cond-mat0511672/PRE.tex
1: \documentstyle[12pt,epsfig]{article}
2: \oddsidemargin .5cm
3: \evensidemargin .5cm
4: \textheight 21truecm 
5: \textwidth 15truecm 
6: 
7: \author{Juli\'an Candia$^{a}$ and Ezequiel V. Albano$^{b}$\\{}\\
8: $^a${\small\it Departamento de F\'{\i}sica, UNLP, 
9: CC67, 1900 La Plata, Argentina}\\
10: $^b${\small\it Instituto de Investigaciones Fisicoqu\'{\i}micas
11: Te\'{o}ricas y Aplicadas}\\{\small\it (INIFTA), UNLP, CONICET, 
12: Suc.4, CC16, 1900 La Plata, Argentina}}
13: 
14: \title{Comparative study of an Eden model 
15: for the irreversible growth of spins and the equilibrium 
16: Ising model}
17: \begin{document}
18: \maketitle
19: 
20: \begin{abstract}
21: The Magnetic Eden Model (MEM) [N. Vandewalle et al., 
22: Phys. Rev. E. {\bf 50}, R635 (1994)]
23: with ferromagnetic interactions between
24: nearest-neighbor spins is studied in $(d+1)-$dimensional rectangular
25: geometries for $d = 1,2$. In the MEM, magnetic clusters are grown
26: by adding spins at the boundaries of the clusters. 
27: The orientation of the added spins
28: depends on both the energetic interaction with already deposited spins and the 
29: temperature, through a Boltzmann factor. 
30: A numerical Monte Carlo investigation of the MEM
31: has been performed and the results of the simulations have been analyzed 
32: using finite-size scaling arguments. 
33: As in the case of the Ising model, the MEM in $d = 1 $ is non-critical 
34: (only exhibits an ordered phase at $T= 0$). 
35: In $d = 2$ the MEM exhibits an order-disorder transition of second-order
36: at a finite temperature. Such transition has been characterized
37: in detail and the 
38: relevant critical exponents have been determined. These exponents are
39: in agreement (within error bars) with those of the Ising model in 
40: 2 dimensions. Further similarities between
41: both models have been found by evaluating the probability distribution 
42: of the order parameter, the magnetization and the susceptibility. 
43: Results obtained by means of extensive computer simulations allow us 
44: to put forward a conjecture which  establishes
45: a nontrivial correspondence between the MEM
46: for the irreversible
47: growth of spins and the equilibrium Ising model. This conjecture
48: is certainly a theoretical challenge and its
49: confirmation will contribute to the development
50: of a framework for the study of irreversible
51: growth processes.
52: \end{abstract}
53: 
54: \section{Introduction}
55: 
56: The study of kinetic growth models such as
57: directed percolation, Eden growth, ballistic
58: deposition, diffusion limited aggregation, random
59: deposition with and without relaxation, cluster-cluster
60: aggregation, etc., is motivated by their interest in
61: many areas of scientific research and technology such
62: as polymer science, crystal and polycrystalline growth,
63: gelation, fracture propagation, epidemic spreading,
64: bacterial and fungi growth colonies, 
65: colloids, etc. \cite{fam,bar,shl1,shl2,mar}.
66: Within this context the Eden model \cite{ede} has become an
67: archetype growth model. Eden clusters are compact but 
68: the self-affinity that characterizes the behavior
69: of the growing interface is of much interest 
70: (see e.g. \cite{ed1,ed2,ed3,ed4,ed5,ed6}). 
71: Few years ago  Ausloos et al. \cite{mem1} 
72: have introduced an additional degree
73: of freedom to the Eden model, namely the spin of the added particles.
74: More recently, the Eden growth of clusters of charged 
75: particles has also been studied \cite{qe}. 
76:  
77: In the Magnetic Eden Model (MEM) \cite{mem1} 
78: with spins having two 
79: orientations (up and down) the growth of the cluster starts
80: from a single seed, e.g.
81: a spin up seed, placed at the center of the two-dimensional square lattice,
82: whose sites are labelled by their rectangular coordinates $(i,j)$.
83: Then, the growth process of the resulting magnetic cluster  
84: consists in adding further spins to the growing cluster
85: taking into account the corresponding interaction energies. By analogy
86: to the Ising model \cite{ising} one takes $J$ 
87: as the coupling constant between
88: nearest-neighbor (NN) spins $S_{ij}$ and the energy $E$ is then given by
89: \begin{equation}
90: E = - \frac{J}{2} \sum_
91: { \langle ij,i^{'}j^{'}  \rangle } S_{ij}S_{i^{'}j^{'}} ,   
92: \end{equation}
93: where $\langle ij,i^{'}j^{'} \rangle$ means 
94: that the summation is 
95: taken over occupied NN sites.
96: The spins can assume two values, namely $S_{ij}= \pm 1$.
97: Throughout this work we set the Boltzmann constant equal
98: to unity ($k_{B} \equiv 1$), we
99: consider $J > 0$
100: (i.e., the ferromagnetic case) and we take the
101: absolute temperature $T$ measured in units of $J$.
102: In the MEM a spin is added to the cluster with a probability
103: proportional to the Boltzmann factor exp$(-\Delta E /T)$, 
104: where $\Delta E$ is the total energy change involved. 
105: It should be noted that at each step all sites of
106: perimeter are considered and the probabilities of adding up
107: and down spins have to be evaluated. After proper normalization
108: of the probabilities the growing site and the orientation of the
109: spin are determined through a pseudo-random number generator.
110:  
111: It is worth mentioning that the MEM has originally been motivated 
112: by the study of the structural properties of magnetically 
113: textured materials \cite{mem1}. While  these previous studies of the 
114: MEM were mainly devoted to determine the
115: lacunarity exponent and the fractal dimension of the set of parallel
116: oriented spins \cite{mem1}, the aim of the 
117: present work is to complement these
118: previous investigations by studying the critical behavior of the
119: MEM using extensive Monte Carlo simulations and applying a
120: finite-size scaling theory. Also, our study is performed
121: in confined (stripped) geometries which resemble
122: recent experiments where the growth of 
123: quasi-one-dimensional strips of Fe on a Cu(111) vicinal surface
124: \cite{iron} and Fe on a W(110) stepped substratum \cite{w110}  
125: have been performed. In fact, the preparation and
126: characterization of magnetized nanowires is of great interest
127: for the development of advanced 
128: microelectronic devices \cite{iron,w110,nw1,nw2}.  
129: Furthermore, the growth of metallic multilayers  
130: of Ni and Co separated by a Cu spacer
131: layer has recently been also studied \cite{cobre}.
132: 
133: Another goal of the present work is to compare the 
134: results obtained for the MEM with the 
135: well known behavior of the classical Ising model \cite{ising,Wu}, 
136: an archetypical  
137: model in the study of thermally driven (reversible)
138: phase transitions in equilibrium systems. The Ising 
139: Hamiltonian $(\bf{H})$ is given by
140: \begin{equation}
141: {\bf H} = - \frac{J}{2} \sum_
142: { \langle ij,i^{'}j^{'} \rangle} S_{ij}S_{i^{'}j^{'}} \ \ ,   
143: \end{equation}
144: where  $\langle ij,i^{'}j^{'} \rangle$ means 
145: that the summation runs over all NN sites,
146: $S_{ij} = \pm 1$ is the state of the spin at the site
147: of coordinates $(i,j)$ and $J$ is the 
148: coupling constant ( $J > 0$).
149: 
150: The MEM is also similar to a family of models for the stochastic growth
151: of crystals generically known as crystal growth models (CGM)
152: \cite{x,y,m,d,w}, for a review see e.g. \cite{q}.
153: As in the MEM, in the case of CGM each atom is adsorbed with 
154: a given probability conditional to the actual configuration of neighboring
155: atoms on the previous layer(s). However, in contrast to the MEM, the 
156: crystal is supposed to grow layer after layer. 
157: It should also be noticed that 
158: relationships established between CGM and a special class 
159: of Ising models \cite{m,w,z} have allowed to derive exact results.
160: Therefore, useful comparisons with the MEM will be also discussed
161: in the presentation of our results.
162: 
163: This paper is organized as follows: in Section 2
164: we give details on the simulation method, Section 3 is devoted
165: to the presentation and discussion of the results obtained for the
166: MEM in $(1 + 1)$-dimensions, while Section 4 refers to results 
167: corresponding to $(2 + 1)$-dimensions. In Sections 3 and 4,
168: detailed discussions comparing our results with the behavior of the 
169: Ising magnet are outlined.
170: Finally our conclusions are stated in Section 5.
171: 
172: \section{Description of the simulation method}
173: 
174: The MEM in $(1 + 1)-$dimensions is studied in the square lattice using a
175: rectangular geometry $L \times M$ with $M \gg L$ 
176: and imposing periodic boundary
177: conditions along the $L-$direction. The location of each site on the
178: lattice is specified through its rectangular coordinates $(i,j)$,
179: ($1 \leq i \leq M$, $1 \leq j \leq L)$.
180: The starting seed for the growing cluster is a column
181: of parallel oriented spins placed at $i=1$. 
182: It should be noticed that previous simulations
183: of the MEM were restricted to rather modest cluster sizes, i.e.
184: containing up to 8000 spins \cite{mem1}, while in the present
185: work clusters having up to $10^{9}$ spins have been typically grown.
186: We have also studied the MEM in $(2+1)-$dimensions employing
187: a $L \times L \times M$ geometry ($M \gg L$) with periodic 
188: boundary conditions along both $L-$directions.
189: 
190: \section{Study of the MEM in $(1+1)-$dimensions: results and discussion}
191: 
192: Magnetic Eden clusters grown on a stripped geometry of
193: finite linear dimension $L$ at low temperatures
194: show an interesting behavior that we call magnetization reversal.
195: In fact, we have observed that long clusters are constituted
196: by a sequence of well ordered magnetic domains. Spins belonging
197: to each domain, of average length $l_{D} \gg L$, have mostly the same
198: orientation and consecutive domains have opposite orientation.
199: Let $l_{R}$ be the characteristic length for the occurrence
200: of the magnetization reversal. Since $l_{R}\sim  L$, 
201: we then conclude that the problem has two characteristic 
202: length scales, namely $l_{D}$ and $l_{R}$,
203: such that $l_{D} \gg l_{R}$. 
204: 
205: In ordinary thermally driven phase transitions, the system
206: changes from a disordered state at high temperatures to a
207: spontaneously ordered state at temperatures below some critical
208: value $T_{c}$ where a second-order phase transition takes place.
209: Regarding the Ising model,
210: one has that, in the absence of an external magnetic
211: field ($H=0$), the low temperature ordered phase is a state with
212: non-vanishing spontaneous magnetization  ($ \pm M_{sp}$).
213: This spontaneous symmetry breaking is possible
214: in the thermodynamic limit only. In fact,
215: it is found that the magnetization $M$ of a finite sample 
216: can pass with a finite probability from a value near
217: $+M_{sp}$ to another
218: near $-M_{sp}$, as well as in the opposite direction. Consequently,
219: the magnetization of a finite system, averaged over a
220: sufficiently large observation time, 
221: vanishes at every positive temperature.
222: The equation $M(T,H=0) \approx 0$ holds if the observation time ($t_{obs}$)
223: becomes larger than the ergodic time ($t_{erg}$), which is defined as
224: the time needed to observe the system passing from $\pm M_{sp}$ 
225: to $\mp M_{sp}$. 
226: Increasing the size of the sample the ergodic time increases too,
227: such that in the thermodynamic limit ergodicity is
228: broken due to the divergence of the ergodic time, yielding
229: broken symmetry.
230: Since Monte Carlo simulations are restricted
231: to finite samples, the standard procedure to avoid the problems
232: treated in the foregoing discussion is to consider the
233: absolute magnetization as an order parameter \cite{kuku}.
234: Turning back to the MEM, we find  that the phenomenon of magnetization
235: reversal also causes the
236: magnetization of the whole cluster to vanish at every non-zero
237: temperature, provided that the length of the cluster $l_{C}$ (which plays the 
238: role of  $t_{obs}$) is much larger than $l_{D}$ (which plays the role
239: of $t_{erg}$ ).
240: Therefore, we have measured the mean absolute column magnetization, given by
241: \begin{equation}
242: |m(i,L,T)| = \frac{1}{L} |\sum_{j = 1}^{L} S_{ij}| \ \ .                       
243: \end{equation}
244: 
245: In the stripped geometry used in this work the bias introduced
246: by the lineal seed (a starting column made up entirely of up spins)
247: can be avoided by calculating relevant properties after 
248: disregarding spins within a distance approximately equal 
249: to few times $L$ from the seed.
250: The procedure of column averaging out from the
251: transient region represents a significant advantage
252: of the stripped geometry used for the
253: simulation of the MEM. In fact, when a single seed 
254: at the center of the sample is used, the definition
255: of the average magnetization of the whole cluster 
256: is strongly biased by the cluster's
257: kernel orientation at the early stages of the growing process.
258: In addition, using several randomly generated seeds
259: we could also establish that the system evolves into
260: a given stationary state independently of the seed employed.
261: 
262: \begin{figure}
263: \centerline{{\epsfxsize=4.0in\epsfysize=3.2in \epsffile{fig1.eps}}}
264: \caption{Plots of the probability distribution of the mean
265: column magnetization $P_{L}(m)$ versus $m$ for the fixed lattice
266: width $L=128$ and different temperatures, as indicated in
267: the figure. The sharp peaks at $m = \pm 1$ for $T=0.45$ have been
268: truncated, in order to allow a detailed observation of the plots
269: corresponding to higher temperatures.
270: This behavior resembles that of the one-dimensional
271: Ising model. More details in the text.}
272: \label{fig1}
273: \end{figure}         
274: 
275: The mean column magnetization
276: is a fluctuating quantity that can assume $L+1$ values. 
277: Then, for given values of both $L$ and $T$,
278: the probability distribution of the mean column
279: magnetization $(P_{L}(m))$ 
280: can be evaluated, since it represents the
281: normalized histogram of $m$ taken over a sufficiently large number
282: of columns in the stationary region \cite{alfa,beta,gamma}.
283: In the thermodynamic limit
284: the probability distribution $(P_{\infty}(m))$
285: of the order parameter of an equilibrium system
286: at criticality is universal (up to rescaling of the 
287: order parameter) and thus it contains
288: very useful and interesting information
289: on the universality class of the system \cite{pd1,pd2,pd3}. 
290: Figure 1 shows the thermal dependence of $P_{L}(m)$
291: for a fixed lattice size ($L=128$)
292: as obtained for the MEM.
293: At high temperatures $P_{L}(m)$ is a Gaussian
294: centered at $m=0$ but when the temperature gets lowered, the distribution
295: broadens and develops two peaks at $m=1$ and $m=-1$.
296: Further decreasing the temperature
297: causes these peaks to become dominant
298: while the distribution turns distinctly non-Gaussian, exhibiting a
299: minimum just at $m=0$. The emergence of the maxima
300: at $m= \pm 1$ is quite abrupt.
301: This behavior reminds us the order parameter probability
302: distribution characteristic of the one dimensional Ising model.
303: In fact, for the well studied $d-$dimensional Ising model \cite{gamma,lali}, 
304: we know that
305: for $T > T_{c}$, $P_{L}(M)$ is a Gaussian centered at $M=0$, given by
306: \begin{equation}
307: P_{L}(M) \propto \exp \left({-M^{2}L^{d}} \over {2T \chi} \right) \ \    ,    
308: \end{equation}
309: where the susceptibility $\chi$ is related to
310: order parameter fluctuations by
311: \begin{equation}
312: \chi = \frac {L^{d}}{T} \left(\langle M^{2} \rangle - \langle M \rangle^{2} \right)  .                 
313: \end{equation} 
314: 
315: \begin{figure}
316: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig2a.eps}}}
317: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig2b.eps}}}
318: \caption{Data for strip widths in the range $16 \leq L \leq 1024$
319: as indicated in the figures.
320: (a) ln-linear plots of $ L  \langle m^{2} \rangle $ versus $T^{-1}$.
321: The slope of the solid line (linear fit to the data) is $a=1.6$.  
322: (b) Plots of $\langle |m| \rangle $ versus $T$.
323: More details in the text.}
324: \label{fig2}
325: \end{figure}
326: 
327: Decreasing temperature the order parameter probability 
328: distribution broadens, it becomes non-Gaussian, and near $T_{c}$ it
329: splits into two peaks that get the more separated the lower the
330: temperature. For $T < T_{c}$ and linear dimensions $L$ much larger
331: than the correlation length $\xi $ of order parameter fluctuations,
332: one may approximate $P_{L}(M)$ near the peaks by a double-Gaussian
333: distribution, i.e.
334: \begin{equation}
335: P_{L}(M) \propto \exp \left( {-(M - M_{sp})^{2}L^{d}} \over {2T \chi} \right) +
336: \exp \left( {-(M + M_{sp})^{2}L^{d}} \over {2T \chi} \right) \ \ ,   
337: \end{equation}                          
338: where $M_{sp}$ is the spontaneous magnetization, while
339: the susceptibility $\chi$ is now given by 
340: \begin{equation}         
341: \chi = \frac{L^{d}}{T} \left(\langle M^{2}\rangle - \langle |M| \rangle ^{2} \right) \ \ .
342: \end{equation}
343: From equation (4) it turns out that 
344: the Gaussian squared width $\sigma^{2}$
345: associated with high temperature distributions is very close to
346: the 2nd moment of the order parameter, i.e.
347: \begin{equation}  
348: \sigma^{2} \approx \langle M^{2} \rangle \ \ .
349: \end{equation}  
350: Equation (8) is a consequence of the Gaussian shape of 
351: the order parameter probability
352: distribution and, thus, it holds for the MEM as well. 
353: From the known one-dimensional exact solution for a
354: chain of $L$ spins \cite{delta,eps} one can obtain
355: \begin{equation}   
356: \chi = \frac {1}{T} \exp (2/T) \ \;
357: \end{equation}   
358: then, equations (5) and (9) lead us to 
359: \begin{equation}  
360: \langle M^{2} \rangle = \frac {1}{L} \exp (2/T) \ \          
361: \end{equation}  
362: (where it has been taken into account that
363: $\langle M \rangle = 0$ due to finite-size effects, irrespective
364: of temperature).
365: From equations (8) and (10) we can see that the high
366: temperature Gaussian probability distribution broadens exponentially
367: as $T$ gets lowered, until it develops delta-like peaks at $M= \pm 1$ 
368: as a consequence
369: of a boundary effect on the widely extended distribution.
370: It should be noticed that for $d \geq 2$ this phenomenon 
371: is prevented by the finite
372: critical temperature which splits the Gaussian, as implied by equation (6).
373: 
374: Turning back to the MEM, figure 1 strongly suggests that an analogous
375: mechanism should be responsible for the thermal dependence exhibited by
376: the MEM's order parameter distribution function. 
377: So, by analogy to equation (9), we assume the relation
378: \begin{equation}   
379: \chi = \frac {1}{T} \exp (a/T) \ \
380: \end{equation}   
381: to hold for the MEM, where we have introduced a
382: phenomenological parameter $a$,
383: and the susceptibility $\chi$ is given by equation (5). We find an excellent
384: agreement to the data by choosing the value $a=1.6$ as
385: observed in figure 2(a), where ln-linear plots
386: of $ L \langle m^2 \rangle$ versus $1/T$ are shown for
387: strip widths varying in the range $16 \leq L \leq 1024$.
388: Figure 2(b) shows plots of $\langle |m|  \rangle$ versus $T$ 
389: for the same lattices.
390: This figure shows that increasing $L$
391: the order parameter curves approach the one that corresponds to
392: the thermodynamic limit (i.e., $\langle |m| \rangle = \theta(T)$, 
393: where $\theta$ is the Heaviside function).
394: 
395: \begin{figure}
396: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig3a.eps}}}
397: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig3b.eps}}}
398: \caption{Comparison of results corresponding to the 
399: $(1+1)-$MEM and the $d=1$ Ising model. 
400: (a) Plots of $\langle |m| \rangle $ versus $T$ obtained for a lattice of side $L=128$. 
401: (b) Linear-log plots of $\langle |M|_{I} \rangle (L,T) - \langle |m|_{MEM}
402: \rangle (L,T)$  versus $L^{-1}$
403: for $T=0.5$ and $T=1.0$. Hence, differences in the magnetization 
404: due to  finite-size effects appear to
405: vanish in the thermodynamic limit.}
406: \label{fig3}
407: \end{figure}
408: 
409: However, it should be pointed out that the
410: results obtained for the $(1+1)$-MEM and the $1$-dimensional Ising
411: model do not exactly coincide for finite lattices, 
412: as figure 3(a) shows for the case of the magnetization. Anyway,
413: this fact should not alarm us, since it can be seen that 
414: differences in the results obtained for both models are a direct 
415: consequence of the
416: finite-size nature of the lattices used in the simulations 
417: and consequently they tend to vanish in the
418: thermodynamic limit. This is actually shown by figure 3(b),
419: where log-linear plots of
420: $ \langle |M|_{Ising} \rangle (L,T) -  \langle |m|_{MEM}  \rangle (L,T)$ 
421: versus $L^{-1}$
422: for two different fixed values of temperature are presented.
423: Thus, we conclude that in view of the full qualitative and
424: quantitative agreement between both models we can safely
425: establish that, as in the 1d Ising model, 
426: the $(1+1)$-MEM is not critical (i.e. it also undergoes a
427: phase transition at $T_{c} = 0$).
428: 
429: \begin{figure}
430: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig4.eps}}}
431: \caption {Plots of $P(n_{NN})$ versus $T$ for $n_{NN}=1,2,3,4$ 
432: as indicated in the figure. 
433: The lines are guides to the eye. More details in the text.}
434: \label{fig4}
435: \end{figure}
436: 
437: We have also computed the number of already occupied NN 
438: sites every time that a
439: new particle was added to the spin system, and thus we have
440: obtained the normalized probability $P(n_{NN})$ of having $n_{NN}$ occupied
441: NN sites. Figure 4 shows the behavior of $P(n_{NN})$ as a function of
442: temperature. Using this probability we have evaluated 
443: $\langle n_{NN} \rangle = 2.0000(1)$ irrespective of the temperature.
444: This result can be understood considering 
445: that the growing process that leads to the assignment of a spin
446: $S_{ij} = \pm 1$ to each lattice site of coordinates $(i,j)$ can
447: be studied by means of a bond model. In fact, we can assign
448: a bond to each pair of neighboring sites, pointing  from the
449: earlier occupied site to the later occupied one. 
450: So, the process that leads
451: to a given spin configuration can be specified by the fields
452: $b_{U}(i,j)$ and $b_{R}(i,j)$, where sub-indexes $U$ and $R$ 
453: refer to the upper bond
454: of $(i,j)$ (i.e., the bond that connects the site of coordinates
455: $(i,j)$ with that of coordinates $(i,j+1)$),
456: and  to the right bond of the site of coordinates $(i,j)$, 
457: respectively \cite{note}. 
458: We take $b(i,j) = +1$ if the bond points outwards and 
459: $b(i,j) = -1$ if it is directed inwards. Therefore, 
460: the net bond flux at
461: a given lattice site $(i,j)$ is given by:
462: \begin{equation}  
463: \phi(i,j) = b_{U}(i,j) + b_{R}(i,j) - b_{U}(i,j-1) - b_{R}(i-1,j)      
464: \end{equation}  
465: and the possible 
466: values that $\phi$ can take are $\phi = -4, -2, 0, 2$. 
467: After some algebra, it follows that
468: $n_{NN} = \frac{1}{2} (4 - \phi)$
469: holds for every site on the lattice. Moreover, it can be seen that,
470: for an arbitrary $d$-dimensional lattice of coordination number $q$,
471: $n_{NN} = \frac{1}{2} (q - \phi)$. Then,
472: \begin{equation}  
473: \langle n_{NN} \rangle = \frac{1}{2} \ \ q                                  
474: \end{equation}
475: is the mean number of occupied NN, since $\langle \phi \rangle = 0$.
476: For the two-dimensional square lattice, $q = 4$ and equation (13) yields
477: $\langle {n_{NN}} \rangle = 2$,
478: in agreement with the result we have already obtained by means
479: of Monte Carlo simulations.
480: 
481: \begin{figure}
482: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig5.eps}}}
483: \caption{Plots of $P(\Delta E)$ versus $T$ for $\Delta E = 
484: 0, \pm 1, \pm 2, \pm 3, \pm4$ (in units of $J$) as indicated in the 
485: figure. The lines are guides to the eye. More details in the text.}
486: \label{fig5}
487: \end{figure}
488: 
489: Further insights into the MEM's growing process can be
490: gained by studying the mean energy change involved in the addition of
491: a new particle to the system. The process of adding a new spin
492: involves an energy change $\Delta E$ and from the definition 
493: of the $(1+1)-$MEM,
494: the possible values that $\Delta E$ can take are 
495: $0, \pm 1, \pm 2, \pm 3, \pm4$ (in units of $J$). Figure 5 shows 
496: plots of the normalized probability $P(\Delta E)$ versus $T$ 
497: for each of these values. 
498: The non-equilibrium nature of the MEM manifests itself
499: through much more complex probability distributions
500: $P(\Delta E)$ (see figure 5) 
501: than the corresponding to the equilibrium Ising model
502: where $\Delta E$ can take only five different values, namely
503: $0, \pm 2, \pm4$ (in units of $J$).
504: The results shown in  figures 4 and 5  
505: confirm the non-trivial nature of the link established between the MEM at
506: stationarity and the Ising model in equilibrium.
507: 
508: It should be noticed that for the case of crystal growth models (CGM)
509: \cite{x,y,m} the growing conditions are quite different than those of
510: the MEM. In fact, in CGM the crystal grows layer by layer   
511: in a given direction \cite{m,w}. Furthermore, the probability
512: distribution of the predecessor spin layer is sampled 
513: from the {\bf equilibrium} distribution, so will be the probability
514: of spins in subsequent layers. This particular growth mechanism
515: allow to establish dual transformations with the kinetic Ising model
516: \cite{d,w} and to extract some exact results. In contrast, the growing
517: interface of the MEM is self-affine and the system is far from equilibrium.
518: So, the link between the 1d Ising model and the $(1 + 1)-$dimensional
519: MEM is quite challenging. 
520: 
521: \section{Study of the MEM in $(2+1)-$dimensions: results and discussion}
522: 
523: \subsection{The order parameter and its probability distribution
524:  function}
525: 
526: \begin{figure}
527: \centerline{{\epsfxsize=4.0in\epsfysize=3.1in \epsffile{fig6.eps}}}
528: \caption{Plots of the probability distribution 
529: $P_{L}(m^{'})$ versus $m^{'}$ for the fixed lattice
530: size $L=16$ and different temperatures, as indicated in
531: the figure.
532: The occurrence of two maxima located at 
533: $m^{'} = \pm M_{sp}$ (for a given value of $M_{sp}$
534: such that $0 < M_{sp} < 1$)
535: is the hallmark of a thermal continuous phase transition
536: that takes place at a finite critical temperature.}
537: \label{fig6}
538: \end{figure}
539: 
540: In order to compare the $(2 + 1)-$dimensional MEM and the 2d Ising model,
541: we have first studied the order parameter probability
542: distribution $P_{L}(m^{'})$, where $m^{'}$ takes now $L^{2}+1$ 
543: possible values (see figure 6).
544: For high temperatures, the probability distribution
545: corresponds to a Gaussian centered at $m^{'}=0$. At lower
546: temperatures we observe the onset of two
547: maxima located at $m^{'} = \pm M_{sp}$ $(0 < M_{sp} < 1)$,
548: which become sharper and approach 
549: $m^{'} = \pm 1$ as $T$ is gradually decreased.
550: 
551: \begin{figure}
552: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig7.eps}}}
553: \caption{Plots showing
554: the location of the maximum of the probability distribution
555: as a function of temperature for both 
556: $(d+1)-$dimensional MEM models ($d=1,2$). The lines are guides to the eye.
557: The smooth transition
558: for d=2 constitutes another evidence of the
559: finite critical point associated with the $(2+1)-$MEM.}
560: \label{fig7}
561: \end{figure}
562: 
563: Figure 7 shows
564: the location of the maximum of the probability distribution
565: as a function of temperature for both 
566: $(d+1)$-dimensional MEM models, with  $d=1,2$.
567: While for the $d=2$ case we observe a smooth
568: transition from the $m^{'}_{max}=0$ value characteristic of high temperatures
569: to nonzero $m^{'}_{max}$ values that correspond to lower temperatures,
570: the curve obtained for $d=1$ shows, in contrast, a Heaviside-like jump.
571: As already discussed, the behavior exhibited by the
572: $(2+1)$-dimensional MEM 
573: (e.g., as displayed by figures 6 and 7) is the signature
574: of a thermal continuous phase transition that takes place at
575: a finite critical temperature.
576: 
577: The broken symmetry at a finite critical temperature $T_{c}$ implied by
578: the thermal continuous phase transition can be explained in terms of the
579: broken ergodicity that occurs in the system when we tend to
580: the thermodynamic limit ($L \rightarrow \infty$)
581: making use of the temperature dependence exhibited by the
582: order parameter distribution function.
583: In fact, if we set the characteristic length of MEM's
584: domains $l_{D}$ equal to an ergodic length $l_{erg}$,
585: we can carry out a complete analogy with the Ising model
586: by associating $l_{erg}$ to $t_{erg}$ (the Ising model ergodic time)
587: and the cluster's total length $l_{C}$ (already defined in Section 3)
588: to the Ising model observation time $t_{obs}$.
589: In this way, we encounter that
590: excursions of $m^{'}$ from $m^{'} = +M_{sp}$ to $m^{'} = -M_{sp}$ and
591: {\it vice versa} occur at
592: length scales of the order of  $l_{erg}$.
593: When the cluster's
594: total length becomes larger and larger ($l_{C} \gg l_{erg}$) the whole
595: cluster's magnetization is averaged to zero.
596: Furthermore, $l_{erg}$ diverges
597: as the strip's width becomes larger and larger, and again broken
598: symmetry arises as the consequence of broken ergodicity.
599: 
600: \subsection{Order-disorder phase transition in the $(2+1)$-dimensional MEM:
601: Finite-size effects and scaling analysis}
602: 
603: As already anticipated and as it follows from figures 6 and 7,
604: the $(2+1)$-dimensional MEM exhibits a 
605: thermally driven order-disorder transition at a finite temperature.
606: In the thermodynamic limit ($L \rightarrow \infty$) 
607: we expect to determine a
608: critical temperature $T_{c}$ such that $\langle |m^{'}| \rangle =0$ for $T > T_{c}$ 
609: while $\langle |m^{'}| \rangle$ remain non-vanishing at temperatures below $T_{c}$.
610: 
611: From the finite-size
612: scaling theory, developed for the treatment of
613: finite-size effects at criticality and under 
614: equilibrium conditions \cite{barba,priv}, 
615: it is well known that if a thermally driven
616: phase transition occurs at a temperature $T_{c} > 0$ in the
617: thermodynamic limit, then in a confined geometry this
618: transition becomes smeared out over the temperature region
619: $\Delta T(L)$ around a shifted effective transition temperature
620: $T_{c}(L)$, and the following relationships hold:
621: \begin{equation} 
622: \Delta T(L) \propto L^{-\theta}   \ \ ,
623: \end{equation} 
624: and 
625: \begin{equation}              
626: |T_{c}(L)-T_{c}| \propto  L^{-\lambda} \ \  ,
627: \end{equation}         
628: where the rounding and shift exponents are given by
629: $\theta = \lambda = \nu^{-1}$, respectively (recalling
630: that $\nu$ is the exponent that characterizes the divergence
631: of the correlation length at criticality).
632: 
633: Furthermore, from well established finite-size scaling relations,
634: the following Ans\"atze hold just at criticality $(T = T_{c})$ :
635: \begin{equation} 
636: \langle |m^{'}(L,T = T_{c})| \rangle \propto L^{-\beta/ \nu} 
637: \end{equation}
638: and
639: \begin{equation}                  
640: \chi_{max}(L)  \propto L^{\gamma/ \nu} \ \ ,   
641: \end{equation} 
642: where $\beta$ and $\gamma$ are the order
643: parameter and the susceptibility critical exponents,
644: respectively. Note that $\chi_{max}(L)$, as given by equation (17),
645: refers to the maximum of $\chi(L,T)$ as a function of $T$ for
646: fixed lattice size $L$.
647: 
648: \begin{figure}
649: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig8.eps}}}
650: \caption {Plots of $T_{cn}(L)$ versus $L^{-1}$ (for $n=1,2$). The
651: solid lines show the linear extrapolations that meet at the
652: critical point given by $T_{c}=0.69 \pm 0.01$.}
653: \label{fig8}
654: \end{figure}
655:  
656: In view of the encountered analogies between the MEM and the 
657: Ising model, it is natural to test the validity of equations 
658: (14-17) for the case of the MEM in $(2 + 1)$-dimensions. 
659: It should be noted that as in the case of equilibrium systems, 
660: in the present  case various ``effective" L-dependent critical 
661: temperatures can also be defined. In particular, we will define $T_{c1}(L)$ 
662: as the value that corresponds to $\langle |m^{'}| \rangle = 0.5$ for fixed $L$,
663: and $T_{c2}(L)$ as the one corresponding to the maximum of the
664: susceptibility for a given $L$, assuming that the susceptibility
665: is related to order parameter fluctuations in the same manner
666: as for equilibrium systems (as given by equations (5) and (7)).
667: Then, we should be able to obtain $T_{c}$ from
668: plots of $T_{cn}(L)$ versus $L^{-1}$  (for $n=1,2$),
669: as it is shown in figure 8.
670: Following this procedure we find that both $T_{c1}(L)$ and $T_{c2}(L)$
671: extrapolate (approximately) to the same value, allowing us 
672: to evaluate the critical temperature
673: $T_{c} = 0.69 \pm 0.01 $ in the thermodynamic limit. 
674: 
675: After determining $T_{c}$,
676: the correlation length exponent $\nu$
677: can be evaluated by means of equation (15), making the replacement
678: $\lambda = 1/ \nu$. In fact, taking $T_c$ 
679: at the mean, maximum and minimum
680: values allowed by the error bars, we obtain six log-log plots
681: of $|T_{cn}(L)-T_{c}|$ versus $L$ for $n=1,2$.
682: The slope of each of these plots, not shown here for the sake of space,
683: yields a value for $\nu$. The obtained values are:
684: $$
685:         \nu = 1.08 (T_{c}=0.68),\ \
686:         \nu = 1.00 (T_{c}=0.69),\ \ 
687: $$
688: \begin{equation} 
689:         \nu = 0.88 (T_{c}=0.70)\ \ \rm{for} \ \ n=1,
690: \end{equation}
691: and
692: $$
693:         \nu = 1.20 (T_{c}=0.68),\ \
694:         \nu = 1.08 (T_{c}=0.69),\ \ 
695: $$
696: \begin{equation}
697:         \nu = 0.95 (T_{c}=0.70)\ \  \rm{for} \ \ n=2.
698: \end{equation} 
699: 
700: Thus our estimate is given by $\nu = 1.04 \pm 0.16 $,
701: where the error bars reflect the
702: error derived from the evaluation of $T_{c}$, as well as the
703: statistical error.
704: 
705: \begin{figure}
706: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig9.eps}}}
707: \caption{Behavior of the susceptibility as a function of temperature.
708: Each curve shows a peak
709: which becomes sharper and shifts towards lower
710: temperatures as $L$ is increased. 
711: The inset shows in greater detail the peaks
712: corresponding to the smaller lattices ($L=16,24,32$).} 
713: \label{fig9}
714: \end{figure}
715: 
716: Figure 9 shows plots of the susceptibility versus $T$ as 
717: obtained using lattices of different side. It is found that the susceptibility
718: exhibits a peak which becomes sharper and shifts towards lower
719: temperatures when $L$ is increased. 
720: This behavior is, in fact,
721: already anticipated by equation (17), and it allows us to evaluate
722: $\gamma/\nu$ from the slope of a log-log plot of
723: $\chi_{max}$ versus $L$,
724: as figure 10 shows. The linear fit yields $\gamma/\nu = 2.02 \pm 0.04$.
725: Using this value and the value formerly obtained for $\nu$ we thus
726: determine $\gamma= 2.10 \pm 0.36$.
727: 
728: \begin{figure}
729: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig10.eps}}}
730: \caption {Log-log plot of $\chi_{max}$ versus $L$.
731: The linear fit (solid line) yields $\gamma/\nu = 2.02 \pm 0.04$.}
732: \label{fig10}
733: \end{figure}
734: 
735: \begin{figure}
736: \centerline{{\epsfxsize=4.0in\epsfysize=3.0in \epsffile{fig11.eps}}}
737: \caption{Log-log plots of 
738: $\langle m^{'}(T=T_{c}) \rangle$ versus $L^{-1}$ for the
739: mean, maximum and minimum allowed values of $T_{c}$.
740: The linear fits (solid lines) yield an estimate 
741: $\beta/\nu = 0.16 \pm 0.05$.}
742: \label{fig11}
743: \end{figure}
744: 
745: Figure 11 shows log-log plots of $\langle |m^{'}| \rangle(T=T_{c})$ versus $L$
746: for the mean, maximum and minimum allowed values of $T_{c}$.
747: Considering only the larger lattices, the linear fits to the data  
748: according to equation (16) yield the following estimates: $\beta/\nu = 0.11$, 
749: $\beta/\nu = 0.16$ and  $\beta/\nu = 0.19$. We then assume the value 
750: $\beta/\nu = 0.15  \pm 0.04$, where the error bars reflect the
751: error derived from the evaluation of $T_{c}$, as well as the
752: statistical error. From this value and the value formerly obtained 
753: for $\nu$ we thus determine $\beta= 0.16 \pm 0.05$. 
754:  
755: The critical exponents of the MEM in $(2 + 1)-$dimensions
756: as obtained using a finite-size scaling analysis are, so far:
757: $\nu = 1.04 \pm 0.16 $, $\gamma = 2.10 \pm 0.36$ and
758: $\beta = 0.16 \pm 0.05$.
759: If we recall the exactly known critical exponents of 
760: the $d = 2$ Ising model, i.e. $\nu = 1$, $\gamma = 7/4$ and
761: $\beta = 1/8$, we find that the $(2+1)-$MEM has the same 
762: critical exponents within error bars.
763: These results further support our conjecture on the connection between
764: the MEM in $(2 + 1)$-dimensions and the Ising model in $2$ dimensions.
765: 
766: As in the case of the MEM in $d = 1$, we have also computed 
767: the number of already occupied NN sites every time that a
768: new particle was added to the spin system. 
769: We found that the value $\langle n_{NN} \rangle = 3.0000(1)$ holds for all
770: temperatures, which is indeed the result given by equation (13),
771: since $q = 6$ for the three-dimensional square lattice.
772: 
773: At this stage, we may recall that for the 
774: $(1+1)-$MEM $\langle n_{NN} \rangle$ equals
775: the coordination number of the $d=1$ Ising model,
776: and that we found that both models
777: have the same critical temperature and exhibit the same critical behavior.
778: Reasoning by analogy, we may expect a coincidence between the critical
779: temperature for the $(2+1)-$MEM and the corresponding one for a
780: $d=2$ Ising model defined on a lattice of coordination number
781: $q=3$. However, this comparison cannot be carried out, since the
782: critical temperature of an Ising model depends on both
783: the coordination number $q$ and the topological structure of the lattice,
784: but for $d \geq 2$ and a given value of $q$ the
785: topological structure is not unique.
786: For instance, for $d=2$ and $q=3$, we can pass from the honeycomb lattice (HL)
787: to the expanded Kagom\'{e} lattice (EKL) through the application 
788: of a star-triangle
789: transformation and obtain the exact values of their critical points,
790: which turn out to be \cite{fisher} 
791: $ T_{c} = 1.5187$ (HL) and $ T_{c} = 1.4530$ (EKL).
792: 
793: \section{Conclusions}
794: 
795: In the present work we have studied the growth of magnetic
796: Eden clusters with ferromagnetic interactions between
797: nearest neighbor spins in a $(d+1)-$dimensional rectangular
798: geometry (for $d=1,2$), using Monte Carlo simulations and applying a
799: finite-size scaling theory.
800: The results obtained allow us to conjecture
801: a nontrivial correspondence between the MEM for the irreversible
802: growth of magnetic materials and the classical Ising model
803: under equilibrium conditions. In fact, we have 
804: found that the $(d+1)-$dimensional
805: MEM and the $d-$dimensional Ising model behave identically 
806: (unless finite-size differences that vanish in the
807: thermodynamic limit) at criticality,
808: i.e. that both models belong to the same universality class.
809: We also conjecture that
810: this correspondence would remain at higher dimensions $(d>2)$. 
811: The results obtained strongly suggest a
812: link between the temporal evolution of equilibrium systems and
813: the stationary growth of nonequilibrium systems.
814: We thus believe that this work will stimulate further developments
815: in the field of nonequilibrium kinetic growth models. A more 
816: precise numerical test of the posed conjecture will certainly 
817: require a considerable computational effort but it will be 
818: of great interest.  Furthermore, analytical
819: developments aimed to establish a theoretical framework
820: for the understanding of far from equilibrium growth phenomena
821: will become stimulated by the reported findings.
822: 
823: \section*{Acknowledgments} This work was supported  financially  by
824: CONICET, UNLP, CIC (Bs. As.), ANPCyT and Fundaci\'on Antorchas 
825: (Argentina) and the Volkswagen Foundation (Germany). The authors
826: thank M. Mu\~noz for helpful discussions and his 
827: proposal of the bond representation for the MEM.
828: 
829: \begin{thebibliography}{99}
830: \bibitem{fam} F. Family and T. Vicsek, Dynamics of Fractal Surfaces.
831: World Scientific, Singapore (1991).
832: \bibitem{bar} A. L. Barabasi and H. E. Stanley, Fractal Concepts in
833: Surface Growth. Cambridge University Press, New York (1995).
834: \bibitem{shl1} Fractals and Disordered Media,
835: Eds. A. Bunde  and S. Havlin, Springer-Verlag, Heidelberg (1991).
836: \bibitem{shl2} Fractals in Science, Eds.  A. Bunde  and S. Havlin, 
837: Springer-Verlag, Heidelberg (1995).
838: \bibitem{mar} M. Marsili, A. Maritan, F. Toigo and J. R. Banava. 
839: Rev. Mod. Phys. {\bf 68}, 963 (1996).
840: \bibitem{ede} M. Eden, in: Symp. on Information Theory in Biology,
841: edited by H. P. Yockey, Pergamon Press, New York (1958) and
842: Proceedings of the Fourth Berkeley Symposium
843: on Mathematics, Statistics and Probability, edited by F. Neyman
844: University of California Press, Berkeley (1961). Vol. IV, p. 223.
845: \bibitem{ed1} D. E. Wolf and J. Kert\'esz. Europhys. Lett.
846: {\bf 4}, 651 (1987).
847: \bibitem{ed2} M. Cieplak, A. Maritan and J. R. Banavar.
848: Phys. Rev. Lett. {\bf 76}, 3754 (1996).
849: \bibitem{ed3} C. S. Ryu and In-mook Kim, Phys. Rev . E. 
850: {\bf 53}, 5643 (1996).
851: \bibitem{ed4} S. S. Manna and D. Dhar, Phys. Rev. E, 
852: {\bf 54}, R3063 (1996).
853: \bibitem{ed5} E. V. Albano, Phys. Rev. E. {\bf 56}, 7301 (1997). 
854: \bibitem{ed6} M. Marsili and M. Vendruscolo. 
855: Europhys. Lett. {\bf 37}, 505 (1997).
856: \bibitem{mem1} M. Ausloos, N. Vandewalle and R. Cloots. Europhys. Lett.,
857: {\bf 24}, 629 (1993), 
858: N. Vandewalle and M. Ausloos. Phys. Rev. E. {\bf 50}, R635 (1994). 
859: \bibitem{qe} Y. V. Ivanenko, N. I. Lebovka and N. V. Vygornitskii,
860: Eur. Phys. J. B. {\bf 11}, 469 (1999).
861: \bibitem{ising} E. Ising, Z. Phys. {\bf 31}, 253 (1925). 
862: \bibitem{iron} J. Shen et al. Phys. Rev. B. {\bf 56}, 2340 (1997).
863: \bibitem{w110} O. Pietzsch, A. Kubetzka, M. Bode and R.
864: Wiesendanger. Phys. Rev . Lett. {\bf 84}, 5212 (2000). 
865: \bibitem{nw1} J. Hauschild, U. Gradmann and H. J. Elmers.
866: Appl. Phys. lett. {\bf 72}, 3211 (1998).
867: \bibitem{nw2} H. J. Elmers, J. Hauschild and U. Gradmann.
868: Phys. Rev. B. {\bf 59}, 3688 (1999).
869: \bibitem{cobre} U. Bovensiepen et al. Phys. Rev. Lett. {\bf 81},
870: 2368 (1998).
871: \bibitem{Wu} B. M. McCoy and T. T. Wu. The two dimensional
872: Ising model. Harvard University Press, Cambridge, MA. (1973). 
873: \bibitem{x} T. R. Welberry and R. Galbraith. J. Appl. Cryst.
874: {\bf 6}, 87 (1973).
875: \bibitem{y} A. M. Verhagen. J. Stat. Phys. {\bf 15}, 219 (1976).
876: \bibitem{m} I. G. Enting. J. Phys. C. (Solid State Physics).
877: {\bf 10}, 1379 (1977).
878: \bibitem{d} P. Ruj\'an. J. Stat. Phys. {\bf 29}, 231 (1982).
879: \bibitem{w} P. Ruj\'an. J. Stat. Phys. {\bf 34}, 615 (1984).
880: \bibitem{q} P. Ruj\'an. J. Stat. Phys. {\bf 49}, 139 (1987).
881: \bibitem{z} J. Stephenson, Phys. Rev. B. {\bf 1}, 4405 (1970)
882: and J. Math. Phys. {\bf 11}, 420 (1970).
883: \bibitem{kuku} K. Binder and D. W. Heermann, in Monte Carlo Simulation
884: in Statistical Physics, Springer Series in Solid State
885: Sciences 80, Springer-Verlag, Berlin (1992).
886: \bibitem{alfa} K. Binder and D. Stauffer, A simple introduction to
887: Monte Carlo simulation and some specialized topics, in Applications of
888: the Monte Carlo Method in Statistical Physics, Ed. K. Binder, 2nd Edition.
889: Springer-Verlag, Berlin (1987).
890: \bibitem{beta} K. Binder, Monte Carlo Methods, in Encyclopedia of Appl. Phys.
891: VCH Publishers Inc., Heidelberg (1994). Vol. 10, 567.
892: \bibitem{gamma} K. Binder, Introduction to Monte Carlo Methods, 
893: Chapter 5, p.124,
894: in Conference Proceedings Vol. 49, Monte Carlo and Molecular Dynamics of
895: Condensed Matter Systems, Eds. K. Binder and G. Ciccotti, SIF, Bologna (1996).
896: \bibitem{pd1} K. Binder. Z. Phys. B {\bf 43}, 119 (1981). 
897: \bibitem{pd2} A. D. Bruce, J. Phys. C. {\bf 14}, 3667 (1981).
898: \bibitem{pd3} M. M. Tsypin and H. W. J. Bl\"ote,
899: Phys. Rev. E. {\bf 62}, 73 (2000).
900: \bibitem{lali} D. P. Landau and E. M. Lifshitz, Statistical Physics,
901: 3rd Edition, Part 1. Pergamon Press, Oxford. (1980).
902: \bibitem{delta} R. J. Baxter, Exactly Solved Models in Statistical
903: Mechanics, Academic Press (1982).
904: \bibitem{eps} K. Huang, Statistical Mechanics, 2nd Edition,
905: John Wiley \& Sons, (1987),  page 363. 
906: \bibitem{note} Since we are dealing with periodic boundary conditions in
907: the $j$-direction, we consider $j-1 = L$ if $j=1$ and analogously 
908: $j+1 = 1$ if $j=L$ throughout.
909: \bibitem{barba} M. N. Barber, in ``Phase transitions and 
910: critical phenomena'', Edited by C. Domb and J. L. Lebowitz. 
911: (Academic, New York). (1983), Vol 8, page 146.  
912: \bibitem{priv} V. Privman(Ed.). ``Finite size scaling and 
913: numerical simulations of statistical systems''. World Scientific,
914: Singapore. (1990).
915: \bibitem{fisher} M. Fisher. Phys. Rev. {\bf 113}, 969 (1959).
916: \end{thebibliography}
917: \end{document}
918: 
919: 
920: 
921: 
922: 
923: 
924: 
925: 
926: 
927: 
928: 
929: 
930: 
931: 
932: 
933: