nlin0308034/pre5.tex
1: %\documentclass[aps,prl,preprint,groupedaddress]{revtex4}
2: \documentclass[aps,pre,twocolumn,english,superscriptaddress,showpacs,floatfix]{revtex4}
3: %\documentclass[aps,prl,twocolumn,groupedaddress]{revtex4}
4: 
5: \usepackage{amsmath,bbm}
6: \usepackage{psfig,graphics,graphicx}
7: \usepackage{babel}
8: 
9: 
10: 
11: \begin{document}
12: 
13: % Use the \preprint command to place your local institutional report
14: % number in the upper righthand corner of the title page in preprint mode.
15: % Multiple \preprint commands are allowed.
16: % Use the 'preprintnumbers' class option to override journal defaults
17: % to display numbers if necessary
18: %\preprint{}
19: 
20: %Title of paper
21: \title{Topological constraints on
22: spiral wave dynamics in spherical geometries with inhomogeneous
23: excitability}
24: 
25: \author{J\"orn Davidsen}
26: \email[]{davidsen@mpipks-dresden.mpg.de}
27: \affiliation{Max-Planck-Institut f\"ur Physik Komplexer Systeme,
28: N\"othnitzer Strasse 38, 01187 Dresden, Germany}
29: \affiliation{Chemical Physics Theory Group, Department of
30: Chemistry, University of Toronto, Toronto, ON M5S 3H6, Canada}
31: \author{Leon Glass}
32: \email[]{glass@cnd.mcgill.ca}
33: \affiliation{Department of Physiology, McGill University,
34: Montreal, Quebec H3G 1Y6, Canada}
35: \author{Raymond Kapral}
36: \email[]{rkapral@chem.utoronto.ca} \affiliation{Max-Planck-Institut
37: f\"ur Physik Komplexer Systeme,
38: N\"othnitzer Strasse 38, 01187 Dresden, Germany}
39: \affiliation{Chemical Physics
40: Theory Group, Department of Chemistry, University of Toronto,
41: Toronto, ON M5S 3H6, Canada}
42: 
43: 
44: \date{\today}
45: 
46: \begin{abstract}
47: We analyze the way topological constraints and inhomogeneity in
48: the excitability influence the dynamics of spiral waves on spheres
49: and punctured spheres of excitable media. We generalize the
50: definition of an index such that it characterizes not only each
51: spiral but also each hole in punctured, oriented, compact,
52: two-dimensional differentiable manifolds and show that the sum of
53: the indices is conserved and zero. We also show that heterogeneity
54: and geometry are responsible for the formation of various spiral
55: wave attractors, in particular, pairs of spirals in which one
56: spiral acts as a source and a second as a sink --- the latter
57: similar to an antispiral. The results provide a basis for the
58: analysis of the propagation of waves in heterogeneous excitable
59: media in physical and biological systems.
60: 
61: 
62: 
63: 
64: 
65: \end{abstract}
66: 
67: % insert suggested PACS numbers in braces on next line
68: \pacs{87.19-j, 05.40.-a, 89.75.-k}
69: % insert suggested keywords - APS authors don't need to do this
70: %\keywords{}
71: 
72: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
73: \maketitle
74: 
75: % body of paper here - Use proper section commands
76: % References should be done using the \cite, \ref, and \label commands
77: 
78: \section{Introduction}
79: Geometry and inhomogeneity influence pattern formation in chemical
80: and biological systems. \cite{murray89,kapral} One example where
81: these two factors play a crucial role is in the experimental
82: observations of distinctive spiral wave dynamics on the surfaces
83: of spherical beads, which are excitable inhomogeneous chemical
84: media \cite{maselko89,maselko90}. A biological example is the
85: origin of abnormal cardiac rhythms in the human heart which depend
86: on the anatomical substrate. The heart possesses a complex
87: nonplanar geometry with multiple chambers, with holes
88: corresponding to valves and blood vessels. Some serious
89: arrhythmias are associated with circulating spiral waves similar
90: to those observed in chemical media \cite{winfree01}. Since an
91: abnormal anatomical substrate is a common finding in patients with
92: some types of cardiac arrhythmias, and interventions that modify
93: the anatomy are an accepted form of therapy \cite{stein02},
94: theoretical analyses of the interplay between geometry of the
95: substrate and dynamics may help in the therapy of cardiac
96: arrhythmias.
97: 
98: In this paper we study spiral wave dynamics on (punctured) spheres
99: with spatially inhomogeneous excitability. We show for punctured
100: spheres that the sum of indices which characterize each spiral has
101: to be zero. Moreover, we demonstrate that topological constraints
102: imposed by the spherical geometry and inhomogeneity in
103: excitability lead to the formation of pairs of spirals, with
104: distinctive transient dynamics or as stable attractors, in which
105: one spiral acts as a source and a second as a sink leading to a
106: source-sink pair under a broad range of conditions. Our results
107: explain the experimental observations of spirals on spherical
108: beads \cite{maselko89,maselko90}.  While we do not consider
109: detailed models of cardiac wave propagation, our results may apply
110: to some generic aspects of atrial arrhythmias because the thin
111: walls of the atria can be described as two-dimensional (2d)
112: inhomogeneous excitable media with specific geometrical features.
113: \cite{3d}
114: 
115: 
116: \section{An Index Theorem for Phase Singularities}
117: 
118: The mathematical description of spiral waves is based on the
119: notion of phase which in turn allows one to characterize spiral
120: waves by an index. From this description, a number of topological
121: results placing restrictions on spiral wave dynamics can be
122: derived. \cite{glass77,winfree83,sumners87,winfree01,cruz-white03}
123: 
124: With the exception of a finite number of singular points, with
125: each point in an orientable and compact two-dimensional
126: differentiable manifold $M$ we identify a unique phase lying on
127: the unit circle, $\Phi \in S^1$. The resulting phase map or phase
128: field is assumed to be continuously differentiable, except at the
129: singular points. The manifold can be triangulated \cite{hocking88}
130: (subdividing it into a set of polygons), where none of the edges
131: or vertices of the polygons pass through a singularity. The index,
132: $I$ (sometimes also called the topological charge or winding
133: number) of a curve $C$ bounding a polygon is found by computing
134: the line integral
135: \begin{equation}\label{top_charge}
136:     2 \pi I = \oint_C \nabla \Phi \cdot d{\bf l},
137: \end{equation}
138: where the polygon is always traversed in a clockwise orientation.
139: By continuity of $\nabla \Phi$, $I$ must be an integer. The index
140: of a singular point is uniquely defined as the index of any curve
141: $C$ provided that $C$ encircles the point but no other singular
142: points. The index of a curve that does not enclose any singular
143: points is obviously zero.
144: 
145: If the manifold $M$ has no boundaries, each edge of the
146: triangulation is an edge of two polygons. Since the line integral
147: adds up the change in phase along the various edges of the
148: polygon, the sum of the indices of the singular points for a phase
149: field in $M$ is
150: \begin{equation}\label{indextheorem}
151: \Sigma_M I =0,
152: \end{equation}
153: where the sum is over all the singular points.  This follows since
154: the contribution of the change in phase of each edge to the total
155: integral is counted twice, but since the edge is traversed in
156: opposite directions each time, the net contribution of each edge
157: is zero. This index theorem is applicable to tori, and other
158: surfaces of genus different from 0. However, unlike the more
159: familiar Poincar\'e index theorem (see Ref. \cite{guillemin74} p.
160: 74 for vector fields) the sum of the indices of the singular
161: points does not depend on the genus of the surface.
162: 
163: This index theorem for manifolds without boundaries can be
164: extended to manifolds with boundaries. In the following, we will
165: consider the case of structures that arise from puncturing
166: orientable and compact two-dimensional differentiable manifolds.
167: The index of a hole can be uniquely defined as the index of a
168: curve $C$ provided that $C$ encircles the hole but no other holes
169: or singular points and $C$ is positively oriented with respect to
170: the domain which contains the hole and is bounded by $C$. Applying
171: this definition and taking the summation in Eq.
172: (\ref{indextheorem}) over the singularities and the hole, or the
173: holes if there is more than one hole, the index theorem can be
174: proven by the same line of arguments as for the case of manifolds
175: without boundaries.
176: 
177: This extension is important in the heart, for example, where the
178: atrium is punctured by valves and veins. In such cases one is led
179: to consider manifolds with holes; for example, a sphere with a
180: hole. A sphere with a hole is topologically equivalent to a disk,
181: and, indeed, the results for disks and for spheres with holes are
182: consistent: For the disk $D^2$, bounded by a curve $C$,
183: \begin{equation}\label{dindextheorem}
184: \Sigma_{D^2} I =\oint_C \nabla \Phi \cdot d{\bf l},
185: \end{equation}
186: so that the sum of the indices of the singular points in the disk
187: is equal to the index of the curve $C$ bounding the disk. If there
188: is a single singular point on the disk, with an index of +1, the
189: index of the curve bounding the disk will also be +1. Imagine now
190: the boundary of the disk to be brought together (like a
191: draw-string bag) so that the boundary of the disk now defines a
192: hole in the sphere. In this geometry the curve $C$ will be
193: traversed in an opposite orientation (the hole is now inside $C$)
194: from the direction it was traversed when it was the boundary of
195: the disk. Now if there is a singular point with an index of +1 on
196: the sphere, the index of the hole is -1, so that the sum of the
197: indices is again zero.
198: 
199: Since it is necessary to conserve the sum of the indices,
200: singularities of index $\pm 1$ usually arise and are destroyed in
201: pairs of opposite sign \cite{winfree83}. An exception occurs when
202: singularities are destroyed by collision into a boundary, so that
203: the index of the singular point and the index of the boundary both
204: change simultaneously. Another exception occurs if there are
205: singularities with index different from one. In such cases
206: interactions between different singularities can lead to
207: destruction or creation of odd numbers of singularities
208: \cite{zaritski02}.
209: 
210: 
211: 
212: \section{The FitzHugh-Nagumo Equation}
213: 
214: The FitzHugh-Nagumo (FHN) equation, \cite{numerics}
215: \begin{eqnarray}
216: \frac{\partial u}{\partial t} &=& \epsilon^{-1} \left( -
217: \frac{u^3}{3} + u - v \right) + D_u \nabla^2 u,\nonumber\\
218: \frac{\partial v}{\partial t} &=& \epsilon \left( u - \alpha v +
219: \beta \right) + D_v \nabla^2 v,\label{equdyn}
220: \end{eqnarray}
221: where $u({\bf r},t)$ and $v({\bf r},t)$ are two scalar fields,
222: $\epsilon^2$ is the ratio of the time scales associated with the
223: two fields, and $D_u$ and $D_v$ are the constant diffusion
224: coefficients, is a
225: prototypical model for excitable media.
226: We choose $D_u = 2$ and $D_v = 0$. The real
227: parameters $\alpha$ and $\beta$ characterize the local dynamics
228: and, hence, the local excitability. Whenever $0<\alpha<1$, $\alpha
229: \epsilon^2 < 1$ and $|\beta| > \beta_H \equiv (1 - \alpha
230: \epsilon^2)^{1/2} (1/3 (2 \alpha + \alpha^2 \epsilon^2) - 1)$, the
231: FHN system is excitable. At $\beta_H$ a Hopf bifurcation occurs
232: such that for $|\beta| < \beta_H$ the system exhibits
233: oscillations. In the following we take $\alpha = 0.2$, $\epsilon=0.2$ and
234: $\beta > \beta_H = 0.863\dots$.
235: 
236: We consider a spherical shell whose outer and inner radii are
237: $R_e$ and $R_i$, respectively, and focus on thin spherical shells
238: where $R_i=40$, $R_e=42$. The radii are large enough to avoid a
239: curvature-dependent loss of excitability \cite{davydov02}, and the
240: shell is sufficiently thin that the dynamics is effectively 2d
241: corresponding to the dynamics on a sphere. \cite{shell} The
242: initial condition is a domain of ``excited" state, adjacent to a
243: domain of the ``refractory" state. Both domains have the forms of
244: slices of the same size oriented from the north to the south pole
245: \cite{details} and yield a pair of counter-rotating spirals.
246: 
247: In order to apply the topological results to the FHN equation, it
248: is necessary to first define the phase. We define a phase,
249: $\Phi({\bf r},t)$, based on the equation $\tan{\Phi({\bf r},t)} =
250: v({\bf r},t)/u({\bf r},t)$ if $v({\bf r},t)\neq 0$ and $u({\bf
251: r},t)\neq 0$. Thus, singular points at given $t$ are points ${\bf
252: r}$ in the medium for which $v({\bf r},t)=0$ and $u({\bf r},t)=0$.
253: For each $t$, we obtain a continuously differentiable phase map
254: ${\cal M}^t=\Phi({\bf :},t)|_{{\cal D}^t}$ that associates to each
255: point in a well-defined domain ${\cal D}^t$ a phase lying on the
256: unit circle, $\Phi \in S^1$. For our FHN medium, the domain is the
257: surface of a (punctured) sphere reduced by a finite number of
258: points where the phase is singular at fixed $t$.
259: 
260: Rotating spiral waves in the FHN equations are obviously
261: associated with a singular point which is called the spiral core.
262: In what follows, we assume that there are only single-armed
263: spirals so that a clockwise rotating source has an index of +1,
264: and a counterclockwise rotating source has index -1. A clockwise
265: rotating sink has an index of -1, and a counterclockwise rotating
266: sink has index +1. From Eq.~(\ref{indextheorem}), it is impossible
267: to have a single rotating spiral wave on a sphere; in addition,
268: there must be at least another singular point or a hole with
269: nonzero index.
270: 
271: For excitable media, a non-zero index of a hole implies that wave
272: fronts travel permanently around the hole such that the numbers of
273: fronts travelling clockwise and counterclockwise are different.
274: This includes the particularly simple case of a single wave front
275: travelling around the hole which can be considered as a spiral
276: wave associated with the hole.
277: 
278: \section{Spiral Wave Dynamics in Spherical Geometries}
279: \subsection{Dynamics in excitability gradients}
280: First consider homogeneous FHN media with a constant
281: $\beta=\beta_{ex}\equiv0.9$. The wave fronts emanating from the
282: spiral cores with opposite index ``annihilate" along the equator
283: such that each spiral determines the dynamics on half of the
284: sphere (see Fig.~\ref{exc_hom}, left panel) --- similar to what
285: has been observed in Ref.~\cite{yagisita98} for a different
286: excitable system.
287:  \begin{figure}
288:  \includegraphics[width=0.75\columnwidth]{comb}
289:  \caption{\label{exc_hom}
290: (Color online) Left: Spiral waves of excitation (light fronts) on
291: the sphere for constant $\beta=\beta_{ex}$ emanating from spiral
292: cores close to the poles on an equator projection. The white
293: arrows show the direction of propagation. One annihilation front
294: along the equator can be identified. Right: Sketch of the constant
295: gradient in the inhomogeneous case. The dashed lines are the
296: equi-$\beta$ lines and we choose $\beta_{max}=1.0$ and
297: $\beta_{min}=\beta_{ex}$. The angle $\phi$ describes the
298: orientation with respect to the axis from pole to pole. The
299: results described in the text do not depend qualitatively on the
300: choice of $\phi$ or $\beta_{min}$ and $\beta_{max}$ as long as
301: they yield stable spirals.}
302:  \end{figure}
303: We have shown that this behavior is robust with respect to
304: disorder in excitability with small amplitude and correlation
305: length. If random, uncorrelated spatial variations in $\beta$
306: exist on length scales much smaller than the diameter of the
307: spiral core meander \cite{random_details}, the dynamics is able to
308: average over such small-scale inhomogeneities. The robustness
309: explains why such states have been experimentally observed in some
310: chemical reactions on spherical beads which are intrinsically
311: inhomogeneous \cite{maselko90}.
312: 
313: Applying a constant gradient in $\beta$ as sketched in the right
314: panel of Fig.~\ref{exc_hom} leads to a different scenario. The time
315: evolution of the spiral pair may be partitioned into four distinct
316: regimes as shown in Figs.~\ref{beta} and \ref{d}.
317:  \begin{figure}
318:  \includegraphics[width=\columnwidth]{beta_ang089_trans_new}
319:  \caption{\label{beta}
320: (Color online) The local excitability $\beta({\bf r}(t))$ at the
321: spiral cores versus time. Gradient-induced motion of the two
322: spiral cores leads to a change in the local excitability at the
323: cores with time. The spiral period in the final state is
324: $T_0=13.2\pm0.1$. Four different regimes can be identified (see
325: text). Inset: The final bound pair of counter-rotating spirals in
326: regime III for $\phi=51.0^\circ$ is shown on an equator projection
327: such that the point of lowest excitability lies on the central
328: longitude. The spiral closer to the equator has index $-1$ while
329: the other one has index $+1$.}
330:  \end{figure}
331:  \begin{figure}
332:  \includegraphics[width=\columnwidth]{d_ang089_trans}
333:  \caption{\label{d}
334: (Color online) The distance between the two spiral cores $d(t)$
335: versus $t$. Gradient-induced motion of the two spiral cores leads
336: to a change in the distance $d$ between the cores with time. The
337: lower and upper curves correspond to the distance in
338: $\mathbbm{R}^3$ and $S^2$, respectively. The dashed lines are the
339: respective upper bounds given by the size of the sphere.}
340:  \end{figure}
341: Because of the gradient, the frequencies of the two spirals differ
342: since a higher value of $\beta$ corresponds to lower excitability,
343: which generally implies a lower spiral frequency \cite{winfree91}.
344: During a short transient T, the spiral with the higher frequency
345: assumes control of the dynamics \cite{krinsky83} on the sphere.
346: The location on the sphere, where wave fronts emanating from the
347: two spiral cores annihilate, moves toward the core of the
348: low-frequency spiral. Finally, the low-frequency spiral core is
349: pushed farther from the high-frequency spiral core
350: \cite{krinsky83,ermakova86}(see Fig.~\ref{d}). After this short
351: transient, the wave fronts travel from pole to pole leading to the
352: creation of a source-sink pair. This (intermediate) state is shown
353: in Fig.~\ref{anti_series} and corresponds to regime I in
354: Figs.~\ref{beta} and \ref{d}. Viewed from the low excitability end
355: of the sphere, the waves wind into a small region about the core,
356: reminiscent of the structure of antispirals, i.e., inward moving
357: spirals seen in oscillatory media \cite{gong03}. However, the
358: origin of this inward spiral motion in oscillatory media differs
359: and is distinct from that observed here. In oscillatory media,
360: either spirals or antispirals are stable depending on system
361: parameters and the wave length diverges on the border in the
362: parameter space between these two regimes. Thus, antispirals exist
363: independently of spirals. This is not the case here because the
364: generation of an inward moving spiral relies on the presence of a
365: spiral source and spherical geometry. For example, consider the
366: FHN system with a disc geometry and a radial gradient in
367: excitability such that the maximum value of $\beta$ is located in
368: the center of the disc. In this case, a source-sink pair cannot
369: occur because the high-frequency spiral, acting as a source, would
370: push the other (low-frequency) spiral out of the system, excluding
371: the presence of any strong random inhomogeneities in excitability
372: which may pin the low-frequency spiral and prevent its motion
373: (see, e.g., \cite{biktashev01}). The lower panel of
374: Fig.~\ref{anti_series} shows that remnants of the low frequency
375: spiral persist in a small area around the core. It has been
376: speculated that antispiral waves might occur in cardiac tissue
377: \cite{vanag01}. While excitable media cannot support a regime of
378: exclusive antispirals due to their excitable character, our
379: results show that source-sink pairs with similar characteristics
380: could form in the heart where the underlying dynamics is
381: excitable, the medium is inhomogeneous and the topology is similar
382: to a (punctured) sphere.
383: 
384: 
385: Spiral dynamics of the type described above has been observed by
386: Maselko and Showalter \cite{maselko89} in experiments on the
387: excitable Belouzov-Zhabotinsky reaction on spherical beads. They
388: attributed the generation of spiral source-sink pairs to
389: inhomogeneities in the medium related to differing chemical
390: environments. This is consistent with our findings for systems
391: with gradients and is further confirmed by the work reported in
392: Ref.~\cite{yagisita98}. While the generation of source-sink pairs
393: due to a gradient in the FHN medium investigated here is only an
394: intermediate state (but one that persists for approximately 240
395: spiral periods in our simulations), random spatial variations of
396: the excitability with a correlation length comparable to the
397: diameter of the spiral core meander or larger can lead to a final
398: state consisting of such a source-sink pair \cite{random_details}.
399: This is due to the fact that the source can be trapped in a region
400: of depressed local excitability.
401:  \begin{figure}
402:  \includegraphics[width=\columnwidth]{anti_series}
403:  \caption{\label{anti_series}
404: (Color online) Waves of excitation on the sphere in regime I of
405: the gradient-induced dynamics shown in Figs.~\ref{beta} and
406: \ref{d}. A source-sink pair has formed. For random spatial
407: variations of excitability with a correlation length comparable to
408: the diameter of the spiral core meander, the final state is very
409: similar to the one shown here \cite{random_details}. Upper panel
410: from left to right: View centered at the north pole, south pole
411: and the equator. The source at the south pole has index $-1$ and
412: the sink at the north pole index $+1$. The black circles show
413: possible choices of $C$. The white arrow shows the direction of
414: wave propagation. Lower panel: Dynamics at the north pole. Time
415: increases from left to right with $\Delta t = 2.5$ between
416: snapshots.}
417:  \end{figure}
418: 
419: Not only does the gradient in the FHN medium change the local
420: excitability but it also induces a drift of the spiral cores
421: \cite{biktashev95}. For our model, the drift is rather slow
422: compared to the transition to the source-sink pair which takes
423: place during the transient regime T. This can be seen in
424: Fig.~\ref{beta}. (The fluctuations in $\beta({\bf r}(t))$ are due
425: to spiral meandering.) In regime I, the dominating spiral drifts
426: toward lower excitability and its wave fronts continuously push
427: the other core in the opposite direction, thus, keeping the
428: distance $d$ between the cores approximately constant (see
429: Fig.~\ref{d}). The fluctuations in $d(t)$ are again due to
430: meandering of the spiral cores. Because of this drift, the local
431: excitabilities at the two spiral cores approach each other until
432: they become equal.
433: 
434: At this point, regime II in Figs.~\ref{beta} and \ref{d} is
435: entered. The dynamics change drastically: the slaved spiral
436: reverses its drift direction and regains control over its own
437: dynamics. Both spirals drift toward lower excitability. Due to the
438: geometric constraints imposed by the spherical geometry, the
439: spirals approach each other until they form a bound pair (at $t
440: \approx 7500$).
441: 
442: They finally reach a stable state (at $t \approx 9500$)
443: corresponding to regime III in Figs.~\ref{beta} and \ref{d}.
444: Neither the \emph{average} distance between the spirals nor the
445: \emph{average} local excitability changes further. Yet, on top of
446: the persisting unsynchronized meandering of the two spiral cores,
447: the bound pair slowly moves along a (closed) equi-$\beta$ curve on
448: the surface of the sphere. The direction of the motion depends on
449: the initial condition, i.e., whether the spiral with positive
450: index was initially closer to the point of lowest excitability or
451: to the point of highest excitability than the spiral with negative
452: index. The wave dynamics generated by this bound pair is shown in
453: the inset of Fig.~\ref{beta}.
454: 
455: While kinematic theory applies only to spirals with large cores,
456: it is instructive to note that this theory predicts that the
457: direction of the drift due to gradients depends on the model
458: system and its parameters \cite{mikhailov94}. Although our
459: simulations have shown the same direction of the drift for a range
460: of parameters in the meander region of the FHN phase diagram, it
461: is conceivable that, under different circumstances favoring a
462: drift toward higher excitability, one spiral could act as a
463: permanent source and a source-sink pair could be the final
464: attractor. Such a scenario would also be consistent with the
465: experimental findings in Ref. \cite{maselko89}.
466: 
467: \subsection{Punctured spheres}
468: Next, we consider a homogeneously excitable sphere with a single
469: hole. Two scenarios can be observed depending on the location of
470: the hole with respect to the spiral pair. If a spiral wave is not
471: permanently attached to the hole, the dynamics is very similar to
472: the case without any hole. If one of the spirals is permanently
473: attached to the hole, the frequency of this spiral is lowered. The
474: size of the hole determines the frequency of the spiral because
475: the wave front has to travel around the hole. The transient
476: dynamics is similar to that in regime T for the case with a
477: gradient; however, no drift of the spiral cores is induced and the
478: final state is a spiral source-sink pair as shown in
479: Fig.~\ref{series_hole}. Not only is the net index conserved during
480: the transition to a spiral source-sink pair but so is the index of
481: the individual spirals: during the transition an outgoing
482: counterclockwise (clockwise) oriented spiral is converted into an
483: ingoing clockwise (counterclockwise) oriented spiral. Thus, the
484: formation of spiral source-sink pairs conforms with the
485: topological constraints.
486: 
487: If a gradient as well as a hole is present, the spiral drift
488: discussed earlier also determines the final state, which depends
489: on the hole's location in the gradient field. For instance, if the
490: point of lowest excitability in the medium is on the hole
491: boundary, simulations show that the spiral which is not attached
492: to the hole will stabilize close to this point. In this final
493: state odd numbers of wavefronts are attached to the hole, again
494: conserving the topological charge of the hole.
495: 
496: 
497: \begin{figure}
498:   \includegraphics[width=\columnwidth]{series_hole}
499:   \caption{\label{series_hole}
500: (Color online) Waves of excitation on the punctured sphere
501: propagating towards the hole. A source-sink spiral wave pair has
502: formed in the final state. Upper panel from left to right: View
503: centered at the north pole, south pole and the equator. Lower
504: panel: Dynamics at the south pole. Time increases from left to
505: right with $\Delta t = 2.5$ between snapshots.}
506: \end{figure}
507: 
508: \section{Concluding Remarks}
509: Inhomogeneities due to spatially-varying excitability on
510: (punctured) spherical shells lead to complex spiral wave dynamics
511: and the formation of source-sink spiral pairs in excitable media.
512: The results presented here are immediately applicable to excitable
513: media in more complicated geometries such as tori or multi-holed
514: tori and to situations in which multi-armed spirals are found.
515: This includes mathematical modelling of cardiac tissue. The
516: approach taken in this paper stresses constraints and aspects that
517: apply to, and must be observed in, all realistic models of the
518: heart satisfying certain criteria of continuity. There are also
519: implications for the treatment of cardiac arrhythmias. In
520: cardiology it is sometimes possible to develop maps showing the
521: timing of the excitation over limited regions of heart
522: \cite{stein02}. In this case, a sink might be confused for a
523: source (of the arrhythmia), and this might have implications for
524: the diagnosis of the mechanism and the choice of therapy. The
525: current work shows how partial knowledge about what is happening
526: in some regions that could be observed, might be helpful in
527: establishing properties of dynamics that could not be observed.
528: While the types of sinks we have described here have only been
529: observed in chemical media \cite{maselko89,maselko90} so far, we
530: certainly expect their existence in the cardiological domain.
531: 
532: 
533: We thank F. Chavez, M. B\"ar, S. G. Whittington and D. Sumners for
534: helpful discussions and G. Rousseau for providing numerical tools.
535: This work was supported in part by a grant from MITACS.
536: 
537: % Specify following sections are appendices. Use \appendix* if there
538: % only one appendix.
539: %\appendix
540: %\section{}
541: 
542: % If you have acknowledgments, this puts in the proper section head.
543: %\begin{acknowledgments}
544: % put your acknowledgments here.
545: %\end{acknowledgments}
546: 
547: % Create the reference section using BibTeX:
548: \begin{thebibliography}{26}
549: \expandafter\ifx\csname
550: natexlab\endcsname\relax\def\natexlab#1{#1}\fi
551: \expandafter\ifx\csname bibnamefont\endcsname\relax
552:   \def\bibnamefont#1{#1}\fi
553: \expandafter\ifx\csname bibfnamefont\endcsname\relax
554:   \def\bibfnamefont#1{#1}\fi
555: \expandafter\ifx\csname citenamefont\endcsname\relax
556:   \def\citenamefont#1{#1}\fi
557: \expandafter\ifx\csname url\endcsname\relax
558:   \def\url#1{\texttt{#1}}\fi
559: \expandafter\ifx\csname
560: urlprefix\endcsname\relax\def\urlprefix{URL }\fi
561: \providecommand{\bibinfo}[2]{#2}
562: \providecommand{\eprint}[2][]{\url{#2}}
563: 
564: 
565: \bibitem[{\citenamefont{Murray}(1989)}]{murray89}
566: \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{Murray}},
567:   \emph{\bibinfo{title}{Mathematical Biology}}
568:   (\bibinfo{publisher}{Springer-Verlag}, \bibinfo{address}{New York},
569:   \bibinfo{year}{1989});
570:   \bibinfo{note}{and references herein.}
571: 
572: \bibitem[{\citenamefont{Kapral and Showalter}(1996)}]{kapral}
573: \bibinfo{editor}{\bibfnamefont{R.}~\bibnamefont{Kapral}} \bibnamefont{and}
574:   \bibinfo{editor}{\bibfnamefont{K.}~\bibnamefont{Showalter}}, eds.,
575:   \emph{\bibinfo{title}{Chemical Waves and Patterns}}
576:   (\bibinfo{publisher}{Kluwer Academic Publishers},
577:   \bibinfo{address}{Dordrecht, Netherlands},
578:   \bibinfo{year}{1996});   \bibinfo{note}{and references herein.}
579: 
580: \bibitem[{\citenamefont{Maselko and Showalter}(1989)}]{maselko89}
581: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Maselko}} \bibnamefont{and}
582:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Showalter}},
583:   \bibinfo{journal}{Nature (London)} \textbf{\bibinfo{volume}{339}},
584:   \bibinfo{pages}{609} (\bibinfo{year}{1989}).
585: 
586: \bibitem[{\citenamefont{Maselko and Showalter}(1990)}]{maselko90}
587: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Maselko}} \bibnamefont{and}
588:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Showalter}},
589:   \bibinfo{journal}{React. Kinet. Catal. Lett.} \textbf{\bibinfo{volume}{42}},
590:   \bibinfo{pages}{263} (\bibinfo{year}{1990}).
591: 
592: \bibitem[{\citenamefont{Winfree}(2001)}]{winfree01}
593: \bibinfo{author}{\bibfnamefont{A.~T.} \bibnamefont{Winfree}},
594:   \emph{\bibinfo{title}{The Geometry of Biological Time}}
595:   (\bibinfo{publisher}{Springer-Verlag}, \bibinfo{address}{New York},
596:   \bibinfo{year}{2001}).
597: 
598: \bibitem[{\citenamefont{Stein et~al.}(2002)\citenamefont{Stein, Markowitz,
599:   Mittal, Slotwiner, Iwai, and Lerman}}]{stein02}
600: \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Stein}},
601:   \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Markowitz}},
602:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Mittal}},
603:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Slotwiner}},
604:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Iwai}}, \bibnamefont{and}
605:   \bibinfo{author}{\bibfnamefont{B.~B.} \bibnamefont{Lerman}},
606:   \bibinfo{journal}{Chaos} \textbf{\bibinfo{volume}{12}}, \bibinfo{pages}{740}
607:   (\bibinfo{year}{2002}).
608: 
609: \bibitem[{num()}]{3d}
610: \bibinfo{note}{Some aspects of cardiac arrhythmias may arise from
611: scroll wave filament instabilities which depend on the 3d nature
612: of cardiac tissue \cite{winfree87,mikhailov03}.}
613: 
614: \bibitem[{\citenamefont{Glass}(1977)}]{glass77}
615: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Glass}},
616:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{198}},
617:   \bibinfo{pages}{321} (\bibinfo{year}{1977}).
618: 
619: \bibitem[{\citenamefont{Winfree and Strogatz}(1983)}]{winfree83}
620: \bibinfo{author}{\bibfnamefont{A.~T.}~\bibnamefont{Winfree}}
621: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~H.} \bibnamefont{Strogatz}},
622:   \bibinfo{journal}{Physica D} \textbf{\bibinfo{volume}{8}},
623:   \bibinfo{pages}{35} (\bibinfo{year}{1983}).
624: 
625: \bibitem[{\citenamefont{Sumners}(1987)}]{sumners87}
626: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Sumners}},
627: \bibinfo{title}{Knots, macromolecules and chemical dynamics},
628: \bibnamefont{in}  \emph{\bibinfo{title}{Graph Theory and
629: Topology in Chemistry}},
630: \bibinfo{editor}{\bibfnamefont{R.~B.}~\bibnamefont{King}} \bibnamefont{and}
631: \bibinfo{editor}{\bibfnamefont{D.}~\bibnamefont{Rouvray}}, eds.,
632: (\bibinfo{publisher}{Elsevier},
633: \bibinfo{address}{Netherlands},
634: \bibinfo{year}{1987}), p. 3.
635: 
636: \bibitem[{\citenamefont{Cruz-White}(2003)}]{cruz-white03}
637: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Cruz-White}},
638: \bibnamefont{in}  \emph{\bibinfo{title}{Topology of Spiral Waves
639: in Excitable Media}}, (\bibinfo{publisher}{Ph.D. Thesis},
640: \bibinfo{address}{Florida State University Department of Mathematics},
641: \bibinfo{year}{2003}).
642: 
643: 
644: \bibitem[{\citenamefont{Hocking}(1988)}]{hocking88}
645: \bibinfo{author}{\bibfnamefont{J.~G.} \bibnamefont{Hocking}}
646: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.~S.} \bibnamefont{Young}},
647:   \emph{\bibinfo{title}{Topology}}
648:   (\bibinfo{publisher}{Dover}, \bibinfo{address}{New York},
649:   \bibinfo{year}{1988}).
650: 
651: 
652: \bibitem[{\citenamefont{Guillemin and Pollack}(1974)}]{guillemin74}
653: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Guillemin}},
654: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Pollack}},
655:   \bibinfo{title}{Differential Topology}
656: (\bibinfo{publisher}{Prentice Hall}, \bibinfo{address}{Englewood Cliffs, NJ},
657:   \bibinfo{year}{1974}).
658: 
659: 
660: \bibitem[{\citenamefont{Winfree}(1987)}]{winfree87}
661: \bibinfo{author}{\bibfnamefont{A.~T.} \bibnamefont{Winfree}},
662:   \emph{\bibinfo{title}{When Time Breaks Down}}
663:   (\bibinfo{publisher}{Princeton University Press},
664: \bibinfo{address}{Princeton, NJ},
665:   \bibinfo{year}{1987}).
666: 
667: \bibitem[{\citenamefont{Alonso et~al. }(2003)}]{mikhailov03}
668: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Alonso}},
669: \bibinfo{author}{\bibfnamefont{F.} \bibnamefont{Sagu\'es}} \bibnamefont{and}
670:   \bibinfo{author}{\bibfnamefont{A.~S.}~\bibnamefont{Mikhailov}},
671:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{299}},
672:   \bibinfo{pages}{1722} (\bibinfo{year}{2003}).
673: 
674: \bibitem[{\citenamefont{Zaritski and Pertsov}(2002)}]{zaritski02}
675: \bibinfo{author}{\bibfnamefont{R.~M.} \bibnamefont{Zaritski}} \bibnamefont{and}
676:   \bibinfo{author}{\bibfnamefont{A.~M.}~\bibnamefont{Pertsov}},
677:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{66}},
678:   \bibinfo{pages}{066120} (\bibinfo{year}{2002}).
679: 
680: 
681: \bibitem[{num()}]{numerics}
682: \bibinfo{note}{We solve Eq. (\ref{equdyn}) numerically using an algorithm that
683:   automatically adjusts the time step to achieve an effcient simulation while
684:   controling the error in the solution \cite{rousseau00}.}
685: 
686: \bibitem[{\citenamefont{Davydov et~al.}(2002)\citenamefont{Davydov, Manz,
687:   Steinbock, and M{\"u}ller}}]{davydov02}
688: \bibinfo{author}{\bibfnamefont{V.~A.} \bibnamefont{Davydov}},
689:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Manz}},
690:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Steinbock}},
691:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~C.}
692:   \bibnamefont{M{\"u}ller}}, \bibinfo{journal}{Europhys. Lett.}
693:   \textbf{\bibinfo{volume}{59}}, \bibinfo{pages}{344} (\bibinfo{year}{2002}).
694: 
695: \bibitem[{det()}]{shell}
696: \bibinfo{note}{It is computationally convenient to carry out the calculations
697: in a thin spherical shell constructed from a cubic rectangular
698: grid. The shell thickness is small compared to characteristic
699: wavelengths and diffusion lengths so that the small shell
700: thickness does not influence the spiral wave dynamics.}
701: 
702: \bibitem[{det()}]{details}
703: \bibinfo{note}{The initial conditions are as in Ref.~\cite{chavez01} with
704:   spherical coordinates $\Delta\theta=\pi/2$ and $\Delta\phi=\pi$.}
705: 
706: \bibitem[{\citenamefont{Yagisita et~al.}(1998)\citenamefont{Yagisita, Mimura,
707:   and Yamada}}]{yagisita98}
708: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Yagisita}},
709:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Mimura}}, \bibnamefont{and}
710:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Yamada}},
711:   \bibinfo{journal}{Physica D} \textbf{\bibinfo{volume}{124}},
712:   \bibinfo{pages}{126} (\bibinfo{year}{1998}).
713: 
714: \bibitem[{ran()}]{random_details}
715: \bibinfo{note}{For a Gaussian distribution of $\beta$ in boxes of volume
716:   $0.5^3$ with mean 0.95 and standard deviation 0.013, the diameter of the
717:   spiral core meander is roughly 10 space units.}
718: 
719: \bibitem[{\citenamefont{Winfree}(1991)}]{winfree91}
720: \bibinfo{author}{\bibfnamefont{A.~T.} \bibnamefont{Winfree}},
721:   \bibinfo{journal}{Chaos} \textbf{\bibinfo{volume}{1}}, \bibinfo{pages}{303}
722:   (\bibinfo{year}{1991}).
723: 
724: %\bibitem[{\citenamefont{Gelfand and Tsetlin}(1960)}]{gelfand60}
725: %\bibinfo{author}{\bibfnamefont{I. M.}~\bibnamefont{Gelfand}} \bibnamefont{and}
726: %  \bibinfo{author}{\bibfnamefont{M. L.}~\bibnamefont{Tsetlin}},
727: %  \bibinfo{journal}{Proc. Acad. Sci. USSR} \textbf{\bibinfo{volume}{131}},
728: %  \bibinfo{pages}{1242} (\bibinfo{year}{1960}).
729: 
730: \bibitem[{\citenamefont{Krinsky and Agladze}(1983)}]{krinsky83}
731: \bibinfo{author}{\bibfnamefont{V. I.}~\bibnamefont{Krinsky}} \bibnamefont{and}
732:   \bibinfo{author}{\bibfnamefont{K. I.}~\bibnamefont{Agladze}},
733:   \bibinfo{journal}{Physica D} \textbf{\bibinfo{volume}{8}},
734:   \bibinfo{pages}{50} (\bibinfo{year}{1983}).
735: 
736: \bibitem[{\citenamefont{Ermakova et al.}(1960)}]{ermakova86}
737: \bibinfo{author}{\bibfnamefont{E. A.}~\bibnamefont{Ermakova}},
738: \bibinfo{author}{\bibfnamefont{V. I.}~\bibnamefont{Krinsky}},
739: \bibinfo{author}{\bibfnamefont{A. V.}~\bibnamefont{Panfilov}},
740:  \bibnamefont{and}
741:   \bibinfo{author}{\bibfnamefont{A. M.}~\bibnamefont{Pertsov}},
742:   \bibinfo{journal}{Biofizika} \textbf{\bibinfo{volume}{31}},
743:   \bibinfo{pages}{318} (\bibinfo{year}{1986}).
744: 
745: 
746: \bibitem[{\citenamefont{Gong and Christini}(2003)}]{gong03}
747: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Gong}} \bibnamefont{and}
748:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Christini}},
749:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{90}},
750:   \bibinfo{pages}{088302} (\bibinfo{year}{2003});
751: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Brusch}},
752:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Nicola}}, \bibnamefont{and}
753:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{B{\"a}r}},
754:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{92}},
755:   \bibinfo{pages}{089801} (\bibinfo{year}{2004}).
756: 
757: \bibitem[{\citenamefont{Biktashev et~al.}(2001)}]{biktashev01}
758:   \bibinfo{author}{\bibfnamefont{V.~N.} \bibnamefont{Biktashev}}
759:   \bibinfo{author}{\bibfnamefont{A.~V.} \bibnamefont{Holden}},
760:   \bibinfo{author}{\bibfnamefont{S.~F.} \bibnamefont{Mironov}},
761:   \bibinfo{author}{\bibfnamefont{A.~M.} \bibnamefont{Pertsov}},
762: \bibnamefont{and}
763:   \bibinfo{author}{\bibfnamefont{A.~V.} \bibnamefont{Zaitsev}},
764:   \bibinfo{journal}{Int. J. Bifurcation Chaos Appl. Sci. Eng.}
765:   \textbf{\bibinfo{volume}{11}},
766:   \bibinfo{pages}{1035} (\bibinfo{year}{2001}).
767: 
768: \bibitem[{\citenamefont{Vanag and Epstein}(2001)}]{vanag01}
769: \bibinfo{author}{\bibfnamefont{V.~K.} \bibnamefont{Vanag}} \bibnamefont{and}
770:   \bibinfo{author}{\bibfnamefont{I.~R.} \bibnamefont{Epstein}},
771:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{294}},
772:   \bibinfo{pages}{835} (\bibinfo{year}{2001}).
773: 
774: \bibitem[{\citenamefont{Biktashev and Holden}(1995)}]{biktashev95}
775: \bibinfo{author}{\bibfnamefont{V.~N.} \bibnamefont{Biktashev}}
776: \bibnamefont{and}
777:   \bibinfo{author}{\bibfnamefont{A.~V.} \bibnamefont{Holden}},
778:   \bibinfo{journal}{Chaos, Sol. Fract.} \textbf{\bibinfo{volume}{5}},
779:   \bibinfo{pages}{575} (\bibinfo{year}{1995}).
780: 
781: \bibitem[{\citenamefont{Mikhailov et~al.}(1994)\citenamefont{Mikhailov,
782:   Davydov, and Zykov}}]{mikhailov94}
783: \bibinfo{author}{\bibfnamefont{A.~S.} \bibnamefont{Mikhailov}},
784:   \bibinfo{author}{\bibfnamefont{V.~A.} \bibnamefont{Davydov}},
785:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{V.~S.} \bibnamefont{Zykov}},
786:   \bibinfo{journal}{Physica D} \textbf{\bibinfo{volume}{70}},
787:   \bibinfo{pages}{1} (\bibinfo{year}{1994});
788: \bibinfo{author}{\bibfnamefont{V.~A.} \bibnamefont{Davydov}},
789:   \bibinfo{author}{\bibfnamefont{V.~S.} \bibnamefont{Zykov}},
790:   \bibinfo{author}{\bibfnamefont{A.~S.} \bibnamefont{Mikhailov}},
791:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.~K.}
792:   \bibnamefont{Brazhnik}}, \bibinfo{journal}{Radiophys. Quantum Electron.}
793:   \textbf{\bibinfo{volume}{31}}, \bibinfo{pages}{574} (\bibinfo{year}{1988}).
794: 
795: \bibitem[{\citenamefont{Rousseau and Kapral}(2000)}]{rousseau00}
796: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Rousseau}} \bibnamefont{and}
797:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Kapral}},
798:   \bibinfo{journal}{Chaos} \textbf{\bibinfo{volume}{10}}, \bibinfo{pages}{812}
799:   (\bibinfo{year}{2000}).
800: 
801: \bibitem[{\citenamefont{Chavez et~al.}(2001)\citenamefont{Chavez, Kapral,
802:   Rousseau, and Glass}}]{chavez01}
803: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Chavez}},
804:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Kapral}},
805:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Rousseau}}, \bibnamefont{and}
806:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Glass}},
807:   \bibinfo{journal}{Chaos} \textbf{\bibinfo{volume}{11}}, \bibinfo{pages}{757}
808:   (\bibinfo{year}{2001}).
809: 
810: 
811: \end{thebibliography}
812: 
813: 
814: \end{document}
815: %
816: % ****** End of file template.aps ******
817: