ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3390
Committed: Wed Apr 30 14:50:53 2008 UTC (16 years, 4 months ago) by xsun
Content type: application/x-tex
File size: 67012 byte(s)
Log Message:
adding refs.

File Contents

# User Rev Content
1 tim 2746 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 gezelter 3352 %\documentclass[aps,prb,preprint,amsmath,amssymb,endfloats]{revtex4}
3 tim 2746 \documentclass[11pt]{article}
4 gezelter 3337 \usepackage{amsmath}
5 tim 2746 \usepackage{amssymb}
6     \usepackage{times}
7     \usepackage{mathptmx}
8     \usepackage{setspace}
9 gezelter 3352 \usepackage{endfloat}
10     %\usepackage{tabularx}
11 tim 2746 \usepackage{graphicx}
12 gezelter 3352 %\usepackage{booktabs}
13     %\usepackage{bibentry}
14     %\usepackage{mathrsfs}
15 tim 2746 \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 gezelter 3352
20 gezelter 3299 \renewcommand\citemid{\ } % no comma in optional referenc note
21 tim 2746
22     \begin{document}
23    
24 gezelter 3352 \title{Langevin dynamics for rigid bodies of arbitrary shape}
25 tim 2746
26 gezelter 3352 \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 tim 2746 University of Notre Dame\\
30     Notre Dame, Indiana 46556}
31    
32     \date{\today}
33    
34    
35 gezelter 3352 \maketitle
36    
37    
38    
39 tim 2746 \begin{abstract}
40 gezelter 3352 We present an algorithm for carrying out Langevin dynamics simulations
41     on complex rigid bodies by incorporating the hydrodynamic resistance
42 gezelter 3367 tensors for arbitrary shapes into an advanced rotational integration
43 gezelter 3352 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 tim 2746 \end{abstract}
49    
50     \newpage
51    
52     %\narrowtext
53    
54     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
55     % BODY OF TEXT
56     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
57    
58 gezelter 3352 \begin{doublespace}
59    
60 tim 2746 \section{Introduction}
61    
62     %applications of langevin dynamics
63 gezelter 3352 Langevin dynamics, which mimics a heat bath using both stochastic and
64 gezelter 3316 dissipative forces, has been applied in a variety of situations as an
65     alternative to molecular dynamics with explicit solvent molecules.
66     The stochastic treatment of the solvent allows the use of simulations
67 gezelter 3352 with substantially longer time and length scales. In general, the
68 gezelter 3316 dynamic and structural properties obtained from Langevin simulations
69     agree quite well with similar properties obtained from explicit
70     solvent simulations.
71    
72     Recent examples of the usefulness of Langevin simulations include a
73     study of met-enkephalin in which Langevin simulations predicted
74     dynamical properties that were largely in agreement with explicit
75     solvent simulations.\cite{Shen2002} By applying Langevin dynamics with
76 gezelter 3352 the UNRES model, Liwo and his coworkers suggest that protein folding
77 gezelter 3316 pathways can be explored within a reasonable amount of
78     time.\cite{Liwo2005}
79    
80     The stochastic nature of Langevin dynamics also enhances the sampling
81     of the system and increases the probability of crossing energy
82     barriers.\cite{Cui2003,Banerjee2004} Combining Langevin dynamics with
83 gezelter 3333 Kramers' theory, Klimov and Thirumalai identified free-energy
84 gezelter 3316 barriers by studying the viscosity dependence of the protein folding
85     rates.\cite{Klimov1997} In order to account for solvent induced
86     interactions missing from the implicit solvent model, Kaya
87     incorporated a desolvation free energy barrier into protein
88     folding/unfolding studies and discovered a higher free energy barrier
89 xsun 3317 between the native and denatured states.\cite{HuseyinKaya07012005}
90 gezelter 3316
91 gezelter 3352 In typical LD simulations, the friction and random ($f_r$) forces on
92 gezelter 3333 individual atoms are taken from Stokes' law,
93 gezelter 3316 \begin{eqnarray}
94 gezelter 3352 m \dot{v}(t) & = & -\nabla U(x) - \xi m v(t) + f_r(t) \notag \\
95     \langle f_r(t) \rangle & = & 0 \\
96     \langle f_r(t) f_r(t') \rangle & = & 2 k_B T \xi m \delta(t - t') \notag
97 gezelter 3316 \end{eqnarray}
98 gezelter 3352 where $\xi \approx 6 \pi \eta \rho$. Here $\eta$ is the viscosity of the
99     implicit solvent, and $\rho$ is the hydrodynamic radius of the atom.
100 tim 2746
101 gezelter 3333 The use of rigid substructures,\cite{Chun:2000fj}
102 gezelter 3352 coarse-graining,\cite{Ayton01,Golubkov06,Orlandi:2006fk,SunX._jp0762020}
103     and ellipsoidal representations of protein side
104     chains~\cite{Fogolari:1996lr} has made the use of the Stokes-Einstein
105     approximation problematic. A rigid substructure moves as a single
106     unit with orientational as well as translational degrees of freedom.
107     This requires a more general treatment of the hydrodynamics than the
108     spherical approximation provides. Also, the atoms involved in a rigid
109     or coarse-grained structure have solvent-mediated interactions with
110     each other, and these interactions are ignored if all atoms are
111     treated as separate spherical particles. The theory of interactions
112     {\it between} bodies moving through a fluid has been developed over
113     the past century and has been applied to simulations of Brownian
114     motion.\cite{FIXMAN:1986lr,Ramachandran1996}
115 tim 2746
116 gezelter 3352 In order to account for the diffusion anisotropy of complex shapes,
117     Fernandes and Garc\'{i}a de la Torre improved an earlier Brownian
118     dynamics simulation algorithm~\cite{Ermak1978,Allison1991} by
119 gezelter 3333 incorporating a generalized $6\times6$ diffusion tensor and
120     introducing a rotational evolution scheme consisting of three
121     consecutive rotations.\cite{Fernandes2002} Unfortunately, biases are
122     introduced into the system due to the arbitrary order of applying the
123     noncommuting rotation operators.\cite{Beard2003} Based on the
124     observation the momentum relaxation time is much less than the time
125     step, one may ignore the inertia in Brownian dynamics. However, the
126     assumption of zero average acceleration is not always true for
127     cooperative motion which is common in proteins. An inertial Brownian
128     dynamics (IBD) was proposed to address this issue by adding an
129 gezelter 3352 inertial correction term.\cite{Beard2000} As a complement to IBD,
130     which has a lower bound in time step because of the inertial
131     relaxation time, long-time-step inertial dynamics (LTID) can be used
132     to investigate the inertial behavior of linked polymer segments in a
133     low friction regime.\cite{Beard2000} LTID can also deal with the
134 gezelter 3333 rotational dynamics for nonskew bodies without translation-rotation
135     coupling by separating the translation and rotation motion and taking
136 gezelter 3352 advantage of the analytical solution of hydrodynamic
137 gezelter 3333 properties. However, typical nonskew bodies like cylinders and
138     ellipsoids are inadequate to represent most complex macromolecular
139 gezelter 3352 assemblies. Therefore, the goal of this work is to adapt some of the
140     hydrodynamic methodologies developed to treat Brownian motion of
141     complex assemblies into a Langevin integrator for rigid bodies with
142     arbitrary shapes.
143 gezelter 3333
144 gezelter 3316 \subsection{Rigid Body Dynamics}
145     Rigid bodies are frequently involved in the modeling of large
146     collections of particles that move as a single unit. In molecular
147     simulations, rigid bodies have been used to simplify protein-protein
148 gezelter 3333 docking,\cite{Gray2003} and lipid bilayer
149 gezelter 3352 simulations.\cite{SunX._jp0762020} Many of the water models in common
150 gezelter 3333 use are also rigid-body
151     models,\cite{Jorgensen83,Berendsen81,Berendsen87} although they are
152 gezelter 3352 typically evolved in molecular dynamics simulations using constraints
153     rather than rigid body equations of motion.
154 gezelter 3316
155 gezelter 3333 Euler angles are a natural choice to describe the rotational degrees
156 gezelter 3337 of freedom. However, due to $\frac{1}{\sin \theta}$ singularities, the
157 gezelter 3333 numerical integration of corresponding equations of these motion can
158     become inaccurate (and inefficient). Although the use of multiple
159     sets of Euler angles can overcome this problem,\cite{Barojas1973} the
160     computational penalty and the loss of angular momentum conservation
161     remain. A singularity-free representation utilizing quaternions was
162     developed by Evans in 1977.\cite{Evans1977} The Evans quaternion
163     approach uses a nonseparable Hamiltonian, and this has prevented
164     symplectic algorithms from being utilized until very
165     recently.\cite{Miller2002}
166 gezelter 3316
167 gezelter 3333 Another approach is the application of holonomic constraints to the
168     atoms belonging to the rigid body. Each atom moves independently
169     under the normal forces deriving from potential energy and constraints
170     are used to guarantee rigidity. However, due to their iterative
171     nature, the SHAKE and RATTLE algorithms converge very slowly when the
172     number of constraints (and the number of particles that belong to the
173     rigid body) increases.\cite{Ryckaert1977,Andersen1983}
174 tim 2746
175 gezelter 3333 In order to develop a stable and efficient integration scheme that
176 gezelter 3367 preserves most constants of the motion in microcanonical simulations,
177     symplectic propagators are necessary. By introducing a conjugate
178     momentum to the rotation matrix ${\bf Q}$ and re-formulating
179     Hamilton's equations, a symplectic orientational integrator,
180     RSHAKE,\cite{Kol1997} was proposed to evolve rigid bodies on a
181     constraint manifold by iteratively satisfying the orthogonality
182     constraint ${\bf Q}^T {\bf Q} = 1$. An alternative method using the
183     quaternion representation was developed by Omelyan.\cite{Omelyan1998}
184     However, both of these methods are iterative and suffer from some
185     related inefficiencies. A symplectic Lie-Poisson integrator for rigid
186     bodies developed by Dullweber {\it et al.}\cite{Dullweber1997} removes
187     most of the limitations mentioned above and is therefore the basis for
188     our Langevin integrator.
189 gezelter 3316
190 tim 2746 The goal of the present work is to develop a Langevin dynamics
191 gezelter 3341 algorithm for arbitrary-shaped rigid particles by integrating an
192     accurate estimate of the friction tensor from hydrodynamics theory
193 gezelter 3367 into a stable and efficient rigid body dynamics propagator. In the
194     sections below, we review some of the theory of hydrodynamic tensors
195     developed primarily for Brownian simulations of multi-particle
196     systems, we then present our integration method for a set of
197     generalized Langevin equations of motion, and we compare the behavior
198     of the new Langevin integrator to dynamical quantities obtained via
199     explicit solvent molecular dynamics.
200 tim 2746
201 gezelter 3316 \subsection{\label{introSection:frictionTensor}The Friction Tensor}
202 gezelter 3341 Theoretically, a complete friction kernel for a solute particle can be
203     determined using the velocity autocorrelation function from a
204     simulation with explicit solvent molecules. However, this approach
205     becomes impractical when the solute becomes complex. Instead, various
206     approaches based on hydrodynamics have been developed to calculate
207     static friction coefficients. In general, the friction tensor $\Xi$ is
208     a $6\times 6$ matrix given by
209 gezelter 3316 \begin{equation}
210 gezelter 3333 \Xi = \left( \begin{array}{*{20}c}
211     \Xi^{tt} & \Xi^{rt} \\
212     \Xi^{tr} & \Xi^{rr} \\
213     \end{array} \right).
214 gezelter 3316 \end{equation}
215     Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
216     rotational resistance (friction) tensors respectively, while
217     $\Xi^{tr}$ is translation-rotation coupling tensor and $\Xi^{rt}$ is
218     rotation-translation coupling tensor. When a particle moves in a
219 gezelter 3341 fluid, it may experience a friction force ($\mathbf{f}_f$) and torque
220     ($\mathbf{\tau}_f$) in opposition to the velocity ($\mathbf{v}$) and
221     body-fixed angular velocity ($\mathbf{\omega}$),
222 gezelter 3316 \begin{equation}
223 tim 2746 \left( \begin{array}{l}
224 xsun 3338 \mathbf{f}_f \\
225 gezelter 3316 \mathbf{\tau}_f \\
226 gezelter 3333 \end{array} \right) = - \left( \begin{array}{*{20}c}
227     \Xi^{tt} & \Xi^{rt} \\
228     \Xi^{tr} & \Xi^{rr} \\
229     \end{array} \right)\left( \begin{array}{l}
230 gezelter 3316 \mathbf{v} \\
231     \mathbf{\omega} \\
232     \end{array} \right).
233     \end{equation}
234 gezelter 3367 For an arbitrary body moving in a fluid, Peters has derived a set of
235     fluctuation-dissipation relations for the friction
236     tensors,\cite{Peters:1999qy,Peters:1999uq,Peters:2000fk}
237     \begin{eqnarray}
238     \Xi^{tt} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
239     F}(0) {\bf F}(-s) \rangle_{eq} - \langle {\bf F} \rangle_{eq}^2
240     \right] ds \\
241     \notag \\
242     \Xi^{tr} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
243     F}(0) {\bf \tau}(-s) \rangle_{eq} - \langle {\bf F} \rangle_{eq}
244     \langle {\bf \tau} \rangle_{eq} \right] ds \\
245     \notag \\
246     \Xi^{rt} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
247     \tau}(0) {\bf F}(-s) \rangle_{eq} - \langle {\bf \tau} \rangle_{eq}
248     \langle {\bf F} \rangle_{eq} \right] ds \\
249     \notag \\
250     \Xi^{rr} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
251     \tau}(0) {\bf \tau}(-s) \rangle_{eq} - \langle {\bf \tau} \rangle_{eq}^2
252     \right] ds
253     \end{eqnarray}
254     In these expressions, the forces (${\bf F}$) and torques (${\bf
255     \tau}$) are those that arise solely from the interactions of the body with
256     the surrounding fluid. For a single solute body in an isotropic fluid,
257     the average forces and torques in these expressions ($\langle {\bf F}
258     \rangle_{eq}$ and $\langle {\bf \tau} \rangle_{eq}$)
259     vanish, and one obtains the simpler force-torque correlation formulae
260     of Nienhuis.\cite{Nienhuis:1970lr} Molecular dynamics simulations with
261     explicit solvent molecules can be used to obtain estimates of the
262     friction tensors with these formulae. In practice, however, one needs
263     relatively long simulations with frequently-stored force and torque
264     information to compute friction tensors, and this becomes
265     prohibitively expensive when there are large numbers of large solute
266     particles. For bodies with simple shapes, there are a number of
267     approximate expressions that allow computation of these tensors
268     without the need for expensive simulations that utilize explicit
269     solvent particles.
270 tim 2746
271 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
272 gezelter 3367 For a spherical body under ``stick'' boundary conditions, the
273     translational and rotational friction tensors can be estimated from
274     Stokes' law,
275 gezelter 3316 \begin{equation}
276 gezelter 3352 \label{eq:StokesTranslation}
277 gezelter 3333 \Xi^{tt} = \left( \begin{array}{*{20}c}
278 gezelter 3352 {6\pi \eta \rho} & 0 & 0 \\
279     0 & {6\pi \eta \rho} & 0 \\
280     0 & 0 & {6\pi \eta \rho} \\
281 gezelter 3333 \end{array} \right)
282 gezelter 3316 \end{equation}
283 tim 2999 and
284 gezelter 3316 \begin{equation}
285 gezelter 3352 \label{eq:StokesRotation}
286 gezelter 3333 \Xi^{rr} = \left( \begin{array}{*{20}c}
287 gezelter 3352 {8\pi \eta \rho^3 } & 0 & 0 \\
288     0 & {8\pi \eta \rho^3 } & 0 \\
289     0 & 0 & {8\pi \eta \rho^3 } \\
290 gezelter 3333 \end{array} \right)
291 gezelter 3316 \end{equation}
292 gezelter 3352 where $\eta$ is the viscosity of the solvent and $\rho$ is the
293     hydrodynamic radius. The presence of the rotational resistance tensor
294     implies that the spherical body has internal structure and
295     orientational degrees of freedom that must be propagated in time. For
296     non-structured spherical bodies (i.e. the atoms in a traditional
297     molecular dynamics simulation) these degrees of freedom do not exist.
298 tim 2999
299     Other non-spherical shapes, such as cylinders and ellipsoids, are
300 gezelter 3341 widely used as references for developing new hydrodynamic theories,
301 tim 2999 because their properties can be calculated exactly. In 1936, Perrin
302 gezelter 3341 extended Stokes' law to general
303     ellipsoids,\cite{Perrin1934,Perrin1936} described in Cartesian
304     coordinates as
305 gezelter 3316 \begin{equation}
306 gezelter 3333 \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1.
307 gezelter 3316 \end{equation}
308 gezelter 3333 Here, the semi-axes are of lengths $a$, $b$, and $c$. Due to the
309     complexity of the elliptic integral, only uniaxial ellipsoids, either
310 gezelter 3341 prolate ($a \ge b = c$) or oblate ($a < b = c$), were solved
311 gezelter 3333 exactly. Introducing an elliptic integral parameter $S$ for prolate,
312 gezelter 3316 \begin{equation}
313     S = \frac{2}{\sqrt{a^2 - b^2}} \ln \frac{a + \sqrt{a^2 - b^2}}{b},
314     \end{equation}
315 gezelter 3333 and oblate,
316 gezelter 3316 \begin{equation}
317     S = \frac{2}{\sqrt {b^2 - a^2 }} \arctan \frac{\sqrt {b^2 - a^2}}{a},
318     \end{equation}
319 gezelter 3341 ellipsoids, it is possible to write down exact solutions for the
320     resistance tensors. As is the case for spherical bodies, the translational,
321     \begin{eqnarray}
322 gezelter 3316 \Xi_a^{tt} & = & 16\pi \eta \frac{a^2 - b^2}{(2a^2 - b^2 )S - 2a}. \\
323     \Xi_b^{tt} = \Xi_c^{tt} & = & 32\pi \eta \frac{a^2 - b^2 }{(2a^2 - 3b^2 )S + 2a},
324 gezelter 3341 \end{eqnarray}
325     and rotational,
326     \begin{eqnarray}
327 gezelter 3316 \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2 - b^2 )b^2}{2a - b^2 S}, \\
328     \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4 - b^4)}{(2a^2 - b^2 )S - 2a}
329 gezelter 3341 \end{eqnarray}
330     resistance tensors are diagonal $3 \times 3$ matrices. For both
331     spherical and ellipsoidal particles, the translation-rotation and
332     rotation-translation coupling tensors are zero.
333 tim 2746
334 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
335 gezelter 3367 Other than the fluctuation dissipation formulae given by
336     Peters,\cite{Peters:1999qy,Peters:1999uq,Peters:2000fk} there are no
337     analytic solutions for the friction tensor for rigid molecules of
338     arbitrary shape. The ellipsoid of revolution and general triaxial
339     ellipsoid models have been widely used to approximate the hydrodynamic
340     properties of rigid bodies. However, the mapping from all possible
341     ellipsoidal spaces ($r$-space) to all possible combinations of
342     rotational diffusion coefficients ($D$-space) is not
343 gezelter 3316 unique.\cite{Wegener1979} Additionally, because there is intrinsic
344 gezelter 3341 coupling between translational and rotational motion of {\it skew}
345     rigid bodies, general ellipsoids are not always suitable for modeling
346     rigid molecules. A number of studies have been devoted to determining
347     the friction tensor for irregular shapes using methods in which the
348     molecule of interest is modeled with a combination of
349     spheres\cite{Carrasco1999} and the hydrodynamic properties of the
350     molecule are then calculated using a set of two-point interaction
351     tensors. We have found the {\it bead} and {\it rough shell} models of
352     Carrasco and Garc\'{i}a de la Torre to be the most useful of these
353     methods,\cite{Carrasco1999} and we review the basic outline of the
354     rough shell approach here. A more thorough explanation can be found
355     in Ref. \citen{Carrasco1999}.
356 gezelter 3333
357 gezelter 3341 Consider a rigid assembly of $N$ small beads moving through a
358     continuous medium. Due to hydrodynamic interactions between the
359     beads, the net velocity of the $i^\mathrm{th}$ bead relative to the
360     medium, ${\bf v}'_i$, is different than its unperturbed velocity ${\bf
361     v}_i$,
362 gezelter 3333 \begin{equation}
363 gezelter 3341 {\bf v}'_i = {\bf v}_i - \sum\limits_{j \ne i} {{\bf T}_{ij} {\bf F}_j }
364 gezelter 3333 \end{equation}
365 gezelter 3341 where ${\bf F}_j$ is the frictional force on the medium due to bead $j$, and
366     ${\bf T}_{ij}$ is the hydrodynamic interaction tensor between the two beads.
367     The frictional force felt by the $i^\mathrm{th}$ bead is proportional to
368     its net velocity
369 tim 2746 \begin{equation}
370 gezelter 3352 {\bf F}_i = \xi_i {\bf v}_i - \xi_i \sum\limits_{j \ne i} {{\bf T}_{ij} {\bf F}_j }.
371 tim 2999 \label{introEquation:tensorExpression}
372 tim 2746 \end{equation}
373 gezelter 3341 Eq. (\ref{introEquation:tensorExpression}) defines the two-point
374     hydrodynamic tensor, ${\bf T}_{ij}$. There have been many proposed
375     solutions to this equation, including the simple solution given by
376     Oseen and Burgers in 1930 for two beads of identical radius. A second
377     order expression for beads of different hydrodynamic radii was
378 gezelter 3352 introduced by Rotne and Prager,\cite{Rotne1969} and improved by
379 tim 2999 Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
380 tim 2746 \begin{equation}
381 gezelter 3341 {\bf T}_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {{\bf I} +
382 gezelter 3352 \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right) + \frac{{\rho
383     _i^2 + \rho_j^2 }}{{R_{ij}^2 }}\left( {\frac{{\bf I}}{3} -
384 gezelter 3341 \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
385 tim 2999 \label{introEquation:RPTensorNonOverlapped}
386 tim 2746 \end{equation}
387 gezelter 3341 Here ${\bf R}_{ij}$ is the distance vector between beads $i$ and $j$. Both
388     the Oseen-Burgers tensor and
389     Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption that
390 gezelter 3352 the beads do not overlap ($R_{ij} \ge \rho_i + \rho_j$).
391 gezelter 3341
392     To calculate the resistance tensor for a body represented as the union
393     of many non-overlapping beads, we first pick an arbitrary origin $O$
394     and then construct a $3N \times 3N$ supermatrix consisting of $N
395     \times N$ ${\bf B}_{ij}$ blocks
396 tim 2746 \begin{equation}
397 gezelter 3341 {\bf B} = \left( \begin{array}{*{20}c}
398     {\bf B}_{11} & \ldots & {\bf B}_{1N} \\
399     \vdots & \ddots & \vdots \\
400     {\bf B}_{N1} & \cdots & {\bf B}_{NN}
401     \end{array} \right)
402 tim 2746 \end{equation}
403 gezelter 3341 ${\bf B}_{ij}$ is a version of the hydrodynamic tensor which includes the
404     self-contributions for spheres,
405 tim 2999 \begin{equation}
406 gezelter 3341 {\bf B}_{ij} = \delta _{ij} \frac{{\bf I}}{{6\pi \eta R_{ij}}} + (1 - \delta_{ij}
407     ){\bf T}_{ij}
408     \end{equation}
409     where $\delta_{ij}$ is the Kronecker delta function. Inverting the
410     ${\bf B}$ matrix, we obtain
411     \begin{equation}
412     {\bf C} = {\bf B}^{ - 1} = \left(\begin{array}{*{20}c}
413     {\bf C}_{11} & \ldots & {\bf C}_{1N} \\
414     \vdots & \ddots & \vdots \\
415     {\bf C}_{N1} & \cdots & {\bf C}_{NN}
416 gezelter 3333 \end{array} \right),
417 tim 2999 \end{equation}
418 gezelter 3341 which can be partitioned into $N \times N$ blocks labeled ${\bf C}_{ij}$.
419     (Each of the ${\bf C}_{ij}$ blocks is a $3 \times 3$ matrix.) Using the
420     skew matrix,
421 gezelter 3333 \begin{equation}
422 gezelter 3341 {\bf U}_i = \left(\begin{array}{*{20}c}
423     0 & -z_i & y_i \\
424     z_i & 0 & - x_i \\
425     -y_i & x_i & 0
426     \end{array}\right)
427     \label{eq:skewMatrix}
428 gezelter 3333 \end{equation}
429 tim 2999 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
430 gezelter 3341 bead $i$ and origin $O$, the elements of the resistance tensor (at the
431     arbitrary origin $O$) can be written as
432 tim 2999 \begin{eqnarray}
433 xsun 3339 \label{introEquation:ResistanceTensorArbitraryOrigin}
434 gezelter 3341 \Xi^{tt} & = & \sum\limits_i {\sum\limits_j {{\bf C}_{ij} } } \notag , \\
435     \Xi^{tr} = \Xi _{}^{rt} & = & \sum\limits_i {\sum\limits_j {{\bf U}_i {\bf C}_{ij} } } , \\
436     \Xi^{rr} & = & -\sum\limits_i \sum\limits_j {\bf U}_i {\bf C}_{ij} {\bf U}_j + 6 \eta V {\bf I}. \notag
437 tim 2999 \end{eqnarray}
438 gezelter 3341 The final term in the expression for $\Xi^{rr}$ is a correction that
439     accounts for errors in the rotational motion of the bead models. The
440     additive correction uses the solvent viscosity ($\eta$) as well as the
441     total volume of the beads that contribute to the hydrodynamic model,
442 gezelter 3310 \begin{equation}
443 gezelter 3352 V = \frac{4 \pi}{3} \sum_{i=1}^{N} \rho_i^3,
444 gezelter 3310 \end{equation}
445 gezelter 3352 where $\rho_i$ is the radius of bead $i$. This correction term was
446 gezelter 3310 rigorously tested and compared with the analytical results for
447 gezelter 3341 two-sphere and ellipsoidal systems by Garc\'{i}a de la Torre and
448 gezelter 3310 Rodes.\cite{Torre:1983lr}
449    
450 gezelter 3341 In general, resistance tensors depend on the origin at which they were
451     computed. However, the proper location for applying the friction
452     force is the center of resistance, the special point at which the
453     trace of rotational resistance tensor, $\Xi^{rr}$ reaches a minimum
454     value. Mathematically, the center of resistance can also be defined
455     as the unique point for a rigid body at which the translation-rotation
456     coupling tensors are symmetric,
457 tim 2999 \begin{equation}
458 gezelter 3341 \Xi^{tr} = \left(\Xi^{tr}\right)^T
459 tim 2999 \label{introEquation:definitionCR}
460     \end{equation}
461 gezelter 3341 From Eq. \ref{introEquation:ResistanceTensorArbitraryOrigin}, we can
462     easily derive that the {\it translational} resistance tensor is origin
463     independent, while the rotational resistance tensor and
464 tim 2999 translation-rotation coupling resistance tensor depend on the
465 gezelter 3341 origin. Given the resistance tensor at an arbitrary origin $O$, and a
466     vector ,${\bf r}_{OP} = (x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we
467     can obtain the resistance tensor at $P$ by
468     \begin{eqnarray}
469     \label{introEquation:resistanceTensorTransformation}
470     \Xi_P^{tt} & = & \Xi_O^{tt} \notag \\
471     \Xi_P^{tr} = \Xi_P^{rt} & = & \Xi_O^{tr} - {\bf U}_{OP} \Xi _O^{tt} \\
472     \Xi_P^{rr} & = &\Xi_O^{rr} - {\bf U}_{OP} \Xi_O^{tt} {\bf U}_{OP}
473     + \Xi_O^{tr} {\bf U}_{OP} - {\bf U}_{OP} \left( \Xi_O^{tr}
474     \right)^{^T} \notag
475     \end{eqnarray}
476     where ${\bf U}_{OP}$ is the skew matrix (Eq. (\ref{eq:skewMatrix}))
477     for the vector between the origin $O$ and the point $P$. Using
478     Eqs.~\ref{introEquation:definitionCR}~and~\ref{introEquation:resistanceTensorTransformation},
479     one can locate the position of center of resistance,
480     \begin{equation*}
481     \left(\begin{array}{l}
482     x_{OR} \\
483     y_{OR} \\
484     z_{OR}
485     \end{array}\right) =
486     \left(\begin{array}{*{20}c}
487     (\Xi_O^{rr})_{yy} + (\Xi_O^{rr})_{zz} & -(\Xi_O^{rr})_{xy} & -(\Xi_O^{rr})_{xz} \\
488     -(\Xi_O^{rr})_{xy} & (\Xi_O^{rr})_{zz} + (\Xi_O^{rr})_{xx} & -(\Xi_O^{rr})_{yz} \\
489     -(\Xi_O^{rr})_{xz} & -(\Xi_O^{rr})_{yz} & (\Xi_O^{rr})_{xx} + (\Xi_O^{rr})_{yy} \\
490     \end{array}\right)^{-1}
491     \left(\begin{array}{l}
492     (\Xi_O^{tr})_{yz} - (\Xi_O^{tr})_{zy} \\
493     (\Xi_O^{tr})_{zx} - (\Xi_O^{tr})_{xz} \\
494     (\Xi_O^{tr})_{xy} - (\Xi_O^{tr})_{yx}
495     \end{array}\right)
496     \end{equation*}
497 xsun 3338 where $x_{OR}$, $y_{OR}$, $z_{OR}$ are the components of the vector
498 tim 2999 joining center of resistance $R$ and origin $O$.
499 tim 2746
500 gezelter 3341 For a general rigid molecular substructure, finding the $6 \times 6$
501     resistance tensor can be a computationally demanding task. First, a
502     lattice of small beads that extends well beyond the boundaries of the
503     rigid substructure is created. The lattice is typically composed of
504     0.25 \AA\ beads on a dense FCC lattice. The lattice constant is taken
505     to be the bead diameter, so that adjacent beads are touching, but do
506     not overlap. To make a shape corresponding to the rigid structure,
507     beads that sit on lattice sites that are outside the van der Waals
508 gezelter 3352 radii of all of the atoms comprising the rigid body are excluded from
509     the calculation.
510 tim 2746
511 gezelter 3341 For large structures, most of the beads will be deep within the rigid
512     body and will not contribute to the hydrodynamic tensor. In the {\it
513     rough shell} approach, beads which have all of their lattice neighbors
514     inside the structure are considered interior beads, and are removed
515     from the calculation. After following this procedure, only those
516     beads in direct contact with the van der Waals surface of the rigid
517     body are retained. For reasonably large molecular structures, this
518     truncation can still produce bead assemblies with thousands of
519     members.
520    
521 gezelter 3352 If all of the {\it atoms} comprising the rigid substructure are
522     spherical and non-overlapping, the tensor in
523 gezelter 3341 Eq.~(\ref{introEquation:RPTensorNonOverlapped}) may be used directly
524     using the atoms themselves as the hydrodynamic beads. This is a
525     variant of the {\it bead model} approach of Carrasco and Garc\'{i}a de
526 gezelter 3352 la Torre.\cite{Carrasco1999} In this case, the size of the ${\bf B}$
527     matrix can be quite small, and the calculation of the hydrodynamic
528     tensor is straightforward.
529 gezelter 3341
530     In general, the inversion of the ${\bf B}$ matrix is the most
531     computationally demanding task. This inversion is done only once for
532 gezelter 3352 each type of rigid structure. We have used straightforward
533     LU-decomposition to solve the linear system and to obtain the elements
534     of ${\bf C}$. Once ${\bf C}$ has been obtained, the location of the
535 gezelter 3341 center of resistance ($R$) is found and the resistance tensor at this
536     point is calculated. The $3 \times 1$ vector giving the location of
537     the rigid body's center of resistance and the $6 \times 6$ resistance
538 gezelter 3352 tensor are then stored for use in the Langevin dynamics calculation.
539     These quantities depend on solvent viscosity and temperature and must
540     be recomputed if different simulation conditions are required.
541 gezelter 3341
542 gezelter 3310 \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
543 gezelter 3337
544 tim 2999 Consider the Langevin equations of motion in generalized coordinates
545 tim 2746 \begin{equation}
546 gezelter 3341 {\bf M} \dot{{\bf V}}(t) = {\bf F}_{s}(t) +
547     {\bf F}_{f}(t) + {\bf F}_{r}(t)
548 tim 2746 \label{LDGeneralizedForm}
549     \end{equation}
550 gezelter 3341 where ${\bf M}$ is a $6 \times 6$ diagonal mass matrix (which
551 gezelter 3337 includes the mass of the rigid body as well as the moments of inertia
552 gezelter 3341 in the body-fixed frame) and ${\bf V}$ is a generalized velocity,
553     ${\bf V} =
554     \left\{{\bf v},{\bf \omega}\right\}$. The right side of
555 gezelter 3333 Eq.~\ref{LDGeneralizedForm} consists of three generalized forces: a
556 gezelter 3352 system force (${\bf F}_{s}$), a frictional or dissipative force (${\bf
557     F}_{f}$) and a stochastic force (${\bf F}_{r}$). While the evolution
558     of the system in Newtonian mechanics is typically done in the lab
559     frame, it is convenient to handle the dynamics of rigid bodies in
560     body-fixed frames. Thus the friction and random forces on each
561 gezelter 3341 substructure are calculated in a body-fixed frame and may converted
562     back to the lab frame using that substructure's rotation matrix (${\bf
563     Q}$):
564 gezelter 3333 \begin{equation}
565 gezelter 3341 {\bf F}_{f,r} =
566 gezelter 3340 \left( \begin{array}{c}
567 gezelter 3341 {\bf f}_{f,r} \\
568     {\bf \tau}_{f,r}
569 gezelter 3340 \end{array} \right)
570     =
571     \left( \begin{array}{c}
572 gezelter 3341 {\bf Q}^{T} {\bf f}^{~b}_{f,r} \\
573     {\bf Q}^{T} {\bf \tau}^{~b}_{f,r}
574 gezelter 3340 \end{array} \right)
575 gezelter 3333 \end{equation}
576 gezelter 3341 The body-fixed friction force, ${\bf F}_{f}^{~b}$, is proportional to
577     the (body-fixed) velocity at the center of resistance
578     ${\bf v}_{R}^{~b}$ and the angular velocity ${\bf \omega}$
579 tim 2746 \begin{equation}
580 gezelter 3341 {\bf F}_{f}^{~b}(t) = \left( \begin{array}{l}
581     {\bf f}_{f}^{~b}(t) \\
582     {\bf \tau}_{f}^{~b}(t) \\
583 gezelter 3333 \end{array} \right) = - \left( \begin{array}{*{20}c}
584 gezelter 3341 \Xi_{R}^{tt} & \Xi_{R}^{rt} \\
585     \Xi_{R}^{tr} & \Xi_{R}^{rr} \\
586 gezelter 3333 \end{array} \right)\left( \begin{array}{l}
587 gezelter 3341 {\bf v}_{R}^{~b}(t) \\
588     {\bf \omega}(t) \\
589 tim 2746 \end{array} \right),
590     \end{equation}
591 gezelter 3341 while the random force, ${\bf F}_{r}$, is a Gaussian stochastic
592     variable with zero mean and variance,
593 tim 2746 \begin{equation}
594 gezelter 3341 \left\langle {{\bf F}_{r}(t) ({\bf F}_{r}(t'))^T } \right\rangle =
595     \left\langle {{\bf F}_{r}^{~b} (t) ({\bf F}_{r}^{~b} (t'))^T } \right\rangle =
596     2 k_B T \Xi_R \delta(t - t'). \label{eq:randomForce}
597 tim 2746 \end{equation}
598 gezelter 3340 $\Xi_R$ is the $6\times6$ resistance tensor at the center of
599 gezelter 3341 resistance. Once this tensor is known for a given rigid body (as
600     described in the previous section) obtaining a stochastic vector that
601     has the properties in Eq. (\ref{eq:randomForce}) can be done
602     efficiently by carrying out a one-time Cholesky decomposition to
603     obtain the square root matrix of the resistance tensor,
604     \begin{equation}
605     \Xi_R = {\bf S} {\bf S}^{T},
606     \label{eq:Cholesky}
607     \end{equation}
608 gezelter 3352 where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
609 gezelter 3341 vector with the statistics required for the random force can then be
610     obtained by multiplying ${\bf S}$ onto a random 6-vector ${\bf Z}$ which
611     has elements chosen from a Gaussian distribution, such that:
612 gezelter 3340 \begin{equation}
613 gezelter 3341 \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
614     {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
615 gezelter 3340 \end{equation}
616 gezelter 3341 where $\delta t$ is the timestep in use during the simulation. The
617     random force, ${\bf F}_{r}^{~b} = {\bf S} {\bf Z}$, can be shown to have the
618     correct properties required by Eq. (\ref{eq:randomForce}).
619 gezelter 3333
620 gezelter 3341 The equation of motion for the translational velocity of the center of
621     mass (${\bf v}$) can be written as
622 tim 2746 \begin{equation}
623 gezelter 3341 m \dot{{\bf v}} (t) = {\bf f}_{s}(t) + {\bf f}_{f}(t) +
624     {\bf f}_{r}(t)
625 tim 2746 \end{equation}
626 gezelter 3341 Since the frictional and random forces are applied at the center of
627     resistance, which generally does not coincide with the center of mass,
628     extra torques are exerted at the center of mass. Thus, the net
629     body-fixed torque at the center of mass, $\tau^{~b}(t)$,
630     is given by
631 tim 2746 \begin{equation}
632 gezelter 3341 \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)
633 tim 2746 \end{equation}
634 gezelter 3341 where ${\bf r}_{MR}$ is the vector from the center of mass to the center of
635     resistance. Instead of integrating the angular velocity in lab-fixed
636     frame, we consider the equation of motion for the angular momentum
637     (${\bf j}$) in the body-fixed frame
638 tim 2746 \begin{equation}
639 gezelter 3352 \frac{\partial}{\partial t}{\bf j}(t) = \tau^{~b}(t)
640 tim 2746 \end{equation}
641 gezelter 3341 Embedding the friction and random forces into the the total force and
642     torque, one can integrate the Langevin equations of motion for a rigid
643     body of arbitrary shape in a velocity-Verlet style 2-part algorithm,
644     where $h = \delta t$:
645 tim 2746
646 gezelter 3341 {\tt move A:}
647 tim 2746 \begin{align*}
648 tim 2999 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
649     + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
650     %
651     {\bf r}(t + h) &\leftarrow {\bf r}(t)
652     + h {\bf v}\left(t + h / 2 \right), \\
653     %
654     {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
655 gezelter 3341 + \frac{h}{2} {\bf \tau}^{~b}(t), \\
656 tim 2999 %
657 gezelter 3341 {\bf Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
658 tim 2999 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
659 tim 2746 \end{align*}
660 gezelter 3341 In this context, $\overleftrightarrow{\mathsf{I}}$ is the diagonal
661     moment of inertia tensor, and the $\mathrm{rotate}$ function is the
662     reversible product of the three body-fixed rotations,
663 tim 2746 \begin{equation}
664     \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
665     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
666     / 2) \cdot \mathsf{G}_x(a_x /2),
667     \end{equation}
668     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
669 gezelter 3341 rotates both the rotation matrix ($\mathbf{Q}$) and the body-fixed
670 tim 2999 angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
671     axis $\alpha$,
672 tim 2746 \begin{equation}
673     \mathsf{G}_\alpha( \theta ) = \left\{
674     \begin{array}{lcl}
675 gezelter 3341 \mathbf{Q}(t) & \leftarrow & \mathbf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
676 tim 2746 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
677     j}(0).
678     \end{array}
679     \right.
680     \end{equation}
681     $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
682     rotation matrix. For example, in the small-angle limit, the
683     rotation matrix around the body-fixed x-axis can be approximated as
684     \begin{equation}
685     \mathsf{R}_x(\theta) \approx \left(
686     \begin{array}{ccc}
687     1 & 0 & 0 \\
688     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
689     \theta^2 / 4} \\
690     0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
691     \theta^2 / 4}
692     \end{array}
693     \right).
694     \end{equation}
695 tim 2999 All other rotations follow in a straightforward manner. After the
696     first part of the propagation, the forces and body-fixed torques are
697 gezelter 3341 calculated at the new positions and orientations. The system forces
698     and torques are derivatives of the total potential energy function
699     ($U$) with respect to the rigid body positions (${\bf r}$) and the
700     columns of the transposed rotation matrix ${\bf Q}^T = \left({\bf
701     u}_x, {\bf u}_y, {\bf u}_z \right)$:
702 tim 2746
703 gezelter 3341 {\tt Forces:}
704 tim 2999 \begin{align*}
705 gezelter 3341 {\bf f}_{s}(t + h) & \leftarrow
706     - \left(\frac{\partial U}{\partial {\bf r}}\right)_{{\bf r}(t + h)} \\
707 tim 2999 %
708 gezelter 3341 {\bf \tau}_{s}(t + h) &\leftarrow {\bf u}(t + h)
709     \times \frac{\partial U}{\partial {\bf u}} \\
710 tim 2999 %
711 gezelter 3341 {\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) \\
712     %
713     {\bf f}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tt} \cdot
714     {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rt} \cdot {\bf \omega}(t+h) \\
715     %
716     {\bf \tau}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tr} \cdot
717     {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rr} \cdot {\bf \omega}(t+h) \\
718     %
719     Z & \leftarrow {\tt GaussianNormal}(2 k_B T / h, 6) \\
720     {\bf F}_{R,r}^{b}(t+h) & \leftarrow {\bf S} \cdot Z \\
721     %
722     {\bf f}(t+h) & \leftarrow {\bf f}_{s}(t+h) + \mathbf{Q}^{T}(t+h)
723     \cdot \left({\bf f}_{R,f}^{~b} + {\bf f}_{R,r}^{~b} \right) \\
724     %
725     \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) \\
726     \tau^{~b}(t+h) & \leftarrow \mathbf{Q}(t+h) \cdot \tau(t+h) \\
727 tim 2999 \end{align*}
728 gezelter 3341 Frictional (and random) forces and torques must be computed at the
729     center of resistance, so there are additional steps required to find
730     the body-fixed velocity (${\bf v}_{R}^{~b}$) at this location. Mapping
731     the frictional and random forces at the center of resistance back to
732     the center of mass also introduces an additional term in the torque
733     one obtains at the center of mass.
734    
735 tim 2746 Once the forces and torques have been obtained at the new time step,
736     the velocities can be advanced to the same time value.
737    
738 gezelter 3341 {\tt move B:}
739 tim 2746 \begin{align*}
740 tim 2999 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
741     \right)
742     + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
743     %
744     {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
745     \right)
746 gezelter 3341 + \frac{h}{2} {\bf \tau}^{~b}(t + h) .
747 tim 2746 \end{align*}
748    
749 gezelter 3310 \section{Validating the Method\label{sec:validating}}
750 gezelter 3302 In order to validate our Langevin integrator for arbitrarily-shaped
751 gezelter 3305 rigid bodies, we implemented the algorithm in {\sc
752     oopse}\cite{Meineke2005} and compared the results of this algorithm
753     with the known
754 gezelter 3302 hydrodynamic limiting behavior for a few model systems, and to
755     microcanonical molecular dynamics simulations for some more
756     complicated bodies. The model systems and their analytical behavior
757     (if known) are summarized below. Parameters for the primary particles
758     comprising our model systems are given in table \ref{tab:parameters},
759     and a sketch of the arrangement of these primary particles into the
760 gezelter 3305 model rigid bodies is shown in figure \ref{fig:models}. In table
761     \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
762     ellipsoidal (Gay-Berne) particles. For spherical particles, the value
763     of the Lennard-Jones $\sigma$ parameter is the particle diameter
764     ($d$). Gay-Berne ellipsoids have an energy scaling parameter,
765     $\epsilon^s$, which describes the well depth for two identical
766     ellipsoids in a {\it side-by-side} configuration. Additionally, a
767     well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
768     describes the ratio between the well depths in the {\it end-to-end}
769     and side-by-side configurations. For spheres, $\epsilon^r \equiv 1$.
770     Moments of inertia are also required to describe the motion of primary
771     particles with orientational degrees of freedom.
772 gezelter 3299
773 gezelter 3302 \begin{table*}
774     \begin{minipage}{\linewidth}
775     \begin{center}
776     \caption{Parameters for the primary particles in use by the rigid body
777     models in figure \ref{fig:models}.}
778     \begin{tabular}{lrcccccccc}
779     \hline
780     & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
781     & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
782     $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
783 gezelter 3308 Sphere & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
784 gezelter 3302 Ellipsoid & & 4.6 & 13.8 & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
785 gezelter 3308 Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
786 gezelter 3302 Banana &(3 identical ellipsoids)& 4.2 & 11.2 & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
787     Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
788     & Ellipsoidal Tail & 4.6 & 13.8 & 0.8 & 0.2 & 760 & 45000 & 45000 & 9000 \\
789     Solvent & & 4.7 & $= d$ & 0.8 & 1 & 72.06 & & & \\
790     \hline
791     \end{tabular}
792     \label{tab:parameters}
793     \end{center}
794     \end{minipage}
795     \end{table*}
796    
797 gezelter 3305 \begin{figure}
798     \centering
799     \includegraphics[width=3in]{sketch}
800 xsun 3390 \caption[A sketch of the model systems used in evaluating the behavior
801     of the rigid body Langevin integrator]{} \label{fig:models}
802 gezelter 3305 \end{figure}
803    
804 gezelter 3302 \subsection{Simulation Methodology}
805     We performed reference microcanonical simulations with explicit
806     solvents for each of the different model system. In each case there
807     was one solute model and 1929 solvent molecules present in the
808 gezelter 3352 simulation box. All simulations were equilibrated for 5 ns using a
809 gezelter 3302 constant-pressure and temperature integrator with target values of 300
810     K for the temperature and 1 atm for pressure. Following this stage,
811 gezelter 3352 further equilibration (5 ns) and sampling (10 ns) was done in a
812     microcanonical ensemble. Since the model bodies are typically quite
813     massive, we were able to use a time step of 25 fs.
814 gezelter 3310
815     The model systems studied used both Lennard-Jones spheres as well as
816     uniaxial Gay-Berne ellipoids. In its original form, the Gay-Berne
817     potential was a single site model for the interactions of rigid
818 gezelter 3352 ellipsoidal molecules.\cite{Gay1981} It can be thought of as a
819 gezelter 3310 modification of the Gaussian overlap model originally described by
820     Berne and Pechukas.\cite{Berne72} The potential is constructed in the
821     familiar form of the Lennard-Jones function using
822     orientation-dependent $\sigma$ and $\epsilon$ parameters,
823     \begin{equation*}
824 gezelter 3341 V_{ij}({{\bf \hat u}_i}, {{\bf \hat u}_j}, {{\bf \hat
825     r}_{ij}}) = 4\epsilon ({{\bf \hat u}_i}, {{\bf \hat u}_j},
826     {{\bf \hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u
827 gezelter 3310 }_i},
828 gezelter 3341 {{\bf \hat u}_j}, {{\bf \hat r}_{ij}})+\sigma_0}\right)^{12}
829     -\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u}_i}, {{\bf \hat u}_j},
830     {{\bf \hat r}_{ij}})+\sigma_0}\right)^6\right]
831 gezelter 3310 \label{eq:gb}
832     \end{equation*}
833    
834     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
835     \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
836     \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
837     are dependent on the relative orientations of the two ellipsoids (${\bf
838     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
839     inter-ellipsoid separation (${\bf \hat{r}}_{ij}$). The shape and
840     attractiveness of each ellipsoid is governed by a relatively small set
841     of parameters: $l$ and $d$ describe the length and width of each
842     uniaxial ellipsoid, while $\epsilon^s$, which describes the well depth
843     for two identical ellipsoids in a {\it side-by-side} configuration.
844     Additionally, a well depth aspect ratio, $\epsilon^r = \epsilon^e /
845     \epsilon^s$, describes the ratio between the well depths in the {\it
846     end-to-end} and side-by-side configurations. Details of the potential
847 gezelter 3352 are given elsewhere,\cite{Luckhurst90,Golubkov06,SunX._jp0762020} and an
848 gezelter 3310 excellent overview of the computational methods that can be used to
849     efficiently compute forces and torques for this potential can be found
850     in Ref. \citen{Golubkov06}
851    
852     For the interaction between nonequivalent uniaxial ellipsoids (or
853     between spheres and ellipsoids), the spheres are treated as ellipsoids
854     with an aspect ratio of 1 ($d = l$) and with an well depth ratio
855     ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of the
856     Gay-Berne potential we are using was generalized by Cleaver {\it et
857     al.} and is appropriate for dissimilar uniaxial
858     ellipsoids.\cite{Cleaver96}
859    
860     A switching function was applied to all potentials to smoothly turn
861     off the interactions between a range of $22$ and $25$ \AA. The
862     switching function was the standard (cubic) function,
863 gezelter 3302 \begin{equation}
864     s(r) =
865     \begin{cases}
866     1 & \text{if $r \le r_{\text{sw}}$},\\
867     \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
868     {(r_{\text{cut}} - r_{\text{sw}})^3}
869     & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
870     0 & \text{if $r > r_{\text{cut}}$.}
871     \end{cases}
872     \label{eq:switchingFunc}
873     \end{equation}
874 gezelter 3310
875 gezelter 3302 To measure shear viscosities from our microcanonical simulations, we
876     used the Einstein form of the pressure correlation function,\cite{hess:209}
877     \begin{equation}
878 gezelter 3310 \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
879     \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \right\rangle_{t_0}.
880 gezelter 3302 \label{eq:shear}
881     \end{equation}
882 gezelter 3352 which converges much more rapidly in molecular dynamics simulations
883     than the traditional Green-Kubo formula.
884 gezelter 3302
885     The Langevin dynamics for the different model systems were performed
886     at the same temperature as the average temperature of the
887     microcanonical simulations and with a solvent viscosity taken from
888 gezelter 3305 Eq. (\ref{eq:shear}) applied to these simulations. We used 1024
889     independent solute simulations to obtain statistics on our Langevin
890     integrator.
891 gezelter 3302
892     \subsection{Analysis}
893    
894     The quantities of interest when comparing the Langevin integrator to
895     analytic hydrodynamic equations and to molecular dynamics simulations
896     are typically translational diffusion constants and orientational
897     relaxation times. Translational diffusion constants for point
898     particles are computed easily from the long-time slope of the
899     mean-square displacement,
900     \begin{equation}
901 gezelter 3310 D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \left\langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \right\rangle,
902 gezelter 3302 \end{equation}
903     of the solute molecules. For models in which the translational
904 gezelter 3305 diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
905     (i.e. any non-spherically-symmetric rigid body), it is possible to
906     compute the diffusive behavior for motion parallel to each body-fixed
907     axis by projecting the displacement of the particle onto the
908     body-fixed reference frame at $t=0$. With an isotropic solvent, as we
909 gezelter 3352 have used in this study, there may be differences between the three
910     diffusion constants at short times, but these must converge to the
911     same value at longer times. Translational diffusion constants for the
912     different shaped models are shown in table \ref{tab:translation}.
913 gezelter 3302
914 gezelter 3305 In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
915 gezelter 3302 diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
916     {\it around} a particular body-fixed axis and {\it not} the diffusion
917     of a vector pointing along the axis. However, these eigenvalues can
918     be combined to find 5 characteristic rotational relaxation
919 gezelter 3305 times,\cite{PhysRev.119.53,Berne90}
920 gezelter 3302 \begin{eqnarray}
921 gezelter 3305 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
922     1 / \tau_2 & = & 6 D_r - 2 \Delta \\
923     1 / \tau_3 & = & 3 (D_r + D_1) \\
924     1 / \tau_4 & = & 3 (D_r + D_2) \\
925     1 / \tau_5 & = & 3 (D_r + D_3)
926 gezelter 3302 \end{eqnarray}
927     where
928     \begin{equation}
929     D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
930     \end{equation}
931     and
932     \begin{equation}
933 gezelter 3305 \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
934 gezelter 3302 \end{equation}
935 gezelter 3305 Each of these characteristic times can be used to predict the decay of
936     part of the rotational correlation function when $\ell = 2$,
937 gezelter 3302 \begin{equation}
938 gezelter 3305 C_2(t) = \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
939 gezelter 3302 \end{equation}
940 gezelter 3305 This is the same as the $F^2_{0,0}(t)$ correlation function that
941     appears in Ref. \citen{Berne90}. The amplitudes of the two decay
942     terms are expressed in terms of three dimensionless functions of the
943     eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
944     2\Delta)$, and $N = 2 \sqrt{\Delta b}$. Similar expressions can be
945     obtained for other angular momentum correlation
946     functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
947     studied, only one of the amplitudes of the two decay terms was
948     non-zero, so it was possible to derive a single relaxation time for
949     each of the hydrodynamic tensors. In many cases, these characteristic
950     times are averaged and reported in the literature as a single relaxation
951     time,\cite{Garcia-de-la-Torre:1997qy}
952 gezelter 3302 \begin{equation}
953 gezelter 3305 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
954     \end{equation}
955     although for the cases reported here, this averaging is not necessary
956     and only one of the five relaxation times is relevant.
957    
958     To test the Langevin integrator's behavior for rotational relaxation,
959     we have compared the analytical orientational relaxation times (if
960     they are known) with the general result from the diffusion tensor and
961     with the results from both the explicitly solvated molecular dynamics
962     and Langevin simulations. Relaxation times from simulations (both
963     microcanonical and Langevin), were computed using Legendre polynomial
964     correlation functions for a unit vector (${\bf u}$) fixed along one or
965     more of the body-fixed axes of the model.
966     \begin{equation}
967 gezelter 3310 C_{\ell}(t) = \left\langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
968     u}_{i}(0) \right) \right\rangle
969 gezelter 3302 \end{equation}
970     For simulations in the high-friction limit, orientational correlation
971     times can then be obtained from exponential fits of this function, or by
972     integrating,
973     \begin{equation}
974 gezelter 3305 \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
975 gezelter 3302 \end{equation}
976 gezelter 3305 In lower-friction solvents, the Legendre correlation functions often
977     exhibit non-exponential decay, and may not be characterized by a
978     single decay constant.
979 gezelter 3302
980     In table \ref{tab:rotation} we show the characteristic rotational
981     relaxation times (based on the diffusion tensor) for each of the model
982     systems compared with the values obtained via microcanonical and Langevin
983     simulations.
984    
985 gezelter 3305 \subsection{Spherical particles}
986 gezelter 3299 Our model system for spherical particles was a Lennard-Jones sphere of
987     diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
988     4.7 \AA). The well depth ($\epsilon$) for both particles was set to
989 gezelter 3302 an arbitrary value of 0.8 kcal/mol.
990 gezelter 3299
991     The Stokes-Einstein behavior of large spherical particles in
992 gezelter 3352 hydrodynamic flows with ``stick'' boundary conditions is well known,
993     and is given in Eqs. (\ref{eq:StokesTranslation}) and
994     (\ref{eq:StokesRotation}). Recently, Schmidt and Skinner have
995     computed the behavior of spherical tag particles in molecular dynamics
996     simulations, and have shown that {\it slip} boundary conditions
997     ($\Xi_{tt} = 4 \pi \eta \rho$) may be more appropriate for
998     molecule-sized spheres embedded in a sea of spherical solvent
999     particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
1000 gezelter 3299
1001     Our simulation results show similar behavior to the behavior observed
1002 gezelter 3302 by Schmidt and Skinner. The diffusion constant obtained from our
1003 gezelter 3299 microcanonical molecular dynamics simulations lies between the slip
1004     and stick boundary condition results obtained via Stokes-Einstein
1005     behavior. Since the Langevin integrator assumes Stokes-Einstein stick
1006     boundary conditions in calculating the drag and random forces for
1007     spherical particles, our Langevin routine obtains nearly quantitative
1008     agreement with the hydrodynamic results for spherical particles. One
1009     avenue for improvement of the method would be to compute elements of
1010     $\Xi_{tt}$ assuming behavior intermediate between the two boundary
1011 gezelter 3302 conditions.
1012 gezelter 3299
1013 gezelter 3310 In the explicit solvent simulations, both our solute and solvent
1014     particles were structureless, exerting no torques upon each other.
1015     Therefore, there are not rotational correlation times available for
1016     this model system.
1017 gezelter 3299
1018 gezelter 3310 \subsection{Ellipsoids}
1019     For uniaxial ellipsoids ($a > b = c$), Perrin's formulae for both
1020 gezelter 3299 translational and rotational diffusion of each of the body-fixed axes
1021     can be combined to give a single translational diffusion
1022 gezelter 3302 constant,\cite{Berne90}
1023 gezelter 3299 \begin{equation}
1024 gezelter 3352 D = \frac{k_B T}{6 \pi \eta a} G(s),
1025 gezelter 3299 \label{Dperrin}
1026     \end{equation}
1027     as well as a single rotational diffusion coefficient,
1028     \begin{equation}
1029 gezelter 3352 \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - s^2)
1030     G(s) - 1}{1 - s^4} \right\}.
1031 gezelter 3299 \label{ThetaPerrin}
1032     \end{equation}
1033 gezelter 3352 In these expressions, $G(s)$ is a function of the axial ratio
1034     ($s = b / a$), which for prolate ellipsoids, is
1035 gezelter 3299 \begin{equation}
1036 gezelter 3352 G(s) = (1- s^2)^{-1/2} \ln \left\{ \frac{1 + (1 - s^2)^{1/2}}{s} \right\}
1037 gezelter 3299 \label{GPerrin}
1038     \end{equation}
1039     Again, there is some uncertainty about the correct boundary conditions
1040     to use for molecular-scale ellipsoids in a sea of similarly-sized
1041     solvent particles. Ravichandran and Bagchi found that {\it slip}
1042 gezelter 3302 boundary conditions most closely resembled the simulation
1043     results,\cite{Ravichandran:1999fk} in agreement with earlier work of
1044     Tang and Evans.\cite{TANG:1993lr}
1045 gezelter 3299
1046 gezelter 3305 Even though there are analytic resistance tensors for ellipsoids, we
1047     constructed a rough-shell model using 2135 beads (each with a diameter
1048 gezelter 3310 of 0.25 \AA) to approximate the shape of the model ellipsoid. We
1049 gezelter 3305 compared the Langevin dynamics from both the simple ellipsoidal
1050     resistance tensor and the rough shell approximation with
1051     microcanonical simulations and the predictions of Perrin. As in the
1052     case of our spherical model system, the Langevin integrator reproduces
1053     almost exactly the behavior of the Perrin formulae (which is
1054     unsurprising given that the Perrin formulae were used to derive the
1055 gezelter 3299 drag and random forces applied to the ellipsoid). We obtain
1056     translational diffusion constants and rotational correlation times
1057     that are within a few percent of the analytic values for both the
1058     exact treatment of the diffusion tensor as well as the rough-shell
1059     model for the ellipsoid.
1060    
1061 gezelter 3352 The translational diffusion constants from the microcanonical
1062     simulations agree well with the predictions of the Perrin model,
1063     although the {\it rotational} correlation times are a factor of 2
1064     shorter than expected from hydrodynamic theory. One explanation for
1065     the slower rotation of explicitly-solvated ellipsoids is the
1066     possibility that solute-solvent collisions happen at both ends of the
1067     solute whenever the principal axis of the ellipsoid is turning. In the
1068     upper portion of figure \ref{fig:explanation} we sketch a physical
1069     picture of this explanation. Since our Langevin integrator is
1070     providing nearly quantitative agreement with the Perrin model, it also
1071     predicts orientational diffusion for ellipsoids that exceed explicitly
1072     solvated correlation times by a factor of two.
1073 gezelter 3299
1074 gezelter 3310 \subsection{Rigid dumbbells}
1075 gezelter 3302 Perhaps the only {\it composite} rigid body for which analytic
1076     expressions for the hydrodynamic tensor are available is the
1077     two-sphere dumbbell model. This model consists of two non-overlapping
1078     spheres held by a rigid bond connecting their centers. There are
1079     competing expressions for the 6x6 resistance tensor for this
1080 gezelter 3341 model. The second order expression introduced by Rotne and
1081     Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la Torre and
1082     Bloomfield,\cite{Torre1977} is given above as
1083 gezelter 3299 Eq. (\ref{introEquation:RPTensorNonOverlapped}). In our case, we use
1084     a model dumbbell in which the two spheres are identical Lennard-Jones
1085     particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
1086 gezelter 3302 a distance of 6.532 \AA.
1087 gezelter 3299
1088     The theoretical values for the translational diffusion constant of the
1089     dumbbell are calculated from the work of Stimson and Jeffery, who
1090     studied the motion of this system in a flow parallel to the
1091 gezelter 3302 inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
1092     motion in a flow {\it perpendicular} to the inter-sphere
1093     axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
1094     orientational} correlation times for this model system (other than
1095 gezelter 3341 those derived from the 6 x 6 tensor mentioned above).
1096 gezelter 3299
1097 gezelter 3305 The bead model for this model system comprises the two large spheres
1098     by themselves, while the rough shell approximation used 3368 separate
1099     beads (each with a diameter of 0.25 \AA) to approximate the shape of
1100     the rigid body. The hydrodynamics tensors computed from both the bead
1101     and rough shell models are remarkably similar. Computing the initial
1102     hydrodynamic tensor for a rough shell model can be quite expensive (in
1103     this case it requires inverting a 10104 x 10104 matrix), while the
1104     bead model is typically easy to compute (in this case requiring
1105 gezelter 3308 inversion of a 6 x 6 matrix).
1106 gezelter 3305
1107 gezelter 3308 \begin{figure}
1108     \centering
1109 gezelter 3310 \includegraphics[width=2in]{RoughShell}
1110 xsun 3390 \caption[The model rigid bodies (left column) used to test this
1111     algorithm and their rough-shell approximations (right-column) that
1112     were used to compute the hydrodynamic tensors. The top two models
1113     (ellipsoid and dumbbell) have analytic solutions and were used to test
1114     the rough shell approximation. The lower two models (banana and
1115     lipid) were compared with explicitly-solvated molecular dynamics
1116     simulations]{}
1117 gezelter 3308 \label{fig:roughShell}
1118     \end{figure}
1119    
1120    
1121 gezelter 3305 Once the hydrodynamic tensor has been computed, there is no additional
1122     penalty for carrying out a Langevin simulation with either of the two
1123     different hydrodynamics models. Our naive expectation is that since
1124     the rigid body's surface is roughened under the various shell models,
1125     the diffusion constants will be even farther from the ``slip''
1126     boundary conditions than observed for the bead model (which uses a
1127     Stokes-Einstein model to arrive at the hydrodynamic tensor). For the
1128     dumbbell, this prediction is correct although all of the Langevin
1129     diffusion constants are within 6\% of the diffusion constant predicted
1130     from the fully solvated system.
1131    
1132 gezelter 3308 For rotational motion, Langevin integration (and the hydrodynamic tensor)
1133     yields rotational correlation times that are substantially shorter than those
1134     obtained from explicitly-solvated simulations. It is likely that this is due
1135     to the large size of the explicit solvent spheres, a feature that prevents
1136     the solvent from coming in contact with a substantial fraction of the surface
1137     area of the dumbbell. Therefore, the explicit solvent only provides drag
1138     over a substantially reduced surface area of this model, while the
1139     hydrodynamic theories utilize the entire surface area for estimating
1140     rotational diffusion. A sketch of the free volume available in the explicit
1141     solvent simulations is shown in figure \ref{fig:explanation}.
1142 gezelter 3305
1143 gezelter 3310
1144     \begin{figure}
1145     \centering
1146     \includegraphics[width=6in]{explanation}
1147     \caption[Explanations of the differences between orientational
1148     correlation times for explicitly-solvated models and hydrodynamic
1149 xsun 3390 predictions. For the ellipsoids (upper figures), rotation of the
1150 gezelter 3310 principal axis can involve correlated collisions at both sides of the
1151     solute. In the rigid dumbbell model (lower figures), the large size
1152     of the explicit solvent spheres prevents them from coming in contact
1153     with a substantial fraction of the surface area of the dumbbell.
1154     Therefore, the explicit solvent only provides drag over a
1155     substantially reduced surface area of this model, where the
1156     hydrodynamic theories utilize the entire surface area for estimating
1157 xsun 3390 rotational diffusion]{} \label{fig:explanation}
1158 gezelter 3310 \end{figure}
1159    
1160     \subsection{Composite banana-shaped molecules}
1161     Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids have
1162     been used by Orlandi {\it et al.} to observe mesophases in
1163     coarse-grained models for bent-core liquid crystalline
1164     molecules.\cite{Orlandi:2006fk} We have used the same overlapping
1165 gezelter 3299 ellipsoids as a way to test the behavior of our algorithm for a
1166     structure of some interest to the materials science community,
1167     although since we are interested in capturing only the hydrodynamic
1168 gezelter 3310 behavior of this model, we have left out the dipolar interactions of
1169     the original Orlandi model.
1170 gezelter 3308
1171 gezelter 3341 A reference system composed of a single banana rigid body embedded in
1172     a sea of 1929 solvent particles was created and run under standard
1173     (microcanonical) molecular dynamics. The resulting viscosity of this
1174     mixture was 0.298 centipoise (as estimated using
1175     Eq. (\ref{eq:shear})). To calculate the hydrodynamic properties of
1176     the banana rigid body model, we created a rough shell (see
1177     Fig.~\ref{fig:roughShell}), in which the banana is represented as a
1178     ``shell'' made of 3321 identical beads (0.25 \AA\ in diameter)
1179     distributed on the surface. Applying the procedure described in
1180     Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1181     identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0
1182     \AA).
1183 gezelter 3299
1184 gezelter 3341 The Langevin rigid-body integrator (and the hydrodynamic diffusion
1185     tensor) are essentially quantitative for translational diffusion of
1186     this model. Orientational correlation times under the Langevin
1187     rigid-body integrator are within 11\% of the values obtained from
1188     explicit solvent, but these models also exhibit some solvent
1189     inaccessible surface area in the explicitly-solvated case.
1190 gezelter 3308
1191 gezelter 3310 \subsection{Composite sphero-ellipsoids}
1192 gezelter 3341
1193 gezelter 3299 Spherical heads perched on the ends of Gay-Berne ellipsoids have been
1194 xsun 3312 used recently as models for lipid
1195 gezelter 3367 molecules.\cite{SunX._jp0762020,Ayton01} A reference system composed
1196     of a single lipid rigid body embedded in a sea of 1929 solvent
1197     particles was created and run under a microcanonical ensemble. The
1198     resulting viscosity of this mixture was 0.349 centipoise (as estimated
1199     using Eq. (\ref{eq:shear})). To calculate the hydrodynamic properties
1200     of the lipid rigid body model, we created a rough shell (see
1201 xsun 3312 Fig.~\ref{fig:roughShell}), in which the lipid is represented as a
1202     ``shell'' made of 3550 identical beads (0.25 \AA\ in diameter)
1203 gezelter 3367 distributed on the surface. Applying the procedure described by
1204     Eq. (\ref{introEquation:ResistanceTensorArbitraryOrigin}), we
1205 xsun 3312 identified the center of resistance, ${\bf r} = $(0 \AA, 0 \AA, 1.46
1206     \AA).
1207 gezelter 3310
1208 gezelter 3341 The translational diffusion constants and rotational correlation times
1209     obtained using the Langevin rigid-body integrator (and the
1210     hydrodynamic tensor) are essentially quantitative when compared with
1211     the explicit solvent simulations for this model system.
1212 gezelter 3315
1213 gezelter 3352 \subsection{Summary of comparisons with explicit solvent simulations}
1214     The Langevin rigid-body integrator we have developed is a reliable way
1215     to replace explicit solvent simulations in cases where the detailed
1216     solute-solvent interactions do not greatly impact the behavior of the
1217     solute. As such, it has the potential to greatly increase the length
1218     and time scales of coarse grained simulations of large solvated
1219     molecules. In cases where the dielectric screening of the solvent, or
1220     specific solute-solvent interactions become important for structural
1221     or dynamic features of the solute molecule, this integrator may be
1222     less useful. However, for the kinds of coarse-grained modeling that
1223     have become popular in recent years (ellipsoidal side chains, rigid
1224     bodies, and molecular-scale models), this integrator may prove itself
1225     to be quite valuable.
1226 xsun 3298
1227 gezelter 3315 \begin{figure}
1228     \centering
1229     \includegraphics[width=\linewidth]{graph}
1230 xsun 3390 \caption[The mean-squared displacements ($\langle r^2(t) \rangle$) and
1231 gezelter 3315 orientational correlation functions ($C_2(t)$) for each of the model
1232     rigid bodies studied. The circles are the results for microcanonical
1233     simulations with explicit solvent molecules, while the other data sets
1234     are results for Langevin dynamics using the different hydrodynamic
1235     tensor approximations. The Perrin model for the ellipsoids is
1236     considered the ``exact'' hydrodynamic behavior (this can also be said
1237     for the translational motion of the dumbbell operating under the bead
1238 xsun 3390 model). In most cases, the various hydrodynamics models reproduce each
1239     other quantitatively]{}
1240 gezelter 3315 \label{fig:results}
1241     \end{figure}
1242    
1243 xsun 3298 \begin{table*}
1244     \begin{minipage}{\linewidth}
1245     \begin{center}
1246 gezelter 3305 \caption{Translational diffusion constants (D) for the model systems
1247     calculated using microcanonical simulations (with explicit solvent),
1248     theoretical predictions, and Langevin simulations (with implicit solvent).
1249 gezelter 3352 Analytical solutions for the exactly-solved hydrodynamics models are obtained
1250     from: Stokes' law (sphere), and Refs. \citen{Perrin1934} and \citen{Perrin1936}
1251 gezelter 3305 (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1252     (dumbbell). The other model systems have no known analytic solution.
1253 gezelter 3352 All diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1254 gezelter 3305 $10^{-4}$ \AA$^2$ / fs). }
1255     \begin{tabular}{lccccccc}
1256 xsun 3298 \hline
1257 gezelter 3305 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1258     \cline{2-3} \cline{5-7}
1259     model & $\eta$ (centipoise) & D & & Analytical & method & Hydrodynamics & simulation \\
1260 xsun 3298 \hline
1261 xsun 3312 sphere & 0.279 & 3.06 & & 2.42 & exact & 2.42 & 2.33 \\
1262 gezelter 3305 ellipsoid & 0.255 & 2.44 & & 2.34 & exact & 2.34 & 2.37 \\
1263     & 0.255 & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1264 xsun 3312 dumbbell & 0.308 & 2.06 & & 1.64 & bead model & 1.65 & 1.62 \\
1265     & 0.308 & 2.06 & & 1.64 & rough shell & 1.59 & 1.62 \\
1266 gezelter 3305 banana & 0.298 & 1.53 & & & rough shell & 1.56 & 1.55 \\
1267 gezelter 3341 lipid & 0.349 & 1.41 & & & rough shell & 1.33 & 1.32 \\
1268 xsun 3298 \end{tabular}
1269     \label{tab:translation}
1270     \end{center}
1271     \end{minipage}
1272     \end{table*}
1273    
1274     \begin{table*}
1275     \begin{minipage}{\linewidth}
1276     \begin{center}
1277 gezelter 3305 \caption{Orientational relaxation times ($\tau$) for the model systems using
1278     microcanonical simulation (with explicit solvent), theoretical
1279     predictions, and Langevin simulations (with implicit solvent). All
1280     relaxation times are for the rotational correlation function with
1281     $\ell = 2$ and are reported in units of ps. The ellipsoidal model has
1282     an exact solution for the orientational correlation time due to
1283     Perrin, but the other model systems have no known analytic solution.}
1284     \begin{tabular}{lccccccc}
1285 xsun 3298 \hline
1286 gezelter 3305 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1287     \cline{2-3} \cline{5-7}
1288     model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic & simulation \\
1289 xsun 3298 \hline
1290 xsun 3312 sphere & 0.279 & & & 9.69 & exact & 9.69 & 9.64 \\
1291 gezelter 3305 ellipsoid & 0.255 & 46.7 & & 22.0 & exact & 22.0 & 22.2 \\
1292     & 0.255 & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1293 xsun 3312 dumbbell & 0.308 & 14.1 & & & bead model & 50.0 & 50.1 \\
1294     & 0.308 & 14.1 & & & rough shell & 41.5 & 41.3 \\
1295 gezelter 3305 banana & 0.298 & 63.8 & & & rough shell & 70.9 & 70.9 \\
1296     lipid & 0.349 & 78.0 & & & rough shell & 76.9 & 77.9 \\
1297     \hline
1298 xsun 3298 \end{tabular}
1299     \label{tab:rotation}
1300     \end{center}
1301     \end{minipage}
1302     \end{table*}
1303    
1304 gezelter 3310 \section{Application: A rigid-body lipid bilayer}
1305    
1306 gezelter 3352 To test the accuracy and efficiency of the new integrator, we applied
1307     it to study the formation of corrugated structures emerging from
1308     simulations of the coarse grained lipid molecular models presented
1309     above. The initial configuration is taken from earlier molecular
1310     dynamics studies on lipid bilayers which had used spherical
1311     (Lennard-Jones) solvent particles and moderate (480 solvated lipid
1312     molecules) system sizes.\cite{SunX._jp0762020} the solvent molecules
1313     were excluded from the system and the box was replicated three times
1314     in the x- and y- axes (giving a single simulation cell containing 4320
1315     lipids). The experimental value for the viscosity of water at 20C
1316     ($\eta = 1.00$ cp) was used with the Langevin integrator to mimic the
1317     hydrodynamic effects of the solvent. The absence of explicit solvent
1318     molecules and the stability of the integrator allowed us to take
1319     timesteps of 50 fs. A simulation run time of 30 ns was sampled to
1320     calculate structural properties. Fig. \ref{fig:bilayer} shows the
1321     configuration of the system after 30 ns. Structural properties of the
1322     bilayer (e.g. the head and body $P_2$ order parameters) are nearly
1323     identical to those obtained via solvated molecular dynamics. The
1324     ripple structure remained stable during the entire trajectory.
1325     Compared with using explicit bead-model solvent molecules, the 30 ns
1326     trajectory for 4320 lipids with the Langevin integrator is now {\it
1327     faster} on the same hardware than the same length trajectory was for
1328     the 480-lipid system previously studied.
1329 xsun 3298
1330 gezelter 3310 \begin{figure}
1331     \centering
1332     \includegraphics[width=\linewidth]{bilayer}
1333 xsun 3390 \caption[A snapshot of a bilayer composed of 4320 rigid-body models
1334     for lipid molecules evolving using the Langevin integrator described
1335     in this work]{} \label{fig:bilayer}
1336 gezelter 3310 \end{figure}
1337    
1338 tim 2746 \section{Conclusions}
1339    
1340 gezelter 3352 We have presented a new algorithm for carrying out Langevin dynamics
1341     simulations on complex rigid bodies by incorporating the hydrodynamic
1342 gezelter 3367 resistance tensors for arbitrary shapes into a stable and efficient
1343 gezelter 3352 integration scheme. The integrator gives quantitative agreement with
1344     both analytic and approximate hydrodynamic theories, and works
1345     reasonably well at reproducing the solute dynamical properties
1346     (diffusion constants, and orientational relaxation times) from
1347     explicitly-solvated simulations. For the cases where there are
1348     discrepancies between our Langevin integrator and the explicit solvent
1349     simulations, two features of molecular simulations help explain the
1350     differences.
1351 tim 2999
1352 gezelter 3352 First, the use of ``stick'' boundary conditions for molecular-sized
1353     solutes in a sea of similarly-sized solvent particles may be
1354     problematic. We are certainly not the first group to notice this
1355     difference between hydrodynamic theories and explicitly-solvated
1356     molecular
1357     simulations.\cite{Schmidt:2004fj,Schmidt:2003kx,Ravichandran:1999fk,TANG:1993lr}
1358     The problem becomes particularly noticable in both the translational
1359     diffusion of the spherical particles and the rotational diffusion of
1360     the ellipsoids. In both of these cases it is clear that the
1361     approximations that go into hydrodynamics are the source of the error,
1362     and not the integrator itself.
1363 tim 2999
1364 gezelter 3352 Second, in the case of structures which have substantial surface area
1365     that is inaccessible to solvent particles, the hydrodynamic theories
1366     (and the Langevin integrator) may overestimate the effects of solvent
1367     friction because they overestimate the exposed surface area of the
1368     rigid body. This is particularly noticable in the rotational
1369     diffusion of the dumbbell model. We believe that given a solvent of
1370     known radius, it may be possible to modify the rough shell approach to
1371     place beads on solvent-accessible surface, instead of on the geometric
1372     surface defined by the van der Waals radii of the components of the
1373     rigid body. Further work to confirm the behavior of this new
1374     approximation is ongoing.
1375    
1376 tim 2746 \section{Acknowledgments}
1377     Support for this project was provided by the National Science
1378     Foundation under grant CHE-0134881. T.L. also acknowledges the
1379 gezelter 3352 financial support from Center of Applied Mathematics at University of
1380     Notre Dame.
1381    
1382     \end{doublespace}
1383 tim 2746 \newpage
1384    
1385 gezelter 3352 \bibliographystyle{jcp2}
1386 tim 2746 \bibliography{langevin}
1387     \end{document}