1: % loop.tex
2:
3: % add editing date at the top whenever any file.tex is modified
4:
5: % PC Aug 3 2003
6: % YL Jul 31 2003
7: % PC Jul 23 2003
8: % YL Jul 18 2003
9: % PC Jul 10 2003
10: % YL old/loop3-070503.zip Jul 5 2003
11: % PC May 13 2003
12: % YL old/newtonrap3.tex Apr 29 2003
13: % YL old/newtonrap.tex Sep 27 2002
14:
15:
16: % this style for submission, web version:
17: \documentclass[pre,twocolumn,groupedaddress,showpacs,showkeys]{revtex4}
18:
19: % this style while editing:
20: %\documentclass[pre,preprint,groupedaddress,showpacs,showkeys]{revtex4}
21:
22: \usepackage{amssymb}
23: \usepackage{amsmath}
24: \usepackage{graphicx}
25: \usepackage{subfigure}
26: \usepackage{dsfont}
27: \usepackage{mathrsfs}
28: \usepackage{natbib} % required by apsrev.bst
29: \usepackage[dvips]{color} % dvips allows for colors
30: \input loopDefs % all definitions in loopDefs.tex
31: \bibliographystyle{apsrev}
32: \begin{document}
33: \title{
34: Variational method for finding periodic orbits in a general flow
35: }\author{
36: Yueheng Lan }\email{gte158y@prism.gatech.edu}\author{
37: Predrag Cvitanovi\'{c} }\email{predrag.cvitanovic@physics.gatech.edu}
38: \affiliation{
39: Center for Nonlinear Science, School of Physics,\\
40: Georgia Institute of Technology, Atlanta 30332-0430, U.S.A}
41: \date{\today}
42:
43: \begin{abstract}
44: A variational principle for determining unstable periodic orbits of
45: flows as well as unstable spatio-temporally periodic solutions of
46: extended systems is proposed and implemented. An initial loop approximating
47: a periodic solution is evolved in the space of loops toward a true periodic solution
48: by a minimization of local errors along the loop. The
49: ``\descent'' partial differential
50: equation that governs this evolution is an infinitesimal step version of
51: the damped Newton-Raphson iteration. The feasibility of the method is demonstrated by its application to
52: the H\'enon-Heiles system, the circular restricted three-body problem,
53: and the Kuramoto-Sivashinsky system in a weakly turbulent regime.
54: \end{abstract}
55: \pacs{95.10.Fh, 02.70.Bf, 47.52.+j, 05.45.+a}
56: \keywords{
57: periodic orbits, variational methods,
58: spatio-temporal chaos, turbulence,
59: cost function minimization,
60: Kuramoto-Sivashinsky system,
61: restricted three-body problem, H\'enon-Heiles system
62: }
63: \maketitle
64: \section{Introduction}
65:
66: The periodic orbit theory of classical and quantum chaos\rf{ruelle,gutbook}
67: is one of the major advances in the study of long-time behavior of chaotic
68: dynamical systems.
69: The theory expresses all long time averages over chaotic
70: dynamics in terms of cycle expansions\rf{cexp},
71: sums over
72: periodic orbits (cycles) ordered hierarchically according to the orbit length, stability,
73: or action.
74: If the symbolic dynamics is known, and the flow is
75: hyperbolic, longer cycles are shadowed by the shorter ones, and
76: cycle expansions converge exponentially or even
77: super-exponentially with the cycle length\rf{hhrugh92}.
78:
79:
80: A variety of
81: methods for determining all periodic orbits up to a given length
82: have been devised and successfully implemented
83: for low-dimensional systems\rf{DasBuch,decar2,lfind,afind,mfind}.
84: For more complex dynamics, such as turbulent flows\rf{frisch}, nonlinear
85: waves\rf{cgl}, or quantum fields\rf{gauge,chfield} with high
86: (or infinite) dimensional phase spaces and complicated dynamical behavior,
87: many of the existing methods become unfeasible in practice.
88: In the most computationally demanding calculation carried out so far, Kawahara
89: and Kida\rf{KawKida01} have found two periodic solutions in a $15,422$-dimensional
90: discretization of a turbulent plane Couette flow.
91: The topology of high-dimensional flows is hard to visualize, and
92: even with a decent starting guess for the shape and location of a periodic
93: orbit, methods like the Newton-Raphson method are likely to fail. In
94: \refref{CvitLanCrete02} we have argued that variational, cost-function minimization
95: methods offer a robust alternative.
96: Here we derive, implement and discuss in detail one
97: such new variational method for finding periodic orbits in
98: general flows, and specifically high-dimensional flows.
99:
100: In essence, any numerical algorithm for finding periodic orbits is based on
101: devising a new dynamical system which possesses the desired orbit as an
102: attracting fixed point with a sizable basin of attraction. Beyond that, there is
103: much freedom in constructing such system.
104:
105: For example, the multipoint shooting method eliminates the
106: long-time exponential
107: instability of unstable orbits by splitting an orbit into a number of short segments, each with a
108: controllable expansion rate.
109: The multiple shooting
110: combined with the Newton-Raphson method is an efficient tool for
111: locating periodic orbits of maps\rf{ChristiansenDasBuch}.
112: A search for periodic orbits of a continuous time flow can be reduced to
113: a multiple shooting search for periodic orbits of a set of maps
114: by constructing
115: a set of phase space Poincar\'{e} sections such that an orbit leaving
116: one section reaches the next one in
117: a qualitatively predictable manner, without traversing other sections
118: along the way. In turbulent, high-dimensional
119: flows such sequences of sections are hard to come by.
120: One solution might be a large set of Poincar\'{e} sections,
121: with the intervening flight segments short and controllable.
122:
123: Here
124: we follow a different strategy, and discard Poincar\'{e} sections altogether;
125: we replace maps between spatially
126: fixed Poincar\'{e} sections, by maps induced by discretizing the time
127: evolution into small time steps. For sufficiently small
128: time steps such maps are small deformations of identity.
129: We distribute many points along a smooth loop $\Loop$, our
130: initial guess of
131: a cycle location and its topological layout. If both
132: the time steps and the loop deformations are taken infinitesimal,
133: a partial differential equation
134: governs the ``\descent'', a fictitious time flow of a trial loop $\Loop$
135: into a genuine
136: cycle $p$, with exponential convergence in the fictitious time
137: variable.
138: We then use methods developed for solving
139: PDEs to get the solution. Stated succinctly, the idea of our method
140: is to make an informed rough guess of what the desired cycle
141: looks like globally, and then use a variational method to drive the
142: initial guess towards the exact solution. For robustness, we replace the
143: guess of a single orbit point by a guess of an entire orbit. For
144: numerical safety we replace the Newton-Raphson iteration by the
145: ``\descent'', a differential flow that minimizes a {\costFct} computed as
146: deviation of the approximate flow from the true flow along a smooth loop
147: approximation to a cycle.
148:
149: In \refsect{sec:der} we derive the partial differential equation
150: which governs the evolution of an initial guess loop toward a cycle and
151: the corresponding {\costFct}. An extension of the method to Hamiltonian
152: systems and systems with higher time derivatives is presented in \refsect{sec:ext}.
153: Simplifications due to symmetries and details of our
154: numerical implementation of the method are discussed in \refsect{sec:nm}.
155: In \refsect{sec:applt} we test
156: the method on the H\'{e}non-Heiles system, the
157: restricted three body problem, and a weakly turbulent
158: Kuramoto-Sivashinsky system.
159: We summarize our results and
160: discuss possible improvements of the method in \refsect{sec:sum}.
161:
162:
163: \section{The \descent\ method in loop space}
164: \label{sec:der}
165:
166:
167:
168: \subsection{A variational equation for the loop evolution}
169:
170: A periodic orbit is a
171: solution $(\pSpace,\period{})$, $\pSpace \in \reals^{d}$,
172: $\period{} \in \reals$ of the {\em periodic orbit condition}
173: \beq
174: f^{\period{}}(\pSpace) = \pSpace
175: \,,\qquad \period{} > 0
176: \label{e:periodic}
177: \eeq
178: for a given flow or discrete time mapping $x \mapsto f^t({x})$.
179: Our goal is to determine periodic
180: orbits of flows defined by first order ODEs
181: \beq
182: \frac{d\pSpace}{dt}=v(\pSpace)
183: \,,\qquad
184: \pSpace \in \pS \subset \mathbb{R}^d
185: \,,\qquad
186: (\pSpace,v) \in \bf{T}\pS
187: \label{fl}
188: \eeq
189: in many (even infinitely many) dimensions $d$. Here $\pS$ is the phase space
190: (or state space) in which evolution takes place,
191: $\bf{T}\pS$ is the tangent bundle\rf{arnold92},
192: and the vector field $v(\pSpace)$ is assumed to be smooth (sufficiently
193: differentiable) almost everywhere.
194:
195: We make our initial guess at the shape and the location of a cycle $p$ by
196: drawing a loop $\Loop$, a smooth, differentiable
197: closed curve $\lSpace(s)\in \Loop \subset \pS$, where
198: $s$ is a loop parameter. As the loop is periodic, we find it convenient
199: to restrict $s$ to $[0,2\pi]$, with the periodic condition
200: $\lSpace(s)=\lSpace(s+2\pi)$.
201: % The loop parameter $s$ is otherwise unspecified at this stage.
202: Assume that $\Loop$ is close to the true cycle $p$,
203: pick $N$ pairs of nearby points
204: %$\{\lSpace_n\}_{n=1,\ldots,N}$
205: along the loop and along the cycle
206: %$\Loop$ and write
207: \bea
208: \lSpace_n &=& \lSpace(s_n)
209: \,,\qquad
210: 0 \leq s_1 < \ldots < s_N < 2\pi
211: \,, \continue
212: \pSpace_n &=& \pSpace(t_n)
213: \,,\qquad
214: 0 \leq t_1 < \ldots < t_N < \period{p}
215: \,,
216: \label{LoopDiscret}
217: \eea
218: and denote by $\delta \lSpace_n$ the deviation of a point $\pSpace_n$ on the
219: periodic orbit $p$ from the nearby point $\lSpace_n$,
220: %$\pSpace(t_n)$
221: %could be written as $\{
222: \[
223: \pSpace_n = \lSpace_n+\delta \lSpace_n
224: % \,,\qquad
225: % {n=1,\ldots,N}
226: \,.
227: \]
228: The deviations $\delta \lSpace$ are assumed small, vanishing as
229: $\Loop$ approaches $p$.
230: % represent of the loop $\Loop$ from the true periodic orbit $p$.
231:
232: The orientation of the $s$-velocity vector tangent to the loop $\Loop$
233: \[
234: \lVeloc(\lSpace)=\frac{d \lSpace}{ds}\,
235: \]
236: is intrinsic to the loop, but its
237: magnitude depends on the (still to be specified)
238: parametrization $s$ of the loop.
239:
240: At each loop point $\lSpace_n \in \Loop$ we thus have two vectors,
241: the loop tangent
242: $\lVeloc_n=\lVeloc(\lSpace_n)$ and the flow velocity
243: $v_n=v(\lSpace_n)$. Our goal is to deform $\Loop$ until
244: the directions of $\lVeloc_n$ and $v_n$ coincide for all $n=1,\ldots,N$,
245: $\; N\to \infty$, {\ie} $\Loop = p$. To match their magnitude, we introduce a
246: local time scaling factor
247: %$\lambda(\lSpace)$,
248: \begin{equation}
249: \lambda(s_n) \equiv \Delta t_n/\Delta s_n\,,
250: \label{dt}
251: \end{equation}
252: where $\Delta s_n=s_{n+1}-s_n, n=1,\ldots,N-1 \,,\Delta s_N=2\pi-(s_N-s_1)$,
253: and likewise for $\Delta t_n$. The scaling factor
254: $\lambda(s_n)$
255: %\YL{The only grid we use is the $\{s_n\}_{n=1}^N \subset [0,2 \pi]$. The phase space
256: %coordate $\lSpace$ ($\pSpace$) and the scaling function $\lambda$ are defined on
257: %this grid. In the ``\descent'' equation, they are functions of $s$ and $\tau$. Other
258: %variables like $v$ are functions of $\lSpace$ and $\lambda$ in this paper.}
259: ensures that the loop increment $\Delta s_n$ is
260: proportional to its counterpart $\Delta t_n + \delta t_n$
261: on the cycle when the loop $\Loop$ is close to the
262: cycle $p$, with $\delta t_n \to 0$ as $\Loop \to p$.
263:
264: Let
265: $ \pSpace(t) = \flow{t}{\pSpace} $
266: be the state of the system at time $t$ obtained by
267: integrating \refeq{fl}, and
268: $\jMps(\pSpace,t) = d \pSpace(t)/d \pSpace(0)$ be
269: the corresponding {\jacobianM} obtained by integrating
270: \beq
271: \frac{d\jMps}{dt} = \Mvar \jMps
272: \,,\quad
273: \Mvar_{ij}=\frac{\partial v_i}{\partial x_j}
274: \,,\qquad
275: \mbox{with } \jMps(x,0)=\mathds{1}
276: \,.
277: \label{doa}
278: \eeq
279: Since the point
280: $\pSpace_n=\lSpace_n+\delta \lSpace_n$ is on the cycle,
281: \begin{equation}
282: \flow{\Delta t_n+\delta t_n}{\lSpace_n+\delta \lSpace_n}=\lSpace_{n+1}+
283: \delta \lSpace_{n+1} \label{bd}
284: \,.
285: \end{equation}
286: Linearization
287: \[
288: \flow{\delta t}{\pSpace} \approx \pSpace + v(\pSpace) \delta t
289: \,, \qquad
290: \flow{t}{\pSpace+\delta\pSpace} \approx \pSpace(t) + \jMps(\pSpace,t) \delta\pSpace
291: \,,
292: \]
293: of \refeq{bd} about the loop point $\lSpace_n$ and the time interval
294: $\Delta t_n$ to the next cycle point
295: leads to the multiple shooting
296: Newton-Raphson equation, for any step size $\Delta t_n$:
297: \begin{equation}
298: \delta \lSpace_{n+1}-\jMps(\lSpace_n,\Delta t_n) \delta \lSpace_n -v_{n+1}
299: \delta t_n=\flow{\Delta t_n}{\lSpace_n}-\lSpace_{n+1}
300: \,.
301: \label{bd1}
302: \end{equation}
303:
304: Provided that the initial guess is sufficiently good, the Newton-Raphson iteration
305: of \refeq{bd1} generates a sequence of loops $\Loop$
306: with a decreasing {\costFct}\rf{CvitLanCrete02}
307: %\PC{Lan, recheck factor $ {2\pi / N^2}$}
308: \begin{equation}
309: \costF(\lSpace) \equiv \frac{N}{(2\pi)^2}
310: \sum_{i=1}^{N}(\flow{\Delta t_n}{\lSpace_n}-\lSpace_{n+1})^2,
311: \qquad \lSpace_{N+1}=\lSpace_1
312: \,.
313: \label{cost}
314: \end{equation}
315: The prefactor ${N/(2\pi)^2}$ makes the definition of $\costF$ consistent with
316: \refeq{funal} in the $N \to \infty$ limit.
317: If the flow is locally strongly unstable,
318: the neighborhood in which the linearization is valid
319: could be so small that the full Newton step
320: would overshoot, rendering ${\costF}$ bigger rather than smaller. In this
321: case the step-reduced, damped Newton method is needed. As proved
322: in~\refref{kellbv}, under conditions
323: satisfied here, $\costF$ decreases monotonically if
324: appropriate step size is taken.
325: If infinitesimal steps are taken, decrease of ${\costF}$ is ensured.
326: % \YL{For sufficiently small step size, if the Newton-Raphson equation \refeq{bd1}
327: % is invertible, $\costF$ is always deceasing. Actually, later we showed that
328: % $\costF$ decreases exponentially to a local minimum. In \refsect{subsec:init},
329: % we discussed the convergence problem. If the local minimum is not zero, the operator
330: % $\bar{A}$ (defined later) is not invertible and our method does not converge. We have
331: % to restart from a new initial configuration. This is the main drawback of our method
332: % as well as of many other methods. It can be circumvented by choosing proper initial
333: % conditions.}
334: We parametrize such continuous deformations of the loop by a
335: {\em fictitious time} $\tau$.
336:
337: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
338: \begin{figure}[t] %[h]
339: \centering
340: %\subfigure[annulus]{
341: (a)~~\includegraphics[width=4.0cm]{tube.eps}
342: %}
343: \hspace{0.2in}
344: %\subfigure[velocity field]{
345: (b)~\includegraphics[width=5cm]{velocField.eps}
346: %}
347: \caption{
348: (a) An annulus $\Loop(\tau)$ swept by the {\descent} flow $d{\lSpace}/d\tau$,
349: connecting
350: smoothly the
351: initial loop $\Loop(0)$ to the periodic orbit $p=\Loop(\infty)$. (b) In general the loop
352: velocity field $\lVeloc(\lSpace)$ does not coincide with $\lambda v(\lSpace)$; for a periodic
353: orbit $p$, it does so at every $x \in p$. }
354: \label{f:velocField}
355: \end{figure}
356: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
357:
358: We fix $\Delta s_n$ and proceed by $\delta \tau$ each step of the iteration,
359: {\ie}, multiply the right hand side of %Eq.
360: \refeq{bd1} by $\delta \tau$.
361: According to \refeq{dt}, the change
362: of $\Delta t_n$ with respect to $\tau$ is equal to $\delta
363: t_n=\frac{\partial \lambda}{\partial \tau}(s_n,\tau) \delta \tau \Delta
364: s_n$. As $\delta \lSpace_n=\frac{\partial}{\partial \tau}
365: \lSpace(s_n,\tau) \delta \tau$, % Eq.
366: dividing both sides of \refeq{bd1} by $\delta \tau$ yields
367: \bea
368: &&
369: \frac{d \lSpace_{n+1}}{d \tau}-\jMps(\lSpace_n,\Delta t_n) \frac{d \lSpace_n}
370: {d \tau} -v_{n+1}\frac{\partial \lambda}{\partial \tau}(s_n,\tau)
371: \Delta s_n
372: \continue
373: && \qquad\qquad\qquad\qquad\qquad
374: =\flow{\Delta t_n}{\lSpace_n}-\lSpace_{n+1}
375: \,.\label{bd2}
376: \eea
377: In the $N\rightarrow \infty$ limit, the stepsizes $\Delta s_n, \Delta
378: t_n= O(1/N)\rightarrow 0$, and we have
379: \begin{eqnarray*}
380: v_{n+1}&\!\approx\!&v_n
381: \,, \qquad \qquad \qquad
382: \lSpace_{n+1} \approx \lSpace _n+\lVeloc_n \Delta s_n
383: \,,\\
384: \jMps(\lSpace_n,\Delta t_n)\!&\!\approx \!&\! 1+\Mvar(\lSpace_n) \Delta t_n
385: \,,\quad
386: \flow{\Delta t_n}{\lSpace_n} \approx \lSpace_n+v_n \Delta t_n
387: \,.
388: \end{eqnarray*}
389: Substituting into %Eq.
390: \refeq{bd2} and using the scaling relation \refeq{dt},
391: % fact that $\Delta t_n = \lambda \, \Delta s_n$
392: we obtain
393: \begin{equation}
394: \frac{\partial^2 \lSpace}{\partial s \partial \tau}-\lambda \Mvar
395: \frac{\partial \lSpace}{\partial \tau}-v\frac{\partial \lambda}{\partial \tau}
396: =\lambda v-\lVeloc.
397: \label{bd3}
398: \end{equation}
399: This PDE, which describes the evolution
400: of a loop $\Loop(\tau)$ % $\lSpace(s,\tau)$
401: toward a periodic
402: orbit $p$, is the central result of this paper.
403: The family of loops so generated is parametrized
404: by $\lSpace=\lSpace(s,\tau)\in \Loop (\tau)$, where $s$ denotes the
405: position along
406: the loop, and the fictitious time $\tau$ parametrizes the deformation of the loop,
407: see \reffig{f:velocField}(a).
408: We refer to this infinitesimal step
409: version of the damped Newton-Raphson method as the ``\descent''.
410:
411: The important feature of this equation is that a decreasing cost functional exists.
412: %\PC{do we need ``Lyapunov functional''? We have the cost function, why two
413: %names for the same function?}
414: Rewriting \refeq{bd3} as
415: \begin{equation}
416: \frac{\partial}{\partial \tau}(\lVeloc-\lambda v)
417: =-(\lVeloc-\lambda v) \,,
418: \label{cc0}
419: \end{equation}
420: we have
421: \begin{equation}
422: \lVeloc-\lambda v= %B(s)
423: e^{-\tau}(\lVeloc-\lambda v)|_{\tau=0}
424: \,,
425: \label{cc}
426: \end{equation}
427: so the fictitious time $\tau$ flow decreases the {\costFct}al
428: \beq
429: \costF[\lSpace]=\frac{1}{2 \pi}\oint_{\Loop(\tau)} ds \,
430: \left(\lVeloc(\lSpace)-\lambda v(\lSpace)\right)^2
431: \label{funal}
432: \eeq
433: monotonically as the loop evolves toward the cycle.
434:
435: At each iteration step the differences of the loop tangent velocities and
436: the dynamical flow velocities are reduced by the {\descent}.
437: As $\tau \rightarrow \infty$, the fictitious time flow alignes
438: the loop tangent $\lVeloc$ with
439: the dynamical flow vector $\lVeloc=\lambda v$, and the loop
440: $\lSpace(s,\tau) \in \Loop(\tau)$, see \reffig{f:velocField}(b),
441: converges to a genuine periodic orbit $p = \Loop(\infty)$ of
442: the dynamical flow $\dot{\pSpace}=v(\pSpace)$.
443: Once the cycle $p$ is reached, by \refeq{dt},
444: $\lambda(s,\infty)=\frac{dt}{ds}(\lSpace(s,\infty))$, and the cycle period is given by
445: \[
446: \period{p}=\int_0^{2\pi} \lambda(\lSpace(s,\infty))ds
447: \,.
448: \]
449: Of course, as at this stage we have already identified the cycle, we may
450: pick instead an initial point
451: on $p$ and calculate the period by a direct integration of the
452: dynamical equations \refeq{fl}.
453:
454:
455: \subsection{Marginal directions and accumulation of loop points}
456: \label{sect:marg}
457:
458: Numerically, two perils lurk in a direct implementation of the {\descent}
459: \refeq{bd3}.
460:
461: First, when a cycle is reached, it remains a cycle under a
462: cyclic permutation of the representative points, so on the cycle
463: the operator
464: \begin{eqnarray*}
465: \bar{A}=\frac{\partial}{\partial s}-\lambda A
466: \end{eqnarray*}
467: has a marginal eigenvector $v(\lSpace(s))$ with
468: eigenvalue $0$.
469: If $\lambda$ is fixed, as the loop approaches the cycle, \refeq{bd3}
470: approaches its limit
471: \[
472: \bar{A}\,\frac{\partial x}{\partial \tau}=0 \,.
473: \]
474: Therefore, on the cycle, the operator $\bar{A}^{-1}$ becomes
475: singular and the numerical woes arise.
476:
477: The second potential peril hides in the freedom of choosing the loop
478: (re-)parametrization. Since $s$ is related to the time $t$
479: by the yet unspecified factor $\lambda(s,\tau)$, uneven distributions of the sampling
480: points over the loop $\Loop$ could arise, with the numerical discretization
481: points $\lSpace_n$
482: clumping densely along some segments of $\Loop$ and leaving big gaps elsewhere,
483: thus degrading the numerical smoothness of the loop.
484:
485: We remedy these difficulties by imposing constraints on
486: \refeq{bd3}. In our calculation for Kuramoto-Sivashinsky system
487: of \refsect{sec:applt}, the first
488: difficulty is dealt with by introducing one Poincar\'{e} section, for example,
489: by fixing one coordinate of one of the sampling points,
490: $\lSpace_1(s_1,\tau)=const$.
491: This breaks the translational invariance along the cycle.
492: Other types of constraints might be better suited to a specific
493: problem at hand. For example, we can demand that the average
494: displacement of the sampling points along the loop vanishes, thus avoiding
495: a spiraling descent towards the desired cycle.
496:
497: We deal with the second potential difficulty by choosing a particularly
498: simple loop parametrization.
499: So far, the
500: parametrization $s$ is arbitrary and there is much freedom in choosing the best
501: one for our purposes.
502: We pick $s-$ and $\tau-$independent constant scaling
503: $\lambda(s, \tau) = \lambda$. % = {\period{p}}/{2\pi}$.
504: With uniform grid size $\Delta s_n=\Delta s$ and fixed $\lambda$,
505: the loop parameter $s=t/\lambda$ is proportional to time $t$, and
506: the discretization \refeq{bd3} distributes the sampling points
507: along the loop evenly in time.
508: As the loop approaches a cycle,
509: $\frac{\partial \lSpace}{\partial \tau}$ is numerically obtainable from
510: \refeq{bd3}, and on the cycle the period is given by $\period{p}=2\pi \lambda$.
511:
512: Even though this paper focuses on searches for periodic orbits, the {\descent} is
513: a general method. With appropriate modifications of boundary conditions and
514: scaling of time,
515: \refeq{bd3} can be adapted to determination of homoclinic or heteroclinic
516: orbits
517: between equilibrium points or periodic orbits of a flow, or more
518: general boundary value
519: problems. Applied to 2-point boundary value problems, {\descent} is similar
520: to the quasilinearization\rf{bl} but has
521: the advantage that the free parameter $\lambda(s,\tau)$ is available
522: for adjusting scales in the problem and that searches
523: can be restricted to phase space submanifolds of interest.
524: A simple example of a restriction to a submanifold are searches for cycles of
525: a given energy,
526: constrained to the $H(q,p)=E$ energy shell
527: in the phase space of a Hamiltonian system. Furthermore,
528: as we shall show now, the symplectic structure of Hamilton's equations
529: greatly reduces the dimensionality of the submanifold that we need to consider.
530:
531: \section{Extensions of \descent}
532: \label{sec:ext}
533:
534: In classical mechanics
535: particle trajectories are also solutions of a variational principle,
536: the Hamilton's variational principle.
537: For example, one can determine a periodic orbit of a
538: billiard by wrapping around a rubber band of a roughly correct topology, and then
539: moving the points along the billiard walls until the length ({\ie}, the action)
540: of the
541: rubber band is extremal (maximal or minimal under infinitesimal changes of the
542: boundary points). Note that the
543: extremization of action requires only
544: $D$ configuration coordinate variations,
545: not the full $2D$-dimensional phase space variations.
546:
547: Can we exploit this property of the Newtonian mechanics to reduce the dimenionality
548: of our variational calculations?
549: The answer is yes, and easiest to understand in terms of the Hamilton's
550: variational principle which states that classical trajectories are extrema
551: of the Hamilton's principal function (or, for fixed energy $E$, the action $S=R+Et$)
552: \[
553: R(q_1,t_1;q_0,t_0) = \int_{t_0}^{t_1} \! dt \,{\cal L}(q(t), {\dot q}(t),t)
554: \,,
555: \]
556: where
557: $ {\cal L}(q, {\dot q},t)$ is the Lagrangian.
558: Given a loop $\Loop(\tau)$ we can compute
559: not only the tangent ``velocity'' vector $\tilde{v}$, but also
560: the local loop curvature or ``acceleration'' vector
561: \[
562: \tilde{a} =
563: \frac{\partial^2 \lSpace}{\partial s^2}
564: \,,
565: \]
566: and indeed, as many $s$ derivatives as needed.
567: Matching the dynamical acceleration $a(\lSpace)$ (assumed to be functions of
568: $\lSpace$ and $v(\lSpace)$)
569: with the loop ``acceleration'' $\tilde{a}(\lSpace)$
570: results in a new {\costFct} and the corresponding PDE \refeq{cc0} for the evolution of the loop
571: \[
572: \frac{\partial}{\partial \tau}(\tilde{a}-\lambda^2 a)=-(\tilde{a}-\lambda^2 a)
573: \,.
574: \]
575: We use $\lambda^2$ instead of $\lambda$ in order to keep the notation
576: consistent with \refeq{dt}, {\ie} $t=\lambda \,s$.
577: % \PC{I think you are assuming gobally constant $\lambda$, otherwise I
578: % would expect:
579: % \[ {d^2 \lSpace(s) \over d s^2} - {d^2 \pSpace(t) \over d s^2} =
580: % {d^2 \lSpace(s) \over d s^2}
581: % - \left({d t \over d s}\right)^2{d^2 \pSpace \over d t^2}
582: % - {d^2 t \over d s^2}{d \pSpace \over d t}
583: % \,.
584: % \]
585: % }
586: %\YL{Actually, $\lambda$ is not a global constant here. See the footnote after \refeq{dt}.
587: %It is a function of $s$ and $\tau$, but not a function of $\lSpace$. At each grid point
588: %$s$ for fixed $\tau$, there is only one phase space coordinate and two velocities or
589: %accelerations defined. That is why I prefer to merge the notation $\lSpace$
590: %with $\pSpace$ but give credit to $\lVeloc$ and $v$.}
591: Expressed
592: in terms of the loop variables $\lSpace(s)$, the above equation becomes
593: \bea
594: &&
595: \frac{\partial^3
596: \lSpace}{\partial^2 s \partial \tau}-\lambda \frac{\partial a}{\partial v} %\cdot
597: \frac{\partial^2 \lSpace}{\partial s \partial \tau}-\lambda^2 \frac{\partial a}
598: {\partial \lSpace} %\cdot
599: \frac{\partial \lSpace}{\partial \tau}
600: +\left(\frac{\partial a}{\partial v} %\cdot
601: \frac{\partial \lSpace}{\partial s}-2\lambda a
602: \right)\!
603: \frac{\partial \lambda}{\partial \tau}
604: \continue
605: && \qquad\qquad\qquad\qquad\qquad
606: =\lambda^2 a-\tilde{a} \, ,
607: \label{hfl}
608: \eea
609: where $v=\frac{\partial \lSpace}{\lambda \partial s}$. Although \refeq{hfl}
610: looks more complicated than \refeq{bd3}, in numerical fictitious time integrations, we
611: are rewarded by having to keep only half of the phase space variables.
612:
613: More generally, if a differential equation has the form:
614: \begin{equation}
615: \pSpaceDer{m}=f(\pSpace,\pSpaceDer{1},\cdots,\pSpaceDer{m-1}) \, , \label{gf}
616: \end{equation}
617: where $\pSpaceDer{k}=\frac{d^k \pSpace}{dt^k},\, k=1,\cdots,m$ and $\pSpace \in
618: \mathbb{R}^d$,
619: the same technique can be used to match the highest derivatives
620: $\lambda^m \pSpaceDer{m}$ and $\lSpaceDer{m}$,
621: \[
622: \frac{\partial}{\partial \tau}(\lSpaceDer{m}-\lambda^m \pSpaceDer{m})
623: =-(\lSpaceDer{m}-\lambda^m \pSpaceDer{m}) \, ,
624: \]
625: with
626: $\lSpaceDer{m}=\frac{\partial^m}{\partial s^m}\lSpace(s)$ calculated directly from
627: $\lSpace(s)$ on the loop by differentiation.
628: In loop variables $\lSpace(s)$ we have,
629: \bea
630: &&
631: \frac{\partial^{m+1} \lSpace}{\partial s^m \partial \tau}-\lambda^m
632: \sum_{k=0}^m
633: \frac{\partial f}{\partial \pSpaceDer{k}} \cdot \frac{\partial}{\partial \tau}
634: \frac{\partial^{k} \lSpace}
635: {\lambda^k \partial s^{k}}-m \lambda^{m-1} \lSpaceDer{m} \frac{\partial \lambda}
636: {\partial \tau}
637: \continue
638: && \qquad\qquad\qquad\qquad\qquad
639: =\lambda^m \pSpaceDer{m}-\lSpaceDer{m}
640: \, , \label{gen}
641: \eea
642: where $\pSpace=\pSpaceDer{0}$ and $\lSpaceDer{k}=\frac{\partial^k
643: \lSpace}{\lambda^k \partial^k s},\,k=1,\cdots,m-1$ are assumed.
644: % \YL{If $f$ depends
645: % on $x^{(i)}$ with $i \neq 0$, in the coordinate $x$ space, we have no idea of
646: % what $\pSpace^{(i)}$ is generally. On a loop, we can calculate $\lSpace^{(i)}$ but
647: % $\pSpace{(i)}$ is not really defined (they are only defined on a true trajectory). If we
648: % make the foregoing assumption about $\pSpace^{(i)}$ for $i=1,\cdots,m-1$, they
649: % converge to the well-defined correspondents on a periodic orbit when
650: % $\lambda \lSpace^{(m)}$ converges to $\pSpace^{(m)}$.}
651: Conventionally, % Eq.
652: \refeq{gf} is
653: converted to a system of $m d$ first order differential equations,
654: whose discretized derivative (see \refeq{dd} below) are banded matrices
655: with band width of $5md$. Using \refeq{gen},
656: we only need $d$ equations for the same accuracy and the corresponding band
657: width is $(m+4)d$. The computing load has been greatly reduced, the
658: more so the larger $m$ is.
659: Nevertheless, choice of a good initial loop guess and
660: visualization of the dynamics are always aided by a plot of
661: the orbit in the full $md$-dimensional phase space, where loops cannot
662: self-intersect and topological features of the flow is exhibited
663: more clearly.
664:
665:
666: \section{Implementation of \descent}
667: \label{sec:nm}
668:
669: As the loop points satisfy a periodic boundary condition,
670: it is natural to employ truncated discrete Fast Fourier Transforms (FFT)
671: in numerical integrations of % Eq.
672: \refeq{bd3}. Since we are
673: interested only in the final, stationary cycle $p$, the accuracy of the
674: fictitious time integration is not crucial; all we have to ensure is the smoothness of the loop
675: throughout the integration.
676: The Euler integration with fairly large time steps $\delta \tau$ suffices.
677: The computationally most onerous step in implementation of
678: the {\descent} is the inversion of large matrix $\bar{A}$ in \refeq{bd3}.
679: When the dimension of the dynamical phase
680: space of \refeq{fl} is high, the inversion of
681: $\bar{A}$ needed to get $\frac{\partial \lSpace}{\partial \tau}$
682: takes most of the integration time, making the evolution extremely slow.
683: This problem is partially
684: solved if the finite difference methods are used. The large matrix $\bar{A}$
685: then becomes sparse and the inversion can be done far more quickly.
686:
687: \subsection{Numerical implementation}
688:
689: In a discretization of a loop, numerical
690: stability requires accurate
691: discretization of loop derivatives such as
692: \[
693: \lVeloc_n \equiv
694: \left.\frac{\partial \lSpace}{\partial s}\right|_{\lSpace = \lSpace(s_n)}
695: \approx (\hat{D}\lSpace)_n
696: \,.
697: \]
698: In our numerical work we use the four-point approximation\rf{brand03},
699: {\small
700: \beq
701: \hat{D} =
702: \frac{1}{12h}
703: \!
704: \left( \begin{array}{ccccccccccc} 0&8&-1&&&\qquad
705: &&&&1&-8 \\ -8&0&8&-1&&\qquad &&&&&1 \\ 1&-8&0&8&-1&\qquad &&&&&\\
706: &&&&&\cdots&&&&&\\
707: &&&&&\qquad &1&-8&0&8&-1 \\
708: -1&&&&&\qquad &&1&-8&0&8 \\
709: 8&-1&&&&\qquad &&&1&-8&0
710: \end{array}\right)
711: \label{dd}
712: \eeq
713: }%end {\small
714: where $h={2 \pi}/{N}$.
715: Here, each entry represents a $[d\! \times\! d ]$ matrix, $8 \to 8\mathds{1}$,
716: {\em etc.},
717: with blank spaces are filled with zeros.
718: The two $[2d\! \times\! 2d]$ matrices
719: \[
720: M_1=\MatrixII{\mathds{1}}{-8\mathds{1}}{0}{\mathds{1}} \,, \qquad
721: M_2=\MatrixII{-\mathds{1}}{0}{8\mathds{1}}{-\mathds{1}} \,,
722: \]
723: located at the top-right and bottom-left corners take care of
724: the periodic boundary condition.
725:
726: The discretized version
727: of \refeq{bd3} with a fictitious time Euler step $\delta \tau$ is
728: \beq
729: \MatrixII
730: {\hat{\Mvar}}{\hat{v}}
731: {\hat{a}}{0}
732: \VectorII {\delta \hat{x}}{\delta \lambda}
733: =\delta \tau \VectorII{\lambda \hat{v}-\hat{\lVeloc}} {0}
734: \,,
735: \label{mform}
736: \eeq
737: where
738: \bea
739: \hat{\Mvar} &=& \hat{D}+\mathrm{diag}[\Mvar_1,\Mvar_2,\cdots,\Mvar_N]
740: \,,
741: \nnu
742: \eea
743: with $\Mvar_n=\Mvar(\lSpace(s_n))$ defined in \refeq{doa}, and
744: \bea
745: \hat{v} &=& (v_1,v_2,\cdots,v_N)
746: \,,\qquad \mbox{ with } v_n=v(\lSpace(s_n))\,,
747: \continue
748: \hat{\lVeloc} &=& (\lVeloc_1,\lVeloc_2,\cdots,\lVeloc_N)
749: \,,\qquad \mbox{ with }
750: \lVeloc_n=\lVeloc(\lSpace(s_n)) \,,
751: \nnu
752: \eea
753: are the two vector fields that we want to match everywhere along the loop.
754: $\hat{a}$ is an $Nd$ dimensional row vector which imposes the constraint on
755: the coordinate variations
756: $\delta\hat{x}=(\delta \lSpace_1,\delta \lSpace_2,\cdots,\delta \lSpace_N)$.
757: The discretized {\descent} \refeq{mform} is
758: an infinitesimal time step variant of the multipoint (Poincar\'{e} section)
759: shooting equation for flows\rf{ChristiansenDasBuch}.
760: In order to solve for the deformation of the loop coordinates and period, $\delta \hat{x}$
761: and $\delta \lambda$,
762: we need to invert the $[(N\, d+1)\! \times\! (N\, d+1)]$ matrix on the left hand side
763: of \refeq{mform}.
764:
765: In our numerical work, this matrix is inverted using the banded LU decomposition
766: on the embedded
767: band-diagonal matrix, and the Woodbury formula\rf{nr} on the cyclic, border
768: terms. The LU decomposition takes most of the computational
769: time and considerably slows down the fictitious time integration. We speed
770: up the integration
771: by a new inversion scheme which relies on
772: the smoothness of the flow in the loop space.
773: It goes as follows. Once we have the
774: LU decomposition at one step, we use it to approximately invert the
775: matrix in the next step, with accurate inversion achieved
776: by the iterative approximate inversions\rf{nr}. In
777: our applications we find that a single LU decomposition can be used
778: for many $\delta \tau$ evolution steps.
779: The further we go, the more
780: iterations at each step are needed to implement the inversion. After
781: the number of such iterations exceeds some given fixed maximum
782: number, we perform another LU decomposition and proceed as before. The
783: number of integration steps following one decomposition is an
784: indication of the smoothness of the evolution, and we adjust accordingly the
785: integration step size $\delta \tau$: the greater the number, the bigger the
786: step size. As the loop approaches a cycle, the
787: evolution becomes so smooth that the step size can be brought all the
788: way up to $\delta \tau = 1$, the full undamped Newton-Raphson iteration step.
789: In practice, one can
790: % \YL{Of course you may start directly with the required accuracy.}
791: start with a small but reasonable number of points,
792: in order to get a coarse solution of relatively low accuracy. After
793: achieving that, the refined guess loop can be constructed
794: by interpolating more points, and proceed with
795: for a more accurate calculation in which $\delta \tau$ can be set as
796: large as the full Newton step $\delta \tau = 1$,
797: recovering the rapid quadratic convergence
798: of the Newton-Raphson method.
799:
800: It is essential that
801: the smoothness of the loop is maintained throughout
802: the calculation. We monitor the smoothness by checking the
803: Fourier spectrum of $\lSpace(\cdot,\tau)$. An unstable difference scheme
804: for loop derivatives might lead to unbounded sawtooth oscillations\rf{ns}.
805: A heuristic local linear stability
806: analysis (described in \rf{lls}) indicates that our scheme is stable, and that
807: the high frequency components do not generate instabilities.
808:
809:
810: \subsection{Initialization and convergence}
811: \label{subsec:init}
812:
813: As in any other method, a qualitative understanding of the dynamics is
814: a prerequisite to successful cycle searches. We start by numerical
815: integration with the dynamical system \refeq{fl}. Numerical experiments reveal
816: regions where a trajectory spends most of its life, giving us
817: the first hunch as to how to initialize a loop. We take the FFT of some nearly
818: recurred orbit segment and keep only the lowest
819: frequency components. The inverse Fourier transform back to the phase
820: space yields a smooth loop that we use as our initial guess.
821: Since any generic orbit segment is not closed and might exhibit large
822: gaps, the Gibbs phenomenon can take the initial loop so constructed
823: quite far away from the region of interest. We deal with this problem by
824: manually deforming the orbit segment into a closed loop
825: before performing the FFT. Searching for longer
826: cycles with multiple circuits requires more delicate initial
827: conditions. The hope is that a few short cycles can help us establish
828: an approximate symbolic dynamics, and guess for longer cycles
829: can be constructed by cutting and glueing
830: the short, known ones. For low dimensional systems, such methods yield
831: quite good systematic initial guesses for longer cycles \rf{ks}.
832:
833: An alternative way to initialize the search is by utilizing adiabatic
834: deformations of dynamics, or the homotopy
835: evolution\rf{cconley}. If the dynamical system \refeq{fl} depends on a
836: parameter $\mu$,
837: short cycles might survive as $\mu$ varies passing through a family of
838: dynamical systems, giving in the process birth to
839: new cycles through sequences of bifurcations. Most short
840: unstable cycles vary little for small changes
841: of $\mu$. So, a cycle existing for parameter value $\mu_1$
842: can be chosen as the initial
843: trial loop for a nearby cycle surviving a small change $\mu_1 \rightarrow \mu_2$.
844: In practice, one or two iterations often suffice to find the new cycle.
845:
846: A good choice of the initial loop significantly expedites the
847: computation, but there are more reasons why good initial loops are crucial.
848: First of all, if we break the translational invariance by
849: imposing a constraint such as $\lSpace_1(s_1,\tau)=c$, we have to
850: make sure that both the initial loop and
851: the desired cycle intersects this Poincar\'{e} plane. Hence,
852: the initial loop cannot be wildly different from the desired cycle.
853: Second, in view of %Eq.
854: \refeq{cc}, the loop
855: always evolves towards a local minimum of the {\costFct}al \refeq{funal},
856: %along a curve over which
857: with discretization points moving along the $\lVeloc-\lambda v$
858: fixed direction, determined by the
859: initial condition. If the local minimum corresponds to a zero of
860: the {\costFct}al, we obtain a true cycle of the dynamical flow \refeq{fl}.
861: However, if the value of {\costFct}al is not equal to zero at the
862: minimum while the gradient is zero, \refeq{mform} yields a singular matrix
863: $\hat{\Mvar}$. In such cases the search has to be abandoned and restarted with a
864: new initial loop guess. In the periodic orbit searches of \refsect{sec:applt}
865: starting with blind
866: initial guesses (guesses unaided by a symbolic dynamics partition), such local
867: minima were encountered in about $30\%$ of cases.
868:
869: \subsection{Symmetry considerations}
870:
871: The system under consideration often possesses certain
872: symmetries. If this is the case, the symmetry should be
873: both be feared for possible marginal eigen-directions, and be embraced as a guide
874: to possible simplifications of the numerical calculation.
875:
876: If the dynamical system equations \refeq{fl} are invariant under a discrete symmetry,
877: the concept of fundamental domain\rf{CvitaEckardt,DasBuch} can be
878: utilized to reduce the length of the initial loop when searching for a
879: cycle of a given symmetry.
880: In this case, we need discretize only an irreducible segment of the
881: loop, decreasing significantly the dimensionality of the loop representation. Other
882: parts of the loop are replicated by symmetry operations,
883: with the full loop tiled by copies of the fundamental domain segment.
884: The boundary conditions are not periodic any longer,
885: but all that we need to do is modify the cyclic terms. Instead of using
886: $M_1$ and $M_2$ in \refeq{dd}, we use $M_1 Q$ and $M_2 Q^{-1}$, where $Q$ is
887: the relevant symmetry operation that maps
888: the fundamental segment to the neighbor that precedes it.
889: In this way, a fraction of the points represent the
890: cycle with the same accuracy, speeding up the search considerably.
891:
892: If a continuous symmetry is present, it may complicate the situation at
893: first glance but becomes something that we can take advantage of after careful
894: checking.
895: For example, for a Hamiltonian system unstable cycles may form continuous
896: families\rf{henonrtb2,condisfam1}, with one or more members of a family belonging
897: to a given constant energy surface. In order to cope with the marginal eigendirection
898: associated with such continous family,
899: we search for a cycle on a particular energy surface by
900: replacing the last row of equation \refeq{mform} by an energy shell
901: constraint\rf{ChristiansenDasBuch}. We
902: put one point of the loop, say $\lSpace_2$, on the constant energy
903: surface $H(\lSpace)=E$, and impose the constraint
904: $\bigtriangledown H(\lSpace_2) \cdot \delta \lSpace_2=0$, so as to
905: keep $\lSpace_2$ on the surface for all $\tau$. The integration of % Eq.
906: \refeq{bd3} then automatically
907: brings all other loop points to the same energy surface. Alternatively, we
908: can look for a cycle of given fixed period $T$ by fixing $\lambda$ and
909: dropping the constraint in the bottom line of % Eq.
910: \refeq{mform}. These two approaches are conjugate to each other, both
911: needed in applications.
912: In most cases, they are equivalent. One exception is
913: the harmonic oscillator for which the oscillations have identical
914: period but different energy. Note that in both cases the translational
915: invariance is restored, as we have discarded the Poincar\'{e} section condition
916: of \refsect{sect:marg}. As explained in\rf{decar2}, this causes no trouble in
917: numerical calculations.
918:
919:
920: \section{Applications}
921: \label{sec:applt}
922:
923: We have checked that the iteration of \refeq{mform} yields quickly and robustly
924: the short unstable cycles for standard models of low-dimensional dissipative flows such as
925: the R\"{o}ssler system\rf{ross}. A more daunting challenge are searches for cycles
926: in Hamiltonian flows, and searches for spatio-temporally periodic solutions of PDEs.
927: In all numerical examples that follow, the convergence condition is $\costF< 10^{-5}$.
928:
929: \subsection{H\'enon-Heiles system and restricted three-body problem}
930:
931: First, we test the
932: Hamiltonian version of the \descent\ derived in \refSect{sec:ext}
933: by applying the method to two Hamiltonian systems, both with two degrees of freedom.
934: In both cases, our initial loop guesses are rather arbitrary
935: combinations of trigonometric functions. Nevertheless, the
936: observed convergence is fast.
937:
938: The H\'enon-Heiles system\rf{hhsys} is a standard model in celestial
939: mechanics, described by the Hamiltonian
940: \begin{equation}
941: H=\frac{1}{2}(p_x^2+p_y^2+x^2+y^2)+x^2 y-\frac{y^3}{3}
942: \,.
943: \end{equation}
944: It has a time reversal symmetry and a three-fold discrete spatial symmetry.
945: \refFig{f:hh} shows a typical application
946: of \refeq{hfl}, with the \descent\ search restricted to the configuration space.
947: The initial loop, \reffig{f:hh}(a), is a rather coarse initial guess. We fix arbitrarily
948: the scaling $\lambda=2.1$, {\ie}, we search for a cycle $p$ of
949: the fixed period $\period{p}=13.1947$, with no constraint on the energy.
950: \refFig{f:hh}(b) shows the cycle found by the \descent, with energy $E=0.1794$,
951: and the full discrete symmetry of the Hamiltonian.
952: This cycle persists adiabatically for a small range of values of $\lambda$;
953: with $\lambda$ changed
954: much, the \descent\ takes the same initial loop into other cycles.
955: \refFig{f:hh}(c) verifies that the {\costFct}al
956: $\costF$ decreases exponentially with slope -2 throughout the $\tau=[0,10]$
957: integration interval, as predicted by \refeq{cc}.
958: The points get more and more sparse as $\tau$ increases, because
959: our numerical implementation adaptively chooses bigger and bigger step
960: sizes $\delta \tau$.
961: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
962: \begin{figure}[t] %[h]
963: %\centering
964: (a)~\includegraphics[width=1.8in]{heils1.eps}
965: (b)~\includegraphics[width=1.8in]{heils2.eps}
966: (c)~\includegraphics[width=1.8in]{con1.eps}
967: \caption{
968: The H\'enon-Heiles system in a chaotic region:
969: (a) An initial loop $L(0)$, and
970: (b) the unstable periodic orbit $p$ of period $\period{}=13.1947$ reached by
971: the {\descent} \refeq{hfl}.
972: (c) The exponential decrease of the \costFct, $\ln (\costF) \approx -2.0502\,\tau+6.0214$.
973: }
974: \label{f:hh}
975: \end{figure}
976: %converges in 601 steps.
977: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
978:
979: In the H\'{e}non-Heiles case, the accelerations $a_x,a_y$ depend only on the
980: configuration variables $x,y$. More generally, the accelerations could also
981: depend on $\dot{x},\dot{y}$. Consider as an example the equations of motion for the
982: restricted three-body problem\rf{rtb},
983: \begin{eqnarray}
984: \ddot{x} &=& 2\dot{y}+x
985: -(1-\mu)\frac{x+\mu}{r_1^3}-\mu \frac{x-1+\mu}{r_2^3} \,,\nonumber \\
986: \ddot{y} &=& -2\dot{x}+y
987: -(1-\mu)\frac{y}{r_1^3}-\mu \frac{y}{r_2^3} \label{rtbeq}
988: \,,
989: \end{eqnarray}
990: where $r_1=\sqrt{(x+\mu)^2+y^2},\, r_2=\sqrt{(x-1+\mu)^2+y^2}$. These equations
991: describe the motion of a test particle in a rotating frame under the
992: influence of the gravitational force of two heavy bodies with masses $1$ and
993: $\mu \ll 1$ fixed at $(-\mu,0)$ and $(1-\mu,0)$ in the $(x,y)$ coordinate frame. The
994: stationary solutions of % Eq.
995: \refeq{rtbeq} are called the Lagrange points,
996: corresponding to a circular motion of the test particle in phase
997: with the rotation of the heavy bodies. The periodic solutions in the rotating
998: frame correspond
999: to periodic or quasi-periodic motion of the test particle in the inertial
1000: frame. \refFig{rtb} shows an initial loop and the cycle to
1001: which it converges, in the rotating frame. Although the cycle looks simple,
1002: the {\descent} requires advancing in small $\delta \tau$ steps in
1003: order for the initial loop to converge to it.
1004: %\YL{Viswanath's solutions are different as they are
1005: %periodic solutions in both the rotating and inertial frame.}
1006: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1007: \begin{figure}[t]
1008: (a)~\includegraphics[width=2.0in]{rtb1.eps}
1009: \hspace{0.2in}
1010: (b)~\includegraphics[width=2.0in]{rtb2.eps}
1011: \caption{
1012: (a) An initial loop $L(0)$, and
1013: (b) the unstable periodic orbit $p$ of period $\period{p}=2.7365$ reached by
1014: the {\descent} \refeq{hfl}, the restricted
1015: three body problem \refeq{rtbeq} in the chaotic regime, $\mu=0.04$.
1016: }
1017: \label{rtb}
1018: \end{figure}
1019: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1020:
1021: In order to successfully
1022: apply the Hamiltonian version of the {\descent} \refeq{hfl}, we
1023: have to ensure that the test
1024: particle keeps a finite distance from the origin. If
1025: a cycle passes very close to one of the heavy bodies,
1026: the acceleration can become so large that
1027: our scheme of uniformly distributing the loop points in time might fail to
1028: represent the loop faithfully. Another distribution scheme is required in
1029: this case, for example, making the density of points proportional to the
1030: magnitude of acceleration.
1031:
1032: \subsection{Periodic orbits of Kuramoto-Sivashinsky system}
1033:
1034: The Kuramoto-Sivashinsky equation arises as an amplitude equation
1035: for interfacial instability in a variety of contexts\rf{ku,siv}.
1036: In 1-dimensional space, it reads
1037: \begin{equation}
1038: u_t=(u^2)_x-u_{xx}-\nu u_{xxxx}, \label{kseq}
1039: \end{equation}
1040: where $\nu$ is a ``super-viscosity'' parameter which controls the rate of
1041: dissipation and $(u^2)_x$ is the nonlinear convection term.
1042: As $\nu$ decreases, the system undergoes a series of bifurcations,
1043: leading to increasingly turbulent, spatio-temporally chaotic dynamics.
1044:
1045: If we impose the periodic boundary condition
1046: $u(t,x+2\pi)=u(t,x)$ and choose to study only the odd solutions
1047: $u(-x,t)=-u(x,t)$, $u(x,t)$ can be expanded in a discrete spatial
1048: Fourier series\rf{ks},
1049: \begin{equation}
1050: u(x,t)=i\sum_{k=-\infty}^{\infty} a_k(t) e^{ikx}, \label{expan}
1051: \end{equation}
1052: where $a_{-k}=-a_k \in \mathbb{R}\,$. In terms of the Fourier components,
1053: PDE \refeq{kseq} becomes an infinite ladder of ODEs,
1054: \begin{equation}
1055: \dot{a_k}=(k^2-\nu k^4)a_k-k\sum_{m=-\infty}^{\infty}a_m a_{k-m} \,. \label{ksf}
1056: \end{equation}
1057: In numerical simulations we work with the Galerkin truncations of the
1058: Fourier series since in the neighborhood of the strange attractor the
1059: magnitude of $a_k$ decreases very fast with
1060: $k$, high frequency modes playing a negligible role
1061: in the asymptotic dynamics. In this way Galerkin truncations reduce the
1062: dynamics to a finite but large number of ODEs. We work with $d=32$ dimensions
1063: in our numerical
1064: calculations. In \refref{ks}, multipoint shooting
1065: has been successfully applied to obtain periodic orbits close to the
1066: onset of spatiotemporal chaos ($\nu=0.03$). In this regime, our method
1067: is so stable that big time steps $\delta \tau$ can
1068: be employed even at the initial guesses, leading to extremely fast convergence.
1069: We attribute this robustness to the simplicity of
1070: the structure of the attractor at high viscosity values.
1071: %
1072: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1073: \begin{figure}[t!]
1074: (a)~\includegraphics[width=1.8in]{ks11.eps}
1075: \hspace{0.2in}
1076: (b)~\includegraphics[width=1.8in]{ks12.eps}\\
1077: (c)~\includegraphics[width=1.8in]{ks21.eps}
1078: \hspace{0.2in}
1079: (d)~\includegraphics[width=1.8in]{ks22.eps}
1080:
1081: \caption{
1082: The Kuramoto-Sivashinsky system in a spatio-temporally
1083: turbulent regime (viscosity parameter
1084: $\nu=0.015$, $d=32$ Fourier modes truncation).
1085: (a) An initial guess $L_1$, and
1086: (b) the periodic orbit $p_1$ of period $\period{1}=0.744892$ reached by the \descent.
1087: (c) Another initial guess $L_2$, and
1088: (d) the resulting periodic orbit $p_2$ of period $\period{2}=1.184668$.}
1089: \label{f:ks1}
1090: \end{figure}
1091: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1092:
1093: The challenge comes with decreasing $\nu$, with the dynamics turning more and
1094: more turbulent.
1095: Already at $\nu=0.015$, the system is moderately turbulent and
1096: the phase space portraits of the dynamics reveal a complex labyrinth of ``eddies''
1097: of different scales and orientations. While the highly unstable nature of orbits
1098: and intricate structure of the invariant set hinder
1099: applications of conventional cycle-search routines, in this setting
1100: our variational method shines through. We design rather arbitrary initial loops
1101: from numerical trajectory segments, and the calculation proceeds
1102: as before, except that now
1103: a small $\delta \tau$ has to be used initially to ensure numerical stability.
1104: Topologically different loops are very likely to result in different cycles,
1105: while some initial loop guesses my lead to local nonzero minima of the {\costFct}al $\costF$.
1106: As explained in \refSect{sec:nm}, in such cases the method diverges, and
1107: the search is restarted with a new inital loop guess.
1108: %\YL{Actually, I choose the initial loops from different parts of {\em one} typical
1109: %trajectory, so $L_1$ and $L_2$ in \refFig{f:ks1} are very close to some parts of
1110: %the typical run. The same figure shows that the resulting periodic orbits $p_1$
1111: %and $p_2$ are quite similar to those parts. So, indeed both $p_1$ and $p_2$ belongs
1112: %to the family of patterns of which we catch a glimpse now and then in the turbulent
1113: %flow. To keep the topological similarity is also one advantage of our method.}
1114:
1115: Two initial loop guesses are displayed in
1116: \reffig{f:ks1}, along with the two periodic orbits
1117: % $p_1, p_2$
1118: detected by
1119: the \descent. In discretization of the initial loops, each point has to be specified in
1120: all $d$~dimensions;
1121: here the coordinates $\{a_1,a_2\}$ are picked so that topological
1122: similarity between initial and final loops is visually easy to identify.
1123: Other projections from $d=32$ dimensions to subsets of 2 coordinates
1124: appear to make the identification harder, if not impossible.
1125: In both calculations, we molded segments of typical trajectories into
1126: smooth closed loops by the Fourier filtering method of
1127: \refSect{sec:nm}. As the desired orbit
1128: becomes longer and more complex, more sampling points are needed to
1129: represent the loop. We use $N=512$ points to represent the loop in the (a)-(b) case and
1130: $N=1024$ points in the (c)-(d) case. The space-time evolution of $u(x,t)$
1131: for these two unstable spatio-temporally periodic solutions is displayed
1132: in \reffig{f:tevol}.
1133: % \YL{These two solutions are topologically quite different, so we
1134: % should not expect similarity of their spatial profiles. A typical trajectory on
1135: % the chaotic set will follow them in different time intervals. We can see this from
1136: % the resemblance between $L$ and $p$ in \refFig{f:ks1}.}
1137: As $u(x,t)$ is antisymmetric on $[-\pi,\pi]$, it suffices to
1138: display the solutions on the $x \in [0,\pi]$ interval.
1139: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1140: \begin{figure}[t!]
1141: \centering
1142: (a)~\includegraphics[width=2.2in]{kscc1.eps}
1143: \hspace{0.2in}
1144: (b)~\includegraphics[width=2.2in]{kscc2.eps}
1145: \caption{
1146: Level plot of the space-time evolution $u(x,t)$ for the
1147: two spatiotemporally periodic solutions of \reffig{f:ks1}:
1148: (a) the evolution of $p_1$, with start of a repeat after the cycle period
1149: $\period{1}=0.744892$, and
1150: (b) one full period $\period{2}=1.184668$ in
1151: the evolution of $p_2$.
1152: }
1153: \label{f:tevol}
1154: \end{figure}
1155: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1156:
1157:
1158:
1159: \section{Discussion}
1160: \label{sec:sum}
1161:
1162: In order to cope with the difficulty of finding periodic orbits in
1163: high-dimensional chaotic flows, we have devised the {\em \descent\ method},
1164: an infinitesimal step variant of
1165: the damped Newton-Raphson method.
1166: Our main result is the PDE \refeq{bd3} which solves the variational problem of
1167: minimizing the {\costFct}al
1168: %$\costF[\lSpace]$
1169: \refeq{funal}.
1170: This equation describes the fictitious
1171: time $\tau$ flow in the space of loops which decreases the {\costFct}al
1172: at uniform exponential rate (see \refeq{cc}).
1173: % functional which ensures the global convergence.
1174: Variants of the method are presented for special classes of systems,
1175: such as Hamiltonian systems. An efficient integration scheme for the PDE
1176: is devised and tested on the Kuramoto-Sivashinsky system,
1177: the H\'{e}non-Heiles system and the restricted three-body
1178: problem.
1179:
1180: Our method uses information from a large number of points in phase
1181: space, with the global topology of the desired cycle protected by insistence on
1182: smoothness and a uniform discretization of the loop.
1183: The method is quite robust in practice.
1184:
1185: The numerical results presented here are only a proof of principle. We do not
1186: know to what periodic orbit the flow \refeq{bd3} will evolve for a
1187: given dynamical system and a given initial loop.
1188: Empirically, the flow goes toward the ``nearest'' periodic orbit,
1189: with the largest topological resemblance. Each particular application will still require
1190: much work in order
1191: to elucidate and enumerate relevant topological structures. The hope is that the short
1192: spatio-temporally periodic solutions revealed by the {\descent} searches will serve as
1193: the basic building blocks for systematic investigations of chaotic and perhaps even
1194: ``turbulent'' dynamics.
1195:
1196: \begin{acknowledgments}
1197:
1198: We would like to thank Cristel Chandre for careful reading of the
1199: manuscript and numerous suggestions. %and corrections.
1200:
1201: \end{acknowledgments}
1202:
1203: \bibliography{../bibtex/nonlind}
1204:
1205: \end{document}
1206: