1: %\documentclass[preprint]{revtex4}
2: \documentclass[prl,aps]{revtex4}
3: \usepackage{graphicx}
4: %\documentclass[aps,preprint]{revtex4}
5: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: %%%%%%%%%%%% DRAFT SETUP %%%%%%%%%%%%%%%%%%%%%%
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8: %\draftcopyFirstPage{1}
9: %\draftcopyLastPage{2}
10: %\draftcopySetGrey{0.95}
11:
12: %\usepackage[first,outline,bottomafter]{draftcopy}
13: \def\n{\langle n \rangle}
14:
15: \begin{document}
16: \title{Critical behavior of a bounded Kardar-Parisi-Zhang equation}
17: %\draftcopyName{\today}{130}
18: %\draftcopySetScale{80}
19:
20: %\draftcopyFirstPage{1}
21: %\draftcopyLastPage{2}
22: %\draftcopySetGrey{0.95}
23:
24: \author{Miguel A. Mu\~ noz}
25: \affiliation{Departamento de Electromagnetismo y F{\'\i}sica de la Materia,
26: Universidad de Granada,
27: Fuentenueva s/n, 18071 Granada, Spain}
28:
29: \author{Francisco de los Santos}
30: \affiliation{Center for Polymer Studies and Department of Physics \\
31: Boston University, Boston, MA 02215, USA}
32:
33: \author{Abdelfattah Achahbar}
34: \affiliation{Departement de Physique, Faculte des Sciences, \\ B.P. 2121
35: M'hannech, 93002 Tetouan, Morocco}
36:
37: \date{today}
38:
39: \begin{abstract}
40: A host of spatially extended systems, both in physics and in other
41: disciplines, are well described at a coarse-grained
42: scale by a Langevin equation with multiplicative-noise.
43: Such systems may exhibit non-equilibrium phase transitions, which
44: can be classified into universality classes.
45: Here we study in detail one of such classes that
46: can be mapped into a Kardar-Parisi-Zhang (KPZ) interface equation
47: with a positive (negative) non-linearity in the presence of a bounding lower
48: (upper) wall.
49: The wall limits the possible values taken by the height variable,
50: introducing a lower (upper) cut-off, and induce a phase transition
51: between a pinned (active) and a depinned (absorbing) phase.
52: This transition is studied here using mean field and field theoretical arguments,
53: as well as from a numerical point of view.
54: Its main properties and critical features, as well as some
55: challenging theoretical difficulties, are reported.
56: The differences with other multiplicative noise and bounded-KPZ universality
57: classes are stressed, and the effects caused by the introduction of
58: ``attractive'' walls, relevant in some physical contexts, are also analyzed.
59:
60: \end{abstract}
61: \maketitle
62:
63: \section{Introduction}
64: Non-equilibrium phase transitions occurring in systems amenable to be
65: described by Langevin equations including a multiplicative noise (MN) term
66: are the subject of current intense studies. This embraces a broad variety
67: of systems both in physics and in other disciplines.
68: A phenomenology much richer and complex than that appearing in equilibrium systems,
69: including counterintuitive behaviors,
70: has been reported to appear in these, typically non-equilibrium, situations.
71: See \cite{Review,Sancho} for detailed introductions to this growing field,
72: including many different realizations.
73:
74: The interest in MN problems is enlarged even further,
75: because of the existing mappings between them and
76: other prototypical non-equilibrium problems \cite{Review}.
77: A well known instance is the Kardar-Parisi-Zhang (KPZ) equation,
78: describing the kinetic roughening transition of generic interfaces under
79: non-equilibrium conditions \cite{KPZ,Barabasi,HZ}, which
80: can be mapped onto a MN Langevin equation, by
81: performing the so called Cole-Hopf transformation linking the interface
82: height at each point with the activity field of the MN equation.
83: If an interface under consideration is described by
84: the KPZ equation and it is physically limited by a wall, {\it i.e.}
85: if the heights cannot be larger of smaller than a certain value,
86: then this problem, {\it bounded KPZ}, can
87: be mapped into a multiplicative noise equation by employing
88: the abovementioned transformation \cite{MN1,MN2,Review}.
89: The bounded KPZ equation may experience, as parameters
90: are varied, a phase transition from a depinned phase in which the
91: interface escapes with probability one from the wall,
92: to a pinned phase characterized by a finite expectation value
93: of the stationary averaged height (measured from the wall).
94: In the MN language ({\it i.e.} after employing the Cole-Hopf transformation)
95: the pinning-depinning transition corresponds to a critical point
96: separating an {\it absorbing phase} in which the order parameter goes
97: exponentially to zero (depinned phase)
98: to an {\it active phase} in which the order parameter takes a
99: non-vanishing average value.
100:
101: Surprisingly enough, it was shown a few years ago that the introduction
102: of ``upper'' or ``lower'' walls into a given KPZ equation
103: (with a fixed non-linearity
104: sign) lead to quite different phenomenologies. The origin of this can be tracked
105: down to the fact that the KPZ equation is not invariant upon inverting
106: the height (see \cite{MN3} or \cite{Review} for a more detailed explanation).
107: Taking, for instance, the sign of the coefficient of the KPZ non-linearity
108: to be positive, the introduction
109: of an upper wall leads to a (well established by now) set of critical exponents
110: characterizing the, so called, multiplicative noise 1 (MN1)
111: universality class \cite{MN1,MN2,MN3,MN4}.
112: On the other hand, a wall limiting negative values of the
113: interface height (lower wall) leads
114: to a different type of phase transition as shown by T. Hwa and one of us
115: some years back \cite{MN3}.
116: In what follows, and following the nomenclature introduced in \cite{Review},
117: we will use the term MN2 to name this class.
118:
119: It can be easily shown that a KPZ equation with positive non-linearity
120: and a lower wall is completely equivalent to a KPZ with
121: a negative non-linearity coefficient and an upper wall \cite{Review}:
122: one just have to change the sign of the height variable in the KPZ-like equation
123: to verify this.
124:
125: The MN2 class is of great importance in the context of alignment of DNA and other
126: biological sequences. It has been argued by Hwa and collaborators, that the
127: phase transition appearing upon changing the, so called, scoring parameter in the
128: commonly used alignment algorithm can be mapped into the MN2
129: critical point \cite{Hwa}. It is also related to some instances
130: of non-equilibrium wetting \cite{Haye1}. For some other applications
131: and physical instances within this class see \cite{Review} and
132: references therein.
133:
134:
135: While the MN1 class has been extensively studied, specially after its connections
136: with non-equilibrium wetting \cite{Haye1,Haye2,Lisboa}
137: and with the problem of synchronization
138: in extended systems were established \cite{synchro,Review},
139: the MN2 class remains poorly studied.
140: Furthermore, recent numerical analysis have revealed that the
141: preliminary critical exponent values reported in \cite{Hwa} might be
142: far from their true asymptotic values.
143:
144: Aimed at clarifying these issues
145: it is the purpose of this paper to analyze the MN2 phase transition
146: in one dimensional systems using:
147: i) mean-field and field-theoretical techniques and
148: ii) numerical (Monte Carlo) analysis of different models claimed to belong
149: to this class.
150:
151: Finally let us stress that if the wall becomes attractive, rather than simply
152: bounding, new phenomenology might appear. This is particularly
153: interesting in the context of synchronization \cite{Review,synchro}.
154: This possibility will also be discussed along the paper.
155:
156: \section{The MN2 class}
157:
158: Let us consider a KPZ equation with a positive non-linearity coefficient,
159: $\lambda >0$, in the presence of a lower wall,
160: %
161: \begin{equation}
162: \partial_t h(x,t) = a + b ~e^{-ph} + D \nabla^2 h + \lambda (\nabla h)^2
163: + \sigma \eta(x,t).
164: \label{BKPZ}
165: \end{equation}
166: %
167: where $h$ is a height variable, $a$ represents a constant drift
168: while $ b ~e^{-ph}$ is a bounding wall.
169: The parameter $p>0$ controls the wall penetrability, the limit
170: $p \to \infty$ corresponding to a perfectly rigid (impenetrable) wall.
171: It has been shown previously that for both the MN1 and the MN2 classes
172: the magnitude of $p$ does not influence the asymptotic properties at
173: criticality \cite{MN1,MN2} (its sign, however, is important, as it determines
174: whether the wall is a lower or an upper one).
175: The same property applies to equilibrium systems ({\it i.e.},
176: for $\lambda =0$) \cite{lipowskyfisher}.
177: $\eta$ is a stochastic white noise with
178: $\langle \eta \rangle =0$ and
179: $\langle \eta(x,t) \eta(x',t') \rangle = 2 \delta(t-t') \delta (x-x')$,
180: where $\langle \cdot \rangle$ denotes an average over the distribution
181: of the noise.
182:
183: For a fixed value of $b$ the interface experiences a pinning-depinning
184: transition at some value $a=a_c$.
185:
186: Some remarks concerning the connection of the previous equation
187: with wetting problems follow.
188: When $h(x,t)$ is viewed as the distance separating a liquid-gas
189: interface from a solid wall, Eq. (\ref{BKPZ}) can then be
190: interpreted as a dynamic model for nonequilibrium wetting. Under this
191: perspective $a$ is the chemical potential difference between the
192: liquid and the gas phases, $\langle h \rangle$ is the thickness of the
193: wetting layer, and the wall is a rigid, physical substrate.
194: At {\em bulk phase coexistence}, {\it i.e.} for the value of $a=a_c$ for
195: which in the absence of the wall the interface does not move on average,
196: $\langle h \rangle$ diverges at all temperatures above certain
197: wetting temperature, $T_W$, while for $a \not= a_c$ the thickness
198: of the liquid film can be big, but finite (pinned interface).
199: The temperature here is controlled by the parameter $b$, which
200: vanishes linearly with the mean-field wetting temperature as $T-T_W$.
201: Thus, on approaching coexistence for $T>T_W$
202: ($b>0$ at the mean-field level), $\langle h \rangle$
203: diverges as $\langle h \rangle \sim |a-a_c|^{\beta_h}$.
204: This transition, termed {\it complete wetting},
205: is always continuous and the value of the $\beta_h$ exponent depends on
206: the nature of the forces between the particles in the fluid phases
207: and the wall. In this paper only short-range, exponentially decaying
208: interactions between all the particles and the substrate (as described
209: by Eq. (\ref{BKPZ})) are considered.
210:
211: The change of variables $n=\exp(-h)$ transforms Eq. (\ref{BKPZ}) into
212: the MN2 Langevin equation:
213: %
214: \begin{equation}
215: {\partial}_{t} n(x,t) =
216: {\nabla}^{2} n - 2 {(\nabla n)^2 \over n}
217: -(a+1)n-bn^{p+1} + n \eta,
218: \label{mnminus}
219: \end{equation}
220: %
221: where for the sake of simplicity we have
222: set $\lambda=D=\sigma=1$ and Ito calculus \cite{Ito} has
223: been used (different coefficients for the Laplacian and the KPZ non-linear
224: term could be reabsorbed
225: using $n=\exp(-\alpha h)$ with a proper choice of $\alpha$).
226: This transformation maps the depinning from the wall
227: $\langle h \rangle \to \infty$ to a transition into an absorbing state
228: $\n \to 0$.
229: The physical equivalence between the cases $\lambda>0$ with a
230: lower-wall ($p>0$), and $\lambda<0$ with an upper wall ($p<0$),
231: reflects in the fact that the same equation is obtained
232: using $n=\exp(-h)$ and $n=\exp h$, respectively.
233:
234: Observe that Eq.(\ref{mnminus}) is identical to the MN equation
235: describing the MN1 class \cite{MN1,MN2,Review}, except for
236: the presence of an extra term $ (\nabla n)^2 /n $.
237: This term can also be written as
238: $(\nabla n) \cdot (\nabla \ln(n)) = (\nabla n ) \cdot (\nabla h)$
239: suggesting that the interface language is the natural one for
240: this class.
241: In fact, except for the factor 2 in front of $(\nabla n)^2 /n$
242: Eq. (\ref{mnminus}) coincides with the Cole-Hopf transform
243: of
244: %
245: \begin{equation}
246: \partial_t h (x,t) = \nabla^2 h + a +b e^{-ph} + \eta(x,t)
247: \end{equation}
248: %
249: that describes the growth of wetting layers toward their
250: equilibrium state \cite{lipowsky}
251: (observe that this is just the equilibrium, Edwards-Wilkinson model,
252: in the presence of a bounding wall). Note also that the factor $2$
253: in Eq. (\ref{mnminus}) cannot be readsorbed by reparametrizing.
254:
255: Finally, let us underline that in the regime where $b<0$
256: the wall becomes attractive (which might be necessary to
257: describe some physical situations as, for instance, synchronization problems
258: as said in the introduction)
259: and a new term, say, $c \exp(-2h)$ (equivalently $c n^{2p+1}$) with $c>0$
260: has to be added to stabilize the equation.
261:
262: Having presented the equations defining the model, in the forthcoming sections
263: we study the associated physics by using i) mean field approaches, ii)
264: field theory, and iii) numerical, Monte Carlo simulations
265: combined with scaling arguments.
266:
267:
268: \section{Mean-field approaches}
269: Mean-field approaches to Eq. (\ref{BKPZ}) can be implemented with
270: several degrees of sophistication. A crude approximation consists of
271: ignoring the noise and spatial variations. At this level one trivially gets
272: that the order parameter $\langle h \rangle$ vanishes as $a \to 0$ with an
273: exponent $\beta_h=0$. When applied to Eq.(\ref{mnminus}), this
274: approximation yields a shifted critical point at $a_c=-1$ and the
275: usual result for the order parameter critical exponent, $\beta_n=1/p$.
276: But, as experience with other systems with multiplicative-noise dictates,
277: neglecting completely the noise is a too crude approximation, that eliminates
278: most of the characteristic traits of MN physics.
279:
280: Now the effect of allowing a spatially varying order
281: parameter and taking the noise into consideration
282: are examined.
283: The Laplacian is discretized as
284: %
285: \begin{equation}
286: \nabla^2 n_i ={1 \over 2d} \sum_{j} (n_j-n_i) \approx \n
287: -n_i,
288: \end{equation}
289: %
290: where the sum runs over the nearest neighbors of $i$ and a large
291: system dimensionality has been assumed. Similarly, the square gradient term
292: can be written as
293: %
294: \begin{equation}
295: {(\nabla n)_i^2 \over n_i} \approx {\langle n^2 \rangle \over n_i}
296: -2\n + n_i.
297: \end{equation}
298: %
299: The one-site stationary probability distribution is then readily
300: obtained from the associated Fokker-Planck equation,
301: \begin{eqnarray}
302: P_{st}\Big(n,\n\Big) &\propto& {1 \over n^2} \exp \int^n {
303: (\nabla x)_i^2-2(\nabla x)_i^2/ x_i -(a+1)x_i-bx_i^{p+1} \over x_i^2} \ dx_i,
304: \nonumber \\
305: &\approx &{1 \over n^{a+6}} \exp \Bigg(-5{\n \over n}-b {n^p\over p}
306: +{\langle n^2 \rangle \over n^2}\Bigg).
307: \label{pmnminus}
308: \end{eqnarray}
309: %
310: where $\n$ and $\langle n^2 \rangle$ have to be calculated self-consistently.
311: For $a< a_c=-5$, $P_{st}$ is not normalizable what means that the
312: stationary state is the absorbing phase $\n =0$. For
313: $\n \not=0$, however, the non-analyticity of
314: $\exp[\langle n^2 \rangle / n^2]$ at $n=0$ again renders $P_{st}$
315: non-normalizable. As a result, there is no well-defined active phase
316: at this mean-field level for Eq. (\ref{pmnminus}).
317: Similar problems are found if the same type of approach is
318: applied to Eq. (\ref{BKPZ}) instead of Eq. (\ref{mnminus}).
319:
320:
321: To avoid the presence of the square gradient term in Eq. (\ref{mnminus}),
322: and the complications of its, somewhat arbitrary, discretization, we resort to
323: a different change of variables. After $n=\exp h$,
324: %
325: \begin{eqnarray}
326: \partial_t n &=& \nabla^2 n +(a+1)n +bn^{3-p} +n \eta, \nonumber \\
327: P_{st} &\sim& {1 \over n^{2-a}} \exp\bigg[{b n^{2-p} \over 2-p}-{\n\over n}\bigg].
328: \label{mnplus}
329: \end{eqnarray}
330: %
331: These equations are simpler than (\ref{mnminus}) and (\ref{pmnminus}),
332: but at the cost that $\n$ is no longer an order parameter
333: as, at the transition point $\n \to \infty$ rather than going to $0$.
334: Furthermore, Landau expansions only make sense when
335: the order parameter vanishes at the critical point, making the whole approach
336: inconsistent.
337: One possible way to circumvent this problem is to monitor $m \equiv 1/n$,
338: and study $\tilde{P}_{st}(m) dm =P_{st}(n) dn$,
339: but given the non-Gaussian nature of the probability distribution
340: the substitution $\n = 1/\langle m \rangle$ is likely to be incorrect,
341: and there seem to be no safe way to proceed.
342:
343: Summing up, non-trivial mean-field approaches detect some problems
344: with the model under consideration, and are not able to predict a phase
345: correct phase diagram. A sound mean-field approximation, needs therefore
346: to be found. Notice that none of these problems occur in the case of a
347: negative KPZ non-linearity, {\it i.e.} in the MN1 class, where a
348: standard mean-field approximation yields qualitatively
349: correct results (see \cite{Review,Lisboa} and references therein).
350:
351: \section{Field theoretical considerations}
352:
353: The Langevin equation for MN1 is known to be super-renormalizable,
354: {\it i.e.} Feynman diagrams can be computed to all orders and resummated.
355: This does not imply that critical exponents can be computed in all the
356: cases, as the renormalization group flow-equation has runaway
357: trajectories supposed to converge to a {\it strong coupling fixed point}.
358: But at least, the correct phase diagram, including strong and weak coupling
359: fixed points can be obtained.
360:
361: For the MN2 the situation is far more complicated, as can be {\it
362: a priori} anticipated given the failure of mean-field approaches.
363: The extra term, $ (\nabla n)^2 /n $, being singular in $n$ precludes
364: the use of perturbative expansions around $n=0$.
365: Given the lack of a non-perturbative approach to the KPZ and MN
366: strong coupling fixed points, there is not much we can add to
367: this section, except that there is a promising attempt to tackle this
368: and related KPZ-like problems.
369: There is a formalism, developed by Fogedby,
370: aimed at developing a strong coupling theory for KPZ based
371: on a semiclassical or WKB approximation applied upon the Martin-Siggia-Rose
372: generating functional \cite{fogedby}. Its main advantage is that it does not
373: involve expansions around classical noiseless solutions, but around
374: classical (extremal) noisy solutions.
375: It would be very interesting to extend these ideas to KPZ problems
376: in the presence walls,
377: namely, to the multiplicative noise universality classes MN1 and MN2.
378:
379: \section{Numerical results}
380:
381: Owing to the failure of standard (non-trivial) mean-field approaches and lacking
382: so far of an alternative analytical route, numerical methods are required to glean
383: insight into the system properties. We have carried out simulations
384: of two surface growth models. Both of them, in the absence of walls,
385: are known to belong to the KPZ universality class.
386: An extra rule is then added to generate a bounding wall, as described above.
387:
388: \subsection{Model 1}
389:
390: Our first model was introduced in \cite{krugs} and is defined
391: as follows: the surface position at time $t$ above a site $x$ on a
392: one-dimensional lattice of size $L$ is given by a continuous height
393: variable $h_t(x)$. A new height configuration is then generated in
394: a three-step process.
395: \begin{enumerate}
396: \item Each lattice site $h_t(x)$
397: is updated according to $h'_t(x)=h_t(x)+a+\eta_t(x)$. $\eta_t(x)$ is
398: a random number uniformly distributed in [0,1] and $a$ is a constant drift
399: term analogous to that of (\ref{BKPZ}).
400: \item The configuration is changed to
401: $h_{t+1}(x)=\min[h'_t(x\pm 1)+\gamma,h'_t(x)]$, where $\gamma$ is
402: a constant whose precise value is not essential for the final results.
403: We have set without loss of generality $\gamma=0.1$ as in previous
404: numerical analyses.
405: \item A hard wall at $h=0$ is introduced by way of the
406: additional rule $h_t(x)=\min[h_t(x),0]$ \cite{MN3}.
407: \end{enumerate}
408: Finally, periodic boundary conditions are imposed and
409: $h_t(x)$ is initially set to $0$.
410:
411: The continuum counterpart of this model is known to be a KPZ equation \cite{krugs}
412: with $\lambda<0$ in the presence of an upper wall \cite{note}
413: which, as remarked before, is
414: equivalent to the case $\lambda >0$ and a lower wall, and corresponds to
415: the MN2 class.
416: Numerical results for this model were first presented in \cite{MN3}
417: and seemed to be consistent with mean-field ({\it i.e.} single-site) like behavior.
418: However, a similar problem recently studied in the context of synchronization
419: has revealed inconsistencies probably due to insufficient
420: statistics \cite{Review,synchro}.
421: In this subsection improved simulation results are provided upon revisiting the
422: analysis reported in \cite{MN3} for larger system sizes and longer
423: sampling times.
424: The case of an attractive wall, not included in \cite{MN3},
425: is also considered.
426:
427: First, we take up the case of a simple (non-attractive) wall, corresponding to
428: $b>0$ in Eqs. (\ref{BKPZ}) and (\ref{mnminus}).
429: Figure (1) shows how the steady order parameter $\n$ changes
430: with the system size $L$. Within the active phase, it saturates to
431: a constant value, while in the absorbing one it bends down and decays
432: exponentially. Our best estimate for the critical point is
433: $a_c=1.57433(2)$ and from
434: the slope of the curve we get $\beta_n/\nu_\bot \approx 0.33(2)$ \cite{note2}.
435: This value of $a_c$ corresponds to the point where a free interface
436: (far from the wall) has zero average velocity.
437: The time evolution of $\n$ at the critical point for different system
438: sizes (Figure (2)) behaves like
439: $\langle n(t) \rangle \sim t^{-\theta_n}$, with $\theta_n =0.215(15)$ or,
440: equivalently, $\langle h(t) \rangle \sim t^{\theta_h}$,
441: with $\theta_h =0.355(15)$.
442: As for the exponents $\beta$, which govern the saturation of the order
443: parameter within the active phase, they have been computed using
444: the largest available system sizes ($L=1600$ and $3200$). The best fit
445: to $ \langle X \rangle \sim |a-a_c|^{\pm \beta_X}$ yields
446: $\beta_n \approx 0.32 (3)$ (for $\langle n \rangle$) the and $\beta_h =0.52(2)$
447: (for $\langle h\rangle$), respectively.
448: The error margin is typically larger here than for other exponents due
449: to the sensitivity to the uncertainty in the determination of the
450: critical point.
451:
452: It was proved in \cite{MN1,MN2} that these exponents must satisfy
453: the scaling relation $z= \beta/(\nu_\bot \theta)$, where
454: $z$ is the dynamic exponent of the KPZ equation.
455: The exact value for $z$ in $d=1$ is $3/2$, thereby
456: $z = 0.33 / 0.215 = 1.5(1)$ in agreement with the prediction.
457: In addition, also from \cite{MN1,MN2},
458: $\nu_\bot =1$, and from our direct measurements we get
459: $(\beta_n/\nu_\bot)/ \beta_n = 0.33/0.32$, implying $\nu_\bot \approx 1$.
460: In terms of $h$ and assuming $\nu_\bot=1$,
461: $z=0.52/.355= 1.5(1)$ which, again, is compatible with $3/2$
462: within error bars.
463:
464: The two alternative, but equivalent, mathematical descriptions of the
465: MN2 class in terms of $h$ and $n=\exp(-h)$ can be
466: related noting that the latter is essentially the
467: density of sites at zero hight, $n(x,t)=\delta_{h(x,t),0}$ \cite{Haye3}.
468: We have verified that $n$ and $\delta_{h(x,t),0}$
469: exhibit the same asymptotic scaling behavior.
470:
471:
472:
473:
474: \begin{figure}
475: \vspace{0.5cm}
476: \includegraphics[width=8cm]{Fig1.eps}
477: \label{betanusin}
478: \caption{Model 1: steady-state values of the oder parameter, $\n =\langle
479: \exp h \rangle$, as function of the system size, $L$, for drifts
480: (top to bottom) 1.57450, 1.57440, 1.57435, 1.57433,
481: 1.57430, 1.57425, 1.57420. $\langle \cdot \rangle$
482: denotes both spatial and temporal averages, as well as
483: averages over independent runs. The straight line corresponds
484: to the critical point $a_c=1.57433(2)$ and from its slope
485: $\beta_n/\nu_\bot \approx 0.33(2)$.}
486: \end{figure}
487:
488:
489: \begin{figure}
490: \vspace{0.5cm}
491: \includegraphics[width=8cm]{Fig2.eps}
492: \label{theta}
493: \caption{Model 1: time evolution of the order parameter, $\langle n(t) \rangle$,
494: at criticality, $a=a_c$, for system sizes 100, 400,
495: 800, and 1600. The saturation values are those of Figure 1.
496: The best fit gives $\theta_n=0.215(15)$. The straight line
497: is a guide for the eye and has a slope -0.215.}
498: \end{figure}
499:
500:
501:
502: \begin{figure}
503: \vspace{0.5cm}
504:
505: \includegraphics[width=8cm]{Fig3.eps}
506: \label{betanucon}
507: \caption{
508: Model 1: steady-state values of the oder parameter for an
509: attractive wall, $b=-0.3$, as function of the system size
510: $L$ for drifts (top to bottom) 1.57450, 1.57447, 1.57445, 1.57440,
511: 1.57435, 1.57433, 1.57430, 1.57425, 1.57420. The best fit to a
512: straight line again corresponds to $a_c=1.57433$.}
513: \end{figure}
514:
515: Attractive walls can also be simulated within this model
516: by simply substituting $a$ by $a-b\delta_{h,0}$,
517: where $b<0$ and the sign convention is
518: chosen to keep the analogy with Eq. (\ref{BKPZ}).
519: This means that wherever the interface in attached to the wall
520: it experiences an additional (``sticky'') force pushing it against
521: the wall.
522: Extensive Monte Carlo simulations for $b=-0.3$ show
523: that the previous results, in what respect universal features,
524: carry over without change, the only difference being that the approach to
525: asymptotics is slower. Upon increasing the attractiveness of the wall
526: the transients become longer.
527: The estimate for the critical point is the same one
528: as before (this is due to the fact that the free interface is not
529: affected by variations of the attractiveness parameter). Again, our best
530: estimates for the critical exponents are $\beta_n/\nu_\bot =0.32$
531: and $\theta_n=0.215$ (see Figure (3)).
532: For $b=-0.4$ we still observe a second-order phase
533: transition with a crossover to the mentioned exponents.
534: For $b=-0.5$ transition becomes first-order but it still
535: occurs at $a=a_c$. Within the active phase all the sites
536: are closed to the wall and the order parameter is $1$, but
537: it suddenly changes to $0$ upon decreasing $a$ and hysteresis
538: is observed for slightly subcritical values $a$ (the interface
539: is pinned for up to long times). Therefore, a tricritical point
540: must exist between $0.4$ and $0.5$. We have identified
541: it at $b=-0.42(1)$.
542: The tricritical behavior has not been analyzed.
543:
544: %This is analogous to a critical wetting
545: %transition with a critical temperature depressed from its mean-field
546: %value $b=0$ or, in other words, the renormalized value of $b$ is positive
547: %even for negative bare values of $b$.
548:
549:
550:
551: Let us stress that, contrarily to what happens for the MN1 class, where the presence
552: of an attractive wall induces a new and rich phenomenology (including
553: a broad region of phase coexistence and directed-percolation type of
554: transitions \cite{Review}), the addition of ``attractiveness'' has
555: a very mild effect here. Basically, it just shifts the position of
556: the critical point
557: and induces a first-order phase transition for very strong attractions.
558:
559:
560: \begin{table}[t]
561: \begin{tabular}{lcccccc}
562: & \multicolumn{6}{c}{} \\
563: \cline{2-7}
564: & $\beta_n$ & $\nu_\bot$ & $\beta_n/\nu_\perp$ & $z$ & $\theta_n$
565: & $\eta$ \\
566: \hline
567: Model 1 ~~~ &$0.32(2)$& $0.97(5)$ & $0.34(2)$ & $1.55(5)$ & $0.215(15)$
568: & not measured \\
569: Model 2 ~~~ &$0.325(5)$& $\approx 1$ & $0.33(2)$ & $\approx 1.5$ & $0.215(5)$
570: &$\eta = 0.8$
571: \\
572: \hline
573: \end{tabular}
574: \caption{Critical exponents for the MN2 class in $d=1$.}
575: \end{table}
576:
577:
578:
579:
580: \subsection{Model 2}
581:
582: In order to verify the robustness and eventual universality of the
583: previous results, we have performed a second study of a different
584: model. It is a restricted solid-on-solid (RSOS) model,
585: a variant of the single-step model introduced in \cite{krugs}, in the
586: presence of a wall.
587: A similar model
588: has been recently studied in the context
589: of synchronization transitions \cite{Ahlers}, to study MN1 type of transitions.
590: Initially the wall is located at $h_w =0$ and a grooved interface
591: is placed beneath it, {\it i.e.} the interface has negative height at all the
592: positions.
593:
594: The dynamics proceeds as follows:
595: At each time step, a site is randomly picked from a one-dimensional lattice of
596: length $L$ and its height decreased two units, $h(i) \to h(i)-2$,
597: provided that $h(i)>h(i+1)$ and $h(i)>h(i-1)$, {\it i.e.} provided
598: that it is a local maximum. Should the
599: the RSOS constraint be violated, the trial is discarded and repeated.
600: Every $2(L-1)/[1-2\delta v (1-L^{-1})]$ steps the wall retreats one
601: unit and, simultaneously, the interface is moved downwards by two units
602: wherever it lies above the wall \cite{Ahlers}.
603: The difference between the wall and interface velocities,
604: $\delta v$, acts as the control parameter: if $\delta v$ is negative,
605: the interface eventually depins from the wall, while for
606: $\delta v>0$ it remains pinned). It can be easily shown, using random walk
607: arguments that for the chosen wall velocity the system seats at its
608: critical point. By varying it, we have a control parameter.
609: The possibility of computing analytically the critical point largely simplifies
610: the numerical analysis, and makes of this an efficient
611: discrete model.
612:
613: The quantities monitored are $\langle \exp(h_w-h) \rangle$
614: and $h_w-h$. We have measured the exponents $\beta_n$ and $\theta_n$
615: for a system of $L=2^{20}$ sites. Our results lead to
616: $\beta_n=0.325(5)$ and $\theta_n=0.215(5)$, in excellent agreement
617: with those of model 1 (Figure (4)).
618: Once more, assuming $\nu_\bot=1$, we get $z \approx 1.51$ in good
619: agreement with the scaling laws.
620: In addition, we have also measured the spreading exponent, $\eta$,
621: that characterizes the number of pinned sites.
622: It is computed averaging over all the runs and starting with an
623: initial condition with a single point attached to the wall.
624: Our best estimate is $\eta =0.80(2)$ (Figure (5)).
625: Measuring the surviving probability, and therefore the exponents $\delta$
626: and $\zeta'$, is a delicate technical point because it is hard to
627: decide when the activity of a run has ceased. We have not tackled
628: this problem here.
629: Lastly, we obtain $\theta_h =0.34(5)$, which is again in good agreement
630: with the value reported for our Model 1.
631: \begin{figure}
632: \vspace{0.5cm}
633: \includegraphics[width=8cm]{Fig4.eps}
634: \label{sidebyside}
635: \caption{
636: Model 2: (A) order parameter behavior, $\langle \exp(h_w-h) \rangle$,
637: in the vicinity of the critical point $\delta v_c =0$. From
638: the slope of the line we get $\beta_n =0.325(5)$.
639: (B) Decay of the order parameter, $\langle \exp(h_w-h) \rangle$,
640: at the critical point yields $\theta_n=0.215(5)$
641: (cf. with Figure (2). The straight line is a guide for the
642: eye and has a slope $-0.215$.}
643: \end{figure}
644:
645: \begin{figure}
646: \vspace{0.5cm}
647: \includegraphics[width=6cm]{Fig5.eps}
648: \label{fattaheta}
649: \caption{
650: Model 2: number of pinned sites as a function of time for an
651: initial condition with a single point attached to the wall.
652: A fit for late times yields $\eta=0.80(2)$. The straight line is a
653: guide for the eye and has slope $0.8$.}
654: \end{figure}
655:
656: We have also considered different variations of the model in which the
657: interface can penetrate the wall at some points, {\it i.e.} the wall is not
658: perfectly rigid. None of the universal features seem to be affected by this
659: change.
660:
661: In conclusion,
662: all these results support strongly the existence of robust
663: universality in the MN2 class.
664:
665:
666: As a matter of consistency, we have modified the algorithm of the
667: model to simulate a lower wall, and therefore a case expected to be
668: in the MN1 class. We obtain the set of exponents
669: $ \beta_n=1.69, \theta_n= 1.19$ and $\eta = -0.4$, all of them
670: in agreement we previously reported results and showing
671: that the upper and lower problem belong to different universality
672: classes \cite{Review,MN2,MN4}.
673:
674: \subsection{Numeric integration of stochastic differential equations}
675:
676:
677: As a further test for universality,
678: we have numerically integrated Eq. (\ref{BKPZ}) using
679: a Milshtein's algorithm \cite{Maxi}. A system size of $L=2000$ was considered
680: and the time step and mesh size were set to $\Delta t =0.001$
681: and $\Delta x=1$, respectively. For simulation times up to $t=10^{6}$
682: ($10^9$ trials per site) our results lie far from the asymptotic regime.
683: We do not discard numerical instabilities in the integration scheme,
684: as it is known that results from numerical integrations
685: may not agree with the predictions from the continuum KPZ \cite{KPZproblem}.
686: The Cole-Hopf transform is a standard way to account numerically
687: for the integration of bounded KPZ equations and, indeed, Langevin equations
688: with MN are by far more stable than their interface-language KPZ-like counterparts
689: \cite{MN2,MN4}. Nevertheless, we have also found
690: numerical problems when integrating (\ref{mnminus}), either when the
691: extra term is written in logarithmic form or averaging for smoothing
692: the gradient. All numerical attempts are unstable nearby the absorbing state,
693: owing to the presence of the extra singular term. We leave, therefore,
694: the numerical integration of a continuous Langevin equation, representative
695: of the MN2 class as an open, challenging problem.
696:
697:
698: \section{Discussion}
699:
700: We have characterized the MN2 universality class, or analogously its
701: bounded KPZ counterpart,
702: which, as commented above, accommodates different physical phenomena.
703: We have studied it from, somehow deceptive mean-field and field theoretical
704: approaches, as well as by numerical studies.
705: None of the analytical methods provides a satisfactory description
706: of the phase transition present in this class.
707: On the other hand, Monte Carlo simulations of two different discrete
708: interface models, argued to belong to this universality class,
709: give a firm evidence for the existence of a robust universality class.
710: Contrarily to what previous simulations seemed to indicate \cite{MN3}, our
711: results are not a simple extension of the ones obtained for one-site,
712: implying that spatial correlations play an important role.
713: Table 1 gathers the values of the critical exponents in terms of
714: $n$ for the two discrete models considered in this paper.
715: From the Monte Carlo estimates,
716: it cannot be discarded that they adopt the rational values $\beta_n=1/3$ and
717: $\theta_n=2/9$, which combined with the exact values
718: derived in \cite{MN1,MN2},
719: $\nu_\bot=1$ and $z=3/2$, would lead to $\beta_n/\nu_\bot= 1/3$.
720: Note that our results do not compare well with those of
721: the nonequilibrium wetting model reported in \cite{Haye2}.
722: We believe this is probably due to the extremely long transients known
723: to be present in that model.
724: We have verified that the transition
725: point is located at the same value of the control parameter
726: for any value of the ``attractiveness'' parameter ($b$ in Model 1), either representing
727: an attractive or a non attractive wall.
728: For strong enough attractive walls, {\it i.e.}
729: $b$ sufficiently negative, the transition becomes first order as
730: in \cite{Haye2}, while if the wall is weakly attractive then it
731: remains in the MN2 class.
732:
733: The problem of reaching a satisfactory analytical understanding,
734: and even that of obtaining sound results from numerical integrations
735: of the continuous Langevin-equation (in either the interface or
736: the density language) representative of this class remains
737: an open challenge.
738:
739: Summing up, even though strong evidence is provided confirming
740: the existence of a universality class ({\it i.e. } the corresponding critical
741: exponents are computed with good precision in one dimension and they are
742: universal in two different discrete models),
743: its theoretical description in terms of Langevin equations,
744: contrarily to what happens for the closely related MN1 class,
745: is far from satisfactory. In particular, the Langevin equation does not seem
746: to admit sound mean-field solutions, nor is amenable to be treated by means
747: of standard perturbative field theoretical tools,
748: nor it admits a stable numerical integration.
749: Identifying
750: the physical causes at the root of these difficulties is a challenge
751: for future research.
752:
753:
754:
755: {\centerline {\bf ACKNOWLEDGMENTS}}
756:
757: \vspace{0.5cm}
758: F.S. acknowledges financial support from the Funda\c c\~ao para a
759: Ci\^encia e a Tecnologia, contract SFRH/BPD/5654/2001.
760: Financial support from the Spanish MCyT (FEDER) under project BFM2001-2841,
761: and from the AECI,
762: are also acknowledged.
763:
764: \begin{thebibliography}{}
765:
766:
767:
768: \bibitem{Sancho}
769: See J. Garc{\'\i}a-Ojalvo, and J. M. Sancho,
770: {\it Noise in Spatially Extended Systems},
771: Springer, New York, 1999; and references therein. See also, J. M. Sancho
772: and J. Garc{\'\i}a-Ojalvo, in Lecture Notes in Physics {\bf 557}, p.235, ed.
773: J. A. Freund and T. P\"oschel, Springer-Verlag, Berlin (2000).
774:
775: \bibitem{Review}
776: M.A. Mu\~noz, preprint 2003, cond-mat/0303650.
777:
778: \bibitem{KPZ}
779: M. Kardar, G. Parisi and Y. C. Zhang,
780: Phys. Rev. Lett. {\bf 56}, 889 (1986).
781:
782: \bibitem{HZ}
783: T. Halpin-Healy and Y.-C. Zhang, Phys. Rep. {\bf 254}, 215 (1995); and
784: references therein.
785:
786: \bibitem{Barabasi}
787: A. L. Barab\'{a}si, H. E. Stanley,
788: {\it Fractal Concepts in Surface Growth}
789: Cambridge University Press, Cambridge, 1995; and references therein.
790:
791: \bibitem{MN1} G. Grinstein, M.A. Mu\~noz, and Y. Tu
792: Phys. Rev. Lett. {\bf 76}, 4376 (1996).
793:
794: \bibitem{MN2} Y. Tu, G. Grinstein and M.A. Mu\~noz,
795: Phys. Rev. Lett. {\bf 78}, 274 (1997).
796:
797: \bibitem{MN3} M.A. Mu\~noz and T. Hwa,
798: Europhys. Lett. {\bf 41}, 147 (1998).
799:
800: \bibitem{MN4}
801: W. Genovese and M.A. Mu\~noz,
802: Phys. Rev. E {\bf 60}, 69 (1999).
803:
804:
805: \bibitem{Haye1}
806: H. Hinrichsen, R. Livi, D. Mukamel, and A. Politi,
807: Phys. Rev. Lett. {\bf 79}, 2710 (1997).
808:
809: \bibitem{Haye2}
810: H. Hinrichsen, R. Livi, D. Mukamel, and A. Politi,
811: Phys. Rev. E {\bf 61}, R1032 (2000).
812:
813: \bibitem{Lisboa}
814: F. de los Santos, M.M. Telo da Gama, and M.A. Mu\~noz,
815: Europhys. Lett. {\bf 57}, 803 (2002);
816: Phys. Rev. E {\bf 67}, 021607 (2003);
817: Proceedings of the 7th Granada Seminar on Computational Physics.
818: Ed. J. Marro and P. L. Garrido; Am. Inst. of Phys. 661 (2003).
819: Cond-mat/0211124.
820:
821: \bibitem{synchro} M.A. Mu{\~n}oz and R. Pastor Satorras,
822: Preprint. Cond-mat/0301059.
823:
824: \bibitem{Hwa} T. Hwa and M. Lassig, Phys. Rev. Lett. {\bf 76}, 2591 (1996).
825: See also, T. Hwa and M. Lassig, ``Optimal Detection of Sequence Similarity by Local Alignment"
826: in Proceedings of the Second Annual Int. Conf. on Computational
827: Molecular Biology (RECOMB98),
828: S. Istrail, P. Pevzner, and M.S. Waterman eds, 109-116 (ACM Press, 1998); and
829: references therein.
830: R. Olsen, T. Hwa and M. Lassig,
831: "Optimizing Smith-Waterman Alignments"
832: in Pacific Symposium on Biocomputing 4, 302-313 (1999).
833:
834: \bibitem{lipowskyfisher}
835: R. Lipowsky and M.E. Fisher, Phys. Rev. B {\bf 36}, 2126 (1987).
836:
837: \bibitem{Ito} The sole difference between utilizing the Ito or the
838: Stratonovich conventions, in this case, is a trivial shift in
839: $a$ \cite{VK,Gardiner}.
840:
841: \bibitem{VK} N.G. van Kampen, {\it Stochastic Processes in Physics
842: and Chemistry}, North Holland, Amsterdam, 1981.
843:
844: \bibitem{Gardiner} C.W. Gardiner,
845: {\it Handbook of Stochastic Methods}, Springer Verlag,
846: Berlin and Heidelberg, 1985.
847:
848: \bibitem{lipowsky}
849: R. Lipowsky, J. Phys. A {\bf 18}, L585 (1985).
850:
851: \bibitem{fogedby} H. D. Fogedby, Phys. Rev. E. {\bf 57}, 4943 (1998).
852: H. D. Fogedby, Cond-mat/0303632.
853:
854: \bibitem{krugs}
855: J. Krug, Adv. in Phys. {\bf 46}, 139 (1997).
856: J. Krug and H. Spohn, in {\it Solids far from equilibrium}, Ed. C. Godr\`eche,
857: Cambridge University Press, (1991).
858:
859: \bibitem{note} Observe that the presence of the function ``$\min$'' is the second
860: step of the algorithm induces a negative average velocity, implying that in
861: the continuum counterpart $\lambda$ has to be negative. On the other
862: hand the last step $\min{h(x,t),0}$ obviously generates an upper wall.
863:
864: \bibitem{note2} We use the subscript $_n$ to denote exponents related to the order
865: parameter $n$ and $_h$ for exponents associated with the average height.
866:
867: \bibitem{Haye3}
868: H. Hinrichsen, cond-mat/0302381.
869:
870: \bibitem{Ahlers} V. Ahlers and A. Pikovsky, Phys. Rev. Lett.
871: {\bf 88}, 254101 (2002). V. Ahlers, Ph. D. thesis.
872: http:// www.stat.physik.uni-potsdam.de/~volker/publ.html.
873: F. Ginelli, {\it et al.}, cond-mat/0302588.
874:
875: \bibitem{Maxi} M. San Miguel and R. Toral, {\it Stochastic Effects in
876: Physical Systems}, to be published in {\it Instabilities and
877: Nonequilibrium Structures}, VI, E. Tirapegui and W. Zeller,
878: eds. Kluwer Academic Pub. (1997). (Cond-mat/9707147).
879:
880: \bibitem{KPZproblem} T. J. Newman and A. J. Bray,
881: J. Phys. A {\bf 29}, 7917 (1996). C.H. Lam and F. G. Shin,
882: Phys. Rev. E {\bf 58}, 5592 (1998).
883:
884: \end{thebibliography}
885:
886: \end{document}