1: \documentclass{elsart}
2:
3: \usepackage{natbib}
4: \usepackage{graphicx}
5: \usepackage{amssymb}
6:
7:
8: \newcommand{\cc}{\mbox{\boldmath$c$}}
9: \newcommand{\ee}{\mbox{\boldmath$e$}}
10: \newcommand{\ff}{\mbox{\boldmath$f$}}
11: \newcommand{\hh}{\mbox{\boldmath$h$}}
12: \newcommand{\bl}{\mbox{\boldmath$l$}}
13: \newcommand{\yy}{\mbox{\boldmath$y$}}
14: \newcommand{\xxx}{\mbox{\boldmath$x$}}
15: \newcommand{\xx}{\mbox{\boldmath$x$}}
16: \newcommand{\zz}{\mbox{\boldmath$z$}}
17: \newcommand{\bb}{\mbox{\boldmath$b$}}
18: \newcommand{\bg}{\mbox{\boldmath$g$}}
19: \newcommand{\mm}{\mbox{\boldmath$m$}}
20: \newcommand{\vl}{\mbox{\boldmath$l$}}
21: \newcommand{\BB}{\mbox{\boldmath$B$}}
22: \newcommand{\CC}{\mbox{\boldmath$C$}}
23: \newcommand{\DD}{\mbox{\boldmath$D$}}
24: \newcommand{\FF}{\mbox{\boldmath$F$}}
25: \newcommand{\VV}{\mbox{\boldmath$V$}}
26: \newcommand{\JJ}{\mbox{\boldmath$J$}}
27: \newcommand{\PP}{\mbox{\boldmath$P$}}
28: \newcommand{\HH}{\mbox{\boldmath$H$}}
29: \newcommand{\LL}{\mbox{\boldmath$L$}}
30: \newcommand{\MM}{\mbox{\boldmath$M$}}
31: \newcommand{\NN}{\mbox{\boldmath$N$}}
32: \newcommand{\UU}{\mbox{\boldmath$U$}}
33: \newcommand{\ZZ}{\mbox{\boldmath$Z$}}
34: \newcommand{\bA}{\mbox{\boldmath$A$}}
35: \newcommand{\bOne}{\mbox{\boldmath$1$}}
36: \newcommand{\bZERO}{\mbox{\boldmath$0$}}
37: \newcommand{\bOmega}{\mbox{\boldmath$\Omega$}}
38: \newcommand{\bPi}{\mbox{\boldmath$\Pi$}}
39: \newcommand{\bPsi}{\mbox{\boldmath$\Psi$}}
40: \newcommand{\bDelta}{\mbox{\boldmath$\Delta$}}
41: \newcommand{\bgamma}{\mbox{\boldmath$\gamma$}}
42: \newcommand{\bmu}{\mbox{\boldmath$\mu$}}
43: \newcommand{\bnabla}{\mbox{\boldmath$\nabla$}}
44: \newcommand{\balpha}{\mbox{\boldmath$\alpha$}}
45: \newcommand{\bbeta}{\mbox{\boldmath$\beta$}}
46:
47: \begin{document}
48:
49: \begin{frontmatter}
50:
51:
52: \title{Method of invariant manifold \\for chemical kinetics}
53:
54:
55:
56: \author{Alexander N.\ Gorban}
57: %\ead{agorban@ifp.mat.ethz.ch}
58: \address{Department of Materials, Institute of Polymer Physics\\
59: Swiss Federal Institute of Technology, CH-8092 Z\"urich, Switzerland
60: \\Institute of Computational Modeling RAS, Krasnoyarsk 660036, Russia}
61:
62: \author{Iliya V.\ Karlin}%\corauthref{cor1}}
63: %\ead{ikarlin@ifp.mat.ethz.ch}
64: \address{Department of Materials, Institute of Polymer Physics\\
65: Swiss Federal Institute of Technology, CH-8092 Z\"urich, Switzerland}
66:
67: %\corauth[cor1]{Corresponding author: ETH-Zentrum, Sonneggstrasse 3, ML J 19,
68: %CH-8092 Z\"urich, Switzerland;
69: %Fax: +41 1 632 10 76}
70: %\address{Department of Materials, Institute of Polymer Physics\\
71: % Swiss Federal Institute of Technology, CH-8092 Z\"urich, Switzerland}
72:
73:
74:
75:
76: \begin{abstract}
77: In this paper, we review the construction of low-dimensional manifolds of reduced description
78: for equations of chemical kinetics from the standpoint of the method of invariant
79: manifold (MIM). MIM is based on a formulation of the condition of invariance as
80: an equation, and its solution by Newton iterations.
81: A review of existing alternative methods is extended by
82: a thermodynamically consistent version of the method of intrinsic low-dimensional manifolds.
83: A grid-based version of MIM is developed, and model extensions of low-dimensional
84: dynamics are described. Generalizations to open systems are suggested.
85: The set of methods covered makes it possible to effectively reduce description
86: in chemical kinetics.
87:
88:
89:
90: \end{abstract}
91:
92: \begin{keyword}
93: Kinetics
94: \sep Model Reduction
95: \sep Invariant Manifold
96: \sep Entropy
97: \sep Nonlinear Dynamics
98: \sep Mathematical Modeling
99:
100:
101: \end{keyword}
102:
103: \end{frontmatter}
104: \section{Introduction}
105: In this paper, we present a general method of
106: constructing the reduced description for dissipative systems of reaction
107: kinetics. Our approach is based on the method of invariant manifold which
108: was developed in end of 1980th - beginning of 1990th by
109: \cite{GK92,GK92a,GK92b}.
110: Its realization for a generic
111: dissipative systems was discussed by \cite{GK94,GKIOe01}.
112: This
113: method was applied to a set of problems of classical kinetic
114: theory based on the Boltzmann kinetic equation
115: \citep{GK94,KDN97,KGDN98}.
116: The method of invariant manifold was successfully applied to a derivation of
117: reduced description for kinetic equations of polymeric solutions
118: \citep{ZKD00}.
119: It was also been tested on systems of chemical kinetics \citep{GKZD00}.
120:
121:
122:
123: The goal of nonequilibrium statistical physics is the
124: understanding of how a system with many degrees of freedom
125: acquires a description with a few degrees of freedom.
126: This should lead to reliable methods of extracting
127: the macroscopic description from a detailed microscopic description.
128:
129: Meanwhile this general problem is still far from the final solution, it is
130: reasonable to study simplified models, where, on the one hand, a
131: detailed description is accessible to numerics, on the other hand,
132: analytical methods designed to the solution of problems in real systems
133: can be tested.
134:
135: In this paper we address the well known class of finite-dimensional
136: systems known from the theory of reaction kinetics.
137: These are equations governing a complex relaxation in perfectly stirred closed
138: chemically active mixtures.
139: Dissipative properties of such systems are characterized with a global
140: convex Lyapunov function $G$ (thermodynamic potential) which implements the second law of
141: thermodynamics: As the time $t$ tends to infinity, the system reaches the
142: unique equilibrium state while in the course of the transition the Lyapunov
143: function decreases monotonically.
144:
145: While the limiting behavior of the dissipative systems just described
146: is certainly very simple, there are still interesting questions to be
147: asked about. One of these questions is closely related to the above
148: general problem of nonequilibrium statistical physics. Indeed,
149: evidence of numerical integration of such systems
150: often demonstrates that the relaxation
151: has a certain geometrical structure in the phase space.
152: Namely, typical individual trajectories
153: tend to manifolds of lower dimension, and further proceed to
154: the equilibrium essentially along these manifolds.
155: Thus, such systems demonstrate a dimensional reduction, and therefore
156: establish a more macroscopic description after some time since
157: the beginning of the relaxation.
158:
159: There are two intuitive ideas behind our approach, and we shall now
160: discuss them informally. Objects to be considered below are manifolds
161: (surfaces) $\bOmega$ in the phase space
162: of the reaction kinetic system
163: (the phase space is usually a convex polytope in a finite-dimensional real space).
164: The `ideal' picture of the reduced description we have in mind is as
165: follows: A typical phase trajectory, ${\cc}(t)$,
166: where $t$ is the time, and $\cc$ is an element of the phase space,
167: consists of two pronounced segments. The first segment connects the
168: beginning of the trajectory, $\cc(0)$, with a certain point, $\cc(t_1)$,
169: on the manifold $\Omega$ (rigorously speaking, we should think
170: of $\cc(t_1)$ not on $\bOmega$ but in a small neighborhood of $\bOmega$
171: but this is inessential for the ideal picture). The second segment belongs to $\bOmega$,
172: and connects the point $\cc(t_1)$ with the equilibrium
173: $\cc^{\rm eq}={\cc}(\infty)$, $\cc^{\rm eq}\in\bOmega$.
174: Thus, the manifolds appearing in our ideal picture are
175: ``patterns'' formed by the segments of individual trajectories,
176: and the goal of the reduced description is to ``filter out'' this manifold.
177:
178:
179: There are two important features behind this ideal picture. The first feature
180: is the {\it invariance} of the manifold $\bOmega$: Once the individual
181: trajectory has started on $\bOmega$, it does not leaves $\bOmega$
182: anymore. The second feature is the {\it projecting}: The phase points
183: outside $\bOmega$ will be projected onto $\bOmega$.
184: Furthermore, the dissipativity of the system provides an
185: {\it additional} information about this ideal picture:
186: Regardless of what happens on the manifold $\bOmega$,
187: the function $G$ was decreasing along each individual trajectory
188: before it reached $\bOmega$. This ideal picture is the guide to extract slow invariant
189: manifolds.
190:
191: The paper is organized as follows. In the section \ref{OUTLINE}, we review
192: the reaction kinetics (section \ref{KINETICS}),
193: and discuss the main methods of model reduction in chemical kinetics
194: (section \ref{reduction_review}).
195: In particular, we present two general
196: versions of extending partially equilibrium manifolds
197: to a single relaxation time model in the whole
198: phase space, and develop a thermodynamically consistent version
199: of the intrinsic low-dimensional manifold (ILDM) approach.
200: In the section \ref{MIMG} we introduce the method of invariant manifold
201: in the way appropriate to this class of nonequilibrium systems.
202: In the sections \ref{THERMO} and \ref{DC} we give some
203: details on the two relatively independent parts of the method,
204: the thermodynamic projector, and the iterations for solving the invariance
205: equation. We also introduce a general
206: symmetric linearization procedure
207: for the invariance equation, and discuss its relevance to the picture
208: of decomposition of motions. In the section \ref{MIM}, these two
209: procedures are combined into an unique algorithm. In the section \ref{EX},
210: we demonstrate an illustrative example of computations for a model catalytic
211: reaction.
212: In the section \ref{PARAMETERIZATION}
213: we demonstrate how the thermodynamic
214: projector is constructed without the a priori parameterization of the manifold.
215: This result is essentially used
216: in the section \ref{grid} where we introduce
217: a computationally effective grid-based method to construct invariant manifolds.
218: In the section \ref{open} we describe an extension of the method
219: of invariant manifold to open systems.
220: Finally, results are discussed in the section \ref{conclusion}.
221:
222: \section{Equations of chemical kinetics and their reduction}
223: \label{OUTLINE}
224: \subsection{Outline of the dissipative reaction kinetics}
225: \label{KINETICS}
226: We begin with an outline of the
227: reaction kinetics (for details see e.\ g.\ the book of \cite{YBGE91}).
228: Let us consider a closed system
229: with $n$ chemical species ${\rm A}_1,\dots,{\rm A}_n$,
230: participating in a complex reaction. The complex reaction is
231: represented by the following stoichiometric mechanism:
232: \begin{equation}
233: \label{stoi}
234: \alpha_{s1}{\rm A}_1+\ldots+\alpha_{sn}{\rm A}_n\rightleftharpoons
235: \beta_{s1}{\rm A}_1+\ldots+\beta_{sn}{\rm A}_n,
236: \end{equation}
237: where the index $s=1,\dots,r$ enumerates the reaction steps, and
238: where integers, $\alpha_{si}$ and $\beta_{si}$, are stoichiometric
239: coefficients. For each reaction step $s$, we introduce
240: $n$--component vectors $\balpha_s$ and $\bbeta_s$ with components
241: $\alpha_{si}$ and $\beta_{si}$. Notation
242: $\mbox{\boldmath$\gamma$}_s$ stands for the vector
243: with integer components
244: $\gamma_{si}=\beta_{si}-\alpha_{si}$ (the stoichiometric vector).
245: We adopt an abbreviated notation for the
246: standard scalar product of the $n$-component vectors:
247: \[\langle\xx,\yy\rangle=\sum_{{i=1}}^{n}x_iy_i.\]
248:
249: The system is described by the $n$-component concentration vector
250: $\cc$, where the component $c_i\ge0$ represents the
251: concentration of the specie ${\rm A}_i$. Conservation laws impose linear
252: constraints on admissible vectors $\cc$ (balances):
253: \begin{equation}
254: \label{conser}
255: \langle\bb_i, \cc\rangle=B_i,\ i=1,\dots,l,
256: \end{equation}
257: where $\bb_i$ are fixed and linearly independent vectors,
258: and $B_i$ are given scalars. Let us denote as $\BB$ the set of vectors
259: which satisfy the conservation laws (\ref{conser}):
260: \[
261: \BB=\left\{\cc|\langle\bb_1,\cc\rangle=B_1,\dots,
262: \langle\bb_l,\cc\rangle=B_l\right\}.
263: \]
264: The phase space $\VV$ of the system is the intersection
265: of the cone of $n$-dimensional vectors with nonnegative components,
266: with the set $\BB$, and ${\rm dim}\VV=d=n-l$.
267: In the sequel, we term a vector $\cc\in\VV$ the state
268: of the system. In addition, we assume that each of the conservation
269: laws is supported by each elementary reaction step, that is
270: \begin{equation}
271: \label{sep}
272: \langle\bgamma_s,\bb_i\rangle=0,
273: \end{equation}
274: for each pair of vectors $\bgamma_s$ and $\bb_i$.
275:
276: Reaction kinetic equations describe variations of the states in time.
277: Given the stoichiometric mechanism (\ref{stoi}),
278: the reaction kinetic equations read:
279: \begin{equation}
280: \label{reaction}
281: \dot{\cc}=\JJ(\cc),\
282: \JJ(\cc)=\sum_{s=1}^{r}\bgamma_sW_s(\cc),
283: \end{equation}
284: where dot denotes the time derivative, and $W_s$ is the
285: reaction rate function of the step $s$.
286: In particular, the mass action law suggests the polynomial form of the
287: reaction rates:
288: \begin{equation}
289: \label{MAL}
290: W_s=k^+_s \prod_{i=1}^{n}c_i^{\alpha_i} -
291: k^-_s \prod_{i=1}^{n}
292: c_i^{\beta_i},
293: \end{equation}
294: where $k^+_s$ and $k^-_s$ are the constants of the direct and of
295: the inverse reactions rates of the $s$th reaction step.
296: The phase space $\VV$ is positive-invariant of the system
297: (\ref{reaction}): If $\cc(0)\in\VV$, then $\cc(t)\in\VV$
298: for all the times $t>0$.
299:
300: In the sequel, we assume that the kinetic equation (\ref{reaction})
301: describes evolution towards the unique equilibrium state, $\cc^{\rm eq}$,
302: in the interior of the phase space $\VV$. Furthermore, we assume
303: that there exists a strictly convex function
304: $G(\cc)$ which decreases monotonically in time due to Eq.\
305: (\ref{reaction}):
306: \begin{equation}
307: \label{Htheorem}
308: \dot{G}=
309: \langle\bnabla G(\cc),\JJ(\cc)\rangle
310: \leq 0,
311: \end{equation}
312: Here $\bnabla G$ is the
313: vector of partial derivatives $\partial G/\partial c_i$, and the convexity
314: assumes that the $n\times n$ matrices
315: \begin{equation}
316: \label{MATRIX}
317: \HH_{\cc}=\|\partial^2G(\cc)/\partial c_i\partial c_j\|,
318: \end{equation}
319: are positive definite for all $\cc\in\VV$. In addition, we assume that the
320: matrices (\ref{MATRIX}) are invertible if $\cc$ is taken in
321: the interior of the phase space.
322:
323: The function $G$ is the Lyapunov function of the system
324: (\ref{reaction}), and $\cc^{\rm eq}$ is the point of global minimum of
325: the function $G$ in the phase space $\VV$. Otherwise stated,
326: the manifold of equilibrium states $\cc^{\rm eq}(B_1,\dots,B_l)$
327: is the solution to the variational problem,
328: \begin{equation}
329: \label{EQUILIBRIUM}
330: G\to{\rm min}\ {\rm for\ }\langle\bb_i,\cc\rangle=B_i,\ i=1,\dots,l.
331: \end{equation}
332: For each fixed value of the conserved quantities $B_i$, the solution
333: is unique. In many cases, however, it is convenient to consider
334: the whole equilibrium manifold, keeping the conserved quantities
335: as parameters.
336:
337: For example, for perfect systems in a constant volume under a constant
338: temperature, the Lyapunov function $G$ reads:
339: \begin{equation}
340: \label{gfun}
341: G=\sum_{i=1}^{n}c_i[\ln(c_i/c^{\rm eq}_i)-1].
342: \end{equation}
343:
344:
345: It is important to stress that $\cc^{\rm eq}$ in Eq.\ (\ref{gfun})
346: is an {\it arbitrary} equilibrium of the system,
347: under arbitrary values of the balances. In order to compute $G(\cc)$,
348: it is unnecessary to calculate
349: the specific equilibrium $\cc^{\rm eq}$ which corresponds to the initial state $\cc$.
350: Moreover, for ideal systems, function $G$ is constructed from the thermodynamic
351: data of individual
352: species, and, as the result of this construction, it turns out that it has the form of
353: Eq.\ (\ref{gfun}).
354: Let us mention here the classical formula for the free energy $F=RTVG$:
355: \begin{equation}
356: F=VRT\sum_{i=1}^{n}c_i[(\ln(c_iV_{{\rm Q}\ i})-1)+F_{{\rm int}\ i}(T)],
357: \end{equation}
358: where $V$ is the volume of the system, $T$ is the temperature,
359: $V_{{\rm Q}\ i}= N_0(2\pi\hbar^2/m_ikT)^{3/2}$
360: is the quantum volume of one mole of the specie $A_i$, $N_0$ is the Avogadro number,
361: $m_i$ is the mass of the molecule of ${\rm A}_i$, $R=kN_0$, and $F_{{\rm int}\ i}(T)$ is the
362: free energy of the internal degrees of freedom per mole of ${\rm A}_i$.
363:
364:
365: Finally, we recall an important generalization of the mass action
366: law (\ref{MAL}), known as the Marcelin-De Donder kinetic function.
367: This generalization was developed by \cite{Feinberg72} based on
368: ideas of the thermodynamic theory of affinity \citep{DeDonder36}.
369: We use
370: the kinetic function suggested by \cite{BGY82}.
371: Within this approach, the functions $W_s$ are
372: constructed as follows: For a given strictly convex function $G$,
373: and for a given stoichiometric mechanism (\ref{stoi}), we define the gain
374: ($+$) and the loss ($-$) rates of the $s$th step,
375: \begin{equation}
376: \label{MDD}
377: W_s^{+}=\varphi_s^+
378: \exp[\langle \nabla G,\mbox{\boldmath$\alpha$}_s\rangle],\quad
379: W_s^{-}=\varphi_s^-\exp[\langle \nabla G,\mbox{\boldmath$\beta$}_s\rangle],
380: \end{equation}
381: where $\varphi_s^{\pm}>0$ are kinetic factors.
382: The Marcelin-De Donder kinetic function reads: $W_s=W_s^+-W_s^-$,
383: and the right hand side of the kinetic equation (\ref{reaction}) becomes,
384: \begin{equation}
385: \label{KINETIC MDD}
386: \JJ=\sum_{s=1}^{r}\bgamma_s
387: \{\varphi_s^+
388: \exp[\langle\bnabla G,\balpha_s\rangle]\!-
389: \varphi_s^-\exp[\langle\bnabla G,\bbeta_s\rangle]\}.
390: \end{equation}
391: For the Marcelin-De Donder reaction rate (\ref{MDD}), the dissipation
392: inequality (\ref{Htheorem}) reads:
393: \begin{equation}
394: \label{HMDD}
395: \dot{G}=\sum_{s=1}^{r}
396: [\langle\bnabla G,\bbeta_s\rangle -
397: \langle\bnabla G,\balpha_s\rangle]
398: \left\{\varphi_s^+
399: e^{\langle\bnabla G,\balpha_s\rangle}-
400: \varphi_s^-e^{\langle\bnabla G,\bbeta_s\rangle}\right\}\le 0.
401: \end{equation}
402: The kinetic factors $\varphi_s^{\pm}$ should satisfy certain conditions
403: in order to make valid the dissipation inequality (\ref{HMDD}).
404: A well known sufficient condition is the detail balance:
405: \begin{equation}
406: \label{DB}
407: \varphi_s^+=\varphi_s^-,
408: \end{equation}
409: other sufficient conditions are discussed in detail elsewhere
410: \citep{YBGE91,G84,K89,K93}.
411: For the function $G$ of the form (\ref{gfun}),
412: the Marcelin-De Donder equation casts into the more familiar mass action
413: law form (\ref{MAL}).
414:
415:
416: \subsection{The problem of reduced description in chemical kinetics}
417: \label{reduction_review}
418: What does it mean, ``to reduce the description of a chemical system''?
419: This means the following:
420: \begin{enumerate}
421: \item To shorten the list of species.
422: This, in turn, can be achieved in two ways:
423:
424: (i) To eliminate inessential components from the list;
425:
426: (ii) To lump some of the species into integrated components.
427:
428: \item To shorten the list of reactions. This also can be done in several ways:
429:
430: (i) To eliminate inessential reactions, those which do not significantly influence the reaction
431: process;
432:
433: (ii) To assume that some of the reactions ``have been already completed'', and that
434: the equilibrium has been reached along their paths (this leads to dimensional reduction
435: because the rate constants of the ``completed'' reactions are not used thereafter, what one
436: needs are equilibrium constants only).
437:
438: \item To decompose the motions into fast and slow, into independent (almost-independent)
439: and slaved etc.
440: As the result of such a decomposition, the system admits a study ``in parts''.
441: After that, results of this study
442: are combined into a joint picture. There are several approaches which fall into this category:
443: The famous method of the quasi-steady state (QSS), pioneered by
444: Bodenstein and Semenov and
445: explored in considerable detail by many authors, in particular, by
446: \cite{Bowen63,Chen88,Segel89,Fraser88,Roussel90},
447: and many others;
448: the quasi-equilibrium approximation \citep{Orlov84,G84,Volpert85,Fraser88,K89,K93};
449: methods of sensitivity analysis \citep{Rabitz83,Lam94};
450: methods based on the derivation of the so-called intrinsic low-dimensional manifolds (ILDM,
451: as suggested by \cite{Maas92}). Our method of invariant manifold
452: (MIM, \citep{GK92,GK92a,GK92b,GK94,GKZD00,GKIOe01}) also belongs to this kind of methods.
453:
454: \end{enumerate}
455: Why to reduce description in the times of supercomputers?
456:
457: First, in order to gain understanding. In the process of reducing the description one is often able
458: to extract the essential, and the mechanisms of the processes under study become
459: more transparent. Second, if one is given the detailed description of the system, then one
460: should be able also to solve the initial-value problem for this system. But what should one do
461: in the case where the the system is representing just a point in a three-dimensional flow?
462: The problem of reduction becomes particularly important for modeling the spatially distributed
463: physical and chemical processes. Third, without reducing the kinetic model, it is impossible to
464: construct this model. This statement seems paradoxal only at the first glance: How come,
465: the model is first simplified, and is constructed only after the simplification is done?
466: However, in practice, the typical for a mathematician statement of the problem,
467: (Let the system of differential equations be {\it given}, then ...) is rather rarely applicable
468: in the chemical engineering science for detailed kinetics. Some reactions are known
469: precisely, some other - only hypothetically. Some intermediate species are well studied,
470: some others - not, it is not known much about them. Situation is even worse with the
471: reaction rates. Quite on the contrary, the thermodynamic data
472: (energies, enthalpies, entropies, chemical
473: potentials etc) for sufficiently rarefied systems are quite reliable. Final identification
474: of the model is always done on the basis of comparison with the experiment and with a help
475: of fitting. For this purpose, it is extremely important to reduce the dimension of the system,
476: and to reduce the number of tunable parameters. The normal logics of modeling for
477: the purpose of chemical engineering science is the following: Exceedingly detailed but coarse
478: with respect to parameters system $\to$ reduction $\to$ fitting $\to$ reduced model with
479: specified parameters (cycles are allowed in this scheme, with returns
480: from fitting to more detailed models etc).
481: A more radical viewpoint is also possible:
482: In the chemical engineering science, detailed kinetics is impossible,
483: useless, and it does not exist. For a recently published discussion on this topic see
484: \cite{Levenspiel99,Levenspiel00}; \cite{Y00}.
485:
486: Alas, with a mathematical statement of the problem
487: related to reduction, we all have to begin with the usual:
488: Let the system of differential equations be given ... . Enormous difficulties related to the question
489: of how well the original system is modeling the real kinetics remain out of focus of these studies.
490:
491: Our present work is devoted to studying reductions in a given system of kinetic equations to
492: invariant manifolds of slow motions. We begin with a brief discussion of existing approaches.
493:
494: \subsection{Partial equilibrium approximations}\label{partial_eq}
495:
496: {\it Quasi-equilibrium with respect to reactions} is constructed as follows:
497: From the list of reactions (\ref{stoi}), one selects those which are assumed to equilibrate first.
498: Let they be indexed with the numbers $s_1,\dots,s_k$.
499: The quasi-equilibrium manifold is defined
500: by the system of equations,
501: \begin{equation}
502: \label{st1}
503: W^+_{s_i}=W^-_{s_i},\ i=1,\dots,k.
504: \end{equation}
505: This system of equations looks particularly elegant when written in terms of conjugated (dual)
506: variables, $\bmu=\bnabla G$:
507: \begin{equation}
508: \label{st2}
509: \langle \bgamma_{s_i},\bmu\rangle=0,\ i=1,\dots,k.
510: \end{equation}
511: In terms of conjugated variables,
512: the quasi-equilibrium manifold
513: forms a linear subspace. This subspace, $L^{\perp}$, is the orthogonal
514: completement to the linear
515: envelope of vectors, $L={\rm lin}\{\bgamma_{s_1},\dots,\bgamma_{s_k}\}$.
516:
517: {\it Quasi-equilibrium with respect to species} is constructed practically in the same
518: way but without
519: selecting the subset of reactions. For a given set of species, $A_{i_1}, \dots, A_{i_k}$,
520: one assumes
521: that they evolve fast to equilibrium, and remain there. Formally, this means that in the
522: $k$-dimensional subspace of the space of concentrations with the coordinates
523: $c_{i_1},\dots,c_{i_k}$, one constructs the subspace $L$ which is defined by the balance
524: equations, $\langle \bb_i,\cc\rangle=0$. In terms of the conjugated variables,
525: the quasi-equilibrium manifold, $L^{\perp}$, is defined by equations,
526: \begin{equation}
527: \label{qe1}
528: \bmu\in L^{\perp},\ (\bmu=(\mu_1,\dots,\mu_n)).
529: \end{equation}
530: The same quasi-equilibrium manifold can be also defined with the help of fictitious reactions:
531: Let $\bg_1,\dots,\bg_q$ be a basis in $L$. Then Eq.\ (\ref{qe1}) may be rewritten as follows:
532: \begin{equation}
533: \label{qe2}
534: \langle \bg_i,\bmu\rangle=0,\ i=1,\dots,q.
535: \end{equation}
536:
537: {\it Illustration:} Quasi-equilibrium with respect to reactions in hydrogen oxidation:
538: Let us assume equilibrium with respect to dissociation reactions,
539: ${\rm H}_2\rightleftharpoons 2{\rm H}$, and, ${\rm O}_2\rightleftharpoons 2{\rm O}$,
540: in some subdomain of
541: reaction conditions. This gives:
542: \[k_1^+c_{{\rm H}_2}=k_1^-c^2_{\rm H},\ k_2^+c_{{\rm O}_2}=k_2^-c_{\rm O}^2.\]
543: Quasi-equilibrium with respect to species: For the same reaction, let us assume
544: equilibrium over ${\rm H}$, ${\rm O}$, ${\rm OH}$, and ${\rm H}_2{\rm O}_2$,
545: in a subgomain of reaction conditions.
546: Subspace $L$ is defined by balance constraints:
547: \[ c_{\rm H}+c_{\rm OH}+2c_{{\rm H}_2{\rm O}_2}=0,\ c_{\rm O}+c_{\rm OH}
548: +2c_{{\rm H}_2{\rm O}_2}=0.\]
549: Subspace $L$ is two-dimensional. Its basis, $\{\bg_1,\bg_2\}$ in the coordinates
550: $c_{\rm H}$, $c_{\rm O}$, $c_{\rm OH}$, and $c_{{\rm H}_2{\rm O}_2}$ reads:
551: \[
552: \bg_1=(1,1,-1,0),\quad
553: \bg_2=(2,2,0,-1).
554: \]
555: Corresponding Eq.\ (\ref{qe2}) is:
556: \[ \mu_{\rm H}+\mu_{\rm O}=\mu_{\rm OH},\ 2\mu_{\rm H}+2\mu_{\rm O}=
557: \mu_{{\rm H}_2{\rm O}_2}.\]
558:
559: {\it General construction of the quasi-equilibrium manifold}:
560: In the space of concentration, one defines a subspace $L$ which satisfies
561: the balance constraints:
562: \[ \langle \bb_i,L\rangle\equiv0.\]
563: The orthogonal complement of $L$ in the space with coordinates
564: $\bmu=\bnabla G$ defines then the quasi-equilibrium manifold $\bOmega_{L}$.
565: For the actual computations, one requires the inversion from $\bmu$ to $\cc$.
566: Duality structure $\bmu\leftrightarrow\cc$ is well studied by many authors
567: \citep{Orlov84,DKN97}.
568:
569: {\it Quasi-equilibrium projector.} It is not sufficient to just derive the manifold,
570: it is also required to define a {\it projector} which would transform
571: the vector field defined on the space of concentrations to a vector field on the manifold.
572: Quasi-equilibrium manifold consists of points which minimize $G$ on the affine
573: spaces of the form $\cc+L$. These affine planes are hypothetic planes of fast motions
574: ($G$ is decreasing in the course of the fast motions). Therefore, the quasi-equilibrium
575: projector maps the whole space of concentrations on $\bOmega_L$ parallel to $L$.
576: The vector field is also projected onto the tangent space of $\bOmega_L$ parallel to
577: $L$.
578:
579:
580: Thus, the quasi-equilibrium approximation implies the decomposition of motions
581: into the fast - parallel to $L$, and the slow - along the quasi-equilibrium manifold.
582: In order to construct the quasi-equilibrium approximation, knowledge of reaction rate constants
583: of ``fast'' reactions is not required (stoichiometric vectors of all these fast reaction are
584: in $L$,
585: $\bgamma_{{\rm fast}}\in L$, thus, knowledge of $L$ suffices), one only needs some
586: confidence in that they all are sufficiently fast \citep{Volpert85}.
587: The quasi-equilibrium manifold itself is constructed based on the knowledge of
588: $L$ and of $G$. Dynamics on the quasi-equilibrium manifold is defined as the quasi-equilibrium projection of the ``slow component'' of kinetic equations (\ref{reaction}).
589:
590: \subsection{Model equations}
591:
592: The assumption behind the quasi-equilibrium is the hypothesis of the decomposition of
593: motions into fast and slow. The quasi-equilibrium approximation itself describes slow motions.
594: However, sometimes it becomes necessary to restore to the whole system, and to
595: take into account the fast motions as well. With this, it is desirable to keep intact
596: one of the important advantages of the quasi-equilibrium approximation -
597: its independence of the rate constants of fast reactions.
598: For this purpose, the detailed fast kinetics is replaced by
599: a model equation ({\it single relaxation time approximation}).
600:
601: {\it Quasi-equilibrium models} (QEM) are constructed as follows: For each concentration vector
602: $\cc$, consider the affine manifold, $\cc+L$. Its intersection with the quasi-equilibrium
603: manifold $\bOmega_L$ consists of one point. This point delivers the minimum
604: to $G$ on $\cc+L$. Let us denote this point as $\cc^*_L(\cc)$. The equation of the
605: quasi-equilibrium model reads:
606: \begin{equation}
607: \label{QEmodel}
608: \dot{\cc}=-\frac{1}{\tau}[\cc-\cc^*_L(\cc)]+\sum_{{\rm slow}}\bgamma_{s}W_s(\cc^*_L(\cc)),
609: \end{equation}
610: where $\tau>0$ is the relaxation time of the fast subsystem. Rates of slow reactions are
611: computed in the points $\cc^*_L(\cc)$ (the second term in the
612: right hand side of Eq.\ (\ref{QEmodel}), whereas the rapid motion is taken into account
613: by a simple relaxational term (the first term in the right hand side of Eq.\ (\ref{QEmodel}).
614: The most famous model kinetic equation is the BGK equation in the theory of the
615: Boltzmann equation \citep{BGK}. The general theory
616: of the quasi-equilibrium models, including proofs of their thermodynamic consistency,
617: was constructed by \cite{GK92c,GK94a}.
618:
619:
620: {\it Single relaxation time gradient models} (SRTGM)
621: were considered by \cite{AK00,AK02,AK02a}
622: in the context of the lattice Boltzmann method for hydrodynamics. These models
623: are aimed at improving the obvious drawback of quasi-equilibrium models (\ref{QEmodel}):
624: In order to construct the QEM, one needs to compute
625: the function,
626: \begin{equation}
627: \label{QEA}
628: \cc^*_L(\cc)=\arg\min_{\xxx\in \cc+L,\ \xxx>0}G(\xxx).
629: \end{equation}
630: This is a convex programming problem. It does not always has a closed-form solution.
631:
632: Let $\bg_1,\dots,\bg_k$ is the orthonormal basis of $L$. We denote
633: as $\DD(\cc)$ the $k\times k$ matrix with the elements
634: $\langle \bg_i,\HH_{\cc}\bg_j\rangle$, where $\HH_{\cc}$ is the matrix of second derivatives
635: of $G$ (\ref{MATRIX}). Let $\CC(\cc)$ be the inverse of $\DD(\cc)$. The
636: single relaxation time gradient model has the form:
637: \begin{equation}
638: \dot{\cc}=-\frac{1}{\tau}\sum_{i,j}\bg_i\CC(\cc)_{ij}\langle \bg_j,\bnabla G\rangle
639: +\sum_{{\rm slow}}\bgamma_{s}W_s(\cc).\label{SRTGM}
640: \end{equation}
641: The first term drives the system to the minimum of $G$ on $\cc+L$,
642: it does not require solving the problem (\ref{QEA}), and its
643: spectrum in the quasi-equilibrium is the same as in the quasi-equilibrium model (\ref{QEmodel}).
644: Note that the slow component is evaluated in the ``current'' state $\cc$.
645:
646: The models (\ref{QEmodel}) and (\ref{SRTGM})
647: lift the quasi-equilibrium approximation to a kinetic equation
648: by approximating the fast dynamics with a single ``reaction rate constant'' -
649: relaxation time $\tau$.
650:
651: \subsection{Quasi-steady state approximation}\label{QSS}
652: The quasi-steady state approximation (QSS) is a tool used in a huge amount of works.
653: Let us split the list of species in two groups: The basic and the intermediate (radicals etc).
654: Concentration vectors are denoted accordingly, $\cc^{\rm s}$ (slow, basic species),
655: and $\cc^{\rm f}$ (fast, intermediate species). The concentration vector
656: $\cc$ is the direct sum, $\cc=\cc^{\rm s}\oplus\cc^{\rm f}$.
657: The fast subsystem is Eq.\ (\ref{reaction}) for the component $\cc^{\rm f}$ at fixed
658: values of $\cc^{{\rm s}}$. If it happens that this way defined fast subsystem
659: relaxes to a stationary state, $\cc^{\rm f}\to\cc^{\rm f}_{\rm qss}(\cc^{\rm s})$, then the
660: assumption that $\cc^{\rm f}=\cc^{\rm f}_{\rm qss}(\cc)$ is precisely the QSS assumption.
661: The slow subsystem is the part of the system (\ref{reaction}) for $\cc^{\rm s}$, in the
662: right hand side of which the component $\cc^{\rm f}$ is replaced with
663: $\cc^{\rm f}_{\rm qss}(\cc)$. Thus, $\JJ=\JJ_{\rm s}\oplus\JJ_{\rm f}$, where
664: \begin{eqnarray}
665: \dot{\cc}^{\rm f}&=&\JJ_{\rm f}(\cc^{\rm s}\oplus\cc^{\rm f}), \ \cc^{\rm s}={\rm const};
666: \quad \cc^{\rm f}\to\cc^{\rm f}_{\rm qss}(\cc^{\rm s});\label{fast}\\
667: \dot{\cc}^{\rm s}&=&\JJ_{\rm s}(\cc^{\rm s}\oplus\cc_{\rm qss}^{\rm f}(\cc^{\rm s})).
668: \label{slow}
669: \end{eqnarray}
670: Bifurcations in the system (\ref{fast}) under variation of $\cc^{\rm s}$ as a parameter are
671: confronted to kinetic critical phenomena. Studies of more complicated dynamic phenomena in
672: the fast subsystem (\ref{fast}) require various techniques of averaging, stability analysis of the
673: averaged quantities etc.
674:
675: Various versions of the QSS method are well possible, and are actually used widely,
676: for example, the hierarchical QSS method.
677: There, one defines not a single fast subsystem but a hierarchy of them,
678: $\cc^{{\rm f}_1},\dots,\cc^{{\rm f}_k}$. Each subsystem $\cc^{{\rm f}_i}$ is regarded as
679: a slow system for all the foregoing subsystems, and it is regarded as a fast subsystem for the
680: following members of the hierarchy. Instead of one system of equations (\ref{fast}),
681: a hierarchy of systems of lower-dimensional equations is considered, each of these
682: subsystem is easier to study analytically.
683:
684: Theory of singularly perturbed systems of ordinary differential equations
685: is used to provide a mathematical background and further development of the
686: QSS approximation \citep{Bowen63,Segel89}.
687: In spite of a broad literature on this subject, it remains, in general, unclear, what
688: is the smallness parameter that separates the intermediate (fast) species from
689: the basic (slow). Reaction rate constants cannot be such a parameter (unlike in the
690: case of the quasi-equilibrium). Indeed, intermediate species participate
691: in the {\it same} reactions, as the basic species (for example,
692: ${\rm H}_2\rightleftharpoons 2{\rm H}$, ${\rm H}+{\rm O}_2\rightleftharpoons {\rm OH}+{\rm O}$).
693: It is therefore incorrect to state that $\cc^{\rm f}$ evolve faster than $\cc^{\rm s}$.
694: In the sense of reaction rate constants, $\cc^{\rm f}$ is not faster.
695:
696: For catalytic reactions, it is not difficult to figure out what is the smallness parameter
697: that separates the intermediate species from the basic, and which allows to
698: upgrade the QSS assumption to a singular perturbation theory rigorously
699: \citep{YBGE91}. This smallness parameter is the ratio of balances:
700: Intermediate species include the catalyst, and their total amount is simply
701: significantly less than the amount of all the $\cc_i$'s. After renormalizing to
702: the variables of one order of magnitude, the small parameter appears explicitly.
703:
704: For usual radicals, the origin of the smallness parameter is quite similar.
705: There are much less radicals than the basic species (otherwise, the QSS
706: assumption is inapplicable). In the case of radicals, however, the smallness
707: parameter cannot be extracted directly from balances $B_i$ (\ref{conser}).
708: Instead, one can come up with a thermodynamic estimate: Function $G$
709: decreases in the course of reactions, whereupon we obtain the limiting estimate
710: of concentrations of any specie:
711: \begin{equation}
712: \label{TDlim}
713: c_i\le \max_{G(\cc)\le G(\cc(0))} c_i,
714: \end{equation}
715: where $\cc(0)$ is the initial composition. If the concentration $c_{\rm R}$ of the
716: radical R is small both initially and in the equilibrium, then it should remain small
717: also along the path to the equilibrium. For example, in the case of ideal $G$ (\ref{gfun})
718: under relevant conditions, for any $t>0$, the following inequality is valid:
719: \begin{equation}
720: \label{INEQ_R}
721: c_{\rm R}[\ln(c_{\rm R}(t)/c_{\rm R}^{\rm eq})-1]\le G(\cc(0)).
722: \end{equation}
723: Inequality (\ref{INEQ_R}) provides the simplest (but rather coarse) thermodynamic
724: estimate of $c_{\rm R}(t)$ in terms of $G(\cc(0))$ and $c_{\rm R}^{\rm eq}$
725: {\it uniformly for $t>0$}. Complete theory of thermodynamic estimates
726: of dynamics has been developed by \cite{G84}.
727: One can also do computations without a priori estimations, if one accepts
728: the QSS assumption until the values $\cc^{\rm f}$ stay sufficiently small.
729:
730: Let us assume that an a priori estimate has been found, $c_i(t)\le c_{i\ {\rm max}}$,
731: for each $c_i$. These estimate may depend on the initial conditions, thermodynamic data etc.
732: With these estimates, we are able to renormalize the variables in the kinetic equations
733: (\ref{reaction}) in such a way that renormalized variables take their values
734: from the unit segment $[0,1]$: $\tilde{c}_i=c_i/c_{i\ {\rm max}}$. Then the system (\ref{reaction})
735: can be written as follows:
736: \begin{equation}
737: \label{reduced}
738: \frac{d\tilde{c}_i}{dt}=\frac{1}{c_{i\ {\rm max}}}J_i(\cc).
739: \end{equation}
740: The system of dimensionless parameters, $\epsilon_i=c_{i\ {\rm max}}/\max_i c_{i\ {\rm max}}$
741: defines a hierarchy of relaxation times, and with its help one can establish various
742: realizations of the QSS approximation. The simplest version is the standard QSS assumption:
743: Parameters $\epsilon_i$ are separated in two groups, the smaller ones, and of the order $1$.
744: Accordingly, the concentration vector is split into $\cc^{\rm s}\oplus\cc^{\rm f}$.
745: Various hierarchical QSS are possible, with this, the problem becomes more tractable
746: analytically.
747:
748: Corrections to the QSS approximation can be addressed in various ways
749: (see, e.\ g., \cite{Vasil'eva95,Strygin88}).
750: There exist a variety of ways to introduce the smallness parameter into kinetic equations,
751: and one can find applications to each of the realizations. However, the two particular
752: realizations remain basic for chemical kinetics:
753: (i) Fast reactions (under a given thermodynamic data);
754: (ii) Small concentrations. In the first case, one is led to the quasi-equilibrium approximation,
755: in the second case - to the classical QSS assumption. Both of these approximations allow
756: for hierarchical realizations,
757: those which include not just two but many relaxation time scales. Such a multi-scale approach
758: {\it essentially simplifies} analytical studies of the problem.
759:
760:
761: The method of invariant manifold which we present
762: below in the section \ref{MIM} allows to use both the QE and the QSS as initial approximations
763: in the iterational process of seeking slow invariant manifolds.
764: It is also possible
765: to use a different initial ansatz chosen by a physical intuition, like, for example, the
766: Tamm--Mott-Smith approximation in the theory of strong shock waves
767: \citep{GK92}.
768:
769: \subsection{Methods based on spectral decomposition of Jacobian fields}\label{Jacobi}
770:
771: The idea to use the spectral decomposition of Jacobian fields in the problem
772: of separating the motions into fast and slow originates from methods of analysis of
773: stiff systems \citep{Gear71}, and from methods of sensitivity analysis in control theory
774: \citep{Rabitz83}. There are two basic statements of the problem for
775: these methods:
776: (i) The problem of the slow manifold, and
777: (ii) The problem of a complete decomposition (complete integrability) of kinetic equations.
778: The first of these problems consists in constructing the slow manifold $\bOmega$, and
779: a decomposition of motions into the fast one - towards $\bOmega$, and the slow
780: one - along $\bOmega$ \citep{Maas92}. The second of these problems consists
781: in a transformation of kinetic equations (\ref{reaction}) to a diagonal form,
782: $\dot{\zeta}_i=f_i(\zeta_i)$ (so-called
783: {\it full nonlinear lumping}
784: or {\it modes decoupling}, \cite{Lam94,Li94,Toth97}).
785: Clearly, if one finds a sufficiently explicit solution
786: to the second problem, then the system (\ref{reaction}) is completely integrable, and nothing
787: more is needed, the result has to be simply used. The question is only to what extend
788: such a solution can be possible, and how difficult it would be as compared to the first
789: problem to find it.
790:
791: One of the currently most popular methods is the construction of the so-called
792: {\it intrinsic low-dimensional manifold} (ILDM, \cite{Maas92}).
793: This method is based on the following geometric picture: For each point $\cc$,
794: one defines the Jacobian matrix of
795: Eq.\ (\ref{reaction}), $\FF_{\cc}\equiv \partial \JJ(\cc)/\partial\cc$. One assumes that, in the
796: domain of interest, the eigenvalues of $\FF_{\cc}$ are separated into two groups,
797: $\lambda_i^{\rm s}$ and $\lambda_j^{\rm f}$, and that the following inequalities are valid:
798: \[ {\rm Re}\ \lambda_i^{\rm s}\ge a > b\ge {\rm Re} \lambda_j^{\rm f},\ a\gg b,\ b<0.\]
799: Let us denote as $L_{\cc}^{\rm s}$ and $L_{\cc}^{\rm f}$ the invariant subspaces corresponding
800: to $\lambda^{\rm s}$ and $\lambda^{\rm f}$, respectively, and let $\ZZ_{\cc}^{\rm s}$
801: and $\ZZ_{\cc}^{\rm f}$ be the corresponding spectral projectors,
802: $\ZZ_{\cc}^{\rm s}L_{\cc}^{\rm s}=L_{\cc}^{\rm s}$,
803: $\ZZ_{\cc}^{\rm f}L_{\cc}^{\rm f}=L_{\cc}^{\rm f}$,
804: $\ZZ_{\cc}^{\rm s}L_{\cc}^{\rm f}=\ZZ_{\cc}^{\rm f}L_{\cc}^{\rm s}=\{0\}$,
805: $\ZZ_{\cc}^{\rm s}+\ZZ_{\cc}^{\rm f}=1$.
806: Operator $\ZZ_{\cc}^{\rm s}$ projects onto the subspace of ``slow modes'' $L_{\cc}^{\rm s}$,
807: and it annihilates the ``fast modes'' $L_{\cc}^{\rm f}$. Operator
808: $\ZZ_{\cc}^{\rm f}$ does the opposite, it projects onto fast modes, and it annihilates
809: the slow modes. The basic equation of the ILDM reads:
810: \begin{equation}
811: \label{ILDM}
812: \ZZ_{\cc}^{\rm f}\JJ(\cc)=0.
813: \end{equation}
814: In this equation, the unknown is the concentration vector $\cc$. The set of solutions
815: to Eq.\ (\ref{ILDM}) is the ILDM manifold $\bOmega_{{\rm ildm}}$.
816:
817: For linear systems, $\FF_{\cc}$, $\ZZ_{\cc}^{\rm s}$, and $\ZZ_{\cc}^{\rm f}$, do not depend
818: on $\cc$, and $\bOmega_{{\rm ildm}}=\cc^{\rm eq}+L^{\rm s}$. On the other hand, obviously,
819: $\cc^{\rm eq}\in\bOmega_{{\rm ildm}}$. Therefore, procedures of solving of Eq.\ (\ref{ILDM})
820: can be initiated by choosing the linear approximation,
821: $\bOmega_{{\rm ildm}}^{(0)}=\cc^{\rm eq}+L^{\rm s}_{\cc^{\rm eq}}$,
822: in the neighborhood of the equilibrium
823: $\cc^{\rm eq}$, and then continued parametrically into the nonlinear domain.
824: Computational technologies of a continuation of solutions with respect to
825: parameters are well developed (see, for example, \cite{Khibnik93,Roose90}).
826: The problem of the relevant parameterization is solved locally: In the neighborhood of
827: a given point $\cc^0$ one can choose $\ZZ_{\cc}^{\rm s}(\cc-\cc^0)$ for a characterization
828: of the vector $\cc$. In this case, the space of parameters is $L_{\cc}^{\rm s}$.
829: There exist other, physically motivated ways to parameterize manifolds
830: (\cite{GK92}; see also section \ref{TDP} below).
831:
832: There are two drawbacks of the ILDM method which call for its refinement:
833: (i) {\it ``Intrinsic'' does not imply ``invariant''.} Eq.\ (\ref{ILDM})
834: is not invariant of the dynamics (\ref{reaction}). If one differentiates Eq.\ (\ref{ILDM})
835: in time due to Eq.\ (\ref{reaction}), one obtains a new equation which is the
836: implication of Eq.\ (\ref{ILDM}) {\it only} for linear systems. In a general case, the
837: motion $\cc(t)$ takes off the $\bOmega_{{\rm ildm}}$.
838: Invariance of a manifold $\bOmega$ means that $\JJ(\cc)$ touches $\bOmega$ in every
839: point $\cc\in\bOmega$. It remains unclear how the ILDM (\ref{ILDM}) corresponds with
840: this condition. Thus, from the dynamical perspective, the status of the
841: ILDM remains not well defined, or ``ILDM is ILDM'', defined self-consistently
842: by Eq.\ (\ref{ILDM}), and that is all what can be said about it.
843: (ii) From the geometrical standpoint, spectral decomposition of Jacobian fields
844: is not the most attractive way to compute manifolds.
845: If we are interested in the behavior of trajectories, how they converge or diverge,
846: then one should consider the symmetrized part of $\FF_{\cc}$, rather than
847: $\FF_{\cc}$ itself.
848:
849:
850: Symmetric part, $\FF_{\cc}^{\rm sym}=(1/2)(\FF_{\cc}^{\dag}+\FF_{\cc})$,
851: defines the dynamics of the distance between two solutions, $\cc$ and $\cc'$,
852: in a given local Euclidean metrics.
853: Skew-symmetric part defines rotations.
854: If we want to study manifolds based on the argument about convergence/divergence
855: of trajectories, then we should use in Eq.\ (\ref{ILDM})
856: the spectral projector $\ZZ_{\cc}^{\rm f sym}$ for the
857: operator $\FF_{\cc}^{\rm sym}$. This, by the way, is also a significant simplification
858: from the standpoint of computations.
859: It remains to choose the metrics.
860: This choice is unambiguous from the thermodynamic perspective.
861: In fact, there is only one choice which fits into the physical meaning of the problem,
862: this is the metrics associated with the thermodynamic (or entropic) scalar product,
863: \begin{equation}
864: \label{ESP}
865: \langle\langle\xxx,\yy\rangle\rangle=\langle\xxx,\HH_{\cc}\yy\rangle,
866: \end{equation}
867: where $\HH_{\cc}$ is the matrix of second-order derivatives
868: of $G$ (\ref{MATRIX}). In the equilibrium, operator $\FF_{\cc^{\rm eq}}$ is
869: selfajoint with respect to this scalar product (Onsager's reciprocity relations).
870: Therefore, the behavior of the ILDM in the vicinity of the equilibrium does
871: not alter under the replacement, $\FF_{\cc^{\rm eq}}=\FF_{\cc^{\rm eq}}^{\rm sym}$.
872: In terms of usual matrix representation, we have:
873: \begin{equation}
874: \FF_{\cc}^{\rm sym}=\frac{1}{2}(\FF_{\cc}+\HH_{\cc}^{-1}\FF_{\cc}^{T}\HH_{\cc}),\label{Fsym}
875: \end{equation}
876: where $\FF_{\cc}^{T}$ is the ordinary transposition.
877:
878: The ILDM constructed with the help of the symmetrized Jacobian field will be termed
879: the {\it symmetric entropic intrinsic low-dimensional manifold} (SEILDM).
880: Selfadjointness of $\FF_{\cc}^{\rm sym}$ (\ref{Fsym}) with respect to the
881: thermodynamic scalar product (\ref{ESP}) simplifies considerably computations of
882: spectral decomposition. Moreover, it becomes necessary to do spectral decomposition
883: in only one point - in the equilibrium. Perturbation theory for selfadjoint operators
884: is a very well developed subject \citep{Kato76}, which makes it possible to easily extend the
885: spectral decomposition with respect to parameters.
886: A more detailed discussion of the selfadjoint linearization will be given below in section
887: \ref{SA}.
888:
889: Thus, when the geometric picture behind the decomposition of motions is specified,
890: the physical significance of the ILDM becomes more transparent, and it leads to
891: its modification into the SEILDM. This also gains simplicity in the implementation
892: by switching from non-selfadjoint spectral problems to selfadjoint. The
893: quantitative estimate of this simplification is readily available: Let $d$ be the dimension
894: of the phase space, and $k$ the dimension of the ILDM ($k={\rm dim} L_{\cc}^{\rm s}$).
895: The space of all the projectors $\ZZ$ with the $k$-dimensional image
896: has the dimension $D=2k(d-k)$. The space of all the selfadjoint projectors with the
897: $k$-dimensional image has the dimension $D^{\rm sym}=k(d-k)$.
898: For $d=20$ and $k=3$, we have $D=102$ and $D^{\rm sym}=51$.
899: When the spectral decomposition by means of parametric extension is addressed,
900: one considers equations of the form:
901: \begin{equation}
902: \label{parametric}
903: \frac{d\ZZ_{\cc(\tau)}^{\rm s}}{d\tau}=\bPsi^{\rm s}\left(\frac{d\cc}{d\tau},
904: \ZZ_{\cc(\tau)}^{\rm s}, \FF_{\cc(\tau)}, \bnabla\FF_{\cc(\tau)}\right),
905: \end{equation}
906: where $\tau$ is the parameter, and $\bnabla\FF_{\cc}=\bnabla\bnabla\JJ(\cc)$
907: is the differential of the Jacobian field.
908: For the selfadjoint case, where we use $=\FF_{\cc}^{\rm sym}$ instead of
909: $\FF_{\cc}$, this system of equations
910: has twice less independent variables, and also the right hand is of a simpler structure.
911:
912: It is more difficult to improve on the first of the remarks (ILDM is not invariant).
913: The following naive approach may seem possible:
914:
915: (i) Take $\bOmega_{\rm ildm}=\cc^{\rm eq}+L^{\rm s}_{\cc^{\rm eq}}$ in a neighborhood
916: $U$ of the equilibrium $\cc^{\rm eq}$. [This is also a useful initial approximation
917: for solving Eq.\ (\ref{ILDM})].
918:
919: (ii) Instead of computing the solution to Eq.\ (\ref{ILDM}), integrate the kinetic
920: equations (\ref{reaction}) {\it backwards in the time}. It is sufficient to take
921: initial conditions $\cc(0)$ from a dense set on the boundary,
922: $\partial U\cap (\cc^{\rm eq}+L^{\rm s}_{\cc^{\rm eq}})$, and to compute
923: solutions $\cc(t)$, $t<0$.
924:
925: (iii) Consider the obtained set of trajectories as an approximation of the slow invariant manifold.
926:
927: This approach will guarantee invariance, by construction, but it is prone to pitfalls in what
928: concerns the slowness. Indeed, the integration backwards in the time will see
929: exponentially divergent trajectories, if they were exponentially converging in the
930: normal time progress. This way one finds {\it some} invariant manifold which touches
931: $\cc^{\rm eq}+L^{\rm s}_{\cc^{\rm eq}}$ in the equilibrium. Unfortunately, there are infinitely
932: many such manifolds, and they fill out almost all the space of concentrations.
933: However, we must select the slow component of motions. Such a regularization is possible.
934: Indeed, let us replace in Eq.\ (\ref{reaction}) the vector field $\JJ(\cc)$
935: by the vector field $\ZZ_{\cc}^{{\rm s sym}}\JJ(\cc)$, and obtain a regularized
936: kinetic equation,
937: \begin{equation}
938: \label{regreaction}
939: \dot{c}=\ZZ_{\cc}^{{\rm s sym}}\JJ(\cc).
940: \end{equation}
941: Let us replace integration backwards in time of the kinetic equation (\ref{reaction})
942: in the naive approach described above by integration backwards in time of the
943: regularized kinetic equation (\ref{regreaction}). With this, we obtain a rather
944: convincing version of the ILDM (SEILDM).
945: Using Eq.\ (\ref{parametric}), one also can write down an
946: equation for the projector $\ZZ_{\cc}^{{\rm s sym}}$, putting $\tau=t$.
947: Replacement of Eq.\ (\ref{reaction}) by Eq.\ (\ref{regreaction})
948: also makes the integration backwards in time in the naive approach more stable.
949: However, {\it regularization will again conflict with invariance}.
950: The ``naive refinement'' after the regularization (\ref{regreaction})
951: produces just a slightly different version of the ILDM (or SEILDM) but
952: it does not construct the slow invariant manifold.
953: So, where is the way out? We believe that the ILDM and its version SEILDM
954: are, in general, good initial approximations of the slow manifold. However,
955: if one is indeed interested in finding the invariant manifold, one has to write
956: out the true condition of invariance and solve it. As for the initial approximation
957: for the method of invariant manifold one can use any ansatz, in particular,
958: the SEILDM.
959:
960:
961: {\it The problem of a complete decomposition} of kinetic equations can be solved indeed in
962: some cases. The first such solution was the spectral decomposition for linear systems
963: \citep{Wei62}. Decomposition is sometimes possible also for
964: nonlinear systems (\cite{Li94}; \cite{Toth97}).
965: The most famous example of a complete decomposition
966: of infinite-dimensional kinetic equation is the complete integrability
967: of the space-independent Boltzmann equation for Maxwell`s molecules found
968: by \cite{Bobylev88}.
969: However, in a general case, there exist no analytical, not even a twice differentiable
970: transformation which would decouple modes. The well known Grobman-Hartman
971: theorem \citep{Hartman63,Hartman82}
972: states only the existence of a continuous transform which decomposes
973: modes in a neighborhood of the equilibrium.
974: For example, the analytic planar system,
975: $dx/dt=-x$, $dy/dt=-2y+x^2$, is not $C^2$ linearizable. These problems remain of interest
976: \citep{Chicone00}.
977: Therefore, in particular, it becomes quite ineffective to construct such a transformation in a
978: form of a series.
979: It is more effective to solve a simpler problem of extraction of a slow invariant manifold
980: \citep{Beyn98}.
981:
982: {\it Sensitivity analysis} \citep{Rabitz83,Rabitz87,Lam94}
983: makes it possible to select essential variables and reactions, and to decompose motions into
984: fast and slow. In a sense, the ILDM method is a development
985: of the sensitivity analysis. Recently, a further step in this direction was done by
986: \cite{Zhu99}. In this work, the authors use a {\it nonlocal in time criterion of
987: closeness of solutions} of the full and of the reduced systems of chemical kinetics.
988: They require not just a closeness of derivatives but a true closeness of the dynamics.
989:
990: Let us be interested in the dynamics of the concentrations of just a few species,
991: ${\rm A}_1,\dots, {\rm A}_{p}$, whereas the rest of the species, ${\rm A}_{p+1},
992: \dots, {\rm A}_{n}$ are used for building the kinetic equation, and for understanding
993: the process. Let $\cc_{\rm goal}$ be the concentration vector with components
994: $c_1,\dots, c_p$, $\cc_{\rm goal}(t)$ be the corresponding components of the
995: solution to Eq.\ (\ref{reaction}),
996: and $\cc^{\rm red}_{\rm goal}$ be the solution to the simplified model with
997: corresponding initial conditions.
998: \cite{Zhu99} suggest to minimize the difference between
999: $\cc_{\rm goal}(t)$ and $\cc^{\rm red}_{\rm goal}$ on the
1000: segment $t\in[0,T]$: $\|\cc_{\rm goal}(t)-\cc^{\rm red}_{\rm goal}\|\to\min$.
1001: In the course of the optimization under certain restrictions one
1002: selects the optimal (or appropriate) reduced model.
1003: The sequential quadratic programming method and heuristic rules of
1004: sorting the reactions, substances etc were used.
1005: In the result, for some stiff systems studied, one avoids typical pitfalls
1006: of the local sensitivity analysis. In simpler situations this method should give
1007: similar results as the local methods.
1008:
1009:
1010:
1011: \subsection{Thermodynamic criteria for selection of important reactions}
1012: One of the problems addressed by the sensitivity analysis is the selection of the important
1013: and discarding the unimportant reactions. \cite{BYA77}
1014: suggested a simple principle to compare importance of different reactions according
1015: to their contribution to the entropy production (or, which is the same, according to
1016: their contribution to $\dot{G}$). Based on this principle, \cite{Dimitrov82} described domains
1017: of parameters in which the reaction of hydrogen oxidation, ${\rm H}_2+{\rm O}_2+{\rm M}$,
1018: proceeds due to different mechanisms. For each elementary reaction, he has derived the domain
1019: inside which the contribution of this reaction is essential (nonnegligible).
1020: Due to its simplicity, this entropy production principle is especially well suited
1021: for analysis of complex problems.
1022: In particular, recently, a version of the entropy production principle
1023: was used in the problem of selection of boundary conditions for Grad's moment equations
1024: \citep{Struchtrup98,GKZ02}. For ideal systems (\ref{gfun}), the contribution of the
1025: $s$th reaction to $\dot{G}$ has a particularly simple form:
1026: \begin{equation}
1027: \label{dotGs}
1028: \dot{G}_{s}=-W_s\ln\left(\frac{W_s^+}{W_s^-}\right),\ \dot{G}=\sum_{s=1}^{r}\dot{G}_s.
1029: \end{equation}
1030: For nonideal systems, the corresponding expressions (\ref{HMDD})
1031: are also not too complicated.
1032:
1033:
1034: \section{Outline of the method of invariant manifold}
1035: \label{MIMG}
1036: In many cases, dynamics of the $d$-dimensional system (\ref{reaction})
1037: leads to a manifold of a lower dimension. Intuitively, a typical
1038: phase trajectory behaves as follows: Given the initial state
1039: $\cc(0)$ at $t=0$, and after some period of time, the trajectory comes
1040: close to some low-dimensional manifold $\bOmega$, and after that
1041: proceeds towards the equilibrium essentially along this manifold.
1042: The goal is to construct this manifold.
1043:
1044: The starting point of our approach is based on a formulation of
1045: the two main requirements:
1046:
1047: (i). {\it Dynamic invariance}: The manifold $\bOmega$
1048: should be (positively) invariant under the dynamics of the originating system
1049: (\ref{reaction}): If $\cc(0)\in\bOmega$, then $\cc(t)\in\bOmega$ for
1050: each $t>0$.
1051:
1052: (ii). {\it Thermodynamic consistency of the reduced dynamics}:
1053: Let {\it some} (not obligatory invariant)
1054: manifold $\bOmega$ is considered as a manifold of
1055: reduced description. We should define a set of linear operators,
1056: $\PP_{\cc}$, labeled by the states $\cc\in\bOmega$, which
1057: project the vectors $\JJ(\cc)$, $\cc\in\bOmega$ onto the
1058: tangent bundle of the manifold $\bOmega$, thereby
1059: generating the induced vector field,
1060: $\PP_{\cc}\JJ(\cc)$, $\cc\in\bOmega$.
1061: This induced vector field on the tangent bundle of the manifold
1062: $\bOmega$ is identified with the reduced dynamics
1063: along the manifold $\bOmega$.
1064: The thermodynamicity requirement for this induced vector field reads
1065: \begin{equation}
1066: \label{thermo}
1067: \langle\bnabla G(\cc),\PP_{\cc}\JJ(\cc)\rangle\leq 0,\mbox{ for\ each\ }\cc\in\bOmega.
1068: \end{equation}
1069:
1070: In order to meet these requirements, the method of invariant
1071: manifold suggests two complementary procedures:
1072:
1073: (i). To treat the condition of dynamic invariance as an equation,
1074: and to solve it iteratively by a Newton method.
1075: This procedure is geometric in its nature, and it does not
1076: use the time dependence and small parameters.
1077:
1078: (ii). Given an approximate manifold of reduced description,
1079: to construct the projector satisfying the condition
1080: (\ref{thermo}) in a way which does not depend on the vector
1081: field $\JJ$.
1082:
1083: We shall now outline both these procedures starting with the second.
1084: The solution consists, in the first place, in formulating the
1085: {\it thermodynamic condition} which should
1086: be met by the projectors $\PP_{\cc}$:
1087: For each $\cc\in\bOmega$, let us consider the linear functional
1088: \begin{equation}
1089: \label{FUNC}
1090: M^*_{\cc}(\xx)=\langle\bnabla G(\cc),\xx\rangle.
1091: \end{equation}
1092: Then the thermodynamic condition for the projectors reads:
1093: \begin{equation}
1094: \label{therm1}
1095: {\rm ker}\PP_{\cc}\subseteq{\rm ker}M^*_{\cc},\
1096: {\rm for\ each}\ \cc\in\bOmega.
1097: \end{equation}
1098: Here ${\rm ker}\PP_{\cc}$ is the null space of the projector, and
1099: ${\rm ker}M^*_{\cc}$ is the hyperplane orthogonal to the vector
1100: $M^*_{\cc}$.
1101: It has been shown \citep{GK92,GK94}
1102: that the condition (\ref{therm1}) is
1103: the necessary and sufficient condition to establish the thermodynamic
1104: induce vector field on the given manifold $\bOmega$ for all possible
1105: dissipative vector fields $\JJ$ simultaneously.
1106:
1107: Let us now turn to the requirement of invariance.
1108: By a definition, the manifold $\bOmega$ is invariant
1109: with respect to the vector field $\JJ$ if and only if
1110: the following equality is true:
1111: \begin{equation}
1112: \label{INVARIANCE}
1113: \left[1-\PP\right]\JJ(\cc)=0,\ {\rm for\ each\ } \cc\in\bOmega.
1114: \end{equation}
1115: In this expression $\PP$ is an {\it arbitrary}
1116: projector on the tangent
1117: bundle of the manifold $\bOmega$.
1118: It has been suggested to consider the condition (\ref{INVARIANCE}) as an
1119: {\it equation} to be solved iteratively starting with some appropriate initial
1120: manifold.
1121:
1122: Iterations for the invariance equation (\ref{INVARIANCE})
1123: are considered in the section \ref{DC}.
1124: The next section presents construction of the thermodynamic projector
1125: using a specific parameterization of manifolds.
1126:
1127: \section{Thermodynamic projector}
1128: \label{THERMO}
1129: \subsection{Thermodynamic parameterization}
1130: \label{TDP}
1131: In this section, $\bOmega$ denotes a generic $p$--dimensional
1132: manifold. First, it should be mentioned that {\it any}
1133: parameterization of $\bOmega$ generates a certain projector, and
1134: thereby a certain reduced dynamics.
1135: Indeed, let us consider a set of $m$ independent functionals
1136: $M(\cc)=\{M_1(\cc),\dots,M_p(\cc)\}$,
1137: and let us assume that they form a
1138: coordinate system on $\bOmega$ in such a way that
1139: $\bOmega=\cc(M)$, where $\cc(M)$ is a vector function of the parameters
1140: $M_1,\dots,M_p$. Then the projector associated with
1141: this parameterization reads:
1142: \begin{equation}
1143: \label{proj}
1144: \PP_{\cc(M)}\xx=\sum_{i,j=1}^p\frac{\partial\cc(M)}{\partial M_i}
1145: N^{-1}_{ij}(M)
1146: \langle\bnabla M_j\bigm|_{\cc(M)},\xx\rangle,
1147: \end{equation}
1148: where $N^{-1}_{ij}$ is the inverse to the $p\times p$
1149: matrix:
1150: \begin{equation}
1151: \label{NORMALIZATION}
1152: \NN(M)=\|\langle \bnabla M_i,\partial\cc/\partial M_j\rangle\|.
1153: \end{equation}
1154: This somewhat involved notation is intended to stress that the projector
1155: (\ref{proj}) is dictated by the choice of the parameterization.
1156: Subsequently, the induced vector field of the reduced dynamics
1157: is found by applying projectors (\ref{proj}) on the vectors
1158: $\JJ(\cc(M))$, thereby inducing the reduced dynamics in terms
1159: of the parameters $M$ as follows:
1160: \begin{equation}
1161: \label{dyn}
1162: \dot{M}_i=\sum_{j=1}^pN^{-1}_{ij}(M)
1163: \langle\bnabla M_j\bigm|_{\cc(M)},\JJ(\cc(M))\rangle,
1164: \end{equation}
1165: Depending on the choice of the parameterization,
1166: dynamic equations (\ref{dyn}) are (or are not) consistent
1167: with the thermodynamic requirement (\ref{thermo}).
1168: The {\it thermodynamic parameterization} makes use of the condition
1169: (\ref{therm1}) in order to establish the thermodynamic projector.
1170: Specializing to the case (\ref{proj}), let us consider the linear functionals,
1171: \begin{equation}
1172: \label{derivative}
1173: DM_i\bigm|_{\cc(M)}(\xx)=
1174: \langle\bnabla M_i\bigm|_{\cc(M)},\xx\rangle.
1175: \end{equation}
1176: Then the condition (\ref{therm1}) takes the form:
1177: \begin{equation}
1178: \label{therm2}
1179: \bigcap_{i=1}^p{\rm ker}DM_i\bigm|_{\cc(M)}\subseteq{\rm ker}
1180: M^*_{\cc(M)},
1181: \end{equation}
1182: that is, the intersection of null spaces of the functionals (\ref{derivative})
1183: should belong to the null space of the differential of the
1184: Lyapunov function $G$, in each point of the manifold $\bOmega$.
1185:
1186: In practice, in order to construct the thermodynamic
1187: parameterization, we take the following set of
1188: functionals in each point $\cc$ of the manifold $\bOmega$:
1189: \begin{eqnarray}
1190: \label{param_thermo}
1191: M_1(\xx)&=&M_{\cc}^*(\xx),\ \cc\in\bOmega\\
1192: M_i(\xx)&=&\langle\mm_i,\xx\rangle,\ i=2,\dots,p
1193: \label{param_other}
1194: \end{eqnarray}
1195: It is required that vectors $\bnabla G(\cc), \mm_2,\dots,\mm_p$
1196: are linearly independent in each state $\cc\in\bOmega$.
1197: Inclusion of the functionals (\ref{FUNC}) as a part of the
1198: system (\ref{param_thermo}) and (\ref{param_other})
1199: implies the thermodynamic condition
1200: (\ref{therm2}). Also, any linear combination of the parameter
1201: set (\ref{param_thermo}), (\ref{param_other}) will meet the thermodynamicity requirement.
1202:
1203: It is important to notice here that the thermodynamic
1204: condition is satisfied whatsoever the functionals $M_2,\dots,M_p$ are.
1205: This is very convenient for it gives an opportunity to take into account
1206: the conserved quantities correctly.
1207: The manifolds we are going to deal with should be consistent
1208: with the conservation laws (\ref{conser}). While the explicit
1209: characterization of the phase space $\VV$ is a problem on its own,
1210: in practice, it is customary to work in the $n$--dimensional space
1211: while keeping the constraints (\ref{conser}) explicitly on each step
1212: of the construction. For this technical reason, it is convenient to
1213: consider manifolds of the dimension $p>l$, where $l$ is the number of conservation
1214: laws, in the $n$--dimensional space rather than in the phase space
1215: $\VV$. The thermodynamic parameterization is then concordant also
1216: with the conservation laws if $l$ of the linear functionals
1217: (\ref{param_other}) are identified with the conservation laws.
1218: In the sequel, only projectors consistent
1219: with conservation laws are considered.
1220:
1221:
1222: Very frequently, the manifold $\bOmega$ is represented as a
1223: $p$-parametric family $\cc(a_1,\dots,a_p)$, where
1224: $a_i$ are coordinates on the manifold. The thermodynamic
1225: {\it re-parameterization}
1226: suggests a representation of the coordinates $a_i$ in terms of
1227: $M_{\cc}^*, M_2,\dots,M_p$ (\ref{param_thermo}), (\ref{param_other}).
1228: While the explicit construction of these functions may be
1229: a formidable task, we notice that the construction
1230: of the thermodynamic projector of the form (\ref{proj}) and of the
1231: dynamic equations (\ref{dyn}) is relatively easy because only
1232: the derivatives $\partial\cc/\partial M_i$ enter these expressions.
1233: This point was discussed in a detail by \cite{GK92,GK94}.
1234:
1235: \subsection{Decomposition of motions: Thermodynamics}
1236: \label{DEC_THERM}
1237: Finally, let us discuss how the thermodynamic projector is related to the
1238: decomposition of motions.
1239: {\it Assuming} that the decomposition of motions near the manifold
1240: $\bOmega$ is true indeed, let us consider states which were initially
1241: close enough to the manifold $\bOmega$. Even without knowing
1242: the details about the evolution of the states towards $\bOmega$,
1243: we know that the Lyapunov function $G$ was decreasing in the
1244: course of this evolution. Let us consider a set of states $\UU_{\cc}$
1245: which contains all those vectors $\cc'$ that
1246: have arrived (in other words, have been projected) into the point
1247: $\cc\in\bOmega$. Then we observe that the
1248: state $\cc$ furnishes the minimum of the function $G$ on the set $\UU_{\cc}$.
1249: If a state $\cc'\in\UU_{\cc}$, and if it deviates small enough
1250: from the state $\cc$ so that the linear
1251: approximation is valid, then $\cc'$ belongs to the affine hyperplane
1252: \begin{equation}
1253: \label{hyperplane}
1254: \Gamma_{\cc}=\cc+{\rm ker\ }M_{\cc}^*,\ \cc\in\bOmega.
1255: \end{equation}
1256: This hyperplane actually participates in the condition (\ref{therm1}).
1257: The consideration was entitled `thermodynamic' \citep{GK92}
1258: because it describes the states $\cc\in\bOmega$ as points of minimum
1259: of the function $G$ over the corresponding hyperplanes
1260: (\ref{hyperplane}).
1261:
1262: \section{Corrections}
1263: \label{DC}
1264: \subsection{Preliminary discussion}
1265: The thermodynamic projector is needed to induce the dynamics on a
1266: given manifold in such a way that the dissipation inequality (\ref{thermo})
1267: holds. Coming back to the issue of constructing corrections, we should
1268: stress that the projector participating in the invariance condition
1269: (\ref{INVARIANCE}) is arbitrary. It is convenient to make
1270: use of this point: When Eq.\ (\ref{INVARIANCE}) is solved iteratively, the
1271: projector may be kept non--thermodynamic unless the induced dynamics
1272: is explicitly needed.
1273:
1274: Let us assume that we have chosen
1275: the initial manifold, $\bOmega_0$, together with the associated
1276: projector $\PP_0$, as the first approximation to the
1277: desired manifold of reduced description. Though the choice
1278: of the initial approximation $\bOmega_0$ depends on the specific
1279: problem, it is often reasonable to consider quasi-equilibrium or
1280: quasi steady-state approximations.
1281: In most cases, the manifold $\bOmega_0$ is not
1282: an invariant manifold. This means that $\bOmega_0$ does not
1283: satisfy the invariance condition
1284: (\ref{INVARIANCE}):
1285: \begin{equation}
1286: \label{DEFECT}
1287: \bDelta_0=[1-\PP_0]\JJ(\cc_0)\ne0,\
1288: {\rm for\ some\ } \cc_0\in\bOmega_0.
1289: \end{equation}
1290: Therefore, we seek a correction $\cc_1=\cc_0+\delta\cc$.
1291: Substituting $\PP=\PP_0$ and $\cc=\cc_0+\delta\cc$ into the
1292: invariance equation (\ref{INVARIANCE}), and after the linearization in
1293: $\delta\cc$, we derive the following linear equation:
1294: \begin{equation}
1295: \label{METHOD}
1296: \left[1-\PP_{0}\right]\left[\JJ(\cc_0)+
1297: \LL_{\cc_0}\delta\cc\right]=0,
1298: \end{equation}
1299: where $\LL_{c_0}$ is the matrix of first derivatives of the
1300: vector function $\JJ$, computed in the state $\cc_0\in\bOmega_0$.
1301: The system of linear algebraic equations (\ref{METHOD}) should be
1302: supplied with the additional condition.
1303: \begin{equation}
1304: \label{uni_con}
1305: \PP_0\delta\cc=0.
1306: \end{equation}
1307:
1308: In order to illustrate the nature of the Eq.\ (\ref{METHOD}), let us consider
1309: the case of linear manifolds for linear systems.
1310: Let a linear evolution equation is given in the finite-dimensional real space:
1311: $\dot{\cc}=\LL\cc$, where $\LL$ is negatively definite symmetric matrix
1312: with a simple spectrum.
1313: Let us further assume the quadratic Lyapunov function,
1314: $G(\cc)=\langle\cc,\cc\rangle$. The manifolds we consider are lines,
1315: $\vl(a)=a\ee$, where $\ee$ is the unit vector, and $a$ is a scalar.
1316: The invariance equation for such manifolds reads:
1317: $\ee\langle\ee,\LL\ee\rangle-\LL\ee=0$,
1318: and is simply a form of the eigenvalue problem for the operator $\LL$.
1319: Solutions to the latter equation are eigenvectors $\ee_{i}$,
1320: corresponding to eigenvalues $\lambda_{i}$.
1321:
1322: Assume that we have chosen a line, $\vl_0=a\ee_0$, defined by the unit
1323: vector $\ee_0$, and that $e_0$ is not an eigenvector of $\LL$.
1324: We seek another line, $\vl_1=a\ee_1$, where $\ee_1$ is another unit
1325: vector, $\ee_1=\yy_1/\|\yy_1\|$, $\yy_1=\ee_0+\delta\yy$. The additional
1326: condition (\ref{uni_con}) now reads: $\langle\delta\yy,\ee_0\rangle=0$.
1327: Then the Eq.\ (\ref{METHOD}) becomes
1328: $[1-\ee_0\langle\ee_0,\cdot\rangle]L[\ee_0+\delta\yy]=0$.
1329: Subject to the additional condition, the
1330: unique solution is as follows: $\ee_0+\delta\yy=
1331: \langle\ee_0,\LL^{-1}\ee_0\rangle^{-1}\LL^{-1}\ee_0$.
1332: Rewriting the latter expression in the eigen--basis of $\LL$,
1333: we have: $\ee_0+\delta\yy\propto
1334: \sum_{i}\lambda_i^{-1}\ee_i\langle\ee_i,\ee_0\rangle$.
1335: The leading term in this sum corresponds to the
1336: eigenvalue with the minimal absolute value.
1337: The example indicates that
1338: the method of linearization (\ref{METHOD}) seeks
1339: the direction of the {\it slowest relaxation}.
1340: For this reason, the method (\ref{METHOD}) can be
1341: recognized as the basis of an iterative method for
1342: constructing the manifolds of slow motions.
1343:
1344: For the nonlinear systems, the matrix $\LL_{c_0}$ in the Eq.\ (\ref{METHOD})
1345: depends nontrivially on $\cc_0$. In this case
1346: the system (\ref{METHOD}) requires a further specification
1347: which will be done now.
1348:
1349: \subsection{Symmetric linearization}
1350: \label{SA}
1351: The invariance condition
1352: (\ref{INVARIANCE}) supports a lot of invariant manifolds, and not all of
1353: them are relevant to the reduced description
1354: (for example, any individual trajectory is itself an invariant manifold).
1355: This should be carefully taken into
1356: account when deriving a relevant equation for the correction in
1357: the states of the initial manifold $\bOmega_0$ which are located far from
1358: equilibrium. This point concerns the procedure of the linearization of
1359: the vector field $\JJ$, appearing in the
1360: equation (\ref{METHOD}).
1361: We shall return to the explicit form of the Marcelin--De Donder kinetic
1362: function (\ref{MDD}). Let $\cc$ is an arbitrary fixed element of
1363: the phase space. The linearization of the vector function $\JJ$
1364: (\ref{KINETIC MDD}) about $\cc$ may be written
1365: $\JJ(\cc+\delta\cc)\approx\JJ(\cc)+\LL_{\cc}\delta\cc$
1366: where the linear operator $\LL_{\cc}$ acts as follows:
1367: \begin{equation}
1368: \label{LINEAR}
1369: \LL_{\cc}\xx=\sum_{s=1}^{r}
1370: \bgamma_s[W^{+}_s(\cc)
1371: \langle\balpha_s,\HH_{\cc}\xx
1372: \rangle-
1373: W^{-}_s(\cc)\langle\bbeta_s,\HH_{\cc}
1374: \xx\rangle].
1375: \end{equation}
1376: Here $\HH_{\cc}$ is the matrix of second derivatives of the
1377: function $G$ in the state $\cc$ [see Eq.\ (\ref{MATRIX})].
1378: The matrix $\LL_{\cc}$ in the Eq.\ (\ref{LINEAR}) can be decomposed
1379: as follows:
1380: \begin{equation}
1381: \label{DECOMPOSITION}
1382: \LL_{\cc}=\LL'_{\cc}+
1383: \LL''_{\cc}.
1384: \end{equation}
1385: Matrices $\LL'_{\cc}$ and $\LL''_{\cc}$ act as follows:
1386: \begin{eqnarray}
1387: \label{SYMMETRIC}
1388: \LL'_{\cc}\xx&=&-\frac{1}{2}\sum_{s=1}^{r}
1389: [W^{+}_s(\cc)+W^{-}_s(\cc)]
1390: \bgamma_s\langle\bgamma_s,
1391: \HH_{\cc}\xx\rangle,\\
1392: \label{NONSYMMETRIC}
1393: \LL''_{\cc}\xx&=&\frac{1}{2}\sum_{s=1}^{r}
1394: [W^{+}_s(\cc)-W^{-}_s(\cc)]\bgamma_s
1395: \langle\balpha_s+\bbeta_s,
1396: \HH_{\cc}\xx\rangle.
1397: \end{eqnarray}
1398: Some features of this decomposition are best seen when we
1399: use the thermodynamic scalar product (\ref{ESP}):
1400: The following properties of the matrix $\LL'_{\cc}$ are verified
1401: immediately:
1402:
1403:
1404: (i) The matrix $\LL'_{\cc}$ is symmetric in the scalar product
1405: (\ref{ESP}):
1406: \begin{equation}
1407: \label{sym}
1408: \langle\!\langle\xx,\LL'_{\cc}\yy\rangle\!\rangle=
1409: \langle\!\langle\yy,\LL'_{\cc}\xx\rangle\!\rangle.
1410: \end{equation}
1411: (ii) The matrix $\LL'_{\cc}$ is nonpositive definite in the scalar product
1412: (\ref{ESP}):
1413: \begin{equation}
1414: \label{pos}
1415: \langle\!\langle\xx,\LL'_{\cc}\xx\rangle\!\rangle\le 0.
1416: \end{equation}
1417: (iii) The null space of the matrix $\LL'_{\cc}$
1418: is the linear envelope of the vectors $\HH^{-1}_{\cc}\bb_i$
1419: representing the complete system of conservation laws:
1420: \begin{equation}
1421: \label{bal}
1422: {\rm ker}\LL'_{\cc}={\rm Lin}\{\HH^{-1}_{\cc}\bb_i, i=1,\dots,l\}
1423: \end{equation}
1424: (iv) If $\cc=\cc^{\rm eq}$, then
1425: $W_s^+(\cc^{\rm eq})=W_s^-(\cc^{\rm eq})$,
1426: and
1427: \begin{equation}
1428: \label{prop4}
1429: \LL'_{\cc^{\rm eq}}=\LL_{\cc^{\rm eq}}.
1430: \end{equation}
1431:
1432:
1433: Thus, the decomposition Eq.\ (\ref{DECOMPOSITION}) splits the
1434: matrix $\LL_{\cc}$ in two parts: one part, Eq.\ (\ref{SYMMETRIC})
1435: is symmetric and nonpositive
1436: definite, while the other part, Eq.\ (\ref{NONSYMMETRIC}),
1437: vanishes in the equilibrium. The
1438: decomposition Eq.\ (\ref{DECOMPOSITION}) explicitly takes into account
1439: the Marcelin-De Donder form of the kinetic function.
1440: For other dissipative systems, the decomposition
1441: (\ref{DECOMPOSITION}) is possible as soon as
1442: the relevant kinetic operator is written in a gain--loss form [for instance,
1443: this is straightforward for the Boltzmann collision operator].
1444:
1445: In the sequel, we shall make use of the properties of the operator
1446: $\LL'_{\cc}$ (\ref{SYMMETRIC}) for constructing the dynamic correction
1447: by extending the picture of the decomposition of motions.
1448:
1449: \subsection{Decomposition of motions: Kinetics}
1450: \label{DEC_KIN}
1451: The assumption about the existence of the decomposition of motions
1452: near the manifold of reduced description $\bOmega$ has led to the
1453: {\it thermodynamic} specifications of the states $\cc\in\bOmega$.
1454: This was accomplished in the section \ref{DEC_THERM}, where the
1455: thermodynamic projector was backed by an appropriate
1456: variational formulation, and this helped us to establish
1457: the induced dynamics consistent with the dissipation property.
1458: Another important feature of the decomposition of motions is that the
1459: states $\cc\in\bOmega$ can be specified {\it kinetically}.
1460: Indeed, let us do it again as if the decomposition of motions
1461: were valid in the neighborhood of the manifold $\bOmega$, and let us
1462: `freeze' the slow dynamics along the $\bOmega$, focusing on the
1463: fast process of relaxation
1464: towards a state $\cc\in\bOmega$. From the thermodynamic perspective,
1465: fast motions take place on the affine hyperplane
1466: $\cc+\delta\cc\in\Gamma_{\cc_0}$, where $\Gamma_{\cc_0}$ is given by Eq.\ (\ref{hyperplane}).
1467: From the kinetic perspective, fast motions on this hyperplane should be treated as
1468: a {\it relaxation} equation,
1469: equipped with the quadratic Lyapunov function
1470: $\delta G=\langle\!\langle\delta\cc,\delta\cc\rangle\!\rangle$,
1471: Futhermore, we require that the linear operator of this evolution equation
1472: should respect Onsager's symmetry
1473: requirements (selfadjointness with respect to the entropic scalar product).
1474: This latter crucial requirement describes fast motions under
1475: the frozen slow evolution in the similar way, as {\it all} the motions near the equilibrium.
1476:
1477:
1478:
1479:
1480: Let us consider now the manifold $\bOmega_0$ which is not the invariant
1481: manifold of the reduced description but, by our assumption, is located
1482: close to it. Consider a state $\cc_0\in\bOmega_0$, and the states
1483: $\cc_0+\delta\cc$ close to it. Further, let us consider an equation
1484: \begin{equation}
1485: \label{RELAX_SYMM}
1486: \dot{\delta\cc}=\LL'_{\cc_0}\delta\cc.
1487: \end{equation}
1488: Due to the properties of the operator $\LL'_{\cc_0}$ (\ref{SYMMETRIC}),
1489: this equation can be regarded as a model of the
1490: assumed true relaxation equation near the true manifold of the reduced
1491: description. For this reason, we shall use the symmetric operator
1492: $\LL'_{\cc}$ (\ref{SYMMETRIC}) {\it instead} of the linear operator $\LL_{\cc}$
1493: when constructing the corrections.
1494:
1495: \subsection{Symmetric iteration}
1496: Let the manifold $\bOmega_0$ and the corresponding projector
1497: $\PP_0$ are the initial approximation to the invariant
1498: manifold of the reduced description.
1499: The dynamic correction $\cc_1=\cc_0+\delta\cc$
1500: is found upon solving the following system of linear algebraic
1501: equations:
1502: \begin{equation}
1503: \label{ITERATION}
1504: \left[1-\PP_0\right]\left[\JJ(\cc_0) +
1505: \LL'_{\cc_0}\delta\cc\right]=0,\
1506: \PP_0\delta\cc=0.
1507: \end{equation}
1508: Here $\LL'_{\cc_0}$ is the matrix (\ref{SYMMETRIC})
1509: taken in the states on the manifold $\bOmega_0$. An important
1510: technical point here is that the linear system
1511: (\ref{ITERATION}) always
1512: has the unique solution for any choice of the manifold $\bOmega$.
1513: This point is crucial since it guarantees the opportunity of carrying
1514: out the correction process for arbitrary number of steps.
1515:
1516: \section{The method of invariant manifold}
1517: \label{MIM}
1518: We shall now combine together the two procedures discussed above.
1519: The resulting method of invariant manifold intends to seek iteratively
1520: the reduced description, starting with an initial approximation.
1521:
1522: (i). {\it Initialization}. In order to start the procedure, it is
1523: required to choose the initial manifold $\bOmega_0$,
1524: and to derive corresponding thermodynamic projector $\PP_0$.
1525: In the majority of cases, initial manifolds are available in two
1526: different ways. The first case are the quasi-equilibrium manifolds described
1527: in the section \ref{partial_eq}.
1528: The macroscopic parameters are $M_i=c_i=\langle\mm_i,\cc\rangle$,
1529: where $\mm_i$ is the unit vector corresponding to the specie $A_i$.
1530: The quasi-equilibrium manifold, $\cc_0(M_1,\dots,M_k,B_1,\dots,B_l)$,
1531: compatible with the conservation laws, is the solution to the variational
1532: problem:
1533: \begin{eqnarray}
1534: \label{QE}
1535: G\to{\rm min\ }, && \langle\mm_i,\cc\rangle=c_i,
1536: \ i=1,\dots,k,\\\nonumber
1537: &&\langle\bb_j,\cc\rangle=B_j, \ j=1,\dots,l .
1538: \end{eqnarray}
1539: In the case of quasi--equilibrium approximation, the corresponding
1540: thermodynamic projector can be written most straightforwardly in
1541: terms of the variables $M_i$:
1542: \begin{equation}
1543: \label{PROJ-QE}
1544: \PP_0\xx=\sum_{i=1}^{k}\frac{\partial\cc_0}{\partial c_i}
1545: \langle\mm_i,\xx\rangle+
1546: \sum_{i=1}^{l}\frac{\partial\cc_0}{\partial B_i}
1547: \langle\bb_i,\xx\rangle.
1548: \end{equation}
1549: For quasi-equilibrium manifolds, a reparameterization with the set
1550: (\ref{param_thermo}), (\ref{param_other}) is {\it not} necessary (\cite{GK92}; \cite{GK94}).
1551:
1552: The second source of initial approximations are quasi-stationary
1553: manifolds (section \ref{QSS}). Unlike
1554: the quasi-equilibrium case, the quasi-stationary manifolds must be
1555: reparameterized in order to construct the thermodynamic projector.
1556:
1557:
1558: (ii). {\it Corrections.} Iterations are organized in accord with the rule:
1559: If $\cc_m$ is the $m$th approximation to the invariant manifold, then
1560: the correction $\cc_{m+1}=\cc_{m}+\delta\cc$ is found from the linear
1561: algebraic equations,
1562: \begin{eqnarray}
1563: \label{corr_eq}
1564: [1-\PP_{m}](\JJ(\cc_m)+\LL'_{\cc_m}\delta\cc)&=&0,\\
1565: \PP_m\delta\cc&=&0.\label{corr_cond}
1566: \end{eqnarray}
1567: Here $\LL'_{\cc_m}$ is the symmetric matrix (\ref{SYMMETRIC})
1568: evaluated at the $m$th approximation. The projector
1569: $\PP_{m}$ is not obligatory thermodynamic at that step, and it is taken as follows:
1570: \begin{equation}
1571: \label{proj_nonthermo}
1572: \PP_m\xx=\sum_{i=1}^{k}\frac{\partial\cc_m}{\partial c_i}
1573: \langle\mm_i,\xx\rangle+
1574: \sum_{i=1}^{l}\frac{\partial\cc_m}{\partial B_i}
1575: \langle\bb_i,\xx\rangle.
1576: \end{equation}
1577: (iii). {\it Dynamics.} Dynamics on the $m$th manifold is obtained with
1578: the thermodynamic re-parameterization.
1579:
1580: In the next section we shall illustrate how this all works.
1581:
1582:
1583:
1584: \section{Illustration: Two-step catalytic reaction}
1585: \label{EX}
1586: Here we consider a two-step four-component reaction with one
1587: catalyst $A_2$:
1588: \begin{equation}
1589: \label{mech_ex}
1590: A_1+A_2\rightleftharpoons\ A_3\rightleftharpoons A_2 + A_4.
1591: \end{equation}
1592: We assume the Lyapunov function of the form (\ref{gfun}),
1593: $G=\sum_{i=1}^{4}c_i[\ln(c_i/c_i^{\rm eq})-1]$.
1594: The kinetic equation for the four--component vector of concentrations,
1595: $\cc=(c_1,c_2,c_3,c_4)$, has the form
1596: \begin{equation}
1597: \label{1exequ}
1598: \dot{\cc}=\bgamma_1W_1+\bgamma_2W_2.
1599: \end{equation}
1600: Here $\bgamma_{1,2}$ are stoichiometric vectors,
1601: \begin{equation}
1602: \label{stochio_ex}
1603: \bgamma_1=(-1,-1,1,0),\ \bgamma_2=(0,1,-1,1),
1604: \end{equation}
1605: while functions $W_{1,2}$ are reaction rates:
1606: \begin{equation}
1607: \label{rates_ex}
1608: W_1=k^+_1c_1c_2-k^-_1c_3,\
1609: W_2=k^+_2c_3- k^-_2c_2 c_4.
1610: \end{equation}
1611: Here $k_{1,2}^{\pm}$ are reaction rate constants.
1612: The system under consideration has two conservation laws,
1613: \begin{equation}
1614: \label{cons_ex}
1615: c_1+c_3+c_4=B_1,\ c_2 +c_3=B_2,
1616: \end{equation}
1617: or $\langle\bb_{1,2},\cc\rangle=B_{1,2}$, where
1618: $\bb_1=(1,0,1,1)$ and $\bb_2=(0,1,1,0)$. The nonlinear system
1619: (\ref{1exequ}) is effectively two-dimensional, and we consider a
1620: one-dimensional reduced description.
1621:
1622: We have chosen the concentration of the specie
1623: $A_1$ as the variable of reduced description:
1624: $M=c_1$, and
1625: $c_1=\langle \mm,\cc\rangle$, where $\mm=(1,0,0,0)$.
1626: The initial manifold $\cc_0(M)$ was taken as the quasi-equilibrium
1627: approximation, i.e. the vector function $\cc_0$ is the solution to
1628: the problem:
1629: \begin{equation}
1630: \label{qe}
1631: G\to{\rm min}\ {\rm for}\
1632: \langle\mm,\cc\rangle=c_1,\ \langle\bb_1,\cc\rangle=B_1,\ \langle\bb_2,\cc\rangle=B_2.
1633: \end{equation}
1634: The solution to the problem (\ref{qe}) reads:
1635: \begin{eqnarray}
1636: \label{1ex1}
1637: c_{01}&=&c_1,\\\nonumber
1638: c_{02}&=&B_2-\phi(c_1), \\\nonumber
1639: c_{03}&=&\phi(c_1),\\\nonumber
1640: c_{04}&=&B_1-c_1-\phi(c_1),\\\nonumber
1641: \phi(M)&=&A(c_1)-\sqrt{A^2(c_1)-B_2(B_1-c_1)},\\\nonumber
1642: A(c_1)&=&\frac{B_2(B_1-c^{\rm eq}_1)+c_3^{\rm eq}(c_1^{\rm eq}+
1643: c_3^{\rm eq}-c_1)}{2c_3^{\rm eq}} .
1644: \end{eqnarray}
1645: The thermodynamic projector associated with the manifold
1646: (\ref{1ex1}) reads:
1647: \begin{equation}
1648: \label{proj_0}
1649: \PP_0\xx=\frac{\partial\cc_0}{\partial c_1}\langle\mm,\xx\rangle
1650: +\frac{\partial\cc_0}{\partial B_1}\langle\bb_1,\xx\rangle+
1651: \frac{\partial\cc_0}{\partial B_2}\langle\bb_2,\xx\rangle.
1652: \end{equation}
1653: Computing
1654: $\bDelta_0=[1-\PP_0]\JJ(\cc_0)$
1655: we find that the inequality (\ref{DEFECT}) takes place, and thus
1656: the manifold $\cc_0$ is not invariant. The first correction, $\cc_1=\cc_0+\delta\cc$,
1657: is found from the linear algebraic system (\ref{corr_eq})
1658: \begin{eqnarray}
1659: \label{LIN_SYSTEM}
1660: (1-\PP_0)\LL'_{0}\delta\cc&=&-[1-\PP_0]\JJ(\cc_0),\\\nonumber
1661: \delta c_1&=&0 \nonumber \\
1662: \delta c_1+\delta c_3+\delta c_4&=&0 \nonumber \\
1663: \delta c_3+\delta c_2&=&0,
1664: \label{correction1}
1665: \end{eqnarray}
1666: where the symmetric $4\times 4$ matrix $\LL'_{0}$ has the form
1667: (we write $0$ instead of $\cc_0$ in the subscript in order to simplify
1668: notations):
1669: \begin{equation}
1670: \label{SL}
1671: L'_{0, kl}=-\gamma_{1k}
1672: \frac{W_1^+(\cc_0)+W_1^-(\cc_0)}{2}
1673: \frac{\gamma_{1l}}{c_{0l}}
1674: -
1675: \gamma_{2k}
1676: \frac{W_2^+(\cc_0)+W_2^-(\cc_0)}{2}
1677: \frac{\gamma_{2l}}{c_{0l}}
1678: \end{equation}
1679: The explicit solution $\cc_1(c_1,B_1,B_2)$ to the linear system
1680: (\ref{LIN_SYSTEM}) is easily
1681: found, and we do not reproduce it here.
1682: The process was iterated. On the $k+1$ iteration, the following projector
1683: $\PP_{k}$ was used:
1684: \begin{equation}
1685: \label{proj_k}
1686: \PP_k\xx=\frac{\partial\cc_k}{\partial c_1}\langle\mm,\xx\rangle+
1687: \frac{\partial\cc_k}{\partial B_1}\langle\bb_1,\xx\rangle+
1688: \frac{\partial\cc_k}{\partial B_2}\langle\bb_2,\xx\rangle.
1689: \end{equation}
1690: Notice that projector $\PP_k$ (\ref{proj_k}) is the thermodynamic
1691: projector only if $k=0$. As we have already mentioned it above,
1692: in the process of finding the corrections to the manifold,
1693: the non-thermodynamic projectors are allowed.
1694: The linear equation at the $k+1$ iteration is thus
1695: obtained by replacing $\cc_0$, $\PP_0$, and $\LL'_0$ with
1696: $\cc_k$, $\PP_k$, and $\LL'_k$ in all the entries of the Eqs.\
1697: (\ref{LIN_SYSTEM}) and (\ref{SL}).
1698:
1699: Once the manifold $\cc_k$ was obtained on the $k$th iteration,
1700: we derived the corresponding dynamics by introducing the thermodynamic
1701: parameterization (and the corresponding thermodynamic projector)
1702: with the help of the function (\ref{param_thermo}).
1703: The resulting dynamic equation for the variable $c_1$
1704: in the $k$th approximation has the form:
1705: \begin{equation}
1706: \label{1ex3}
1707: \langle\bnabla G\bigm|_{\cc_k},\partial\cc_k/\partial c_1\rangle\dot{c_1}=
1708: \langle\bnabla G\bigm|_{\cc_k},\JJ(\cc_k)\rangle.
1709: \end{equation}
1710: Here $[\bnabla G\bigm|_{\cc_k}]_i=\ln[c_{ki}/ c^{\rm eq}_i]$.
1711:
1712: Analytic results were compared with the results of the numerical
1713: integration. The following set of parameters was used:
1714: \begin{eqnarray*}
1715: k^+_1=1.0,\ k^-_1=0.5,\ k^+_2=0.4 ,\ k^-_2=1.0;\\
1716: c_1^{\rm eq}=0.5,\ c_2^{\rm eq}=0.1,\ c_3^{\rm eq}=0.1,\ c_4^{\rm eq}=0.4,\\
1717: B_1=1.0 ,\ \ B_2=0.2 .
1718: \end{eqnarray*}
1719: Direct numerical integration of the system has demonstrated
1720: that the manifold $c_3= c^{\rm eq}_3$ in the plane $(c_1,c_3)$
1721: attracts all individual trajectories. Thus, the reduced description in this
1722: example should extract this manifold.
1723:
1724: Fig.\ \ref{Fig1} demonstrates the quasi--equilibrium manifold
1725: (\ref{1ex1}) and first two corrections found analytically. It is apparent that
1726: while the initial quasi-equilibrium
1727: approximation is in a poor agreement with the reduced description, the
1728: corrections {\it rapidly} improve the situation. This confirms our expectation
1729: of an advantage of using iteration methods in comparison to methods
1730: based on a small parameter expansions.
1731:
1732:
1733:
1734:
1735:
1736: \section{Method of invariant manifold without a priori parameterization}
1737: \label{PARAMETERIZATION}
1738: Formally, the method of invariant manifold does not require a global parameterization of the
1739: manifolds. However, in most of the cases, one makes use of a priori defined
1740: ``macroscopic'' variables
1741: $M$. This is motivated by the choice of quasi-equilibrium initial approximations.
1742:
1743: Let a manifold $\bOmega$ be defined in the phase space of the system, its tangent
1744: space in the point $\cc$ be $T_{\cc}\bOmega$. How to define the projector
1745: of the whole concentrations space onto $T_{\cc}\bOmega$ without using any
1746: a priori parameterization of $\bOmega$?
1747:
1748: The basis of the answer to this question is the condition of thermodynamicity
1749: (\ref{therm1}). Let us denote $E$ as the concentration space, and consider the
1750: problem of the choice of the projector in the quadratic approximation
1751: to the thermodynamic potential $G$:
1752:
1753: \begin{equation}
1754: G_{\rm q}=\langle \bg,\HH_{\cc}\Delta\cc\rangle
1755: +\frac{1}{2}\langle\Delta\cc,\HH_{\cc}\Delta\cc\rangle
1756: =\langle\!\langle\bg,\Delta\cc\rangle\!\rangle+
1757: \frac{1}{2}\langle\!\langle\Delta\cc,\Delta\cc\rangle\!\rangle,
1758: \label{Gquadratic}
1759: \end{equation}
1760: where $\HH_{\cc}$ is the matrix of the second-order derivatives of $G$ (\ref{MATRIX}),
1761: $\bg=\HH^{-1}_{\cc}\bnabla G$, $\Delta\cc$ is the deviation of the concentration
1762: vector from the expansion point.
1763:
1764: Let a linear subspace $T$ be given in the concentrations space $E$.
1765: {\it Problem:} For every $\Delta\cc+T$, and for every $\bg\in E$, define a subspace
1766: $L_{\Delta\cc}$ such that:
1767: (i) $L_{\Delta\cc}$ is a complement of $T$ in $E$:
1768: \[L_{\Delta\cc}+T=E,\ L_{\Delta\cc}\cap T=\{\bZERO\}.\]
1769: (ii) $\Delta\cc$ is the point of minimum of
1770: $G_{\rm q}$ on $L_{\Delta\cc}+\Delta\cc$:
1771: \begin{equation}
1772: \label{minQ}
1773: \Delta\cc=\arg\min_{\xx-\Delta\cc\in L_{\Delta\cc}}G_{\rm q}(\xx).
1774: \end{equation}
1775: Besides (i) and (ii), we also impose the requirement of
1776: a {\it maximal smoothness} (analyticity) on $L_{\Delta\cc}$ as a function of
1777: $\bg$ and $\Delta\cc$. Requirement (\ref{minQ}) implies that $\Delta\cc$ is the quasi-equilibrium
1778: point for the given $L_{\Delta\cc}$, while the problem in a whole is the
1779: {\it inverse} quasi-equilibrium problem: We construct $L_{\Delta\cc}$ such that
1780: $T$ will be the quasi-equilibrium manifold. Then
1781: subspaces $L_{\Delta\cc}$ will actually be the kernels of the quasi-equilibrium projector.
1782:
1783: Let $\ff_1,\dots,\ff_k$ be the orthonormalized with respect to
1784: $\langle\!\langle\cdot,\cdot\rangle\!\rangle$ scalar product basis of $T$,
1785: vector $\hh$ be orthogonal to $T$, $\langle\!\langle\hh,\hh\rangle\!\rangle=1$,
1786: $\bg=\alpha\ff_1+\beta\hh$. Condition (\ref{minQ})
1787: implies that the vector $\bnabla G$ is orthogonal to
1788: $L_{\Delta\cc}$ in the point $\Delta\cc$.
1789:
1790: Let us first consider the case $\beta=0$. The requirement of analyticity
1791: of $L_{\Delta\cc}$ as the function of $\alpha$ and $\Delta\cc$ implies
1792: $L_{\Delta\cc}=L_{\bZERO}+o(1)$, where
1793: $L_{\bZERO}=T^{\perp}$ is the orthogonal completement of $T$ with respect
1794: to scalar product $\langle\langle\cdot,\cdot\rangle\rangle$.
1795: The constant solution, $L_{\Delta\cc}\equiv L_{\bZERO}$ also
1796: satisfies (\ref{minQ}). Let us fix $\alpha\ne0$, and extend this latter solution
1797: to $\beta\ne0$. With this, we obtain a basis, $\bl_1,\dots,\bl_{n-k}$.
1798: Here is the simplest construction of this basis:
1799: \begin{equation}
1800: \label{ort1}
1801: \bl_1=\frac{\beta\ff_1-(\alpha+\Delta c_1)\hh}{(\beta^2+(\alpha+\Delta c_1)^2)^{1/2}},
1802: \end{equation}
1803: where $\Delta c_1=\langle\langle\Delta\cc,\ff_1\rangle\rangle$
1804: is the first component in the expansion, $\Delta\cc=\sum_i\Delta c_i\ff_i$.
1805: The rest of the basis elements, $\bl_2,\dots,\bl_{n-k}$ form
1806: the orthogonal completement of $T\oplus(\hh)$ with respect to scalar
1807: product $\langle\langle\cdot,\cdot\rangle\rangle$, $(\hh)$ is the line spanned by
1808: $\hh$.
1809:
1810: Dependence $L_{\Delta\cc}$ (\ref{ort1}) on $\Delta\cc$, $\alpha$ and $\beta$
1811: is singular: At $\alpha+\Delta c_1$, vector $\bl_1\in T$, and then $L_{\Delta\cc}$
1812: is not the completement of $T$ in $E$ anymore.
1813: For $\alpha\ne0$, dependence $L_{\Delta\cc}$ gives one of the solutions to the inverse
1814: quasi-equilibrium problem in the neighborhood of zero in $T$. We are interested only in the limit,
1815: \begin{equation}
1816: \label{lim1}
1817: \lim_{\Delta\cc\to\bZERO}L_{\Delta\cc}={\rm Lin}\left\{
1818: \frac{\beta\ff_1-\alpha\hh}{\sqrt{\alpha^2+\beta^2}},\bl_2,\dots,\bl_{n-k}\right\}.
1819: \end{equation}
1820:
1821: Finally, let us define now the projector $\PP_{\cc}$ of the space $E$ onto $T_{\cc}\bOmega$.
1822: If $\HH^{-1}_{\cc}\bnabla G\in T_{\cc}\bOmega$,
1823: then $\PP_{\cc}$ is the orthogonal projector with respect
1824: to the scalar product
1825: $\langle\langle\cdot,\cdot\rangle\rangle$:
1826:
1827: \begin{equation}
1828: \PP_{\cc}\zz=\sum_{i=1}^k\ff_i\langle\langle \ff_i,\zz\rangle\rangle.
1829: \end{equation}
1830: If $\HH^{-1}_{\cc}\bnabla G \notin T_{\cc}\bOmega$, then, according to
1831: Eq.\ (\ref{lim1}),
1832: \begin{equation}
1833: \label{result_projector}
1834: \PP_{\cc}\zz=\frac{\langle\langle \ff_1,\zz\rangle\rangle
1835: -\langle\langle \bl_1,\zz\rangle\rangle
1836: \langle\langle \ff_1,\bl_1\rangle\rangle}
1837: {1-\langle\langle \ff_1,\bl_1\rangle\rangle^2}\ff_1+
1838: \sum_{i=2}^k\ff_i\langle\langle \ff_i,\zz\rangle\rangle,
1839: \end{equation}
1840: where
1841: $\{\ff_1,\dots,\ff_k\}$ is the orthonormal with respect to
1842: $\langle\langle\cdot,\cdot \rangle\rangle$ basis of $T_{\cc}\bOmega$,
1843: $\hh$ is orthogonal to $T$, $\langle\langle \hh,\hh\rangle\rangle=1$,
1844: $\HH^{-1}_{\cc}\bnabla G=\alpha\ff_1+\beta\hh$,
1845: $\bl_1=(\beta\ff_1-\alpha\hh)/\sqrt{\alpha^2+\beta^2}$,
1846: $\langle\langle \ff_1,\bl_1\rangle\rangle=\beta/\sqrt{\alpha^2+\beta^2}$.
1847:
1848: Thus, for solving the invariance equation iteratively, one needs only projector $\PP_{\cc}$
1849: (\ref{result_projector}), and one does not need a priori parameterization of $\bOmega$
1850: anymore.
1851:
1852:
1853: \section{Method of invariant grids}
1854: \label{grid}
1855:
1856: Grid-based approximations of manifolds are attractive
1857: from the computational perspective. Since no a priori parameterization
1858: is required in the method of invariant manifold, in this section
1859: we develop its grid-based realization.
1860: Let us consider a regular grid $Q$ in $R^k$, and its
1861: mapping $F$ into the concentrations space $E$.
1862: It makes sense to consider only
1863: $F$ which map a finite part of the grid into
1864: the phase space $V$. This part of the map is termed {\it essential}.
1865: Extension of the map $F$ onto the rest of the nodes is done by a
1866: simple (for example, linear) extrapolation of the essential part
1867: (in practice, one needs to extrapolate only onto the next
1868: neighbors of the essential nodes).
1869:
1870:
1871: Let operators of grid differentiation $D_i$
1872: where be defined for functions on the grid, where $i=1,\dots,k$
1873: label grid coordinates $x_i$.
1874: With this, the tangent space to the image of the grid in the
1875: point $\cc(x)=F(x)$ is defined for each node of the grid $x$:
1876: \begin{eqnarray}
1877: T_x={\rm Lin}\{\varphi_1,\dots,\varphi_k\},\nonumber\\
1878: \varphi_i=D_i\cc(x)=(D_ic_1(x),\dots,D_ic_n(x)).
1879: \end{eqnarray}
1880: The grid is termed invariant if, for each essential node,
1881: \[ \JJ(\cc(x))\in T_x.\]
1882: For the essential nodes, we write down the invariance equation
1883: with the projector, $P_{\cc(x)}: E\to T_x$:
1884: This equation is solved using the Newton method as it was described
1885: above in the section \ref{MIM}. A good initial approximation
1886: is a linear map of the grid on the affine manifold corresponding
1887: to slow relaxation in the vicinity of the equilibrium.
1888: It is convenient to take this map isometric with respect
1889: to the metrics generated by the entropic scalar product
1890: in the equilibrium.
1891:
1892: If the vector field of the reduced model,
1893: $\dot{\cc}=\PP_{\cc(x)}\JJ(\cc(x))$,
1894: is defined on the nodes $F(x)$, then one can define the
1895: dynamics $\dot{x}_i$ on the nodes. In order to do this,
1896: we expand $\dot{\cc}$ over $\varphi_i$:
1897: $\dot{\cc}=\sum_{i=1}^ka_i\varphi_i$.
1898: The dynamics on the nodes is then defined by equations,
1899: $\dot{x}_i=a_i$.
1900: Using interpolation, we can define the vector field
1901: $\dot{x}$ within the essential cells of the grid (those cells for
1902: which the all the nodes are essential). The system of equations thus
1903: obtained models the dynamics on the invariant manifold.
1904:
1905: The essence of this construction is that, by solving a set
1906: of uncomplicated linear equations arising from linearization
1907: of the invariance equations on the nodes one gets a reliable
1908: numerical scheme for constructing invariant manifolds.
1909: The use of the grid differentiation rather than a differentiable
1910: approximation to the manifold makes the scheme suited for
1911: parallel realizations. We stress it once again that
1912: such realizations are only possible if no a priori global
1913: parameterization of manifolds is required.
1914: Further refinements of the scheme,
1915: taking into account the process of moving the inessential nodes
1916: into the phase space, and the opposite process of essential
1917: nodes leaving the phase space can be done based in the same way
1918: as for grid-based data analysis \citep{GR99,GZ01}.
1919:
1920:
1921:
1922:
1923: \section{Method of invariant manifold for open systems}
1924: \label{open}
1925: One of the problems to be focused on when studying closed
1926: systems is to prepare extensions of the result for
1927: open or driven by flows systems.
1928: External flows are usually taken into account
1929: by addidinal terms in the kinetic equations (\ref{reaction}):
1930: \begin{equation}
1931: \label{external}
1932: \dot{\cc}=\JJ(\cc)+\bPi.
1933: \end{equation}
1934: {\it Zero-order approximation} assumes that the flow does
1935: not change the invariant manifold.
1936: Equations of the reduced dynamics, however, do change:
1937: Instead of $\JJ(\cc(M))$ we substitute
1938: $\JJ(\cc(M))+\bPi$ into Eq.\ (\ref{dyn}):
1939: \begin{equation}
1940: \label{zero}
1941: \dot{M}_i=\sum_{j=1}^pN^{-1}_{ij}\langle \bnabla M_j\big|_{\cc(M)},
1942: \JJ(\cc(M))+\bPi\rangle.
1943: \end{equation}
1944: Zero-order approximation assumes that the fast dynamics in the
1945: closed system strongly couples the variables $\cc$, so that
1946: flows cannot influence this coupling.
1947:
1948: {\it First-order approximation} takes into account the shift
1949: of the invariant manifold by $\delta\cc$. Equations for Newton's iterations have
1950: the same form (\ref{ITERATION}) but instead of the vector field
1951: $\JJ$ they take into account the presence of the flow:
1952: \begin{equation}
1953: \label{one}
1954: [1-\PP_{\cc}](\bPi+\LL'_{\cc}\delta\cc)=0,\ \PP_{\cc}\delta\cc=0,
1955: \end{equation}
1956: where projector $\PP_{\cc}$ corresponds to the unperturbed manifold.
1957:
1958: The first-order approximation means that fluxes change the coupling
1959: between the variables (concentrations). It is assumed that
1960: these new coupling is also set instantaneously (neglect of inertia).
1961:
1962: {\it Remark.} Various realizations of the first-order approximation
1963: in physical and chemical dynamics implement the viewpoint
1964: of an infinitely small chemical reactor driven by the flow.
1965: In other words, this approximation is applicable
1966: in the Lagrangian system of coordinates \citep{KGDN98,ZKD00}.
1967: Transition to Eulerian coordinates is possible
1968: but the relations between concentrations and the flow will change
1969: its form. In a contrast, the more simplistic zero-order approximation
1970: is equally applicable in both the coordinate system, if it is valid.
1971:
1972:
1973: \section{Conclusion}
1974: \label{conclusion}
1975:
1976: In this paper, we have
1977: presented the method for constructing the invariant manifolds
1978: for reducing systems of chemical kinetics.
1979: Our approach to computations of invariant manifolds
1980: of dissipative systems is close in spritit to the
1981: Kolmogorov-Arnold-Moser theory of invariant tori of Hamiltonian systems
1982: \citep{Arnold63,Arnold83}: We also base our consideration on the Newton
1983: method instead of Taylor series expansions \citep{Beyn98}, and systematically
1984: use duality structures.
1985: Recently, a version of an approach based on the invarinace equations
1986: was rediscovered by \cite{Kazantzis00}. He was solving the invariance
1987: equation by a Taylor series expansion. A counterpart of Taylor series expansions
1988: for constructing the slow invariant manifolds in the classical kinetic theory
1989: is the famous Chapman-Enskog method. The question of how this
1990: compares to iteration methods was studied extensively for
1991: certain classes of Grad moment equations \citep{GK96a,KDN97a,K00}.
1992:
1993: The thermodynamic parameterization and the selfadjoint linearization arise
1994: in a natural way in the problem of finding slowest invariant manifolds for
1995: closed systems. This also leads to various applications
1996: in different approaches to reducing the description, in particular, to
1997: a thermodynamically consistent version
1998: of the intrinsic low-dimensional manifold, and to model kinetic equations for lifting
1999: the reduced dynamics. Use of the thermodynamic projector
2000: makes it unnecessary global parameterizations of manifolds, and
2001: thus leads to computationally promising grid-based realizations.
2002:
2003: Invariant manifolds are constructed for closed space-independent chemical systems.
2004: We also describe how to use these manifolds for modeling open and distributed
2005: systems.
2006:
2007:
2008:
2009:
2010:
2011:
2012:
2013:
2014: \begin{thebibliography}{}
2015:
2016:
2017:
2018: \bibitem[Ansumali \& Karlin(2000)]{AK00} Ansumali, S., \& Karlin, I.\ V. (2000). Stabilization of
2019: the Lattice Boltzmann method by the $H$ theorem: A numerical test.
2020: {\it Phys.\ Rev.\ E,} {\bf 62(6)}, 7999-8003.
2021:
2022:
2023: \bibitem[Ansumali \& Karlin(2002)]{AK02}Ansumali S.,\ \& Karlin, I.\ V. (2002). Single relaxation time
2024: model for entropic Lattice Boltzmann methods. {\it Phys.\ Rev.\ E,} {\bf 65} 056312(1-9).
2025:
2026: \bibitem[Ansumali \& Karlin(2002a)]{AK02a}
2027: Ansumali, S., \& Karlin, I.V. (2002a). Entropy function approach
2028: to the lattice Boltzmann method. {\it J.\ Stat.\ Phys.,} {\bf 107(1/2)} 291-
2029: 308.
2030:
2031: \bibitem[Arnold(1963)]{Arnold63}
2032: Arnold, V.\ I.\ (1963). Proof of a theorem of A.\ N.\ Kolmogorov on the
2033: invariance of quasi-periodic motions under small perturbations of the
2034: Hamiltonian. (English translation). {\it Russian Mathematical Surveys,} {\bf 18}, 9-36.
2035:
2036: \bibitem[Arnold(1983)]{Arnold83}Arnold, V.\ I.\ (1983).
2037: {\it Geometrical methods in the theory of ordinary differential equations.}
2038: New York: Springer.
2039:
2040: \bibitem[Beyn \& Kless(1998)]{Beyn98}
2041: Beyn, W.-J., \& Kless, W. (1998). Numerical Taylor expansions of invariant
2042: manifolds in large dynamical systems. {\it Numerische Mathematik,} {\bf 80}, 1-38.
2043:
2044: \bibitem[Bhatnagar, Gross \& Krook(1954)]{BGK}Bhatnagar, P.\ L.,
2045: Gross, E.\ P., \& Krook, M. (1954).
2046: A model for collision processes in gases.
2047: Small amplitude processes in charged and neutral one-component systems.
2048: {\it Phys.\ Rev.,}
2049: {\bf 94}, 511-525.
2050:
2051:
2052: \bibitem[Bobylev(1988)]{Bobylev88}Bobylev, A.\ V.\ (1988).
2053: The theory of the nonlinear spatially uniform Boltzmann equation for Maxwell molecules.
2054: {\it Mathematical physics reviews,} {\bf 7}, 111-233.
2055:
2056: \bibitem[Bowen, Acrivos \& Oppenheim(1963)]{Bowen63}
2057: Bowen, J.\ R., Acrivos, A., \& Oppenheim, A.\ K. (1963). Singular Perturbation
2058: Refinement to Quasi-Steady State Approximation in Chemical Kinetics,
2059: {\it Chemical Engineering Science,} {\bf 18}, 177-188.
2060:
2061: \bibitem[Bykov, Gorban \& Yablonskii(1982)]{BGY82}
2062: Bykov, V.\ I., Gorban, A.\ N., \& Yablonskii, G.\ S. (1982). Description of
2063: nonisothermal reactions in terms of Marcelin-de Donder kinetics and its
2064: generalizations. {\it React.\ Kinet.\ Catal.\ Lett.,} {\bf 20}, 261-265.
2065:
2066: \bibitem[Bykov, Yablonskii \& Akramov(1977)]{BYA77}
2067: Bykov, V.\ I., Yablonskii, G.\ S., \& Akramov, T.\ A. (1977).
2068: The rate of the free energy decrease in the course of the complex chemical reaction.
2069: {\it Dokl.\ Akad.\ Nauk USSR,} {\bf 234 (3)}, 621-634.
2070:
2071: \bibitem[Chen(1988)]{Chen88}Chen, J.-Y. (1988). A general procedure for constructing
2072: reduced reaction
2073: mechanisms with given independent relations. {\it Combustion Science and
2074: Technology,} {\bf 57}, 89-94.
2075:
2076: \bibitem[Chicone(2000)]{Chicone00}
2077: Chicone, C., \& Swanson, R. (2000). Linearization via the Lie derivative.
2078: {\it Electron.\ J.\ Diff.\ Eqns.,} Monograph, 02,
2079: http://ejde.math.swt.edu or http://ejde.math.unt.edu
2080: ftp ejde.math.swt.edu or ejde.math.unt.edu (login: ftp)
2081:
2082: \bibitem[Dimitrov(1982)]{Dimitrov82}Dimitrov, V.I. (1982).
2083: {\it Prostaya kinetika [Simple Kinetics]}. Novosibirsk: Nauka.
2084:
2085:
2086: \bibitem[De Donder \& Van Rysselberghe(1936)]{DeDonder36}
2087: De Donder, T., \& Van Rysselberghe, P. (1936). {\it Thermodynamic theory of
2088: affnity. A book of principles.} Stanford: University Press.
2089:
2090: \bibitem[Dukek, Karlin \& Nonnenmacher(1997)]{DKN97}
2091: Dukek, G., Karlin, I.\ V., \& Nonnenmacher, T.\ F. (1997). Dissipative brackets
2092: as a tool for kinetic modeling. {\it Physica A}, {\bf 239(4)}, 493-508.
2093:
2094: \bibitem[Feinberg(1972)]{Feinberg72}Feinberg, M. (1972).
2095: On chemical kinetics of a certain class. {\it Arch.\ Rational
2096: Mech. Anal.,} {\bf 46(1)}, 1-41.
2097:
2098: \bibitem[Fraser(1988)]{Fraser88}Fraser, S.\ J. (1988).
2099: The steady state and equilibrium approximations: A
2100: geometrical picture. {\it J.\ Chem.\ Phys.}, {\bf 88(8)}, 4732-4738.
2101:
2102: \bibitem[Gear(1971)]{Gear71}Gear, C.\ W. (1971).
2103: {\it Numerical initial value problems in ordinary differential equations.}
2104: Prentice-Hall, Englewood Cliffs, NJ.
2105:
2106: \bibitem[Gorban(1984)]{G84}
2107: Gorban, A.N. (1984). {\it Obkhod ravnovesiya [Equilibrium encircling].}
2108: Novosibirsk: Nauka.
2109:
2110:
2111: \bibitem[Gorban \& Karlin(1992)]{GK92}Gorban, A.\ N., \& Karlin, I.\ V. (1992).
2112: Thermodynamic parameterization. {\it Physica A}, {\bf 190}, 393-404 .
2113:
2114: \bibitem[Gorban \& Karlin(1992a)]{GK92a} Gorban, A.\ N., \& Karlin, I.\ V. (1992a).
2115: The constructing of
2116: invariant manifolds for the Boltzmann equation, {\it Adv.\ Model. and
2117: Analysis C}, {\bf 33(3)}, 39-54.
2118:
2119: \bibitem[Gorban \& Karlin(1992b)]{GK92b}
2120: Gorban, A.\ N., \& Karlin, I.\ V. (1992b). Coarse-grained quasi-
2121: equilibrium approximations for kinetic equations. {\it Adv.\ Model. and
2122: Analysis C}, {\bf 35(1)}, 17-27.
2123:
2124: \bibitem[Gorban \& Karlin(1992c)]{GK92c}Gorban, A.\ N., \& Karlin, I.\ V. (1992c).
2125: H-theorem for generalized
2126: models of the Boltzmann equation. {\it Adv.\ Model. and Analysis C}, {\bf
2127: 33(3)}, 33-38.
2128:
2129: \bibitem[Gorban \& Karlin(1994)]{GK94}Gorban, A.\ N., \& Karlin, I.\ V. (1994).
2130: Method of invariant
2131: manifolds and regularization of acoustic spectra. {\it Transport Theory and
2132: Stat.\ Phys.}, {\bf 23}, 559-632.
2133:
2134: \bibitem[Gorban \& Karlin(1994a)]{GK94a}
2135: Gorban, A.\ N., \& Karlin, I.\ V. (1994a). General approach to
2136: constructing models of the Boltzmann equation. {\it Physica A}, {\bf 206},
2137: 401-420.
2138:
2139: \bibitem[Gorban \& Karlin(1996)]{GK96}Gorban, A.\ N., \& Karlin, I.\ V., (1996).
2140: Scattering rates versus moments: Alternative Grad equations. {\it Phys.\ Rev. E},
2141: {\bf 54(4)}, R3109-R3113.
2142:
2143: \bibitem[Gorban \& Karlin(1996a)]{GK96a}Gorban, A.\ N., \& Karlin, I.\ V., (1996a)
2144: Short-wave limit
2145: of hydrodynamics: A soluble example. {\it Phys.\ Rev.\ Lett.,} {\bf
2146: 77}, 282-285.
2147:
2148: \bibitem[Gorban, Karlin, Zmievskii \& Nonnenmacher(1996)]{GKZN96}
2149: Gorban, A.\ N., Karlin, I.\ V., Zmievskii, V.\ B., \& Nonnenmacher, T.\ F.
2150: (1996). Relaxational trajectories: global approximations. {\it Physica A},
2151: {\bf 231}, 648-672.
2152:
2153: \bibitem[Gorban, Karlin \& Zmievskii(1999)]{GKZ99}
2154: Gorban, A.\ N., Karlin, I.\ V., \& Zmievskii, V.\ B. (1999). Two-
2155: step approximation of space-independent relaxation. {\it Transp.\ Theory
2156: Stat.\ Phys.}, {\bf 28(3)}, 271-296.
2157:
2158: \bibitem[Gorban, Karlin, Zmievskii \& Dymova(2000)]{GKZD00}
2159: Gorban, A.\ N., Karlin, I.\ V., Zmievskii, V.\ B., \& Dymova S.\ V.
2160: (2000). Reduced description in reaction kinetics.
2161: {\it Physica A}, {\bf 275(3-4)}, 361-379.
2162:
2163: \bibitem[Gorban, Karlin, Ilg \& \"Ottinger(2001)]{GKIOe01}
2164: Gorban, A.\ N., Karlin, I.\ V., Ilg, P., \& \"{O}ttinger, H.\ C. (2001).
2165: Corrections and enhancements of quasi-equilibrium states. {\it J.\ Non-
2166: Newtonian Fluid Mech.}, {\bf 96(1-2)}, 203-219.
2167:
2168:
2169: \bibitem[Gorban \& Rossiev(1999)]{GR99}
2170: Gorban, A.\ N., \& Rossiev, A.\ A. (1999). Neural network iterative method of
2171: principal curves for data with gaps. {\it Journal of Computer and System
2172: Sciences Intrnational}, {\bf 38(5)}, 825-831.
2173:
2174: \bibitem[Gorban \& Zinovyev(2001)]{GZ01}
2175: Gorban, A.\ N., Zinovyev, A.\ Yu. (2001). Visualization of data by method of
2176: elastic maps and its applications in genomics, economics and sociology.
2177: Institut des Hautes Etudes Scientifiques, Preprint. IHES M/01/36. Online-
2178: version: http://www.ihes.fr/PREPRINTS/M01/Resu/resu-M01-36.html.
2179:
2180:
2181: \bibitem[Grmela, Karlin \& Zmievski(2002)]{GKZ02}
2182: Grmela, M., Karlin, I.\ V., \& Zmievski, V.\ B. (2002). Boundary
2183: layer minimum entropy principles: A case study.
2184: {\it Phys.\ Rev.\ E} (in press).
2185:
2186: \bibitem[Hartman(1963)]{Hartman63}
2187: Hartman, P. (1963). On the local linearization of differential equations. {\it Proc.
2188: Amer.\ Math.\ Soc.}, {\bf 14}, 568-573.
2189:
2190: \bibitem[Hartman(1982)]{Hartman82}Hartman, P. (1982).
2191: {\it Ordinary differential equations}. Boston: Birkh\"auser.
2192:
2193: \bibitem[Karlin(1989)]{K89} Karlin, I.\ V. (1989). On the relaxation of the chemical reaction
2194: rate. In: {\it Mathematical Problems of Chemical Kinetics}, eds. K.\ I.\
2195: Zamaraev and G.\ S.\ Yablonskii, (Nauka, Novosibirsk), 7-42.
2196:
2197: \bibitem[Karlin(1993)]{K93} Karlin, I.\ V. (1993). The problem of reduced description in
2198: kinetic theory of chemically reacting gas. {\it Modeling, Measurement and
2199: Control C}, {\bf 34(4)}, 1-34.
2200:
2201: \bibitem[Karlin(2000)]{K00} Karlin, I.\ V. (2000).
2202: Exact summation of the Chapman-Enskog expansion
2203: from moment equations. {\it J.\ Phys.\ A: Math.\ Gen.},
2204: {\bf 33}, 8037-8046.
2205:
2206: \bibitem[Karlin, Dukek \& Nonnenmacher(1997)]{KDN97}Karlin, I.\ V., Dukek, G.,
2207: Nonnenmacher, T.\ F. (1997). Invariance principle
2208: for extension of hydrodynamics: Nonlinear viscosity. {\it Phys.\ Rev.\ E},
2209: {\bf 55(2)}, 1573-1576.
2210:
2211: \bibitem[Karlin, Dukek \& Nonnenmacher(1997a)]{KDN97a}
2212: Karlin, I.\ V., Dukek, G., \& Nonnenmacher, T.\ F. (1997a)
2213: Gradient expansions in kinetic theory of phonons.
2214: {\it Phys.\ Rev.\ B,} {\bf 55}, 6324-6329.
2215:
2216: \bibitem[Karlin, Gorban, Dukek \& Nonnenmacher(1998)]{KGDN98}
2217: Karlin, I.\ V., Gorban, A.\ N., Dukek, \& G., Nonnenmacher, T.\ F.
2218: (1998). Dynamic correction to moment approximations. {\it Phys.\ Rev.\ E},
2219: {\bf 57}, 1668--1672.
2220:
2221:
2222:
2223: \bibitem[Kato(1976)]{Kato76}Kato, T. (1976).
2224: {\it Perturbation theory for Linear operators.} Berlin: Springer.
2225:
2226: \bibitem[Kazantzis(2000)]{Kazantzis00}Kazantzis N, (2000) Singular PDEs and the problem of finding invariant manifolds for nonlinear dynamical systems. {\it Physics Letters}, {\bf A272(4)},
2227: 257-263.
2228:
2229: \bibitem[Khibnik, Kuznetsov, Levitin \& Nikolaev(1993)]
2230: {Khibnik93}Khibnik, A., Kuznetsov, Y., Levitin, V., \& Nikolaev, E. (1993). Continuation
2231: techniques and interactive software for bifurcation analysis of ODEs and
2232: iterated maps. {\it Physica D}, {\bf 62}, 360-370.
2233:
2234:
2235: \bibitem[Lam \& Goussis(1994)]{Lam94}
2236: Lam, S.H., Goussis, D. A. (1994). The CSP Method for Simplifying Ki-
2237: netics. {\it International Journal of Chemical Kinetics}, {\bf 26}, 461-486.
2238:
2239: \bibitem[Levenspiel(1999)]{Levenspiel99}
2240: Levenspiel, O. (1999). Chemical Reaction Engineering. {\it Ind.\ Eng.\ Chem.\ Res.},
2241: {\bf 38}, 4139-4143.
2242:
2243: \bibitem[Levenspiel(2000)]{Levenspiel00}
2244: Levenspiel, O. (2000). Response to Professor Yablonsky. {\it Ind.\ Eng.\ Chem.\ Res.},
2245: {\bf 39}, 3120.
2246:
2247: \bibitem[Li, Rabitz \& T\'oth(1994)]{Li94}Li, G., Rabitz, H., \& T\'oth, J. (1994).
2248: A general analysis of exact nonlinear
2249: lumping in chemical kinetics. {\it Chem.\ Eng.\ Sci.}, {\bf 49(3)}, 343-361.
2250:
2251: \bibitem[Maas \& Pope(1992)]{Maas92}Maas, U., \& Pope, S.B. (1992).
2252: Simplifying chemical kinetics: intrinsic low-
2253: dimensional manifolds in composition space. {\it Combustion and Flame}, {\bf 88},
2254: 239-264.
2255:
2256: \bibitem[Orlov \& Rozonoer(1984)]{Orlov84}Orlov, N. N., \& Rozonoer, L. I. (1984).
2257: The macrodynamics of open systems
2258: and the variational principle of the local potential. {\it J. Franklin Inst.}, {\bf 318}, 283-
2259: 314 and 315-347.
2260:
2261: \bibitem[Rabitz(1987)]{Rabitz87}Rabitz, H. (1987).
2262: Chemical Dynamics and Kinetics Phenomena as revealed
2263: by Sensitivity Analysis Techniques. {\it Chem.\ Rev.}, {\bf 87}, 101-112.
2264:
2265: \bibitem[Rabitz, Kramer \& Dacol(1983)]{Rabitz83}Rabitz, H., Kramer, M., \& Dacol, D. (1983).
2266: Sensitivity analysis in chemical
2267: kinetics. {\it Ann.\ Rev.\ Phys.\ Chem.}, {\bf 34}, 419-461.
2268:
2269: \bibitem[Roose \& Spence(1990)]{Roose90}Roose, De Dier, B., \& Spence, A., eds. (1990).
2270: {\it Continuation and bifurcations
2271: numerical techniques and applications}. Dordrecht: Kluwer.
2272:
2273: \bibitem[Roussel \& Fraser(1990)]{Roussel90}Roussel, M.R., \& Fraser, S.J. (1990).
2274: Geometry of the steady-state
2275: approximation: Perturbation and accelerated convergence methods. {\it J.\ Chem.\
2276: Phys.}, {\bf 93}, 1072-1081.
2277:
2278: \bibitem[Roussel \& Fraser(1991)]{Roussel91}Roussel, M.R., \& Fraser, S.J. (1991).
2279: On the geometry of transient relaxation.
2280: {\it J.\ Chem.\ Phys.}, {\bf 94}, 7106-711.
2281:
2282: \bibitem[Segel \& Slemrod(1989)]{Segel89}Segel, L.A., \& Slemrod, M. (1989).
2283: The quasi-steady-state assumption: A case
2284: study in perturbation. {\it SIAM Rev.,} {\bf 31}, 446-477.
2285:
2286: \bibitem[Struchtrup \& Weiss(1998)]{Struchtrup98}
2287: Struchtrup, H., \& Weiss, W. (1998). Maximum of the local entropy production becomes
2288: minimal in stationary processes. {\it Phys.\ Rev.\ Lett.,} {\bf 80}, 5048-5051.
2289:
2290: \bibitem[Strygin \& Sobolev(1988)]{Strygin88}
2291: Strygin V.\ V., \& Sobolev, V.\ A. (1988). {\it
2292: Spliting of motion by means of integral manifolds}. Moscow: Nauka.
2293:
2294: \bibitem[T\'oth, Li, Rabitz \& Tomlin(1997)]{Toth97}
2295: T\'oth, J., Li, G., Rabitz, H., \& Tomlin, A. S. (1997). The effect of lumping and
2296: expanding on kinetic differential equations. {\it SIAM J. Appl. Math.,} {\bf 57(6)},
2297: 1531-1556.
2298:
2299: \bibitem[Vasil'eva, Butuzov \& Kalachev(1995)]{Vasil'eva95}
2300: Vasil'eva A.B., Butuzov V.F., \& Kalachev L.V. (1995). {\it The boundary
2301: function method for singular perturbation problems}, Philadelphia: SIAM.
2302:
2303: \bibitem[Volpert \& Hudjaev(1985)]{Volpert85}Volpert, A. I., \& Hudjaev, S. I. (1985).
2304: {\it Analysis in classes of discontinuous
2305: functions and the equations of mathematical physics}. Dordrecht: Nijhoff.
2306:
2307:
2308: \bibitem[Wei \& Prater(1962)]{Wei62}Wei, J., \& Prater, C. (1962).
2309: The structure and analysis of complex reaction
2310: systems. {\it Adv.\ Catalysis,} {\bf 13}, 203-393.
2311:
2312: \bibitem[Yablonsky(2000)]{Y00}Yablonsky, G.\ S. (2000).
2313: Comments on a commentary by Professor
2314: Levenspiel, {\it Ind.\ Eng.\ Chem.\ Res.}, {\bf 39}, 3120.
2315:
2316: \bibitem[Yablonskii, Bykov, Gorban \& Elokhin(1991)]{YBGE91}
2317: Yablonskii, G.S., Bykov, V.I., Gorban, A.N., \& Elokhin, V.I. (1991). {\it Kinetic
2318: models of catalytic reactions.} Comprehensive Chemical Kinetics, Vol. 32,
2319: Compton, R.\ G., ed. Amsterdam: Elsevier.
2320:
2321: \bibitem[Zhu \& Petzold(1999)]{Zhu99}Zhu, W., \& Petzold, L. (1999).
2322: Model reduction for chemical kinetics: An
2323: optimization approach. {\it AIChE Journal}, (April 1999), 869-886.
2324:
2325: \bibitem[Zmievskii, Kalin \& Deville(2000)]{ZKD00}
2326: Zmievskii, V.\ B., Karlin, I.\ V., \& Deville, M. (2000). The
2327: universal limit in dynamics of dilute polymeric solutions. {\it Physica A},
2328: {\bf 275(1-2)}, 152-177.
2329:
2330:
2331:
2332:
2333:
2334: \end{thebibliography}
2335:
2336: \newpage
2337:
2338:
2339:
2340: \begin{figure}
2341:
2342:
2343:
2344: %\includegraphics[scale=0.2]{invchemFig1.ps}
2345:
2346: %\vspace*{14cm}
2347: \caption{Images of the initial quasi-equilibrium manifold (bold line)
2348: and the first two corrections
2349: (solid normal lines) in the phase plane
2350: $[c_1,c_3]$ for two-step catalytic reaction (\ref{mech_ex}).
2351: Dashed lines are individual trajectories.}
2352: \label{Fig1}
2353: \end{figure}
2354:
2355: \end{document}
2356:
2357:
2358:
2359:
2360:
2361:
2362:
2363:
2364:
2365:
2366:
2367: