: $$\xymatrix{W = z_0(t)\dot{z}_1(t) - z_1(t)\dot{z}_0(t)\ \ar[rr]^{t \to
713: \infty} && \ \alpha a_0a_1 \sin(\phi_1-\phi_0),}$$ by the
714: same argument as above. Therefore $I/W = \alpha \cot(B)$.
715: \end{proof}
716:
717:
718: \medskip
719: \noindent{\bf Remarks.} (1) The angle $B$ has the following
720: interpretation which emerges from the proof of lemma
721: (\ref{lem:zero}). The zeros of solutions to (\ref{eq:dez}) are
722: asymptotically $\pi/\alpha$ apart, and the zeros of linearly
723: independent solutions are interlaced. The angle $B$ is defined so that
724: $B/\alpha \bmod \pi/\alpha$ is asymptotically the time between
725: consecutive zeros of the linearly independent solutions. Figure
726: \ref{fig:1}, left, plots $B$ as a function of $\alpha$ for the
727: solutions $z_j=Y_j$ to the initial-value problem (\ref{eq:deY}). One
728: expects that as $\alpha \to \infty$, the solutions should converge
729: quite quickly to $\cos$ and $\sin$, whence $B$ should approach
730: $\frac{\pi}{2}$. The figure captures this behaviour quite nicely. (2)
731: The function $I = W\alpha \cot(B)$ from lemma~\ref{lem:zero} can be
732: computed numerically. The Sturm comparison theorem implies that the
733: $n$-th zero $t_n$ of a solution $z$ satisfies $\pi/\alpha < t_{n+1} -
734: t_n < \pi/\alpha \times (1 + q/\alpha^2)$ if $|q(t)| < \alpha^2/2$ for
735: all $t > t_n$. If $q$ goes to zero sufficiently fast, one can
736: numerically compute the first several zeros and obtain a reasonably
737: accurate estimate of $B$. Figure \ref{fig:1}, right, shows the graph
738: of $I$ for $q=2\sech(t)^2$.
739: \medskip
740:
741: \begin{lemma} \label{lem:I0}
742: If $z_0, z_1$ are solutions to (\ref{eq:dez}) such that $\dot{z}_0$
743: and $z_1$ vanish at $t=0$, then the integral
744: $$
745: I = \int_{0}^{\infty} \dot{q}(t)\, z_0(t)\, z_1(t)\, \d t \leqno({\rm \ref{eq:Iz}})
746: $$
747: exists and equals $W \alpha\cot(B)$ where the angle $B$ is described
748: in Lemma \ref{lem:zero}, and $W$ is the Wronskian of the solutions
749: $z_0,z_1$.
750: \end{lemma}
751:
752: \begin{proof}
753: By lemma \ref{lem:bdd}, both solutions are bounded, so one can apply
754: integration by parts to the integral. This yields
755: \begin{align*}
756: I &= q(0) z_0(0) z_1(0) - \int_{0}^{\infty} q(t)\, \left[
757: \dot{z}_0(t)\, z_1(t)\, + z_0(t)\, \dot{z}_1(t)\, \right]\, \d t,\\
758: &= -\int_{0}^{\infty} q(t)\, \left[
759: \dot{z}_0(t)\, z_1(t)\, + z_0(t)\, \dot{z}_1(t)\, \right]\, \d t
760: \end{align*}
761: since $z_1$ vanishes at $t=0$.
762:
763: From (\ref{eq:dez}), it is known that $q(t)z_0(t) = \ddot{z}_0(t)+\alpha^2
764: z_0(t)$ and similarly for $z_1$. Therefore
765: \begin{align*}
766: I &= -\int_{0}^{\infty} \frac{\d\ }{\d t}\,
767: \left[ \dot{z}_0(t)\, \dot{z}_1(t)\, + \alpha^2 z_0(t)\, z_1(t)\,
768: \right]\, \d t,\\
769: &= - \lim_{t\to \infty} \left[ \dot{z}_0(t)\,
770: \dot{z}_1(t)\, + \alpha^2 z_0(t)\, z_1(t) \right], \hspace{5mm}
771: \textrm{since}\ \dot{z}_0(0)=0=z_1(0),\\
772: &= W \alpha \cot(B) \hspace{45mm} {\rm by\ lemma\ (\ref{lem:zero}).}
773: \end{align*}
774: \end{proof}
775:
776:
777:
778: \begin{lemma} \label{lem:nov}
779: Assume that there exists $C, \lambda > 0$ such that $|q(t)| < C
780: e^{\lambda t}$ for all $t > 0$. Then the integral $I=I(\alpha)$ is a
781: holomorphic function of $\alpha$ on the strip $|\im \alpha| < \lambda$
782: about the real line.
783:
784: Consequently, if $q$ is an even, monotone function, then $I$ vanishes
785: countably many times at most.
786: \end{lemma}
787:
788: \begin{proof}
789: A solution $z=z(t;\alpha)$ to (\ref{eq:dez}) is a holomorphic function
790: of $\alpha$ for each fixed $t$ \cite{Zettl}. For large $t$ and $|\im
791: \alpha| < \lambda$, the solution $z=z(t;\alpha)$ is equal to
792: $\cos(\alpha t + \phi)$ plus a term that grows slower than $e^{\lambda
793: t}$. This implies, by the residue formula, that $I=I(\alpha)$ is
794: holomorphic provided that $|\im \alpha| < \lambda$.
795:
796: When $\alpha=0$ and $q$ is even, the even and odd solutions to
797: (\ref{eq:dez}) do not change sign. Therefore, if $q$ is monotone, then
798: the integrand defining $I(0)$ does not change sign, so $I(0)\neq
799: 0$. Thus, $I$ can vanish at most countably many times on the strip
800: $|\im \alpha| < \lambda$.
801: \end{proof}
802:
803: \begin{proof}[Theorem \ref{thm:main}]
804: If $a_{13} \neq 0$ -- whence $c \neq 0$ in Lemma \ref{lem:h2} --, then
805: lemma \ref{lem:nov} shows that the hamiltonian flow of $H$ (equation
806: \ref{eq:h}) on all but countably many coadjoint orbits in $\t^*_4$ has
807: a horseshoe. This proves the main result, Theorem \ref{thm:main}.
808: \end{proof}
809:
810: \section{The degenerate case when $\alpha \equiv 1$} \label{sec:deg}
811: If $a_{13} = 0$, as occurs for the Carnot subriemannian metric of
812: \cite{Montgomeryetal}, then $\alpha \equiv 1$ and lemma \ref{lem:nov}
813: cannot be applied. We investigate two distinct ways to address this
814: problem. The first is direct and numerical; the second leads to some
815: further insight into the integral $I$.
816:
817: \subsection{Numerical evidence} \label{ssec:num}
818: In this case, figure \ref{fig:1} indicates that $I(1)$ is
819: approximately $-2.75$. Table \ref{tab:1} shows the results of a
820: numerical computation of $I(1)$ with varying step sizes. It is clear
821: from this table that $I(1)=-2.76$ to two decimal places.
822:
823: To estimate the error in the computations, one uses the fact that the
824: differential equations (\ref{eq:dez}) are hamiltonian with the
825: hamiltonian
826: \begin{equation} \label{eq:Hz}
827: \H = \frac{1}{2}p^2 + \frac{1}{2}\left[ \alpha^2 - q(\tau) \right]\,
828: z^2 + u,
829: \end{equation}
830: where $p,z$ and $u,\tau$ are canonically conjugate variables (along
831: solutions, $\tau = \tau_0+t$, so it is a pseudo-time). Since $\H$ is
832: preserved along solutions to (\ref{eq:dez}), the maximum deviation of
833: $\H$ along a numerical solution provides an estimate of the upper
834: bound of the error in the solutions $z_0, z_1$.
835:
836: \subsection{Qualitative evidence} \label{ssec:qual}
837: As explained in the Remark in subsection \ref{ssec:melfn0}, one may
838: compute $I$ as a function of $\alpha$ by computing the phase angle
839: $B$. Figure \ref{fig:1} graphs $B$ and $I$ versus $\alpha$. This
840: figure shows that $I(1)$ does not vanish.
841:
842:
843: \begin{sidewaystable}
844: \centering
845: \caption{The numerical calculation of $I$ with $\alpha=1$. } \label{tab:1}
846: \begin{tabular}{|llllllll|}
847: \hline\hline $h$ & $I$ & $\H_{0,min}$ & $\H_{0,max}$ & $\H_{1,min}$ &
848: $\H_{1,max}$ & $\H_{0,max}-\H_{0,min}$ & $\H_{1,max}-\H_{1,min}$
849: \\ \hline\hline
850:
851: \rmt{ 0.5 } & \rmt{ -2.76812630 } & \rmt{ -0.5 } & \rmt{ -0.49025150 } & \rmt{ 0.49833857 } & \rmt{ 0.51067514 } & \rmt{ 0.00974849 } & \rmt{ 0.01233657}\\
852: \rmt{ 0.25 } & \rmt{ -2.76366763 } & \rmt{ -0.5 } & \rmt{ -0.49944022 } & \rmt{ 0.49992738 } & \rmt{ 0.50063022 } & \rmt{ 0.00055977 } & \rmt{ 0.00070283}\\
853: \rmt{ 0.125 } & \rmt{ -2.76340793 } & \rmt{ -0.5 } & \rmt{ -0.49996571 } & \rmt{ 0.49999582 } & \rmt{ 0.50003878 } & \rmt{ 3.42847126 $\times 10^{-5}$ } & \rmt{ 4.29572646 $\times 10^{-5}$ }\\
854: \rmt{ 0.0625 } & \rmt{ -2.76339200 } & \rmt{ -0.5 } & \rmt{ -0.49999786 } & \rmt{ 0.49999974 } & \rmt{ 0.50000241 } & \rmt{ 2.13207876 $\times 10^{-6}$ } & \rmt{ 2.67223559 $\times 10^{-6}$ }\\
855: \rmt{ 0.03125 } & \rmt{ -2.76339101 } & \rmt{ -0.5 } & \rmt{ -0.49999986 } & \rmt{ 0.49999998 } & \rmt{ 0.50000015 } & \rmt{ 1.33088170 $\times 10^{-7}$ } & \rmt{ 1.66835298 $\times 10^{-7}$ }\\
856: \rmt{ 0.015625 } & \rmt{ -2.76339095 } & \rmt{ -0.5 } & \rmt{ -0.49999999 } & \rmt{ 0.49999999 } & \rmt{ 0.50000000 } & \rmt{ 8.31541421 $\times 10^{-9}$ } & \rmt{ 1.04238692 $\times 10^{-8}$ }\\
857: \rmt{ 0.0078125 } & \rmt{ -2.76339094 } &
858: \rmt{ -0.5 } & \rmt{ -0.49999999 } &
859: \rmt{ 0.49999999 } & \rmt{
860: 0.50000000 } & \rmt{ 5.19672801 $\times 10^{-10}$ } &
861: \rmt{ 6.51456639 $\times 10^{-10}$} \\\hline
862: \multicolumn{8}{p{16cm}}{$ $}\\
863: \multicolumn{8}{p{16cm}}{Solutions to the hamiltonian equations of $\H$ are computed
864: with the Forest-Ruth $4$-th order symplectic integrator
865: \cite{SanzSerna} and initial conditions
866: $z(0)=j,\dot{z}(0)=1-j,\tau(0)=0,u(0)=0$ for $j=0,1$. The maximum
867: (resp. minimum) value of $\H$ along the $j$-th numerical solution
868: over the interval $[0,35]$ is indicated by $\H_{j,max}$
869: (resp. $\H_{j,min}$). The integral $I$ is computed by
870: \[
871: \hspace{-4cm} I^h = h \times \sum_{i=0}^N \dot{q}(t_i)\, z^h_0(t_i)\, z^h_1(t_i)
872: \]
873: where $z^h_j$ is the computed solution with step size $h$, $N=35/h$,
874: $t_i = i \times h$ and $q(t) = 2\sech(t)^2$.
875: Files at \url{http://www.maths.ed.ac.uk/~lbutler/t4.html}}
876: \end{tabular}
877: \end{sidewaystable}
878:
879:
880:
881:
882:
883: {
884: \def\figa{{
885: \def\orbitpica{\resizebox{5cm}{!}{\includegraphics{phaseangle.eps}}}
886: \xy
887: \xyimport(4,4){\orbitpica}
888: ,(2.,3.)*!L\txt{$B(\alpha)$}*=0{},{\ar-(.5,.5)}
889: ,(1.2,1.1)*!L\txt{$\frac{\pi}{2}$}*=0{},{\ar-(.1,.4)}
890: ,(2.2,.1)*!L\txt{$\alpha$}*=0{}
891: ,(-0.1,2.5)*!L\txt{$B$}*=0{}
892: \endxy
893: }}
894: \def\figb{{
895: \def\orbitpicb{\resizebox{5cm}{!}{\includegraphics{I.eps}}}
896: \xy
897: \xyimport(4,4){\orbitpicb}
898: ,(2.,3.)*!L\txt{$\alpha\cot(B)$}*=0{},{\ar-(0.5,0.0)}
899: ,(2.2,.0)*!L\txt{$\alpha$}*=0{}
900: ,(-0.1,2.5)*!L\txt{$I$}*=0{}
901: \endxy
902: }}
903: \begin{center}
904: \begin{figure}[htb]
905: $$
906: \xymatrix{
907: \figa & \figb
908: }
909: $$
910: \caption{Left: Phase angle $B$ vs. $\alpha$; Right: $I=\alpha\cot(B)$
911: vs. $\alpha$. Both plots use the solutions $z_j = Y_j$ of
912: (\ref{eq:deY}) with $q(t) = 2\sech(t)^2$. These solutions are
913: computed numerically in Maple using the $4$-th order Runge-Kutta
914: method; the zeros are located by interval halving.} \label{fig:1}
915: \end{figure}
916: \end{center}
917: }
918:
919:
920:
921:
922:
923:
924:
925: \begin{thebibliography}{aa}
926:
927:
928: \bibitem{Butler:2000a} L. T. Butler, {\it Integrable Geodesic Flows on
929: $n$-step Nilmanifolds.} J. Geom. Phys. {\bf 36}(3-4) (2000) 315--323.
930:
931: \bibitem{Butler:2003a} L. T. Butler, {\it Integrable Geodesic Flows
932: with Wild First Integrals: the case of two-step nilmanifolds}, Ergodic
933: Theory Dynamical Systems. 23(3) (2003) 771--797.
934:
935:
936: \bibitem{Butler:2003b} L. T. Butler, {\it Invariant metrics on
937: nilmanifolds with positive topological entropy.} Geom. Dedicata 100
938: (2003), 173--185.
939:
940: \bibitem{Gruendler} J. Gruendler, {\em The existence of homoclinic
941: orbits and the method of Melnikov for systems in ${\bf R}\sp n$.} SIAM
942: J. Math. Anal. 16 (1985), no. 5, 907--931.
943:
944: \bibitem{GS} V. Guillemin and S. Sternberg, {\em Symplectic techniques
945: in physics, 2nd ed.}, {Cambridge University Press}, {Cambridge},
946: (1990).
947:
948: \bibitem{Kar} R. Karidi, {\it Geometry of balls in nilpotent {L}ie groups}, Duke Math. J. {\bf 74} (1994) 301--317.
949:
950: \bibitem{MarsdenHolmes:1983a}
951: P. Holmes, J. Marsden,
952: {\it Horseshoes and {A}rnol$'$d diffusion for {H}amiltonian
953: systems on {L}ie groups}, Indiana Univ. Math. J. {\bf 32} (1983) 273--309.
954:
955:
956: \bibitem{Man} R. Ma\~{n}\'{e}, {\it On the topological entropy of
957: geodesic flows}, J. Diff. Geom. {\bf 45} (1997) 74--93.
958:
959: \bibitem{Manning} A. Manning, ``More topological entropy for geodesic
960: flows,'' in {\it Dynamical systems and turbulence, Warwick 1980
961: (Coventry, 1979/1980)} Springer Lecture Notes in Mathematics 898,
962: 243--249.
963:
964: \bibitem{Montgomeryetal} R. Montgomery, M. Shapiro and A. Stolin, {\it
965: A nonintegrable sub-{R}iemannian geodesic flow on a {C}arnot group},
966: J. Dynam. Control Systems, {\bf 3}(4) (1998) 519--530.
967:
968: \bibitem{Robinson:1988a}
969: C. Robinson,
970: {\it Horseshoes for autonomous {H}amiltonian systems using the
971: {M}elnikov integral,}
972: Ergodic Theory Dynam. Systems, {\bf 8}$\sp *$ (Charles Conley Memorial
973: Issue) (1988) 395--409.
974:
975: \bibitem{SanzSerna}
976: J.M. Sanz-Serna and M.P. Calvo,
977: {\it Numerical Hamiltonian problems.}
978: Applied Mathematics and Mathematical Computation, 7. Chapman \& Hall, London, 1994.
979:
980: \bibitem{Zettl} A. Zettl, {\bf Sturm-Liouville theory.} Mathematical
981: Surveys and Monographs, 121. American Mathematical Society,
982: Providence, RI, 2005.
983:
984: \bibitem{Zig1} S. L. Ziglin, {\it Bifurcation of solutions and the
985: nonexistence of first integrals in {H}amiltonian mechanics. {I}},
986: Funktsional. Anal. i Prilozhen. {\bf 16}(3) (1982) 30--41.
987:
988:
989: \bibitem{Zig2} S. L. Ziglin, {\it Bifurcation of solutions and the
990: nonexistence of first integrals in {H}amiltonian mechanics. {I}{I}},
991: Funktsional. Anal. i Prilozhen. {\bf 17}(1) (1983) 8--23.
992:
993: \end{thebibliography}
994:
995: \end{document}
996:
997:
998:
999: \rmt{ 0.5 } & \rmt{ -2.76812630719696 } & \rmt{ -0.5 } & \rmt{ -0.4902515057504894 } & \rmt{ 0.49833857313872 } & \rmt{ 0.5106751468509393 } & \rmt{ 0.009748494249510559 } & \rmt{ 0.01233657371221933}\\
1000: \rmt{ 0.25 } & \rmt{ -2.763667635603004 } & \rmt{ -0.5 } & \rmt{ -0.4994402289011139 } & \rmt{ 0.4999273875356691 } & \rmt{ 0.5006302219289245 } & \rmt{ 0.000559771098886072 } & \rmt{ 0.0007028343932553842}\\
1001: \rmt{ 0.125 } & \rmt{ -2.763407937260531 } & \rmt{ -0.5 } & \rmt{ -0.4999657152873865 } & \rmt{ 0.4999958242863702 } & \rmt{ 0.5000387815509778 } & \rmt{ 3.428471261350508e-05 } & \rmt{ 4.295726460764075e-05}\\
1002: \rmt{ 0.0625 } & \rmt{ -2.763392003969109 } & \rmt{ -0.5 } & \rmt{ -0.4999978679212389 } & \rmt{ 0.4999997421321118 } & \rmt{ 0.5000024143677053 } & \rmt{ 2.132078761114885e-06 } & \rmt{ 2.672235593571484e-06}\\
1003: \rmt{ 0.03125 } & \rmt{ -2.763391012815484 } & \rmt{ -0.5 } & \rmt{ -0.4999998669118291 } & \rmt{ 0.4999999839147126 } & \rmt{ 0.5000001507500108 } & \rmt{ 1.330881709366233e-07 } & \rmt{ 1.66835298217921e-07}\\
1004: \rmt{ 0.015625 } & \rmt{ -2.763390950941556 } & \rmt{ -0.5 } & \rmt{ -0.4999999916845858 } & \rmt{ 0.4999999989956971 } & \rmt{ 0.5000000094195664 } & \rmt{ 8.315414211151548e-09 } & \rmt{ 1.042386924110625e-08}\\
1005: \rmt{ 0.0078125 } & \rmt{ -2.76339094707558 } &
1006: \rmt{ -0.5 } & \rmt{ -0.4999999994803272 } &
1007: \rmt{ 0.4999999999372302 } & \rmt{
1008: 0.5000000005886868 } & \rmt{ 5.196728017527186e-10 } &
1009: \rmt{ 6.514566393387052e-10} \\\hline
1010:
1011:
1012: \rmt{ 0.5 } & \rmt{ -2.76812630 } & \rmt{ -0.5 } & \rmt{ -0.49025150 } & \rmt{ 0.49833857 } & \rmt{ 0.51067514 } & \rmt{ 0.00974849 } & \rmt{ 0.01233657}\\
1013: \rmt{ 0.25 } & \rmt{ -2.76366763 } & \rmt{ -0.5 } & \rmt{ -0.49944022 } & \rmt{ 0.49992738 } & \rmt{ 0.50063022 } & \rmt{ 0.00055977 } & \rmt{ 0.00070283}\\
1014: \rmt{ 0.125 } & \rmt{ -2.76340793 } & \rmt{ -0.5 } & \rmt{ -0.49996571 } & \rmt{ 0.49999582 } & \rmt{ 0.50003878 } & \rmt{ 3.42847126e-05 } & \rmt{ 4.29572646e-05}\\
1015: \rmt{ 0.0625 } & \rmt{ -2.76339200 } & \rmt{ -0.5 } & \rmt{ -0.49999786 } & \rmt{ 0.49999974 } & \rmt{ 0.50000241 } & \rmt{ 2.13207876e-06 } & \rmt{ 2.67223559e-06}\\
1016: \rmt{ 0.03125 } & \rmt{ -2.76339101 } & \rmt{ -0.5 } & \rmt{ -0.49999986 } & \rmt{ 0.49999998 } & \rmt{ 0.50000015 } & \rmt{ 1.33088170e-07 } & \rmt{ 1.66835298e-07}\\
1017: \rmt{ 0.015625 } & \rmt{ -2.76339095 } & \rmt{ -0.5 } & \rmt{ -0.49999999 } & \rmt{ 0.49999999 } & \rmt{ 0.50000000 } & \rmt{ 8.31541421e-09 } & \rmt{ 1.04238692e-08}\\
1018: \rmt{ 0.0078125 } & \rmt{ -2.76339094 } &
1019: \rmt{ -0.5 } & \rmt{ -0.49999999 } &
1020: \rmt{ 0.49999999 } & \rmt{
1021: 0.50000000 } & \rmt{ 5.19672801e-10 } &
1022: \rmt{ 6.51456639e-10} \\\hline
1023:
1024:
1025: \hline
1026: \begin{multicolumn}
1027: \\
1028: \begin{center}
1029: The numerical calculation of $I$ with $\alpha=1$.
1030: \end{center}
1031: \\
1032: \end{multicolumn}
1033: