ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3341
Committed: Wed Feb 6 22:34:50 2008 UTC (16 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 62754 byte(s)
Log Message:
*** empty log message ***

File Contents

# User Rev Content
1 tim 2746 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2     %\documentclass[aps,prb,preprint]{revtex4}
3     \documentclass[11pt]{article}
4     \usepackage{endfloat}
5 gezelter 3337 \usepackage{amsmath}
6 tim 2746 \usepackage{amssymb}
7     \usepackage{times}
8     \usepackage{mathptmx}
9     \usepackage{setspace}
10     \usepackage{tabularx}
11     \usepackage{graphicx}
12     \usepackage{booktabs}
13     \usepackage{bibentry}
14     \usepackage{mathrsfs}
15     \usepackage[ref]{overcite}
16     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18     9.0in \textwidth 6.5in \brokenpenalty=10000
19     \renewcommand{\baselinestretch}{1.2}
20 gezelter 3299 \renewcommand\citemid{\ } % no comma in optional referenc note
21 tim 2746
22     \begin{document}
23    
24 gezelter 3205 \title{An algorithm for performing Langevin dynamics on rigid bodies of arbitrary shape }
25 tim 2746
26 gezelter 3299 \author{Xiuquan Sun, Teng Lin and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
27 tim 2746 gezelter@nd.edu} \\
28     Department of Chemistry and Biochemistry\\
29     University of Notre Dame\\
30     Notre Dame, Indiana 46556}
31    
32     \date{\today}
33    
34     \maketitle \doublespacing
35    
36     \begin{abstract}
37    
38     \end{abstract}
39    
40     \newpage
41    
42     %\narrowtext
43    
44     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45     % BODY OF TEXT
46     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
47    
48     \section{Introduction}
49    
50     %applications of langevin dynamics
51 gezelter 3316 Langevin dynamics, which mimics a simple heat bath with stochastic and
52     dissipative forces, has been applied in a variety of situations as an
53     alternative to molecular dynamics with explicit solvent molecules.
54     The stochastic treatment of the solvent allows the use of simulations
55     with substantially longer time and length scales. In general, the
56     dynamic and structural properties obtained from Langevin simulations
57     agree quite well with similar properties obtained from explicit
58     solvent simulations.
59    
60     Recent examples of the usefulness of Langevin simulations include a
61     study of met-enkephalin in which Langevin simulations predicted
62     dynamical properties that were largely in agreement with explicit
63     solvent simulations.\cite{Shen2002} By applying Langevin dynamics with
64     the UNRES model, Liow and his coworkers suggest that protein folding
65     pathways can be explored within a reasonable amount of
66     time.\cite{Liwo2005}
67    
68     The stochastic nature of Langevin dynamics also enhances the sampling
69     of the system and increases the probability of crossing energy
70     barriers.\cite{Cui2003,Banerjee2004} Combining Langevin dynamics with
71 gezelter 3333 Kramers' theory, Klimov and Thirumalai identified free-energy
72 gezelter 3316 barriers by studying the viscosity dependence of the protein folding
73     rates.\cite{Klimov1997} In order to account for solvent induced
74     interactions missing from the implicit solvent model, Kaya
75     incorporated a desolvation free energy barrier into protein
76     folding/unfolding studies and discovered a higher free energy barrier
77 xsun 3317 between the native and denatured states.\cite{HuseyinKaya07012005}
78 gezelter 3316
79     Because of its stability against noise, Langevin dynamics has also
80     proven useful for studying remagnetization processes in various
81     systems.\cite{Palacios1998,Berkov2002,Denisov2003} [Check: For
82 tim 2746 instance, the oscillation power spectrum of nanoparticles from
83 gezelter 3316 Langevin dynamics has the same peak frequencies for different wave
84     vectors, which recovers the property of magnetic excitations in small
85     finite structures.\cite{Berkov2005a}]
86 tim 2746
87 gezelter 3316 In typical LD simulations, the friction and random forces on
88 gezelter 3333 individual atoms are taken from Stokes' law,
89 gezelter 3316 \begin{eqnarray}
90     m \dot{v}(t) & = & -\nabla U(x) - \xi m v(t) + R(t) \\
91     \langle R(t) \rangle & = & 0 \\
92     \langle R(t) R(t') \rangle & = & 2 k_B T \xi m \delta(t - t')
93     \end{eqnarray}
94     where $\xi \approx 6 \pi \eta a$. Here $\eta$ is the viscosity of the
95     implicit solvent, and $a$ is the hydrodynamic radius of the atom.
96 tim 2746
97 gezelter 3333 The use of rigid substructures,\cite{Chun:2000fj}
98     coarse-graining,\cite{Ayton01,Golubkov06,Orlandi:2006fk,SunGezelter08}
99     and ellipsoidal representations of protein side chains~\cite{Fogolari:1996lr}
100     has made the use of the Stokes-Einstein approximation problematic. A
101     rigid substructure moves as a single unit with orientational as well
102     as translational degrees of freedom. This requires a more general
103 gezelter 3316 treatment of the hydrodynamics than the spherical approximation
104     provides. The atoms involved in a rigid or coarse-grained structure
105     should properly have solvent-mediated interactions with each
106     other. The theory of interactions {\it between} bodies moving through
107     a fluid has been developed over the past century and has been applied
108     to simulations of Brownian
109 gezelter 3333 motion.\cite{FIXMAN:1986lr,Ramachandran1996}
110 tim 2746
111 gezelter 3333 In order to account for the diffusion anisotropy of arbitrarily-shaped
112     particles, Fernandes and Garc\'{i}a de la Torre improved the original
113     Brownian dynamics simulation algorithm~\cite{Ermak1978,Allison1991} by
114     incorporating a generalized $6\times6$ diffusion tensor and
115     introducing a rotational evolution scheme consisting of three
116     consecutive rotations.\cite{Fernandes2002} Unfortunately, biases are
117     introduced into the system due to the arbitrary order of applying the
118     noncommuting rotation operators.\cite{Beard2003} Based on the
119     observation the momentum relaxation time is much less than the time
120     step, one may ignore the inertia in Brownian dynamics. However, the
121     assumption of zero average acceleration is not always true for
122     cooperative motion which is common in proteins. An inertial Brownian
123     dynamics (IBD) was proposed to address this issue by adding an
124     inertial correction term.\cite{Beard2000} As a complement to IBD which
125     has a lower bound in time step because of the inertial relaxation
126     time, long-time-step inertial dynamics (LTID) can be used to
127     investigate the inertial behavior of linked polymer segments in a low
128     friction regime.\cite{Beard2000} LTID can also deal with the
129     rotational dynamics for nonskew bodies without translation-rotation
130     coupling by separating the translation and rotation motion and taking
131     advantage of the analytical solution of hydrodynamics
132     properties. However, typical nonskew bodies like cylinders and
133     ellipsoids are inadequate to represent most complex macromolecular
134     assemblies. There is therefore a need for incorporating the
135     hydrodynamics of complex (and potentially skew) rigid bodies in the
136     library of methods available for performing Langevin simulations.
137    
138 gezelter 3316 \subsection{Rigid Body Dynamics}
139     Rigid bodies are frequently involved in the modeling of large
140     collections of particles that move as a single unit. In molecular
141     simulations, rigid bodies have been used to simplify protein-protein
142 gezelter 3333 docking,\cite{Gray2003} and lipid bilayer
143     simulations.\cite{SunGezelter08} Many of the water models in common
144     use are also rigid-body
145     models,\cite{Jorgensen83,Berendsen81,Berendsen87} although they are
146     typically evolved using constraints rather than rigid body equations
147     of motion.
148 gezelter 3316
149 gezelter 3333 Euler angles are a natural choice to describe the rotational degrees
150 gezelter 3337 of freedom. However, due to $\frac{1}{\sin \theta}$ singularities, the
151 gezelter 3333 numerical integration of corresponding equations of these motion can
152     become inaccurate (and inefficient). Although the use of multiple
153     sets of Euler angles can overcome this problem,\cite{Barojas1973} the
154     computational penalty and the loss of angular momentum conservation
155     remain. A singularity-free representation utilizing quaternions was
156     developed by Evans in 1977.\cite{Evans1977} The Evans quaternion
157     approach uses a nonseparable Hamiltonian, and this has prevented
158     symplectic algorithms from being utilized until very
159     recently.\cite{Miller2002}
160 gezelter 3316
161 gezelter 3333 Another approach is the application of holonomic constraints to the
162     atoms belonging to the rigid body. Each atom moves independently
163     under the normal forces deriving from potential energy and constraints
164     are used to guarantee rigidity. However, due to their iterative
165     nature, the SHAKE and RATTLE algorithms converge very slowly when the
166     number of constraints (and the number of particles that belong to the
167     rigid body) increases.\cite{Ryckaert1977,Andersen1983}
168 tim 2746
169 gezelter 3333 In order to develop a stable and efficient integration scheme that
170     preserves most constants of the motion, symplectic propagators are
171     necessary. By introducing a conjugate momentum to the rotation matrix
172     $Q$ and re-formulating Hamilton's equations, a symplectic
173     orientational integrator, RSHAKE,\cite{Kol1997} was proposed to evolve
174     rigid bodies on a constraint manifold by iteratively satisfying the
175     orthogonality constraint $Q^T Q = 1$. An alternative method using the
176     quaternion representation was developed by Omelyan.\cite{Omelyan1998}
177     However, both of these methods are iterative and suffer from some
178     related inefficiencies. A symplectic Lie-Poisson integrator for rigid
179     bodies developed by Dullweber {\it et al.}\cite{Dullweber1997} removes
180     most of the limitations mentioned above and is therefore the basis for
181     our Langevin integrator.
182 gezelter 3316
183 tim 2746 The goal of the present work is to develop a Langevin dynamics
184 gezelter 3341 algorithm for arbitrary-shaped rigid particles by integrating an
185     accurate estimate of the friction tensor from hydrodynamics theory
186     into a symplectic rigid body dynamics propagator. In the sections
187     below, we review some of the theory of hydrodynamic tensors developed
188 gezelter 3333 primarily for Brownian simulations of multi-particle systems, we then
189     present our integration method for a set of generalized Langevin
190     equations of motion, and we compare the behavior of the new Langevin
191     integrator to dynamical quantities obtained via explicit solvent
192     molecular dynamics.
193 tim 2746
194 gezelter 3316 \subsection{\label{introSection:frictionTensor}The Friction Tensor}
195 gezelter 3341 Theoretically, a complete friction kernel for a solute particle can be
196     determined using the velocity autocorrelation function from a
197     simulation with explicit solvent molecules. However, this approach
198     becomes impractical when the solute becomes complex. Instead, various
199     approaches based on hydrodynamics have been developed to calculate
200     static friction coefficients. In general, the friction tensor $\Xi$ is
201     a $6\times 6$ matrix given by
202 gezelter 3316 \begin{equation}
203 gezelter 3333 \Xi = \left( \begin{array}{*{20}c}
204     \Xi^{tt} & \Xi^{rt} \\
205     \Xi^{tr} & \Xi^{rr} \\
206     \end{array} \right).
207 gezelter 3316 \end{equation}
208     Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
209     rotational resistance (friction) tensors respectively, while
210     $\Xi^{tr}$ is translation-rotation coupling tensor and $\Xi^{rt}$ is
211     rotation-translation coupling tensor. When a particle moves in a
212 gezelter 3341 fluid, it may experience a friction force ($\mathbf{f}_f$) and torque
213     ($\mathbf{\tau}_f$) in opposition to the velocity ($\mathbf{v}$) and
214     body-fixed angular velocity ($\mathbf{\omega}$),
215 gezelter 3316 \begin{equation}
216 tim 2746 \left( \begin{array}{l}
217 xsun 3338 \mathbf{f}_f \\
218 gezelter 3316 \mathbf{\tau}_f \\
219 gezelter 3333 \end{array} \right) = - \left( \begin{array}{*{20}c}
220     \Xi^{tt} & \Xi^{rt} \\
221     \Xi^{tr} & \Xi^{rr} \\
222     \end{array} \right)\left( \begin{array}{l}
223 gezelter 3316 \mathbf{v} \\
224     \mathbf{\omega} \\
225     \end{array} \right).
226     \end{equation}
227 tim 2746
228 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
229 gezelter 3341 For a spherical body under ``stick'' boundary conditions, the
230 gezelter 3316 translational and rotational friction tensors can be calculated from
231 gezelter 3341 Stokes' law,\cite{stokes}
232 gezelter 3316 \begin{equation}
233 gezelter 3333 \Xi^{tt} = \left( \begin{array}{*{20}c}
234 tim 2999 {6\pi \eta R} & 0 & 0 \\
235     0 & {6\pi \eta R} & 0 \\
236     0 & 0 & {6\pi \eta R} \\
237 gezelter 3333 \end{array} \right)
238 gezelter 3316 \end{equation}
239 tim 2999 and
240 gezelter 3316 \begin{equation}
241 gezelter 3333 \Xi^{rr} = \left( \begin{array}{*{20}c}
242 tim 2999 {8\pi \eta R^3 } & 0 & 0 \\
243     0 & {8\pi \eta R^3 } & 0 \\
244     0 & 0 & {8\pi \eta R^3 } \\
245 gezelter 3333 \end{array} \right)
246 gezelter 3316 \end{equation}
247 tim 2999 where $\eta$ is the viscosity of the solvent and $R$ is the
248     hydrodynamic radius.
249    
250     Other non-spherical shapes, such as cylinders and ellipsoids, are
251 gezelter 3341 widely used as references for developing new hydrodynamic theories,
252 tim 2999 because their properties can be calculated exactly. In 1936, Perrin
253 gezelter 3341 extended Stokes' law to general
254     ellipsoids,\cite{Perrin1934,Perrin1936} described in Cartesian
255     coordinates as
256 gezelter 3316 \begin{equation}
257 gezelter 3333 \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1.
258 gezelter 3316 \end{equation}
259 gezelter 3333 Here, the semi-axes are of lengths $a$, $b$, and $c$. Due to the
260     complexity of the elliptic integral, only uniaxial ellipsoids, either
261 gezelter 3341 prolate ($a \ge b = c$) or oblate ($a < b = c$), were solved
262 gezelter 3333 exactly. Introducing an elliptic integral parameter $S$ for prolate,
263 gezelter 3316 \begin{equation}
264     S = \frac{2}{\sqrt{a^2 - b^2}} \ln \frac{a + \sqrt{a^2 - b^2}}{b},
265     \end{equation}
266 gezelter 3333 and oblate,
267 gezelter 3316 \begin{equation}
268     S = \frac{2}{\sqrt {b^2 - a^2 }} \arctan \frac{\sqrt {b^2 - a^2}}{a},
269     \end{equation}
270 gezelter 3341 ellipsoids, it is possible to write down exact solutions for the
271     resistance tensors. As is the case for spherical bodies, the translational,
272     \begin{eqnarray}
273 gezelter 3316 \Xi_a^{tt} & = & 16\pi \eta \frac{a^2 - b^2}{(2a^2 - b^2 )S - 2a}. \\
274     \Xi_b^{tt} = \Xi_c^{tt} & = & 32\pi \eta \frac{a^2 - b^2 }{(2a^2 - 3b^2 )S + 2a},
275 gezelter 3341 \end{eqnarray}
276     and rotational,
277     \begin{eqnarray}
278 gezelter 3316 \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2 - b^2 )b^2}{2a - b^2 S}, \\
279     \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4 - b^4)}{(2a^2 - b^2 )S - 2a}
280 gezelter 3341 \end{eqnarray}
281     resistance tensors are diagonal $3 \times 3$ matrices. For both
282     spherical and ellipsoidal particles, the translation-rotation and
283     rotation-translation coupling tensors are zero.
284 tim 2746
285 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
286 gezelter 3341 There is no analytical solution for the friction tensor for rigid
287     molecules of arbitrary shape. The ellipsoid of revolution and general
288     triaxial ellipsoid models have been widely used to approximate the
289 gezelter 3316 hydrodynamic properties of rigid bodies. However, the mapping from all
290 gezelter 3341 possible ellipsoidal spaces ($r$-space) to all possible combinations
291     of rotational diffusion coefficients ($D$-space) is not
292 gezelter 3316 unique.\cite{Wegener1979} Additionally, because there is intrinsic
293 gezelter 3341 coupling between translational and rotational motion of {\it skew}
294     rigid bodies, general ellipsoids are not always suitable for modeling
295     rigid molecules. A number of studies have been devoted to determining
296     the friction tensor for irregular shapes using methods in which the
297     molecule of interest is modeled with a combination of
298     spheres\cite{Carrasco1999} and the hydrodynamic properties of the
299     molecule are then calculated using a set of two-point interaction
300     tensors. We have found the {\it bead} and {\it rough shell} models of
301     Carrasco and Garc\'{i}a de la Torre to be the most useful of these
302     methods,\cite{Carrasco1999} and we review the basic outline of the
303     rough shell approach here. A more thorough explanation can be found
304     in Ref. \citen{Carrasco1999}.
305 gezelter 3333
306 gezelter 3341 Consider a rigid assembly of $N$ small beads moving through a
307     continuous medium. Due to hydrodynamic interactions between the
308     beads, the net velocity of the $i^\mathrm{th}$ bead relative to the
309     medium, ${\bf v}'_i$, is different than its unperturbed velocity ${\bf
310     v}_i$,
311 gezelter 3333 \begin{equation}
312 gezelter 3341 {\bf v}'_i = {\bf v}_i - \sum\limits_{j \ne i} {{\bf T}_{ij} {\bf F}_j }
313 gezelter 3333 \end{equation}
314 gezelter 3341 where ${\bf F}_j$ is the frictional force on the medium due to bead $j$, and
315     ${\bf T}_{ij}$ is the hydrodynamic interaction tensor between the two beads.
316     The frictional force felt by the $i^\mathrm{th}$ bead is proportional to
317     its net velocity
318 tim 2746 \begin{equation}
319 gezelter 3341 {\bf F}_i = \zeta_i {\bf v}_i - \zeta _i \sum\limits_{j \ne i} {{\bf T}_{ij} {\bf F}_j }.
320 tim 2999 \label{introEquation:tensorExpression}
321 tim 2746 \end{equation}
322 gezelter 3341 Eq. (\ref{introEquation:tensorExpression}) defines the two-point
323     hydrodynamic tensor, ${\bf T}_{ij}$. There have been many proposed
324     solutions to this equation, including the simple solution given by
325     Oseen and Burgers in 1930 for two beads of identical radius. A second
326     order expression for beads of different hydrodynamic radii was
327 tim 2999 introduced by Rotne and Prager\cite{Rotne1969} and improved by
328     Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
329 tim 2746 \begin{equation}
330 gezelter 3341 {\bf T}_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {{\bf I} +
331     \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right) + \frac{{\sigma
332     _i^2 + \sigma _j^2 }}{{R_{ij}^2 }}\left( {\frac{{\bf I}}{3} -
333     \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
334 tim 2999 \label{introEquation:RPTensorNonOverlapped}
335 tim 2746 \end{equation}
336 gezelter 3341 Here ${\bf R}_{ij}$ is the distance vector between beads $i$ and $j$. Both
337     the Oseen-Burgers tensor and
338     Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption that
339     the beads do not overlap ($R_{ij} \ge \sigma _i + \sigma _j$).
340    
341     To calculate the resistance tensor for a body represented as the union
342     of many non-overlapping beads, we first pick an arbitrary origin $O$
343     and then construct a $3N \times 3N$ supermatrix consisting of $N
344     \times N$ ${\bf B}_{ij}$ blocks
345 tim 2746 \begin{equation}
346 gezelter 3341 {\bf B} = \left( \begin{array}{*{20}c}
347     {\bf B}_{11} & \ldots & {\bf B}_{1N} \\
348     \vdots & \ddots & \vdots \\
349     {\bf B}_{N1} & \cdots & {\bf B}_{NN}
350     \end{array} \right)
351 tim 2746 \end{equation}
352 gezelter 3341 ${\bf B}_{ij}$ is a version of the hydrodynamic tensor which includes the
353     self-contributions for spheres,
354 tim 2999 \begin{equation}
355 gezelter 3341 {\bf B}_{ij} = \delta _{ij} \frac{{\bf I}}{{6\pi \eta R_{ij}}} + (1 - \delta_{ij}
356     ){\bf T}_{ij}
357     \end{equation}
358     where $\delta_{ij}$ is the Kronecker delta function. Inverting the
359     ${\bf B}$ matrix, we obtain
360     \begin{equation}
361     {\bf C} = {\bf B}^{ - 1} = \left(\begin{array}{*{20}c}
362     {\bf C}_{11} & \ldots & {\bf C}_{1N} \\
363     \vdots & \ddots & \vdots \\
364     {\bf C}_{N1} & \cdots & {\bf C}_{NN}
365 gezelter 3333 \end{array} \right),
366 tim 2999 \end{equation}
367 gezelter 3341 which can be partitioned into $N \times N$ blocks labeled ${\bf C}_{ij}$.
368     (Each of the ${\bf C}_{ij}$ blocks is a $3 \times 3$ matrix.) Using the
369     skew matrix,
370 gezelter 3333 \begin{equation}
371 gezelter 3341 {\bf U}_i = \left(\begin{array}{*{20}c}
372     0 & -z_i & y_i \\
373     z_i & 0 & - x_i \\
374     -y_i & x_i & 0
375     \end{array}\right)
376     \label{eq:skewMatrix}
377 gezelter 3333 \end{equation}
378 tim 2999 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
379 gezelter 3341 bead $i$ and origin $O$, the elements of the resistance tensor (at the
380     arbitrary origin $O$) can be written as
381 tim 2999 \begin{eqnarray}
382 xsun 3339 \label{introEquation:ResistanceTensorArbitraryOrigin}
383 gezelter 3341 \Xi^{tt} & = & \sum\limits_i {\sum\limits_j {{\bf C}_{ij} } } \notag , \\
384     \Xi^{tr} = \Xi _{}^{rt} & = & \sum\limits_i {\sum\limits_j {{\bf U}_i {\bf C}_{ij} } } , \\
385     \Xi^{rr} & = & -\sum\limits_i \sum\limits_j {\bf U}_i {\bf C}_{ij} {\bf U}_j + 6 \eta V {\bf I}. \notag
386 tim 2999 \end{eqnarray}
387 gezelter 3341 The final term in the expression for $\Xi^{rr}$ is a correction that
388     accounts for errors in the rotational motion of the bead models. The
389     additive correction uses the solvent viscosity ($\eta$) as well as the
390     total volume of the beads that contribute to the hydrodynamic model,
391 gezelter 3310 \begin{equation}
392     V = \frac{4 \pi}{3} \sum_{i=1}^{N} \sigma_i^3,
393     \end{equation}
394     where $\sigma_i$ is the radius of bead $i$. This correction term was
395     rigorously tested and compared with the analytical results for
396 gezelter 3341 two-sphere and ellipsoidal systems by Garc\'{i}a de la Torre and
397 gezelter 3310 Rodes.\cite{Torre:1983lr}
398    
399 gezelter 3341 In general, resistance tensors depend on the origin at which they were
400     computed. However, the proper location for applying the friction
401     force is the center of resistance, the special point at which the
402     trace of rotational resistance tensor, $\Xi^{rr}$ reaches a minimum
403     value. Mathematically, the center of resistance can also be defined
404     as the unique point for a rigid body at which the translation-rotation
405     coupling tensors are symmetric,
406 tim 2999 \begin{equation}
407 gezelter 3341 \Xi^{tr} = \left(\Xi^{tr}\right)^T
408 tim 2999 \label{introEquation:definitionCR}
409     \end{equation}
410 gezelter 3341 From Eq. \ref{introEquation:ResistanceTensorArbitraryOrigin}, we can
411     easily derive that the {\it translational} resistance tensor is origin
412     independent, while the rotational resistance tensor and
413 tim 2999 translation-rotation coupling resistance tensor depend on the
414 gezelter 3341 origin. Given the resistance tensor at an arbitrary origin $O$, and a
415     vector ,${\bf r}_{OP} = (x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we
416     can obtain the resistance tensor at $P$ by
417     \begin{eqnarray}
418     \label{introEquation:resistanceTensorTransformation}
419     \Xi_P^{tt} & = & \Xi_O^{tt} \notag \\
420     \Xi_P^{tr} = \Xi_P^{rt} & = & \Xi_O^{tr} - {\bf U}_{OP} \Xi _O^{tt} \\
421     \Xi_P^{rr} & = &\Xi_O^{rr} - {\bf U}_{OP} \Xi_O^{tt} {\bf U}_{OP}
422     + \Xi_O^{tr} {\bf U}_{OP} - {\bf U}_{OP} \left( \Xi_O^{tr}
423     \right)^{^T} \notag
424     \end{eqnarray}
425     where ${\bf U}_{OP}$ is the skew matrix (Eq. (\ref{eq:skewMatrix}))
426     for the vector between the origin $O$ and the point $P$. Using
427     Eqs.~\ref{introEquation:definitionCR}~and~\ref{introEquation:resistanceTensorTransformation},
428     one can locate the position of center of resistance,
429     \begin{equation*}
430     \left(\begin{array}{l}
431     x_{OR} \\
432     y_{OR} \\
433     z_{OR}
434     \end{array}\right) =
435     \left(\begin{array}{*{20}c}
436     (\Xi_O^{rr})_{yy} + (\Xi_O^{rr})_{zz} & -(\Xi_O^{rr})_{xy} & -(\Xi_O^{rr})_{xz} \\
437     -(\Xi_O^{rr})_{xy} & (\Xi_O^{rr})_{zz} + (\Xi_O^{rr})_{xx} & -(\Xi_O^{rr})_{yz} \\
438     -(\Xi_O^{rr})_{xz} & -(\Xi_O^{rr})_{yz} & (\Xi_O^{rr})_{xx} + (\Xi_O^{rr})_{yy} \\
439     \end{array}\right)^{-1}
440     \left(\begin{array}{l}
441     (\Xi_O^{tr})_{yz} - (\Xi_O^{tr})_{zy} \\
442     (\Xi_O^{tr})_{zx} - (\Xi_O^{tr})_{xz} \\
443     (\Xi_O^{tr})_{xy} - (\Xi_O^{tr})_{yx}
444     \end{array}\right)
445     \end{equation*}
446 xsun 3338 where $x_{OR}$, $y_{OR}$, $z_{OR}$ are the components of the vector
447 tim 2999 joining center of resistance $R$ and origin $O$.
448 tim 2746
449 gezelter 3341 For a general rigid molecular substructure, finding the $6 \times 6$
450     resistance tensor can be a computationally demanding task. First, a
451     lattice of small beads that extends well beyond the boundaries of the
452     rigid substructure is created. The lattice is typically composed of
453     0.25 \AA\ beads on a dense FCC lattice. The lattice constant is taken
454     to be the bead diameter, so that adjacent beads are touching, but do
455     not overlap. To make a shape corresponding to the rigid structure,
456     beads that sit on lattice sites that are outside the van der Waals
457     radii of any atoms comprising the rigid body are excluded from the
458     calculation.
459 tim 2746
460 gezelter 3341 For large structures, most of the beads will be deep within the rigid
461     body and will not contribute to the hydrodynamic tensor. In the {\it
462     rough shell} approach, beads which have all of their lattice neighbors
463     inside the structure are considered interior beads, and are removed
464     from the calculation. After following this procedure, only those
465     beads in direct contact with the van der Waals surface of the rigid
466     body are retained. For reasonably large molecular structures, this
467     truncation can still produce bead assemblies with thousands of
468     members.
469    
470     If all of the atoms comprising the rigid substructure are spherical
471     and non-overlapping, the tensor in
472     Eq.~(\ref{introEquation:RPTensorNonOverlapped}) may be used directly
473     using the atoms themselves as the hydrodynamic beads. This is a
474     variant of the {\it bead model} approach of Carrasco and Garc\'{i}a de
475     la Torre.\cite{Carrasco1999} In this case, the size of the ${\bf B}$ matrix
476     can be quite small, and the calculation of the hydrodynamic tensor is
477     straightforward.
478    
479     In general, the inversion of the ${\bf B}$ matrix is the most
480     computationally demanding task. This inversion is done only once for
481     each type of rigid structure. We have been using straightforward
482     LU-decomposition to solve the linear system and obtain the elements of
483     ${\bf C}$. Once ${\bf C}$ has been obtained, the location of the
484     center of resistance ($R$) is found and the resistance tensor at this
485     point is calculated. The $3 \times 1$ vector giving the location of
486     the rigid body's center of resistance and the $6 \times 6$ resistance
487     tensor are stored for use in the Langevin dynamics calculation. Note
488     that these quantities depend on solvent viscosity and temperature and
489     must be recomputed if different simulation conditions are required.
490    
491 gezelter 3310 \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
492 gezelter 3337
493 tim 2999 Consider the Langevin equations of motion in generalized coordinates
494 tim 2746 \begin{equation}
495 gezelter 3341 {\bf M} \dot{{\bf V}}(t) = {\bf F}_{s}(t) +
496     {\bf F}_{f}(t) + {\bf F}_{r}(t)
497 tim 2746 \label{LDGeneralizedForm}
498     \end{equation}
499 gezelter 3341 where ${\bf M}$ is a $6 \times 6$ diagonal mass matrix (which
500 gezelter 3337 includes the mass of the rigid body as well as the moments of inertia
501 gezelter 3341 in the body-fixed frame) and ${\bf V}$ is a generalized velocity,
502     ${\bf V} =
503     \left\{{\bf v},{\bf \omega}\right\}$. The right side of
504 gezelter 3333 Eq.~\ref{LDGeneralizedForm} consists of three generalized forces: a
505 gezelter 3341 system force (${\bf F}_{s}$), a frictional or dissipative force
506     (${\bf F}_{f}$) and stochastic force (${\bf F}_{r}$). While the
507 gezelter 3340 evolution of the system in Newtonian mechanics is typically done in
508 gezelter 3341 the lab frame, it is convenient to handle the dynamics of rigid bodies
509     in body-fixed frames. Thus the friction and random forces on each
510     substructure are calculated in a body-fixed frame and may converted
511     back to the lab frame using that substructure's rotation matrix (${\bf
512     Q}$):
513 gezelter 3333 \begin{equation}
514 gezelter 3341 {\bf F}_{f,r} =
515 gezelter 3340 \left( \begin{array}{c}
516 gezelter 3341 {\bf f}_{f,r} \\
517     {\bf \tau}_{f,r}
518 gezelter 3340 \end{array} \right)
519     =
520     \left( \begin{array}{c}
521 gezelter 3341 {\bf Q}^{T} {\bf f}^{~b}_{f,r} \\
522     {\bf Q}^{T} {\bf \tau}^{~b}_{f,r}
523 gezelter 3340 \end{array} \right)
524 gezelter 3333 \end{equation}
525 gezelter 3341 The body-fixed friction force, ${\bf F}_{f}^{~b}$, is proportional to
526     the (body-fixed) velocity at the center of resistance
527     ${\bf v}_{R}^{~b}$ and the angular velocity ${\bf \omega}$
528 tim 2746 \begin{equation}
529 gezelter 3341 {\bf F}_{f}^{~b}(t) = \left( \begin{array}{l}
530     {\bf f}_{f}^{~b}(t) \\
531     {\bf \tau}_{f}^{~b}(t) \\
532 gezelter 3333 \end{array} \right) = - \left( \begin{array}{*{20}c}
533 gezelter 3341 \Xi_{R}^{tt} & \Xi_{R}^{rt} \\
534     \Xi_{R}^{tr} & \Xi_{R}^{rr} \\
535 gezelter 3333 \end{array} \right)\left( \begin{array}{l}
536 gezelter 3341 {\bf v}_{R}^{~b}(t) \\
537     {\bf \omega}(t) \\
538 tim 2746 \end{array} \right),
539     \end{equation}
540 gezelter 3341 while the random force, ${\bf F}_{r}$, is a Gaussian stochastic
541     variable with zero mean and variance,
542 tim 2746 \begin{equation}
543 gezelter 3341 \left\langle {{\bf F}_{r}(t) ({\bf F}_{r}(t'))^T } \right\rangle =
544     \left\langle {{\bf F}_{r}^{~b} (t) ({\bf F}_{r}^{~b} (t'))^T } \right\rangle =
545     2 k_B T \Xi_R \delta(t - t'). \label{eq:randomForce}
546 tim 2746 \end{equation}
547 gezelter 3340 $\Xi_R$ is the $6\times6$ resistance tensor at the center of
548 gezelter 3341 resistance. Once this tensor is known for a given rigid body (as
549     described in the previous section) obtaining a stochastic vector that
550     has the properties in Eq. (\ref{eq:randomForce}) can be done
551     efficiently by carrying out a one-time Cholesky decomposition to
552     obtain the square root matrix of the resistance tensor,
553     \begin{equation}
554     \Xi_R = {\bf S} {\bf S}^{T},
555     \label{eq:Cholesky}
556     \end{equation}
557     where ${\bf S}$ is a lower triangular matrix.\cite{SchlickBook} A
558     vector with the statistics required for the random force can then be
559     obtained by multiplying ${\bf S}$ onto a random 6-vector ${\bf Z}$ which
560     has elements chosen from a Gaussian distribution, such that:
561 gezelter 3340 \begin{equation}
562 gezelter 3341 \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
563     {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
564 gezelter 3340 \end{equation}
565 gezelter 3341 where $\delta t$ is the timestep in use during the simulation. The
566     random force, ${\bf F}_{r}^{~b} = {\bf S} {\bf Z}$, can be shown to have the
567     correct properties required by Eq. (\ref{eq:randomForce}).
568 gezelter 3333
569 gezelter 3341 The equation of motion for the translational velocity of the center of
570     mass (${\bf v}$) can be written as
571 tim 2746 \begin{equation}
572 gezelter 3341 m \dot{{\bf v}} (t) = {\bf f}_{s}(t) + {\bf f}_{f}(t) +
573     {\bf f}_{r}(t)
574 tim 2746 \end{equation}
575 gezelter 3341 Since the frictional and random forces are applied at the center of
576     resistance, which generally does not coincide with the center of mass,
577     extra torques are exerted at the center of mass. Thus, the net
578     body-fixed torque at the center of mass, $\tau^{~b}(t)$,
579     is given by
580 tim 2746 \begin{equation}
581 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)
582 tim 2746 \end{equation}
583 gezelter 3341 where ${\bf r}_{MR}$ is the vector from the center of mass to the center of
584     resistance. Instead of integrating the angular velocity in lab-fixed
585     frame, we consider the equation of motion for the angular momentum
586     (${\bf j}$) in the body-fixed frame
587 tim 2746 \begin{equation}
588 gezelter 3341 \dot{\bf j}(t) = \tau^{~b}(t)
589 tim 2746 \end{equation}
590 gezelter 3341 Embedding the friction and random forces into the the total force and
591     torque, one can integrate the Langevin equations of motion for a rigid
592     body of arbitrary shape in a velocity-Verlet style 2-part algorithm,
593     where $h = \delta t$:
594 tim 2746
595 gezelter 3341 {\tt move A:}
596 tim 2746 \begin{align*}
597 tim 2999 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
598     + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
599     %
600     {\bf r}(t + h) &\leftarrow {\bf r}(t)
601     + h {\bf v}\left(t + h / 2 \right), \\
602     %
603     {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
604 gezelter 3341 + \frac{h}{2} {\bf \tau}^{~b}(t), \\
605 tim 2999 %
606 gezelter 3341 {\bf Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
607 tim 2999 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
608 tim 2746 \end{align*}
609 gezelter 3341 In this context, $\overleftrightarrow{\mathsf{I}}$ is the diagonal
610     moment of inertia tensor, and the $\mathrm{rotate}$ function is the
611     reversible product of the three body-fixed rotations,
612 tim 2746 \begin{equation}
613     \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
614     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
615     / 2) \cdot \mathsf{G}_x(a_x /2),
616     \end{equation}
617     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
618 gezelter 3341 rotates both the rotation matrix ($\mathbf{Q}$) and the body-fixed
619 tim 2999 angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
620     axis $\alpha$,
621 tim 2746 \begin{equation}
622     \mathsf{G}_\alpha( \theta ) = \left\{
623     \begin{array}{lcl}
624 gezelter 3341 \mathbf{Q}(t) & \leftarrow & \mathbf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
625 tim 2746 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
626     j}(0).
627     \end{array}
628     \right.
629     \end{equation}
630     $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
631     rotation matrix. For example, in the small-angle limit, the
632     rotation matrix around the body-fixed x-axis can be approximated as
633     \begin{equation}
634     \mathsf{R}_x(\theta) \approx \left(
635     \begin{array}{ccc}
636     1 & 0 & 0 \\
637     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
638     \theta^2 / 4} \\
639     0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
640     \theta^2 / 4}
641     \end{array}
642     \right).
643     \end{equation}
644 tim 2999 All other rotations follow in a straightforward manner. After the
645     first part of the propagation, the forces and body-fixed torques are
646 gezelter 3341 calculated at the new positions and orientations. The system forces
647     and torques are derivatives of the total potential energy function
648     ($U$) with respect to the rigid body positions (${\bf r}$) and the
649     columns of the transposed rotation matrix ${\bf Q}^T = \left({\bf
650     u}_x, {\bf u}_y, {\bf u}_z \right)$:
651 tim 2746
652 gezelter 3341 {\tt Forces:}
653 tim 2999 \begin{align*}
654 gezelter 3341 {\bf f}_{s}(t + h) & \leftarrow
655     - \left(\frac{\partial U}{\partial {\bf r}}\right)_{{\bf r}(t + h)} \\
656 tim 2999 %
657 gezelter 3341 {\bf \tau}_{s}(t + h) &\leftarrow {\bf u}(t + h)
658     \times \frac{\partial U}{\partial {\bf u}} \\
659 tim 2999 %
660 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) \\
661     %
662     {\bf f}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tt} \cdot
663     {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rt} \cdot {\bf \omega}(t+h) \\
664     %
665     {\bf \tau}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tr} \cdot
666     {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rr} \cdot {\bf \omega}(t+h) \\
667     %
668     Z & \leftarrow {\tt GaussianNormal}(2 k_B T / h, 6) \\
669     {\bf F}_{R,r}^{b}(t+h) & \leftarrow {\bf S} \cdot Z \\
670     %
671     {\bf f}(t+h) & \leftarrow {\bf f}_{s}(t+h) + \mathbf{Q}^{T}(t+h)
672     \cdot \left({\bf f}_{R,f}^{~b} + {\bf f}_{R,r}^{~b} \right) \\
673     %
674     \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) \\
675     \tau^{~b}(t+h) & \leftarrow \mathbf{Q}(t+h) \cdot \tau(t+h) \\
676 tim 2999 \end{align*}
677 gezelter 3341 Frictional (and random) forces and torques must be computed at the
678     center of resistance, so there are additional steps required to find
679     the body-fixed velocity (${\bf v}_{R}^{~b}$) at this location. Mapping
680     the frictional and random forces at the center of resistance back to
681     the center of mass also introduces an additional term in the torque
682     one obtains at the center of mass.
683    
684 tim 2746 Once the forces and torques have been obtained at the new time step,
685     the velocities can be advanced to the same time value.
686    
687 gezelter 3341 {\tt move B:}
688 tim 2746 \begin{align*}
689 tim 2999 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
690     \right)
691     + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
692     %
693     {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
694     \right)
695 gezelter 3341 + \frac{h}{2} {\bf \tau}^{~b}(t + h) .
696 tim 2746 \end{align*}
697    
698 gezelter 3310 \section{Validating the Method\label{sec:validating}}
699 gezelter 3302 In order to validate our Langevin integrator for arbitrarily-shaped
700 gezelter 3305 rigid bodies, we implemented the algorithm in {\sc
701     oopse}\cite{Meineke2005} and compared the results of this algorithm
702     with the known
703 gezelter 3302 hydrodynamic limiting behavior for a few model systems, and to
704     microcanonical molecular dynamics simulations for some more
705     complicated bodies. The model systems and their analytical behavior
706     (if known) are summarized below. Parameters for the primary particles
707     comprising our model systems are given in table \ref{tab:parameters},
708     and a sketch of the arrangement of these primary particles into the
709 gezelter 3305 model rigid bodies is shown in figure \ref{fig:models}. In table
710     \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
711     ellipsoidal (Gay-Berne) particles. For spherical particles, the value
712     of the Lennard-Jones $\sigma$ parameter is the particle diameter
713     ($d$). Gay-Berne ellipsoids have an energy scaling parameter,
714     $\epsilon^s$, which describes the well depth for two identical
715     ellipsoids in a {\it side-by-side} configuration. Additionally, a
716     well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
717     describes the ratio between the well depths in the {\it end-to-end}
718     and side-by-side configurations. For spheres, $\epsilon^r \equiv 1$.
719     Moments of inertia are also required to describe the motion of primary
720     particles with orientational degrees of freedom.
721 gezelter 3299
722 gezelter 3302 \begin{table*}
723     \begin{minipage}{\linewidth}
724     \begin{center}
725     \caption{Parameters for the primary particles in use by the rigid body
726     models in figure \ref{fig:models}.}
727     \begin{tabular}{lrcccccccc}
728     \hline
729     & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
730     & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
731     $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
732 gezelter 3308 Sphere & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
733 gezelter 3302 Ellipsoid & & 4.6 & 13.8 & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
734 gezelter 3308 Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
735 gezelter 3302 Banana &(3 identical ellipsoids)& 4.2 & 11.2 & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
736     Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
737     & Ellipsoidal Tail & 4.6 & 13.8 & 0.8 & 0.2 & 760 & 45000 & 45000 & 9000 \\
738     Solvent & & 4.7 & $= d$ & 0.8 & 1 & 72.06 & & & \\
739     \hline
740     \end{tabular}
741     \label{tab:parameters}
742     \end{center}
743     \end{minipage}
744     \end{table*}
745    
746 gezelter 3305 \begin{figure}
747     \centering
748     \includegraphics[width=3in]{sketch}
749     \caption[Sketch of the model systems]{A sketch of the model systems
750     used in evaluating the behavior of the rigid body Langevin
751     integrator.} \label{fig:models}
752     \end{figure}
753    
754 gezelter 3302 \subsection{Simulation Methodology}
755     We performed reference microcanonical simulations with explicit
756     solvents for each of the different model system. In each case there
757     was one solute model and 1929 solvent molecules present in the
758     simulation box. All simulations were equilibrated using a
759     constant-pressure and temperature integrator with target values of 300
760     K for the temperature and 1 atm for pressure. Following this stage,
761     further equilibration and sampling was done in a microcanonical
762 gezelter 3305 ensemble. Since the model bodies are typically quite massive, we were
763 gezelter 3310 able to use a time step of 25 fs.
764    
765     The model systems studied used both Lennard-Jones spheres as well as
766     uniaxial Gay-Berne ellipoids. In its original form, the Gay-Berne
767     potential was a single site model for the interactions of rigid
768     ellipsoidal molecules.\cite{Gay81} It can be thought of as a
769     modification of the Gaussian overlap model originally described by
770     Berne and Pechukas.\cite{Berne72} The potential is constructed in the
771     familiar form of the Lennard-Jones function using
772     orientation-dependent $\sigma$ and $\epsilon$ parameters,
773     \begin{equation*}
774 gezelter 3341 V_{ij}({{\bf \hat u}_i}, {{\bf \hat u}_j}, {{\bf \hat
775     r}_{ij}}) = 4\epsilon ({{\bf \hat u}_i}, {{\bf \hat u}_j},
776     {{\bf \hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u
777 gezelter 3310 }_i},
778 gezelter 3341 {{\bf \hat u}_j}, {{\bf \hat r}_{ij}})+\sigma_0}\right)^{12}
779     -\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u}_i}, {{\bf \hat u}_j},
780     {{\bf \hat r}_{ij}})+\sigma_0}\right)^6\right]
781 gezelter 3310 \label{eq:gb}
782     \end{equation*}
783    
784     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
785     \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
786     \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
787     are dependent on the relative orientations of the two ellipsoids (${\bf
788     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
789     inter-ellipsoid separation (${\bf \hat{r}}_{ij}$). The shape and
790     attractiveness of each ellipsoid is governed by a relatively small set
791     of parameters: $l$ and $d$ describe the length and width of each
792     uniaxial ellipsoid, while $\epsilon^s$, which describes the well depth
793     for two identical ellipsoids in a {\it side-by-side} configuration.
794     Additionally, a well depth aspect ratio, $\epsilon^r = \epsilon^e /
795     \epsilon^s$, describes the ratio between the well depths in the {\it
796     end-to-end} and side-by-side configurations. Details of the potential
797     are given elsewhere,\cite{Luckhurst90,Golubkov06,SunGezelter08} and an
798     excellent overview of the computational methods that can be used to
799     efficiently compute forces and torques for this potential can be found
800     in Ref. \citen{Golubkov06}
801    
802     For the interaction between nonequivalent uniaxial ellipsoids (or
803     between spheres and ellipsoids), the spheres are treated as ellipsoids
804     with an aspect ratio of 1 ($d = l$) and with an well depth ratio
805     ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of the
806     Gay-Berne potential we are using was generalized by Cleaver {\it et
807     al.} and is appropriate for dissimilar uniaxial
808     ellipsoids.\cite{Cleaver96}
809    
810     A switching function was applied to all potentials to smoothly turn
811     off the interactions between a range of $22$ and $25$ \AA. The
812     switching function was the standard (cubic) function,
813 gezelter 3302 \begin{equation}
814     s(r) =
815     \begin{cases}
816     1 & \text{if $r \le r_{\text{sw}}$},\\
817     \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
818     {(r_{\text{cut}} - r_{\text{sw}})^3}
819     & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
820     0 & \text{if $r > r_{\text{cut}}$.}
821     \end{cases}
822     \label{eq:switchingFunc}
823     \end{equation}
824 gezelter 3310
825 gezelter 3302 To measure shear viscosities from our microcanonical simulations, we
826     used the Einstein form of the pressure correlation function,\cite{hess:209}
827     \begin{equation}
828 gezelter 3310 \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
829     \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \right\rangle_{t_0}.
830 gezelter 3302 \label{eq:shear}
831     \end{equation}
832     A similar form exists for the bulk viscosity
833     \begin{equation}
834 gezelter 3310 \kappa = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
835 gezelter 3302 \int_{t_0}^{t_0 + t}
836 gezelter 3310 \left(P\left(t'\right)-\left\langle P \right\rangle \right)dt'
837     \right)^2 \right\rangle_{t_0}.
838 gezelter 3302 \end{equation}
839     Alternatively, the shear viscosity can also be calculated using a
840     Green-Kubo formula with the off-diagonal pressure tensor correlation function,
841     \begin{equation}
842 gezelter 3310 \eta = \frac{V}{k_B T} \int_0^{\infty} \left\langle P_{xz}(t_0) P_{xz}(t_0
843     + t) \right\rangle_{t_0} dt,
844 gezelter 3302 \end{equation}
845     although this method converges extremely slowly and is not practical
846     for obtaining viscosities from molecular dynamics simulations.
847    
848     The Langevin dynamics for the different model systems were performed
849     at the same temperature as the average temperature of the
850     microcanonical simulations and with a solvent viscosity taken from
851 gezelter 3305 Eq. (\ref{eq:shear}) applied to these simulations. We used 1024
852     independent solute simulations to obtain statistics on our Langevin
853     integrator.
854 gezelter 3302
855     \subsection{Analysis}
856    
857     The quantities of interest when comparing the Langevin integrator to
858     analytic hydrodynamic equations and to molecular dynamics simulations
859     are typically translational diffusion constants and orientational
860     relaxation times. Translational diffusion constants for point
861     particles are computed easily from the long-time slope of the
862     mean-square displacement,
863     \begin{equation}
864 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,
865 gezelter 3302 \end{equation}
866     of the solute molecules. For models in which the translational
867 gezelter 3305 diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
868     (i.e. any non-spherically-symmetric rigid body), it is possible to
869     compute the diffusive behavior for motion parallel to each body-fixed
870     axis by projecting the displacement of the particle onto the
871     body-fixed reference frame at $t=0$. With an isotropic solvent, as we
872     have used in this study, there are differences between the three
873 gezelter 3302 diffusion constants, but these must converge to the same value at
874     longer times. Translational diffusion constants for the different
875 gezelter 3305 shaped models are shown in table \ref{tab:translation}.
876 gezelter 3302
877 gezelter 3305 In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
878 gezelter 3302 diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
879     {\it around} a particular body-fixed axis and {\it not} the diffusion
880     of a vector pointing along the axis. However, these eigenvalues can
881     be combined to find 5 characteristic rotational relaxation
882 gezelter 3305 times,\cite{PhysRev.119.53,Berne90}
883 gezelter 3302 \begin{eqnarray}
884 gezelter 3305 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
885     1 / \tau_2 & = & 6 D_r - 2 \Delta \\
886     1 / \tau_3 & = & 3 (D_r + D_1) \\
887     1 / \tau_4 & = & 3 (D_r + D_2) \\
888     1 / \tau_5 & = & 3 (D_r + D_3)
889 gezelter 3302 \end{eqnarray}
890     where
891     \begin{equation}
892     D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
893     \end{equation}
894     and
895     \begin{equation}
896 gezelter 3305 \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
897 gezelter 3302 \end{equation}
898 gezelter 3305 Each of these characteristic times can be used to predict the decay of
899     part of the rotational correlation function when $\ell = 2$,
900 gezelter 3302 \begin{equation}
901 gezelter 3305 C_2(t) = \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
902 gezelter 3302 \end{equation}
903 gezelter 3305 This is the same as the $F^2_{0,0}(t)$ correlation function that
904     appears in Ref. \citen{Berne90}. The amplitudes of the two decay
905     terms are expressed in terms of three dimensionless functions of the
906     eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
907     2\Delta)$, and $N = 2 \sqrt{\Delta b}$. Similar expressions can be
908     obtained for other angular momentum correlation
909     functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
910     studied, only one of the amplitudes of the two decay terms was
911     non-zero, so it was possible to derive a single relaxation time for
912     each of the hydrodynamic tensors. In many cases, these characteristic
913     times are averaged and reported in the literature as a single relaxation
914     time,\cite{Garcia-de-la-Torre:1997qy}
915 gezelter 3302 \begin{equation}
916 gezelter 3305 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
917     \end{equation}
918     although for the cases reported here, this averaging is not necessary
919     and only one of the five relaxation times is relevant.
920    
921     To test the Langevin integrator's behavior for rotational relaxation,
922     we have compared the analytical orientational relaxation times (if
923     they are known) with the general result from the diffusion tensor and
924     with the results from both the explicitly solvated molecular dynamics
925     and Langevin simulations. Relaxation times from simulations (both
926     microcanonical and Langevin), were computed using Legendre polynomial
927     correlation functions for a unit vector (${\bf u}$) fixed along one or
928     more of the body-fixed axes of the model.
929     \begin{equation}
930 gezelter 3310 C_{\ell}(t) = \left\langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
931     u}_{i}(0) \right) \right\rangle
932 gezelter 3302 \end{equation}
933     For simulations in the high-friction limit, orientational correlation
934     times can then be obtained from exponential fits of this function, or by
935     integrating,
936     \begin{equation}
937 gezelter 3305 \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
938 gezelter 3302 \end{equation}
939 gezelter 3305 In lower-friction solvents, the Legendre correlation functions often
940     exhibit non-exponential decay, and may not be characterized by a
941     single decay constant.
942 gezelter 3302
943     In table \ref{tab:rotation} we show the characteristic rotational
944     relaxation times (based on the diffusion tensor) for each of the model
945     systems compared with the values obtained via microcanonical and Langevin
946     simulations.
947    
948 gezelter 3305 \subsection{Spherical particles}
949 gezelter 3299 Our model system for spherical particles was a Lennard-Jones sphere of
950     diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
951     4.7 \AA). The well depth ($\epsilon$) for both particles was set to
952 gezelter 3302 an arbitrary value of 0.8 kcal/mol.
953 gezelter 3299
954     The Stokes-Einstein behavior of large spherical particles in
955     hydrodynamic flows is well known, giving translational friction
956     coefficients of $6 \pi \eta R$ (stick boundary conditions) and
957 gezelter 3302 rotational friction coefficients of $8 \pi \eta R^3$. Recently,
958     Schmidt and Skinner have computed the behavior of spherical tag
959     particles in molecular dynamics simulations, and have shown that {\it
960     slip} boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
961 gezelter 3299 appropriate for molecule-sized spheres embedded in a sea of spherical
962 gezelter 3310 solvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
963 gezelter 3299
964     Our simulation results show similar behavior to the behavior observed
965 gezelter 3302 by Schmidt and Skinner. The diffusion constant obtained from our
966 gezelter 3299 microcanonical molecular dynamics simulations lies between the slip
967     and stick boundary condition results obtained via Stokes-Einstein
968     behavior. Since the Langevin integrator assumes Stokes-Einstein stick
969     boundary conditions in calculating the drag and random forces for
970     spherical particles, our Langevin routine obtains nearly quantitative
971     agreement with the hydrodynamic results for spherical particles. One
972     avenue for improvement of the method would be to compute elements of
973     $\Xi_{tt}$ assuming behavior intermediate between the two boundary
974 gezelter 3302 conditions.
975 gezelter 3299
976 gezelter 3310 In the explicit solvent simulations, both our solute and solvent
977     particles were structureless, exerting no torques upon each other.
978     Therefore, there are not rotational correlation times available for
979     this model system.
980 gezelter 3299
981 gezelter 3310 \subsection{Ellipsoids}
982     For uniaxial ellipsoids ($a > b = c$), Perrin's formulae for both
983 gezelter 3299 translational and rotational diffusion of each of the body-fixed axes
984     can be combined to give a single translational diffusion
985 gezelter 3302 constant,\cite{Berne90}
986 gezelter 3299 \begin{equation}
987     D = \frac{k_B T}{6 \pi \eta a} G(\rho),
988     \label{Dperrin}
989     \end{equation}
990     as well as a single rotational diffusion coefficient,
991     \begin{equation}
992     \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - \rho^2)
993     G(\rho) - 1}{1 - \rho^4} \right\}.
994     \label{ThetaPerrin}
995     \end{equation}
996     In these expressions, $G(\rho)$ is a function of the axial ratio
997     ($\rho = b / a$), which for prolate ellipsoids, is
998     \begin{equation}
999     G(\rho) = (1- \rho^2)^{-1/2} \ln \left\{ \frac{1 + (1 -
1000     \rho^2)^{1/2}}{\rho} \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 gezelter 3302 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 gezelter 3299
1010 gezelter 3305 Even though there are analytic resistance tensors for ellipsoids, we
1011     constructed a rough-shell model using 2135 beads (each with a diameter
1012 gezelter 3310 of 0.25 \AA) to approximate the shape of the model ellipsoid. We
1013 gezelter 3305 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 gezelter 3299 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 gezelter 3308 The translational diffusion constants from the microcanonical simulations
1026     agree well with the predictions of the Perrin model, although the rotational
1027     correlation times are a factor of 2 shorter than expected from hydrodynamic
1028     theory. One explanation for the slower rotation
1029     of explicitly-solvated ellipsoids is the possibility that solute-solvent
1030     collisions happen at both ends of the solute whenever the principal
1031     axis of the ellipsoid is turning. In the upper portion of figure
1032     \ref{fig:explanation} we sketch a physical picture of this explanation.
1033     Since our Langevin integrator is providing nearly quantitative agreement with
1034     the Perrin model, it also predicts orientational diffusion for ellipsoids that
1035     exceed explicitly solvated correlation times by a factor of two.
1036 gezelter 3299
1037 gezelter 3310 \subsection{Rigid dumbbells}
1038 gezelter 3302 Perhaps the only {\it composite} rigid body for which analytic
1039     expressions for the hydrodynamic tensor are available is the
1040     two-sphere dumbbell model. This model consists of two non-overlapping
1041     spheres held by a rigid bond connecting their centers. There are
1042     competing expressions for the 6x6 resistance tensor for this
1043 gezelter 3341 model. The second order expression introduced by Rotne and
1044     Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la Torre and
1045     Bloomfield,\cite{Torre1977} is given above as
1046 gezelter 3299 Eq. (\ref{introEquation:RPTensorNonOverlapped}). In our case, we use
1047     a model dumbbell in which the two spheres are identical Lennard-Jones
1048     particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
1049 gezelter 3302 a distance of 6.532 \AA.
1050 gezelter 3299
1051     The theoretical values for the translational diffusion constant of the
1052     dumbbell are calculated from the work of Stimson and Jeffery, who
1053     studied the motion of this system in a flow parallel to the
1054 gezelter 3302 inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
1055     motion in a flow {\it perpendicular} to the inter-sphere
1056     axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
1057     orientational} correlation times for this model system (other than
1058 gezelter 3341 those derived from the 6 x 6 tensor mentioned above).
1059 gezelter 3299
1060 gezelter 3305 The bead model for this model system comprises the two large spheres
1061     by themselves, while the rough shell approximation used 3368 separate
1062     beads (each with a diameter of 0.25 \AA) to approximate the shape of
1063     the rigid body. The hydrodynamics tensors computed from both the bead
1064     and rough shell models are remarkably similar. Computing the initial
1065     hydrodynamic tensor for a rough shell model can be quite expensive (in
1066     this case it requires inverting a 10104 x 10104 matrix), while the
1067     bead model is typically easy to compute (in this case requiring
1068 gezelter 3308 inversion of a 6 x 6 matrix).
1069 gezelter 3305
1070 gezelter 3308 \begin{figure}
1071     \centering
1072 gezelter 3310 \includegraphics[width=2in]{RoughShell}
1073 gezelter 3308 \caption[Model rigid bodies and their rough shell approximations]{The
1074     model rigid bodies (left column) used to test this algorithm and their
1075     rough-shell approximations (right-column) that were used to compute
1076     the hydrodynamic tensors. The top two models (ellipsoid and dumbbell)
1077     have analytic solutions and were used to test the rough shell
1078     approximation. The lower two models (banana and lipid) were compared
1079     with explicitly-solvated molecular dynamics simulations. }
1080     \label{fig:roughShell}
1081     \end{figure}
1082    
1083    
1084 gezelter 3305 Once the hydrodynamic tensor has been computed, there is no additional
1085     penalty for carrying out a Langevin simulation with either of the two
1086     different hydrodynamics models. Our naive expectation is that since
1087     the rigid body's surface is roughened under the various shell models,
1088     the diffusion constants will be even farther from the ``slip''
1089     boundary conditions than observed for the bead model (which uses a
1090     Stokes-Einstein model to arrive at the hydrodynamic tensor). For the
1091     dumbbell, this prediction is correct although all of the Langevin
1092     diffusion constants are within 6\% of the diffusion constant predicted
1093     from the fully solvated system.
1094    
1095 gezelter 3308 For rotational motion, Langevin integration (and the hydrodynamic tensor)
1096     yields rotational correlation times that are substantially shorter than those
1097     obtained from explicitly-solvated simulations. It is likely that this is due
1098     to the large size of the explicit solvent spheres, a feature that prevents
1099     the solvent from coming in contact with a substantial fraction of the surface
1100     area of the dumbbell. Therefore, the explicit solvent only provides drag
1101     over a substantially reduced surface area of this model, while the
1102     hydrodynamic theories utilize the entire surface area for estimating
1103     rotational diffusion. A sketch of the free volume available in the explicit
1104     solvent simulations is shown in figure \ref{fig:explanation}.
1105 gezelter 3305
1106 gezelter 3310
1107     \begin{figure}
1108     \centering
1109     \includegraphics[width=6in]{explanation}
1110     \caption[Explanations of the differences between orientational
1111     correlation times for explicitly-solvated models and hydrodynamics
1112     predictions]{Explanations of the differences between orientational
1113     correlation times for explicitly-solvated models and hydrodynamic
1114     predictions. For the ellipsoids (upper figures), rotation of the
1115     principal axis can involve correlated collisions at both sides of the
1116     solute. In the rigid dumbbell model (lower figures), the large size
1117     of the explicit solvent spheres prevents them from coming in contact
1118     with a substantial fraction of the surface area of the dumbbell.
1119     Therefore, the explicit solvent only provides drag over a
1120     substantially reduced surface area of this model, where the
1121     hydrodynamic theories utilize the entire surface area for estimating
1122     rotational diffusion.
1123     } \label{fig:explanation}
1124     \end{figure}
1125    
1126     \subsection{Composite banana-shaped molecules}
1127     Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids have
1128     been used by Orlandi {\it et al.} to observe mesophases in
1129     coarse-grained models for bent-core liquid crystalline
1130     molecules.\cite{Orlandi:2006fk} We have used the same overlapping
1131 gezelter 3299 ellipsoids as a way to test the behavior of our algorithm for a
1132     structure of some interest to the materials science community,
1133     although since we are interested in capturing only the hydrodynamic
1134 gezelter 3310 behavior of this model, we have left out the dipolar interactions of
1135     the original Orlandi model.
1136 gezelter 3308
1137 gezelter 3341 A reference system composed of a single banana rigid body embedded in
1138     a sea of 1929 solvent particles was created and run under standard
1139     (microcanonical) molecular dynamics. The resulting viscosity of this
1140     mixture was 0.298 centipoise (as estimated using
1141     Eq. (\ref{eq:shear})). To calculate the hydrodynamic properties of
1142     the banana rigid body model, we created a rough shell (see
1143     Fig.~\ref{fig:roughShell}), in which the banana is represented as a
1144     ``shell'' made of 3321 identical beads (0.25 \AA\ in diameter)
1145     distributed on the surface. Applying the procedure described in
1146     Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1147     identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0
1148     \AA).
1149 gezelter 3299
1150 gezelter 3341 The Langevin rigid-body integrator (and the hydrodynamic diffusion
1151     tensor) are essentially quantitative for translational diffusion of
1152     this model. Orientational correlation times under the Langevin
1153     rigid-body integrator are within 11\% of the values obtained from
1154     explicit solvent, but these models also exhibit some solvent
1155     inaccessible surface area in the explicitly-solvated case.
1156 gezelter 3308
1157 gezelter 3310 \subsection{Composite sphero-ellipsoids}
1158 gezelter 3341
1159 gezelter 3299 Spherical heads perched on the ends of Gay-Berne ellipsoids have been
1160 xsun 3312 used recently as models for lipid
1161 gezelter 3341 molecules.\cite{SunGezelter08,Ayton01} A reference system composed of
1162     a single lipid rigid body embedded in a sea of 1929 solvent particles
1163     was created and run under a microcanonical ensemble. The resulting
1164     viscosity of this mixture was 0.349 centipoise (as estimated using
1165 xsun 3312 Eq. (\ref{eq:shear})). To calculate the hydrodynamic properties of
1166     the lipid rigid body model, we created a rough shell (see
1167     Fig.~\ref{fig:roughShell}), in which the lipid is represented as a
1168     ``shell'' made of 3550 identical beads (0.25 \AA\ in diameter)
1169     distributed on the surface. Applying the procedure described in
1170     Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1171     identified the center of resistance, ${\bf r} = $(0 \AA, 0 \AA, 1.46
1172     \AA).
1173 gezelter 3310
1174 gezelter 3341 The translational diffusion constants and rotational correlation times
1175     obtained using the Langevin rigid-body integrator (and the
1176     hydrodynamic tensor) are essentially quantitative when compared with
1177     the explicit solvent simulations for this model system.
1178 gezelter 3315
1179 gezelter 3310 \subsection{Summary}
1180 gezelter 3341 According to our simulations, the Langevin rigid-body integrator we
1181     have developed is a reliable way to replace explicit solvent
1182     simulations in cases where the detailed solute-solvent interactions do
1183     not greatly impact the forces on the solute. In cases where the
1184     dielectric screening of the solvent, or specific solute-solvent
1185     interactions become important for structural or dynamic features of
1186     the solute molecule, this integrator may be less useful. However, for
1187     the kinds of coarse-grained modeling that have become popular in
1188     recent years, this integrator may prove itself to be quite valuable.
1189 xsun 3298
1190 gezelter 3315 \begin{figure}
1191     \centering
1192     \includegraphics[width=\linewidth]{graph}
1193     \caption[Mean squared displacements and orientational
1194     correlation functions for each of the model rigid bodies.]{The
1195     mean-squared displacements ($\langle r^2(t) \rangle$) and
1196     orientational correlation functions ($C_2(t)$) for each of the model
1197     rigid bodies studied. The circles are the results for microcanonical
1198     simulations with explicit solvent molecules, while the other data sets
1199     are results for Langevin dynamics using the different hydrodynamic
1200     tensor approximations. The Perrin model for the ellipsoids is
1201     considered the ``exact'' hydrodynamic behavior (this can also be said
1202     for the translational motion of the dumbbell operating under the bead
1203     model). In most cases, the various hydrodynamics models reproduce
1204     each other quantitatively.}
1205     \label{fig:results}
1206     \end{figure}
1207    
1208 xsun 3298 \begin{table*}
1209     \begin{minipage}{\linewidth}
1210     \begin{center}
1211 gezelter 3305 \caption{Translational diffusion constants (D) for the model systems
1212     calculated using microcanonical simulations (with explicit solvent),
1213     theoretical predictions, and Langevin simulations (with implicit solvent).
1214     Analytical solutions for the exactly-solved hydrodynamics models are
1215     from Refs. \citen{Einstein05} (sphere), \citen{Perrin1934} and \citen{Perrin1936}
1216     (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1217     (dumbbell). The other model systems have no known analytic solution.
1218     All diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1219     $10^{-4}$ \AA$^2$ / fs). }
1220     \begin{tabular}{lccccccc}
1221 xsun 3298 \hline
1222 gezelter 3305 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1223     \cline{2-3} \cline{5-7}
1224     model & $\eta$ (centipoise) & D & & Analytical & method & Hydrodynamics & simulation \\
1225 xsun 3298 \hline
1226 xsun 3312 sphere & 0.279 & 3.06 & & 2.42 & exact & 2.42 & 2.33 \\
1227 gezelter 3305 ellipsoid & 0.255 & 2.44 & & 2.34 & exact & 2.34 & 2.37 \\
1228     & 0.255 & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1229 xsun 3312 dumbbell & 0.308 & 2.06 & & 1.64 & bead model & 1.65 & 1.62 \\
1230     & 0.308 & 2.06 & & 1.64 & rough shell & 1.59 & 1.62 \\
1231 gezelter 3305 banana & 0.298 & 1.53 & & & rough shell & 1.56 & 1.55 \\
1232 gezelter 3341 lipid & 0.349 & 1.41 & & & rough shell & 1.33 & 1.32 \\
1233 xsun 3298 \end{tabular}
1234     \label{tab:translation}
1235     \end{center}
1236     \end{minipage}
1237     \end{table*}
1238    
1239     \begin{table*}
1240     \begin{minipage}{\linewidth}
1241     \begin{center}
1242 gezelter 3305 \caption{Orientational relaxation times ($\tau$) for the model systems using
1243     microcanonical simulation (with explicit solvent), theoretical
1244     predictions, and Langevin simulations (with implicit solvent). All
1245     relaxation times are for the rotational correlation function with
1246     $\ell = 2$ and are reported in units of ps. The ellipsoidal model has
1247     an exact solution for the orientational correlation time due to
1248     Perrin, but the other model systems have no known analytic solution.}
1249     \begin{tabular}{lccccccc}
1250 xsun 3298 \hline
1251 gezelter 3305 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1252     \cline{2-3} \cline{5-7}
1253     model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic & simulation \\
1254 xsun 3298 \hline
1255 xsun 3312 sphere & 0.279 & & & 9.69 & exact & 9.69 & 9.64 \\
1256 gezelter 3305 ellipsoid & 0.255 & 46.7 & & 22.0 & exact & 22.0 & 22.2 \\
1257     & 0.255 & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1258 xsun 3312 dumbbell & 0.308 & 14.1 & & & bead model & 50.0 & 50.1 \\
1259     & 0.308 & 14.1 & & & rough shell & 41.5 & 41.3 \\
1260 gezelter 3305 banana & 0.298 & 63.8 & & & rough shell & 70.9 & 70.9 \\
1261     lipid & 0.349 & 78.0 & & & rough shell & 76.9 & 77.9 \\
1262     \hline
1263 xsun 3298 \end{tabular}
1264     \label{tab:rotation}
1265     \end{center}
1266     \end{minipage}
1267     \end{table*}
1268    
1269 gezelter 3310 \section{Application: A rigid-body lipid bilayer}
1270    
1271     The Langevin dynamics integrator was applied to study the formation of
1272     corrugated structures emerging from simulations of the coarse grained
1273     lipid molecular models presented above. The initial configuration is
1274 xsun 3298 taken from our molecular dynamics studies on lipid bilayers with
1275 gezelter 3310 lennard-Jones sphere solvents. The solvent molecules were excluded
1276     from the system, and the experimental value for the viscosity of water
1277     at 20C ($\eta = 1.00$ cp) was used to mimic the hydrodynamic effects
1278     of the solvent. The absence of explicit solvent molecules and the
1279     stability of the integrator allowed us to take timesteps of 50 fs. A
1280     total simulation run time of 100 ns was sampled.
1281     Fig. \ref{fig:bilayer} shows the configuration of the system after 100
1282     ns, and the ripple structure remains stable during the entire
1283     trajectory. Compared with using explicit bead-model solvent
1284     molecules, the efficiency of the simulation has increased by an order
1285 xsun 3298 of magnitude.
1286    
1287 gezelter 3310 \begin{figure}
1288     \centering
1289     \includegraphics[width=\linewidth]{bilayer}
1290     \caption[Snapshot of a bilayer of rigid-body models for lipids]{A
1291     snapshot of a bilayer composed of rigid-body models for lipid
1292     molecules evolving using the Langevin integrator described in this
1293     work.} \label{fig:bilayer}
1294     \end{figure}
1295    
1296 tim 2746 \section{Conclusions}
1297    
1298 tim 2999 We have presented a new Langevin algorithm by incorporating the
1299     hydrodynamics properties of arbitrary shaped molecules into an
1300 gezelter 3308 advanced symplectic integration scheme. Further studies in systems
1301     involving banana shaped molecules illustrated that the dynamic
1302     properties could be preserved by using this new algorithm as an
1303     implicit solvent model.
1304 tim 2999
1305    
1306 tim 2746 \section{Acknowledgments}
1307     Support for this project was provided by the National Science
1308     Foundation under grant CHE-0134881. T.L. also acknowledges the
1309     financial support from center of applied mathematics at University
1310     of Notre Dame.
1311     \newpage
1312    
1313 gezelter 3305 \bibliographystyle{jcp}
1314 tim 2746 \bibliography{langevin}
1315    
1316     \end{document}