math0701417/ms.tex
1: \documentclass[11pt, epsfig]{article}
2: \usepackage{epsfig, amsmath, amssymb, amsthm}
3: \parindent=1.5em
4: \parskip10pt
5: \textwidth=6.5in
6: \topmargin= 0.0in
7: \oddsidemargin=0in
8: \textheight=8.4in
9: 
10: \def\theequation{\arabic{section}.\arabic{equation}}\def\theequation{\arabic{section}.\arabic{equation}}
11: \renewcommand{\theequation}{\thesection.\arabic{equation}}
12: 
13: \def\p{\partial}
14: \def\Re{{\rm I}\!{\rm R}}
15: \def\Na{{\rm I}\!{\rm N}}
16: \def\lbl{\label}
17: \def\be{\begin{equation}}
18: \def\ee{\end{equation}}
19: \def\lbl{\label}
20: \def\qed{\sq}
21: \def\bu{\bf u}
22: \def\bv{\bf v}
23: \def\tl{\tilde}
24: \def\mG{\mathcal{G}}
25: \def\mI{\mathcal{I}}
26:                                                                         
27: \title{Multimodal oscillations in systems with strong contraction}
28: \author{ 
29: Georgi Medvedev and Yun Yoo \thanks{
30: Department of Mathematics, Drexel University, 3141 Chestnut Street,
31: Philadelphia, PA 19104, {\it medvedev@drexel.edu}
32:  }
33:  }
34: 
35: \begin{document}
36: \maketitle
37: 
38: \begin{abstract}
39: One- and two-parameter families of flows in $\Re^3$
40: near an Andronov-Hopf bifurcation (AHB) are investigated in this work. 
41: We identify conditions on the 
42: global vector field, which yield a rich family of multimodal orbits 
43: passing close to a weakly unstable
44: saddle-focus and  perform a detailed asymptotic analysis of the
45: trajectories in the vicinity of the saddle-focus. Our analysis covers both
46: cases of sub- and supercritical AHB. For the supercritical case, we find
47: that the periodic orbits born from the AHB are bimodal when viewed in the
48: frame of coordinates generated by the linearization about the bifurcating
49: equilibrium. If the AHB is subcritical, it
50: is accompanied by the  appearance of multimodal orbits, which consist
51: of long series of nearly harmonic oscillations separated by large
52: amplitude spikes. We analyze the dependence of the interspike intervals
53: (which can be extremely long) on the control parameters. In particular,
54: we show that the interspike intervals grow logarithmically as the boundary
55: between regions of sub- and supercritical AHB is approached in the parameter
56: space. We also identify a window of complex and possibly chaotic oscillations
57: near the boundary between the regions of sub- and supercritical AHB
58: and explain the mechanism generating these oscillations. This work is motivated 
59: by the numerical results for a finite-dimensional approximation of a free 
60: boundary problem modeling solid fuel combustion.
61: \end{abstract}
62: 
63: 
64: \section{Introduction}
65: 
66: Systems of differential equations in both finite- and infinite-dimensional settings
67: close to an AHB have been subject to intense research 
68: due to their dynamical complexity and importance in applications. The latter
69: range from models in fluid dynamics \cite{MarsdenMcCracken} to those in the life sciences, 
70: in particular, in computational neuroscience 
71: \cite{BER, DK05, DRSE, GW, KR83, MC, RE89, ROWK}.
72: When the proximity to the AHB 
73: coincides with certain global properties of the vector field,
74: it may result in a very complex dynamics \cite{BS, DENG90, DENG95, GW, Hirschberg_Knobloch}.
75: The formation of the Smale horseshoes in systems 
76: with a homoclinic orbit to a saddle-focus equilibrium  
77: provides one of the most representative examples of this type 
78: \cite{AAIS}.
79: Canard explosion in relaxation systems affords another example \cite{CDD, KS01b}. 
80: Recent studies of relaxation systems,
81: motivated mainly by applications in the life sciences, have revealed that
82: the proximity to an AHB has a significant impact on the system
83: dynamics. It manifests itself as a family of multimodal periodic solution
84: that are composed of large-amplitude relaxation oscillations 
85: (generated by the global structure of the vector field)
86: and small-amplitude nearly harmonic oscillations (generated by the vector field near the
87: equilibrium close to the AHB)
88: \cite{DRSE, KOP, MC, MSLG, PSS, VV} (see Figure \ref{f.1}). 
89: These families of solutions possess rich bifurcation structure.
90: A remarkable example of an infinite-dimensional system close to the AHB
91: has been recently studied by Frankel and Roytburd 
92: \cite{FR05a, FR05b, FR03, FR95, FR94}. They derived and systematically studied 
93: a model of solid fuel combustion 
94: in the form of a free boundary problem for a $1D$ heat equation with
95: nonlinear conditions imposed at the free boundary modeling the interface between 
96: solid fuel mixture and a solid product.
97: The investigations of this model revealed a wealth of spatial-temporal patterns
98: ranging from a uniform front propagation to periodic and aperiodic front oscillations. 
99: The transitions between different dynamical regimes involve a variety of nontrivial bifurcation
100: phenomena including period-doubling cascades, period-adding sequences, 
101: and windows of chaotic dynamics. To elucidate the mechanisms responsible for 
102: different dynamical regimes and transitions between them, Frankel and Roytburd
103: employed pseudo-spectral techniques to derive a finite-dimensional approximation
104: for the interface dynamics in the free boundary problem \cite{FR95}. 
105: As shown in \cite{FKRT, FR95},
106: a system of three ordinary differential equations captured the essential features of 
107: the bifurcation structure of the infinite-dimensional problem.
108: The numerical bifurcation analysis of the finite-dimensional approximation revealed a rich
109: family of multimodal periodic solutions similar to those reported in the context of relaxation
110: systems near the AHB \cite{FKRT}. The bifurcation diagrams presented in \cite{FKRT}
111: and in \cite{MC} share a striking similarity, despite the absence of any apparent common 
112: structures in the underlying models (except to the proximity to the AHB). 
113: In particular, in both models, topologically distinct multimodal periodic solutions
114: are located on isolas, closed curves in the parameter space.
115: The methods of analysis of the mixed-mode solutions in \cite{KOP, MC, MSLG, WESCH} used
116: in an essential way the relaxation structure present in these problems. These approaches
117: can not be applied directly to analyzing the model in \cite{FKRT}, because it is not a priori clear 
118: what creates the separation of the time scales in this model, in spite of the evident fast-slow
119: character of the numerical solutions. This is partly due to the spectral method, which
120: was used to derive the system of equations in \cite{FKRT}: while it has captured well
121: the finite-dimensional attractor of the interface dynamics, it has disguised 
122: the structure of the physical model. One of the goals of the present paper is to identify 
123: the structure responsible for the generation of the multimodal oscillations 
124: in a finite-dimensional model for the interface dynamics and 
125: to relate it to those studied in the context of relaxation oscillations.
126: 
127: The family of flows in \cite{FKRT} includes 
128: in a natural way two types of the AHBs. Depending on the parameter values,
129: the equilibrium of the system of ordinary differential equations in \cite{FKRT}  
130: undergoes either a sub- or a supercritical AHB. A similar
131: situation is encountered in certain neuronal models (see, e.g., \cite{DK05,OP}).
132: In either case, the global multimodal periodic solutions are created after
133: the AHB. However, 
134: in the case of a supercritical bifurcation, they are preceded by a series of 
135: period-doubling bifurcations of small amplitude limit cycles, arising from
136: the AHB. 
137: On the other hand, in the subcritical case, the AHB gives rise to 
138: multimodal solutions, whose lengths and time intervals between successive large
139: amplitude oscillations can be very long. In the present paper,
140: we perform a detailed asymptotic analysis of the trajectories in a class of systems motivated
141: by the problem in \cite{FKRT}. Our analysis includes both cases of the sub- and supercritical AHBs. 
142: We also investigate the dynamical regimes arising near the border between the regions of sub- and
143: supercritical AHB. This region in the parameter space contains a number of nontrivial oscillatory
144: patterns including multimodal trajectories with substantial time intervals between successive
145: spikes, irregular, and possibly chaotic oscillations, as well as a family of periodic orbits 
146: undergoing a cascade of period-doubling bifurcations. Our analysis shows that these dynamical
147: patterns and the order in which they appear under the variation of the control parameters are
148: independent on the details of the model, but are characteristic to the transition from sub-
149: to supercritical AHB.
150: 
151: The outline of the paper is as follows. After introducing the model and rewriting it in the
152: normal coordinates, we present a set of the numerical experiments to be explained in the remainder of the paper.
153: Then we state our results for each of the following cases: supercritical AHB,
154: subcritical AHB, and the transition layer between the regions of sub- and supercritical AHB. In Section 3, we analyze
155: the local behavior of trajectories near a weakly unstable saddle-focus. 
156: The local expansions used in this section are similar to those used in \cite{CMP} for analyzing 
157: the AHB using the method of averaging. However, rather than establishing existence of periodic solutions, the goal
158: of the present section is to approximate the trajectories near the saddle-focus. For this,
159: after rescaling the variables and recasting the system into cylindrical coordinates, we reduce
160: the system dynamics to a $2D$ slow manifold. By integrating the leading order approximation of the
161: reduced system, we obtain necessary information about the local behavior of trajectories. The results
162: of this section are summarized in Theorem 3.1. In Section 4, we  study oscillatory patterns generated by the
163: class of systems under investigation. It is divided into three subsections devoted to the oscillations
164: triggered by the supercritical AHB (\S 4.1), subcritical AHB (\S 4.2), and those found in the transition region
165: between sub- and supercritical AHB (\S 4.3). In the supercritical case, we show that the oscillations just
166: after the AHB are already bimodal. Generically, one needs to use two harmonics to describe the
167: limit cycle born at the supercritical AHB in $\Re^3$. We give a geometric interpretation of this
168: effect, which we call the frequency doubling, due to the fact that the second frequency is twice as large
169: as that predicted by the Hopf Bifurcation Theorem \cite{MarsdenMcCracken}. 
170: We also compute the curvature and the torsion of the periodic orbit as a curve
171: in $\Re^3$. The latter is useful for the geometric explanation of the frequency doubling. In Subsection 4.2,
172: we study the subcritical case. We show that under two general assumptions on the global vector field,
173: the presence of the return mechanism ({\bf G1}) and the strong contraction property ({\bf G2}), the
174: subcritical AHB results in sustained multimodal oscillations. Even though the oscillations may not be periodic
175: (our assumptions do not warrant periodicity), the time intervals between consecutive spikes of the resultant motion comply to
176: the uniform bounds given in Theorem 4.1. This subsection also contains the definitions of the multimodal oscillations
177: and the precise formulation of the assumptions on the global vector field. The proof of Theorem 4.1
178: is relegated to Section 5. Subsection 4.3 describes the
179: transition between sub- and supercritical AHB. This transition contains a distinct bifurcation scenario, when
180: a small positive real part of the complex conjugate pair of eigenvalues is positive and fixed and the first 
181: Lyapunov coefficient changes sign. The numerical simulations show that this results 
182: in the formation of the chaotic attractor followed
183: by the reverse period-doubling cascade. To explain this bifurcation sequence, we derive a $1D$ first return map.
184: The asymptotic analysis in this subsection is complemented by numerical extension of the map to the region
185: inaccessible by local asymptotic expansions. The first return map obtained by the combination of the analytic
186: and numerical techniques reveals the principal traits of the transition from sub- to supercritical AHB.
187: Finally, the discussion of the results of the present paper and their relation to the previous work is
188: given in Section 6.  
189: 
190: \section{The model and numerical results}
191: In the present section, we formulate the model and present a set of
192: numerical results, which motivated our study. The following system of three ODEs was derived in 
193: \cite{FKRT}, as a finite-dimensional
194: approximation for the interface dynamics in a free boundary problem modeling solid fuel 
195: combustion:
196: \begin{eqnarray} \lbl{6.1}
197: \dot v_1 &=& \frac{3(v_3+v_2-v_1)-\nu k(v_1)-v_1^2}{ \nu k^\prime(v_1)},\\ \lbl{6.2}
198: \dot v_2 &=& v_3-v_1,\\ \lbl{6.3}
199: \dot v_3 &=& 9(v_1-v_3)-6v_2+\nu(v_1+1)k(v_1)+2v_1^2.
200: \end{eqnarray}
201: Here, $v_1(t)$ approximates the velocity of the interface between the solid fuel and the
202: burnt material. Functions $v_2(t)$ and $v_3(t)$ are
203: the first two coefficients in the series expansion for the spatial temperature profile
204: with respect to the basis of Chebyshev-Laguerre polynomials. Equations (\ref{6.1})-(\ref{6.3})
205: are obtained by projecting the original infinite-dimensional problem onto a finite-dimensional
206: function space and using the method of collocations for determining the unknown coefficients.
207: An important ingredient of the model, nonlinear kinetic function $k(v)$ is given by
208: \be\lbl{6.4}
209: k(v)={(1-v)^{p}-(1-v)^{-1}\over p+1}.
210: \ee
211: \begin{figure}
212: \begin{center}
213: {\bf a}\epsfig{figure=f1a.eps, height=1.8in, width=2.5in, angle=0} 
214: {\bf b}\epsfig{figure=f1b.eps, height=1.8in, width=2.5in, angle=0}\\
215: {\bf c}\epsfig{figure=f1c.eps, height=1.8in, width=2.5in, angle=0}
216: {\bf d}\epsfig{figure=f1d.eps, height=1.8in, width=2.5in, angle=0}\\
217: {\bf e}\epsfig{figure=f1e.eps, height=1.8in, width=2.5in, angle=0}
218: {\bf f}\epsfig{figure=f1f.eps, height=1.8in, width=2.5in, angle=0}\\
219: {\bf g}\epsfig{figure=f1g.eps, height=1.8in, width=2.5in, angle=0}
220: {\bf h}\epsfig{figure=f1h.eps, height=1.8in, width=2.5in, angle=0}
221: \end{center}
222: \caption{Stable periodic solutions of system of differential equations (\ref{6.5}). Plots
223: in a-d show solutions near a supercritical AHB ($p=2.2$) and those in e-h
224: correspond to the subcritical case ($p=3$). The time series of $x_1$ are presented in the
225: left column and those of $x_2$ are given in the right column. The time series for 
226: $x_3$ and $x_2$ are qualitatively similar. 
227: }\label{f.1}
228: \end{figure} 
229: It reflects the dependence of the velocity of propagation on the temperature in the front. This 
230: relation is very complex and is not completely understood at present. To account for a range of 
231: possible kinetic mechanisms, the model includes two control parameters: $\nu$ and $p$.
232: Both the interface velocity and the temperature profile are calculated in the uniformly
233: moving frame of reference. Therefore, system of equations (\ref{6.1})-(\ref{6.3}) describes
234: the deviations of the interface dynamics from that of a front traveling with constant
235: speed. In particular, periodic and aperiodic oscillations generated by (\ref{6.1})-(\ref{6.3})
236: correspond to complex spatio-temporal patterns in the infinite-dimensional model. The numerical
237: results presented in \cite{FKRT} show a remarkable similarity between the complex patterns 
238: generated by the infinite- and finite-dimensional models 
239: and between the scenarios for transitions between different regimes in both models.
240: For more information about the derivation of the free-boundary problem for solid fuel combustion
241: and its finite-dimensional approximation, we refer the reader to \cite{FKRT} and bibliography
242: therein.  
243: 
244: 
245: A simple inspection of (\ref{6.1})-(\ref{6.3}) shows that it has an equilibrium at the origin $O=(0,0,0)^T$ for all
246: values of $\nu$ and $p>0$. We linearize (\ref{6.1})-(\ref{6.3}) about the equilibrium at the origin
247: $$
248: \dot v = A(\nu)v+ \dots,\quad\mbox{where}\quad
249: A(\nu) =\left( \begin{array}{ccc}
250: \ \frac{3-\nu}{\nu} &  \frac {-3}{\nu}  &  \frac {-3}{\nu} \\
251: -1 & 0 & 1 \\
252: 9-\nu & -6  &  -9 \end{array} \right),
253: \quad v=\left(v_1, v_2, v_3\right)^T.
254: $$
255: Near $\nu_{AH}={1\over 3}$, Jacobian matrix $A(\nu)$ has a negative eigenvalue $-1$ and a pair of complex
256: conjugate eigenvalues $\lambda=a(\nu) + ib(\nu)$ and $\bar\lambda$. The latter crosses the imaginary
257: axis transversally at $\nu=\nu_{AH}$:
258: $$
259: a\left(\nu_{AH}\right)=0\qquad\mbox{and}\qquad a^\prime\left(\nu_{AH}\right)\ne 0.
260: $$
261: Therefore, the equilibrium of (\ref{6.1})-(\ref{6.3}) undergoes an AHB. In the  neighborhood of $\nu_{AH}$,
262: $\alpha=a(\nu)$ defines a smooth invertible function. We shall use $\alpha$ as a new control parameter.
263: After a linear coordinate transformation,
264: system of equations (\ref{6.1})-(\ref{6.3}) has the following form:  
265: \be\lbl{6.5} 
266: \dot x = 
267: \left(\begin{array}{ccc}
268:                       -1 & 0 & 0\\ 
269:                        0 & \alpha & -b(\alpha)\\ 
270:                        0 & b(\alpha) &  \alpha
271:       \end{array}\right) x + h(x,\alpha),
272: \quad x=\left(x_1, x_2, x_3\right)^T, \quad h=\left(h_1, h_2, h_3\right)^T.
273: \ee
274: Here, by $b(\alpha):=b\left(a^{-1}(\alpha)\right)$ we denote the imaginary part of $\lambda$ as a function of
275: the new control parameter. Nonlinear function $h:\Re^3\times \Re\to \Re^3$  is a smooth function 
276: such that $h\left(0,\alpha\right)=0$ and ${\p h (0,\alpha)\over \p x}=0$
277: for values of $\alpha$ close to $0$. More precisely, $h$ stands for a family of functions parametrized by $p$
278: (see (\ref{6.4})). To keep the notation simple, we omit the dependence of $h$ on the second control parameter $p$.
279: \begin{figure}
280: \begin{center}
281: {\bf a}\epsfig{figure=f2a.eps, height=2.75in, width=3.0in, angle=0}
282: {\bf b}\epsfig{figure=f2b.eps, height=2.75in, width=3.0in, angle=0}
283: \end{center} 
284: \caption{{\bf a}. A periodic trajectory of the original system in the regime close to a subcritical AHB.
285: Cross-sections $\Sigma^+$ and $\Sigma^-$ are defined as follows:
286: $\Sigma^+=\left\{\left(x_1, \rho\cos\theta, \rho\sin\theta\right):\; \rho=0.1,\; x_1\in [-0.02, 0.02],\;
287: \theta\in [0,2\pi)\right\}$. $\Sigma^-$ is orthogonal to $x_1-$axis and is set at 
288: $x_1=-0.65$.
289: The numerical results shown in {\bf b} test the return mechanism and the strong contraction
290: property for (\ref{6.5}) (see {\bf (G1)} and {\bf (G2)} in the text). For this, we
291: covered $\Sigma^+$ with
292: a uniform mesh  $\Sigma_{N_i,N_j}^+=\left\{ \xi_{ij}\right\},$ $i=1,2,\dots,N_i,$ $j=1,2,\dots, N_j$. 
293: Using points $\xi_{ij}$ as initial conditions, we integrated
294: (\ref{6.5}) until the corresponding trajectories hit the second cross-section, $\Sigma^-$,
295: transverse to the stable manifold. The top plots in {\bf b} shows the image of $\Sigma_{N_i,N_j}$ 
296: in $\Sigma^-$ under the flow-defined map.
297: To test the strong contraction property, we repeated this experiment by taking 
298: $\Sigma^-$ progressively closer to the origin (the location of $\Sigma^-$ is 
299: indicated in each plot). The top five images in {\bf b} clearly indicate
300: that the projection of  the vector field onto $\Sigma^-$ is strongly contracting. The bottom plot
301: shows that the trajectories starting from $\Sigma^+$ reach a small neighborhood of the 
302: unstable equilibrium.
303: }\label{f.2}
304: \end{figure} 
305: It turns out that in the range of parameters of interest, the AHB of the equilibrium for $\alpha=0$ 
306: can be either subcritical, or 
307: supercritical, or degenerate depending on the value of $p$.  This gives rise to several 
308: qualitatively distinct oscillatory regimes 
309: generated by (\ref{6.5}) for small values of $\alpha> 0$.  The corresponding bifurcation scenarios 
310: for (\ref{6.1})-(\ref{6.3}) 
311: for fixed values of $p$ and varying $\nu$ were studied numerically in \cite{FKRT}. 
312: Below, we reproduce some of these numerics for the system in new coordinates and supplement 
313: them with a set of new numerical experiments relevant to the analysis 
314: of this paper. 
315: After that we state our results for each of the following cases: supercritical AHB,
316: subcritical AHB, and the transition layer between the regions of sub- and supercritical 
317: bifurcations. 
318: 
319: \begin{figure}
320: \begin{center}
321: {\bf a}\epsfig{figure=f3a.eps, height=2.25in, width=3.0in, angle=0}
322: {\bf b}\epsfig{figure=f3b.eps, height=2.25in, width=3.0in, angle=0}
323: \end{center}
324: \caption{
325: The duration of the ISIs is plotted for:
326: ({\bf a}) $\gamma=0.778$ and ({\bf b}) $\alpha=0.045$.
327: }\label{f.3}
328: \end{figure}
329: 
330: We start with discussing the supercitical case. It is well known that a supercritical AHB
331: produces a stable periodic orbit in a small neighborhood 
332: of the bifurcating equilibrium. The period of the nascent orbit is approximately 
333: equal to $2\pi\beta^{-1},$ where $\beta=b(0)$ is the imaginary part of the pair of complex conjugate  eigenvalues 
334: of the matrix of the linearized system at the bifurcation.  The numerical time series of $x_1$ and $x_{2,3}$ show 
335: that, while $x_{2,3}$ oscillate with the frequency prescribed by the Andronov-Hopf Bifurcation Theorem 
336: \cite{KUZ, MarsdenMcCracken}, the former oscillates 
337: with twice that frequency (see Figure \ref{f.1}a,b). 
338: The asymptotic analysis of Section 3 explains this counter-intuitive effect and shows that 
339: this is, in fact, a generic property of the periodic orbits born from a supercritical bifurcation in $R^n$, $n\ge3$.  
340: Note that the difference in frequencies of oscillations in $x_1$ and $x_{2,3}$ can not be understood from the 
341: topological normal form from the AHB \cite{GEO}, because the latter does not contain the information about 
342: the geometry of the periodic orbit. In Section 4.1,
343: we compute two geometric invariants of the periodic orbits as a curve in $\Re^3$: the curvature and the torsion.  
344: The latter shows that generically the orbit born from a supercritical AHB is not planar.  This together with certain 
345: symmetry properties of the orbit explains the frequency doubling of the oscillations in $x_1$.
346: As was noted in \cite{FKRT}, for increasing values of $\alpha > 0$, the small periodic orbit born from the supercritical 
347: AHB undergoes a cascade of period-doubling bifurcations, the first of which is shown in Figure \ref{f.1}c,d. Already 
348: at the first period-doubling bifurcation, the periodic orbit lies outside the 
349: region of validity of the local power series expansions, developed in the present paper.  
350: Therefore, our analysis does not explain the period-doubling
351: bifurcations for increasing values of $\alpha>0$.
352: In Section 4.3 we complement our analytical results with the numerical construction of the $1D$ first return map. 
353: The latter explains the mechanism for period-doubling cascade and the window of complex dynamics reported in 
354: \cite{FKRT}. 
355: \begin{figure}
356: \begin{center}
357: {\bf a}\epsfig{figure=f4a.eps, height=1.6in, width=3.0in, angle=0}\quad
358: {\bf b}\epsfig{figure=f4b.eps, height=1.6in, width=3.0in, angle=0}
359: {\bf c}\epsfig{figure=f4c.eps, height=1.6in, width=3.0in, angle=0}
360: {\bf d}\epsfig{figure=f4d.eps, height=1.6in, width=3.0in, angle=0}\quad
361: \end{center}
362: \caption{The time series of $x_2$ are plotted for decreasing values of $\gamma>0$.
363: In plots {\bf a} and {\bf b}, the number of small amplitude oscillations separating the spikes increases,
364: while the qualitative form of the solutions remains the same.
365: For smaller values of $\gamma$, the system exhibits intermittency:
366: series of regular small amplitude oscillations are followed by those of irregular
367: oscillations of the intermediate amplitude ({\bf c} and {\bf d}). 
368: }\label{f.4}
369: \end{figure}
370: \begin{figure}
371: \begin{center}
372: {\bf a}\epsfig{figure=f5a.eps, height=1.6in, width=3.0in, angle=0}\quad
373: {\bf b}\epsfig{figure=f5b.eps, height=1.6in, width=3.0in, angle=0}
374: {\bf c}\epsfig{figure=f5c.eps, height=1.6in, width=3.0in, angle=0}\quad
375: {\bf d}\epsfig{figure=f5d.eps, height=1.6in, width=3.0in, angle=0}
376: {\bf e}\epsfig{figure=f5e.eps, height=1.6in, width=3.0in, angle=0}\quad
377: {\bf f}\epsfig{figure=f5f.eps, height=1.6in, width=3.0in, angle=0}
378: {\bf g}\epsfig{figure=f5g.eps, height=1.6in, width=3.0in, angle=0}\quad
379: {\bf h}\epsfig{figure=f5h.eps, height=1.6in, width=3.0in, angle=0}
380: \end{center}
381: \caption{
382: The time series of $x_2$ (left column) and the corresponding trajectories of (\ref{6.1})-(\ref{6.3})
383: (right column) plotted for negative values of $\gamma$ close to $0$. The transition from the 
384: complex dynamics in {\bf a,b} to regular approximately harmonic oscillations in {\bf g,h}
385: contains a reverse cascade of period-doubling bifurcations ({\bf c}-{\bf f}).
386: }\label{f.5}
387: \end{figure}
388:  	
389: We next turn to the subcritical case.  
390: The dynamics resulting from the subcritical AHB depends on the properties of the vector field outside 
391: of a small neighborhood of the equilibrium. We conducted a series of numerical experiments to study
392: the global properties of the vector field (see caption of Figure \ref{f.2} for details). Based 
393: on these numerical observations, we identify two essential features of the vector field:
394: {\bf (G1)} the return mechanism and {\bf (G2)} the strong contraction property. Specifically,
395: \begin{description}
396: \item[(G1)]
397: For a suitably chosen cylindrical crossection, $\Sigma^+$, placed sufficiently close to the origin, and
398: another crossection, $\Sigma^-$, transverse to $W^s(O)$ (see Figure \ref{f.2}a), the flow-defined map
399: $Q:\Sigma^+\to\Sigma^-$ is well-defined for $\alpha\ge 0$ and depends smoothly on the parameters of the
400: system.
401: \item[(G2)] There is a region $\Pi$, adjacent to $\Sigma^-$ and containing a subset of $W^s(O)$
402: (see Figure \ref{f.2}a), in which the projection of the vector field in direction transverse to 
403: $W^s(O)$ is sufficiently stronger than that in the tangential direction.
404: \end{description} 
405: In Section 4.2, conditions {\bf (G1)} and {\bf (G2)} are made precise.
406: Properties {\bf (G1)} and {\bf (G2)} guarantee that for small values of $\alpha>0$, the trajectories leaving
407: a small neighborhood of the saddle-focus, after some relatively short time enter $\Pi$. In $\Pi$,
408: the trajectories approach $W^s(O)$ closely and then follow it to a sufficiently small neighborhood of the 
409: unstable equilibrium. For small $\alpha>0$, the trajectories starting from a sufficiently small neighborhood
410: of the saddle-focus, remain in some larger (but still small) neighborhood of the origin for a long time.
411: Eventually, they hit $\Sigma^+$ and the dynamics described above repeats. Therefore, under conditions
412: {\bf (G1)} and {\bf (G2)}, the
413: subcritical AHB bifurcation results in a 
414: sustained motion consisting of long intervals of time that the trajectory spends near a weakly unstable saddle-focus 
415: and relatively brief excursions outside of a small neighborhood of the origin (see Figure \ref{f.2}a).  
416: In terms of the time series, the dynamical variables undergo series of small amplitude oscillations 
417: alternating with large spikes (Figure \ref{f.1}e-h). For increasing values of $\alpha > 0$, the number of small 
418: amplitude oscillations decreases and so do the time intervals between consecutive spikes, so-called 
419: interspike intervals (ISIs).  
420: The multimodal solutions arising near the subcritical AHB are not necessarily periodic. Nonetheless, 
421: the ISIs may be characterized in terms of the control parameters present in the system, regardless of whether the
422: underlying trajectories 
423: are periodic or not.  Specifically, in Section 4.2, we derive an asymptotic relation for the 
424: duration of the ISIs: 
425: \be\lbl{6.6}
426: \tau \sim {1\over 2\alpha} \ln\left(1+ {\alpha\over\gamma(p)}O\left(\epsilon^{-\varsigma}\right)\right),\;\;
427: \varsigma\ge 4,
428: \ee
429: where the first Lyapunov coefficient $\gamma(p)$ reflects the distance of the system from the transition 
430: from sub- to supercritical AHB, with $\gamma(p)$ 
431: being positive when the AHB is subcritical and equal to zero at the transition point. 
432: Finally, $\epsilon>0$ is a small parameter and positive $\alpha=O(\epsilon^2)$.
433: The role of $\epsilon$ will become clear later.  We illustrate (\ref{6.6}) with 
434: numerically computed plots of the ISIs under the variation of control parameters in Figure \ref{f.3}.  
435: The ISIs provide a 
436: convenient and important characteristics of the oscillatory patterns involving pronounced spikes.  For example, it 
437: is widely used in both theoretical and experimental neuroscience for description of patterns of electric activity 
438: in neural cells.  An important aspect of the ISIs, is that the dependence of the ISIs on the control parameters 
439: in the system (which often can be directly established experimentally) reveals the bifurcation structure of the system.  
440: Therefore, the analytical characterization of the ISIs in terms of the bifurcation parameters, 
441: such as given in (\ref{6.6}), is important in applications. For the problem at hand, 
442: the ISIs depend on the interplay of the two parameters 
443: $\alpha$ and $\gamma(p)$, which reflect the proximity of the system to a codimension 2 bifurcation.  
444: The dependence of the ISIs on $\alpha$ was studied in \cite{GW} 
445: for a model problem near a Hopf-homoclinic bifurcation.  Our analysis extends the formula 
446: for the ISIs obtained in \cite{GW} to a wide class of systems and emphasizes the role 
447: of $\gamma(p)$, which is important for the present problem.  
448: We also note that the structure of the global vector field suggests that (\ref{6.5}) is 
449:  also close to  a homoclinic bifurcation, which can also influence the ISIs.  However, under 
450: the variation of the control parameters $\alpha$ and $p$ the system remains 
451: bounded away from the homoclinic bifurcation, so that the latter effectively does not affect the ISIs. 
452: In fact, (\ref{6.6}) can be easily extended to include the distance from the homoclinic bifurcation as well.  
453: As follows from (\ref{6.6}), the ISIs increase for decreasing values of $\gamma>0$ (see Figure \ref{f.3}b).  
454: This observation prompted our interest in investigating the transition in the system 
455: dynamics as $\gamma$ crosses $0$. Numerical simulations show that as $\gamma$ approaches $0$ from the positive side, the 
456: oscillations become very irregular and very likely to be chaotic (Figure \ref{f.4}).  
457: When $\alpha >0$ is fixed and $\gamma$ crosses 0, 
458: the irregular dynamics is followed by the reverse period-doubling cascade and terminates with the creation of a 
459: regular small amplitude periodic orbit, which can be then tracked 
460: down to a nondegenerate supercritical AHB (Figure \ref{f.5}a-h).
461: This scenario presents a certain interest as it suggests the formation of 
462: the chaotic attractor in a well-defined bifurcation setting.  Namely, the transition 
463: from the region of subcritical to supercritical AHB for small $\alpha >0$ leads 
464: to the formation of a chaotic attractor and a reverse period-doubling cascade.
465: As in the case of the complex dynamics appearing for 
466: increasing values of $\alpha$ for supercritical AHB, the irregular oscillations in the present 
467: case can not be understood using the local analysis alone.  Therefore, 
468: we complemented our analysis with the study of numerically constructed first return map.  
469: The latter gives a clear geometric picture of the origins of the chaotic dynamics and the 
470: period-doubling cascade of periodic orbits in this parameter regime.   
471: 
472: 
473: \section{The local expansions}
474: \setcounter{equation}{0}
475: In the present section, we study trajectories of (\ref{6.5}) in a small neighborhood of the 
476: equilibrium at the origin. Assume that $h(x,\alpha)$ is a smooth function in a small neighborhood
477: of $(0,0)$, so that all power series expansions below are justified. 
478: We are interested in the dynamics of (\ref{6.5}) for values of $\alpha$ close to $0$. 
479: The role of $\alpha$ in our analysis is twofold. On the one hand, $\alpha$ is a control
480: parameter in this problem, on the other hand, the smallness of $\alpha$ is used
481: in the asymptotic analysis below. To separate these two roles, we use the following
482: rescaling
483: \be\lbl{5.1a}
484: \alpha=\mu\epsilon^2,
485: \ee
486: where $\epsilon>0$ is a fixed sufficiently small constant and $\mu$ varies in a certain interval
487: of size $O(1)$. We introduce cylindrical coordinates in $\Re^3$
488: \be\lbl{cl}
489: \left(x_1,x_2,x_3\right)\mapsto\left(x_1, \rho,\theta\right)\in\Re\times\Re^+\times S^1,\quad
490: x_2=\rho\cos\theta,\;x_3=\rho\sin\theta,
491: \ee
492: and define
493: \be\lbl{5.2}
494: D=\left\{ \left(x_1,\rho, \theta \right):\; \left| x_1 \right| \le M^2\epsilon^2, \rho\in(0,2M\epsilon]\right\},\;
495: D_0=\left\{ \left(x_1,\rho, \theta \right):\; \left| x_1 \right| \le M^2\epsilon^2, \rho\in(0,M\epsilon]\right\},
496: \ee
497: where $M>0$ is sufficiently large.
498: 
499: To characterize the trajectories of (\ref{6.5}) in $D$ we will use an exponentially stable slow manifold, $S$,
500: whose leading order approximation is given by
501: \be\lbl{5.3}
502: S_0= \left\{ 
503: \left( x_1,\rho,\theta\right):\; x_1=U(\theta)\rho^2,\; 0\le\rho\le 2\epsilon M\right\},
504: \ee
505: where
506: \be\lbl{5.4}
507: U(\theta)= a + A\cos\left(2\theta-\vartheta\right).
508: \ee
509: and $a, A$ and $\vartheta$ are computable constants (see Appendix A). 
510: 
511: 
512: Below we show that  the trajectories of (\ref{6.5}) with initial data from $D_0$ approach an
513: $O(\epsilon)$  neighborhood of $S_0$ in time $O\left(\left| \ln \epsilon \right|\right)$
514: and stay in this neighborhood as long as they remain in $D$. The reduction of the system dynamics to $S$
515: yields a complete description of the trajectories of (\ref{6.5}) in $D$. The qualitative character of the
516: solution behavior in $D$ depends on the sign of {\it the first Lyapunov coefficient}
517: \be\lbl{5.5}
518: \gamma={1\over 2\pi}\int_0^{2\pi} \left( U\left(\theta\right) \bar Q_2\left(\theta\right)+
519: \bar Q_3\left(\theta\right)\right) d\theta,
520: \ee
521: where $2\pi-$periodic trigonometric polynomial $\bar Q_{2,3}$ are given in Appendix A.
522: Finally, by $\beta=b(0)$ we denote the absolute value of the imaginary parts of the
523: complex conjugate pair of eigenvalues at the AHB (see (\ref{6.5})).
524: The results of the analysis of this section are summarized in
525: 
526: \noindent
527: {\bf Theorem 3.1} 
528: {\sc
529: Suppose $\beta>0$ and $\gamma\ne 0$. Let $\epsilon>0$ denote a small parameter.
530: Then for $\alpha=O(\epsilon^2)$ and for sufficiently large (independent of $\epsilon$)
531: $M>0$, the trajectories
532: with initial conditions in $D_0$ enter an $O(\epsilon)$ neighborhood of $S_0$ in time
533: $O\left(\left|\ln\epsilon\right|\right)$ and remain there as long as they stay in $D$.
534: In the neighborhood of $S_0$, the trajectories can be uniformly approximated on any
535: interval of time $[t_0,\bar t]$ of length $O(1)$:
536: \be\lbl{5.7}
537: x_1(t)=\rho^2 U(\theta)+O(\epsilon^3), \quad x_2(t)=\rho\cos\theta +O(\epsilon^2),\quad\mbox{and}\quad
538: x_3(t)=\rho\sin\theta +O(\epsilon^2),
539: \ee
540: where $U(\theta)$ and $\gamma$ are defined in (\ref{5.4}) and (\ref{5.5}), respectively,  and
541: $$
542: \rho(t) = \left(\left(\rho(t_0)^{-2}+{\gamma\over\alpha}\right) e^{-2 \alpha(t-t_0)}-
543: {\gamma\over\alpha}\right)^{-1\over 2}+O(\epsilon^2),\quad
544: \dot\theta=\beta,\quad
545: \tan\left(\theta(0)\right)= {x_3(t_0)\over x_2(t_0)}.
546: $$
547: } 
548: 
549: \noindent {\bf Remark 3.1}
550: \begin{description}
551: \item[a)] From Theorem 3.1, one can easily deduce the existence of a periodic orbit $O_\alpha$,
552: when $\alpha\gamma<0$. $O_\alpha$ is stable  (unstable) if $\alpha>0$ ($\alpha<0$). The leading order
553: approximation for $O_\alpha$ follows from (\ref{5.7}):
554: \be\lbl{5.8}
555: x_1=\bar\rho^2 U(\theta)+O(\epsilon^3),\;\; x_2=\bar\rho\cos\theta +O(\epsilon^2),\;\;
556: x_3=\bar\rho\sin\theta +O(\epsilon^2),\;\;
557: \mbox{and}\;\;
558: \theta\in [0, 2\pi),
559: \ee
560: where $\bar\rho=\sqrt{\alpha\over -\gamma}$. Equations (\ref{5.4}) and (\ref{5.8})
561: imply that the frequency of oscillations in $x_1$ is twice
562: that of oscillations in $x_{2,3}$, provided $A\ne 0$ in (\ref{5.4}) 
563: (see Figure \ref{f.1}a,b). In Section 4.1, we give a geometric interpretation of this frequency doubling 
564: effect.
565: \item[b)] If $\alpha>0$ and $\gamma>0$, Theorem 3.1 describes the trajectories with initial conditions
566: near a weakly unstable equilibrium. It shows that along these trajectories, $x_{2,3}$ undergo  approximately
567: harmonic oscillations, whose amplitude grows at the rate $O(\epsilon^2)$. In addition, they satisfy the
568: following scaling relation:
569: \be\lbl{5.9}
570: \bar x_1(t) \sim a\rho^2(t),
571: \ee
572: where $\bar x_1(t)$ stands for the average value of $x_1(t)$ over one cycle of oscillations.
573: \item[c)]
574: The domain of validity of the asymptotic expansions in (\ref{5.7}) extends much farther than
575: $D$. As will follow from the proof of Theorem 3.1, $M$ in the definition of $D$ may be taken 
576: up to $o\left(\epsilon^{-1\over 3}\right)$. This
577: implies that, with the error term $o(1)$, (\ref{5.7}) remains valid for the ranges of $x_1$ and $x_{2,3}$ up to 
578: $o(\epsilon^{4\over 3})$ and $o(\epsilon^{2\over 3})$, respectively.
579: \end{description}
580: 
581: In the remainder of this section, we prove Theorem 3.1. By expanding
582: $h(x,\alpha)$ into finite Taylor sum with the reminder term, we have  
583: \begin{eqnarray}\nonumber
584: \dot x_1 &=& -x_1+\sum a_{ij}(\alpha)x_i x_j +\sum a_{ijk}(\alpha) x_ix_jx_k + O(4),\\ \lbl{5.1}
585: \dot x_2 &=& a(\alpha) x_2 -b(\alpha) x_3 +\sum b_{ij}(\alpha)x_i x_j +\sum b_{ijk}(\alpha) x_ix_jx_k +O(4),\\ \nonumber
586: \dot x_3 &=& b(\alpha) x_2 +a(\alpha) x_3 +\sum c_{ij}(\alpha)x_i x_j +\sum c_{ijk}(\alpha) x_ix_jx_k +O(4),
587: \end{eqnarray}
588: where
589: \begin{eqnarray*}
590: \sum \sigma_{ij}(\alpha) x_ix_j &=& \sum_{i=1}^3 \sum_{j=i}^3 \sigma_{ij}(\alpha) x_ix_j,\\
591: \sum \sigma_{ijk}(\alpha) x_ix_jx_k &=& \sum_{i=1}^3 \sum_{j=i}^3 \sum_{k=j}^3 \sigma_{ijk}(\alpha) x_ix_jx_k,\quad
592: \sigma\in\left\{a,b,c\right\}.
593: \end{eqnarray*}
594: To study system of equations (\ref{5.1}) in a small neighborhood of the origin, 
595: we rewrite it in cylindrical coordinates (\ref{cl}) and rescale variables
596: \be\lbl{5.10}
597:  x_1=\xi\epsilon^2\;\mbox{and}\; \rho=r\epsilon.
598: \ee
599: In new coordinates, we have
600: \begin{eqnarray}
601: \dot \xi &=&-\xi+ r^2 T_1(\theta) +O\left(\epsilon\right),\nonumber\\ \lbl{5.11}
602: \dot r &=& \epsilon r^2 R_1(\theta)+
603: \epsilon^2 r\left(\mu+\xi R_2(\theta)+r^2R_3(\theta)\right)+O(\epsilon^3),\\
604: \nonumber
605: \dot\theta &=& \beta+\epsilon r L_1(\theta)+ O\left(\epsilon^2\right),
606: \end{eqnarray}
607: where $T_1, L_1$ and $R_{1,2,3}$ are homogeneous trigonometric polynomials.
608: 
609: Using $\phi=\beta^{-1}\theta$ as a new variable, from (\ref{5.11}), we obtain
610: \begin{eqnarray}\lbl{5.11a}
611: \frac{d\xi}{d\phi}&=&-\xi+r^2 P_1(\phi)+\epsilon \Xi\left(\xi,r,\phi,\epsilon\right),\\ 
612: \lbl{5.12}
613: \frac{dr}{d\phi}&=&\epsilon Q_1(\phi) r^2 +\epsilon^2 r\left(\mu+\xi Q_2(\phi)+r^2 Q_3(\phi)\right)+
614: \epsilon^3 \Psi\left(\xi,r,\phi,\epsilon\right),
615: \end{eqnarray}
616: where $Q_{1,2,3}(\phi)$ and $P_1(\phi)$ are trigonometric polynomials of period $\omega=2\pi\beta^{-1}$.
617: Functions $\Xi$ and $\Psi$ are bounded and continuous in $D'\times\Re\times [0, 1]$.
618: Here $D'$ stands for the domain of the rescaled variable $(\xi,r)$ when the original variable
619: $x\in D\subset \Re^3$:
620: \be\lbl{Dprime} 
621: D^\prime= \left\{ \left( \xi,r \right):\; \left| \xi \right| \le 4M^2, \; 0<r\le 2 M \right\}.
622: \ee
623: Similarly, we define
624: \be\lbl{D0prime}
625: D^\prime_0=\left\{ \left(\xi,r\right):\; \left| \xi \right| \le M^2, \; 0<r\le M\right\}.
626: \ee
627: 
628: To determine the slow manifold for (\ref{5.1}), we introduce the following notation. 
629: By $\bar p_1$ and $\tilde P_1(\phi)$ we denote the mean value and the oscillating part 
630: of the trigonometric polynomial $P_1(\phi)$ in (\ref{5.11a}), respectively:
631: \be\lbl{5.13}
632: \bar p_1={\beta\over 2\pi}\int_0^{2\pi\over\beta} P_1(\phi)d\phi
633: \quad\mbox{and}\quad \tilde P_1(\phi)= P_1(\phi)-\bar p_1,
634: \ee
635: and define
636: \be\lbl{5.16}
637: \xi_0(\phi)=\bar p_1 +p_1(\phi) \quad\mbox{and}\quad
638: p_1(\phi)=
639: {\tilde P_1(\phi)-{\tilde P_1}^\prime (\phi)\over 1+4\beta^2}.
640: \ee
641: 
642: We first prove an auxiliary lemma, which characterizes the trajectories of (\ref{5.11a}) and (\ref{5.12})
643: in $D^\prime$.
644: In the following two lemmas, we will keep track of how the remainder terms in the asymptotic expansions
645: depend on the size of $D^\prime$. This will be used later to determine the domain of validity for the 
646: asymptotic approximation of the slow manifold. Below $\kappa=O\left(M\epsilon\right)$ means that there
647: exist positive constants $\epsilon_0$ and $C$ (independent of $\epsilon$) such that
648: $\left|\kappa\right| \le CM\epsilon$ for $\epsilon\in\left[0, \epsilon_0\right]$.
649: In the remainder of this section, all estimates are valid for sufficiently large $M>0$ independent 
650: of $\epsilon$ (as stated in the Theorem 3.1), but also allow the possibility that $M>1$ grows as 
651: $\epsilon\to 0$.
652: 
653: \noindent
654: {\bf Lemma 3.1} 
655: {\sc Suppose 
656: $\left(\xi, r\right)$ remains in $D^\prime$ for $\phi\in\left[\phi_0,\bar\phi\right]$. Then
657: \be\lbl{5.15}
658: \xi(\phi)= \xi_0(\phi)r^2(\phi) +
659: e^{-(\phi-\phi_0)}\left\{ \xi(\phi_0)-\xi_0(\phi_0) r^2(\phi_0)\right\} +O\left(M^3\epsilon\right),
660: \quad\phi\in\left[\phi_0, \bar\phi\right],
661: \ee
662: where $\xi_0(\phi)$ is defined in (\ref{5.16}).
663: }
664: 
665: \noindent{\bf Proof:}
666: Using (\ref{5.13}), we integrate (\ref{5.11a}) over $\left[\phi_0,\phi\right]$
667: \be\lbl{5.17}
668: e^\phi\xi(\phi)=e^{\phi_0}\xi(\phi_0)+
669: \bar p_1 \int_{\phi_0}^\phi e^s r^2(s)ds  + \int_{\phi_0}^\phi e^s r^2(s) \tilde P_1(s) ds+ 
670: \epsilon\int_{\phi_0}^\phi e^s \Xi\left(\xi(s), r(s),s,\epsilon\right) ds.
671: \ee
672: Using integration by parts, we have
673: $$
674: \int_{\phi_0}^\phi e^s r^2(s)ds=e^\phi r^2(\phi)-e^{\phi_0} r^2(\phi_0)
675: -2\int_{\phi_0}^\phi e^s r(s)\dot r(s)ds.
676: $$
677: From (\ref{5.12}) and (\ref{Dprime}), we find that for $\left(\xi,r\right)\in D^\prime$
678: $$
679: \left|\int_{\phi_0}^\phi e^s r(s)\dot r(s)ds\right| \le C_1 e^{\phi-\phi_0}M^3\epsilon,
680: $$ 
681: where $C_1>0$ is independent of $\epsilon$.
682: Therefore,
683: \be\lbl{5.18}
684: \int_{\phi_0}^\phi e^s r^2(s)ds=e^\phi r^2(\phi)-e^{\phi_0} r^2(\phi_0)
685: +e^{\phi-\phi_0}O\left(M^3\epsilon\right).
686: \ee
687: Similarly, using integration by parts in the second integral on the right hand
688: side of (\ref{5.17}), we obtain
689: \begin{eqnarray}\nonumber
690: \int_{\phi_0}^\phi e^s r^2(s) \tilde P_1(s) ds &=& e^\phi r^2(\phi) \tilde P_1(\phi)-e^{\phi_0} r^2(\phi_0) \tilde P_1(\phi_0)\\
691:    & &-2\int_{\phi_0}^\phi e^s r(s) \dot r(s) \tilde P_1(s)ds - \int_{\phi_0}^\phi e^s r^2(s) \tilde P_1^\prime(s)ds.\lbl{A.2}
692: \end{eqnarray}
693: Applying integration by parts in the last integral on the right hand side of (\ref{A.2}),
694: we have
695: \begin{eqnarray}\nonumber
696: \int_{\phi_0}^\phi e^s r^2(s) \tilde P_1^\prime(s) ds &=& e^\phi r^2(\phi) \tilde P_1^\prime(\phi)-e^{\phi_0}r^2(\phi_0) \tilde P_1^\prime(\phi_0)\\
697: && -2\int_{\phi_0}^\phi e^s r(s) \dot r(s) \tilde P_1^\prime (s)ds - \int_{\phi_0}^\phi e^s r^2(s) \tilde P_1^{\prime\prime}(s)ds.
698: \lbl{A.3}
699: \end{eqnarray}
700: By direct verification, one finds that 
701: $
702: \tilde P_1^{\prime\prime}=\left(1+4\beta^2\right)\tilde P_1.
703: $
704: The integrals involving $\dot r$ on the right hand sides of (\ref{A.2}) and
705: (\ref{A.3}) are bounded by $e^{\phi-\phi_0}O(M^3\epsilon)$ since, by (\ref{5.12}), $\dot r=O(\epsilon M^2)$.
706: Using these observations, from (\ref{A.2}) and (\ref{A.3}) we obtain
707: \be\lbl{A.4}
708: \int_{\phi_0}^\phi e^s r^2(s) \tilde P_1(s)ds = e^\phi r^2(\phi)p_1(\phi)-e^{\phi_0} r^2(\phi_0)p_1(\phi_0)+ e^{\phi-\phi_0}O(M^3\epsilon),
709: \ee
710: where $p_1(\phi)$ is given by (\ref{5.16}).
711: The combination of (\ref{5.17}), (\ref{5.18}), and  (\ref{A.4}) yields (\ref{5.15}).
712: \\
713: $\Box$
714: 
715: The following lemma shows that the trajectories of (\ref{5.11a}) and (\ref{5.12}),
716: which stay in $D^\prime$ for sufficiently long time 
717: enter a small neighborhood of $S$ 
718: no later than in time $O(\left|\ln\epsilon\right|)$
719: and remain in this neighborhood as long as they stay in $D^\prime$ (see (\ref{Dprime})).
720: 
721: \noindent
722: {\bf Lemma 3.2} 
723: {\sc 
724: Let $\left(\xi(\phi),r(\phi)\right)$ denote a trajectory of (\ref{5.11a}) and (\ref{5.12}),
725: which stays in $D^\prime$ for $\phi\in \left[\phi_0, \bar\phi\right]$, $\bar \phi>\phi_1$,
726: $\phi_1=\phi_0+\left|\ln\epsilon\right|$. Then
727: \be\lbl{5.20}
728: \xi(\phi)=\xi_0(\phi) r^2(\phi) +O\left(M^3\epsilon\right),
729: \ee
730: for $\phi\ge\phi_1$ and as long as $\left(\xi(\phi),r(\phi)\right)\in D^\prime$.
731: }
732: 
733: \noindent{\bf Proof:} Denote $\phi_1=\phi_0+\left|\ln\epsilon\right|$.
734: By plugging in $\phi=\phi_1$ into (\ref{5.15}), we have
735: \be\lbl{5.21}
736: \left|\xi(\phi_1) - \xi_0(\phi_1)r^2(\phi_1)\right|\le C_2\epsilon,
737: \ee
738: for some $C_2>0$ independent of $\epsilon>0$.
739: Using Lemma 3.1 again with $\phi_0:=\phi_1$, we obtain
740: $$
741: \xi(\phi)= \xi_0(\phi)r^2(\phi) +
742: e^{-(\phi-\phi_1)}\left\{\xi(\phi_1)-\xi_0(\phi_1)r^2(\phi_1)\right\}
743: +O(M^3\epsilon),\;\phi\ge \phi_1.
744: $$
745: The expression in the curly brackets is $O(\epsilon)$ by (\ref{5.21}).
746: \\
747: $\Box$
748: 
749: \noindent {\bf Remark 3.2}
750: Lemmas 3.1 and 3.2 show that the trajectories of (\ref{5.11a}) and (\ref{5.12})
751: converge to an exponentially stable manifold $S$, whose  
752: leading order approximation is given in the definition of $S_0$. The method, 
753: which we used in Lemmas 3.1 and 3.2 to obtain the
754: leading order approximation of the slow manifold, can be extended to calculate the
755: higher order terms in the expansion for $S$.
756:  
757: Having shown that the trajectories approach an $O(\epsilon)$ neighborhood of $S_0$ in time 
758: $O\left(\left|\ln\epsilon\right|\right)$, next we
759: reduce the dynamics of (\ref{5.11a}) and (\ref{5.12}) to the slow manifold. 
760: For this, we define
761: \be\lbl{5.22}
762: I(\phi)=\left({1\over r(\phi)}+\epsilon q_1(\phi)\right)^2,
763: \ee
764: where $q_1^\prime(\phi)=Q_1(\phi)$ and $Q_1(\phi)$ is a trigonometric polynomial on 
765: the right hand side of (\ref{5.12}). For the sake of definiteness, 
766: we choose $q_1^\prime(\phi)$ such that 
767: $$
768: \int_0^{2\pi\over\beta}q_1(s)ds=0.
769: $$
770: The following lemma provides the desired reduction.
771: 
772: \noindent
773: {\bf Lemma 3.3} 
774: {\sc For $\phi\ge\phi_1$ and as long as $(\xi,r)\in D^\prime$, $(I,\phi)$ satisfies the 
775: following system of equations
776: \begin{eqnarray}\lbl{5.24}
777: \dot I &=& -2\epsilon^2\left(\mu I+\gamma+Q(\phi)\right)+O(M^3\epsilon^3),\\ \lbl{5.25}
778: \dot \phi &=& 1+O(M\epsilon),
779: \end{eqnarray}
780: where $Q(\phi)$ is a trigonometric polynomial with period $\omega={2\pi\beta^{-1}}$ and zero mean 
781: $\int_0^\omega Q(\phi) d\phi =0$. The expression for $\gamma$ is given in
782: (\ref{5.5}).
783: }
784: 
785: \noindent{\bf Proof:}
786: The change of variables
787: \be\lbl{5.26}
788: r(\phi)={1\over J(\phi)-\epsilon q_1\left(\phi\right)},\quad \mbox{where}\quad q_1^\prime(\phi)=Q_1(\phi),
789: \ee
790: in Equation (\ref{5.12}) yields
791: \be\lbl{5.27}
792: \dot J ={-\epsilon^2\over J}\left(\left(\mu+\xi Q_2(\phi) \right) J^2 +Q_3(\phi)\right)+O(M^2\epsilon^3).
793: \ee
794: After another change of variables, $I=J^2$, we obtain
795: \be\lbl{5.28}
796: {d I\over d\phi} = -2\epsilon^2 \left( \left(\mu+ \xi Q_2(\phi)\right) I +Q_3(\phi)\right)
797: + O(M^3\epsilon^3),
798: \ee
799: By Lemma 3.2, for $\phi\ge\phi_1$
800: \be\lbl{5.29}
801: \xi(\phi)=\xi_0(\phi)r^2(\phi) +O(M^3\epsilon).
802: \ee
803: By plugging in (\ref{5.29}) into (\ref{5.28}), we obtain
804: \be\lbl{5.30}
805: {d I\over d\phi} = -2\epsilon^2\left(
806: \mu I +\xi_0(\phi)Q_2(\phi)+Q_3(\phi)\right)+
807: O(M^3\epsilon^3).
808: \ee
809: Equation (\ref{5.24}) follows by rewriting  (\ref{5.30}) 
810: \be\lbl{5.31}
811: {d I\over d\phi} = -2\epsilon^2\left(
812: \mu I +\gamma+ Q(\phi)\right)+
813: O(M^3\epsilon^3),
814: \ee
815: where
816: \be\lbl{5.32}
817: \gamma = {1\over\omega}\int_0^\omega\left(\xi_0(\phi)Q_2(\phi)+Q_3(\phi)\right)d\phi\;\;\mbox{and}\;\; 
818: \int_0^{\omega} Q(\phi)d\phi =0,
819: \ee
820: and $\xi_0(\phi)$ is given in (\ref{5.16}).
821: Equation (\ref{5.25})  follows from the last equation in (\ref{5.11}) and the definition of $\phi$.\\
822: $\Box$
823: 
824: The statements in Theorem 3.1 can now be deduced from Equations (\ref{5.24}) and (\ref{5.25}).
825: We only need to show that a trajectory with initial condition in $D_0^\prime$ remains in $D^\prime$
826: for times longer than $O\left(\left|\ln\epsilon\right|\right)$. This follows from the fact that
827: $\dot I=O\left(\epsilon^2\right)$ in $D^\prime$. Consequently, it takes time $O(\epsilon^{-2})$ for
828: $I$ to undergo $O(1)$ change necessary for leaving $D^\prime$ from $D^\prime_0$.
829: We omit any further details. In conclusion, we  note that the domain of validity of the asymptotic 
830: analysis of this section extends
831: much further beyond $D^\prime$. Indeed, from (\ref{5.20}) we observe that the remainder term 
832: tends to $0$ with $\epsilon\to 0$ provided $M=o\left(\epsilon^{-1\over 3}\right)$. This means that
833: the expansions for $r$ and $\xi$ can be controlled in regions of size up to $O\left(\epsilon^{-1\over 3}\right)$
834: and $O\left(\epsilon^{-2\over 3}\right)$ respectively. For the original variables $\rho,$
835: $x_1$, these estimates translate into  $O\left(\epsilon^{2\over 3}\right)$
836: and $O\left(\epsilon^{4\over 3}\right)$ respectively.
837: 
838: We end this section by deriving a useful estimate for the time of flight of trajectories
839: passing close to the saddle-focus. For this, consider an initial value problem for system of equations
840: (\ref{5.11a}) and (\ref{5.12}). Suppose that the initial condition implies that $I\left(\phi_0\right)=I_0>0$.
841: We would like to know how long it takes for $I$ to reach a given value $\bar I=I\left(\phi_0+\Delta\right)$,
842: $\Delta>0$.
843: We assume that $I_0$ and $\bar I$ are sufficiently separated, e.g., $2\bar I\le I_0$.
844: To estimate $\Delta$, note that
845: \be\lbl{5.51}
846: \Delta=\Delta_1+\Delta_2,
847: \ee
848: where $\Delta_1=O\left(\left|\ln\epsilon\right|\right)$ is time necessary to reach 
849: an $O(\epsilon)$ neighborhood of $S_0$ from $I_0$. Denote 
850: \be\lbl{5.52}
851:   I_1:= I\left(\phi_1\right)=I_0+O\left(\epsilon^2\left|\ln\epsilon\right|\right),\;
852:   \mbox{where}\; \phi_1=\phi_0+\Delta_1.
853: \ee
854: The second term on the right hand side of (\ref{5.51}) can be estimated by integrating (\ref{5.24})
855: over $\left[ \phi_1, \phi_1+\Delta_2\right]:$
856: \be\lbl{5.53}
857: \bar I= I_1 e^{-2\alpha\Delta_2}+{\gamma\over\mu}\left(1- e^{-2\alpha\Delta_2}\right)
858: -2\epsilon^2\int_0^{\Delta_2} e^{-2\alpha\left(\Delta_2-s\right)} Q\left(\phi_0+s\right)ds +
859: O\left(\epsilon^3\right).
860: \ee
861: The integral on the right hand side of (\ref{5.53}) is bounded uniformly for $\Delta_2>0$.
862: This observation combined with (\ref{5.52}) implies 
863: \be\lbl{5.54}
864: \bar I= I_0 e^{-2\alpha\Delta_2}-{\gamma\over\mu}\left(1- e^{-2\alpha\Delta_2}\right)
865: +O\left(\epsilon^2\right).
866: \ee
867: Note that the contribution of $O\left(\epsilon^2 \left| \ln\epsilon \right|\right)$ term in (\ref{5.52})
868: to (\ref{5.54}) is negligible provided $\Delta_2$ is sufficiently large.
869: From  (\ref{5.54}), we obtain the desired estimate
870: \be\lbl{5.55}
871: \Delta={1\over 2\mu\epsilon^2}\ln \left({I_0 +{\gamma\over \mu}\over \bar I +{\gamma\over\mu}+O(\epsilon^2)}\right).
872: \ee
873: 
874: 
875: \section{The oscillations}
876: \setcounter{equation}{0}
877: 
878: In the present section, we use the local analysis of Section 3 to study certain oscillatory patterns
879: arising in the model of solid fuel combustion (\ref{6.1})-(\ref{6.3}). By plugging in the values of the 
880: parameters of (\ref{6.1})-(\ref{6.3}) into the expression for $\gamma$ (\ref{5.5}), we find that $\gamma(p)$
881: is a quadratic function with two zeros at $p_1\approx 0.34$ and $p_2\approx 2.73$. 
882: These values are in a good agreement with those obtained by numerical bifurcation analysis in \cite{FKRT}.
883: The quadratic character of $\gamma(p)$ is explained by the fact that $\gamma$ is determined by the
884: second order terms in the Taylor expansions of $h_{2,3}(0,0)$.
885: We concentrate on the parameter region around $p=p_2$, where $\gamma(p)$ changes its sign.
886: For small positive $\alpha$, there are two dynamical regimes: for values of $p$ lying 
887: to the left and to the right of some $O(\epsilon)$ neighborhood
888: of $p_2$ corresponding to the supercritical and subcritical AHBs. The transition region between these two
889: parameter regimes adds to the repertoire of qualitatively distinct dynamical behaviors. Below, we study
890: these three cases in more detail.
891: 
892: \subsection{The supercritcal AHB: geometry of the periodic orbit}
893: \setcounter{equation}{0}
894: It follows from Theorem 3.1 that for $\gamma(p)<0$ and sufficiently small $\alpha>0$,
895: (\ref{6.5}) has a stable limit cycle $O_\alpha$, whose leading order approximation is given by
896: \be\lbl{4.1}
897: x(\theta)=\left(\bar\rho^2\left(a+A \cos 2\theta\right), \bar\rho\cos\theta, \bar\rho\sin\theta\right),
898: \quad\bar\rho=\sqrt{\alpha\over -\gamma},\quad \theta\in[0, 2\pi)
899: \ee
900: (see Figure \ref{f.6}a).
901: Moreover, the analysis of Section 3 shows that all trajectories starting from a sufficiently small 
902: neighborhood of the origin and not belonging to $W^s(O)$, converge to $O_\alpha$. The leading
903: order approximation of $O_\alpha$ in (\ref{4.1}) reveals a remarkable property of the oscillations 
904: generated by the limit cycle born from the supercritical AHB: the frequency of oscillations in $x_1$
905: is twice as large as that of the oscillations in $x_{2,3}$. To explain the frequency doubling effect,
906: we recall that 
907: \be\lbl{4.2}
908: \dot\theta=\beta+O(\epsilon).
909: \ee
910: Therefore, (\ref{4.1}) implies that, unless $A=0$,
911: the frequency of oscillations in $x_1$ is $\beta\pi^{-1}$, while that of oscillations in $x_{2,3}$
912: is ${\beta\over 2\pi}$. The latter coincides with the frequency of the bifurcating periodic solution.
913: Below, we complement the analytical explanation of the frequency doubling with the geometric interpretation.
914: 
915: The oscillations in $x_{2,3}$ can be well understood using the topological normal form of the AHB
916: \cite{GEO,GH, KUZ}. Indeed, since at the bifurcation the center manifold at the origin is tangent to the $x_2-x_3$
917: plane, a standard treatment of the AHB using the center manifold reduction \cite{Carr, GH} shows that the projection
918: of the bifurcating limit cycle onto $x_2-x_3$ plane to leading order is a circle (Figure \ref{f.6}b)
919: and the projection
920: of the vector field is given by the equation of the angular variable (\ref{4.2}). Therefore, the oscillations in
921: $x_{2,3}$ are approximately harmonic with the period equal approximately to $2\pi\beta^{-1}$. The topological
922: normal from, however, does not describe the oscillations in $x_1$. For this, one needs to take into account the
923: geometry of the bifurcating periodic orbit as a curve in $\Re^3$. The geometry of $O_\alpha$ is fully determined by 
924: the two geometric invariants: the curvature, $\mbox{k}(\theta)$, and the torsion, $\kappa(\theta)$. For the purposes
925: of the present discussion, we need only the latter, but we compute both invariants for completeness.
926: After some algebra, the parametric equation for the periodic orbit (\ref{4.1}) yields  
927: \begin{eqnarray}\nonumber 
928: \mbox{k}(\theta)&=& {\left|\left[\dot x, \ddot x\right]\right|\over \left|\dot x\right|^3}=
929: \sqrt{1+2A^2\left(3\cos 4\theta +5\right)}+\mbox{h.o.t},\\
930: \lbl{4.3} 
931: \kappa(\theta) &=& {\left(\dot x, \ddot x, \dddot x\right)\over\left|\left[\dot x, \ddot x\right]\right|^2}=
932: {-\gamma\over\alpha} {6A\sin 2\theta\over 1+2A^2\left(3\cos 4\theta +5\right)}
933: +\mbox{h.o.t},\quad \theta\in[0, 2\pi).
934: \end{eqnarray}
935: The geometry of the leading order approximation of the periodic orbit is determined by the three
936: parameters $\alpha,$ $\gamma,$ and $A$. The former two parameters are the same as the parameters in the  
937: topological normal form of the nondegenerate AHB \cite{KUZ}; while the latter captures the geometry of the slow manifold (or unstable
938: manifold) near the origin. From the geometrical viewpoint the bifurcation is degenerate if either $\gamma$ or $A$ is equal
939: to zero. The latter condition holds if and only if the slow manifold is either a plane or a circular paraboloid
940: near the origin. In this case, Equation (\ref{4.3})
941: implies that (to leading order) the bifurcating orbit lies in a plane. Generically, $O_\alpha$ is not planar 
942: (see Figure \ref{f.6}a). 
943: The fact that the periodic orbit is generically not planar combined with the
944: symmetry of the orbit about the $x_1-$axis implies that its projection onto any plane containing $x_1-$axis has to have
945: a self-intersection (see Figure \ref{f.6}c). From  Figure \ref{f.6}c, it is clear that as the phase point goes 
946: around $O_\alpha$ once, $x_1$
947: has to trace its range at least twice. Therefore, the frequency of oscillations in $x_1$ has to be at least twice
948: as high as that of the oscillations in $x_{2,3}$. The above discussion implies that the frequency doubling of oscillations in
949: $x_1$ is a generic  geometric property of the AHB.  
950: \begin{figure}
951: \begin{center}
952: {\bf a}\epsfig{figure=f6a.eps, height=1.8in, width=2.0in, angle=0}
953: {\bf b}\epsfig{figure=f6b.eps, height=1.8in, width=2.0in, angle=0}
954: {\bf c}\epsfig{figure=f6c.eps, height=1.8in, width=2.0in, angle=0}
955: \end{center}
956: \caption{ A periodic orbit born from the supercritical AHB ({\bf a}), and
957: its projections onto $x_2-x_3$({\bf b}) and $x_1-x_2$ planes ({\bf c}).
958: Note that the periodic orbit shown in ({\bf a}) is not planar. This 
959: accounts for the presence of the self-intersection in the its projection
960: in ({\bf c}) (see text for details).
961: }\label{f.6}
962: \end{figure}
963: 
964: 
965: \subsection{The subcritcal AHB: multimodal oscillations}
966: 
967: If the AHB is subcritical ($\gamma>0$) the loss of stability of the equilibrium at the origin of (\ref{6.5}) 
968: results in the creation of multimodal trajectories which spend a considerable amount of time near a weakly 
969: unstable equilibrium. To describe the resultant dynamics we give the following definition.
970: 
971: \noindent
972: {\bf Definition 4.1} 
973: {\sc We say that a trajectory of (\ref{6.5}) undergoes multimodal oscillations for $t\ge 0$ if
974: there exist positive constants $r_1$ and $r_2$ independent of $\epsilon$ and an unbounded sequence of times
975: $$
976: 0\le t_1<t_2<\dots,\quad \lim_{i\to\infty} t_i=\infty,
977: $$ 
978: such that 
979: \begin{description}
980: \item[a.] $\left|x\left(t_{2i}\right)\right|\le r_1\epsilon^2\;\;\&\;\;\left|x\left(t_{2i-1}\right)\right|> r_2,\quad i=1,2,\dots$,
981: \item[b.] $\left(\exists t^\prime> t_1:\;\left| x(t^\prime)\right| \le r_1\epsilon^2\right)\;\Rightarrow\; 
982: \exists i\in\Na\;:\;
983: \left|x(t)\right|\le r_1\epsilon^2, 
984: t\in\left[ \min(t^\prime, t_{2i}), \max(t^\prime, t_{2i})\right]\;\;\forall t^\prime \in \left(t_{2i-1}, t_{2i+1}\right).$
985: \end{description}
986: The time intervals $\tau_i=t_{2i+1}-t_{2i-1},\; i=1,2,\dots$ are called ISIs.
987: }
988: 
989: The proximity of (\ref{6.5}) to the AHB alone is clearly not sufficient to account for  the appearance of 
990: the multimodal oscillations in (\ref{6.5}).
991: Below, we formulate two additional assumptions on the vector field outside of the small neighborhood of the equilibrium, which
992: are relevant to (\ref{6.1})-(\ref{6.3}).
993: Under these conditions, we show that the subcritical AHB results in sustained multimodal
994: oscillations. In addition, we determine the asymptotics of the ISIs for positive $\alpha=O(\epsilon^2)$. 
995: For (\ref{6.5}) to generate multimodal oscillations, it is necessary that the trajectories leaving an $O(\epsilon^2)$
996: neighborhood of the origin reenter it after some interval of time. Therefore, the vector field in $O(1)$ neighborhood of the
997: origin must provide a return mechanism. Our second assumption on the global vector field of (\ref{6.5})  
998: is that the trajectories approaching the origin
999: along a $1D$ stable manifold, $W^s(O)$, are subject to a strong contraction toward $W^s(O)$, i.e., in a small neighborhood
1000: of an $O(1)$ segment of $W^s(O)$, the projection of the vector field onto a plane transversal to $W^s(O)$ is sufficiently stronger than
1001: that along $W^s(O)$ (Figure \ref{f.2}a). This guarantees that the trajectories entering such 
1002: region of strong contraction approach $W^s(O)$ very 
1003: closely and follow it to an $O(\epsilon^2)$ neighborhood of the unstable equilibrium, from where they are propelled away along 
1004: the unstable manifold, $W^u(O)$. Due to the proximity of (\ref{6.5}) to the AHB, the motion away from the origin is very slow.
1005: This results in the pronounced intervals of the small amplitude oscillations. Below, we  summarize these observations into 
1006: two formal assumptions on the global vector field {\bf (G1)} and {\bf(G2)}. For this, we first need to introduce some
1007: auxiliary notation.
1008: For analytical convenience, we assume that in an $O(1)$ neighborhood of the origin, $W^s(O)$ can be and has been straightened by  
1009: a smooth change of coordinates. More specifically, the nonlinear terms in (\ref{6.5}) satisfy
1010: \be\lbl{4.3a}
1011: h_{2,3} \left(x_1,0,0,\alpha\right)=0,\quad w(x_1,\alpha)=-x_1+h_1\left(x_1,0,0,\alpha\right)>0,\; 
1012: x_1\in [d_1,0),
1013: \ee
1014: for some $d_1<0$. Note that $\left(d_1, 0,0\right)$ is assumed to be sufficiently far away from the origin (see Figure \ref{f.8}).  
1015: To describe the mechanism of return, we introduce two crossections:
1016: \be\lbl{4.4}
1017:  \Sigma^+=[-c_1\delta^2, c_1\delta^2]\times D_\delta\quad\mbox{and}\quad\Sigma^-=\left\{d_1\right\}\times D_{\delta_1},
1018: \ee
1019: where $D_\delta \subset\Re^2$ denotes a disk of radius $\delta$ centered at the origin:
1020: \be\lbl{4.4a}
1021: D_\delta=\left\{ y\in\Re^2:\;\left|y\right|\le \delta\right\}.
1022: \ee
1023: Positive constants $\delta$ and $\delta_1$ are sufficiently small (see Figure \ref{f.2}a) and
1024: $c_1>0$ is chosen so that $\Sigma^+$ intersects the slow manifold $S$ transversally.
1025: In addition, we require that $\delta=o(\epsilon^{2/3})$ to guarantee that $\Sigma^+$
1026: belongs to the region of validity of the local analysis of Section 3 (see Remark 3.1c).
1027: Let $x_0\in \Sigma^+$ and  consider a trajectory of (\ref{6.5}) 
1028: starting from $x_0$.
1029: We assume that for sufficiently small $\epsilon>0$ and $\alpha=O(\epsilon^2)$,
1030: every such trajectory  intersects $\Sigma^-$ from the left. Denote the point of the first 
1031: intersection by $Q(x_0)\in\Sigma^-$.
1032: We assume that 
1033: 
1034: \noindent
1035: {\bf(G1)} the first return map $Q:\Sigma^-\to\Sigma^+$ depends smoothly on $\epsilon$ and 
1036: $\min_{x\in Q\left(\Sigma^+\right)} \left|x\right|\ge\zeta>0$.
1037: 
1038: To measure the rate of contraction toward $W^s(O)$, we consider a $2\times 2$ matrix
1039: \be\lbl{4.5}
1040: A\left(x_1,\alpha\right)=\left(\begin{array}{cc}
1041: {\partial f_2\over \partial x_2} &{\partial f_2\over \partial x_3} \\
1042:  {\partial f_3\over \partial x_2} &{\partial f_3\over \partial x_3} 
1043: \end{array}
1044: \right)_{(x_1,0,0,\alpha)}.
1045: \ee
1046: Let $\lambda_{1}(x_1,\alpha)\le \lambda_2(x_1,\alpha)$ denote the eigenvalues of the symmetric matrix 
1047: $$
1048: A^s\left(x_1,\alpha\right)={1\over 2}\left(A\left(x_1,\alpha\right)+A^T\left(x_1,\alpha\right)\right).
1049: $$
1050: Denote $\underline\lambda(x_1,\alpha)=-\lambda_1(x_1,\alpha)$ and 
1051: $\bar\lambda(x_1,\alpha)=-\lambda_2(x_1,\alpha)$. 
1052: We assume that for sufficiently small $\epsilon>0$ and $\alpha=O(\epsilon^2)$,
1053: 
1054: \noindent
1055: {\bf(G2)} 
1056: \begin{eqnarray}\lbl{4.6}
1057: {\p \over \p x_1} \bar\lambda(x_1,\alpha) <0, & x_1\in[d_1,0];&\\
1058: \lbl{4.7}
1059: \exists\; d_2 \in (d_1,0):\; \bar\lambda(x_1,\alpha)>0,& x_1\in[d_1,d_2]&\quad \& 
1060: \quad \min_{x_1\in[d_1, d_2]} {\bar\lambda(x_1,\alpha)\over w(x_1,\alpha)} = 
1061: O\left(\left|\ln\epsilon\right|\right),
1062: \end{eqnarray}
1063: 
1064: \noindent
1065: {\bf(G3)} 
1066: $$
1067: \max_{x_1\in[d_1, 0)} {\underline\lambda(x_1,\alpha)\over w(x_1,\alpha)} = 
1068:            O\left(\left|\ln\epsilon\right|\right).
1069: $$
1070: 
1071: Under these conditions, we have
1072: 
1073: \noindent
1074: {\bf Theorem 4.1}
1075: {\sc
1076: Let $\gamma>0$, conditions in (\ref{4.3a}), {\bf (G1)} and {\bf (G2)} hold. 
1077: Then for sufficiently small $\epsilon>0$ and $\alpha=O\left(\epsilon^2\right),$ a trajectory
1078: of (\ref{6.5}) with initial condition from an $O(\epsilon^2)$ neighborhood of the
1079: origin and not belonging to $W^s(O)$ undergoes multimodal oscillations.
1080: The ISIs are uniformly bounded from below
1081: \be\lbl{4.8b}
1082: \tau_i\ge \tau^-={1\over 2\alpha}\ln\left(1+{\alpha\over\gamma\epsilon^4}C^-\right),\;\; i=2,3, \dots.
1083: \ee
1084: If, in addition, {\bf (G3)} holds then the ISIs satisfy two-sided bounds
1085: \be\lbl{4.8a}
1086: \tau^-\le \tau_i \le \tau^+={1\over 2\alpha}\ln\left(1+{\alpha\over\gamma\epsilon^{4+\chi}}C^+\right),\;\; i=2,3, \dots,
1087: \ee
1088: for some $\chi>0$.
1089: Positive constants $C^\pm$ do not depend on $\epsilon,$ $\alpha,$ and $\gamma$.
1090: }
1091: \newpage
1092: 
1093: \noindent
1094: {\bf Remark 4.1}
1095: \begin{description}
1096: \item[(a)]
1097: The principal assumptions on the global vector field are formulated in {\bf (G1)} and
1098: (\ref{4.7}). The condition in (\ref{4.6}) makes the derivation of certain  estimates
1099: in the proof of the theorem easier and is used for analytical convenience. Likewise,
1100: {\bf (G3)} is not essential for the proposed mechanism. However, without this condition
1101: obtaining the two-sided estimates for the ISIs requires an additional argument in the 
1102: proof. Condition {\bf (G3)} means that the two eigenvalues of $A$ have the same order
1103: of magnitude. This condition is not restrictive.
1104: \item[(b)]
1105: For fixed $\gamma>0$, the estimates in (\ref{4.8a}) and (\ref{4.8b}) can be rewrittten as 
1106: \be\lbl{4.8c}
1107: {\ln(1+\tilde C^-\alpha)\over 2\alpha} \le \tau_i\le {\ln(1+\tilde C^+\alpha)\over 2\alpha},
1108: \ee
1109: with constants $\tilde C^\pm$ independent from $\alpha$. 
1110: Similarly, for fixed $\alpha$ and varying $\gamma>0$ 
1111: inequalities in (\ref{4.8a}) and (\ref{4.8b}) imply
1112: \be\lbl{4.8d}
1113: \bar C^- -{1\over 2\alpha}\ln\gamma \le \tau_i \le \bar C^+ -{1\over 2\alpha}\ln\gamma,
1114: \ee
1115: where constants $\bar C^\pm=O\left(\left|\ln\epsilon\right|\right)$ do not depend on $\gamma>0$.
1116: \end{description}
1117: 
1118: 
1119: \begin{figure}
1120: \begin{center}
1121: \epsfig{figure=f7.eps, height=1.8in, width=2.0in, angle=0}
1122: \end{center}
1123: \caption{ 
1124: The interpretation of the bifurcation scenario arising in (\ref{6.1})-(\ref{6.3}) for
1125: fixed positive value of $\alpha=O(\epsilon^2)$ and varying $\gamma$ near $0$ using
1126: a one-parameter family of $1D$ maps. For negative values of $\gamma=O(1)$, the map
1127: in (\ref{9.4}) has a stable fixed point (upper graph), which corresponds to a 
1128: stable limit cycle born from the supercritical AHB. For increasing values of $\gamma$
1129: the fixed point looses stability via a period-doubling bifurcation (lower graph),
1130: which initiates a period-doubling cascade leading to the formation of the chaotic attractor.
1131: }\label{f.7}
1132: \end{figure}
1133: 
1134: \subsection{The transition from subcritical to supercritical AHB}
1135: 
1136: The numerical experiments presented in Section 2 show that the transition from subcritical to
1137: supercritical AHB contains a distinct bifurcation scenario involving the formation
1138: of chaotic attractor via a period-doubling cascade. The analytical explanation of these 
1139: phenomena is outside the scope of the present paper.
1140: Below we comment on the difficulties arising in the analytical treatment of
1141: this problem. In the present section, we use a combination of the analytic and numerical techniques
1142: to elucidate the origins of the complex dynamics arising in the parameter regime near
1143: the border between regions of sub- and supercritical AHB.
1144: The principal features of this bifurcation scenario are 
1145: summarized in Figures \ref{f.4} and \ref{f.5}. In these numerical experiments, we
1146: kept $\alpha$ fixed at a small positive value and varied $\gamma$.
1147: By taking progressively smaller values of $\gamma>0$, one first observes that
1148: the multimodal patterns exhibit an increase in the ISIs (Figure \ref{f.4}a,b; see also Figure \ref{f.3}b).
1149: We consider these oscillatory patterns regular (even if they are not periodic),
1150: because the timings of the spikes remain within narrow bounds in accord with (\ref{4.8a}).
1151: For smaller values of $\gamma>0$, the oscillatory patterns become irregular and
1152: are characterized by long intervals of oscillations of small and intermediate amplitudes
1153: between successive spikes (Figure \ref{f.4} c,d).
1154: As $\gamma$ becomes negative, the trajectories lose spikes and
1155: consist of irregular oscillations 
1156: (Figure \ref{f.5} a,b). Further decrease of $\gamma$ leads the system through
1157: the reverse cascade of period-doubling bifurcations (Figure \ref{f.5} c-h).
1158: The period-doubling cascade terminates with the creation of the limit cycle,
1159: which can be followed to a nondegenerate supercritical AHB by letting $\alpha\to 0$
1160: (Figure \ref{f.5} g,h).
1161: To account for the bifurcation scenario described above, we construct the first
1162: return map. Since near the origin the trajectories spend most of the time in the vicinity of
1163: the $2D$ slow manifold, the first return map is effectively one-dimensional. The distinct
1164: unimodal structure of the $1D$ first return map affords a lucid geometric interpretation
1165: for the bifurcation scenario in the $3D$ systems of differential equations near the
1166: transition from sub- to supercritical AHB. Specifically, we show that the mechanism
1167: for generating complex dynamics in the continuous system in this parameter regime is
1168: the same as in the classical scenario of the period-doubling transition to chaos in
1169: the one-parameter families of unimodal maps \cite{GH}. 
1170: \begin{figure}
1171: \begin{center}
1172: {\bf a}\epsfig{figure=f8a.eps, height=1.8in, width=1.8in, angle=0}\quad
1173: {\bf b}\epsfig{figure=f8b.eps, height=1.8in, width=1.8in, angle=0}\\
1174: {\bf c}\epsfig{figure=f8c.eps, height=1.8in, width=1.8in, angle=0}\quad
1175: {\bf d}\epsfig{figure=f8d.eps, height=1.8in, width=1.8in, angle=0}
1176: \end{center}
1177: \caption{ The plots of first return map for $\mI$ (\ref{9.6}) computed for
1178: ({\bf a}) $\alpha= 0.0445$, $\gamma=-0.0486$,
1179: ({\bf b}) $\alpha= 0.0445$, $\gamma=-0.0126$,
1180: ({\bf c}) $\alpha= 0.0445$, $\gamma=0.0290$,
1181: ({\bf d}) $\alpha=0.12 $, $\gamma=-1.0741$.
1182: The plots in ({\bf a}-{\bf c}) illustrate the bifurcations in the family of the first return maps
1183: for the values of $\gamma$ near zero. The map shown in ({\bf d}) corresponds to the 
1184: irregular oscillations following the period-doubling cascade for increasing values
1185: of $\alpha$ in the case of supercritical AHB (see Remark 4.2). 
1186: }\label{f.8}
1187: \end{figure}
1188: 
1189: We next turn to the derivation of the first return map.
1190: We start with extending the asymptotic analysis of Section 3 to cover the case of $\gamma=o(1)$. 
1191: This requires computing additional terms on the right hand side of the reduced equation
1192: (\ref{5.24}), because already for 
1193: $\left|\gamma\right|=O(\epsilon)$, the term involving $\gamma$ on the right hand side of (\ref{5.24})
1194: is comparable with the $O(\epsilon^3)$ remainder term.
1195: We then use the more accurate approximation for the equation for $I$ to compute a
1196: $1D$ mapping:
1197: \be\lbl{9.0}
1198: G:\; I(\phi)\mapsto I\left(\phi + 2\pi\beta^{-1}\right),
1199: \ee
1200: which describes how  $I$ changes after one cycle of oscillations. 
1201: As noted above, the reduced equation (\ref{5.24}) is not suitable for analyzing
1202: the case of small $\left|\gamma\right|$ and one needs to include more terms in the expansion on the right hand side of (\ref{5.24}).  
1203: In analogy with the topological normal form for the degenerate AHB 
1204: \cite{KUZ, GL}, we expect that terms up to $O(\epsilon^5)$) are needed in (\ref{5.24}) to resolve the dynamics 
1205: for small $\left|\gamma\right|$.  This can be achieved by a straightforward albeit tedious calculation.  Below we describe the 
1206: formal procedure of obtaining the required expansions.  The justification of these expansions can be done in complete analogy 
1207: with the analysis of Section 3.  First, we compute terms on the right hand sides of (\ref{5.11a}) and (\ref{5.12}) up to $O(\epsilon^3)$ and $O(\epsilon^6)$ respectively.  Then we look for a solution of (\ref{5.11a}) in the following form:
1208: \be\lbl{9.1}
1209: \xi(\phi)=\xi_0(r,\phi)+\epsilon\xi_1(r,\phi)+\epsilon^2\xi_2(r,\phi)+O(\epsilon^3).
1210: \ee
1211: By taking an initial  condition from $D_0$ and using Anzats (\ref{9.1}) in (\ref{5.11a}), we recover $\xi_0(r,\phi)$ 
1212: (see (\ref{5.16})) and find the next 
1213: two terms in the expansion of $\xi(r,\phi)$: $\xi_1(r,\phi)$ and $\xi_2(r,\phi)$.  Thus, we obtain the approximation of the slow manifold 
1214: with the accuracy $O(\epsilon^3)$.  Next, we plug in (\ref{9.1}) into (\ref{5.12}) and collect terms multiplying equal powers of $\epsilon$
1215: to obtain 
1216: \be\lbl{9.2}
1217: \dot r = \epsilon R_1(r,\phi)+\dots+\epsilon^5 R_5(r,\phi)+O(\epsilon^6),
1218: \ee
1219: where $R_k(r,\phi)$ are $2\pi\beta^{-1}$– periodic functions of the second argument.
1220: Finally, by rewriting (\ref{9.2}) in terms of $I$ (see (3.25)), integrating it over one period of oscillations, 
1221: $\omega$, and disregarding $O(\epsilon^5)$ terms, we obtain a map for the change in $I$ after one cycle of oscillations
1222: \be\lbl{9.4}
1223: I_{n+1}=G(I_n),\;\; G(I)\equiv I-2\epsilon^2\omega\left(\mu I +\gamma +{\epsilon^2c\over I}\right),\; \omega=2\pi\beta^{-1}.
1224: \ee
1225: Except for the last term in the definition of $G$, the map in (\ref{9.4}) follows from the reduced system (\ref{5.24}) and (\ref{5.25}).  
1226: The last term is only needed if the value of $\left|\gamma\right|$ does not exceed $O(\epsilon)$.  Our calculations show that 
1227: the second Lyapunov coefficient, $c$, is negative for the values of parameters used in (\ref{6.1})-(\ref{6.3}).
1228: Therefore, for sufficiently small $\epsilon>0$, (\ref{9.4}) defines a unimodal map.  Away from an $O(\epsilon^2)$ neighborhood of 
1229: $0$, the graph of  $\tilde I=G(I)$ is almost linear with a weakly attracting slope $1-2\omega\mu\epsilon^2$ for $\mu>0$ (Figure \ref{f.7}).
1230: For $\mu>0$, $G$ has a unique fixed point:
1231: \be\lbl{9.5}
1232: \bar I=\left\{
1233: \begin{array}{cc}
1234: {-\gamma\over\mu} +{c\over\gamma}\epsilon^2+O(\epsilon^3),& \gamma<0,\\
1235: \sqrt{-c\over\mu}\epsilon +O(\epsilon^2),&\gamma=0,\\
1236: {-c\over\gamma}\epsilon^2+O(\epsilon^3),& \gamma>0.
1237: \end{array}
1238: \right.
1239: \ee
1240: For negative values of $\gamma=O(1)$, $I=\bar I$ is a stable fixed point of $G$, which corresponds to a stable limit cycle  born
1241: from a  supercritical AHB (see Figure \ref{f.5} g,h). It follows from (\ref{9.4}) that for increasing values of $\gamma$, 
1242: the graph of the map moves down. Consequently, 
1243: the fixed point moves to the left and eventually looses stability via a period-doubling bifurcation (Figure \ref{f.5} e,f).  
1244: Further increase in $\gamma$, yields more period-doubling bifurcations (Figure \ref{f.5} c,d).  Given the unimodal character of 
1245: $G$, one expects that this sequence of period-doubling bifurcations eventually leads to the formation of a chaotic attractor 
1246: (Figure \ref{f.5} a,b). The unimodal character of the map combined with the manner of its
1247: dependence on $\gamma$ suggest a clear geometric mechanism for the formation of the chaotic attractor and the subsequent
1248: period-doubling cascade arising near the transition from sub- to supercritical AHB (Figure \ref{f.5}a-h). 
1249: We now outline the  limitations of the asymptotic analysis of this section and difficulties
1250: arising in the justification of (\ref{9.4}).
1251: According to (\ref{9.5}) already at the moment of the first period-doubling bifurcation, fixed point $\bar I$
1252: belongs to an $O(\epsilon^2)$ neighborhood of $0$. In this neighborhood,
1253: $\rho=O(1)$ (see (\ref{5.22}))  and, therefore, $\bar I$ lies outside of the region of validity of
1254: the asymptotic analysis. In this region, (\ref{9.4}) may only be considered as a formal 
1255: asymptotic expression. A rigorous justification of the from of the map in the 
1256: boundary layer meets substantial analytical difficulties: it requires introducing an additional set 
1257: of intermediate asymptotic expansions and matching them with those obtained in Section 3. 
1258: We do not address this problem in the present work. 
1259: Below, we resort to using numerical techniques to verify the principal features
1260: of the first return map suggested by the asymptotic analysis: the unimodality of $G$ and its
1261: dependence on $\gamma$. The numerics confirmed our predictions about the form
1262: of the first return map and it also revealed certain additional features.
1263: 
1264: For numerical construction of the first-return map, we fix angle $\bar \theta \in [0, 2\pi)$
1265: and define a crossection in cylindrical coordinates:
1266: $
1267: \Sigma=\{(\rho, \xi, \theta): \theta=\bar \theta\}.
1268: $
1269: Let $(\rho(t), \xi(t), \theta(t))$ be a trajectory of (\ref{6.5}) and $0<t_1<t_2<t_3<\dots<t_k<\dots$
1270: denote a sequence of times, at which $(\rho(t_k),\xi(t_k),\theta(t_k))\in \Sigma,\;k=1,2,3,\dots$.
1271: Then, we define 
1272: \be\lbl{9.6}
1273: \mG: \mI(t_k) \mapsto \mI(t_{k+1})\quad\mbox{and}\quad \mI(t) =\frac{1}{\rho^2(t)}.
1274: \ee
1275: By a suitable choice of  $\bar\theta$, we can achieve that a trajectory of
1276: (\ref{6.5}), intersects $\Sigma$ once during one cycle of oscillations. Therefore,
1277: for a trajectory lying in a $2D$ slow manifold, (\ref{9.6}) defines a 
1278: $1D$ first-return map. Note that maps (\ref{9.4}) and (\ref{9.6}) are 
1279: related via rescaling of the coefficients, since to leading order $ \mI=\epsilon^{-2}I$.
1280: Therefore, the graphs of $G(I)$ and $\mG(\mI)$ are
1281: similar in the domain, where (\ref{9.4}) is valid. We computed the first return maps
1282: for fixed $\alpha=0.0445$ and for several values of $\gamma$. 
1283: The representative plots are shown in Figure \ref{f.8} a-c.
1284: The numerically computed maps confirmed  our expectations about the graph of
1285: $\mG$ in the boundary layer near $0$: in this region, the map is decreasing
1286: with a strongly expanding slope. 
1287: Away from $0$, the graph of $\mG$ contains an almost linear branch,
1288: whose slope is very close to $1-2\alpha\omega,$ as predicted by (\ref{9.4})
1289: (see lower branches in Figure \ref{f.8}b,c).
1290: These numerics also confirm that under the variation of $\gamma$, the graph of the first return
1291: map is translated in vertical direction (Figure \ref{f.8} a-c). Therefore, the family of the first return maps 
1292: possesses two principal ingredients (the unimodality and the additive dependence on $\gamma$),
1293: which are necessary for the qualitative explanation of the bifurcation scenario 
1294: given in Figure \ref{f.7}. In addition, our numerical experiments reveal
1295: a new feature of the first return map: for small $|\gamma|$, the graph of the map has 
1296: another almost linear branch in the outer region away from the origin (see Figure \ref{f.8}b).
1297: This branch of the graph of the map also has an attracting slope.
1298: The presence of this branch in the first return map indicates the existence of another (branch of the) 
1299: slow manifold different from that described in Section 3. 
1300: A possible explanation for the appearance
1301: of the upper branch in the first return map is due to the unstable manifold of the periodic orbit.
1302: For small $|\gamma|$, the first-return map (\ref{9.6}) is multivalued 
1303: in the outer region. However, since both the upper and the lower branches have positive slopes less than $1$, 
1304: the qualitative dynamics for the map shown in Figure \ref{f.8}b does not depend on the exact mechanism for
1305: the selection between the branches in (\ref{9.6}). Although explaining the multivaluedness of 
1306: the first return maps shown in Figures \ref{f.8}b,c presents an interesting problem, it is not 
1307: critical for the qualitative explanation of the bifurcation scenario arising during the transition from 
1308: the subcritical to supercritical AHB. The latter was the main goal of the present subsection.
1309: 
1310: \noindent
1311: {\bf Remark 4.2 }
1312: In addition, to the region in the parameter space containing the border between the regions of
1313: sub- and supercritical AHB, there is another parameter regime resulting in
1314: the complex dynamics.
1315: In \cite{FKRT}, it was shown numerically that the limit cycle born from the supercritical
1316: AHB in (\ref{6.5}) undergoes a period-doubling cascade leading to the formation of the chaotic attractor
1317: for increasing values of $\alpha$. The first period-doubling bifurcation in this cascade is shown in Figure \ref{f.1}c,d.
1318: This bifurcation scenario is consistent with the form of the first-return map constructed 
1319: in this subsection. Indeed, it is easy to see from (\ref{9.5}) that for increasing values of $\mu$ the fixed
1320: point, $\bar I$, moves to the left. Therefore, the explanation given above for the period-doubling cascade
1321: resulting from the variation of $\gamma$ near $0$ also applies to the case of increasing $\alpha$ and
1322: fixed $\gamma <0$ (see Figure \ref{f.8}d).
1323: 
1324: 
1325: \section{The proof of Theorem 4.1}
1326: \setcounter{equation}{0}
1327: 
1328: In the present section, we show that under the assumptions of Theorem 4.1
1329: the trajectories of (\ref{2.5}) with initial conditions in $\Sigma^-$ enter $D_0$.
1330: The local analysis in Section 3 describes the behavior of trajectories from
1331: the moment they reach $D_0$ and until they leave $D$. In particular,
1332: it shows that the dynamics near the origin has two phases: the fast approach to 
1333: the slow manifold and the slow drift away from the origin along the slow manifold.
1334: Upon leaving $D$, the trajectories are reinjected back to $\Sigma^-$ by the  
1335: return mechanism postulated in {\bf (G1)}. This scenario implies that the 
1336: system undergoes multimodal oscillations as stated in Theorem 4.1.
1337: \begin{figure}
1338: \begin{center}
1339: \epsfig{figure=f9.eps, height=3.0in, width=5.0in, angle=0}
1340: \end{center}
1341: \caption{ A schematic representation of the trajectory of (\ref{6.1})-(\ref{6.3})  approaching
1342: a weakly unstable saddle-focus. The trajectory crossing $\Sigma^-$ is subject to strong vector
1343: field transverse to $W^s(O)$ in $\Pi^1_{\delta_1}$. This guarantees that it remains close to $W^s(O)$
1344: until it enters an $O(\epsilon^2)$ neighborhood of the origin. From that moment and until the trajectory
1345: reaches $\Sigma^+$ its behavior is described by the local analysis of Section 3.
1346: }
1347: \label{f.9}
1348: \end{figure}
1349: 
1350: We start with presenting several auxiliary estimates, which will be needed for the
1351: proof. By (\ref{4.3a}), one can choose $C_1>0$ such that
1352: \be\lbl{7.1}
1353: w(x_1, \alpha)\ge C_1 \left|x_1\right|,\quad x_1\in\left[ d_1,0\right]
1354: \ee 
1355: and for sufficiently small $\alpha\ge 0$.
1356: Next, we note that $\bar\lambda(0,0)=0$. This equation and (\ref{4.6}), by  the Implicit
1357: Function Theorem, imply that there exists $0>x^*(\alpha)=O(\epsilon^2)$ such that
1358: $$
1359: \bar\lambda(x_1,\alpha)\left\{\begin{array}{ll}
1360: <0,& x_1\in[d_1, x_1^*(\alpha)),\\
1361: >0,& x_1\in(x_1^*(\alpha),0].
1362: \end{array}
1363: \right.
1364: $$
1365: 
1366: Using (\ref{4.6}) and (\ref{7.1}), we have 
1367: $$
1368: {\bar\lambda(x_1,\alpha)\over w(x_1,\alpha)}\ge {\bar\lambda(x_1^*,\alpha)\over w(x_1,\alpha)}=0,\quad x_1\in [d_1, x^*_1(\alpha)],
1369: $$
1370: where $w(x_1,\alpha)$ is defined in (\ref{4.3a}).
1371: 
1372: Let $\bar\mu>0$ be such that ({\bf G1}) and ({\bf G2}) hold for $\alpha\in [0,\bar\mu\epsilon^2]$. 
1373: In the remainder of this section, unless stated otherwise,  it is assumed that $\alpha\in 
1374: \left[0,\bar\mu\epsilon^2 \right]$. To simplify notation, we will often omit the dependence 
1375: of various functions on $\alpha$.
1376: Using (\ref{4.3a}), we rewrite (\ref{6.5}) in the following form:
1377: \begin{eqnarray}\lbl{7.2}
1378: \dot x_1 &=& w(x_1)+y_1\phi_1(x_1,y)+y_2\phi_2(x_1,y),\\ 
1379: \lbl{7.3}
1380: \dot y &=& A(x_1)y+\psi(x_1,y),
1381: \end{eqnarray}
1382: where $w(x_1)$ and $A(x_1)$ are defined in (\ref{4.3a}) and (\ref{4.5}) respectively;
1383: $y=(y_1,y_2):=(x_2,x_3)$, and 
1384: \be\lbl{7.4}
1385: w(0)=0,\; \psi(x_1,y)=O\left(\left|y\right|^2\right),\; \phi_{1,2}(x_1,y)=\phi_{1,2}(x_1)+O(\left|y|\right),\quad
1386: x\in \left[d_1, 0\right].
1387: \ee
1388: Let 
1389: \be\lbl{7.4a}
1390: d_3=-M^2\epsilon^2
1391: \ee
1392: denote the $x_1-$coordinate on the left lateral boundary of $D_0$ (see (\ref{5.2})) and $\Pi_{\delta_1}=[d_1,d_3]\times D_{\delta_1}$
1393: (see Figure \ref{f.9}).
1394: From (\ref{6.5}) and (\ref{4.3a}), one can see that  for ${\delta_1}>0$ sufficiently small (independent of $\epsilon$),
1395: the right hand side of (\ref{7.2})
1396: $$
1397: w(x_1)+y_1\phi_1(x_1,y)+y_2\phi_2(x_1,y)>0,\quad (x_1,y)\in\Pi_{\delta_1}.
1398: $$
1399: Therefore, in $\Pi_{\delta_1}$, (\ref{7.2}) and
1400: (\ref{7.3}) may be rewritten as follows
1401: \be\lbl{7.5}
1402: {d y(x_1)\over d x_1}=\tilde A(x_1) y +\tilde\psi(x_1,y),
1403: \ee
1404: where 
1405: $$
1406: \tilde A(x_1)={1\over w(x_1)} A(x_1)\quad\mbox{ and}\quad \tilde\psi(x_1,y)={\psi(x_1,y)\over w(x_1,y)}\left(1+O\left(\left|y\right|\right)\right).
1407: $$
1408: By (\ref{7.1}) and (\ref{7.4}), we have
1409: \be\lbl{7.6}
1410: \left|\tilde\psi(x_1,y)\right|\le {C_2\left|y\right|^2\over \left|x_1\right|},\quad (x_1,y)\in\Pi_{\delta_1},
1411: \ee
1412: for some $C_2>0$ independent of $\epsilon$.
1413: 
1414: To follow the trajectories from $\Sigma^-$ to $D_0$, we introduce two regions:
1415: $$
1416: \Pi_{\delta_1}^1=[d_1,d_2]\times D_{\delta_1}\quad\mbox{and}\quad\Pi_{\delta_2}^2=[d_2,d_3]\times D_{\delta_2}\;\;\mbox{(see Figure \ref{f.9})}.
1417: $$
1418: Recall that $D_{\delta_{1,2}}$ denotes the disk of radius $\delta_{1,2}$ (see (\ref{4.4a})).
1419: Positive constant $\delta_1$ is the same as in (\ref{4.4}) and $0<\delta_2\ll \delta_1$ will be specified later (Figure \ref{f.9}). 
1420: Recall that $d_3=-M\epsilon^2$ denotes the value of the $x_1-$coordinate on the the left lateral boundary of $D_0$ (see (\ref{5.2})).
1421: By taking $M>0$ large enough, we can arrange  $d_3<x^*(\alpha)<0$ for $\alpha\in \left[0,\bar\mu\epsilon^2 \right]$.
1422: By {\bf(G1)}, the vector field in $\Pi^1_{\delta_1}$ is sufficiently strong so that the trajectories entering $\Pi_{\delta_1}^1$
1423: through $\Sigma^-$ get into a narrow domain $\Pi_{\delta_2}^2$ and remain there until they reach $D_0$.
1424: The following lemma allows to control the trajectories in $\Pi_{\delta_1}=[d_1,d_3]\times D_{\delta_1}$.
1425: 
1426: \noindent{\bf Lemma 5.1}
1427: {\sc Let $[\ell_1,\ell_2]\subseteq[d_1,d_3],\; 0<\bar\delta\le {\delta_1},$ and
1428: \be\lbl{7.7}
1429: -\upsilon=\max_{x_1\in[\ell_1,\ell_2]}\sup_{\left|y\right|=1}\left(\tilde A(x_1)y,y \right)<0.
1430: \ee
1431: Then
1432: \be\lbl{7.8}
1433: \left|y(\ell_1)\right|\le\bar\delta\quad\Rightarrow\quad \left|y(x_1)\right|\le 2\bar\delta e^{-\upsilon(x_1-\ell_1)},\; x_1\in[\ell_1,\ell_2],
1434: \ee
1435: provided
1436: \be\lbl{7.9}
1437: {4 C_2\bar\delta \left(1- e^{-\upsilon(\ell_2-\ell_1})\right)\over\left|\ell_2\right|\upsilon}\le 1.
1438: \ee
1439: }
1440: \noindent
1441: {\bf Proof:}
1442: We may assume that $|y(x_1)|\not=0,x \in [\ell_1,\ell_2]$, since, otherwise, 
1443: by the uniqueness of solution of the initial value problem for (\ref{7.5}), $y(x_1)=0,$ $x_1\in[\ell_1,\ell_2]$,
1444: and (\ref{7.8}) holds. From (\ref{7.5}), (\ref{7.6}), and (\ref{7.7}), we have
1445: \be\lbl{7.10} 
1446: \frac{d|y(x_1)|}{dx_1} \le -\upsilon \left|y(x_1)\right| + \bar\psi\left(x_1,\left|y\right|\right),\;\;
1447: \bar\psi\left(x_1,\left|y\right|\right)={ C_2 \left|y \right|^2 \over \left|x_1 \right|},\;\; x_1\in[\ell_1,\ell_2],\;
1448: \left|y\right|\le\left|\delta_1\right|.
1449: \ee
1450: where $\upsilon$ is defined in (\ref{7.7}), and $\left|y\right|$ denotes the Euclidean norm of $y\in\Re^2$.
1451: Let $\bar y(x_1)$ denote the solution of the initial value problem 
1452: \be\lbl{7.11}
1453: \bar y^\prime=-\upsilon \bar y +\bar\psi(x_1,\bar y),\;
1454: \bar y(\ell_1)=\bar\delta.
1455: \ee
1456: It is sufficient to show that (\ref{7.8}) holds for
1457: $\bar y(x_1)$. We represent $\bar y(x_1)$ as the limit of the sequence
1458: of successive approximations
1459: \be\lbl{7.12}
1460: \bar y^{(n+1)}(x_1)=\bar\delta e^{-\upsilon(x_1-\ell_1)}+\int_{\ell_1}^{x_1} e^{-\upsilon(x_1-s)}
1461: \bar\psi\left(s,\bar y^{(n)}(s)\right) ds,\quad n=0,1,2,\dots
1462: \ee
1463: and $\bar y^{(0)}(x_1)=\bar \delta e^{-\upsilon(x_1-\ell_1)}.$
1464: We use induction to show 
1465: \be\lbl{7.13}
1466: \left|\bar y^{(n)}(x_1)\right| \le 2 \bar\delta e^{-\upsilon(x_1-\ell_1)},\quad x \in [\ell_1,\ell_2],\quad
1467: n=0,1,2,\dots .
1468: \ee 
1469: Inequality (\ref{7.13}) holds for $n=0$. We show that $(\ref{7.13})_{n=k} \Rightarrow (\ref{7.13})_{n=k+1}$.
1470: Using the definition of $\bar\psi$ in (\ref{7.10}) and $(\ref{7.13})_{n=k}$, we have
1471: \begin{eqnarray}\nonumber
1472: \int_{\ell_1}^{x_1} e^{-\upsilon(x_1-s)}\bar\psi(s,\bar y^{(k)}(s) ds &\le&
1473: \frac{C_2}{|\ell_2|}\int_{\ell_1}^{x_1} e^{-\upsilon(x_1-s)}\left|\bar y^{(k)}(s)\right|^2 ds\le
1474: \frac{C_2}{|\ell_2|}\int_{\ell_1}^{x_1}e^{-\upsilon (x_1-s)} 4 {\bar\delta}^2 e^{-2\upsilon(s-\ell_1)} ds\\ 
1475: \lbl{7.14}
1476: &=&\bar\delta e^{-\upsilon(x_1-\ell_1)} {4 C_2\bar\delta \left( 1- e^{-\upsilon(\ell_2-\ell_1)}\right)\over\left|\ell_2\right|\upsilon}.
1477: \end{eqnarray}
1478: Using (\ref{7.14}) and (\ref{7.9}),
1479: from $(\ref{7.13})_{n=k}$ we obtain  $(\ref{7.13})_{n=k+1}$. By induction, (\ref{7.13}) holds. By taking $n\to\infty$ 
1480: in (\ref{7.13}), we have 
1481: \be\lbl{7.15}
1482: \left|\bar y(x_1)\right|\le 2\bar\delta e^{-\upsilon(x_1-\ell_1)},\; x_1\in[\ell_1,\ell_2].
1483: \ee
1484: The statement in (\ref{7.8}) then follows from (\ref{7.15}) and the standard theory for differential inequalities
1485: \cite{hartman}.\\
1486: $\Box$
1487: 
1488: In the following lemma, we determine the size of $\Pi^2_{\delta_2}$.
1489: 
1490: \noindent{\bf Lemma 5.2}
1491: {\sc There exists positive $\delta_2=O(\epsilon^2)$, such that the trajectories of (\ref{7.2}) and (\ref{7.3}) 
1492: entering $\Pi^2_{\delta_2}$ from $\Pi^1_{{\delta_1}}$ remain in $\Pi^2_{\delta_2}$ until they reach $D_0$ (see Figure \ref{f.9}.)
1493: }
1494: 
1495: \noindent {\bf Proof:} 
1496: Denote
1497: $$
1498: -\upsilon_2=\max_{x_1\in[d_2,d_3]}\sup_{\left|y\right|=1}\left(\tilde A(x_1)y, y\right).
1499: $$
1500: Recall 
1501: $$
1502: \bar\lambda (x^*)=0 \quad \mbox{and}\quad O(\epsilon^2)= d_3< x^*=O(\epsilon^2).
1503: $$
1504: Therefore, by (\ref{4.6}), $\bar\lambda(d_3)=O(\epsilon^2)$ is positive.
1505: In addition,
1506: \begin{eqnarray*}
1507: \upsilon_2&=&-\max_{x_1\in[d_2,d_3]} \sup_{\left|y\right|=1}\left(\tilde A(x_1)y, y\right)=
1508: \min_{x_1\in[d_2,d_3]}{-\sup_{\left|y\right|=1}\left( A(x_1)y, y\right)\over w(x_1)}\\
1509: &=& {\min_{x_1\in[d_2,d_3]} \bar\lambda(x_1)\over\max_{x_1\in[d_2,d_3]} w(x_1)}= 
1510: {\bar\lambda(d_3)\over\max_{x_1\in[d_2,d_3]} w(x_1)}= O(\epsilon^2). 
1511: \end{eqnarray*}
1512: and $\upsilon_2>0$. 
1513: Next we apply Lemma 5.1 with $\bar\delta:=\delta_2,$ $\upsilon:=\upsilon_2$, and $\ell_{1,2}:=d_{2,3}$. Note that for small $\upsilon_2$
1514: the inequality (\ref{7.9}) can be rewritten as 
1515: \be\lbl{7.16}
1516: {4 C_2\delta_2 \left(d_3-d_2 +O(\upsilon_2^2)\right)\over\left|d_3\right|}\le 1.
1517: \ee
1518: Since $d_3=O(\epsilon^2)$  (see (\ref{7.4a})) and $\upsilon_2=O(\epsilon^2)$, one can choose $\delta_2=O(\epsilon^2)$ so 
1519: that (\ref{7.16}) holds.\\
1520: $\Box$
1521: 
1522: Having found the size of $\Pi^2_{\delta_2}$, we now determine the rate of contraction in $\Pi^1_{\delta_1}$ sufficient
1523: to funnel the trajectories entering $\Pi^1_{\delta_1}$ through $\Sigma^-$ to $\Pi^2_{\delta_2}$.
1524: 
1525: \noindent{\bf Lemma 5.3}
1526: {\sc The trajectories of (\ref{7.2}) and (\ref{7.3}) 
1527: entering $\Pi^1_{\delta_1}$ through $\Sigma^-$ remain in $\Pi^1_{\delta_1}$ until they reach 
1528: $\Pi^2_{\delta_2}$.
1529: }
1530: 
1531: \noindent {\bf Proof:} 
1532: Denote
1533: $$
1534: -\upsilon_1=\max_{x_1\in[d_1,d_2]}\sup_{\left|y\right|=1}\left(\tilde A(x_1)y, y\right).
1535: $$
1536: From {\bf (G2)}, one finds that $\upsilon_1=O(\left|\ln \epsilon\right|)$ is positive.
1537: Lemma 5.3 now follows from Lemma 5.1 with  $\bar\delta:={\delta_1},$ $\upsilon:=\upsilon_1$, and $\ell_{1,2}:=d_{1,2}$.  
1538: Indeed, for $\upsilon_1=O\left(\left|\ln\epsilon\right|\right)$, we have 
1539: \be\lbl{7.17}
1540: \upsilon_1\ge {4C_5{\delta_1}\over\left|d_2\right|}.
1541: \ee
1542: Inequality (\ref{7.17}) is sufficient for (\ref{7.9}) to hold. By Lemma 5.1, we have
1543: \be\lbl{7.19}
1544: \left|y(d_2)\right|\le 2{\delta_1} e^{-\upsilon_1(d_2-d_1)}.
1545: \ee
1546: With $\upsilon_1=O\left(\left|\ln\epsilon\right|\right)$, by (\ref{7.19}), we can achieve 
1547: $\left|y(d_2)\right|<\delta_2=O(\epsilon^2)$.\\
1548: $\Box$
1549: 
1550: Lemmas 5.2 and 5.3 imply that the trajectories entering $\Pi^1_{\delta_1}$ through $\Sigma^-$
1551: stay in $\Pi^1_{\delta_1}\bigcup\Pi^2_{\delta_2}$ until they reach $D_0$. Moreover, the
1552: inequality in {\bf (G1)} guarantees that such trajectories are bounded away from $W^s(O)$.
1553: Thus, we can use the analysis of Section 3 to describe the evolution of trajectories 
1554: from the moment they reach $D_0$ until they leave $D$. By Remark 3.1c, this description extends
1555: to any region where $x_{2,3}=o\left(\epsilon^{2/3}\right),$ i.e., the trajectories can be controlled 
1556: until they hit $\Sigma^+$. This is followed by the return to $\Pi_{\delta_1}^1$ according to {\bf (G1)},
1557: and the next cycle of the multimodal oscillations begins. The analysis of this section applies to 
1558: any trajectory starting from $D_0$ and not belonging to $W^s(O)$.
1559: 
1560: It remains to estimate the ISIs. For this, we compute the time needed for the trajectory starting
1561: in an $O(\epsilon^2)$ neighborhood to return back to this neighborhood after making one global 
1562: excursion. Since the time of flight of the trajectory outside a small neighborhood of the 
1563: origin depends regularly on the control parameters, the duration of the very long ISIs is determined
1564: by the time spent in the neighborhood of the origin. To estimate the latter, 
1565: we note that at the moment a trajectory enters $D_0$, we have  
1566: \be\lbl{7.20a}
1567: \left|y(d_3)\right|\le C_3 \epsilon^2.
1568: \ee
1569: This follows from Lemma 5.3, since $\left|y(d_3)\right|\le \delta_2$ and $\delta_2=O\left(\epsilon^2\right).$
1570: To obtain the lower bound on $\left|y(d_3)\right|$, recall that by {\bf (G1)}, we have 
1571: \be\lbl{7.20b}
1572: \left|y(d_1)\right|\ge \zeta>0.
1573: \ee
1574: As follows from {\bf (G3)}, the maximal rate of contraction in $\Pi_{\delta_1}$ does not exceed 
1575: $O\left(\left|\ln\epsilon\right|\right)$ in absolute value. This combined with (\ref{7.20b}) implies that
1576: \be\lbl{7.20c}
1577: \left|y(d_3)\right|\ge C_4\epsilon^{2+{\chi\over 2}}
1578: \ee
1579: for some $\chi\ge 0$ and $C_4$ independent of $\epsilon$. We omit the proof of (\ref{7.20c}), because it is 
1580: completely analogous to that of Lemma 5.1. 
1581: 
1582: Let $t=t^-$ denote the moment 
1583: of time when a trajectory 
1584: of (\ref{6.5}) enters $D_0$ from $\Pi^2_{\delta_2}$. After switching back to the original parametrization 
1585: of $y$ by time, $t$, we rewrite (\ref{7.20a}) and (\ref{7.20c}):
1586: \be\lbl{7.20}
1587: C_3\epsilon^{2+{\chi \over 2}}\le\left|y(t^-)\right|\le C_4\epsilon^2.
1588: \ee
1589: For $t>t^-$ the trajectory approaches and remains close to the slow manifold as long as 
1590: $\left|y\right|=o(\epsilon^{2/3})$
1591: (see Remark 3.1c). Let ${2\over3}<j<1$ and denote 
1592: \be\lbl{7.21} 
1593: t^+=\min\left\{t>t^-:\;\left|y(t)\right|=\epsilon^j\right\}.
1594: \ee
1595: From (\ref{7.20}) and (\ref{7.21}), we have  the following bounds for $I_0:=I(t^-)$ 
1596: and $I_1:=I(t^+)$:
1597: \be\lbl{7.22}
1598: C_6\epsilon^{-2-\chi}\le I_0\le C_7\epsilon^{-2}
1599: \quad\mbox{and}\quad
1600: I_1=O(\epsilon^{2(1-j)}),\; j\in\left({2\over 3}, 1\right).
1601: \ee
1602: For these ranges of values of $I_0$ and $I_1$, from (\ref{5.55}) we have 
1603: \be\lbl{7.23}
1604: \tau_{in}=t^+-t^-={1\over 2\mu\epsilon^2}\ln\left(\left(1+{\mu\over\gamma}I_0\right)
1605: \left(1+{\mu\over\gamma}o(1)\right)\right).
1606: \ee
1607: The combination of (\ref{7.22}) and (\ref{7.23}) yields two-sided bounds for $\tau_{in}$:
1608: \be\lbl{7.24}
1609: \tau^-_{in}\le \tau_{in}\le \tau_{in}^+,
1610: \ee
1611: where
1612: $$
1613: \tau_{in}^- ={1\over 2\mu\epsilon^2}
1614: \ln\left(1+{\mu\over\gamma\epsilon^2}C_8^-\right)\quad\mbox{and}\quad
1615: \tau_{in}^+ ={1\over 2\mu\epsilon^2}
1616: \ln\left(1+{\mu\over\gamma\epsilon^{2+\chi}}C_8^+\right),
1617: $$
1618: where positive constants $C^\mp_8$ can be chosen independent of $\epsilon$.
1619: Inequalities (\ref{7.24}) provide bounds for the time that a multimodal trajectory spends
1620: near the unstable equilibrium. On the other hand, the time of flight outside a small neighborhood 
1621: of the origin, $\tau_{out}$, depends regularly on $\epsilon$, by {\bf (G2)}.
1622: Therefore, by adjusting constants $C_8^\mp$ in (\ref{7.24}) if necessary, one can obtain 
1623: uniform bounds on the ISIs $\tau=\tau_{in}+\tau_{out}$ for sufficiently small $\epsilon>0$,
1624: positive $\gamma$ and $\mu$ from bounded intervals, as stated in Theorem 4.1.
1625: 
1626: \section{Discussion}
1627: \setcounter{equation}{0}
1628: 
1629: In the present paper, we investigated a mechanism for generation of multimodal oscillations
1630: in a class of systems of differential equations close to an AHB. 
1631: Our analysis covers both cases of sub- and supercritical AHB. 
1632: For the supercritical case, we identified a novel geometric feature of the bifurcating limit cycle, 
1633: the frequency doubling effect.
1634: It turns out that generically in the normal system of coordinates the oscillations 
1635: in one of the variables are twice faster than in the remaining two variables.
1636: Therefore, the leading order approximation of the limit cycle bifurcating from the supercritical AHB
1637: requires two harmonics.
1638: The asymptotic analysis of the present paper explains the frequency-doubling. In addition, we provide
1639: a complementary geometric interpretation to this counterintuitive effect. In particular,
1640: we showed that it is a consequence of the geometry of the limit cycle. 
1641: The latter is not captured by the topological normal form of the AHB.
1642: The analysis of the multimodal oscillations arising from the subcritical AHB requires
1643: additional assumptions on the global behavior of trajectories. 
1644: We identified two principal properties of the global vector field:
1645: the mechanism of return and the strong contraction property.
1646: In the presence of this global structure, the subcritical AHB produces
1647: sustained multimodal oscillations combining the small amplitude oscillations near
1648: the unstable equilibrium with large amplitude spikes. 
1649: The resultant motion is recurrent in a weak sense:
1650: it may not be periodic but nevertheless the timings of the spikes possess certain
1651: regularity. We have shown that the ISIs have well-defined asymptotics near the AHB
1652: and comply to the two-sided bounds, which depend on the principal bifurcation
1653: parameters. Our estimates show that near the AHB, the ISIs can be extremely long
1654: and can change greatly under relatively small variation of the bifurcation parameters.
1655: The ability of the system to exhibit such extreme variability in the ISI duration is
1656: important in many applications, in particular, in the context of neuronal dynamics.
1657: Previous studies investigated different possible mechanisms for generating multimodal
1658: patterns with very long ISIs \cite{BER,DK05,DRSE,MC,RE89}.
1659: For the finite dimensional approximation of the model
1660: of solid-fuel combustion (\ref{6.1})-(\ref{6.3}), the main motivating example for our work, 
1661: the proximity to the homoclinic bifurcation was suggested in \cite{FKRT} as a possible mechanism
1662: for prolonged ISIs. 
1663: Our conclusions confirm the importance of the proximity to the homoclinic bifurcation for explaining
1664: the oscillatory patterns in (\ref{6.1})-(\ref{6.3}). The proximity to the homoclinic bifurcation
1665: is implicitly reflected in our assumptions
1666: on the global vector field.
1667: However, we emphasize the critical role of the AHB:
1668: the duration of the ISIs can be effectively controlled by the parameters associated with the AHB
1669: without changing the distance of the system to the homoclinic bifurcation.
1670: In all our numerical experiments, the system remained bounded away from the homoclinic bifurcation,
1671: nevertheless it exhibited patterns with very long ISIs whose duration was amenable to control.
1672: Our analysis extends the estimate for the ISIs obtained in \cite{GW} to a wide class of problems. 
1673: It also emphasizes the proximity of the system to the border between 
1674: sub- and supercritical AHB, as another factor in creating oscillatory patterns with long ISIs.
1675: We show that as this border is approached from the subcritical side
1676: the ISI grow logarithmically. This observation is important for explaining the oscillatory patterns
1677: generated by (\ref{6.1})-(\ref{6.3}) since in this model the AHB can change its type under the variation
1678: of the second control parameter $p$. This situation is not specific to the model of 
1679: solid fuel combustion. Recent studies suggest 
1680: that there is a class of neuronal models close to the AHB whose type may change with the values of parameters 
1681: \cite{DK05, OP}.
1682: Therefore, it is important to understand the principal features of the transition from sub- to 
1683: supercritical AHB. We found that when the border between the regions of sub- and supercritical
1684: bifurcation is approached from the subcritical side (while the distance from the AHB remains fixed), 
1685: the oscillations become chaotic. The regime of irregular oscillations is then followed by the 
1686: reverse period doubling cascade. To understand the nature of this bifurcation scenario 
1687: we used a combination of analytic and numerical techniques.
1688: Using the insights gained from the asymptotic analysis, we constructed a $1D$ first-return map.
1689: The map provides a clear geometric interpretation for the bifurcation scenario near the 
1690: transition from sub- to supercritical AHB. Our study suggests that the formation of the chaotic attractor
1691: via a period-doubling cascade is a universal feature of this transition. For example, the bifurcation
1692: scenarios reported for the Hodgkin-Huxley model in \cite{DK05} are very similar to those studied in the
1693: present paper and are likely to share the same mechanism.
1694: 
1695: Mixed-mode oscillations similar to those studied in the present paper, have been studied in for a class
1696: of the slow-fast systems in $\Re^3$ near the AHB. Although, the work toward developing a complete mathematical 
1697: theory for such oscillations is still in progress, the general mechanism for their generation and the bifurcation
1698: structure of the problem have been greatly elucidated  recently \cite{KOP, MC, MSLG, WESCH}.
1699: The present paper shows the relation between the mechanisms for the mixed-mode oscillations in the model
1700: in \cite{FKRT} and for those in the slow-fast systems. The latter possess a well-defined structure of the
1701: global vector field due to the presence of the disparate timescales in the governing equations \cite{JON}. The analyses
1702: of the mixed-mode oscillations in \cite{KOP, MC, MSLG, WESCH} use in an essential way the relaxation structure 
1703: of the problem. The model in \cite{FKRT} is an example of the mixed-mode generating system, which does not 
1704: possess an explicit relaxation structure. In fact, it is hard to expect such structure in a system obtained via
1705: projecting an infinite-dimensional system onto a finite-dimensional subspace. In formulating the assumptions on the
1706: global vector-field ({\bf G1}) and ({\bf G2}), we were looking for the minimal requirements on the system near
1707: an AHB that guarantee the existence of the mixed-mode solutions. Due to the lack of the information about the
1708: global vector field of (\ref{6.1})-(\ref{6.3}), it appears impossible to verify these conditions analytically.
1709: However, the numerical simulations clearly show that system of equations (\ref{6.1})-(\ref{6.3}) possesses
1710: the qualitative structure required by ({\bf G1}) and ({\bf G2}) (Figure \ref{f.2}b). On the other hand, we expect that conditions 
1711: ({\bf G1}) and ({\bf G2}) should be possible to verify analytically for a wide class of slow-fast systems.
1712: Therefore, we believe that our results will be useful for understanding mixed-mode oscillations in such systems. 
1713: In particular, it would be interesting to apply this approach to the modified Hodgkin-Huxley system \cite{DK05}.
1714: The numerical results reported in \cite{DK05} strongly suggest that the mechanism proposed in the present paper
1715: is responsible for the generation of the very slow rhythms and chaotic dynamics in the Hodgkin-Huxley model.
1716: 
1717: \noindent
1718: {\bf Acknowledgments.}
1719: We thank Victor Roytburd for introducing us to this problem and to Michael Frankel and 
1720: Victor Roytburd for helpful conversations. This work  was
1721: partially supported through National Science Foundation Award No 0417624.
1722: 
1723: \renewcommand{\theequation}{A.\arabic{equation}}
1724: \section*{Appendix A.} 
1725: \setcounter{equation}{0}
1726: \label{sec:C}
1727: In this appendix, we list 
1728: explicit expressions of various constants and trigonometric polynomials, which appear
1729: in the definitions in of the slow manifold, $S$, and the first Lyapunov coefficient, $\gamma$.
1730: All expressions are given in terms
1731: of the coefficients of the power expansions on the right  hand side of (\ref{5.1}).
1732: By $\sigma_{ijk}$ we denote $\sigma_{ijk}(0),$ $\sigma\in\{a,b,c\}$.
1733: The following constants are used to define the leading order approximation of the slow
1734: manifold in (\ref{5.3}) and (\ref{5.4}):
1735: $$
1736: a={a_{22}+a_{33}\over 2},\;
1737: A=\sqrt{a^2_{23}+\left(a_{22}-a_{33}\right)^2\over 4\left(1+ 4\beta^2\right)},\;\mbox{and}\;
1738: \vartheta=\arctan {2\beta a_{23}+\left(a_{22}-a_{33}\right)\over  a_{23}+2\beta \left(a_{22}-a_{33}\right)} .
1739: $$
1740: The following trigonometric polynomials enter the right hand sides of 
1741: (\ref{5.11a}) and (\ref{5.12}): 
1742: \begin{eqnarray*}
1743: P_{1}(\phi)&=& a_{22}\cos^2\beta\phi+a_{23}\cos\beta\phi\sin\beta\phi+a_{33}\sin^2\beta\phi,\\
1744: Q_{1}(\phi)&=&b_{22}\cos^3\beta\phi+(b_{23}+c_{22})\cos^2\beta\phi\sin\beta\phi
1745: +(b_{33}+c_{23})\cos\beta\phi\sin^2\beta\phi+c_{33}\sin^3\beta\phi,\\
1746: Q_{2}(\phi)&=&b_{12}\cos^2\beta\phi+(b_{13}+c_{12})\cos\beta\phi \sin\beta\phi
1747: +c_{13}\sin^2\beta\phi,\\
1748: Q_{3}(\phi)&=&b_{222}\cos^4\beta\phi+(b_{223}+c_{222})\cos^3\beta\phi \sin\beta\phi
1749:                   +(b_{233}+c_{223})\cos^2\beta\phi \sin^2\beta\phi\\
1750:                  &&+(b_{333}+c_{233})\cos\beta\phi \sin^3\beta\phi+c_{333}\sin^4\beta\phi,\\
1751: \end{eqnarray*}
1752: Functions $\bar Q_{2,3}(\theta)=Q_{2,3}(\beta^{-1}\theta)$ are used in the calculation
1753: of the first Lyapunov coefficient $\gamma$ (see (\ref{5.5})).
1754: 
1755: \begin{thebibliography} {99}
1756: 
1757: \bibitem{GEO} V.I. Arnold, 
1758: {\it Geometrical Methods in the Theory of Ordinary Differential 
1759: Equations}, Springer, 1983.
1760: 
1761: \bibitem{AAIS}
1762: V. Arnold, V. Afraimovich, Y. Iliyashenko, L. Shilnikov, 
1763: {\it Bifurcation Theory}, Springer, New York, 1994.
1764: 
1765: \bibitem{BER}
1766: S.M. Baer, T. Erneux, and J. Rinzel,
1767: The slow passage through a Hopf bifurcation: delay, memory
1768: effects, and resonance, {\it SIAM J. Appl. Math.},
1769: $\bf 49$ 55--71, 1989.
1770: 
1771: \bibitem{BS}
1772: M. Bosch and C. Simo, Attractors in a Shilnikov-Hopf scenario
1773: and a related one-dimensional map, {\it Physica D}, $\bf 62$, 217-229, 1993.
1774: 
1775: \bibitem{Brai_Siva}
1776: I. Brailovski and G. Sivashinsky, Chaotic dynamics in solid fuel combustion, {\it Physica D},
1777: $\bf 65$, 191-198, 1992.
1778: 
1779: \bibitem{CDD}
1780: J.L. Callot, F. Diener, and M. Diener,
1781: Le probleme de la "chasse au canard",
1782: {\it C. R. Acad. Sci. Paris (ser. I)}, $\bf 286$, 1059--1061, 1978.
1783: 
1784: \bibitem{Carr} 
1785: J. Carr, {\it Applications of Centre Manifold Theory}, Springer, New York, 1981.
1786: 
1787: %\bibitem{CH} 
1788: %S.-N. Chow and J.K. Hale, {\it Methods of Bifurcation Theory}, Springer, New York, 1982.
1789: %
1790: \bibitem{CMP} 
1791: S.-N. Chow and J Mallet-Paret, Integral averaging and bifurcation, {\it JDE}, $\bf 26$, 112­
1792: 159, 1977.
1793: 
1794: \bibitem{DENG90} 
1795: B. Deng, Homoclinic bifurcations with nonhyperbolic equilibria,
1796: {\it SIAM J. Math. Anal.}, $\bf 21$(3), 693-720, 1990.
1797: 
1798: \bibitem{DENG95}
1799: B. Deng, Shilnikov-Hopf bifurcations, {\it JDE}, $\bf 119$, 1-23, 1995. 
1800: 
1801: \bibitem{DK05}
1802: S. Doi and S. Kumagai, 
1803: Generation of very slow neuronal rhythms and chaos near the Hopf bifurcation in single neuron models,
1804: {\it J. Comp. Neurosc.}, 325--356, 2005.
1805: 
1806: \bibitem{DRSE}
1807: J. Drover, J. Rubin, J. Su, and B. Ermentrout, 
1808: Analysis of a canard mechanism by which excitatory synaptic coupling
1809: can synchronize neurons at low firing frequencies,
1810: {\it SIAM J. Appl. Math.}, $\bf 65$,  69--92, 2004.
1811: 
1812: \bibitem{FKRT}
1813: M. Frankel, G.Kovacic, V. Roytburd, and I. Timofeyev,
1814: Finite-dimensional dynamical system modeling thermal instabilites, {\it Physica D},
1815: $\bf 137$, 295--315, 2000.
1816: 
1817: \bibitem{FR05a}
1818: M. Frankel and V. Roytburd, 
1819: Dynamical structure of one-phase model of solid combustion,
1820: Preprint, 2005.
1821: 
1822: \bibitem{FR05b}
1823: M. Frankel and V. Roytburd, 
1824: Frequency locking for combustion synthesis in periodic medium, {\it Physics Letters A},
1825: $\bf 329$, 2004.
1826: 
1827: \bibitem{FR03}
1828: M. Frankel and V. Roytburd, 
1829: Finite-dimensional attractor for a 1-phase Stefan problem with kinetics, 
1830: {\it J. Dynamics Diff. Equations}, 2003. 
1831: 
1832: \bibitem{FR95}
1833: M. Frankel and V. Roytburd, Finite-dimensional model of thermal
1834: instabilities, {\it Appl. Math. Let.}, $\bf 8$, 39--44, 1995.
1835: 
1836: \bibitem{FR94}
1837: M. Frankel and V. Roytburd,
1838: A free boundary problem modeling thermal instabilities:
1839: stability and bifurcation,
1840: {\it J. of Dynamics and Differential Equations},
1841: $\bf 6$, 447--486, 1994.
1842: 
1843: \bibitem{GL}
1844:  M. Golubitsky and W.F. Langford,
1845:  Classification and unfoldings of degenerate Hopf bifurcations, {\it JDE}, $\bf 41$,  375-415, 1981.
1846: 
1847: \bibitem{GH}
1848: J. Guckenheimer and  P. Holmes,
1849: {\it  Nonlinear Oscillations,
1850: Dynamical Systems, and Bifurcations of Vector Fields}, Springer, 1983.
1851: 
1852: \bibitem{GW}
1853: J. Guckenheimer and A.R. Willms,
1854: Asymptotic analysis of subcritical Hopf-homoclinic bifurcation,
1855: {\it Physica D}, $\bf 139$, 195--216, 2000.
1856: 
1857: \bibitem{hartman}
1858: P. Hartman, {\it Ordinary Differential Equations}, 
1859: John Willey \& Sons, Baltimore, 1973.
1860: 
1861: \bibitem{Hirschberg_Knobloch}
1862: P. Hirschberg and E. Knobloch, Shilnikov-Hopf bifurcation, {\it Physica D}, $\bf 62$,
1863: 202-216, 1993. 
1864: 
1865: \bibitem{JON}
1866: C.K.R.T. Jones, Geometric singular perturbation theory, in {\it CIME Lectures 
1867: in Dynamical Systems}, Lecture Notes in Mathematics, Springer-Verlag, 1994.
1868: 
1869: \bibitem{KR83}
1870: J.P. Keener and J. Rinzel, 
1871: Hopf Bifurcation to Repetitive Activity in Nerve, 
1872: {\it SIAM J. Appl. Math.}, $\bf 43$, 907-922, 1983.
1873: 
1874: \bibitem{KOP}
1875: M.T.M. Koper, 
1876: Bifurcations of mixed-mode oscillations in a three-variable autonomous
1877: van der Pol-Duffing model with a cross-shaped phase diagram,
1878: {\it Phys. D},  $\bf 80$, 72--94, 1995.
1879: 
1880: \bibitem{KS01b}
1881: M. Krupa and P. Szmolyan,
1882: Relaxation oscillation and canard explosion, {\it JDE}, 
1883: $\bf174$, 312--368, 2001.
1884: 
1885: \bibitem{KUZ}
1886: A.Yu. Kuznetsov, 
1887: {\em Elements of Applied Bifurcation Theory}, Springer, 1998.
1888: 
1889: \bibitem{MarsdenMcCracken}
1890: J. Marsden and M. McCracken,
1891: {\it Hopf Bifurcation and Its Applications}, Springer-Verlag, New York, 1976.
1892: 
1893: \bibitem{MC}
1894: G. Medvedev, J.E. Cisternas, 
1895: Multimodal regimes in a compartmental model of the dopamine neuron , 
1896: {\it Physica D}, $\bf 194$(3-4), 333-356, 2004.
1897: 
1898: \bibitem{MSLG}
1899: A. Milik, P. Szmolyan, H. L$\o$ffelman, and E. Gr$\o$ller,
1900: Geometry of mixed-mode oscillations in the 3-D autocatalator, 
1901: {\it Intern. J. of Bifurc. and Chaos}, $\bf 8$, 505--519, 1998.
1902: 
1903: \bibitem{MKKR}
1904: E.F. Mischenko, Yu.S. Kolesov, A.Yu. Kolesov, and N.Kh. Rozov,
1905: {\it Asymptotic Methods in Singularly Perturbed Systems}, Consultants Bureau,
1906: New York, 1994.
1907: 
1908: \bibitem{OP} 
1909: V.V. Osipov and E.V. Ponizovskaya, Multivalued 
1910: stochastic resonance in a model of an excitable neuron,
1911: {\it Physic Letters A}, $\bf 271$, 191--197, 2000.
1912: 
1913: \bibitem{PSS}
1914: V. Petrov, S.K. Scott, and K. Showalter,
1915: Mixed-mode oscillations in chemical systems,
1916: {\it J. Chem. Phys.}, $\bf 97$(9), 6191--6198.
1917:  
1918: \bibitem{RE89}
1919: J. Rinzel and G.B. Ermentrout,
1920: Analysis of neural excitability and oscillations,
1921: in C. Koch and I. Segev, eds
1922: {\it Methods in Neuronal Modeling}, MIT Press,
1923: Cambridge, MA, 1989.
1924: 
1925: \bibitem{ROWK}
1926: H. G. Rotstein, T. Oppermann, J. A. White, and N. Kopell,
1927: A reduced model for medial entorhinal cortex stellate cell: subthreshold oscillations, 
1928: spiking and synchronization, Preprint, 2005.
1929: 
1930: \bibitem{VV}
1931: E.I. Volkov and D.V. Volkov,
1932: Multirhythmicity generated by slow variable diffusion in a ring of
1933: relaxation oscillators and noise-induced abnormal interspike
1934: variability, {\it Phys. Rev. E}, $\bf 65$,  2002.
1935: 
1936: \bibitem{WESCH}
1937: M. Wechselberger,
1938: Existence and Bifurcation of Canards in $R^3$ in the case of a Folded Node,
1939: {\it SIAM J. Applied Dynamical Systems}, $\bf 4$(1), 101-139, 2005.
1940: 
1941: \end{thebibliography}
1942: 
1943: 
1944: \end{document}
1945: