ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3352
Committed: Fri Feb 29 22:02:46 2008 UTC (17 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 65334 byte(s)
Log Message:
newest version

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