math0211382/fa.tex
1: \NeedsTeXFormat{LaTeX2e}
2: % LaTeX 2.09 can't be used (nor non=LaTeX) [1994/12/01]
3: % LaTeX date must December 1994 or later
4: 
5: %%%%%%%%%%%%%%% Macros and Definitions %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: %\newcommand{\revision}{kryolit}
7: %\newcommand{\pending}[1]{ \par\medskip{\sc #1}
8: %\marginpar{\sc {\Writinghand} } \par\medskip }
9: %\newcommand{\pig}{{\Writinghand}\marginpar{{\Writinghand}}}
10: 
11: \newif\ifpdf \ifx\pdfoutput\undefined \pdffalse \else \pdfoutput=1
12: \pdftrue \fi
13: 
14:  \newenvironment{fig}[2] { \begin{center}
15:  \begin{figure}
16:  \def\myfigcaptionlar{#1} \def\myfigcaptionvar{{#2}}}
17:  {\caption{\myfigcaptionvar}\label{\myfigcaptionlar}
18:  \end{figure}\end{center}}
19:  
20:  \newcommand{\psfig}[3] {\begin{fig}{#1}{#2}\if 0#3 \else
21:  \begin{center}\includegraphics{#3}\end{center}\fi
22:  \end{fig} }
23:  
24:  \newcommand{\texfig}[3] {\begin{fig}{#1}{#2} \if 0#3 \else
25:  \begin{center}
26:  \input{#3}
27:  \end{center}
28:  \fi
29:  \end{fig}
30:  \if 0#3 \fi  }
31:  
32:  \newcommand{\citez}[1]{\cite{#1}} \newcommand{\In}[1]{{\tt
33:  In$[$#1$]$}}
34: 
35: %%%%%%%%%%%%%%% Document Openings %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
36: \documentclass[openany,oneside,fleqn,11pt]{article}
37: \usepackage{amsmath,amssymb,amscd,amsthm} 
38: % \usepackage{marvosym}
39: % \usepackage{makeidx}
40: \usepackage{url} 
41: \usepackage[numbers]{natbib}
42: % \usepackage[small,compact,sc]{titlesec} 
43:  \usepackage{float}
44: \ifpdf \usepackage[pdftex]{graphicx}
45: \else 
46: % \usepackage{pandora} 
47:  \usepackage{graphicx} 
48: \fi
49: %\usepackage[mathcal]{euscript}
50: %\usepackage{showkeys}  
51: \usepackage[bookmarks=true]{hyperref} %this one must be the last one
52: %\usepackage[a4paper, left=15mm, right=15mm]{geometry} 
53: \newcommand{\hyref}[1]{\href{#1}{\tt{#1}}}
54: \newcommand{\hyreff}[2]{\href{#1}{\tt{#2}}}
55:  
56: \begin{document}
57: %\DeclareGraphicsExtensions{.pdf} 
58: \DeclareGraphicsExtensions{.mps}
59: 
60: 
61: %%%%%%%%%%%%%%% Textual Macros %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62: \newcommand{\Ito}{It\^o }
63: 
64: \newcommand{\Mathematica}{{\it
65:  Mathematica }}
66: \newcommand{\MathReader}{{\it
67:  MathReader }}
68: 
69: \newcommand{\abbrev}[1]{#1. }
70: \newcommand{\ie}{\abbrev{i.e}}
71: \newcommand{\pg}{\abbrev{p}}
72: \newcommand{\eg}{\abbrev{eg}}
73: \newcommand{\etc}{\abbrev{etc}}
74: \newcommand{\vs}{\abbrev{vs}}
75: \newcommand{\propan}{\abbrev{Propos}}
76: \newcommand{\theopan}{Theorem}
77: 
78: %%%%%%%%%%%%%%% Propositions %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
79: \theoremstyle{theorem} 
80: \newtheorem{prop}{Proposition}
81: \newtheorem{lemma}[prop]{Lemma} 
82: \newtheorem{theo}[prop]{Theorem}
83: \newtheorem{coroll}[prop]{Corollary}
84: \newcommand{\refprop}[1]{Proposition~\ref{#1}}
85: \newcommand{\refppropag}[1]{\propan~\ref{#1} \pg~\pageref{#1}}
86: \newcommand{\reftpropag}[1]{\theopan~\ref{#1} \pg~\pageref{#1}}
87: \newcommand{\refcorol}[1]{Corollary~\ref{#1} \pg~\pageref{#1}}
88: \theoremstyle{definition} 
89: \newtheorem{adefn}[prop]{Definition}
90: \newenvironment{defn}{\begin{adefn}}{\hfill$\blacksquare$\end{adefn}}
91: \theoremstyle{remark}
92: \newtheorem{arem}[prop]{Remark}
93: \newtheorem{aexample}[prop]{Example}
94: \newcommand{\defbf}[1]{{\bf #1}}
95: \newenvironment{solution}{\begin{proof}[Solution]}{\end{proof}}
96: \newenvironment{rem}{\begin{arem}}{\hfill$\triangledown$\end{arem}}
97: \newenvironment{example}{\begin{aexample}}{\hfill$\vartriangle$\end{aexample}}
98: 
99: \newcommand{\algitem}[1]{\item #1}
100: 
101: \floatstyle{boxed} \newfloat{Algorithm}{H}{lop}
102: \newenvironment{alg}[5]{ \def\myalgcaptionvar{{\bf #2}$\quad[$#1$]$}
103:  \if 0#5 \def\myalgcaptionvarl{\relax} \else
104:  \def\myalgcaptionvarl{\noindent{#5}} \fi
105:  \begin{Algorithm}
106:  \noindent{\bf input: }#3
107:  \noindent{\bf output: }#4
108:  \begin{itemize}
109:  } {
110:  \end{itemize}
111:  \myalgcaptionvarl
112: % \hrule
113:  \caption{\myalgcaptionvar}
114:  \end{Algorithm}
115: }
116:  
117: \newenvironment{romanenu}
118: {\begin{enumerate}\renewcommand{\labelenumi}{(\roman{enumi})}}
119:  {\end{enumerate}}
120: 
121: 
122: 
123: %%%%%%%%%%%%%%% Notation %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
124: % common ==========================
125: \newcommand{\rfor}[1]{\quad\text{for}\quad#1}
126: \newcommand{\firstfor}{\qquad&&\text{for}\quad}
127: \newcommand{\nextfor}{&&\text{for}\quad}
128: 
129: \newcommand{\alterbar}{\vec}
130: 
131: \newcommand{\origi}[1]{{#1}_0} \newcommand{\define}{:=}
132: \newcommand{\definer}{=:} \newcommand{\compose}{\circ}
133: \newcommand{\frahalf}{{\frac{1}{2}}}
134: 
135: % iteration =======================
136: \newcommand{\iva}[1]{0 \le i \le #1} \newcommand{\jva}[1]{0 \le j \le
137:  #1}
138: %\newcommand{\ssum}[3]{{\sum_{#2 \le #1 \le #3}}}
139: \newcommand{\ssum}[3]{{\sum_{#1 = #2}^{#3}}}
140: \newcommand{\ssumi}[1]{{\ssum{i}{1}{#1}}}
141: 
142: % sets and spaces =================
143: \newcommand{\RR}{{\mathbb R}} \newcommand{\NN}{{\mathbb N}}
144: \newcommand{\ZZ}{{\mathbb Z}} \newcommand{\CC}{{\mathbb C}}
145: 
146: \newcommand{\RntoRn}{{\colon\RR^n \to \RR^n}}
147: \newcommand{\RntoR}{{\colon\RR^n \to \RR}}
148: \newcommand{\RtoRn}{{\colon\RR \to \RR^n}}
149: \newcommand{\RtoR}{{\colon\RR \to \RR}}
150: 
151: % systems =========================
152: \newcommand{\dsys}[3]{\left(#1,#2,U,\origi{#3}\right)}
153: \newcommand{\ssys}[4]{\left(#1,#2,#3,U,\origi{#4}\right)}
154: \newcommand{\osys}[4]{\left(#1,#2,#3,U,\origi{#4}\right)}
155: % \newcommand{\dsys}[3]{\left(#1(#3),#2(#3),U,\origi{#3}\right)}
156: % \newcommand{\ssys}[4]{\left(#1(#4),#2(#4),#3(#4),U,\origi{#4}\right)}
157: % \newcommand{\osys}[4]{\left(#1(#4),#2(#4),#3(#4),U,\origi{#4}\right)}
158: \newcommand{\dsysfgx}{\dsys{f}{g}{x}}
159: \newcommand{\dosysfghx}{\osys{f}{g}{h}{x}}
160: \newcommand{\ssysfgx}{\ssys{f}{g}{\sigma}{x}}
161: 
162: \newcommand{\bigdsys}[3]{\left(#1\, , \, #2\, , \, U\, , \,
163:  \origi{#3}\right)} \newcommand{\bigssys}[4]{\left(#1\, , \, #2\, ,
164:  \, #3\, , \, U\, , \, \origi{#4}\right)}
165: 
166: \newcommand{\clssys}{{\mathbb X}} \newcommand{\clssysdet}{\clssys_D}
167: \newcommand{\clssysito}{\clssys_I}
168: \newcommand{\clssysstrat}{\clssys_S}
169: 
170: \newcommand{\clsys}{\clssys(n,1,1)}
171: \newcommand{\clsysdet}{\clssysdet(n,1)}
172: \newcommand{\clsysito}{\clssysito(n,1,1)}
173: \newcommand{\clsysstrat}{\clssysstrat(n,1,1)}
174: 
175: \newcommand{\clmisys}{\clssys(n,m,k)}
176: \newcommand{\clmisysdet}{\clssysdet(n,m)}
177: \newcommand{\clmisysito}{\clssysito(n,m,k)}
178: \newcommand{\clmisysstrat}{\clssysstrat(n,m,k)}
179: 
180: \newcommand{\dsdef}{$\Theta_D = \dsysfgx \in \clsysdet$ }
181: \newcommand{\isdef}{$\Theta_I = \ssysfgx \in \clsysito$ }
182: \newcommand{\ssdef}{$\Theta_S = \ssysfgx \in \clsysstrat$ }
183: 
184: \newcommand{\dsdefmiuu}{$\Theta = \ssysfgx \in \clmisys$ }
185: \newcommand{\dsdefmi}{$\Theta_D = \dsysfgx \in \clmisysdet$ }
186: \newcommand{\isdefmi}{$\Theta_I = \ssysfgx \in \clmisysito$ }
187: \newcommand{\ssdefmi}{$\Theta_S = \ssysfgx \in \clmisysstrat$ }
188: 
189: \newcommand{\smooth}{C^{\infty}} \newcommand{\analytic}{C^{\omega(x)}}
190: \newcommand{\kdifferentiable}{C^{k}}
191: 
192: % operations
193: 
194: % differentiation ============
195: \newcommand{\parby}[2]{\frac{\partial #1}{\partial #2}}
196: \newcommand{\parbyx}[1]{\parby{#1}{x}}
197: \newcommand{\parbyz}[1]{\parby{#1}{z}}
198: 
199: \newcommand{\parbysec}[2]{\frac{\partial^2 #1}{\partial #2^2}}
200: 
201: \newcommand{\parbyxx}[3]{\frac{\partial^2 #1}{\partial #2 \partial
202:  #3}} \newcommand{\parbyxxx}[4]{\frac{\partial^3 #1}{\partial #2
203:  \partial #3 \partial #4}}
204: \newcommand{\parbyxxxx}[5]{\frac{\partial^4 #1}{\partial #2 \partial
205:  #3 \partial #4 \partial #5 }}
206: 
207: \newcommand{\lie}[2]{{\mathcal L}_{#1} {#2}}
208: \newcommand{\multilie}[3]{{\mathcal L}_{#1}^{#2} {#3}}
209: \newcommand{\biglie}[2]{\langle d{#2},{#1}\rangle}
210: %\newcommand{\biglie}[2]{\lie{#1}{#2}}
211: 
212: \newcommand{\ad}[3]{\operatorname{ad}_{{#1}}^{{#2}} {{#3}}}
213: \newcommand{\adfg}[1]{\ad{f}{{#1}}{g}}
214: \newcommand{\lighk}[1]{\lie{g}{\multilie{f}{#1}{h}}}
215: \newcommand{\distrofg}[1]{\left\{\ad{f}{i}{g},\,\iva{{#1}}\right\}}
216: \newcommand{\distroabafg}[1]{\left\{\ad{{\alterbar
217:  f}}{i}{g},\,\iva{{#1}}\right\}}
218: 
219: \newcommand{\compoit}{\compose T^{-1} (z)}
220: \newcommand{\travf}[1]{\parbyx{T} #1 \compoit}
221: \newcommand{\trav}[1]{\parbyx{T} #1}
222: 
223: % operators ==================
224: \newcommand{\flow}[1]{Fl^{#1}}
225: \newcommand{\trace}{{\operatorname{trace}}}
226: \newcommand{\grad}{{\operatorname{grad}}}
227: \newcommand{\kernel}{{\operatorname{kernel}}}
228: \newcommand{\annihilator}{{\operatorname{annihilator}}}
229: \newcommand{\rank}{{\operatorname{rank}}}
230: \newcommand{\abs}{{\operatorname{abs}}}
231: \newcommand{\sgn}{{\operatorname{sgn}}}
232: \newcommand{\sspan}{{\operatorname{span}}} \newcommand{\ito}[2]{P_#1
233:  #2} \newcommand{\corr}[2]{\operatorname{corr}_#1(#2)}
234: \newcommand{\Corr}[2]{\operatorname{Corr}_#1 #2}
235: 
236: \newcommand{\corrxx}[2]{{\frahalf \parby{#1}{#2} #1}}
237: \newcommand{\corrx}{\corrxx{\sigma}{x}}
238: 
239: \newcommand{\itox}{{\frahalf \sigma^2 \parbysec{T}{x} \compose T^{-1}
240:  (z)}} \newcommand{\itoxs}{{\frahalf \sigma^2 \parbysec{T}{x} }}
241: 
242: 
243: % coordinate transformation ======
244: \newcommand{\tantra}[1]{#1_\ast} \newcommand{\cotantra}[1]{#1^\ast}
245: 
246: \newcommand{\sct}[1]{{\mathcal T}_{#1}} \newcommand{\sctt}{\sct{T}}
247: 
248: \newcommand{\scti}[1]{\sct{#1}^{I}}
249: \newcommand{\scts}[1]{\sct{#1}^{S}}
250: 
251: \newcommand{\feedback}[2]{{\mathcal F}_{#1,#2}}
252: \newcommand{\feedbackab}{\feedback{{\alpha}}{{\beta}}}
253: 
254: \newcommand{\combinedz}{{\mathcal J}}
255: \newcommand{\combined}[3]{\combinedz_{#1,#2,#3}}
256: \newcommand{\combinedtab}{\combined{T}{{\alpha}}{{\beta}}}
257: 
258: 
259: % stochastic
260: \newcommand{\expected}[1]{{\cal E}\left\{#1\right\}}
261: \newcommand{\var}[1]{\operatorname{var}\left\{#1\right\}}
262: \newcommand{\cov}[1]{\operatorname{cov}\left\{#1\right\}}
263: 
264: \newcommand{\wrong}{^\#}
265: 
266: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
267: \newcommand{\tpDate}{2002/10/09}
268: \newcommand{\tpAuthor}{Ladislav Sl\'ade\v cek}
269: \newcommand{\tpeAddress}{\v R\'\i{}kovice 18, CZ 751 18, Czech Republic}
270: 
271: \newcommand{\tpEmail}{\hyreff{mailto:lsla@post.cz}{lsla@post.cz}}
272: \newcommand{\tpeTitle}{Exact Feedback Linearization of Stochastic
273: Control Systems}
274: % #3
275: \newcommand{\tpeAbstractN}{Abstract}
276: % #4
277: \newcommand{\tpeAbstract}{This paper studies exact linearization
278:  methods for stochastic SISO affine controlled dynamical systems. The
279:  systems are defined as vectorfield triplets in Euclidean space. The
280:  goal is to find, for a given nonlinear stochastic system, a
281:  combination of invertible transformations which transform the system
282:  into a controllable linear form. Of course, for most nonlinear
283:  systems  such transformation does not exist.
284: 
285:   We are focused on linearization by state coordinate transformation
286:  combined with feedback.  The difference between \Ito and Stratonovich
287:  systems is emphasized. Moreover, we define three types of linearity
288:  of stochastic systems --- $g$-linearity, $\sigma$-linearity, and
289:  $g\sigma$-linearity.
290:  
291:  Six variants of the stochastic exact linearization problem are
292:  studied. The most useful problem --- the \Ito-~$g\sigma$
293:  linearization is solved using the correcting term, which proved to be
294:  a very useful tool for \Ito systems.  The results are illustrated on
295:  a numerical example solved with help of symbolic algebra.  }
296: % #5
297: \newcommand{\tpeKeywordsN}{Keywords}
298: % #6
299: \newcommand{\tpeKeywords}{ exact linearization, feedback
300:   linearization, nonlinear dynamical system, \Ito integral,
301:   Stratonovich integral, correcting term 
302: {\bf MCS classification:} 93B18, 93E03}
303: 
304: 
305: 
306: % ============================================================
307: \title{\tpeTitle} \author{\tpAuthor}
308: 
309: \newcommand{\tp}[6]{ \vfill
310:  \begin{titlepage}
311:  \begin{center} 
312:  {~\\} \vspace{25mm}
313:  {\Large #1}\\
314:  \vspace{1cm}
315:  {\large \tpAuthor}\\
316:  {\tpEmail}\\
317:  {#2}\\
318:  
319:  \vspace{0.5cm}
320:  {\large \tpDate}\\
321:  \end{center}
322:  \begin{quote}\small
323:  \vspace{15mm}
324:  {\sc\bf #3:} #4\\
325:  {\sc\bf #5:} #6\\
326:  \end{quote}
327:  \end{titlepage}
328:  }
329:  
330:  \newcommand{\tpe}{ \tp {\tpeTitle} {\tpeAddress} {\tpeAbstractN}
331:  {\tpeAbstract} {\tpeKeywordsN} {\tpeKeywords} }
332: 
333:  
334: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
335: \tpe 
336: 
337: \tableofcontents
338: %%%%%%%%%%%%%%% Section %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
339: 
340: 
341: \section{Introduction}
342: \pagenumbering{arabic}\count1=0
343: \label{sec:intro}
344: 
345: The theory of exact linearization of deterministic dynamical systems
346: has been thoroughly studied since seventies. This paper attempts to
347: apply some of the results to the stochastic area. We emphasize the
348: exact linearization by state coordinate transformation combined with
349: feedback (further abbreviated as SFB linearization). Our main goal is
350: to identify the main difficulties of this approach and to consider
351: applicability of the methods known from the deterministic systems.
352: 
353: The task of SFB linearization is following: given a dynamical
354: systems~$\Theta$ we are looking for a combination of coordinate
355: transformation~$\sctt$ and feedback~$\feedbackab$ which will make the
356: resulting system~$\feedbackab \compose \sctt (\Theta)$ linear and
357: controllable. One can also define the feedback-less linearization by
358: coordinate transformation only (here abbreviated as SCT) or several
359: variants of the input-output linearization. These variants are not
360: considered here.
361: \begin{rem}
362: The notation for composition of mappings sometimes differs;
363: right-to-left convention is used here:~$f \compose g(x) \define
364: f(g(x))$.
365: \end{rem}
366: 
367: 
368: The subject of exact linearization of stochastic controlled dynamical
369: systems lies on the intersection of three branches of science:
370: differential geometry, control theory, and the theory of stochastic
371: processes. Each of them is very broad and it is virtually impossible
372: to cover all details of their combination. Hence it is necessary to
373: choose a minimalistic simplified model for our problem and to refrain
374: from most technical details. {\em We decided to represent dynamical
375: systems under investigation by triplets of smooth vectorfields and to
376: concentrate on transformation rules for these triplets\/}. The detailed
377: interpretation of the vectorfield systems (\ie solvability of
378: underlying differential equations, properties of flows and
379: trajectories) will be considered only on an informal, motivational level.
380: 
381: For simplicity, we shall confine all the definitions of geometrical
382: object to the Euclidean space; we will work in a fixed coordinate
383: system using explicit local coordinates, which may be considered to be
384: local coordinates of some manifold. This is mainly because we are
385: unable to capture all consequences  of the modern,
386: coordinate-free differential geometry to the stochastic calculus (see
387: \eg~\citet{kendall86},~\citet{malliavin},~\citet{emery89}). We believe
388: that this approach is quite satisfactory for the majority of practical
389: applications.
390: 
391: 
392: \subsection{Dynamical systems}
393: 
394: \begin{defn}
395: In this paper, a stochastic dynamical system~$\Theta \define
396:  \ssysfgx$ is defined to be a triplet of  smooth and bounded
397:  vectorfields~$f$,~$g$, and~$\sigma$ defined on an open
398:  neighborhood~$U$ of a point~$\origi{x} \in \RR^n$.  We usually
399:  call~$U\in \RR^n$ the \defbf{state space},~$f$ the \defbf{drift
400:  vectorfield},~$g$ the \defbf{control vectorfield}, and ~$\sigma$ the
401:  \defbf{dispersion vectorfield}.
402: \end{defn}
403: 
404: From now on, let's assume that all functions, vectorfields, forms, and
405: distributions are smooth and bounded on~$U$.
406: 
407: In this paper, we will study almost only SISO systems, but in the case
408: of stochastic MIMO systems with~$m$ control inputs and~$k$-dimensional noise the
409: symbols~$g$ and~$\sigma$ stand for~$n\times m$ ($n \times k$
410: respectively) matrix of smooth vectorfields having its rank equal
411: to~$m$ ($k$ respectively). The class of all
412: deterministic~$n$-dimensional dynamical systems with~$m$ inputs will
413: be called~$\clmisysdet$ and the class of stochastic systems
414: with~$k$-dimensional noise will be denoted with~$\clmisys$.
415: 
416: Similarly, autonomous deterministic dynamical system
417: corresponds to a single vectorfield and a controlled
418: deterministic dynamical system corresponds to a vectorfield pair. It
419: is obvious that this approach is limited to time invariant, affine
420: systems.
421: 
422: 
423: \begin{rem}
424:   The acronyms SISO and MIMO are used in the usual meaning even for
425:   systems without outputs, where the wording ``scalar-input''and
426:   ``vector-input'' will be appropriate. Stochastic systems with~$m =1$
427:   and~$k=1$ will be considered SISO.
428: \end{rem}
429: 
430: The definition may be interpreted as follows: there is a stochastic
431:  process~$x_t$ defined on~$\RR^n$ which is a strong solution of the
432:  stochastic differential equation $dx_t = f(x_t)\,dt + g(x_t) u(t)\,dt
433:  + \sigma(x_t)\,dw_t$, with initial condition~$\origi{x}$,
434:  where~$u(t)$ is a smooth function with bounded derivatives and~$w_t$
435:  is a one-dimensional Brownian motion. The differential $dw_t$ is just
436:  a notational shortcut for the stochastic integral.
437:  
438:  Details of the theory of stochastic processes are beyond the scope of
439:  this article. The reader is referred to \citet{wong84},
440:  \citet{oksendal}, \citet{sagemelsa}, \citet{malliavin},
441:  \citet{kendall86}, \citet{karatzas}; the text of \citet{kohlman} is
442:  freely available on the Internet.
443:  
444:  Theory of stochastic processes offers several alternative definitions
445:  of the stochastic integral, among them the \Ito integral and the
446:  Stratonovich integral; each of them is used to model different
447:  physical problems. Consequently there are two classes of differential
448:  equations and two alternative definitions of a stochastic dynamical
449:  system --- \Ito dynamical systems defined by \Ito integrals and
450:  Stratonovich systems defined by Stratonovich integrals.
451: 
452:  Serious differences between these integrals exists but from out point
453:  of view there is a single important one: {\em the rules for
454:  coordinate transformations of dynamical systems defined by \Ito
455:  stochastic integral are quite different from the transformation rules
456:  which are valid for Stratonovich systems\/}.
457:  
458:  The definition of the \Ito dynamical system used by us is formally
459:  equivalent to the definition of the Stratonovich system; the only
460:  difference will be in the corresponding coordinate transformation.
461: 
462: If necessary, \Ito and Stratonovich dynamical
463:  systems will be distinguished by a subscript: $\Theta_I \in
464:  \clmisysito$ and~$\Theta_S \in \clmisysstrat$.
465: 
466: \begin{rem}
467: In this paper we will use the adjectives {\em \Ito\/} and {\em
468:   Stratonovich} rather freely. For example we will speak of
469:   'Stratonovich linearization' instead of `exact linearization of
470:   stochastic dynamical system defined by Stratonovich integral'.
471: \end{rem}
472: 
473: 
474: 
475: \subsection{Transformations}
476: 
477: Furthermore, we will study two transformations of dynamical systems:
478: the coordinate transformation~$\sctt$ and the
479: feedback~$\feedbackab$. The definition of these transformation should
480: be in accord with their common interpretation. This can be illustrated
481: on the definition of the \defbf{coordinate transformation of a
482: deterministic dynamical system} $\sctt: \clmisysdet \to \clmisysdet$
483: which is induced by a diffeomorphism~$T \colon U \to \RR^n$ between two
484: coordinate systems on an open set~$U \subset \RR^n$.
485: The mapping~$\sctt$  is defined by:
486: \begin{align}
487:  \label{eq:83}
488: \sctt \dsys{f}{g}{x} \define  \left(
489: \tantra{T} f , \tantra{T} g ,T(U),T(\origi{x}) 
490: \right) 
491: .\end{align}
492: 
493: Recall that the symbol~$\tantra{T}$ stands for the contravariant
494: transformation~$(\tantra{T} f)_i = \ssum{j}{0}{n} f_j
495: \parby{T_i}{x_j}$. Moreover, we will require
496: that the coordinate transformation~$T$ preserves the equilibrium state
497: of the system \ie~$T(\origi{x}) = 0$.
498: 
499: The definition captures the contravariant transformation
500: rules for differential equations known from the basic calculus.
501: 
502: Note that the words ``coordinate transformation'' are used in two
503: different meanings: first as the
504: diffeomorphism~$T\colon U\to \RR^n$ between coordinates;
505: second as the mapping between systems~$\sctt: \clmisysdet \to
506: \clmisysdet$.
507: 
508: Coordinate transformation of stochastic systems distinguish between
509: \Ito and Stratonovich systems.  One of the major complications of the
510: linearization problems for \Ito systems is the second-order term in
511: the transformation rules for \Ito systems:
512: \begin{defn} \label{def:ctito}
513:  Let~$U\in \RR^n$ be an open set and let~$T\colon U\to \RR^n$ be a
514:  diffeomorphism from~$U$ to~$\RR^n$ with bounded
515:  first derivative on $U$ such that~$T(\origi{x})=0$. The mapping
516:  $\sctt\colon \clmisysito \to \clmisysito$ will be called a
517:  \defbf{coordinate transformation of an \Ito dynamical system} induced by
518:  diffeomorphism~$T$ if the systems $\Theta_1
519:  \define \ssys{f}{g}{\sigma}{x}$ and $\Theta_2 \define
520:  \left(\tilde f,\tilde g,\tilde \sigma,T(U),\origi{x}\right)$; $\Theta_2
521:  = \sctt\left(\Theta_1\right)$ are related by:
522:  \begin{align}
523:  \label{eq:86}
524:  \tilde f &= \tantra{T}f + \ito{\sigma}{T}  \\
525:  \label{eq:87}
526:  \tilde g_i &= \tantra{T}g_i \firstfor 1 \le i \le m\\
527:  \label{eq:89}
528:  \tilde \sigma_i &= \tantra{T}\sigma_i  \firstfor 1 \le i \le
529:  k
530: .\end{align}
531: \end{defn}
532: 
533: The symbol~$\ito{\sigma}{T}$ stands for the \defbf{\Ito term} which is
534: a second order linear operator defined by the following relation for
535: the~$m$-th component of~$\ito{\sigma}{T}$, $1\le m\le n$
536: \begin{equation}
537: \label{eq:16}
538: \ito{\sigma} T_m \define \frahalf \ssum{{i,j}}{1}{n}
539: \frac{\partial^2 T_m}{\partial x_i x_j} \ssum{l}{1}{k} \sigma_{il}
540: \sigma_{jl}
541: .\end{equation}
542: 
543: The transformation rules for Stratonovich
544: systems~$\sctt\colon \clmisysstrat \to \clmisysstrat$,
545: $(f,g,\sigma,U,\origi{x}) \mapsto (\tantra{T}f, \tantra{T}g, \tantra{T}\sigma,
546: T(U), T(\origi{x}))$ are 
547: equivalent to rules valid for the deterministic systems; only the
548: rule\eqref{eq:89} for the drift vectorfield must be added.
549: 
550: The difference between the coordinate transformation of \Ito and
551: Stratonovich systems should be emphasized: in the Stratonovich case
552: all the vectorfields transform contravariantly; on the other hand, in
553: the \Ito case, the \Ito term~$\ito{\sigma}{T}$ is added to the drift
554: vectorfield  of the resulting system.
555: 
556: \psfig{fig:introsfee}{Regular State Feedback}{introsfee} Another
557: important transformation of dynamical systems is the regular feedback
558: transformation. A feedback transformation is determined by two smooth
559: nonlinear functions $\alpha \colon \RR^n \to \RR^m$
560: and~$\beta \colon \RR^n \to \RR^m \times \RR^m$ defined on~$U$
561: with~$\beta $ nonsingular for every~$x \in U$ (see
562: Figure~\ref{fig:introsfee}). Usually, $\alpha $ is written as a column~$m \times
563: 1$ matrix; $\beta $ as a square $m \times m$ matrix.
564: 
565: \begin{defn} \label{def:feedback}
566:   Let~\dsdefmiuu be a stochastic dynamical system. A \defbf{regular
567:   state feedback} is the transformation $\feedbackab \colon \clmisys
568:   \to \clmisys~$, $ (f,g,\sigma,U,\origi{x}) \mapsto \ssys{f + g\alpha
569:   }{g\beta }{\sigma}{x} $.
570: \end{defn}
571: 
572: A new input variable~$v$ is introduced by the
573: relation~$u=\alpha +\beta v$.  Given the feedback~$\feedbackab$
574: with nonsingular~$\beta $, we can always construct an inverse
575: relation~$\feedback{a}{b} \define \feedbackab^{-1}$ such
576: that~$\feedbackab \compose \feedback{a}{b} = \feedback{a}{b} \compose
577: \feedbackab$ is the identity. The coefficients are related as
578: follows:~$\beta  = b^{-1}$, $\alpha  = - b^{-1}a$, and~$a = -\beta ^{-1}
579: \alpha $.
580: 
581: This definition of feedback transformation can be used also for
582: deterministic systems provided that the drift vectorfield~$\sigma$ is
583: assumed to be zero.
584: 
585: The symbol~$\combinedtab$ is used to indicate the combination of
586:  coordinate transformation with feedback~$\combinedtab \define 
587:  \feedbackab\compose\sctt$ can be interchanged.
588: 
589: \begin{rem}
590: \label{rem:orderinv}
591: Observe that the order of feedback and
592: coordinate transformation in the
593: composed transformation~$\combinedtab \define \feedbackab \compose
594: \sctt$
595: \begin{multline}
596:  \combinedtab = \sctt \compose \feedbackab \bigdsys{f}{g}{x} =
597:  \sctt \bigdsys{f+g\alpha }{g\beta }{x}\\
598:  = \bigdsys{\tantra{T} f + \tantra{T} g\alpha }{\tantra{T} g\beta }{z}
599:  = \bigdsys{\tantra{T} f + (\tantra{T} g) \alpha  }{ (\tantra{T} g)
600:  \beta  }{z} \\
601: = \feedback{\alpha '}{\beta '} \compose \sctt
602: .\end{multline}
603: The functions~$\alpha(z)'$, $\beta(z)'$ are equal to~$\alpha(x) $ 
604: and~$\beta(x) $ written in the $z$~ coordinates $\alpha(z) '=\alpha(x) 
605: \compose T^{-1}(z)$, $\beta(z) '=\beta(x) \compose T^{-1}(z)$. 
606: \end{rem}
607: 
608: 
609: %======== Subsection ================================================
610: \subsection{Linearity}
611: 
612: \label{sub:linearity}
613: 
614: The definition of linearity is straightforward in the deterministic
615: case. In contrast, the stochastic case is more complex, because there
616: are two ``input'' vectorfields and thereby several degrees of
617: linearity can be specified.
618: 
619: \begin{defn} \label{def:linear} The deterministic dynamical
620:   system~$\Theta_D = (f,g,U,0) \in \clmisysdet$ is \defbf{linear} if
621:   the vectorfield~$f$ is a linear mapping without no constant term and
622:   the vectorfields~$g_i$ are constant; that is they can be written as
623:   $f(x)= Ax$, $g(x)= B$ with~$A$ a square~$n \times n$ matrix and~$B$
624:   an~$n \times m$ matrix. The matrices must be constant on whole~$U$.
625: \end{defn}
626: \begin{defn} \label{def:linearsto}
627:   The stochastic dynamical system~$\Theta=(f,g,\sigma,U,0)$ is:
628: \begin{itemize}
629: \item \defbf{$g$-linear} if the mapping~$f(x) = Ax$ is linear without 
630:   constant term and~$g(x) = B$ is constant on~$U$.
631: \item \defbf{$\sigma$-linear} if the mapping~$f(x) = Ax$ is linear
632:   without constant term and~$\sigma(x) = S$ is constant on~$U$.
633: \item \defbf{$g\sigma$-linear} if it is both~$g$-linear and
634:   $\sigma$-linear.
635: \end{itemize}
636: The matrices~$A$ and~$B$ have the same dimensions as in
637: Definition~\ref{def:linear}; $S$ is an~$n \times k$ matrix.
638: \end{defn}
639: 
640: 
641: \begin{rem}
642:  The vectorfield~$g$ is \defbf{constant} on~$U$ if the value
643:  of~$g(x)$ is the same for every~$x \in U$. The vectorfield~$f$
644:  on~$U$ is \defbf{linear without constant term} if~$f(\origi{x})=0$
645:  and the superposition principle~$f(x_1+x_2)=f(x_1)+f(x_2)$
646:  holds for every~$x_1$, $x_2 \in U$.
647: \end{rem}
648: 
649: We study systems at equilibrium \ie we require that~$f(\origi{x})=0$
650: and that all transformations preserve the equilibrium:~$T(\origi{x})
651: =0$,~$\alpha (\origi{x})=0$, and~$\beta(\origi{x})$ is nonsingular.
652: The \Ito systems require an additional
653: condition~$f(\origi{x})+\corr{\sigma}{\origi{x}}=0$.  The non-equilibrium case can
654: be easily handled by extending the linear model with a constant term.
655: 
656: Moreover we require that the resulting linear systems are
657: controllable.  A controlled deterministic dynamical linear
658: system~$\Theta_D = (Ax,B,\RR^n,0) \in \clmisysdet$ is
659: \defbf{controllable} if its first $n$ repeated brackets form an
660: $n$-dimensional space
661: \begin{equation}
662:  \label{eq:68}
663:  \dim \left\{A^kB, 0\le k\le n-1 \right\} = n
664: .\end{equation}
665: Other definitions of controllability of linear
666: systems exist. For example Theorem 3.1 of~\citet{zhou} gives six
667: definitions with proofs of equivalence.
668: 
669: The controllability property deserves some attention in the stochastic
670: case. The linear stochastic dynamical system is characterized by two input
671: vectorfields $g(x)=B$ and $\sigma(x)=S$. 
672: \begin{enumerate}
673: \item The definition of the controllability for the control
674:   vectorfield~$g$ is identical to the deterministic case;
675:   \ie~\eqref{eq:68} must be satisfied. This property will be called
676:   \defbf{$g$-controllability}.
677: \item We will also define \defbf{$\sigma$-controllability} as the
678:   requirement that the repeated brackets $S,AS,A^2S,\dots,A^{n-1}S$
679:   form an $n$-dimensional space.
680: \item Finally, the linear system is \defbf{$g\sigma$-controllable} if
681: \begin{equation}
682: \label{eq:3}
683:  \dim \left( \left\{A^kB, 0\le k\le n-1 \right\} \bigcup \left\{A^kS, 0\le
684:      k\le n-1 \right\} \right)= n
685: .\end{equation}
686: \end{enumerate}
687: In this paper, we do not deal with the reachability, controllability,
688: accessibility, observability, and similar properties of nonlinear
689: systems.
690: 
691: 
692: \begin{defn} \label{def:itogssisosfb}
693:   Let~\dsdefmiuu be a dynamical system such that~$f(\origi{x}) =0$. We
694:   call the combination of a coordinate transformation~$\sctt$ and a
695:   regular feedback~$\feedbackab$ such that $T(\origi{x})=0$,
696:   $\alpha (\origi{x})=0$, and $\beta (\origi{x})$ is nonsingular the
697:   \defbf{linearizing transformation } of $\Theta$ if the
698:   transformation~$\feedbackab \compose \sctt $ converts~$\Theta$ into
699:   a~{\em controllable\/} linear system.
700:  
701: For stochastic system we distinguish:
702: \begin{romanenu}
703: \item \defbf{$g$-linearizing transformation} which transforms~$\Theta$ into
704:   a~$g$-linear and~$g$-controllable system
705: \item \defbf{$\sigma$-linearizing transformation} which transforms~$\Theta$ into
706:   a~$\sigma$-linear and~$\sigma$-controllable system 
707: \item \defbf{$g\sigma$-linearizing 
708:  transformation}  which transforms~$\Theta$ into
709:   a~$g\sigma$-linear and~$g$-controllable system 
710: 
711: Note that for $g\sigma$-linearization we
712: require~$g$-controllability. This is slightly stricter requirement
713: than~$g\sigma$ controllability but it should be naturally fulfilled by
714: the majority of practical control systems. This requirement cancels
715: many ``uncomfortable'' linear forms. Consider for example the system
716: with prefilter of Figure~\ref{fig:prefilter} which
717: is~$g\sigma$-controllable but~$g$-uncontrollable.
718: \end{romanenu}\psfig{fig:prefilter}{Dynamical System with
719: a Prefilter.}{prefilter}
720:  The system~$\Theta$ is \defbf{linearizable} if there exists 
721: linearizing transformation of~$\Theta$.
722: \end{defn}
723: 
724: 
725: 
726: \subsection{Computational Issues}
727: 
728: In most practical circumstances, computational issues are the limiting
729: factor of any application of differential geometric methods in
730: control.
731: 
732: The equations of exact linearization algorithms must be dealt in a
733: symbolic form. Even the simplest exact linearization problems are
734: extremely complex from the computational point of view.  Therefore,
735: the computer algebra tools are often employed. The results presented
736: in this paper were tested by the author on  few simulations of
737: control systems in the symbolic  system \Mathematica.
738: 
739: Of course, the computer algebra has apparently serious limitations and
740: drawbacks. Viability of the symbolic computational approach to the
741: problems of the nonlinear control is studied by \citet{jager95}. Some
742: very useful theoretical notes about the symbolic computation can be
743: found in~\citet{winkler}. Unfortunately, the limited scope
744: of this article does not allow deeper discussion of these subjects.
745: 
746: \subsection{Applications}
747: 
748: We propose, very briefly, two applications of the theory presented here: 
749: \begin{romanenu}
750: \item \defbf{Control ---}
751:  a dynamical systems~$\Theta$ obtained by exact
752: linearization will be controlled using the linear
753: feedback law:
754: \begin{equation}
755:  \label{eq:141}
756:  v = Kz + \kappa \nu
757: ,\end{equation}
758: where~$K$ is a row matrix of feedback gains,~$\kappa$ is an input
759: gain, $z$~is the state vector, $v$~is the original control input,
760:  and~$\nu$ is a new control input. 
761: 
762: Two approaches can be studied --- classical linear control methods and
763: the more sophisticated stochastic optimal control  approach studied
764: for example by \citet{oksendal}.
765: 
766: The~$g\sigma$-linear systems are natural candidates for such
767: approach because  the other linear forms leave certain
768: part of the resulting system nonlinear.
769: 
770: \item \defbf{Filtering} --- the filtering problem is probably the most useful
771: application of the theory of stochastic processes. We want
772: to give the best estimate to the state of a dynamical system defined
773: by the stochastic differential equation:
774: \begin{align}
775: \label{eq:118} dx_t = f(x_t)\,dt + \sigma_f(x_t)\,dw_tf;
776: \end{align} based on observations of the from:
777: \begin{align}
778: \label{eq:119}y_t = h(x_t) +
779: \sigma_h(x_t)\,dw_th .\end{align} $x_t$ is an $n$-dimensional stochastic process, $f$, $\sigma_f$,
780: $\sigma_h$ are smooth vectorfields and $h$ is a
781: smooth function.
782: 
783: It would be interesting to use exact linearization of the nonlinear
784: system to design an exact linear filter. Unfortunately, this idea has
785: no direct association with the linearization results presented below,
786: because it requires {\em output\/} exact linearization or
787: linearization of an autonomous system. Therefore, it would be helpful
788: to extend our results to these cases in the future.
789: 
790: 
791: \end{romanenu}
792: 
793: 
794: \subsection{Previous Work}
795: \label{ssub:laal}
796: 
797: The problem of SFB~$g$-linearization of SISO dynamical system defined in the
798:  \Ito formalism has been 
799: studied by~\citet{lahdhiri}.  The authors derive
800: equations corresponding to~\eqref{eq:50},~\eqref{eq:51} (eq 14, 15, 16
801: in~\citez{lahdhiri}). These equations are combined and then
802: reduced to a set of PDEs of a single
803: unknown function~$T_1$. Because there is no commuting relation
804: similar to~\eqref{eq:103} the equations contain partial derivatives
805: of~$T_1$ up to the~$2n$-th order (eq 23 in~\citez{lahdhiri}).
806: Next, the authors propose a lemma (Lemma 1) that identifies the
807: linearity conditions with non-singularity and involutivness
808: of~$\distrofg{n-2}$. Unfortunately, we disagree with this result.
809: 
810:  It can be easily verified that for~$\sigma=0$ this
811: statement does not correspond to the deterministic conditions
812: (Proposition~\ref{prop:d1sfb2}), because the deterministic case
813: requires non-singularity up to the~$(n-1)$-th bracket, not only up to
814: the~$(n-2)$-th one. Second, although the method of finding~$T_1$
815: was given (solving PDE), we do not think
816: that the existence of~$T_1$ was proved.
817: 
818: After this paper was finished, we discovered recent works
819: of~\citet{pan02} and~\citez{pan01}. In the article~\citez{pan01}
820: Pan defines and solves the problem of {\em feedback complete
821: linearization of stochastic nonlinear systems}. In our terminology,
822: this problem is equivalent to SFB MIMO input--output \Ito $g\sigma$
823: linearization which was not studied by us. 
824: 
825: In~\citez{pan02} Pan declares and proves so called \index{invariance
826:   under transformation rule}{\em invariance under transformation rule}
827:   which is exactly equivalent to our Theorem~\ref{prop:corr} which is
828:   probably the most important result of our paper.
829: 
830: Althougth the problems solved by Pan were slightly different, he uses
831: the same equivalence --- Theorem~\ref{prop:corr}.  This proves that
832: our conclusions about applicability of the \index{correcting
833: term}correcting term are perfectly valid.
834: 
835: In~\citez{pan02} Pan
836: consider three other \index{canonical form}canonical forms of
837: stochastic nonlinear systems, namely the \index{noise-prone strict
838: feedback form}noise-prone strict feedback form, \index{zero dynamics
839: canonical form}zero dynamics canonical form and \index{and observer
840: canonical form}observer canonical form also not studied by us. 
841: 
842: %%%%%%%%%%%%%%% Section %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
843: \section{Deterministic Case}
844: \label{sec:detcase}
845: 
846: In this section we recapitulate the results of the SFB and SCT exact linearization
847: theory for SISO systems. For detailed treatment and proofs we refer to
848: existing literature, above all  the
849: classical monographs of~\citet{isidori85} and~\citet{nijmeijer94}. For a
850: very readable introduction to the field we refer to the seventh
851: chapter of~\citet{vidyasagar93}. The books also contain extensive
852: bibliography. The monograph of~\citet{isidori85} builds mainly on the
853: concept of relative degree. In contrast we will emphasize the approach
854: of~\citet{vidyasagar93} because the method is  more suitable for the
855: stochastic case.
856: 
857: \subsection{Useful Relations}
858: 
859: The solution of the SFB linearization problem as presented
860: here  uses  the Leibniz rule
861: \begin{align}
862: \label{eq:63}
863: \lie{\lbrack f,g\rbrack }{\alpha } =
864: \lie{f}{(\lie{g}{\alpha })} - \lie{g}{ (\lie{f}{\alpha })}
865: \end{align} 
866: with~$f,g$ smooth vectorfields on~$U$; $\alpha\colon U\to \RR $ is
867: a smooth function. The recursive form of the Leibniz rule allows to
868: simplify the chains of differential equations for the
869: transformation~$T$. This can be expressed in the form of the following
870: statement:
871: 
872: For all $x \in U$, $k \ge 0$ these two sets of conditions are
873: equivalent:
874: 
875: \begin{align}
876: \label{eq:2}
877: \text{(i)}&\qquad&\lie{g}{\alpha } = \lie{g}{\lie{f}{\alpha }} = 
878: \cdots = \lie{g}{\multilie{f}{k}{\alpha }}=0 \\
879: \label{eq:10}
880: \text{(ii)}&&\lie{g}{\alpha } = \lie{\adfg{}}{\alpha } = \cdots =
881: \lie{\adfg{k}}{\alpha }=0 .\end{align}
882: 
883: Recall that the symbol~$\lie{f}{h}$ stands for the Lie derivative 
884:  defined by~$\lie{f}{h} = \langle f,\grad\,h \rangle = \ssumi{n} f_i(x)
885:  \parby{}{x_i} h(x)$. Higher order Lie derivatives
886:  can be defined recursively
887: $\multilie{f}{0}{h} = h$, $\multilie{f}{k+1}{h} =
888:  \lie{f}{\multilie{f}{k}{h}}$ for~$k \ge 0$. The Lie Bracket is defined as
889: ~$[f,g] \define \parbyx{g}f-\parbyx{f}g$; there is also a recursive
890:  definition:
891: \begin{equation}
892:  \ad{f}{0}{g} \define g; \quad \ad{f}{k+1}{g} \define
893:  \left[f,\ad{f}{k}{g}\right] \quad \text{for } k \ge 0 
894: .\end{equation}
895: 
896: 
897: Another very important result of the differential geometry is
898: invariance of the Lie bracket under the tangent
899: transformation~$\tantra{T}$ 
900: (see~\citet{nijmeijer94} Proposition 2.30 \pg~50):
901: 
902: Let~$T\colon U\to \RR^n$ be a diffeomorphic coordinate transformation, and~$f$ and~$g$
903: be smooth vectorfields. Then
904: \begin{align}
905:  \label{eq:103}
906:  \tantra{T} [f,g] = [\tantra{T} f, \tantra{T} g] .\end{align}
907: 
908: \subsection{SFB Linearization}
909: 
910: Every controllable linear system may be, by a linear coordinate
911: transformation, transformed to the controllable canonical form
912: (\citet{kalman}). Furthermore, this controllable canonical form can
913: be always transformed into the integrator chain by a linear
914: regular feedback.  Therefore, the integrator chain is a canonical
915: form for all feedback linearizable systems. (See \citet{vidyasagar93}
916: section 7.4). Consequently, the  equations of the integrator chain can be compared
917: with the equations of the nonlinear systems and the following
918: proposition can be proved:
919: 
920: \begin{prop} \label{prop:d1sfb1a}
921:  There is a SFB linearizing transformation~$\combinedtab$ of a SISO
922:  deterministic dynamical system~\dsdef into a controllable linear
923:  system if and only if there is a solution~$T_1, T_2, \dots,
924:  T_n\RntoR$ to the set of partial differential equations
925:  defined on~$U$
926: \begin{alignat}{2}
927: \label{eq:150}
928: \lie{f}{}{T_i} &= T_{i+1} \firstfor 1 \le i \le {n-1}\\
929: \label{eq:151}
930: \lie{g}{}{T_i} &= 0 \nextfor 1 \le i \le {n-1}\\
931: \label{eq:152}
932: \lie{g}{}{T_n} &\ne 0 .\end{alignat} Then the
933: feedback is defined as follows:
934: \begin{align}
935: \label{eq:57}
936: \alpha &=-\frac{\lie{f}{T_n}}{\lie{g}{{T_n}}}\qquad\qquad
937: \beta =\frac{1}{\lie{g}{T_n}} .\end{align}
938: \end{prop}
939: \begin{proof}
940: See \citet{vidyasagar93} equations 7.4.20--21.
941: \end{proof}
942: 
943: 
944: \begin{prop} \label{prop:d1sfb1}
945:  The SFB linearizing transformation~$\combinedtab$ of a SISO
946:  deterministic dynamical system~\dsdef into a controllable linear
947:  system  exists if and only if there is
948:  a solution~$\lambda \RntoR$ to the set of partial differential
949:  equations:
950: \begin{alignat}3
951: \label{eq:8}
952: \biglie{\adfg{i}}{\lambda }&=0\firstfor{\iva{n-2}}\\
953: \label{eq:9}
954: \biglie{\adfg{{n-1}}}{\lambda }&\not=0 .\end{alignat} The
955: linearizing transformation~$T(x)$ is given by:
956: \begin{alignat}3
957: \label{eq:11}
958: T_{i}&=\multilie{f}{{i-1}}\lambda  \firstfor 1\le i\le n\\
959: \alpha &=\frac{-\multilie{f}{{n}}\lambda 
960: }{\lie{g}{}\multilie{f}{{n-1}}\lambda }&\qquad\qquad&
961: \label{eq:12}
962: \beta =\frac{1}{\lie{g}{}\multilie{f}{{n-1}}\lambda }
963: .\end{alignat}
964: \begin{proof}
965: See \citet{vidyasagar93} equations 7.4.23--33 and \citet{nijmeijer94} Corollary 6.16.
966: \end{proof}
967: 
968: \end{prop}
969: Finally, the  geometrical conditions for the
970: existence of the linearizing transformation  are studied.
971: 
972: \begin{theo} \label{prop:d1sfb2}
973:  A deterministic SFB linearizing transformation of~\dsdef into a
974:  controllable linear system exists if and only if the
975:  distribution~$\Delta_{n} \define \sspan\left\{{\adfg{i},
976:  \iva{n-1}}\right\}$ is nonsingular on~$U$ and the
977:  distribution~$\Delta_{n-1} \define \sspan\left\{{\adfg{i},
978:  \iva{n-2}}\right\}$ is involutive on~$U$.
979: \end{theo}
980: \begin{proof}
981: See \citet{nijmeijer94} Corollary 6.17, \citet{vidyasagar93} Theorem
982: 7.4.16, \citet{isidori89} Theorem 4.2.3 .
983: \end{proof}
984: 
985: \subsection{SCT Linearization}
986: 
987: \begin{theo} \label{prop:s1ctt2t}
988:   There is a SCT~linearizing transformation~$\sctt$ of a deterministic
989:   MIMO system~\dsdefmi into a controllable linear system if  and only
990:   if there exists a reordering of the vectorfields~$g_1 \dots g_m$ and
991:   an~$m$-tuple of integers~$\kappa_1 \le \kappa_2 \le \dots \kappa_m$
992:   with~$\ssumi{m} \kappa_i = n$ called the \defbf{controllability
993:     indexes} such that the following conditions are satisfied for
994:   all~$x \in U$:
995:  \begin{align}
996:  \label{eq:43}
997:  \text{(i)}&\qquad\dim\left(\sspan\left\{( \ad{f}{j}{g_i}(x),
998:  \iva{m}, \jva{\kappa_i-1})\right\}\right) = n\\
999:  \label{eq:44}
1000:  \text{(ii)}&\qquad[\ad{f}{k}{g_i},\ad{f}{l}{g_j}]=0
1001:  \qquad\text{for}\qquad 0 \le k+l \le \kappa_i+\kappa_j-1,\, 0 \le
1002:  i,j \le m . \end{align}
1003: \end{theo}
1004: \begin{proof}
1005: See  \citet{nijmeijer94} Theorem 5.3 and Corollary 5.6.
1006: \end{proof}
1007: 
1008: The following corollary can be verified for SISO systems:
1009: 
1010: \begin{coroll} \label{prop:s1ctt2a}
1011:  For a SISO system with~$m=1$ the
1012:  condition (ii) of~\ref{prop:s1ctt2t} can be simplified as
1013:  follows:
1014: \begin{align}
1015: \label{eq:45}
1016: [g,ad^l_f g] = 0, \quad l = 1, 3, 5, \dots, 2n-1, \forall x \in U
1017: .\end{align}
1018: \end{coroll}
1019: \begin{proof}
1020: See  \citet{nijmeijer94} Corollary 5.6 and the text which follows.
1021: \end{proof}
1022: 
1023: 
1024: 
1025: 
1026: %%%%%%%%%%%%%%% Section %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1027: \section{Transformations of \Ito Dynamical Systems}
1028: \label{sec:stotrans}
1029: 
1030: The transformation rules of \Ito systems are motivated by the  \Ito
1031: differential rule (see \eg~\citet{wong84} Section 3.3), which defines
1032: the influence of nonlinear coordinate transformations on  \Ito stochastic
1033: processes.
1034: 
1035: The \Ito differential rule applies to the situation where a scalar
1036: valued stochastic process~$x_t$ defined by a stochastic differential
1037: equation~$dx_t = f(x_t)\,dt + \sigma(x_t)\,dw_t$ with~$f\RtoR$ and~$\sigma\RtoR$
1038: smooth real functions and~$w_t$  a Brownian motion, is transformed
1039: by a a diffeomorphic coordinate transformation~$T\colon \RR \to \RR$.
1040: Then the stochastic process~$z_t\define T(x_t)$ exists and is an \Ito
1041: process. Further, the process~$z_t$ is the solution of the stochastic
1042: differential equation
1043: \begin{equation} \label{eq:126}
1044: dz_t = \parbyx{T} {f(x_t)}\,dt + \parbyx{T} {\sigma(x_t)}\,dw_t + \itoxs\,dt
1045: .\end{equation}
1046: All details together with a proof are
1047: available for example in~\citet{karatzas}.
1048: 
1049: The \Ito rule can be also derived for the multidimensional case: for
1050: the~$m$-th component of an~$n$-dimensional stochastic process the \Ito
1051: rule can be expressed as follows:
1052: \begin{multline}
1053:  dz_m = \ssumi{n} \parby{T_m}{x_i} f_i\,dt + \frahalf \ssum{i}{1}{n}
1054:  \ssum{j}{1}{k} \parby{T_m}{x_i} \sigma_{ij}\,dw_j + \frahalf
1055:  \ssum{{i,j}}{1}{n} \frac{\partial^2 T_m}{\partial x_i x_j}
1056:  \ssum{l}{1}{k} \sigma_{il}
1057:  \sigma_{jl}\,dt \\
1058:  =\lie{f}T_m\,dt + \ssum{j}{1}{k} \lie{\sigma_j}{T_m}\,dw_j +
1059:  \ito{\sigma}T_m\,dt
1060: .\end{multline}
1061: 
1062: For the most common case with scalar noise~$k=1$ the equation
1063:  can be further simplified to:
1064: \begin{equation}
1065: dz_m =\lie{f}{T_m}\,dt + \lie{\sigma}{T_m}\,dw +
1066: \frahalf \ssum{{i,j}}{1}{n}
1067: \parbyxx{T_m}{x_i}{x_j} \sigma_{i} \sigma_{j}\,dt 
1068: .\end{equation}
1069: The operator~$\ito{\sigma} T_m$ is sometimes written using
1070: matrix notation as:
1071: \begin{equation}
1072: \ito{\sigma} T_m = \frahalf \trace \left(\sigma^T\sigma \parbysec{T_m}{x}\right)
1073: .\end{equation}
1074: Generally, $\ito{\sigma}$ vanishes for linear~$T$ or zero~$\sigma$. 
1075: 
1076: 
1077: %======== Subsection ================================================
1078: \subsection{The Correcting Term}
1079: \label{sub:corr}
1080: 
1081: 
1082: In this section we introduce an extremely useful equivalence
1083: between \Ito and Stratonovich systems, which allows to use some
1084: Stratonovich linearization techniques for \Ito problems. The
1085: motivation is following: let~\isdefmi be an  \Ito system. We
1086: are looking for a Stratonovich system $\Theta_S = \ssys{\alterbar
1087:  f}{\alterbar g}{\alterbar \sigma}{x}$ such that the trajectories
1088: of~$\Theta_I$ and~$\Theta_S$ are identical. The aim is to find
1089: equations relating the quantities~$\alterbar f$,~$\alterbar
1090: g$, and~$\alterbar \sigma$ with~$f$,~$g$, and~$\sigma$.
1091: 
1092: \begin{defn} \label{def:corr}
1093:  Let~$\Theta_{1I} = \ssys{f}{g}{\sigma}{x} \in \clmisysito$ be
1094:  an~$n$-dimensional \Ito dynamical
1095:  system with~$k$-dimensional Brownian
1096:  motion~$w$. The vectorfield $\corr{\sigma}{x}$ whose ~$r$-th
1097:  coordinate is equal to
1098: \begin{alignat}3 
1099: \label{eq:5}
1100:  (\corr{\sigma}{x})_r &= -\frahalf \ssum{i}{1}{n} \ssum{j}{1}{k}
1101:  \parby{{\sigma_{rj}}}{{x_i}} \sigma_{ij} \firstfor 1 \le r \le n
1102: \end{alignat}
1103: is called the \defbf{correcting term}.
1104: Note that the derivative is always evaluated in the corresponding
1105: coordinate system.
1106: Further, let us define the \defbf{correcting mapping} $\Corr{\sigma} \colon \clmisysito \to \clmisysstrat$ by
1107: \begin{align}
1108: \label{eq:21}
1109:  \Corr{\sigma} (f,g,\sigma,U,\origi{x}) \define
1110:  (f+\corr{\sigma}{x},g,\sigma,U,\origi{x}) .\end{align}
1111:  \end{defn}
1112:  The general treatment of the subject can be found for example
1113: in~\citet{wong84} \pg~160 or in \citet{sagemelsa}.  The following
1114: theorem describes the
1115: behavior of the correcting term under the coordinate transformation.
1116: \begin{theo} \label{prop:corr}
1117:   Let~\isdef be a one-dimensional \Ito dynamical system. Let $T$ be a
1118:   diffeomorphism defined on~$U$ and the symbols~$\scti{T}$ and
1119:   $\scts{T}$ denote a \Ito coordinate transformation and a
1120:   Stratonovich coordinate transformation induced by the same
1121:   diffeomorphism~$T$ and~$\tilde\sigma = \tantra{T}{\sigma}$.  Then the following diagram commutes:
1122: \begin{equation}
1123: \begin{CD}
1124: \label{eq:4}
1125: \Theta_{1I} @>\scti{T}>> \Theta_{2I}\\
1126: @V{\Corr{\sigma}}VV @AA{\Corr{{\tantra{T} \sigma}}^{-1}}A\\
1127: \Theta_{1S} @>\scts{T}>> \Theta_{2S}\\
1128: \end{CD}
1129: .\end{equation}
1130: In other words:
1131:  \begin{align}
1132:  \label{eq:19}
1133:   \scti{T} &= {\Corr{{\tilde\sigma}}{}}^{-1} \compose \scts{T} 
1134:   \compose \Corr{\sigma}{} \quad\text{and} \\
1135:  \label{eq:20}
1136:   \scts{T} &=  {\Corr{\sigma}{}}^{-1}  \compose \scti{T}
1137:   \compose \Corr{{\tilde\sigma}}{}  
1138:  . \end{align}
1139: 
1140: The notation~$\Corr{{\sigma}}{}^{-1}$ is used to denote the
1141: inverse mapping 
1142: \begin{align}
1143: \label{eq:23}
1144:  {\Corr{\sigma}{}}^{-1} (f,g,\sigma,U,\origi{x}) \define
1145:  (f-\corr{\sigma}{x},g,\sigma,U,\origi{x}) 
1146: .\end{align}
1147: \end{theo}
1148: 
1149: \begin{proof}
1150:  The correcting term~$\corr{\sigma}{x}\RntoRn$
1151:  is equal to
1152: \begin{equation}
1153: \corr{\sigma}{x} = -\corrx
1154: .\end{equation}
1155: The transformations specified in the diagram~\eqref{eq:4} will be
1156: evaluated in the following order:
1157: \begin{equation}
1158: \begin{CD}
1159: \label{eq:6}
1160: \Theta_{1I} @>({\operatorname a})>> \Theta_{2I} \\
1161: @V({\operatorname b})VV \\
1162: \Theta_{1S} @>({\operatorname c})>> \Theta_{2S}
1163: @>({\operatorname d})>> \Theta_{3I}\\
1164: \end{CD}
1165: \end{equation}
1166: We want to prove the equivalence of~$\Theta_{2I}$ and~$\Theta_{3I}$.
1167: The symbols (a) and (c) denote \Ito
1168: coordinate transformations; the
1169: symbols (b) (d) stand for the correcting
1170: mapping and its inverse. Note that the systems~$\Theta_{2I}$,
1171: $\Theta_{2S}$, and $\Theta_{3I}$ are defined in the~$z$-coordinate
1172: systems. Further, let
1173: \begin{align}
1174:  \label{eq:243}
1175:  \tilde\sigma&\define\trav{\sigma}\\
1176:  \kappa&\define \tantra{T} \corr{\sigma}{x} = -\trav{(\corrx)}\\
1177:  \origi{z}&\define T(\origi{x}) .\end{align} Then
1178: \begin{align}
1179:  \text{(a)}&\qquad \Theta_{2I} =\bigssys{\trav{f} + \itoxs }{
1180:  \trav{g}}{ \tilde\sigma }{z}\\
1181:  \text{(b)}&\qquad \Theta_{1S} =\bigssys{f - \corrx }{ g }{
1182:  \sigma }{x} \\
1183:  \text{(c)}&\qquad \Theta_{2S} =\bigssys{\trav{\left(f-\corrx\right)}
1184:  }{ \trav{g} }{
1185:  \trav{\sigma} }{z}\\
1186:  \text{(d)}&\qquad \Theta_{3I} =\bigssys{\trav{f} + \kappa + \frahalf
1187:  \parbyz{\tilde\sigma}{\tilde\sigma} }{ \trav{g} }{ \tilde\sigma
1188:  }{z}
1189: .\end{align}
1190: 
1191: All the terms in (a) are equivalent to the respective terms in (d)
1192: except for the drift terms containing functions of~$\sigma$. Therefore,
1193: we continue comparing these terms only. For (a):
1194: 
1195: \begin{align}
1196:  \label{eq:37}
1197:  L &\define \itoxs .\end{align} For (d):
1198: \begin{multline}
1199: \label{eq:38}
1200: R \define \kappa + \corrxx{{\tilde\sigma}}{z} = \kappa + \frahalf
1201: \parbyz{} \left( \parbyx{T} \sigma \right) \parbyx{T}\sigma = \kappa +
1202: \frahalf \parbyz{x} \parbyx{} \left( \parbyx{T} \sigma \right)
1203: \parbyx{T}\sigma = \\
1204: \kappa + \frahalf \left( \parbysec{T}{x}\sigma +
1205:  \parbyx{T}\parbyx{\sigma} \right) \sigma = \kappa + \frahalf
1206: \parbysec{T}{x}\sigma^2 - \kappa= \itoxs .\end{multline} Thus~$L=R$
1207:  and $\Theta_{2I} = \Theta_{3I}$.
1208: \end{proof}
1209: 
1210: Theorem~\ref{prop:corr} is valid also for combined
1211:  transformations:
1212: \begin{coroll} \label{prop:corrcorol}
1213:   Let~\isdef, $T$, $\scti{T}$, and
1214:   $\scts{T}$ have the same meaning as in Theorem~\ref{prop:corr}. 
1215:  Then the following diagram commutes for arbitrary regular feedback~$\feedbackab$:
1216: \begin{equation}
1217: \begin{CD}
1218: \label{eq:4000}
1219: \Theta_{1I} @>\scti{T}>> \Theta_{2I} @>\feedbackab>> \Theta_{4I}\\
1220: @V{\Corr{\sigma}}VV @AA{\Corr{{\tantra{T} \sigma}}^{-1}}A @AA{\Corr{{\tantra{T} \sigma}}^{-1}}A\\
1221: \Theta_{1S} @>\scts{T}>> \Theta_{2S}  @>\feedbackab>> \Theta_{4S}\\
1222: \end{CD}
1223: .\end{equation}
1224: \end{coroll}
1225: 
1226: \begin{proof}
1227:   We want to prove equivalence of~$\Theta_{4I}$
1228:   and~${\Corr{{\tantra{T} \sigma}}^{-1}} \Theta_{4S}$. 
1229:   
1230:   The control and dispersion vectorfields of~$\Theta_{4I}$
1231:   and~$\Theta_{4S}$ are identical and they are not influenced by the
1232:   correcting mapping.
1233: 
1234:   Using the notation of Theorem~\ref{prop:corr} we can express the
1235:   drift term of~$\Theta_{4I}$ as~$\tantra{T}f + L + g\alpha $. The
1236:   drift term of~${\Corr{{\tantra{T} \sigma}}^{-1}} \Theta_{4S}$ 
1237:   is~$\tantra{T}f + R + g\alpha  $. The effect of feedback is purely
1238:   additive and both the systems are equal.
1239: \end{proof}
1240: 
1241: 
1242: 
1243: At first glance the correcting term is rather
1244: surprising. How can the second derivative of~$T$ in (\ref{eq:37}) be
1245: compensated by the correcting term, which does
1246: not contain the~$T$ at all? The answer is quite simple: the second
1247: derivative is hidden in the correcting term
1248: implicitly because the correcting term depends
1249: on the coordinate system in which the system~$\Theta_{1I}$ is defined.
1250: The derivative~$\parbyx{\sigma}$ contained in the
1251: correcting term is always taken in the
1252: appropriate coordinate system. To emphasize the dependence of the
1253: correcting term on the coordinate system, we
1254: will never omit the independent variable (\eg $x$ or $z$) from the
1255: symbol~$\corr{\sigma}{x}$.
1256: 
1257:  Proposition~\ref{prop:corr} is valid for general multidimensional
1258:  systems~\isdefmi; the proof is purely mechanical and is not presented
1259:  here.
1260: 
1261: Let us now turn our attention to several special cases of the
1262: correcting mapping.
1263: \begin{coroll} \label{prop:corrvectr}
1264:  Let~$\Theta_I = \ssysfgx \in \clssysito(n,m,1)$ be
1265:  an~$n$-dimensional stochastic dynamical system with an
1266:  one-dimensional Brownian motion~$w$.
1267:  The~$r$-th coordinate of the correcting
1268:  term~$(\corr{\sigma}{x})_r$ is equal to
1269: \begin{equation}
1270: \label{eq:214}
1271: (\corr{\sigma}{x})_r = \frahalf \ssum{i}{1}{n}
1272: \parby{{\sigma_{r}}}{{x_i}} \sigma_{i} = \frahalf
1273: \lie{\sigma}{\sigma_r}
1274: \rfor{1\le r\le n}
1275: .\end{equation}
1276: \end{coroll}
1277: \begin{proof}
1278:  Substitute~$k = 1$ into~\eqref{eq:5}.
1279: \end{proof}
1280: \begin{coroll}
1281:  \label{prop:ic1}
1282:  For systems with one-dimensional noise ($k=1$) define the matrix
1283:  valued \Ito term~$\ito{\sigma}{T}$ for~$T\RntoRn$
1284:  with components~$T_i$, $1\le i\le n$, as a column $n\times 1$
1285:  matrix $\ito{\sigma}{T} \define \left[\ito{\sigma}{T_1},
1286:  \ito{\sigma}{T_2}, \dots,\ito{\sigma}{T_n}\right]^T$.
1287:  
1288:  Then the relations~\eqref{eq:214} can be expressed as
1289: \begin{align}
1290: \label{eq:215}
1291: \ito{\sigma}{T} = \tantra{T} \left(\corr{\sigma}{x}\right) -
1292: \corr{{\tilde\sigma}}{z} .\end{align}
1293: \end{coroll}
1294: \begin{proof}
1295:  The proof is almost identical to that of
1296:  the multidimensional variant of Corollary~\eqref{prop:corr}.The symbols can be
1297:  identified as follows:
1298:  \begin{alignat}2
1299:  L_i &= \left(\ito{\sigma}{T}\right)_i \firstfor 1\le i\le n \\
1300:  \kappa_i &= \left( \tantra{T} \left(\corr{\sigma}{x}\right)\right)_i \\
1301:  R_i &= \left(\tantra{T} \left(\corr{\sigma}{x}\right) \right)_i -
1302:  \left(\corr{{\tilde\sigma}}{z}\right)_i .\end{alignat}
1303: \end{proof}
1304: \begin{coroll} 
1305:  \label{prop:ic2}
1306:  Assume that the conditions of Proposition~\ref{prop:ic1} hold. The
1307:  relation~\eqref{eq:215} can be written as:
1308:  \begin{alignat}3
1309:  \label{eq:140}
1310:  \frahalf \lie{\sigma}{\lie{\sigma}{T_i}} &=
1311:  \ito{\sigma}{T_i} - \lie{\corr{\sigma}{x}}{T_i} \firstfor 1
1312:  \le i \le n
1313: . \end{alignat} 
1314: \end{coroll}
1315: \begin{proof}
1316:  The formula can be expressed as:
1317:  \begin{multline}
1318:  \label{eq:139}
1319:  \frahalf \lie{\sigma}{{\lie{\sigma}{T_i}}} = \frahalf
1320:  \lie{\sigma}{ \left( \ssum{j}{1}{n} \parby{T_i}{x_j} \sigma_j
1321:  \right) } = \frahalf \ssum{l}{1}{n} \sigma_k \parby{}{x_k}
1322:  \left( \ssum{j}{1}{n} \parby{T_i}{x_j} \sigma_j
1323:  \right) =\\
1324:  \frahalf \ssum{k,j}{1}{n} \left( \sigma_j\sigma_k
1325:  \parbyxx{T_i}{x_k}{x_j} +\parby{T_i}{x_j}
1326:  \parby{\sigma_j}{x_k} \sigma_k \right) = \ito{\sigma}{T_i} -
1327:  \lie{{\corr{\sigma}{x}}}{T_i} . \end{multline}
1328: \end{proof}
1329: 
1330: %======== Subsection ================================================
1331: \subsection{Composition of Coordinate Transformations of \Ito Systems}
1332: \label{sub:stgroup}
1333: 
1334: The set of all deterministic coordinate transformations~$\sctt$
1335: together with composition~$\sct{RS} \define \sct{S} \compose \sct{R}$
1336: forms a group. Obviously, this fact is a straightforward result of the
1337: behavior of the contravariant transformation and therefore an
1338: analogous statement must hold for Stratonovich systems.  Surprisingly,
1339: this is valid also for \Ito systems as will be shown here. This has an
1340: important consequence: we may always find the inverse transformation
1341: to a given coordinate transformation of \Ito systems. We will prove
1342: the following assertion:
1343: \begin{theo}
1344:  \label{prop:compogr}
1345: 
1346:  Let~$\scti{R}$, $\scti{S} \in \clsysito$ be
1347:  coordinate transformations of
1348:  one-dimensional \Ito systems  induced by
1349:  diffeomorphisms~$R$ and~$S$. Then
1350: \begin{equation}
1351:  \label{eq:123}
1352:  \scti{S} \compose \scti{R} = \scti{S \compose R} 
1353: .\end{equation}
1354: \end{theo}
1355: \begin{proof}
1356:  We will transform the system in two
1357:  different ways and show that the results are equal.
1358: \begin{enumerate}
1359: \item In the first method the system~$A=(0,0,a,U,\origi{x})$ which
1360:   corresponds to a differential equation~$dx = a(x)\,dw$ will be
1361:   transformed twice:
1362: \begin{enumerate} \item first,  by~$y=R(x)$ to~$y$
1363:  coordinates \item and then the result~$B=(g,0,b,U,\origi{x})$ which
1364:   corresponds to~$dy = g(y)\,dt + b(y)\,dw$
1365:   by~$z=S(y)$ to~$z$ coordinates. 
1366: \end{enumerate}
1367: \item The other method
1368:  transforms the system~$A$ only once by~$z=T(x) = S(R(x)) = (S
1369:  \compose R)(x)$.
1370: \end{enumerate}
1371: Without loss of generality, the equation~$dx = a(x)\,dw$ is assumed
1372:  to have no drift term because the drift term transforms in the
1373:  contravariant fashion. The derivatives will be
1374:  denoted by~$\parbyx{T(x)} \definer T'$,~$\parbyx{R(x)} \definer
1375:  R'$,~$\parby{S(y)}{y} \definer S'$ and similarly for~$T''$,
1376:  $R''$ and $S''$. Note that the prime is always used to denote
1377:  derivatives by the argument of the function.
1378:  
1379:  The transformation by~$R$ gives:
1380:  \begin{align}
1381:  \label{eq:64} 
1382:  dx &= a \, dw\\
1383:  \label{eq:266} 
1384:  dy &= R' a\,dw + \frahalf a^2 R'' \,dt .\end{align} Thus the
1385:  coefficients of the second SDE are defined
1386:  by
1387: \begin{align}
1388:  \label{eq:267}
1389:  b(y) &\define (R'a) \compose R^{-1}(y)\\
1390:  \label{eq:268}
1391:  g(y) &\define (\frahalf a^2 R'') \compose R^{-1}(y) .\end{align}
1392: 
1393: The second transformation (by~$S$) gives
1394: \begin{align}
1395: \label{eq:269}
1396: dz &= \left( S' g + \frahalf b^2 S'' \right)\,dt + S' b \,dw =\\
1397: \label{eq:270}
1398: &= \left(\frahalf a^2 S' R'' + \frahalf (R')^2 a^2 S'' \right) \,dt +
1399: S' R' a \,dw = \\
1400: \label{eq:272}
1401: &= \frahalf a^2 \left(S'R'' + (R')^2 S'' \right)\,dt + S'R' a\,dw =\\
1402: \label{eq:273}
1403: &= \frahalf a^2 T''\,dt + T' a\,dw .\end{align} The last equality
1404: follows from the fact that~$T''=S''(R')^2+S'R''$.
1405: \end{proof}
1406: One can verify the multidimensional case in the same
1407: spirit. 
1408: %======== Subsection ================================================
1409: \subsection{Invariants}
1410: \label{sub:stoinvar}
1411: 
1412: 
1413: In the deterministic case, some useful propositions about the
1414: invariant properties for example the Leibniz rule~\eqref{eq:63} and
1415: the relation~\eqref{eq:2} were employed. 
1416: 
1417: Unfortunately, we have not found any analogy for the \Ito systems yet.
1418: To point out the main complications, we will analyze the \Ito
1419: equivalent of the Leibniz rule~\eqref{eq:63}, which is essential for
1420: reducing the order of partial differential equations in the
1421: deterministic exact linearization.
1422: 
1423: If the Lie derivative~$\lie{g}{}$ is interpreted
1424: as a general first order operator
1425: \begin{align}
1426:  \lie{g}{} = \ssumi{n}g_i \parby{}{x_i}
1427: \end{align}
1428: then the commutator of two such first order
1429: operators $\lie{f}{\lie{g}{}} - \lie{g}{\lie{f}{}}$ is also a first
1430: order operator~$\lie{\lbrack f,g\rbrack }{}$ (see~\eqref{eq:63}).
1431: 
1432: Similarly, define the general second order operator as
1433: \begin{align}
1434: \label{eq:17}
1435: O(g,G) &\define \ssumi{n}g_i \parby{}{x_i} + \ssum{i,j}{0}{n}
1436: G_{ij}\parbyxx{}{x_i}{x_j}
1437: \end{align}
1438: where $g_i,G_{ij}\RntoR$ for $1 \le i,j \le n$. We can compute
1439: the commutator
1440: \begin{align}
1441: \label{eq:213}
1442: C(f,F,g,G) \define O(f,F)O(g,G) - O(g,G)O(f,F)
1443: \end{align} of such
1444: second order operators. If this commutator was also
1445: a second order operator (\ie there were~$\varphi$ and $\Phi$
1446: such that the operator~$C(f,F,g,G) = O(\varphi,\Phi)$), then we would
1447: be able to simplify any PDEs of stochastic
1448: transformations. (See Proposition~\ref{prop:sfbooo}).
1449: 
1450: Because the operator~$O$ is linear, \ie~$O(f,F) = O(f,0) + O(0,F)$, we
1451:  can split the computation into four independent, reusable parts:
1452: \begin{multline}
1453:  C(f,F,g,G) = O(f,F)O(g,G) - O(g,G)O(f,F) = \\
1454:  =\lie{[f,g]}{} + C(0,F,0,G) + C(f,0,0,G) + C(0,F,g,0)
1455:  .\end{multline}
1456: 
1457: The first term is already a first order operator. Only the second
1458:  and the third terms need to be computed because the fourth term
1459:  can be obtained from the third one by formal substitution. For
1460: the third term:
1461: \begin{multline}
1462:  O(f,0) O(0,G) = \ssum{i}{1}{n} f_{i} \parby{}{x_i} \left(
1463:  \ssum{k,l}{1}{n} G_{kl}
1464:  \parbyxx{}{x_k}{x_l} \right) = \\
1465:  \ssum{i,k,l}{1}{n} \left( f_{i} \parby{G_{kl}}{x_i}
1466:  \parbyxx{}{x_k}{x_l} + f_{i} G_{kl} \parbyxxx{}{x_i}{x_k}{x_l}
1467:  \right) 
1468: .\end{multline} 
1469: Further,
1470: \begin{multline}
1471:  O(0,G)0(f,0) = \ssum{k,l}{1}{n} G_{kl} \parbyxx{}{x_k}{x_l} \left(
1472:  \ssum{i}{1}{n} f_i \parby{}{x_i}
1473:  \right) =\\
1474:  \ssum{i,k,l}{1}{n} G_{kl} \left( \parbyxx{f_i}{x_k}{x_l}
1475:  \parby{}{x_i} + 2 \parby{f_i}{x_l} \parbyxx{}{x_i}{x_k} + f_i
1476:  \parbyxxx{}{x_i}{x_k}{x_l} \right) .\end{multline}
1477: 
1478: The intermediate results for the third and the fourth terms
1479:  can be combined into
1480: \begin{multline}
1481:  O(f,0) O(0,G) + O(0,F) O(g,0) - O(g,0) O(0,F) - O(0,G) O(f,0) =\\
1482:  \ssum{i,k,l}{1}{n} \left( f_{i} \parby{G_{kl}}{x_i} - g_{i}
1483:  \parby{F_{kl}}{x_i} + 2 F_{ki} \parby{g_l}{x_i} - 2 G_{ki}
1484:  \parby{f_l}{x_i}
1485:  \right) \parbyxx{}{x_k}{x_l} \\
1486:  \left( F_{kl} \parbyxx{f_i}{x_k}{x_l} - G_{kl}
1487:  \parbyxx{g_i}{x_k}{x_l} \right) \parby{}{x_i} .\end{multline} All
1488: of them are first and second order operators. Now let's evaluate the
1489: second term
1490: \begin{multline}
1491:  O(0,F) O(0,G) = \ssum{i,j}{1}{n} F_{ij} \parbyxx{}{x_i}{x_j} \biggl(
1492:  \ssum{k,l}{1}{n} G_{kl} \parbyxx{}{x_k}{x_l} \biggr) =\\
1493:  \ssum{i,j,k,l}{1}{n} \biggl( F_{ij} \parbyxx{G_{kl}}{x_i}{x_j}
1494:  \parbyxx{}{x_k}{x_l} + (F_{ij} + F_{ji}) \parby{G_{kl}}{x_j}
1495:  \parbyxxx{}{x_i}{x_k}{x_l} + F_{ij} G_{kl}
1496:  \parbyxxxx{}{x_i}{x_j}{x_k}{x_l} \biggr) .\end{multline}
1497: \begin{multline}
1498:  O(0,F) O(0,G) - O(0,G) O(0,F) = \ssum{i,j,k,l}{1}{n} \biggl( \biggl(
1499:  F_{ij} \parbyxx{G_{kl}}{x_i}{x_j} - G_{ij}
1500:  \parbyxx{F_{kl}}{x_i}{x_j}
1501:  \biggr) \parbyxx{}{x_k}{x_l} +\\
1502:  \biggl( (F_{ij} + F_{ji}) \parby{G_{kl}}{x_j} - (G_{ij} + G_{ji})
1503:  \parby{F_{kl}}{x_j} \biggr) \parbyxxx{}{x_i}{x_k}{x_l} \biggr)
1504:  .\end{multline} Unfortunately the last term
1505: \begin{align}
1506:  \label{eq:60}
1507:  \biggl( (F_{ij} + F_{ji}) \parby{G_{kl}}{x_j} - (G_{ij} + G_{ji})
1508:  \parby{F_{kl}}{x_j} \biggr) \parbyxxx{}{x_i}{x_k}{x_l}
1509: \end{align}
1510: is of third order and, in general, it does not vanish.  Thus we have
1511: shown that the commutator of two general second order operators is of
1512: third order. Consequently, the Leibniz rule simplifications used in
1513: the deterministic case cannot be applied to the general stochastic
1514: linearization problem.
1515: 
1516: 
1517: 
1518: 
1519: %======== Subsection ================================================
1520: \section{Stochastic Case } 
1521: \label{sub:stoclass}
1522: 
1523: Since there are two  definitions of coordinate transformation of stochastic
1524: differential equations (\Ito, Stratonovich) and three
1525: definitions of linearity ($g,\sigma,g\sigma$), we face at least six
1526: stochastic problems per a deterministic one. In this section we will
1527: discuss all of them giving at least partial solutions to the feedback
1528: linearization problem. We consider mainly the SISO problem except for
1529: cases where the MIMO extension is trivial.
1530: 
1531: %======== Subsection ================================================
1532: \subsection{Stratonovich $g$-linearization}
1533: \label{sub:stotech}
1534: 
1535: 
1536: We show that the method known for deterministic systems can
1537: be applied without modifications.
1538: 
1539: \begin{prop} \label{prop:stratgprop}
1540:  The Stratonovich dynamical system~\ssdefmi is $g$-linearizable if and
1541:  only if the deterministic system~\dsdefmi is linearizable. These two
1542:  linearizing transformation{}s are equal. This holds both for SISO and
1543:  MIMO systems and both for SFB and SCT linearization.
1544: \end{prop}
1545: \begin{proof}
1546:  The comparison of transformation laws for deterministic and
1547:  Stratonovich systems shows that the coefficients of $f$ and
1548:  $g$ transform in the same way.  The
1549:  controllability conditions and the definition of linearity are
1550:  also identical (compare Definition~\ref{def:linear} with
1551:  Definition~\ref{def:linearsto}). Identical problems have
1552:  identical  solutions.
1553: \end{proof}
1554: 
1555: %======== Subsection ================================================
1556: \subsection{Stratonovich $g\sigma$-linearization}
1557: \label{sub:ssgsigma}
1558: 
1559: The Stratonovich problems are not complicated by the second order \Ito
1560: term. The transformation laws for Stratonovich systems are the same as
1561: the deterministic transformation laws, therefore many results of the
1562: deterministic linearization theory can be used. 
1563: 
1564: For example, the stochastic SCT $g\sigma$-linearization of a
1565: Stratonovich system is equivalent to the linearization of a
1566: deterministic, non-square MIMO system with two inputs and a single
1567: output. The SFB problem, which is studied in this section, is not as
1568: simple as the SCT one because the feedback influences only the control
1569: input~$u$ (Figure~\ref{fig:picassym}). The ``dispersion input'' is not
1570: a part of the feedback.  Consequently, in order to solve the
1571: Stratonovich SFB $g\sigma$-linearization, we have to deal with a
1572: combined deterministic SFB-SCT problem.
1573: \psfig{fig:picassym}{Asymmetry of  SFB $g\sigma$-Linearization}{picassym}
1574: 
1575: \subsubsection{Canonical Form}
1576: Recall that we require~$g$-controllability of the resulting system
1577: 
1578: Since this is a Stratonovich problem, the transformed
1579: vectorfields~$\tilde f$ and~$\tilde g$ do not depend on the dispersion
1580: vectorfield~$\sigma$. Therefore, the {\em control part\/} and the {\em
1581:   dispersion part\/} can be studied independently.
1582: 
1583: Any~$g$-linear system can be transformed into integrator chain by a
1584: combination of a linear coordinate transformation and linear
1585: feedback. Therefore, if we set~$\sigma=0$, the canonical form is the
1586: integrator chain. 
1587: 
1588: In general, the dispersion vectorfield~$\tilde \sigma$ is assumed to
1589: be arbitrary constant vectorfield~$\tilde \sigma(x)_i=s_i$, $1\le i\le n$
1590: (See Definition~\ref{def:linearsto}) and this form is preserved by arbitrary linear
1591: transformations. Therefore the canonical form can be written as:
1592: \begin{align}
1593: \label{eq:1000}
1594: \tilde f_i(x) &= x_{i+1} \firstfor 1\le i\le n-1\\
1595: \tilde f_n(x) &= 0\\
1596: \tilde g_i(x) &= 0 \firstfor 1\le i\le n-1\\
1597: \tilde g_n(x) &= 1\\
1598: \label{eq:1077}
1599: \tilde \sigma_i(x) &= s_i \firstfor 1\le i\le n
1600: .\end{align}
1601: 
1602: We can compare this canonical form with the equations which
1603: define the transformed system~$\tilde \Theta$.
1604: 
1605: \begin{prop} \label{prop:s1sfb1a}
1606: There is a SFB $g\sigma$-linearizing transformation
1607:  of the SISO Stratonovich system
1608:  \ssdef into a~$g$-controllable linear system if and only if
1609:  there is a solution~$\lambda \RntoR$ of the set of partial differential equations:
1610: \begin{alignat}3
1611: \label{eq:7}
1612: \biglie{\adfg{i}}{\lambda }&=0\firstfor 0 \le i \le n-2\\
1613: \label{eq:13}
1614: \biglie{\adfg{{n-1}}}{\lambda }&\ne 0\\
1615: \label{eq:14}
1616: \biglie{\ad{f}{i}{\sigma}}{\lambda }&=s'_{i+1} \nextfor 0 \le i \le
1617: n-1
1618: \end{alignat}
1619: such that~$s'_i \in \RR$ are constants on~$U$ for $1 \le i \le n$.
1620: Then the linearizing transformation is given by:
1621: \begin{alignat}3
1622: \label{eq:22}
1623: T_i&=\multilie{f}{{i-1}}\lambda  \firstfor 1\le i\le n\\
1624: \alpha &=\frac{-\multilie{f}{{n}}\lambda 
1625: }{\lie{g}{}\multilie{f}{{n-1}}\lambda } &\qquad\qquad&
1626: \label{eq:28}
1627: \beta =\frac{1}{\lie{g}{}\multilie{f}{{n-1}}\lambda }
1628: .\end{alignat}
1629: \end{prop}
1630: \begin{proof}
1631: Assume that~$\Theta_S$ is transformed by~$\combinedtab$ into 
1632: $\tilde \Theta \define \left(\tilde f,\tilde g,\tilde
1633:   \sigma,T(U),T(\origi{x})\right)$ where the~$i$-th components
1634: of~$f$,$g$, and~$\sigma$ can be expressed as:~$\tilde f_i =
1635: \lie{f}{T_i}$, $\tilde g_i =
1636: \lie{g}{T_i}$, $\tilde \sigma_i =
1637: \lie{\sigma}{T_i}$. Moreover, the feedback is defined by~$ u = \alpha  + \beta  v $.
1638: The equations of~$\Theta$ can be compared to
1639:  the equation of the canonical  form~\eqref{eq:1000}-\eqref{eq:1077}.
1640: \begin{alignat}{3}
1641:  \label{eq:160}
1642:  \lie{f}{T_i} &= T_{i+1} \firstfor 1 \le i \le n-1\\
1643:  \label{eq:161}
1644:  \lie{g}{T_i} &= 0 \nextfor 1 \le i \le n-1\\
1645:  \label{eq:162}
1646:  \lie{g}{T_n} &= 1/\beta  \ne 0 \\
1647:  \label{eq:163}
1648:  \lie{f}{T_n} &= -\alpha /\beta  .\end{alignat} 
1649: 
1650: The
1651: relations~\eqref{eq:7}, \eqref{eq:13}, \eqref{eq:22}, and~\eqref{eq:28} are equivalent to
1652: relations~\eqref{eq:8}-\eqref{eq:12} from 
1653: Proposition~\ref{prop:d1sfb1}. The relation~\eqref{eq:14} can be
1654: verified in a similar way:
1655: \begin{alignat}{3}
1656:  \label{eq:165}
1657:  \lie{\sigma}{T_i} &= s_{i} \firstfor 1 \le i \le n
1658: \end{alignat}
1659: thus by ~\eqref{eq:160}
1660: \begin{alignat}{3}
1661:  \label{eq:254}
1662:  \lie{\sigma}{\lie{f}{T_i}} &= s_{i+1} \firstfor 1 \le i \le n-1
1663: \end{alignat}
1664:  and by ~\eqref{eq:63}
1665: \begin{alignat}{3}
1666:  \label{eq:255}
1667:  \lie{\sigma}{\lie{f}{T_i}} &= \lie{f}{\lie{\sigma}{T_i}}
1668:  -\lie{[f,\sigma]}{T_i} \firstfor 1 \le i \le n-1
1669: \end{alignat}
1670: since the Lie derivative of a constant is zero:
1671: \begin{alignat}{3}
1672:  \label{eq:256} \lie{f}{\lie{\sigma}{T_i}} &= \lie{f}{s_i} = 0\\
1673:  s_{i+1} \define \lie{\sigma}{\lie{f}{T_i}} &=
1674:  -\lie{[f,\sigma]}{T_i} \firstfor 1 \le i \le n-1 .\end{alignat}
1675: The equations~\eqref{eq:14} are obtained by successive application of
1676: this relation. The symbols~$s_i$ are equal to~$s'_{i}$ except for the signs.
1677: \end{proof}
1678: 
1679: \subsubsection{Conditions for the Control Part}
1680: 
1681: The necessary conditions for linearizability of the {\em control
1682:   part\/} of~$\Theta$ (\ie the system~$\left(f,g,0,U,\origi{x}
1683: \right)$) can be expressed in geometrical form. We intentionally omit
1684: the dispersion part, using the fact that the resulting system must be
1685: linear when the noise is zero.
1686: 
1687: Further, the class of all solutions of this subproblem will be
1688: called~$C$. This class can be studied to find  if some member
1689: of~$C$ linearizes the {\em dispersion part\/} of the system.
1690: 
1691: We can find a geometrical criterion similar to the conditions of
1692: Proposition~\ref{prop:d1sfb2}. In this case these conditions are
1693: necessary but not sufficient since also \eqref{eq:14} must be
1694: satisfied.
1695: 
1696: \begin{prop}
1697:  \label{prop:s1sfb1}
1698:  SFB~$g\sigma$-linearizing transformation of the Stratonovich system
1699:  $\Theta_S$ into a~$g\sigma$-controllable system 
1700:  linear system exists only if the distribution~$\distrofg{n-2}$ is
1701:  involutive and the distribution~$\distrofg{n-1}$ is $n$-dimensional.
1702: \end{prop}
1703: \begin{proof}
1704:  This theorem is equivalent to Proposition~\ref{prop:d1sfb2} which is
1705: a direct consequence of Proposition~\ref{prop:d1sfb1a} which
1706: corresponds to Proposition~\ref{prop:s1sfb1a}. 
1707: \end{proof}
1708: 
1709: 
1710: \subsubsection{Condition for the Dispersion Part}
1711: \label{prop:cftdp}
1712: The conditions of Proposition~\ref{prop:s1sfb1} can be written
1713: in matrix form. We are looking for~$T_1 = \lambda\RntoR$ such that
1714: \begin{align}
1715:  \label{eq:189}
1716:  \left[\,\,\begin{matrix}
1717:  \adfg{0}\\
1718:  \adfg{1}\\
1719:  \vdots\\
1720:  \adfg{n-2}\\
1721:  \end{matrix}\,\,\right]
1722: \quad \left[\,\,\begin{matrix}
1723:  \parby{\lambda }{x_1}\\
1724:  \parby{\lambda }{x_2}\\
1725:  \vdots\\
1726:  \parby{\lambda }{x_n}\\
1727:  \end{matrix}\,\,\right]
1728: = \left[0\right]
1729: .\end{align}
1730: 
1731: 
1732: The vectors~$ \adfg{0} \dots \adfg{n-2}$ are written in coordinates
1733: as~$1\times n$ rows. The first matrix is~$(n-1)\times n$. Moreover it is
1734: required that
1735: \begin{equation}
1736: \label{eq:1020}
1737: \biglie{\adfg{n-1}}{\lambda}
1738: \end{equation} is nonzero. 
1739: 
1740: We will use the algorithm for SFB deterministic linearization (see
1741: Section~\ref{sec:detcase}) to find such a transformation~$\lambda$.Then we will
1742: verify if the conditions for linearity of the dispersion part of the
1743: system~\eqref{eq:14} are also valid. There are~$n$ additional linearity
1744: conditions ($s_i$ are constants):
1745: 
1746: \begin{align}
1747: \label{eq:1001}
1748:  \left[\,\,\begin{matrix}
1749:  \ad{f}{0}{\sigma}\\
1750:  \ad{f}{1}{\sigma}\\
1751:  \vdots\\
1752:  \ad{f}{n-1}{\sigma}\\
1753:  \end{matrix}\,\,\right]
1754: \quad \left[\,\,\begin{matrix}
1755:  \parby{\lambda }{x_1}\\
1756:  \parby{\lambda }{x_2}\\
1757:  \vdots\\
1758:  \parby{\lambda }{x_n}\\
1759:  \end{matrix}\,\,\right]=
1760: \left[\,\,\begin{matrix}
1761:  s_1\\
1762:  s_2\\
1763:  \vdots\\
1764:  s_n\\
1765:  \end{matrix}\,\,\right]
1766: .\end{align}
1767: 
1768: In the deterministic case we were satisfied with {\em arbitrary\/}
1769: solution~$\lambda $ to the equations~\eqref{eq:189},
1770: and~\eqref{eq:1020} . In this stochastic case we must find the class
1771: of {\em all\/} solutions and then check if this class contains the
1772: solution for the~$\sigma$ part~\eqref{eq:1001}. Details depend on the
1773: methods used for solving the set of PDEs.
1774: 
1775: This result is summarized in the following
1776: algorithm:
1777: 
1778: \begin{itemize}
1779: \item{Find $\Delta_k \define \adfg{i}$ for $0\le i \le k-1$.}
1780: \item{Verify that $\dim(\Delta_{n})$ is $n$.}
1781: \item{Verify that $\Delta_{n-1}$ is
1782:  involutive (see \citet{nijmeijer94} Remark following Definition 2.39)
1783:  otherwise no linearizing 
1784: transformation exists.}
1785:  \item{Find all~$\lambda$ satisfying~\eqref{eq:192} by solving 
1786: PDEs~\eqref{eq:192}; denote~$C$ the set of all
1787:  such functions.} 
1788: \item{Verify that there is a~$\lambda_1\in
1789:  C$ such that the conditions~\eqref{eq:1001} are
1790:  satisfied, otherwise no linearizing
1791:  transformation exists.} 
1792: \item{Compute $T$,$\alpha $,$\beta $ from
1793:  \eqref{eq:160} -- \eqref{eq:163}.}
1794: \end{itemize}
1795: 
1796: Now, we can illustrate one possible practical approach  which
1797: worked for several simple problems solved by us (see the example in
1798: Section~\ref{sub:appcrane}).
1799: 
1800: First we can compute the kernel of the matrix~$M_g$ to find the
1801: form~$\omega = \left[ \omega_1, \omega_2, \dots, \omega_n
1802: \right]^T$ which satisfies~$M_g \omega = 0$, \ie~$\omega$ is
1803: perpendicular to~$M_g$. In modern computer algebra systems there is a
1804: single command for this.
1805: 
1806: Proposition~\ref{prop:s1sfb1} assumes that~$n$
1807: vectorfields~$\Delta_{n} \define \distrofg{n-1}$ form an~$n$
1808: dimensional space. The vectorfields~$\Delta_{n-1} \define
1809: \distrofg{n-2}$ are chosen from them and consequently must form
1810: an~$(n-1)$-dimensional space. Thus their kernel~$d\lambda $ is
1811: exactly one dimensional and arbitrary~$\omega' = c(x)\omega(x)$ also
1812: belongs to the kernel ($c(x)$ is a scalar). 
1813: 
1814: But not every~$\omega'$ that is perpendicular to~$M_g$ is a
1815: solution to the original linearization problem. The
1816: function~$\omega'$ must be an exact one-form \ie there must be
1817: a scalar function~$\lambda $ such that~$d \lambda =
1818: c(x)\omega(x)$. The Frobenius theorem guarantees that
1819: if~$\Delta_{n-1}$ is involutive, then there is always~$c(x)\in\RR$
1820: such that~$c(x)\omega(x)$ is the exact one-form.
1821: 
1822: A necessary condition for a one-form~$\omega =
1823: \ssumi{n} \omega_i$ to be exact is
1824: \begin{equation}
1825:  \label{eq:102}
1826:  \parby{\omega_i}{x_j} = \parby{\omega_j}{x_i} \qquad \text{for}
1827:  \qquad 1 \le i,j \le n
1828: .\end{equation}
1829: 
1830: Hence for every~$1\le i,j \le
1831: n$
1832: \begin{alignat}{3}
1833:  \parby{}{x_j}\left(c(x)\omega_i \right) &=
1834:  \parby{}{x_i}\left(c(x)\omega_j \right),\end{alignat} thus for
1835: every~$1\le i,j \le n$
1836: \begin{alignat}{3}
1837: \label{eq:192}
1838: \parby{c(x)}{x_i}\omega_j - \parby{c(x)}{x_j}\omega_i + c(x)
1839: \left( \parby{\omega_j}{x_i} - \parby{\omega_i}{x_j} \right) = 0
1840: .\end{alignat} The later condition is a set of linear PDEs, with
1841: unknown~$c(x)$, which are guaranteed to have a solution by the
1842: involutivness of~$\Delta_{n-1}$ (the Frobenius theorem). 
1843: 
1844: In our computations the equation~\eqref{eq:192} was in a simple form
1845: which allowed to determine all the solutions easily. More complicated
1846: cases will require more sophisticated analysis.
1847: 
1848: 
1849: 
1850: %======== Subsection ================================================
1851: \subsection{\Ito $g\sigma$-linearization}
1852: \label{sub:isgsigma}
1853: 
1854: 
1855: In the previous subsection we tried to find~$g\sigma$-linearizations
1856: for Stratonovich dynamical systems. Once this is done, the
1857: correcting mapping can be used to
1858: construct \Ito $g\sigma$-linearizing
1859: transformation. This method works for both the
1860: SFB and the
1861: SCT case.
1862: 
1863: Given an \Ito system~$\Theta_I$, the corresponding
1864: Stratonovich system $\Theta_S$ can be
1865: obtained using the correcting mapping
1866: $\Theta_S = \Corr{\sigma}\left({\Theta_I}\right)$.  Afterward, the Stratonovich
1867: $g\sigma$-linearization algorithm can be applied giving a
1868: linear system~$\Theta_{2S}$. Due to linearity of the
1869: drift
1870: vectorfield~$\tilde \sigma$ of~$\Theta_{2S}$, the
1871: correcting term~$\corr{{\tilde \sigma}}{z}$ of
1872: the backward transformation~${\Corr{{\tilde \sigma}}{}}^{-1}$
1873: vanishes.
1874: 
1875: 
1876: \begin{theo} \label{prop:gsigmaprop}
1877:  The SFB 
1878:  $g\sigma$-linearizing
1879:  transformation~$\combinedz_I$ of the~\Ito dynamical system \isdefmi,
1880:  $f(\origi{x})=0$, $\corr{\sigma}{\origi{x}}=0$, into a $g\sigma$-controllable linear
1881:  system  exists if and
1882:  only if there is a SFB
1883:  $g\sigma$-linearizing
1884:  transformation~$\combinedz_S$ of the Stratonovich dynamical
1885:  system
1886: 
1887: \begin{align}
1888: \label{eq:24}
1889: \Theta_S &= \ssys{\alterbar f}{g}{\sigma}{x} =
1890:  \Corr{\sigma}{(\Theta_I)}\\
1891: \alterbar f &= f + \corr{\sigma}{x}
1892: .\end{align}
1893: Moreover~$\combinedz_I = \combinedz_S \compose \Corr{\sigma}{}$.
1894: 
1895: \end{theo}
1896: \begin{proof}[Proof (sufficiency)]
1897:  We use the properties of the correcting term
1898:  (Subsection \ref{sub:corr}). Assume that there is a
1899:  mapping~$\combinedz_S$ which transform $\Theta_S$ into a linear
1900: ~$g$-controllable system $(Ax,B,S,U,0)$  . By~(\ref{eq:19})
1901:  \begin{align}
1902:  \label{eq:248}
1903:  \combinedz_I = \Corr{{\tilde\sigma}}{}^{-1} \compose \combinedz_S
1904:  \compose \Corr{\sigma}{} 
1905: .\end{align}
1906:  The backward correcting
1907: transformation $\Corr{{\tilde \sigma}}^{-1}$ is identity because the correcting term
1908: of a linear mapping $\corr{{\tilde \sigma}}{x}$ is zero. Thus~$
1909: \combinedz_I =  \combinedz_S \compose \Corr{\sigma}{} $
1910: and ~$\combinedz_I (\Theta_{I})$ equals $(Ax,B,S,U,0)$, which is
1911: linear and~$g$-controllable by assumption.
1912: \end{proof}
1913: \begin{proof}[Proof (necessity)]
1914:   Assume that there is the \Ito transformation~$\combinedz_I$ which
1915:   linearizes $\Theta_I$ and by~(\ref{eq:24}) $\Theta_I =
1916:   {\Corr{\sigma}{}}^{-1}(\Theta_S)$. Construct Stratonovich
1917:   linearization by~$\combinedz_S = \combinedz_I \compose
1918:   {\Corr{\sigma}{}}^{-1} $.  Hence~$\combinedz_I$ linearizes
1919:   ${\Corr{\sigma}{}}^{-1}(\Theta_S)$ and~$\combinedz_S$
1920:   linearizes~$\Theta_S$ into the same linear and controllable system
1921:   as~$\combinedz_I$.
1922: \end{proof}
1923:  
1924: 
1925: 
1926: %======== Subsection ================================================
1927: \subsection{\Ito~$g$-linearization}
1928: \label{sub:sfbig}
1929: 
1930: The \Ito $g$-linearization problem is probably the most complicated
1931: variant of exact linearization studied in this paper. The dispersion
1932: vectorfield of an \Ito dynamical system transformed by a coordinate
1933: transformation~$\sctt$ consists of two terms: the transformed
1934: vectorfield~$\tantra{T} f$ and the \Ito term~$\ito{\sigma}$. We
1935: require that the sum of these  terms is linear, thus the
1936: nonlinearity of the drift~$\tantra{T}f$ must compensate for the \Ito
1937: term. Since the \Ito term behaves to~$T$ as a second order
1938: differential operator, this problems generates a set of second order
1939: partial differential equations. One can attempt to use simplifications
1940: as in the deterministic linearization, namely, the recursive Leibniz
1941: rule~\eqref{eq:10}. Unfortunately, this approach does not work for the
1942: stochastic case. In general, the \Ito equations cannot be easily
1943: simplified by commutators because the commutator of second order
1944: operators is not a second order operator but a third order operator
1945: (see Subsection~\ref{sub:stoinvar}).
1946: 
1947: Nevertheless, there are special cases for which simpler conditions
1948:  can be found. The most important special case (commuting $g$
1949:  and $\sigma$) will be studied here.
1950: 
1951: \subsubsection{Canonical Form ---$n$ unknowns}
1952: 
1953: The canonical form for the $g$-linearization
1954:   is the integrator chain with a nonlinear drift
1955: \begin{align}
1956: \label{eq:1007}
1957: \tilde f_i(x) &= x_{i+1} \firstfor 1\le i\le n-1\\
1958: \label{eq:1008}
1959: \tilde f_n(x) &= 0\\
1960: \label{eq:1009}
1961: \tilde g_i(x) &= 0 \firstfor 1\le i\le n-1\\
1962: \label{eq:1010}
1963: \tilde g_n(x) &= 1
1964: .\end{align}
1965: Assume that there is a~$g$-linear
1966: system~$\Theta_I = (Ax,B,\sigma(x),U,\origi{x})$. Then the drift part
1967: of~$\Theta_I$ can be transformed by a {\em linear} transformation into
1968: the integrator chain. This is because the \Ito term of a linear
1969: transformation vanishes. 
1970: 
1971: The equations which define~$T$ can be obtained by comparing  this
1972: canonical form with the equations of~$\tilde \Theta$.
1973: 
1974: \begin{prop} \label{prop:s1sfb2}
1975:   Let~\isdef be an \Ito dynamical system with~$f(\origi{x})=0$ such
1976:   that~$\corr{\sigma}{\origi{x}}=0$. There is a SFB~$g$-linearizing
1977:   transformation~$\combinedtab$ of the system~$\Theta_{I}$ into a
1978:   ~$g$-controllable linear system if and only if there is a solution
1979:   $T_i\RntoR$, $1\le i \le n$, to the set of partial differential
1980:   equations defined on~$U$:
1981: \begin{alignat}{3}
1982: \label{eq:50}
1983: T_{i+1}&=\lie{f}{T_i}+\ito{\sigma}{T_i} \firstfor 1\le i\le{n-1}\\
1984: \label{eq:51}
1985: \lie{g}{T_i}&=0 \nextfor 1\le i\le{n-1}\\
1986: \label{eq:52}
1987: \lie{g}{T_n}&\not=0 .\end{alignat} The symbol~$\ito{\sigma}$
1988: denotes the \Ito operator (see Definition~\ref{eq:16}). The
1989: feedback can be constructed as:
1990: \begin{align}
1991: \label{eq:53}
1992: \alpha &=-\frac{(\lie{f}{T_n}+\ito{\sigma}{T_n})}{\lie{g}{}T_n}
1993: \qquad\qquad \beta =\frac{1}{\lie{g}{T_n}}
1994: .\end{align}
1995: \end{prop}
1996: \begin{proof}
1997: 
1998: The~$i$-th components
1999: of~$f$,$g$, and~$\sigma$ are:~$\tilde f_i =
2000: \lie{f}{T_i} + \ito{\sigma}{T_i}$, $\tilde g_i =
2001: \lie{g}{T_i}$, $\tilde \sigma_i =
2002: \lie{\sigma}{T_i}$. 
2003:  The partial differential equations \eqref{eq:50}, \eqref{eq:51}
2004:  and \eqref{eq:52} are obtained by comparison of \eqref{eq:86}-\eqref{eq:89}
2005:  with the equations~\eqref{eq:1007}-\eqref{eq:1010}.
2006: \end{proof}
2007: \subsubsection{PDEs of single unknown}
2008: 
2009: One can attempt to reduce the equations~\eqref{eq:50},
2010: \eqref{eq:51}, and~\eqref{eq:52}, to a set of equations of a
2011: single unknown, similarly to the results of Proposition \ref{prop:d1sfb1}.
2012: 
2013: \begin{coroll}
2014:  \label{prop:sfbooo} Define the general second order
2015:  operator~$O(f,F)$ as in~\eqref{eq:17}. The exponential notation
2016:  for~$O$ will be defined recursively: $O^0(f,F)T \define T$ and
2017:  $O^{l+1}(f,F)T \define O(f,F)O^l(f,F)T$ for~$l \ge 0$. Next, define
2018:  \begin{align} F_{ij} \define \frahalf \sigma_i \sigma_j .
2019:  \end{align} Then the set of partial differential equations~\eqref{eq:50}-\eqref{eq:52} of~$n$ unknowns has a 
2020:  solution if and only if there is a 
2021:  solution~$\lambda \RntoR$ defined on~$U$ to the set of
2022:  PDEs of single unknown:
2023:  \begin{alignat}{3} \label{eq:153}
2024:  O(g,0)O^i(f,F)\lambda &=0\firstfor 0\le i\le n-2\\ 
2025:  \label{eq:154} O(g,0)O^{n-1}(f,F)\lambda &\ne 0 . \end{alignat}
2026:  The original solution and the feedback
2027:  can be found as \begin{alignat}{3}
2028:  T_i&=O^{i-1}(f,F)\lambda \firstfor 1 \le i \le n\\
2029:  \label{eq:58} \alpha &= -
2030:  \frac{O(f,F)^{n-1}\lambda }{O(g,0)^{n-1}\lambda } \quad\quad \beta =
2031:  \frac{1}{O(g,0)^{n-1}\lambda } . \end{alignat}
2032: \end{coroll}
2033: \begin{proof}
2034:  Since~$T_1=\lambda $ and by definition
2035:  of~$\ito{\sigma}{}$ \eqref{eq:16} and~$O$~\eqref{eq:17}:
2036:  \begin{alignat}3
2037: \label{eq:264}
2038: O(f,F)T_i &= \lie{f}{T_i} + \ito{\sigma}{T_i} \firstfor 1\le i\le n\\
2039: \label{eq:265}
2040: O(g,0)T_i &=\lie{g}{T_i} \nextfor 1\le i\le n ;
2041:  \end{alignat} then~$T_{i+1} = O(f,F)T_i$ by~\eqref{eq:50} and
2042:  $T_i = O^{i-1}(f,F)T_1 = O^{i-1}(f,F)\lambda $. Similarly
2043:  the equation \eqref{eq:53} implies~\refprop{eq:58}.
2044: \end{proof}
2045: 
2046: Note, that for the deterministic case~$\sigma=0$, the
2047: operators~$O(g,0)$ and~$O(g,0)O^i(f,F)$ degenerate
2048: to~$\lie{g}{}$ and to~$\lie{g}{}\multilie{f}{i}{}$ respectively,
2049: thus the result is the same as that of Proposition~\ref{prop:d1sfb1}.
2050: 
2051: In general, the equations of the system are of an order up to~$2n$ and
2052: cannot be reduced to a lower order. The
2053: commutator of two second order operators is of third
2054: order as was pointed out by~\eqref{eq:60}. In particular, for~$i=1$ we
2055: have to evaluate~$C(g,0,f,F) \definer O(a,A)$, which {\em is\/} of
2056: second order due to the fact that~$G=0$. But starting from~$i=2$ the
2057: commutator~$C(f,F,a,A) = C(f,F,C(g,0,f,F))$ is of
2058: third order.
2059: 
2060: \subsubsection{Correcting Term}
2061: 
2062: The same problem can be reformulated using conversion to the
2063: Stratonovich formalism. One can compute the drift vectorfield~$\alterbar f$
2064: of the equivalent Stratonovich system by applying the
2065: correcting term~$\alterbar f \define f +
2066: \corr{\sigma}{x}$. Then the Stratonovich system~$\ssys{\alterbar
2067:  f}{g}{\sigma}{x}$ may be transformed, by a suitable transformation,
2068: to such a form~$(\tilde f,\tilde g,\tilde \sigma,T(U),0)$ that after
2069: applying the backward correcting
2070: term~$-\corr{{\tilde \sigma}}{z}$ the resulting \Ito system will be
2071: linear.
2072: 
2073: Compare this formulation with the~$g\sigma$-linearization where the
2074: backward correcting term vanished due to
2075: linearity of~$\tilde \sigma$. This does not happen with
2076: the~$g$-linearization, and the backward correcting term is a part of the equations.
2077: 
2078: In general, it may be difficult to solve these equations. Nevertheless
2079: there are special cases when the solution can be obtained. See for example
2080: Section~\ref{sub:appcrane}. Another important case (commuting $g$~
2081: and~$\sigma$) is studied below.
2082: 
2083: \begin{coroll} \label{prop:s1sfb3}
2084:  The equations of Proposition~\ref{prop:s1sfb2} are equivalent to~$n$
2085:  partial differential equations:
2086: \begin{alignat}3
2087: \label{eq:143}
2088: T_{i+1}&=\lie{\alterbar f}{T_i} -\corr{{\alterbar \sigma}}{z}=\\
2089: &\lie{\alterbar f}{T_i} + \frahalf \lie{\sigma}{\lie{\sigma}{T_i}} \firstfor 1 \le i \le{n-1}\\
2090: \label{eq:144}
2091: \lie{g}{T_i}&=0\nextfor 1 \le i \le{n-1}\\
2092: \label{eq:145}
2093: \lie{g}{T_n}&\not=0 .\end{alignat} Then the
2094: feedback can be constructed as
2095: \begin{align}
2096:   \alpha &=-\frac{\lie{\alterbar f}{T_n} + \frahalf
2097:     \lie{\sigma}{\lie{\sigma}{T_n}}}{\lie{g}{T_n}} \qquad\qquad
2098:   \beta =\frac{1}{\lie{g}{T_n}} .\end{align} Where $\alterbar f
2099: \define f + \corr{\sigma}{x}$.
2100: \end{coroll}
2101: \begin{proof}
2102:  The equations~\eqref{eq:143} can be obtained
2103:  from~\eqref{eq:50} by applying~\eqref{eq:215} 
2104:  and~\eqref{eq:140}:
2105: \begin{multline}
2106:  \label{eq:216}
2107:  T_{i+1} = \lie{f}{T_i}+\ito{\sigma}{T_i} = \lie{\left(
2108:  \alterbar f - \corr{{\alterbar \sigma}}{x} \right)}{T_i}+
2109:  \ito{\sigma}{T_i} = \\
2110:  \lie{\alterbar f}{T_i} - \lie{\corr{{\alterbar
2111:  \sigma}}{x}}{T_i} + \ito{\sigma}{T_i} = \lie{\alterbar
2112:  f}{T_i} + \frahalf\lie{\sigma}{\lie{\sigma}{T_i}}
2113:  .\end{multline}
2114: 
2115: The other equations are adopted from Proposition~\ref{prop:s1sfb2}.
2116: The set of PDEs of~$n$ unknowns can be transformed into a set
2117: of PDEs of a single unknown~$\lambda =T_1$, but
2118: the order of the equation will be~$2n-1$.
2119: \end{proof}
2120: 
2121: \begin{rem} \label{prop:s1sfb4}
2122:   Observe that the set of~$n$ second order partial differential
2123:   equations~\eqref{eq:143}-~\eqref{eq:145} defined in
2124:   Proposition~\ref{prop:s1sfb3} can be transformed, by introducing 
2125:   new variables~$S_i = \lie{\sigma}{T_i}$,  to the following system
2126:   of~$2n-1$ first order partial differential equations for~$1 \le i
2127:   \le {n-1}$:
2128: \begin{alignat}3
2129: \label{eq:146}
2130: \lie{g}{T_i}&=0\\
2131: \label{eq:147}
2132: \lie{\sigma}{T_i}-S_i&=0\\
2133: \label{eq:148}
2134: \lie{{\alterbar f}}{T_i} + \frahalf\lie{\sigma}{S_i}&=T_{i+1}\\
2135: \label{eq:149}
2136: \lie{g}{T_n}&\not=0 .\end{alignat} $T_i$ 
2137: and~$S_i\RntoR$ are unknown real valued functions defined on $U$.
2138: \end{rem}
2139: 
2140: \subsubsection{Systems with Commuting $g$ and $\sigma$}
2141: There is a special case of \Ito dynamical systems for which the 
2142: solution is completely known and can be computed using
2143: only first order PDEs.
2144: 
2145: \begin{theo}
2146:  \label{prop:ssfbco}
2147:  Let~\isdef be a SISO \Ito dynamical system. If the
2148:  vectorfield~$\sigma$ commutes with all vectorfields~$\ad{{\alterbar
2149:  f}}{i}{g}$ for~$0\le i\le n-1$, \ie ~$[\ad{{\alterbar
2150:  f}}{i}{g},\sigma]=0$, where $\alterbar f
2151: \define f + \corr{\sigma}{x}$, then the \Ito system is $g$-linearizable if and
2152:  only if the distribution
2153: \begin{align}
2154: \alterbar \Delta_{n} \define
2155:  \sspan\left\{{\ad{{\alterbar f}} {i}{g}, \iva{n-1}}\right\}
2156: \end{align}
2157: is
2158:  nonsingular on~$U$ and the distribution
2159: \begin{align}
2160: \alterbar \Delta_{n-1}
2161:  \define \sspan\left\{{\ad{{\alterbar f}}{i}{g}, \iva{n-2}}\right\}
2162: \end{align}
2163: is involutive on~$U$. If these conditions hold, then a
2164:  solution~$\lambda \RntoR$ to the set of partial differential
2165:  equations exists
2166: \begin{align}
2167: \label{eq:77}
2168: \biglie{\ad{{\alterbar f}}{i}{g}}{\lambda }&=0\rfor{\iva{n-2}}\\
2169: \label{eq:78}
2170: \biglie{\ad{{\alterbar f}}{{n-1}}{g}}{\lambda }&\not=0 .\end{align}
2171: the linearizing transformation is given by:
2172: \begin{alignat}3
2173: \label{eq:79}
2174: T_1&=\lambda \\
2175: \label{eq:80}
2176: T_{i+1}&= \lie{\alterbar f}{T_i} + \frahalf
2177: \lie{\sigma}{\lie{\sigma}{T_i}} \firstfor 1
2178: \le i \le{n-1}\\
2179: \alpha &=\frac{-\multilie{\alterbar f}{{n}}\lambda 
2180: }{\lie{g}{}\multilie{\alterbar f}{{n-1}}\lambda } &\qquad\qquad&
2181: \label{eq:81}
2182: \beta =\frac{1}{\lie{g}{}\multilie{\alterbar f}{{n-1}}\lambda }
2183: .\end{alignat}
2184: \end{theo}
2185: \begin{proof}
2186:  We will apply the Leibniz rule to the relation of~(\ref{eq:143})
2187: Corollary~\ref{prop:s1sfb3} to expand the
2188:  term~$\lie{g}{T_{i+1}}$ for~$1\le i\le n-1$ 
2189: \begin{multline}
2190:  \label{eq:195}
2191:  \lie{g}{T_{i+1}} = \lie{g}{ \left( \lie{{\alterbar f}}{T_i}+
2192:  \frahalf \lie{\sigma}{\lie{\sigma}{T_i}} \right) } =
2193:  \lie{g}{\lie{{\alterbar f}}{T_i}} + \frahalf
2194:  \lie{g}{\lie{\sigma}{\lie{\sigma}{T_i}}}
2195:  =\\
2196:  \lie{{\alterbar f}}{\lie{g}{T_i}} - \lie{[\alterbar f,g]}{T_i}
2197:  + \frahalf \lie{\sigma}{\lie{g}{\left(\lie{\sigma}{T_i}\right)}}
2198:  - \frahalf \lie{[\sigma,g]}{\left(\lie{\sigma}{T_i}\right)}
2199:  =\\
2200:  0 - \lie{[\alterbar f,g]}{T_i} + \frahalf \lie{\sigma}{ \left(
2201:  \lie{\sigma}{\lie{g}{T_i}} - \lie{[\sigma,g]}{T_i} \right)
2202:  } - \frahalf \left( \lie{\sigma}{\lie{[\sigma,g]}{T_i}} -
2203:  \lie{[\sigma,[\sigma,g]]}{T_i} \right)
2204:  =\\
2205:  -\lie{[\alterbar f,g]}{T_i} + 0 -  \frahalf
2206:  \lie{\sigma}{\lie{[\sigma,g]}{T_i}} + \frahalf
2207:  \lie{[\sigma,[\sigma,g]]}{T_i} .\end{multline} If the
2208: vectorfields~$g$ and~$\sigma$ commute, then the second and
2209: third terms vanish. If, moreover, $[\sigma,[\alterbar f,g]] = 0$ then
2210: \begin{alignat}{3}
2211:  \label{eq:257}
2212:  \lie{g}{{T_{i+2}}} &= - \lie{{[\alterbar f,g]}}{ \left(
2213:  \lie{{\alterbar f}}{T_i} + \frahalf
2214:  \lie{\sigma}{\lie{\sigma}{T_i}} \right)} = \lie{{[\alterbar
2215:  f,[\alterbar f,g]]}}{T_i} .\end{alignat} In general
2216: if~$[\sigma,\ad{{\alterbar f}}{i}{g}]=0$ for~$0\le i\le n-1$ then 
2217: \begin{alignat}{3}
2218:  \label{eq:258}
2219:  \lie{g}{T_{k}} &= (-1)^k \lie{{[\ad{{\alterbar f}}{k}{g},]}}{T_1}
2220:  .\end{alignat}
2221: 
2222: Thus the equations~\eqref{eq:143} and \eqref{eq:144} will be
2223: equivalent to
2224: \begin{alignat}{3}
2225:  \label{eq:193}
2226:  \biglie{\ad{{\alterbar f}}{i}{g}}{\lambda }&=0\firstfor{\iva{n-2}}\\
2227:  \label{eq:194}
2228:  \biglie{\ad{{\alterbar f}}{{n-1}}{g}}{\lambda }&\not=0
2229: ,\end{alignat} which are of the same form as the equations of
2230: Proposition~\ref{prop:d1sfb1} and consequently the conditions
2231: from Proposition \ref{prop:d1sfb2} can be used.
2232: \end{proof}
2233: 
2234: 
2235:  
2236: %======== Subsection ================================================
2237: \subsection{\Ito and Stratonovich $\sigma$-linearization}
2238: \label{sub:sfbchsigma}
2239: 
2240: 
2241: The stochastic SFB $\sigma$-linearization problem is similar to
2242: deterministic SCT linearization.
2243: The dispersion
2244: vectorfield~$\sigma$ transforms in the same way as deterministic
2245: drift vectorfields do. Consequently, no \Ito term
2246: complicates the transformation. Moreover, the \Ito and
2247: Stratonovich cases are equivalent.
2248: 
2249: On the other hand, in the SFB $\sigma$-linearization we are free to
2250: choose the feedback~$\feedbackab$ that perturbs the
2251: drift
2252: vectorfield~$f$ into~$\tilde f = f + g \alpha $.
2253: 
2254: \psfig{fig:sigmasfb}{Ito and Stratonovich
2255:  $\sigma$-linearization}{sigmasfb}
2256: \begin{prop} \label{prop:sfsigma}
2257:  Let~$\Theta$ be a SISO stochastic
2258:  system~$\Theta = \ssysfgx$. There is a
2259:  SFB $\sigma$-linearizing
2260:  transformation~$\combinedtab$ into a $\sigma$-controllable linear
2261:  system if and only if there is a smooth
2262:  feedback function~$\alpha \RntoR$ such that the
2263:  deterministic system~$\dsys{f+g\alpha }{\sigma}{x}$ has a
2264:  SCT
2265:  linearizing
2266:  transformation~$\sctt$. Equivalently,  there must be
2267: such~$\alpha $ that the the modified odd bracket condition:
2268: \begin{equation}
2269:  \label{eq:59}
2270:  [\sigma,\ad{f+g \alpha  }{l}{\sigma}]=0 \qquad\text{for}\qquad l = 1,\dots,2n-1 
2271: \end{equation}
2272: is satisfied (see~\eqref{eq:45}). The resulting combined
2273: transformation consists of composition of the coordinate transformation~$\sctt$ and the
2274: feedback~$\feedbackab$ where~$\beta $ is arbitrary
2275: function of~$x$; for instance~$\beta  = 1$.
2276: \end{prop}
2277: \begin{proof}
2278: Compare definition of linearity of a deterministic system with
2279:  definition of 
2280:  $\sigma$-linearity. The system is is $\sigma$-linearizable if and only if the
2281:  deterministic systems~$\dsys{f+g\alpha }{\sigma}{x}$ is
2282:  SCT linearizable (see Corollary~\ref{prop:s1ctt2a}).
2283:  
2284:  It is evident that the function~$\beta $ (see
2285:  Figure~\ref{fig:sigmasfb}) has no effect on the dispersion part and
2286:  can be chosen arbitrarily. (Probably nonzero for otherwise the system will
2287:  be $g$-uncontrollable).
2288: \end{proof}
2289: 
2290: 
2291: The condition~\eqref{eq:59} can be expressed in terms of
2292: derivatives of $\alpha $ using bracket
2293: relations known from differential geometry. For example, for~$l=1$:
2294: \begin{align}
2295: \label{eq:156}
2296: [\sigma,[f+g \alpha ,\sigma ]] &= [\sigma, [f,\sigma ]+[g \alpha
2297: ,\sigma ]] = [\sigma,[f,\sigma ]]+[\sigma,[g \alpha ,\sigma ]]=\\
2298: &=[\sigma,g+[f+g,\sigma ]] + g \lie{\sigma}{\lie{\sigma}{\alpha }}
2299: .\end{align}
2300: The other conditions for~$k=3,5,7,\dots$ can be expressed in a similar
2301: way giving the set of~$n$ partial
2302: differential equations of the order up to~$2n$ for example by a
2303: computer using symbolic algebra tools. The
2304: problem is not very interesting from the practical point of view.
2305: 
2306: 
2307: %======== Subsection ================================================
2308: \section{Example---Crane}
2309: \label{sub:appcrane}
2310: 
2311: 
2312: In this section the methods of stochastic exact linearization are
2313: demonstrated on an example --- control of a crane under
2314: the influence of random disturbances. The description of the plant was
2315: adopted from~\citet{ackermann93} where the model of a crane linearized
2316: by approximative methods was studied. Unlike Ackermann, we control the
2317: same system using the exact model. Moreover the influence of random
2318: disturbances is added. \psfig{fig:crane1}{Crane}{crane1} Consider the
2319: crane of Figure~\ref{fig:crane1}, which can be used for example for
2320: loading containers into a ship. The hook must
2321: be automatically placed to a given position.
2322: Feedback control is needed in order to dampen
2323: the motion before the hook is lowered into the ship. The input signal
2324: is the force~$u$ that accelerates the crab. The crab mass is~$m_C$,
2325: the mass of the load~$m_L$, the rope length is~$l$, and the
2326: gravity acceleration~$g$.
2327: 
2328: We assume that the driving motor has no nonlinearities, there is no
2329: friction or slip, no elasticity of the rope and no damping
2330: of the pendulum (\eg from air drag). We will define four state
2331: variables: the rope angle $x_1$ (in radian), the angular velocity~$x_2
2332: = \dot x_1$, the position of the crab~$x_3$, and the velocity of
2333: the crab~$x_4 = \dot x_3$. As shown in~\citez{ackermann93}, the
2334: plant is described by two second order differential equations:
2335: \begin{align}
2336:  \label{eq:72}
2337:  u &= (m_L + m_C) \ddot x_3 + m_L l ( \ddot x_1 \cos x_1 - \dot x_1^2
2338:  \sin
2339:  x_1) \\
2340:  \label{eq:73}
2341:  0 &=m_L \ddot x_3 \cos x_1 + m_L l \ddot x_1 + m_L g \sin x_1
2342: .\end{align}
2343: Additionally, we assume that the load is under influence of random
2344: disturbance, which can be modeled as a white noise
2345: process. The disturbance (wind) is horizontal, has zero mean and
2346:  can be described by the \Ito differential~$dw$:
2347: \begin{align}
2348:  \label{eq:74}
2349:  dx_2 = \frac{F \cos x_1}{m_L l}\,dw,\end{align} where~$F$ is a
2350: constant having the physical unit of force.
2351: 
2352: We used symbolic algebraic system \Mathematica to handle the
2353: computations. The complete \Mathematica worksheet can be downloaded
2354: from the web page of the author  \hyref{http://www.tenzor.cz/sladecek}.
2355: 
2356: \Mathematica was used to solve the equations of the system for
2357: unknown values~$\dot x_2$ and~$\dot x_4$ (angular and
2358: positional acceleration). Values of vectorfields~$f$, $g$, and
2359: $\sigma$ were derived as follows:
2360: 
2361: \begin{align}
2362:  \label{eq:76}
2363:  f &= \left[ x_2, -\frac{\sin x_1 \left( g (m_L+m_C) + l m_L x_2 \cos
2364:  x_1 \right) }
2365:  {l (m_C + m_L - m_L \cos^2 x_1) }, x_4, 0\right]^T\\
2366:  g &= \left[ 0, -\frac{\cos x_1 }
2367:  {l (m_C + m_L - m_L \cos^2 x_1) }, 0, u\right]^T\\
2368:  \sigma &= \left[0, \frac{F \cos x_1}{m_L l},0,0 \right]^T
2369: .\end{align}
2370: 
2371: 
2372: The state space model is shown in Figure~\ref{fig:crane2}. We can see
2373: that the positional state variables~$x_3$ and~$x_4$ are isolated from
2374: the angular state variables~$x_1$ and~$x_2$. Later, we will
2375: concentrate on the angular variables pretending that the load will be
2376: stabilized no matter where the crane is. Consequently, we obtain only
2377: two-dimensional system on which the exact linearization techniques can
2378: be demonstrated. \psfig{fig:crane2}{The State Space Model of
2379:   Crane}{crane2}
2380: 
2381: Next, consider the random disturbances. Because the
2382: correcting term~$\corr{\sigma}{x}$ is zero, 
2383: there is no difference in using either the
2384: \Ito or the Stratonovich integral. In case of more ``nonlinear'' noise, one of
2385: the integrals must be selected. If the \Ito model is chosen,
2386: Theorem~\ref{prop:gsigmaprop} must be applied.
2387: 
2388: Now we evaluate the conditions of Proposition~\ref{prop:s1sfb1}
2389: to check that the system is linearizable. In fact, we must only
2390: evaluate the non-singularity condition because every one-dimensional
2391: distribution is involutive, and the
2392: integrability is satisfied automatically. To this
2393: end, we will compute the null space (kernel) of the
2394: matrix~$[[f,g],g]$, which is empty and therefore the matrix is
2395: nonsingular. We conclude, that the {\em deterministic\/}
2396: SFB problem is solvable. 
2397: 
2398: Notice, that the system is already in the integrator chain form and
2399: hence~$\lambda = x_1$ satisfies this condition. Therefore, the
2400: {\em deterministic\/} system is linearizable by feedback only, with no
2401: state space transformation at all, \ie $z=T(x)=x$.
2402: 
2403: This choice of the output function~$\lambda $ is natural but does
2404: not cancel the nonlinearity in the dispersion coefficient~$\sigma$.
2405: For this purpose, we must use the algorithm of Section~\ref{prop:cftdp} to construct
2406: another nontrivial coordinate transformation~$T$.
2407: 
2408: To obtain this transformation, we must find the space of all
2409: functions~$\lambda $ satisfying conditions for
2410: feedback linearity~\eqref{eq:7}. Observe
2411: that~$\lie{g}{\lambda }$ must be zero hence
2412: \begin{align}
2413:  \label{eq:75}
2414:  \parby{\lambda }{x_1} g_1 + \parby{\lambda }{x_2} g_2 &= 0
2415:  .\end{align}
2416: 
2417: Since~$g_1=0$ and~$g_2 \ne 0$ in
2418: neighborhood of~$x_0$, then
2419: $\parby{\lambda }{x_2} = 0$ and $\lambda  = c_1(x_1)$ is a
2420: function of~$x_1$ only (\ie without~$x_2$). The coordinate transformation is $T =
2421: \left[\lambda ,\lie{f}{\lambda } \right]^T$. We want to select
2422: such $c_1(x_1)$ that the dispersion
2423: vectorfield~$\tilde \sigma \define \tantra{T}\sigma$ in the new
2424: coordinate system~$z=T(x)$ will be constant:
2425: \begin{align}
2426:  \label{eq:70}
2427:  \parby{c_1(x_1)}{x_1} \frac{F \cos x_1}{m_L l} = \text{constant}
2428:  .\end{align}
2429: 
2430: We decided to define the constant as~${F}/{(m_L l)}$, therefore
2431: \begin{align}
2432:  \label{eq:221}
2433:  \parby{c_1}{x_1} &= \frac{1}{\cos x_1}
2434: \end{align}
2435:  and
2436: \begin{multline}
2437:  \label{eq:71}
2438:  T_1 = \lambda  = c_1(x_1) = \int \frac{1}{\cos x_1} \, d x_1 =\\
2439:  -\ln \left( \cos \frahalf{x_1} - \sin \frahalf{x_1} \right) +\ln
2440:  \left( \cos \frahalf{x_1} + \sin \frahalf{x_1} \right)
2441:  .\end{multline}
2442: \begin{align}
2443:  \label{eq:222}
2444:  T_2 = \lie{f}{\lambda } &= x_2 \sec x_1
2445: .\end{align}
2446: 
2447: Finally, we can compute the feedback from~\eqref{eq:28}. In the
2448: \Mathematica worksheet we validate the results by computing~$\tilde \Theta
2449: = \combinedtab \ssysfgx $, The computation showed that the
2450: system~$\ssys{\hat f}{\hat g}{\hat \sigma}{x}$ is in the integrator
2451: chain form in the~$z$ coordinate chart.  
2452: 
2453: 
2454: \begin{multline}
2455: b=\frac{1}{l \left( m_c + m_l \left( \sin (x_1) \right)^2 \right)}\\
2456: a=\tan ({x_1}) \left( \sec (x_1)\,x_2^2 -
2457:          b\,g\, ( m_c +  m_l ) + l\,m_l\,{x_2}\cos ({x_1}) \right)
2458: .\end{multline}
2459: 
2460: 
2461: \section{Conclusion}
2462: 
2463: \subsection{Main Results}
2464: 
2465: \begin{romanenu} 
2466:  
2467: \item The structure of the stochastic linearization problem is much richer
2468:  than the structure  of the deterministic one. Two definitions of coordinate
2469:  transformation exist and there are differences between the~$g$,~$\sigma$,
2470:  and~$g\sigma$-linearization.
2471: 
2472: \item In the case of \Ito integrals, the coordinate transformation
2473:  laws are of second order (the \Ito rule). 
2474: 
2475: There is a large difference between
2476: $g\sigma$-linearization and $g$-linearization. In the former case the
2477: effect of the \Ito term can be reduced to the first order operator and
2478: consequently the problem is solvable by differential geometry. On the
2479: other hand, in the later case there is no easy method to elliminate
2480: the \Ito term and the a set of second order partial differential
2481: equations must be solved to get the linearizing transformation.
2482: 
2483: 
2484: \item We have given (at least partial) solutions to all
2485:  SISO SFB problems. The results are listed in Table~\ref{tab:sfb}.
2486: 
2487: \begin{table}[htpb]
2488:  \begin{center}
2489: \begin{tabular}{llll}
2490: \hline
2491: Linearization&$g$&$g\sigma$&$\sigma$\\
2492: \hline
2493: Deterministic&\reftpropag{prop:d1sfb2}&&\\ 
2494: Stratonovich &\refppropag{prop:stratgprop}
2495: &\reftpropag{prop:s1sfb1} 
2496: &\refppropag{prop:sfsigma}\\ 
2497: \Ito&\refcorol{prop:sfbooo}
2498: &\reftpropag{prop:gsigmaprop}
2499: &\refppropag{prop:sfsigma}\\ 
2500: \hline
2501: 
2502: \end{tabular}
2503: \caption{Overview of results --- SISO SFB case}
2504: \label{tab:sfb}
2505: \end{center}
2506: \end{table}
2507: 
2508: \item \Ito linearization problems can be approached by means of the
2509:  correcting term. The \Ito differential
2510:  equation can be converted to the Stratonovich equation
2511:  whose behavior under coordinate transformations is simpler.
2512:  This method is only partially applicable to the $g$-linearization.
2513: \item An important special case was identified for the \Ito
2514:   $g$-linearization. The case is characterized by commuting control
2515:   vectorfields~$g$ and dispersion
2516:   vectorfields~$\sigma$.  Solutions can be found
2517:   using  first order methods.
2518: \item Computer algebra proved to be a useful tool for solving exact
2519:  linearization problems. 
2520: \item Industrial applications of the exact linearization in general
2521:  are still unlikely, mainly due to complexity, sensitivity, and
2522:  limited robustness of the control laws designed by the method. 
2523: \end{romanenu}
2524: 
2525: \subsection{Future Research}
2526: 
2527: \begin{romanenu}
2528: \item Find a solution to the \Ito~$g$-linearization problem in general
2529:  case, including geometric criteria, using second order geometry.
2530: \item Analyze the computability issues; implement a universal symbolic
2531:   algebra toolbox for the problem.
2532: \item Solve the SCT problem.
2533: \item Extend the results to the MIMO systems.
2534: \item Extend the results to the input-output problems and
2535:   linearization of autonomous systems. Work out the applications of
2536:   nonlinear filtering.
2537: \item Perhaps, some of the results can be used as a starting
2538:  point for approaching more general class of problems as the problems
2539:  of disturbance decoupling, input invariance of stochastic
2540:  non--linear systems, or problems of reachability and
2541:  observability.
2542: 
2543: \end{romanenu}
2544: 
2545: 
2546: \bibliography{sladecek}
2547: \bibliographystyle{plainnat}
2548: 
2549: 
2550: \end{document}
2551: \endinput
2552: 
2553: 
2554: % LocalWords: ftn Appendices
2555: 
2556: 
2557: 
2558: