nlin0510025/mech.tex
1: \documentclass[12pt, epsf]{article}
2: %\input eps.tex
3: 
4: % -\documentclass[12pt]{article}
5: \usepackage{latexsym}
6: \usepackage{amsmath}
7: \usepackage{amssymb}
8: \usepackage{amsfonts}
9: %\usepackage{times}
10: \usepackage{graphicx}
11: \usepackage{epsfig}
12: \usepackage{array}
13: \usepackage{flafter}
14: %\usepackage[all]{xy}
15: 
16: %\addtolength{\textwidth}{.9in}
17: %\addtolength{\voffset}{-0.6in}
18: %\addtolength{\textheight}{0.9in}
19: 
20: %-------------------------------------------------
21: % ------------ Some LaTeX definitions ------------
22: %-------------------------------------------------
23: 
24: % -- CHANGE SPACING --
25: 
26: % this line condenses the spacing between 
27: % consecutive lines of text
28: % \renewcommand{\baselinestretch}{0.9}
29: 
30: % change the spacing below figures and tables and text
31: % \setlength{\floatsep}{6pt}
32: % \setlength{\textfloatsep}{6pt}
33: % \setlength{\intextsep}{6pt}
34: 
35: % insert an 8pt space before a paragraph
36: \setlength{\parskip}{8pt}
37: 
38: % -- CUSTOMIZATIONS --
39: 
40: 
41: % Change equation environment
42: \newcommand{\Eq}[2]{\begin{equation} #1 \label{eq#2} \end{equation}}
43: \newcommand{\Eqn}[1]{\[ #1 \]}
44: 
45: % Refer to examples
46: \newcommand{\Ex}[1]{Example~\ref{ex#1}}
47: \newcommand{\ex}[1]{example~\ref{ex#1}}
48: 
49: % Refer to equations 
50: \newcommand{\eq}[1]{equation~\ref{eq#1}}
51: \newcommand{\refeqn}[1]{equation~(\ref{eq:#1})}
52: \newcommand{\refeqns}[2]{equations~(\ref{eq:#1})-(\ref{eq:#2})}
53: \newcommand{\Refeqn}[1]{Eqn.~(\ref{eq:#1})}
54: 
55: % Refer to figures
56: \newcommand{\reffig}[1]{Fig.~\ref{fig:#1}}
57: \newcommand{\Reffig}[1]{Figure~\ref{fig:#1}}
58: \newcommand{\reffigs}[2]{Figs.~\ref{fig:#1} to \ref{fig:#2}}
59: 
60: %Refer to table
61: \newcommand{\reftbl}[1]{Table~(\ref{tbl:#1})}
62: 
63: % ------------------ Defining the \Example environment ---------------------
64: %Syntax:
65: %  \Example{title}{body}{label}
66: %  \Ex{label} ...  will produce Example xxx ..
67: %  ... \ex{label} ..... will produce: ... example xxx ....
68: %  the numbering of the examples (xxx) is done according to the section
69: %  number.
70: 
71: \newtheorem{ExampleDef}{Example}[section]
72: 
73: \newcommand{\Example}[3]{
74:   \begin{list}{}{
75:       \setlength{\leftmargin}{1em}}     % Indent everything by this amount
76:     \item                               % Group everything in one item
77:     \small                              % Use a smaller font size
78:     \begin{ExampleDef} \rm              % Theorems are italic - select roman
79:       {\bf \hspace{-1ex}: #1}           % The name, use \\[1ex] to break line
80:       #2                                % The actual stuff
81:       \hfill {\large \boldmath $\Box$}  % The box
82:       \label{ex#3}                      % Label the example
83:     \end{ExampleDef}
84:   \end{list}}
85: 
86: %-----------------------------------------------------------
87: 
88: 
89: \begin{document}
90: \rightline{\small NSL-050901. September, 2005}
91: \vspace{1.em}
92: \begin{center}
93: {\Large \bf Higher-Order Nonlinear Contraction Analysis \par}
94: \vspace{1.5em}
95: {\large Winfried Lohmiller and Jean-Jacques E. Slotine \par}
96: {Nonlinear Systems Laboratory \\
97: Massachusetts Institute of Technology \\
98: Cambridge, Massachusetts, 02139, USA\\
99: {\sl wslohmil@mit.edu, jjs@mit.edu} \par}
100: \vspace{3em}
101: \end{center}
102: 
103: \begin{abstract}
104: 
105: Nonlinear contraction theory is a comparatively recent dynamic control
106: system design tool based on an exact differential analysis of
107: convergence, in essence converting a nonlinear stability problem into
108: a linear time-varying stability problem. Contraction analysis relies
109: on finding a suitable {\it metric} to study a generally nonlinear and
110: time-varying system.  This paper shows that the computation of the
111: metric may be largely simplified or indeed avoided altogether by
112: extending the exact differential analysis to the higher-order dynamics
113: of the nonlinear system. Simple applications in economics, classical
114: mechanics, and process control are described.
115: 
116: \end{abstract}
117: 
118: %Discussion of mechanical systems is moved after end{document}
119: 
120: \section{Introduction}
121: 
122: Nonlinear contraction theory is a comparatively recent dynamic control
123: system design tool based on an exact differential analysis of
124: convergence~\cite{Lohm1}. In essence, it allows one to convert a
125: nonlinear stability problem into a linear time-varying stability
126: problem.  Contraction analysis relies on finding a suitable {\it
127: metric} to study a generally nonlinear and time-varying system.
128: Depending on the application, the metric may be trivial (identity or
129: rescaling of states), or obtained from physics, combination of
130: contracting subsystems~\cite{Lohm1}, semi-definite
131: programming~\cite{Lohm4}, or recently sums-of-squares
132: programming~\cite{Parrilo}. 
133: 
134: The goal of this paper is to show that the computation of the metric
135: may be largely simplified or avoided altogether by extending the exact
136: differential analysis to the higher-order dynamics of the nonlinear
137: system.  Intuitively this is not surprising, since, as an elementary
138: instance, a scalar linear time-invariant system would require in the
139: original approach a non-identity metric (obtained from a Lyapunov
140: Matrix Equation).
141: 
142: After a brief review of contraction theory in Section 2, the main
143: results are presented in Section 3, first in the discrete-time case
144: (with a simple application to price dynamics in economics) and then in
145: the continuous-time case. Simple examples and applications are
146: discussed in Section 4, in the contexts of classical mechanics,
147: process control, and observer design (see
148: \cite{Rouchon,Nguyen,Jouff,inertial} for other recent applications of
149: contraction theory to observer design). Hamiltonian systems are
150: studied in section 5. Concluding remarks are offered in section 6.
151: 
152: \section{Contraction theory}
153: 
154: The basic theorem of contraction analysis~\cite{Lohm1} can be
155: stated as
156: 
157: \newtheorem{theorem}{Theorem}
158: \begin{theorem} Consider the deterministic system $ \ \dot{\bf x} =
159: {\bf f}({\bf x},t) \ $, where ${\bf f}$ is a smooth nonlinear
160: function. If there exist a uniformly positive definite metric
161: 
162: ${\bf M(\bf x}, t) \ = \ {\bf \Theta}({\bf x}, t)^T \ {\bf
163: \Theta}^{\ast}({\bf x}, t)$
164: 
165: \noindent such that the Hermitian part of the associated generalized
166: Jacobian
167: ${\bf F} \ = \ \left(\dot{\bf \Theta} + {\bf \Theta}
168: \frac{\partial {\bf
169:   f}} {\partial {\bf \bf x}} \right){\bf \Theta}^{-\ast}$
170: 
171: \noindent is uniformly negative definite, then all system
172: trajectories converge exponentially to a single trajectory,
173: with convergence rate $| \lambda_{max} |$, where $\lambda_{max}$
174: is the largest eigenvalue of the Hermitian part of $\ {\bf F}$. The
175: system is said to be contracting. \label{th:theoremF}
176: \end{theorem}
177: 
178: In the above, $\ ^{\ast}$ denotes complex conjugation, $\ ^{-\ast}$
179: for a matrix denotes the inverse of the conjugate matrix, and the
180: state-space is $R^n$ (in this paper) or $C^n$. The system is said to
181: be {\it semi-contracting} (for the metric ${\bf M(\bf x}, t)$) if
182: ${\bf F}$ is always negative semi-definite, and {\it indifferent} if
183: ${\bf F}$ is always zero.
184: 
185: It can be shown conversely that the existence of a uniformly positive
186: definite metric with respect to which the system is contracting is
187: also a necessary condition for global exponential convergence of
188: trajectories.  In the linear time-invariant case, a system is globally
189: contracting if and only if it is strictly stable, with ${\bf F}$
190: simply being a normal Jordan form of the system and ${\bf \Theta}$ the
191: coordinate transformation to that form. Conceptually, approaches
192: closely related to contraction, although not based on differential
193: analysis, can be traced back to \cite{Hart} and even to \cite{Lew}.
194: 
195: Similarly, a discrete system 
196: $$ {\bf x}_{i+1} = {\bf f}_i({\bf x}_i, i)
197: $$ 
198: will be contracting in a metric ${\bf \Theta}^T{\bf \Theta}^{\ast}$ if the
199: largest singular value of the discrete Jacobian ${\bf \Theta}_{i+1}
200: \frac{\partial {\bf f}_i}{\partial {\bf x}_i} {\bf \Theta}_i^{-\ast}$ is
201: strictly smaller than 1.  In the particular case of real autonomous
202: systems with identity metric, the basic contraction theorem
203: corresponds in the continuous-time case to Krasovkii's theorem~\cite{Sl91}, and
204: in the discrete-time case to the contraction mapping theorem~\cite{Bert}.
205: 
206: Contraction theory proofs make extensive use of {\it virtual
207: displacements}, which are differential displacements at fixed time
208: borrowed from mathematical physics and optimization theory. Formally,
209: if we view the position of the system at time $t$ as a smooth function
210: of the initial condition ${\bf x}_o$ and of time, $\ {\bf x} = {\bf
211: x}({\bf x}_o ,t)\ $, then $\ \delta {\bf x} = \frac{\partial {\bf
212: x}}{\partial {\bf x}_o} \ d {{\bf x}_o}\ $. For instance~\cite{Lohm1},
213: for the system of Theorem 1, one easily computes
214: \begin{equation}\label{differential}
215: \frac{d}{dt} ({\bf \Theta}  \delta {\bf x}) \ = \ {\bf F} ({\bf \Theta}  \delta {\bf x})
216: \end{equation}
217: 
218: An appropriate metric to show that the system is contracting may be
219: obtained from physics, combination of contracting
220: subsystems~\cite{Lohm1}, semi-definite programming~\cite{Lohm4}, or
221: sums-of-squares programming~\cite{Parrilo}. The goal of this paper is
222: to show that the computation of the metric may be largely simplified
223: or avoided altogether by considering the system's {\it higher-order}
224: virtual dynamics (rather than merely its first-order virtual dynamics,
225: as in equation (\ref{differential})).
226: 
227: \section{Higher-order contraction}
228: 
229: \subsection{The discrete-time case}
230: 
231: Technically, the extension to higher-order contraction is simplest in
232: the discrete-time case, which we discuss first. The main idea is to
233: construct an exponential bound on the virtual displacement $\delta
234: {\bf x}$ over $n$ time-steps, rather than over a single time-step as
235: in \cite{Lohm1}.
236: 
237: Consider for $i \ge 0$ the $n$-dimensional ($n \ge 1$) virtual
238: dynamics
239: $$ 
240:  \delta {\bf x}_{i+n} = {\bf A}^{n-1}_i \delta {\bf x}_{i+n-1} +
241:  \ldots + {\bf A}^o_i \delta {\bf x}_i
242: $$ 
243: Taking the norm (denoted by $|\ \ |$\ ) on both sides, and bounding, yields 
244: \begin{equation} 
245: |\delta {\bf x}_{i+n}| \ \le \ \ |A^{n-1}_i| \ |\delta {\bf x}_{i+n-1}|\ + \
246: \ldots \ + \ |A^o_i| \ |\delta {\bf x}_i| \nonumber
247: \end{equation}
248: where the norm of a matrix is the largest singular value of that
249: matrix. Let us bound for $i=0$ the initial conditions using real
250: positive constants $\lambda$ and $K$ as
251: $$
252:   |\delta {\bf x}_{j}| \le K \lambda^{j}, \ 0 \le j < n 
253: $$
254: Assume now that the following characteristic equation is verified,
255: $$  
256: \lambda^n \ge |A^{n-1}_i| \lambda^{n-1} + \ldots + |A^o_i|, \ \forall
257: i \ge 0 \nonumber
258: $$
259: We then get
260: $$ 
261: |\delta {\bf x}_{i+n}| \le |A^{n-1}_i| K \lambda^{i+n-1} + \ldots +
262: |A^o_i| K \lambda^i \le K \lambda^{i+n}
263: $$ 
264: \begin{figure}
265: \begin{picture}(200,200)(0,-100)
266: \put(-2,0){\vector(1,0){210}}
267: \put(3,-100){\vector(0,1){200}}
268: \put(215,0){$i$}
269: \put(0,105){$\delta x_i$}
270: \put(-10,0){$0$}
271: \put(25,-2){$:$}
272: \put(50,-2){$:$}
273: \put(75,-2){$:$}
274: \put(100,-2){$:$}
275: \put(125,-2){$:$}
276: \put(150,-2){$:$}
277: \put(175,-2){$:$}
278: \put(0,30){$\bullet$}
279: \put(25,-30){$\bullet$}
280: \put(50,40){$\bullet$}
281: \put(75,-35){$\bullet$}
282: \put(100,-15){$\bullet$}
283: \put(125,-28){$\bullet$}
284: \put(150,17){$\bullet$}
285: \put(175,-15){$\bullet$}
286: \qbezier(0,90)(100,27)(200,13)
287: \qbezier(0,-90)(100,-27)(200,-13)
288: \end{picture}
289: \caption{$\pm K \lambda^i$ defined by $\delta x_i$ over $i$}
290: \label{fig:discretecontraction}
291: \end{figure}
292: Repeating the above recursively for $ i \ge 0$ we get by complete
293: induction
294: $$ 
295: |\delta {\bf x}_i| \le K \lambda^i
296: $$
297: and hence exponential convergence of $|\delta {\bf x}_i|$, as
298: illustrated in Figure~\ref{fig:discretecontraction}.
299: 
300: \begin{theorem}
301: Consider for $i \ge 0$ the $n$-dimensional ($n \ge 1$) virtual dynamics  
302: $$ 
303:  \delta {\bf x}_{i+n}     = {\bf A}^{n-1}_i \delta {\bf x}_{i+n-1} +
304:  \ldots + {\bf A}^o_i \delta {\bf x}_i
305: $$ 
306: Let us define $\forall i \ge 0$ a constant $\lambda$ with the
307: characteristic equation
308: \begin{eqnarray} 
309: \lambda^n &\ge& |A^{n-1}_i| \lambda^{n-1} + \ldots + |A^o_i|, \
310: \forall i \ge 0 \ \label{eq:discretecharacteristic}
311: \end{eqnarray}
312: We can then conclude
313: $$
314: |\delta {\bf x}_{i+n}| \le K \lambda^{i+n}
315: $$  
316: where $K$ is defined by
317: \begin{equation}
318:  |\delta {\bf x}_{j}| \le K \lambda^{j}, \ 0 \le j < n \
319:   \label{eq:discretedefinitionofK} 
320: \end{equation}
321: Thus, the system is contracting if $\ \lambda < 1$.
322: \label{th:higherorderdiscrete}
323: \end{theorem}
324: \Example{}{Consider first a second-order linear time invariant (LTI)
325:   dynamics
326: $$
327: x_{i+2} \ + \ 2\ \gamma \ x_{i+1} \ + \ \alpha \gamma^2 x_i \ = \ u_i
328: $$ where $u_i$ is an input, and $\gamma$ and $\alpha$ are
329: constants. The virtual dynamics is
330: $$
331: \delta x_{i+2} = - \ 2\ \gamma \ \delta x_{i+1}\  - \ \alpha \gamma^2
332: \ \delta x_i  
333: $$
334: The characteristic equation (\ref{eq:discretecharacteristic}) for
335: $\lambda \ge 0$ is then given by
336: \begin{equation}
337: \lambda^2 \ \ge \ 2\ |\gamma| \ \lambda + |\alpha| \gamma^2 \ \ \ \ \
338: \ \ \ \ i.e. \ \ \ \ \ \ \ \ \ \lambda \ \ge \ |\gamma| (1 + \sqrt{
339:   1 + |\alpha |\ }) \nonumber
340: \end{equation}
341: Thus, the contraction condition $\lambda < 1$, or
342: $$
343: |\gamma| ( 1 + \sqrt{ 1 + |\alpha | \ }) \ < \ 1
344: $$
345: simply means that both eigenvalues of the system have to lie for the
346: conjugate complex case ($\alpha > 1$) within the red half circles in
347: (\ref{fig:discretecirecle}) or on the green line for the real case 
348: ($\alpha \le 1$). Note that Theorem \ref{th:higherorderdiscrete}
349: simply bounds the possibly oscillating discrete system with a
350: non-oscillating system of the same convergence rate for the real case.
351: 
352: Consider now the virtual dynamics of an arbitrary second-order
353: nonlinear time-varying system,
354: $$
355: \delta x_{i+2} + \ 2\ \gamma(i)\ \delta x_{i+1} + \alpha(i) \
356: \gamma^2(i) \delta x_i \ = \ 0
357: $$
358: The characteristic equation and the contraction condition are the same as
359: above, except that $\ \gamma\ $ and $\ \alpha\ $ are now
360: time-dependent.
361: \begin{figure}[h]
362: \begin{center}
363: \epsfig{figure=discretecircle.eps,height=90mm,width=90mm}
364: \end{center}
365: \caption{Contraction region in the complex plane of second-order
366:   discrete system} \label{fig:discretecirecle}
367: \end{figure}
368: }{LTVdiscrete}
369: %% Nonlinear example. Perhaps observer for time series prediction
370: %% using the logistic map
371: %% \ \ \ \ \  $x_{i+1} = a_i \ x_i \ (1 - x_i)$
372: \Example{}{In economics, consider the price dynamics
373: \begin{eqnarray}
374: {\bf n}_{i+1} &=& {\bf f}_i({\bf p}_i, i) \nonumber \\
375: {\bf p}_{i+1} &=& {\bf g}_i({\bf n}_i, i) \nonumber
376: \end{eqnarray}
377: with ${\bf n}_i$ the number of sold products at time $i$ and
378: corresponding price ${\bf p}_i$. 
379: 
380: The first line above defines the customer demand as a reaction to a
381: given price. The second line defines the price, given by the
382: production cost under competition, as a reaction to the number of sold
383: items.  The dynamics above corresponds to the second-order economic
384: growth cycle dynamics
385: $$
386: {\bf n}_{i+2} = {\bf f}_{i+1} \left( {\bf g}_i ({\bf n}_i, i) \right)
387: $$
388: 
389: Contraction  behavior of this economic behavior with contraction rate
390: $\lambda$ can then be concluded with Theorem
391: \ref{th:higherorderdiscrete} for
392: \begin{equation}
393: \lambda^2 \ge | \frac{\partial {\bf f}_{i+1}}{\partial {\bf p}_{i+1}}
394: \frac{\partial {\bf g}_i}{\partial {\bf n}_i} | \label{eq:game}
395: \end{equation}
396: That means we get stable (contraction) behavior if the product of
397: customer demand sensitivity to price and production cost sensitivity
398: to number of sold items has singular values less than 1. We can get
399: unstable (diverging) behavior for the opposite case. 
400: 
401: Note that this result even holds when no precise model of the
402: sensitivity is known, which is usually the case in economic or game
403: situations. 
404: 
405: Whereas the above is well known for LTI economic models we can see
406: that the economic behavior is unchanged for a non-linear, time-varying
407: economic environment.
408: 
409: Let as assume now that the above corresponds to a game situation (see
410: e.g. \cite{Shamma} or \cite{Bryson}) between two players with
411: strategic action ${\bf p}_i$ and ${\bf n}_i$. Both players optimize
412: their reaction ${\bf g}$ and ${\bf f}$ with respect to the opponent's
413: action.
414: 
415: We can then again conclude for (\ref{eq:game}) to global contraction
416: behavior to a unique time-dependent trajectory (in the
417: autonomous case, the Nash equilibrium).}{economics}
418: 
419: Of course, and throughout this paper, in some cases the analysis may
420: yet be further streamlined by {\it first} applying a simplifying metric
421: transformation of the form $\ \delta {\bf z} = {\bf \Theta} \delta
422: {\bf x}\ $, and then applying the results to $\ \delta {\bf z}$.
423: 
424: \subsection{The continuous-time case}
425: 
426: Let us now derive the continuous-time version of the previous
427: results. Consider for $t\ge0$ the $n$-dimensional ($n \ge 1$) virtual dynamics  
428: $$ 
429:  \delta {\bf x}^{(n)} = - {\bf A}_{n-1} \delta {\bf x}^{(n-1)} - \
430:  ... \ - {\bf A}_o \delta {\bf x}
431: $$ 
432: The proof is based on splitting up the dynamics into a stable part,
433: described by a block diagonal matrix composed of identical negative
434: definite blocks ${\bf F}$ which we select, and an unstable
435: higher-order part. Let $\delta {\bf x}_o = \delta {\bf x}$, and define
436: recursively
437: \begin{eqnarray}
438: \delta \dot{\bf x}_o     &=& {\bf F} \delta {\bf x}_o + \delta {\bf
439:   x}_1 \nonumber \\
440:                         &...& \nonumber \\
441: \delta \dot{\bf x}_{n-2} &=& {\bf F} \delta {\bf x}_{n-2} +
442: \delta {\bf x}_{n-1} \nonumber \\ 
443: \delta \dot{\bf x}_{n-1} &=& - {\bf A}_{n-1} \delta {\bf x}_o^{(n-1)} -
444: {\bf A}_{n-2} \delta {\bf x}_o^{(n-2)} - \ ... \ -  {\bf A}_o \delta
445: {\bf x}_o \nonumber \\ 
446: &&- \left( {\bf F} \delta{\bf x}_o \right)^{(n-1)} - ... - \left( {\bf
447:     F} \delta {\bf x}_{n-2} \right)^{(1)} \nonumber \\
448: &=& -{\bf A}_{n-1} \left( {\bf F} \delta {\bf x}_o + \delta {\bf
449:   x}_1 \right)^{(n-2)} - ... - {\bf A}_o \delta {\bf x}_o \nonumber \\
450: &&- \left( L^1 {\bf F} \delta {\bf x}_o + L^o {\bf F} \delta {\bf x}_1
451: \right)^{(n-2)} - ... - \left( L^1 {\bf F} \delta {\bf x}_{n-2} + L^o
452:   {\bf F} \delta {\bf x}_{n-1} \right) \nonumber \\
453: &=& {\bf F} \delta {\bf x}_{n-1} 
454: - {\bf A}_{n-1}^{\ast} \delta {\bf x}_{n-1} 
455: - {\bf A}_{n-2}^{\ast} \delta {\bf x}_{n-2}
456: - {\bf A}_{n-3}^{\ast} \delta {\bf x}_{n-3} -
457: ... \label{eq:superimposed}
458: \end{eqnarray}
459: where
460: \begin{eqnarray}
461: L^o {\bf F} &=& {\bf F} \nonumber \\
462: L^j {\bf F} &=& \frac{d}{dt} L^{j-1} {\bf F} + L^{j-1} {\bf F} \
463: {\bf F} \ \ \ \ \ \ \ \ \ j \ge 1 \nonumber
464: \end{eqnarray}
465: and
466: \begin{eqnarray}
467: {\bf A}_{n-1}^{\ast} &=& {\bf A}_{n-1} + \left( \begin{array}{c} n \\
468:     1 \end{array} \right) L^o {\bf F} \label{eq:Aast} \\
469: {\bf A}_{n-2}^{\ast} &=& {\bf A}_{n-2} + \left( \begin{array}{c} n-1
470:     \\ 1 \end{array} \right) {\bf A}_{n-1} L^o {\bf F} + \left(
471:   \begin{array}{c} n \\ 2 \end{array} \right) L^1 {\bf F} \nonumber \\
472: {\bf A}_{n-3}^{\ast} &=& {\bf A}_{n-3} + \left( \begin{array}{c} n-2
473:     \\ 1 \end{array} \right) {\bf A}^{n-2} L^o {\bf F} + \left(
474:   \begin{array}{c} n-1 \\ 2 \end{array} \right) {\bf A}^{n-1} L^1 {\bf
475:   F} + \left( \begin{array}{c} n \\ 3 \end{array} \right) L^2 {\bf F}
476: \nonumber \\ &...& \nonumber
477: \end{eqnarray}
478: Equation (\ref{eq:superimposed}) represents the superposition of a
479: higher-order-system and a block diagonal dynamics in the chosen ${\bf
480: F}$. Let us assess the contraction behavior of the higher-order part
481: by taking the norm
482: $$
483: |\delta {\bf x}^{(n)}| \le |{\bf A}_{n-1}^{\ast}| |\delta {\bf
484: x}^{(n-1)}| + |{\bf A}_{n-2}^{\ast}| |\delta {\bf x}^{(n-2)}| + \
485: \ldots \ 
486: $$
487: where the norm of a matrix is the largest singular value
488: of that matrix. Let us bound for $t=0$ the initial
489: conditions with real and constant $\lambda, K \ge 0$ and assume the
490: following characteristic equation
491: \begin{eqnarray} 
492: |\delta {\bf x}^{(j)}| &\le& K \lambda^j e^{\lambda t}, \ \ \ \ \ \ \ \
493: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 0 \le j < n \ \label{eq:definitionofK}
494: \\ \lambda^n &\ge& |{\bf A}_{n-1}^{\ast}| \lambda^{n-1} + \ \ldots \ +
495: |{\bf A}_o^{\ast}|, \ \forall t \ge 0 \label{eq:characteristic}
496: \end{eqnarray}
497: Figure \ref{fig:discretecontraction} shows how $K$ has to be selected
498: for a given $\lambda$ for a second-order system ($n=2$).
499: \begin{figure}
500: \begin{picture}(200,200)(0,-100)
501: \put(0,0){\vector(1,0){200}}
502: \put(103,-100){\vector(0,1){200}}
503: \put(205,0){t}
504: \put(200,20){$\delta x(t)$}
505: \put(200,100){$K e^{\lambda t}$}
506: \qbezier(0,-50)(100,107)(200,30)
507: \qbezier(0,55)(100,72)(200,115)
508: \end{picture}
509: \caption{$K e^{\lambda t}$ bounding the first and second derivative of
510:   $\delta x$ for a given $\lambda$ at $t=0$}
511: \label{fig:discretecontraction}
512: \end{figure}
513: With (\ref{eq:characteristic}) we can bound the $n$'th derivative of
514: $\delta {\bf x}$ as
515: $$
516: |\delta {\bf x}^{(n)}| \le |{\bf A}_{n-1}^{\ast}| K \lambda^{n-1}
517: e^{\lambda t} + \ \ldots \  + |{\bf A}_o^{\ast}| K e^{\lambda t} \le K
518: \lambda^n e^{\lambda t}
519: $$
520: Integrating the above for $t \ge 0$ we can exponentially bound the
521: higher-order dynamics $\delta {\bf x}$ as 
522: $$ 
523: |\delta {\bf x}| \le K e^{\int_o^t \lambda d \tau}
524: $$ 
525: Using the above this allow to conclude:
526: \begin{theorem}
527: Consider for $t\ge0$ the $n$-dimensional ($n \ge 1$) virtual dynamics  
528: $$ 
529:  \delta {\bf x}^{(n)} = - {\bf A}^{n-1} \delta {\bf x}^{(n-1)} - \
530:  \ldots \  - {\bf A}^o \delta {\bf x}
531: $$  
532: Let us define a constant $\lambda \ge 0$ such that $\forall t \ge 0$
533: we fulfill the characteristic equation
534: \begin{equation} 
535:  \lambda^n \ge {\bf A}_{n-1}^{\ast} \lambda^{n-1} + \ \ldots \  + {\bf
536:    A}_o^{\ast} \label{eq:theoremcharacteristic}
537: \end{equation}
538: where ${\bf A}_j^{\ast}$ is defined in (\ref{eq:Aast}) for a given choice 
539: of the matrix ${\bf F}$. 
540: 
541: We can then conclude on contraction rate (i.e., the largest eigenvalue
542: of the symmetric part of) ${\bf F} + \lambda {\bf I}$, where
543: $|\delta {\bf x}|$ is initially bounded with $K$, defined in
544: (\ref{eq:definitionofK}). 
545: \label{th:higherordercontinuous}
546: \end{theorem}
547: One specific choice of ${\bf F}$ is $-\frac{{\bf A}_{n-1}}{n}$, which
548: cancels the highest time-derivative on the right-hand side, and is
549: known for LTV systems as the reduced or unstable form~\cite{Kailath}
550: of the original higher-order dynamics. We will use this definition of
551: ${\bf F}$ in most of the following examples. Also note that more
552: general forms could be chosen for the stable part.
553: 
554: \section{Examples and Applications}
555: 
556: In this section, we discuss simple examples (section 4.1),
557: applications to nonlinear observer design (section 4.2), and adding
558: an indifferent system (section 4.3).
559: 
560: \subsection{Some simple examples}
561: 
562: \Example{}{Consider the second-order LTI dynamics
563: $$
564: \ddot{x} = - 2 \zeta \omega \dot{x} - \omega^2 x
565: $$
566: with constant $\zeta$ and $\omega \ge 0$. The virtual dynamics is
567: $$
568: \delta \ddot{x} = - 2 \zeta \omega \delta \dot{x} - \omega^2 \delta x  
569: $$
570: The characteristic equation (\ref{eq:theoremcharacteristic}) is then
571: given with $F = \zeta \omega$ for constant, positive $\lambda$ by 
572: \begin{eqnarray}
573:  \lambda^2 \ge | - \omega^2 + \frac{(2 \zeta \omega)^2}{4} | =
574: | \zeta^2 - 1 | \omega^2 \ \ \ \ \ \ \ \ \ \rm{i.e.} \ \ \ \ \ \ \ \ \ 
575: \lambda \ge \omega \sqrt{| \zeta^2 - 1 |} \nonumber
576: \end{eqnarray}
577: Using Theorem \ref{th:higherordercontinuous}
578: we can then conclude on contraction behavior with convergence rate 
579: $$
580: \omega (-\zeta  + \sqrt{| \zeta^2 - 1 |})
581: $$
582: This means that we require the poles to lie within the $\pm45^o$ 
583: quadrant of the left-half complex plane.}{LTI}
584: 
585: While Theorem \ref{th:higherordercontinuous} can thus be
586: overly conservative for LTI systems, this is not the case for general
587: nonlinear time-varying systems, as we now illustrate.
588: 
589: \Example{}{Consider the second-order LTV dynamics
590: $$
591: \ddot{x} + a_1 \ \dot{x} + a_o(t)\ \ x = u(t)
592: $$
593: with $a_1, a_o \ge 0$, which would be sufficient conditions for LTI
594: stability. Let us assume a small damping gain $a_1$ and strong spring
595: gains $a_o$ such that the system oscillates.
596: 
597: If now the time-varying gain $a_o(t)$ is chosen to be very large when
598: the system oscillates back to $0$ and small otherwise then the energy
599: is constantly increased, which makes the system unstable (Figure
600: \ref{fig:example_34}). This is precisely what is excluded by Theorem
601: \ref{th:higherordercontinuous}.}{LTV}
602: 
603: \begin{figure}[h]
604: \begin{center}
605: \epsfig{figure=example_33c.eps,height=60mm,width=90mm}
606: \end{center}
607: \caption{\small LTV system with $a_0=900+850\sin t$ and
608: $a_1=0.01$} 
609: \label{fig:example_34}
610: \end{figure}
611: \Example{}{Consider the second-order LTV dynamics
612: $$
613: \ddot{x} + a_1(t)\ \dot{x} + a_o(t)\ x = u(t)
614: $$
615: The virtual dynamics is
616: $$
617: \delta \ddot{x} = - a_1 \delta \dot{x} - a_o \delta x  
618: $$
619: The characteristic equation (\ref{eq:theoremcharacteristic}) is then
620: given with $F = - \frac{a_1}{2}$ by 
621: $$
622:  \lambda^2 \ge | - a_o + \frac{a_1^2}{4} + \frac{\dot{a_1}}{2}|
623: $$ 
624: for constant positive $\lambda$. Using Theorem
625: \ref{th:higherordercontinuous} we can then conclude on contraction
626: behavior with convergence rate $-\frac{a_1}{2} + \lambda\ $ (Figure
627: \ref{fig:example_35}).}{LTV}
628: 
629: \begin{figure}[h]
630: \begin{center}
631: \epsfig{figure=example_34.eps,height=60mm,width=90mm}
632: \end{center}
633: \caption{\small LTV system with $a_0=2+\cos t$ and $a_1=8+\sin
634: t$} 
635: \label{fig:example_35}
636: \end{figure}
637: 
638: \Example{}{Let us illustrate a case where choosing ${\bf F}$ other
639: than $\ - \frac{{\bf A}_{n-1}}{n}\ $ can simplify the result.  Consider
640: the generalized Van der Pol or Lienard dynamics
641: $$
642: \ddot{x} = - a_1(x) \dot{x} - a_o(x, t)
643: $$
644: with $a_1, \frac{\partial a_o}{\partial x} \ge 0$. The virtual
645: dynamics is
646: $$
647: \delta \ddot{x} = - a_1 \delta \dot{x} - \left( \frac{\partial
648: a_o}{\partial x} + \dot{a_1} \right) \delta x  
649: $$
650: The characteristic equation (\ref{eq:theoremcharacteristic}) is then
651: given with $F = - a_1 + \frac{\min(a_1)}{2}$ by 
652: $$
653: \lambda^2 = ( a_1 - \min(a_1) ) \lambda + |-\frac{ \min(a_1)^2}{4} +
654: \frac{a_1 \min( a_1)}{2} - \frac{\partial a_o}{\partial x}|   
655: $$ 
656: and hence $\lambda \ge 0.5 \left( a_1 - \min(a_1) + \sqrt{ ( a_1 -
657: \min(a_1) )^2 + |-\min(a_1)^2 + 2 a_1 \min(a_1) - 4 \frac{\partial
658: a_o}{\partial x}|} \right)$. Thus the contraction behavior of
659: $\delta x$ is then given by
660: $$
661: \lambda + F \le - 0.5 a_1 + 0.5 \sqrt{ ( a_1 - \min(a_1) )^2 +
662: |-( a_1 - \min(a_1) )^2 + a_1^2 - 4 \frac{\partial a_o}{\partial x}|} 
663: $$
664: which is negative if the argument of the square root is less than
665: $a_1^2$, i.e. for $\min(a_1)^2 - 2 a_1 \min(a_1) + 4 \frac{\partial
666: a_o}{\partial x} \le 0$. 
667: 
668: Hence we can conclude on contraction behavior if the poles of the
669: minimal damping $\min(a_1)$ with the actual spring gain
670: $\frac{\partial a_o}{\partial x}$ lie within the $\pm 45^o$ quadrant of
671: the left-half complex plane (Figure
672: \ref{fig:example_37}). Note that this result can be extended to
673: the vector case ${\bf x}$ if the corresponding matrix $A_1({\bf x})$
674: is integrable.}{VanderPol}
675: 
676: \begin{figure}[h]
677: \begin{center}
678: \epsfig{figure=example_37.eps,height=60mm,width=90mm}
679: \end{center}
680: \caption{\small Van der Pol with $a_1=8+\sin x$ and $u(t)=\cos
681: t$} 
682: \label{fig:example_37}
683: \end{figure}
684: 
685: \Example{}{Consider the second-order nonlinear vector system
686: $$
687: \ddot{\bf x} + {\bf D} \ \dot{\bf x} + \frac{\partial V}{\partial {\bf
688:     x}} = u(t)
689: $$
690: with potential energy $V = x_1^2+x_2^2+x_1x_2\sin t$ and constant
691: damping gain ${\bf D} =diag(1,4)$.
692: 
693: The corresponding variational dynamics is
694: $$
695: \delta \ddot{\bf x} + {\bf D} \ \delta \dot{\bf x} + \frac{\partial^2
696:   V}{\partial {\bf x}^2} \ \delta {\bf x} = 0
697: $$
698: The characteristic equation (\ref{eq:theoremcharacteristic}) for ${\bf
699:   F} = - \frac{\bf D}{2}$ is then given by 
700: $$
701:  \lambda^2 \ge | \frac{\partial^2
702:  V}{\partial {\bf x}^2} - \frac{\bf D D}{4}\ |
703: $$ 
704: for constant positive $\lambda$. Using Theorem
705: \ref{th:higherordercontinuous} we can then conclude on contraction
706: behavior with convergence rate $\ \lambda {\bf I}\ -\frac{\bf D}{2}$ in
707: Figure \ref{fig:example_36}.}{second-order-system}
708: 
709: \begin{figure}[h]
710: \begin{center}
711: \epsfig{figure=example_35.eps,height=60mm,width=90mm}
712: \end{center}
713: \caption{\small Second-order system with $ {\bf D}=diag(1,4)$
714: and $V=x_1^2+x_2^2+x_1x_2\sin t$} 
715: \label{fig:example_36}
716: \end{figure}
717: 
718: \Example{}{Let us consider two
719: coupled systems of the same dimensions, with the virtual dynamics
720: $$
721: \frac{d}{dt}
722: \left(
723: \begin{array}{c}
724: \delta {\bf x}_1 \\
725: \delta {\bf x}_2
726: \end{array}
727: \right) \ = \
728: \left(
729: \begin{array}{cc}
730: {\bf F}_{11} & {\bf F}_{12}  \\
731: {\bf F}_{21} & {\bf F}_{22}
732: \end{array}
733: \right)
734: \left(
735: \begin{array}{c}
736: \delta {\bf x}_1 \\
737: \delta {\bf x}_2
738: \end{array}
739: \right)
740: $$
741: Let us transform this dynamics in the following second-order
742: dynamics
743: \begin{eqnarray}
744: \frac{d^2}{dt^2} \ \delta {\bf x}_1 &=& \dot{\bf F}_{11} \delta
745: {\bf x}_1 + {\bf F}_{11} \ \frac{d}{dt}\ \delta {\bf x}_1 +
746: (\dot{\bf F}_{12} + {\bf F}_{12} {\bf F}_{22}) \delta {\bf x}_2 +
747: {\bf F}_{12} {\bf F}_{21} \delta {\bf x}_1 \nonumber \\ &=& \left(
748: {\bf F}_{11} + {\bf F}_{22}^{\ast} \right) \frac{d}{dt} \ \delta
749: {\bf x}_1 + \left( \dot{\bf F}_{11} + {\bf F}_{12} {\bf F}_{21} -
750: {\bf F}_{22}^{\ast} {\bf F}_{11} \right) \delta {\bf x}_1 \nonumber
751: \end{eqnarray}
752: with the generalized Jacobian ${\bf F}_{22}^{\ast} = (\dot{\bf
753: F}_{12} + {\bf F}_{12} {\bf F}_{22}) {\bf F}_{12}^{-1}$. The
754: shrinking rate of this system is now the average of ${\bf F}_{11}$
755: and ${\bf F}_{22}^{\ast}$. Using a direct contraction approach with
756: e.g. ${\bf F}_{12}\ = -\ k \ {\bf F}_{21}^T$ the guaranteed
757: contraction rate would be a more conservative value, namely the
758: largest of the individual contraction rates of ${\bf F}_{11}$ and
759: ${\bf F}_{22}^{\ast}$.}{second-order-state-space}
760: 
761: \subsection{Higher-order observer design}
762: 
763: While a controller for an $n^{\rm th}$ order system simply has to add
764: stabilizing feedback in ${\bf x}^{(n-1)}, \ \ldots \ , {\bf x}, t$
765: according to Theorem \ref{th:higherordercontinuous}, the situation is
766: not such straightforward for observers since here only a part of the
767: state is measured. Motivated by the linear Luenberger observer and the
768: linear reduced-order Luenberger observer, we derive such an observer
769: design for higher-order nonlinear systems.
770: 
771: Consider the $n$-dimensional nonlinear system dynamics
772: $$
773: {\bf x}^{(n)} = {\bf a}_o({\bf x}, \dot{\bf x}, t) + \ \ldots \  + {\bf
774: a}_{n-2}^{(n-2)} ({\bf x}, \dot{\bf x}, t) + {\bf
775: a}_{n-1}^{(n-1)}({\bf x}, t)
776: $$
777: with the measurement ${\bf y}({\bf x}^{(n-1)}, \ \ldots \ , {\bf x},
778: t)$. Note that for a linear Luenberger observer ${\bf y}$ is
779: equivalent to ${\bf x}$ and all ${\bf a}_i$ are linear functions of
780: ${\bf x}$.
781: 
782: Consider now the corresponding nonlinear observer 
783: \begin{eqnarray}
784: \dot{\hat{\bf x}}_{n-1} &=& {\bf a}_o - {\bf e}_{o}(\hat{\bf y}) +
785: {\bf e}_{o}({\bf y}) \nonumber \\ \dot{\hat{\bf x}}_{n-2} &=& \hat{\bf
786: x}_{n-1} + {\bf a}_1 - {\bf e}_{1}(\hat{\bf y}) + {\bf e}_{1}({\bf y})
787: \nonumber \\ &\ \ldots \ & \nonumber \\ \dot{\hat{\bf x}}_o &=&
788: \hat{\bf x}_1 + {\bf a}_{n-1}- {\bf e}_{n-1}(\hat{\bf y}) + {\bf
789: e}_{n-1}({\bf y})\nonumber
790: \end{eqnarray}
791: with ${\bf a}_i(\hat{\bf x}_o, \hat{\bf x}_1 + {\bf a}_{n-1}- {\bf
792: e}_{n-1}(\hat{\bf y}) + {\bf e}_{n-1}({\bf y}), t)$ and ${\bf x}_o =
793: {\bf x}$.  Note that the coordinate transformation in the bracket is a
794: nonlinear generalization of the reduced Luenberger observer.  The
795: above dynamics is equivalent to
796: \begin{eqnarray}
797: \hat{\bf x}^{(n)} &=& {\bf a}_o(\hat{\bf x}, \dot{\hat{\bf x}}, t) + \
798: \ldots \ + {\bf a}_{n-2}^{(n-2)} (\hat{\bf x}, \dot{\hat{\bf x}}, t) +
799: {\bf a}_{n-1}^{(n-1)}(\hat{\bf x}, t) \nonumber \\ &-& {\bf
800: e}_o(\hat{\bf y}) + {\bf e}_o({\bf y}) -\ \ldots \ - {\bf
801: e}_{n-1}^{(n-1)}(\hat{\bf y}) + {\bf e}_{n-1}^{(n-1)}({\bf y})
802: \nonumber
803: \end{eqnarray}
804: whose variational dynamics is
805: $$
806: \delta \hat{\bf x}^{(n)} = \delta {\bf a}_o + \ \ldots \  + \delta
807: {\bf a}_{n-1}^{(n-1)} - \delta {\bf e}_o -\ \ldots \ - \delta {\bf
808:   e}_{n-1}^{(n-1)}
809: $$
810: where the varation is performed on $\hat{\bf x}^{n-1}, \ \ldots \ ,
811: \hat{\bf x}$. Hence the feedback is not only performed in $\hat{\bf
812: y}$, but implicitly also up to the $(n-1)^{\rm th}$ time-derivative of
813: $\hat{\bf y}$.
814: 
815: \Example{}{Consider the second-order nonlinear system
816: $$
817: \ddot{x} + \frac{\partial a_1}{\partial x} (x)\ \dot{x} + a_o(x) = 0
818: $$
819: where $x$ is measured. Consider now the corresponding nonlinear 
820: observer 
821: \begin{eqnarray}
822: \dot{\hat{x}}_1 &=& a_o(\hat{x})       - e_o (\hat{x} - x) 
823: \nonumber \\ 
824: \dot{\hat{x}}_o &=& \hat{x}_1 + a_1(x) - e_1 (\hat{x} - x) 
825: \nonumber
826: \end{eqnarray}
827: with constant $e_o$ and $e_1$ and where we have replaced with the
828: feedback $a_1$ as a function of $x$. The corresponding second-order 
829: variational dynamics is
830: $$
831: \delta \ddot{\hat{x}} + e_1 \ \delta \dot{\hat{x}} + (e_o -
832: \frac{\partial a_o}{\partial \hat{x}})\ \delta \hat{x} = 0
833: $$
834: Contraction behavior can then be shown with Theorem
835: \ref{th:higherordercontinuous}.}{second-order-system-observer} 
836: 
837: \Example{}{Consider the temperature-dependent reaction
838: $A \rightarrow B$ in a closed tank 
839: $$
840: \frac{d}{dt}
841: \left( 
842: \begin{array}{c}
843: c_A \\
844: T 
845: \end{array}
846: \right) =
847: \left( 
848: \begin{array}{c}
849: -1 \\
850: -10
851: \end{array}
852: \right)
853: e^{-\frac{E}{T}} c_A
854: $$
855: with $c_A$ the concentration of A, $T$ the measured temperature, and
856: $E$ the specific activation energy. This reaction dynamics is
857: equivalent to the following second-order dynamics in temperature
858: $$
859: \ddot{T} + \frac{-E}{T^2} \dot{T}^2 = -e^{-\frac{E}{T}} \dot{T} 
860: $$
861: Letting $\tau = \int_o^T e^{\frac{-E}{T}} dT$ yields
862: $$
863: \ddot{\tau} = -e^{-\frac{E}{T}} \dot{\tau} 
864: $$
865: Contraction can then be shown as in \Ex{VanderPol}.
866: The observer dynamics
867: \begin{eqnarray}
868: \dot{\hat{T}}_1 &=& -e^{-\frac{E}{\hat{T}_o}} \dot{\hat{T}}_o +
869: \frac{E}{\hat{T}_o^2} \dot{\hat{T}}_o^2 - e_o e^{\frac{E}{\hat{T}_o}} 
870: \int^{\hat{T}}_T e^{\frac{-E}{T}} dT \nonumber \\
871: \dot{\hat{T}}_o &=& \hat{T}_1 - e_1 (\hat{T}_o - T_o) \nonumber
872: \end{eqnarray}
873: with $\hat{T}_o = \hat{T}$ and constant $e_o$ and $e_1$ leads to
874: $$
875: \ddot{\hat{\tau}} = (-e^{-\frac{E}{T}} - e_1) \dot{\hat{\tau}} - e_o \
876: \hat{\tau} + e_1 \dot{\tau}(t) + e_o \tau(t) 
877: $$ 
878: whose contraction behavior can now be tuned as in
879: \Ex{VanderPol}.}{chemical}
880: \Example{}{
881: Consider the system
882: $$
883: \dddot{x}\ = \ \sin (\dot{x})\ +\ 0.1 \sin t\ - \ 0.015
884: $$ 
885: with measurement $y=x$. Letting $e_0(y)=y$, $e_1(y)=4y$, and
886: $e_2(y)=3y$, the variational equation is 
887: $$
888: \delta \dddot{\hat{x}}+3\ \delta \ddot{\hat x}+(4-\cos(\dot{\hat x}))\
889: \delta \dot{\hat x}+\delta \hat{x}\ =\ 0
890: $$
891: The corresponding observer is illustrated in Figure
892: \ref{fig:example_4}.}{}
893: 
894: \begin{figure}[h]
895: \begin{center}
896: \epsfig{figure=example_4.eps,height=60mm,width=90mm}
897: \end{center}
898: \caption{\small $\dddot{x}\ = \ \sin (\dot{x})+0.1\sin
899: t-0.015\ $, with $\ y=x$.}\label{fig:example_4}
900: \end{figure}
901: 
902: \subsection{Adding an indifferent system}
903: 
904: The analysis may be further simplified by using superposition to
905: compare the system to one whose contraction behavior is known
906: analytically. We illustrate this idea on second-order systems using
907: an indifferent added dynamics.  In principle, the approach can be
908: extended to higher-order systems as well as other types of added
909: dynamics.
910: 
911: Consider the indifferent system \cite{Lohm1}
912: $$
913: \delta \dot{\bf x} = i \Omega \delta {\bf x}
914: $$
915: with real and invertible $\Omega(\dot {\bf x}, {\bf x}, t)$. The above
916: corresponds to the second-order dynamics
917: $$
918: \delta \ddot{\bf x} = \dot{\Omega} \Omega^{-1} \delta \dot{\bf x} -
919: \Omega \Omega \delta {\bf x}
920: $$ 
921: which is thus itself indifferent.  We can write the reduced form of
922: Theorem \ref{th:higherordercontinuous} as
923: \begin{eqnarray} 
924: \frac{d}{dt}
925: \left(
926: \begin{array}{c}
927:   \delta \dot{\bf x} \\
928:   \delta     {\bf x}
929: \end{array} 
930: \right) &=&
931: \left(
932: \begin{array}{cc}
933:   - \Theta({\bf x}, \dot{\bf x}, t) & - \Omega \Omega(\dot {\bf x}, {\bf
934:     x}, t)  \\
935:   {\bf I} & {\bf 0}
936: \end{array} 
937: \right)
938: \left(
939: \begin{array}{c}
940:   \delta \dot{\bf x} \\
941:   \delta     {\bf x}
942: \end{array} 
943: \right) \nonumber \\ &=&
944: \left( \left(
945: \begin{array}{cc}
946:   \dot{\Omega} \Omega^{-1} & - \Omega \Omega \\
947:   {\bf I} & {\bf 0}
948: \end{array} 
949: \right)
950: +
951: \left(
952: \begin{array}{cc}
953:   - \dot{\Omega} \Omega^{-1} - \Theta & \ {\bf 0} \\
954:   {\bf 0} & \ {\bf 0}
955: \end{array} 
956: \right)
957: \right)
958: \left(
959: \begin{array}{c}
960:   \delta \dot{\bf x} \\
961:   \delta     {\bf x}
962: \end{array} 
963: \right) \nonumber
964: \end{eqnarray}
965: which thus corresponds to the superposition of an indifferent system with a
966: semi-contracting system of rate $-\Theta - \dot{\Omega} \Omega^{-1}$.
967: 
968: \begin{theorem}
969: The reduced form
970: \begin{equation} 
971: \delta \ddot{\bf x} + \Theta({\bf x}, \dot{\bf x}, t) \delta
972: \dot{\bf x} + \Omega \Omega(\dot {\bf x}, {\bf x}, t)\ \delta{\bf x} =
973: 0 \nonumber
974: \end{equation}
975: is semi-contracting with rate $-\Theta - \dot{\Omega}
976: \Omega^{-1}$. The corresponding unreduced form (see Theorem
977: \ref{th:higherordercontinuous}) has an additional contraction rate
978: $-\frac{\bf F}{2}$.
979: \label{th:variationalenergy}
980: \end{theorem}
981: 
982: \Example{}{Consider the restricted Three-Body Problem \cite{ThreeBody}
983: $$
984: \ddot{\bf x} = 
985: 2 \left(
986: \begin{array}{ccc}
987:   0  & 1 & 0 \\
988:   -1 & 0 & 0 \\
989:   0  & 0 & 0
990: \end{array}
991: \right)
992: \dot{\bf x} +
993: \left(
994: \begin{array}{ccc}
995:   0  & 1 & 0 \\
996:   1 & 0 & 0 \\
997:   0  & 0 & 0
998: \end{array}
999: \right)
1000: {\bf x}
1001: -\nabla V({\bf x})
1002: $$ 
1003: with state ${\bf x} = (x, y, z)^T$, potential energy $V({\bf x})
1004: = -\nu \sqrt{y^2 + z^2 + (-1+x+y)^2}^{-1} - (1-\nu)
1005: \sqrt{y^2+z^2+(x+y)^2}^{-1}$, $\nu$ the ratio of the smaller
1006: mass to the larger mass, and the third body of zero mass.
1007: 
1008: Using ${\bf F} =  \left(
1009: \begin{array}{ccc}
1010:   0  & 1 & 0 \\
1011:   -1 & 0 & 0 \\
1012:   0  & 0 & 0
1013: \end{array}
1014: \right)$ in Theorem \ref{th:higherordercontinuous}, the reduced
1015: variational dynamics is
1016: $$
1017: \delta \ddot{\bf x} = - \Delta V \delta {\bf x}
1018: $$
1019: Using Theorem \ref{th:variationalenergy}, the contraction behavior of
1020: the three-body problem is the superposition of 
1021: 
1022: $\bullet$ $\ i \sqrt{\Delta V}$, which is indifferent for $\Delta V \ge
1023: 0$ and unstable otherwise.
1024: 
1025: $\bullet$ $\ -\frac{d \sqrt{\Delta V}}{dt} \sqrt{\Delta V}^{-1}$, where
1026: a tightening (relaxing) potential force adds semi-contracting
1027: (diverging)  behavior.
1028: }{ThreeBodyProblem}
1029: 
1030: 
1031: \section{Hamiltonian system dynamics} \label{Hamilton}
1032: 
1033: Consider the general $n$-dimensional Hamiltonian dynamics
1034: \cite{Lovelock} with sum convention over the free index
1035: $$
1036: \ddot{x}^j + \gamma_{kh}^j \dot{x}^k \dot{x}^h = - H^{jh} f_h - D_h^j
1037: \dot{x}^h
1038: $$
1039: with external forces $f_h(x^l, t)$, damping gain $D_h^j(x^l)$ and the
1040: Christoffel term $\gamma^j_{kh} = \frac{1}{2} H^{mj} \left(
1041: \frac{\partial H_{km}}{\partial x^h} + \frac{\partial H_{hm}}{\partial
1042: x^k} - \frac{\partial H_{kh}}{\partial x^m} \right)$ (section 7.2. in
1043: \cite{Lovelock}) of the symmetric inertia tensor $H_{lh}(x^m)$ and
1044: where we use the convention $H^{mj} = H_{mj}^{-1}$. The variational
1045: dynamics of the above is
1046: $$
1047: \delta \ddot{x}^j + \frac{\partial \gamma_{kh}^j}{\partial x^l} \delta
1048: x^l \dot{x}^k \dot{x}^h + 2 \gamma_{kh}^j \dot{x}^k \delta \dot{x}^h =
1049: - \frac{\partial (H^{jh} f_h)}{\partial x^l} \delta x^l - \frac{\partial
1050:   D_h^j}{\partial x^l} \delta x^l \dot{x}^h - D_h^j \delta \dot{x}^h
1051: $$
1052: Let us now compute with the above the covariant time-derivative
1053: (section 3.6. in \cite{Lovelock} and generalized contraction with metric
1054: in \cite{Lohm1}) of the covariant velocity
1055: variation $\delta \dot{x}^j + \gamma_{lh}^j \dot{x}^h \delta x^l$
1056: with respect to the metric $H_{lh}$ as
1057: \begin{eqnarray}
1058: \frac{d}{dt} \left( \delta \dot{x}^j + \gamma_{lh}^j
1059: \dot{x}^h \delta x^l \right) + \gamma_{ih}^j \dot{x}^h \left( \delta
1060: \dot{x}^i + \gamma_{kl}^i \dot{x}^k \delta x^l \right) &=&
1061: \nonumber \\ \delta \ddot{x}^j + \frac{\partial
1062: \gamma_{lh}^j}{\partial x^k} \dot{x}^k \dot{x}^h \delta x^l +
1063: \gamma_{lh}^j \ddot{x}^h \delta x^l + \gamma^j_{lh} \dot{x}^h \delta
1064: \dot{x}^l + \gamma_{ih}^j \dot{x}^h \left( \delta \dot{x}^i +
1065: \gamma_{kl}^i \dot{x}^k \delta x^l \right) &=& \nonumber \\
1066: K^j_{lkh} \dot{x}^k \dot{x}^h \delta x^l - H^{jh} \left(
1067: \frac{\partial f_h}{\partial x^l} - \gamma^k_{hl} f_k \right) \delta
1068: x^l - D_h^j \left( \delta \dot{x}^h + \gamma^h_{kl} \dot{x}^k \delta
1069: x^l \right)  \label{eq:varacceldynamics}
1070: \end{eqnarray}
1071: with curavture tensor $K^j_{lkh} = \frac{\partial
1072: \gamma_{lh}^j}{\partial x^k} - \frac{\partial \gamma_{kh}^j}{\partial
1073: x^l} + \gamma_{ih}^j \gamma_{kl}^i - \gamma_{li}^j \gamma_{kh}^i$
1074: (section 7.3. in \cite{Lovelock}), covariant derivative of the
1075: external forces $\frac{\partial f_h}{\partial x^l} - \gamma^k_{hl}
1076: f_k$, and where we have assumed the covariant derivative of $D_h^j$,
1077: that is $\frac{\partial D_h^j}{\partial x^l} + \gamma_{lk}^j D_h^k -
1078: \gamma_{lh}^k D_k^j$ (see section 3.6 in \cite{Lovelock}), to vanish
1079: which is usually the case for mechanical damping.
1080: 
1081: Note that according section 7.3 in \cite{Lovelock} corresponds
1082: $K^i_{lkh} H_{ji} Y^k Y^h X^j X^l$ to the Riemannian or Gaussian
1083: curvature of the 2-D subspace span by $X$ and $Y$. Hence the convexity
1084: of $H$ orthogonal to $\dot{x}$ acts as a spring, whose gain
1085: increases linearly with velocity. 
1086: 
1087: Also note that (\ref{eq:varacceldynamics}) can be used directly in
1088: combination with Theorem 2 in \cite{Lohm5} to show under which
1089: condition the covariant Hessian of the action $\phi$ becomes convex,
1090: which implies contraction behavior.
1091: 
1092: However we use here Theorem \ref{th:higherordercontinuous} for ${\bf
1093: F} = {\bf D}$, which is more general than the above (i.e. it allows to
1094: assess complex solutions of the Hessian dynamics in Theorem 2 in
1095: \cite{Lohm5}), to conclude:
1096: 
1097: \begin{theorem}
1098: Consider for $t\ge0$ the $n$-dimensional ($n \ge 1$) dynamics  
1099: with sum convention over the free index
1100: \begin{equation}
1101: \ddot{x}^j + \gamma_{kh}^j \dot{x}^k \dot{x}^h = - H^{jh} f_h - D_h^j
1102: \dot{x}^h
1103: \label{eq:Hamiltonian}
1104: \end{equation}
1105: with external forces $f^j(x^l, t)$, damping gain $D_h^j(x^l)$ with
1106: covariant derivative $\frac{\partial D_h^j}{\partial x^l} +
1107: \gamma_{lk}^j D_h^k - \gamma_{lh}^k D_k^j = 0$, and the Christoffel
1108: term $\gamma^j_{kh} = \frac{1}{2} H^{mj} \left( \frac{\partial
1109: H_{km}}{\partial x^h} + \frac{\partial H_{hm}}{\partial x^k} -
1110: \frac{\partial H_{kh}}{\partial x^m} \right)$ of the symetric,
1111: u.p.d. and invertible inertia tensor $H_{lh}(x^m)$.
1112: 
1113: Let us define a constant $\lambda \ge 0$ as the largest singular value
1114: $\forall t \ge 0$ as
1115: \begin{small}
1116: \begin{equation} 
1117: \left( -K_{ipkh} \dot{x}^k \dot{x}^h + \frac{\partial f_i}{\partial
1118: x^p} - \gamma^h_{ip} f_h + H_{ih} \frac{D_k^h D_p^k}{4} \right) H^{ij}
1119: \left( -K_{jlkh} \dot{x}^k \dot{x}^h + \frac{\partial f_j}{\partial
1120: x^l} - \gamma^h_{jh} f_h + H_{jh} \frac{D_k^h D_l^k}{4} \right) \le
1121: \lambda^2 H_{pl} \label{eq:characteristicHamiltonian}
1122: \end{equation}
1123: \end{small}
1124: with curavture tensor $K_{jlkh} = H_{jo} \frac{\partial
1125: \gamma_{lh}^o}{\partial x^k} - \frac{\partial \gamma_{kh}^o}{\partial
1126: x^l} + \gamma_{ih}^o \gamma_{kl}^i - \gamma_{li}^o \gamma_{kh}^i$.
1127: 
1128: We can then conclude on contraction rate $- \frac{D_h^j}{2} +
1129: \lambda$. \label{th:Hamiltoniancontinuous}
1130: \end{theorem}
1131: 
1132: Also note that taking the double integral of the dynamics leads to an
1133: exponential Lyaponov energy stability proof for the autonomous
1134: case. In this sense Theorem 5 represents a variational extension of
1135: the classical energy based Lyaponov proofs for autonomous systems.
1136: Note that at a variational energy approach a tightening spring
1137: (positive $\dot{\Omega}$) leads to semi-contraction behavior.
1138: 
1139: \Example{}{Let us now illustrate the simplicity of the stability
1140: results using the inertia tensor as metric. Consider the Euler
1141: dynamics of a rigid body, with Euler angles ${\bf x} = (\psi, \theta,
1142: \phi)^T$ and measured rotation vector ${\bf \omega}$ in body
1143: coordinates \cite{Goldstein}
1144: \begin{equation}
1145: \dot{\bf x} = \left(
1146: \begin{array}{ccc}
1147:   1 &          0 & - \sin \theta \\
1148:   0 &  \cos \psi &   \cos \theta \sin \psi \\
1149:   0 & -\sin \psi &   \cos \theta \cos \psi
1150: \end{array}
1151: \right)^{-1} {\bf \omega} \label{eq:rotationdynamics}
1152: \end{equation}
1153: The underlying energy is  
1154: \begin{eqnarray}
1155: h &=& \frac{1}{2} {\bf \omega}^T {\bf \omega} \nonumber \\
1156: &=& \frac{1}{2}\dot{\bf x}^T 
1157: \left(
1158: \begin{array}{ccc}
1159:   1             & 0 & -\sin \theta \\
1160:               0 & 1 & 0 \\
1161:   - \sin \theta & 0 & 1
1162: \end{array}
1163: \right)
1164: \dot{\bf x} \nonumber
1165: \end{eqnarray}
1166: After a straightforward but tedious calculation we can compute
1167: $$
1168: \frac{d}{dt} \left( \delta {\bf x}^T {\bf H} \delta {\bf x} \right) = 0
1169: $$
1170: Thus, the Euler dynamics (\ref{eq:rotationdynamics}) is globally
1171: indifferent. Note that this can also be seen from the
1172: quaternion angular dynamics, whose Jacobian is
1173: skew-symmetric~\cite{inertial}.}{Eulerdynamics}
1174:  
1175: \Example{}{The convexity of a the inertia tensor can act similarly to
1176: a stabilizing potential force in the variational Hamiltonian dynamics
1177: (\ref{eq:varacceldynamics}). Consider a rotating point mass of mass
1178: $m$ on a ball with radius $R$. The Hamiltonian energy is
1179: $$
1180: h = \frac{m R^2}{2} \dot{\bf x}^T
1181: \left(
1182: \begin{array}{cc}
1183:   1 &  0  \\
1184:   0 &  \sin^2 \phi 
1185: \end{array}
1186: \right)
1187: \dot{\bf x}
1188: $$
1189: with latitude $\phi$ and longitude $\psi$ in ${\bf x} = (\phi,
1190: \psi)^T$. The curvature tensor can be computed e.g. with MAPLE as
1191: $$
1192: K_{jlkh} \dot{x}^k \dot{x}^h =
1193: \frac{m R^2 \sin^2 \phi}{2} 
1194: \left( \begin {array}{cc}  
1195: -\dot{\psi}^2         & \dot{\psi} \dot{\phi} \\
1196: \dot{\psi} \dot{\phi} & -\dot{\phi}^2  
1197: \end {array} 
1198: \right)
1199: $$ which scales as the inertia tensor with $\frac{m R^2}{2}$ and is
1200: negative orthogonal to the velocity and indifferent along the
1201: velocity. Hence the convexitiy in Theorem
1202: \ref{th:Hamiltoniancontinuous} acts under motion as a stabilizing
1203: spring, that lets two moving neighboring trajectories oscillate around
1204: each other.}{ball}
1205: 
1206: \Example{}{Theorem \ref{th:Hamiltoniancontinuous} can be used to
1207: define observers or tracking controllers for time-varying Hamiltonian
1208: systems. Consider a two-link robot manipulator, with kinetic energy
1209: $$
1210: \frac{1}{2}
1211: \left( \begin{array}{cc} \dot{q}_1 & \dot{q}_2 \end{array} \right)
1212: \left( \begin{array}{cc} 
1213: a_1 + 2 a_2 \cos q_2 & a_2 \cos q_2 + a_3 \\
1214: a_2 \cos q_2 + a_3 & a_3 
1215: \end{array} \right)
1216: \left( \begin{array}{c} \dot{q}_1 \\ \dot{q}_2 \end{array} \right)
1217: $$
1218: with $a_1 = m_1 l_{c1}^2 + I_1 + m_2(l_1^2+l_{c2}^2) + I_2$, $a_2 =
1219: m_2 l_1 l_{c2}$ and $a_3 = m_2 L_{c2}^2 + I_2$ and $-\pi \le q_1, q_2
1220: \le \pi$.  Let us assume that $q^j$ is measured and define the
1221: observer
1222: \begin{eqnarray}
1223: \dot{\hat{\omega}}^j + \gamma_{kh}^j(\hat{q}^j) \dot{\hat{q}}^k
1224: \dot{\hat{q}}^h &=& - H^{hj} f_h \nonumber \\ \dot{\hat{q}}^j &=&
1225: \hat{\omega}^j + d_h^j ( q^h - \hat{q}^h ) \nonumber
1226: \end{eqnarray}
1227: with the external forces
1228: $$
1229: f_h = 
1230: \left( 
1231: \begin{array}{c}
1232: g (m_1 l_{c1} + m_e l_1) \cos q_1 + g m_e l_{ce} \cos(q_1 + q_2) +
1233: \tau_1 \\ g m_e l_{ce} \cos(q_1 + q_2) + \tau_2
1234: \end{array}
1235: \right)
1236: + k_h^i \left(\hat{q}^i -q^i \right)
1237: $$  
1238: and external torques $\tau_1$, $\tau_2$. The above is equivalent to
1239: (\ref{eq:Hamiltonian}) 
1240: $$
1241: \ddot{\hat{q}}^j + \gamma_{kh}^j(\hat{q}^j) \dot{\hat{q}}^k
1242: \dot{\hat{q}}^h = - H^{jh} \left( f_h - d_{hk} \dot{q}^k \right)
1243: - d_h^j \dot{\hat{q}}^h  
1244: $$
1245: where the covariant derivative of a constant scalar $d_h^j$
1246: vanishes. The curvature tensor can be computed e.g. with MAPLE as
1247: $$
1248: K_{jlkh} \dot{x}^k \dot{x}^h =
1249: \frac{ \left( a_1 a_3 - a_3^2 - a_2^2 \right) a_2 \cos q_2} {a_1 a_3
1250:   -a_2^2 \cos^2 q_2 -a_3^2}
1251: \left( \begin {array}{cc}  
1252: -\dot{q}_2^2 & \dot{q}_1 \dot{q}_2 \\
1253: \dot{q}_1 \dot{q}_2 & -\dot{q}_1^2  
1254: \end {array} 
1255: \right)
1256: $$
1257: The curvature is for $a_1 a_3 \ge a_2^2 + a_3^2$ convex (concave) for
1258: $-\frac{\pi}{2} \le q_2 \le \frac{\pi}{2}$ ($-\frac{\pi}{2} \ge q_2 \
1259: or \ q_2 \ge \frac{\pi}{2}$) and accordingly (de)-stabilizes the
1260: dynamics when the arm is retracted (extended). Let us now compute the
1261: covariant derivative of the external forces
1262: $$
1263: \frac{\partial f_i}{\partial x^p} + \gamma^h_{ip} \left( f_h - d_{hk}
1264: \dot{q}^k \right) = -k_p^i + \gamma^h_{ip} \left( f_h - d_{hk}
1265: \dot{q}^k \right)
1266: $$
1267: with $-\gamma^1_{11} = \gamma^2_{12} = \gamma^2_{21} = \gamma^2_{22} =
1268: \frac{ \left( a_2 \cos q_2 + a_3 \right) a_2 \sin q_2} {a_1 a_3 -a_2^2
1269: \cos^2 q_2 -a_3^2}$, $\gamma^1_{12} = \gamma^1_{21} = \gamma^1_{22} =
1270: - \frac{a_3 a_2 \sin q_2} {a_1 a_3 -a_2^2 \cos^2 q_2 -a_3^2}$ and
1271: $\gamma^2_{11} = \frac{ \left( a_1 + 2 a_2 \cos q_2 \right) a_2 \sin
1272: q_2} {a_1 a_3 -a_2^2 \cos^2 q_2 -a_3^2}$.
1273: 
1274: The spring gain $k_p^i$ stabilizes the system, whereas the supporting
1275: force $f_h - d_{hk} \dot{q}^k$ can (de)-stabilize the system
1276: proportional to the magnitude of the supporting force. Note that $f_h
1277: - d_{hk} \dot{q}^k$ can (de)-stabilize a curved system is unavoidable
1278: since here no constant or parallel (force) vectors exist, whose
1279: covariant derivative vanishes \cite{Lovelock}.
1280: 
1281: Computing $\lambda$ from (\ref{eq:characteristicHamiltonian}) then
1282: allows with Theoreom \ref{th:Hamiltoniancontinuous} to compute bounds
1283: on the velocity $\dot{q}^j$ and external forces $f_h - d_{hk}
1284: \dot{q}^k$ for which global contraction behavior can be concluded.
1285: 
1286: System responses to a control input $\tau_i = (\sin t, \cos 5 t)$,
1287: initial conditions $q^i(0) = (-\frac{\pi}{2}, \pi)$ rad, $\dot{q}^i(0)
1288: = (3, -3)$ rad/s, $\hat{q}^i(0) = (- \frac{\pi}{2}, \pi)$ rad,
1289: $\hat{\dot{q}}^i(0) = (-5, 5)$ rad/s and parameters $m_1 = 1$ kg, $l_1
1290: = 1$ m, $m_e = 2$ kg, $I_1 = 0.12$ kgm$^2$, $l_{c1} = 0.5$ m, $I_e =
1291: 0.25$ kgm$^2$, $l_{ce} = 0.6$ m, $d_h^j = 5$ Nms/rad, $k_h^i = 5$
1292: Nm/rad are illustrated in figure \ref{fig:robotqhat1} and
1293: \ref{fig:robotwhat1}. The solid lines represent the real plant, and
1294: the dashed lines the observer estimate.
1295: 
1296: \begin{figure}
1297: \begin{center}
1298: \includegraphics[scale=0.3]{robotqhat1.eps}
1299: \end{center}
1300: \caption{Positions of two-link robot}
1301: \label{fig:robotqhat1}
1302: \end{figure}
1303: 
1304: \begin{figure}
1305: \begin{center}
1306: \includegraphics[scale=0.3]{robotwhat1.eps}
1307: \end{center}
1308: \caption{Velocities of two-link robot}
1309: \label{fig:robotwhat1}
1310: \end{figure}
1311: The above observers provide a simple alternative to current design
1312: methods [see e.g., Berghuis and Nijmeyer, 1993; Marino and Tomei,
1313: 1995], and guarantees local (for bounded velocities and time-varing
1314: inputs) exponential convergence.
1315: 
1316: Note that Theoreom \ref{th:Hamiltoniancontinuous} can also be used to
1317: bound the diverging behavior, caused by the concave inertia when the
1318: robot arm is pointing inwards, of the double inverted pendulum, when
1319: no damping or stabilizing potential force is applied.}{robot} 
1320: 
1321: \Example{}{Biology found a solution to the problem that a time-varying
1322: supporting torque can destabilize a system. 
1323: 
1324: Recently, there has been considerable interest in analyzing feedback
1325: controllers for biological motor control systems as combinations of
1326: simpler elements, or motion primitives.  For instance \cite{Bizzi} and
1327: \cite{Mussa} stimulate a small number of areas (A, B, C, and D) in a
1328: frog's spinal cord and measure the resulting torque angle relations.
1329: 
1330: Force fields seem to add when different areas are stimulated at the
1331: same time so that \cite{Bizzi} and \cite{Mussa} propose the following
1332: biological control inputs
1333: $$
1334:   f_i = - \sum_{l=1}^n k_l(t) f_{il} (q^j)
1335: $$
1336: where each single torque $k_l(t) f_{il} (q^j)$ results from the
1337: stimulation of area $l$ in the spinal cord. With force measurements
1338: the above authors did show that the covariant derivative
1339: $\frac{\partial f_{il}}{\partial x^p} + \gamma^h_{ip} f_{hl}$ of
1340: $f_{il}$ with respect to the inertia tensor of the frog's leg or body
1341: is uniformly positive definite.
1342: 
1343: Likely candidates for $k_i(t)$ are positive upper and lower bounded
1344: sigmoids and pulses and periodic activation patterns.
1345: 
1346: Using Theorem \ref{th:Hamiltoniancontinuous} or the discussion in
1347: \Ex{robot} with sufficient damping then allows to compute a maximal
1348: $\dot{q}^j$ for which exponential convergence to a single motion is
1349: guaranteed.
1350: 
1351: Note that the achievement of tracking control with a proportional gain
1352: $k_l(t)$ rather than an additional supporting force as in \Ex{robot}
1353: has the advantage that the supporting force has no impact on the
1354: contraction behavior anymore.}{motionprimitives.}
1355: 
1356: \section{Concluding remarks}
1357: 
1358: The research in this paper can be extended in several directions, as
1359: the development suggests.
1360: 
1361: Some of the extensions will likely require the combination of the
1362: above results with a simplifying metric pre-transformation, as
1363: mentioned in section 3.1.  In particular, classical transformation
1364: ideas in nonlinear control such as feedback linearization and
1365: flatness~\cite{flatness} typically use linear time-invariant target
1366: dynamics, while the framework provided in this paper should allow
1367: considerably more flexibility. This, combined with the fact that a
1368: metric transformation such as $\ \delta {\bf z} = {\bf \Theta} \delta
1369: {\bf x}\ $ need not be integrable (i.e. does not require an explicit
1370: ${\bf z}$ to exist), could potentially lead to useful generalisations
1371: of these methods.
1372: 
1373: {\bf Acknowledgement}\ \ The authors are grateful to Yong Zhao for
1374: performing the simulations and for stimulating discussions, and to Wei
1375: Wang for thoughtful comments and suggestions.
1376: 
1377: \begin{thebibliography}{XX}
1378: \begin{small}
1379: 
1380: \bibitem{Rouchon} Aghannan, N., Rouchon, P.,
1381: An Intrinsic Observer for a Class of Lagrangian Systems, {\em
1382: I.E.E.E. Trans. Aut. Control}, {\bf 48(6)} (2003).
1383: \bibitem{Parrilo} Aylward E., Parrilo P., and J.J.E. Slotine,
1384: Algorithmic search for contraction metrics via SOS programming,
1385: {\em submitted to the 2006 American Control Conference}. 
1386: \bibitem{Bert} Bertsekas, D., and Tsitsiklis, J., Parallel and
1387: distributed computation: numerical methods, {\it Prentice-Hall}, 1989.
1388: \bibitem{Bizzi} Bizzi E., Giszter S.F., Loeb E., Mussa-Ivaldi F.A.,
1389: Saltiel P., {\it Trends in Neurosciences. Review 18:442}, 1995.
1390: \bibitem{Bryson} Bryson A., Ho, Y., Applied Optimal Control, {\em
1391: Taylor and Francis}, 1975.
1392: \bibitem{flatness} Fliess M., Levine J., Martin Ph., and
1393: Rouchon P., Flatness and defect of nonlinear systems: introductory
1394: theory and examples. {\it Int. J. Control, 61(6)}, 1995. 
1395: \bibitem{Goldstein} Goldstein H., Classical Mechanics, {\em Addison
1396: Wesley}, 1980.
1397: \bibitem{Hart} Hartmann, P. Ordinary differential equations,
1398: {\em John Wiley $\&$ Sons, New York}, 1964.
1399: \bibitem{Jouff} Jouffroy J. and J. Opderbecke. Underwater vehicle trajectory
1400: estimation using contracting PDE-based observers, {\em American
1401: Control Conference}, Boston, Ma, 2004.
1402: \bibitem{Kailath} Kailath, T., Linear Systems, {\em Prentice Hall},
1403: 1980.
1404: \bibitem{Lew} Lewis, D.C., Metric properties of differential equations,
1405: {\it American Journal of Mathematics,} {\bf 71}, pp. 294-312, 1949.
1406: \bibitem{Lohm1} Lohmiller, W., and Slotine, J.J.E., On Contraction
1407: Analysis for Nonlinear Systems, {\em Automatica, 34(6)}, 1998.
1408: \bibitem{Lohm4} Lohmiller, W., and Slotine, J.J.E., Nonlinear Process
1409: Control Using Contraction Theory, {\em A. I. Che. Journal}, March
1410: 2000.
1411: \bibitem{Lohm5} Lohmiller, W., and Slotine, J.J.E., Contraction
1412: Analysis of Nonlinear Distributed Systems, {\em International Journal
1413: Of Control, 78(9)}, 2005.
1414: \bibitem{Lovelock} Lovelock D., and Rund, H., Tensors, Differential
1415: Forms, and Variational Principles, {\em Dover}, 1989.
1416: \bibitem{Mussa} Mussa-Ivaldi, F.A., {\it I.E.E.E.  International
1417: Symposium on Computational Intelligence in Robotics and Automation},
1418: 1997.
1419: \bibitem{Nguyen} Nguyen, T.D., and Egeland, O. Observer Design for a
1420: Towed Seismic Cable, {\it American Control Conference, Boston} (2004)
1421: \bibitem{Shamma} Shamma, J., and Gurdal, A., Dynamic Fictitious Play,
1422: Dynamic Gradient Play and Distributed Convergence to Nash Equilibra,
1423: {\em IEEE Transactions on Automatic Control}, March 2005
1424: \bibitem{Sl91} Slotine and Li, Applied Nonlinear Control, {\em Prentice Hall},
1425: 1991.
1426: \bibitem{inertial} Zhao, Y., and Slotine, J.J.E., Discrete Nonlinear
1427: Observers for Inertial Navigation, {\it Systems and Control Letters,
1428: 54(8)}, 2005.
1429: \bibitem{ThreeBody} The Restricted Three Body Problem,
1430: http://www.physics.cornell.edu/ sethna/teaching/sss/jupiter/Web/Rest3Bdy.htm
1431: \end{small}
1432: \end{thebibliography}
1433: 
1434: \newpage
1435: 
1436: \end{document}
1437: 
1438: 
1439: For discrete $n^{\rm th}$-order systems we will give a necessary and
1440: sufficient condition to bound the exponential discrete stability over
1441: $n$ iterations.
1442: 
1443: 
1444: For reduced continuous $n^{\rm th}$-order systems we will give a necessary
1445: and sufficient condition to bound the exponential continous stability
1446: over $n$ time derivatives.
1447: 
1448: 
1449: In addition, this leads to more accurate convergence rates estimates,
1450: as the following example motivates.
1451: 
1452: \Example{}{Consider two first order LTI systems in a classical
1453:   feedback combination
1454: \begin{eqnarray} 
1455: \dot{x}_1 &=& - \lambda_1 \ {x}_1 - k\ x_2 \nonumber \\
1456: \dot{x}_2 &=& - \lambda_2 \ {x}_2 + x_1 \nonumber
1457:  \end{eqnarray}
1458: where $k$ and the $\lambda_i$ are strictly positive constants. The
1459: system is contracting with metric $\ {\rm diag}(1,k)$, leading to a
1460: contraction rate $\ {\rm min}(\lambda_1,\lambda_2)$. However, we see
1461: (e.g., from root locus) that this estimate of the exponential rate of
1462: convergence with this specific metric can be unduly conservative,
1463: particularly if the $\lambda_i$ are very different.
1464: 
1465: This situation is similar to the classical case of deriving Lyapunov
1466: functions for LTI systems, where the Lyapunov matrix equation
1467: generally gives conservative rate estimates~\cite{Sl91}. The same
1468: situation occurs for discrete systems.}{ex:3.1}
1469: 
1470: 
1471: 
1472: &=& -{\bf A}_{n-1} \left( \left( \begin{array}{c} n-1 \\ 0 \end{array}
1473:   \right) L^{n-2} {\bf F} \delta {\bf x}_o + ... +
1474: \left( \begin{array}{c} n-1 \\ n-2 \end{array} \right) L^o {\bf F}
1475: \delta {\bf x}_{n-2} +
1476: \left( \begin{array}{c} n-1 \\ n-1 \end{array} \right) \delta {\bf
1477:   x}_{n-1} \right) \nonumber \\ 
1478: &&- ... - {\bf A}_o \delta {\bf x}_o \nonumber \\
1479: &&- \left( \begin{array}{c} n-1 \\ 0 \end{array} \right) L^{n-1}
1480:   {\bf F} \delta {\bf x}_o - ... -
1481: \left( \begin{array}{c} n-1 \\ n-2 \end{array} \right) L^{1} {\bf F}
1482: \delta {\bf x}_{n-2} - 
1483: \left( \begin{array}{c} n-1 \\ n-1 \end{array} \right) L^o {\bf F} 
1484: \delta {\bf x}_{n-1} \nonumber \\
1485: &&- ... - \left( \begin{array}{c} 1 \\ 0 \end{array} \right) L^1
1486: {\bf F} \delta {\bf x}_{n-2} -
1487: \left( \begin{array}{c} 1 \\ 1 \end{array} \right) L^o {\bf F}
1488: \delta {\bf x}_{n-1} \nonumber \\
1489: 
1490: 
1491: 
1492: Note that a corresponding separation principle between contracting
1493: observer and controller is shown in \cite{Lohm4}.
1494: 
1495: 
1496: 
1497: Typically ${\bf F}$ will be chosen to cancel the damping
1498: gain $\Theta$. 
1499: \Example{}{Consider the Van der Pol dynamics
1500: $$
1501: \ddot{x} + a_1(x) \dot{x} + x \ = \ u(t)
1502: $$
1503: with $a_1 \ge 0$. The virtual dynamics is
1504: $$
1505: \delta \ddot{x} = - a_1 \delta \dot{x} - \left( 1 + \dot{a}_1 \right)
1506: \delta x
1507: $$
1508: Choosing $F = -a_1$ in Theorem \ref{th:higherordercontinuous} leads to
1509: the reduced form
1510: $$
1511: \delta \ddot{x} = a_1 (x)\ \delta \dot{x} + \delta x  
1512: $$
1513: whose divergence rate is less than $a_1$. Hence from Theorem
1514: \ref{th:variationalenergy} the overall system is semi-contracting with
1515: contraction rate $F + a_1 = 0$ (Figure
1516: \ref{fig:example_37}).}{VanderPol}
1517: 
1518: \begin{figure}[h]
1519: \begin{center}
1520: \epsfig{figure=example_37.eps,height=60mm,width=90mm}
1521: \end{center}
1522: \caption{\small Van der Pol with $a_1=8+\sin x$ and $u(t)=\cos
1523: t$} 
1524: \label{fig:example_37}
1525: \end{figure}
1526: 
1527: \Example{}{Consider again the second-order LTV dynamics
1528: $$
1529: \ddot{x} + a_1(t)\ \dot{x} + a_o(t)\ x \ = \ u(t)
1530: $$
1531: Its virtual dynamics is
1532: $$
1533: \delta \ddot{x} = - a_1 \delta \dot{x} - a_o \delta x  
1534: $$
1535: Using the above and Barbalat's lemma~\cite{Sl91} shows asymptotic
1536: convergence of $\delta x$ to zero if both $a_o\ $ and $ \ a_1 +
1537: \frac{1}{2} \ \frac{d}{dt} \ \rm{ln}\ a_o$ are uniformly positive
1538: definite.}{LTV}
1539: 
1540: 
1541: 
1542: While the stability of LTI systems simple requires the poles to be
1543: negative for continuous systems or magnitude less than $1$ for
1544: discrete systems this paper extends this result to general nonlinear,
1545: time-varying systems if only a reduced set of poles is accepted. 
1546: 
1547: E.g. for continuous second-order systems the poles simply have to lie in
1548: $\pm 45^o$ quadrant of the left-half complex plane. For the discrete
1549: second-order case the poles either have to be real or to lie in the
1550: circle segment through $\pm i$.
1551: 
1552: Indeed this is equivalent to a necessary and sufficient stability
1553: principle for a system to converge exponentially at least in $n$
1554: iterations for the discrete case or to be bounded exponentially in the
1555: $n$'th time derivative for the continuous case.
1556: 
1557: In addition a variational energy conservation principle was
1558: derived. Here  increasing potential energy Hessians, as in the case of
1559: the Three Body Problem, lead to an additional semi-contraction behavior
1560: on top to the contraction behavior of the damping.
1561: 
1562: 
1563: