0803.4397/text.tex
1: \documentclass[twocolumn]{revtex4}
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \usepackage{amssymb}
4: \usepackage{amsmath}
5: \usepackage{graphicx}
6: \usepackage{color}
7: \newcommand{\comment}[1]{\emph{\color{red}#1}}
8: 
9: \begin{document}
10: 
11: \title{Stochastic thermostats: comparison of local and global schemes}
12: \author{Giovanni Bussi}
13: \email{gbussi@ethz.ch}                                 
14: \author{Michele Parrinello}                            
15: \affiliation{Computational Science, Department of Chemistry and Applied Biosciences,                                                   
16: ETH Z\"urich, USI Campus, Via Giuseppe Buffi 13, CH-6900 Lugano, Switzerland}
17: %\date{\today}                                         
18:                                                        
19: \begin{abstract}                                       
20: We show that a recently introduced stochastic thermostat
21: [J.~Chem.~Phys.~{\bf 126}, 014101 (2007)] can be considered as
22: a global version of the Langevin thermostat.           
23: We compare the global scheme and the local one (Langevin)
24: from a formal point of view and through practical calculations
25: on a model Lennard-Jones liquid.                       
26: At variance with the local scheme, the global thermostat preserves
27: the dynamical properties for a wide range of coupling parameters,
28: and allows for a faster sampling of the phase-space.
29: \end{abstract}
30: 
31: \maketitle
32: 
33: The most common approaches to isothermal molecular dynamics
34: are perhaps those based on the introduction of an extended
35: Lagrangian. The root of all these schemes is the Nos\'e{} algorithm
36: \cite{nose84jcp}, often used in the Hoover formulation \cite{hoov85pra}.
37: This scheme can be rigorously shown to provide the correct Boltzmann
38: distribution and has a conserved quantity,
39: which can be used to check the integration timestep.
40: A major drawback of the Nos\'e-Hoover method is that it is
41: not ergodic in some difficult cases, such as harmonic systems.
42: Several different extensions of the Nos\'e-Hoover method
43: have been introduced, the most notable one being the
44: so-called Nos\'e-Hoover chains \cite{mart92jcp},
45: which addresses the ergodicity issue at the price of an increased complexity
46: in the algorithm.
47: Although the Nos\'e-Hoover scheme was originally written as a global thermostat,
48: i.e. coupled only to the
49: total kinetic energy of the system, it is sometimes implemented
50: in a local manner (also called massive Nos\'e-Hoover),
51: i.e. using an independent thermostat on each degree of freedom
52: \cite{tobi+93jpc}.
53: 
54: Another common choice is the weak-coupling method, introduced by
55: Berendsen \emph{et al} \cite{bere+84jcp}. This scheme is a continuous version of
56: the velocity-rescaling scheme, thus it is a global thermostat.
57: It is deterministic, stable and easy to implement,
58: but it does not produce configurations in the
59: canonical ensemble.
60: 
61: An alternative approach to canonical sampling
62: is to use stochastic molecular dynamics.
63: The most common form is Langevin dynamics \cite{schn-stol78prb}.
64: The Langevin thermostat is local, and
65: its major feature is that ergodicity can be
66: proven also in pathological cases.
67: However, since the friction and noise terms alter significantly the
68: Hamiltonian dynamics, it
69: cannot be used to compute dynamical properties, unless an extremely
70: small friction is used.
71: Moreover, the effect of the friction and noise terms on the
72: sampling efficiency is non trivial. Even in applications where dynamical
73: properties are not relevant it can be difficult to properly tune
74: the friction in order to achieve an efficient sampling.
75: 
76: In a recent paper~\cite{buss+07jcp} we proposed a stochastic velocity
77: rescaling which can be considered as Berendsen thermostat plus a stochastic
78: correction leading to canonical sampling. We also
79: showed that, in spite of its stochastic nature,
80: one can define a conserved quantity.
81: This scheme does not suffer of ergodicity problems in solids~\cite{buss+07jcp},
82: has been used in practical applications
83: for equilibration purposes~\cite{dona-gall07prl} or to perform
84: ensemble averages~\cite{brun+07jpcb,bard+08prl}
85: and can be combined with variable-cell dynamics to perform
86: simulations in the isothermal-isobaric ensemble~\cite{zyko+08jcp}.
87: In the present paper we present an alternative derivation of the same scheme,
88: where stochastic velocity rescaling is obtained starting from Langevin
89: dynamics and minimizing the disturbance of the thermostat on the
90: Hamiltonian trajectory, nevertheless retaining the same thermalization
91: speed of Langevin dynamics.
92: This idea was also used by Berendsen \emph{et al} to derive their
93: algorithm \cite{bere+84jcp}.
94: Moreover, we show how stochastic
95: velocity rescaling can be considered as a global version of Langevin
96: dynamics. Thus the relationship between the two schemes is similar to that
97: between standard Nos\'e-Hoover and massive Nos\'e-Hoover.
98: Finally, we compare in practical
99: situations the efficiency of the local (Langevin) and global (rescaling)
100: versions, and show that the disruption of Hamiltonian dynamics
101: observed using Langevin thermostat
102: is not due to the the stochastic nature of the algorithm but to the use
103: of a local thermostat.
104: 
105: \section{Continuous equations of motion}
106: 
107: We consider a system described by coordinates $q_i$ and momenta $p_i$,
108: where $i$ runs over the $N_f$ degrees of freedom,
109: and with $q$ and $p$ we indicate the set of coordinates $q_i$ and $p_i$.
110: We associate a mass $m_i$ to each degree of freedom, and
111: we define a Hamiltonian $H(p,q)=K(p)+U(q)$,
112: where $U(q)$ is the potential energy,
113: and $K(p)=\sum_i \frac{p^2_i}{2m_i}$ is the kinetic energy.
114: We want to sample the canonical distribution
115: $P(p,q)dpdq\propto e^{-\beta (K(p)+U(q))}$,
116: where $\beta$ is the inverse temperature,
117: by means of equations of motion in the form
118: \begin{subequations}
119: \label{eq-langevin}
120: \begin{align}
121: dp_i(t) & = -\frac{\partial U}{\partial q_i}dt + g_i(t)dt \\
122: dq_i(t) & = \frac{p_i(t)}{m_i}dt.
123: \end{align}
124: \end{subequations}
125: Equations~\eqref{eq-langevin}
126: are Hamilton equations
127: plus a correction force $g_i(t)$ which artificially modifies the
128: dynamics of the system.
129: Since the total energy $H$ is conserved in Hamiltonian dynamics,
130: only $g_i(t)$ is responsible for its variations
131: and leads to the system thermalization.
132: 
133: In standard Langevin dynamics, the correction force is
134: \begin{equation}
135: \label{eq-langevin-force}
136: g_i(t)dt=- \gamma p_i(t)dt + \sqrt{\frac{2m_i\gamma}{\beta}}dW_i(t),
137: \end{equation}
138: where $\gamma$ is the friction coefficient,
139: and $dW_i(t)$ is a vector of $N_f$ independent Wiener noises,
140: normalized as $\langle \frac{dW_i(t)}{dt} \frac{dW_j(t')}{dt}\rangle = \delta(t-t')\delta_{ij}$.
141: The thermalization speed can be quantified calculating the
142: time derivative of the total energy from Eqs.~\eqref{eq-langevin}
143: and~\eqref{eq-langevin-force}.
144: Using the Itoh chain rule \cite{gard03book} we obtain
145: \begin{multline}
146: \label{eq-dh}
147: dH(t)=
148: \sum_i\left(-\frac{\gamma p_i^2(t)}{m_i}dt
149: + \sqrt{\frac{2\gamma }{\beta}\frac{p_i^2(t)}{m_i}}dW_i(t)\right) \\
150: + \frac{\gamma N_f}{\beta}dt.
151: \end{multline}
152: This expression can be further simplified defining
153: the average kinetic energy $\bar{K}=N_f(2\beta)^{-1}$,
154: a relaxation time $\tau=(2\gamma)^{-1}$,
155: and exploiting
156: the fact that the noise terms on different degrees of freedom are independent
157: of each other:
158: \begin{equation}
159: \label{eq-dh2}
160: dH(t)=-\frac{K(t)-\bar{K}}{\tau}dt + 2\sqrt{\frac{\bar{K}K(t)}{N_f\tau}} dW(t).
161: \end{equation}
162: It is worth noting that while in Eqs.~\eqref{eq-langevin} and \eqref{eq-dh} there are
163: $N_f$ independent noise terms, in Eq.~\eqref{eq-dh2} a single noise term is present.
164: 
165: \begin{figure}
166: \begin{center}
167: \includegraphics[clip,width=0.3\textwidth]{fig1.eps}
168: \end{center}
169: \caption{
170: \label{fig-scheme}
171: Schematic representation of the momentum components at time $t$
172: and at time $t+\Delta t$. $g(t)$ is a generic force applied to the system.
173: $\tilde{g}(t)$ is a force which leads to the same change in the kinetic
174: energy as $g(t)$, but minimizes the disturbance.
175: }
176: \end{figure}
177: We now want to design a new correction force $\tilde{g}_i(t)$ which gives
178: the same variation of the total energy as Langevin dynamics,
179: thus the same thermalization speed,
180: but minimizes the disturbance on the trajectory.
181: This procedure is exactly the same used by Berendsen \emph{et al}
182: \cite{bere+84jcp}, the only difference
183: being that Eq.~\eqref{eq-dh2} on the total energy now
184: contains a stochastic term.
185: We first notice that, since the force only acts on the momenta,
186: fixing a value for $H$ is equivalent to fixing a value for $K$.
187: Following Ref.~\cite{bere+84jcp} we quantify
188: the disturbance as $\sum_i m_i^{-1}(\tilde{g}_i(t)dt)^2$.
189: As it is seen in Fig.~\ref{fig-scheme},
190: the minimal disturbance for a fixed kinetic energy increment
191: is obtained
192: with a force $\tilde{g}_i(t)$ which is proportional to $p_i(t)$.
193: Thus $\tilde{g}_i(t)=\lambda(t) p_i(t)$, where $\lambda(t)$
194: is chosen so as to enforce a given variation of the total energy.
195: Since $\lambda(t)$ includes
196: a stochastic part, and since the variation of the total energy depends
197: on the momenta only through the kinetic energy $K$,
198: this last relation can be written as
199: \begin{equation}
200: \label{eq-g-global}
201: \tilde{g}_i(t)dt=p_i(t)\left[A(K(t))dt+B(K(t))dW(t)\right],
202: \end{equation}
203: where $A(K)$ and $B(K)$ are arbitrary functions of the kinetic energy.
204: The change in the total energy is then
205: \begin{equation}
206: \label{eq-dh3}
207: dH=2A(K)K dt+2B(K)KdW+B^2(K)Kdt.
208: \end{equation}
209: Expressions for $A(K)$ and $B(K)$ can be obtained setting
210: Eq.~\eqref{eq-dh2} equal to Eq.~\eqref{eq-dh3},
211: resulting in the correction force
212: \begin{multline}
213: \label{eq-global-force}
214: \tilde{g}_idt=\frac{1}{2\tau}\left[
215: \left(1-\frac{1}{N_f}\right)\frac{\bar{K}}{K}
216: -1
217: \right]p_i dt \\
218: +\sqrt{\frac{\bar{K}}{N_fK\tau}}p_idW.
219: \end{multline}
220: This equation is stochastic, with the same noise term used on all
221: the particles. It is also very similar to the expression of the
222: force in the Berendsen algorithm.
223: 
224: The combination of Eqs.~\eqref{eq-langevin} and~\eqref{eq-global-force}
225: results in a continuous, stochastic dynamics which can be shown to
226: sample exactly the canonical ensemble.
227: The effect of a $\tilde{g}_i$ parallel to $p_i$
228: is the same of a rescaling procedure and
229: the enforced increment of the total energy
230: in Eq.~\eqref{eq-dh2} is the same as in Ref.~\cite{buss+07jcp}.
231: Thus, Eqs.~\eqref{eq-langevin} and~\eqref{eq-global-force}
232: represents the continuous version
233: of the velocity rescaling described in Ref.~\cite{buss+07jcp}.
234: 
235: Notably, if $N_f=1$, Eq.~\eqref{eq-global-force} becomes equivalent to
236: Eq.~\eqref{eq-langevin-force}. Thus when
237: the thermostat is applied to a single degree of freedom,
238: it is completely equivalent to a Langevin thermostat.
239: One can perform Langevin molecular dynamics by applying
240: a thermostat per degree of freedom, or stochastic rescaling
241: by applying a single thermostat to the total kinetic energy.
242: Intermediate schemes can be designed, where a thermostat is applied
243: on each molecule or group of atoms.
244: 
245: \section{Finite timestep algorithm}
246: In the practical implementation,
247: time is incremented in discrete steps, and
248: the Trotter decomposition scheme
249: \cite{tuck+92jcp,buss-parr07pre}
250: can be used to separate the integration of Hamilton equations and
251: the update of the momenta due to $\tilde{g}$.
252: The former is then integrated using standard velocity-Verlet,
253: while for the latter we need to integrate
254: Eq.~\eqref{eq-global-force}.
255: A possible approach is to consider the propagation of
256: kinetic energy when the momenta evolve according
257: to Eq.~\eqref{eq-global-force}, as it is done
258: in the appendix of Ref.~\cite{buss+07jcp}.
259: The analytical solution of Eq.~\eqref{eq-global-force}
260: for a finite time $\Delta t$ is
261: \begin{subequations}
262: \label{eq-propagator}
263: \begin{equation}
264: p_i(t+\Delta t)
265: =
266: \alpha(t) p_i(t),
267: \end{equation}
268: where
269: \begin{multline}
270: \alpha^2(t) = 
271: c+\frac{(1-c)(S_{N_f-1}(t)+R^2(t))\bar{K}}{N_fK(t)} \\
272: + 2R(t)\sqrt{\frac{c(1-c)\bar{K}}{N_fK(t)}}.
273: \end{multline}
274: \end{subequations}
275: Here $c=e^{-2\gamma \Delta t}=e^{-\Delta t /\tau}$, $R(t)$ is a Gaussian number
276: with unitary variance and $S_{N_f-1}$ is the sum of $N_f-1$
277: independent, squared, Gaussian numbers.
278: Equation~\eqref{eq-propagator}
279: has been obtained enforcing the evolution of the kinetic energy, thus,
280: strictly speaking, it does not fix the sign of $\alpha$.
281: A more rigorous analysis shows that the sign of $\alpha$ should be chosen as
282: \begin{equation}
283: \label{eq-alpha-sign}
284: \text{sign} (\alpha(t))=\text{sign}\left(R(t)+\sqrt{\frac{cN_fK(t)}{(1-c)\bar{K}}}\right),
285: \end{equation}
286: to keep into account the finite probability to observe a flip
287: of the momenta $p_i$ when the force in Eq.~\eqref{eq-global-force} is applied.
288: The Gaussian number in Eq.~\eqref{eq-alpha-sign} needs to be the same that
289: is used in Eq.~\eqref{eq-propagator}.
290: The probability to observe the flip is extremely small
291:  if $N_f$ is large
292: and $c\approx 1$, which is the usual case when the thermostat is used as global
293: and $\tau>\Delta t$.
294: This is the case in Ref.~\cite{buss+07jcp}, where we set $\alpha>0$.
295: On the other hand, when the thermostat acts on a few
296: degrees of freedom,
297: the sign of $\alpha$ needs to be
298: calculated by means of Eq.~\eqref{eq-alpha-sign}.
299: This is always the case for Langevin dynamics.
300: With simple manipulation,
301: it can be shown that for $N_f=1$ the integration scheme
302: given by Eqs.~\eqref{eq-propagator}
303: and~\eqref{eq-alpha-sign}, combined with velocity-Verlet,
304: is completely equivalent to the integration
305: scheme for Langevin dynamics introduced in Ref.~\cite{buss-parr07pre}.
306: 
307: \section{Examples}
308: 
309: Up to now we simply established a theoretical relationship between Langevin dynamics
310: and stochastic rescaling. The outcome is that the effect of the two algorithms on the
311: total energy should be equivalent if the friction in the Langevin dynamics
312: and the relaxation time in the stochastic scale are related by
313: $\tau=(2\gamma)^{-1}$. However, the stochastic rescaling is expected to give better
314: dynamical properties, since only the component of the force which changes
315: the total energy is retained. We test here this affirmations on a simple test-case,
316: namely a Lennard-Jones fluid with density $\rho=.8442$ and temperature $\beta^{-1}=0.722$,
317: which is close to the triple point.
318: Throughout this section we use
319: reduced Lennard-Jones units for temperature, distance and time.
320: We simulate a box containing 108 particles,
321: with periodic boundary conditions, and we cut the interaction at
322: distance 2.5. We set the timestep to $\Delta t=0.005$, which leads to
323: a reasonable conservation of the effective energy~\cite{buss+07jcp,buss-parr07pre}.
324: We then perform runs of $10^{7}$ steps, with both the local
325: scheme (Langevin dynamics) and the global one (stochastic rescaling),
326: using a broad range of values of the thermostat
327: relaxation time $\tau$.
328: 
329: \begin{figure}
330: \includegraphics[clip,width=0.45\textwidth]{fig2.eps}
331: \caption{
332: \label{fig-diffusion}
333: Diffusion coefficient as obtained from thermostated simulations
334: as a function of the thermostat relaxation time $\tau$, 
335: for the local and global thermostats as indicated. The statistical error
336: is smaller than the symbol size.
337: As a reference, the diffusion coefficient of a free particle
338: subject to the same Langevin equation is plotted (dashed line).
339: All the quantities are in Lennard-Jones reduced units.
340: }
341: \end{figure}
342: To quantify the disturbance on Hamiltonian dynamics,
343: in Fig.~\ref{fig-diffusion} we show
344: the diffusion coefficient as a function of $\tau$, as
345: obtained from the Einstein relations.
346: When the local thermostat is used with a short
347: relaxation time, the diffusion is strongly quenched.
348: This happens when the typical collision
349: time with the external bath $\gamma^{-1}$ is shorter than the
350: typical collision time between the particles, so that the former
351: becomes the real bottleneck for the diffusion process.
352: In the limit of short $\tau$, the equations
353: of motion tend to a high-friction Langevin dynamics.
354: In this case, the observed diffusion coefficient
355: $D$ is proportional to $(\beta m \gamma)^{-1}$,
356: which is the value of the diffusion coefficient for a free particle
357: subject to the same Langevin dynamics.
358: The prefactor is related to the difficulty in crossing the barriers between
359: different liquid configurations.
360: On the contrary,
361: with the global thermostat $D$ is almost independent on $\tau$,
362: indicating that the disturbance on the dynamics is very small and that
363: good estimates of $D$ in the canonical ensemble
364: can be obtained also with a thermostated simulation.
365: 
366: To quantify the equilibration speed we calculate the autocorrelation
367: time of a few global observables.
368: The efficiency of a sampling algorithm is optimal
369: when the autocorrelation time $\tau_X$ of the desired quantity
370: $X$ is minimal.
371: If $X$ is the total energy,
372: $\tau_X$ also indicates how fast
373: a simulation started from
374: an unlikely configuration is equilibrated.
375: We integrate the autocorrelation function
376: using a windowing function
377: \begin{equation}
378: \label{eq-autocorrelation}
379: \tau_X=
380: \int_0^Tdt \frac{\langle \delta X(0) \delta X(t) \rangle}{\langle \delta X(0) \delta X(0) \rangle}\left(1-\frac{t}{T}\right),
381: \end{equation}
382: where $\delta X=X-\langle X \rangle$ and $T$ is a large value.
383: The windowing function in parethesis helps the convergence
384: because it weights less the points with larger statistical error.
385: Moreover, Equation~\eqref{eq-autocorrelation}
386: is exactly equal to $T\epsilon^2(T)/(2\langle\delta X^2\rangle)$,
387: where $\epsilon^2(T)$ is the mean square error from a time
388: average of length $T$.
389: The relative accuracy in evaluation of $\tau_X$ is approximately
390: $\sqrt{2T/T'}$, where $T'=5\times10^4$ is the total run length.
391: In the following we choose $T=50$, and we expect a relative
392: accuracy on $\tau_X$ on the order of 5\%.
393: Since $T\gg\tau_X$, $\tau_X$ is also a good approximation for
394: the autocorrelation time.
395: 
396: \begin{figure}
397: \includegraphics[clip,width=0.45\textwidth]{fig3.eps}
398: \caption{
399: \label{fig-autocorrelation}
400: Autocorrelation time of the kinetic ($\tau_K$), potential ($\tau_U$),
401: and total ($\tau_H$) energy, as a function of the thermostat relaxation time $\tau$, 
402: for the local and global thermostats as indicated. The statistical error
403: is smaller than the symbol size.
404: All the quantities are in Lennard-Jones reduced units.
405: }
406: \end{figure}
407: 
408: In Figure~\ref{fig-autocorrelation} we plot the autocorrelation time
409: of the kinetic energy $\tau_K$, of the potential energy $\tau_U$
410: and of the total energy $\tau_H$, as a function of the
411: thermostat relaxation time $\tau$.
412: The autocorrelation time of the kinetic energy is
413: completely dictated by the thermostat relaxation time, and independent
414: on the choice of a local or a global scheme. On the contrary,
415: the autocorrelation time of the total energy and of the potential energies
416: are proportional to $\tau$ only in the limit of large $\tau$.
417: In the local scheme, when $\tau$ is smaller than 0.2 the disturbance
418: of the trajectory becomes so large that the phase-space exploration turns out to
419: be slower. Comparing Figs.~\ref{fig-diffusion}
420: and~\ref{fig-autocorrelation}, it is seen that the optimal value for
421: $\tau$ is the smallest one that still does not
422: affect the diffusion coefficient.
423: When the global scheme is adopted, even small values of $\tau$ can be safely
424: used, resulting in a faster decorrelation of the total energy.
425: 
426: \section{Conclusions}
427: 
428: In conclusion, we have presented an alternative derivation of the global
429: thermostat introduced in Ref.~\cite{buss+07jcp}. This derivation
430: allows to write continuous equations of motion and
431: shows the analogy between this scheme and the
432: standard Langevin thermostat. Namely, the new scheme can be considered
433: as a global version of the Langevin thermostat,
434: that minimizes the disturbance of the original Hamiltonian dynamics.
435: Finally, we have discussed these properties on a simple test case,
436: showing that the global scheme preserves the dynamical properties.
437: Moreover, using as a measure the autocorrelation time of the total
438: energy, we have shown that the global scheme allows for a faster
439: sampling.
440: 
441: \begin{thebibliography}{10}
442: 
443: \bibitem{nose84jcp}
444: S.~Nos{\'e},
445: \newblock J.~Chem.~Phys. {\bf 81}, 511 (1984).
446: 
447: \bibitem{hoov85pra}
448: W.~G. Hoover,
449: \newblock Phys.~Rev.~A {\bf 31}, 1695 (1985).
450: 
451: \bibitem{mart92jcp}
452: G.~J. Martyna, M.~L. Klein, and M.~Tuckerman,
453: \newblock J.~Chem.~Phys. {\bf 97}, 2635 (1992).
454: 
455: \bibitem{tobi+93jpc}
456: D.~J. Tobias, G.~J. Martyna, and M.~L. Klein,
457: \newblock J.~Phys.~Chem. {\bf 97}, 12959 (1993).
458: 
459: \bibitem{bere+84jcp}
460: H.~J.~C. Berendsen, J.~P.~M. Postma, W.~F. van Gunsteren, A.~DiNola, and J.~R.
461:   Haak,
462: \newblock J.~Chem.~Phys. {\bf 81}, 3684 (1984).
463: 
464: \bibitem{schn-stol78prb}
465: T.~Schneider and E.~Stoll,
466: \newblock Phys.~Rev.~B {\bf 17}, 1302 (1978).
467: 
468: \bibitem{buss+07jcp}
469: G.~Bussi, D.~Donadio, and M.~Parrinello,
470: \newblock J.~Chem.~Phys. {\bf 126}, 014101 (2007).
471: 
472: \bibitem{dona-gall07prl}
473: D.~Donadio and G.~Galli,
474: \newblock Phys.~Rev.~Lett. {\bf 99}, 255502 (2007).
475: 
476: \bibitem{brun+07jpcb}
477: F.~Bruneval, D.~Donadio, and M.~Parrinello,
478: \newblock J.~Phys.~Chem.~B {\bf 111}, 12219 (2007).
479: 
480: \bibitem{bard+08prl}
481: A.~Barducci, G.~Bussi, and M.~Parrinello,
482: \newblock Phys.~Rev.~Lett. {\bf 100}, 020603 (2008).
483: 
484: \bibitem{zyko+08jcp}
485: T.~Zykova, G.~Bussi, and M.~Parrinello,
486: \newblock In preparation.
487: 
488: \bibitem{gard03book}
489: C.~W. Gardiner,
490: \newblock {\em Handbook of Stochastic Methods},
491: \newblock Springer, Berlin, third edition, 2003.
492: 
493: \bibitem{tuck+92jcp}
494: M.~Tuckerman, B.~J. Berne, and G.~J. Martyna,
495: \newblock J.~Chem.~Phys. {\bf 97}, 1990 (1992).
496: 
497: \bibitem{buss-parr07pre}
498: G.~Bussi and M.~Parrinello,
499: \newblock Phys.~Rev.~E {\bf 75}, 056707 (2007).
500: 
501: \end{thebibliography}
502: \end{document}
503: