physics0008245/la.tex
1: \documentstyle[twoside,multicol,aps,epsf]{revtex}
2: \topmargin -30pt \headsep 0.0in
3: \textwidth=18.2cm
4: \textheight=58\baselineskip   
5: \evensidemargin = -.35in
6: \oddsidemargin = -.3in
7: \begin{document}
8: \author{J. P. Neirotti \\
9: and\\
10: David L. Freeman \\ Department of Chemistry, University of Rhode
11: Island\\
12: 51 Lower College Road, Kingston, RI 02881-0809\\
13: and\\
14: J. D. Doll\\
15: Department of Chemistry, Brown University\\
16: Providence, RI 02912}
17: \title{The Approach to Ergodicity in Monte Carlo Simulations}
18: \date{\today}
19: \maketitle
20: 
21: \makeatletter
22: \global\@specialpagefalse
23: \def\@evenfoot{\thepage\hfill{\small\sf The Approach to Ergodicity in Monte Carlo Simulations}} 
24: \def\@oddfoot{
25: \ifnum\c@page>1
26:   {\small\sf J. P. Neirotti, D. L. Freeman. and J. D. Doll}\hfill\thepage
27: \fi
28: \ifnum\c@page=1
29:   \hfill\thepage\hfill\llap{\protect{PREPRINT}}
30: \fi
31: } 
32: \makeatother
33: \vspace{.5cm}
34: \begin{center}
35: {\bf Abstract}
36: \end{center}
37: 
38: \begin{abstract}
39: The approach to the ergodic limit in Monte Carlo simulations is
40: studied using both analytic and numerical methods.
41: With the help of a stochastic model, a metric is defined that enables the
42: examination of a simulation in both the ergodic and non-ergodic regimes.  
43: In the non-ergodic regime, the model implies how the simulation is expected
44: to approach ergodic behavior analytically, and the analytically inferred
45: decay law of the metric
46: allows the monitoring of the onset of ergodic behavior. The
47: metric is related to previously defined measures developed for molecular
48: dynamics simulations, and the metric
49: enables the comparison of the relative efficiencies of different
50: Monte Carlo schemes.
51: Applications to Lennard-Jones 13-particle clusters are shown
52: to match the model for
53: Metropolis, J-walking and parallel tempering based approaches.  The
54: relative efficiencies of these three Monte Carlo approaches are
55: compared, and the decay law is shown to be useful
56: in determining needed high temperature parameters in parallel tempering
57: and J-walking studies of atomic clusters.
58: 
59: 
60: \noindent{\bf PACS} numbers: 05.10.Ln, 02.70.Lq
61: \end{abstract}
62: 
63: \begin{multicols}{2}
64: 
65: \section{Introduction}
66: 
67: A goal of Monte Carlo (MC) simulations in statistical 
68: mechanics \cite{valleau} is the calculation of ensemble mean values of 
69: thermodynamic quantities. Ensemble mean values are multidimensional 
70: integrals over configuration space
71: 
72: \begin{equation}
73: \label{eq0}
74: \langle U\rangle = \int\!\!d{\bf x}\,P({\bf x})\,U({\bf x})\,\,,
75: \end{equation}
76: 
77: \noindent where $P({\bf x})$ is the probability of finding a system 
78: in the state defined by ${\bf x}$, and the functional form 
79: of $P({\bf x})$ depends on 
80: the ensemble investigated. MC simulations usually generate a sampling 
81: of configuration space $\left\{{\bf x}_k\right\}_{k=1}^K$ by the use of 
82: a stochastic process with stationary probability $P({\bf x}_k)$. The 
83: quantity $U$ evaluated at ${\bf x}_k$ is the output of the simulation 
84: $U({\bf x}_k)=U_k$, and its arithmetic mean value $\overline{U}$, 
85: in principle, must approach the ensemble mean value.\cite{valleau} 
86: In this paper we refer to the set of configurations generated in a Monte
87: Carlo simulation as a time sequence, and
88: we study the behavior of these temporal sequences $\left\{U_k\right\}$ and 
89: their arithmetic mean, to understand better how MC simulations approach 
90: ergodic behavior.  It
91: is important to emphasize that there are two time variables to consider.
92: The time variable $k$ labels the separate configurations generated in a
93: Monte Carlo walk.  Variations of properties with $k$ provide information
94: about the short-time behavior of a MC simulation.  The time
95: variable $K$ labels the total length of the MC walk, and variations of
96: computed properties with $K$ provide information about the convergence
97: of the simulation on a long time scale.
98: 
99: Given an infinite time, the stochastic walker in a MC simulation visits 
100: every allowed point in configuration space.\cite{wood} Ergodic behavior is reached 
101: when the length of the walk is sufficiently long to sample configuration 
102: space appropriately.\cite{thiru1} In practice, this does not mean that 
103: the space has been densely covered but that every region with non-negligible 
104: probability has been reached. In such a case
105: we can say that the simulation is effectively 
106: ergodic or that it has reached the ergodic limit.
107: 
108: For a finite walk, in the event of broken ergodicity \cite{palmer}, 
109: phase space is effectively disconnected. The different disconnected 
110: regions (called components) are separated by barriers of zero
111: effective probability. If a stochastic walker starts its walk in one 
112: of these regions, it may not cross the barriers within the time of the 
113: simulation. If the simulation length is increased, some barriers may become 
114: accessible for the walker and phase space is better sampled. We can 
115: conclude that a time $\tau$ exists such that, for simulation lengths shorter 
116: than $\tau$, the walker becomes trapped in one of the phase space components. 
117: For simulation lengths much larger than $\tau$, phase space is effectively 
118: covered by the walker.
119: 
120: In this study we imagine a system having more than one time scale
121: $\tau_1\ll\tau_2\ll\dots\ll\tau_{\Lambda}$. In a Monte Carlo simulation
122: each scale comes
123: from stochastic processes with different correlation times.\cite{gardiner}
124: A precise definition of the correlation times for Monte Carlo
125: processes is given in Section III, but for the moment we can think of
126: these correlation times as identical to physical time scales of the system
127: under study.  To
128: understand these time scales more fully,
129: it is useful to focus on an example.
130: Prototypical of systems having such disparate time scales are
131: atomic and molecular clusters.  Typical cluster potential surfaces have
132: many local minima separated by significant energy barriers. 
133: \cite{freeman1,wales1,wales2}
134: The local
135: minima can be grouped into basins of similar energies, with each basin
136: separated from other basins again by energy barriers.
137: At short
138: Monte Carlo times a cluster system executes small amplitude oscillations
139: about one of its potential minima.  We can think of these vibrational
140: time scales as the shortest time scales that define a cluster system. 
141: As the simulation time is extended the system eventually hops between different
142: local minima within the same basin.  
143: The time scale for the first hops between local minima
144: can be considered the next shortest time scale for the simulation.  At
145: still longer Monte Carlo times, the system hops between different energy
146: basins defining yet another time scale for the simulation.  This
147: grouping of time scales continues until the longest time scale for a
148: given system is reached.  At Monte Carlo times that are long compared to
149: this longest time scale, the simulation is ergodic.
150: 
151: Consider a system with several time scales as mentioned above. 
152: If the length of the simulation is smaller than the smallest correlation 
153: time, the walker may become trapped in an effectively disconnected region 
154: and the sampling of phase space is incomplete. By increasing the time, 
155: the memory of the initial condition in the sampling decreases as 
156: the walker crosses to other previously unreachable regions. These oscillations
157: and hoppings can be modeled by a superposition of stochastic processes with 
158: different correlation
159: times. These processes with non-zero correlation times are known as {\em 
160: colored noise} processes (as opposed to zero correlation time {\em white 
161: noise} processes). \cite{gardiner} From the study of the autocorrelation 
162: functions of a stochastic model defined using these colored noise processes, 
163: we can verify that, at a fixed run length $K$, there exist two different 
164: groups of processes; those that contribute to the autocorrelation function with
165: terms that decay like
166: $1/k$ (called {\em diffusive} processes), and those that contribute to the 
167: autocorrelation function with terms that decay slower than $1/k$ (called {\em 
168: non-diffusive} processes). When the time of the simulation is increased, some
169: non-diffusive processes at shorter run lengths, start to contribute to the
170: autocorrelation function like diffusive processes. After the walk length reaches
171: the largest correlation time $\tau_{\Lambda}$, all processes contribute to the
172: autocorrelation function with terms that decay like $1/k$. At this point, the
173: simulation is at the diffusive regime and effective ergodicity has been reached.
174: A principal
175: goal of this work is 
176: to investigate the way in which the MC output $\left\{U_k\right\}$ reaches 
177: the diffusive limit (i.e. the ergodic limit) by studying the properties 
178: of autocorrelation functions under changes of scale in time, 
179: $K\to bK$ with $b>1$. 
180: By time scaling 
181: it is possible to infer the decay law of the non-diffusive contributions
182: with respect to the total simulation
183: time $K$.  The functional dependence of the non-diffusive contributions
184: on the parameter $b$ that is used to scale $K$ is determined empirically.
185: We have found the decay law so determined 
186: to be a particularly valuable method of concluding when a simulation
187: can be considered ergodic.  Unlike previous studies
188: \cite{thiru1,thiru2,thiru3,thiru4}
189: that only have
190: investigated the behavior of certain autocorrelation functions in the
191: ergodic regime, by focusing on the approach to ergodic behavior we have
192: a more careful monitor of the onset of ergodicity.
193: Once the non-diffusive 
194: contributions have decayed to a point where they are too small to be
195: distinguished from zero to within the fluctuations of the calculation,
196: we can say that the ergodic limit has been reached. 
197: 
198: The autocorrelation functions we use to measure the approach to the ergodic
199: limit are based on
200: one of the probes of ergodicity developed by Thirumalai and 
201: co-workers \cite{thiru1,thiru2,thiru3,thiru4}, 
202: and is often called the {\em energy metric}. 
203: The energy metric has been proposed as an alternative to other 
204: techniques \cite{thiru1} 
205: (like the study of the Lyapunov exponents \cite{lieberman}) for the study 
206: of ergodic properties in molecular dynamics (MD) simulations.  The
207: metric has been used to study the relative efficiency of MC simulation
208: methods as well. \cite{straub}
209: The MC metric as used in the current work 
210: can easily be extended from the energy to
211: other scalar observables of the system. 
212: 
213: We present two key issues in this paper. First, from the knowledge of the 
214: decay law of the non-diffusive contributions to the MC metric, we infer how 
215: long a simulation must be to be considered effectively ergodic. Second, once 
216: the ergodic limit is reached, we 
217: can compare the results from different numerical algorithms 
218: to measure relative efficiencies. Because the outcomes of MC simulations are 
219: noisy, we have found it useful to separate diffusive and 
220: non-diffusive terms in the MC metric with a Fourier analysis so that we
221: can neglect
222: the high frequency components of the noise.
223: This technique has given reproducible
224: results.
225: 
226: To test the match between the stochastic model and actual Monte Carlo
227: simulations, we examine the approach to ergodic behavior in simulations of
228: Lennard-Jones clusters.  Recently
229: \cite{juan1,juan2}
230: we have studied the thermodynamic
231: properties of Lennard-Jones clusters as a function of temperature using
232: both J-walking 
233: \cite{frantz}
234: and parallel tempering methods.  
235: \cite{p1,p2,p3}
236: Both simulation
237: techniques require an initial high temperature that must be ergodic when
238: Metropolis Monte Carlo methods 
239: \cite{metropolis}
240: are used.  If the Metropolis method does
241: not give ergodic results at the initial high temperature, systematic
242: errors propagate to the lower temperatures in J-walking and parallel
243: tempering simulations, and the results can be flawed or meaningless.  In
244: most Monte Carlo simulations of clusters at finite temperatures, 
245: \cite{lee,labastie}
246: the
247: clusters are defined by enclosing the atoms within a constraining
248: potential about the center of mass of the system.  The constraining
249: potential is necessary because clusters at finite temperatures have
250: finite vapor pressures, and the association of any one atom with the
251: cluster can be ill-defined.  From experience 
252: \cite{juan1,juan2,juan3}
253: we have found that if the
254: radius of the constraining potential and the initial high temperature
255: are not both carefully chosen,
256: it can be difficult to
257: attain ergodicity with Metropolis methods.  A key concern then is the
258: choice of constraining radius and the choice of initial temperature.  We
259: verify the stochastic model by investigating Monte Carlo simulation
260: results as a function of the temperature and the size of the
261: constraining potential.
262: 
263: The contents of the remainder of this paper are as follows.  In Section
264: II we motivate the studies that follow by examining numerally the
265: behavior of a set of Monte Carlo simulations of a 13-particle
266: Lennard-Jones cluster.  This cluster system is used to illustrate the
267: results throughout this paper.
268: In Section
269: III we introduce the stochastic model based on a continuous time
270: sequence.  In Section IV we extend the model to discrete time sequences
271: characteristic of actual Monte Carlo simulations.  In Section V we test
272: the discrete stochastic model with applications to Lennard-Jones
273: clusters and in Section VI we summarize our conclusions.  Many of the key
274: derivations needed for the developments are found in two appendices.
275: 
276: 
277: \section{An Example Calculation}
278: 
279: Before discussing the major developments of this work, it is useful to
280: understand the nature of the problem we are attempting to solve by
281: examining some numerical results on a prototypical system.  We take the
282: 13-particle Lennard-Jones cluster defined by the potential function
283: 
284: \begin{equation}
285: V({\bf x}) = 4\varepsilon 
286: \sum_{i=2}^N\sum_{j=1}^{i-1}\left[\left(\frac{\sigma}{r_{ij}}\right)^{12} - 
287: \left(\frac{\sigma}{r_{ij}}\right)^{6}\right]
288: + \sum_{i=1}^NV_C( \vec{x}_i,R_c)
289: \label{pot}
290: \end{equation}
291: 
292: 
293: \noindent where $\varepsilon$ and $\sigma$ 
294: are the standard Lennard-Jones energy and length parameters,
295: $N$ is the number of particles in the cluster (13 in the present case), 
296: $r_{ij}$ is the distance 
297: between particles $i$ and $j$
298: 
299: \begin{equation}
300: \label{rij}
301: r_{ij} = |\vec{x}_i-\vec{x}_j|,
302: \end{equation}
303: 
304: \noindent and $V_C$ is the constraining potential discussed in Sec. I
305: 
306: \begin{equation}
307: \label{cp}
308: V_C(\vec{x}_i,R_c) = \left\{
309: \begin{array}{ll}
310: 0&\left|\vec{x}_i-\vec{X}_c\right| < R_c\\
311: \infty&R_c<\left|\vec{x}_i-\vec{X}_c\right|  
312: \end{array}
313: \right.\,\,,
314: \end{equation}
315: 
316: \noindent where $\vec{X}_c$ is the coordinate of the 
317: center of mass of the cluster and
318: $R_c$ is the radius of the constraining sphere. 
319: The 13-particle Lennard-Jones cluster has a complex
320: potential surface with many minima separated by significant energy
321: barriers, 
322: \cite{freeman1,wales1,wales2}
323: and ergodicity problems associated with the simulation of
324: properties of this system are well-known.
325: \cite{frantz}
326: We now consider a Metropolis
327: MC simulation of the average potential energy of the system in the
328: canonical ensemble at temperature $k_BT/\varepsilon=0.393$($k_B$ is the 
329: Boltzmann constant).  This average
330: potential energy $\overline{V}_k$ is defined by
331: \begin{equation}
332: \overline{V}_k=\frac{1}{k} \sum_{k'=1}^k V_{k'}
333: \end{equation}
334: 
335: \noindent and is displayed in the upper panel of Fig. 1 as a function
336: of the walk length $k$ for 20 independent simulations each
337: initialized from a random configuration.  Over the maximum time scale 
338: $K$ of the
339: walks, it apparent that the potential energy averaged over each
340: independent walk has not converged to the same result.  Such
341: unreproducible behavior is symptomatic of a simulation not yet at the
342: ergodic limit.
343: 
344: \begin{figure}
345: \epsfxsize=.48\textwidth
346: \epsfbox{f1.ps}
347: \end{figure}
348: 
349: {\footnotesize
350: {\bf Fig. 1:} The upper panel shows the ``time evolution" of 
351: $\overline{V}_k$ (in units of $\varepsilon$) 
352: for $M=20$ independent experiments. The lower panel 
353: shows $d_k$ (in units of $\varepsilon^2$) vs. $k$ for the experiments of the 
354: upper panel. 
355: $R_c$ has been set to $4\sigma$ and  $k_BT/\varepsilon=0.393$. At least 
356: two basins with different energies are present. Clearly, 
357: $d_k$ goes to a constant when $k$ is increased within the total time
358: scale of the simulation.}
359: \vspace{.5cm}
360: 
361: At the ergodic limit (i.e. for the maximum walk length $K$ greater than
362: that included in Fig. 1) the averages displayed in the upper panel of Fig. 1
363: must approach the same value for each walker.  Using related ideas
364: developed elsewhere,
365: \cite{thiru1,thiru2,thiru3}
366: the extent to which the walks approach the same limit can be measured in
367: terms of a metric $d_k$ defined by
368: 
369: \begin{equation}
370: \label{omega1}
371: d_k = \frac2{M(M-1)} \sum_{i=2}^M \sum_{j=1}^{i-1} 
372: \left[\,\overline{V}^{(i)}_k - \overline{V}^{(j)}_k\right]^2\,\,,
373: \end{equation}
374: 
375: \noindent In Eq. (\ref{omega1}) $M$ represents the number of independent walks, 
376: and
377: $\overline{V}^{(i)}_k$ is the 
378: average potential energy computed in walk $i$ at MC 
379: time $k$.  The metric measures the energy fluctuations in the walk as a
380: function of the walk length.  For an ergodic simulation, the metric must
381: decay to zero.
382: For the 20 simulations of the 13-particle
383: Lennard-Jones cluster, the metric as a function of $k$ is plotted in the
384: lower panel of Fig. 1.  Rather than asymptotically approaching zero, over
385: the short length of the walk displayed here, $d_k$ has decayed to a
386: constant, and as discussed later in this paper, over the time scale of
387: this simulation, $d_k$ can be qualitatively represented by the function
388: 
389: \begin{equation}
390: d_k=\frac{A_K}{k}+B_K
391: \end{equation}
392: 
393: \noindent where $A_K$ and $B_K$ are coefficients that are dependent
394: on the total walk
395: length $K$. As $K$ is increased to a time
396: where the walk is ergodic,
397: $B_K$ must decay to zero.  Major goals of this work are to understand
398: how $B_K$ decays and to use the discovered decay law to determine the
399: onset of ergodic behavior.
400: Our approach is to introduce first a continuous
401: stochastic model of a simulation followed by a discrete model more
402: clearly linked to actual MC studies.
403: 
404: \section{Stochastic Model}
405: 
406: We have discussed in the introduction how the output of MC simulations can be 
407: considered to be 
408: a combination of stochastic processes with different time scales, 
409: and how the contributions to autocorrelation function from these
410: processes can vary when the length of the 
411: simulation is enlarged. Here we present a continuous time model for the 
412: stochastic processes that occur in a simulation. Even though a MC simulation 
413: occurs 
414: in a discrete time (each MC point represents a time unit), we find that the 
415: continuous model helps to understand better the ideas used in the 
416: modeling of the MC output.
417: 
418: In this section the ensemble mean value is used to find the expression 
419: for the autocorrelation functions of the model. Although in actual numerical 
420: calculations the ensemble mean is replaced by a mean over a finite number of 
421: independent experiments, the results obtained here give information about the 
422: limit of an infinite sample.
423: 
424: The stationary process used to sample space is a stochastic 
425: process. We assume the output of the MC simulation can be modeled by a linear 
426: superposition of stochastic processes with different 
427: correlation times  $\tau_{\ell}\ge 0$, 
428: 
429: \begin{equation}
430: \label{eq:2}
431: U(t) = U_c +  \sqrt{\Gamma_0} \,\xi(t)
432:         + \sum_{\ell=1}^{\Lambda} 
433: \sqrt{\Gamma_{\ell}}\,g_{\ell}(t/\tau_{\ell})\,\,,
434: \end{equation}
435: 
436: \noindent where $U_c$ is a constant,
437: the random variable $\xi(t)$ represents
438: white noise processes with
439: zero correlation time ($\tau_0=0$), and the $\{g_{\ell}(t/\tau_{\ell})\}$ 
440: are stochastic processes 
441: with correlation times $\tau_{\ell}>0$. 
442: $\xi(t)$ and $g_{\ell}(t/\tau_{\ell})$ 
443: have 
444: units of the square root of time, and $\Gamma_0$ and $\Gamma_{\ell}$ are constants
445: with units of $U^2/t$. If $U$ is chosen to be the the $x$-coordinate
446: of a particle, $\Gamma_0$ and $\Gamma_{\ell}$ have units of a
447: diffusion constant.  Consequently we refer to these constants as
448: generalized diffusion coefficients.  The white noise process has the following 
449: properties \cite{gardiner}
450: 
451: \begin{eqnarray}
452: \langle \xi(t) \rangle &=& 0 \label{noise1}\\
453: \langle \xi(t)\, \xi(t') \rangle &=& \delta(t-t')\label{noise2}\,\,,
454: \end{eqnarray}
455: 
456: \noindent and the remaining colored noise processes are assumed to satisfy
457: 
458: \begin{eqnarray}
459: \langle g_{\ell}(t/\tau_{\ell}) \rangle &=& 0 \label{g1}\\
460: \langle g_{\ell}(t/\tau_{\ell})\, g_{\ell}(t'/\tau_{\ell}) 
461: \rangle &=&  \frac1{\tau_{\ell}}\label{g2}\,f_{\ell}
462: \left(\frac{|t-t'|}{\tau_{\ell}}\right)
463: \label{ggg}
464: \end{eqnarray}
465: 
466: \noindent so that they represent processes with a memory $f_{\ell}$. 
467: Even though correlations between processes with different 
468: correlation times may be non-zero, 
469: we assume the processes to be independent, i.e.
470: 
471: \begin{eqnarray}
472: \langle g_{\ell}(t/\tau_{\ell})\,g_{\ell'}(t'/\tau_{\ell'})\rangle
473: &=&\langle g_{\ell}(t/\tau_{\ell})\rangle\langle g_{\ell'}(t'/\tau_{\ell'})
474: \rangle\nonumber\\
475: &=&0\,\,\,\,\forall\,\ell\neq\ell'
476: \label{a1}\\
477: \langle \xi(t)\, g_{\ell}(t'/\tau_{\ell})\rangle&=&\langle \xi(t)
478: \rangle\langle  g_{\ell}(t'/\tau_{\ell})\rangle\nonumber\\
479: &=&0\,\,\,\,\forall\,\,t\,{\rm and}\,t'
480: \label{a2}\,\,.
481: \end{eqnarray}
482: 
483: \noindent The memory function is assumed to be a continuous function 
484: that depends only on the distance between $t$ and $t'$ disregarding 
485: the time origin (stationary condition). The memory function represents 
486: the correlation between two times of the process $g_{\ell}$. In our model 
487: we impose the condition
488: 
489: \begin{equation}
490: \frac{t}{\tau_{\ell}}f_{\ell}\left( \frac{t}{\tau_{\ell}}\right)
491: <\int_0^t\!\!dt'\,\,\frac1{\tau_{\ell}}f_{\ell}\left( \frac{t'}
492: {\tau_{\ell}}\right)<\infty\label{F}\,\,.
493: \end{equation}
494: 
495: \noindent The scope and implications
496: of the leftmost inequality are explored in Appendix A. In 
497: Appendix A we also examine the conditions $f_{\ell}$ must satisfy in order to 
498: yield contributions to autocorrelation function that decay more weakly
499: than $1/t$.
500: We now assume that this inequality can be taken as a bound to possible 
501: maxima of $f_{\ell}$ appearing at $t>0$. The rightmost inequality enables 
502: us to assume $f_{\ell}$ is normalized
503: 
504: \begin{equation}
505: \label{norm}
506: \int_{-\infty}^{\infty}\!\!dt\,\,\frac1{\tau_{\ell}}f_{\ell}
507: \left( \frac{|t|}{\tau_{\ell}}\right)=1\,\,.
508: \end{equation}
509: 
510: \noindent We have identified here the time scale $\tau_{\ell}$ with the 
511: correlation time of the stochastic process $g_{\ell}$. This identification is 
512: valid if
513: 
514: \begin{equation}
515: \int_{-\infty}^{\infty}\!\!dt\,\,\frac{|t|}{\tau_{\ell}}f_{\ell}
516: \left( \frac{|t|}{\tau_{\ell}}\right)= \tau_{\ell},
517: \end{equation}
518: 
519: \noindent which implies that the behavior of $f_{\ell}$ at large $t$ must be
520: ${\cal O}\left(t^{-(2+\epsilon)}\right)$, or smaller.
521: 
522: In addition, by the properties of the ensemble mean value, 
523: we have that for all real $\lambda$
524: 
525: \begin{eqnarray}
526: 0&\leq& \left\langle \left[ g_{\ell}(t/\tau_{\ell})\,+\,\lambda\, g_{\ell}(t'/\tau_{\ell})\right]^2\right\rangle \nonumber\\
527: &\leq& 
528: \left\langle g_{\ell}(t/\tau_{\ell})^2\right\rangle + 2\lambda\,\left\langle g_{\ell}(t/\tau_{\ell})\, g_{\ell}(t'/\tau_{\ell}) \right\rangle
529: + \lambda^2\left\langle g_{\ell}(t'/\tau_{\ell})^2\right\rangle
530: \nonumber\\
531: &\leq& \frac1{\tau_{\ell}}\left\{ f_{\ell}(0) + 2\lambda\, f_{\ell}\left(\frac{|t-t'|}{\tau_{\ell}}\right) + \lambda^2\, f_{\ell}(0) \right\}
532: \label{a3}\,\,.
533: \end{eqnarray}
534: 
535: \noindent Equation (\ref{a3}) must be true for all $\lambda$. 
536: Therefore, the discriminant of the polynomial in $\lambda$ must be 
537: non-positive
538: 
539: \begin{equation}
540: \label{schwartz}
541: 4\,\left[f_{\ell}\left(\frac{|t-t'|}{\tau_{\ell}}\right) - 
542: \, f_{\ell}(0)\right]
543: \left[f_{\ell}\left(\frac{|t-t'|}{\tau_{\ell}}\right) + 
544: \, f_{\ell}(0)\right]\leq0\,\,.
545: \end{equation}
546: 
547: \noindent Consequently, $f_{\ell}(0)=\max\{f_{\ell}(x)\,\forall\,x\geq0\}$. 
548: Other properties of $f_{\ell}$ are studied in Appendix A.
549: 
550: The ensemble mean value $\langle U\rangle$ is time independent. The ensemble
551: mean value of
552: the noise processes is zero. Therefore, $U_c$ must be equal
553: to $\langle U\rangle$.
554: Processes defined by Eq. (\ref{eq:2}) have 
555: two different components, uncorrelated white noise and correlated processes 
556: with correlation time $\tau_{\ell}$.
557: Because the goal of the simulation is the calculation of the ensemble 
558: mean $\langle U \rangle$ by the analysis of the time series, we study the 
559: behavior of the temporal mean $\overline{U}(t)$
560: 
561: \begin{eqnarray}
562: \overline{U}(t) &=& \frac1t \int_0^t\!\!dt'\,\,U(t')\nonumber\\
563: &=& \langle U\rangle + \frac1t \sqrt{\Gamma_0} \,W(t) +
564: \frac1t \sum_{\ell=1}^{\Lambda} 
565: \sqrt{\Gamma_{\ell}}\,G_{\ell}(t/\tau_{\ell})\,\,,
566: \label{eq:3}
567: \end{eqnarray}
568: 
569: \noindent where $W(t)$ is a Wiener process, \cite{gardiner}
570: 
571: \begin{eqnarray}
572: W(t)&=&\int_0^t\!\!dt'\,\,\xi(t')\\
573: \langle W(t)\rangle &=& 0 \label{wiener1}\\
574: \langle W(t)\,W(t')\rangle &=& t_<\,\,,\label{wiener2}
575: \end{eqnarray}
576: 
577: \noindent with $t_<=\min(t,t')$, and 
578: $G_{\ell}(t/\tau_{\ell})=\int_0^t\!dt'\,\, g_{\ell}(t'/\tau_{\ell})$. 
579: 
580: Equation (\ref{eq:3}) implies that the evolution of the temporal 
581: mean $\overline{U}(t)$ has the same structure as $U$, with an uncorrelated 
582: term and terms with tailed correlation functions. 
583: 
584: The autocorrelation function of the process $\overline{U}$ at 
585: times $t$ and $t'$ is defined by
586: 
587: \end{multicols}
588: 
589: \begin{eqnarray}
590: \kappa(t,t')&=&\left\langle\left(\overline{U}(t)-\langle U\rangle\right)
591: \left(\overline{U}(t')-\langle U\rangle\right)
592: \right\rangle \nonumber\\
593: &=& \frac{\Gamma_0}{tt'}\,\left\langle W(t)\,W(t')\right\rangle +
594: \frac1{tt'}\sum_{\ell=1}^{\Lambda}\Gamma_{\ell}\,\left\langle 
595: G_{\ell}(t/\tau_{\ell})\,G_{\ell}(t'/\tau_{\ell})\right\rangle\,\,,
596: \label{c1}
597: \end{eqnarray}
598: 
599: \noindent where we have used Eqs. (\ref{a1}) and (\ref{a2}) to 
600: neglect terms involving processes with different correlation times.
601: 
602: Because we have assumed $f_{\ell}$ is a continuous function, 
603: $f_{\ell}$ reaches its maximum and minimum value within any closed 
604: interval considered. The $\ell$th non-diffusive contribution 
605: to $\kappa(t,t')$
606:  
607: \begin{equation}
608: \frac1{tt'}\,\langle G_{\ell}(t/\tau_{\ell})\,G_{\ell}
609: (t'/\tau_{\ell})\rangle =\frac1{tt'}\int_0^t\!\! dt_1
610: \int_0^{t'}\!\! dt_2 \,\,\frac1{\tau_{\ell}}f_{\ell}
611: \left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)\,\,,
612: \end{equation}
613: 
614: \noindent is bounded
615: 
616: \begin{eqnarray}
617: \frac1{t_<t_>}\int_0^{t_<}\!\!dt_1\,\,\int_0^{t_>}\!\!dt_2\,
618: \,\frac1{\tau_{\ell}}f_{\ell}\left( \frac{t_{min}}{\tau_{\ell}}\right)
619: \leq \frac1{tt'}&\langle G_{\ell}(t/\tau_{\ell})\,G_{\ell}(t'/\tau_{\ell})
620: \rangle\;&\leq 
621: \frac1{t_<t_>}\int_0^{t_<}\!\!dt_1\,\,\int_0^{t_>}\!\!dt_2\,\,
622: \frac1{\tau_{\ell}}f_{\ell}(0)
623: \nonumber\\
624: \frac1{\tau_{\ell}}f_{\ell}\left(\frac{t_{min}}{\tau_{\ell}}\right)
625: \leq
626: \frac1{tt'}&\langle G_{\ell}(t/\tau_{\ell})\,G_{\ell}(t'/\tau_{\ell})
627: \rangle\;&\leq 
628: \frac1{\tau_{\ell}}f_{\ell}(0)\label{bb}\,\,,
629: \end{eqnarray}
630: 
631: \begin{multicols}{2}
632: 
633: \noindent where $t_>=\max(t,t')$, and $t_{min}$ is the time at which 
634: $f_{\ell}$ reaches its minimum value in the closed interval $[0,t_>]$. 
635: There exists a  $t^*_{\ell}(t_>)\in[0,t_{min}]$ \cite{spivak} such that,
636: 
637: \begin{equation}
638: \frac1{tt'}\,\langle G_{\ell}(t/\tau_{\ell})\,G_{\ell}(t'/\tau_{\ell})\rangle 
639: = \frac1{\tau_{\ell}}f_{\ell}\left(\frac{t_{\ell}^*(t_>)}{\tau_{\ell}}
640: \right)\,\,.
641: \label{tl}
642: \end{equation}
643: 
644: \noindent
645: Using Eqs. (\ref{wiener2}) and (\ref{tl}) in (\ref{c1}), we find that
646: 
647: \begin{equation}
648: \label{correl}
649: \kappa(t,t')= \frac{\Gamma_0}{t_>} + \sum_{\ell=1}^{\Lambda}
650: \frac{\Gamma_{\ell}}{\tau_{\ell}} \, 
651: f_{\ell}\left(\frac{t_{\ell}^*(t_>)}{\tau_{\ell}}\right)\,\,.
652: \end{equation}
653: 
654: \noindent For all times shorter than $\tau_1$ the autocorrelation function 
655: is the sum of {\em diffusive contributions} (proportional to $1/t$) plus 
656: {\em non-diffusive contributions}. These contributions implicitly depend on 
657: $t_>$ through $t^*_{\ell}(t_>)$. We assume that $f_{\ell}$ satisfies the 
658: conditions stated in Appendix A, so that the
659: dependence of $f_{\ell}$ on $t$ is {\em weaker than} $1/t$ 
660: (for total time scales shorter than $\tau_{\ell}$;  see Appendix A).
661: 
662: We next consider the behavior of Eq. (\ref{correl}) for time scales
663: greater than $\tau_1$.
664: Under the scale change $t \to bt$ such that $\tau_1\ll bt_{>} \ll \tau_2$, 
665: the contributions to the correlation function from the process with 
666: correlation time $\tau_1$ can be considered diffusive [in other words, 
667: by virtue of Eqs. (\ref{noise2}) and (\ref{ggg}), $f_1/\tau_1$ has become a 
668: delta function]. With $bt_{>} \ll \tau_2$, the other processes preserve 
669: their old properties. Then, the autocorrelation function can be expressed
670: 
671: \begin{equation}
672: \label{urenor}
673: \kappa(bt,bt') =
674: \frac{\Gamma_0+\Gamma_1}{bt_>}+\sum_{\ell=2}^{\Lambda}
675: \frac{\Gamma_{\ell}}{\tau_{\ell}} \, f_{\ell}
676: \left(\frac{t_{b\ell}^*(t_>)}{\tau_{b\ell}}\right)\,\,.
677: \end{equation}
678: 
679: \noindent The complete derivation of Eq. (\ref{urenor}) can be found in 
680: Appendix B.
681: For a times larger than the correlation time $\tau_{\Lambda}$, all contributions
682: to the autocorrelation function are diffusive,
683: the simulation can be considered ergodic, the sampling complete, 
684: and the temporal mean is equal to the ensemble mean within ${\cal O}(1/t)$ 
685: mean square fluctuations.
686: 
687: \section{Discrete time sequences and the MC metric}
688: 
689: Monte Carlo simulations generate discrete sequences $U_k$ of values of the 
690: quantity under study. Additionally,
691: in actual calculations the ensemble of sequences is represented by a
692: finite rather than an infinite set.  In this section, the model developed in 
693: the previous section is extended to finite sets of discrete sequences. 
694: We express the
695: $M$ sequences $\,\left\{U_k^{(m)}\right\}_{k=1}^K\,$, where the label $(m)$ 
696: ranges
697: from 1 to $M$. The exact ensemble mean value $\langle U\rangle$ can 
698: be obtained
699: in the limit that $M$ becomes infinite.  In analogy with the model developed 
700: in Section III, each output is assumed to have the form
701: 
702: \begin{equation}
703: \label{seq}
704: U_k^{(m)} = \langle U\rangle + \sqrt{\Gamma_0}\,\xi_k^{(m)} + \sum_{\ell=1}^
705: {\Lambda}\sqrt{\Gamma_{\ell}}\,g_{\ell;\,k/\tau_{\ell}}^{(m)}\,\,,
706: \end{equation}
707: 
708: \noindent where
709: 
710: \begin{eqnarray}
711: \langle \xi_k^{(m)}\rangle&=&0\\
712: \langle \xi_k^{(m)}\,\,\xi_{k'}^{(n)}\rangle&=&\delta_{m,n}\,\delta_{k,k'}\\
713: \langle g_{\ell;\,k/\tau_{\ell}}^{(m)}\rangle&=&0\\
714: \langle g_{\ell;\,k/\tau_{\ell}}^{(m)}\,\,
715: g_{\ell';\,k'/\tau_{\ell'}}^{(n)}\rangle&=&
716: \delta_{m,n}\,\delta_{\ell,\ell'}\,f_{\ell}\left(\frac{|k-k'|}{\tau_{\ell}}
717: \right)\,\,.
718: \end{eqnarray}
719: 
720: \noindent The true ensemble
721: average $\langle U\rangle$
722: does not depend on the index $m$. 
723: 
724: In the discrete case we define a metric
725: 
726: \begin{equation}
727: \label{omega}
728: d_k = \frac2{M(M-1)} \sum_{i=2}^M \sum_{j=1}^{i-1} 
729: \left[\,\overline{U}^{(i)}_k - \overline{U}^{(j)}_k\right]^2\,\,,
730: \end{equation}
731: 
732: \noindent where the bars represent the temporal mean value
733: 
734: \begin{eqnarray}
735: \overline{U}^{(m)}_k &=& \frac1k \sum_{k'=1}^k U^{(m)}_{k'}\nonumber\\
736: &=&\langle U\rangle + \frac{\sqrt{\Gamma_0}}k\,W_k^{(m)}+ \sum_{\ell=1}^{\Lambda}\frac{\sqrt{\Gamma_{\ell}}}k\,G_{\ell;\,k/\tau_{\ell}}
737: ^{(m)} \,\,,
738: \end{eqnarray}
739: 
740: \noindent with
741: 
742: \begin{eqnarray}
743: W_k^{(m)} &=& \sum_{k'=1}^k \xi_{k'}^{(m)}\\
744: G_{\ell;\,k/\tau_{\ell}}^{(m)} &=& 
745: \sum_{k'=1}^k g_{\ell;\,k'/\tau_{\ell}}^{(m)}\,\,.
746: \end{eqnarray}
747: 
748: \noindent Observe that in the present case, our finite sample 
749: of the infinite ensemble is the set of outcomes from $M$ independent 
750: numerical experiments. The metric we have defined in Eq. (\ref{omega})
751: can be contrasted with alternative metrics 
752: \cite{thiru1,thiru2,thiru3}
753: previously defined for molecular dynamics simulations.
754: These alternative metrics examine
755: the fluctuations of two simulations initialized from different components
756: of configuration space averaged with respect to all the
757: particles in the system.  The metric we use in this work is determined
758: using an
759: average with respect to $M$ independent 
760: simulations that represent a subset of the full
761: ensemble.
762: 
763: Using the model presented in Eq. (\ref{seq}), we now develop 
764: a way
765: to predict the behavior of the MC simulation in the 
766: non-ergodic and the ergodic regimes.
767: We first consider the case that the total simulation time 
768: $K$ is larger than the first 
769: correlation time $\tau_1$ but shorter than $\tau_2$, i.e. 
770: $\tau_1\ll K\ll\tau_2$. The expression for 
771: $d_k$ is given by
772: 
773: \end{multicols}
774: 
775: \begin{eqnarray}
776: d_k &=& \frac2{M(M-1)} \sum_{i=2}^M \sum_{j=1}^{i-1} 
777: \left[ \left(\overline{U}^{(i)}_k-\langle U\rangle\right) - 
778: \left(\overline{U}^{(j)}_k-\langle U\rangle\right)\right]^2\nonumber\\
779: &=& \frac2M \sum_{i=1}^M\left(\overline{U}^{(i)}_k-\langle U\rangle\right)^2 - 
780: \frac4{M(M-1)} \sum_{i=2}^M \sum_{j=1}^{i-1} 
781: \left(\overline{U}^{(i)}_k-\langle U\rangle\right) \, 
782: \left(\overline{U}^{(j)}_k-\langle U\rangle\right)\nonumber\\
783: &=& 2\,\Gamma_0 \,\frac1M \sum_{i=1}^M
784: \left(\frac{W_k^{(i)}}k\right)^2 + 
785: 2\,\Gamma_1 \,\frac1M \sum_{i=1}^M
786: \left(\frac{G_{1;\,k/\tau_1}^{(i)}}k\right)^2 +
787: 2\,\sum_{\ell=2}^{\Lambda}\Gamma_{\ell} \,\frac1M \sum_{i=1}^M
788: \left(\frac{G_{\ell;\,k/\tau_{\ell}}^{(i)}}k\right)^2 \nonumber\\
789: &&+ \,4\,\sum_{\ell=1}^{\Lambda}\sqrt{\Gamma_0 
790: \Gamma_{\ell}} \,\,\frac1M \sum_{i=1}^M\frac{W_k^{(i)}
791: \,G_{\ell;\,k/\tau_{\ell}} ^{(i)}}{k^2}
792:  + \,4\,\sum_{\ell=2}^{\Lambda}\sum_{\ell'=1}^{\ell-1}\sqrt{\Gamma_{\ell} 
793: \Gamma_{\ell'}} \,\,\frac1M \sum_{i=1}^M\frac{G_{\ell;\,k/\tau_{\ell}}^{(i)}
794: \,G_{\ell';\,k/\tau_{\ell'}} ^{(i)}}{k^2} \nonumber\\
795: &&- \frac4{M(M-1)} \sum_{i=2}^M 
796: \sum_{j=1}^{i-1} \left(\overline{U}^{(i)}_k-\langle U\rangle\right) \, 
797: \left(\overline{U}^{(j)}_k-\langle U\rangle\right)\,\,.  
798: \end{eqnarray}
799: 
800: \noindent If the number of experiments $M$ is sufficiently large, 
801: we can neglect terms involving processes with different correlation times, 
802: and products of sequences belonging to different experiments. 
803: Under these assumptions we obtain
804: 
805: \begin{equation}
806: \label{om}
807: d_k = 2\,\frac{\Gamma_0}k \,\frac1M \sum_{i=1}^M
808: \frac{{W_k^{(i)}}^2}k +
809: 2\,\frac{\Gamma_1}k \,\frac1M \sum_{i=1}^M
810: \frac{{G_{1;\,k/\tau_1}^{(i)}}^2}k +
811:     2\,\sum_{\ell=2}^{\Lambda}\Gamma_{\ell} \,\frac1M \sum_{i=1}^M
812: \left(\frac{G_{\ell;\,k/\tau_{\ell}}^{(i)}}k\right)^2\,.
813: \end{equation}
814: 
815: \begin{multicols}{2}
816: 
817: \noindent Equation (\ref{om}) preserves the form of Eq. (\ref{correl}). 
818: To make this statement explicit, let us rewrite Eq. (\ref{om}) as
819: 
820: \begin{equation}
821: d_k = 2\,\frac{\Gamma_k}k + 2\,\Upsilon_k\,\,,
822: \end{equation}
823: 
824: \noindent where
825: 
826: \begin{eqnarray}
827: \Gamma_k   &=& \Gamma_0\,\frac1M \sum_{i=1}^M
828: \frac{{W_k^{(i)}}^2}k +
829: \Gamma_1\,\frac1M \sum_{i=1}^M
830: \frac{{G_{1;\,k/\tau_1}^{(i)}}^2}k\label{g} \label{eq:gmmg} \\
831: \Upsilon_k &=& \sum_{\ell=2}^{\Lambda}\Gamma_{\ell} \,\frac1M \sum_{i=1}^M
832: \left(\frac{G_{\ell;\,k/\tau_{\ell}}^{(i)}}k\right)^2\,\,.
833: \label{ups}
834: \end{eqnarray}
835: 
836: \noindent In Appendix B we present a study of the way non-diffusive contribution become diffusive under time scale 
837: changes. If $M$ is sufficiently large and $\tau_1\ll K \ll \tau_2$, 
838: by virtue of Appendix 
839: B, $\Gamma_k$ must roughly be a constant.
840: By {\em roughly a constant} we mean a 
841: constant $C$ plus some rapidly fluctuating function $\zeta_k$, with the 
842: following properties: a) $\langle\zeta_k\rangle=0$ and b) 
843: $|C|\gg\max_{k=1,2,\dots,K}(|\zeta_k|)$. Then
844: 
845: \begin{equation}
846: \label{gz}
847: \Gamma_k \simeq \Gamma_K + \zeta_k
848: \end{equation}
849:  
850: \noindent If $K$ is enlarged, we expect to 
851: have a larger 
852: value of $\Gamma_K$. 
853: $\Upsilon_k$ is a quantity related to the memory functions 
854: $f_{\ell}$ with 
855: correlation times $\tau_{\ell}\gg K$. In the continuous 
856: time model, the colored noise 
857: processes contribute to the autocorrelation function 
858: with terms proportional to $f_{\ell}(t^*_{\ell}(t_>)/\tau_{\ell})$, which are 
859: weakly dependent on $t$ (see Appendix A). 
860: We can expect $\Upsilon_k$ to be
861: weakly dependent on $k$, and for sequences of length $K$ and for $M$ 
862: sufficiently large, we consider this quantity roughly to be a constant
863: 
864: \begin{equation}
865: \label{uz} 
866: \Upsilon_k\simeq\Upsilon_K+\beta_k\,\,.
867: \end{equation}
868: 
869: \noindent where $\beta_k$ represents additional random noise.
870: Then, for a given length 
871: $k\leq K$, the MC metric $d_k$ can be approximated by 
872: 
873: \begin{equation}
874: \label{d}
875: d_k = 2\,\frac{\Gamma_K}k + 2\,\Upsilon_K+\gamma_k\,\,,
876: \end{equation}
877: 
878: \noindent where $\gamma_k=2(\zeta_k/k+\beta_k)$ represents remaining stochastic noise 
879: from
880: both contributions.  In  
881: this approximation, $\Gamma_K$ and $\Upsilon_K$ are the quantities 
882: that 
883: carry the long time dependence. Short time features appear in the $1/k$ 
884: dependence and in the remaining noise $\gamma_k$. If the sequences considered 
885: are increased in size by a factor of $b$, such that 
886: $\tau_{\lambda-1}\ll K\ll\tau_{\lambda}\ll bK$ for a given 
887: $1\leq\lambda\leq\Lambda$, $\Gamma_K$ ($\Upsilon_K$) is 
888: increased (decreased) (see Appendix B). Then,
889: 
890: \begin{equation}
891: \label{di}
892: d_{bk} = 2\,\frac{\Gamma_{bK}}{bk} + 2\,\Upsilon_{bK}+\gamma_{bk}\,\,,
893: \end{equation}
894: 
895: \noindent where $\Upsilon_{bK}$ must go to zero and 
896: $\Gamma_{bK}$ must approach a constant when $b$ is increased. 
897: By virtue of the expected behavior of the non-diffusive contributions (see 
898: Appendix A), we propose the following expression for $\Upsilon_{bK}$
899: 
900: \begin{equation}
901: \label{upa}
902: \Upsilon_{bK}=\Upsilon_K\,\phi(b)\,\,,
903: \end{equation}
904: 
905: \noindent where $\phi(b)$ is a decreasing function of $b$. Moreover, 
906: $\Upsilon_{bK}$ is a sum of non-diffusive contributions. As presented in 
907: Appendix A, each non-diffusive 
908: contribution to the autocorrelation function has a relative variation smaller 
909: than the relative variation of the diffusive contribution, namely $1-1/b$. If 
910: this inequality
911: is applicable to the sum of non-diffusive contributions, we have that
912: 
913: \begin{eqnarray}
914: 1-\frac1b &>& 1-\frac{\Upsilon_{bK}}{\Upsilon_K}\\
915: 1-\frac1b &>& 1-\phi(b)\\
916: 1&<&b\,\phi(b)\label{bphi}\,\,,
917: \end{eqnarray}
918: 
919: \noindent for all $b>1$. Then, $\phi$ must be either
920: 
921: \begin{equation}
922: \phi(b)=b^{-\upsilon}\label{upa1}
923: \end{equation}
924: 
925: \noindent or
926: 
927: \begin{equation}
928: \phi(b)=\frac1{\eta\,\ln(b)+1}\label{upa2}\,\,,
929: \end{equation}
930: 
931: \noindent with $0<\upsilon<1$ and $0<\eta \le 1$. Equation (\ref{upa2}) can be 
932: thought as the limit
933: of Eq. (\ref{upa1}) when the exponent goes to zero. We know of no 
934: {\em a priori}
935: argument to 
936: justify Eq. (\ref{upa}). However, as is discussed in Section V, 
937: our numerical experience has shown Eq. (\ref{upa}) to be obeyed in all
938: cases we have examined.
939: 
940: Our goal is to develop a criterion to decide when the 
941: simulation can be 
942: considered ergodic. From the previous considerations it is clear that the 
943: ergodic limit is reached when $\Upsilon_K$ is indistinguishable from zero. 
944: The output from a MC simulation is usually noisy. Therefore, $\gamma_k$ 
945: can not 
946:  be neglected. A useful way to separate diffusive 
947: and non-diffusive 
948: contributions and to eliminate the stochastic noise from Eq. (\ref{d}), is to 
949: perform a Fourier analysis of the function $kd_k$. Let us define the 
950: frequencies $\omega_n=(2\pi/K)\,n$, with $n=0,1,\dots,K-1$. The discrete 
951: Fourier transform of the function $kd_k$ is the signal $Y_K(\omega_n)$
952: 
953: \end{multicols}
954: 
955: \begin{eqnarray}
956: Y_K(\omega_n)&=&\widehat{kd_k}(\omega_n)=\frac1K\sum_{k=1}^K 
957: \exp(-i\omega_nk)\,kd_k \label{eq:ft}\\
958: &=& \frac2K\sum_{k=1}^K \exp(-i\omega_nk) \,\,\Gamma_K\,+\,
959: \frac2K\sum_{k=1}^K k\,\exp(-i\omega_nk) \,\,\Upsilon_K+
960: \widehat{k\gamma_k}(\omega_n)\nonumber\\
961: &=& 2\,\delta_{n,0}\,\Gamma_K + \left\{\delta_{n,0} (K+1) + 
962: \left(1-\delta_{n,0}\right)\left(1+i\,\cot(\omega_n/2)\right)\right\}
963: \Upsilon_K
964: +\widehat{k\gamma_k}(\omega_n)\nonumber\\
965: &=& 2\,\delta_{n,0}\,\Gamma_K + \left(K\,\delta_{n,0}+1\right)\,
966: \Upsilon_K
967: + i\,\left(1-\delta_{n,0}\right)\,\cot(\omega_n/2)\,\Upsilon_K+
968: \widehat{k\gamma_k}(\omega_n)
969: \label{fs}
970: \end{eqnarray}
971: 
972: \begin{multicols}{2}
973: 
974: \noindent In general, $\widehat{k\gamma_k}(\omega_n)$ 
975: is negligible except at high 
976: frequencies. For small positive values of the frequency we can make 
977: the approximation $\cot(\omega_n/2)\simeq 2/\omega_n$.  From this
978: approximation we have
979: 
980: \begin{equation}
981: \label{eq:10}
982: {\rm Im}\left(Y_K(\omega_n)\right)\simeq \frac2{\omega_n}\, 
983: \Upsilon_K\,\,.
984: \end{equation}
985: 
986: \noindent The real part of Eq. (\ref{fs}) for positive frequencies is
987: 
988: \begin{equation}
989: \label{ns}
990: {\rm Re}\left(Y_K(\omega_n)\right)=\Upsilon_K\,\,.
991: \end{equation}
992: 
993: \noindent Even though simpler than Eq. (\ref{eq:10}), we have
994: found Eq. (\ref{ns}) is more
995: sensitive to the deviations of $d_k$ from the approximation Eq. (\ref{d}).
996: Therefore, the data obtained from the real part is of poorer quality than
997: the data obtained from the imaginary part.
998: 
999: Equation (\ref{eq:10}) implies that for a given simulation 
1000: length $K$, the contributions to the MC metric from the non-diffusive 
1001: process can be determined from a simple relationship involving the 
1002: Fourier transform of the function $kd_k$ at low frequencies. By increasing 
1003: the length of the run $K$ by a factor of $b$, it is possible to observe the 
1004: dependence of $\Upsilon_{bK}$ on $bK$. 
1005: 
1006: \section{Applications}
1007: 
1008: The concepts developed in the previous sections are sufficiently 
1009: general to be applied to any kind of MC simulation.
1010: We devote the present section to the application of the developments of
1011: this paper
1012: to the study of the Lennard-Jones 
1013: 13-particle cluster in the canonical
1014: ensemble. This system has been introduced previously in Sec. II.
1015: 
1016: Some thermodynamic properties of clusters as a function of temperature
1017: exhibit rapid changes that are reminiscent of similar changes that occur
1018: for the same properties in bulk systems at phase transitions.  In a bulk
1019: system a phase transition occurs at a single temperature.  For clusters
1020: the rapid changes in thermodynamic properties occur over a finite
1021: temperature interval.  To distinguish the temperature range where
1022: thermodynamic properties change rapidly in clusters from a true phase
1023: transition,  we follow Berry {\em et al.} 
1024: \cite{berry}
1025: and refer to such changes in
1026: physical properties as associated with a phase change.  A common
1027: property that has been found to be useful in monitoring these phase
1028: change intervals of temperature is the heat capacity at constant 
1029: volume \cite{clusters}
1030: 
1031: \begin{equation}
1032: \label{cv}
1033: C_V(T) = \frac1{k_BT^2}\,\left\langle(V-\langle V\rangle_T)^2
1034: \right\rangle_T + \frac32Nk_B\,\,,
1035: \end{equation}
1036: 
1037: \noindent where
1038: $\langle\cdot\rangle_T$ represents the classical canonical mean value.
1039: 
1040: In this work we consider the bare Metropolis (Met), 
1041: \cite{metropolis}
1042: J-walking (Jw), \cite{frantz} and
1043: parallel tempering (PT) \cite{p1,p2,p3} approaches to Monte Carlo simulations.
1044: The free variable of all these methods is the reduced temperature 
1045: $k_BT/\varepsilon$. In PT and Jw simulations, the highest temperature 
1046: used ($T_h$) must be sufficiently large to ensure that Met is 
1047: ergodic.\cite{frantz} 
1048: From experience simulating a variety of systems, we have found
1049: that $T_h$ must also be lower than a temperature $T_b$ where cluster
1050: evaporation events become frequent.  It is useful to think of $T_b$ as
1051: the cluster analogue of a boiling temperature.  We have found that Met
1052: is unable to sample the boiling phase change region for clusters
1053: ergodically, using total time scales accessible to current simulations.
1054: 
1055: For the results that follow, $U_k^{(m)}$ is chosen to be represented by
1056: the potential energy of the system.  
1057: In general $U_k^{(m)}$ can be any scalar property of the system.
1058: We define a pass to
1059: represent a set of single particle MC moves taken sequentially over the 13
1060: particles in the cluster. We take $U_k^{(m)}$ to be the
1061: potential energy at the $k'$th pass, in the $m'$th experiment. Using 
1062: Eq. (\ref{fs}) we can write
1063: 
1064: \begin{equation}
1065: \label{o1}
1066: Y_K(0) = 2\,\Gamma_K+(K+1)\,\Upsilon_K\,\,.
1067: \end{equation}
1068: 
1069: \noindent In the non-ergodic regime, $Y_K(0)$ grows 
1070: with $K$, while in the ergodic regime, the signal $Y_K(0)$ 
1071: approaches a constant. 
1072: 
1073: We begin by displaying results obtained for a calculation that has not
1074: attained ergodicity over the time scale of the simulation.  We examine
1075: the 13-particle Lennard-Jones cluster with the Met algorithm setting
1076: $R_c=4\sigma$ at a temperature of $k_BT_h/\varepsilon=0.393$.  The
1077: temperature is chosen to be that typically used as the initial high
1078: temperature in Jw and PT studies of LJ$_{13}$.  By choosing a large
1079: constraining radius, the evaporation events are so frequent at the
1080: chosen temperature that attaining ergodicity proves to be quite
1081: difficult.  We demonstrate the effect of reducing the constraining
1082: radius shortly.  
1083: 
1084: \begin{figure}
1085: \epsfxsize=.48\textwidth
1086: \epsfbox{f2.ps}
1087: \end{figure}
1088: 
1089: {\footnotesize
1090: {\bf Fig. 2:} The upper panel is the signal $Y_K(0)$ (in units
1091: of $\varepsilon^2$) vs. $K$ for 
1092: $R_c=4\sigma$ at $k_BT_h/\varepsilon=0.393$. from $M=40$ independent 
1093: experiments, of LJ$_{13}$. The length of the simulation is $10^4$ 
1094: MC passes. The lower panel shows the ``time evolution" of $\overline{U}_K$ 
1095: (in units of $\varepsilon$) for 15 independent experiments. At 
1096: least three basins with different energies are present. Clearly, the 
1097: simulation at this scale of time, is not ergodic.}
1098: \vspace{.5cm}
1099: 
1100: \noindent The 
1101: number of replicas used in the calculation is $M=40$, and $K=10^4$. 
1102: The upper panel of Fig. 2
1103: shows the signal $Y_K(0)$ [evaluated using Eq. (\ref{eq:ft})], 
1104: which grows along the entire
1105: simulation. This is the behavior expected in the non-ergodic regime. In the 
1106: lower panel we can see 
1107: the ``time evolution" of the temporal mean values of 15 experiments. 
1108: 
1109: \begin{figure}
1110: \epsfxsize=.48\textwidth
1111: \epsfbox{f3.ps}
1112: \end{figure}
1113: 
1114: {\footnotesize
1115: {\bf Fig. 3:} Upper panel: $\Upsilon_{bK}$ (in units
1116: of $\varepsilon^2$) as a 
1117: function of $\log_2(b)$ for $R_c=4\sigma$, $3\sigma$, $2.5\sigma$, and 
1118: $2\sigma$. For the
1119: two larger radii the full line is the best fit to the data points, 
1120: according
1121: to Eq. (\ref{upa}) with $\phi$ defined in Eq. (\ref{upa2}).
1122: The lower panel shows the linear behavior of 
1123: $\Upsilon_K/\Upsilon_{bK}$ vs. $\log_2(b)$, for 
1124: $R_c=4\sigma$ and $3\sigma$. $K$ has been set to $10^4$.}
1125: \vspace{.5cm}
1126: 
1127: \noindent There are three sets of curves, each of which is indicative of 
1128: sampling of at least 
1129: three different energy basins. At low values of $K$ the curves in the
1130: lower panel differ significantly.
1131: At $K\simeq4\,000$ the high energy basin curves begin to decrease
1132: in energy. For a value of $K$ larger than the data displayed in Fig. 2,
1133: the curves can be expected to coalesce with
1134: the low energy basin curves. It is clear that for $K\leq10\,000$, the 
1135: simulation is not ergodic.
1136: 
1137: In PT and Jw studies it is essential that the initial high temperature
1138: walk be ergodic.  Ergodicity can be attained for LJ$_{13}$ by reducing
1139: the radius of the constraining potential so that evaporation events are
1140: rare.  We now present a study of $\Upsilon_K$ as a function of $K$ for
1141: several values of $R_c$.
1142: To determine $\Upsilon_K$, we have calculated the
1143: Fourier transform function $Y_K(\omega_n)$ using Eq. (\ref{eq:ft}) 
1144: at a
1145: series of frequencies $\omega_n=2\pi n/K$ where $n$ has ranged from 1 to
1146: $\min(\sqrt{12}bK/20\pi,100)$.  This range of frequencies ensures the linear
1147: approximation used in Eq. (\ref{fs}) is valid while including sufficient
1148: numbers of points for accuracy.\cite{citfreq}
1149: Using Eq. (\ref{eq:10}), we have 
1150: calculated the slope of the imaginary part of 
1151: $1/Y_K(\omega_n)$ as a function of $\omega_n$, for 
1152: these frequencies. The data points appearing in Fig. 3 are 
1153: the mean value over twenty independent calculations of the slope of 
1154: $1/Y_K(\omega_n)$.
1155: 
1156: \begin{figure}
1157: \epsfxsize=.48\textwidth
1158: \epsfbox{f4.ps}
1159: \end{figure}
1160: 
1161: {\footnotesize
1162: {\bf Fig. 4:} $\Upsilon_{bK}$ (in units
1163: of $\varepsilon^2$) and its error vs. $\log_2(b)$ for 
1164: $R_c=2.5\sigma$ and $2\sigma$, with $K=10^4$. When $\Upsilon_{bK}$ is on 
1165: the order of its 
1166: own error, the simulation can be 
1167: considered ergodic. For $R_c = 2\sigma$ the simulation becomes ergodic at 
1168: $\log_2(b)\simeq4$ ($bK\simeq16\times 10^4$). For $R_c=2.5\sigma$ a longer 
1169: simulation is needed to reach ergodicity.}
1170: \vspace{.5cm}
1171: 
1172: Starting from random configurations, we have performed $5\times10^4$ 
1173: Met passes at $k_BT_h/\varepsilon=0.393$. After this warmup process, we 
1174: have created sequences of sizes $bK=10^4$, $2\times10^4$, $4\times10^4$,
1175: $\dots$, $64\times10^4$. 
1176: The results are presented in Fig. 3 for $R_c=4\sigma$, $3\sigma$, 
1177: $2.5\sigma$, and $2\sigma$.
1178: The upper panel shows $\Upsilon_{bK}$ as a function 
1179: of $\log_2(b)$, for fixed $K=10^4$. We have chosen to present the data
1180: using base 2 logarithms for clarity (each increase by 1 unit of $\log_2(b)$
1181: represents a factor of 2 scale increase).
1182: All the data decrease with increasing
1183: $b$, but only $R_c=2 \sigma$ and $R_c=2.5 \sigma$ appear to vanish
1184: to within the error bars over the time scale of the current simulation.
1185: In the lower panel we present
1186: $\Upsilon_K/\Upsilon_{bK}$ as a function of
1187: $\log_2(b)$ for $R_c=4$ and $3\sigma$. 
1188: The decay law suggested in Eq. (\ref{upa}) with $\phi$ 
1189: given by Eq. (\ref{upa2}) is satisfied for both radii.
1190: 
1191: \begin{figure}
1192: \epsfxsize=.48\textwidth
1193: \epsfbox{f5.ps}
1194: \end{figure}
1195: 
1196: {\footnotesize
1197: {\bf Fig. 5:} The upper panel shows the decay behavior of $\Upsilon_{bK}$
1198: (in units of $\varepsilon^2$)
1199: as a function of $\log_2(b)$ for PT and Met, at the temperature of the 
1200: melting peak
1201: of the heat capacity, $k_BT_m/\varepsilon=0.282$. From Eq. (\ref{upa1}), we 
1202: plot $\log_2{(\Upsilon_{K}/\Upsilon_{bK})}$ vs. $\log_2{(b)}$, to extract the 
1203: value of the exponent $\upsilon$ (the slope of the linear fit). We have 
1204: found $\upsilon=0.93\pm0.03$ for PT, and $\upsilon=0.94\pm0.02$ for Met. 
1205: The straight lines are the best linear fits of the data points.}
1206: \vspace{.5cm}
1207: 
1208: We have stated that the simulation can be considered effectively 
1209: ergodic when $\Upsilon_K$ is indistinguishable from zero. 
1210: In Fig. 4 we have plotted $\Upsilon_{bK}$ and its statistical 
1211: error as a function
1212: of
1213: $\log_2(b)$ for $R_c=2.5$ and $2\sigma$. For $R_c=2\sigma$ the 
1214: crossing point of $\Upsilon_{bK}$ and its error is at $bK\simeq16\times10^4$. 
1215: For $R_c=2.5\sigma$ the crossing point is at a $bK>64\times10^4$. We can 
1216: conclude that for $k_BT_h/\varepsilon=0.393$ and $R_c=2\sigma$ the 
1217: simulation can be considered effectively ergodic after $16\times10^4$ Met 
1218: passes.
1219: 
1220: Once a constraining radius is chosen, PT and Jw simulations require the
1221: highest temperature $T_h$ be chosen so that Met is ergodic.  For a given
1222: $R_c$, the extent of ergodicity can be tested using the same metric that
1223: has been used for determining the optimum value of $R_c$, but by varying
1224: the temperature.
1225: For the parameters $k_BT_h/\varepsilon=0.393$ and $R_c=2\sigma$ the simulation 
1226: is ergodic even at very short sequence lengths.  We have 
1227: found that for $k_BT_h/\varepsilon<0.393$ the simulations are not ergodic. 
1228: To be sure that the parameters are appropriate, we have performed a short PT 
1229: simulation ($10^4$ passes, ten PT passes consists of nine Met passes
1230: plus an exchange attempt) with 40 equally spaced temperatures in the 
1231: range $k_BT/\varepsilon=$[0.028,0.393] in order to obtain a first estimate of the position of the 
1232: melting and boiling temperature regions. The boiling peak in the specific 
1233: heat appears to be located at a higher temperature than $k_BT/\varepsilon=$
1234: 0.393. Moreover, 
1235: the value of $C_V$ at $k_BT/\varepsilon=0.393$ is about one-half the value of $C_V$ at 
1236: the temperature of the melting peak $k_BT_m/\varepsilon=0.282$. From 
1237: these results we
1238: feel it is safe to choose $R_c=2\sigma$ and $k_BT_h/\varepsilon=0.393$ for
1239: the calculations that follow.
1240: 
1241: We now illustrate the convergence characteristics of $\Upsilon_K$
1242: when we increase the total time scale of the calculation by a
1243: factor $b$.  We illustrate this behavior using a PT simulation of
1244: LJ$_{13}$, and we focus on results at the temperature of the melting
1245: peak in the heat capacity ($k_BT_m/\varepsilon=0.282$).  
1246: We choose this temperature, because from
1247: experience 
1248: \cite{juan1,juan2,juan3}
1249: we know the statistical fluctuations are large at the melting
1250: heat capacity maximum.  The large statistical fluctuations make it
1251: possible to emphasize the behavior of $\Upsilon_K$.
1252: We have run the PT simulation at 40 equally spaced temperatures in the 
1253: range $k_BT/\varepsilon=$ [0.028,0.393]. The initial warmup time has been 
1254: set to $10^4$ 
1255: Met passes, followed by $2\times10^4$ PT passes. Following the warm-up
1256: period, we perform 
1257: simulations of $10^5$, $2\times10^5$, $4\times10^5$, $8\times10^5$, 
1258: $16\times10^5$, and $32\times10^5$ PT passes.  In each case the initial 
1259: configuration has been taken to be the last configuration of the previous run.
1260: The 
1261: output of the simulation are sequences of the potential energy. 
1262: $\Upsilon_K$ has been determined in the same way
1263: as in the calculation of the high temperature parameters (presented in
1264: Fig. 3 and Fig. 4). The data points appearing in the upper panel of Fig. 5 are 
1265: the mean value over twenty independent calculations of the slope of 
1266: $1/Y_K(\omega_n)$. In the lower panel of Fig. 5 we have plotted 
1267: $\log_2{(\Upsilon_{K}/\Upsilon_{bK})}$ as a function of 
1268: $\log_2{(b)}$, where $K=10^4$ 
1269: and $b=1,2,4,\dots,32$. The slope of the linear fit is the exponent 
1270: $\upsilon$, according to Eq. (\ref{upa1}). At the temperature of the 
1271: melting peak, $\upsilon=0.93\pm0.03$. 
1272: 
1273: It is of interest to perform a similar study of the behavior of
1274: $\Upsilon_K$ as a function of the time scaling for an Met calculation. 
1275: We have taken the final configuration of the PT simulation at
1276: $k_BT_m/\varepsilon=0.282$ as an initial configuration, and we have 
1277: performed a simple Met simulation at that melting temperature.  
1278: A graph of $\Upsilon_{bK}$ and $\log_2{(\Upsilon_{K}/\Upsilon_{bK})}$ 
1279: as a function of $\log_2{(b)}$ for Met is also presented in Fig. 5. 
1280: From the upper panel of Fig. 5, it is evident that Met results are not
1281: ergodic within the same scaled time as the PT results.  It is also
1282: evident that the power law exponent for both Met and PT 
1283: are not distinguishable. 
1284: Similar studies of the power law using the Jw method also give the same
1285: exponent.  Neither an increase in the number of temperatures nor
1286: changing the distribution of temperatures in both Jw and PT simulations
1287: has any effect on the calculated exponent.
1288: 
1289: \begin{figure}
1290: \epsfxsize=.48\textwidth
1291: \epsfbox{f6.ps}
1292: \end{figure}
1293: 
1294: {\footnotesize
1295: {\bf Fig. 6:} Comparison of the Met and Jw diffusion coefficients
1296: with the PT diffusion coefficient as a function of the reduced temperature. 
1297: The dashed line represents equivalence between methods.}
1298: \vspace{.5cm}
1299: 
1300: By using the results to compare the relative
1301: efficiencies of Met, Jw and PT simulations for the LJ$_{13}$ system.
1302: We have found that
1303: PT and Jw simulations can be considered ergodic if the run length is on the 
1304: order of $2\times10^5$ passes, while Met simulations that are initialized 
1305: from configurations
1306: generated from an ergodic PT study are ergodic when the total run length
1307: consists of $2\times10^6$ passes or more.
1308: 
1309: In order to compare approaches, 
1310: we have calculated $\Gamma$ as a function of the reduced temperature, 
1311: for the three methods. The comparison of diffusion coefficients from different
1312: algorithms has also been used by Andricioaei and
1313: Straub
1314: \cite{straub}.
1315: The comparison of Jw and Met 
1316: with PT is presented in Fig. 6. The Jw and PT simulations are found
1317: to have comparable efficiencies using $\Gamma$ as a measure for all
1318: calculated temperatures.  At intermediate temperatures, Met is
1319: significantly less efficient.  We have chosen to truncate the Jw study
1320: at $k_BT/\varepsilon=0.12$.  For temperatures below $k_BT/\varepsilon=0.12$,
1321: Jw simulations require significant effort, because a large set of
1322: external distributions must be generated.  Because at temperatures below
1323: $k_BT/\varepsilon=0.12$ LJ$_{13}$ is dominated by structures close to the
1324: lowest energy icosahedral isomer, we expect the Jw and PT methods to
1325: have similar efficiencies (as measured by $\Gamma$) for all
1326: temperatures.
1327: 
1328: \section{Conclusions}
1329: 
1330: In this paper we have presented a study of the approach to
1331: the ergodic limit in MC simulations. In all the cases examined, 
1332: the behavior of the MC metric $d_k$ can be approximated by Eq. (\ref{d}), 
1333: and the behavior of $\Upsilon_{bK}$ satisfies Eq. (\ref{upa}). Because the 
1334: exponent $\upsilon$ is smaller than one for all the cases studied, the 
1335: dependence of the non-diffusive contributions on $d_k$ is 
1336: weaker (in the sense of Appendix A) than the diffusive contributions. 
1337: The assumption on which we have built the stochastic model 
1338: have been verified numerically for a system having a sufficiently
1339: complex potential surface to be viewed as prototypical of a large set of
1340: many-particle systems.
1341: 
1342: The MC metric used in this work appears to be a valuable tool to 
1343: study the ergodicity properties of MC simulations. The non-ergodic components
1344: of the MC metric enable the prediction of the minimum length 
1345: a MC simulation must have in order to be considered ergodic. 
1346: The comparison of $\Gamma$ from 
1347: different algorithms gives a reasonable estimate of their relative
1348: efficiencies.
1349: 
1350: From the study of the melting region of 13 particle clusters, we have 
1351: found that the exponent $\upsilon$ depends both on the method used and the 
1352: nature of the potential energy function. We have performed calculations,
1353: not discussed in this work, where the functional form of the potential
1354: energy is modified.  These studies have shown $\upsilon$ to be dependent
1355: on the details of the potential. We have not found the exponent $\upsilon$ 
1356: to be a strong function of method.  Although PT and Met have
1357: significantly different efficiencies as measured by their relative
1358: diffusion coefficients, $\upsilon$ is nearly the same in the two
1359: methods.  The difference in the decay of $\Upsilon_K$ appears to be
1360: dominated by the coefficient in Eqs.(\ref{upa}) and (\ref{upa1})
1361: rather than the exponent.
1362: 
1363: 
1364: As discussed in the text, parallel tempering and J-walking studies of
1365: many-particle systems must have an initial high temperature 
1366: component that is chosen so that a Met simulation is known to be ergodic.
1367: For cluster simulations that require an external constraining potential
1368: to define the cluster, the radius of the constraining potential must be
1369: carefully chosen in order to achieve ergodic results.  We have found the
1370: metric and associated decay laws developed in this work to be a
1371: particularly valuable method of choosing these initial parameters in
1372: both parallel tempering and J-walking simulations.
1373: 
1374: We also remark that the metric introduced here may be a more
1375: sensitive probe of ergodicity than may be required in some applications. 
1376: For example in previous J-walking studies\cite{clusters} of 
1377: the 13-particle Lennard-Jones cluster,
1378: the heat capacity curve determined with a constraining radius of 4$\sigma$
1379: is nearly indistinguishable from the curve obtained with a constraining
1380: radius of 2$\sigma$.  From the results of this work, we know the initial
1381: high temperature walk is not ergodic when a constraining 
1382: radius of 4$\sigma$ is used.  It is striking that the non-ergodicity as
1383: measured by the energy metric is not apparent in the heat capacity
1384: curve.
1385: 
1386: We have constructed a metric based on an ensemble of MC trajectories. 
1387: By using an ensemble we attempt to cover sufficient portions of space
1388: so that all components are accessible.  In practice only a finite subset of
1389: a full ensemble can be included, and it is always possible that
1390: components of space are missed.  In such a case $\Upsilon_K$ may decay
1391: to zero numerically within the subspace, and the behavior may give
1392: misleading evidence that the simulation is ergodic.  Because 
1393: components of space may be missed in any finite simulation, it is impossible to
1394: guarantee ergodicity.  It is hoped by using a sufficiently large 
1395: ensemble of trajectories to define the metric, the possibility of
1396: missing components is minimized.
1397: 
1398: \section*{Acknowledgments}
1399: 
1400: We would like to thank Dr. O. Osenda for helpful comments.
1401: This work has been supported in part by the National Science Foundation under
1402: grant numbers CHE-9714970 and CDA-9724347.
1403: This research has 
1404: been supported in
1405: part by the Phillips Laboratory, Air Force Material Command, USAF,
1406: through the use of the MHPCC under cooperative agreement number
1407: F29601-93-0001.  The views and conclusions contained in this document
1408: are those of the authors and should not be interpreted as necessarily
1409: representing the official policies or endorsements, either expressed or
1410: implied, of Phillips Laboratory or the U.S. Government.
1411: 
1412: 
1413: \appendix
1414: \section{Weak dependence of the non-diffusive contributions}
1415: 
1416: We have considered two overall time scales for a MC simulation. 
1417: Properties calculated at short times (labeled $k$ in the discrete case)
1418: provide information about each step of the MC process, and properties
1419: averaged over the total simulation time (labeled $K$ in the discrete
1420: case) give information about the approach to ergodic behavior.  When $K$
1421: is sufficiently short we have
1422: both diffusive and non-diffusive contributions as a function of $k$.
1423: In this Appendix we explain the relative time dependence of the diffusive and 
1424: non-diffusive contributions to the autocorrelation function.
1425: 
1426: It has been assumed that the autocorrelation function 
1427: Eq. (\ref{correl}) can be expressed as the sum of diffusive terms plus 
1428: non-diffusive terms, i.e.
1429: 
1430: \begin{equation}
1431: \label{h1}
1432: \kappa(t,t')= \kappa_d(t,t') + \sum_{\ell=\lambda+1}^{\Lambda}\kappa_{nd\,,\,\ell}(t,t')\,\,,
1433: \end{equation}
1434: 
1435: \noindent where
1436: 
1437: \begin{eqnarray}
1438: \kappa_d(t,t')&=&\frac{\Gamma_0+\Gamma_1+\Gamma_2+\ldots+\Gamma_{\lambda
1439: }}
1440: {t_>}\label{h2}\\
1441: \kappa_{nd\,,\,\ell}(t,t')&=& \frac{\Gamma_{\ell}}{\tau_{\ell}} \, f_{\ell}\left(\frac{t_{\ell}^*}{\tau_{\ell}}\right)\label{h3}\,\,.
1442: \end{eqnarray}
1443: 
1444: Increasing the time variables by a factor $b>1$, such that 
1445: $\tau_{\lambda}\ll bt_>\ll\tau_{\lambda+1}$, with $\lambda\geq1$, 
1446: we can study the relative variations of each contribution to the correlation 
1447: function, diffusive and non-diffusive (labeled by $\ell>\lambda$).  In this
1448: Appendix we only consider values of $b$ such that
1449: the transformation $t\to bt$ does not increase the time scale beyond the
1450: local correlation time.  In Appendix B values of $b$ are considered that
1451: do cross such time scales.
1452: 
1453: By relative variations we mean
1454: 
1455: \begin{eqnarray}
1456: \Delta_d(t,t';b) &=&\left|  \frac{\kappa_d(bt,bt')-\kappa_d(t,t')}
1457: {\kappa_d(t,t')}\right|\label{h4}\\
1458: \Delta_{nd\,,\,\ell}(t,t';b) &=&\left|  \frac{\kappa_{nd\,,\,\ell}(bt,bt')
1459: -\kappa_{nd\,,\,\ell}(t,t')}{\kappa_{nd\,,\,\ell}(t,t')}
1460: \right|\label{h5}\,\,.
1461: \end{eqnarray}
1462: 
1463: \noindent The relative variation of each non-diffusive contribution is
1464: 
1465: \begin{equation}
1466: \label{h7}
1467: \Delta_{nd\,,\,\ell}(t,t';b) = \left|1-\frac1{b^2} \frac{\int_0^{bt}\! dt_1
1468: \int_0^{bt'}\! dt_2 \,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)}{\int_0^t\! dt_1
1469: \int_0^{t'}\! dt_2 \,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)}\right|\,.
1470: \end{equation}
1471: 
1472: \noindent whereas the relative variation of the diffusive contribution is
1473: 
1474: \begin{equation}
1475: \label{h6}
1476: \Delta_d(t,t';b) = 1-\frac1b\,\,,
1477: \end{equation}
1478: 
1479: \noindent If $\Delta_d(t,t';b)>\Delta_{nd\,,\,\ell}(t,t';b)$ for 
1480: all pair of times $t$ and $t'$ and for all $b>1$ such that 
1481: $bt_>\ll\tau_{\ell}$, we say that the  non-diffusive contributions are weaker 
1482: than the diffusive contribution in their dependence on $t$. We explore, 
1483: in the remainder of this appendix, the properties $f_{\ell}$ must have in 
1484: order that the inequality 
1485: $\Delta_d(t,t';b)>\Delta_{nd\,,\,\ell}(t,t';b)$
1486: is satisfied.
1487: 
1488: \vspace{.5cm}
1489: {\bf Lemma:} If the function $H_{\ell}(t;\tau)$
1490: 
1491: \begin{equation}
1492: \label{lemma1}
1493: H_{\ell}(t;\tau) = \int_0^t\!\!dt'\,\,f_{\ell}
1494: \left(\frac{t'}{\tau}\right)>0\,\,
1495: \end{equation}
1496: 
1497: \noindent satisfies the inequality
1498: 
1499: \begin{equation}
1500: \label{lemma2}
1501: H_{\ell}(t;\tau) >t\,f_{\ell} \left(\frac{t}{\tau}\right)\,\,\forall\,\,t\,
1502: {\rm and}\,\tau\,\,,
1503: \end{equation}
1504: 
1505: \noindent then, $H_{\ell}(t;\tau)$ is an increasing function of $\tau$.
1506: 
1507: \vspace{.5cm}
1508: {\bf Demonstration:} For $\ell$ and $t$ fixed, the function $H_{\ell}(t,\tau)$ 
1509: evaluated in $\tau'$ is
1510: 
1511: \begin{eqnarray}
1512: H_{\ell}(t;\tau')&=&\int_0^t\!\!dt'\,\,f_{\ell}\left(\frac{t'}{\tau'}
1513: \right)\\
1514: &=&\int_0^t\!\!dt'\,\,f_{\ell}\left(\frac{\tau t'}{\tau'\tau}\right)\\
1515: &=&\frac{\tau'}{\tau} \int_0^{\tau t/\tau'}\!\!du\,\,f_{\ell}
1516: \left(\frac{u}{\tau}\right)\\
1517: &=&\frac{\tau'}{\tau} H_{\ell}(\tau t/\tau';\tau) \label{n0}\,\,,
1518: \end{eqnarray}
1519: 
1520: \noindent then, for $\Delta\tau>0$
1521: 
1522: \end{multicols}
1523: \renewcommand{\thesection}{\Alph{section}}
1524: \renewcommand{\theequation}{\thesection\arabic{equation}}
1525: 
1526: \begin{eqnarray}
1527: \frac{H_{\ell}(t;\tau+\Delta\tau)-H_{\ell}(t;\tau)}{\Delta\tau}&=& 
1528: \frac1{\Delta\tau} \left\{\frac{\tau+\Delta\tau}{\tau} \int_0^
1529: {\tau t/(\tau+\Delta\tau)}\!\!dt'\,\,f_{\ell}\left(\frac{t'}{\tau}\right) -
1530: \int_0^t\!\!dt'\,\,f_{\ell}\left(\frac{t'}{\tau}\right)\right\}\\
1531: &=&\frac1{\Delta\tau} \left\{
1532: \frac{\Delta\tau}{\tau}\,\int_0^{\tau t/(\tau+\Delta\tau)}
1533: \!\!dt'\,\,f_{\ell}\left(\frac{t'}{\tau}\right) -
1534: \int_{\tau t/(\tau+\Delta\tau)}^t\!\!dt'\,\,f_{\ell}\left(\frac{t'}
1535: {\tau}\right)
1536: \right\}\\
1537: &=&\frac1{\Delta\tau} \left\{
1538: \frac{\Delta\tau}{\tau}\,\int_0^{\tau t/(\tau+\Delta\tau)}\!
1539: \!dt'\,\,f_{\ell}\left(\frac{t'}{\tau}\right) -
1540: \frac{t\Delta\tau}{\tau+\Delta\tau}\,f_{\ell}\left(\frac{t^*}{\tau}\right) 
1541: \right\}\label{app1}\,\,,
1542: \end{eqnarray}
1543: 
1544: \begin{multicols}{2}
1545: 
1546: \noindent where $t^*\in[t\tau/(\tau+\Delta\tau),t]$. In the limit 
1547: $\Delta\tau\to0$, and by virtue of the continuity of $f_{\ell}$, 
1548: the derivative takes the form
1549: 
1550: \begin{equation}
1551: \label{derivada}
1552: \frac{\partial H_{\ell}(t;\tau)}{\partial\tau}=\frac1{\tau}
1553: \left\{H_{\ell}(t;\tau)-tf_{\ell} \left(\frac t{\tau}\right)\right\}
1554: \label{der}\,\,.
1555: \end{equation}
1556: 
1557: \noindent Then, $\partial H_{\ell}(t;\tau)/\partial\tau>0$, and 
1558: $H_{\ell}(t;\tau)$ is an increasing function of $\tau.\;\;\Box$ 
1559: 
1560: Here we have presented the two first conditions 
1561: $f_{\ell}$ must have, namely Eqs. (\ref{lemma1}) and (\ref{lemma2}). 
1562: From Eq. (\ref{schwartz}) $f_{\ell}(0)$ is a global maximum, and
1563: the memory functions must have a positive peak at zero. The area below that 
1564: peak must be sufficiently large to satisfy Eq. (\ref{lemma1}). 
1565: Moreover, $f_{\ell}(0)$ must be sufficiently large to 
1566: satisfy Eq. (\ref{lemma2}), even at points where
1567: $f_{\ell}(t/\tau)$ is a local maximum. Then, to satisfy this Lemma, 
1568: we need a memory function with a sufficiently large global maximum at $t=0$.
1569: 
1570: \vspace{.5cm}
1571: {\bf Corollary:} Suppose $H_{\ell}(t;\tau_{\ell})>tf_{\ell}(t/\tau_{\ell})$. If $b>1$, then $0<\Delta_{nd\,,\,\ell}(t,t';b)<1$ for all pair of times $t$ and $t'$.
1572: \vspace{.5cm}
1573: 
1574: {\bf Demonstration:} Under the change of scale in time $t\to bt$, 
1575: $\kappa_{nd\,,\,\ell}(t,t')$ can be written
1576: 
1577: \end{multicols}
1578: 
1579: \begin{eqnarray}
1580: \kappa_{nd\,,\,\ell}(bt,bt')&=&\frac1{b^2tt'}\int_0^{bt}\!\! dt_1
1581: \int_0^{bt'}\!\! dt_2 \,\,\frac1{\tau_{\ell}}f_{\ell}
1582: \left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)\\
1583: &=&\frac1{tt'}\int_0^t\!\! dt_1
1584: \int_0^{t'}\!\! dt_2 \,\,\frac1{\tau_{\ell}}f_{\ell}
1585: \left(\frac{b|t_1-t_2|}{\tau_{\ell}}\right)\,\,,
1586: \end{eqnarray}
1587: 
1588: \noindent then, the quotient 
1589: $\kappa_{nd\,,\,\ell}(bt,bt')/\kappa_{nd\,,\,\ell}(t,t')$ is
1590: 
1591: \begin{eqnarray}
1592: \frac{\kappa_{nd\,,\,\ell}(bt,bt')}{\kappa_{nd\,,\,\ell}(t,t')} &=& 
1593: \frac{\int_0^t\!\! dt_1\left\{
1594: \int_0^{t_1}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}/b}\right)
1595: + \int_0^{t_>-t_1}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}/b}\right)
1596: \right\}} {\int_0^t\!\! dt_1
1597: \left\{
1598: \int_0^{t_1}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)+
1599: \int_0^{t_>-t_1}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)
1600: \right\}}\\
1601: &=&\frac{\int_0^t\!\! dt_1\left\{H_{\ell}(t_1;\tau_{\ell}/b)
1602: +H_{\ell}(t_>-t_1;\tau_{\ell}/b)\right\}}
1603: {\int_0^t\!\! dt_1\left\{H_{\ell}(t_1;\tau_{\ell})
1604: +H_{\ell}(t_>-t_1;\tau_{\ell})\right\}}\,\,.
1605: \label{n1}
1606: \end{eqnarray}
1607: 
1608: \begin{multicols}{2}
1609: 
1610: \noindent By Eq. (\ref{lemma1}), $H_{\ell}(t;\tau)>0\,\,\forall$ $t$ and 
1611: $\tau$. By the {\bf Lemma} the numerator is smaller than the denominator. 
1612: Then $0<\kappa_{nd\,,\,\ell}(bt,bt')/\kappa_{nd\,,\,\ell}(t,t')<1$ and 
1613: then, $0<\Delta_{nd\,,\,\ell}(t,t';b)<1.\;\;\Box$
1614: 
1615: \vspace{.5cm}
1616: {\bf Theorem:} Suppose that $b>1$ is such that 
1617: $\tau_{\ell-1}\ll bt_>\ll\tau_{\ell}$, 
1618:  $H_{\ell}(t;\tau_{\ell})>tf_{\ell}(t/\tau_{\ell})$, and
1619: all $f_{\ell}$ satisfy the Lipschitz condition 
1620: \cite{fomin}
1621: (for all
1622: closed interval ${\cal A}$ exists a real positive number $C_{\ell}$ such that
1623: 
1624: \begin{equation}
1625: \left|f_{\ell}(x) - f_{\ell}(y)\right|\leq C_{\ell}\,\left|x-y\right|
1626: \label{lipschitz}
1627: \end{equation}
1628: 
1629: for all $x$ and $y$ in ${\cal A}$). 
1630: Then $\Delta_{nd\,,\,\ell}(t,t';b)<\Delta_d(t,t';b)$ 
1631: if and only if $f_{\ell}$ is non-negative in the interval $[0,t_>)$.
1632: 
1633: \vspace{.5cm}
1634: {\bf Demonstration:} If $\Delta_{nd\,,\,\ell}(t,t';b)<\Delta_d(t,t';b)$,
1635: then
1636: 
1637: \begin{eqnarray}
1638: 1-\frac1b &>& 1-\frac1{b^2} \frac{\int_0^{bt}\! dt_1
1639: \int_0^{bt'}\! dt_2 \,\,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)}
1640: {\int_0^t\! dt_1
1641: \int_0^{t'}\! dt_2 \,\,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)}\\
1642: 1&<&\frac1b \frac{\int_0^{bt}\! dt_1
1643: \int_0^{bt'}\! dt_2 \,\,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)}
1644: {\int_0^t\! dt_1
1645: \int_0^{t'}\! dt_2 \,\,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)}
1646: \label{ineq}
1647: \end{eqnarray}
1648: 
1649: \noindent where the operations to reach Eq. (\ref{ineq}) are valid by 
1650: using {\bf Corollary}. Then
1651: 
1652: \end{multicols}
1653: 
1654: \begin{eqnarray}
1655: 0&<& \int_0^{bt}\!\! dt_1
1656: \int_0^{bt'}\!\! dt_2\,\, \frac1b\,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}
1657: \right)-
1658: \int_0^{t}\!\! dt_1
1659: \int_0^{t'}\!\! dt_2 \,\,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)\\
1660: 0&<& \int_0^{t}\!\! dt_1
1661: \int_0^{t'}\!\! dt_2 \left\{b\,f_{\ell}\left(\frac{b\,|t_1-t_2|}{\tau_{\ell}}
1662: \right)-
1663: \,f_{\ell}\left(\frac{|t_1-t_2|}{\tau_{\ell}}\right)\right\}\\
1664: 0&<& \int_0^{t_<}\!\! dt_1\left\{  
1665: \int_0^{t_1}\!\! dt_2 \left[b\,f_{\ell}\left(\frac{b\,(t_1-t_2)}{\tau_{\ell}}
1666: \right)-
1667: \,f_{\ell}\left(\frac{t_1-t_2}{\tau_{\ell}}\right)\right]+
1668: \int^{t_>}_{t_1}\!\! dt_2 \left[b\,f_{\ell}\left(\frac{b\,(t_2-t_1)}
1669: {\tau_{\ell}}\right)-
1670: \,f_{\ell}\left(\frac{t_2-t_1}{\tau_{\ell}}\right)\right]
1671: \right\}\\
1672: 0&<& \int_0^{t_<}\!\! dt_1\left\{  
1673: \int_0^{t_1}\!\! dt \left[b\,f_{\ell}\left(\frac{bt}{\tau_{\ell}}\right)-
1674: \,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)\right]
1675: +
1676: \int_0^{t_>-t_1}\!\! dt \left[b\,f_{\ell}\left(\frac{bt}{\tau_{\ell}}\right)-
1677: \,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)\right]
1678: \right\}\\
1679: 0&<& \int_0^{t_<}\!\! dt_1\left\{  
1680: \int_0^{bt_1}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)-
1681: \int_0^{t_1}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)
1682: +
1683: \int_0^{b(t_>-t_1)}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)-
1684: \int_0^{t_>-t_1}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)
1685: \right\}\\
1686: 0&<& \int_0^{t_<}\!\! dt_1\left\{  
1687: \int_{t_1}^{bt_1}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)
1688: +
1689: \int_{t_>-t_1}^{b(t_>-t_1)}\!\! dt \,\,f_{\ell}\left(\frac{t}{\tau_{\ell}}
1690: \right)\right\}\,\,.\label{k1}
1691: \end{eqnarray}
1692: 
1693: \noindent Using the intermediate 
1694: value theorem, \cite{spivak} we have
1695: 
1696: \begin{eqnarray}
1697: \int_t^{bt}\!\!dt'\,\,f_{\ell}\left(\frac{t'}{\tau_{\ell}}
1698: \right)&=&(b-1)\,t \,f_{\ell}\left(\frac{t^*(t)}{\tau_{\ell}}\right)\\
1699: &=&(b-1)\,t\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right) + (b-1)\,t\,
1700: \left[f_{\ell}\left(\frac{t^*(t)}{\tau_{\ell}}\right) - 
1701: f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)\right]\label{propx}\,\,,
1702: \end{eqnarray}
1703: 
1704: \noindent where $t^*(t)\in[t,bt]$. Let be $t^*_{\alpha}(t)$ and
1705: $t^*_{\beta}(t)$ the values at 
1706: which the intermediate value theorem is satisfied, in the intervals $[t,bt]$ 
1707: and $[t_>-t,b(t_>-t)]$ respectively
1708: 
1709: \begin{eqnarray}
1710: (b-1)\,t\,f_{\ell}\left(\frac{t^*_{\alpha}(t)}{\tau_{\ell}}\right) &=&
1711: \int_t^{bt}\!\!dt'\,\,f_{\ell}\left(\frac{t'}{\tau_{\ell}}\right)\\
1712: (b-1)\,(t_>-t)\,f_{\ell}\left(\frac{t^*_{\beta}(t)}{\tau_{\ell}}\right) 
1713: &=&
1714: \int_{t_>-t}^{b(t_>-t)}\!\!dt'\,\,f_{\ell}\left(\frac{t'}
1715: {\tau_{\ell}}\right)\,\,,
1716: \end{eqnarray}
1717: 
1718: \noindent then, the remainder can be written as
1719: 
1720: \begin{equation}
1721: \label{rem}
1722: R_{\ell}(t_<,t_>;b) = \int_0^{t_<}\!\!dt\,\,\left\{t\,\left[f_{\ell}
1723: \left(\frac{t^*_{\alpha}(t)}{\tau_{\ell}}\right) 
1724: - f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)\right]+(t_>-t)\,\left[f_{\ell}
1725: \left(\frac{t^*_{\beta}(t)}{\tau_{\ell}}\right) 
1726: - f_{\ell}\left(\frac{t_>-t}{\tau_{\ell}}\right)\right]\right\}\,\,.
1727: \end{equation}
1728: 
1729: \noindent By the Lipschitz condition, we have that
1730: 
1731: \begin{eqnarray}
1732: R_{\ell}(t_<,t_>;b) &\leq& \int_0^{t_<}\!\!dt\,\,\left\{t\,\left|f_{\ell}
1733: \left(\frac{t^*_{\alpha}(t)}{\tau_{\ell}}\right) 
1734: - f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)\right|+(t_>-t)\,\left|f_{\ell}
1735: \left(\frac{t^*_{\beta}(t)}{\tau_{\ell}}\right) 
1736: - f_{\ell}\left(\frac{t_>-t}{\tau_{\ell}}\right)\right|\right\}\\
1737: &<& \int_0^{t_<}\!\!dt\,\,\left\{t\,C_{\ell}\,\left|
1738: \frac{t^*_{\alpha}(t)-t}{\tau_{\ell}}\right|+(t_>-t)\,C_{\ell}\,\left|
1739: \frac{t^*_{\beta}(t)-(t_>-t)}{\tau_{\ell}}\right|\right\}\\
1740: &<& \frac{C_{\ell}}{\tau_{\ell}}\,\int_0^{t_<}\!\!dt\,\,\left\{t\,\left|
1741: bt-t\right|+(t_>-t)\,\left|
1742: b(t_>-t)-(t_>-t)\right|\right\}\\
1743: &<& \frac{C_{\ell}}{\tau_{\ell}}\,(b-1)\,\int_0^{t_<}\!\!dt\,\,\left[t^2
1744: +(t_>-t)^2\right]\\
1745: &<& \frac{C_{\ell}}{\tau_{\ell}}\,(b-1)\,\left(\frac23\,t_<^3+t_<t_>\,
1746: (t_>-t_<)\right)\\
1747: &<& \frac23\,t_>^3\,\frac{C_{\ell}}{\tau_{\ell}}\,(b-1)\label{cota}\,\,,
1748: \end{eqnarray}
1749: 
1750: \noindent where $C_{\ell}$ is a suitable positive real constant. Using Eqs. 
1751: (\ref{propx}) and (\ref{rem}) in Eq. (\ref{k1}) we have
1752: 
1753: \begin{eqnarray}
1754: 0&<& \int_0^{t_<}\!\! dt\,\,(b-1)\,\left\{ t\,f_{\ell}\left(\frac{t}
1755: {\tau_{\ell}}\right) + (t_>-t)\,f_{\ell}\left(\frac{t_>-t}{\tau_{\ell}}
1756: \right)\right\}+(b-1)\,R_{\ell}(t_<,t_>;b) \\
1757: 0&<& \int_0^{t_<}\!\! dt\,\,t\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right) +
1758: \int_{t_>-t_<}^{t_>}\!\! dt\,\,t\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)+
1759: R_{\ell}(t_<,t_>;b)\\
1760: 0&<& \int_0^{t_<}\!\! dt\,\,t\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right) +
1761: \int_0^{t_>}\!\! dt\,\,t\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)-
1762: \int_0^{t_>-t_<}\!\! dt\,\,t\,f_{\ell}\left(\frac{t}{\tau_{\ell}}\right)
1763: +R_{\ell}(t_<,t_>;b)\\
1764: 0&<& F_{\ell}(t_<) + F_{\ell}(t_>) - F_{\ell}(t_> - t_<)+
1765: \frac23\,t_>^3\,\frac{C_{\ell}}{\tau_{\ell}}\,(b-1)
1766: \label{xx}\,\,,
1767: \end{eqnarray}
1768: 
1769: \noindent where
1770: 
1771: \begin{equation}
1772: F_{\ell}(t) = \int_0^t\!\! dt'\,\,t'\,f_{\ell}\left(\frac{t'}{\tau_{\ell}}
1773: \right)\,\,,
1774: \end{equation}
1775: 
1776: \noindent is a continuous and differentiable function of $t$. 
1777: The inequality (\ref{xx}) holds for any $b>1$. Suppose that 
1778: $F_{\ell}(t_<) + F_{\ell}(t_>) - F_{\ell}(t_> - t_<)<0$. Then, if
1779: $b$ is such that
1780: 
1781: \begin{equation}
1782: b = 1 + \frac3{2L}\,\frac{\tau_{\ell}}{t_>^3\,C_{\ell}}\,
1783: \left|F_{\ell}(t_<) + F_{\ell}(t_>) - F_{\ell}(t_> - t_<)\right|\,\,,
1784: \end{equation}
1785: 
1786: \noindent where $L>2$, we have that
1787: 
1788: \begin{eqnarray}
1789: 0&<& F_{\ell}(t_<) + F_{\ell}(t_>) - F_{\ell}(t_> - t_<)+
1790: \frac1L\,\left|F_{\ell}(t_<) + F_{\ell}(t_>) - F_{\ell}(t_> - t_<)\right|\\
1791: 0&<&\frac{L-1}L\,\left[F_{\ell}(t_<) + F_{\ell}(t_>) - F_{\ell}(t_> - t_<)\right]
1792: \end{eqnarray}
1793: 
1794: \noindent in contradiction with the hypothesis that 
1795: $F_{\ell}(t_<) + F_{\ell}(t_>) - F_{\ell}(t_> - t_<)$ is negative. Then
1796: 
1797: \begin{equation}
1798: 0\leq F_{\ell}(t_<) + F_{\ell}(t_>) - F_{\ell}(t_> - t_<)\label{xxx}\,\,.
1799: \end{equation}
1800: 
1801: Let us define the function 
1802: 
1803: \begin{equation}
1804: \Delta F_{\ell}(t) = F_{\ell}(t) - F_{\ell}(t_> - t)\,\,,
1805: \end{equation}
1806: 
1807: \noindent where $t\in(0,t_>)$. The right derivative at $t=0$ of 
1808: $\Delta F_{\ell}(t)$ is
1809: 
1810: \begin{eqnarray}
1811: \lim_{\Delta t\to0^+}\frac{\Delta F_{\ell}(\Delta t) - \Delta F_{\ell}(0)}
1812: {\Delta t}&=&\lim_{\Delta t\to0^+}\frac{F_{\ell}(\Delta t)-F_{\ell}(0)
1813: +F_{\ell}(t_>)-F_{\ell}(t_>-\Delta t)}{\Delta t}\\
1814: &=&
1815: \lim_{\Delta t\to0^+}\frac1{\Delta t}
1816: \left\{\int_0^{\Delta t}\!\!dt\,\,t\,f_{\ell}\left(\frac{t}{\tau_{\ell}}
1817: \right)+\int^{t_>}_{t_>-\Delta t}\!\!dt\,\,t\,f_{\ell}\left(\frac{t}
1818: {\tau_{\ell}}\right)
1819: \right\}\\
1820: &=&\lim_{\Delta t\to0^+}\frac1{\Delta t}
1821: \left\{\Delta t \,\,t_1^* f_{\ell}\left(\frac{t_1^*}{\tau_{\ell}}\right) 
1822: + \Delta t \,\,t_2^* f_{\ell}\left(\frac{t_2^*}{\tau_{\ell}}\right)\right\}
1823: \label{n3}
1824: \end{eqnarray}
1825: 
1826: \begin{multicols}{2}
1827: 
1828: \noindent where $t_1^*\in[0,\Delta t]$ and $t_2^*\in[t_>-\Delta t,t_>]$. Thus
1829: 
1830: \begin{equation}
1831: \label{n5}
1832: \left.\frac{\partial\Delta F_{\ell}(t)}{\partial t}\right|_{t\to 0^+}
1833: = t_>\,f_{\ell}\left(\frac{t_>}{\tau_{\ell}}\right)\,\,.
1834: \end{equation}
1835: 
1836: \noindent If the right derivative at 0 of $\Delta F_{\ell}(t)$ is 
1837: negative, $\Delta F_{\ell}(t)$ approaches $-F_{\ell}(t_>)$ from below, 
1838: when $t\to0$. There exists a time $0<\tilde{t}<t_>$, such that 
1839: $0> F_{\ell}(t_>) + \Delta F_{\ell}(\tilde{t})$, in contradiction with 
1840: Eq. (\ref{xxx}). Then, $f_{\ell}$ must be non-negative for $t\in(0,t_>)$. 
1841: By the property Eq. (\ref{schwartz}) $f_{\ell}(0)$ must be positive.
1842: This proves that $\Delta_{nd\,,\,1}(t,t';b) < \Delta_d(t,t';b) 
1843: \Rightarrow f_{\ell}(t)\geq0$ for $0 \le t < t_{>}$.
1844: To demonstrate that if $f_{\ell}$ is positive yields 
1845: $\Delta_{nd\,,\,1}(t,t';b)<\Delta_d(t,t';b)$ (i.e. the converse), 
1846: follow the argument 
1847: backwards, from Eq. (\ref{k1}). $\Box$
1848: 
1849: In conclusion, if the memory functions are positive, 
1850: satisfy the Lipschitz condition, and satisfy the 
1851: condition Eqs. (\ref{lemma1}) and (\ref{lemma2}), 
1852: the non-diffusive 
1853: contributions are more weakly dependent on time than $1/t$. 
1854: 
1855: The results of the present appendix are valid in the 
1856: limit of a complete 
1857: ensemble. In our numerical experiments only partial samples of the ensemble can
1858: be considered. The memory functions that appear in our numerical calculations 
1859: come from partial mean values of the 
1860: product of discontinuous
1861: functions (every noise process is a discontinuous function). These
1862: memory functions are discontinuous. The behavior of the 
1863: non-diffusive contributions observed in our
1864: numerical experiments is in agreement with these analytic 
1865: (infinite ensemble limit) results. We can infer that there might be
1866: a version of the theorem applied to discontinuous memory functions, but
1867: we have been unable to develop such a theorem.
1868: 
1869: \section{Consequences of the time scale change in the non-diffusive 
1870: contributions}
1871: 
1872: In this appendix we show the behavior of the function $f_1$ when its 
1873: correlation time is changed according to $\tau_1\to\tau_{b1}=\tau_1/b$, 
1874: with $b\gg1$; i.e. when the total simulation time is scaled to exceed
1875: the correlation time of the first colored noise process.
1876: 
1877: We multiply the time variables by a number $b$, such that 
1878: $\tau_1\ll bt_> \ll \tau_2$. We have that the $g_1$ 
1879: process contributes to the autocorrelation function with
1880: 
1881: \end{multicols}
1882: 
1883: \begin{eqnarray}
1884: \frac1{b^2tt'}\langle G_1(bt/\tau_{\ell})\,G_1(bt'/\tau_{\ell})
1885: \rangle &=&\frac1{b^2tt'}\, \int_0^{bt_<}\!\! dt_1\int_0^{bt_>}\!\! 
1886: dt_2\,\,\frac1{\tau_1}f_1\left(\frac{|t_1-t_2|}{\tau_1}\right)\\
1887: &=& \frac1{btt'}\,\int_0^{t_<}\!\! dt_1'\int_0^{t_>}\!\! dt_2'\,\,
1888: \frac1{\tau_{b1}}f_1\left(\frac{|t_1'-t_2'|}{\tau_{b1}}\right) \label{eq:6}
1889: \end{eqnarray}
1890: 
1891: \noindent where $t'=t/b$ and $\tau_{b1}=\tau_1/b$. We want to 
1892: compute this contribution both within the neighborhood $t_1=t_2$ as well
1893: as outside
1894: such a region. To do so, we can split the integral in Eq. (\ref{eq:6}) 
1895: in three parts
1896: 
1897: \begin{equation}
1898: \frac1{b^2tt'}\langle G_1(bt/\tau_{\ell})\,G_1(bt'/\tau_{\ell})\rangle = 
1899: I_1+I_2+I_3
1900: \end{equation}
1901: 
1902: \noindent where
1903: 
1904: \begin{eqnarray}
1905: I_1 &=& \frac1{btt'}\,\int_0^{t_<}\!\! dt_1
1906: \int_0^{\max(0,t_1-\epsilon/2)}\!\! dt_2\,\,\frac1{\tau_{b1}}f_1
1907: \left(\frac{t_1-t_2}{\tau_{b1}}\right)\label{i1}\\
1908: I_2 &=& \frac1{btt'}\,\int_0^{t_<}\!\! dt_1
1909: \int_{\max(0,t_1-\epsilon/2)}^{\min(t_>,t_1+\epsilon/2)}\!\! dt_2\,\,
1910: \frac1{\tau_{b1}}f_1\left(\frac{|t_1-t_2|}{\tau_{b1}}\right)\label{i2}\\
1911: I_3 &=& \frac1{btt'}\,\int_0^{t_<}\!\! dt_1
1912: \int_{\min(t_>,t_1+\epsilon/2)}^{t_>}\!\! dt_2\,\,
1913: \frac1{\tau_{b1}}f_1\left(\frac{t_2-t_1}{\tau_{b1}}\right)\label{i3}
1914: \end{eqnarray}
1915: 
1916: \noindent with $t_<>\epsilon>0$ 
1917: (observe that the only integral involving $t_1=t_2$ is $I_2$). 
1918: Consider $I_1$. If $t_1<\epsilon/2$ the inner integral is zero. 
1919: Therefore, $t_1$ must be bigger than $\epsilon/2$ and
1920: 
1921: \begin{equation}
1922: I_1 = \frac1{bt_<t_>}\,\int_{\epsilon/2}^{t_<}\!\! dt_1
1923: \int_0^{t_1-\epsilon/2}\!\! dt_2\,\,\frac1{\tau_{b1}}f_1
1924: \left(\frac{t_1-t_2}{\tau_{b1}}\right)\,\,,
1925: \end{equation}
1926: 
1927: \noindent which, by virtue of the continuity of $f_1$, can be bounded as
1928: follows
1929: 
1930: \begin{eqnarray}
1931: \frac1{bt_<t_>}\,\int_{\epsilon/2}^{t_<}\!\! dt_1\,\,
1932: \frac b{\tau_1}\,\left(t_1-\frac{\epsilon}2\right)f_1
1933: \left(\frac{bt_{min}}{\tau_1}\right)
1934: \leq &I_1&\leq \frac1{bt_<t_>}\,\int_{\epsilon/2}^{t_<}\!\! dt_1\,\,
1935: \frac b{\tau_1}\,\left(t_1-\frac{\epsilon}2\right)f_1\left(\frac{bt_{max}}
1936: {\tau_1}\right)\nonumber\\
1937: \frac12\,\frac{(t_<-\epsilon/2)^2}{t_<t_>}\,\frac1{\tau_1}\,f_1
1938: \left( \frac{bt_{min}}{\tau_1}\right)\leq&I_1&\leq
1939: \frac12\,\frac{(t_<-\epsilon/2)^2}{t_<t_>}\,\frac1{\tau_1}\,f_1
1940: \left( \frac{bt_{max}}{\tau_1}\right)
1941: \label{i111}
1942: \end{eqnarray}
1943: 
1944: \noindent where $t_{max}$ ($t_{min}$) is the time in the interval 
1945: $[\epsilon/2,t_<]$ at which the function $f_1$ 
1946: reaches its maximum (minimum) value. Because $f_1$ is continuous, 
1947: there exists $t_1^*\in[t_{min},t_{max}]$ at which
1948: 
1949: \begin{equation}
1950: \label{i11}
1951: I_1=
1952: \frac12\,\frac{(t_<-\epsilon/2)^2}{t_<t_>}\,\frac1{\tau_1}\,f_1
1953: \left( \frac{bt_1^*}{\tau_1}\right)\,\,.
1954: \end{equation}
1955: 
1956: Consider now $I_3$. If $t_1+\epsilon/2>t_>$, the inner integral is zero. 
1957: Therefore, $0<t_1<\min(t_<,t_>-\epsilon/2)$ and
1958: 
1959: \begin{eqnarray}
1960: I_3&=& \frac1{btt'}\,\int_0^{\min(t_<,t_>-\epsilon/2)}\!\! dt_1
1961: \int_{t_1+\epsilon/2}^{t_>}\!\! dt_2\,\,\frac1{\tau_{b1}}f_1
1962: \left(\frac{t_2-t_1}{\tau_{b1}}\right)\nonumber\\
1963: &=& \frac{\min(t_<,t_>-\epsilon/2)}{t_<t_>}\left[t_>-\frac{\epsilon}2
1964: -\frac12 \min(t_<,t_>-\epsilon/2) \right]\,\frac1{\tau_1}\,f_1\left( 
1965: \frac{bt_3^*}{\tau_1}\right)\,\,,
1966: \label{i33}
1967: \end{eqnarray}
1968: 
1969: \noindent where $t^*_3\in[t_{min},t_{max}]$, and now $t_{max}$ ($t_{min}$) 
1970: is the time in $[\epsilon/2,t_>]$ at which the function $f_1$ reaches its 
1971: maximum (minimum) value.
1972: 
1973: \begin{figure}
1974: \epsfxsize=.48\textwidth
1975: \epsfbox{f7.ps}
1976: \end{figure}
1977: 
1978: {\footnotesize
1979: {\bf Fig. 7:} The area under the curve represents the 
1980: first integral in Eq. (\ref{f1}). The darker piece is 
1981: half of the integral in the interval $[-t_1,t_1]$, the lighter is 
1982: half of the integral in $[-\epsilon/2,\epsilon/2]$.}
1983: \vspace{.5cm}
1984: 
1985: Let us consider now $I_2$. First observe that for the integral in $t_1$, 
1986: if $0\leq t_1\leq\epsilon/2$, $\max(0,t_1-\epsilon/2)=0$ 
1987: and $\min(t_>,t_1+\epsilon/2)=t_1+\epsilon/2$. 
1988: If $\epsilon/2\leq t_1\leq t_<$ then $\max(0,t_1-\epsilon/2)=
1989: t_1-\epsilon/2$. Then
1990: 
1991: \begin{equation}
1992: I_2 = \frac1{bt_<t_>}\,\left\{ \int_0^{\epsilon/2}\!\!dt_1 
1993: \int_{0}^{t_1+\epsilon/2}\!\!dt_2\,\,\frac1{\tau_{b1}} f_1
1994: \left(\frac{|t_1-t_2|}{\tau_{b1}}\right) +
1995: \int_{\epsilon/2}^{t_<}\!\!dt_1 
1996: \int_{t_1-\epsilon/2}^{\min(t_>,t_1+\epsilon/2)}\!\!dt_2\,\,
1997: \frac1{\tau_{b1}} f_1\left(\frac{|t_1-t_2|}{\tau_{b1}}\right)\right\}
1998: \label{f1}\,\,.
1999: \end{equation}
2000: 
2001: \noindent The integral in $t_2$ between 0 and $t_1+\epsilon/2$ can be 
2002: evaluated with the help of Fig. 7
2003: 
2004: \begin{equation}
2005: \int_{0}^{t_1+\epsilon/2}\!\!dt_2\,\,\frac1{\tau_{b1}} f_1
2006: \left(\frac{|t_1-t_2|}{\tau_{b1}}\right)=\frac12 \,\,
2007: \int_{-\epsilon/2}^{\epsilon/2}\!\!dt\,\,\frac1{\tau_{b1}} 
2008: f_1\left(\frac{|t|}{\tau_{b1}}\right) + \frac12 \,\,
2009: \int_{-t_1}^{t_1}\!\!dt\,\,\frac1{\tau_{b1}} f_1\left(\frac{|t|}
2010: {\tau_{b1}}\right)\,\,.
2011: \label{int1}
2012: \end{equation}
2013: 
2014: The second integral in $t_1$ can be separated in two parts; the first for 
2015: $\epsilon /2 \leq t_1 \leq \min(t_<,t_>-\epsilon/2)$ and the 
2016: second for $\min(t_<,t_>-\epsilon/2)\leq t_1 \leq t_<$. If 
2017: $t_>-t_<<\epsilon/2$ the second term is zero. Then
2018: 
2019: \begin{eqnarray}
2020: \int_{\epsilon/2}^{t_<}\!\!dt_1 
2021: \int_{t_1-\epsilon/2}^{\min(t_>,t_1+\epsilon/2)}&&\!\!dt_2\,\,
2022: \frac1{\tau_{b1}} f_1\left(\frac{|t_1-t_2|}{\tau_{b1}}\right) = 
2023: \int_{\epsilon/2}^{\min(t_<,t_>-\epsilon/2)}\!\!dt_1 
2024: \int_{t_1-\epsilon/2}^{\min(t_>,t_1+\epsilon/2)}\!\!dt_2\,\,
2025: \frac1{\tau_{b1}} f_1\left(\frac{|t_1-t_2|}{\tau_{b1}}\right) + 
2026: \nonumber\\ 
2027: &&\;\;\;\;\;\;\;\;\;\;\;\;\;
2028: \Theta\left(\frac{\epsilon}2+t_<-t_>\right)\,\,
2029: \int_{t_>-\epsilon/2} ^{t_<}\!\!dt_1
2030: \int_{t_1-\epsilon/2}^{\min(t_>,t_1+\epsilon/2)}\!\!dt_2\,\,
2031: \frac1{\tau_{b1}} f_1\left(\frac{|t_1-t_2|}{\tau_{b1}}\right)\,\,,
2032: \label{int2}
2033: \end{eqnarray}
2034: 
2035: \noindent where $\Theta$ is the step function. If 
2036: $t_1\leq\min(t_<,t_> -\epsilon/2)$ then $\min(t_>,t_1+\epsilon/2)=
2037: t_1+\epsilon/2$. The last integral in 
2038: $t_2$ can be rearranged in the same way as 
2039: Eq. (\ref{int1}). Then
2040: 
2041: \begin{eqnarray}
2042: \int_{\epsilon/2}^{t_<}\!\!dt_1&& 
2043: \int_{t_1-\epsilon/2}^{\min(t_>,t_1+\epsilon/2)}\!\!dt_2\,\,
2044: \frac1{\tau_{b1}} f_1\left(\frac{|t_1-t_2|}{\tau_{b1}}\right) = 
2045: \int_{\epsilon/2}^{\min(t_<,t_>-\epsilon/2)}\!\!dt_1 
2046: \int_{-\epsilon/2}^{\epsilon/2}\!\!dt\,\,\frac1{\tau_{b1}} 
2047: f_1\left(\frac{|t|}{\tau_{b1}}\right) + \nonumber\\ 
2048: &&\;\;\;\;\;\;\;\;\;\;\;\;\;\;\frac12\,
2049: \Theta\left(\frac{\epsilon}2+t_<-t_>\right)\,\,
2050: \int_{t_>-\epsilon/2} ^{t_<}\!\!dt_1\left[
2051: \int_{-\epsilon/2}^{\epsilon/2}\!\!dt\,\,\frac1{\tau_{b1}} 
2052: f_1\left(\frac{|t|}{\tau_{b1}}\right)+
2053: %\right.\\&&\left. 
2054: \int_{t_1-t_>}^{t_>-t_1}\!\!dt\,\,\frac1{\tau_{b1}} f_1\left(\frac{|t|}
2055: {\tau_{b1}}\right)\right]\,\,,
2056: \label{int3}
2057: \end{eqnarray}
2058: 
2059: We can observe that the correlation time $\tau_{b1}$ goes to zero when 
2060: $b$ is increased. The function $(1/\tau_1)\,f_1(bt/\tau_{1})$ 
2061: becomes negligible outside a neighborhood of $t=0$ [observe Eqs. (\ref{i11}) 
2062: and (\ref{i33})]. Equation (\ref{norm}) holds, then, if $b$ is sufficiently 
2063: large, $(1/\tau_{b1})\,f_1(t/\tau_{b1})$ can be considered a delta function. 
2064: The integrals $I_1$ and $I_3$ become zero, and the integrals involving $t=0$ 
2065: in the expression of $I_2$ converge to one. $I_2$ becomes
2066: 
2067: \begin{equation}
2068: I_2 = \frac1{bt_<t_>} \,\,\left\{ \min\left(t_<,t_>-\frac{\epsilon}2\right) + \Theta\left
2069: (\frac{\epsilon}2+t_<-t_>\right)\, \left(\frac{\epsilon}2+t_<-t_> \right) \right\}=\frac1{bt_>}\,\,,
2070: \label{i22}
2071: \end{equation}
2072: 
2073: \begin{multicols}{2}
2074: 
2075: \noindent which is a diffusive contribution to the autocorrelation function. 
2076: The autocorrelation function becomes then
2077: 
2078: \begin{equation}
2079: \kappa(bt,bt') =
2080: \frac{\Gamma_0+\Gamma_1}{bt_>}+\sum_{\ell=2}^{\Lambda}
2081: \frac{\Gamma_{\ell}}{\tau_{\ell}} \, f_{\ell}
2082: \left(\frac{t_{b\ell}^*(t_>)}{\tau_{b\ell}}\right)\,\,.
2083: \end{equation}
2084: 
2085: The same argument can be used when $b$ is such that $\tau_2\ll bt_>\ll\tau_3$. After such changes in the time scale, the diffusion coefficient $\Gamma=\Gamma_0+\Gamma_1$ is enlarged, and the non-diffusive contributions are reduced. There is an ultimate scale change, such that $\tau_{\Lambda}\ll bt_>$.
2086: Beyond this maximum time scale the process can be considered diffusive.
2087: 
2088: \begin{references}
2089: 
2090: \bibitem{valleau} J. P. Valleau and S. G. Whittington, A guide to Monte Carlo
2091: for Statistical Mechanics: 1 Highways, in {\em Statistical Mechanics, Part A:
2092: Equilibrium Techniques}, Modern Theoretical Chemistry Series, Vol. 5, Chap. 4,
2093: B. Berne Ed. (Plenum, New York, 1976).
2094: 
2095: \bibitem{wood} W. W. Wood and F. R. Parker, J. Chem. Phys. {\bf27}, 720 (1957).
2096: 
2097: \bibitem{thiru1} D. Thirumalai, R. D. Mountain, and T. R. Kirpatrick, Phys. Rev. A {\bf 39}, 3563 (1989).
2098: 
2099: \bibitem{palmer} R. G. Palmer, Adv. Phys. {\bf 31}, 669 (1982).
2100: 
2101: \bibitem{gardiner} C. W. Gardiner, Handbook of Stochastic Methods for Physics, Chemistry and the Natural Sciences (Springer-Verlag. Berlin, Heidelberg, New York, Tokyo, 1983).
2102: 
2103: \bibitem{freeman1} D. L. Freeman and J. D. Doll, Annu. Rev. Phys. Chem. {\bf47},
2104: 43 (1996).
2105: 
2106: \bibitem{wales1} R. M. Lynden-Bell and D. J. Wales, J. Chem. Phys. {\bf 101}, 1460 (1994).
2107: 
2108: \bibitem{wales2} J. P. K. Doye, D. J. Wales, and M. A. Miller, J. Chem. Phys. {\bf 109}, 8143 (1998).
2109: 
2110: \bibitem{thiru2} R. D. Mountain and D. Thirumalai, J. Chem. Phys. {\bf 93}, 6975 (1989).
2111: 
2112: \bibitem{thiru3} D. Thirumalai and R. D. Mountain, Phys. Rev. A {\bf 42}, 4574 (1990).
2113: 
2114: \bibitem{thiru4} J. E. Straub and D. Thirumalai, Proc. Nat. Acad. Sci. USA {\bf 90}, 809 (1993).
2115: 
2116: \bibitem{lieberman} A. J. Lichtenberg and M. A. Lieberman, {\em Regular and 
2117: Stochastic Motion} (Springer-Verlag, New York, 1983).
2118: 
2119: \bibitem{straub} I. Andricioaei and J. E. Straub, J. Chem. Phys. {\bf107},
2120: 9117 (1997).
2121: 
2122: \bibitem{juan1} J. P. Neirotti, F. Calvo, D. L. Freeman and J. D. Doll, J. Chem.
2123: Phys. {\bf112}, 10340 (2000).
2124: 
2125: \bibitem{juan2} F. Calvo, J. P. Neirotti, D. L. Freeman and J. D. Doll, J. Chem.
2126: Phys. {\bf112}, 10350 (2000).
2127: 
2128: 
2129: \bibitem{frantz} D. D. Frantz, D. L. Freeman, and J. D. Doll, J. Chem. Phys. 
2130: {\bf93}, 2769 (1990).
2131: 
2132: \bibitem{p1} E. Marinari and G. Parissi, Europhys. Lett. {\bf19}, 451 (1992).
2133: 
2134: \bibitem{p2} C. J. Geyer and E. A. Thompson, J. Am. Stat. Assoc. {\bf90},
2135: 909 (1995).
2136: 
2137: \bibitem{p3} M. Falcioni and M. W. Deem, J. Chem. Phys. {\bf110}, 1754 (1999).
2138: 
2139: \bibitem{metropolis} N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. 
2140: Teller, and E. Teller, J. Chem. Phys. {\bf21}, 1087 (1953).
2141: 
2142: \bibitem{lee} J. K. Lee, J. A. Barker, and F. F. Abraham, J. Chem. Phys. 
2143: {\bf58}, 3166 (1973).
2144: 
2145: \bibitem{labastie} P. Labastie and R. L. Whetten, Phys. Rev. Lett. {\bf65}, 1567
2146: (1990).
2147: 
2148: \bibitem{juan3} J. P. Neirotti, D. L. Freeman, and J. D. Doll, J. Chem. Phys. 
2149: {\bf112}, 3990 (2000).
2150: 
2151: \bibitem{spivak} M. Spivak, Calculus (Publish or Perish, 3ed.,
2152: 1994).
2153: 
2154: \bibitem{berry} R. S. Berry, T. L. Beck, H. L. Davis, and J. Jellinek, Adv. 
2155: Chem. Phys. {\bf 70B}, 75 (1988).
2156: 
2157: \bibitem{clusters} D. D. Frantz, J. Chem. Phys. {\bf102}, 3747 (1995).
2158: 
2159: \bibitem{citfreq}Assuming the tolerable error to be on the order of 1\%,
2160: we set $0.01 \simeq |[\cot(\omega_n/2)-2/\omega_n]/(2/\omega_n)| =
2161: \omega_n^2/12+{\cal O}(\omega_n^4)$.  Then $n_{max}=bK\sqrt{12}/20\pi$.
2162: 
2163: \bibitem{fomin} A. N. Kolmogorov and S. V. Fomin, {\em Introductory Real 
2164: Analysis} (Dover, New York, 1970).
2165: 
2166: \end{references}
2167: 
2168: \end{multicols}
2169: 
2170: \end{document}
2171: 
2172: