ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3337
Committed: Thu Jan 31 14:03:06 2008 UTC (16 years, 7 months ago) by gezelter
Content type: application/x-tex
File size: 57623 byte(s)
Log Message:
Editing

File Contents

# User Rev Content
1 tim 2746 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2     %\documentclass[aps,prb,preprint]{revtex4}
3     \documentclass[11pt]{article}
4     \usepackage{endfloat}
5 gezelter 3337 \usepackage{amsmath}
6 tim 2746 \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 gezelter 3299 \renewcommand\citemid{\ } % no comma in optional referenc note
21 tim 2746
22     \begin{document}
23    
24 gezelter 3205 \title{An algorithm for performing Langevin dynamics on rigid bodies of arbitrary shape }
25 tim 2746
26 gezelter 3299 \author{Xiuquan Sun, Teng Lin and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
27 tim 2746 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 gezelter 3316 Langevin dynamics, which mimics a simple heat bath with stochastic and
52     dissipative forces, has been applied in a variety of situations as an
53     alternative to molecular dynamics with explicit solvent molecules.
54     The stochastic treatment of the solvent allows the use of simulations
55     with substantially longer time and length scales. In general, the
56     dynamic and structural properties obtained from Langevin simulations
57     agree quite well with similar properties obtained from explicit
58     solvent simulations.
59    
60     Recent examples of the usefulness of Langevin simulations include a
61     study of met-enkephalin in which Langevin simulations predicted
62     dynamical properties that were largely in agreement with explicit
63     solvent simulations.\cite{Shen2002} By applying Langevin dynamics with
64     the UNRES model, Liow and his coworkers suggest that protein folding
65     pathways can be explored within a reasonable amount of
66     time.\cite{Liwo2005}
67    
68     The stochastic nature of Langevin dynamics also enhances the sampling
69     of the system and increases the probability of crossing energy
70     barriers.\cite{Cui2003,Banerjee2004} Combining Langevin dynamics with
71 gezelter 3333 Kramers' theory, Klimov and Thirumalai identified free-energy
72 gezelter 3316 barriers by studying the viscosity dependence of the protein folding
73     rates.\cite{Klimov1997} In order to account for solvent induced
74     interactions missing from the implicit solvent model, Kaya
75     incorporated a desolvation free energy barrier into protein
76     folding/unfolding studies and discovered a higher free energy barrier
77 xsun 3317 between the native and denatured states.\cite{HuseyinKaya07012005}
78 gezelter 3316
79     Because of its stability against noise, Langevin dynamics has also
80     proven useful for studying remagnetization processes in various
81     systems.\cite{Palacios1998,Berkov2002,Denisov2003} [Check: For
82 tim 2746 instance, the oscillation power spectrum of nanoparticles from
83 gezelter 3316 Langevin dynamics has the same peak frequencies for different wave
84     vectors, which recovers the property of magnetic excitations in small
85     finite structures.\cite{Berkov2005a}]
86 tim 2746
87 gezelter 3316 In typical LD simulations, the friction and random forces on
88 gezelter 3333 individual atoms are taken from Stokes' law,
89 gezelter 3316 \begin{eqnarray}
90     m \dot{v}(t) & = & -\nabla U(x) - \xi m v(t) + R(t) \\
91     \langle R(t) \rangle & = & 0 \\
92     \langle R(t) R(t') \rangle & = & 2 k_B T \xi m \delta(t - t')
93     \end{eqnarray}
94     where $\xi \approx 6 \pi \eta a$. Here $\eta$ is the viscosity of the
95     implicit solvent, and $a$ is the hydrodynamic radius of the atom.
96 tim 2746
97 gezelter 3333 The use of rigid substructures,\cite{Chun:2000fj}
98     coarse-graining,\cite{Ayton01,Golubkov06,Orlandi:2006fk,SunGezelter08}
99     and ellipsoidal representations of protein side chains~\cite{Fogolari:1996lr}
100     has made the use of the Stokes-Einstein approximation problematic. A
101     rigid substructure moves as a single unit with orientational as well
102     as translational degrees of freedom. This requires a more general
103 gezelter 3316 treatment of the hydrodynamics than the spherical approximation
104     provides. The atoms involved in a rigid or coarse-grained structure
105     should properly have solvent-mediated interactions with each
106     other. The theory of interactions {\it between} bodies moving through
107     a fluid has been developed over the past century and has been applied
108     to simulations of Brownian
109 gezelter 3333 motion.\cite{FIXMAN:1986lr,Ramachandran1996}
110 tim 2746
111 gezelter 3333 In order to account for the diffusion anisotropy of arbitrarily-shaped
112     particles, Fernandes and Garc\'{i}a de la Torre improved the original
113     Brownian dynamics simulation algorithm~\cite{Ermak1978,Allison1991} by
114     incorporating a generalized $6\times6$ diffusion tensor and
115     introducing a rotational evolution scheme consisting of three
116     consecutive rotations.\cite{Fernandes2002} Unfortunately, biases are
117     introduced into the system due to the arbitrary order of applying the
118     noncommuting rotation operators.\cite{Beard2003} Based on the
119     observation the momentum relaxation time is much less than the time
120     step, one may ignore the inertia in Brownian dynamics. However, the
121     assumption of zero average acceleration is not always true for
122     cooperative motion which is common in proteins. An inertial Brownian
123     dynamics (IBD) was proposed to address this issue by adding an
124     inertial correction term.\cite{Beard2000} As a complement to IBD which
125     has a lower bound in time step because of the inertial relaxation
126     time, long-time-step inertial dynamics (LTID) can be used to
127     investigate the inertial behavior of linked polymer segments in a low
128     friction regime.\cite{Beard2000} LTID can also deal with the
129     rotational dynamics for nonskew bodies without translation-rotation
130     coupling by separating the translation and rotation motion and taking
131     advantage of the analytical solution of hydrodynamics
132     properties. However, typical nonskew bodies like cylinders and
133     ellipsoids are inadequate to represent most complex macromolecular
134     assemblies. There is therefore a need for incorporating the
135     hydrodynamics of complex (and potentially skew) rigid bodies in the
136     library of methods available for performing Langevin simulations.
137    
138 gezelter 3316 \subsection{Rigid Body Dynamics}
139     Rigid bodies are frequently involved in the modeling of large
140     collections of particles that move as a single unit. In molecular
141     simulations, rigid bodies have been used to simplify protein-protein
142 gezelter 3333 docking,\cite{Gray2003} and lipid bilayer
143     simulations.\cite{SunGezelter08} Many of the water models in common
144     use are also rigid-body
145     models,\cite{Jorgensen83,Berendsen81,Berendsen87} although they are
146     typically evolved using constraints rather than rigid body equations
147     of motion.
148 gezelter 3316
149 gezelter 3333 Euler angles are a natural choice to describe the rotational degrees
150 gezelter 3337 of freedom. However, due to $\frac{1}{\sin \theta}$ singularities, the
151 gezelter 3333 numerical integration of corresponding equations of these motion can
152     become inaccurate (and inefficient). Although the use of multiple
153     sets of Euler angles can overcome this problem,\cite{Barojas1973} the
154     computational penalty and the loss of angular momentum conservation
155     remain. A singularity-free representation utilizing quaternions was
156     developed by Evans in 1977.\cite{Evans1977} The Evans quaternion
157     approach uses a nonseparable Hamiltonian, and this has prevented
158     symplectic algorithms from being utilized until very
159     recently.\cite{Miller2002}
160 gezelter 3316
161 gezelter 3333 Another approach is the application of holonomic constraints to the
162     atoms belonging to the rigid body. Each atom moves independently
163     under the normal forces deriving from potential energy and constraints
164     are used to guarantee rigidity. However, due to their iterative
165     nature, the SHAKE and RATTLE algorithms converge very slowly when the
166     number of constraints (and the number of particles that belong to the
167     rigid body) increases.\cite{Ryckaert1977,Andersen1983}
168 tim 2746
169 gezelter 3333 In order to develop a stable and efficient integration scheme that
170     preserves most constants of the motion, symplectic propagators are
171     necessary. By introducing a conjugate momentum to the rotation matrix
172     $Q$ and re-formulating Hamilton's equations, a symplectic
173     orientational integrator, RSHAKE,\cite{Kol1997} was proposed to evolve
174     rigid bodies on a constraint manifold by iteratively satisfying the
175     orthogonality constraint $Q^T Q = 1$. An alternative method using the
176     quaternion representation was developed by Omelyan.\cite{Omelyan1998}
177     However, both of these methods are iterative and suffer from some
178     related inefficiencies. A symplectic Lie-Poisson integrator for rigid
179     bodies developed by Dullweber {\it et al.}\cite{Dullweber1997} removes
180     most of the limitations mentioned above and is therefore the basis for
181     our Langevin integrator.
182 gezelter 3316
183 tim 2746 The goal of the present work is to develop a Langevin dynamics
184 tim 2999 algorithm for arbitrary-shaped rigid particles by integrating the
185 gezelter 3316 accurate estimation of friction tensor from hydrodynamics theory into
186     a symplectic rigid body dynamics propagator. In the sections below,
187 gezelter 3333 we review some of the theory of hydrodynamic tensors developed
188     primarily for Brownian simulations of multi-particle systems, we then
189     present our integration method for a set of generalized Langevin
190     equations of motion, and we compare the behavior of the new Langevin
191     integrator to dynamical quantities obtained via explicit solvent
192     molecular dynamics.
193 tim 2746
194 gezelter 3316 \subsection{\label{introSection:frictionTensor}The Friction Tensor}
195     Theoretically, a complete friction kernel can be determined using the
196 tim 2999 velocity autocorrelation function. However, this approach becomes
197 gezelter 3333 impractical when the solute becomes complex. Instead, various
198 gezelter 3316 approaches based on hydrodynamics have been developed to calculate the
199     friction coefficients. In general, the friction tensor $\Xi$ is a
200     $6\times 6$ matrix given by
201     \begin{equation}
202 gezelter 3333 \Xi = \left( \begin{array}{*{20}c}
203     \Xi^{tt} & \Xi^{rt} \\
204     \Xi^{tr} & \Xi^{rr} \\
205     \end{array} \right).
206 gezelter 3316 \end{equation}
207     Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
208     rotational resistance (friction) tensors respectively, while
209     $\Xi^{tr}$ is translation-rotation coupling tensor and $\Xi^{rt}$ is
210     rotation-translation coupling tensor. When a particle moves in a
211     fluid, it may experience friction force ($\mathbf{F}_f$) and torque
212     ($\mathbf{\tau}_f$) in opposition to the directions of the velocity
213     ($\mathbf{v}$) and body-fixed angular velocity ($\mathbf{\omega}$),
214     \begin{equation}
215 tim 2746 \left( \begin{array}{l}
216 gezelter 3316 \mathbf{F}_f \\
217     \mathbf{\tau}_f \\
218 gezelter 3333 \end{array} \right) = - \left( \begin{array}{*{20}c}
219     \Xi^{tt} & \Xi^{rt} \\
220     \Xi^{tr} & \Xi^{rr} \\
221     \end{array} \right)\left( \begin{array}{l}
222 gezelter 3316 \mathbf{v} \\
223     \mathbf{\omega} \\
224     \end{array} \right).
225     \end{equation}
226 tim 2746
227 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
228 gezelter 3316 For a spherical particle under ``stick'' boundary conditions, the
229     translational and rotational friction tensors can be calculated from
230 gezelter 3333 Stokes' law,
231 gezelter 3316 \begin{equation}
232 gezelter 3333 \Xi^{tt} = \left( \begin{array}{*{20}c}
233 tim 2999 {6\pi \eta R} & 0 & 0 \\
234     0 & {6\pi \eta R} & 0 \\
235     0 & 0 & {6\pi \eta R} \\
236 gezelter 3333 \end{array} \right)
237 gezelter 3316 \end{equation}
238 tim 2999 and
239 gezelter 3316 \begin{equation}
240 gezelter 3333 \Xi^{rr} = \left( \begin{array}{*{20}c}
241 tim 2999 {8\pi \eta R^3 } & 0 & 0 \\
242     0 & {8\pi \eta R^3 } & 0 \\
243     0 & 0 & {8\pi \eta R^3 } \\
244 gezelter 3333 \end{array} \right)
245 gezelter 3316 \end{equation}
246 tim 2999 where $\eta$ is the viscosity of the solvent and $R$ is the
247     hydrodynamic radius.
248    
249     Other non-spherical shapes, such as cylinders and ellipsoids, are
250 gezelter 3316 widely used as references for developing new hydrodynamics theories,
251 tim 2999 because their properties can be calculated exactly. In 1936, Perrin
252 gezelter 3333 extended Stokes' law to general ellipsoids which are given in
253     Cartesian coordinates by~\cite{Perrin1934,Perrin1936}
254 gezelter 3316 \begin{equation}
255 gezelter 3333 \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1.
256 gezelter 3316 \end{equation}
257 gezelter 3333 Here, the semi-axes are of lengths $a$, $b$, and $c$. Due to the
258     complexity of the elliptic integral, only uniaxial ellipsoids, either
259     prolate ($a \ge b = c$) or oblate ($a < b = c$), can be solved
260     exactly. Introducing an elliptic integral parameter $S$ for prolate,
261 gezelter 3316 \begin{equation}
262     S = \frac{2}{\sqrt{a^2 - b^2}} \ln \frac{a + \sqrt{a^2 - b^2}}{b},
263     \end{equation}
264 gezelter 3333 and oblate,
265 gezelter 3316 \begin{equation}
266     S = \frac{2}{\sqrt {b^2 - a^2 }} \arctan \frac{\sqrt {b^2 - a^2}}{a},
267     \end{equation}
268 gezelter 3333 ellipsoids, one can write down the translational and rotational
269     resistance tensors:
270 tim 2999 \begin{eqnarray*}
271 gezelter 3316 \Xi_a^{tt} & = & 16\pi \eta \frac{a^2 - b^2}{(2a^2 - b^2 )S - 2a}. \\
272     \Xi_b^{tt} = \Xi_c^{tt} & = & 32\pi \eta \frac{a^2 - b^2 }{(2a^2 - 3b^2 )S + 2a},
273 tim 2999 \end{eqnarray*}
274 gezelter 3333 for oblate, and
275 tim 2999 \begin{eqnarray*}
276 gezelter 3316 \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2 - b^2 )b^2}{2a - b^2 S}, \\
277     \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4 - b^4)}{(2a^2 - b^2 )S - 2a}
278 tim 2999 \end{eqnarray*}
279 gezelter 3333 for prolate ellipsoids. For both spherical and ellipsoidal particles,
280     the translation-rotation and rotation-translation coupling tensors are
281 gezelter 3316 zero.
282 tim 2746
283 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
284     Unlike spherical and other simply shaped molecules, there is no
285     analytical solution for the friction tensor for arbitrarily shaped
286     rigid molecules. The ellipsoid of revolution model and general
287     triaxial ellipsoid model have been used to approximate the
288 gezelter 3316 hydrodynamic properties of rigid bodies. However, the mapping from all
289     possible ellipsoidal spaces, $r$-space, to all possible combination of
290     rotational diffusion coefficients, $D$-space, is not
291     unique.\cite{Wegener1979} Additionally, because there is intrinsic
292     coupling between translational and rotational motion of rigid bodies,
293     general ellipsoids are not always suitable for modeling arbitrarily
294     shaped rigid molecules. A number of studies have been devoted to
295 tim 2999 determining the friction tensor for irregularly shaped rigid bodies
296 gezelter 3316 using more advanced methods where the molecule of interest was modeled
297     by a combinations of spheres\cite{Carrasco1999} and the hydrodynamics
298     properties of the molecule can be calculated using the hydrodynamic
299 gezelter 3333 interaction tensor.
300    
301     Consider a rigid assembly of $N$ beads immersed in a continuous
302     medium. Due to hydrodynamic interaction, the ``net'' velocity of $i$th
303     bead, $v'_i$ is different than its unperturbed velocity $v_i$,
304     \begin{equation}
305 tim 2999 v'_i = v_i - \sum\limits_{j \ne i} {T_{ij} F_j }
306 gezelter 3333 \end{equation}
307     where $F_i$ is the frictional force, and $T_{ij}$ is the hydrodynamic
308     interaction tensor. The frictional force on the $i^\mathrm{th}$ bead
309     is proportional to its ``net'' velocity
310 tim 2746 \begin{equation}
311 tim 2999 F_i = \zeta _i v_i - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
312     \label{introEquation:tensorExpression}
313 tim 2746 \end{equation}
314 tim 2999 This equation is the basis for deriving the hydrodynamic tensor. In
315     1930, Oseen and Burgers gave a simple solution to
316     Eq.~\ref{introEquation:tensorExpression}
317 tim 2746 \begin{equation}
318 tim 2999 T_{ij} = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
319     R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
320 tim 2746 \end{equation}
321 tim 2999 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
322     A second order expression for element of different size was
323     introduced by Rotne and Prager\cite{Rotne1969} and improved by
324     Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
325 tim 2746 \begin{equation}
326 tim 2999 T_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
327     \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
328     _i^2 + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
329     \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
330     \label{introEquation:RPTensorNonOverlapped}
331 tim 2746 \end{equation}
332 tim 2999 Both of the Eq.~\ref{introEquation:oseenTensor} and
333     Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption
334     $R_{ij} \ge \sigma _i + \sigma _j$. An alternative expression for
335     overlapping beads with the same radius, $\sigma$, is given by
336 tim 2746 \begin{equation}
337 tim 2999 T_{ij} = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
338     \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
339     \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
340     \label{introEquation:RPTensorOverlapped}
341 tim 2746 \end{equation}
342 tim 2999 To calculate the resistance tensor at an arbitrary origin $O$, we
343     construct a $3N \times 3N$ matrix consisting of $N \times N$
344     $B_{ij}$ blocks
345     \begin{equation}
346 gezelter 3333 B = \left( \begin{array}{*{20}c}
347     B_{11} & \ldots & B_{1N} \\
348 tim 2999 \vdots & \ddots & \vdots \\
349 gezelter 3333 B_{N1} & \cdots & B_{NN} \\
350     \end{array} \right),
351 tim 2999 \end{equation}
352     where $B_{ij}$ is given by
353 gezelter 3333 \begin{equation}
354 tim 2999 B_{ij} = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
355     )T_{ij}
356 gezelter 3333 \end{equation}
357 tim 2999 where $\delta _{ij}$ is the Kronecker delta function. Inverting the
358     $B$ matrix, we obtain
359 tim 2746 \[
360 tim 2999 C = B^{ - 1} = \left( {\begin{array}{*{20}c}
361     {C_{11} } & \ldots & {C_{1N} } \\
362     \vdots & \ddots & \vdots \\
363     {C_{N1} } & \cdots & {C_{NN} } \\
364     \end{array}} \right),
365 tim 2746 \]
366 tim 2999 which can be partitioned into $N \times N$ $3 \times 3$ block
367     $C_{ij}$. With the help of $C_{ij}$ and the skew matrix $U_i$
368 tim 2746 \[
369 tim 2999 U_i = \left( {\begin{array}{*{20}c}
370     0 & { - z_i } & {y_i } \\
371     {z_i } & 0 & { - x_i } \\
372     { - y_i } & {x_i } & 0 \\
373     \end{array}} \right)
374 tim 2746 \]
375 tim 2999 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
376     bead $i$ and origin $O$, the elements of resistance tensor at
377     arbitrary origin $O$ can be written as
378     \begin{eqnarray}
379     \Xi _{}^{tt} & = & \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
380     \Xi _{}^{tr} & = & \Xi _{}^{rt} = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
381 gezelter 3310 \Xi _{}^{rr} & = & - \sum\limits_i {\sum\limits_j {U_i C_{ij} } }
382     U_j + 6 \eta V {\bf I}. \notag
383 tim 2999 \label{introEquation:ResistanceTensorArbitraryOrigin}
384     \end{eqnarray}
385 gezelter 3310 The final term in the expression for $\Xi^{rr}$ is correction that
386     accounts for errors in the rotational motion of certain kinds of bead
387     models. The additive correction uses the solvent viscosity ($\eta$)
388     as well as the total volume of the beads that contribute to the
389     hydrodynamic model,
390     \begin{equation}
391     V = \frac{4 \pi}{3} \sum_{i=1}^{N} \sigma_i^3,
392     \end{equation}
393     where $\sigma_i$ is the radius of bead $i$. This correction term was
394     rigorously tested and compared with the analytical results for
395     two-sphere and ellipsoidal systems by Garcia de la Torre and
396     Rodes.\cite{Torre:1983lr}
397    
398    
399 tim 2999 The resistance tensor depends on the origin to which they refer. The
400     proper location for applying the friction force is the center of
401     resistance (or center of reaction), at which the trace of rotational
402     resistance tensor, $ \Xi ^{rr}$ reaches a minimum value.
403     Mathematically, the center of resistance is defined as an unique
404     point of the rigid body at which the translation-rotation coupling
405     tensors are symmetric,
406     \begin{equation}
407     \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
408     \label{introEquation:definitionCR}
409     \end{equation}
410     From Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
411     we can easily derive that the translational resistance tensor is
412     origin independent, while the rotational resistance tensor and
413     translation-rotation coupling resistance tensor depend on the
414     origin. Given the resistance tensor at an arbitrary origin $O$, and
415     a vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
416     obtain the resistance tensor at $P$ by
417     \begin{equation}
418     \begin{array}{l}
419     \Xi _P^{tt} = \Xi _O^{tt} \\
420     \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
421     \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{{tr} ^{^T }} \\
422     \end{array}
423     \label{introEquation:resistanceTensorTransformation}
424     \end{equation}
425     where
426 tim 2746 \[
427 tim 2999 U_{OP} = \left( {\begin{array}{*{20}c}
428     0 & { - z_{OP} } & {y_{OP} } \\
429     {z_i } & 0 & { - x_{OP} } \\
430     { - y_{OP} } & {x_{OP} } & 0 \\
431     \end{array}} \right)
432 tim 2746 \]
433 tim 2999 Using Eq.~\ref{introEquation:definitionCR} and
434     Eq.~\ref{introEquation:resistanceTensorTransformation}, one can
435     locate the position of center of resistance,
436     \begin{eqnarray*}
437     \left( \begin{array}{l}
438     x_{OR} \\
439     y_{OR} \\
440     z_{OR} \\
441 gezelter 3333 \end{array} \right) & = &\left( \begin{array}{*{20}c}
442 tim 2999 {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
443     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
444     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
445 gezelter 3333 \end{array} \right)^{ - 1} \\
446 tim 2999 & & \left( \begin{array}{l}
447     (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
448     (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
449     (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
450     \end{array} \right) \\
451     \end{eqnarray*}
452     where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
453     joining center of resistance $R$ and origin $O$.
454 tim 2746
455    
456 gezelter 3310 \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
457 gezelter 3337
458 tim 2999 Consider the Langevin equations of motion in generalized coordinates
459 tim 2746 \begin{equation}
460 gezelter 3337 \mathbf{M} \dot{\mathbf{V}}(t) = \mathbf{F}_{s}(t) +
461     \mathbf{F}_{f}(t) + \mathbf{R}(t)
462 tim 2746 \label{LDGeneralizedForm}
463     \end{equation}
464 gezelter 3337 where $\mathbf{M}$ is a $6 \times 6$ diagonal mass matrix (which
465     includes the mass of the rigid body as well as the moments of inertia
466     in the body-fixed frame) and $\mathbf{V}$ is a generalized velocity,
467     $\mathbf{V} =
468     \left\{\mathbf{v},\mathbf{\omega}\right\}$. The right side of
469 gezelter 3333 Eq.~\ref{LDGeneralizedForm} consists of three generalized forces: a
470 gezelter 3337 system force $\mathbf{F}_{s}$, a frictional or dissipative force
471     $\mathbf{F}_{f}$ and stochastic force $\mathbf{R}$. While the
472 gezelter 3333 evolution of the system in Newtownian mechanics is typically done in the
473     lab-fixed frame, it is convenient to handle the rotation of rigid
474     bodies in the body-fixed frame. Thus the friction and random forces are
475     calculated in body-fixed frame and converted back to lab-fixed frame
476 gezelter 3337 using the rigid body's rotation matrix ($Q$):
477 gezelter 3333 \begin{equation}
478 tim 2746 \begin{array}{l}
479 gezelter 3337 \mathbf{F}_{f}(t) = Q^{T} \mathbf{F}_{f}^b (t), \\
480     \mathbf{R}(t) = Q^{T} \mathbf{R}^b (t). \\
481 tim 2999 \end{array}
482 gezelter 3333 \end{equation}
483     Here, the body-fixed friction force $\mathbf{F}_{f,i}^b$ is proportional to
484     the body-fixed velocity at the center of resistance $\mathbf{v}_{R,i}^b$ and
485     angular velocity $\mathbf{\omega}_i$
486 tim 2746 \begin{equation}
487 gezelter 3337 \mathbf{F}_{f}^b (t) = \left( \begin{array}{l}
488     \mathbf{f}_{f}^b (t) \\
489     \mathbf{\tau}_{f}^b (t) \\
490 gezelter 3333 \end{array} \right) = - \left( \begin{array}{*{20}c}
491     \Xi_{R,t} & \Xi_{R,c}^T \\
492     \Xi_{R,c} & \Xi_{R,r} \\
493     \end{array} \right)\left( \begin{array}{l}
494 gezelter 3337 \mathbf{v}_{R}^b (t) \\
495     \mathbf{\omega} (t) \\
496 tim 2746 \end{array} \right),
497     \end{equation}
498 gezelter 3337 while the random force $\mathbf{R}^l$ is a Gaussian stochastic variable
499 tim 2746 with zero mean and variance
500     \begin{equation}
501 gezelter 3337 \left\langle {\mathbf{R}^l (t) (\mathbf{R}^l (t'))^T } \right\rangle =
502     \left\langle {\mathbf{R}^b (t) (\mathbf{R}^b (t'))^T } \right\rangle =
503 gezelter 3333 2 k_B T \Xi_R \delta(t - t'). \label{randomForce}
504 tim 2746 \end{equation}
505 gezelter 3333 Once the $6\times6$ resistance tensor at the center of resistance
506     ($\Xi_R$) is known, obtaining a stochastic vector that has the
507     properties in Eq. (\ref{eq:randomForce}) can be done efficiently by
508     carrying out a one-time Cholesky decomposition to obtain the square
509     root matrix of $\Xi_R$.\cite{SchlickBook} Each time a random force
510     vector is needed, a gaussian random vector is generated and then the
511     square root matrix is multiplied onto this vector.
512    
513 gezelter 3337 The equation of motion for $\mathbf{v}$ can be written as
514 tim 2746 \begin{equation}
515 gezelter 3337 m \dot{\mathbf{v}} (t) = \mathbf{f}_{s} (t) + \mathbf{f}_{f}^l (t) +
516     \mathbf{R}^l (t)
517 tim 2746 \end{equation}
518     Since the frictional force is applied at the center of resistance
519     which generally does not coincide with the center of mass, an extra
520     torque is exerted at the center of mass. Thus, the net body-fixed
521 gezelter 3337 frictional torque at the center of mass, $\tau_{f}^b (t)$, is
522 tim 2746 given by
523     \begin{equation}
524 gezelter 3337 \tau_{f}^b \leftarrow \tau_{f}^b + \mathbf{r}_{MR} \times \mathbf{f}_{r}^b
525 tim 2746 \end{equation}
526     where $r_{MR}$ is the vector from the center of mass to the center
527 tim 2999 of the resistance. Instead of integrating the angular velocity in
528     lab-fixed frame, we consider the equation of angular momentum in
529     body-fixed frame
530 tim 2746 \begin{equation}
531 gezelter 3337 \dot j(t) = \tau_{s} (t) + \tau_{f}^b (t) + \mathbf{R}^b(t)
532 tim 2746 \end{equation}
533 gezelter 3333 Embedding the friction terms into force and torque, one can integrate
534     the Langevin equations of motion for rigid body of arbitrary shape in
535     a velocity-Verlet style 2-part algorithm, where $h= \delta t$:
536 tim 2746
537 tim 2999 {\tt moveA:}
538 tim 2746 \begin{align*}
539 tim 2999 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
540     + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
541     %
542     {\bf r}(t + h) &\leftarrow {\bf r}(t)
543     + h {\bf v}\left(t + h / 2 \right), \\
544     %
545     {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
546     + \frac{h}{2} {\bf \tau}^b(t), \\
547     %
548     \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
549     (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
550 tim 2746 \end{align*}
551     In this context, the $\mathrm{rotate}$ function is the reversible
552 tim 2999 product of the three body-fixed rotations,
553 tim 2746 \begin{equation}
554     \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
555     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
556     / 2) \cdot \mathsf{G}_x(a_x /2),
557     \end{equation}
558     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
559 tim 2999 rotates both the rotation matrix ($\mathsf{Q}$) and the body-fixed
560     angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
561     axis $\alpha$,
562 tim 2746 \begin{equation}
563     \mathsf{G}_\alpha( \theta ) = \left\{
564     \begin{array}{lcl}
565 tim 2999 \mathsf{Q}(t) & \leftarrow & \mathsf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
566 tim 2746 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
567     j}(0).
568     \end{array}
569     \right.
570     \end{equation}
571     $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
572     rotation matrix. For example, in the small-angle limit, the
573     rotation matrix around the body-fixed x-axis can be approximated as
574     \begin{equation}
575     \mathsf{R}_x(\theta) \approx \left(
576     \begin{array}{ccc}
577     1 & 0 & 0 \\
578     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
579     \theta^2 / 4} \\
580     0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
581     \theta^2 / 4}
582     \end{array}
583     \right).
584     \end{equation}
585 tim 2999 All other rotations follow in a straightforward manner. After the
586     first part of the propagation, the forces and body-fixed torques are
587     calculated at the new positions and orientations
588 tim 2746
589 tim 2999 {\tt doForces:}
590     \begin{align*}
591     {\bf f}(t + h) &\leftarrow
592     - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
593     %
594     {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
595     \times \frac{\partial V}{\partial {\bf u}}, \\
596     %
597     {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{Q}(t + h)
598     \cdot {\bf \tau}^s(t + h).
599     \end{align*}
600 tim 2746 Once the forces and torques have been obtained at the new time step,
601     the velocities can be advanced to the same time value.
602    
603 tim 2999 {\tt moveB:}
604 tim 2746 \begin{align*}
605 tim 2999 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
606     \right)
607     + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
608     %
609     {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
610     \right)
611     + \frac{h}{2} {\bf \tau}^b(t + h) .
612 tim 2746 \end{align*}
613    
614 gezelter 3310 \section{Validating the Method\label{sec:validating}}
615 gezelter 3302 In order to validate our Langevin integrator for arbitrarily-shaped
616 gezelter 3305 rigid bodies, we implemented the algorithm in {\sc
617     oopse}\cite{Meineke2005} and compared the results of this algorithm
618     with the known
619 gezelter 3302 hydrodynamic limiting behavior for a few model systems, and to
620     microcanonical molecular dynamics simulations for some more
621     complicated bodies. The model systems and their analytical behavior
622     (if known) are summarized below. Parameters for the primary particles
623     comprising our model systems are given in table \ref{tab:parameters},
624     and a sketch of the arrangement of these primary particles into the
625 gezelter 3305 model rigid bodies is shown in figure \ref{fig:models}. In table
626     \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
627     ellipsoidal (Gay-Berne) particles. For spherical particles, the value
628     of the Lennard-Jones $\sigma$ parameter is the particle diameter
629     ($d$). Gay-Berne ellipsoids have an energy scaling parameter,
630     $\epsilon^s$, which describes the well depth for two identical
631     ellipsoids in a {\it side-by-side} configuration. Additionally, a
632     well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
633     describes the ratio between the well depths in the {\it end-to-end}
634     and side-by-side configurations. For spheres, $\epsilon^r \equiv 1$.
635     Moments of inertia are also required to describe the motion of primary
636     particles with orientational degrees of freedom.
637 gezelter 3299
638 gezelter 3302 \begin{table*}
639     \begin{minipage}{\linewidth}
640     \begin{center}
641     \caption{Parameters for the primary particles in use by the rigid body
642     models in figure \ref{fig:models}.}
643     \begin{tabular}{lrcccccccc}
644     \hline
645     & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
646     & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
647     $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
648 gezelter 3308 Sphere & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
649 gezelter 3302 Ellipsoid & & 4.6 & 13.8 & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
650 gezelter 3308 Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
651 gezelter 3302 Banana &(3 identical ellipsoids)& 4.2 & 11.2 & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
652     Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
653     & Ellipsoidal Tail & 4.6 & 13.8 & 0.8 & 0.2 & 760 & 45000 & 45000 & 9000 \\
654     Solvent & & 4.7 & $= d$ & 0.8 & 1 & 72.06 & & & \\
655     \hline
656     \end{tabular}
657     \label{tab:parameters}
658     \end{center}
659     \end{minipage}
660     \end{table*}
661    
662 gezelter 3305 \begin{figure}
663     \centering
664     \includegraphics[width=3in]{sketch}
665     \caption[Sketch of the model systems]{A sketch of the model systems
666     used in evaluating the behavior of the rigid body Langevin
667     integrator.} \label{fig:models}
668     \end{figure}
669    
670 gezelter 3302 \subsection{Simulation Methodology}
671     We performed reference microcanonical simulations with explicit
672     solvents for each of the different model system. In each case there
673     was one solute model and 1929 solvent molecules present in the
674     simulation box. All simulations were equilibrated using a
675     constant-pressure and temperature integrator with target values of 300
676     K for the temperature and 1 atm for pressure. Following this stage,
677     further equilibration and sampling was done in a microcanonical
678 gezelter 3305 ensemble. Since the model bodies are typically quite massive, we were
679 gezelter 3310 able to use a time step of 25 fs.
680    
681     The model systems studied used both Lennard-Jones spheres as well as
682     uniaxial Gay-Berne ellipoids. In its original form, the Gay-Berne
683     potential was a single site model for the interactions of rigid
684     ellipsoidal molecules.\cite{Gay81} It can be thought of as a
685     modification of the Gaussian overlap model originally described by
686     Berne and Pechukas.\cite{Berne72} The potential is constructed in the
687     familiar form of the Lennard-Jones function using
688     orientation-dependent $\sigma$ and $\epsilon$ parameters,
689     \begin{equation*}
690     V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
691     r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
692     {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u
693     }_i},
694     {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
695     -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
696     {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
697     \label{eq:gb}
698     \end{equation*}
699    
700     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
701     \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
702     \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
703     are dependent on the relative orientations of the two ellipsoids (${\bf
704     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
705     inter-ellipsoid separation (${\bf \hat{r}}_{ij}$). The shape and
706     attractiveness of each ellipsoid is governed by a relatively small set
707     of parameters: $l$ and $d$ describe the length and width of each
708     uniaxial ellipsoid, while $\epsilon^s$, which describes the well depth
709     for two identical ellipsoids in a {\it side-by-side} configuration.
710     Additionally, a well depth aspect ratio, $\epsilon^r = \epsilon^e /
711     \epsilon^s$, describes the ratio between the well depths in the {\it
712     end-to-end} and side-by-side configurations. Details of the potential
713     are given elsewhere,\cite{Luckhurst90,Golubkov06,SunGezelter08} and an
714     excellent overview of the computational methods that can be used to
715     efficiently compute forces and torques for this potential can be found
716     in Ref. \citen{Golubkov06}
717    
718     For the interaction between nonequivalent uniaxial ellipsoids (or
719     between spheres and ellipsoids), the spheres are treated as ellipsoids
720     with an aspect ratio of 1 ($d = l$) and with an well depth ratio
721     ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of the
722     Gay-Berne potential we are using was generalized by Cleaver {\it et
723     al.} and is appropriate for dissimilar uniaxial
724     ellipsoids.\cite{Cleaver96}
725    
726     A switching function was applied to all potentials to smoothly turn
727     off the interactions between a range of $22$ and $25$ \AA. The
728     switching function was the standard (cubic) function,
729 gezelter 3302 \begin{equation}
730     s(r) =
731     \begin{cases}
732     1 & \text{if $r \le r_{\text{sw}}$},\\
733     \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
734     {(r_{\text{cut}} - r_{\text{sw}})^3}
735     & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
736     0 & \text{if $r > r_{\text{cut}}$.}
737     \end{cases}
738     \label{eq:switchingFunc}
739     \end{equation}
740 gezelter 3310
741 gezelter 3302 To measure shear viscosities from our microcanonical simulations, we
742     used the Einstein form of the pressure correlation function,\cite{hess:209}
743     \begin{equation}
744 gezelter 3310 \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
745     \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \right\rangle_{t_0}.
746 gezelter 3302 \label{eq:shear}
747     \end{equation}
748     A similar form exists for the bulk viscosity
749     \begin{equation}
750 gezelter 3310 \kappa = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
751 gezelter 3302 \int_{t_0}^{t_0 + t}
752 gezelter 3310 \left(P\left(t'\right)-\left\langle P \right\rangle \right)dt'
753     \right)^2 \right\rangle_{t_0}.
754 gezelter 3302 \end{equation}
755     Alternatively, the shear viscosity can also be calculated using a
756     Green-Kubo formula with the off-diagonal pressure tensor correlation function,
757     \begin{equation}
758 gezelter 3310 \eta = \frac{V}{k_B T} \int_0^{\infty} \left\langle P_{xz}(t_0) P_{xz}(t_0
759     + t) \right\rangle_{t_0} dt,
760 gezelter 3302 \end{equation}
761     although this method converges extremely slowly and is not practical
762     for obtaining viscosities from molecular dynamics simulations.
763    
764     The Langevin dynamics for the different model systems were performed
765     at the same temperature as the average temperature of the
766     microcanonical simulations and with a solvent viscosity taken from
767 gezelter 3305 Eq. (\ref{eq:shear}) applied to these simulations. We used 1024
768     independent solute simulations to obtain statistics on our Langevin
769     integrator.
770 gezelter 3302
771     \subsection{Analysis}
772    
773     The quantities of interest when comparing the Langevin integrator to
774     analytic hydrodynamic equations and to molecular dynamics simulations
775     are typically translational diffusion constants and orientational
776     relaxation times. Translational diffusion constants for point
777     particles are computed easily from the long-time slope of the
778     mean-square displacement,
779     \begin{equation}
780 gezelter 3310 D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \left\langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \right\rangle,
781 gezelter 3302 \end{equation}
782     of the solute molecules. For models in which the translational
783 gezelter 3305 diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
784     (i.e. any non-spherically-symmetric rigid body), it is possible to
785     compute the diffusive behavior for motion parallel to each body-fixed
786     axis by projecting the displacement of the particle onto the
787     body-fixed reference frame at $t=0$. With an isotropic solvent, as we
788     have used in this study, there are differences between the three
789 gezelter 3302 diffusion constants, but these must converge to the same value at
790     longer times. Translational diffusion constants for the different
791 gezelter 3305 shaped models are shown in table \ref{tab:translation}.
792 gezelter 3302
793 gezelter 3305 In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
794 gezelter 3302 diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
795     {\it around} a particular body-fixed axis and {\it not} the diffusion
796     of a vector pointing along the axis. However, these eigenvalues can
797     be combined to find 5 characteristic rotational relaxation
798 gezelter 3305 times,\cite{PhysRev.119.53,Berne90}
799 gezelter 3302 \begin{eqnarray}
800 gezelter 3305 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
801     1 / \tau_2 & = & 6 D_r - 2 \Delta \\
802     1 / \tau_3 & = & 3 (D_r + D_1) \\
803     1 / \tau_4 & = & 3 (D_r + D_2) \\
804     1 / \tau_5 & = & 3 (D_r + D_3)
805 gezelter 3302 \end{eqnarray}
806     where
807     \begin{equation}
808     D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
809     \end{equation}
810     and
811     \begin{equation}
812 gezelter 3305 \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
813 gezelter 3302 \end{equation}
814 gezelter 3305 Each of these characteristic times can be used to predict the decay of
815     part of the rotational correlation function when $\ell = 2$,
816 gezelter 3302 \begin{equation}
817 gezelter 3305 C_2(t) = \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
818 gezelter 3302 \end{equation}
819 gezelter 3305 This is the same as the $F^2_{0,0}(t)$ correlation function that
820     appears in Ref. \citen{Berne90}. The amplitudes of the two decay
821     terms are expressed in terms of three dimensionless functions of the
822     eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
823     2\Delta)$, and $N = 2 \sqrt{\Delta b}$. Similar expressions can be
824     obtained for other angular momentum correlation
825     functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
826     studied, only one of the amplitudes of the two decay terms was
827     non-zero, so it was possible to derive a single relaxation time for
828     each of the hydrodynamic tensors. In many cases, these characteristic
829     times are averaged and reported in the literature as a single relaxation
830     time,\cite{Garcia-de-la-Torre:1997qy}
831 gezelter 3302 \begin{equation}
832 gezelter 3305 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
833     \end{equation}
834     although for the cases reported here, this averaging is not necessary
835     and only one of the five relaxation times is relevant.
836    
837     To test the Langevin integrator's behavior for rotational relaxation,
838     we have compared the analytical orientational relaxation times (if
839     they are known) with the general result from the diffusion tensor and
840     with the results from both the explicitly solvated molecular dynamics
841     and Langevin simulations. Relaxation times from simulations (both
842     microcanonical and Langevin), were computed using Legendre polynomial
843     correlation functions for a unit vector (${\bf u}$) fixed along one or
844     more of the body-fixed axes of the model.
845     \begin{equation}
846 gezelter 3310 C_{\ell}(t) = \left\langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
847     u}_{i}(0) \right) \right\rangle
848 gezelter 3302 \end{equation}
849     For simulations in the high-friction limit, orientational correlation
850     times can then be obtained from exponential fits of this function, or by
851     integrating,
852     \begin{equation}
853 gezelter 3305 \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
854 gezelter 3302 \end{equation}
855 gezelter 3305 In lower-friction solvents, the Legendre correlation functions often
856     exhibit non-exponential decay, and may not be characterized by a
857     single decay constant.
858 gezelter 3302
859     In table \ref{tab:rotation} we show the characteristic rotational
860     relaxation times (based on the diffusion tensor) for each of the model
861     systems compared with the values obtained via microcanonical and Langevin
862     simulations.
863    
864 gezelter 3305 \subsection{Spherical particles}
865 gezelter 3299 Our model system for spherical particles was a Lennard-Jones sphere of
866     diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
867     4.7 \AA). The well depth ($\epsilon$) for both particles was set to
868 gezelter 3302 an arbitrary value of 0.8 kcal/mol.
869 gezelter 3299
870     The Stokes-Einstein behavior of large spherical particles in
871     hydrodynamic flows is well known, giving translational friction
872     coefficients of $6 \pi \eta R$ (stick boundary conditions) and
873 gezelter 3302 rotational friction coefficients of $8 \pi \eta R^3$. Recently,
874     Schmidt and Skinner have computed the behavior of spherical tag
875     particles in molecular dynamics simulations, and have shown that {\it
876     slip} boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
877 gezelter 3299 appropriate for molecule-sized spheres embedded in a sea of spherical
878 gezelter 3310 solvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
879 gezelter 3299
880     Our simulation results show similar behavior to the behavior observed
881 gezelter 3302 by Schmidt and Skinner. The diffusion constant obtained from our
882 gezelter 3299 microcanonical molecular dynamics simulations lies between the slip
883     and stick boundary condition results obtained via Stokes-Einstein
884     behavior. Since the Langevin integrator assumes Stokes-Einstein stick
885     boundary conditions in calculating the drag and random forces for
886     spherical particles, our Langevin routine obtains nearly quantitative
887     agreement with the hydrodynamic results for spherical particles. One
888     avenue for improvement of the method would be to compute elements of
889     $\Xi_{tt}$ assuming behavior intermediate between the two boundary
890 gezelter 3302 conditions.
891 gezelter 3299
892 gezelter 3310 In the explicit solvent simulations, both our solute and solvent
893     particles were structureless, exerting no torques upon each other.
894     Therefore, there are not rotational correlation times available for
895     this model system.
896 gezelter 3299
897 gezelter 3310 \subsection{Ellipsoids}
898     For uniaxial ellipsoids ($a > b = c$), Perrin's formulae for both
899 gezelter 3299 translational and rotational diffusion of each of the body-fixed axes
900     can be combined to give a single translational diffusion
901 gezelter 3302 constant,\cite{Berne90}
902 gezelter 3299 \begin{equation}
903     D = \frac{k_B T}{6 \pi \eta a} G(\rho),
904     \label{Dperrin}
905     \end{equation}
906     as well as a single rotational diffusion coefficient,
907     \begin{equation}
908     \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - \rho^2)
909     G(\rho) - 1}{1 - \rho^4} \right\}.
910     \label{ThetaPerrin}
911     \end{equation}
912     In these expressions, $G(\rho)$ is a function of the axial ratio
913     ($\rho = b / a$), which for prolate ellipsoids, is
914     \begin{equation}
915     G(\rho) = (1- \rho^2)^{-1/2} \ln \left\{ \frac{1 + (1 -
916     \rho^2)^{1/2}}{\rho} \right\}
917     \label{GPerrin}
918     \end{equation}
919     Again, there is some uncertainty about the correct boundary conditions
920     to use for molecular-scale ellipsoids in a sea of similarly-sized
921     solvent particles. Ravichandran and Bagchi found that {\it slip}
922 gezelter 3302 boundary conditions most closely resembled the simulation
923     results,\cite{Ravichandran:1999fk} in agreement with earlier work of
924     Tang and Evans.\cite{TANG:1993lr}
925 gezelter 3299
926 gezelter 3305 Even though there are analytic resistance tensors for ellipsoids, we
927     constructed a rough-shell model using 2135 beads (each with a diameter
928 gezelter 3310 of 0.25 \AA) to approximate the shape of the model ellipsoid. We
929 gezelter 3305 compared the Langevin dynamics from both the simple ellipsoidal
930     resistance tensor and the rough shell approximation with
931     microcanonical simulations and the predictions of Perrin. As in the
932     case of our spherical model system, the Langevin integrator reproduces
933     almost exactly the behavior of the Perrin formulae (which is
934     unsurprising given that the Perrin formulae were used to derive the
935 gezelter 3299 drag and random forces applied to the ellipsoid). We obtain
936     translational diffusion constants and rotational correlation times
937     that are within a few percent of the analytic values for both the
938     exact treatment of the diffusion tensor as well as the rough-shell
939     model for the ellipsoid.
940    
941 gezelter 3308 The translational diffusion constants from the microcanonical simulations
942     agree well with the predictions of the Perrin model, although the rotational
943     correlation times are a factor of 2 shorter than expected from hydrodynamic
944     theory. One explanation for the slower rotation
945     of explicitly-solvated ellipsoids is the possibility that solute-solvent
946     collisions happen at both ends of the solute whenever the principal
947     axis of the ellipsoid is turning. In the upper portion of figure
948     \ref{fig:explanation} we sketch a physical picture of this explanation.
949     Since our Langevin integrator is providing nearly quantitative agreement with
950     the Perrin model, it also predicts orientational diffusion for ellipsoids that
951     exceed explicitly solvated correlation times by a factor of two.
952 gezelter 3299
953 gezelter 3310 \subsection{Rigid dumbbells}
954 gezelter 3302 Perhaps the only {\it composite} rigid body for which analytic
955     expressions for the hydrodynamic tensor are available is the
956     two-sphere dumbbell model. This model consists of two non-overlapping
957     spheres held by a rigid bond connecting their centers. There are
958     competing expressions for the 6x6 resistance tensor for this
959     model. Equation (\ref{introEquation:oseenTensor}) above gives the
960     original Oseen tensor, while the second order expression introduced by
961     Rotne and Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la
962     Torre and Bloomfield,\cite{Torre1977} is given above as
963 gezelter 3299 Eq. (\ref{introEquation:RPTensorNonOverlapped}). In our case, we use
964     a model dumbbell in which the two spheres are identical Lennard-Jones
965     particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
966 gezelter 3302 a distance of 6.532 \AA.
967 gezelter 3299
968     The theoretical values for the translational diffusion constant of the
969     dumbbell are calculated from the work of Stimson and Jeffery, who
970     studied the motion of this system in a flow parallel to the
971 gezelter 3302 inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
972     motion in a flow {\it perpendicular} to the inter-sphere
973     axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
974     orientational} correlation times for this model system (other than
975 gezelter 3305 those derived from the 6 x 6 tensors mentioned above).
976 gezelter 3299
977 gezelter 3305 The bead model for this model system comprises the two large spheres
978     by themselves, while the rough shell approximation used 3368 separate
979     beads (each with a diameter of 0.25 \AA) to approximate the shape of
980     the rigid body. The hydrodynamics tensors computed from both the bead
981     and rough shell models are remarkably similar. Computing the initial
982     hydrodynamic tensor for a rough shell model can be quite expensive (in
983     this case it requires inverting a 10104 x 10104 matrix), while the
984     bead model is typically easy to compute (in this case requiring
985 gezelter 3308 inversion of a 6 x 6 matrix).
986 gezelter 3305
987 gezelter 3308 \begin{figure}
988     \centering
989 gezelter 3310 \includegraphics[width=2in]{RoughShell}
990 gezelter 3308 \caption[Model rigid bodies and their rough shell approximations]{The
991     model rigid bodies (left column) used to test this algorithm and their
992     rough-shell approximations (right-column) that were used to compute
993     the hydrodynamic tensors. The top two models (ellipsoid and dumbbell)
994     have analytic solutions and were used to test the rough shell
995     approximation. The lower two models (banana and lipid) were compared
996     with explicitly-solvated molecular dynamics simulations. }
997     \label{fig:roughShell}
998     \end{figure}
999    
1000    
1001 gezelter 3305 Once the hydrodynamic tensor has been computed, there is no additional
1002     penalty for carrying out a Langevin simulation with either of the two
1003     different hydrodynamics models. Our naive expectation is that since
1004     the rigid body's surface is roughened under the various shell models,
1005     the diffusion constants will be even farther from the ``slip''
1006     boundary conditions than observed for the bead model (which uses a
1007     Stokes-Einstein model to arrive at the hydrodynamic tensor). For the
1008     dumbbell, this prediction is correct although all of the Langevin
1009     diffusion constants are within 6\% of the diffusion constant predicted
1010     from the fully solvated system.
1011    
1012 gezelter 3308 For rotational motion, Langevin integration (and the hydrodynamic tensor)
1013     yields rotational correlation times that are substantially shorter than those
1014     obtained from explicitly-solvated simulations. It is likely that this is due
1015     to the large size of the explicit solvent spheres, a feature that prevents
1016     the solvent from coming in contact with a substantial fraction of the surface
1017     area of the dumbbell. Therefore, the explicit solvent only provides drag
1018     over a substantially reduced surface area of this model, while the
1019     hydrodynamic theories utilize the entire surface area for estimating
1020     rotational diffusion. A sketch of the free volume available in the explicit
1021     solvent simulations is shown in figure \ref{fig:explanation}.
1022 gezelter 3305
1023 gezelter 3310
1024     \begin{figure}
1025     \centering
1026     \includegraphics[width=6in]{explanation}
1027     \caption[Explanations of the differences between orientational
1028     correlation times for explicitly-solvated models and hydrodynamics
1029     predictions]{Explanations of the differences between orientational
1030     correlation times for explicitly-solvated models and hydrodynamic
1031     predictions. For the ellipsoids (upper figures), rotation of the
1032     principal axis can involve correlated collisions at both sides of the
1033     solute. In the rigid dumbbell model (lower figures), the large size
1034     of the explicit solvent spheres prevents them from coming in contact
1035     with a substantial fraction of the surface area of the dumbbell.
1036     Therefore, the explicit solvent only provides drag over a
1037     substantially reduced surface area of this model, where the
1038     hydrodynamic theories utilize the entire surface area for estimating
1039     rotational diffusion.
1040     } \label{fig:explanation}
1041     \end{figure}
1042    
1043    
1044    
1045     \subsection{Composite banana-shaped molecules}
1046     Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids have
1047     been used by Orlandi {\it et al.} to observe mesophases in
1048     coarse-grained models for bent-core liquid crystalline
1049     molecules.\cite{Orlandi:2006fk} We have used the same overlapping
1050 gezelter 3299 ellipsoids as a way to test the behavior of our algorithm for a
1051     structure of some interest to the materials science community,
1052     although since we are interested in capturing only the hydrodynamic
1053 gezelter 3310 behavior of this model, we have left out the dipolar interactions of
1054     the original Orlandi model.
1055 gezelter 3308
1056     A reference system composed of a single banana rigid body embedded in a
1057     sea of 1929 solvent particles was created and run under standard
1058     (microcanonical) molecular dynamics. The resulting viscosity of this
1059     mixture was 0.298 centipoise (as estimated using Eq. (\ref{eq:shear})).
1060     To calculate the hydrodynamic properties of the banana rigid body model,
1061 gezelter 3310 we created a rough shell (see Fig.~\ref{fig:roughShell}), in which
1062 gezelter 3308 the banana is represented as a ``shell'' made of 3321 identical beads
1063 gezelter 3310 (0.25 \AA\ in diameter) distributed on the surface. Applying the
1064 gezelter 3308 procedure described in Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1065 gezelter 3310 identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0 \AA), as
1066     well as the resistance tensor,
1067     \begin{equation*}
1068     \Xi =
1069 gezelter 3308 \left( {\begin{array}{*{20}c}
1070     0.9261 & 0 & 0&0&0.08585&0.2057\\
1071     0& 0.9270&-0.007063& 0.08585&0&0\\
1072     0&-0.007063&0.7494&0.2057&0&0\\
1073 gezelter 3310 0&0.0858&0.2057& 58.64& 0&0\\0.08585&0&0&0&48.30&3.219&\\0.2057&0&0&0&3.219&10.7373\\\end{array}} \right),
1074     \end{equation*}
1075     where the units for translational, translation-rotation coupling and
1076     rotational tensors are (kcal fs / mol \AA$^2$), (kcal fs / mol \AA\ rad),
1077     and (kcal fs / mol rad$^2$), respectively.
1078 gezelter 3299
1079 gezelter 3308 The Langevin rigid-body integrator (and the hydrodynamic diffusion tensor)
1080     are essentially quantitative for translational diffusion of this model.
1081     Orientational correlation times under the Langevin rigid-body integrator
1082     are within 11\% of the values obtained from explicit solvent, but these
1083     models also exhibit some solvent inaccessible surface area in the
1084     explicitly-solvated case.
1085    
1086 gezelter 3310 \subsection{Composite sphero-ellipsoids}
1087 gezelter 3299 Spherical heads perched on the ends of Gay-Berne ellipsoids have been
1088 xsun 3312 used recently as models for lipid
1089     molecules.\cite{SunGezelter08,Ayton01}
1090 gezelter 3310 MORE DETAILS
1091 xsun 3298
1092 xsun 3312 A reference system composed of a single lipid rigid body embedded in a
1093     sea of 1929 solvent particles was created and run under standard
1094     (microcanonical) molecular dynamics. The resulting viscosity of this
1095     mixture was 0.349 centipoise (as estimated using
1096     Eq. (\ref{eq:shear})). To calculate the hydrodynamic properties of
1097     the lipid rigid body model, we created a rough shell (see
1098     Fig.~\ref{fig:roughShell}), in which the lipid is represented as a
1099     ``shell'' made of 3550 identical beads (0.25 \AA\ in diameter)
1100     distributed on the surface. Applying the procedure described in
1101     Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1102     identified the center of resistance, ${\bf r} = $(0 \AA, 0 \AA, 1.46
1103     \AA).
1104 gezelter 3310
1105 gezelter 3315
1106 gezelter 3310 \subsection{Summary}
1107 xsun 3298 According to our simulations, the langevin dynamics is a reliable
1108     theory to apply to replace the explicit solvents, especially for the
1109     translation properties. For large molecules, the rotation properties
1110     are also mimiced reasonablly well.
1111    
1112 gezelter 3315 \begin{figure}
1113     \centering
1114     \includegraphics[width=\linewidth]{graph}
1115     \caption[Mean squared displacements and orientational
1116     correlation functions for each of the model rigid bodies.]{The
1117     mean-squared displacements ($\langle r^2(t) \rangle$) and
1118     orientational correlation functions ($C_2(t)$) for each of the model
1119     rigid bodies studied. The circles are the results for microcanonical
1120     simulations with explicit solvent molecules, while the other data sets
1121     are results for Langevin dynamics using the different hydrodynamic
1122     tensor approximations. The Perrin model for the ellipsoids is
1123     considered the ``exact'' hydrodynamic behavior (this can also be said
1124     for the translational motion of the dumbbell operating under the bead
1125     model). In most cases, the various hydrodynamics models reproduce
1126     each other quantitatively.}
1127     \label{fig:results}
1128     \end{figure}
1129    
1130 xsun 3298 \begin{table*}
1131     \begin{minipage}{\linewidth}
1132     \begin{center}
1133 gezelter 3305 \caption{Translational diffusion constants (D) for the model systems
1134     calculated using microcanonical simulations (with explicit solvent),
1135     theoretical predictions, and Langevin simulations (with implicit solvent).
1136     Analytical solutions for the exactly-solved hydrodynamics models are
1137     from Refs. \citen{Einstein05} (sphere), \citen{Perrin1934} and \citen{Perrin1936}
1138     (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1139     (dumbbell). The other model systems have no known analytic solution.
1140     All diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1141     $10^{-4}$ \AA$^2$ / fs). }
1142     \begin{tabular}{lccccccc}
1143 xsun 3298 \hline
1144 gezelter 3305 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1145     \cline{2-3} \cline{5-7}
1146     model & $\eta$ (centipoise) & D & & Analytical & method & Hydrodynamics & simulation \\
1147 xsun 3298 \hline
1148 xsun 3312 sphere & 0.279 & 3.06 & & 2.42 & exact & 2.42 & 2.33 \\
1149 gezelter 3305 ellipsoid & 0.255 & 2.44 & & 2.34 & exact & 2.34 & 2.37 \\
1150     & 0.255 & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1151 xsun 3312 dumbbell & 0.308 & 2.06 & & 1.64 & bead model & 1.65 & 1.62 \\
1152     & 0.308 & 2.06 & & 1.64 & rough shell & 1.59 & 1.62 \\
1153 gezelter 3305 banana & 0.298 & 1.53 & & & rough shell & 1.56 & 1.55 \\
1154     lipid & 0.349 & 0.96 & & & rough shell & 1.33 & 1.32 \\
1155 xsun 3298 \end{tabular}
1156     \label{tab:translation}
1157     \end{center}
1158     \end{minipage}
1159     \end{table*}
1160    
1161     \begin{table*}
1162     \begin{minipage}{\linewidth}
1163     \begin{center}
1164 gezelter 3305 \caption{Orientational relaxation times ($\tau$) for the model systems using
1165     microcanonical simulation (with explicit solvent), theoretical
1166     predictions, and Langevin simulations (with implicit solvent). All
1167     relaxation times are for the rotational correlation function with
1168     $\ell = 2$ and are reported in units of ps. The ellipsoidal model has
1169     an exact solution for the orientational correlation time due to
1170     Perrin, but the other model systems have no known analytic solution.}
1171     \begin{tabular}{lccccccc}
1172 xsun 3298 \hline
1173 gezelter 3305 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1174     \cline{2-3} \cline{5-7}
1175     model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic & simulation \\
1176 xsun 3298 \hline
1177 xsun 3312 sphere & 0.279 & & & 9.69 & exact & 9.69 & 9.64 \\
1178 gezelter 3305 ellipsoid & 0.255 & 46.7 & & 22.0 & exact & 22.0 & 22.2 \\
1179     & 0.255 & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1180 xsun 3312 dumbbell & 0.308 & 14.1 & & & bead model & 50.0 & 50.1 \\
1181     & 0.308 & 14.1 & & & rough shell & 41.5 & 41.3 \\
1182 gezelter 3305 banana & 0.298 & 63.8 & & & rough shell & 70.9 & 70.9 \\
1183     lipid & 0.349 & 78.0 & & & rough shell & 76.9 & 77.9 \\
1184     \hline
1185 xsun 3298 \end{tabular}
1186     \label{tab:rotation}
1187     \end{center}
1188     \end{minipage}
1189     \end{table*}
1190    
1191 gezelter 3310 \section{Application: A rigid-body lipid bilayer}
1192    
1193     The Langevin dynamics integrator was applied to study the formation of
1194     corrugated structures emerging from simulations of the coarse grained
1195     lipid molecular models presented above. The initial configuration is
1196 xsun 3298 taken from our molecular dynamics studies on lipid bilayers with
1197 gezelter 3310 lennard-Jones sphere solvents. The solvent molecules were excluded
1198     from the system, and the experimental value for the viscosity of water
1199     at 20C ($\eta = 1.00$ cp) was used to mimic the hydrodynamic effects
1200     of the solvent. The absence of explicit solvent molecules and the
1201     stability of the integrator allowed us to take timesteps of 50 fs. A
1202     total simulation run time of 100 ns was sampled.
1203     Fig. \ref{fig:bilayer} shows the configuration of the system after 100
1204     ns, and the ripple structure remains stable during the entire
1205     trajectory. Compared with using explicit bead-model solvent
1206     molecules, the efficiency of the simulation has increased by an order
1207 xsun 3298 of magnitude.
1208    
1209 gezelter 3310 \begin{figure}
1210     \centering
1211     \includegraphics[width=\linewidth]{bilayer}
1212     \caption[Snapshot of a bilayer of rigid-body models for lipids]{A
1213     snapshot of a bilayer composed of rigid-body models for lipid
1214     molecules evolving using the Langevin integrator described in this
1215     work.} \label{fig:bilayer}
1216     \end{figure}
1217    
1218 tim 2746 \section{Conclusions}
1219    
1220 tim 2999 We have presented a new Langevin algorithm by incorporating the
1221     hydrodynamics properties of arbitrary shaped molecules into an
1222 gezelter 3308 advanced symplectic integration scheme. Further studies in systems
1223     involving banana shaped molecules illustrated that the dynamic
1224     properties could be preserved by using this new algorithm as an
1225     implicit solvent model.
1226 tim 2999
1227    
1228 tim 2746 \section{Acknowledgments}
1229     Support for this project was provided by the National Science
1230     Foundation under grant CHE-0134881. T.L. also acknowledges the
1231     financial support from center of applied mathematics at University
1232     of Notre Dame.
1233     \newpage
1234    
1235 gezelter 3305 \bibliographystyle{jcp}
1236 tim 2746 \bibliography{langevin}
1237    
1238     \end{document}