ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3298
Committed: Wed Jan 2 21:06:31 2008 UTC (16 years, 7 months ago) by xsun
Content type: application/x-tex
File size: 40320 byte(s)
Log Message:
Add the text about langevin simulation on different shaped molecules.

File Contents

# Content
1 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 %\documentclass[aps,prb,preprint]{revtex4}
3 \documentclass[11pt]{article}
4 \usepackage{endfloat}
5 \usepackage{amsmath,bm}
6 \usepackage{amssymb}
7 \usepackage{times}
8 \usepackage{mathptmx}
9 \usepackage{setspace}
10 \usepackage{tabularx}
11 \usepackage{graphicx}
12 \usepackage{booktabs}
13 \usepackage{bibentry}
14 \usepackage{mathrsfs}
15 \usepackage[ref]{overcite}
16 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18 9.0in \textwidth 6.5in \brokenpenalty=10000
19 \renewcommand{\baselinestretch}{1.2}
20 \renewcommand\citemid{\ } % no comma in optional reference note
21
22 \begin{document}
23
24 \title{An algorithm for performing Langevin dynamics on rigid bodies of arbitrary shape }
25
26 \author{Teng Lin, Xiuquan Sun and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
27 gezelter@nd.edu} \\
28 Department of Chemistry and Biochemistry\\
29 University of Notre Dame\\
30 Notre Dame, Indiana 46556}
31
32 \date{\today}
33
34 \maketitle \doublespacing
35
36 \begin{abstract}
37
38 \end{abstract}
39
40 \newpage
41
42 %\narrowtext
43
44 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45 % BODY OF TEXT
46 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
47
48 \section{Introduction}
49
50 %applications of langevin dynamics
51 As alternative to Newtonian dynamics, Langevin dynamics, which
52 mimics a simple heat bath with stochastic and dissipative forces,
53 has been applied in a variety of studies. The stochastic treatment
54 of the solvent enables us to carry out substantially longer time
55 simulations. Implicit solvent Langevin dynamics simulations of
56 met-enkephalin not only outperform explicit solvent simulations for
57 computational efficiency, but also agrees very well with explicit
58 solvent simulations for dynamical properties.\cite{Shen2002}
59 Recently, applying Langevin dynamics with the UNRES model, Liow and
60 his coworkers suggest that protein folding pathways can be possibly
61 explored within a reasonable amount of time.\cite{Liwo2005} The
62 stochastic nature of the Langevin dynamics also enhances the
63 sampling of the system and increases the probability of crossing
64 energy barriers.\cite{Banerjee2004, Cui2003} Combining Langevin
65 dynamics with Kramers's theory, Klimov and Thirumalai identified
66 free-energy barriers by studying the viscosity dependence of the
67 protein folding rates.\cite{Klimov1997} In order to account for
68 solvent induced interactions missing from implicit solvent model,
69 Kaya incorporated desolvation free energy barrier into implicit
70 coarse-grained solvent model in protein folding/unfolding studies
71 and discovered a higher free energy barrier between the native and
72 denatured states. Because of its stability against noise, Langevin
73 dynamics is very suitable for studying remagnetization processes in
74 various systems.\cite{Palacios1998,Berkov2002,Denisov2003} For
75 instance, the oscillation power spectrum of nanoparticles from
76 Langevin dynamics simulation has the same peak frequencies for
77 different wave vectors, which recovers the property of magnetic
78 excitations in small finite structures.\cite{Berkov2005a}
79
80 %review rigid body dynamics
81 Rigid bodies are frequently involved in the modeling of different
82 areas, from engineering, physics, to chemistry. For example,
83 missiles and vehicle are usually modeled by rigid bodies. The
84 movement of the objects in 3D gaming engine or other physics
85 simulator is governed by the rigid body dynamics. In molecular
86 simulation, rigid body is used to simplify the model in
87 protein-protein docking study{\cite{Gray2003}}.
88
89 It is very important to develop stable and efficient methods to
90 integrate the equations of motion for orientational degrees of
91 freedom. Euler angles are the natural choice to describe the
92 rotational degrees of freedom. However, due to $\frac {1}{sin
93 \theta}$ singularities, the numerical integration of corresponding
94 equations of these motion is very inefficient and inaccurate.
95 Although an alternative integrator using multiple sets of Euler
96 angles can overcome this difficulty\cite{Barojas1973}, the
97 computational penalty and the loss of angular momentum conservation
98 still remain. A singularity-free representation utilizing
99 quaternions was developed by Evans in 1977.\cite{Evans1977}
100 Unfortunately, this approach used a nonseparable Hamiltonian
101 resulting from the quaternion representation, which prevented the
102 symplectic algorithm from being utilized. Another different approach
103 is to apply holonomic constraints to the atoms belonging to the
104 rigid body. Each atom moves independently under the normal forces
105 deriving from potential energy and constraint forces which are used
106 to guarantee the rigidness. However, due to their iterative nature,
107 the SHAKE and Rattle algorithms also converge very slowly when the
108 number of constraints increases.\cite{Ryckaert1977, Andersen1983}
109
110 A break-through in geometric literature suggests that, in order to
111 develop a long-term integration scheme, one should preserve the
112 symplectic structure of the propagator. By introducing a conjugate
113 momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
114 equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
115 proposed to evolve the Hamiltonian system in a constraint manifold
116 by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
117 An alternative method using the quaternion representation was
118 developed by Omelyan.\cite{Omelyan1998} However, both of these
119 methods are iterative and inefficient. In this section, we descibe a
120 symplectic Lie-Poisson integrator for rigid bodies developed by
121 Dullweber and his coworkers\cite{Dullweber1997} in depth.
122
123 %review langevin/browninan dynamics for arbitrarily shaped rigid body
124 Combining Langevin or Brownian dynamics with rigid body dynamics,
125 one can study slow processes in biomolecular systems. Modeling DNA
126 as a chain of rigid beads, which are subject to harmonic potentials
127 as well as excluded volume potentials, Mielke and his coworkers
128 discovered rapid superhelical stress generations from the stochastic
129 simulation of twin supercoiling DNA with response to induced
130 torques.\cite{Mielke2004} Membrane fusion is another key biological
131 process which controls a variety of physiological functions, such as
132 release of neurotransmitters \textit{etc}. A typical fusion event
133 happens on the time scale of a millisecond, which is impractical to
134 study using atomistic models with newtonian mechanics. With the help
135 of coarse-grained rigid body model and stochastic dynamics, the
136 fusion pathways were explored by many
137 researchers.\cite{Noguchi2001,Noguchi2002,Shillcock2005} Due to the
138 difficulty of numerical integration of anisotropic rotation, most of
139 the rigid body models are simply modeled using spheres, cylinders,
140 ellipsoids or other regular shapes in stochastic simulations. In an
141 effort to account for the diffusion anisotropy of arbitrary
142 particles, Fernandes and de la Torre improved the original Brownian
143 dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
144 incorporating a generalized $6\times6$ diffusion tensor and
145 introducing a simple rotation evolution scheme consisting of three
146 consecutive rotations.\cite{Fernandes2002} Unfortunately, unexpected
147 errors and biases are introduced into the system due to the
148 arbitrary order of applying the noncommuting rotation
149 operators.\cite{Beard2003} Based on the observation the momentum
150 relaxation time is much less than the time step, one may ignore the
151 inertia in Brownian dynamics. However, the assumption of zero
152 average acceleration is not always true for cooperative motion which
153 is common in protein motion. An inertial Brownian dynamics (IBD) was
154 proposed to address this issue by adding an inertial correction
155 term.\cite{Beard2000} As a complement to IBD which has a lower bound
156 in time step because of the inertial relaxation time, long-time-step
157 inertial dynamics (LTID) can be used to investigate the inertial
158 behavior of the polymer segments in low friction
159 regime.\cite{Beard2000} LTID can also deal with the rotational
160 dynamics for nonskew bodies without translation-rotation coupling by
161 separating the translation and rotation motion and taking advantage
162 of the analytical solution of hydrodynamics properties. However,
163 typical nonskew bodies like cylinders and ellipsoids are inadequate
164 to represent most complex macromolecule assemblies. These intricate
165 molecules have been represented by a set of beads and their
166 hydrodynamic properties can be calculated using variants on the
167 standard hydrodynamic interaction tensors.
168
169 The goal of the present work is to develop a Langevin dynamics
170 algorithm for arbitrary-shaped rigid particles by integrating the
171 accurate estimation of friction tensor from hydrodynamics theory
172 into the sophisticated rigid body dynamics algorithms.
173
174 \section{Computational Methods{\label{methodSec}}}
175
176 \subsection{\label{introSection:frictionTensor}Friction Tensor}
177 Theoretically, the friction kernel can be determined using the
178 velocity autocorrelation function. However, this approach becomes
179 impractical when the system becomes more and more complicated.
180 Instead, various approaches based on hydrodynamics have been
181 developed to calculate the friction coefficients. In general, the
182 friction tensor $\Xi$ is a $6\times 6$ matrix given by
183 \[
184 \Xi = \left( {\begin{array}{*{20}c}
185 {\Xi _{}^{tt} } & {\Xi _{}^{rt} } \\
186 {\Xi _{}^{tr} } & {\Xi _{}^{rr} } \\
187 \end{array}} \right).
188 \]
189 Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are $3 \times 3$
190 translational friction tensor and rotational resistance (friction)
191 tensor respectively, while ${\Xi^{tr} }$ is translation-rotation
192 coupling tensor and $ {\Xi^{rt} }$ is rotation-translation coupling
193 tensor. When a particle moves in a fluid, it may experience friction
194 force or torque along the opposite direction of the velocity or
195 angular velocity,
196 \[
197 \left( \begin{array}{l}
198 F_R \\
199 \tau _R \\
200 \end{array} \right) = - \left( {\begin{array}{*{20}c}
201 {\Xi ^{tt} } & {\Xi ^{rt} } \\
202 {\Xi ^{tr} } & {\Xi ^{rr} } \\
203 \end{array}} \right)\left( \begin{array}{l}
204 v \\
205 w \\
206 \end{array} \right)
207 \]
208 where $F_r$ is the friction force and $\tau _R$ is the friction
209 torque.
210
211 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
212
213 For a spherical particle with slip boundary conditions, the
214 translational and rotational friction constant can be calculated
215 from Stoke's law,
216 \[
217 \Xi ^{tt} = \left( {\begin{array}{*{20}c}
218 {6\pi \eta R} & 0 & 0 \\
219 0 & {6\pi \eta R} & 0 \\
220 0 & 0 & {6\pi \eta R} \\
221 \end{array}} \right)
222 \]
223 and
224 \[
225 \Xi ^{rr} = \left( {\begin{array}{*{20}c}
226 {8\pi \eta R^3 } & 0 & 0 \\
227 0 & {8\pi \eta R^3 } & 0 \\
228 0 & 0 & {8\pi \eta R^3 } \\
229 \end{array}} \right)
230 \]
231 where $\eta$ is the viscosity of the solvent and $R$ is the
232 hydrodynamic radius.
233
234 Other non-spherical shapes, such as cylinders and ellipsoids, are
235 widely used as references for developing new hydrodynamics theory,
236 because their properties can be calculated exactly. In 1936, Perrin
237 extended Stokes's law to general ellipsoids, also called a triaxial
238 ellipsoid, which is given in Cartesian coordinates
239 by\cite{Perrin1934, Perrin1936}
240 \[
241 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
242 }} = 1
243 \]
244 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
245 due to the complexity of the elliptic integral, only the ellipsoid
246 with the restriction of two axes being equal, \textit{i.e.}
247 prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
248 exactly. Introducing an elliptic integral parameter $S$ for prolate
249 ellipsoids :
250 \[
251 S = \frac{2}{{\sqrt {a^2 - b^2 } }}\ln \frac{{a + \sqrt {a^2 - b^2
252 } }}{b},
253 \]
254 and oblate ellipsoids:
255 \[
256 S = \frac{2}{{\sqrt {b^2 - a^2 } }}arctg\frac{{\sqrt {b^2 - a^2 }
257 }}{a},
258 \]
259 one can write down the translational and rotational resistance
260 tensors
261 \begin{eqnarray*}
262 \Xi _a^{tt} & = & 16\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - b^2 )S - 2a}}. \\
263 \Xi _b^{tt} & = & \Xi _c^{tt} = 32\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - 3b^2 )S +
264 2a}},
265 \end{eqnarray*}
266 and
267 \begin{eqnarray*}
268 \Xi _a^{rr} & = & \frac{{32\pi }}{3}\eta \frac{{(a^2 - b^2 )b^2 }}{{2a - b^2 S}}, \\
269 \Xi _b^{rr} & = & \Xi _c^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^4 - b^4 )}}{{(2a^2 - b^2 )S - 2a}}.
270 \end{eqnarray*}
271
272 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
273
274 Unlike spherical and other simply shaped molecules, there is no
275 analytical solution for the friction tensor for arbitrarily shaped
276 rigid molecules. The ellipsoid of revolution model and general
277 triaxial ellipsoid model have been used to approximate the
278 hydrodynamic properties of rigid bodies. However, since the mapping
279 from all possible ellipsoidal spaces, $r$-space, to all possible
280 combination of rotational diffusion coefficients, $D$-space, is not
281 unique\cite{Wegener1979} as well as the intrinsic coupling between
282 translational and rotational motion of rigid bodies, general
283 ellipsoids are not always suitable for modeling arbitrarily shaped
284 rigid molecules. A number of studies have been devoted to
285 determining the friction tensor for irregularly shaped rigid bodies
286 using more advanced methods where the molecule of interest was
287 modeled by a combinations of spheres\cite{Carrasco1999} and the
288 hydrodynamics properties of the molecule can be calculated using the
289 hydrodynamic interaction tensor. Let us consider a rigid assembly of
290 $N$ beads immersed in a continuous medium. Due to hydrodynamic
291 interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
292 than its unperturbed velocity $v_i$,
293 \[
294 v'_i = v_i - \sum\limits_{j \ne i} {T_{ij} F_j }
295 \]
296 where $F_i$ is the frictional force, and $T_{ij}$ is the
297 hydrodynamic interaction tensor. The friction force of $i$th bead is
298 proportional to its ``net'' velocity
299 \begin{equation}
300 F_i = \zeta _i v_i - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
301 \label{introEquation:tensorExpression}
302 \end{equation}
303 This equation is the basis for deriving the hydrodynamic tensor. In
304 1930, Oseen and Burgers gave a simple solution to
305 Eq.~\ref{introEquation:tensorExpression}
306 \begin{equation}
307 T_{ij} = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
308 R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
309 \end{equation}
310 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
311 A second order expression for element of different size was
312 introduced by Rotne and Prager\cite{Rotne1969} and improved by
313 Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
314 \begin{equation}
315 T_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
316 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
317 _i^2 + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
318 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
319 \label{introEquation:RPTensorNonOverlapped}
320 \end{equation}
321 Both of the Eq.~\ref{introEquation:oseenTensor} and
322 Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption
323 $R_{ij} \ge \sigma _i + \sigma _j$. An alternative expression for
324 overlapping beads with the same radius, $\sigma$, is given by
325 \begin{equation}
326 T_{ij} = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
327 \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
328 \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
329 \label{introEquation:RPTensorOverlapped}
330 \end{equation}
331 To calculate the resistance tensor at an arbitrary origin $O$, we
332 construct a $3N \times 3N$ matrix consisting of $N \times N$
333 $B_{ij}$ blocks
334 \begin{equation}
335 B = \left( {\begin{array}{*{20}c}
336 {B_{11} } & \ldots & {B_{1N} } \\
337 \vdots & \ddots & \vdots \\
338 {B_{N1} } & \cdots & {B_{NN} } \\
339 \end{array}} \right),
340 \end{equation}
341 where $B_{ij}$ is given by
342 \[
343 B_{ij} = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
344 )T_{ij}
345 \]
346 where $\delta _{ij}$ is the Kronecker delta function. Inverting the
347 $B$ matrix, we obtain
348 \[
349 C = B^{ - 1} = \left( {\begin{array}{*{20}c}
350 {C_{11} } & \ldots & {C_{1N} } \\
351 \vdots & \ddots & \vdots \\
352 {C_{N1} } & \cdots & {C_{NN} } \\
353 \end{array}} \right),
354 \]
355 which can be partitioned into $N \times N$ $3 \times 3$ block
356 $C_{ij}$. With the help of $C_{ij}$ and the skew matrix $U_i$
357 \[
358 U_i = \left( {\begin{array}{*{20}c}
359 0 & { - z_i } & {y_i } \\
360 {z_i } & 0 & { - x_i } \\
361 { - y_i } & {x_i } & 0 \\
362 \end{array}} \right)
363 \]
364 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
365 bead $i$ and origin $O$, the elements of resistance tensor at
366 arbitrary origin $O$ can be written as
367 \begin{eqnarray}
368 \Xi _{}^{tt} & = & \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
369 \Xi _{}^{tr} & = & \Xi _{}^{rt} = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
370 \Xi _{}^{rr} & = & - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
371 \label{introEquation:ResistanceTensorArbitraryOrigin}
372 \end{eqnarray}
373 The resistance tensor depends on the origin to which they refer. The
374 proper location for applying the friction force is the center of
375 resistance (or center of reaction), at which the trace of rotational
376 resistance tensor, $ \Xi ^{rr}$ reaches a minimum value.
377 Mathematically, the center of resistance is defined as an unique
378 point of the rigid body at which the translation-rotation coupling
379 tensors are symmetric,
380 \begin{equation}
381 \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
382 \label{introEquation:definitionCR}
383 \end{equation}
384 From Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
385 we can easily derive that the translational resistance tensor is
386 origin independent, while the rotational resistance tensor and
387 translation-rotation coupling resistance tensor depend on the
388 origin. Given the resistance tensor at an arbitrary origin $O$, and
389 a vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
390 obtain the resistance tensor at $P$ by
391 \begin{equation}
392 \begin{array}{l}
393 \Xi _P^{tt} = \Xi _O^{tt} \\
394 \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
395 \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{{tr} ^{^T }} \\
396 \end{array}
397 \label{introEquation:resistanceTensorTransformation}
398 \end{equation}
399 where
400 \[
401 U_{OP} = \left( {\begin{array}{*{20}c}
402 0 & { - z_{OP} } & {y_{OP} } \\
403 {z_i } & 0 & { - x_{OP} } \\
404 { - y_{OP} } & {x_{OP} } & 0 \\
405 \end{array}} \right)
406 \]
407 Using Eq.~\ref{introEquation:definitionCR} and
408 Eq.~\ref{introEquation:resistanceTensorTransformation}, one can
409 locate the position of center of resistance,
410 \begin{eqnarray*}
411 \left( \begin{array}{l}
412 x_{OR} \\
413 y_{OR} \\
414 z_{OR} \\
415 \end{array} \right) & = &\left( {\begin{array}{*{20}c}
416 {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
417 { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
418 { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
419 \end{array}} \right)^{ - 1} \\
420 & & \left( \begin{array}{l}
421 (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
422 (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
423 (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
424 \end{array} \right) \\
425 \end{eqnarray*}
426 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
427 joining center of resistance $R$ and origin $O$.
428
429 \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
430
431 Consider the Langevin equations of motion in generalized coordinates
432 \begin{equation}
433 M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)} + F_{r,i} (t)
434 \label{LDGeneralizedForm}
435 \end{equation}
436 where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
437 and moment of inertial) matrix and $V_i$ is a generalized velocity,
438 $V_i = V_i(v_i,\omega _i)$. The right side of
439 Eq.~\ref{LDGeneralizedForm} consists of three generalized forces in
440 lab-fixed frame, systematic force $F_{s,i}$, dissipative force
441 $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
442 system in Newtownian mechanics typically refers to lab-fixed frame,
443 it is also convenient to handle the rotation of rigid body in
444 body-fixed frame. Thus the friction and random forces are calculated
445 in body-fixed frame and converted back to lab-fixed frame by:
446 \[
447 \begin{array}{l}
448 F_{f,i}^l (t) = Q^T F_{f,i}^b (t), \\
449 F_{r,i}^l (t) = Q^T F_{r,i}^b (t). \\
450 \end{array}
451 \]
452 Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
453 the body-fixed velocity at center of resistance $v_{R,i}^b$ and
454 angular velocity $\omega _i$
455 \begin{equation}
456 F_{r,i}^b (t) = \left( \begin{array}{l}
457 f_{r,i}^b (t) \\
458 \tau _{r,i}^b (t) \\
459 \end{array} \right) = - \left( {\begin{array}{*{20}c}
460 {\Xi _{R,t} } & {\Xi _{R,c}^T } \\
461 {\Xi _{R,c} } & {\Xi _{R,r} } \\
462 \end{array}} \right)\left( \begin{array}{l}
463 v_{R,i}^b (t) \\
464 \omega _i (t) \\
465 \end{array} \right),
466 \end{equation}
467 while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
468 with zero mean and variance
469 \begin{equation}
470 \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle =
471 \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle =
472 2k_B T\Xi _R \delta (t - t'). \label{randomForce}
473 \end{equation}
474 The equation of motion for $v_i$ can be written as
475 \begin{equation}
476 m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
477 f_{r,i}^l (t)
478 \end{equation}
479 Since the frictional force is applied at the center of resistance
480 which generally does not coincide with the center of mass, an extra
481 torque is exerted at the center of mass. Thus, the net body-fixed
482 frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
483 given by
484 \begin{equation}
485 \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
486 \end{equation}
487 where $r_{MR}$ is the vector from the center of mass to the center
488 of the resistance. Instead of integrating the angular velocity in
489 lab-fixed frame, we consider the equation of angular momentum in
490 body-fixed frame
491 \begin{equation}
492 \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
493 + \tau _{r,i}^b(t)
494 \end{equation}
495 Embedding the friction terms into force and torque, one can
496 integrate the langevin equations of motion for rigid body of
497 arbitrary shape in a velocity-Verlet style 2-part algorithm, where
498 $h= \delta t$:
499
500 {\tt moveA:}
501 \begin{align*}
502 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
503 + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
504 %
505 {\bf r}(t + h) &\leftarrow {\bf r}(t)
506 + h {\bf v}\left(t + h / 2 \right), \\
507 %
508 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
509 + \frac{h}{2} {\bf \tau}^b(t), \\
510 %
511 \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
512 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
513 \end{align*}
514 In this context, the $\mathrm{rotate}$ function is the reversible
515 product of the three body-fixed rotations,
516 \begin{equation}
517 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
518 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
519 / 2) \cdot \mathsf{G}_x(a_x /2),
520 \end{equation}
521 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
522 rotates both the rotation matrix ($\mathsf{Q}$) and the body-fixed
523 angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
524 axis $\alpha$,
525 \begin{equation}
526 \mathsf{G}_\alpha( \theta ) = \left\{
527 \begin{array}{lcl}
528 \mathsf{Q}(t) & \leftarrow & \mathsf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
529 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
530 j}(0).
531 \end{array}
532 \right.
533 \end{equation}
534 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
535 rotation matrix. For example, in the small-angle limit, the
536 rotation matrix around the body-fixed x-axis can be approximated as
537 \begin{equation}
538 \mathsf{R}_x(\theta) \approx \left(
539 \begin{array}{ccc}
540 1 & 0 & 0 \\
541 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
542 \theta^2 / 4} \\
543 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
544 \theta^2 / 4}
545 \end{array}
546 \right).
547 \end{equation}
548 All other rotations follow in a straightforward manner. After the
549 first part of the propagation, the forces and body-fixed torques are
550 calculated at the new positions and orientations
551
552 {\tt doForces:}
553 \begin{align*}
554 {\bf f}(t + h) &\leftarrow
555 - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
556 %
557 {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
558 \times \frac{\partial V}{\partial {\bf u}}, \\
559 %
560 {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{Q}(t + h)
561 \cdot {\bf \tau}^s(t + h).
562 \end{align*}
563 Once the forces and torques have been obtained at the new time step,
564 the velocities can be advanced to the same time value.
565
566 {\tt moveB:}
567 \begin{align*}
568 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
569 \right)
570 + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
571 %
572 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
573 \right)
574 + \frac{h}{2} {\bf \tau}^b(t + h) .
575 \end{align*}
576
577 \section{Results and Discussion}
578
579 The Langevin algorithm described in previous section has been
580 implemented in {\sc oopse}\cite{Meineke2005} and applied to studies
581 of the static and dynamic properties in several systems.
582
583 \subsection{Temperature Control}
584
585 As shown in Eq.~\ref{randomForce}, random collisions associated with
586 the solvent's thermal motions is controlled by the external
587 temperature. The capability to maintain the temperature of the whole
588 system was usually used to measure the stability and efficiency of
589 the algorithm. In order to verify the stability of this new
590 algorithm, a series of simulations are performed on system
591 consisiting of 256 SSD water molecules with different viscosities.
592 The initial configuration for the simulations is taken from a 1ns
593 NVT simulation with a cubic box of 19.7166~\AA. All simulation are
594 carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
595 with reference temperature at 300~K. The average temperature as a
596 function of $\eta$ is shown in Table \ref{langevin:viscosity} where
597 the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
598 1$ poise. The better temperature control at higher viscosity can be
599 explained by the finite size effect and relative slow relaxation
600 rate at lower viscosity regime.
601 \begin{table}
602 \caption{AVERAGE TEMPERATURES FROM LANGEVIN DYNAMICS SIMULATIONS OF
603 SSD WATER MOLECULES WITH REFERENCE TEMPERATURE AT 300~K.}
604 \label{langevin:viscosity}
605 \begin{center}
606 \begin{tabular}{lll}
607 \hline
608 $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
609 \hline
610 1 & 300.47 & 10.99 \\
611 0.1 & 301.19 & 11.136 \\
612 0.01 & 303.04 & 11.796 \\
613 \hline
614 \end{tabular}
615 \end{center}
616 \end{table}
617
618 Another set of calculations were performed to study the efficiency of
619 temperature control using different temperature coupling schemes.
620 The starting configuration is cooled to 173~K and evolved using NVE,
621 NVT, and Langevin dynamic with time step of 2 fs.
622 Fig.~\ref{langevin:temperature} shows the heating curve obtained as
623 the systems reach equilibrium. The orange curve in
624 Fig.~\ref{langevin:temperature} represents the simulation using
625 Nos\'e-Hoover temperature scaling scheme with thermostat of 5 ps
626 which gives reasonable tight coupling, while the blue one from
627 Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
628 scaling to the desire temperature. When $ \eta = 0$, Langevin dynamics becomes normal
629 NVE (see orange curve in Fig.~\ref{langevin:temperature}) which
630 loses the temperature control ability.
631
632 \begin{figure}
633 \centering
634 \includegraphics[width=\linewidth]{temperature.pdf}
635 \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
636 temperature fluctuation versus time.} \label{langevin:temperature}
637 \end{figure}
638
639 \subsection{Langevin Dynamics simulation {\it vs} NVE simulations}
640
641 To validate our langevin dynamics simulation. We performed several NVE
642 simulations with explicit solvents for different shaped
643 molecules. There are one solute molecule and 1929 solvent molecules in
644 NVE simulation. The parameters are shown in table
645 \ref{tab:parameters}. The force field between spheres is standard
646 Lennard-Jones, and ellipsoids interact with other ellipsoids and
647 spheres with generalized Gay-Berne potential. All simulations are
648 carried out at 300 K and 1 Atm. The time step is 25 ns, and a
649 switching function was applied to all potentials to smoothly turn off
650 the interactions between a range of $22$ and $25$ \AA. The switching
651 function was the standard (cubic) function,
652 \begin{equation}
653 s(r) =
654 \begin{cases}
655 1 & \text{if $r \le r_{\text{sw}}$},\\
656 \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
657 {(r_{\text{cut}} - r_{\text{sw}})^3}
658 & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
659 0 & \text{if $r > r_{\text{cut}}$.}
660 \end{cases}
661 \label{eq:switchingFunc}
662 \end{equation}
663 We have computed translational diffusion constants for lipid molecules
664 from the mean-square displacement,
665 \begin{equation}
666 D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
667 \end{equation}
668 of the solute molecules. Translational diffusion constants for the
669 different shaped molecules are shown in table
670 \ref{tab:translation}. We have also computed orientational correlation
671 times for different shaped molecules from fits of the second-order
672 Legendre polynomial correlation function,
673 \begin{equation}
674 C_{\ell}(t) = \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
675 \mu}_{i}(0) \right)
676 \end{equation}
677 the results are shown in table \ref{tab:rotation}. We used einstein
678 format of the pressure correlation function,
679 \begin{equation}
680 C_{\ell}(t) = \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
681 \mu}_{i}(0) \right)
682 \end{equation}
683 to estimate the viscosity of the systems from NVE simulations. The
684 viscosity can also be calculated by Green-Kubo pressure correlaton
685 function,
686 \begin{equation}
687 C_{\ell}(t) = \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
688 \mu}_{i}(0) \right)
689 \end{equation}
690 However, this method converges slowly, and the statistics are not good
691 enough to give us a very accurate value. The langevin dynamics
692 simulations for different shaped molecules are performed at the same
693 conditions as the NVE simulations with viscosity estimated from NVE
694 simulations. To get better statistics, 1024 non-interacting solute
695 molecules are put into one simulation box for each langevin
696 simulation, this is equal to 1024 simulations for single solute
697 systems. The diffusion constants and rotation relaxation times for
698 different shaped molecules are shown in table \ref{tab:translation}
699 and \ref{tab:rotation} to compare to the results calculated from NVE
700 simulations. The theoretical values for sphere is calculated from the
701 Stokes-Einstein law, the theoretical values for ellipsoid is
702 calculated from Perrin's fomula, the theoretical values for dumbbell
703 is calculated from StinXX and Davis theory. The exact method is
704 applied to the langevin dynamics simulations for sphere and ellipsoid,
705 the bead model is applied to the simulation for dumbbell molecule, and
706 the rough shell model is applied to ellipsoid, dumbbell, banana and
707 lipid molecules. The results from all the langevin dynamics
708 simulations, including exact, bead model and rough shell, match the
709 theoretical values perfectly for all different shaped molecules. This
710 indicates that our simulation package for langevin dynamics is working
711 well. The approxiate methods ( bead model and rough shell model) are
712 accurate enough for the current simulations. The goal of the langevin
713 dynamics theory is to replace the explicit solvents by the friction
714 forces. We compared the dynamic properties of different shaped
715 molecules in langevin dynamics simulations with that in NVE
716 simulations. The results are reasonable close. Overall, the
717 translational diffusion constants calculated from langevin dynamics
718 simulations are very close to the values from the NVE simulation. For
719 sphere and lipid molecules, the diffusion constants are a little bit
720 off from the NVE simulation results. One possible reason is that the
721 calculation of the viscosity is very difficult to be accurate. Another
722 possible reason is that although we save very frequently during the
723 NVE simulations and run pretty long time simulations, there is only
724 one solute molecule in the system which makes the calculation for the
725 diffusion constant difficult. The sphere molecule behaves as a free
726 rotor in the solvent, so there is no rotation relaxation time
727 calculated from NVE simulations. The rotation relaxation time is not
728 very close to the NVE simulations results. The banana and lipid
729 molecules match the NVE simulations results pretty well. The mismatch
730 between langevin dynamics and NVE simulation for ellipsoid is possibly
731 caused by the slip boundary condition. For dumbbell, the mismatch is
732 caused by the size of the solvent molecule is pretty large compared to
733 dumbbell molecule in NVE simulations.
734
735 According to our simulations, the langevin dynamics is a reliable
736 theory to apply to replace the explicit solvents, especially for the
737 translation properties. For large molecules, the rotation properties
738 are also mimiced reasonablly well.
739
740 \begin{table*}
741 \begin{minipage}{\linewidth}
742 \begin{center}
743 \caption{}
744 \begin{tabular}{llccccccc}
745 \hline
746 & & Sphere & Ellipsoid & Dumbbell(2 spheres) & Banana(3 ellpsoids) &
747 Lipid(head) & lipid(tail) & Solvent \\
748 \hline
749 $d$ (\AA) & & 6.5 & 4.6 & 6.5 & 4.2 & 6.5 & 4.6 & 4.7 \\
750 $l$ (\AA) & & $= d$ & 13.8 & $=d$ & 11.2 & $=d$ & 13.8 & 4.7 \\
751 $\epsilon^s$ (kcal/mol) & & 0.8 & 0.8 & 0.8 & 0.8 & 0.185 & 0.8 & 0.8 \\
752 $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 & 0.2 & 1 & 0.2 & 1 \\
753 $m$ (amu) & & 190 & 200 & 190 & 240 & 196 & 760 & 72.06 \\
754 %$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
755 %\multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
756 %\multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
757 %\multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\
758 %$\mu$ (Debye) & & varied & 0 & 0 \\
759 \end{tabular}
760 \label{tab:parameters}
761 \end{center}
762 \end{minipage}
763 \end{table*}
764
765 \begin{table*}
766 \begin{minipage}{\linewidth}
767 \begin{center}
768 \caption{}
769 \begin{tabular}{lccccc}
770 \hline
771 & & & & &Translation \\
772 \hline
773 & NVE & & Theoretical & Langevin & \\
774 \hline
775 & $\eta$ & D & D & method & D \\
776 \hline
777 sphere & 3.480159e-03 & 1.643135e-04 & 1.942779e-04 & exact & 1.982283e-04 \\
778 ellipsoid & 2.551262e-03 & 2.437492e-04 & 2.335756e-04 & exact & 2.374905e-04 \\
779 & 2.551262e-03 & 2.437492e-04 & 2.335756e-04 & rough shell & 2.284088e-04 \\
780 dumbell & 2.41276e-03 & 2.129432e-04 & 2.090239e-04 & bead model & 2.148098e-04 \\
781 & 2.41276e-03 & 2.129432e-04 & 2.090239e-04 & rough shell & 2.013219e-04 \\
782 banana & 2.9846e-03 & 1.527819e-04 & & rough shell & 1.54807e-04 \\
783 lipid & 3.488661e-03 & 0.9562979e-04 & & rough shell & 1.320987e-04 \\
784 \end{tabular}
785 \label{tab:translation}
786 \end{center}
787 \end{minipage}
788 \end{table*}
789
790 \begin{table*}
791 \begin{minipage}{\linewidth}
792 \begin{center}
793 \caption{}
794 \begin{tabular}{lccccc}
795 \hline
796 & & & & &Rotation \\
797 \hline
798 & NVE & & Theoretical & Langevin & \\
799 \hline
800 & $\eta$ & $\tau_0$ & $\tau_0$ & method & $\tau_0$ \\
801 \hline
802 sphere & 3.480159e-03 & & 1.208178e+04 & exact & 1.20628e+04 \\
803 ellipsoid & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & exact & 2.21507e+04 \\
804 & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & rough shell & 2.21714e+04 \\
805 dumbell & 2.41276e-03 & 1.42974e+04 & & bead model & 7.12435e+04 \\
806 & 2.41276e-03 & 1.42974e+04 & & rough shell & 7.04765e+04 \\
807 banana & 2.9846e-03 & 6.38323e+04 & & rough shell & 7.0945e+04 \\
808 lipid & 3.488661e-03 & 7.79595e+04 & & rough shell & 7.78886e+04 \\
809 \end{tabular}
810 \label{tab:rotation}
811 \end{center}
812 \end{minipage}
813 \end{table*}
814
815 Langevin dynamics simulations are applied to study the formation of
816 the ripple phase of lipid membranes. The initial configuration is
817 taken from our molecular dynamics studies on lipid bilayers with
818 lennard-Jones sphere solvents. The solvent molecules are excluded from
819 the system, the experimental value of water viscosity is applied to
820 mimic the heat bath. Fig. XXX is the snapshot of the stable
821 configuration of the system, the ripple structure stayed stable after
822 100 ns run. The efficiency of the simulation is increased by one order
823 of magnitude.
824
825 \subsection{Langevin Dynamics of Banana Shaped Molecules}
826
827 In order to verify that Langevin dynamics can mimic the dynamics of
828 the systems absent of explicit solvents, we carried out two sets of
829 simulations and compare their dynamic properties.
830 Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
831 made of 256 pentane molecules and two banana shaped molecules at
832 273~K. It has an equivalent implicit solvent system containing only
833 two banana shaped molecules with viscosity of 0.289 center poise. To
834 calculate the hydrodynamic properties of the banana shaped molecule,
835 we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
836 in which the banana shaped molecule is represented as a ``shell''
837 made of 2266 small identical beads with size of 0.3 \AA on the
838 surface. Applying the procedure described in
839 Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
840 identified the center of resistance at (0 $\rm{\AA}$, 0.7482 $\rm{\AA}$,
841 -0.1988 $\rm{\AA}$), as well as the resistance tensor,
842 \[
843 \left( {\begin{array}{*{20}c}
844 0.9261 & 0 & 0&0&0.08585&0.2057\\
845 0& 0.9270&-0.007063& 0.08585&0&0\\
846 0&-0.007063&0.7494&0.2057&0&0\\
847 0&0.0858&0.2057& 58.64& 0&0\\
848 0.08585&0&0&0&48.30&3.219&\\
849 0.2057&0&0&0&3.219&10.7373\\
850 \end{array}} \right).
851 \]
852 where the units for translational, translation-rotation coupling and rotational tensors are $\frac{kcal \cdot fs}{mol \cdot \rm{\AA}^2}$, $\frac{kcal \cdot fs}{mol \cdot \rm{\AA} \cdot rad}$ and $\frac{kcal \cdot fs}{mol \cdot rad^2}$ respectively.
853 Curves of the velocity auto-correlation functions in
854 Fig.~\ref{langevin:vacf} were shown to match each other very well.
855 However, because of the stochastic nature, simulation using Langevin
856 dynamics was shown to decay slightly faster than MD. In order to
857 study the rotational motion of the molecules, we also calculated the
858 auto-correlation function of the principle axis of the second GB
859 particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
860 probably due to the reason that we used the experimental viscosity directly instead of calculating bulk viscosity from simulation.
861
862 \begin{figure}
863 \centering
864 \includegraphics[width=\linewidth]{roughShell.pdf}
865 \caption[Rough shell model for banana shaped molecule]{Rough shell
866 model for banana shaped molecule.} \label{langevin:roughShell}
867 \end{figure}
868
869 \begin{figure}
870 \centering
871 \includegraphics[width=\linewidth]{twoBanana.pdf}
872 \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
873 256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
874 molecules and 256 pentane molecules.} \label{langevin:twoBanana}
875 \end{figure}
876
877 \begin{figure}
878 \centering
879 \includegraphics[width=\linewidth]{vacf.pdf}
880 \caption[Plots of Velocity Auto-correlation Functions]{Velocity
881 auto-correlation functions of NVE (explicit solvent) in blue and
882 Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
883 \end{figure}
884
885 \begin{figure}
886 \centering
887 \includegraphics[width=\linewidth]{uacf.pdf}
888 \caption[Auto-correlation functions of the principle axis of the
889 middle GB particle]{Auto-correlation functions of the principle axis
890 of the middle GB particle of NVE (blue) and Langevin dynamics
891 (red).} \label{langevin:uacf}
892 \end{figure}
893
894 \section{Conclusions}
895
896 We have presented a new Langevin algorithm by incorporating the
897 hydrodynamics properties of arbitrary shaped molecules into an
898 advanced symplectic integration scheme. The temperature control
899 ability of this algorithm was demonstrated by a set of simulations
900 with different viscosities. It was also shown to have significant
901 advantage of producing rapid thermal equilibration over
902 Nos\'{e}-Hoover method. Further studies in systems involving banana
903 shaped molecules illustrated that the dynamic properties could be
904 preserved by using this new algorithm as an implicit solvent model.
905
906
907 \section{Acknowledgments}
908 Support for this project was provided by the National Science
909 Foundation under grant CHE-0134881. T.L. also acknowledges the
910 financial support from center of applied mathematics at University
911 of Notre Dame.
912 \newpage
913
914 \bibliographystyle{jcp2}
915 \bibliography{langevin}
916
917 \end{document}