q-bio0607047/mn.tex
1: \documentclass[a4paper,12pt]{article}
2: \usepackage{a4wide}
3: \usepackage{amsfonts}          % |
4: \usepackage{amssymb}           % | AmS-LaTeX packages
5: \usepackage{amsmath}           % |
6: \usepackage{color}
7: \usepackage{epsfig}
8: \usepackage{psfrag}
9: \usepackage[numbers,sort&compress]{natbib}
10: 
11: \allowdisplaybreaks
12: 
13: % Keystroke saving macros
14: \input{abbr.tex}
15: 
16: % Define the notations
17: \newcommand{\+}[5]{\def#1{{#3}}}
18: \input{nots.tex}
19: 
20: \begin{document}
21: \title{Asymptotic analysis and analytical solutions of a  model of cardiac excitation}
22: \author{
23:   V.~N.~Biktashev$^{1}$,
24:   R.~Suckley$^{1}$,
25:   Y.~E.~Elkin$^{2}$ and
26:   R.~D.~Simitev$^{1,3}$
27: }
28: \maketitle
29: 
30: $^1$ Department of Mathematical Sciences,
31:      University of Liverpool, 
32:      Liverpool L69 7ZL, UK 
33: 
34: $^2$ Deceased 25/03/2007. Last affiliation: Institute of Mathematical Problems of Biology 
35:      of the Russian Academy of Sciences, Pushchino,
36:      and the Pushchino branch of the Moscow State University, Russia
37: 
38: $^3$ Current address: 
39:      Department of Mathematics,
40:      University of Glasgow, 
41:      Glasgow G12 8QW, UK
42: 
43: 
44: \begin{abstract}
45: %
46:   We describe an asymptotic approach to gated ionic  models of
47:   single-cell cardiac excitability. It has a form  essentially  
48:   different from the Tikhonov fast-slow form assumed in standard asymptotic
49:   reductions of excitable systems. This is of
50:   interest since the standard approaches have been previously found
51:   inadequate to describe phenomena such as the dissipation of cardiac
52:   wave fronts and the shape of action potential at repolarization. The
53:   proposed asymptotic description overcomes these deficiencies by
54:   allowing, among other non-Tikhonov features, that a dynamical
55:   variable may change its character from fast to slow within a single
56:   solution. The general asymptotic approach is best demonstrated on an
57:   example which should be both simple and generic. 
58:   The classical model of Purkinje fibers (Noble, 1962) has the
59:   simplest functional form of all cardiac models but according to the current
60:   understanding it assigns a physiologically incorrect role to the Na
61:   current. This leads us to suggest an ``Archetypal Model'' with the simplicity of the Noble
62:   model but with a  structure more typical to contemporary cardiac
63:   models. We demonstrate that 
64:   the Archetypal Model admits a complete asymptotic solution in quadratures.
65:   To validate our asymptotic approach, we proceed  to consider an
66:   exactly solvable ``caricature'' of the Archetypal Model and 
67:   demonstrate that the asymptotic of its exact solution
68:   coincides with the solutions obtained by substituting the ``caricature''
69:   right-hand sides
70:   into the asymptotic solution of the generic Archetypal Model.
71:   This is necessary, because, unlike in standard asymptotic
72:   descriptions, no general results exist which can guarantee the
73:   proximity of the non-Tikhonov asymptotic solutions to the solutions
74:   of the corresponding detailed ionic model.   
75: % 258 words
76: \end{abstract}
77: 
78: \textbf{Keywords} 
79: excitability,
80: action potential,
81: asymptotic methods,
82: singular perturbations.
83: 
84: \pagestyle{myheadings}
85: \thispagestyle{plain}
86: \markboth%
87:   {V.~N.~BIKTASHEV \etal{} (2007/04/04)}%
88:   {Asymptotic analysis of cardiac excitation (2007/04/04)}
89: 
90: 
91: \tableofcontents
92: 
93: \section{Introduction}
94: \label{S:Intro}
95: 
96: \subsection{Physiological and mathematical motivation}
97: \label{S:Intro:Motivation}
98: 
99: Mechanical activity of the heart is controlled by electrical
100: excitation of cardiac cells, characterized by
101: ``action potentials'' (APs) across their membranes.
102: Abnormalities of cardiac rhythm are a major public health hazard
103: \cite{Zipes-Wellens-1998}, and great efforts are directed to the
104: mathematical modelling of APs. Investigations at various levels of
105: membrane, cellular and myocardial organisation have lead to the
106: development of a large number of detailed ionic
107: cardiac models (for reviews, see \eg{}~\cite{%
108:   Glass-etal-1991,%
109:   Holden-Panfilov-1997,%
110:   Kohl-etal-2000,%
111:   Winslow-etal-2000,%
112:   Clayton-2001%
113: }) for  various types of cardiac cells. These detailed models are
114: realistic in the sense that they demonstrate good agreement
115: with experimental data available to their authors at the time 
116: of creating the model. There even exists an
117: optimistic view that with the help of detailed cardiac computational
118: models ``it will soon be possible to do \textit{in silico} experiments that
119: would be impossible, difficult or unethical in animals or patients''
120: \cite{Clayton-2001}.   
121: 
122: In reality however, detailed ionic models of cardiac excitation
123: are immensely complicated. Their
124: analytical solution is impossible while their simulations are expensive
125: especially in three dimensions. Thus, numerous attempts have been made
126: to construct simplified mathematical models of cardiac action potentials (APs); for
127: examples, see \cite{%
128:   vanCapelle-Durrer-1980,%
129:   Karma-1993,%
130:   Aliev-Panfilov-1996,%
131:   Fenton-Karma-1998,%
132:   Rogers-2000,%
133:   Duckett-Barkley-2000,%
134:   Hinch-2002,%
135:   Echebarria-Karma-2002,%
136:   Bernus-etal-2002%
137: }.
138: Detailed ionic cardiac models were initially constructed as variations
139: of the spectacularly successful Hodgkin-Huxley model of nerve
140: excitability \cite{Hodgkin-Huxley-1952}. In a similar way the most
141: simplified cardiac models are often based on the elegant
142: FitzHugh-Nagumo two-variable reduction
143: \cite{FitzHugh-1961,Nagumo-etal-1962} of the Hodgkin-Huxley model.  A
144: typical way to create a simplified model is either to make an
145: appropriate functional generalization of the FitzHugh-Nagumo system,
146: or truncate a detailed ionic models. Then the modellers proceed to fit
147: the parameters of their simplified models to reproduce the AP shape or
148: other selected properties of the chosen detailed ionic model. 
149: 
150: This modelling approach has an obvious flaw, as such simplified models
151: are not reliable outside the range of phenomena which they have been
152: fitted to reproduce. So it would be desirable to \emph{derive} a
153: simplified model from a detailed model, based on a well defined set of
154: verifiable assumptions. One possible way to do that is via asymptotic
155: methods, which would utilize small parameters available in the
156: detailed model.
157: 
158: More importantly, reducing the \emph{number of equations} in the model often
159: delivers only minor reduction of its computational complexity.
160: Paradoxically, the better a simplified model is, the better it
161: reproduces another difficult feature of realistic models, their
162: \emph{stiffness}. 
163: That is,
164: detailed ionic models typically
165: have small parameters which considerably complicate their simulation.
166: However it would be natural to try and eliminate those parameters by
167: asymptotic methods, and resulting problems without small parameters
168: should be much easier for computational study.
169: 
170: Yet another paradoxical property of ionic models is that increasing
171: the number of physiological details does not necessarily make the
172: models more reliable, as experimental data are not always sufficient
173: for unequivocal identification of model parameters.  As argued by
174: Cherry and Fenton~\cite{Cherry-Fenton-2006}, detailed ionic models of
175: the same types of cells in the same species, but developed by
176: different authors, may disagree significantly. Thus there is a demand
177: in modelling practice for models that would be realistic
178: physiologically but independent on experimentally unreliable details,
179: i.e. depend on fewer parameters. Again, one possible way to create
180: such a model is to simplify a ``too detailed'' model by eliminating
181: exceeding details by some sort of asymptotic procedure.
182: 
183: So there is more than one serious reason to look more carefully into
184: the small parameters in detailed ionic models. The progress in
185: this direction was hampered for some time by implicit assumption,
186: induced by the success of the FitzHugh-Nagumo system, 
187: that
188: the essential properties of cardiac APs can be captured by dynamical
189: systems of the type
190: \begin{gather}
191: \label{tikhonov}
192:   \eps \Df{\x}{\t} = \f(\x,\y), \quad  \Df{\y}{\t} = \g(\x,\y), \qquad
193:    \x \in\Real^k, \;
194:    \y \in \Real^l, \;
195:   \eps \ll 1,
196: \end{gather}
197: in which some of the dynamic variables are fast ($\x$) while others are
198: slow ($\y$) and where $\eps>0$ is a small parameter.  The asymptotic
199: structure of \eqref{tikhonov} is mathematically very
200: convenient. Indeed, the presence of the small parameter $\eps$ at the
201: derivatives  allows to ``dissect'' equations 
202: \eqref{tikhonov}, in the leading order in $\eps$, into a slow-time (degenerate) subsystem
203: \begin{gather}
204: \label{tikhonovslow}
205:   0=\f(\x,\y), \quad   \Df{\y}{\t}=\g(\x,\y),
206: \end{gather}
207: and a fast-time subsystem 
208: \begin{gather}
209: \label{tikhonovfast}
210:   \Df{\x}{\T}=\f(\x,\y), \quad   \Df{\y}{\T}=0,
211: \end{gather}
212: with  $\T=\eps\,\t$, which are much easier to study. 
213: It is assumed that the fast subsystem \eqref{tikhonovfast} has, for all relevant values of $v$,
214: no attractors but isolated, asymptotically stable equilibria. 
215: There exist a classical theory stating the general conditions
216: which guarantee that, in the limit of $\eps\to0$, 
217: any finite segment of the graph of solution
218: of the full equations \eqref{tikhonov} 
219: approaches uniformly to that of the degenerate subsystem
220: \eqref{tikhonovslow} which, in turn, consists of slow-motion parts and
221: fast-motion parts separated by junction points. The slow-motion parts
222: are on the $l$-dimensional ``slow manifold'' determined by the first $k$
223: equations in \eqref{tikhonovslow} and the fast-motion parts
224: are arcs of trajectories of \eqref{tikhonovfast} within leaves
225: of the ``fast foliation'' $\y=\const$ \cite[p.173]{Mishchenko-Rozov}.
226: A central result of the theory of slow-fast systems is the Tikhonov
227: theorem \cite{Tikhonov-1952}. 
228: 
229: A paradigm for applying the theory of fast-slow systems to biological
230: excitability has been laid down by Zeeman \cite{Zeeman-1972}. His
231: fundamental idea is a generalization of the FitzHugh-Nagumo view and
232: states that the resting state is the unique and globally stable
233: equilibrium of the full system, but in the fast subsystem, it is only
234: one of three equilibria. There is another stable equilibria
235: corresponding to the excited state, and the two stable equilibria are
236: separated by an unstable equilibrium which thereby represents the
237: threshold of excitation. The repolarization from the excited state, which is
238: an equilibrium in the fast subsystem but not in the full system, to
239: the resting state which is the true equilibrium, happens via the slow
240: system, and its details, i.e. the shape of the action potential,
241: depends on the structure of the slow manifold.
242: 
243: Thus, we refer to excitable systems with the asymptotic structure of
244: equations \eqref{tikhonov} as FitzHugh-Nagumo type or Tikhonov-Zeeman
245: systems.
246: 
247: Extension of this ideology to spatially-extended excitable systems
248: produces a very attractive and promising asymptotic theory, see
249: e.g. \cite{Tyson-Keener-1988} for a review.  In this theory,
250: description of excitation waves is decoupled into description of fast
251: motion of their sharp ``fronts'' and ``backs'', and description of slow parts of
252: APs outside fronts and backs. After appropriate rescaling, both the
253: fast and the slow subsystems do not have the small parameter in them,
254: so their numerical simulation could be implemented efficiently.
255: 
256: However, this attractive theory was never really applied to cardiac
257: ionic models. It appears that FitzHugh-Nagumo type systems fail, in
258: principle, to reproduce some \emph{qualitative} features of ionic models.
259: Here are some examples.
260: 
261: \subsubsection*{Features of cardiac excitability:}
262: 
263: \begin{enumerate} 
264: %
265: \item \label{slow-repolarization} \textbf{Slow repolarization.}
266: Cardiac APs tend to have very fast upstrokes and much slower
267: other phases, including repolarization (``downstroke'').  In the
268: asymptotic limit $\eps\to0$, a FitzHugh-Nagumo system of the form
269: \eqref{tikhonov} with $l=1$ will have a fast, i.e. duration
270: $\t\sim \eps$, upstroke of the action potential will have
271: also fast, $\sim\eps$ downstroke. In principle, this can be avoided
272: if the slow manifold, defined by $\f(\x,\y)=0$, has a cusp singularity
273: with respect to foliation $\y=\const$, which is theoretically possible
274: if $l\geq2$ \cite{Zeeman-1972}. However, our attempts to identify such
275: a cusp singularity in cardiac equations have not been successful
276: \cite{Suckley-Biktashev-2003b,Biktasheva-etal-2006}
277: %
278: \item \label{slow-subthreshold} \textbf{Slow subthreshold response.}
279: An excitable system reacts to a subthreshold stimulus by an
280: immediate return to the resting state. In the asymptotic limit
281: $\eps\to0$, a FitzHugh-Nagumo system of the form
282: \eqref{tikhonov} will have both subthreshold return and super-threshold
283: upstroke very fast, $\t\sim\eps$. However, subthreshold
284: return in real cells and realistic models has speed comparable to the
285: slow stages of the AP, i.e. much slower than the upstroke.
286: %
287: \item \label{accommodation} \textbf{Fast accommodation.}
288: If the stimulus is applied not instantly but gradually, then the
289: threshold of excitation may increase, and if the perturbation is too
290: slow, the system may fail to generate AP altogether (phenomenon known
291: as accommodation). FitzHugh-Nagumo systems demonstrate accommodation
292: but only if the time scale of the stimulus is comparable to the
293: duration of the AP. In real cells and realistic models, accommodation
294: is observed for much faster stimuli, comparable more to the upstroke
295: duration than to the AP duration.
296: %
297: \item \label{variable-peak} \textbf{Variable peak voltage.}
298: The maximum of the AP in a single cell does not, in the first
299: approximation, depend on the way the AP has been elicited. Moreover,
300: the asymptotic theory of \cite{Tyson-Keener-1988} predicts that the
301: maximum of the AP in a propagating wave will be the same as in a
302: single cell.  However, in real cells and in realistic models the
303: maximum of AP does depend on the mode of excitation, and in the
304: propagating AP it could be significantly different than in a single
305: cell.
306: %
307: \item \label{front-dissipation} \textbf{Front dissipation.}
308: The fast accommodation of realistic models has an interesting
309: consequence for propagating waves. If excitability ahead of the wave
310: is temporarily blocked by a transient process, the wave will expire if
311: the block lasts too long. In a FitzHugh-Nagumo type system, this will
312: happen if the duration of the block equals the duration of the wave;
313: that is, the wave will cease when its 
314: ``length'', understood as the spatial size of the excited zone, has
315: decreased to zero \cite{Weiss-etal-2000}.  In contrast, in ionic models the
316: front looses its sharpness, ``dissipates'', much sooner than AP finishes,
317: and after that fails to propagate even though AP continues in
318: many cells and the excitability of the tissue ahead of the wave has
319: fully recovered 
320: \cite{Biktashev-2002,Biktasheva-2003,Biktashev-Biktasheva-2005}.
321: %
322: \end{enumerate}
323: 
324: In this article, we describe a non-Tikhonov asymptotic approach to
325: cardiac excitation. Its purpose is to overcome the limitations of
326: simplified models of FitzHugh-Nagumo type discussed above. The
327: approach has already been used in some form in
328: \cite{Simitev-Biktashev-2006} to achieve numerically accurate
329: prediction of the front propagation velocity (within 16\,\%) and its
330: profile (within $0.7$\,mV) for a realistic model of human atrial
331: tissue \cite{CRN}. The asymptotic reduction was sufficiently simple to
332: allow the derivation of an analytical condition for propagation block
333: in a re-entrant wave which was in an excellent agreement with results
334: of direct numerical simulations of the realistic atrial ionic
335: model. This has been achieved by considering only the ``fast
336: subsystem'' of the full model. 
337: 
338: 
339: Here we take a further step and present a complete asymptotic
340: description of a simple cardiac excitation model, including both fast
341: subsystem and slow subsystem, i.e. describing the whole AP rather than
342: the upstroke only. A brief sketch of this description has been
343: outlined in \cite{Biktashev-Suckley-2004}. That sketch has left
344: several questions unanswered, and the purpose of the present article
345: is to fill in the gaps.
346: The asymptotic reductions we propose are based on a well-defined and
347: verifiable set of assumptions and in this sense they are ``derived''
348: from a detailed ionic model.  We do not include arbitrary fitting
349: parameters which limit the model to the reproduction of few
350: hand-picked AP properties. All arbitrariness is restricted to choice
351: of small parameters in the model, which is the key stage in any
352: asymptotic approach. In our approach, the main small parameter
353: occurs in an essentially different way from \eqref{tikhonov}.
354: Consequently, the results of the classical
355: theory of slow-fast systems \cite{Mishchenko-Rozov} are not
356: applicable. In other words, there is no existing rigorous theory which can
357: guarantee that the asymptotic solutions are close to the true
358: solutions of the corresponding detailed ionic model. 
359: To deal with this issue, we formulate a caricature model which exactly
360: duplicates the asymptotic structure of a detailed ionic model but
361: allows exact analytical solution. We proceed to apply the asymptotic
362: procedure to this caricature and to compare the exact and the
363: asymptotic solutions in order to validate our approach.
364: 
365: To demonstrate our approach, we need to chose an appropriate gated
366: ionic model. Such a selected model must satisfy two criteria. Firstly
367: it should be as simple as possible. Secondly, it should have the
368: generic structure similar to all or most of contemporary ionic
369: models. The first requirement is best satisfied by the classical Noble
370: model of excitability of cardiac Purkinje fibers \cite{N62}. This
371: model has spawned generations of models of increasing accuracy and
372: complexity up to modern models with more than sixty differential
373: equations per single cell~\cite{Iyer-etal-2004}. The Noble model,
374: however, does not satisfy the second requirement. Indeed, while that
375: model reproduces the AP of Purkinje fibers in detail, it does not
376: correctly reflect their physiology.  As the model was constructed
377: before sufficient experimental data on the ionic currents became
378: available \cite{Noble-Rudy-2001}, the inward sodium current was given
379: the dual role of generating the upstroke and maintaining the
380: plateau. To avoid this peculiarity and to ensure that our asymptotic
381: procedure is sufficiently general, we propose an ``Archetypal Model'' (AM) which has
382: the generic structure of modern cardiac models but keeps the
383: functional simplicity of the the Noble model and is identical to it in
384: the asymptotic limit.  The asymptotic procedure is then demonstrated
385: on the example of this Archetypal Model.
386: 
387: The structure of the paper is as follows. We conclude the
388: Introduction with a short discussion of the procedure of parametric
389: embedding which is an important instrument in our work.  In
390: section~\ref{S:Noble}, we use a set of numerical observations to
391: postulate a system of axioms for a non-Tikhonov parametric embedding
392: of the Noble model.  In section~\ref{S:AM}, we describe the AM and
393: discuss its similarities with contemporary ionic models
394: and then study the asymptotic limits of the AM and obtain analytical
395: solutions in quadratures in these limits.
396: In section~\ref{S:Caricature}, we formulate the exactly solvable
397: ``caricature'' simplification of the AM.  In
398: section~\ref{S:Validation}, we validate our general asymptotic results
399: by a comparison of the limits of the exact solution of the caricature
400: model to its solution in the asymptotic limit. We conclude by
401: outlining the most essential features of our approach and discussing
402: possibilities for its application.
403: 
404: In all asymptotics, we restrict consideration to the leading order,
405: with the exception of Appendix~\ref{S:Accurate} which is about a
406: first-order correction.
407: 
408: Some parts of this work (subsections %
409:   \ref{S:Intro:Embedding},
410:   \ref{S:Noble:Formulation},
411:   \ref{S:Noble:Axiomatic},
412:   \ref{S:AM:Formulation},
413:   \ref{S:AM:Asymptotics}%
414: )
415: appeared in a very brief form in
416: \cite{Biktashev-Suckley-2004} 
417: and are reproduced here in the full form, with additional details, for completeness
418: and convenience of reference in the rest of the article;
419: other parts (subsections %
420:   \ref{S:Noble:Naive},
421:   \ref{S:Noble:Explicit},
422: sections
423:   \ref{S:Caricature}
424: and
425:   \ref{S:Validation}
426: and appendices A, B and C
427: ) are entirely new.
428: 
429: 
430: \subsection{The procedure of parametric embedding}
431: \label{S:Intro:Embedding} % BS04: yes
432: 
433: The nonlinear problems of physical, 
434: chemical and biological applications do not normally have 
435: parameters which are literally approaching zero within their normal
436: range relevant to the application. Hence, a typical practical approach
437: to asymptotic reduction is to identify a ``small constant'', say $a$,
438: in the model and to replace it by a parameter $\epsilon$, so that the
439: original problem corresponds to $\epsilon=a$, whereas the asymptotic
440: formulae are obtained in the limit $\epsilon\rightarrow0$. A
441: mathematically equivalent modification of this procedure is based on
442: the following
443: 
444: \begin{definition}
445: A \emph{parametric embedding} 
446: with parameter $\epsilon$ of a function $f(x)$
447: is any function $\emb{f}(x;\epsilon)$ such that
448: $\emb{f}(x,1)= f(x)$ for all $x\in \dom(f)$. A parametric embedding in
449: the context of $\epsilon\rightarrow0$ is called \emph{asymptotic
450: embedding}. An embedding of a dynamical system corresponds to an
451: embedding of its generating vector field or map.
452: \end{definition}
453: 
454: 
455: Thus, the ``small constant'' $a$ is replaced by $\epsilon a$, and
456: then the original problem corresponds to $\epsilon=1$ while the
457: asymptotic analysis is performed in the limit
458: $\epsilon\rightarrow0$. As long as $a$ is a nonzero constant, the
459: limits $\epsilon\to$ and $\epsilon a\to$ are
460: mathematically equivalent. The purpose of the above definition is
461: uniformity, especially when the small parameters appear in more than
462: one place in the equations.   
463: From this perspective, the algorithm of obtaining an approximation
464: using, say, a partial sum of an asymptotic series is: the power
465: series in $\epsilon$ is truncated to a selected number of terms, and
466: then $\epsilon=1$ substituted in the result. When applied formally,
467: this may look counter-intuitive, and yet for reasons explained above,
468: this is precisely equivalent to what is always done when ``small
469: quantities up to a certain order'' are taken into account in any
470: asymptotic approach.
471: 
472: So, asymptotic analysis of a mathematical model by
473: necessity implies introduction of artificial small parameters, which
474: is  equivalent to drawing a curve in functional space,
475: $\emb{f}(x;\epsilon)$, with the only condition that at $\epsilon=1$
476: this curve passes through the given point $f(x)$. There are
477: infinitely many ways to do  this, and the question arises, which
478: of these embeddings is ``the correct'' or ``the better'' one. Unless
479: the resulting asymptotic series for the solutions are converging and
480: the error terms can be estimated, there is no obvious answer to this
481: question within a purely mathematical context. So the choice should be
482: based on practical considerations. 
483: The embedding should be such as to allow efficient asymptotic analysis. 
484: A better embedding should also
485: provide a better quantitative approximation for a selected
486: class of solutions to the original problem, and/or preserve better
487: their qualitative features of interest. For a given embedding, these aspects
488: can be verified by comparing solutions at $\epsilon=1$ with solutions at smaller
489: $\epsilon$. In terms of the more conventional ``small constant'' $a$
490: approach, the procedure is to verify if $a$ is indeed small enough to be
491: used for asymptotics and try to reduce it further and see how the
492: solutions behave. Note this can be done numerically, prior to any
493: analytical work.   
494: 
495: In this paper, we are interested in AP solutions. Typically, we will propose an embedding
496: and assess its quality by comparing the numerical AP solutions at
497: $\epsilon=1$ and $\epsilon=10^{-3}$. If the embedding is found
498: reasonable, we proceed to study the limit $\epsilon\rightarrow0$
499: analytically.   
500: 
501: \section{The Noble model}
502: \label{S:Noble}
503: 
504: \subsection{Formulation}
505: \label{S:Noble:Formulation} % BS04: yes
506: 
507: The original Noble \cite{N62} model may be simplified by an
508: adiabatic elimination of the super-fast $\m$-gate,  
509: \begin{subequations}
510: \label{N62}
511: \begin{align}
512: & \Df{\E}{\t} = \gone(\E)\,\mbarcub(\E)\,\h + \gtwo(\E)\,\n^4 + \gthree(\E),\\
513: & \Df{\h}{\t} = \fone(\E)\,\left(\hbar(\E) - \h \right), \\
514: & \Df{\n}{\t} = \ftwo(\E)\,\left( \nbar(\E) - \n \right),  
515: \end{align}
516: \end{subequations}
517: %
518: where
519: \begin{equation}
520: \begin{split}
521: & \gone(\E) = \CM^{-1} \gNa\left(\ENa-\E\right), \quad \gtwo(\E) = \CM^{-1} \gK\left(\EK-\E\right), \\
522: & \gthree(\E) = \CM^{-1} \left[\gNap\left(\ENa-\E\right) + \gKp(\E)\left(\EK-\E\right) \right], \\
523: & \gKp(\E)=1.2\exp\left((-\E-90)/50\right)+0.015\,\exp\left((\E+90)/60\right), \\
524: & \ybar(\E)=\alpha_y(\E)/\left(\alpha_y(\E)+\beta_y(\E)\right), \quad y=\h,\n,\m, \\
525: & f_y(\E) = \alpha_y(\E)+\beta_y(\E), \quad y=\h,\n, \\
526: & \alpha_m(\E)=\frac{0.1\,\left(-\E-48\right)}{\exp\left((-\E-48)/15\right)-1} ,  \quad
527:   \beta_m(\E) =\frac{0.12\,\left(\E+8\right)}{\exp\left((\E+8)/5\right)-1}, \\
528: & \alpha_h(\E)=0.17\,\exp\left((-\E-90)/20\right), \quad
529:   \beta_h(\E) =\frac{1}{\exp\left((-\E-42)/10\right)+1}, \\
530: & \alpha_n(\E)=\frac{0.0001\,\left(-\E-50\right)}{\exp\left((-\E-50)/10\right)-1}, \quad
531:   \beta_n(\E) =0.002\,\exp\left((-\E-90)/80\right),
532: \end{split}
533: \label{N62-funs}
534: \end{equation}
535: %
536: and 
537: \begin{equation}
538: \CM=12, \quad \gNa=400, \quad \gK=1.2, \quad \gNap=0.14, \quad \EK=-110 
539:         \quad \textrm{and} \quad \ENa=40.
540:                                                         \label{N62-consts}
541: \end{equation}
542: %
543: Here $\E$ is the transmembrane voltage with values $\E\in[\EK,\ENa]$,
544: and $\h$ and $\n$ are ``gating variables'' with ranges $\h\in[0,1]$,
545: $\n\in[0,1]$; more specifically, $\h$ is the inactivation gate of the fast sodium current $\INa$
546: (the first term in the equation for $\d\E/\d\t$)
547: and $\n$ is the activation gate of the slow potassium current $\IK$
548: (the second term in the equation for $\d\E/\d\t$).
549: Notice that $\gone(\E)\ge0$ (this represents an ``inward current'') 
550: and $\gtwo\le0$ (this represents an ``outward'' current).
551: Our choice for the value of the parameter $\EK$ differs from
552: the original value of $\EK=-100$ used in \cite{N62}. This transforms the
553: Noble system  \cite{N62} from a self-oscillatory to an excitable one.
554: Another possibility to achieve this effect is to increase the value of the 
555: coefficient at the first exponent in $\gKp$, as suggested  in
556: \cite{N62,Krinsky-Kokoz-1973}, which leads to similar
557: results \cite{Biktashev-Suckley-2004}.
558: Equations \eqref{N62} represent a good simplification of the Noble
559: model \cite{N62} as can be seen by the insignificant difference in the
560: solutions plotted in \fig{01}(a).  
561: \input{fig1.tex}
562: 
563: \subsection{Naive embedding}
564: \label{S:Noble:Naive} % BS04: no
565: 
566: Seeking further simplification, we note that the functions
567: $\mbarcub(\E)$ and $\hbar(\E)$ are approximately  
568: stepwise as illustrated in \fig{02}(a). Thus, as suggested in
569: \cite{Biktashev-2002,Biktashev-2003}, 
570: the crudest approximation of \eqref{N62} is given by 
571: % 
572: \begin{equation}
573: \mbarcub(\E) \approx \Heav(\E-\Em), \qquad \hbar(\E) \approx \Heav(\Eh-\E),
574:                 \label{PureHeaviside}
575: \end{equation}
576: %
577: where $\Em$, $\Eh$ satisfy $\mbarcub(\Em)=1/2$ and $\hbar(\Eh)=1/2$,
578: respectively and $\Heav(\cdot)$ is the Heaviside step function.
579: A similar approximation was done in several earlier simplified models,
580: e.g.~\cite{Fenton-Karma-1998,Echebarria-Karma-2002,Hinch-2002}.
581: They typically took, for simplicity, that $\Eh=\Em$. 
582: This however leads to unsatisfactory description of the 
583: front dissipation phenomenon~\cite{Biktashev-2003}.
584: 
585: The naive approximation \eqref{PureHeaviside} turns out to be unsuccessful as
586: shown in \fig{01}(a). The AP produced when \eqref{PureHeaviside}
587: is used does not have a plateau but returns immediately to the resting
588: state. The reason for this behaviour is revealed by an analysis of the
589: individual currents which are illustrated in \fig{01}(b) both for
590: the detailed system \eqref{N62} and for the approximation
591: \eqref{PureHeaviside}. In the detailed model, the sodium current
592: remains significant during the plateau phase, successfully
593: counteracting the potassium current for some time. In the 
594: approximated model, this current is virtually absent after the
595: initial upstroke. This is due to the fact that during the slow
596: decrease of the voltage, both the $\m$ and $\h$ gates remain close to
597: their quasi-stationary values $\mbar$, $\hbar$, and their product
598: $\W(\E)\equiv \mbarcub\hbar$ is exactly zero in the approximated model. In
599: contrast, in the detailed model, this product remains
600: significant and although it is much smaller than unity, when
601: multiplied by the large factor $\gNa$, produces a sodium current which
602: is comparable to the potassium and the leakage currents. The current
603: $\W(\E)$ is called the ``window'' sodium current, because it runs in the region of
604: voltages between $\Eh$ and $\Em$ where the gates are supposed to be
605: ``almost closed''~\cite{Attwell-etal-1979}. 
606: 
607: \subsection{Axiomatic embedding}
608: \label{S:Noble:Axiomatic} % BS04: yes
609: 
610: In this section we consider a more elaborate parametric embedding, 
611: the asymptotic limit of which dissects equations \eqref{N62} into simpler
612: subsystems.
613: It 
614: is based on a number of  observations  of the properties of
615: the Noble model, which will be discussed below, and may be formalized
616: in the following
617: 
618: \begin{axioms}
619: The functions $\gNa$, $\fone(\E)$, $\hbar(\E)$ and $\mbarcub(\E)$ are
620: parametrically embedded in the functions $\emb{\gNa}(\eps)$,
621: $\emb{\fone}(\E;\eps)$, $\emb{\hbar}(\E;\eps)$ and
622: $\emb{\mbarcub}(\E;\eps)$, $\eps>0$,  such that  
623: % 
624: \begin{axiomlist}{\rule{5em}{0mm}}
625:   \item[Axiom I.] $\emb{\gNa}(\epsilon)=\eps^{-1}\gNa$,
626:   \item[Axiom II.] $\emb{\fone}(\E;\eps)=\eps^{-1}\fone(\E)$,
627:   \item[Axiom III.] $\lim\limits_{\eps\rightarrow0} \; \emb{\mbarcub}(\E;\eps)
628:  =\M(\E)\,\Heav(\E-\Em)$, \\
629: where  the function $\M(\E)$ is close to $\mbarcub(\E)$ for $\E>\Em$, 
630:   \item[Axiom IV.] $\lim\limits_{\eps\rightarrow0} \;
631:     \emb{\hbar}(\E;\eps)=\H(\E)\,\Heav(\Eh-\E)$  \\
632:     where the function $\H(\E)$ is close to $\hbar(\E)$ for $\E<\Eh$,
633:   \item[Axiom V.] $\Em>\Eh$,
634:   \item[Axiom VI.] $\lim\limits_{\eps\rightarrow 0} \;
635: \left(\eps^{-1}\emb{\mbarcub}(\E;\eps)\,\emb{\hbar}(\E;\eps)\right)
636: \equiv\Wtilde(\E) > 0$, \\
637: where  $\Wtilde(\E)$ is close to 
638: the window current $\W(\E)\equiv\mbarcub(\E)\hbar(\E)$,
639:   \item[Axiom VII.]  $\displaystyle \lim\limits_{\eps\rightarrow0}
640:     \emb{\S}(\E;\eps) \equiv
641: \lim\limits_{\eps\rightarrow0}
642:     \left(\emb{\mbarcub}(\E;\eps) \df{}{\E}\emb{\hbar}(\E;\eps)\right) =0$.
643: \end{axiomlist}
644: \end{axioms}
645: 
646: Indeed, the permittivity of the Na current $\gNa=400$ is large compared to
647: those of the other currents  $\gK=1.2$, $\max(\gKp)=1.8$ and
648: $\gNap=0.14$ and thus the values of associated small 
649: constants $g_x/\gNa$, $g_x=\gK, \gKp,\gNap$ are of the order of
650: $10^{-2}$. This observation is formalized in Axiom I by an introduction of
651: a small parameter $\eps$ which divides $\gNa$. 
652: %
653: Axiom II is postulated on the basis of the observation that
654: the function $\fone(\E)$ is large compared to $\ftwo(\E)$ as
655: illustrated in \fig{02}(b). These functions are reciprocal to the
656: time-scale functions of the gates $\h$ and $\n$ and therefore $\h$ is
657: a fast variable while $n$ is slow. The speed of $\h$ is even comparable to
658: $\E$ during the upstroke. These observations are justified in
659: \cite{Biktashev-2002,Biktashev-2003}, where we argued that although in healthy tissue  $\E$
660: can be, or at least seems, faster than $\h$, there are important
661: applications where the two variables should be considered of
662: comparable speed. 
663: %
664: Axioms III (IV) are suggested by the fact that functions
665: $\mbarcub(\E)$ ($\hbar(\E)$) are close to $1$ above (below) some
666: switch voltages $\Em$ ($\Eh$), and almost vanish
667: otherwise as seen in \fig{02}(a). The two Axioms do 
668: not give a precise definition of the functions $\H(\E)$ and $\M(\E)$, they
669: only  require that these are reasonably close to $\hbar(\E)$ and
670: $\mbarcub(\E)$ for those values of $\E$ where these functions are not
671: small. Here ``reasonably close'' means that replacement of  
672: $\hbar(\E)$ with $\H(\E)\,\Heav(\Eh-\E)$ and
673: $\mbarcub(\E)$ with $\M(\E)\,\Heav(\E-\Em)$ should not change
674: significantly the properties of the solution.
675: A possible choice for $\Em$ and $\Eh$ is so that they satisfy
676: $\mbarcub(\Em)=1/2$ and $\hbar(\Eh)=1/2$.
677: %
678: Axiom V is clearly satisfied for equations \eqref{N62} as shown in
679: \fig{02}(a). Some simplified models take $\Em=\Eh$ in similar situations~\cite{%
680:   Fenton-Karma-1998,%
681:   Hinch-2002,%
682:   Echebarria-Karma-2002%
683: }, however we already know that such simplification affects the
684: ``front dissipation'' feature of realistic models \cite{Biktashev-2003}.
685: Axioms III--V have a corollary that
686: \begin{equation}
687:   \lim\limits_{\eps\rightarrow0}\left(\emb{\mbarcub}(\E;
688:   \eps)\,\emb{\hbar}(\E; \eps)\right)=0.
689:   \label{Wsmall}  
690: \end{equation}
691: %
692: This indicates that the permittivity of the window current $\W(\E)$
693: is small of the order $\eps$. However,
694: as discussed in subsection~\ref{S:Noble:Naive},
695: it is a particular property
696: of the Noble model that the window current is finite since
697: it is multiplied by the large factor $\gNa$ which is of order
698: $\eps^{-1}$. To ensure this we postulate Axiom VI.
699: %
700: Finally, \fig{02}(b) demonstrates the plausibility of Axiom VII
701: where the graph of the function $\S(\E)\equiv\emb{\S}(\E;1)$ is shown.
702: \input{fig2.tex}
703: 
704: Thus, according to Axioms I--VII the parametric embedding of the model
705: \eqref{N62} is 
706: %
707: \begin{subequations} 
708: \begin{align}       
709: & \Df{\E}{\t} = \eps^{-1} \gone(\E)\, \emb{\mbarcub}(\E;\eps)\, \h + \gtwo(\E)\, \n^4 + \gthree(\E),	\label{embed01} \\
710: & \Df{\h}{\t} = \eps^{-1} \fone(\E)\, \left( \emb{\hbar}(\E;\eps) - \h \right), 			\label{embed02} \\
711: & \Df{\n}{\t} = \ftwo(\E)\, \left( \nbar(\E) - \n \right).  						\label{embed03}
712: \end{align}												\label{embed0}
713: \end{subequations}
714: %
715: First, we consider this system in the fast time $\T=\t/\eps$. Changing
716: the independent variable from $\t$ to  $\T$, taking the limit
717: $\eps\rightarrow0$ and using Axioms III and IV, we obtain the fast-time system,\footnote{
718:   Technically, the dynamic variables $\E$, $\h$ and $\n$ as
719:   functions of $\T$ ought to be denoted by different letters than the
720:   same variables as functions of $\t$. 
721:   We follow the tradition, however, and
722:   neglect this subtlety. 
723:   Later we comment on some cases to avoid
724:   ambiguity caused by this compromise.
725: }
726: %
727: \begin{equation}
728: \begin{split}
729: & \Df{\E}{\T} = \gone(\E)\, \M(\E)\, \Heav(\E-\Em)\, \h, \\
730: & \Df{\h}{\T} = \fone(\E)\, \left(\H(\E)\,\Heav(\Eh-\h) - \h \right),  \\
731: & \Df{\n}{\T} = 0. 							
732: \end{split}												\label{embed0fast} 
733: \end{equation}
734: %
735: As intended, the right-hand sides of the first two equations are
736: nonzero, thus we have two fast variables, $\E$ and $\h$.  The slow
737: manifold is the set of equilibria of this system and it is defined by
738: the finite equations 
739: %
740: \begin{align}
741: & \M(\E)\, \Heav(\E-\Em)\, \h = 0, \label{Econdition} \\
742: & \H(\E)\, \Heav(\Eh-\E) - \h = 0, \label{hcondition} 
743: \end{align}
744: %
745: since $\gone(\E)>0$, $\fone(\E)>0$ in the physiological range of the voltage,
746: $\E\in[\EK,\ENa]$.  Substitution of \eqref{hcondition} into
747: \eqref{Econdition} with account of Axiom V turns \eqref{Econdition}
748: into an identity as the product of the two Heaviside functions
749: vanishes. Thus, we have a codimension-1 slow manifold, defined by
750: equation \eqref{hcondition}. This is a non-Tikhonov feature since in
751: Tikhonov systems \cite{Mishchenko-Rozov}
752: the
753: codimension of  the slow manifold is equal to the number of fast variables. This 
754: peculiar feature results from \eqref{Econdition} becoming an identity if
755: \eqref{hcondition} is satisfied, which, in turn, is due to a near-perfect
756: switch behaviour of $\hbar(\E)$ and $\mbarcub(\E)$, becoming 
757: perfect switches in the limit $\eps\rightarrow0$. A consequence of this
758: feature is that the equilibria of the fast system are not
759: isolated. Therefore, Tikhonov's theorem \cite{Tikhonov-1952} is not
760: applicable as it requires asymptotic stability of equilibria of the
761: fast system which does take place here.
762: The fact that our parametric embedding is a non-Tikhonov one was 
763: already obvious from the dependence of \eqref{embed0} on the 
764: small parameter which is not of the form allowed by Tikhonov's
765: theorem, namely the large parameter  $\eps^{-1}$ in the first equation
766: multiplies only one of the terms in the right-hand side, rather than
767: the whole right-hand side.
768: 
769: Since Tikhonov's theorem cannot be used to describe the slow motion in
770: the standard way, we discuss it in more detail, however without
771: attempts of rigorous treatment. We consider system \eqref{embed0} in
772: the original (slow) time, and restrict our attention to trajectories
773: near the slow manifold, \ie{} ones for which $\h\approx\emb{\hbar}(\E;0)$.  
774: By re-arranging equation \eqref{embed02}, and assuming that when moving
775: along the slow manifold the derivative $\d{\h}/\d{\t}$ is of order
776: unity (this assumption is confirmed by the following result), we obtain
777: %
778: \begin{equation}
779:   \h = \emb{\hbar}(\E;\eps) - \frac{\eps}{\fone(\E)} \Df{\h}{\t} =
780:   \emb{\hbar}(\E;\eps) + O(\eps).
781: \label{h-mfd}
782: \end{equation}
783: From here we deduce that $\h=\emb{\hbar}(\E;\eps) + O(\eps)$.
784: Differentiating this with respect to time and substituting the result
785: back into the right-hand side of \eqref{h-mfd}, we obtain
786: %
787: \begin{equation}
788:   \h = \emb{\hbar}(\E;\eps) - \frac{\eps}{\fone(\E)}\,
789:   \df{\emb{\hbar}}{\E} \Df{\E}{\t} + O(\eps^2),
790: \label{h-mfd-2}
791: \end{equation}
792: which substituted in equation \eqref{embed01} yields
793: %
794: \begin{gather}
795:   \Df{\E}{\t} =
796:   \eps^{-1} \gone(\E)\, \emb{\mbarcub}(\E;\eps)\, \emb{\hbar}(\E;\eps) 
797:   -  \frac{\gone(\E)}{\fone(\E)}\, \emb{\S}(\E;\eps)\,\Df{\E}{\t}
798:   + O(\eps) 
799: %\nonumber\\
800: %  \mbox{} 
801: + \gtwo(\E)\, \n^4 + \gthree(\E),
802: \label{h-mfd-subs}
803: \end{gather}
804: %
805: where the function $\emb{\S}(\E;\eps)$ is defined in Axiom VII. 
806: Using Axiom VI for the first term and Axiom VII for the second term
807: and taking the limit $\eps\rightarrow0$ we obtain the slow-time system,
808: %
809: \begin{equation}
810: \begin{split}
811: & \Df{\E}{\t} = \gone(\E)\, \Wtilde(\E) + \gtwo(\E)\, \n^4 + \gthree(\E), \\
812: & \h = \H(\E)\, \Heav(\Eh-\E), \\
813: & \Df{\n}{\t} = \ftwo(\E)\, \left( \nbar(\E) - \n \right).
814: \end{split} 										\label{embed0slow}
815: \end{equation}
816: %
817: This is a system of two differential equations for the slow variables
818: $\E$ and $\n$ plus a finite equation for $\h$ defining the slow
819: manifold. Note that we have two slow variables in agreement with the
820: two dimensions of the slow manifold. Thus, $\E$ is both a fast and a
821: slow variable which is yet another non-Tikhonov feature.   
822: \input{fig3.tex}
823: 
824: \subsection{Explicit embedding}
825: \label{S:Noble:Explicit} % BS04: no
826: 
827: To show that Axioms I--VII are consistent and usable we need to
828: demonstrate that there exist embedding functions $\emb{\gNa}$, 
829: $\emb{\fone}$, $\emb{\hbar}$ and $\emb{\mbarcub}$ which satisfy these
830: axioms. The first two functions are already 
831: defined by Axioms I and II. Thus, in this section we suggest an explicit
832: dependence of $\emb{\mbarcub}$ and $\emb{\hbar}$ on $\eps$ which
833: satisfies Axioms III--VI, namely
834: %
835: \begin{subequations}
836: \begin{align}
837: & \emb{\mbarcub}(\E;\eps) \equiv \M(\E)\; \left(
838:    \eps^\p\, \Heav(\Eh-\E) + \eps^\r\, \Heav(\E-\Eh)\,\Heav(\Em-\E) + \Heav(\E-\Em)
839:   \right), 										\label{m-emb} \\
840: & \emb{\hbar}(\E;\eps) \equiv \H(\E)\; \left(
841:     \Heav(\Eh-\E) + \eps^{1-\r}\, \Heav(\E-\Eh)\,\Heav(\Em-\E) + \eps^\q\,\Heav(\E-\Em)
842:   \right), 										\label{h-emb} 
843: \end{align}										\label{mh-emb}
844: \end{subequations}
845: %
846: where $\p\in[1,+\infty)$, $\q\in[1,+\infty)$, $\r\in(0,1)$ and, of
847: course, we must have $\M(\E)=\mbarcub(\E)$, $\H(\E)=\hbar(\E)$ to
848: ensure these functions coincide with $\mbarcub(\E)$ and $\hbar(\E)$ at
849: $\eps=1$. It straightforward to verify, that Axioms III--VI are then
850: satisfied. To verify Axiom VII is more complicated: obviously, the above
851: derivation of the slow system \eqref{embed0slow} is not technically
852: valid for discontinuous $\emb{\hbar}$ as defined by \eqref{h-emb}, and
853: Axiom VII does not make sense literally. Still, it will be
854: satisfied if we assume that $\delta(\E-\Eh)\Heav(\E-\Em)=0$,
855: i.e. infinity times zero at $\E=\Eh$ equals zero. If $\p=\q=1$, we
856: have the asymptotic window current exactly equal to the window current
857: of the Noble model, $\Wtilde(\E)=\W(\E)$; if $\p>1$ and $\q>1$,
858: then the asymptotic window current is a cut-off version of the original,
859: $\Wtilde(\E)=\W(\E)\,\Heav(\E-\Eh)\,\Heav(\Em-\E)$. The adequacy of this
860: embedding for $\p=\q=1$, $\r=0.5$ is demonstrated in \fig{03}.
861: 
862: The explicit embedding \eqref{mh-emb} is rather
863: complicated. Formally, there are infinitely many embeddings;
864: \eg{}~equations \eqref{m-emb} and \eqref{h-emb} define a 
865: three-parameter family of embeddings, all satisfying the Axioms and
866: all leading to the same fast and slow systems and having the
867: same asymptotic properties.  And it is not possible to infer from the
868: original problem, which of the embeddings is ``correct''. The
869: discontinuity of $\emb{\hbar}$ and $\emb{\mbarcub}$ is also a cause for
870: concern. It is true that their limits in $\eps\rightarrow0$ have
871: to be discontinuous, or at least non-analytical, according to Axioms III
872: and IV, but with \eqref{m-emb} and \eqref{h-emb} they are discontinuous
873: already for any $\eps\ne1$, which is inconvenient even in this heuristic
874: treatment. Moreover, this is likely to cause serious technical difficulties in
875: any attempt of rigorous treatment of the problem. This difficulty can be
876: overcome, for instance, by using in \eqref{m-emb} and \eqref{h-emb},
877: instead of Heaviside functions, functions which are
878: smooth for all $\eps>0$ and only tend to Heaviside functions in the limit
879: $\eps=0$, \eg{}
880: %
881: \begin{equation}
882:   \emb{\Heav}(x;\eps)=1/(1+e^{-x/\eps}).
883: \end{equation}
884: %
885: Indeed, in that case the peak value of
886: $\left|\@\emb{\hbar}(\E;\eps)/\@\E\right|$ is attained at $\E=\Eh$
887: and is $(4\eps)^{-1}\hbar(\Eh)$ in the leading order, whereas the
888: value of $\emb{\mbarcub}(\Eh;\eps)$ is exponentially small in
889: $\eps$, thus their product is uniformly small and Axiom VII is
890: satisfied. 
891: 
892: \section{The Archetypal Model}
893: \label{S:AM}
894: 
895: \subsection{Formulation and parametric embedding}
896: \label{S:AM:Formulation} % BS04: yes
897: 
898: \input{fig4.tex}
899: 
900: The complications arising in the construction of a possible embedding
901: of the Noble model \eqref{N62}, as discussed in the preceding section,
902: are not essential. In fact, they arise only due to the fact that
903: insufficient experimental data was available at the time of the
904: construction of the Noble model and in particular the existence of a
905: Ca current was not yet discovered. Thus, the Na current was given a
906: dual role to produce  the upstroke and to keep voltage elevated during
907: the long plateau stage leading to the large Na window illustrated in \fig{01}(b).
908: Such a large window current is not present in other 
909: cardiac models. This is demonstrated in \fig{04}(b) in the
910: case of the  currently-accepted detailed ionic model of human atrial
911: cells of Courtemanche \etal{}~\cite{CRN}. 
912: Bearing in mind the possibility of extending our present analysis and results to
913: models of other types of cardiac cells, we propose a modified version
914: of the Noble model \eqref{N62}. It is similar in many aspects to
915: \eqref{N62} in but has a more ``generic cardiac'' structure and admits a
916: straightforward asymptotic embedding:
917: %
918: \begin{equation}
919: \begin{split}
920:  \Df{\E}{\t} & = \gone(\E)\, \M(\E)\, \Heav(\E-\Em)\, \h + \gone(\E)\,\W(\E) + \gtwo(\E)\, \n^4 + \gthree(\E) \\
921:              & = \gone(\E)\, \M(\E)\, \Heav(\E-\Em)\, \h + \gtwo(\E)\, \n^4 + \G(\E),\\
922:  \Df{\h}{\t} & = \fone(\E)\, \left(\H(\E)\, \Heav(\Eh-\E) - \h\right),\\
923:  \Df{\n}{\t} & = \ftwo(\E)\, \left(\nbar(\E) - \n\right),
924: \end{split}
925: \label{N62mod}
926: \end{equation}
927: %
928: where the functions $\mbar$, $\hbar$, $\fone$, $\ftwo$, $\gone$, $\gtwo$,
929: $\gthree$ are defined as in equations \eqref{N62}, and
930: \(
931:   \M(\E) = \mbarcub(\E), 
932: \) \(
933:   \H(\E) = \hbar(\E), 
934: \) \(
935:   \G(\E) = \gone(\E)\W(\E) + \gthree(\E),
936: \) \(
937:   \W(\E) = \mbarcub(\E) \hbar(\E).
938: \)
939: In this formulation, the perfect-switch behaviour of the Na gates is
940: represented by the Heaviside functions multiplying $\M$ and $\H$.
941: The deviation from the perfect-switch behaviour, due to the
942: window current component $\W=\mbarcub\hbar$, is separated from the fast
943: Na dynamics and appears as an additional ``time-independent'' current with a
944: role similar to $\gthree(\E)$. 
945: 
946: We shall call \eqref{N62mod} the ``Archetypal Model'' (AM).
947: The AP of this model is very similar
948: to the AP of the Noble model \eqref{N62} as shown in 
949: \fig{04}(a). Its simplest asymptotic embedding consistent with
950: Axioms I--VII can be done with $\eps$ introduced linearly and only in
951: two places, 
952: %
953: \begin{equation}
954: \begin{split}
955: & \Df{\E}{\t} = \eps^{-1} \gone(\E)\, \M(\E)\, \Heav(\E-\Em)\, \h + \gone(\E)\,\W(\E) + \gtwo(\E)\, \n^4 + \gthree(\E), \\
956: & \Df{\h}{\t} = \eps^{-1} \fone(\E)\, \left(\H(\E)\, \Heav(\Eh-\E) - \h\right),\\
957: & \Df{\n}{\t} = \ftwo(\E)\, \left( \nbar(\E) - \n \right).
958: \end{split}										\label{N62mod-emb}
959: \end{equation}
960: %
961: The quality of this embedding is illustrated in \fig{05}. 
962: The limit of the fast-time system is 
963: %
964: \begin{equation}
965: \begin{split}
966: & \Df{\E}{\T} = \gone(\E)\, \M(\E)\, \Heav(\E-\Em)\, \h, \\
967: & \Df{\h}{\T} = \fone(\E)\, \left( \H(\E)\, \Heav(\Eh-\E) - \h \right),\\
968: & \Df{\n}{\T} = 0,
969: \end{split}									\label{N62mod-embfast}
970: \end{equation}
971: %
972: and the limit of the slow-time system is 
973: %
974: \begin{subequations}
975: \label{N62mod-embslow}
976: \begin{align}
977: & \Df{\E}{\t} = \gone(\E)\, \W(\E) + \gtwo(\E)\, \n^4 + \gthree(\E),			\label{N62mod-embslow1}\\
978: & \h = \H(\E)\, \Heav(\Eh-\E), 								\label{N62mod-embslow2}\\
979: & \Df{\n}{\t} = \ftwo(\E)\, \left( \nbar(\E) - \n \right). 				\label{N62mod-embslow3}
980: \end{align}
981: \end{subequations}
982: %
983: These systems coincide with \eqref{embed0fast} and \eqref{embed0slow}, if
984: $\Wtilde(\E)=\W(\E)$. Their solutions and phase portraits are shown in
985: \fig{05} and \fig{06}, respectively.
986: 
987: \input{fig5.tex}
988: 
989: The AM \eqref{N62mod} produces an AP similar to
990: the Noble model \eqref{N62}. In fact, the agreement between the two
991: can be improved further as demonstrated in Appendix~\ref{S:Accurate}.
992: Moreover, the AM and the Noble model have the same asymptotic
993: limits. The AM has two advantages. Firstly, the relevant 
994: asymptotic embedding is much simpler: the right-hand sides of \eqref{N62mod-emb} 
995: linearly depend on $\eps^{-1}$, 
996: and it appears only in two places. Secondly and more importantly,
997: judging from \fig{04}(b), the asymptotic properties of the fast Na
998: current in modern detailed models 
999: are likely to be similar to those of the AM
1000: \eqref{N62mod} rather than to those of the Noble model
1001: \eqref{N62}. Therefore, we adopt the AM \eqref{N62mod} as the
1002: example on which the non-Tikhonov asymptotic procedure is demonstrated.
1003: 
1004: \subsection{Asymptotic analysis}
1005: \label{S:AM:Asymptotics} % BS04: yes
1006: 
1007: \input{fig6.tex}
1008: 
1009: \subsubsection{The fast-time subsystem}
1010: \label{S:AM:Asymptotics:Fast}
1011: 
1012: The fast-time subsystem \eqref{N62mod-embfast} of the AM
1013: governs the evolution of the two fast variables $\E$ and $\h$ on 
1014: a time scale $\t \sim \eps$ or equivalently $\T\sim 1$.
1015: Its solution does not depend on the third variable $\n$, which is slow
1016: and thus stays close to its initial value $\nini$ on this time scale. 
1017: 
1018: System \eqref{N62mod-embfast} admits exact analytical solution.
1019: If $\Eini<\Em$ the solution of the
1020: initial-value problem with $\E(0)=\Eini$, $\h(0)=\hini$ is  
1021: %
1022: \begin{equation}
1023: \begin{split}
1024: & \E = \Eini,\\
1025: & \h = \H(\Eini)\, \Heav(\Eh-\Eini) + \left(\hini - \H(\Eini)\, \Heav(\Eh-\Eini)\right)\, \exp\left(-\fone(\Eini)\, \T\right).
1026: \end{split}										\label{N62mod-fastsubthr}
1027: \end{equation}
1028: %
1029: If $\Eini>\Em$
1030: the solution of the same initial-value problem is given in quadratures by 
1031: \begin{subequations}
1032: \begin{align}
1033: & \h = \hini + \J(\Eini)-\J(\E),					\label{N62mod-fastupstrokesol1}\\
1034: & \T = \int\limits_{\Eini}^{\E} \frac{\d\Emute}{
1035:     \gone(\Emute)\, \M(\Emute)\, \left(\hini+\J(\Eini)-\J(\Emute)\right)
1036:   },
1037: 									\label{N62mod-fastupstrokesol2}
1038: \end{align}								\label{N62mod-fastupstrokesol}
1039: \end{subequations} 
1040: %
1041: where
1042: %
1043: \begin{equation}
1044: \J(\Emute) = \int\frac{\fone(\Emute)}{\gone(\Emute)\, \M(\Emute)} \,\d{\Emute}.
1045: \label{N62mod-Q}
1046: \end{equation}
1047: 
1048: Note that each trajectory of the fast-time subsystem
1049: \eqref{N62mod-embfast} crosses the slow manifold
1050: \eqref{N62mod-embslow2} only once as shown in \fig{06}(a). Thus,
1051: singularity of the slow manifold with respect to the trajectories of
1052: the fast system, which in Zeeman's \cite{Zeeman-1972} interpretation of Tikhonov systems
1053: determines the threshold of excitation, does not exist in the AM.
1054: To shed light on the threshold properties of our
1055: fast-time subsystem, let us consider  the maximal overshoot voltage 
1056: $\Einf\equiv\E(+\infty)$ in the system as a function of the initial 
1057: conditions. Taking into account that $\h(+\infty)=0$ we can use
1058: \eqref{N62mod-fastupstrokesol1} to find $\Einf$ as a
1059: solution of the finite equation $\J\left(\Einf\right)
1060: =\J(\Eini)+\hini$, provided that $\Eini>\Em$ and, of course, 
1061: $  \Einf = \Eini$, otherwise.
1062: Thus, the function $\Einf(\Eini,\hini)$ has a
1063: discontinuity along the line $\{(\Eini,\hini)\}=\{\Em\}\times(0,1]$ as
1064: shown in \fig{06}(a). This
1065: discontinuity is the mathematical equivalent of the physiological
1066: notion of excitability: the
1067: upstroke does take place if and only if $\Eini$ is above the threshold,
1068: $\Em$. Note that excitability in Tikhonov-Zeeman systems has this
1069: property, too, but is related to unstable branches of the slow
1070: manifold, rather than discontinuities in the fast flow.  
1071: 
1072: The numerical values of parameters of the Noble model from which our
1073: Archetypal Model descends, are such that $\fone(\E)/\gone(\E)$ is a
1074: uniformly small function of $\E$ in the physiological range. That is,
1075: condition $\hini=\J(\Einf)-\J(\Eini)\sim1$ requires
1076: relatively large values of the quadrature \eqref{N62mod-Q}, the
1077: integrand of which is generally a relatively small quantity. This is
1078: achievable if the integration interval goes close to the pole of the
1079: integrand, i.e. zero of $\M(\Emute)$. In other words, the noted smallness
1080: of $\fone/\gone$ necessitates that $\Einf\approx\ENa$ for typical 
1081: $\Eini$ and $\hini$. A more detailed and formal analysis of this aspect
1082: is given in Appendix~\ref{S:HighExc}.
1083: 
1084: \subsubsection{The slow-time subsystem}
1085: \label{S:AM:Asymptotics:Slow}
1086: 
1087: The slow-time subsystem \eqref{N62mod-embslow} of the AM
1088: governs the evolution of the two slow variables $\E$ and $\n$ on a
1089: time scale $\t\sim 1$ or $\T \sim \eps^{-1}$. In fact, 
1090: equations \eqref{N62mod-embslow1} and \eqref{N62mod-embslow3} form a
1091: closed subsystem. Gate $\h$ described by \eqref{N62mod-embslow2} can
1092: be evaluated if $\E(\t)$ is known but does not affect the dynamics of other variables.
1093: The slow-time system has been studied by Krinsky and Kokoz
1094: \cite{Krinsky-Kokoz-1973-3}. The slow-time subsystem \eqref{N62mod-embslow} is, in turn,
1095: a fast-slow system, this time in the classical Tikhonov sense.
1096: This is illustrated by \fig{07}(a) which compares the characteristic
1097: time-scale functions of the dynamical variables $\E$ and $\n$, defined
1098: as $\tau_y=\left|\@\dot{y}/\@y\right|^{-1}$ for $y=\E,\n$. It can be seen
1099: that $\E$ is a fast variable and $\n$ is a slow variable, which 
1100: motivates the introduction of a second small parameter $\epstwo>0$ in
1101: the following standard Tikhonov way, 
1102: %
1103: \begin{subequations}
1104: \begin{align}
1105: & \Df{\E}{\t} = \G(\E)+ \gtwo(\E)\, \n^4,  				\label{N62mod-slowsubemb-fast}\\
1106: & \Df{\n}{\t} = \epstwo \ftwo(\E)\, \left(\nbar(\E) - \n \right). 	\label{N62mod-slowsubemb-slow}
1107: \end{align}								\label{N62mod-slowsubemb}
1108: \end{subequations}
1109: 
1110: \input{fig7.tex}
1111: 
1112: System \eqref{N62mod-slowsubemb} is a fast-time system. In the limit
1113: $\epstwo \rightarrow 0$ the lines $\n=\n(\tini)=\const$ form the leaves
1114: of the fast foliation. On every such leaf the solution for the voltage
1115: is given by
1116: \begin{equation}
1117:   \t - \tini  = \int\limits_{\E(\tini)}^{\E(\t)}
1118:   \frac{\d{\Emute}}{\G(\Emute)+ \gtwo(\Emute)\, \n(\tini)^4}.
1119: \label{N62mod-slowfastsol}
1120: \end{equation}
1121: The corresponding slow-time system is obtained by the change of variable
1122: $\tslow=\epstwo \t$ and in the limit $\epstwo\rightarrow 0$ becomes
1123: %
1124: \begin{subequations}
1125: \begin{align}
1126: & 0 = \G(\E)+ \gtwo(\E)\, \n^4, 					\label{N62mod-slowslow1}\\
1127: & \Df{\n}{\tslow} = \ftwo(\E)\, \left( \nbar(\E) - \n \right).		\label{N62mod-slowslow2}
1128: \end{align}								\label{N62mod-slowslow}
1129: \end{subequations}
1130: %
1131: Equation \eqref{N62mod-slowslow1} defines the super-slow
1132: manifold,
1133: %
1134: \begin{equation}
1135:   \n = \Nss(\E) = \left( - \G(\E)/\gtwo(\E) \right)^{1/4},  		\label{N62mod-slowslowmfd}
1136: \end{equation}
1137: and equation \eqref{N62mod-slowslow2} describes the motion along this manifold.
1138: As illustrated in \fig{06}(b) the super-slow manifold is split into two
1139: parts by the condition $\n^4\ge0$: the 
1140: ``diastolic'' branch $\E\in(-\infty,\Eone]$ and the ``systolic'' branch for
1141: $\E\in[\Etwo,\Ethree]$, where $\Eone\approx-95.75$, $\Etwo\approx-61.81$
1142: and $\Ethree\approx1.86$ are roots of the equation $\G(\E)=0$. 
1143: The stability of the fast equilibria is determined by the sign of
1144: $\@\dot{\E}/\@\E$ which coincides with the sign of
1145: $\Nss'(\E)=\d{\Nss}/\d{\E}$: the stable branches of the super-slow
1146: manifold correspond to regions in $(\n,\E)$ plane where its graph has
1147: a negative slope, \ie{}~$\Nss'(\E)<0$. These are the regions of the
1148: entire diastolic branch and the upper part of the systolic branch,
1149: corresponding to $\E\in(\East,\Ethree]$, where $\East\approx-17.05$ is
1150: the root of the equation  $\Nss'(\East)=0$. These considerations
1151: determine the excitability properties  in terms of the super-slow subsystem
1152: \eqref{N62mod-slowslow}. As seen in \fig{06}(b) the
1153: threshold of excitability is 
1154: $\Etwo$ since a trajectory starting from $E(\tini)> \Etwo$ will be repelled  by the
1155: lower systolic branch and attracted by the upper one, thus making a
1156: relatively large excursion. This will be followed by a slow movement
1157: along the upper systolic branch and a jump to the diastolic branch at $\East$.
1158: On every monotonic branch of the super-slow manifold the equation
1159: \eqref{N62mod-slowslow1} can, in principle, be resolved with respect
1160: to $\E$ to give $\E=\Nss^{-1}(\n)$ and the result can be substituted
1161: into \eqref{N62mod-slowslow2} leading to the following quadrature
1162: solution of equations \eqref{N62mod-slowslow},
1163: %
1164: \begin{equation}
1165: \begin{split}
1166: & \tslow-\tslowC = \int\limits_{\nC=0}^{\n} \frac{\d\nmute}{
1167:     \ftwo\left(\Nss^{-1}(\nmute)\right)\,\left(\nbar\left(\Nss^{-1}(\nmute)\right)-\nmute\right)
1168:   }, \\
1169: & \E=\Nss^{-1}\left(\n(\tslow)\right),
1170: \end{split}						\label{N62mod-slowslow.sol1}
1171: \end{equation}
1172: %
1173: where $\tslowC$ is the slow time of the beginning of this asymptotic
1174: stage. Alternatively, we may use, as suggested in \cite{N62},  $\E$
1175: rather than $\n$ as a coordinate on the super-slow manifold. To do that we substitute
1176: $\n=\Nss(\E)$ into the second equation 
1177: \[
1178:   \Df{\E}{\tslow}=\frac{\ftwo(\E)\,\left(\nbar(\E)-\Nss(\E)\right)}{\Nss'(\E)}, 
1179: \]
1180: where from solution of the slow-time system in quadratures follows
1181: without the need to invert the function $\Nss(\E)$.
1182: In particular, the
1183: time between the points $(\E,\h)=(\Ethree,0)$ and $(\East,\Nast)$,
1184: \ie{}~the duration of the plateau of an AP starting from
1185: $\n=0$, is
1186: \begin{equation}
1187: \tslow(\East)-\tslow(\Ethree)  
1188: = \int\limits_{\Ethree}^{\East} \frac{ \Nss'(\E)\,
1189:     \d{\E}}{\ftwo(\E)\,\left( \nbar(\E)-\Nss(\E)\right)},
1190: \label{N62-slowslowsol}
1191: \end{equation}
1192: where   $\Nast=\Nss(\East)\approx 0.6512$ for the standard  parameter values.
1193: 
1194: This completes a brief overview of the leading-order asymptotics of the
1195: Archetypal Model.  An inventory of resulting formulas describing all the stages of
1196: typical solutions corresponding to various sorts of initial conditions,
1197: is given in Appendix~\ref{S:Synthesis}.
1198: 
1199: 
1200: \section{The caricature model}
1201: \label{S:Caricature} % BS04: no
1202: 
1203: \subsection{Motivation}
1204: \label{S:Caricature:Motivation}
1205: 
1206: As pointed out in the Introduction and throughout the article, the
1207: asymptotic structure of the embeddings of both the Noble and the
1208: Archetypal Model is non-Tikhonov and the results of the
1209: classical theory of slow-fast systems are not applicable. 
1210: The fact that we are able to demonstrate a good numerical agreement in
1211: \figs{03}, \figref{05} and \figref{07} is
1212: reassuring but far from reliable since the numerical solutions are, by their
1213: nature, only approximate. Developing an alternative to the classical
1214: slow-fast asymptotic theory is beyond the scope ot this paper.
1215: Instead, in this section we propose and study a simple caricature model of cardiac
1216: excitation. One can think of the caricature model as a detailed ionic model which
1217: allows an exact solution and which has been embedded in a non-Tikhonov
1218: system. Once exact solutions are available, their the proximity
1219: to asymptotic solutions can be proved in this particular
1220: case. We have to emphasize that, the caricature is not ``derived''
1221: from the Noble or from the Archetypal Model. They are,
1222: once again, used only as starting point for their simple functional forms.
1223: The caricature is constructed so that it has an asymptotic structure
1224: identical to that of 
1225: the AM of section \ref{S:AM} and differs from it in that the
1226: functions in its the right-hand side are chosen so as to make
1227: possible to {\bf(a)} evaluate the  quadratures of section
1228: \ref{S:AM:Asymptotics} and thus obtain explicit asymptotic solutions and
1229: {\bf(b)} to go a step further and find an exact analytical solution of
1230: the simple model. We then proceed to to demonstrate that the
1231: appropriate limits of the exact analytical solution of this caricature
1232: model coincide with  its solution in the asymptotic limits as given by
1233: quadratures 
1234: \eqref{N62mod-fastsubthr}, \eqref{N62mod-fastupstrokesol},
1235: \eqref{N62mod-slowfastsol},
1236: \eqref{N62mod-slowslow.sol1}.  Such an effort is still not a proof of
1237: our asymptotic analysis in general, but makes a good
1238: justification. 
1239: 
1240: \subsection{Formulation} 
1241: \label{S:Caricature:Formulation}
1242: 
1243: In order to formulate a caricature of the AM
1244: \eqref{N62mod-emb} we keep the asymptotic structure of the
1245: latter and replace (``approximate'') the functions forming its right-hand
1246: side with simpler ones. We make the following simplifications.
1247: \begin{enumerate}
1248:  \renewcommand{\labelenumi}{(\alph{enumi})}
1249: %
1250: \item We replace the functions $\M(\E)$ and $\H(\E)$ by unity. Thus,
1251: the function $\mbarcub(\E)$ is replaced by $\Heav(\E-\Em)$ and the
1252: function $\hbar(\E)$ by
1253: $\Heav(\Eh-\E)$.  Analogously, we replace the function $\nbar(\E)$ by
1254: $\Heav(\E-\En)$. The values of the constants $\Em$, $\Eh$ and $\En$
1255: will be discussed in item (e) below.
1256: \item We replace the function $\fone(\E)$ by the constant
1257: $\Fone\equiv1/2$. Analogously,  we replace the
1258: function $\ftwo(\E)$ by the constant $\Ftwo\equiv 1/270$.
1259: %
1260: \item We replace the function $\G(\E)$ by the continuous
1261: piecewise linear function, 
1262: \begin{equation}
1263:   \Gtilde(\E) = \left\{\begin{array}{lll}
1264:   \kone (\Eone-\E), &  \E\in(-\infty,\Edagger), \\
1265:   \ktwo (\E-\Etwo), &  \E\in[\Edagger,\East), \\
1266:   \kthree (\Ethree-\E), &  \E\in[\East,+\infty),
1267:   \end{array}\right.                             
1268: \label{Gtilde}
1269: \end{equation}
1270: where the constants $\kone\equiv0.075$, $\ktwo\equiv1/25$,
1271: $\kthree\equiv1/10$ while the constants $\Eone\equiv-280/3$, $\Etwo\equiv-55$ and
1272: $\Ethree\equiv1$ are close
1273: to the roots of $\G(\E)$ (%
1274:   which are $\approx-95.75$, $-61.81$ and   $1.86$ respectively; we prefer
1275:   to avoid multi-digit decimal values%
1276: ). 
1277: The constants
1278: $\East\equiv-15$ and $\Edagger\equiv-80$ are determined by the
1279: intersection points of the three linear pieces of $\Gtilde(\E)$.
1280: Note that at $\East$ the function $\Gtilde(\E)$ reaches its local
1281: maximum and thus  $\East$ corresponds to the point on the systolic branch of the
1282: super-slow manifold  which forms the boundary between its stable and
1283: unstable parts. 
1284: %
1285: \item We replace the function $\gtwo(\E)$ by
1286:   $\Gtwoh\Heav(\E-\Edagger)$, where $\Gtwoh=-9$.
1287: %
1288: \item Finally, we set $\Em\equiv\East$,
1289: $\Eh=\En\equiv\Edagger$.  A more accurate approximation would be to choose
1290: these constants as the solutions of the equations
1291: $\mbarcub(\Em)=1/2$, $\hbar(\Eh)=1/2$, and $\nbar(\En)=1/2$,
1292: respectively. However, this would lead to a caricature model, the
1293: right-hand side of which would consist of six pieces instead of only
1294: three as with the present choice. Obviously, this is not a principal
1295: difficulty but would be rather technically inconvenient. Moreover, the
1296: present choice preserves the relatively good quantitative agreement
1297: with the original models without the additional complications.
1298: %
1299: \end{enumerate}
1300: A justification for the simplifications of the functions $\mbarcub(\E)$ and
1301: $\hbar(\E)$ can be found in \fig{02}(a) while the rest of the
1302: replacements are illustrated in  \fig{08}.
1303: 
1304: \input{fig8.tex}
1305: 
1306: On these grounds, we postulate the following caricature of the AM
1307: \eqref{N62mod},
1308: %
1309: \begin{subequations}
1310: \begin{align}
1311: & \Df{\E}{\t} =
1312:   \eps^{-1} \GNa\,(\ENa-\E)\,\Heav(\E-\East)\,\h+\Gtwoh\Heav(\E-\Edagger)\,\n^4+\tilde{\G}(\E), \label{caric-lin-eq.E}\\
1313: & \Df{\h}{\t} = \eps^{-1}\Fone\,\left(\Heav(\Edagger-\E)-\h\right),				\label{caric-lin-eq.h}\\
1314: & \Df{\n}{\t} = \epstwo\,\Ftwo\,\left(\Heav(\E-\Edagger)-\n\right),				\label{caric-lin-eq.n}
1315: \end{align}											\label{caric-lin-eq}
1316: \end{subequations}
1317: %
1318: where we have included the artificial small parameters $\eps$ and
1319: $\epstwo$ to ensure the correct asymptotic structure of the system.
1320: 
1321: 
1322: \subsection{Exact solution} 
1323: \label{S:Caricature:Exact}
1324: 
1325: It is possible to obtain an exact analytical solution of the caricature model
1326: \eqref{caric-lin-eq}. Indeed, equations \eqref{caric-lin-eq.h} and 
1327: \eqref{caric-lin-eq.n} are separable and simple enough to be easily solvable. 
1328: After their solutions are substituted in equation \eqref{caric-lin-eq.E} 
1329: it, too, becomes a readily-solvable first-order linear ODE. 
1330: Therefore, assuming the initial conditions,
1331: %
1332: \begin{equation}
1333:  \E(0)=\Eini>\East,\quad \h(0)=1,\quad \n(0)=0,
1334: \label{IC}
1335: \end{equation}
1336: %
1337: of a fast-upstroke AP and natural continuity conditions 
1338: at the ends of the three intervals separated by  $\Edagger$ and $\East$ 
1339: or, equivalently, at the ends of the corresponding time intervals $\t\in[0,\tast]$,
1340: $\t\in[\tast,\tdagger]$ and $\t\in[\tdagger,\infty)$,
1341: system \eqref{caric-lin-eq} has the following exact analytical solution,
1342: \begin{subequations}
1343: \begin{align}
1344: &
1345: \n(\t) = \begin{cases}
1346:     1-\exp(-\epstwo \Ftwo\t),  &       \t\in[0,\tdagger]\\[1ex]
1347:     \left(\exp(\epstwo \Ftwo \tdagger)-1 \right) \exp(-\epstwo \Ftwo\t), &      \t\in[\tdagger,\infty]
1348: \end{cases} 											\label{exactsol.n}\\[0.5ex]
1349: &
1350: \h(\t) = \begin{cases}
1351:     \exp(-\Fone\t/\eps), & \t\in[0,\tdagger]\\[1ex]
1352:     1 - \left(1+\exp(\Fone\tdagger/\eps) \right) \exp(-\Fone\t/\eps), &
1353:     \t\in[\tdagger,\infty]
1354: \end{cases} 											\label{exactsol.h}\\[0.5ex]
1355: &
1356: \E(\t) = \begin{cases}
1357: \displaystyle 
1358: 	\Eeins(\t)=\exp\Bigg(\frac{\GNa}{\Fone}\,\exp\bigg(-\frac{\Fone\t}{\eps}\bigg)-\kthree\t\Bigg)\,\times	& \\
1359: \displaystyle \hspace*{2em} 
1360: 	 \Bigg[
1361:            \Eini \exp\bigg(-\frac{\GNa}{\Fone}\bigg)
1362:           -\kthree\Ethree\, \u(-\kthree,\t) 			& \\
1363: \displaystyle \hspace*{2em}
1364:            -\Gtwoh\sum\limits_{\l=0}^4
1365:            (-1)^{\l}\,\dbinom{4}{\l}\,\u\left((4-\l)\,\epstwo \Ftwo-\kthree,\t\right) & \\
1366: \displaystyle \hspace*{2em}
1367:           -\frac{\GNa\ENa}{\eps}\,\u\bigg(\frac{\Fone}{\eps}-\kthree,\t\bigg)
1368:           \Bigg], 								& \t\in[0,\t_*]\\[0.5ex]
1369:   \Ezwei(\t) = \left(\East-\w(\tast)\right)\,\exp\left(\ktwo\,(\t-\tast)\right)+
1370:    \w(\t), 									& \t\in[\t_*,\tdagger]\\[1ex] 
1371:   \Edrei(\t)=    (\Edagger-\Eone)\exp\left(-\kone(\t-\tdagger)\right)+\Eone,	& \t\in[\tdagger,\infty]
1372: \end{cases} 											\label{exactsol.E}
1373: \end{align}											\label{exactsol}
1374: \end{subequations}
1375: %
1376: where
1377: %
1378: \begin{subequations}
1379: \begin{align}
1380: &
1381:   \u(\qq,\t)\equiv\frac{\eps}{\Fone}\bigg(\frac{\GNa}{\Fone}\bigg)^{-\frac{\qq\eps}{\Fone}} 
1382:             \Bigg[
1383:           \Gamma\Bigg(\frac{\qq\eps}{\Fone},\frac{\GNa}{\Fone}\Bigg)-
1384:           \Gamma\Bigg(\frac{\qq\eps}{\Fone},\frac{\GNa}{\Fone}\,\exp\bigg({-\frac{\Fone\t}{\eps}}\bigg)\Bigg)
1385:             \Bigg],										\label{exactsolfun:u}\\
1386: &
1387:  \w(\t)\equiv \Etwo -\Gtwoh \sum\limits_{\l=0}^{4}(-1)^{\l} \dbinom{4}{\l}
1388:  \frac{\exp\left(-\l\,\epstwo \Ftwo \t\right)}{\ktwo+\l\,\epstwo \Ftwo},			\label{exactsolfun:w}
1389: \end{align}											\label{exactsolfun}
1390: \end{subequations}
1391: and 
1392: $\Gamma(a,x)$ is the upper incomplete gamma function,
1393: $\Gamma(a,x)\equiv\int_x^{\infty} z^{a-1}e^{-z}\,\d{z}$ for $\Re{a}>0$ and
1394: $\Gamma(a+1,x)=a\Gamma(a,x)+x^a\,e^{-x}$
1395: as defined in \cite{Abramowitz-Stegun}.
1396: The exact analytical solution is plotted in \fig{09} where it is
1397: compared with the numerical solutions of the AM
1398: \eqref{N62mod} and the internal boundary points $\Edagger$ and $\East$
1399: are also indicated.  
1400: The parameters $\tast$ and $\tdagger$ can be found
1401: as solutions  of
1402: \begin{gather}
1403: \label{ee:0005}
1404: \Eeins(\tast)=\East, \qquad
1405: \Ezwei(\tdagger)=\Edagger.
1406: \end{gather}
1407: Both equations are transcendental and not solvable analytically. 
1408: It is possible to solve them by using perturbation expansions 
1409: about the small parameters $\eps$ and $\epstwo$. This, however, 
1410: will contribute little beyond its value as a technical exercise 
1411: and will lead us away from the main point of this part, which is to
1412: validate the asymptotic solutions for $\E(\t)$, $\h(\t)$ and
1413: $\n(\t)$. So here we omit these formulae, assuming that $\tast$ and
1414: $\tdagger$ are known where they are needed. Numerically, for
1415: the standard values of parameters and $\Eini=-10$, we obtain
1416: $\tast\approx292.815$  and $\tdagger\approx345.241$.  
1417: \input{fig9.tex}
1418: 
1419: \section{Validation of the general asymptotic analysis}
1420: \label{S:Validation} % BS04: no
1421: 
1422: In the following we consider the stages of a normal fast-upstroke
1423: AP. For each stage we {\bf(a)}
1424: evaluate the appropriate 
1425: quadratures of section \ref{S:AM:Asymptotics} to obtain explicit
1426: asymptotic solutions for this stage and {\bf(b)} evaluate the
1427: limit of the exact analytical solution \eqref{exactsol} as the small
1428: parameters tend to zero as appropriate for the same stage. The
1429: asymptotic theory will be validated if the results from these two
1430: steps are identical. 
1431: 
1432: \subsection[Fast upstroke]{Fast upstroke, stage A--B}
1433: Here and below, letters A--F refer to the labels in \fig{05}(a).
1434: The fast upstroke occurs during the interval
1435: $\t\in[0,\tB]\subset[0,\tast]$, where $\tB\to0$ but
1436: $\tB/\eps\to\infty$ as $\eps\to0$. During this stage $\E$
1437: and the $h$-gate change together on a time scale 
1438: $\sim\eps$ from the point $(\E,\h)=(\Eini,1)$ to $(\EB,0)$ while $\n$
1439: is a slow variable and remains approximately at its initial
1440: value $\n\approx\nini=0$. 
1441: 
1442: \noindent{\bf (a) Asymptotic solution}.
1443: Asymptotically, this stage is described by quadratures
1444: \eqtwo(N62mod-fastupstrokesol,N62mod-Q).
1445: Substituting there the functions defined in section \ref{S:Caricature:Formulation}, we obtain
1446: %
1447: \begin{subequations}
1448: \begin{align}
1449: & \h(\E)= 1+\frac{\Fone}{\GNa}\,\ln\left(\frac{\ENa-\E}{\ENa-\Eini}\right),		\label{ee:0010.1}\\
1450: & \T(\E)= -\frac{1}{\Fone}\,
1451:   \ln\Bigg(1+\frac{\Fone}{\GNa}\,\ln\left(\frac{\ENa-\E}{\ENa-\Eini}\right)\Bigg). 	\label{ee:0010.2}
1452: \end{align}										\label{ee:0010}
1453: \end{subequations}
1454: %
1455: Solving \eqref{ee:0010.2} for the voltage $\E$ and substituting
1456: the result  in \eqref{ee:0010.1}, we arrive at an explicit
1457: asymptotic solution,
1458: %
1459: \begin{equation}
1460: \begin{split}
1461: & \E=\ENa-(\ENa-\Eini)\, \exp\left(\frac{\GNa}{\Fone}\left(e^{-\Fone\T}-1\right)\right),\\
1462: & \h=e^{-\Fone \T}.
1463: \end{split}						\label{ee:0020}
1464: \end{equation}
1465: %
1466: The maximal overshoot voltage $\EB$ is obtained as the fast time $\T$
1467: tends to infinity,
1468: \begin{gather}
1469: \EB=\Einf=\lim\limits_{\T\rightarrow+\infty} \E(\T)
1470: 	= \ENa-(\ENa-\Eini)\, \exp\left(-\frac{\GNa}{\Fone}\right). \label{EB}
1471: \end{gather}
1472: Since $\exp(-\GNa/\Fone) \approx 10^{-29}$, the  maximal overshoot
1473: voltage $\EB$ is extremely close to $\ENa$ and $\h(\tB)=\h(\T\to\infty)=0$ as expected.
1474: 
1475: \noindent{\bf (b) Limit of the exact solution.}
1476: We change to fast time, $\T=\t/\eps$, and
1477: take the limit of \eqref{exactsol.h} and \eqref{exactsol.E} as 
1478: $\eps$ tends to  zero, for fixed $\T$.
1479: Since $\lim\limits_{a\to1}\Gamma(a,x)=\exp(-x)$, it follows that 
1480: $\lim\limits_{\eps\to0}\u(\qq,\eps\T)=0$, for $\qq \sim 1$ and therefore 
1481: %
1482: \begin{align*}
1483: & \lim_{\eps\to0} \u(-\kthree,\eps\T)=0, \\
1484: & \lim_{\eps\to0} \u\left((4-\l)\,\epstwo \Ftwo-\kthree,\eps\T\right)=0.
1485: \end{align*}
1486: %
1487: Further,
1488: \[
1489: \u\left(\frac{\Fone}{\eps}-\kthree,\eps\T\right) \approx
1490:    \frac{\eps}{\GNa}\left[\exp\left(-\frac{\GNa}{\Fone}\right)-  
1491:   \exp\left(-\frac{\GNa}{\Fone}e^{-\Fone\T}\right)\right]. 
1492: \]
1493: Therefore, taking the limit
1494: $\eps\to0$ of expression $\Eeins$ in \eqref{exactsol.E} we obtain,\footnote{
1495:   Here we stress
1496:   that $\E$ and $\h$ are considered as  functions of $\t$ rather than $\T$.
1497: }
1498: %
1499: \begin{equation}
1500: \begin{split}
1501: & \lim\limits_{\eps\to0}\E(\eps\T)=\ENa-(\ENa-\Eini)\, \exp\left(\frac{\GNa}{\Fone}\left(e^{-\Fone \T}-1\right)\right),\\
1502: & \lim\limits_{\eps\to0}\h(\eps\T)=e^{-\Fone \T}.
1503: \end{split}									\label{ee:0030}
1504: \end{equation}
1505: %
1506: Since equations \eqref{ee:0020} and \eqref{ee:0030} are identical, the
1507: asymptotic theory of the fast upstroke of the AP is validated.
1508: 
1509: \subsection[Post-overshoot drop]{Post-overshoot drop of the voltage, stage B--C}
1510: This corresponds to the time interval $\t\in[\tB,\tC]\subset[0,\tast]$,
1511: where $\tC\to+\infty$ but $\epstwo\tC\to0$ as $\epstwo\to0$.
1512: We keep $\tB$ in some of the formulae rather than replacing it with its limit,
1513: $\lim_{\eps\to0}\tB=0$,
1514: as a symbolic reminder of the beginning of this asymptotic stage.
1515: The voltage $\E$ decreases on a time scale $\sim 1$ towards
1516: the stable systolic branch, the $\h$-gate remains at $\h\approx 0$ and
1517: the slow $\n$-gate also remains approximately unchanged at $\n\approx \nini=0$.
1518: 
1519: \noindent{\bf (a) Asymptotic solution}.
1520: The asymptotics of this stage are given by quadrature
1521: \eqref{N62mod-slowfastsol} which, with account of the approximations of 
1522: section \ref{S:Caricature:Formulation} and of the fact that asymptotically $\n=0$,
1523: evaluates to
1524: \begin{equation}
1525: \label{ee:0040}
1526: \t-\tB = \frac{1}{\kthree}\,\ln\left(\frac{\Ethree-\EB}{\Ethree-\E}\right).
1527: \end{equation}
1528: Solving for $\E(\t)$, the explicit asymptotic solution for this stage is, 
1529: \begin{equation}
1530: \label{ee:0050}
1531: \E(\t) = \Ethree+(\EB-\Ethree) e^{-\kthree(\t-\tB)}.
1532: \end{equation}
1533: 
1534: \noindent{\bf (b) Limit of the exact solution.}
1535: We are still within the interval $\t\in[\tB,\tC]\subset[0,\tast]$ and so use
1536: $\Eeins(\t)$ of \eqref{exactsol.E}. Taking the limit of \eqref{exactsol.n} and
1537: \eqref{exactsol.h} as $\eps$ and $\epstwo$  tend to  zero, we obtain
1538: %
1539: \begin{subequations}
1540: \begin{align}
1541: & \lim\limits_{\epstwo\to0}\n(\t)=0,					\label{ee:0061}\\
1542: & \lim\limits_{\eps\to0}\h(\t)=0,					\label{ee:0062}
1543: \end{align}								\label{ee:0060}
1544: \end{subequations}
1545: %
1546: in agreement with the asymptotic analysis.
1547: 
1548: Now we consider the limit of $\Eeins(\t)$ given by \eqref{exactsol.E}
1549: for $\eps\to0$ and $\epstwo\to0$ at a fixed $\t\in(0,\tast)$. 
1550: In the limit $\epstwo\to0$ the terms $\u\left((4-\l)\,\epstwo
1551: \Ftwo-\kthree,\t\right)$ become independent of the index $\l$ and the
1552: sum over $\l$ in the expression for $\Eeins(\t)$ vanishes since the
1553: binomial coefficients cancel each other. 
1554: For the remaining upper incomplete gamma functions, we use the
1555: recurrence relation, 
1556: %
1557: \[
1558: \Gamma(a+1,x)=a\,\Gamma(a,x)+x^a\,e^{-x} \nonumber
1559: \]
1560: %
1561: and the  following asymptotics in the limit $a\searrow0$, 
1562: %
1563: \begin{equation}
1564: \begin{split}
1565: & \Gamma(-a,x) = \Ei(1,x) + O(a), \\
1566: & \Gamma\left(-a,A\exp(-B/a)\right) = -a^{-1}\left(1-e^{B}\right)+O(1),
1567: \end{split}								\label{aux.BC}
1568: \end{equation}
1569: %
1570: for fixed $x, A, B>0$ and where $\Ei(\nu, x) =\int_1^\infty
1571: z^{-\nu}\exp(-xz)\, \d{z}$, $\Re{x} > 0$  is the exponential  integral.
1572: Hence, as $\eps\to0$, we have
1573: %
1574: \begin{align}  
1575: & \u(-\kthree,\t) \approx \kthree^{-1}\left(1-e^{\kthree\t}\right), \\
1576: & \u\left(\frac{\Fone}{\eps}-\kthree,\t\right) \approx
1577:   \frac{\eps}{\GNa}\Bigg(\exp\bigg(-\frac{\GNa}{\Fone}\bigg)-1\Bigg).
1578: \end{align}
1579: %
1580:   Substituting these results into the expression for
1581: $\Eeins$ in \eqref{exactsol.E} and taking the limits $\eps\to0$ and $\epstwo\to0$, we get 
1582: %
1583: \begin{equation}  
1584: \lim_{\epstwo,\eps\to0} \Eeins(t) =
1585: \Ethree+\Bigg(\ENa-(\ENa-\Eini)\exp\bigg(-\frac{\GNa}{\Fone}\bigg)-\Ethree\Bigg) e^{-\kthree\t},
1586: 									\label{stageB-C}
1587: \end{equation}
1588: %
1589: which with account of \eqref{EB}, coincides with the asymptotic
1590: expression  \eqref{ee:0050} where $\tB$ is taken at its limit,
1591: $\tB=0$. Thus the asymptotic
1592: procedure is validated in this stage as well.
1593: 
1594: \subsection[Plateau]{Plateau, stage C--D}
1595: This corresponds to the time interval $\t\in[\tC,\tD]\subset[0,\tast]$.
1596: Again, we keep $\tC$ in some formulae as a symbolic reminder of the
1597: beginning of this asymptotic stage, $1\ll\tC\ll\epstwo^{-1}$, 
1598: and we define $\tD=\tast$ precisely.
1599: The voltage $\E$ and the $\n$-gate change on a time scale of
1600: the order $\tD-\tC=O(\epstwo^{-1})$ along the upper  systolic  branch of the
1601: super-slow manifold  until $\E(\tast)=\East$ and $\n(\tast)\approx\Nast$.
1602: The $\h$-gate remains close to its quasi-stationary value at $\h\approx  0$.  
1603: 
1604: \noindent{\bf (a) Asymptotic solution}.
1605: The general asymptotic results describing this stage are given by
1606: \eqref{N62mod-slowslow.sol1}. Since the functions of the caricature
1607: model in the time interval $\t\in[0,\tast]$ are
1608: $\Gtilde(\E)=\kthree(\Ethree-\E)$, $\nbar(\E)=1$, $\gtwo(\E)=\Gtwoh$, 
1609: $\ftwo(\E)=\Ftwo$, the quadrature \eqref{N62mod-slowslow.sol1}
1610: evaluates to 
1611: \begin{subequations}
1612: \begin{align}
1613: & \n = 1-\exp\left(-\Ftwo(\tslow-\tslowC)\right), 						\label{ee:0091}\\
1614: & \E = \Ethree+ \frac{\Gtwoh}{\kthree}\left(1-\exp\left(-\Ftwo(\tslow-\tslowC)\right)\right)^4,	\label{ee:0092}
1615: \end{align}											\label{ee:0090}
1616: \end{subequations}
1617: where $\tslowC=\epstwo\tC$ and $\lim\limits_{\epstwo\to0}\tslowC=0$. 
1618: 
1619: \noindent{\bf (b) Limit of the exact solution.}
1620: The plateau stage occurs during the time interval
1621: $\t\in[0,\tast]$. In order to compare the limit of the exact
1622: solution \eqref{exactsol} to equations \eqref{ee:0090} we need to
1623: change to the super-slow time, $\tslow=\epstwo \t$. Then the
1624: solution \eqref{exactsol.n} for $\n(\tslow)$ does not contain a small
1625: parameter and is readily comparable with the asymptotics,
1626: \begin{gather}
1627: \label{ee:0100}
1628: \lim\limits_{\epstwo\to0}\n(\tslow) = 1-\exp\left(-\Ftwo(\tslow-\tslowC)\right),
1629: \end{gather}
1630: where 
1631: we have taken into account that the initial value of $\n$ according to \eqref{ee:0061} is
1632: $\n(\tslowC)=0$.
1633: 
1634: For the voltage during this stage, we consider $\Eeins$ from 
1635: \eqref{exactsol.E} with a change to the super-slow time $\tslow=\epstwo\t$,   
1636: \footnote{
1637:   Here and in similar cases further on, we imply that $\Eeins$ and
1638:   $\EBC$ are defined as functions of $\t$, rather than $\tslow$ 
1639:   or any other stage-specific time variable.
1640: }
1641: \begin{equation}
1642: \begin{split}
1643: \displaystyle \Eeins(\epstwo^{-1}\tslow)= & \EBC(\epstwo^{-1}\tslow)  \\
1644: \displaystyle
1645:   &+ \exp\Bigg(\frac{\GNa}{\Fone}\,\exp\bigg(-\frac{\Fone\tslow}{\eps\,\epstwo}\bigg)
1646: 	-\frac{\kthree\tslow}{\epstwo}\Bigg)\,
1647:            \Gtwoh\sum\limits_{\l=0}^4 
1648:            (-1)^\l\,\dbinom{4}{\l}\,\u\left((4-\l)\,\epstwo
1649: 	   \Ftwo-\kthree,\frac{\tslow}{\epstwo}\right),
1650: \end{split}								\label{stage.CD}
1651: \end{equation}
1652: and look for the limit $\eps\to0$, $\epstwo\to0$ at a fixed $\tslow$.
1653: The terms denoted here by $\EBC(\epstwo^{-1}\tslow)$ are precisely those discussed
1654: in connection with the limit of the exact solution during stage B--C
1655: with the only difference that now we consider them in the super-slow
1656: time $\tslow$. For a fixed $\tslow$, and small $\epstwo$, expression
1657: \eqref{stageB-C}  evaluates to
1658: \begin{gather}  
1659: \label{stageCD.01}
1660: \lim_{\epstwo\to0} \EBC(\epstwo^{-1}\tslow) = \Ethree.
1661: \end{gather}
1662: Using expressions \eqref{aux.BC} derived in the previous stage the
1663: remaining $\u(\cdot,\cdot)$-function in \eqref{stage.CD} in the limit
1664: $\eps\to0$ becomes 
1665: \begin{gather}
1666: \label{stageCD.02}
1667: \lim_{\eps\to0}\u\left((4-\l)\,\epstwo \Ftwo-\kthree,\frac{\tslow}{\epstwo}\right)=
1668: -\frac{1-\exp\left( \kthree \tslow\right)\,\exp\left(-(4-\l)\Ftwo\tslow\right)}{(4-\l)\,\epstwo\Ftwo-\kthree}.
1669: \end{gather}
1670: The exponentially growing term is cancelled by
1671: $\exp(-\kthree\tslow/\epstwo)$  in \eqref{stage.CD} and therefore
1672: substituting the above expressions in \eqref{stage.CD} and taking the
1673: limit $\epstwo\to0$ ultimately gives
1674: \begin{gather}
1675: \label{stageCD.03}
1676: \displaystyle \lim_{\epstwo,\eps\to0} \Eeins(\tslow) =
1677: \Ethree+ \frac{\Gtwoh}{\kthree}\left(1-\exp(-\Ftwo\tslow)\right)^4,
1678: \end{gather}
1679: in full accord with the asymptotic result \eqref{ee:0092} for the plateau stage of the
1680: AP, if we replace $\tslowC$ with its limit $\lim\limits_{\epstwo\to0}\tslowC=0$.
1681: 
1682:   Before proceeding to the next stage, we comment on the position of the point D
1683:   on the $(\n,\E)$ plane, which is close to the end of the systolic 
1684:   branch of the super-slow manifold, the point $(\Nast,\East)$. 
1685:   As stated earlier, we define point D by the exact
1686:   condition $\E(\tast)=\East$, hence the condition 
1687:   $\n(\tast)\approx\Nast$ is only approximate, since the AP trajectory
1688:   follows the super-slow manifold only approximately, with precision $O(\epstwo)$.
1689:   We shall see shortly that for the next
1690:   stage it is important that $\n(\tast)\ne\Nast$. 
1691:   The sense of this inequality can be easily seen from
1692:   \eqref{N62mod-slowsubemb-fast}: we know that during the C--D stage
1693:   $\E(\t)$ decreases, $\G(\E)>0$ and 
1694:   $\gtwo(\E)<0$, hence $\n>\left(-\G(\E)/\gtwo(\E) \right)^{1/4}=\Nss(\E)$ 
1695:   during the whole of that stage, which for $\t=\tast$, $\E=\East$ gives
1696:   $\n>\Nss(\East)=\Nast$.
1697:   More accurately the value of $\n(\tast)$ can be estimated using perturbation 
1698:   theory in $\epstwo$ around the super-slow manifold $\n=\Nss(\E)$;
1699:   this would be a further distraction from our main goal, and we omit
1700:   these formulae, like in the problem of determining $\tast$ and $\tdagger$. 
1701: 
1702: \subsection[Repolarization]{Repolarization, stage D--E}
1703: During this stage the voltage $\E$ jumps on a time scale  $\sim 1$
1704: from the systolic to the diastolic branch of the super-slow manifold.
1705: The $\h$-gate changes swiftly on a time scale $\sim \eps$
1706: from its lower quasi-stationary value close to zero to its upper
1707: quasi-stationary value close to unity due to the
1708: discontinuity in the right-hand side of equation \eqref{caric-lin-eq.h}.
1709: The slow $\n$-gate remains approximately unchanged at $\n\approx
1710: \Nast$.
1711: The associated time interval is
1712: $\t\in[\tD,\tE]=[\tast,\tdagger]\cup[\tdagger,\tE]$
1713: where $\tD=\tast$ is the beginning of this stage, 
1714: $\tdagger$ is the time of the inflexion point defined by $\E(\tdagger)=\Edagger$ and
1715: $\tE$ is the time of the end of the stage constrained by
1716: $1\ll\tE-\tdagger\ll\epstwo^{-1}$.
1717: 
1718: \noindent{\bf (a) Asymptotic solution}.
1719: The asymptotics at this stage are given by quadrature
1720: \eqref{N62mod-slowfastsol}. During this stage of the AP, the form of
1721: the caricature equations changes as the solution $\E(\t)$ moves 
1722: through the point $\Edagger$.  
1723: 
1724: In the time interval $\t\in[\tast,\tdagger]$ the relevant functions of the 
1725: caricature model are $\Gtilde(\E)=\ktwo(\E-\Etwo)$ and $\gtwo(\E)=
1726: \Gtwoh$ and therefore quadrature \eqref{N62mod-slowfastsol} evaluates to
1727: %
1728: \begin{subequations}
1729: \begin{align}
1730: & \E = \left(\Etwo-\frac{\Gtwoh \n(\tast)^4}{\ktwo}\right) +
1731: 	\Bigg(\East - \left(\Etwo-\frac{\Gtwoh \n(\tast)^4}{\ktwo}\right)
1732: 	\Bigg)\exp\left(\ktwo(\t-\tast)\right),				\label{ee:0132}\\
1733: & \n = \n(\tast) = 
1734: 	1-\exp{\left(-\Ftwo (\tslowast - \tslowC)\right)} = \const,	\label{ee:0131}
1735: \end{align}								\label{ee:0130}
1736: \end{subequations}
1737: %
1738: with initial conditions $\n(\tast)$ and $\E(\tast)\equiv\East$. 
1739: The function $\E(\t)$ given by expression \eqref{ee:0132}
1740: monotonically decreases since
1741: $\n(\tast)>\Nast=(\ktwo(\East-\Etwo)/\Gtwoh)^{1/4}$.
1742: In the second time interval, $\t\geq\tdagger$, the relevant functions of the
1743: caricature model are $\Gtilde(\E)=\kone(\Eone-\E)$ and $\gtwo(\E)=0$,
1744: and quadrature \eqref{N62mod-slowfastsol} gives 
1745: %
1746: \begin{equation}
1747: \begin{split}
1748: & \E(\t) = (\Edagger-\Eone)\exp\left(-\kone(\t-\tdagger)\right)+\Eone,\\
1749: & \n(\t) = \n(\tast) = \const, 
1750: \end{split}							\label{ee:0150}
1751: \end{equation}
1752: %
1753: with initial conditions $\n(\tast)$ and  $\E(\tdagger)=\Edagger$.
1754: 
1755: Finally, we note that the dynamics of $\h$ gate during this stage have a peculiarity
1756: due to the discontinuity of the right-hand side of \eqref{caric-lin-eq.h}.
1757: The finite constraint $\h=\Heav(\Edagger-\E)$ which according to 
1758: \eqref{N62mod-embslow2}
1759: is supposed to be approximately
1760: observed outside the AP upstroke, cannot be observed when $\E$ crosses
1761: the level $\Edagger$, as this would mean a discontinuity in $\h(\t)$. 
1762: In fact, the jump of $\h(\t)$ from 1 to 0 happens gradually, of course.
1763: We can see from \eqref{caric-lin-eq.h} that 
1764: this jump takes time $\sim\eps$.
1765: This violation of \eqref{N62mod-embslow2}
1766: does not, however, in any way affect the dynamics of $\E$ and $\n$,
1767: as $\Eh=\Edagger<\Em=\East$ and therefore the factor $\h\,\Heav(\E-\East)$
1768: in \eqref{caric-lin-eq.E} remains identically zero throughout this stage.
1769: 
1770: \noindent{\bf (b) Limit of the exact solution.} In order to compare
1771: the asymptotic and the exact solutions in a given AP stage we need to
1772: align them in time.  The beginning of the present stage is
1773: $\tD=\tast=\tslowast/\epstwo$, so we set 
1774: \begin{gather}
1775: \label{timealign1}
1776: \t=\epstwo^{-1}\tslowast+\tDE
1777: \end{gather}
1778: and then consider the limit $\eps,\epstwo\to0$ at a fixed $\tDE$.
1779: As discussed earlier, we assume here that the limit of $\tslowast$ is known (and finite).
1780: With this the  exact solution in the interval $\t\in[\tast,\tdagger]$ becomes
1781: %
1782: \begin{subequations}
1783: \begin{align}
1784: & \n(\epstwo^{-1}\tslowast+\tDE) =  1-\exp(-\Ftwo\tslowast-\epstwo \Ftwo \tDE), \label{DEexact.n}\\
1785: & \Ezwei(\epstwo^{-1}\tslowast+\tDE) =
1786: 	\left(\East-\w(\epstwo^{-1}\tslowast)\right)\,\exp\left(\ktwo\,\tDE\right)
1787: 	+ \w(\epstwo^{-1}\tslowast+\tDE).					\label{DEexact.E}
1788: \end{align}
1789: \end{subequations}
1790: %
1791: We have immediately that in the limit $\epstwo\to0$, expression \eqref{DEexact.n}
1792: coincides with \eqref{ee:0131}, since $\lim\tslowC=0$. Then, we have from
1793: \eqref{exactsolfun:w}
1794: \begin{gather}
1795: \lim\limits_{\epstwo\to0}\w\left(\frac{\tslowast}{\epstwo}+\tDE\right)
1796: =\Etwo-\frac{\Gtwoh}{\ktwo}\,\sum\limits_{\l=0}^{4}(-1)^{\l}\,\dbinom{4}{\l}\,e^{-\l\Ftwo\tslowast}
1797: =\Etwo-\frac{\Gtwoh}{\ktwo}\left(1-\exp(-\Ftwo\tslowast)\right).
1798: \end{gather}
1799: According to \eqref{exactsol.n}, $\n(\tast)=1-\exp(-\Ftwo\epstwo\tast)=1-\exp(-\Ftwo\tslowast)$.
1800: Hence we have that the limit of \eqref{DEexact.E} coincides with \eqref{ee:0132}
1801: since $\tDE=\t-\tast$.
1802: Analogously, substituting \eqref{timealign1} in the exact solution
1803: \eqref{exactsol} for the interval $\t>\tdagger$ and taking the limit
1804: $\epstwo\to0$  we reproduce the asymptotic solution \eqref{ee:0150}
1805: identically. 
1806: 
1807: Finally, the limit of the exact solution \eqref{exactsol.h} for the
1808: $h$-gate as $\eps\to0$ results in 
1809: $h=0$ for $\t\in[\tast,\tdagger]$ and $\h=1$ for
1810: $\t\in[\tdagger,\t_E]$, in accordance with the asymptotic theory. 
1811: 
1812: \subsection[Recovery]{Recovery, stage E--F}
1813: The voltage $\E$ and the $\n$-gate move on a time scale
1814: $\sim\epstwo^{-1}$ close to the diastolic branch of the
1815: super-slow manifold. This continues forever as the system approaches it
1816: stable equilibrium.
1817: The associated time interval is $\t\in[\tE,+\infty)$ with $\tE$ as
1818: defined above. Since $\tE\sim\epstwo^{-1}$, we define 
1819: $\tslowE=\epstwo\tE$ which has a finite limit, coinciding with
1820: that of $\epstwo\tdagger$ and $\epstwo\tast$.
1821: 
1822: 
1823: \noindent{\bf (a) Asymptotic solution}.
1824: The asymptotic solution in the the recovery stage is given by quadrature
1825: \eqref{N62mod-slowslow.sol1}. With the approximations of section
1826: \ref{S:Caricature:Formulation}, the functions of the caricature model in
1827: \eqref{N62mod-slowslow.sol1} are  $\nbar(\E)=0$, $\ftwo(\E)=\Ftwo$
1828: and $\Nss^{-1}(\n)=\Eone$ and so the asymptotic solution is
1829: %
1830: \begin{subequations}
1831: \begin{align}
1832: & \n = \n(\tE) \,\exp\left(-\Ftwo(\tslow-\tslowE)\right),		\label{ee:0191}\\
1833: & \E=\Eone,							\label{ee:0192}\\
1834: & \h=1.								\label{ee:0193}
1835: \end{align}							\label{ee:0190}
1836: \end{subequations}
1837: 
1838: \noindent{\bf (b) Limit of the exact solution.} At a fixed super-slow time
1839: $\tslow=\epstwo\t$, the limit $\epstwo,\eps\to0$ of the exact solution
1840: \eqref{exactsol} is
1841: %
1842: \begin{subequations}
1843: \begin{align}
1844: & \lim_{\epstwo\to0} \n(\tslow/\epstwo)  
1845:   = \left(\exp(\Ftwo \tslowdagger)-1 \right) \exp(-\Ftwo\tslow),\label{ee:0201}\\
1846: & \lim_{\epstwo\to0}\Edrei(\tslow/\epstwo)
1847:   =\lim_{\epstwo\to0}(\Edagger-\Eone)\exp\left(-\kone(\tslow-\tslowdagger)/\epstwo \right)+\Eone 
1848:   = \Eone,							\label{ee:0202}\\
1849: & \lim_{\epstwo,\eps\to0}\h(\tslow/\epstwo)
1850:   = \lim_{\epstwo,\eps\to0} \left( 1 - \left(1+\exp(\Fone\t_{2,\dagger}/(\eps \epstwo)) \right)
1851:   \exp(-\Fone\tslow/(\eps\epstwo))\Eone \right)
1852:   =1.								\label{ee:0203}
1853: \end{align}							\label{ee:0200}
1854: \end{subequations}
1855: %
1856: Finally, we notice that according to \eqref{ee:0201},
1857: $\n(\tdagger)=1-\exp(-\Ftwo\tslowdagger)$. Using this fact 
1858: equation \eqref{ee:0201} can be shown to be identical to equation \eqref{ee:0191},
1859: so there is full agreement between \eqref{ee:0190} and \eqref{ee:0200}.
1860: 
1861: \subsubsection*{}
1862: 
1863: We have demonstrated that at every stage of the AP the explicit
1864: asymptotic solution coincides with the appropriate limit of the exact
1865: analytical solution of the caricature model \eqref{caric-lin-eq}.
1866: We conclude that the asymptotic theory is validated. 
1867: 
1868: \section{Discussion}
1869: 
1870: \subsection{Summary of results}
1871: 
1872: We have started from the classical Noble model \cite{N62} of cardiac Purkinje fibres,
1873: arguing that it is the simplest ionic model based on cardiac
1874: electrophysiology. Using numerical observations, we have postulated a system
1875: of axioms which allowed us to propose a reasonable parametric
1876: embedding \eqref{embed0} of the Noble model. We have also noted that some of the features
1877: of the Noble model are rather peculiar in comparison with other
1878: cardiac models. This have lead us to propose the Archetypal Model
1879: \eqref{N62mod} with the ``generic'' structure of modern cardiac
1880: models, which allows a simple parametric embedding \eqref{N62mod-emb} 
1881: and which, in addition, is still a very accurate approximation of
1882: the Noble model and whose asymptotic limit coincides with that of the Noble model.
1883: Finally, we have obtained analytical
1884: solutions in quadratures, given by formulae
1885: \eqref{N62mod-fastsubthr}, \eqref{N62mod-fastupstrokesol},
1886: \eqref{N62mod-slowfastsol} and \eqref{N62mod-slowslow.sol1}, 
1887: corresponding to the asymptotic limits in the embedded small
1888: parameters in the Archetypal Model.
1889: 
1890: In that sense, we have achieved a fully analytical description of an
1891: action potential in a detailed ionic model of cardiac excitability.
1892: 
1893: The accurate reproduction of the properties of the authentic ionic
1894: model necessitated a number of mathematical features of the parametric
1895: embedding used which made the standard singular perturbation
1896: approaches based on Tikhonov theorem inapplicable:
1897: %
1898: \begin{itemize}
1899: \item a large factor in front of \textit{only} some terms in the
1900: right-hand side of the same equation;
1901: \item non-analytical, perhaps even discontinuous, asymptotic limit of
1902: some right-hand sides, even though the original system is analytical,
1903: \item  non-isolated equilibria in the fast subsystem;
1904: \item  dynamic variables which change their character from fast to slow 
1905: within one solution (remember in Tikhonov's theory, the roles
1906: of ``fast'' and ``slow'' variables are fixed);
1907: \item  the slow set may not even be a manifold, but may consist of
1908: pieces of different dimensionality (see Appendix \ref{S:HighExc}).
1909: \end{itemize}
1910: %
1911: All these features are related to each other and originate from the
1912: biophysics of excitable membranes, namely the fact that ionic gates
1913: work as nearly-perfect switches. Note that in some previous
1914: two-component simplified cardiac models,
1915: e.g. \cite{vanCapelle-Durrer-1980,Aliev-Panfilov,Rogers-2000}, there
1916: are segments where null-clines of both variables are very close to
1917: each other. Perhaps, in an appropriate asymptotic embedding, these
1918: segments would be near continua of non-isolated equilibria in the fast
1919: subsystem, which is one of the key features mentioned above and which
1920: might be related to the success of those models.
1921: 
1922: The asymptotic analysis we have done reproduces, in the limit
1923: $\eps\to0$, all five qualitative phenomenological features of cardiac
1924: excitability, listed in section~\ref{S:Intro:Motivation}, which are
1925: inconsistent with FitzHugh-Nagumo type systems.  Namely,
1926: %
1927: \begin{enumerate} 
1928: %
1929: \item % \label{slow-return} 
1930: \textbf{Slow repolarization.}  The only fast part of
1931: a typical AP is the upstroke A-B, which goes from the upper edge of
1932: the red cross-hatched rectangle on \fig{06}(a) towards dashed blue
1933: line along the vertical axis.  The other parts B-C-D-E-F all go along
1934: the dashed blue line there, which is a set of the equilibria of the
1935: fast subsystem.  In other words, all stages B-C-D-E-F are described in
1936: the slow subsystem \eqref{N62mod-slowsubemb}. As seen in \fig{06}(b),
1937: this includes relatively fast parts B-C and D-E as well as relatively
1938: slow ones C-D and E-F, but these have different speeds due to the
1939: secondary small parameter $\epstwo$. From the viewpoint of the main
1940: small parameter $\eps$ they are all slow, i.e. much slower than the
1941: upstroke.
1942: %
1943: \item % \label{slow-subthreshold} 
1944: \textbf{Slow subthreshold response.}
1945: This corresponds to a trajectory starting from an initial voltage $\Eini<\Em$. For
1946: $\Eini$ above (below) the line $\h=\H(\E)$, this corresponds to the leftward
1947: (rightward) going trajectory in the red cross-hatched region of
1948: \fig{06}(a). That is, the only fast process, if any, following a subthreshold
1949: initial perturbation, is the relaxation of the $\h$ gate. The fast
1950: dynamics of $\E$ are not engaged as $\INa$ channels remained closed.
1951: Thus, any dynamics of $\E$ after such initial perturbation occur only
1952: on the slow time scale.
1953: %
1954: \item % \label{accommodation} 
1955: \textbf{Fast accommodation.}  In
1956: voltage-clamped conditions, i.e. when $\E(\t)$ is prescribed, the fast
1957: subsystem \eqref{N62mod-embfast} gives that $\h(\t)$ tends towards
1958: $\H(\E)\,\Heav(\Eh-\E)$, i.e. towards the blue dashed line on
1959: \fig{06}(a), and of course for $\E$ staying above $\Em$ long enough,
1960: we have eventually $\h=0$.
1961: In other words, if $\E$ raises too slowly, the $\h$ gates have
1962: sufficient time to close, which prevents excitation. For this to
1963: happen, the (prescribed) dynamics of $\E$ should be slow compared to $\h$ dynamics,
1964: which are fast in terms of the $\eps$-embedding.
1965: %
1966: \item % \label{variable-peak} 
1967: \textbf{Variable peak voltage.}
1968: The peak voltage is the voltage at point B. From \fig{06}(a) it
1969: appears that this voltage is always very close to $\ENa$ as long as
1970: point A is above the line $\E=\Em$. However, according to the analysis
1971: of section~\ref{S:AM:Asymptotics:Fast}, the peak voltage $\Einf$
1972: is determined via equation $\hini=\J(\Einf)-\J(\Eini)$, that is
1973: it does depend on initial condition. This paradox is explained in
1974: the end of section~\ref{S:AM:Asymptotics:Fast} and in 
1975: Appendix~\ref{S:HighExc} as a consequence of presence, in the Noble model
1976: and its descendant Archetypal Model, of a yet another small parameter
1977: $\epshi$ which is a factor of $\fone/\gone$. Namely, $\ENa-\Einf$
1978: happens to be exponentially small in $\epshi$. A conclusion follows
1979: from there that if the parameters in the Archetypal Model are changed
1980: in such a way that $\fone/\gone$ is not so small --- e.g. by
1981: decreasing the value of $\GNa$, the peak voltage will demonstrate a
1982: noticeable dependence on $\hini$ and $\Eini$. 
1983: This in fact happens in other cardiac models that are not so stiff as Noble model,
1984: i.e. in Courtemanche \etal{}~\cite{CRN} model of human atrium:
1985: as we have shown in \cite{Biktasheva-etal-2006}, 
1986: the phase portrait of its fast subsystem
1987: is very similar to \fig{06}(a), but less stiff and the peak 
1988: voltage does indeed vary widely in the whole of $(\Em,\ENa)$ 
1989: range. 
1990: %
1991: \item % \label{front-dissipation} 
1992: \textbf{Front dissipation.}  This
1993: feature requires the analysis of the spatially distributed version of
1994: the models and is therefore beyond the scope of this paper.  However,
1995: the phenomenon of front dissipation has been a specific target of a
1996: number of our previous
1997: papers~\cite{Biktashev-2002,Biktashev-2003,Simitev-Biktashev-2006,Biktasheva-etal-2006}.
1998: In~\cite{Biktashev-2002,Biktashev-2003} we considered a piecewise
1999: linear ``caricature'' version of the fast subsystem which allowed an
2000: explanation of front dissipation via establishing existence of a lower
2001: limit for the propagation speed, as opposed to FitzHugh-Nagumo type
2002: system which do not have such
2003: limit. In~\cite{Biktasheva-etal-2006,Simitev-Biktashev-2006} we
2004: demonstrated that the spatially extended version of fast system
2005: in~\cite{CRN}, which as noted above is essentially identical to that
2006: of the AM, demonstrated the lower limit of the conduction velocity
2007: similar to that in the piecewiselinear caricature. Moreover, we have
2008: demonstrated that the understanding of conduction blocks based on this
2009: lower limit has a predictive ability for wavebreaks in complicated
2010: spatiotemporal regimes in the Courtemanche et al. model.
2011: %
2012: \end{enumerate}
2013: 
2014: As no rigorous theory of slow-fast
2015: systems of non-Tikhonov type exists at present, we have formulated a
2016: ``caricature''-style simplification of the AM, which is a less
2017: accurate approximation of the Noble model but has the same asymptotic
2018: structure as AM and admits exact solution. Using this exact solution we
2019: have been able to prove the asymptotic results in the particular case.
2020: 
2021: 
2022: \subsection{Further directions}
2023: 
2024: We believe that the described procedure is generic and can be
2025: applied to typical detailed cardiac models including the
2026: complicated contemporary models. For instance, part of the same procedure
2027: has already been successfully applied to the human atrial tissue
2028: model of Courtemanche \etal{}~\cite{CRN} for which an analytical
2029: condition for propagation block in a re-entrant wave has been derived
2030: and a satisfactory quantitative agreement with
2031: results of direct numerical simulations have been demonstrated
2032: \cite{Biktashev-2002,Biktashev-2003,Simitev-Biktashev-2006}. It is of
2033: even greater interest to investigate break-up and self-termination of
2034: AP fronts in models of ventricular myocytes such as 
2035: \cite{Beeler-Reuter-1977,Luo-Rudy-1991,tenTuscher-etal-2004,Iyer-etal-2004} 
2036: since cases of
2037: ventricular fibrillation have more serious health consequences than
2038: those of atrial fibrillation. Another direction in which the proposed
2039: asymptotic description might be useful is the derivation of action
2040: potential restitution curves and conduction velocity dispersion
2041: curves for realistic cardiac models. These two curves are the
2042: most popular and widely available experimental characteristics of cardiac tissue
2043: upon which various interpretations of cardiac dynamics are based
2044: \cite{Shiferaw-etal-2006}.
2045: 
2046: Finally, a simplified model, like the Archetypal Model or its
2047: caricature suggested here, can be a useful tool for large-scale
2048: numerics. Confidence in such a tool will increase if the simplified
2049: model has been derived by a controlled asymptotic procedure from a
2050: detailed model and preserves its predictive power.  Such a model can
2051: also be useful for theoretical studies.  For example, the difference
2052: between ``slow over-threshold response'' and ``normal fast upstroke'',
2053: discussed in Appendix~\ref{S:Synthesis} only appears in the
2054: asymptotic limit $\eps\to0$ and cannot be mathematically identified in
2055: the model at $\eps=1$, even if the exact analytical solution like
2056: \eqref{exactsol} is available.  This difference may be important
2057: physiologically. E.g.~it creates a possibility for ``slow'' and
2058: ``fast'' propagating waves in the system, a feature that has been
2059: observed in other models and in electro-physiological experiments as
2060: ``Na'' and ``Ca'' excitation
2061: waves~\cite{Romashko-Starmer-1995,Rohr-Kucera-1997}. We believe that
2062: asymptotic analysis is the most adequate tool for mathematical
2063: description of this sort of ``qualitative'' phenomena, and cannot be
2064: replaced with numerical or even exact analytical solutions.
2065: 
2066: %%%%%%%%%%%%%%%%%%%%%%%%
2067: \appendix
2068: 
2069: \section*{Appendices}
2070: 
2071: \section{A more accurate archetypal model}
2072: \label{S:Accurate}
2073: 
2074: As one can see from \fig{04}(a), the AP of the
2075: AM \eqref{N62mod} is a bit longer, and its repolarization is a bit
2076: slower than that of the Noble model \eqref{N62}. Although this
2077: difference may seem relatively minor, it is in fact surprisingly
2078: large, considering that the small parameter $\eps$ used to derive equations
2079: \eqref{N62mod} from equations \eqref{N62} is related to small quantities in
2080: \eqref{N62} of the order of $10^{-2}$. Thus, we would expect
2081: an accuracy of the order of $1\%$ in all results, which is clearly not
2082: the case in \fig{04}(a). The reason for this is that the asymptotic
2083: structure of Noble model \eqref{N62} is even more complicated than
2084: that summarized in the Axioms I--VII. Here we argue that the observed
2085: discrepancy is mainly due to a deficiency of Axiom VII. However, we believe that the
2086: complication in question is idiosyncratic for Noble model 
2087: and is not actually observed in later more realistic models.
2088: Still, in order to demonstrate the validity of our approach, we show
2089: here how the AM can be improved by an appropriate correction.
2090: 
2091: \input{fig10.tex}
2092: 
2093: \Fig{02}(b) shows that the function $\S(\E)$ is relatively
2094: small. However, it appears in \eqref{h-mfd-subs} multiplied by the
2095: large factor $\gone(\E)/\fone(\E)$ and thus the values of the product
2096: \begin{equation}
2097:   \Q(\E) \equiv \gone(\E)\, \S(\E)/ \fone(\E),
2098: \label{Q}
2099: \end{equation}
2100: are in fact of the order unity over a significant range of
2101: voltages as shown in \fig{10}(a). It is then clear that 
2102: neglecting this essentially non-zero term in equation
2103: \eqref{h-mfd-subs} leads to the low accuracy of the AM
2104: \eqref{N62mod}. 
2105: At first glance, this means that in the Noble model \eqref{N62},
2106: during the repolarization phase of the AP, gate $\h$ is, in fact, not
2107: fast compared to other two 
2108: variables. This would mean that reduction to a two-variable model in that
2109: region is not possible. However, as we have discussed in Introduction
2110: after the definition of embedding, a replacement of $1$ with a small
2111: parameter can, in fact, be a reasonable embedding, and its quality can be
2112: assessed by a comparison of the numerical solutions of the original and
2113: the embedded problem. Such comparison is provided in \fig{03} and
2114: shows that the embedding \eqref{N62mod} is indeed reasonable although
2115: of a not very high accuracy. 
2116: Thus we suppose that a higher accuracy can be achieved by 
2117: taking a first-order approximation of the term $\Q$ in the parameter
2118: $\eps$, rather than zero-order as in the other terms. Let us denote
2119: the small parameter associated with the function $\Q(\E)$ by
2120: $\epsQ$, $\epsQ>0$. Then the error term 
2121: after the first-order approximation in $\epsQ$ will be
2122: $O(\epsQ^2)$. The error term after the zero-order approximation in
2123: $\eps$ will be, of course $O(\eps)$. If we want these two kinds of
2124: error terms to be of the same order, we must therefore consider
2125: $\epsQ=\eps^{1/2}$.  
2126: These arguments are formalized by the following improved version of Axiom VII, 
2127: 
2128: \begin{axiomlist}{\rule{23mm}{0mm}}
2129:   \item[Axiom VIIa.] 
2130:   $\emb{\S}(\E;\eps) = \eps^{1/2} \Stilde(\E) + O(\eps)$, 
2131:  {\it where $\Stilde(\E)\approx\S(\E)$.}
2132: \end{axiomlist}
2133: Besides, for technical reasons the following, stronger version of Axiom VI will
2134: be more convenient for us:
2135: \begin{axiomlist}{\rule{23mm}{0mm}}
2136:   \item[Axiom VIa.] 
2137:   $\emb{\mbarcub}(\E;\eps)\,\emb{\hbar}(\E;\eps)=\eps\Wtilde(\E) +
2138:   O(\eps^2)$,   
2139:   {\it for the same $\Wtilde(\E)$ as in Axiom VI.} 
2140: \end{axiomlist}
2141: 
2142: Substituting Axioms VIa and VIIa into \eqref{h-mfd-subs}, we get
2143: %
2144: \begin{gather}
2145:   \Df{\E}{\t} = \gone\Wtilde(\E) 
2146:   - \eps^{1/2} \frac{\gone(\E)}{\fone(\E)}\, \Stilde(\E)\,\Df{\E}{\t}
2147:   + O(\eps) .				\label{embed0slow-implicit}
2148: \end{gather}
2149: %
2150: From here we deduce that $\d\E/\d\t=\gone\Wtilde+O(\eps^{1/2})$. Substituting
2151: this into the right-hand side of \eqref{embed0slow-implicit} we obtain 
2152: \begin{gather}
2153:   \Df{\E}{\t} = \gone\Wtilde(\E) 
2154:   - \eps^{1/2} \frac{\gone(\E)}{\fone(\E)}\, \Stilde(\E) \gone\Wtilde
2155:   + O(\eps).				\label{embed0slow-explicit}
2156: \end{gather}
2157: %
2158: Thus, after discarding $O(\eps)$ and putting $\eps=1$, we arrive at the
2159: following variant of \eqref{embed0slow},
2160: %
2161: \begin{subequations}
2162: \begin{align}
2163: & \Df{\E}{\t} = \left(\gone(\E)\, \Wtilde(\E) + \gtwo(\E)\, \n^4 +
2164:   \gthree(\E)\right)\,\left(1-\Qtilde(\E)\right),		\label{embed0slowQ.1}\\
2165: & \h = \H(\E)\, \Heav(\Eh-\E),					\label{embed0slowQ.2}\\
2166: & \Df{\n}{\t} = \ftwo(\E)\, \left(\nbar(\E) - \n \right),		\label{embed0slowQ.3}
2167: \end{align}							\label{embed0slowQ}
2168: \end{subequations}
2169: %
2170: where the first and third equations are satisfied with accuracy
2171: $O(\eps)$ and the second equation is satisfied only with accuracy
2172: $O(\eps^{1/2})$. Here $\Qtilde(\E)=\Stilde(\E)\, \gone(\E)/\fone(\E)$
2173: which according to Axiom VIIa should be close to $\Q(\E)$.  
2174: 
2175: 
2176: We see that in terms of the slow subsystem, the modification simply
2177: amounts to multiplying the right-hand side of equation \eqref{embed0slowQ.1} by a
2178: known function of $\E$. By analogy, it is straightforward to propose an
2179: improved version of the AM \eqref{N62mod},
2180: %
2181: \begin{equation}
2182: \begin{split}
2183: & \Df{\E}{\t} = \left(
2184:   \gone(\E)\, \M(\E)\, \Heav(\E-\Em)\, \h + 
2185:   \gone(\E)\,\W(\E) + 
2186:   \gtwo(\E)\, \n^4 + 
2187:   \gthree(\E)\right)\,
2188:   \left(1-\Q(\E)\right),\\
2189: & \Df{\h}{\t} = \fone(\E)\, \left( \H(\E)\, \Heav(\Eh-\E) - \h \right),\\
2190: & \Df{\n}{\t} = \ftwo(\E)\, \left( \nbar(\E) - \n \right).                 
2191: \end{split}								\label{N62modQ}
2192: \end{equation}
2193: %
2194: This improved AM gives solutions very close to those of
2195: the Noble system \eqref{N62}, as illustrated in \fig{10}(b).  
2196: 
2197: Notice that the asymptotic structure of the improved AM
2198: \eqref{N62modQ} is exactly the same as that of \eqref{N62mod}. Indeed,
2199: the difference amounts to redefining the functions $\gone(\E)$,
2200: $\gtwo(\E)$, $\gthree(\E)$ via a factor $1-\Q$, thus the asymptotic
2201: theory discussed in section~\ref{S:AM:Asymptotics} is equally
2202: applicable to both systems.
2203: 
2204: \section{The high-excitability embedding}
2205: \label{S:HighExc}
2206: 
2207: For the numerical values of the parameters corresponding to healthy
2208: tissue, the voltage upstroke at the beginning of the AP is, in
2209: fact, a much faster variable than the $\h$-gate. Indeed, the voltage speed
2210: constant $\gNa/\CM=33\frac13$ is large compared to the typical values of the
2211: $\h$-gate speed function,
2212: $\max\limits_{[\Em,\ENa]}\fone(\E)\approx1$.  
2213: This speed difference is unaccounted for by Axioms I--VII(a)
2214: where these two quantities have the same asymptotic order.
2215: However, some features of a typical AP solution depend on the ratio
2216: of these quantities in an exponential way. 
2217: 
2218: To take this feature into account, here we construct an improved embedding
2219: which takes the $\E$-vs.-$\h$ speed difference into 
2220: account.  We use an additional embedding with one more artificial
2221: small parameter $\epshi>0$. Formally, we replace Axiom I with a new
2222: Axiom,\\[2mm] 
2223: %
2224: \noindent\textsl{Axiom Ia.}\rule{10mm}{0mm} $
2225: \emb{\gNa}(\eps,\epshi)=\epshi^{-1}\eps^{-1}\gNa.$ \\[2mm]
2226: %
2227: This, of course, does not affect the slow-time subsystem
2228: \eqref{embed0slow}. The fast-time subsystem \eqref{embed0fast} now depends
2229: on the new parameter $\epshi$ and becomes
2230: %
2231: \begin{equation}
2232: \begin{split}
2233: & \Df{\E}{\T} = \epshi^{-1} \gone(\E)\, \M(\E)\, \Heav(\E-\Em)\, \h,\\
2234: & \Df{\h}{\T} = \fone(\E)\, \left(\H(\E)\,\Heav(\Eh-\E) - \h \right),
2235: \end{split}							\label{embed0fast-hiex}
2236: \end{equation} 
2237: %
2238: where the trivial equation for $\n$ is omitted as the $\n$-gate neither
2239: changes nor matters on time scales of interest here. The equations
2240: \eqref{embed0fast-hiex} appear to be a standard fast-slow Tikhonov
2241: system. 
2242: However, this is deceptive since the specific properties of the right-hand
2243: sides lead to a number of nonstandard features. Let us consider the limit
2244: $\epshi\rightarrow0$ in this system. In the super-fast time $\thi=\T/\epshi$,
2245: the $\h$-gate is a first integral, $\d{\h}/\d{\thi}=0$, and $\E$
2246: satisfies the super-fast equation, 
2247: %
2248: \begin{equation}
2249: \Df{\E}{\thi} = \gone(\E)\, \M(\E)\, \Heav(\E-\Em)\,\h,
2250: \label{embed0fast-fast}
2251: \end{equation}
2252: %
2253: depending on $\h$ as a parameter.  The equilibria of this system for
2254: which the the right-hand side vanishes, form an unusual set consisting of
2255: the lines $\h=0$  and $\E=\ENa$ and the semi-stripe
2256: $\{(\E,\h)\}=(-\infty,\Em)\times[0,1]$. Hence, the slow set is not a
2257: manifold, but consists of pieces of different dimensionalities.  This
2258: feature is even ``more non-Tikhonov'' than the many non-standard
2259: properties observed in the main embedding.
2260: One consequence is that the slow-time subsystem of \eqref{embed0fast-hiex}
2261: changes form as the trajectory passes through the various pieces of the slow
2262: set. This dependence is substantial: even the dimensionality of the
2263: slow system changes. For instance, for a typical AP
2264: solution in the beginning of the plateau, the trajectory crosses the
2265: $\E=\ENa$ piece, which can be parameterized by $\h$. Then we have a
2266: one-dimensional slow subsystem of \eqref{embed0fast-fast} (which is in
2267: the ``fast time'' $\T$) in the form of an equation for $\h$, 
2268: %
2269: \begin{equation}
2270: \Df{\h}{\T} = - \fone(\ENa)\, \h.
2271: \label{embed0fast-slow1} 
2272: \end{equation}
2273: %
2274: Then, during the later part of the plateau, which proceeds along the
2275: one-dimensional piece $\h=0$, the slow subsystem becomes 
2276: ``zero-dimensional'' in the sense that all right-hand sides vanish and
2277: all trajectories are fixed points. This corresponds to the fact that the
2278: movement along this piece occurs slowly in terms of the parameter
2279: $\eps$, \ie{}~is infinitely slow not only in terms of time $\thi$ but
2280: in terms of time $\T$ too. Finally, during the repolarization phase of the AP
2281: when the voltage $\E$ drops below $\Em$, the slow subsystem is on the
2282: two-dimensional piece but is in fact foliated to
2283: one-dimensional pieces, as the right-hand side of the equation for
2284: $\E$ vanishes and the voltage $\E$ is a first integral,
2285: %
2286: \begin{equation}
2287: \begin{split}
2288: & \Df{\E}{\T} = 0,\\
2289: & \Df{\h}{\T} = \fone(\E)\, \left( \H(\E)\,\Heav(\Eh-\E) - \h \right).
2290: \end{split}							\label{embed0fast-slow3}
2291: \end{equation} 
2292: 
2293: Since the equations of this ``high-excitability'' embedding are at most
2294: one-dimensional systems, obviously all of them can be solved in quadratures. 
2295: This embedding is rather instructive since it demonstrates that a
2296: non-Tikhonov embedding might lead to rather untypical consequences.
2297: It might also be useful for a number of applications, especially when 
2298: healthy, well-excitable tissues are concerned. However, 
2299: the most important applications are those related to the failure of 
2300: excitation, or of excitation propagation in the case of
2301: spatially-extended systems. These processes are observed, however,
2302: exactly when the excitability, represented here by formal parameter
2303: $\epshi^{-1}$, is not so high. 
2304: 
2305: \section{Asymptotic synthesis}
2306: \label{S:Synthesis}
2307: 
2308: We assume, for simplicity, that the initial values of the gating
2309: variables are given by $\hini=1$, $\nini=0$. Then, depending on the
2310: initial trans-membrane voltage $\Eini$, there exist three
2311: types of solutions of \eqref{N62mod} as visualized by the phase
2312: portraits in \fig{06} and described below: 
2313: 
2314: \subsection{Sub-threshold response}
2315: If the initial value of $\E$ is less that the threshold value
2316: of the super-slow subsystem \eqref{N62mod-slowslow} \ie{}~$\Eini<\Etwo$,
2317: the voltage decays towards its global equilibrium  $\Eone$:
2318: \begin{nlist}{\rule{3mm}{0mm}} 
2319: %
2320: \item[Relaxation of the $\h$-gate.] The $\h$-gate relaxes on a time scale
2321:   $\sim\eps$ towards its quasi-stationary value $\h\approx
2322:   \H(\E)\,\Heav(\Eh-\E)$ according to \eqref{N62mod-fastsubthr}. The
2323:   variables $\E$ and $\n$  remain close to their original values of
2324:   $\E\approx \Eini$,  $\n\approx \nini=0$. 
2325: %
2326: \item[Relaxation of the voltage $\E$.] The voltage decays on a time scale
2327:   $\sim1$ towards its equilibrium value $\Eone$  according to
2328:   \eqref{N62mod-slowfastsol} (with $\n=0$).  The $\h$-gate remains close
2329:   to its quasi-stationary value except,  possibly, swiftly moving on
2330:   a time scale $\sim\eps$ along its  discontinuity as $\E$
2331:   passes through $\Eh$. The   slow $\n$-gate remains approximately
2332:   at $\n\approx \nini=0$.
2333: %
2334: \end{nlist}
2335: 
2336: \subsection{Slow over-threshold response}
2337: 
2338: If the initial value of the voltage is bigger than the threshold value
2339: of the super-slow subsystem \eqref{N62mod-slowslow} but smaller
2340: than $\Em$ \ie{}~$\Etwo<\Eini<\Em$ then  the  slow subsystem alone
2341: is sufficient to describe the AP evolution and the fast Na current is
2342: not involved. The voltage makes a relatively small excursion towards
2343: the upper systolic branch of the super-slow manifold and approaches it from 
2344: below:
2345: \begin{nlist}{\rule{3mm}{0mm}}
2346: \item[Relaxation of the $\h$-gate.] The $\h$-gate relaxes on a time
2347:   scale $\sim\eps$ in the same way as in the case of
2348:   sub-threshold response, except this time the quasi-stationary value
2349:   of $\h$ is zero since $\Etwo>\Eh$.  
2350: \item[Rise of the voltage $\E$.] The voltage increases on 
2351:   time scale $\sim1$ towards $\Ethree$ \ie{}~towards the
2352:   upper part of the systolic branch of the super-slow  manifold according to
2353:   \eqref{N62mod-slowfastsol} (with $\n=0$). The slow $\n$-gate remains
2354:   approximately unchanged at   $\n\approx \nini=0$.   
2355: \item[Plateau.] The variables $\E$ and $\n$ move on a time scale
2356:   $\sim\epstwo^{-1}$ along the upper  systolic  branch of the
2357:   super-slow manifold $\n=\Nss(\E)$, according to \eqref{N62mod-slowslow}
2358:   until they reach the  point $(\East,\Nast)$. The $\h$-gate remains
2359:   close to its quasi-stationary value of $\h\approx  0$. 
2360: \item[Repolarization.] The voltage $\E$ jumps on a time scale
2361:   $\sim1$ according to \eqref{N62mod-slowfastsol}, towards
2362:   the diastolic branch of the super-slow manifold.  The $\h$-gate remains
2363:   close to its quasi-stationary value except, possibly, for a swift movement
2364:   along the discontinuity of $\H(\E)\Heav(\Eh-\E)$ at $\E=\Eh$ on a
2365:   time scale $\sim\eps$. The slow $\n$-gate remains
2366:   approximately at $\n\approx \Nast$.
2367: \item[Recovery.] The variables $\E$ and $\n$ move on a time scale
2368:   $\sim\epstwo^{-1}$ along the
2369:   diastolic branch of the super-slow manifold $\n=\Nss(\E)$, according to
2370:   \eqref{N62mod-slowslow}. This continues forever, with $(\E,\n)$
2371:   asymptotically approaching the true equilibrium of the system
2372:   $(\Eone,0)$. Gate $\h$ stays close to its quasi-stationary
2373:   value $\h\approx 0$.
2374: \end{nlist}
2375: 
2376: \subsection{Normal fast-upstroke action potential}
2377: 
2378:   If the initial value of the voltage exceeds the threshold of the
2379:   primary fast-time system \eqref{N62mod-embfast} \ie{}~$\Eini>\Em$ the
2380:   fast Na current is activated and a normal fast-upstroke AP,
2381:   is initiated:
2382: \begin{nlist}{\rule{3mm}{0mm}}
2383: \item[Fast upstroke.] The variables $\h$ and $\E$ change together
2384:     on a time scale $\sim\eps$ according to
2385:     \eqref{N62mod-fastupstrokesol} from the point $(\E,\h)=(\Eini,1)$
2386:     asymptotically (in $\T\rightarrow\infty$) approaching the point
2387:     $(\E,\h)=\left(\Einf(\Eini,1),0\right)$. The slow $\n$-gate remains
2388:     approximately unchanged at $\n\approx \nini=0$. 
2389: \item[Post-overshoot drop of the voltage $\E$.] The voltage $\E$
2390:   descends, inasmuch as $\Einf(\Eini,1)>\Ethree$,
2391:   on a time scale $\sim1$ towards its  higher
2392:   equilibrium value $\Ethree$ according to  \eqref{N62mod-slowfastsol}
2393:   (with $\n=0$).  Variable $\h$ remains close to its  quasi-stationary
2394:   value of $\h\approx 0$. The slow $\n$-gate remains approximately
2395:   unchanged at $\n\approx \nini=0$.   
2396: \item[Plateau, repolarization and recovery] stages follow which are 
2397:    similar to the corresponding stages in the case of slow
2398:   over-threshold response. 
2399: \end{nlist}
2400: \vspace{2mm}
2401: 
2402: 
2403: %%%%%%%%%%%%%%%%%%%%%%%%
2404: 
2405: \section*{Acknowledgement} 
2406:   We are saddened by the recent sudden death of Yury E. Elkin who has
2407:   made a decisive contribution to this work. This paper is dedicated
2408:   to his memory. This study has been supported in part by EPSRC grant
2409:   GR/S75314/01 (UK).
2410: 
2411: \bibliographystyle{agsm}
2412: \bibliography{mn}
2413: 
2414: \end{document}
2415: