1: \documentclass[12pt]{iopart}
2: \usepackage{graphicx}
3: \begin{document}
4: \title{Optimum Monte Carlo Simulations: Some Exact Results}
5: \author{J. Talbot$^1$, G. Tarjus$^2$ and P. Viot$^2$}
6: \address{$^1$Department of Chemistry and Biochemistry,
7: Duquesne University, Pittsburgh, PA 15282-1530\\
8: $^2$Laboratoire de Physique Th\'eorique des Liquides, Universit\'e
9: Pierre et Marie Curie, 4, place Jussieu, 75252 Paris Cedex, 05 France}
10: %\bibliographystyle{apsrev}
11: %\bibliographystyle{apalike}
12:
13: \begin{abstract}
14: We obtain exact results for the acceptance ratio and mean squared
15: displacement in Monte Carlo simulations of the simple harmonic
16: oscillator in $D$ dimensions. When the trial displacement is made
17: uniformly in the radius, we demonstrate that the results are
18: independent of the dimensionality of the space. We also study the
19: dynamics of the process via a spectral analysis and we obtain an
20: accurate description for the relaxation time.
21: \end{abstract}
22: \pacs{05.70.Ln,81.05.Rm,75.10.Nr,64.60.My, 68.43.Mn, 75.40.Gb}
23: %\maketitle
24: \section{Introduction}
25:
26: Since the original Metropolis algorithm appeared five decades ago,
27: countless studies have employed the technique to evaluate the
28: thermodynamic properties of model systems\cite{AT87,B97,FS02}. The
29: essence of the method is to generate a sequence of configurations that
30: represent a given thermodynamic ensemble, often the canonical
31: ensemble. Properties of interest are then obtained as averages over
32: the configurations. At each step of the simulation a trial
33: configuration is obtained from the current one by making a random
34: displacement in the configuration space. This might correspond to, for
35: example, displacing a randomly selected particle. The trial
36: configuration is either accepted or rejected with a probability given
37: by the appropriate Boltzmann factor for the ensemble. In case of
38: rejection, the current configuration is retained for use in evaluating
39: the properties of interest.
40:
41: Since many applications of MC are computationally intensive, a much
42: addressed issue has been the optimization of the simulation with
43: respect to one or more control parameters so that the configuration
44: space is sampled in the most efficient way. Bouzida, Kumar and
45: Swendsen (BKS)\cite{BKS92,S02}, as part of a program aimed at
46: improving the efficiency of MC simulations of biomolecules, performed
47: numerical studies of the simple harmonic oscillator (SHO) where the
48: convergence of the simulation depends on the maximum
49: displacement. There is no unique measure of efficiency, but two simple
50: choices are the mean squared and mean absolute displacements. As the
51: maximum displacement tends to zero or infinity it is clear that the
52: average value of both of these quantities tends to zero since, in the
53: first case the particle does not move, while in the second all
54: attempted moves are rejected. Thus both quantities have a maximum for
55: some intermediate value of the maximum displacement.
56:
57: It is also useful to consider the dynamical process associated with an
58: MC simulation, even though it does not correspond to the actual
59: dynamics of the system. One can, for example, calculate various time
60: correlation functions that can be used to develop alternative
61: efficiency criteria\cite{K88,MT94}.
62:
63:
64: BKS\cite{BKS92} performed numerical studies of the SHO in one, two and
65: three dimensions and examined the acceptance ratio $P_{acc}$ (i.e.,
66: the fraction of accepted trial configurations), mean squared, $<(\Delta
67: x)^2)>$, and mean absolute, $<|\Delta x|>$, displacements as a function of
68: the maximum displacement, $\delta$. They found that the acceptance ratio
69: decreases approximately exponentially for small to intermediate values
70: of $\delta$ and then inversely for larger values. In one dimension they
71: found that the maxima in $<(\Delta x)^2>$ and $<|\Delta x|>$ occur at
72: $P_{acc}=0.42$ and $P_{acc}=0.56$, respectively. In higher dimensions
73: the results depend on how the jump is made. BKS considered two cases:
74: in one the jumps are performed uniformly to any point in a spherical
75: volume of radius $\delta$ centered on the current position, a choice that
76: favors larger radial displacements for dimensions $D>1$. In a second
77: method, the jumps were sampled uniformly in the radius (and randomly
78: in the orientation) so that all radial displacements are equally
79: probable. In the former case BKS observed that, for a given $\delta$, $P_{acc}$
80: decreases as a function of $D$, while for uniform radius sampling the
81: numerical results suggested that $P_{acc}$ is independent of $D$.
82:
83: In addition to these static properties, BKS also examined the
84: correlation time, $\tau$, of the energy-energy correlation function. They
85: observed a minimum correlation time for an acceptance ratio of
86: approximately $50\%$.
87:
88: Here we present an analytical study of the SHO in arbitrary dimension.
89: We obtain exact expressions for the acceptance ratio and the mean
90: squared and mean absolute displacements as functions of the maximum
91: displacement $\delta$. We show that when the trial jump is selected
92: uniformly in the radius, the results are independent of the dimension.
93: We also present an analysis of the dynamics of the process.
94:
95: \section{ONE DIMENSION}
96: We first investigate the case of a SHO in one dimension, whose
97: potential energy is given by $V(x)=kx^2/2$ where $x$ is the position
98: and $k$ is the stiffness constant.
99:
100: In a standard Metropolis Monte Carlo Simulation one makes a trial move
101: with a uniform random displacement selected between $-\delta$ and $\delta$.
102: The dynamical process generated by the successive trial moves of the
103: Monte Carlo simulation can be written
104: \begin{eqnarray}\label{eq:1}
105: \fl\frac{dP(x,t)}{dt}&=-\frac{1}{2\delta}\int_{-\delta}^{\delta}dhW(x\to x+h)P(x,t)+
106: \frac{1}{2\delta}\int_{-\delta}^{\delta}dhW(x+h\to x)P(x+h,t)
107: \end{eqnarray}
108: where $W(x\to x+h)$ denotes the transition rate from the state $x$ to
109: the state $x+h$ and $P(x,t)$ is the probability to find the oscillator
110: at position $x$ at time $t$. To ensure the convergence towards
111: equilibrium, a sufficient condition is given by detailed balance,
112: which is expressed as
113: \begin{equation}\label{eq:2}
114: \frac{W(x\to x+h)}{W(x+h\to x)}=\frac{P_{eq}(x+h)}{P_{eq}(x)}
115: \end{equation}
116: where
117: \begin{equation}\label{eq:3}
118: P_{eq}(x)=c \exp(-\beta V(x))
119: \end{equation}
120: and $c=\sqrt{\beta k/2\pi}$ is a normalization constant ensuring that
121: $\int_{-\infty}^{\infty} dx
122: P_{eq}(x)=1$. One solution of Eq.~(\ref{eq:2}) is the Metropolis
123: rule,
124: \begin{equation}\label{eq:4}
125: W(x\to x+h)=\min(1,\exp(-\beta(V(x+h)-V(x))).
126: \end{equation}
127:
128: \begin{figure}
129: \begin{center}
130: \resizebox{10cm}{!}{\includegraphics{1d.ps}}
131:
132: \caption{Acceptance ratio $P_{acc}$ versus $\xi=\sqrt {\frac{\beta k}{2}}\delta$
133: for a harmonic oscillator with uniform volume sampling in $D=1,2$, and
134: $3$ dimensions (full, dashed, and dotted lines respectively). For a
135: uniform radius sampling, all curves coincide with the one-dimensional
136: result (full line).}\label{fig:1}
137: \end{center}
138: \end{figure}
139:
140:
141: Most of the properties of the $1D$ SHO can be obtained analytically.
142: For instance, the acceptance ratio, which is the number of accepted
143: trials over the total number of trials can be expressed as
144: \begin{equation}\label{eq:5}
145: P_{acc}(\delta)=c\int_{-\infty}^{+\infty} dxe^{-\beta V(x)} \frac{1}{2\delta}\int_{-\delta}^{\delta}dhW(x\to
146: x+h).
147: \end{equation}
148: In this equation $\exp(-\beta V(x))dx$ is the probability that the
149: oscillator is between $x$ and $x+dx$, $dh/2\delta$ is the probability of
150: selecting a random displacement between $h$ and $h+dh$ and $W(x\to x+h)$
151: is the probability of accepting the trial displacement (given by
152: Eq.~(\ref{eq:4})). Integration over the allowed values of $x$ and
153: $h$ then gives the average acceptance probability. Since
154: displacements to the left and right are symmetric, we need consider
155: only one direction. For displacements to the right, Eq.~(\ref{eq:5}),
156: can be written as
157: \begin{eqnarray}\label{eq:6}
158: \frac{d(\delta P_{acc}(\delta ))}{d\delta }&=c\int_{-\infty }^{+\infty }dx e^{-\beta V(x)} W(x\to x+\delta).
159: \end{eqnarray}
160: For $x<-\delta/2$, $W(x\to x+\delta)=1$ and for $x>-\delta/2$, $W(x\to x+\delta
161: )=e^{-\beta k((x+\delta)^2-x^2)/2}$. One thus obtains
162: \begin{eqnarray}\label{eq:7}
163: \frac{d(\xi P_{acc}(\xi ))}{d\xi }&=1-erf\left(\frac{\xi}{2} \right)
164: \end{eqnarray}
165: where $erf(x)$ is the error function and $\xi=\sqrt{\frac{\beta k}{2}}\delta $. Using the
166: initial condition, i.e., $P_{acc}(0)=1$, the solution of the
167: differential equation~(\ref{eq:7}) is
168: \begin{equation}
169: P_{acc}(\xi )=1-erf\left(\frac{\xi}{2} \right)+\frac{2}{\sqrt{\pi}\xi }\left(1-e^{-\xi^2/4}\right).
170: \end{equation}
171: The function is plotted in Fig.~\ref{fig:1}.
172:
173: \begin{figure}
174: \begin{center}
175:
176: \resizebox{9cm}{!}{\includegraphics{meanrms.ps}}
177: \caption{Mean squared displacement $<(\Delta r)^2>\frac{\beta k}{2}$
178: versus the acceptance ratio $P_{acc}$ for uniform volume sampling in
179: one, two and three dimensions (full curve, dashed, and dotted lines,
180: respectively). For a uniform radius sampling all curves coincide with
181: the one-dimensional result (full line). }\label{fig:2}
182: \end{center}
183: \end{figure}
184:
185:
186: \begin{figure}
187: \begin{center}
188: \resizebox{9cm}{!}{\includegraphics{meana.ps}}
189:
190: \caption{Mean absolute displacement $<|\Delta r|>\sqrt{\frac{\beta k}{2}}$
191: versus the acceptance ratio $P_{acc}$ for a uniform volume sampling in
192: one, two and three dimensions (full curve, dashed, and dotted lines,
193: respectively). For a uniform radius sampling all curves coincide with
194: the one-dimensional result (full line).}\label{fig:3}
195: \end{center}
196: \end{figure}
197: The mean squared displacement, $<(\Delta x)^2>$, and the mean absolute
198: displacement, $<|\Delta x|>$, defined as
199: \begin{eqnarray}
200: <(\Delta x)^2>&=c\int_{-\infty}^{+\infty} dx\exp(-\beta V(x))
201: \frac{1}{2\delta}\int_{-\delta}^{\delta}dhh^2W(x\to x+h),\\
202: <|\Delta x|>&=c\int_{-\infty}^{+\infty} dx\exp(-\beta V(x))
203: \frac{1}{2\delta}\int_{-\delta}^{\delta}dh|h|W(x\to x+h).
204: \end{eqnarray}
205: can be quite simply obtained from the generating function
206: \begin{equation}\label{eq:30}
207: Z(\lambda)=c\int_{-\infty}^{+\infty} dx\exp(-\beta V(x))
208: \frac{1}{2\delta}\int_{-\delta}^{\delta}dh\exp(-\lambda|h|)W(x\to x+h),
209: \end{equation}
210: by the derivatives with respect to $\lambda$ and set $\lambda=0$, i.e., $<|\Delta
211: x|>=-(\partial Z(\lambda)/\partial\lambda)_{\lambda=0}$, $<(\Delta x)^2>=(\partial^2Z(\lambda)/\partial\lambda^2)_{\lambda=0}$ (and, of
212: course, $P_{acc}=Z(\lambda=0)$).
213:
214:
215: By multiplying both sides of Eq.~(\ref{eq:30}) by $\xi$ and next
216: differentiating with respect to $\xi$, one obtains
217:
218: \begin{equation}\label{eq:34}
219: \frac{\partial(\xi Z(\lambda, \xi))}{\partial\xi}=\exp\left(-\sqrt{\frac{2}{\beta k}}\lambda \xi\right)\left(1-erf\left(\frac{\xi}{2}\right)\right)
220: \end{equation}
221: which, after defining $\tilde{\lambda}=\sqrt{\frac{2}{\beta k}}\lambda$, leads to
222: \begin{equation}\label{eq:8}
223: \fl Z(\lambda, \xi)=\frac{1}{\tilde{\lambda}\xi }\left[1- \exp(-{
224: \tilde{\lambda}\xi})\left(1- erf\left(\frac{\xi}{2}
225: \right)\right)-\exp(\tilde{\lambda}^2)\left(
226: erf\left(\frac{\xi}{2}+\tilde{\lambda}\right)-erf(\tilde{\lambda})\right)\right]
227: \end{equation}
228:
229: By taking the derivatives of the above formula with respect to $\lambda$ and
230: evaluating the resulting expressions at $\lambda=0$, it is easy to show that
231: the mean squared and mean absolute displacements are given by
232:
233: \begin{equation}\label{eq:33}
234: \fl \sqrt{\frac{\beta k}{2}} <|\Delta x|>=\left[\frac{\xi}{2}\left(1- erf\left(\frac{\xi}{2} \right)\right)
235: -\frac{1}{\sqrt{\pi} }\exp\left(-\frac{\xi^2}{4}\right)+\frac{erf\left(\frac{\xi}{2}
236: \right)}{\xi}\right],
237: \end{equation}
238: \begin{equation}\label{eq:32}
239: \fl\frac{\beta k}{2} <(\Delta x)^2>=\frac{1}{3}\left[\xi^2\left(1- erf\left(\frac{\xi}{2} \right)\right)
240: +\frac{8}{\sqrt{\pi}\xi
241: }\left(1-\left(1+\frac{\xi^2}{4}\right)\exp\left(-\frac{\xi^2}{4}
242: \right)\right)\right].
243: \end{equation}
244:
245:
246: The maximum in the mean squared and mean absolute displacements occur
247: for $\xi=2.61648$ and $\xi=1.76332$, which corresponds to acceptance ratio
248: values of $P_{acc}=0.41767$ and $P_{acc}=0.558239$ in agreement with
249: the numerical results of BKS\cite{BKS92}: see Figs.~\ref{fig:2} and
250: ~\ref{fig:3}.
251:
252:
253:
254:
255: \section{D dimensions}
256: We show here that the acceptance ratio, mean squared displacement and
257: other quantities of interest can be obtained exactly in any
258: dimension. Note that the term ``volume'' should be interpreted as the
259: hypervolume in $D$ dimensions, e.g., area in 2D and volume in 3D.
260: For simplicity, we derive exact expressions in odd
261: dimensions, but similar results can be obtained in even
262: dimensions.
263: \subsection{Uniform volume sampling}
264: In $D$ dimensions, the acceptance ratio is expressed as
265: \begin{equation}\label{eq:9}
266: \fl P_{acc}(\delta)=\frac{c_D}{\delta^DV_D}\int
267: d^D{\bf r}\exp(-\beta k {r}^2/2) \int_{|h|\leq\delta}d^D{\bf h} \min(1,\exp(-
268: \beta k(|{\bf r} +{\bf h}|^2-{r}^2)/2))
269: \end{equation}
270: where $V_D=\pi^{D/2}/\Gamma(D/2+1)$ is volume of the sphere of unit radius in $D$
271: dimensions and
272: \begin{eqnarray}
273: c_D&=\left(DV_D\int_0^\infty dr r^{D-1}e^{-\beta kr^2/2}\right)^{-1}=
274: \left(\frac{\beta k}{2\pi}\right)^{D/2}.
275: \end{eqnarray}
276:
277: In odd dimensions, the derivative of Eq.~(\ref{eq:9}) with respect to
278: $\delta$ can be written explicitly by using generalized spherical
279: coordinates
280: \begin{eqnarray}\label{eq:10}
281: \frac{d(\delta^D P_{acc}(\delta))}{d\delta} &=\frac{c_D}{ V_D} \int d^D{\bf r}\exp(-\beta k {r}^2/2) \nonumber\\
282: &\delta^{D-1} \int d\Omega \min(1,\exp(- \beta k(r\delta\cos\phi_1+\delta^2/2))),
283: \end{eqnarray}
284: where $d\Omega =(\prod_{j=1}^{D-2}(\sin(\phi_j))^{D-1-j}d\phi_j)d\phi_{D-1}$ such that
285: $\int d\Omega=DV_D$. The first $D-2$ variables $\phi_j$ are integrated from $0$
286: to $\pi$, whereas $\phi_{D-1}$ is integrated from $0$ to $2\pi$. If we
287: denote $u=\cos(\phi_1)$, perform integration over $\phi_2\ldots\phi_{D-1}$,
288: and introduce the variable $v=r\sqrt{\frac{\beta k}{2}} $
289: Eq. (\ref{eq:10}) can be rewritten as
290: \begin{eqnarray}\label{eq:11}
291: \fl\frac{d(\xi^D P_{acc}(\xi))}{d\xi }&= D\xi^{D-1} \frac{2}{\Gamma\left(\frac{D}{2}\right)\int_{-1}^1du (1-u^2)^{(D-3)/2}}\times \nonumber\\
292: \lo \times & \int_0^{+\infty } v^{D-1}
293: dv e^{-v^2}
294: \int_{-1}^1du (1-u^2)^{(D-3)/2} \min(1,\exp(- (2\xi vu+ \xi^2)))
295: \end{eqnarray}
296: where $\min(1,\exp(-(2\xi vu+ \xi^2)))=\exp((-(2\xi vu+ \xi^2))$
297: for $v<\xi /2$ with $-1<u<1$ and for $v>\xi /2$ with $ -\xi /(2v)<u<1$ and $
298: \min(1,\exp(- (2\xi vu+ \xi^2)))=1$ for $r>\xi /2$ with $-1<u<
299: -\xi /(2v)$. Using that $\int_{-1}^1du (1-u^2)^{(D-3)/2}=\frac{\Gamma \left(\frac{D-1}{2}\right)}{\Gamma \left(\frac{D}{2}
300: \right)}\sqrt \pi$, Eq.~(\ref{eq:11}) then becomes
301: \begin{eqnarray}\label{eq:12}
302: \fl\frac{d(\xi^D P_{acc}(\xi))}{d\xi }=&\frac{2D\xi^{D-1}}{\sqrt{\pi }\Gamma\left(\frac{(D-1)}{2}\right)}\left( \int_0^{\xi /2} dv v^{D-1}
303: e^{- v^2} \int_{-1}^1du (1-u^2)^{(D-3)/2} \exp(- ( \xi^2+2\xi
304: vu))\right.\nonumber\\
305: \lo+&\int_{\xi /2}^{+\infty} v^{D-1} dv e^{- v^2}\left[\int_{-1}^{-\xi /(2v)}du
306: (1-u^2)^{(D-3)/2} \right.+\nonumber\\
307: \lo +&\left.\left.\int_{-\xi /(2v)}^1du (1-u^2)^{1/2}\exp((-( \xi^2+2\xi vu)))
308: \right]\right).
309: \end{eqnarray}
310:
311: After some calculation (see Appendix A) one obtains,
312: \begin{eqnarray}\label{eq:13}
313: \frac{d(\xi^D P_{acc}(\xi))}{d\xi }&=D\xi^{D-1}\left(1-erf\left(\frac{\xi}{2}\right)\right)
314: \end{eqnarray}
315: which gives, for instance, in three dimensions
316: \begin{equation}\label{eq:37}
317: P_{acc}(\xi)= 1-erf\left(\frac{\xi}{2}\right) + \frac
318: {8}{\sqrt{\pi}\xi^3}\left(1-\left(1+\frac{\xi^2}{4}\right)
319: \exp\left(-\frac{\xi^2}{4}\right)\right)
320: \end{equation}
321: Figure \ref{fig:1} shows the acceptance ratio $P_{acc}$ versus $\delta$ in one, two
322: and three dimensions.
323:
324: A similar calculation for generating function $Z_D(\lambda,\xi)$ leads to
325: \begin{equation}\label{eq:36}
326: \frac{\partial(\xi^D Z_D(\tilde{\lambda},\xi))}{\partial\xi }=D\xi^{D-1}e^{-\tilde{\lambda}\xi}\left(1-erf\left(\frac{\xi}{2}\right)\right),
327: \end{equation}
328: and to
329: \begin{equation}\label{eq:35}
330: Z_D(\tilde{\lambda},\xi)=D(-\xi)^{1-D}\frac{\partial^{D-1}}{\partial\tilde{\lambda}^{D-1}}Z_{D=1}(\tilde{\lambda},\xi),
331: \end{equation}
332: where the expression for $Z_{D=1}(\tilde{\lambda},\xi)$ is given in
333: Eq.~(\ref{eq:8}). Since $P_{acc}(\xi)=Z_{D}(\tilde{\lambda}=0,\xi)$, it follows
334: from Eq.~(\ref{eq:35}) that the acceptance ratio in $D$ dimensions is
335: equal, up a factor $D(-\xi)^{1-D}\sqrt{\frac{\beta k}{2}}$, to the mean
336: squared displacement $<(\Delta x)^2>$ in one dimension: compare
337: Eq.~(\ref{eq:37}) to Eq.~(\ref{eq:32}).
338:
339: Although straightforward, the algebra rapidly becomes tedious, and we
340: only illustrate the results by giving the expression of the mean
341: squared displacement $<(\Delta r)^2>$ in three dimensions
342: \begin{eqnarray}
343: \fl\frac{\beta k}{2}\ <(\Delta r)^2>&=\frac{\partial^2}{\partial\tilde{\lambda}^2}Z_{D=3}(\tilde{\lambda},\xi)|_{\tilde{\lambda}=0}
344: \nonumber\\&= \frac{3}{\xi^2 }\left(\frac{\beta k}{2}\right)^2<(\Delta x)^4>_{D=1}\nonumber\\
345: &=\frac{12}{5
346: }\left[\frac{\xi^2}{4}\left(1-erf\left(\frac{\xi}{2}\right)\right)
347: +\frac{16}{\sqrt{\pi}\xi^2 }\left(1-\left(1+\frac{\xi^2}{4}
348: +\frac{\xi^4}{32} \right)e^{\left(-\frac{\xi^2}{4}\right)}\right)\right].
349: \end{eqnarray}
350: The mean squared and mean absolute displacement are plotted versus the
351: acceptance ratio $P_{acc}$ in one, two and three dimensions in
352: Figs.~\ref{fig:2} and ~\ref{fig:3}. Note that the maximum is shifted
353: to the left, i.e., to the smallest values of the acceptance ratio,
354: when the space dimension increases.
355: \subsection{Uniform radius sampling}
356: The acceptance ratio $P_{acc,w}$ in $D$ dimensions can be
357: expressed as
358: \begin{eqnarray}\label{eq:15}
359: P_{acc,w}(\delta)&=c_D\int_D d^D{\bf r}\exp(-\beta k {\bf r}^2/2) \nonumber\\
360: &\int_{|h|\leq\delta}d^D{\bf h} P_w(h)\min(1,\exp(- \beta k(|{\bf r} +{\bf
361: h}|^2-{r}^2)/2))
362: \end{eqnarray}
363: where $P_w(h)$ is the weighted probability. For a uniform
364: distribution in radius, $h^{D-1}P_w(h)=(DV_D\delta)^{-1}$. Using the method
365: developed in the above section, it is straightforward to obtain that
366: \begin{equation}\label{eq:16}
367: \frac{d(\xi P_{acc,w}(\delta))}{d\xi }=\left(1-erf\left(\frac{\xi}{2}\right)\right)
368: \end{equation}
369: which shows that the acceptance ratio is the same whatever the
370: dimension and explains the data collapse observed in \cite{BKS92}.
371: Similarly, the generating function $Z(\lambda,\xi)$ can be shown to obey to
372: the differential equation
373: \begin{equation}
374: \frac{\partial Z(\lambda,\xi)}{\partial\xi }=\exp(-\sqrt{\frac{2}{\beta k}}\lambda \xi) \left(1-erf\left(\frac
375: \xi 2\right )\right),
376: \end{equation}
377: independently of the dimension $D$. This proves that $Z(\lambda,\xi)$ and all
378: moments such as $<|\Delta r|>$ and $<(\Delta r)^2>$ are independent of
379: dimension, as numerically found by BKS\cite{BKS92}.
380:
381:
382:
383:
384:
385: \section{Dynamic behavior}
386: In addition to the exact results for the static properties presented
387: above, we have investigated the dynamic behavior of the SHO by using
388: numerical and analytical approaches. For simplicity, we discuss only
389: the unidimensional case, but the approach can be generalized to $D$
390: dimensions.
391:
392: The master equation describing the dynamical evolution of the system
393: during the Monte Carlo simulation, Eq.~(\ref{eq:1}), can be formally
394: written
395: \begin{equation}
396: \frac{d{\bf P}(t)}{dt}=-L{\bf P}(t)
397: \end{equation}
398: where $L$ is a linear operator acting on $P$ and the Metropolis
399: rule, Eq~(\ref{eq:4}), is used for the transition rate. We consider the
400: spectrum of eigenvalues $\lambda$ of $L$. Denoting $P_\lambda(x)$ the
401: eigenfunction associated with $\lambda$ and introducing $f_\lambda (x)$ via
402: $P_\lambda(x)=P_{eq}(x)f_\lambda (x)$, where $P_{eq}(x)$ is given in
403: Eq.~(\ref{eq:3}), one can express the eigenvalue equation
404: \begin{equation}
405: \lambda P_{eq}(x)f_\lambda (x)=L( P_{eq}(x)f_\lambda (x))
406: \end{equation}
407: as
408: \begin{equation}\label{eq:17}
409: \lambda f_\lambda (x)=\frac{1}{2\delta}\int_{-\delta}^{\delta}dh Min(1,e^{-\beta (V(x+h)-V(x))})(f_\lambda (x)-f_\lambda (x+h)).
410: \end{equation}
411:
412: Multiplying both sides by $\exp(-\beta V(x))f^*_\lambda(x)$, where the star
413: denotes a complex conjugate, and integrating over $x$ then gives
414:
415: \begin{equation}\label{eq:18}
416: \lambda =\frac{1}{4\delta}\frac{\int_{-\infty}^{+\infty}dx\int_{-\delta}^{\delta }dh
417: Min(e^{-\beta V(x)},e^{-\beta V(x+h)})|f_\lambda (x+h)-f_\lambda(x)|^2}{
418: \int_{-\infty}^{+\infty}dx e^{-\beta V(x)}|f_\lambda (x)|^2 }.
419: \end{equation}
420:
421: As anticipated for a Markov process satisfying detailed balance, one
422: deduces from the above formula that all eigenvalues are real and
423: positive; the smallest eigenvalue is $\lambda_0=0$ and it is associated with
424: $f_0(x)=constant\neq 0$. One need consider only real
425: eigenfunctions. Moreover, the eigenvalues can be sorted according
426: to the symmetry of the associated eigenfunctions: it is easy to check
427: that the eigenfunctions are either even or odd functions of $x$, due
428: to the fact that the potential $V(x)$ is an even function of $x$.
429:
430: Any solution of the master equation can be expanded as
431: \begin{equation}\label{eq:19}
432: P(x,t)=P_{eq}(x)(1+\sum_{\lambda >0 }c_\lambda f_\lambda (x)e^{-\lambda t}),
433: \end{equation}
434: and a similar expansion applies to the conditional probability
435: $P(x,t|x_0,0)$ from which one can compute any time-dependent
436: correlation function. The long-time kinetics governing the approach
437: to equilibrium in $P(x,t)$ and in any correlation function is
438: characterized by the smallest non-zero eigenvalue for which the
439: amplitude, i.e., the projection of $P(x,0)$, or of the dynamic
440: observable, onto the relevant eigenfunction, does not vanish.
441:
442: Since, according to Eq.~(\ref{eq:18}), the eigenvalues are expressed
443: as the ratio of two positive quadratic functionals (that in the
444: denominator being also definite), one can use the Rayleigh-Ritz
445: procedure to find a variational upper-bound for the
446: eigenvalues\cite{D62,A85}. Consider first the smallest non-zero
447: eigenvalue $\lambda_1$. For any real function $\phi(x)$ which is both
448: normalized and orthogonal to $f_0(x)$, i.e., satisfies
449: \begin{eqnarray}
450: \int_{-\infty}^\infty P_{eq}(x)\phi (x)^2dx&=1\\
451: \int_{-\infty}^\infty P_{eq}(x)\phi (x)dx&=0,
452: \end{eqnarray}
453: one has the inequality
454: \begin{equation}
455: \fl \lambda_1\leq \lambda_1[\phi]=\frac{1}{4\delta}\int_{-\infty}^{+\infty}dx\int_{-\delta}^{\delta }dh
456: Min(e^{-\beta V(x)},e^{-\beta V(x+h)})|\phi (x+h)-\phi (x)|^2.
457: \end{equation}
458:
459: A convenient choice of trial functions is provided by linear
460: combinations of Hermite polynomials $H_n(\xi )$ (where, as in the
461: previous sections $\xi=\sqrt{\frac{\beta k}{2}}\delta$), with $n> 0$, since these
462: form, up to a trivial multiplicative factors, an orthonormal basis
463: with respect to the weight function $\exp(-\beta V(x))$ with
464: $V(x)=(1/2)kx^2$. For $\lambda_1$, which is associated with an odd
465: eigenfunction, one need consider only the odd polynomials $H_{2n+1}(\xi
466: )$, $n\geq 0$.
467:
468: The simplest estimate of $\lambda_1$ is provided by taking
469: \begin{equation}
470: \phi(\xi )=\frac{H_1(\xi)}{\sqrt{2\sqrt{\pi}}}=\sqrt{\frac{2}{\sqrt{\pi}}}\xi ,
471: \end{equation}
472: which gives
473: \begin{equation}\label{eq:20}
474: \lambda_1[\phi]=\frac{4}{3}
475: \left(\frac{\xi^2}{4}\left(1-erf\left(\frac{\xi}2\right)\right)\right)
476: +\frac{2}{\sqrt{\pi}\xi}(1-\left(1+\frac{\xi^2}{4}\exp\left(\frac{-\xi^2} 4\right)\right)
477: \end{equation}
478: With this
479: choice of $\phi(\xi )$, $ \lambda_1[\phi]$ simply reduces to the mean squared
480: displacement $<(\Delta x)^2>$ multiplied by $\left(\frac{\beta k}{2}\right)$
481: (see Eq.~(\ref{eq:8})). One then derives from section 2 that $
482: \lambda_1[\phi]$ versus $\xi$ passes though a maximum for $\xi\simeq 2.611648$, which
483: corresponds to an acceptance ratio of $P_{acc}=0.41767$.
484:
485: A better estimate of $\lambda_1$ can be obtained by using a linear
486: combination of $H_1(\xi )$ and $H_3(\xi )$:
487: \begin{equation}
488: \phi(\xi ;\theta)=\left[\frac{\cos(\theta)}{\sqrt{2\sqrt{\pi}}}H_1(\xi )+
489: \frac{\sin(\theta)}{\sqrt{48\sqrt{\pi}}}H_3(\xi )\right]
490: \end{equation}
491: where only one independent parameter $\theta$ appears due to the
492: normalization condition. The best bound is determined by minimizing
493: the expression $ \lambda_1[\phi]$ with respect to $\theta$:
494: \begin{equation}\label{eq:21}
495: \frac{\partial\lambda_1[\phi]}{\partial\theta}=0.
496: \end{equation}
497:
498:
499:
500: \begin{figure}
501: \begin{center}
502: \resizebox{10cm}{!}{\includegraphics{exact.ps}}
503:
504: \caption{$\lambda_1$ versus $\xi $ . The full curve was obtained by
505: numerical diagonalization of the master equation. The dashed curve
506: corresponds to the zeroth-order estimate, Eq.~(\ref{eq:20}), the
507: dotted curve corresponds to the solution of the first-order trial
508: function, Eq.~(\ref{eq:21})), and the dash-dot curve corresponds to the
509: exact asymptotic behavior, Eq.~(\ref{eq:31}). }\label{fig:4}
510: \end{center}
511: \end{figure}
512:
513:
514: The result is a lengthly algebraic formula that is plotted in
515: Figs.~\ref{fig:4} and \ref{fig:5}, together with the expression in
516: Eq.~(\ref{eq:20}).
517:
518: An improved estimate of $\lambda_1$ can be derived by noting that at
519: large $\xi $, $\lambda_1$ is inversely proportional to $\xi $. Actually, one can
520: show that this is true for all eigenvalues except $\lambda_0=0$. By
521: considering Eq.~(\ref{eq:17}) in the limit where $x$ goes to zero, one
522: arrives at the result (see Appendix B)
523: \begin{equation}\label{eq:31}
524: \lambda(\xi)\sim \frac{\sqrt{\pi}}{2\xi }+ 0(e^{-\xi^2}),
525: \end{equation}
526: valid for large $\xi $. Note that the correction terms are very small as
527: soon as $\xi\geq 3$. One can build an estimate of $\lambda_1(\xi)$ by using the
528: piecewise function that is equal to $ \frac{\sqrt{\pi}}{2\xi }$ for $\xi\geq
529: \xi^*$ and is equal to $ \lambda_1[\phi]$ obtained for a linear combination of
530: $H_1$ and $H_3$ (see above) for $\xi \leq \xi^*$, where $ \xi^*$ is the value
531: at which $\lambda_1[\phi]= \frac{\sqrt{\pi}}{2\xi }$. $\lambda(\xi)$ is then maximum for
532: $\xi=\xi^*\simeq 2.35$; the corresponding value of the acceptance ratio is
533: $P_{acc}\simeq 0.56$. The estimate is shown in Figs.~\ref{fig:4} and
534: \ref{fig:5}.
535:
536:
537:
538:
539:
540: \begin{figure}
541: \begin{center}
542: \resizebox{10cm}{!}{\includegraphics{exact2.ps}}
543:
544: \caption{Same as Figure \ref{fig:4} except $\xi \lambda$ versus $\xi$.}\label{fig:5}
545: \end{center}
546: \end{figure}
547: In order to compare our results to the BKS paper \cite{BKS92}, it is
548: necessary to calculate the second eigenvalue $\lambda_2$. The trial function
549: is then chosen in a subspace orthogonal not only to $f_0(\xi )=constant$
550: but also to the eigenfunction associated with $\lambda_1$. A convenient
551: choice is provided by (normalized) linear combinations of the even
552: Hermite polynomials $H_{2n}(\xi )$ with $n\geq 1$. An
553: estimate of $\lambda_2$ can be obtained by using a linear combination of
554: $H_2(\xi )$ and $H_4(\xi )$:
555: \begin{equation}\label{eq:22}
556: \phi(\xi ;\theta)=\left[\frac{\cos(\theta)}{2\sqrt{2\sqrt{\pi}}}H_2(\xi )+
557: \frac{\sin(\theta)}{8\sqrt{6\sqrt{\pi}}}H_4(\xi )\right]
558: \end{equation}
559: and the result is shown in Figs.~\ref{fig:6} and ~\ref{fig:7},
560: together with the zeroth-order approximation obtained with only
561: $H_2(\xi)$ and the improved estimate taking into
562: account the large-$\xi$ behavior.
563:
564: \begin{figure}
565: \begin{center}
566: \resizebox{10cm}{!}{\includegraphics{exact3.ps}}
567:
568: \caption{$\lambda_2$ versus $\xi $ . The full curve was obtained by
569: numerical diagonalization of the master equation. The dashed curve
570: corresponds to the zeroth order estimate, ($\theta=0$ in Eq.~(\ref{eq:22})),
571: the dotted curve correspond to the solution of the first order trial
572: function, Eq.~(\ref{eq:22}) and the dash-dot curve corresponds to the
573: exact asymptotic behavior, Eq.~(\ref{eq:31}). }\label{fig:6}
574: \end{center}
575: \end{figure}
576:
577: \begin{figure}
578: \begin{center}
579: \resizebox{10cm}{!}{\includegraphics{exact4.ps}}
580:
581: \caption{Same as Figure \ref{fig:6} except $\xi \lambda_2$ versus $\xi$}\label{fig:7}
582: \end{center}
583: \end{figure}
584:
585: In addition to the above analytical estimates, we have also performed
586: a numerical study of the spectrum of eigenvalues of the master
587: equation, Eq.~(\ref{eq:1}). The latter has been discretized in
588: $x$-space by taking a constant step size $\Delta$, which leads to a matrix
589: form,
590:
591: \begin{eqnarray}
592: \frac{d{\bf P}(t)}{dt}=-{\bf W}{\bf P}(t)
593: \end{eqnarray}
594: where ${\bf P}$ is a vector with components $ P_i(t)=P(x_i,t)$ and the
595: elements $W_{ij}$ of the matrix ${\bf W}$ are such that
596: \begin{eqnarray}
597: W_{ii}&= \Delta \sum_{j=i-Nh,j\neq i}^{i+Nh}W(i\to j)\nonumber
598: \\W_{ij}&= -\Delta W(j\to i)
599: \end{eqnarray}
600: and $W(i\to j)=Min(1,e^{-\beta(V(x_j)-V(x_i))})$. The eigenvalues $\lambda$ can
601: then be obtained via an exact numerical diagonalization of the matrix
602: $W$. In practice, convergence is obtained for a unidimensional lattice
603: of $400$ sites, where the $x$-range is $[-10,10]$, and $\xi $ goes from
604: $0$ to $5$. One checks that the lowest eigenvalue is equal to zero and
605: corresponds to the equilibrium state, and that all other eigenvalues
606: are real and strictly positive and behave as $\sqrt{\pi}/(2\xi)$ for large
607: enough $\xi$. The results for the first non-zero eigenvalue $\lambda_1$ and
608: $\lambda_2$ are displayed in Figs.~\ref{fig:4} and ~\ref{fig:6},
609: respectively. One can see that the best analytical estimate described
610: above (linear combination of two Hermite polynomials plus exact
611: asymptotic behavior at large $\xi$), is in excellent agreement with the
612: numerical value in both cases.
613:
614:
615:
616:
617: \begin{figure}
618: \begin{center}
619: \resizebox{10cm}{!}{\includegraphics{invl.ps}}
620:
621: \caption{Inverse of the first nonzero eigenvalues, $1/ \lambda_1$ (dotted
622: curve) and $1/ \lambda_2$ (full curve), versus the acceptance
623: probability $P_{acc}$}\label{fig:8}
624: \end{center}
625: \end{figure}
626:
627:
628: The above analysis allows us to derive by analytical means the
629: simulation result obtained by BKS\cite{BKS92} for the dependence of
630: the characteristic time $\tau$ of the energy-energy correlation function on
631: the acceptance ratio: approximating $\tau$ by $1/ \lambda_2$ (since the energy
632: $V(x)$ is an even function of $x$, its projection on the first
633: eigenfunction associated with $\lambda_1$ vanishes), using for $\lambda_2$ our
634: best analytical estimate, and combining this with the exact result for
635: the acceptance ratio $P_{acc}$ in section 2 lead to the full curve
636: plotted in Fig.~\ref{fig:8}; the time $\tau$ is minimum for the
637: acceptance ratio close to $0.47$, as found by BKS\cite{BKS92} ($\simeq
638: 0.50$).
639:
640: In Fig.~\ref{fig:8}, we have also plotted $1/\lambda_1$ versus the
641: acceptance ratio: it is minimum for $P_{acc}=0.45$. Note that using
642: the results of this and the preceding sections, one can rigorously
643: show that the correlation time $\tau$, no matter how it is precisely
644: defined, diverges as $1/P_{acc}$ when $P_{acc}\to0$ and as
645: $1/(1-P_{acc})$ when $P_{acc}\to1$.
646:
647: \section{Conclusion}
648: We have obtained exact results concerning Metropolis Algorithms for
649: the displacement of a particle in the simple harmonic potential. Our
650: analysis provides a theoretical explanation of the numerical results
651: obtained by BKS\cite{BKS92}. In particular, we show that the results
652: become independent of the space dimension when the successive trial
653: moves are sampled according to a Metropolis algorithm with a uniform
654: distribution in radius (instead of volume). This rationalizes the
655: search for efficient Monte Carlo methods for the simulation of systems
656: with intrinsic inhomogeneity and anisotropy such as biological
657: molecules\cite{BKS92,S02}
658: \appendix
659: \section{Acceptance probability}
660: The derivation of Eq.~(\ref{eq:13}) can be done from Eq.~(\ref{eq:12})
661: after some manipulations whose details are given here. Let us denote
662: $I_D$ and $J_D$ as
663: \begin{equation}
664: I_D= \int_0^{\xi /2} dv v^{D-1}
665: e^{- v^2} \int_{-1}^1du (1-u^2)^{(D-3)/2} \exp(- ( \xi^2+2\xi
666: vu))
667: \end{equation}
668: \begin{eqnarray}
669: J_D&=\int_{\xi /2}^{+\infty} dv v^{D-1} e^{- v^2}\left[\int_{-1}^{-\xi /(2v)}du
670: (1-u^2)^{(D-3)/2} \right.+\nonumber\\
671: &+\left.\int_{-\xi /(2v)}^1du (1-u^2)^{(D-3)/2}\exp((-( \xi^2+2\xi vu)))
672: \right].
673: \end{eqnarray}
674: Changing the variable $u$ to $y=vu$ and $v$ to $t=v^2-y^2$ leads to
675: the following relations
676: \begin{equation}
677: I_D=\int_0^{\xi/2}dy\frac{\exp(-(\xi+y)^2)+\exp(-(y-\xi)^2)}{2}\int_0^{\xi^2/4-y^2}dt\, t^{(D-3)/2}e^{-t}
678: \end{equation}
679: \begin{eqnarray}
680: J_D&=\int_{\xi/2}^{+\infty }dy\frac{\exp(-y^2)}{2}\int_{0}^{+\infty}dt\,t^{(D-3)/2}e^{-t}
681: \nonumber\\
682: &+\int_0^{\xi/2}dy \frac{\exp(-(y-\xi)^2)}{2}\int_{\xi^2/4-y^2}^{+\infty}dt\,t^{(D-3)/2}e^{-t}
683: \nonumber\\
684: &+\int_0^{\xi/2}dy
685: \frac{\exp(-(y+\xi)^2)}{2}\int_{\xi^2/4-y^2}^{+\infty}dt\,t^{(D-3)/2}e^{-t}
686: \nonumber\\
687: &+\int_{\xi/2}^{+\infty }dy
688: \frac{\exp(-(y+\xi)^2)}{2}\int_{0}^{+\infty}dt\,t^{(D-3)/2}e^{-t}
689: \end{eqnarray}
690: Using that $\int_{0}^{+\infty}dt\,t^{(D-3)/2}e^{-t}=\Gamma((D-1)/2)$, one obtains
691: that
692: \begin{equation}\label{eq:14}
693: I_D+J_D=\frac{\sqrt{\pi}}{2}\Gamma\left(\frac{D-1}{2}\right)
694: \left(1-erf\left(\frac{\xi}{2}\right)\right)
695: \end{equation}
696: Inserting Eqs.~(\ref{eq:14}) in Eq.~(\ref{eq:12})
697: leads to Eq.~(\ref{eq:13}).
698:
699: \section{Asymptotic behavior of the eigenvalues for large $\delta$}
700: Consider Eq.~(\ref{eq:17}) in the limit $x\to0$ ; one
701: obtains after rearranging the various terms:
702: \begin{eqnarray}\label{eq:23}
703: \fl\lambda f_\lambda (x)&=\frac{1}{2\delta}\int_{-\delta}^{\delta}dh \exp(\frac{-\beta k}{2} ((x+h)^2-x^2))
704: (f_\lambda (x)-f_\lambda (x+h))\nonumber\\
705: &+\frac{1}{2\delta}\int_{-2x}^{0}dh(1-\exp(\frac{-\beta k}{2} ((x+h)^2-x^2)))(f_\lambda (x)-f_\lambda (x+h)).
706: \end{eqnarray}
707: The second term of the right-hand side of Eq.~(\ref{eq:23}) is at most of
708: order $x^3|f_\lambda (x)|$ and is always negligible so that one can rewrite
709: Eq.~(\ref{eq:23}) as
710: \begin{equation}\label{eq:24}
711: \fl\left(\lambda\delta - \int_{0}^{\delta }dh e^{\frac{-\beta k}{2}h^2}+O(x^2)\right) f_\lambda (x)\simeq -\frac{e^{\frac{-\beta k}{2}x^2}}{2}\int_{-\delta}^{\delta}dh \exp(\frac{-\beta k}{2} (x+h)^2)f_\lambda (x+h).
712: \end{equation}
713: Shifting the variable from $h$ to $x+h$ in the integral of the r.h.s
714: of Eq.~(\ref{eq:24}) and using the orthogonality of $f_\lambda (x)$ to
715: $f_0(x)=constant$ for all non-zero eigenvalues $\lambda$, i.e.,
716: $\int_{-\infty }^{+\infty }dx \exp(\frac{-\beta k}{2} x^2)f_\lambda (x)=0$, leads to the
717: following expression,
718: \begin{equation}\label{eq:25}
719: \fl\left(\lambda\delta - \sqrt{\frac{\pi }{2\beta k}}+ O(x^2)\right) f_\lambda (x)\simeq \frac{1+
720: O(x^2)}{2}\left[\int_{x+\delta}^{+\infty }dh e^{\frac{-\beta kh^2}{2}} f_\lambda (h)+\int^{x-\delta}_{-\infty }dh e^{\frac{-\beta kh^2}{2}} f_\lambda (h)\right]
721: \end{equation}
722: for any non-zero $\lambda$.
723: The r.h.s. of Eq.~(\ref{eq:25}) can be Taylor expanded, which gives
724: \begin{eqnarray}\label{eq:26}
725: \fl \left(\lambda\delta - \sqrt{\frac{\pi }{2\beta k}}+ O(x^2)\right) f_\lambda (x)&\simeq \frac{1}{2}
726: \left[\int_{\delta}^{+\infty }dh e^{\frac{-\beta kh^2}{2}}( f_\lambda (h)+f_\lambda(-h))\right.\nonumber\\
727: &\left. -x
728: e^{\frac{-\beta kh^2}{2}}( f_\lambda (h)-f_\lambda(-h))+ 0(x^2)\right].
729: \end{eqnarray}
730: If the eigenfunctions $f_\lambda$ is an even function of $x$, one then
731: derives that $ f_\lambda (x)=f_\lambda(0)+ O(x^2)$ with $f_\lambda(0)\neq 0$ and
732: \begin{equation}\label{eq:27}
733: \left(\lambda\delta - \sqrt{\frac{\pi }{2\beta k}}\right)=\int_{\delta}^{+\infty }dh
734: \exp(\frac{-\beta kh^2}{2}) \frac{ f_\lambda (h)}{ f_\lambda (0)},
735: \end{equation}
736: which after introducing $\xi=\sqrt{\frac{\beta k}{2}}\delta$ can be rewritten as
737: \begin{equation}\label{eq:28}
738: \lambda= \frac{\sqrt{\pi }}{2\xi }+\sqrt{\frac{2}{\beta k}}\int_{\xi }^{+\infty }dh
739: \exp(\frac{-\beta kh^2}{2}) \frac{ f_\lambda (\sqrt{\frac{2}{\beta k}}h)}{ f_\lambda (0)}.
740: \end{equation}
741: If the eigenfunction is an odd function of $x$, one has that $f_\lambda
742: (x)=f'_\lambda(0)x(1 + O(x^2)$ with $f'_\lambda(0)\neq 0$ and
743: \begin{equation}\label{eq:29}
744: \lambda= \frac{\sqrt{\pi }}{2\xi }
745: \frac{ f_\lambda (\sqrt{\frac{2}{\beta k}}\xi )}{ f_\lambda (0)}\exp(-\xi^2).
746: \end{equation}
747: From Eqs.~(\ref{eq:28}) and (\ref{eq:29}), one immediately obtains
748: that all non-zero eigenvalues behave as
749: \begin{equation}
750: \lambda\sim \frac{\sqrt{\pi }}{2\xi }
751: + O(\exp(-\xi^2))
752: \end{equation}
753: when $\xi \to+\infty$, since $f_\lambda(x)$ diverges more slowly than $e^{x^2}$ when
754: $x\to+\infty$.
755: \section*{References}
756:
757: %\bibliography{mc}
758: \begin{thebibliography}{}
759:
760: \bibitem[Allen and Tildesley, 1987]{AT87}
761: Allen, M.~P. and Tildesley, D.~J. (1987).
762: \newblock {\em Computer Simulation of Liquids}.
763: \newblock Clarendon Press, London.
764:
765: \bibitem[Arfken, 1985]{A85}
766: Arfken, G. (1985).
767: \newblock {\em Mathematical Methods for Physicists, 3rd ed.}
768: \newblock Academic Press, Orlando, FL.
769:
770: \bibitem[Binder, 1997]{B97}
771: Binder, K. (1997).
772: \newblock Applications of monte-carlo methods to statistical physics.
773: \newblock {\em Rep. Prog. Phys.}, 60:487.
774:
775: \bibitem[Bouzida et~al., 1992]{BKS92}
776: Bouzida, D., Kumar, S., and Swendsen, R. (1992).
777: \newblock Efficient monte carlo methods for the computer simulation of
778: biological molecules.
779: \newblock {\em Phys. Rev. A}, 45:8894.
780:
781: \bibitem[Dettman, 1962]{D62}
782: Dettman, J. (1962).
783: \newblock {\em Mathematical Methods in Physics and Engineering}.
784: \newblock Dover, New York.
785:
786: \bibitem[Frenkel and Smit, 2002]{FS02}
787: Frenkel, D. and Smit, B. (2002).
788: \newblock {\em Understanding molecular simulation: From algorithms to
789: applications}.
790: \newblock Academic Press, San Diego.
791:
792: \bibitem[Kolafa, 1988]{K88}
793: Kolafa, J. (1988).
794: \newblock On optimization of monte carlo simulations.
795: \newblock {\em Mol. Phys.}, 63:559.
796:
797: \bibitem[Mountain and Thirumalai, 1994]{MT94}
798: Mountain, R. and Thirumalai, D. (1994).
799: \newblock Quantitative measure of efficiency of monte carlo simulations.
800: \newblock {\em Physica A}, 210:453.
801:
802: \bibitem[Swendsen, 2002]{S02}
803: Swendsen, R. (2002).
804: \newblock {\em private communication}.
805:
806: \end{thebibliography}
807:
808:
809:
810: \end{document}
811:
812: