ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3391
Committed: Wed Apr 30 16:16:25 2008 UTC (16 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 67647 byte(s)
Log Message:
Nearing the end

File Contents

# Content
1 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 %\documentclass[aps,prb,preprint,amsmath,amssymb,endfloats]{revtex4}
3 \documentclass[11pt]{article}
4 \usepackage{amsmath}
5 \usepackage{amssymb}
6 \usepackage{times}
7 \usepackage{mathptmx}
8 \usepackage{setspace}
9 \usepackage{endfloat}
10 \usepackage{caption}
11 %\usepackage{tabularx}
12 \usepackage{graphicx}
13 %\usepackage{booktabs}
14 %\usepackage{bibentry}
15 %\usepackage{mathrsfs}
16 \usepackage[ref]{overcite}
17 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
18 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
19 9.0in \textwidth 6.5in \brokenpenalty=10000
20
21 % double space list of tables and figures
22 \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
23 \setlength{\abovecaptionskip}{20 pt}
24 \setlength{\belowcaptionskip}{30 pt}
25
26 \renewcommand\citemid{\ } % no comma in optional referenc note
27
28 \begin{document}
29
30 \title{Langevin dynamics for rigid bodies of arbitrary shape}
31
32 \author{Xiuquan Sun, Teng Lin and J. Daniel
33 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
34 Department of Chemistry and Biochemistry,\\
35 University of Notre Dame\\
36 Notre Dame, Indiana 46556}
37
38 \date{\today}
39
40
41 \maketitle
42
43
44 \begin{doublespace}
45
46 \begin{abstract}
47 We present an algorithm for carrying out Langevin dynamics simulations
48 on complex rigid bodies by incorporating the hydrodynamic resistance
49 tensors for arbitrary shapes into an advanced rotational integration
50 scheme. The integrator gives quantitative agreement with both
51 analytic and approximate hydrodynamic theories for a number of model
52 rigid bodies, and works well at reproducing the solute dynamical
53 properties (diffusion constants, and orientational relaxation times)
54 obtained from explicitly-solvated simulations.
55 \end{abstract}
56
57 \newpage
58
59 %\narrowtext
60
61 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62 % BODY OF TEXT
63 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64
65
66 \section{Introduction}
67
68 %applications of langevin dynamics
69 Langevin dynamics, which mimics a heat bath using both stochastic and
70 dissipative forces, has been applied in a variety of situations as an
71 alternative to molecular dynamics with explicit solvent molecules.
72 The stochastic treatment of the solvent allows the use of simulations
73 with substantially longer time and length scales. In general, the
74 dynamic and structural properties obtained from Langevin simulations
75 agree quite well with similar properties obtained from explicit
76 solvent simulations.
77
78 Recent examples of the usefulness of Langevin simulations include a
79 study of met-enkephalin in which Langevin simulations predicted
80 dynamical properties that were largely in agreement with explicit
81 solvent simulations.\cite{Shen2002} By applying Langevin dynamics with
82 the UNRES model, Liwo and his coworkers suggest that protein folding
83 pathways can be explored within a reasonable amount of
84 time.\cite{Liwo2005}
85
86 The stochastic nature of Langevin dynamics also enhances the sampling
87 of the system and increases the probability of crossing energy
88 barriers.\cite{Cui2003,Banerjee2004} Combining Langevin dynamics with
89 Kramers' theory, Klimov and Thirumalai identified free-energy
90 barriers by studying the viscosity dependence of the protein folding
91 rates.\cite{Klimov1997} In order to account for solvent induced
92 interactions missing from the implicit solvent model, Kaya
93 incorporated a desolvation free energy barrier into protein
94 folding/unfolding studies and discovered a higher free energy barrier
95 between the native and denatured states.\cite{HuseyinKaya07012005}
96
97 In typical LD simulations, the friction and random ($f_r$) forces on
98 individual atoms are taken from Stokes' law,
99 \begin{eqnarray}
100 m \dot{v}(t) & = & -\nabla U(x) - \xi m v(t) + f_r(t) \notag \\
101 \langle f_r(t) \rangle & = & 0 \\
102 \langle f_r(t) f_r(t') \rangle & = & 2 k_B T \xi m \delta(t - t') \notag
103 \end{eqnarray}
104 where $\xi \approx 6 \pi \eta \rho$. Here $\eta$ is the viscosity of the
105 implicit solvent, and $\rho$ is the hydrodynamic radius of the atom.
106
107 The use of rigid substructures,\cite{Chun:2000fj}
108 coarse-graining,\cite{Ayton01,Golubkov06,Orlandi:2006fk,SunX._jp0762020}
109 and ellipsoidal representations of protein side
110 chains~\cite{Fogolari:1996lr} has made the use of the Stokes-Einstein
111 approximation problematic. A rigid substructure moves as a single
112 unit with orientational as well as translational degrees of freedom.
113 This requires a more general treatment of the hydrodynamics than the
114 spherical approximation provides. Also, the atoms involved in a rigid
115 or coarse-grained structure have solvent-mediated interactions with
116 each other, and these interactions are ignored if all atoms are
117 treated as separate spherical particles. The theory of interactions
118 {\it between} bodies moving through a fluid has been developed over
119 the past century and has been applied to simulations of Brownian
120 motion.\cite{FIXMAN:1986lr,Ramachandran1996}
121
122 In order to account for the diffusion anisotropy of complex shapes,
123 Fernandes and Garc\'{i}a de la Torre improved an earlier Brownian
124 dynamics simulation algorithm~\cite{Ermak1978,Allison1991} by
125 incorporating a generalized $6\times6$ diffusion tensor and
126 introducing a rotational evolution scheme consisting of three
127 consecutive rotations.\cite{Fernandes2002} Unfortunately, biases are
128 introduced into the system due to the arbitrary order of applying the
129 noncommuting rotation operators.\cite{Beard2003} The effects of these
130 biases can be reduced by decreasing the size of the time step. Based
131 on the observation the momentum relaxation time is much less than the
132 time step, one may ignore the inertia in Brownian dynamics. However,
133 the assumption of zero average acceleration is not always true for
134 cooperative motion which is common in proteins. An inertial Brownian
135 dynamics (IBD) was proposed to address this issue by adding an
136 inertial correction term.\cite{Beard2000} As a complement to IBD,
137 which has a lower bound in time step because of the inertial
138 relaxation time, long-time-step inertial dynamics (LTID) can be used
139 to investigate the inertial behavior of linked polymer segments in a
140 low friction regime.\cite{Beard2000} LTID can also deal with the
141 rotational dynamics for nonskew bodies without translation-rotation
142 coupling by separating the translation and rotation motion and taking
143 advantage of the analytical solution of hydrodynamic
144 properties. However, typical nonskew bodies like cylinders and
145 ellipsoids are inadequate to represent most complex macromolecular
146 assemblies. Therefore, the goal of this work is to adapt some of the
147 hydrodynamic methodologies developed to treat Brownian motion of
148 complex assemblies into a Langevin integrator for rigid bodies with
149 arbitrary shapes.
150
151 \subsection{Rigid Body Dynamics}
152 Rigid bodies are frequently involved in the modeling of large
153 collections of particles that move as a single unit. In molecular
154 simulations, rigid bodies have been used to simplify protein-protein
155 docking,\cite{Gray2003} and lipid bilayer
156 simulations.\cite{SunX._jp0762020} Many of the water models in common
157 use are also rigid-body
158 models,\cite{Jorgensen83,Berendsen81,Berendsen87} although they are
159 typically evolved in molecular dynamics simulations using constraints
160 rather than rigid body equations of motion.
161
162 Euler angles are a natural choice to describe the rotational degrees
163 of freedom. However, due to $\frac{1}{\sin \theta}$ singularities, the
164 numerical integration of corresponding equations of these motion can
165 become inaccurate (and inefficient). Although the use of multiple
166 sets of Euler angles can overcome this problem,\cite{Barojas1973} the
167 computational penalty and the loss of angular momentum conservation
168 remain. A singularity-free representation utilizing quaternions was
169 developed by Evans in 1977.\cite{Evans1977} The Evans quaternion
170 approach uses a nonseparable Hamiltonian, and this has prevented
171 symplectic algorithms from being utilized until very
172 recently.\cite{Miller2002}
173
174 Another approach is the application of holonomic constraints to the
175 atoms belonging to the rigid body. Each atom moves independently
176 under the normal forces deriving from potential energy and constraints
177 are used to guarantee rigidity. However, due to their iterative
178 nature, the SHAKE and RATTLE algorithms converge very slowly when the
179 number of constraints (and the number of particles that belong to the
180 rigid body) increases.\cite{Ryckaert1977,Andersen1983}
181
182 In order to develop a stable and efficient integration scheme that
183 preserves most constants of the motion in microcanonical simulations,
184 symplectic propagators are necessary. By introducing a conjugate
185 momentum to the rotation matrix ${\bf Q}$ and re-formulating
186 Hamilton's equations, a symplectic orientational integrator,
187 RSHAKE,\cite{Kol1997} was proposed to evolve rigid bodies on a
188 constraint manifold by iteratively satisfying the orthogonality
189 constraint ${\bf Q}^T {\bf Q} = 1$. An alternative method using the
190 quaternion representation was developed by Omelyan.\cite{Omelyan1998}
191 However, both of these methods are iterative and suffer from some
192 related inefficiencies. A symplectic Lie-Poisson integrator for rigid
193 bodies developed by Dullweber {\it et al.}\cite{Dullweber1997} removes
194 most of the limitations mentioned above and is therefore the basis for
195 our Langevin integrator.
196
197 The goal of the present work is to develop a Langevin dynamics
198 algorithm for arbitrary-shaped rigid particles by integrating accurate
199 estimates of the friction tensor that can be obtaind from classical
200 hydrodynamics theory (or other recently-developed hydrodynamic
201 procedures) into a stable and efficient rigid body dynamics
202 propagator. In the sections below, we review some of the theory of
203 hydrodynamic tensors developed primarily for Brownian simulations of
204 multi-particle systems, we then present our integration method for a
205 set of generalized Langevin equations of motion, and we compare the
206 behavior of the new Langevin integrator to dynamical quantities
207 obtained via explicit solvent molecular dynamics.
208
209 \subsection{\label{introSection:frictionTensor}The Friction Tensor}
210 Theoretically, a complete friction kernel for a solute particle can be
211 determined using the velocity autocorrelation function from a
212 simulation with explicit solvent molecules. However, this approach
213 becomes impractical when the solute becomes complex. Instead, various
214 approaches based on hydrodynamics have been developed to calculate
215 static friction coefficients. In general, the friction tensor $\Xi$ is
216 a $6\times 6$ matrix given by
217 \begin{equation}
218 \Xi = \left( \begin{array}{*{20}c}
219 \Xi^{tt} & \Xi^{rt} \\
220 \Xi^{tr} & \Xi^{rr} \\
221 \end{array} \right).
222 \end{equation}
223 Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
224 rotational resistance (friction) tensors respectively, while
225 $\Xi^{tr}$ is translation-rotation coupling tensor and $\Xi^{rt}$ is
226 rotation-translation coupling tensor. When a particle moves in a
227 fluid, it may experience a friction force ($\mathbf{f}_f$) and torque
228 ($\mathbf{\tau}_f$) in opposition to the velocity ($\mathbf{v}$) and
229 body-fixed angular velocity ($\mathbf{\omega}$),
230 \begin{equation}
231 \left( \begin{array}{l}
232 \mathbf{f}_f \\
233 \mathbf{\tau}_f \\
234 \end{array} \right) = - \left( \begin{array}{*{20}c}
235 \Xi^{tt} & \Xi^{rt} \\
236 \Xi^{tr} & \Xi^{rr} \\
237 \end{array} \right)\left( \begin{array}{l}
238 \mathbf{v} \\
239 \mathbf{\omega} \\
240 \end{array} \right).
241 \end{equation}
242 For an arbitrary body moving in a fluid, Peters has derived a set of
243 fluctuation-dissipation relations for the friction
244 tensors,\cite{Peters:1999qy,Peters:1999uq,Peters:2000fk}
245 \begin{eqnarray}
246 \Xi^{tt} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
247 F}(0) {\bf F}(-s) \rangle_{eq} - \langle {\bf F} \rangle_{eq}^2
248 \right] ds \\
249 \notag \\
250 \Xi^{tr} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
251 F}(0) {\bf \tau}(-s) \rangle_{eq} - \langle {\bf F} \rangle_{eq}
252 \langle {\bf \tau} \rangle_{eq} \right] ds \\
253 \notag \\
254 \Xi^{rt} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
255 \tau}(0) {\bf F}(-s) \rangle_{eq} - \langle {\bf \tau} \rangle_{eq}
256 \langle {\bf F} \rangle_{eq} \right] ds \\
257 \notag \\
258 \Xi^{rr} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
259 \tau}(0) {\bf \tau}(-s) \rangle_{eq} - \langle {\bf \tau} \rangle_{eq}^2
260 \right] ds
261 \end{eqnarray}
262 In these expressions, the forces (${\bf F}$) and torques (${\bf
263 \tau}$) are those that arise solely from the interactions of the body with
264 the surrounding fluid. For a single solute body in an isotropic fluid,
265 the average forces and torques in these expressions ($\langle {\bf F}
266 \rangle_{eq}$ and $\langle {\bf \tau} \rangle_{eq}$)
267 vanish, and one obtains the simpler force-torque correlation formulae
268 of Nienhuis.\cite{Nienhuis:1970lr} Molecular dynamics simulations with
269 explicit solvent molecules can be used to obtain estimates of the
270 friction tensors with these formulae. In practice, however, one needs
271 relatively long simulations with frequently-stored force and torque
272 information to compute friction tensors, and this becomes
273 prohibitively expensive when there are large numbers of large solute
274 particles. For bodies with simple shapes, there are a number of
275 approximate expressions that allow computation of these tensors
276 without the need for expensive simulations that utilize explicit
277 solvent particles.
278
279 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
280 For a spherical body under ``stick'' boundary conditions, the
281 translational and rotational friction tensors can be estimated from
282 Stokes' law,
283 \begin{equation}
284 \label{eq:StokesTranslation}
285 \Xi^{tt} = \left( \begin{array}{*{20}c}
286 {6\pi \eta \rho} & 0 & 0 \\
287 0 & {6\pi \eta \rho} & 0 \\
288 0 & 0 & {6\pi \eta \rho} \\
289 \end{array} \right)
290 \end{equation}
291 and
292 \begin{equation}
293 \label{eq:StokesRotation}
294 \Xi^{rr} = \left( \begin{array}{*{20}c}
295 {8\pi \eta \rho^3 } & 0 & 0 \\
296 0 & {8\pi \eta \rho^3 } & 0 \\
297 0 & 0 & {8\pi \eta \rho^3 } \\
298 \end{array} \right)
299 \end{equation}
300 where $\eta$ is the viscosity of the solvent and $\rho$ is the
301 hydrodynamic radius. The presence of the rotational resistance tensor
302 implies that the spherical body has internal structure and
303 orientational degrees of freedom that must be propagated in time. For
304 non-structured spherical bodies (i.e. the atoms in a traditional
305 molecular dynamics simulation) these degrees of freedom do not exist.
306
307 Other non-spherical shapes, such as cylinders and ellipsoids, are
308 widely used as references for developing new hydrodynamic theories,
309 because their properties can be calculated exactly. In 1936, Perrin
310 extended Stokes' law to general
311 ellipsoids,\cite{Perrin1934,Perrin1936} described in Cartesian
312 coordinates as
313 \begin{equation}
314 \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1.
315 \end{equation}
316 Here, the semi-axes are of lengths $a$, $b$, and $c$. Due to the
317 complexity of the elliptic integral, only uniaxial ellipsoids, either
318 prolate ($a \ge b = c$) or oblate ($a < b = c$), were solved
319 exactly. Introducing an elliptic integral parameter $S$ for prolate,
320 \begin{equation}
321 S = \frac{2}{\sqrt{a^2 - b^2}} \ln \frac{a + \sqrt{a^2 - b^2}}{b},
322 \end{equation}
323 and oblate,
324 \begin{equation}
325 S = \frac{2}{\sqrt {b^2 - a^2 }} \arctan \frac{\sqrt {b^2 - a^2}}{a},
326 \end{equation}
327 ellipsoids, it is possible to write down exact solutions for the
328 resistance tensors. As is the case for spherical bodies, the translational,
329 \begin{eqnarray}
330 \Xi_a^{tt} & = & 16\pi \eta \frac{a^2 - b^2}{(2a^2 - b^2 )S - 2a}. \\
331 \Xi_b^{tt} = \Xi_c^{tt} & = & 32\pi \eta \frac{a^2 - b^2 }{(2a^2 - 3b^2 )S + 2a},
332 \end{eqnarray}
333 and rotational,
334 \begin{eqnarray}
335 \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2 - b^2 )b^2}{2a - b^2 S}, \\
336 \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4 - b^4)}{(2a^2 - b^2 )S - 2a}
337 \end{eqnarray}
338 resistance tensors are diagonal $3 \times 3$ matrices. For both
339 spherical and ellipsoidal particles, the translation-rotation and
340 rotation-translation coupling tensors are zero.
341
342 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
343 Other than the fluctuation dissipation formulae given by
344 Peters,\cite{Peters:1999qy,Peters:1999uq,Peters:2000fk} there are no
345 analytic solutions for the friction tensor for rigid molecules of
346 arbitrary shape. The ellipsoid of revolution and general triaxial
347 ellipsoid models have been widely used to approximate the hydrodynamic
348 properties of rigid bodies. However, the mapping from all possible
349 ellipsoidal spaces ($r$-space) to all possible combinations of
350 rotational diffusion coefficients ($D$-space) is not
351 unique.\cite{Wegener1979} Additionally, because there is intrinsic
352 coupling between translational and rotational motion of {\it skew}
353 rigid bodies, general ellipsoids are not always suitable for modeling
354 rigid molecules. A number of studies have been devoted to determining
355 the friction tensor for irregular shapes using methods in which the
356 molecule of interest is modeled with a combination of
357 spheres\cite{Carrasco1999} and the hydrodynamic properties of the
358 molecule are then calculated using a set of two-point interaction
359 tensors. We have found the {\it bead} and {\it rough shell} models of
360 Carrasco and Garc\'{i}a de la Torre to be the most useful of these
361 methods,\cite{Carrasco1999} and we review the basic outline of the
362 rough shell approach here. A more thorough explanation can be found
363 in Ref. \citen{Carrasco1999}.
364
365 Consider a rigid assembly of $N$ small beads moving through a
366 continuous medium. Due to hydrodynamic interactions between the
367 beads, the net velocity of the $i^\mathrm{th}$ bead relative to the
368 medium, ${\bf v}'_i$, is different than its unperturbed velocity ${\bf
369 v}_i$,
370 \begin{equation}
371 {\bf v}'_i = {\bf v}_i - \sum\limits_{j \ne i} {{\bf T}_{ij} {\bf F}_j }
372 \end{equation}
373 where ${\bf F}_j$ is the frictional force on the medium due to bead $j$, and
374 ${\bf T}_{ij}$ is the hydrodynamic interaction tensor between the two beads.
375 The frictional force felt by the $i^\mathrm{th}$ bead is proportional to
376 its net velocity
377 \begin{equation}
378 {\bf F}_i = \xi_i {\bf v}_i - \xi_i \sum\limits_{j \ne i} {{\bf T}_{ij} {\bf F}_j }.
379 \label{introEquation:tensorExpression}
380 \end{equation}
381 Eq. (\ref{introEquation:tensorExpression}) defines the two-point
382 hydrodynamic tensor, ${\bf T}_{ij}$. There have been many proposed
383 solutions to this equation, including the simple solution given by
384 Oseen and Burgers in 1930 for two beads of identical radius. A second
385 order expression for beads of different hydrodynamic radii was
386 introduced by Rotne and Prager,\cite{Rotne1969} and improved by
387 Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
388 \begin{equation}
389 {\bf T}_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {{\bf I} +
390 \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right) + \frac{{\rho
391 _i^2 + \rho_j^2 }}{{R_{ij}^2 }}\left( {\frac{{\bf I}}{3} -
392 \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
393 \label{introEquation:RPTensorNonOverlapped}
394 \end{equation}
395 Here ${\bf R}_{ij}$ is the distance vector between beads $i$ and $j$. Both
396 the Oseen-Burgers tensor and
397 Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption that
398 the beads do not overlap ($R_{ij} \ge \rho_i + \rho_j$).
399
400 To calculate the resistance tensor for a body represented as the union
401 of many non-overlapping beads, we first pick an arbitrary origin $O$
402 and then construct a $3N \times 3N$ supermatrix consisting of $N
403 \times N$ ${\bf B}_{ij}$ blocks
404 \begin{equation}
405 {\bf B} = \left( \begin{array}{*{20}c}
406 {\bf B}_{11} & \ldots & {\bf B}_{1N} \\
407 \vdots & \ddots & \vdots \\
408 {\bf B}_{N1} & \cdots & {\bf B}_{NN}
409 \end{array} \right)
410 \end{equation}
411 ${\bf B}_{ij}$ is a version of the hydrodynamic tensor which includes the
412 self-contributions for spheres,
413 \begin{equation}
414 {\bf B}_{ij} = \delta _{ij} \frac{{\bf I}}{{6\pi \eta R_{ij}}} + (1 - \delta_{ij}
415 ){\bf T}_{ij}
416 \end{equation}
417 where $\delta_{ij}$ is the Kronecker delta function. Inverting the
418 ${\bf B}$ matrix, we obtain
419 \begin{equation}
420 {\bf C} = {\bf B}^{ - 1} = \left(\begin{array}{*{20}c}
421 {\bf C}_{11} & \ldots & {\bf C}_{1N} \\
422 \vdots & \ddots & \vdots \\
423 {\bf C}_{N1} & \cdots & {\bf C}_{NN}
424 \end{array} \right),
425 \end{equation}
426 which can be partitioned into $N \times N$ blocks labeled ${\bf C}_{ij}$.
427 (Each of the ${\bf C}_{ij}$ blocks is a $3 \times 3$ matrix.) Using the
428 skew matrix,
429 \begin{equation}
430 {\bf U}_i = \left(\begin{array}{*{20}c}
431 0 & -z_i & y_i \\
432 z_i & 0 & - x_i \\
433 -y_i & x_i & 0
434 \end{array}\right)
435 \label{eq:skewMatrix}
436 \end{equation}
437 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
438 bead $i$ and origin $O$, the elements of the resistance tensor (at the
439 arbitrary origin $O$) can be written as
440 \begin{eqnarray}
441 \label{introEquation:ResistanceTensorArbitraryOrigin}
442 \Xi^{tt} & = & \sum\limits_i {\sum\limits_j {{\bf C}_{ij} } } \notag , \\
443 \Xi^{tr} = \Xi _{}^{rt} & = & \sum\limits_i {\sum\limits_j {{\bf U}_i {\bf C}_{ij} } } , \\
444 \Xi^{rr} & = & -\sum\limits_i \sum\limits_j {\bf U}_i {\bf C}_{ij} {\bf U}_j + 6 \eta V {\bf I}. \notag
445 \end{eqnarray}
446 The final term in the expression for $\Xi^{rr}$ is a correction that
447 accounts for errors in the rotational motion of the bead models. The
448 additive correction uses the solvent viscosity ($\eta$) as well as the
449 total volume of the beads that contribute to the hydrodynamic model,
450 \begin{equation}
451 V = \frac{4 \pi}{3} \sum_{i=1}^{N} \rho_i^3,
452 \end{equation}
453 where $\rho_i$ is the radius of bead $i$. This correction term was
454 rigorously tested and compared with the analytical results for
455 two-sphere and ellipsoidal systems by Garc\'{i}a de la Torre and
456 Rodes.\cite{Torre:1983lr}
457
458 In general, resistance tensors depend on the origin at which they were
459 computed. However, the proper location for applying the friction
460 force is the center of resistance, the special point at which the
461 trace of rotational resistance tensor, $\Xi^{rr}$ reaches a minimum
462 value. Mathematically, the center of resistance can also be defined
463 as the unique point for a rigid body at which the translation-rotation
464 coupling tensors are symmetric,
465 \begin{equation}
466 \Xi^{tr} = \left(\Xi^{tr}\right)^T
467 \label{introEquation:definitionCR}
468 \end{equation}
469 From Eq. \ref{introEquation:ResistanceTensorArbitraryOrigin}, we can
470 easily derive that the {\it translational} resistance tensor is origin
471 independent, while the rotational resistance tensor and
472 translation-rotation coupling resistance tensor depend on the
473 origin. Given the resistance tensor at an arbitrary origin $O$, and a
474 vector ,${\bf r}_{OP} = (x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we
475 can obtain the resistance tensor at $P$ by
476 \begin{eqnarray}
477 \label{introEquation:resistanceTensorTransformation}
478 \Xi_P^{tt} & = & \Xi_O^{tt} \notag \\
479 \Xi_P^{tr} = \Xi_P^{rt} & = & \Xi_O^{tr} - {\bf U}_{OP} \Xi _O^{tt} \\
480 \Xi_P^{rr} & = &\Xi_O^{rr} - {\bf U}_{OP} \Xi_O^{tt} {\bf U}_{OP}
481 + \Xi_O^{tr} {\bf U}_{OP} - {\bf U}_{OP} \left( \Xi_O^{tr}
482 \right)^{^T} \notag
483 \end{eqnarray}
484 where ${\bf U}_{OP}$ is the skew matrix (Eq. (\ref{eq:skewMatrix}))
485 for the vector between the origin $O$ and the point $P$. Using
486 Eqs.~\ref{introEquation:definitionCR}~and~\ref{introEquation:resistanceTensorTransformation},
487 one can locate the position of center of resistance,
488 \begin{equation*}
489 \left(\begin{array}{l}
490 x_{OR} \\
491 y_{OR} \\
492 z_{OR}
493 \end{array}\right) =
494 \left(\begin{array}{*{20}c}
495 (\Xi_O^{rr})_{yy} + (\Xi_O^{rr})_{zz} & -(\Xi_O^{rr})_{xy} & -(\Xi_O^{rr})_{xz} \\
496 -(\Xi_O^{rr})_{xy} & (\Xi_O^{rr})_{zz} + (\Xi_O^{rr})_{xx} & -(\Xi_O^{rr})_{yz} \\
497 -(\Xi_O^{rr})_{xz} & -(\Xi_O^{rr})_{yz} & (\Xi_O^{rr})_{xx} + (\Xi_O^{rr})_{yy} \\
498 \end{array}\right)^{-1}
499 \left(\begin{array}{l}
500 (\Xi_O^{tr})_{yz} - (\Xi_O^{tr})_{zy} \\
501 (\Xi_O^{tr})_{zx} - (\Xi_O^{tr})_{xz} \\
502 (\Xi_O^{tr})_{xy} - (\Xi_O^{tr})_{yx}
503 \end{array}\right)
504 \end{equation*}
505 where $x_{OR}$, $y_{OR}$, $z_{OR}$ are the components of the vector
506 joining center of resistance $R$ and origin $O$.
507
508 For a general rigid molecular substructure, finding the $6 \times 6$
509 resistance tensor can be a computationally demanding task. First, a
510 lattice of small beads that extends well beyond the boundaries of the
511 rigid substructure is created. The lattice is typically composed of
512 0.25 \AA\ beads on a dense FCC lattice. The lattice constant is taken
513 to be the bead diameter, so that adjacent beads are touching, but do
514 not overlap. To make a shape corresponding to the rigid structure,
515 beads that sit on lattice sites that are outside the van der Waals
516 radii of all of the atoms comprising the rigid body are excluded from
517 the calculation.
518
519 For large structures, most of the beads will be deep within the rigid
520 body and will not contribute to the hydrodynamic tensor. In the {\it
521 rough shell} approach, beads which have all of their lattice neighbors
522 inside the structure are considered interior beads, and are removed
523 from the calculation. After following this procedure, only those
524 beads in direct contact with the van der Waals surface of the rigid
525 body are retained. For reasonably large molecular structures, this
526 truncation can still produce bead assemblies with thousands of
527 members.
528
529 If all of the {\it atoms} comprising the rigid substructure are
530 spherical and non-overlapping, the tensor in
531 Eq.~(\ref{introEquation:RPTensorNonOverlapped}) may be used directly
532 using the atoms themselves as the hydrodynamic beads. This is a
533 variant of the {\it bead model} approach of Carrasco and Garc\'{i}a de
534 la Torre.\cite{Carrasco1999} In this case, the size of the ${\bf B}$
535 matrix can be quite small, and the calculation of the hydrodynamic
536 tensor is straightforward.
537
538 In general, the inversion of the ${\bf B}$ matrix is the most
539 computationally demanding task. This inversion is done only once for
540 each type of rigid structure. We have used straightforward
541 LU-decomposition to solve the linear system and to obtain the elements
542 of ${\bf C}$. Once ${\bf C}$ has been obtained, the location of the
543 center of resistance ($R$) is found and the resistance tensor at this
544 point is calculated. The $3 \times 1$ vector giving the location of
545 the rigid body's center of resistance and the $6 \times 6$ resistance
546 tensor are then stored for use in the Langevin dynamics calculation.
547 These quantities depend on solvent viscosity and temperature and must
548 be recomputed if different simulation conditions are required.
549
550 \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
551
552 Consider the Langevin equations of motion in generalized coordinates
553 \begin{equation}
554 {\bf M} \dot{{\bf V}}(t) = {\bf F}_{s}(t) +
555 {\bf F}_{f}(t) + {\bf F}_{r}(t)
556 \label{LDGeneralizedForm}
557 \end{equation}
558 where ${\bf M}$ is a $6 \times 6$ diagonal mass matrix (which
559 includes the mass of the rigid body as well as the moments of inertia
560 in the body-fixed frame) and ${\bf V}$ is a generalized velocity,
561 ${\bf V} =
562 \left\{{\bf v},{\bf \omega}\right\}$. The right side of
563 Eq.~\ref{LDGeneralizedForm} consists of three generalized forces: a
564 system force (${\bf F}_{s}$), a frictional or dissipative force (${\bf
565 F}_{f}$) and a stochastic force (${\bf F}_{r}$). While the evolution
566 of the system in Newtonian mechanics is typically done in the lab
567 frame, it is convenient to handle the dynamics of rigid bodies in
568 body-fixed frames. Thus the friction and random forces on each
569 substructure are calculated in a body-fixed frame and may converted
570 back to the lab frame using that substructure's rotation matrix (${\bf
571 Q}$):
572 \begin{equation}
573 {\bf F}_{f,r} =
574 \left( \begin{array}{c}
575 {\bf f}_{f,r} \\
576 {\bf \tau}_{f,r}
577 \end{array} \right)
578 =
579 \left( \begin{array}{c}
580 {\bf Q}^{T} {\bf f}^{~b}_{f,r} \\
581 {\bf Q}^{T} {\bf \tau}^{~b}_{f,r}
582 \end{array} \right)
583 \end{equation}
584 The body-fixed friction force, ${\bf F}_{f}^{~b}$, is proportional to
585 the (body-fixed) velocity at the center of resistance
586 ${\bf v}_{R}^{~b}$ and the angular velocity ${\bf \omega}$
587 \begin{equation}
588 {\bf F}_{f}^{~b}(t) = \left( \begin{array}{l}
589 {\bf f}_{f}^{~b}(t) \\
590 {\bf \tau}_{f}^{~b}(t) \\
591 \end{array} \right) = - \left( \begin{array}{*{20}c}
592 \Xi_{R}^{tt} & \Xi_{R}^{rt} \\
593 \Xi_{R}^{tr} & \Xi_{R}^{rr} \\
594 \end{array} \right)\left( \begin{array}{l}
595 {\bf v}_{R}^{~b}(t) \\
596 {\bf \omega}(t) \\
597 \end{array} \right),
598 \end{equation}
599 while the random force, ${\bf F}_{r}$, is a Gaussian stochastic
600 variable with zero mean and variance,
601 \begin{equation}
602 \left\langle {{\bf F}_{r}(t) ({\bf F}_{r}(t'))^T } \right\rangle =
603 \left\langle {{\bf F}_{r}^{~b} (t) ({\bf F}_{r}^{~b} (t'))^T } \right\rangle =
604 2 k_B T \Xi_R \delta(t - t'). \label{eq:randomForce}
605 \end{equation}
606 $\Xi_R$ is the $6\times6$ resistance tensor at the center of
607 resistance. Once this tensor is known for a given rigid body (as
608 described in the previous section) obtaining a stochastic vector that
609 has the properties in Eq. (\ref{eq:randomForce}) can be done
610 efficiently by carrying out a one-time Cholesky decomposition to
611 obtain the square root matrix of the resistance tensor,
612 \begin{equation}
613 \Xi_R = {\bf S} {\bf S}^{T},
614 \label{eq:Cholesky}
615 \end{equation}
616 where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
617 vector with the statistics required for the random force can then be
618 obtained by multiplying ${\bf S}$ onto a random 6-vector ${\bf Z}$ which
619 has elements chosen from a Gaussian distribution, such that:
620 \begin{equation}
621 \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
622 {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
623 \end{equation}
624 where $\delta t$ is the timestep in use during the simulation. The
625 random force, ${\bf F}_{r}^{~b} = {\bf S} {\bf Z}$, can be shown to have the
626 correct properties required by Eq. (\ref{eq:randomForce}).
627
628 The equation of motion for the translational velocity of the center of
629 mass (${\bf v}$) can be written as
630 \begin{equation}
631 m \dot{{\bf v}} (t) = {\bf f}_{s}(t) + {\bf f}_{f}(t) +
632 {\bf f}_{r}(t)
633 \end{equation}
634 Since the frictional and random forces are applied at the center of
635 resistance, which generally does not coincide with the center of mass,
636 extra torques are exerted at the center of mass. Thus, the net
637 body-fixed torque at the center of mass, $\tau^{~b}(t)$,
638 is given by
639 \begin{equation}
640 \tau^{~b} \leftarrow \tau_{s}^{~b} + \tau_{f}^{~b} + \tau_{r}^{~b} + {\bf r}_{MR} \times \left( {\bf f}_{f}^{~b} + {\bf f}_{r}^{~b} \right)
641 \end{equation}
642 where ${\bf r}_{MR}$ is the vector from the center of mass to the center of
643 resistance. Instead of integrating the angular velocity in lab-fixed
644 frame, we consider the equation of motion for the angular momentum
645 (${\bf j}$) in the body-fixed frame
646 \begin{equation}
647 \frac{\partial}{\partial t}{\bf j}(t) = \tau^{~b}(t)
648 \end{equation}
649 Embedding the friction and random forces into the the total force and
650 torque, one can integrate the Langevin equations of motion for a rigid
651 body of arbitrary shape in a velocity-Verlet style 2-part algorithm,
652 where $h = \delta t$:
653
654 {\tt move A:}
655 \begin{align*}
656 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
657 + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
658 %
659 {\bf r}(t + h) &\leftarrow {\bf r}(t)
660 + h {\bf v}\left(t + h / 2 \right), \\
661 %
662 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
663 + \frac{h}{2} {\bf \tau}^{~b}(t), \\
664 %
665 {\bf Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
666 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
667 \end{align*}
668 In this context, $\overleftrightarrow{\mathsf{I}}$ is the diagonal
669 moment of inertia tensor, and the $\mathrm{rotate}$ function is the
670 reversible product of the three body-fixed rotations,
671 \begin{equation}
672 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
673 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
674 / 2) \cdot \mathsf{G}_x(a_x /2),
675 \end{equation}
676 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
677 rotates both the rotation matrix ($\mathbf{Q}$) and the body-fixed
678 angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
679 axis $\alpha$,
680 \begin{equation}
681 \mathsf{G}_\alpha( \theta ) = \left\{
682 \begin{array}{lcl}
683 \mathbf{Q}(t) & \leftarrow & \mathbf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
684 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
685 j}(0).
686 \end{array}
687 \right.
688 \end{equation}
689 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
690 rotation matrix. For example, in the small-angle limit, the
691 rotation matrix around the body-fixed x-axis can be approximated as
692 \begin{equation}
693 \mathsf{R}_x(\theta) \approx \left(
694 \begin{array}{ccc}
695 1 & 0 & 0 \\
696 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
697 \theta^2 / 4} \\
698 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
699 \theta^2 / 4}
700 \end{array}
701 \right).
702 \end{equation}
703 All other rotations follow in a straightforward manner. After the
704 first part of the propagation, the forces and body-fixed torques are
705 calculated at the new positions and orientations. The system forces
706 and torques are derivatives of the total potential energy function
707 ($U$) with respect to the rigid body positions (${\bf r}$) and the
708 columns of the transposed rotation matrix ${\bf Q}^T = \left({\bf
709 u}_x, {\bf u}_y, {\bf u}_z \right)$:
710
711 {\tt Forces:}
712 \begin{align*}
713 {\bf f}_{s}(t + h) & \leftarrow
714 - \left(\frac{\partial U}{\partial {\bf r}}\right)_{{\bf r}(t + h)} \\
715 %
716 {\bf \tau}_{s}(t + h) &\leftarrow {\bf u}(t + h)
717 \times \frac{\partial U}{\partial {\bf u}} \\
718 %
719 {\bf v}^{b}_{R}(t+h) & \leftarrow \mathbf{Q}(t+h) \cdot \left({\bf v}(t+h) + {\bf \omega}(t+h) \times {\bf r}_{MR} \right) \\
720 %
721 {\bf f}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tt} \cdot
722 {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rt} \cdot {\bf \omega}(t+h) \\
723 %
724 {\bf \tau}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tr} \cdot
725 {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rr} \cdot {\bf \omega}(t+h) \\
726 %
727 Z & \leftarrow {\tt GaussianNormal}(2 k_B T / h, 6) \\
728 {\bf F}_{R,r}^{b}(t+h) & \leftarrow {\bf S} \cdot Z \\
729 %
730 {\bf f}(t+h) & \leftarrow {\bf f}_{s}(t+h) + \mathbf{Q}^{T}(t+h)
731 \cdot \left({\bf f}_{R,f}^{~b} + {\bf f}_{R,r}^{~b} \right) \\
732 %
733 \tau(t+h) & \leftarrow \tau_{s}(t+h) + \mathbf{Q}^{T}(t+h) \cdot \left(\tau_{R,f}^{~b} + \tau_{R,r}^{~b} \right) + {\bf r}_{MR} \times \left({\bf f}_{f}(t+h) + {\bf f}_{r}(t+h) \right) \\
734 \tau^{~b}(t+h) & \leftarrow \mathbf{Q}(t+h) \cdot \tau(t+h) \\
735 \end{align*}
736 Frictional (and random) forces and torques must be computed at the
737 center of resistance, so there are additional steps required to find
738 the body-fixed velocity (${\bf v}_{R}^{~b}$) at this location. Mapping
739 the frictional and random forces at the center of resistance back to
740 the center of mass also introduces an additional term in the torque
741 one obtains at the center of mass.
742
743 Once the forces and torques have been obtained at the new time step,
744 the velocities can be advanced to the same time value.
745
746 {\tt move B:}
747 \begin{align*}
748 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
749 \right)
750 + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
751 %
752 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
753 \right)
754 + \frac{h}{2} {\bf \tau}^{~b}(t + h) .
755 \end{align*}
756
757 \section{Validating the Method\label{sec:validating}}
758 In order to validate our Langevin integrator for arbitrarily-shaped
759 rigid bodies, we implemented the algorithm in OOPSE\cite{Meineke2005}
760 and compared the results of this algorithm with the known hydrodynamic
761 limiting behavior for a few model systems, and to microcanonical
762 molecular dynamics simulations for some more complicated bodies. Code
763 to estimate the hydrodynamic tensors using the bead and rough shell
764 models was also implemented in OOPSE, although a number of other
765 programs exist that can perform these
766 tasks.\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002}
767
768 The model systems and their analytical behavior
769 (if known) are summarized below. Parameters for the primary particles
770 comprising our model systems are given in table \ref{tab:parameters},
771 and a sketch of the arrangement of these primary particles into the
772 model rigid bodies is shown in figure \ref{fig:models}. In table
773 \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
774 ellipsoidal (Gay-Berne) particles. For spherical particles, the value
775 of the Lennard-Jones $\sigma$ parameter is the particle diameter
776 ($d$). Gay-Berne ellipsoids have an energy scaling parameter,
777 $\epsilon^s$, which describes the well depth for two identical
778 ellipsoids in a {\it side-by-side} configuration. Additionally, a
779 well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
780 describes the ratio between the well depths in the {\it end-to-end}
781 and side-by-side configurations. For spheres, $\epsilon^r \equiv 1$.
782 Moments of inertia are also required to describe the motion of primary
783 particles with orientational degrees of freedom.
784
785 \begin{table*}
786 \begin{minipage}{\linewidth}
787 \begin{center}
788 \caption[Parameters for the primary particles in use by the rigid body
789 models in figure \ref{fig:models}.]{}
790 \begin{tabular}{lrcccccccc}
791 \hline
792 & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
793 & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
794 $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
795 Sphere & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
796 Ellipsoid & & 4.6 & 13.8 & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
797 Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
798 Banana &(3 identical ellipsoids)& 4.2 & 11.2 & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
799 Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
800 & Ellipsoidal Tail & 4.6 & 13.8 & 0.8 & 0.2 & 760 & 45000 & 45000 & 9000 \\
801 Solvent & & 4.7 & $= d$ & 0.8 & 1 & 72.06 & & & \\
802 \hline
803 \end{tabular}
804 \label{tab:parameters}
805 \end{center}
806 \end{minipage}
807 \end{table*}
808
809 \begin{figure}
810 \centering
811 \includegraphics[width=3in]{sketch}
812 \caption[A sketch of the model systems used in evaluating the behavior
813 of the rigid body Langevin integrator]{} \label{fig:models}
814 \end{figure}
815
816 \subsection{Simulation Methodology}
817 We performed reference microcanonical simulations with explicit
818 solvents for each of the different model system. In each case there
819 was one solute model and 1929 solvent molecules present in the
820 simulation box. All simulations were equilibrated for 5 ns using a
821 constant-pressure and temperature integrator with target values of 300
822 K for the temperature and 1 atm for pressure. Following this stage,
823 further equilibration (5 ns) and sampling (10 ns) was done in a
824 microcanonical ensemble. Since the model bodies are typically quite
825 massive, we were able to use a time step of 25 fs.
826
827 The model systems studied used both Lennard-Jones spheres as well as
828 uniaxial Gay-Berne ellipoids. In its original form, the Gay-Berne
829 potential was a single site model for the interactions of rigid
830 ellipsoidal molecules.\cite{Gay1981} It can be thought of as a
831 modification of the Gaussian overlap model originally described by
832 Berne and Pechukas.\cite{Berne72} The potential is constructed in the
833 familiar form of the Lennard-Jones function using
834 orientation-dependent $\sigma$ and $\epsilon$ parameters,
835 \begin{equation*}
836 V_{ij}({{\bf \hat u}_i}, {{\bf \hat u}_j}, {{\bf \hat
837 r}_{ij}}) = 4\epsilon ({{\bf \hat u}_i}, {{\bf \hat u}_j},
838 {{\bf \hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u
839 }_i},
840 {{\bf \hat u}_j}, {{\bf \hat r}_{ij}})+\sigma_0}\right)^{12}
841 -\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u}_i}, {{\bf \hat u}_j},
842 {{\bf \hat r}_{ij}})+\sigma_0}\right)^6\right]
843 \label{eq:gb}
844 \end{equation*}
845
846 The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
847 \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
848 \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
849 are dependent on the relative orientations of the two ellipsoids (${\bf
850 \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
851 inter-ellipsoid separation (${\bf \hat{r}}_{ij}$). The shape and
852 attractiveness of each ellipsoid is governed by a relatively small set
853 of parameters: $l$ and $d$ describe the length and width of each
854 uniaxial ellipsoid, while $\epsilon^s$, which describes the well depth
855 for two identical ellipsoids in a {\it side-by-side} configuration.
856 Additionally, a well depth aspect ratio, $\epsilon^r = \epsilon^e /
857 \epsilon^s$, describes the ratio between the well depths in the {\it
858 end-to-end} and side-by-side configurations. Details of the potential
859 are given elsewhere,\cite{Luckhurst90,Golubkov06,SunX._jp0762020} and an
860 excellent overview of the computational methods that can be used to
861 efficiently compute forces and torques for this potential can be found
862 in Ref. \citen{Golubkov06}
863
864 For the interaction between nonequivalent uniaxial ellipsoids (or
865 between spheres and ellipsoids), the spheres are treated as ellipsoids
866 with an aspect ratio of 1 ($d = l$) and with an well depth ratio
867 ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of the
868 Gay-Berne potential we are using was generalized by Cleaver {\it et
869 al.} and is appropriate for dissimilar uniaxial
870 ellipsoids.\cite{Cleaver96}
871
872 A switching function was applied to all potentials to smoothly turn
873 off the interactions between a range of $22$ and $25$ \AA. The
874 switching function was the standard (cubic) function,
875 \begin{equation}
876 s(r) =
877 \begin{cases}
878 1 & \text{if $r \le r_{\text{sw}}$},\\
879 \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
880 {(r_{\text{cut}} - r_{\text{sw}})^3}
881 & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
882 0 & \text{if $r > r_{\text{cut}}$.}
883 \end{cases}
884 \label{eq:switchingFunc}
885 \end{equation}
886
887 To measure shear viscosities from our microcanonical simulations, we
888 used the Einstein form of the pressure correlation function,\cite{hess:209}
889 \begin{equation}
890 \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
891 \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \right\rangle_{t_0}.
892 \label{eq:shear}
893 \end{equation}
894 which converges much more rapidly in molecular dynamics simulations
895 than the traditional Green-Kubo formula.
896
897 The Langevin dynamics for the different model systems were performed
898 at the same temperature as the average temperature of the
899 microcanonical simulations and with a solvent viscosity taken from
900 Eq. (\ref{eq:shear}) applied to these simulations. We used 1024
901 independent solute simulations to obtain statistics on our Langevin
902 integrator.
903
904 \subsection{Analysis}
905
906 The quantities of interest when comparing the Langevin integrator to
907 analytic hydrodynamic equations and to molecular dynamics simulations
908 are typically translational diffusion constants and orientational
909 relaxation times. Translational diffusion constants for point
910 particles are computed easily from the long-time slope of the
911 mean-square displacement,
912 \begin{equation}
913 D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \left\langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \right\rangle,
914 \end{equation}
915 of the solute molecules. For models in which the translational
916 diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
917 (i.e. any non-spherically-symmetric rigid body), it is possible to
918 compute the diffusive behavior for motion parallel to each body-fixed
919 axis by projecting the displacement of the particle onto the
920 body-fixed reference frame at $t=0$. With an isotropic solvent, as we
921 have used in this study, there may be differences between the three
922 diffusion constants at short times, but these must converge to the
923 same value at longer times. Translational diffusion constants for the
924 different shaped models are shown in table \ref{tab:translation}.
925
926 In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
927 diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
928 {\it around} a particular body-fixed axis and {\it not} the diffusion
929 of a vector pointing along the axis. However, these eigenvalues can
930 be combined to find 5 characteristic rotational relaxation
931 times,\cite{PhysRev.119.53,Berne90}
932 \begin{eqnarray}
933 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
934 1 / \tau_2 & = & 6 D_r - 2 \Delta \\
935 1 / \tau_3 & = & 3 (D_r + D_1) \\
936 1 / \tau_4 & = & 3 (D_r + D_2) \\
937 1 / \tau_5 & = & 3 (D_r + D_3)
938 \end{eqnarray}
939 where
940 \begin{equation}
941 D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
942 \end{equation}
943 and
944 \begin{equation}
945 \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
946 \end{equation}
947 Each of these characteristic times can be used to predict the decay of
948 part of the rotational correlation function when $\ell = 2$,
949 \begin{equation}
950 C_2(t) = \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
951 \end{equation}
952 This is the same as the $F^2_{0,0}(t)$ correlation function that
953 appears in Ref. \citen{Berne90}. The amplitudes of the two decay
954 terms are expressed in terms of three dimensionless functions of the
955 eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
956 2\Delta)$, and $N = 2 \sqrt{\Delta b}$. Similar expressions can be
957 obtained for other angular momentum correlation
958 functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
959 studied, only one of the amplitudes of the two decay terms was
960 non-zero, so it was possible to derive a single relaxation time for
961 each of the hydrodynamic tensors. In many cases, these characteristic
962 times are averaged and reported in the literature as a single relaxation
963 time,\cite{Garcia-de-la-Torre:1997qy}
964 \begin{equation}
965 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
966 \end{equation}
967 although for the cases reported here, this averaging is not necessary
968 and only one of the five relaxation times is relevant.
969
970 To test the Langevin integrator's behavior for rotational relaxation,
971 we have compared the analytical orientational relaxation times (if
972 they are known) with the general result from the diffusion tensor and
973 with the results from both the explicitly solvated molecular dynamics
974 and Langevin simulations. Relaxation times from simulations (both
975 microcanonical and Langevin), were computed using Legendre polynomial
976 correlation functions for a unit vector (${\bf u}$) fixed along one or
977 more of the body-fixed axes of the model.
978 \begin{equation}
979 C_{\ell}(t) = \left\langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
980 u}_{i}(0) \right) \right\rangle
981 \end{equation}
982 For simulations in the high-friction limit, orientational correlation
983 times can then be obtained from exponential fits of this function, or by
984 integrating,
985 \begin{equation}
986 \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
987 \end{equation}
988 In lower-friction solvents, the Legendre correlation functions often
989 exhibit non-exponential decay, and may not be characterized by a
990 single decay constant.
991
992 In table \ref{tab:rotation} we show the characteristic rotational
993 relaxation times (based on the diffusion tensor) for each of the model
994 systems compared with the values obtained via microcanonical and Langevin
995 simulations.
996
997 \subsection{Spherical particles}
998 Our model system for spherical particles was a Lennard-Jones sphere of
999 diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
1000 4.7 \AA). The well depth ($\epsilon$) for both particles was set to
1001 an arbitrary value of 0.8 kcal/mol.
1002
1003 The Stokes-Einstein behavior of large spherical particles in
1004 hydrodynamic flows with ``stick'' boundary conditions is well known,
1005 and is given in Eqs. (\ref{eq:StokesTranslation}) and
1006 (\ref{eq:StokesRotation}). Recently, Schmidt and Skinner have
1007 computed the behavior of spherical tag particles in molecular dynamics
1008 simulations, and have shown that {\it slip} boundary conditions
1009 ($\Xi_{tt} = 4 \pi \eta \rho$) may be more appropriate for
1010 molecule-sized spheres embedded in a sea of spherical solvent
1011 particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
1012
1013 Our simulation results show similar behavior to the behavior observed
1014 by Schmidt and Skinner. The diffusion constant obtained from our
1015 microcanonical molecular dynamics simulations lies between the slip
1016 and stick boundary condition results obtained via Stokes-Einstein
1017 behavior. Since the Langevin integrator assumes Stokes-Einstein stick
1018 boundary conditions in calculating the drag and random forces for
1019 spherical particles, our Langevin routine obtains nearly quantitative
1020 agreement with the hydrodynamic results for spherical particles. One
1021 avenue for improvement of the method would be to compute elements of
1022 $\Xi_{tt}$ assuming behavior intermediate between the two boundary
1023 conditions.
1024
1025 In the explicit solvent simulations, both our solute and solvent
1026 particles were structureless, exerting no torques upon each other.
1027 Therefore, there are not rotational correlation times available for
1028 this model system.
1029
1030 \subsection{Ellipsoids}
1031 For uniaxial ellipsoids ($a > b = c$), Perrin's formulae for both
1032 translational and rotational diffusion of each of the body-fixed axes
1033 can be combined to give a single translational diffusion
1034 constant,\cite{Berne90}
1035 \begin{equation}
1036 D = \frac{k_B T}{6 \pi \eta a} G(s),
1037 \label{Dperrin}
1038 \end{equation}
1039 as well as a single rotational diffusion coefficient,
1040 \begin{equation}
1041 \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - s^2)
1042 G(s) - 1}{1 - s^4} \right\}.
1043 \label{ThetaPerrin}
1044 \end{equation}
1045 In these expressions, $G(s)$ is a function of the axial ratio
1046 ($s = b / a$), which for prolate ellipsoids, is
1047 \begin{equation}
1048 G(s) = (1- s^2)^{-1/2} \ln \left\{ \frac{1 + (1 - s^2)^{1/2}}{s} \right\}
1049 \label{GPerrin}
1050 \end{equation}
1051 Again, there is some uncertainty about the correct boundary conditions
1052 to use for molecular-scale ellipsoids in a sea of similarly-sized
1053 solvent particles. Ravichandran and Bagchi found that {\it slip}
1054 boundary conditions most closely resembled the simulation
1055 results,\cite{Ravichandran:1999fk} in agreement with earlier work of
1056 Tang and Evans.\cite{TANG:1993lr}
1057
1058 Even though there are analytic resistance tensors for ellipsoids, we
1059 constructed a rough-shell model using 2135 beads (each with a diameter
1060 of 0.25 \AA) to approximate the shape of the model ellipsoid. We
1061 compared the Langevin dynamics from both the simple ellipsoidal
1062 resistance tensor and the rough shell approximation with
1063 microcanonical simulations and the predictions of Perrin. As in the
1064 case of our spherical model system, the Langevin integrator reproduces
1065 almost exactly the behavior of the Perrin formulae (which is
1066 unsurprising given that the Perrin formulae were used to derive the
1067 drag and random forces applied to the ellipsoid). We obtain
1068 translational diffusion constants and rotational correlation times
1069 that are within a few percent of the analytic values for both the
1070 exact treatment of the diffusion tensor as well as the rough-shell
1071 model for the ellipsoid.
1072
1073 The translational diffusion constants from the microcanonical
1074 simulations agree well with the predictions of the Perrin model,
1075 although the {\it rotational} correlation times are a factor of 2
1076 shorter than expected from hydrodynamic theory. One explanation for
1077 the slower rotation of explicitly-solvated ellipsoids is the
1078 possibility that solute-solvent collisions happen at both ends of the
1079 solute whenever the principal axis of the ellipsoid is turning. In the
1080 upper portion of figure \ref{fig:explanation} we sketch a physical
1081 picture of this explanation. Since our Langevin integrator is
1082 providing nearly quantitative agreement with the Perrin model, it also
1083 predicts orientational diffusion for ellipsoids that exceed explicitly
1084 solvated correlation times by a factor of two.
1085
1086 \subsection{Rigid dumbbells}
1087 Perhaps the only {\it composite} rigid body for which analytic
1088 expressions for the hydrodynamic tensor are available is the
1089 two-sphere dumbbell model. This model consists of two non-overlapping
1090 spheres held by a rigid bond connecting their centers. There are
1091 competing expressions for the 6x6 resistance tensor for this
1092 model. The second order expression introduced by Rotne and
1093 Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la Torre and
1094 Bloomfield,\cite{Torre1977} is given above as
1095 Eq. (\ref{introEquation:RPTensorNonOverlapped}). In our case, we use
1096 a model dumbbell in which the two spheres are identical Lennard-Jones
1097 particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
1098 a distance of 6.532 \AA.
1099
1100 The theoretical values for the translational diffusion constant of the
1101 dumbbell are calculated from the work of Stimson and Jeffery, who
1102 studied the motion of this system in a flow parallel to the
1103 inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
1104 motion in a flow {\it perpendicular} to the inter-sphere
1105 axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
1106 orientational} correlation times for this model system (other than
1107 those derived from the 6 x 6 tensor mentioned above).
1108
1109 The bead model for this model system comprises the two large spheres
1110 by themselves, while the rough shell approximation used 3368 separate
1111 beads (each with a diameter of 0.25 \AA) to approximate the shape of
1112 the rigid body. The hydrodynamics tensors computed from both the bead
1113 and rough shell models are remarkably similar. Computing the initial
1114 hydrodynamic tensor for a rough shell model can be quite expensive (in
1115 this case it requires inverting a 10104 x 10104 matrix), while the
1116 bead model is typically easy to compute (in this case requiring
1117 inversion of a 6 x 6 matrix).
1118
1119 \begin{figure}
1120 \centering
1121 \includegraphics[width=2in]{RoughShellReduced}
1122 \caption[The model rigid bodies (left column) used to test this
1123 algorithm and their rough-shell approximations (right-column) that
1124 were used to compute the hydrodynamic tensors. The top two models
1125 (ellipsoid and dumbbell) have analytic solutions and were used to test
1126 the rough shell approximation. The lower two models (banana and
1127 lipid) were compared with explicitly-solvated molecular dynamics
1128 simulations]{}
1129 \label{fig:roughShell}
1130 \end{figure}
1131
1132
1133 Once the hydrodynamic tensor has been computed, there is no additional
1134 penalty for carrying out a Langevin simulation with either of the two
1135 different hydrodynamics models. Our naive expectation is that since
1136 the rigid body's surface is roughened under the various shell models,
1137 the diffusion constants will be even farther from the ``slip''
1138 boundary conditions than observed for the bead model (which uses a
1139 Stokes-Einstein model to arrive at the hydrodynamic tensor). For the
1140 dumbbell, this prediction is correct although all of the Langevin
1141 diffusion constants are within 6\% of the diffusion constant predicted
1142 from the fully solvated system.
1143
1144 For rotational motion, Langevin integration (and the hydrodynamic tensor)
1145 yields rotational correlation times that are substantially shorter than those
1146 obtained from explicitly-solvated simulations. It is likely that this is due
1147 to the large size of the explicit solvent spheres, a feature that prevents
1148 the solvent from coming in contact with a substantial fraction of the surface
1149 area of the dumbbell. Therefore, the explicit solvent only provides drag
1150 over a substantially reduced surface area of this model, while the
1151 hydrodynamic theories utilize the entire surface area for estimating
1152 rotational diffusion. A sketch of the free volume available in the explicit
1153 solvent simulations is shown in figure \ref{fig:explanation}.
1154
1155
1156 \begin{figure}
1157 \centering
1158 \includegraphics[width=6in]{explanation}
1159 \caption[Explanations of the differences between orientational
1160 correlation times for explicitly-solvated models and hydrodynamic
1161 predictions. For the ellipsoids (upper figures), rotation of the
1162 principal axis can involve correlated collisions at both sides of the
1163 solute. In the rigid dumbbell model (lower figures), the large size
1164 of the explicit solvent spheres prevents them from coming in contact
1165 with a substantial fraction of the surface area of the dumbbell.
1166 Therefore, the explicit solvent only provides drag over a
1167 substantially reduced surface area of this model, where the
1168 hydrodynamic theories utilize the entire surface area for estimating
1169 rotational diffusion]{} \label{fig:explanation}
1170 \end{figure}
1171
1172 \subsection{Composite banana-shaped molecules}
1173 Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids have
1174 been used by Orlandi {\it et al.} to observe mesophases in
1175 coarse-grained models for bent-core liquid crystalline
1176 molecules.\cite{Orlandi:2006fk} We have used the same overlapping
1177 ellipsoids as a way to test the behavior of our algorithm for a
1178 structure of some interest to the materials science community,
1179 although since we are interested in capturing only the hydrodynamic
1180 behavior of this model, we have left out the dipolar interactions of
1181 the original Orlandi model.
1182
1183 A reference system composed of a single banana rigid body embedded in
1184 a sea of 1929 solvent particles was created and run under standard
1185 (microcanonical) molecular dynamics. The resulting viscosity of this
1186 mixture was 0.298 centipoise (as estimated using
1187 Eq. (\ref{eq:shear})). To calculate the hydrodynamic properties of
1188 the banana rigid body model, we created a rough shell (see
1189 Fig.~\ref{fig:roughShell}), in which the banana is represented as a
1190 ``shell'' made of 3321 identical beads (0.25 \AA\ in diameter)
1191 distributed on the surface. Applying the procedure described in
1192 Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1193 identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0
1194 \AA).
1195
1196 The Langevin rigid-body integrator (and the hydrodynamic diffusion
1197 tensor) are essentially quantitative for translational diffusion of
1198 this model. Orientational correlation times under the Langevin
1199 rigid-body integrator are within 11\% of the values obtained from
1200 explicit solvent, but these models also exhibit some solvent
1201 inaccessible surface area in the explicitly-solvated case.
1202
1203 \subsection{Composite sphero-ellipsoids}
1204
1205 Spherical heads perched on the ends of Gay-Berne ellipsoids have been
1206 used recently as models for lipid
1207 molecules.\cite{SunX._jp0762020,Ayton01} A reference system composed
1208 of a single lipid rigid body embedded in a sea of 1929 solvent
1209 particles was created and run under a microcanonical ensemble. The
1210 resulting viscosity of this mixture was 0.349 centipoise (as estimated
1211 using Eq. (\ref{eq:shear})). To calculate the hydrodynamic properties
1212 of the lipid rigid body model, we created a rough shell (see
1213 Fig.~\ref{fig:roughShell}), in which the lipid is represented as a
1214 ``shell'' made of 3550 identical beads (0.25 \AA\ in diameter)
1215 distributed on the surface. Applying the procedure described by
1216 Eq. (\ref{introEquation:ResistanceTensorArbitraryOrigin}), we
1217 identified the center of resistance, ${\bf r} = $(0 \AA, 0 \AA, 1.46
1218 \AA).
1219
1220 The translational diffusion constants and rotational correlation times
1221 obtained using the Langevin rigid-body integrator (and the
1222 hydrodynamic tensor) are essentially quantitative when compared with
1223 the explicit solvent simulations for this model system.
1224
1225 \subsection{Summary of comparisons with explicit solvent simulations}
1226 The Langevin rigid-body integrator we have developed is a reliable way
1227 to replace explicit solvent simulations in cases where the detailed
1228 solute-solvent interactions do not greatly impact the behavior of the
1229 solute. As such, it has the potential to greatly increase the length
1230 and time scales of coarse grained simulations of large solvated
1231 molecules. In cases where the dielectric screening of the solvent, or
1232 specific solute-solvent interactions become important for structural
1233 or dynamic features of the solute molecule, this integrator may be
1234 less useful. However, for the kinds of coarse-grained modeling that
1235 have become popular in recent years (ellipsoidal side chains, rigid
1236 bodies, and molecular-scale models), this integrator may prove itself
1237 to be quite valuable.
1238
1239 \begin{figure}
1240 \centering
1241 \includegraphics[width=\linewidth]{graph}
1242 \caption[The mean-squared displacements ($\langle r^2(t) \rangle$) and
1243 orientational correlation functions ($C_2(t)$) for each of the model
1244 rigid bodies studied. The circles are the results for microcanonical
1245 simulations with explicit solvent molecules, while the other data sets
1246 are results for Langevin dynamics using the different hydrodynamic
1247 tensor approximations. The Perrin model for the ellipsoids is
1248 considered the ``exact'' hydrodynamic behavior (this can also be said
1249 for the translational motion of the dumbbell operating under the bead
1250 model). In most cases, the various hydrodynamics models reproduce each
1251 other quantitatively]{}
1252 \label{fig:results}
1253 \end{figure}
1254
1255 \begin{table*}
1256 \begin{minipage}{\linewidth}
1257 \begin{center}
1258 \caption[Translational diffusion constants (D) for the model systems
1259 calculated using microcanonical simulations (with explicit solvent),
1260 theoretical predictions, and Langevin simulations (with implicit solvent).
1261 Analytical solutions for the exactly-solved hydrodynamics models are obtained
1262 from: Stokes' law (sphere), and Refs. \citen{Perrin1934} and \citen{Perrin1936}
1263 (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1264 (dumbbell). The other model systems have no known analytic solution.
1265 All diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1266 $10^{-4}$ \AA$^2$ / fs).]{}
1267 \begin{tabular}{lccccccc}
1268 \hline
1269 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1270 \cline{2-3} \cline{5-7}
1271 model & $\eta$ (centipoise) & D & & Analytical & method & Hydrodynamics & simulation \\
1272 \hline
1273 sphere & 0.279 & 3.06 & & 2.42 & exact & 2.42 & 2.33 \\
1274 ellipsoid & 0.255 & 2.44 & & 2.34 & exact & 2.34 & 2.37 \\
1275 & 0.255 & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1276 dumbbell & 0.308 & 2.06 & & 1.64 & bead model & 1.65 & 1.62 \\
1277 & 0.308 & 2.06 & & 1.64 & rough shell & 1.59 & 1.62 \\
1278 banana & 0.298 & 1.53 & & & rough shell & 1.56 & 1.55 \\
1279 lipid & 0.349 & 1.41 & & & rough shell & 1.33 & 1.32 \\
1280 \end{tabular}
1281 \label{tab:translation}
1282 \end{center}
1283 \end{minipage}
1284 \end{table*}
1285
1286 \begin{table*}
1287 \begin{minipage}{\linewidth}
1288 \begin{center}
1289 \caption[Orientational relaxation times ($\tau$) for the model systems using
1290 microcanonical simulation (with explicit solvent), theoretical
1291 predictions, and Langevin simulations (with implicit solvent). All
1292 relaxation times are for the rotational correlation function with
1293 $\ell = 2$ and are reported in units of ps. The ellipsoidal model has
1294 an exact solution for the orientational correlation time due to
1295 Perrin, but the other model systems have no known analytic solution.]{}
1296 \begin{tabular}{lccccccc}
1297 \hline
1298 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1299 \cline{2-3} \cline{5-7}
1300 model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic & simulation \\
1301 \hline
1302 sphere & 0.279 & & & 9.69 & exact & 9.69 & 9.64 \\
1303 ellipsoid & 0.255 & 46.7 & & 22.0 & exact & 22.0 & 22.2 \\
1304 & 0.255 & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1305 dumbbell & 0.308 & 14.1 & & & bead model & 50.0 & 50.1 \\
1306 & 0.308 & 14.1 & & & rough shell & 41.5 & 41.3 \\
1307 banana & 0.298 & 63.8 & & & rough shell & 70.9 & 70.9 \\
1308 lipid & 0.349 & 78.0 & & & rough shell & 76.9 & 77.9 \\
1309 \hline
1310 \end{tabular}
1311 \label{tab:rotation}
1312 \end{center}
1313 \end{minipage}
1314 \end{table*}
1315
1316 \section{Application: A rigid-body lipid bilayer}
1317
1318 To test the accuracy and efficiency of the new integrator, we applied
1319 it to study the formation of corrugated structures emerging from
1320 simulations of the coarse grained lipid molecular models presented
1321 above. The initial configuration is taken from earlier molecular
1322 dynamics studies on lipid bilayers which had used spherical
1323 (Lennard-Jones) solvent particles and moderate (480 solvated lipid
1324 molecules) system sizes.\cite{SunX._jp0762020} the solvent molecules
1325 were excluded from the system and the box was replicated three times
1326 in the x- and y- axes (giving a single simulation cell containing 4320
1327 lipids). The experimental value for the viscosity of water at 20C
1328 ($\eta = 1.00$ cp) was used with the Langevin integrator to mimic the
1329 hydrodynamic effects of the solvent. The absence of explicit solvent
1330 molecules and the stability of the integrator allowed us to take
1331 timesteps of 50 fs. A simulation run time of 30 ns was sampled to
1332 calculate structural properties. Fig. \ref{fig:bilayer} shows the
1333 configuration of the system after 30 ns. Structural properties of the
1334 bilayer (e.g. the head and body $P_2$ order parameters) are nearly
1335 identical to those obtained via solvated molecular dynamics. The
1336 ripple structure remained stable during the entire trajectory.
1337 Compared with using explicit bead-model solvent molecules, the 30 ns
1338 trajectory for 4320 lipids with the Langevin integrator is now {\it
1339 faster} on the same hardware than the same length trajectory was for
1340 the 480-lipid system previously studied.
1341
1342 \begin{figure}
1343 \centering
1344 \includegraphics[width=\linewidth]{bilayer}
1345 \caption[A snapshot of a bilayer composed of 4320 rigid-body models
1346 for lipid molecules evolving using the Langevin integrator described
1347 in this work]{} \label{fig:bilayer}
1348 \end{figure}
1349
1350 \section{Conclusions}
1351
1352 We have presented a new algorithm for carrying out Langevin dynamics
1353 simulations on complex rigid bodies by incorporating the hydrodynamic
1354 resistance tensors for arbitrary shapes into a stable and efficient
1355 integration scheme. The integrator gives quantitative agreement with
1356 both analytic and approximate hydrodynamic theories, and works
1357 reasonably well at reproducing the solute dynamical properties
1358 (diffusion constants, and orientational relaxation times) from
1359 explicitly-solvated simulations. For the cases where there are
1360 discrepancies between our Langevin integrator and the explicit solvent
1361 simulations, two features of molecular simulations help explain the
1362 differences.
1363
1364 First, the use of ``stick'' boundary conditions for molecular-sized
1365 solutes in a sea of similarly-sized solvent particles may be
1366 problematic. We are certainly not the first group to notice this
1367 difference between hydrodynamic theories and explicitly-solvated
1368 molecular
1369 simulations.\cite{Schmidt:2004fj,Schmidt:2003kx,Ravichandran:1999fk,TANG:1993lr}
1370 The problem becomes particularly noticable in both the translational
1371 diffusion of the spherical particles and the rotational diffusion of
1372 the ellipsoids. In both of these cases it is clear that the
1373 approximations that go into hydrodynamics are the source of the error,
1374 and not the integrator itself.
1375
1376 Second, in the case of structures which have substantial surface area
1377 that is inaccessible to solvent particles, the hydrodynamic theories
1378 (and the Langevin integrator) may overestimate the effects of solvent
1379 friction because they overestimate the exposed surface area of the
1380 rigid body. This is particularly noticable in the rotational
1381 diffusion of the dumbbell model. We believe that given a solvent of
1382 known radius, it may be possible to modify the rough shell approach to
1383 place beads on solvent-accessible surface, instead of on the geometric
1384 surface defined by the van der Waals radii of the components of the
1385 rigid body. Further work to confirm the behavior of this new
1386 approximation is ongoing.
1387
1388 \section{Acknowledgments}
1389 Support for this project was provided by the National Science
1390 Foundation under grant CHE-0134881. T.L. also acknowledges the
1391 financial support from Center of Applied Mathematics at University of
1392 Notre Dame.
1393
1394 \newpage
1395
1396 \bibliographystyle{jcp2}
1397 \bibliography{langevin}
1398 \end{doublespace}
1399 \end{document}