1: \documentstyle[aps,multicol,prl,psfig]{revtex}
2:
3: \draft
4:
5: \begin{document}
6:
7: \newcommand{\be}{\begin{eqnarray}}
8: \newcommand{\ee}{\end{eqnarray}}
9:
10: \title{Kinematic reduction of reaction-diffusion fronts
11: with multiplicative noise: \\
12: Derivation of stochastic sharp-interface equations}
13:
14: \author{A. Rocco$^{1,2}$, L. Ram\'{\i}rez-Piscina$^{3}$ and
15: J. Casademunt$^{2}$}
16: \address{$^{1}$CWI, Postbus 94079, 1090 GB Amsterdam, The Netherlands}
17: \address{$^{2}$Departament d'Estructura i Constituents de la Mat\`{e}ria,
18: Facultat de F\'{\i}sica, Universitat de Barcelona, Av. Diagonal 647,
19: E-08028 Barcelona, Spain}
20: \address{$^{3}$Departament de F\'{\i}sica Aplicada, Universitat
21: Polit\`{e}cnica de Catalunya, Avenida Dr. Gregorio Mara\~n\'on 50,
22: E-08028 Barcelona, Spain}
23:
24:
25: \date{\today}
26:
27: \maketitle
28:
29: \begin{abstract}
30: We study the dynamics of generic reaction-diffusion
31: fronts, including pulses and chemical waves,
32: in the presence
33: of multiplicative noise.
34: We discuss the connection between the reaction-diffusion Langevin-like
35: field equations
36: and the kinematic (eikonal) description in terms of a
37: stochastic moving-boundary or sharp-interface approximation.
38: We find that the effective noise is additive and we relate its
39: strength to the noise parameters in the original field equations, to
40: first order in noise strength, but including a partial resummation to
41: all orders which captures the singular dependence on the microscopic
42: cutoff associated to the spatial correlation of the noise.
43: This dependence is essential for a quantitative and qualitative
44: understanding of fluctuating fronts, affecting both scaling properties
45: and nonuniversal quantities. Our results
46: predict phenomena such as
47: the shift of the transition point between the pushed and pulled
48: regimes of front propagation,
49: in terms of the noise parameters, and the corresponding transition
50: to a non-KPZ universality class.
51: We assess the
52: quantitative validity of the results in several examples
53: including equilibrium fluctuations,
54: kinetic roughening,
55: and the noise-induced pushed-pulled transition, which is predicted and
56: observed for the
57: first time.
58: The analytical predictions
59: are successfully tested against rigorous results and show excellent
60: agreement with
61: numerical simulations of reaction-diffusion field equations with multiplicative
62: noise.
63: \end{abstract}
64:
65: \pacs{PACS number(s): 05.40.-a,05.45.-a,82.40.Ck,47.54.+r}
66:
67:
68: \section{Introduction}
69: The dynamics of localized solutions in the form of fronts or
70: pulses in reaction-diffusion systems has received a great deal of attention
71: for a long time in the context of nonequilibrium extended systems
72: \cite{Cross}. Examples of fronts formed by stable regions
73: invading unstable or metastable ones
74: are found in a large variety of physical, chemical or biological systems,
75: and have been studied in great detail \cite{vanSaarloos,Collet90,Ebert00a}.
76: In the context
77: of excitable media and chemical waves, extended pulses do also exhibit a rich
78: phenomenology\cite{Kapral93,Mikhailov95,tyson,Meron,Mikhailov94}.
79: As opposed to fronts, excitation waves or pulses are such
80: that the region behind them
81: eventually returns to the same linearly stable state that is
82: ahead.
83: In this case, in dimensions higher than one there may be open ends of the
84: pulse region which give rise to spiral (2d) or scroll (3d) waves.
85: There has been much interest in the study of such objects from the fundamental
86: point of view of pattern forming dynamics, but also because of their potential
87: applications in biological systems, such as in cardiac tissue \cite{heart}.
88: In this paper we will not consider the case of
89: open ends, so unless otherwise specified, we will refer indistinctly to
90: fronts and pulses under the common term of 'fronts'.
91:
92: One aspect which has received increasing interest in recent years
93: has been the effect of fluctuations both of internal and external origin
94: in the dynamics and in the
95: roughening properties of fronts
96: \cite{Riordan95,Armero96,Armero98,Kessler98,Rocco00,Tripathy00,Tripathy01,Rocco01,vulpiani}.
97: More recently the effect of noise in pattern
98: forming dynamics of chemical waves has been
99: fostered by the development of the experimental capability to introduce
100: external spatio-temporal
101: noise in a controlled way in different photosensitive
102: nonlinear chemical reactions, through the optical
103: projection of computer-designed spatio-temporal fluctuations in the
104: local illumination conditions, acting as a stochastic
105: control parameter
106: \cite{Sendina-Nadal97,Sendina-Nadal98,Sendina-Nadal00,Showalter98}.
107: From a theoretical point of view, a common starting point to study
108: fluctuations
109: is in terms of master equations defining evolution of reacting and diffusing
110: particles in a lattice\cite{Riordan95}.
111: The connection of this microscopic level of description
112: to the mesoscopic one in terms of Langevin field
113: equations has proven a rather
114: subtle issue, in connection with the distinction of the so-called pushed vs
115: pulled fronts \cite{Ebert00a}. Only recently a complete understanding
116: of the instrinsically
117: different nature of these two types of fronts, and the corresponding
118: consequences
119: concerning the presence of cutoffs \cite{Brunet97,Kessler98b,Panja01,Moro01},
120: and the effects of fluctuations
121: has started
122: to emerge \cite{Armero96,Armero98,Rocco00,Tripathy00,Tripathy01}.
123: It has been shown, for instance, that pulled fronts define a new
124: universality class of kinetic roughnening different
125: from the Kardar-Parisi-Zhang
126: universality class \cite{Tripathy00,Tripathy01}.
127: On the other hand, it has been shown that intrinsic noise
128: at the microscopic level may induce a morphological instability at the
129: macroscopic level of description\cite{Kessler98}.
130: In this paper we will be mostly concerned
131: with the macroscopic description of pushed fronts with fluctuations, but
132: also on how this description incorporates the transition to pulled fronts
133: induced by the noise itself.
134:
135: In the absence of noise, and in the appropriate limit,
136: the description of fronts and pulses defined by reaction-diffusion
137: field equations can be reduced by means of a moving boundary approximation
138: to a kinematic description in terms of much simpler local equations
139: \cite{Mikhailov94,hakim}. This procedure, which is mathematically well grounded
140: in the framework of the so-called inertial manifold reduction,
141: can be carried out systematically, for instance, through asymptotic matching
142: techniques using the front thickness as a small parameter defining a singular
143: perturbation problem \cite{Ebert00b}. This
144: is commonly used to relate macroscopic
145: interface equations to order-parameter or phase-field descriptions in
146: many problems involving interface dynamics, such as in solidification
147: \cite{Wang93,Karma96},
148: viscous fingering \cite{folch} etc.
149: In the case of chemical pulses in excitable media
150: in the limit of weak excitability, this leads to a local equation
151: where the normal velocity of the pulse is a constant plus a correction
152: proportional to curvature \cite{Mikhailov94}.
153: This is often called the eikonal equation.
154: In the case of open ends, a similar local equation can be derived for
155: the motion of the endpoint of the pulse \cite{Mikhailov94,hakim}.
156: For pushed fronts it can be shown that for smooth, long wavelength deformations
157: of the front, the separation of time scales between the soft deformation modes and
158: the internal degrees of freedom of the fields leads naturally to the same
159: eikonal equation. For pulled fronts, however, this separation of time scales
160: does not exist, the relaxation being algebraic instead of exponential,
161: and a local moving-boundary approximation is not justified \cite{Ebert00b}.
162:
163: The kinematic description in terms of eikonal-like equations is a very
164: useful approximation from both a theoretical and a practical point of
165: view. In the context of the study of universality of fluctuation
166: properties \cite{stanley,Krug97}
167: for instance, it leads naturally to the relevant effective universal
168: description of a broad class of systems in terms of the KPZ equation
169: \cite{Kardar86}.
170: It is also very useful for numerical simulation purposes, since it
171: avoids resolving the fine structure of the front and the bulk degrees
172: of freedom, which become
173: irrelevant,
174: dealing only with the kinematic degrees of freedom of an object
175: of reduced dimension.
176:
177: When noise is present in the original field equations, however, the
178: situation is not so clear. Stochastic eikonal equations have
179: proven useful
180: in a phenomenological description of the dynamics of
181: pulses and spiral waves in photosensitive chemical systems with external
182: noise imposed on the illumination conditions
183: \cite{Sendina-Nadal98,Sendina-Nadal00}.
184: Such description, however, relied on some fitting parameters and some
185: uncontrolled hypothesis on the way the noise must enter
186: the kinematic
187: equations. These results, together with the fact that the statistical
188: properties of the noise present in those experiments are fully controlled,
189: clearly call for a more 'microscopic' derivation of
190: stochastic kinematic equations corresponding to Langevin reaction-diffusion
191: field equations with multiplicative noise, with no free parameters.
192: The connection between bulk and interface fluctuations has been worked
193: out so far only for equilibrium fluctuations in coarse-grained,
194: Ginzburg-Landau-like
195: equations\cite{Hohenberg,Safran94,karmaprl,rappel}.
196: In such cases, the sharp-interface
197: limit can be taken at the level of the free-energy, and the existence of a
198: fluctuation-dissipation theorem then yields the proper way to incorporate
199: thermal fluctuations into the effective interface equations. However, in many
200: nonequilibrium systems, for instance in the context of reaction-diffucion problems,
201: a free-energy or generically a Liapunov functional may
202: not exist and no fluctuation-dissipation relation may be invoked for external
203: noise. In such cases the connection between the bulk description and the effective
204: interface description in presence of fluctuations must be worked out at the
205: level of the dynamical equations.
206:
207: The purpose of this paper is to
208: address this point by deriving stochastic
209: eikonal equations, including the complete specification of the noise statistics
210: in connection to that of the noise in the mesoscopic field equations.
211: In particular we shall focus on the singular dependence on the spatial
212: cutoff when noise is multiplicative, and its importance in the quantitative
213: description of the statistical properties of the front fluctuations.
214: The predictions will be tested against numerical simulations of
215: reaction-diffusion equations, and also in cases where exact
216: results are available concerning the spectrum of interface fluctuations.
217: We will also see that the stochastic eikonal
218: equation derived is consistent with
219: the scenario of the pushed-pulled transition, and that, changing the
220: spatial cutoff or the noise intensity may have effects such drastic
221: as changing the universality
222: class of kinetic roughening of the front
223: through that transition.
224:
225: Although our obtention of the stochastic eikonal equation here presented
226: is not a first-principles rigorous derivation, we will provide sufficient
227: evidence to conclude that the result is the correct one to lowest order
228: in noise intensity, including the singular dependence on the noise
229: correlation cutoff (which involves a partial resumation to all orders),
230: within the long time and length scales limits
231: which are inherent to the kinematic descrition itself.
232: However,
233: it does not apply to situations in which the front dynamics is
234: nonlocal, such as solidification fronts or viscous fingers, where a
235: different type of formulation is appropriate even in the deterministic case
236: \cite{Wang93,Karma96,folch,cinca}.
237:
238: \section{Kinematic reduction for generic
239: reaction-diffusion systems}
240:
241: Let us consider a vectorial field ${\boldmath{\mbox{$\phi$}}}({\bf
242: x},t)$ with $N$ components
243: ${\boldmath{\mbox{$\phi$}}}({\bf x}) \equiv \phi_1({\bf x}),...,
244: \phi_N({\bf x})$ in a $d-$dimensional space with
245: ${\bf x} \equiv x_1,...,x_d$, which
246: obeys a reaction-diffusion equation with
247: multiplicative noise of the form:
248: \be
249: \frac{\partial \boldmath{\mbox{$\phi$}}}{\partial t} =
250: \hat{D} \nabla^2 \boldmath{\mbox{$\phi$}} +
251: {\bf f}(\boldmath{\mbox{$\phi$}})
252: + \varepsilon^{1/2} {\bf g}(\boldmath{\mbox{$\phi$}})
253: \eta({\bf x},t), \label{multfront}
254: \ee
255: where $\eta({\bf x},t)$ is a gaussian noise with zero mean and
256: correlation given by
257: \be
258: \langle \eta({\bf x},t) \eta({\bf x}^{\prime},t^{\prime}) \rangle =
259: 2 \lambda^{-d} C(|{\bf x}-{\bf x}^{\prime}|/\lambda) \delta(t-t^{\prime}).
260: \label{noisecorr}
261: \ee
262:
263: We take a one-component noise for simplicity, as the natural case when it
264: originates in fluctuations of a single control parameter. The generalization
265: of the formalism to multicomponent noise is straightforward.
266: Notice the asymmetry with which we treat the spatial and temporal
267: correlator of the noise. As we will see, this reflects a nontrivial issue related
268: to the intrinsically different nature of the
269: white noise limit in space as opposed to time. We have taken in Eq.
270: (\ref{noisecorr}) the noise as delta-correlated in time.
271: This temporal white noise limit is well behaved, once a prescription
272: for the multiplicative noise term in
273: Eq. (\ref{multfront}) has been chosen.
274: For external fluctuations the physically relevant prescription
275: is to consider the white noise as the limit of some properly
276: defined correlated noise. This corresponds to the well-known Stratonovich
277: prescription \cite{Gardiner}.
278: On the other hand, the spatial noise must always be defined as
279: colored, its white (uncorrelated) limit being intrinsically ill-defined. Hence
280: the notation with the function $\lambda^{-d}C(r/\lambda)$ in the correlator,
281: meaning a general correlation function, dependent on some correlation length
282: $\lambda$, which in the limit $\lambda \rightarrow 0$ is such that
283: $\lambda^{-d}C(r/\lambda) \rightarrow \delta(r)$.
284: The fact that the spatial continuum limit
285: $\lambda \rightarrow 0$ does not exist in the Stratonovich interpretation
286: is thus reflected in the fact that, even if $\lambda$ is much smaller than
287: any other length scale in the problem, the existence of such microscopic cutoff
288: always shows up in the quantitative description of the large scale behavior and
289: cannot be reabsorbed in a redefinition of parameters.
290: That is, the noise cannot be considered as effectively white in space,
291: regardless of how small the noise correlation length is.
292:
293: The multiplicative noise term present in Eq. (\ref{multfront}) has an
294: average value different from zero. Applying Novikov theorem
295: \cite{Novikov64}, we get
296: \be
297: \varepsilon^{1/2} \langle {\bf g}(\boldmath{\mbox{$\phi$}}) \eta({\bf x},t)
298: \rangle = \varepsilon \lambda^{-d}
299: C(0) \langle {\bf G(\boldmath{\mbox{$\phi$}})} \rangle
300: \ee
301: where
302: \be
303: G_i(\boldmath{\mbox{$\phi$}}) \equiv \sum_j \frac{\partial
304: g_i(\boldmath{\mbox{$\phi$}})}{\partial
305: \phi_j} g_j(\boldmath{\mbox{$\phi$}}). \label{product}
306: \ee
307: This suggests to separate the average contribution from the
308: multiplicative noise term and rewrite Eq. (\ref{multfront}) in terms of a
309: renormalized potential and a zero average noise,
310: \be
311: \frac{\partial \boldmath{\mbox{$\phi$}}}{\partial t} =
312: \hat{D} \nabla^2 \boldmath{\mbox{$\phi$}} +
313: {\bf h}(\boldmath{\mbox{$\phi$}}) + \varepsilon^{1/2}
314: \boldmath{\mbox{$\Omega$}}(\boldmath{\mbox{$\phi$}},
315: {\bf x},t), \label{main}
316: \ee
317: where
318: \be
319: {\bf h}(\boldmath{\mbox{$\phi$}}) =
320: {\bf f}(\boldmath{\mbox{$\phi$}}) + \varepsilon \lambda^{-d}
321: C(0) {\bf G}(\boldmath{\mbox{$\phi$}}) \label{functh}
322: \ee
323: and
324: \be
325: \boldmath{\mbox{$\Omega$}}(\boldmath{\mbox{$\phi$}}, {\bf x},t) =
326: {\bf G}(\boldmath{\mbox{$\phi$}})\eta({\bf x},t)
327: - \varepsilon^{\frac{1}{2}} \lambda^{-d} C(0) {\bf G}(\boldmath{\mbox{$\phi$}}).
328: \ee
329: Here the new noise $\boldmath{\mbox{$\Omega$}}$ has zero average.
330: Notice that this decomposition corresponds to transform Eq. (\ref{multfront}) into
331: its equivalent It\^o stochastic equation in the white noise limit.
332: Accordingly, the stochastic term $\Omega$ reduces to
333: \be
334: \lim_{\lambda \rightarrow 0} \boldmath{\mbox{$\Omega$}}(\boldmath{\mbox{$\phi$}},
335: {\bf x},t) = {\bf G}(\boldmath{\mbox{$\phi$}})\eta_I({\bf x},t)
336: \ee
337: where $\eta_I({\bf x},t)$ is now a white noise in the It\^o interpretation.
338: The deterministic term ${\bf h}$ includes thus noise effects through the so-called
339: Stratonovich term that has been added to ${\bf f}$. These noise effects on the
340: deterministic part of Eq. (\ref{main}) depend on a 'dressed' noise intensity
341: $\varepsilon_{\lambda}$ which contains the singular dependence on the
342: the spatial cutoff in the form
343: of the 'bare' $\varepsilon$ as $\varepsilon_{\lambda} \equiv \varepsilon C(0)
344: \lambda^{-d}$.
345: The important point is that the spatial white noise limit in
346: the continuum equations is mathematically well-defined for an It\^o noise. This
347: has been proven rigorously for relatively broad classes of equations (see for
348: instance Refs.\cite{Gyongy98,Nualart99}). On the contrary, this cannot be the case
349: for a Stratonovich noise, as it is obvious from the singular
350: dependence on $\lambda$. The practical implications of these fact are that,
351: while the singular contribution of $\lambda$ in the Stratonovich term must be kept
352: explicitly, the dependence on
353: $\lambda$ contained in
354: $\boldmath{\mbox{$\Omega$}}$ will indeed be weak (nonsingular),
355: and indeed negligible if $ \lambda$ is much smaller than the other length
356: scales in the problem, in particular the front thickness. Therefore, the
357: splitting
358: of the Stratonovich noise by means of a deterministic term plus an It\^o noise
359: has the virtue to isolate the singular dependence on the microscopic cutoff.
360: This will also define a useful partial resummation of orders
361: in $\varepsilon_{\lambda}$. After this decomposition, the theory will thus
362: be correct to first order in $\varepsilon$, to all orders in $\varepsilon/\lambda$
363: and to order $\lambda^0$ for the regular dependence of
364: $\boldmath{\mbox{$\Omega$}}$ in the noise correlation length.
365:
366: Eq. (\ref{main}) will be our starting point. We are interested in representing the
367: dynamics of a front obeying Eq. (\ref{main}) as
368: the evolution of a $d-1$ surface.
369: In this effective dynamics we assume the details of the front structure (at scales
370: of order of or smaller than the front thickness) to be unimportant.
371: In what follows we will consider the evolution of a 1-dimensional front embedded
372: in a 2-dimensional system. In the procedure we are going to apply we will write
373: the evolution equation (\ref{main}) in the curvilinear
374: coordinates defined by the
375: 1d front in the sharp-interface limit, and obtain the evolution equation for this
376: front as a solvability problem
377: with the basic assumption that curvature and noise be small perturbations.
378:
379: Before proceeding with the formal derivation, let us first point out some
380: subtleties related to the stochastic case as opposed to the deterministic
381: one. In the latter, it is customary to define a curvilinear coordinate
382: system $(s,r)$ in which $r=0$ stands for the curve representing the front
383: position, which can be associated for instance to a level curve of the appropriate
384: field. The scheme assumes that the front thickness is small compared to the
385: radius of curvature, and that the relaxation of the internal degrees of
386: freedom of the front is fast compared to the time scale of the long wavelength
387: deformations of the front.
388: When noise is present in the field equation, the appropriate curvilinear
389: coordinate system cannot be associated to level curves, which are very rough
390: at length scales smaller than the front thickness. On the other hand, at larger
391: scales, which are the ones we are interested in, a coarse
392: grained description makes perfect sense. This is actually implicit in the
393: very idea of the stochastic eikonal equation. One can think of different
394: schemes to explicitly define such coarse-grained description, all of them
395: equivalent. However, since the rest of the derivation cannot be carried
396: out explicitly in full rigor in any of those, and since the result is
397: expected to be
398: independent of the details of the definition of the coarse-graning, we will
399: proceed more or less heuristically.
400: In essence this is a reformulationn of the approach introduced for the
401: derivation of the diffusive wandering of fronts in one dimension
402: discussed in Ref.\cite{Armero98}. There, the basic idea was that only the
403: low-frequency components of the noise are responsible for the front wandering
404: so the high-frequency components can be implicitly integrated out. The effect
405: of the high-frequency components of the noise is thus to renormalize the
406: mean front profile. As a consequence, they renormalize the front velocity, and
407: also the diffusion coefficient.
408: More precisely, this means in our case that the
409: high-frequency renormalization is carried out by the Stratonovich term in the
410: function ${\bf h}$. Once this term is explictly extracted, the remainder is an
411: It\^o multiplicative noise. Then the high-frequency components
412: of this term are irrelevant and can be averaged out, while the low-frequency
413: components will be responsible for the roughnening of the front at scales
414: larger than the coarse-graining length.
415: In the case of 1d fronts, the resulting
416: diffusion coefficient for the front wandering has been rederived rigorously
417: in Ref.\cite{Rocco00}. Unfortunately, the identification of the collective
418: variable in 1d cannot be immediately generalized to higher dimensions, so
419: we must still rely on a less precise formulation and check the consistency with
420: rigorous results and numerical simulations \it a posteriori\rm.
421:
422: After the above considerations, we can make the following theoretical
423: construction. We assume we have solved the
424: field equations with noise without any approximation. We now
425: coarse grain the fields with some local
426: average procedure, both in time and space, and use the coarse-grained fields to
427: define a curvilinear
428: coordinate system based, for instance, in terms of level curves at any time.
429: At short length and time scales
430: this coordinate system is smooth and, in principle,
431: we could write the full field equation
432: (still with the bare fields) in these curvilinear coordinates.
433: In the absence of noise,
434: an expansion in the front thickness
435: would unambiguosly yield the terms which are dominant in the range where the
436: eikonal equation is devised for, namely for sufficiently long length scales
437: (small curvatures) and long time scales.
438: Terms such as second derivatives on $s$ and the time derivative would
439: automatically drop out of the description.
440: In the presence of noise stochastic case, this is not so automatic unless
441: the fields themselves are coarse-grained so that they are also
442: sufficiently smooth in space and time.
443: We assume that, for the coarse-grained fields,
444: the order of the different terms in the front thickness will be the same than
445: for the deterministic case, when such expansion makes sense (excluding pulled
446: fronts, for instance). We claim that such assumption is the one implicit in the
447: very idea of the existence of a stochastic kinematic formulation of the front
448: dynamics.
449: Then, for the coarse-grained fields, Eq. (\ref{main}) is expected to reduce,
450: in analogy to the deterministic case, to
451: \be
452: \hat{D} \frac{\partial^2 \boldmath{\mbox{$\phi$}}}{\partial r^2} +
453: \hat{D} \kappa(s,t) \frac{\partial\boldmath{\mbox{$\phi$}}}{\partial r}
454: + {\bf h}(\boldmath{\mbox{$\phi$}}) +
455: v_n(s,t) \frac{\partial \boldmath{\mbox{$\phi$}}} {\partial r}
456: &+&
457: \varepsilon^{1/2} \boldmath{\mbox{$\Omega$}}(\boldmath{\mbox{$\phi$}}; r,s;t)
458: = 0,
459: \label{wtho}
460: \ee
461: where $\kappa$ is the local curvature and $v_n$ the normal velocity of the front.
462: This normal velocity provides the evolution of the curvilinear
463: coordinates in which
464: Eq. (\ref{main}) takes the form of Eq. (\ref{wtho}), and is the
465: fundamental quantity we are interested
466: in. The noise term in Eq. (\ref{wtho}) must also be considered as
467: coarse-grained, with the high-frequency high-wavenumber
468: components integrated out.
469:
470: At this point of the derivation it is useful to consider the 1d
471: problem which
472: corresponds to the zeroth order of the eikonal description. This is defined
473: by neglecting the curvature and the fluctuating term in Eq. (\ref{wtho}): \be
474: 0 = \hat{D} \frac{\partial^2 \boldmath{\mbox{$\phi_0$}}}{\partial r^2}
475: + \bar{v}(\varepsilon_\lambda) \frac{\partial \boldmath{\mbox{$\phi_0$}}}{\partial r} +
476: {\bf h}(\boldmath{\mbox{$\phi_0$}}). \label{fi}
477: \ee
478: This is the eigenvalue problem that gives the renormalized velocity $\bar{v}$
479: in the 1d problem as obtained in Refs. \cite{Armero96,Armero98}. Note that this
480: equation does contain noise effects through the high-frequency renormalization
481: provided by the Stratonovich term.
482: In fact, the effective velocity $\bar{v}(\varepsilon_\lambda)$ resulting
483: from Eq. (\ref{fi}) has contributions from all orders in the dressed noise
484: intensity
485: $\varepsilon_\lambda$. An explicit first order in $\varepsilon_\lambda$
486: approximation will be given in the Appendix.
487: In the problem defined by Eq. (\ref{wtho}) curvature and
488: fluctuations will be taken as small perturbations.
489: Hence we assume for the field $\boldmath{\mbox{$\phi$}}$ and the
490: velocity $v_n$ expansions of the form
491: \be
492: \boldmath{\mbox{$\phi$}}(r,s,t) = \boldmath{\mbox{$\phi_0$}}(r)
493: + \delta\boldmath{\mbox{$\phi$}}(r,s,t),
494: \label{pe}
495: \ee
496: \be
497: v_n(s,t) = \bar{v}(\varepsilon_\lambda)
498: + \beta(\varepsilon_\lambda) \kappa(s,t)
499: + \delta v(s,t), \label{scurv}
500: \ee
501: where $\boldmath{\mbox{$\phi_0$}}(r)$ and $\bar{v}(\varepsilon_\lambda)$
502: are the solution of the
503: 1d problem of Eq. (\ref{fi}). The term $\beta \kappa(s,t)$ is a
504: curvature correction and $\delta v(s,t)$ describes fluctuations.
505: Linearizing in both perturbations $\delta{\boldmath{\mbox{$\phi$}}}(r,s)$
506: then verifies
507: \be
508: 0 = \hat{\Gamma} \delta \boldmath{\mbox{$\phi$}}
509: + (\beta \kappa + \delta v) \frac{\partial \boldmath{\mbox{$\phi_0$}}}{\partial r}
510: + \hat{D} \kappa \frac{\partial \boldmath{\mbox{$\phi_0$}}}{\partial r}
511: + \varepsilon^{1/2} \boldmath{\mbox{$\Omega$}}(\boldmath{\mbox{$\phi_0$}}; r,s;t),
512: \label{deltafi}
513: \ee
514: where
515: \be
516: \hat{\Gamma} = \hat{D} \frac{\partial^2}{\partial r^2}
517: + \bar{v}(\varepsilon_\lambda)
518: \frac{\partial}{\partial r} + \left.\frac{\partial {\bf h}}
519: {\partial \boldmath{\mbox{$\phi$}}} \right|_{ \boldmath{\mbox{$\phi$}}
520: = \boldmath{\mbox{$\phi_0$}}}.
521: \label{gamma}
522: \ee
523:
524: Taking the derivative of Eq. (\ref{fi}) with respect to $r$, it is a
525: simple matter to prove that
526: \be
527: {\bf u_0} = \frac{\partial \boldmath{\mbox{$\phi_0$}}}{\partial r}
528: \ee
529: is the right eigenvector of $\hat{\Gamma}$ with zero eigenvalue.
530: Due to the non-hermiticity of $\hat{\Gamma}$, finding
531: an analytic expression for the left eigenvector, ${\bf u^0}$, is not
532: trivial. Notice that because of the vectorial character of the field, the
533: simple expression obtained in Ref. \cite{Armero98}
534: for scalar fields is
535: in general not applicable. Nevertheless, the corresponding eigenvector
536: can always be obtained at least numerically.
537:
538: Now, taking Eq. (\ref{deltafi}) and performing the scalar product with
539: ${\bf u^0}$, we obtain
540: \be
541: \kappa ({\bf u^0},\hat{D}{\bf u_0}) + \beta \kappa ({\bf u^0},
542: {\bf u_0}) + \varepsilon^{1/2} ({\bf u^0},\boldmath{\mbox{$\Omega$}}
543: (\boldmath{\mbox{$\phi_0$}}; r,s;t) ) + \delta v(s,t) ({\bf u^0}, {\bf u_0}) = 0
544: \ee
545: where the scalar product is defined by
546: \be
547: ({\bf f},{\bf g}) = \sum_i \int dr f_i(r) g_i(r)
548: \ee
549: Owing to the independence of the two first-order perturbations
550: (curvature and fluctuations), we get
551: \be
552: \beta(\varepsilon_\lambda) = - \frac{({\bf u^0}, \hat{D}{\bf u_0})}
553: {({\bf u^0},{\bf u_0})} \label{beta}
554: \ee
555: and
556: \be
557: \delta v(s,t)= -
558: \varepsilon^{1/2}
559: \frac{({\bf u^0},\boldmath{\mbox{$\Omega$}}(\boldmath{\mbox{$\phi_0$}}; r,s;t) )}
560: {({\bf u^{0}},{\bf u_0})}.
561: \label{deltav}
562: \ee
563: The stochastic process (\ref{deltav})
564: is not white since the
565: high-frequency components of $\boldmath{\mbox{$\Omega$}}$ have been integrated
566: out by the coarse-graining
567: procedure. However, we can now restore them harmlessly by replacing
568: $\boldmath{\mbox{$\Omega$}}$ with the original multiplicative white
569: noise process. By doing so, we are modifying the part which is
570: not intended to be accounted for by the very eikonal description, while
571: the treatment is simpler. Analogously, once the dependence on the cut-off
572: $\lambda$ has been explicitely worked out,
573: we can take the process (\ref{deltav})
574: as delta-correlated in space. We will explicitly check the limit of validity
575: of the
576: eikonal description here proposed when scales comparable to the front thickness
577: are reached, in the sections below.
578:
579: The resulting stochastic eikonal equation with the explicit dependence
580: on the original noise parameters then takes the form
581: \be
582: v_n (s,t)= \bar{v}(\varepsilon_{\lambda}) +
583: \beta(\varepsilon_{\lambda}) \kappa(s,t) +
584: D_f^{1/2}(\varepsilon,\varepsilon_{\lambda})\zeta(s,t) \label{jy}
585: \ee
586: where $\bar{v}(\varepsilon_{\lambda})$ is defined by Eq.(\ref{fi}),
587: $\beta(\varepsilon_{\lambda})$ is given by Eq.(\ref{beta}),
588: and the noise $\zeta(s,t)$ is a zero mean Gaussian white process with
589: \be
590: \langle \zeta(s,t) \zeta(s^{\prime},t^{\prime}) \rangle =
591: 2 \delta (s-s^{\prime}) \delta(t-t^{\prime}),
592: \ee
593: which follows from the
594: statistical properties of $\Omega$ with
595: \be
596: D_f(\varepsilon,\varepsilon_{\lambda}) =
597: \varepsilon \frac{\int dr \sum_{i,j} u_i^{0} u_{j(0)}
598: g_i(\boldmath{\mbox{$\phi_0$}})
599: g_j(\boldmath{\mbox{$\phi_0$}})}{({\bf u^0},{\bf u_0})^2}.
600: \label{tere}
601: \ee
602: Note that the dependence of $D_f$ on the dressed noise intensity
603: $\varepsilon_\lambda$ comes from the dependence on the same quantity of
604: $\boldmath{\mbox{$\phi_0$}}$ (and hence of ${\bf u_0}$ and ${\bf
605: u^0}$) as solution of the renormalized problem of Eq. (\ref{fi}).
606:
607: The above equations constitute the main result of the first part of the
608: paper.
609: Although the derivation is not rigorous because the coarse-graining could not
610: be carried out explicitly, the result is appealing
611: from the theoretical point of view in that it separates the problem into
612: an effective deterministic one, where the original field equations are
613: modified by additional deterministic terms which depend on noise
614: parameters, plus an additive noise which would be the necessary one to
615: describe the wandering of the problem in one dimension. For the
616: renormalized one-dimensional deterministic problem, therefore, the two present
617: perturbations, curvature and noise, decouple from each other.
618:
619: An important point to emphasize here is the separate dependence of the result on
620: two noise parameters, namely
621: $\varepsilon$ and $\varepsilon_{\lambda}$. While the renormalized velocity
622: and the coefficient of the curvature depend solely on $\varepsilon_{\lambda}$,
623: the effective noise intensity $D_f$ depends separately on both. This
624: clearly illustrates how the ultraviolet cutoff is an additional parameter of
625: the problem when the noise is multiplicative, in correspondence to the fact that
626: the continuum limit is not well defined for noise delta-correlated in
627: space.
628: It is also important to remark that our derivation
629: procedure is expected to be valid for small noise intensity $\varepsilon$,
630: but contains all orders in $\varepsilon_{\lambda}$. This was already the case
631: in the one-dimensional derivation of Ref.\cite{Armero98}, where the small
632: noise approximation was phrased in terms of a separation of time
633: scales. The connection between that scale separation, the coarse-graining
634: procedure and the small noise expansion has been recently clarified in
635: Ref.\cite{Rocco01}, where a rigorous derivation of the result of
636: Ref.\cite{Armero98} has been presented for the case of a single-component field
637: in one dimension, in terms of suitable projection
638: techniques.
639: Unfortunately, that rigorous derivation is based on the identification
640: of a specific collective coordinate which has no simple extension to higher
641: dimensions. Nevertheless, the fact that this approximate procedure has
642: proven
643: correct in $1d$ gives further support to our main result above, which is
644: not claimed to be rigorously proven.
645: In the following sections we will check this prediction
646: against analytical results and
647: numerical simulations of the full reaction-diffusion equations in
648: explicit examples. We will see that, the dependence on the cutoff $\lambda$
649: is essential not only for a quantitative description of the problem, but
650: is crucial to predict nontrivial phenomena such the transition to pulled
651: fronts, in which the whole eikonal description fails.
652: This failure of the
653: present description is signaled by the vanishing of the effective noise
654: intensity $D_f$ when that point is reached. In fact, $D_f$ is linear in
655: $\varepsilon$ to lowest order, but has a complicated dependence on
656: $\varepsilon_{\lambda}$.
657: As we will explicitly see, the partial resummation
658: of orders in $\varepsilon_{\lambda}$ captures important physical features of the
659: problem. For instance, it allows the
660: non-monotonic dependence and eventual vanishing of the front diffusion coefficient
661: $D_f$.
662: We expect the pushed-pulled transition to occur exactly at this point.
663: Another qualitative change captured by the above resummation is the destruction
664: of the front itself, associated to the fact that the front thickness may diverge
665: in some circumstances. This phenomenon is the signaled by a dicergence of $D_f$
666: at some finite value of $\varepsilon_{\lambda}$.
667:
668: Although the predictions above are expected to be correct for
669: $\epsilon_{\lambda} \sim 1$ as long as $\epsilon \ll 1$, in practice this
670: may be limited by the fact that $\bar{v}(\varepsilon_{\lambda})$ and the
671: Goldstone modes are not in general analytically known.
672: In such case one can rely on numerical resolutions of the eigenvalue problem posed
673: by Eq. (\ref{fi}), or alternatively one can find $\bar{v}(\varepsilon_{\lambda})$
674: as a systematic expansion in powers of $\varepsilon_{\lambda}$ as described
675: in the Appendix. On the other hand, it is worth remarking that
676: ${\beta}$ can be a nontrivial function of $\varepsilon_{\lambda}$ only
677: for multicomponent fields, that is for pulses. For one-component fields
678: (fronts) the coefficient $\beta$ is not renormalized by noise.
679:
680: We remark that this derivation is
681: meaningful only when pulses are involved or, in the case of fronts,
682: when the relevant dynamical regime is the
683: pushed one. In the last Section of this paper, we shall give more
684: details about the main differences between pushed and pulled fronts,
685: and about their different response to noise. For the time being, we
686: point out that the lack of time-scale separation between the
687: relaxation of the zero mode and the other eigenmodes of the
688: spectral operator for pulled fronts prevents in general from
689: constructing a local equation for the interface motion \cite{Ebert00b}.
690: Examples of reaction-diffusion systems which do not admit local kinematic
691: descriptions are the phase-field formulations of solidification
692: \cite{Wang93,cinca} or viscous fingering \cite{folch}. In those cases a nonlocal
693: interface equation does exist so an extension of our derivation
694: is in principle feasible. This may be particularly interesting in cases such
695: as in Refs. \cite{cinca,folch} were the relevant fluctuations may be external.
696:
697:
698: \section{Kinetic roughening and connection to equilibrium fluctuations}
699:
700: In the previous Section, we have derived
701: a stochastic sharp interface
702: approximation for a generic RD system with multiplicative noise.
703: A pictorial description of a noisy front
704: is given in Fig. \ref{front}
705: \begin{figure}
706: \centerline{{\psfig{figure=pgplotphinew.ps,angle=-90,height=2.4in}}}
707: \caption{\small{An example of a noisy front with a level curve which
708: defines the precise location of the front.}} \label{front}
709: \end{figure}
710: We see in this figure that noise in the RD system induces
711: fluctuations in the front
712: shape, thus generating roughening of the sharp interface
713: that should emerge at the eikonal
714: level of description. The identification of universality
715: classes of kinetic roughening will come naturally at this level of description.
716: In particular we will establish the connection with the universality classes
717: defined by the Kardar-Parisi-Zhang (KPZ) Equation \cite{Kardar86}
718: and by the Edwards-Wilkinson (EW) Equation \cite{Edwards82}.
719:
720: The stochastic eikonal equation (\ref{jy}) is written in instrinsic, rotation
721: invariant form. For the purposes of scaling theory it is convenient
722: to write it in cartesian coordinates. The front location is then given as
723: $y=h(x,t)$. Retaining only the relevant nonlinear terms in the Renormalization
724: Group (RG) sense, we then recover the KPZ equation
725: \be
726: \frac{\partial h}{\partial t} = \nu \frac{\partial^2 h}{\partial x^2}
727: + \frac{\lambda}{2} \left(\frac{\partial h}{\partial x}\right)^2 + \mu(x,t),
728: \ee
729: with
730: \be
731: \langle \mu(x,t) \mu(x^{\prime},t^{\prime}) \rangle =
732: 2 D_{\rm KPZ} \delta (x-x^{\prime}) \delta(t-t^{\prime}). \label{corrkpz}
733: \ee
734: The KPZ parameters turn out to be related to the eikonal ones as
735: \be
736: D_{\rm KPZ} = D_f, \;\;\;\;\;\;\;\; \nu = \beta, \;\;\;\;\;\;\;\;
737: \lambda = \bar{v}. \label{id}
738: \ee
739:
740: In the special case of $\bar{v}=0$, the EW equation is obtained,
741: \be
742: \frac{\partial h}{\partial t} = \nu \frac{\partial^2 h}{\partial x^2}
743: + \mu(x,t), \label{EW}
744: \ee
745: with
746: \be
747: \langle \mu(x,t) \mu(x^{\prime},t^{\prime}) \rangle =
748: 2 D_{\rm EW} \delta (x-x^{\prime}) \delta(t-t^{\prime}),
749: \ee
750: and again
751: \be
752: D_{\rm EW} = D_f, \;\;\;\;\;\;\;\; \nu = \beta. \label{ident}
753: \ee
754: These equations are well known to be the paradigm for many different
755: growth processes \cite{stanley,Krug97}.
756: Even if the microscopic dynamics of the system under
757: study may correspond to different equations of motion for the
758: respective interfaces or surfaces,
759: nevertheless the KPZ and EW Equations do capture the universal
760: features of the system, namely the scaling properties of fluctuations.
761:
762: Usually, such effective equations cannot be derived from the original
763: microscopic description of the particular systems and are introduced on
764: a phenomenological basis, relying on the claim of universality within
765: a RG framework. Nonuniversal quantities such as prefactors of scaling
766: functions, affected for instance by the noise intensity in the interface
767: equation cannot be derived. In our case, we are able to compute the noise
768: intensity in the eikonal equation so we can also predict the nonuniversal
769: prefactors if the noise is known in the reaction-diffusion level of description.
770: %%@@
771: For instance, not only the scaling of the interface roughness with
772: system size can be predicted, but also the actual values of average interface
773: roughness in terms of the original microscopic parameters of the RD model
774: are worked out.
775:
776: As a test of our derivation we will now check consistency with equilibrium
777: fluctuation theory. The connection between bulk thermal fluctuations
778: and fluctuations of the interface between thermodynamical phases can be
779: established rigorously in the case of equilibrium fluctuations.
780: This is possible because a free energy functional
781: does exist and the sharp-interface limit can be performed at the level of the
782: free energy itself.
783: Then, the fluctuations can be obtained independently from the free energy
784: at each level of description (either bulk or interface fluctuations),
785: consistently with the fluctuation-dissipation theorem.
786: The important difference here is that no dynamical
787: equation must be invoked but only equilibrium properties.
788: On the contrary, in the more general case
789: where
790: there is no fluctuation-dissipation theorem and not even a free energy
791: functional, we must rely on dynamical equations. In the case of equilibrium
792: fluctuations, however, we must reproduce the
793: known correct result.
794: As we will see below, this case falls in the EW universality class.
795:
796: The general solution of the EW Equation, Eq. (\ref{EW}), is known \cite{Krug97}.
797: Consider the interface $h(x,t)$ and its discrete Fourier Transform
798: $\hat{h}_q(t)$, defined through
799: \be
800: h(x,t) = \sum_q \hat{h}_q(t) \exp(iqx).
801: \ee
802: It is possible to show \cite{Krug97} that the long time limit of the spectrum
803: $S(q,t) = \langle |\hat{h}_q(t)|^2 \rangle$ is given by the expression
804: \be
805: \lim_{t \rightarrow \infty} S(q,t) = \frac{1}{L} \frac{D_{\rm EW}}{\nu q^2}.
806: \label{lims}
807: \ee
808: Our strategy now will be to calculate explicitly the spectrum
809: (\ref{lims}) in terms of the coefficients predicted for
810: our eikonal equation, and
811: then show that the resulting expression coincides with the independent
812: result that can be obtained from equilibrium fluctuation theory.
813: Hence we insert an additive noise
814: in the original RD system, that is in setting
815: $g=1$:
816: \be
817: \frac{\partial \phi}{\partial t} = D \nabla^2 \phi +
818: F(\phi) + \varepsilon^{1/2} \eta({\bf x},t),
819: \label{rdv0}
820: \ee
821: with
822: \be
823: \langle \eta({\bf x},t) \eta({\bf x}^{\prime},t^{\prime}) \rangle =
824: 2 \delta({\bf x}-{\bf x}^{\prime}) \delta(t-t^{\prime}).
825: \ee
826: We consider a $F(\phi)$ with a symmetric double-well form,
827: {\it i.e.}
828: the deterministic part of Eq. (\ref{rdv0}) has a $1d$ solution
829: with zero velocity
830: $\phi=\phi_0(x)$. This is what corresponds to the usual time-dependent
831: Gizburg-Landau Langevin equation for a nonconserved order parameter
832: (Model A in the Hohenberg-Halperin classification \cite{Hohenberg}), where
833: noise intensity must be identified as $\varepsilon=k_BT$.
834: Since $g=1$, our expression for the noise intensity at the sharp interface
835: level takes the simpler form
836: \be
837: D_f=\frac{\varepsilon}{\displaystyle \int_{-\infty}^{\infty}dx\left
838: (\frac{\partial \phi_0} {\partial x}\right)^2},
839: \ee
840: which, according to (\ref{lims}) and performing the identifications of Eq.
841: (\ref{ident}) with $\beta=D$, produces
842: \be
843: \lim_{t \rightarrow \infty}
844: \langle |\hat{h}_q(t)|^2 \rangle = \frac{\varepsilon}{L D \int_{-\infty}
845: ^{\infty}dx\left (\frac{\partial \phi_0} {\partial x}\right)^2} \frac{1}{q^2}.
846: \label{resu}
847: \ee
848: On the other hand, we can take the sharp-interface limit on the free
849: Ginzburg-Landau free energy. The calculation is standard
850: (see for instance Ref.\cite{Safran94}) and yields the interface free
851: energy
852: \be
853: F_I=\sigma \int dx \sqrt{1+\left(\frac{\partial h}{\partial x} \right)^2},
854: \ee
855: where the parameter $\sigma$ is identified as the interfacial tension
856: and can be evaluated from the bulk free energy of the system \cite{Safran94}
857: as
858: \be
859: \sigma = D \int dx \left(\frac{\partial \phi_0}{\partial x} \right)^2,
860: \label{sigma}
861: \ee
862: where $\phi_0$ is the corresponding kink solution.
863: For soft (long wavelength) deformations of the interface,
864: the excess free energy reads
865: \be
866: \Delta F_I \approx \frac{\sigma}{2} \int dx \left(\frac{\partial h}
867: {\partial x} \right)^2,
868: \ee
869: and the corresponding stationary spectrum of fluctuations, consistent with
870: the fluctuation-dissipation theorem takes the form
871: \cite{Safran94}
872: \be
873: \langle |\hat{h}_q|^2 \rangle = \frac{k_B T}{L \sigma q^2}.
874: \ee
875: Using Eq.(\ref{sigma}), this expression yields the same
876: result of Eq.(\ref{resu}).
877: This proves that the front roughening obtained from our derivation
878: is exact in the case of equilibrium fluctuations, and by extension
879: in the additive noise case.
880: We have shown this
881: for the case of a non-conserved order parameter. In the conserved case (Model
882: B of Ref.\cite{Hohenberg}), the projection to a sharp interface
883: description yields
884: a non-local equation and therefore lies outside the validity of our theory.
885: Similarly, equilibrium fluctuations have also been studied in both
886: sharp-interface \cite{karmaprl} and phase-field \cite{rappel,cinca}
887: formulation in the context of solidification,
888: which also yields non-local interface dynamics. The universality classes in
889: those cases are not well established.
890:
891:
892:
893: \section{Application to a prototype model of front propagation}
894:
895: To illustrate our general theory we
896: consider here as an example the propagation of a scalar front
897: \be
898: \frac{\partial \phi}{\partial t} = D \nabla^2 \phi +
899: F(\phi,a) + \varepsilon^{1/2} g(\phi) \eta, \label{eqmain}
900: \ee
901: with the noise correlator defined as in Eq. (\ref{noisecorr}).
902: We specify our prototype model through the following definitions:
903: \be
904: &F(\phi)=\phi(1-\phi)(\phi+a) \label{proto} \\
905: &g(\phi)=\phi(1-\phi), \label{protog}
906: \ee
907: and we will consider a front of the $\phi=1$ state invading the $\phi=0$ one.
908: The constant $a$ is a control parameter.
909: As it is well known for fronts without fluctuations,
910: the deterministic force given
911: by (\ref{proto}) leads to different modes of front propagation depending on the
912: value of $a$ \cite{vanSaarloos}. The front velocity depends also on $a$.
913: The choice of the coupling function (\ref{protog}) for the multiplicative noise
914: term of Eq. (\ref{eqmain}) is the simplest one that preserves the
915: stationary states $\phi=0$ and
916: $\phi=1$. The fact that the noise term vanishes in the two asymptotic states
917: prevents nucleation phenomena in the invaded state. The form of $g(\phi)$ is
918: also such that the multiplicative noise term arises naturally as the
919: fluctuation of the control parameter $a$. Moreover, for the prototype model
920: here proposed, the corresponding function $h(\phi)$ defined in
921: Eq.(\ref{functh}) which appears in the renormalized equation takes the same
922: functional form as $f(\phi)$, only with renormalized coefficients. This
923: allows a simpler analytical treatment and intepretation of the
924: results.
925:
926: \subsection{The 1d case revisited}
927:
928: This prototype model has already been analyzed
929: for 1d in Refs.
930: \cite{Armero96,Armero98},
931: where in the regime $ -\varepsilon_\lambda < a < 1/2-\varepsilon_\lambda$
932: it is proven to result in a renormalized average velocity
933: \be
934: \bar{v}=\frac{2a + 1}{\sqrt{2(1-2\varepsilon_\lambda)}} \label{vavera}
935: \ee
936: and in the diffusion coefficient
937: \be
938: D_f = \varepsilon \;\frac{\int_{-\infty}^{\infty} d \xi \;
939: e^{2 \bar{v} \xi}\;(d \phi_0/d\xi)^2 \;g^2(\phi_0)}
940: {\left[\int_{-\infty}^{\infty} d \xi \;e^{\bar{v} \xi}
941: \;(d\phi_0/d\xi)^2\right]^2}. \label{formula}
942: \ee
943: Here $\phi_0$ is the
944: solution of the zeroth order equation:
945: \be
946: \frac{d^2 \phi_0}{d \xi^2} + \bar{v} \frac{d \phi_0}{d \xi} +
947: h(\phi_0)=0. \label{partialeqm}
948: \ee
949: Notice the dependence of both these quantities on $\lambda$ through
950: the $\varepsilon_\lambda$ parameter.
951: Specifically about the diffusion coefficient, the
952: first order dependence in $\varepsilon$ is apparent, while the
953: functions present in the integral and defined through
954: (\ref{partialeqm}) contain all orders in $\varepsilon_\lambda$. This is in
955: contrast to the renormalized velocity, which depends solely on
956: $\varepsilon_\lambda$. This means that, both $\varepsilon$ and $\lambda$
957: can be determined independently from separate measurements
958: of both the ballistic and the
959: diffusive components of the front propagation. This is quite remarkable since
960: it provides indirect means of measurement of the (microscopic) noise,
961: which may not be directly accessible in many cases.
962:
963: As an illustrative and well-controlled example, we have explicitly tested
964: the prediction of the dependence
965: on $\lambda$ of both velocity and diffusion coefficient with direct numerical
966: simulation of the RD equation.
967: The first set of results is shown in Fig. \ref{figlambdavel}. Here
968: data from a simulation of the RD equation with the reaction
969: term (\ref{proto}) are reported and compared with the theoretical
970: prediction (\ref{vavera}) for two different values of $\lambda=1$
971: and $\lambda=5$, in some dimensionless units of the simulation.
972:
973: \begin{figure}
974: \centerline{{\psfig{figure=figlambdavelnew.eps,angle=0,height=2.4in}}}
975: \caption{\small{Change with $\lambda$ of the average renormalized
976: velocity. Values of the parameters are: $a = 0.1$, $\varepsilon=0.1$,
977: $\lambda$=1,5. The theoretical values of the corresponding average
978: velocities as from Eq. (\ref{vavera}), 0.948 and 0.866, are also
979: plotted.}} \label{figlambdavel}
980: \end{figure}
981:
982: Notice that the result is indeed sensitive to the
983: microscopic cutoff $\lambda$ of the noise, even though this length is
984: significantly smaller
985: than the front thickness, of the order of 25, in the same units.
986:
987: Regarding the diffusion coefficient (\ref{formula}), we measured the mean
988: square displacement of the front position. If we define the front
989: position as
990: \be
991: z(t) = \int_{x_0}^{\infty} dx \phi(x,t),
992: \ee
993: then the diffusion coefficient (\ref{formula}) is related to the mean
994: square displacement
995: \be
996: \Delta = \sqrt{\langle z^2 \rangle - \langle z \rangle^2}
997: \ee
998: as
999: \be
1000: \Delta^2(t) \sim 2 D_f t
1001: \ee
1002: In Fig. \ref{figlambdadiff} the quantity $\Delta(t)$ is plotted.
1003: \begin{figure}
1004: \centerline{{\psfig{figure=figlambdadiffnew.eps,angle=0,height=2.4in}}}
1005: \caption{\small{Change with $\lambda$ of the diffusion
1006: coefficient for the front wandering in 1d. Parameters have the same
1007: values as in Fig. \ref{figlambdavel}.}}\label{figlambdadiff}
1008: \end{figure}
1009: The values of the parameters are the same as in Fig. \ref{figlambdavel},
1010: while the values of $\lambda$ run from 1 to 15. The diffusive
1011: behavior and the dependence on the value of $\lambda$ is manifest.
1012:
1013: In this particular case of 1d, a more systematic derivation of the diffusion
1014: coefficient has been reported in Ref.\cite{Rocco00}.
1015: By proper identification of the
1016: natural collective variable which describes the front wandering as strictly
1017: (not just asymptotically) diffusive, it has been
1018: shown that, in fact, the result first found in Ref.\cite{Armero98} is rigorous
1019: to first order in $\varepsilon$. Together with the case of equilibrium
1020: fluctuations, this is the second rigorous test of our general theory.
1021:
1022:
1023: \subsection{The 2d case}
1024:
1025: Let us now consider the propagation of a front in the prototype model of Eqs.
1026: (\ref{eqmain}, \ref{proto}, \ref{protog}) in 2d.
1027: We have already shown that in two dimensions the eikonal equation
1028: reduces to either the EW equation or the KPZ equation, depending on
1029: whether the velocity of a planar front is zero or non-zero respectively.
1030: In order to make a numerical check of our theory we consider then
1031: the simplest case of zero velocity, with expected EW scaling.
1032: This case corresponds to the choice $a=-1/2$ in Eq. (\ref{proto}).
1033: As seen explicitly in Eq.(\ref{vavera}) it turns out that the
1034: renormalized velocity is
1035: also zero, since the noise does not break the symmetry associated to the
1036: double-well form of the deterministic potential. It is important to remark
1037: that, unlike the Ginzburg-Landau model discussed in Section III for the
1038: equilibrium case, the noise is now multiplicative, and no
1039: fluctuation-dissipation theorem can be invoked.
1040: Hence the first-principles derivation of the fluctuation spectrum in the
1041: sharp-interface description is no longer available.
1042:
1043: We thus rely on the dynamical equation
1044: \be
1045: \label{sym}
1046: \frac{\partial \phi}{\partial t} = D \nabla^2 \phi + \phi(1-\phi)
1047: (\phi-\frac{1}{2}) + \varepsilon^{1/2}\phi(1-\phi) \eta({\bf x},t),
1048: \ee
1049: where the noise $\eta$ is defined through the usual correlator of
1050: Eq.(\ref{noisecorr}).
1051:
1052: Now, our basic goal is to connect this level of description with the
1053: eikonal level, determining thereby the noise corrections
1054: to this equation. Therefore we
1055: assume a stochastic eikonal equation of the form
1056: \be
1057: \left\{
1058: \begin{array}{l}
1059: v_n(s) = - \beta \kappa + D_f^{1/2} \zeta(s,t)\\
1060: \langle \zeta(s,t) \zeta(s^{\prime},t^{\prime}) \rangle =
1061: 2 \delta (s-s^{\prime}) \delta(t-t^{\prime})
1062: \end{array}
1063: \right.
1064: \ee
1065: where the coefficients $\beta$ and $D_f$ are given by
1066: \be
1067: \beta = D, \;\;\;\;\;\;\;\;\;\;
1068: D_f = \frac{\int dx u^{0} u_{0} g^2(\phi_0)}{(u^{0},u_{0})^2}.
1069: \label{betaD}
1070: \ee
1071: Notice that we have directly taken into account that the bare as well as
1072: the noise-renormalized velocities are both zero
1073: (see for example \cite{Armero98}). Also notice that the
1074: renormalization of the curvature term is absent for
1075: both $\varepsilon$ and $\varepsilon_\lambda$ expansions due to
1076: the fact that we are now dealing with a front ($N=1$).
1077:
1078: Now, in order to calculate $D_f$, we need to specify the solution of
1079: the 1d model. This is known and for $a=-1/2$ is given by
1080: \be
1081: \phi_0 = \frac{1}{2}(1-\tanh kx), \;\;\;\;\;\;\;\;\;\;\; k=\frac{1}{2 \sqrt{2}},
1082: \ee
1083: and
1084: \be
1085: u^{0} = u_{0} = \frac{d \phi_0}{dx}.
1086: \ee
1087: The integral in (\ref{betaD}) can then be computed exactly and gives:
1088: \be
1089: \label{exact}
1090: D_f = \varepsilon \frac{9}{35} \sqrt{\frac{2}{1-2 \varepsilon/\lambda^2}}.
1091: \ee
1092: The above result clearly illustrates the different treatment of the parameters
1093: $\varepsilon$ and
1094: $\varepsilon/\lambda^2$. The result is first order in $\varepsilon$ and
1095: contains
1096: all orders in $\varepsilon/\lambda^2$. It is interesting to remark that the
1097: partial resummation of all orders in $\epsilon$ captures important
1098: nonpertubative phenomena. For instance, the divergence of $D_f$ at
1099: $\varepsilon/\lambda^2=1/2$, reflects the fact that at this point the
1100: front itself is destroyed, or, equivalently, the front thickness becomes
1101: infinite. This is equivalent to reaching a critical point, except that this
1102: is not the equilibrium one because the noise is multiplicative.
1103: Remarkably,
1104: our result for additive noise case does not capture that feature
1105: because it is only first order in the noise strength.
1106: In the case with an asymmetric double-well potential, which has a finite front
1107: velocity, it was explicitly checked numerically in Ref.\cite{Armero98} that
1108: the diffusion coefficient of the front has a non-monotonic dependence with
1109: $\varepsilon/\lambda^2$. Most importantly it vanishes at a finite value of
1110: $\varepsilon/\lambda^2$ which corresponds exactly to the point were the front
1111: reaches the pushed-pulled transition. Again we see that the resummation of
1112: orders $\varepsilon/\lambda^2$ captures important physical information
1113: (see discussion in Section V).
1114:
1115: We now come back to the numerical test of the EW scaling of our particular
1116: symmetric case, with the identification (\ref{ident}).
1117: Accordingly, we can rewrite the complete power spectrum as it is known
1118: theoretically \cite{Krug97} in terms of the coefficients that we just
1119: calculated, and compare it with data from a direct simulation of the field
1120: model Eq.(\ref{sym}).
1121:
1122: From \cite{Krug97} the power spectrum $S$ reads,
1123: \be
1124: S(q,t) = \frac{1}{L} \frac{D_{\rm EW}}{\nu q^2}(1-\exp(-2 \nu q^2 t)).
1125: \label{powspectr}
1126: \ee
1127: In terms of the parameters in the original field equation this yields
1128: \be
1129: S(q,t) = \frac{1}{L} \frac{D_{f}}{D q^2}(1-\exp(-2 D q^2 t)),
1130: \label{predicted}
1131: \ee
1132: where $D_f$ is given by Eq.(\ref{exact}). Accordingly, the fluctuations of
1133: the front position in the RD model Eq.(\ref{eqmain}) for length scales
1134: larger than the front thickness itself must obbey the spectrum defined
1135: by Eq.(\ref{predicted}). It is worth remarking that the prediction is not
1136: only for the universal features, namely the shape of the scaling function
1137: and the exponents, but for the actual absolute values of the spectrum.
1138: We have performed simulations of the field RD model with a correlator of
1139: the form
1140: \be
1141: \langle \eta({\bf x},t) \eta({\bf x}^{\prime},t^{\prime}) \rangle =
1142: \frac{1}{\lambda^2}
1143: (1-|{\bf x}-{\bf x}^{\prime}|/\lambda)
1144: (1-|\vec{y}-\vec{y}^{\prime}|/\lambda)
1145: \Theta (1-|{\bf x}-{\bf x}^{\prime}|/\lambda)
1146: \Theta (1-|\vec{y}-\vec{y}^{\prime}|/\lambda)
1147: \delta(t-t^{\prime})
1148: \ee
1149: This corresponds to assuming that at any time, the noise takes the same
1150: value in a square of side $\lambda$, uncorrelated with the neighboring
1151: squares. This is done for simplicity but no significant dependence is expected
1152: on the details of the shape of the spatial correlation provided $\lambda$ is
1153: kept smaller than front thickness.
1154:
1155: We have studied the fluctuations of the internal level curve of the front
1156: $\phi=1/2$.
1157: In Fig. \ref{figspectr1} we show the scaling region, with the correct slope
1158: and location of the curve. More remarkably, Fig. \ref{figspectr2} shows
1159: the measured spectra for the simulation of the field equations compared to
1160: the prediction given by Eq.(\ref{predicted}).
1161:
1162: \begin{figure}
1163: \centerline{{\psfig{figure=spectr1new.eps,angle=0,height=2.4in}}}
1164: \caption{\small{Numerical data from 2d simulation of the RD
1165: field equations for the protorype model
1166: with $a=-1/2$ and analytical prediction for the
1167: power spectrum (\ref{powspectr}) in the scaling region.
1168: The time is here 1000, and the parameters of the simulation are
1169: $L=100$, $\varepsilon=5$, $\lambda=5$.}} \label{figspectr1}
1170: \end{figure}
1171:
1172: \begin{figure}
1173: \centerline{{\psfig{figure=spectr2new.eps,angle=0,height=2.4in}}}
1174: \caption{\small{Numerical data from 2d simulation of the RD field
1175: equations for the protorype model with $a=-1/2$ and analytical prediction
1176: for the power spectrum (\ref{powspectr}). The three sets of data points refer
1177: to times 50, 250 and 500. The parameters of the simulation are
1178: $L=500$, $\varepsilon=1$, $\lambda=2$. The value of $q_c$ corresponds to
1179: a wavelength of the order of the front thickness. The analytical prediction
1180: is only intended for $q < q_c$.}} \label{figspectr2}
1181: \end{figure}
1182:
1183: It should be stressed that in this comparison there is no free parameter.
1184: It is also interesting to observe the deviations from the prediction at
1185: length scales smaller than the front thickness. In this high-q region the
1186: data also collapse but not to EW scaling. An estimate of the exponent
1187: $\alpha$ in this region is around $3/2$.
1188:
1189: For values of $a$ such that the front has a finite velocity, one would
1190: expect that the scaling would be given by that of the KPZ universality
1191: class. Although the scaling function in that case is not exactly known,
1192: the prediction of our eikonal equation is expected to fit the data for
1193: the corresponding RD model also without free parameters. We have not checked
1194: this case because it is obviously more involved and less conclusive
1195: because of the practical difficulties to reach the scaling regime already
1196: at the eikonal level of description \cite{Beccaria94}.
1197:
1198: From a practical point of view it is to be remarked that the noise intensity
1199: cannot be increased arbitrarily in a simulation without destroying
1200: the front itself.
1201: This can be easily seen in a numerical simulation.
1202: Although the noise vanishes asymptotically in the
1203: stationary states $\phi = 0$ and $\phi = 1$, if noise is sufficiently strong
1204: it may be capable to nucleate the other state in the region not too far from
1205: the front. We have found that this effect is more pronounced in the
1206: region behind the front $\phi = 1$. For a given $\lambda$ there will typically
1207: be a maximum value of $\varepsilon$ after which the front is essentially
1208: destroyed. If we increase $\lambda$ the effect is milder. In order to
1209: make the front roughening appreciable in not too large system sizes, it is
1210: thus convenient to have a moderately large $\lambda$, which in turn will
1211: allow larger values of $\varepsilon$. Typical values that we have
1212: considered are in the range of $\lambda=2,5$ n units for which the front thickness is of the
1213: order of $25$.
1214:
1215:
1216: \section{The pushed - pulled transition}
1217: As we have being mentioning so far, two classes of fronts must be
1218: distinguished from a dynamical point of view, the so-called 'pushed' and
1219: 'pulled' fronts \cite{Ebert00a,Ebert00b}.
1220: The simplest is pushed case, in which
1221: the front propagation
1222: depends on the full non-linear structure of the equation of
1223: motion and the front is said to be 'pushed' by its internal part. This is
1224: usually the case when the invaded state is locally stable.
1225: On the contrary, if the invaded state is unstable, it can happen that
1226: the relevant dynamics takes place in the semi-infinite
1227: leading edge region ahead of the front itself. Then the
1228: propagation of the front is governed by the growth and spreading of
1229: linear perturbations in that region which 'pulls' the front. In this case
1230: the linearization about the
1231: unstable state accounts for its dynamical behavior \cite{Ebert00a}, but
1232: there is degeneracy of propagating velocities\cite{vanSaarloos}. In the
1233: present context, the most important distinction between the two situations
1234: is that, while for pushed fronts, the relaxation of bulk modes is exponential,
1235: for pulled fronts it is algebraic, as a result from the fact that the
1236: linearized operator describing perturbations around the stationary propagating
1237: mode is gapless. This means that for pulled fronts, there is no natural
1238: time-scale separation which allows for the decoupling of the interface modes
1239: from the bulk ones. Our derivation, and the whole idea of a kinematic
1240: moving-boundary approximation, is not valid for pulled fronts\cite{Ebert00b}.
1241: Nevertheless, we will show that our theory does predict correctly its
1242: failure at the pushed-pulled transition.
1243:
1244: The intrinsic differences between pushed and pulled fronts also show
1245: dramatically in the statistics of fluctuations. In 1d, for instance the
1246: front wandering turns from difussive (pushed fronts) to subdiffusive
1247: (pulled fronts) \cite{Armero98,Rocco00}. In turn, pulled fronts in 2d with
1248: multiplicative noise have been found to belong to a different universality
1249: class than the ordinary KPZ \cite{Tripathy00,Tripathy01}, as opposed to
1250: the KPZ scaling of pushed fronts.
1251:
1252: In Ref.\cite{Armero98} it was already observed that the diffusion coefficient
1253: of fronts in 1d was a nonmonotonic function of $\varepsilon/\lambda$ which
1254: vanished at a finite value of that parameter. After that point, subdiffusive
1255: behavior was found. The picture was confirmed and completed in Ref.
1256: \cite{Rocco00}. Here we want to stress that the point where that transition
1257: occurs corresponds {\it exactly} to the pushed-pulled transition. In fact,
1258: the renormalized equation defined by Eq.(\ref{main}) for the protoype model
1259: takes the same form as the original one, with renormalized coefficients.
1260: The deterministic equation has the transition to the pulled regime at
1261: $a=1/2$. The same transition in the renormalized equation does occur at
1262: $\tilde{a}=1/2$ where $\tilde{a}= a + \varepsilon_\lambda$. It is thus clear
1263: that, for parameter values of the deterministic equation in the pushed regime,
1264: increasing noise intensity $\varepsilon$ or decreasing $\lambda$ will
1265: imply a transition to the pulled regime. While the front velocity will not
1266: experience any dramatic effect, the fluctuations (wandering in 1d or
1267: roughening in 2d) will be dramatically affected, since the universality
1268: class will change. In Fig. \ref{transition} we show the change from
1269: diffusive to subdiffusive behavior induced solely by a change in the
1270: effective noise instensity $\varepsilon_\lambda$.
1271: \begin{figure}
1272: \centerline{{\psfig{figure=transitionnew.eps,angle=0,height=2.4in}}}
1273: \caption{
1274: \small{The pushed-pulled transition in 1d. The system size is $L=2000$ and
1275: averages have been carried out on $3600$ realizations of noise in
1276: the pulled case and about $1000$ in the pushed one.}}
1277: \label{transition}
1278: \end{figure}
1279: Once more it is remarkable the dynamical importance of the
1280: noise correlation length $\lambda$ in the long-wavelength behavior of
1281: the front. The same effect should be expected in 2d, namely, the scaling
1282: could be KPZ or non-KPZ depending solely on noise parameters.
1283: To our knowledge, this is the first time that such a dramatic effect of
1284: noise is reported.
1285: Remarkably enough, while our eikonal description is not able to describe
1286: the second regime, it does predict the transition at the right values of
1287: parameters.
1288:
1289:
1290: \section{Discussion and Conclusions}
1291:
1292: The formulation
1293: in terms of kinematic eikonal-like equations provides a useful
1294: framework for studying front or pulse propagation when one is interested in long
1295: spatial and temporal scales.
1296: This kind of equations have advantages both for numerical
1297: simulations and for theoretical analysis, and has fruitfully been used for instance
1298: in the iphenomenological modeling of excitable wave
1299: propagation in disordered and noisy media, by the \it ad hoc\rm procedure of adding
1300: fluctuations to a generic eikonal equation. In this paper we have derived
1301: stochastic eikonal equations from the more microscopic RD
1302: field equations with noise, which
1303: can be multiplicative.
1304:
1305: The derivation presented here relies on the hypothesis of separation of scales
1306: between the front dynamics at large scales and the internal degrees of freedom of
1307: the front. That means that it is not valid for pulled fronts, which are indeed
1308: known not to have a local, eikonal-like description even in the absence of noise.
1309: However, even for the pushed case, the usual projection techniques for derivation
1310: of sharp-interface equations
1311: cannot be simply extended to the stochastic case due to the fact that the noise
1312: contains the full range of length and time scales. Hence our derviation is formulated
1313: within a coarse-graining scheme, which we claim makes sense for the kind of problems
1314: which admit a local, eikonal-like equation. Derivation of sharp-interface approximations
1315: with noise have been possible so far for thermal noise only,
1316: in systems with local equilibrium.
1317: For multiplicative, generically external noise, however, the absence of a free energy
1318: and a fluctuation-dissipation theorem requires an alternative scheme based on
1319: dynamical equations. We have explicitly checked that our scheme is exact for
1320: the cases of equilibrium fluctuations and also reproduces a rigorous result for
1321: multiplicative noise in d.
1322: For the general case, however,
1323: we rely on numerical tests to fully justify
1324: its validity. In any case, the excellent agreement with numerical and
1325: analytical tests clearly suggests that a more rigorous derivation should be
1326: possible. An extension of our procedure is also conceivable in RD systems for which
1327: a nonlocal sharp-interface description exists for instance in solidification or viscous
1328: fingering.
1329:
1330: One of the main points of this paper
1331: has been to clarify the role of the spatial cutoff of noise
1332: correlations $\lambda$. On the basis of recent rigorous mathematical findings on
1333: the spatio-temporal white-noise limit, we have argued and shown in explicit examples,
1334: that the common splitting
1335: of the Stratonovich white noise in a term which is singular as $\lambda \rightarrow 0$
1336: plus an It\^o noise, does capture the correct dependence
1337: on $\lambda$ of macroscopic
1338: quantities (at scales much larger than $\lambda$). The remaining spatially-correlated
1339: It\^o noise will only carry a weak dependence on $\lambda$, noticeable at length
1340: scales of the order of $\lambda$, and can be treated perturbatively.
1341: The $\lambda \rightarrow 0$ limit can thus be taken safely for the
1342: remaining It\^o noise, once
1343: the singular part has been extracted.
1344: Within this scheme,
1345: macroscopic quantities such as front velocity or front roughening properties,
1346: depend separately of the noise strength $\varepsilon$ and on $\varepsilon/\lambda^d$.
1347: Incidentally, this implies the possibility of measuring the (microscopic)
1348: noise parameters $\varepsilon$ and $\lambda$ from the macroscopic
1349: dynamics of the fronts.
1350: Most remarkably, a partial resummation scheme can be naturally defined for the
1351: parameter $\varepsilon/\lambda^d$
1352: which captures some important, nonperturbative physical phenomena which would be missed
1353: otherwise, such as the transition from pushed to pulles regimes, or the destruction
1354: of the front associated to a divergence of the front thickness. These phenomena,
1355: at the edge of validity of the sharp-interface approximation, are detected respectively
1356: by the vanishing and the divergence of the effective noise-strength on the eikonal
1357: equation.
1358:
1359: Eikonal stochastic equations like the ones derived here are directly related to the
1360: EW and KPZ equations of kinetic roughening. By using known results for these equations and
1361: the results presented here one can predict roughening properties of noisy pulses
1362: or fronts appearing in field equations. We have used such correspondence to check
1363: the predictions in a case of zero velocity and additive noise, for
1364: which the general
1365: theory of equilibrium fluctuations can be directly applied to the RD equation.
1366: Analytical results obtained for the corresponding (Edwards-Wilkinson) eikonal
1367: equation have been identical (including prefactors) to that independent
1368: calculation for the RD system.
1369:
1370: We have also applied our results to a prototype model with multiplicative noise
1371: constructed to represent a variety of different front propagation regimes by
1372: changing a single control parameter. While this model has already been used to
1373: study effects of noise on $1d$ fronts, we have addressed here a more specific $2d$
1374: effect such as the case of front roughening. In a zero-velocity case, simulations
1375: of the reaction-diffusion equation have presented a very good agreement with the
1376: predictions of roughening for the corresponding EW equation with no adjustable parameter.
1377: The results show that the dependence on $\lambda$ is quantitatively important
1378: even if $\lambda$ is significantly smaller
1379: than the front thickness, which may seem counterintuitive.
1380: Although we have not
1381: checked the analogous results for a nonzero velocity, which would correspond to
1382: the KPZ equation, we expect our results to be valid also in this case.
1383:
1384: Most interestingly we have explicitly checked the prediction of qualitative changes
1385: as $\varepsilon/\lambda^d$ is varied, such as the transition from pushed to pulled
1386: propagation regimes. We have directly observed the change in the wandering exponent in
1387: 1d as $\lambda$ is decreased. The immediate extension of this result implies that
1388: a transition from KPZ to non-KPZ scaling is to be expected in higher dimensions.
1389: We thus conclude that the dependence on the spatial cutoff of
1390: the noise may have dramatic effects, not only on the nonuniversal quantities but also
1391: on the universal ones.
1392:
1393: In summary, we have derived stochastic eikonal equations from stochastic RD
1394: equations completely specifying the parameters of the noise. Although this is not
1395: a systematic derivation, we have presented a wealth of independent
1396: evidence on the validity of our results both from analytical
1397: calculations and from numerical simulations of different systems and looking at
1398: distinct noise effects, both qualitative and quantitative.
1399: These results should be of interest for theoretical and
1400: practical purposes both in the study of kinetic roughening and in the context of
1401: propagation of chemical waves.
1402:
1403:
1404: \acknowledgments
1405:
1406: We acknowledge financial support from Direcci\'on
1407: General de Investigaci\'on Cient\'{\i}fica y T\'ecnica (Spain)
1408: (Projects BFM2000-0624-C03-02 and BXX2000-0638-C02-02),
1409: Comissionat per a Universitats i Recerca (Spain)
1410: (Projects 1999SGR00145 and 2000XT-0005),
1411: and European Commision (Project TMR-ERBFMRXCT96-0085). We also
1412: acknowledge computing support from Fundaci\'o Catalana per a la
1413: Recerca-Centre de Supercomputaci\'o de Catalunya (Spain).
1414:
1415: \begin{thebibliography}{99}
1416:
1417: \bibitem{Cross} C. Cross and P. C. Hohenberg, Rev. Mod. Phys. {\bf
1418: 65},
1419: 851 (1993).
1420:
1421: \bibitem{vanSaarloos} W. van Saarloos, Phys. Rev. Lett. {\bf 58}, 2571
1422: (1987);
1423: Phys. Rev. A {\bf 37}, 211 (1988); {\bf 39}, 6367 (1989).
1424:
1425: \bibitem{Collet90}
1426: P. Collet, J.P. Eckmann, {\em Instabilities and Fronts in Extended Systems},
1427: Princeton University Press, Princeton, NJ, 1990.
1428:
1429: \bibitem{Ebert00a} U. Ebert, W. van Saarloos, Physica D {\bf 146}, 1
1430: (2000).
1431:
1432: \bibitem{Kapral93}
1433: {\em Chemical Waves and Patterns}, edited by R. Kapral and K. Showalter,
1434: Kluwer Academic, Dordrecht, 2001.
1435:
1436: \bibitem{Mikhailov95} A.S. Mikhailov, {\it Foundations of Synergetics
1437: I},
1438: (Springer-Verlag, Berlin, 1995)
1439:
1440: \bibitem{tyson} J.J. Tyson and J.P. Keener, Physica D {\bf 32}, 327 (1988).
1441:
1442: \bibitem{Meron} E. Meron, Phys. Rep. {\bf 218}, 1 (1992).
1443:
1444: \bibitem{Mikhailov94}
1445: A.S. Mikhailov, V.A. Davydov, and
1446: V.S. Zykov,
1447: Phys. D {\bf70}, 1 (1994).
1448:
1449: \bibitem{heart} J.M. Davidenko, A.M. Pertsov, R. Salomonz, W. Baxter,
1450: and J. Jalife, Nature (London) {\bf335}, 349 (1992); A.M. Pertsov,
1451: J.M. Davidenko, R. Salomonz, W. Baxter and J. Jalife, Circ. Res. {\bf
1452: 72},
1453: 631 (1992); A. Karma, Proc. Natl. Acad. Sci. USA {\bf 97}, 5687
1454: (2000).
1455:
1456: \bibitem{Riordan95} J. Riordan, C.R. Doering, and D. ben-Avraham,
1457: Phys. Rev. Lett. {\bf 75}, 565 (1995).
1458:
1459: \bibitem{Armero96} J. Armero, J.M. Sancho, J. Casademunt, A.M.
1460: Lacasta, L.
1461: Ram\'{\i}rez-Piscina, and F. Sagues, Phys. Rev. Lett. {\bf 76}, 3045
1462: (1996).
1463:
1464: \bibitem{Armero98} J. Armero, J. Casademunt, L. Ram\'{\i}rez-Piscina,
1465: and J.M.
1466: Sancho, Phys. Rev. E {\bf 58}, 5494 (1998).
1467:
1468: \bibitem{Kessler98} D.A. Kessler and H. Levine, Nature {\bf 394}, 556
1469: (1998).
1470:
1471: \bibitem{Rocco00} A. Rocco, U. Ebert, W. van Saarloos, Phys. Rev. E
1472: {\bf 62}, R13 (2000).
1473:
1474: \bibitem{Tripathy00} G. Tripathy and W. van Saarloos,
1475: Phys. Rev. Lett. {\bf 85}, 3556 (2000).
1476:
1477: \bibitem{Tripathy01} G. Tripathy, A. Rocco. J. Casademunt, W. van
1478: Saarloos,
1479: Phys. Rev. Lett. {\bf 86}, 5215 (2001).
1480:
1481: \bibitem{Rocco01} A. Rocco, J. Casademunt, U. Ebert and W. van
1482: Saarloos,
1483: to appear in Phys. Rev. E, issue of January 1, 2002
1484: (cond-mat/0106081).
1485:
1486: \bibitem{vulpiani} A. Torcini, A. Vulpiani, A. Rocco, to appear in
1487: European Physical Journal B (cond-mat/0107136)
1488:
1489: \bibitem{Sendina-Nadal97}
1490: I. Sendi\~na-Nadal, M. G\'omez-Gesteira, V. P\'erez-Mu\~nuzuri,
1491: V. P\'erez-Villar, J.Armero, L. Ram\'{\i}rez-Piscina, J.Casademunt,
1492: J.M. Sancho
1493: and F.Sagu\'es, Phys. Rev. E {\bf 56}, 6298 (1997)
1494:
1495: \bibitem{Sendina-Nadal98}
1496: I. Sendi\~na-Nadal, A.P.
1497: P\'erez-Mu\~uzuri, D. Vives, V. P\'erez-Mu\~uzuri, J.Casademunt, L.
1498: Ram\'{\i}rez-Piscina, J.M. Sancho and F.Sagu\'es, Phys. Rev. Lett.
1499: {\bf 80},
1500: 5437 (1998)
1501:
1502: \bibitem{Sendina-Nadal00}
1503: I. Sendi\~na-Nadal, S. Alonso, V.P\'erez-Mu\~nuzuri, M.
1504: G\'omez-Gesteira,
1505: V. P\'erez-Villar, L. Ram\'{\i}rez-Piscina, J. Casademunt, J.M. Sancho
1506: and F.
1507: Sagu\'es, Phys. Rev. Lett. {\bf 84}, 2734 (2000).
1508:
1509: \bibitem{Showalter98} S. K\`ad\`ar, J. Wang, and K. Showalter, Nature
1510: (London)
1511: {\bf 391}, 770 (1998).
1512:
1513: \bibitem{Brunet97} E. Brunet and B. Derrida, Phys. Rev. E {\bf 56}, 2597
1514: (1997).
1515:
1516: \bibitem{Kessler98b} D. A. Kessler, Z. Ner, and L.M. Sander, Phys. Rev. E {\bf 58},
1517: 107 (1998).
1518:
1519: \bibitem{Panja01} D. Panja and W. van Saarloos, cond-mat/0110344.
1520:
1521: \bibitem{Moro01}
1522: E. Moro, Phys. Rev. Lett. {\bf 87}, 238303 (2001).
1523:
1524: \bibitem{hakim} V. Hakim, A. Karma, Phys. Rev. E {\bf 60}, 5073
1525: (1999).
1526:
1527: \bibitem{Ebert00b} U. Ebert, W. van Saarloos, Phys. Rep {\bf 337}, 139
1528: (2000).
1529:
1530: \bibitem{Wang93}
1531: S.-L. Wang, R.F. Sekerka, A.A. Wheeler, B.T. Murray, S.R. Coriell,
1532: R.J. Braun and G.B. McFadden, Physica D {\bf 69}, 189 (1993).
1533:
1534: \bibitem{Karma96}
1535: A. Karma and W.-J. Rappel, Phys. Rev. E {\bf 53}, R3017 (1996).
1536:
1537: \bibitem{folch} R. Folch, J. Casademunt, A. Hern\'andez-Machado and L.
1538: Ram\'{\i}rez-Piscina, Phys. Rev. E {\bf 60}, 1724 (1999); Phys. Rev. E
1539: {\bf 60}, 1734 (1999).
1540:
1541: \bibitem{stanley} A.-L. Barab\'asi and H.E. Stanley, {\em Fractal
1542: Concepts
1543: in Surface Growth}, Cambridge University Press, Cambridge, 1995.
1544:
1545: \bibitem{Krug97} J. Krug, Adv. Phys. {\bf 46}, 139 (1997).
1546:
1547: \bibitem{Kardar86} M. Kardar, G. Parisi, and Y.C. Zhang,
1548: Phys. Rev. Lett. {\bf 56}, 889 (1986).
1549:
1550: \bibitem{Hohenberg} P.C. Hohenberg and B.I. Halperin, Rev. Mod. Phys.
1551: {\bf 49}, 435 (1977).
1552:
1553: \bibitem{Safran94} S.A. Safran, {\em Statistical Thermodynamics of
1554: Surfaces, Interfaces, and Membranes}, Addison-Weslesy 1994.
1555:
1556: \bibitem{karmaprl} A. Karma, Phys. Rev. Lett. {\bf 70}, 3439 (1993).
1557:
1558: \bibitem{rappel} A. Karma, W.-J. Rappel, Phys. Rev. E {\bf
1559: 60},
1560: 3614 (1999).
1561:
1562: \bibitem{cinca} R. Gonz\'alez-Cinca, L. Ram\'irez-Piscina, J.
1563: Casademunt and
1564: A. Hern\'andez-Machado, Phys. Rev. E {\bf 63}, 051602 (2001).
1565:
1566: \bibitem{Gardiner} C.W. Gardiner, {\em Handbook of Stochastic Methods
1567: for Physics, Chemistry and the Natural Science}, Springer-Verlag,
1568: Berlin, 1983; N.G. van Kampen, {\em Stochastic Processes in Physics
1569: and Chemistry}, North Holland,
1570: Amsterdam, 1981.
1571:
1572: \bibitem{Novikov64} E.A. Novikov, Zh. Eksp. Teor. Fiz. {\bf 47}, 1919
1573: (1964) [Sov. Phys. JETP {\bf 20}, 1290 (1965)].
1574:
1575: \bibitem{Gyongy98}
1576: S. Gy\"ongy,
1577: Stochastic Processes and their Applications, \bf73\rm (2) 271-299 (1998).
1578:
1579: \bibitem{Nualart99}
1580: E. Al\`os, J. A. Le\'on, D. Nualart,
1581: Probab. Theory Relat. Fields, \bf115\rm (1) 41-94 (1999).
1582:
1583: \bibitem{Edwards82} S.F. Edwards and D.R. Wilkinson,
1584: Proc. R. Soc. Lond. A {\bf 381}, 177 (1982).
1585:
1586: \bibitem{Beccaria94} M. Beccaria and G. Curci, Phys. Rev. E {\bf 50}, 4560 (1994).
1587:
1588: \end{thebibliography}
1589:
1590:
1591:
1592: \section*{Appendix: Systematic Expansion}
1593:
1594: In some circumstances, the eigenvalue problem defined by Eq. (\ref{fi}) may not be
1595: solvable analytically, due to the functional form of
1596: ${\bf h}(\boldmath{\mbox{$\phi_0$}})$, which has been renormalized by noise, even though
1597: the 'bare' problem may be solvable.
1598: In such cases it may be useful
1599: to solve for the case $\varepsilon = 0$ (no noise) and
1600: obtain corrections in a systematic expansion in $\varepsilon$. This is the aim of
1601: this appendix.
1602:
1603: Let us consider then the zero noise eigenvalue problem defined by the
1604: (deterministic) $1d$ equation
1605: \be
1606: 0 = \hat{D} \frac{\partial^2 \boldmath{\mbox{$\phi_{\cal D}$}}}{\partial r^2}
1607: + v_0 \frac{\partial \boldmath{\mbox{$\phi_{\cal D}$}}}{\partial r} +
1608: {\bf f}(\boldmath{\mbox{$\phi_{\cal D}$}}) \label{fi0}
1609: \ee
1610: We now expand the fields of the noisy 2d problem of Eq. (\ref{wtho}) as
1611: perturbations in $\varepsilon$ and $\kappa$ of the solution of Eq. (\ref{fi0}):
1612: \be
1613: \boldmath{\mbox{$\phi$}}(r,s,t) = \boldmath{\mbox{$\phi_{\cal D}$}}(r)
1614: + \delta\boldmath{\mbox{$\phi$}}(r,s,t)
1615: \label{pe0}
1616: \ee
1617: \be
1618: v_n(s,t) = v_0 + \alpha \varepsilon
1619: + \beta \kappa(s,t)
1620: + \delta v(s,t) \label{scurv0}
1621: \ee
1622: We get at the linear order
1623: \be
1624: 0 = \hat{\Gamma}_0 \delta \boldmath{\mbox{$\phi$}} + \varepsilon {\bf G}(\boldmath{\mbox{$\phi_{\cal D}$}})
1625: + (\alpha \varepsilon + \beta k + \delta v)
1626: \frac{\partial \boldmath{\mbox{$\phi_{\cal D}$}}}{\partial r} + \hat{D} k
1627: \frac{\partial \boldmath{\mbox{$\phi_{\cal D}$}}}{\partial r}
1628: + \varepsilon^{1/2} \boldmath{\mbox{$\Omega$}}(\boldmath{\mbox{$\phi_{\cal D}$}}; r,s;t)
1629: \label{deltafi0}
1630: \ee
1631: where
1632: \be
1633: \hat{\Gamma}_0 = \hat{D} \frac{\partial^2}{\partial r^2}
1634: + v_0
1635: \frac{\partial}{\partial r} + \left.\frac{\partial {\bf f}}
1636: {\partial \boldmath{\mbox{$\phi$}}} \right|_{ \boldmath{\mbox{$\phi$}} = \boldmath{\mbox{$\phi_{\cal D}$}}}
1637: \label{gamma0}
1638: \ee
1639:
1640: Now the right eigenvector of $\hat{\Gamma}_0$ reads
1641: \be
1642: {\bf u_{(0)}} = \frac{\partial \boldmath{\mbox{$\phi_{\cal D}$}}}{\partial r}
1643: \ee
1644: which is independent of both $\varepsilon$ and $k$.
1645: Taking (\ref{deltafi0}) and performing the scalar product with the left
1646: eigenvector
1647: ${\bf u^{(0)}}$, we get
1648: \be
1649: &&k ({\bf u^{(0)}},\hat{D}{\bf u_{(0)}}) +
1650: (\alpha \varepsilon + \beta k ) ({\bf u^{(0)}}, {\bf u_{(0)}})
1651: + \varepsilon ({\bf u^{(0)}}, {\bf G}(\boldmath{\mbox{$\phi_{\cal D}$}}))
1652: \nonumber \\
1653: && \qquad \qquad \qquad \qquad + \varepsilon^{1/2}
1654: ({\bf u^{(0)}},\boldmath{\mbox{$\Omega$}}
1655: (\boldmath{\mbox{$\phi_{\cal D}$}}; r,s;t) )
1656: + \delta v(s,t) ({\bf u^{(0)}}, {\bf u_{(0)}})
1657: = 0
1658: \ee
1659: As curvature and fluctuations are linear perturbations we get
1660: \be
1661: \alpha = - \frac{({\bf u^{(0)}},
1662: {\bf G}(\boldmath{\mbox{$\phi_{\cal D}$}}))}{({\bf u^{(0)}},{\bf u_{(0)}})},
1663: \;\;\;\;\;\;\;\;\;\;\;\beta = - \frac{({\bf u^{(0)}}, \hat{D}{\bf u_{(0)}})}
1664: {({\bf u^{(0)}},{\bf u_{(0)}})}
1665: \ee
1666: and
1667: \be
1668: \delta v(s,t)= -
1669: \varepsilon^{1/2}
1670: \frac{({\bf u^{(0)}},\boldmath{\mbox{$\Omega$}}(\boldmath{\mbox{$\phi_{\cal D}$}}; r,s;t) )}
1671: {({\bf u^{(0)}},{\bf u_{(0)}})}.
1672: \ee
1673:
1674: The stochastic eikonal equation is now
1675: \be
1676: v_n (s,t)= v_0 + \alpha \varepsilon +
1677: \beta \kappa(s,t) +
1678: D_f^{1/2}(\varepsilon)\zeta(s,t)
1679: \ee
1680: where the noise $\zeta(s,t)$ is a zero mean Gaussian white process with
1681: \be
1682: \langle \zeta(s,t) \zeta(s^{\prime},t^{\prime}) \rangle =
1683: 2 \delta (s-s^{\prime}) \delta(t-t^{\prime}),
1684: \ee
1685: and
1686: \be
1687: D_f(\varepsilon) =
1688: \varepsilon \frac{\int dr \sum_{i,j} u_i^{(0)} u_{j(0)}
1689: g_i(\boldmath{\mbox{$\phi_{\cal D}$}})g_j(\boldmath{\mbox{$\phi_{\cal D}$}})}{({\bf u^{(0)}},{\bf u_{(0)}})^2}.
1690: \ee
1691:
1692: \end{document}
1693:
1694:
1695:
1696: