ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
(Generate patch)

Comparing trunk/langevin/langevin.tex (file contents):
Revision 3298 by xsun, Wed Jan 2 21:06:31 2008 UTC vs.
Revision 3316 by gezelter, Fri Jan 18 20:43:53 2008 UTC

# Line 17 | Line 17
17   \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18   9.0in \textwidth 6.5in \brokenpenalty=10000
19   \renewcommand{\baselinestretch}{1.2}
20 < \renewcommand\citemid{\ } % no comma in optional reference note
20 > \renewcommand\citemid{\ } % no comma in optional referenc note
21  
22   \begin{document}
23  
24   \title{An algorithm for performing Langevin dynamics on rigid bodies of arbitrary shape }
25  
26 < \author{Teng Lin, Xiuquan Sun and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
26 > \author{Xiuquan Sun, Teng Lin and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
27   gezelter@nd.edu} \\
28   Department of Chemistry and Biochemistry\\
29   University of Notre Dame\\
# Line 48 | Line 48 | As alternative to Newtonian dynamics, Langevin dynamic
48   \section{Introduction}
49  
50   %applications of langevin dynamics
51 < As alternative to Newtonian dynamics, Langevin dynamics, which
52 < mimics a simple heat bath with stochastic and dissipative forces,
53 < has been applied in a variety of studies. The stochastic treatment
54 < of the solvent enables us to carry out substantially longer time
55 < simulations. Implicit solvent Langevin dynamics simulations of
56 < met-enkephalin not only outperform explicit solvent simulations for
57 < computational efficiency, but also agrees very well with explicit
58 < solvent simulations for dynamical properties.\cite{Shen2002}
59 < Recently, applying Langevin dynamics with the UNRES model, Liow and
60 < his coworkers suggest that protein folding pathways can be possibly
61 < explored within a reasonable amount of time.\cite{Liwo2005} The
62 < stochastic nature of the Langevin dynamics also enhances the
63 < sampling of the system and increases the probability of crossing
64 < energy barriers.\cite{Banerjee2004, Cui2003} Combining Langevin
65 < dynamics with Kramers's theory, Klimov and Thirumalai identified
66 < free-energy barriers by studying the viscosity dependence of the
67 < protein folding rates.\cite{Klimov1997} In order to account for
68 < solvent induced interactions missing from implicit solvent model,
69 < Kaya incorporated desolvation free energy barrier into implicit
70 < coarse-grained solvent model in protein folding/unfolding studies
71 < and discovered a higher free energy barrier between the native and
72 < denatured states. Because of its stability against noise, Langevin
73 < dynamics is very suitable for studying remagnetization processes in
74 < various systems.\cite{Palacios1998,Berkov2002,Denisov2003} For
51 > 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 > Kramers's theory, Klimov and Thirumalai identified free-energy
72 > 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 > between the native and denatured states.\cite{XXX}
78 >
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   instance, the oscillation power spectrum of nanoparticles from
83 < Langevin dynamics simulation has the same peak frequencies for
84 < different wave vectors, which recovers the property of magnetic
85 < excitations in small finite structures.\cite{Berkov2005a}
83 > 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  
87 < %review rigid body dynamics
88 < Rigid bodies are frequently involved in the modeling of different
89 < areas, from engineering, physics, to chemistry. For example,
90 < missiles and vehicle are usually modeled by rigid bodies.  The
91 < movement of the objects in 3D gaming engine or other physics
92 < simulator is governed by the rigid body dynamics. In molecular
93 < simulation, rigid body is used to simplify the model in
94 < protein-protein docking study{\cite{Gray2003}}.
87 > In typical LD simulations, the friction and random forces on
88 > individual atoms are taken from the Stokes-Einstein hydrodynamic
89 > approximation,
90 > \begin{eqnarray}
91 > m \dot{v}(t) & = & -\nabla U(x) - \xi m v(t) + R(t) \\
92 > \langle R(t) \rangle & = & 0 \\
93 > \langle R(t) R(t') \rangle & = & 2 k_B T \xi m \delta(t - t')
94 > \end{eqnarray}
95 > where $\xi \approx 6 \pi \eta a$.  Here $\eta$ is the viscosity of the
96 > implicit solvent, and $a$ is the hydrodynamic radius of the atom.
97  
98 < It is very important to develop stable and efficient methods to
99 < integrate the equations of motion for orientational degrees of
100 < freedom. Euler angles are the natural choice to describe the
101 < rotational degrees of freedom. However, due to $\frac {1}{sin
102 < \theta}$ singularities, the numerical integration of corresponding
103 < equations of these motion is very inefficient and inaccurate.
104 < Although an alternative integrator using multiple sets of Euler
105 < angles can overcome this difficulty\cite{Barojas1973}, the
106 < computational penalty and the loss of angular momentum conservation
107 < still remain. A singularity-free representation utilizing
108 < quaternions was developed by Evans in 1977.\cite{Evans1977}
109 < Unfortunately, this approach used a nonseparable Hamiltonian
110 < resulting from the quaternion representation, which prevented the
111 < symplectic algorithm from being utilized. Another different approach
112 < is to apply holonomic constraints to the atoms belonging to the
104 < rigid body. Each atom moves independently under the normal forces
105 < deriving from potential energy and constraint forces which are used
106 < to guarantee the rigidness. However, due to their iterative nature,
107 < the SHAKE and Rattle algorithms also converge very slowly when the
108 < number of constraints increases.\cite{Ryckaert1977, Andersen1983}
98 > The use of rigid substructures,\cite{???}
99 > coarse-graining,\cite{Ayton,Sun,Zannoni} and ellipsoidal
100 > representations of protein side chains~\cite{Schulten} has made the
101 > use of the Stokes-Einstein approximation problematic.  A rigid
102 > substructure moves as a single unit with orientational as well as
103 > translational degrees of freedom.  This requires a more general
104 > treatment of the hydrodynamics than the spherical approximation
105 > provides.  The atoms involved in a rigid or coarse-grained structure
106 > should properly have solvent-mediated interactions with each
107 > other. The theory of interactions {\it between} bodies moving through
108 > a fluid has been developed over the past century and has been applied
109 > to simulations of Brownian
110 > motion.\cite{MarshallNewton,GarciaDeLaTorre} There a need to have a
111 > more thorough treatment of hydrodynamics included in the library of
112 > methods available for performing Langevin simulations.
113  
114 < A break-through in geometric literature suggests that, in order to
114 > \subsection{Rigid Body Dynamics}
115 > Rigid bodies are frequently involved in the modeling of large
116 > collections of particles that move as a single unit.  In molecular
117 > simulations, rigid bodies have been used to simplify protein-protein
118 > docking,\cite{Gray2003} and lipid bilayer simulations.\cite{Sun2008}
119 > Many of the water models in common use are also rigid-body
120 > models,\cite{TIPs,SPC/E} although they are typically evolved using
121 > constraints rather than rigid body equations of motion.
122 >
123 > Euler angles are a natural choice to describe the rotational
124 > degrees of freedom.  However, due to $1 \over \sin \theta$
125 > singularities, the numerical integration of corresponding equations of
126 > these motion can become inaccurate (and inefficient).  Although an
127 > alternative integrator using multiple sets of Euler angles can
128 > overcome this problem,\cite{Barojas1973} the computational penalty and
129 > the loss of angular momentum conservation remain. A singularity-free
130 > representation utilizing quaternions was developed by Evans in
131 > 1977.\cite{Evans1977} Unfortunately, this approach uses a nonseparable
132 > Hamiltonian resulting from the quaternion representation, which
133 > prevented symplectic algorithms from being utilized until very
134 > recently.\cite{Miller2002} Another approach is the application of
135 > holonomic constraints to the atoms belonging to the rigid body. Each
136 > atom moves independently under the normal forces deriving from
137 > potential energy and constraint forces which are used to guarantee the
138 > rigidness. However, due to their iterative nature, the SHAKE and
139 > Rattle algorithms also converge very slowly when the number of
140 > constraints increases.\cite{Ryckaert1977,Andersen1983}
141 >
142 > A breakthrough in geometric literature suggests that, in order to
143   develop a long-term integration scheme, one should preserve the
144   symplectic structure of the propagator. By introducing a conjugate
145   momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
146 < equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
147 < proposed to evolve the Hamiltonian system in a constraint manifold
148 < by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
149 < An alternative method using the quaternion representation was
150 < developed by Omelyan.\cite{Omelyan1998} However, both of these
151 < methods are iterative and inefficient. In this section, we descibe a
152 < symplectic Lie-Poisson integrator for rigid bodies developed by
153 < Dullweber and his coworkers\cite{Dullweber1997} in depth.
146 > equation, a symplectic integrator, RSHAKE,\cite{Kol1997} was proposed
147 > to evolve the Hamiltonian system in a constraint manifold by
148 > iteratively satisfying the orthogonality constraint $Q^T Q = 1$.  An
149 > alternative method using the quaternion representation was developed
150 > by Omelyan.\cite{Omelyan1998} However, both of these methods are
151 > iterative and suffer from some related inefficiencies. A symplectic
152 > Lie-Poisson integrator for rigid bodies developed by Dullweber {\it et
153 > al.}\cite{Dullweber1997} gets around most of the limitations mentioned
154 > above and has become the basis for our Langevin integrator.
155  
156 < %review langevin/browninan dynamics for arbitrarily shaped rigid body
157 < Combining Langevin or Brownian dynamics with rigid body dynamics,
158 < one can study slow processes in biomolecular systems. Modeling DNA
159 < as a chain of rigid beads, which are subject to harmonic potentials
160 < as well as excluded volume potentials, Mielke and his coworkers
161 < discovered rapid superhelical stress generations from the stochastic
162 < simulation of twin supercoiling DNA with response to induced
163 < torques.\cite{Mielke2004} Membrane fusion is another key biological
164 < process which controls a variety of physiological functions, such as
165 < release of neurotransmitters \textit{etc}. A typical fusion event
166 < happens on the time scale of a millisecond, which is impractical to
167 < study using atomistic models with newtonian mechanics. With the help
168 < of coarse-grained rigid body model and stochastic dynamics, the
169 < fusion pathways were explored by many
170 < researchers.\cite{Noguchi2001,Noguchi2002,Shillcock2005} Due to the
171 < difficulty of numerical integration of anisotropic rotation, most of
172 < the rigid body models are simply modeled using spheres, cylinders,
173 < ellipsoids or other regular shapes in stochastic simulations. In an
174 < effort to account for the diffusion anisotropy of arbitrary
175 < particles, Fernandes and de la Torre improved the original Brownian
176 < dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
177 < incorporating a generalized $6\times6$ diffusion tensor and
178 < introducing a simple rotation evolution scheme consisting of three
179 < consecutive rotations.\cite{Fernandes2002} Unfortunately, unexpected
180 < errors and biases are introduced into the system due to the
156 >
157 > \subsection{The Hydrodynamic tensor and Brownian dynamics}
158 > Combining Brownian dynamics with rigid substructures, one can study
159 > slow processes in biomolecular systems.  Modeling DNA as a chain of
160 > beads which are subject to harmonic potentials as well as excluded
161 > volume potentials, Mielke and his coworkers discovered rapid
162 > superhelical stress generations from the stochastic simulation of twin
163 > supercoiling DNA with response to induced torques.\cite{Mielke2004}
164 > Membrane fusion is another key biological process which controls a
165 > variety of physiological functions, such as release of
166 > neurotransmitters \textit{etc}. A typical fusion event happens on the
167 > time scale of a millisecond, which is impractical to study using
168 > atomistic models with newtonian mechanics. With the help of
169 > coarse-grained rigid body model and stochastic dynamics, the fusion
170 > pathways were explored by Noguchi and others.\cite{Noguchi2001,Noguchi2002,Shillcock2005}
171 >
172 > Due to the difficulty of numerically integrating anisotropic
173 > rotational motion, most of the coarse-grained rigid body models are
174 > treated as spheres, cylinders, ellipsoids or other regular shapes in
175 > stochastic simulations.  In an effort to account for the diffusion
176 > anisotropy of arbitrarily-shaped particles, Fernandes and Garc\'{i}a
177 > de la Torre improved the original Brownian dynamics simulation
178 > algorithm~\cite{Ermak1978,Allison1991} by incorporating a generalized
179 > $6\times6$ diffusion tensor and introducing a rotational evolution
180 > scheme consisting of three consecutive rotations.\cite{Fernandes2002}
181 > Unfortunately, biases are introduced into the system due to the
182   arbitrary order of applying the noncommuting rotation
183   operators.\cite{Beard2003} Based on the observation the momentum
184   relaxation time is much less than the time step, one may ignore the
185 < inertia in Brownian dynamics. However, the assumption of zero
186 < average acceleration is not always true for cooperative motion which
187 < is common in protein motion. An inertial Brownian dynamics (IBD) was
188 < proposed to address this issue by adding an inertial correction
185 > inertia in Brownian dynamics.  However, the assumption of zero average
186 > acceleration is not always true for cooperative motion which is common
187 > in proteins. An inertial Brownian dynamics (IBD) was proposed to
188 > address this issue by adding an inertial correction
189   term.\cite{Beard2000} As a complement to IBD which has a lower bound
190   in time step because of the inertial relaxation time, long-time-step
191   inertial dynamics (LTID) can be used to investigate the inertial
192   behavior of the polymer segments in low friction
193   regime.\cite{Beard2000} LTID can also deal with the rotational
194   dynamics for nonskew bodies without translation-rotation coupling by
195 < separating the translation and rotation motion and taking advantage
196 < of the analytical solution of hydrodynamics properties. However,
197 < typical nonskew bodies like cylinders and ellipsoids are inadequate
198 < to represent most complex macromolecule assemblies. These intricate
165 < molecules have been represented by a set of beads and their
166 < hydrodynamic properties can be calculated using variants on the
167 < standard hydrodynamic interaction tensors.
195 > separating the translation and rotation motion and taking advantage of
196 > the analytical solution of hydrodynamics properties. However, typical
197 > nonskew bodies like cylinders and ellipsoids are inadequate to
198 > represent most complex macromolecular assemblies.
199  
200   The goal of the present work is to develop a Langevin dynamics
201   algorithm for arbitrary-shaped rigid particles by integrating the
202 < accurate estimation of friction tensor from hydrodynamics theory
203 < into the sophisticated rigid body dynamics algorithms.
202 > accurate estimation of friction tensor from hydrodynamics theory into
203 > a symplectic rigid body dynamics propagator.  In the sections below,
204 > we review some of the theory of hydrodynamic tensors developed for
205 > Brownian simulations of rigid multi-particle systems, we then present
206 > our integration method for a set of generalized Langevin equations of
207 > motion, and we compare the behavior of the new Langevin integrator to
208 > dynamical quantities obtained via explicit solvent molecular dynamics.
209  
210 < \section{Computational Methods{\label{methodSec}}}
211 <
176 < \subsection{\label{introSection:frictionTensor}Friction Tensor}
177 < Theoretically, the friction kernel can be determined using the
210 > \subsection{\label{introSection:frictionTensor}The Friction Tensor}
211 > Theoretically, a complete friction kernel can be determined using the
212   velocity autocorrelation function. However, this approach becomes
213 < impractical when the system becomes more and more complicated.
214 < Instead, various approaches based on hydrodynamics have been
215 < developed to calculate the friction coefficients. In general, the
216 < friction tensor $\Xi$ is a $6\times 6$ matrix given by
217 < \[
213 > impractical when the solute becomes complex.  Instead, various
214 > approaches based on hydrodynamics have been developed to calculate the
215 > friction coefficients. In general, the friction tensor $\Xi$ is a
216 > $6\times 6$ matrix given by
217 > \begin{equation}
218   \Xi  = \left( {\begin{array}{*{20}c}
219     {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
220     {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
221   \end{array}} \right).
222 < \]
223 < Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are $3 \times 3$
224 < translational friction tensor and rotational resistance (friction)
225 < tensor respectively, while ${\Xi^{tr} }$ is translation-rotation
226 < coupling tensor and $ {\Xi^{rt} }$ is rotation-translation coupling
227 < tensor. When a particle moves in a fluid, it may experience friction
228 < force or torque along the opposite direction of the velocity or
229 < angular velocity,
230 < \[
222 > \end{equation}
223 > Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
224 > rotational resistance (friction) tensors respectively, while
225 > $\Xi^{tr}$ is translation-rotation coupling tensor and $\Xi^{rt}$ is
226 > rotation-translation coupling tensor. When a particle moves in a
227 > fluid, it may experience friction force ($\mathbf{F}_f$) and torque
228 > ($\mathbf{\tau}_f$) in opposition to the directions of the velocity
229 > ($\mathbf{v}$) and body-fixed angular velocity ($\mathbf{\omega}$),
230 > \begin{equation}
231   \left( \begin{array}{l}
232 < F_R  \\
233 < \tau _R  \\
232 > \mathbf{F}_f  \\
233 > \mathbf{\tau}_f  \\
234   \end{array} \right) =  - \left( {\begin{array}{*{20}c}
235 <   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
236 <   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
235 >   \Xi ^{tt} & \Xi ^{rt}  \\
236 >   \Xi ^{tr} & \Xi ^{rr}  \\
237   \end{array}} \right)\left( \begin{array}{l}
238 < v \\
239 < w \\
240 < \end{array} \right)
241 < \]
208 < where $F_r$ is the friction force and $\tau _R$ is the friction
209 < torque.
238 > \mathbf{v} \\
239 > \mathbf{\omega} \\
240 > \end{array} \right).
241 > \end{equation}
242  
243   \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
244 <
245 < For a spherical particle with slip boundary conditions, the
246 < translational and rotational friction constant can be calculated
247 < from Stoke's law,
248 < \[
217 < \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
244 > For a spherical particle under ``stick'' boundary conditions, the
245 > translational and rotational friction tensors can be calculated from
246 > Stoke's law,
247 > \begin{equation}
248 > \Xi^{tt}  = \left( {\begin{array}{*{20}c}
249     {6\pi \eta R} & 0 & 0  \\
250     0 & {6\pi \eta R} & 0  \\
251     0 & 0 & {6\pi \eta R}  \\
252   \end{array}} \right)
253 < \]
253 > \end{equation}
254   and
255 < \[
255 > \begin{equation}
256   \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
257     {8\pi \eta R^3 } & 0 & 0  \\
258     0 & {8\pi \eta R^3 } & 0  \\
259     0 & 0 & {8\pi \eta R^3 }  \\
260   \end{array}} \right)
261 < \]
261 > \end{equation}
262   where $\eta$ is the viscosity of the solvent and $R$ is the
263   hydrodynamic radius.
264  
265   Other non-spherical shapes, such as cylinders and ellipsoids, are
266 < widely used as references for developing new hydrodynamics theory,
266 > widely used as references for developing new hydrodynamics theories,
267   because their properties can be calculated exactly. In 1936, Perrin
268   extended Stokes's law to general ellipsoids, also called a triaxial
269   ellipsoid, which is given in Cartesian coordinates
270 < by\cite{Perrin1934, Perrin1936}
271 < \[
272 < \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
273 < }} = 1
274 < \]
275 < where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
276 < due to the complexity of the elliptic integral, only the ellipsoid
277 < with the restriction of two axes being equal, \textit{i.e.}
278 < prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
279 < exactly. Introducing an elliptic integral parameter $S$ for prolate
280 < ellipsoids :
281 < \[
251 < S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
252 < } }}{b},
253 < \]
270 > by\cite{Perrin1934,Perrin1936}
271 > \begin{equation}
272 > \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1
273 > \end{equation}
274 > where the semi-axes are of lengths $a$, $b$, and $c$. Due to the
275 > complexity of the elliptic integral, only uniaxial ellipsoids,
276 > {\it i.e.} prolate ($ a \ge b = c$) and oblate ($ a < b = c $), can
277 > be solved exactly. Introducing an elliptic integral parameter $S$ for
278 > prolate ellipsoids :
279 > \begin{equation}
280 > S = \frac{2}{\sqrt{a^2  - b^2}} \ln \frac{a + \sqrt{a^2  - b^2}}{b},
281 > \end{equation}
282   and oblate ellipsoids:
283 < \[
284 < S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
285 < }}{a},
258 < \]
283 > \begin{equation}
284 > S = \frac{2}{\sqrt {b^2  - a^2 }} \arctan \frac{\sqrt {b^2  - a^2}}{a},
285 > \end{equation}
286   one can write down the translational and rotational resistance
287 < tensors
287 > tensors for oblate,
288   \begin{eqnarray*}
289 < \Xi _a^{tt}  & = & 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}}. \\
290 < \Xi _b^{tt}  & = & \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S +
264 < 2a}},
289 > \Xi_a^{tt}  & = & 16\pi \eta \frac{a^2  - b^2}{(2a^2  - b^2 )S - 2a}. \\
290 > \Xi_b^{tt} =  \Xi_c^{tt} & = & 32\pi \eta \frac{a^2  - b^2 }{(2a^2 - 3b^2 )S + 2a},
291   \end{eqnarray*}
292 < and
292 > and prolate,
293   \begin{eqnarray*}
294 < \Xi _a^{rr} & = & \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}}, \\
295 < \Xi _b^{rr} & = & \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}}.
294 > \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2  - b^2 )b^2}{2a - b^2 S}, \\
295 > \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4  - b^4)}{(2a^2  - b^2 )S - 2a}
296   \end{eqnarray*}
297 + ellipsoids. For both spherical and ellipsoidal particles, the
298 + translation-rotation and rotation-translation coupling tensors are
299 + zero.
300  
301   \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
302  
# Line 275 | Line 304 | hydrodynamic properties of rigid bodies. However, sinc
304   analytical solution for the friction tensor for arbitrarily shaped
305   rigid molecules. The ellipsoid of revolution model and general
306   triaxial ellipsoid model have been used to approximate the
307 < hydrodynamic properties of rigid bodies. However, since the mapping
308 < from all possible ellipsoidal spaces, $r$-space, to all possible
309 < combination of rotational diffusion coefficients, $D$-space, is not
310 < unique\cite{Wegener1979} as well as the intrinsic coupling between
311 < translational and rotational motion of rigid bodies, general
312 < ellipsoids are not always suitable for modeling arbitrarily shaped
313 < rigid molecules. A number of studies have been devoted to
307 > hydrodynamic properties of rigid bodies. However, the mapping from all
308 > possible ellipsoidal spaces, $r$-space, to all possible combination of
309 > rotational diffusion coefficients, $D$-space, is not
310 > unique.\cite{Wegener1979} Additionally, because there is intrinsic
311 > coupling between translational and rotational motion of rigid bodies,
312 > general ellipsoids are not always suitable for modeling arbitrarily
313 > shaped rigid molecules.  A number of studies have been devoted to
314   determining the friction tensor for irregularly shaped rigid bodies
315 < using more advanced methods where the molecule of interest was
316 < modeled by a combinations of spheres\cite{Carrasco1999} and the
317 < hydrodynamics properties of the molecule can be calculated using the
318 < hydrodynamic interaction tensor. Let us consider a rigid assembly of
319 < $N$ beads immersed in a continuous medium. Due to hydrodynamic
320 < interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
321 < than its unperturbed velocity $v_i$,
315 > using more advanced methods where the molecule of interest was modeled
316 > by a combinations of spheres\cite{Carrasco1999} and the hydrodynamics
317 > properties of the molecule can be calculated using the hydrodynamic
318 > interaction tensor. Let us consider a rigid assembly of $N$ beads
319 > immersed in a continuous medium. Due to hydrodynamic interaction, the
320 > ``net'' velocity of $i$th bead, $v'_i$ is different than its
321 > unperturbed velocity $v_i$,
322   \[
323   v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
324   \]
# Line 367 | Line 396 | arbitrary origin $O$ can be written as
396   \begin{eqnarray}
397   \Xi _{}^{tt}  & = & \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
398   \Xi _{}^{tr}  & = & \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
399 < \Xi _{}^{rr}  & = &  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
399 > \Xi _{}^{rr}  & = &  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } }
400 > U_j  + 6 \eta V {\bf I}. \notag
401   \label{introEquation:ResistanceTensorArbitraryOrigin}
402   \end{eqnarray}
403 + The final term in the expression for $\Xi^{rr}$ is correction that
404 + accounts for errors in the rotational motion of certain kinds of bead
405 + models. The additive correction uses the solvent viscosity ($\eta$)
406 + as well as the total volume of the beads that contribute to the
407 + hydrodynamic model,
408 + \begin{equation}
409 + V = \frac{4 \pi}{3} \sum_{i=1}^{N} \sigma_i^3,
410 + \end{equation}
411 + where $\sigma_i$ is the radius of bead $i$.  This correction term was
412 + rigorously tested and compared with the analytical results for
413 + two-sphere and ellipsoidal systems by Garcia de la Torre and
414 + Rodes.\cite{Torre:1983lr}
415 +
416 +
417   The resistance tensor depends on the origin to which they refer. The
418   proper location for applying the friction force is the center of
419   resistance (or center of reaction), at which the trace of rotational
# Line 426 | Line 470 | joining center of resistance $R$ and origin $O$.
470   where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
471   joining center of resistance $R$ and origin $O$.
472  
429 \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
473  
474 + \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
475   Consider the Langevin equations of motion in generalized coordinates
476   \begin{equation}
477   M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i} (t)
# Line 574 | Line 618 | the velocities can be advanced to the same time value.
618      + \frac{h}{2} {\bf \tau}^b(t + h) .
619   \end{align*}
620  
621 < \section{Results and Discussion}
621 > \section{Validating the Method\label{sec:validating}}
622 > In order to validate our Langevin integrator for arbitrarily-shaped
623 > rigid bodies, we implemented the algorithm in {\sc
624 > oopse}\cite{Meineke2005} and  compared the results of this algorithm
625 > with the known
626 > hydrodynamic limiting behavior for a few model systems, and to
627 > microcanonical molecular dynamics simulations for some more
628 > complicated bodies. The model systems and their analytical behavior
629 > (if known) are summarized below. Parameters for the primary particles
630 > comprising our model systems are given in table \ref{tab:parameters},
631 > and a sketch of the arrangement of these primary particles into the
632 > model rigid bodies is shown in figure \ref{fig:models}. In table
633 > \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
634 > ellipsoidal (Gay-Berne) particles.  For spherical particles, the value
635 > of the Lennard-Jones $\sigma$ parameter is the particle diameter
636 > ($d$).  Gay-Berne ellipsoids have an energy scaling parameter,
637 > $\epsilon^s$, which describes the well depth for two identical
638 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a
639 > well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
640 > describes the ratio between the well depths in the {\it end-to-end}
641 > and side-by-side configurations.  For spheres, $\epsilon^r \equiv 1$.
642 > Moments of inertia are also required to describe the motion of primary
643 > particles with orientational degrees of freedom.
644  
645 < The Langevin algorithm described in previous section has been
646 < implemented in {\sc oopse}\cite{Meineke2005} and applied to studies
581 < of the static and dynamic properties in several systems.
582 <
583 < \subsection{Temperature Control}
584 <
585 < As shown in Eq.~\ref{randomForce}, random collisions associated with
586 < the solvent's thermal motions is controlled by the external
587 < temperature. The capability to maintain the temperature of the whole
588 < system was usually used to measure the stability and efficiency of
589 < the algorithm. In order to verify the stability of this new
590 < algorithm, a series of simulations are performed on system
591 < consisiting of 256 SSD water molecules with different viscosities.
592 < The initial configuration for the simulations is taken from a 1ns
593 < NVT simulation with a cubic box of 19.7166~\AA. All simulation are
594 < carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
595 < with reference temperature at 300~K. The average temperature as a
596 < function of $\eta$ is shown in Table \ref{langevin:viscosity} where
597 < the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
598 < 1$ poise. The better temperature control at higher viscosity can be
599 < explained by the finite size effect and relative slow relaxation
600 < rate at lower viscosity regime.
601 < \begin{table}
602 < \caption{AVERAGE TEMPERATURES FROM LANGEVIN DYNAMICS SIMULATIONS OF
603 < SSD WATER MOLECULES WITH REFERENCE TEMPERATURE AT 300~K.}
604 < \label{langevin:viscosity}
645 > \begin{table*}
646 > \begin{minipage}{\linewidth}
647   \begin{center}
648 < \begin{tabular}{lll}
649 <  \hline
650 <  $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
651 <  \hline
652 <  1    & 300.47 & 10.99 \\
653 <  0.1  & 301.19 & 11.136 \\
654 <  0.01 & 303.04 & 11.796 \\
655 <  \hline
648 > \caption{Parameters for the primary particles in use by the rigid body
649 > models in figure \ref{fig:models}.}
650 > \begin{tabular}{lrcccccccc}
651 > \hline
652 > & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
653 > & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
654 > $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
655 > Sphere   & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75  & 802.75  \\
656 > Ellipsoid & & 4.6 & 13.8  & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
657 > Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1   & 190 & 802.75 & 802.75 & 802.75 \\
658 > Banana  &(3 identical ellipsoids)& 4.2 & 11.2  & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
659 > Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
660 >       & Ellipsoidal Tail & 4.6 & 13.8  & 0.8   & 0.2 & 760 & 45000 & 45000 & 9000 \\
661 > Solvent &  & 4.7 & $= d$ & 0.8 & 1   & 72.06 & & & \\
662 > \hline
663   \end{tabular}
664 + \label{tab:parameters}
665   \end{center}
666 < \end{table}
666 > \end{minipage}
667 > \end{table*}
668  
618 Another set of calculations were performed to study the efficiency of
619 temperature control using different temperature coupling schemes.
620 The starting configuration is cooled to 173~K and evolved using NVE,
621 NVT, and Langevin dynamic with time step of 2 fs.
622 Fig.~\ref{langevin:temperature} shows the heating curve obtained as
623 the systems reach equilibrium. The orange curve in
624 Fig.~\ref{langevin:temperature} represents the simulation using
625 Nos\'e-Hoover temperature scaling scheme with thermostat of 5 ps
626 which gives reasonable tight coupling, while the blue one from
627 Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
628 scaling to the desire temperature. When $ \eta = 0$, Langevin dynamics becomes normal
629 NVE (see orange curve in Fig.~\ref{langevin:temperature}) which
630 loses the temperature control ability.
631
669   \begin{figure}
670   \centering
671 < \includegraphics[width=\linewidth]{temperature.pdf}
672 < \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
673 < temperature fluctuation versus time.} \label{langevin:temperature}
671 > \includegraphics[width=3in]{sketch}
672 > \caption[Sketch of the model systems]{A sketch of the model systems
673 > used in evaluating the behavior of the rigid body Langevin
674 > integrator.} \label{fig:models}
675   \end{figure}
676  
677 < \subsection{Langevin Dynamics simulation {\it vs} NVE simulations}
677 > \subsection{Simulation Methodology}
678 > We performed reference microcanonical simulations with explicit
679 > solvents for each of the different model system.  In each case there
680 > was one solute model and 1929 solvent molecules present in the
681 > simulation box.  All simulations were equilibrated using a
682 > constant-pressure and temperature integrator with target values of 300
683 > K for the temperature and 1 atm for pressure.  Following this stage,
684 > further equilibration and sampling was done in a microcanonical
685 > ensemble.  Since the model bodies are typically quite massive, we were
686 > able to use a time step of 25 fs.
687  
688 < To validate our langevin dynamics simulation. We performed several NVE
689 < simulations with explicit solvents for different shaped
690 < molecules. There are one solute molecule and 1929 solvent molecules in
691 < NVE simulation. The parameters are shown in table
692 < \ref{tab:parameters}. The force field between spheres is standard
693 < Lennard-Jones, and ellipsoids interact with other ellipsoids and
694 < spheres with generalized Gay-Berne potential. All simulations are
695 < carried out at 300 K and 1 Atm. The time step is 25 ns, and a
696 < switching function was applied to all potentials to smoothly turn off
697 < the interactions between a range of $22$ and $25$ \AA.  The switching
698 < function was the standard (cubic) function,
688 > The model systems studied used both Lennard-Jones spheres as well as
689 > uniaxial Gay-Berne ellipoids. In its original form, the Gay-Berne
690 > potential was a single site model for the interactions of rigid
691 > ellipsoidal molecules.\cite{Gay81} It can be thought of as a
692 > modification of the Gaussian overlap model originally described by
693 > Berne and Pechukas.\cite{Berne72} The potential is constructed in the
694 > familiar form of the Lennard-Jones function using
695 > orientation-dependent $\sigma$ and $\epsilon$ parameters,
696 > \begin{equation*}
697 > V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
698 > r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
699 > {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u
700 > }_i},
701 > {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
702 > -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
703 > {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
704 > \label{eq:gb}
705 > \end{equation*}
706 >
707 > The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
708 > \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
709 > \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
710 > are dependent on the relative orientations of the two ellipsoids (${\bf
711 > \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
712 > inter-ellipsoid separation (${\bf \hat{r}}_{ij}$).  The shape and
713 > attractiveness of each ellipsoid is governed by a relatively small set
714 > of parameters: $l$ and $d$ describe the length and width of each
715 > uniaxial ellipsoid, while $\epsilon^s$, which describes the well depth
716 > for two identical ellipsoids in a {\it side-by-side} configuration.
717 > Additionally, a well depth aspect ratio, $\epsilon^r = \epsilon^e /
718 > \epsilon^s$, describes the ratio between the well depths in the {\it
719 > end-to-end} and side-by-side configurations.  Details of the potential
720 > are given elsewhere,\cite{Luckhurst90,Golubkov06,SunGezelter08} and an
721 > excellent overview of the computational methods that can be used to
722 > efficiently compute forces and torques for this potential can be found
723 > in Ref. \citen{Golubkov06}
724 >
725 > For the interaction between nonequivalent uniaxial ellipsoids (or
726 > between spheres and ellipsoids), the spheres are treated as ellipsoids
727 > with an aspect ratio of 1 ($d = l$) and with an well depth ratio
728 > ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$).  The form of the
729 > Gay-Berne potential we are using was generalized by Cleaver {\it et
730 > al.} and is appropriate for dissimilar uniaxial
731 > ellipsoids.\cite{Cleaver96}
732 >
733 > A switching function was applied to all potentials to smoothly turn
734 > off the interactions between a range of $22$ and $25$ \AA.  The
735 > switching function was the standard (cubic) function,
736   \begin{equation}
737   s(r) =
738          \begin{cases}
# Line 660 | Line 744 | We have computed translational diffusion constants for
744          \end{cases}
745   \label{eq:switchingFunc}
746   \end{equation}
747 < We have computed translational diffusion constants for lipid molecules
748 < from the mean-square displacement,
747 >
748 > To measure shear viscosities from our microcanonical simulations, we
749 > used the Einstein form of the pressure correlation function,\cite{hess:209}
750   \begin{equation}
751 < D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
751 > \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
752 > \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \right\rangle_{t_0}.
753 > \label{eq:shear}
754   \end{equation}
755 < of the solute molecules. Translational diffusion constants for the
669 < different shaped molecules are shown in table
670 < \ref{tab:translation}.  We have also computed orientational correlation
671 < times for different shaped molecules from fits of the second-order
672 < Legendre polynomial correlation function,
755 > A similar form exists for the bulk viscosity
756   \begin{equation}
757 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
758 < \mu}_{i}(0) \right)
757 > \kappa = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
758 > \int_{t_0}^{t_0 + t}
759 > \left(P\left(t'\right)-\left\langle P \right\rangle \right)dt'
760 > \right)^2 \right\rangle_{t_0}.
761   \end{equation}
762 < the results are shown in table \ref{tab:rotation}. We used einstein
763 < format of the pressure correlation function,
762 > Alternatively, the shear viscosity can also be calculated using a
763 > Green-Kubo formula with the off-diagonal pressure tensor correlation function,
764   \begin{equation}
765 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
766 < \mu}_{i}(0) \right)
765 > \eta = \frac{V}{k_B T} \int_0^{\infty} \left\langle P_{xz}(t_0) P_{xz}(t_0
766 > + t) \right\rangle_{t_0} dt,
767   \end{equation}
768 < to estimate the viscosity of the systems from NVE simulations. The
769 < viscosity can also be calculated by Green-Kubo pressure correlaton
770 < function,
768 > although this method converges extremely slowly and is not practical
769 > for obtaining viscosities from molecular dynamics simulations.
770 >
771 > The Langevin dynamics for the different model systems were performed
772 > at the same temperature as the average temperature of the
773 > microcanonical simulations and with a solvent viscosity taken from
774 > Eq. (\ref{eq:shear}) applied to these simulations.  We used 1024
775 > independent solute simulations to obtain statistics on our Langevin
776 > integrator.
777 >
778 > \subsection{Analysis}
779 >
780 > The quantities of interest when comparing the Langevin integrator to
781 > analytic hydrodynamic equations and to molecular dynamics simulations
782 > are typically translational diffusion constants and orientational
783 > relaxation times.  Translational diffusion constants for point
784 > particles are computed easily from the long-time slope of the
785 > mean-square displacement,
786   \begin{equation}
787 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
688 < \mu}_{i}(0) \right)
787 > D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \left\langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \right\rangle,
788   \end{equation}
789 < However, this method converges slowly, and the statistics are not good
790 < enough to give us a very accurate value. The langevin dynamics
791 < simulations for different shaped molecules are performed at the same
792 < conditions as the NVE simulations with viscosity estimated from NVE
793 < simulations. To get better statistics, 1024 non-interacting solute
794 < molecules are put into one simulation box for each langevin
795 < simulation, this is equal to 1024 simulations for single solute
796 < systems. The diffusion constants and rotation relaxation times for
797 < different shaped molecules are shown in table \ref{tab:translation}
798 < and \ref{tab:rotation} to compare to the results calculated from NVE
700 < simulations. The theoretical values for sphere is calculated from the
701 < Stokes-Einstein law, the theoretical values for ellipsoid is
702 < calculated from Perrin's fomula, the theoretical values for dumbbell
703 < is calculated from StinXX and Davis theory. The exact method is
704 < applied to the langevin dynamics simulations for sphere and ellipsoid,
705 < the bead model is applied to the simulation for dumbbell molecule, and
706 < the rough shell model is applied to ellipsoid, dumbbell, banana and
707 < lipid molecules. The results from all the langevin dynamics
708 < simulations, including exact, bead model and rough shell, match the
709 < theoretical values perfectly for all different shaped molecules. This
710 < indicates that our simulation package for langevin dynamics is working
711 < well. The approxiate methods ( bead model and rough shell model) are
712 < accurate enough for the current simulations. The goal of the langevin
713 < dynamics theory is to replace the explicit solvents by the friction
714 < forces. We compared the dynamic properties of different shaped
715 < molecules in langevin dynamics simulations with that in NVE
716 < simulations. The results are reasonable close. Overall, the
717 < translational diffusion constants calculated from langevin dynamics
718 < simulations are very close to the values from the NVE simulation. For
719 < sphere and lipid molecules, the diffusion constants are a little bit
720 < off from the NVE simulation results. One possible reason is that the
721 < calculation of the viscosity is very difficult to be accurate. Another
722 < possible reason is that although we save very frequently during the
723 < NVE simulations and run pretty long time simulations, there is only
724 < one solute molecule in the system which makes the calculation for the
725 < diffusion constant difficult. The sphere molecule behaves as a free
726 < rotor in the solvent, so there is no rotation relaxation time
727 < calculated from NVE simulations. The rotation relaxation time is not
728 < very close to the NVE simulations results. The banana and lipid
729 < molecules match the NVE simulations results pretty well. The mismatch
730 < between langevin dynamics and NVE simulation for ellipsoid is possibly
731 < caused by the slip boundary condition. For dumbbell, the mismatch is
732 < caused by the size of the solvent molecule is pretty large compared to
733 < dumbbell molecule in NVE simulations.
789 > of the solute molecules.  For models in which the translational
790 > diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
791 > (i.e. any non-spherically-symmetric rigid body), it is possible to
792 > compute the diffusive behavior for motion parallel to each body-fixed
793 > axis by projecting the displacement of the particle onto the
794 > body-fixed reference frame at $t=0$.  With an isotropic solvent, as we
795 > have used in this study, there are differences between the three
796 > diffusion constants, but these must converge to the same value at
797 > longer times.  Translational diffusion constants for the different
798 > shaped models are shown in table \ref{tab:translation}.
799  
800 + In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
801 + diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
802 + {\it around} a particular body-fixed axis and {\it not} the diffusion
803 + of a vector pointing along the axis.  However, these eigenvalues can
804 + be combined to find 5 characteristic rotational relaxation
805 + times,\cite{PhysRev.119.53,Berne90}
806 + \begin{eqnarray}
807 + 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
808 + 1 / \tau_2 & = & 6 D_r - 2 \Delta \\
809 + 1 / \tau_3 & = & 3 (D_r + D_1)  \\
810 + 1 / \tau_4 & = & 3 (D_r + D_2) \\
811 + 1 / \tau_5 & = & 3 (D_r + D_3)
812 + \end{eqnarray}
813 + where
814 + \begin{equation}
815 + D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
816 + \end{equation}
817 + and
818 + \begin{equation}
819 + \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
820 + \end{equation}
821 + Each of these characteristic times can be used to predict the decay of
822 + part of the rotational correlation function when $\ell = 2$,
823 + \begin{equation}
824 + C_2(t)  =  \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
825 + \end{equation}
826 + This is the same as the $F^2_{0,0}(t)$ correlation function that
827 + appears in Ref. \citen{Berne90}.  The amplitudes of the two decay
828 + terms are expressed in terms of three dimensionless functions of the
829 + eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
830 + 2\Delta)$, and $N = 2 \sqrt{\Delta b}$.  Similar expressions can be
831 + obtained for other angular momentum correlation
832 + functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
833 + studied, only one of the amplitudes of the two decay terms was
834 + non-zero, so it was possible to derive a single relaxation time for
835 + each of the hydrodynamic tensors. In many cases, these characteristic
836 + times are averaged and reported in the literature as a single relaxation
837 + time,\cite{Garcia-de-la-Torre:1997qy}
838 + \begin{equation}
839 + 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
840 + \end{equation}
841 + although for the cases reported here, this averaging is not necessary
842 + and only one of the five relaxation times is relevant.
843 +
844 + To test the Langevin integrator's behavior for rotational relaxation,
845 + we have compared the analytical orientational relaxation times (if
846 + they are known) with the general result from the diffusion tensor and
847 + with the results from both the explicitly solvated molecular dynamics
848 + and Langevin simulations.  Relaxation times from simulations (both
849 + microcanonical and Langevin), were computed using Legendre polynomial
850 + correlation functions for a unit vector (${\bf u}$) fixed along one or
851 + more of the body-fixed axes of the model.
852 + \begin{equation}
853 + C_{\ell}(t)  =  \left\langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
854 + u}_{i}(0) \right) \right\rangle
855 + \end{equation}
856 + For simulations in the high-friction limit, orientational correlation
857 + times can then be obtained from exponential fits of this function, or by
858 + integrating,
859 + \begin{equation}
860 + \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
861 + \end{equation}
862 + In lower-friction solvents, the Legendre correlation functions often
863 + exhibit non-exponential decay, and may not be characterized by a
864 + single decay constant.
865 +
866 + In table \ref{tab:rotation} we show the characteristic rotational
867 + relaxation times (based on the diffusion tensor) for each of the model
868 + systems compared with the values obtained via microcanonical and Langevin
869 + simulations.
870 +
871 + \subsection{Spherical particles}
872 + Our model system for spherical particles was a Lennard-Jones sphere of
873 + diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
874 + 4.7 \AA).  The well depth ($\epsilon$) for both particles was set to
875 + an arbitrary value of 0.8 kcal/mol.  
876 +
877 + The Stokes-Einstein behavior of large spherical particles in
878 + hydrodynamic flows is well known, giving translational friction
879 + coefficients of $6 \pi \eta R$ (stick boundary conditions) and
880 + rotational friction coefficients of $8 \pi \eta R^3$.  Recently,
881 + Schmidt and Skinner have computed the behavior of spherical tag
882 + particles in molecular dynamics simulations, and have shown that {\it
883 + slip} boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
884 + appropriate for molecule-sized spheres embedded in a sea of spherical
885 + solvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
886 +
887 + Our simulation results show similar behavior to the behavior observed
888 + by Schmidt and Skinner.  The diffusion constant obtained from our
889 + microcanonical molecular dynamics simulations lies between the slip
890 + and stick boundary condition results obtained via Stokes-Einstein
891 + behavior.  Since the Langevin integrator assumes Stokes-Einstein stick
892 + boundary conditions in calculating the drag and random forces for
893 + spherical particles, our Langevin routine obtains nearly quantitative
894 + agreement with the hydrodynamic results for spherical particles.  One
895 + avenue for improvement of the method would be to compute elements of
896 + $\Xi_{tt}$ assuming behavior intermediate between the two boundary
897 + conditions.
898 +
899 + In the explicit solvent simulations, both our solute and solvent
900 + particles were structureless, exerting no torques upon each other.
901 + Therefore, there are not rotational correlation times available for
902 + this model system.
903 +
904 + \subsection{Ellipsoids}
905 + For uniaxial ellipsoids ($a > b = c$), Perrin's formulae for both
906 + translational and rotational diffusion of each of the body-fixed axes
907 + can be combined to give a single translational diffusion
908 + constant,\cite{Berne90}
909 + \begin{equation}
910 + D = \frac{k_B T}{6 \pi \eta a} G(\rho),
911 + \label{Dperrin}
912 + \end{equation}
913 + as well as a single rotational diffusion coefficient,
914 + \begin{equation}
915 + \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - \rho^2)
916 + G(\rho) - 1}{1 - \rho^4} \right\}.
917 + \label{ThetaPerrin}
918 + \end{equation}
919 + In these expressions, $G(\rho)$ is a function of the axial ratio
920 + ($\rho = b / a$), which for prolate ellipsoids, is
921 + \begin{equation}
922 + G(\rho) = (1- \rho^2)^{-1/2} \ln \left\{ \frac{1 + (1 -
923 + \rho^2)^{1/2}}{\rho} \right\}
924 + \label{GPerrin}
925 + \end{equation}
926 + Again, there is some uncertainty about the correct boundary conditions
927 + to use for molecular-scale ellipsoids in a sea of similarly-sized
928 + solvent particles.  Ravichandran and Bagchi found that {\it slip}
929 + boundary conditions most closely resembled the simulation
930 + results,\cite{Ravichandran:1999fk} in agreement with earlier work of
931 + Tang and Evans.\cite{TANG:1993lr}
932 +
933 + Even though there are analytic resistance tensors for ellipsoids, we
934 + constructed a rough-shell model using 2135 beads (each with a diameter
935 + of 0.25 \AA) to approximate the shape of the model ellipsoid.  We
936 + compared the Langevin dynamics from both the simple ellipsoidal
937 + resistance tensor and the rough shell approximation with
938 + microcanonical simulations and the predictions of Perrin.  As in the
939 + case of our spherical model system, the Langevin integrator reproduces
940 + almost exactly the behavior of the Perrin formulae (which is
941 + unsurprising given that the Perrin formulae were used to derive the
942 + drag and random forces applied to the ellipsoid).  We obtain
943 + translational diffusion constants and rotational correlation times
944 + that are within a few percent of the analytic values for both the
945 + exact treatment of the diffusion tensor as well as the rough-shell
946 + model for the ellipsoid.
947 +
948 + The translational diffusion constants from the microcanonical simulations
949 + agree well with the predictions of the Perrin model, although the rotational
950 + correlation times are a factor of 2 shorter than expected from hydrodynamic
951 + theory.  One explanation for the slower rotation
952 + of explicitly-solvated ellipsoids is the possibility that solute-solvent
953 + collisions happen at both ends of the solute whenever the principal
954 + axis of the ellipsoid is turning. In the upper portion of figure
955 + \ref{fig:explanation} we sketch a physical picture of this explanation.
956 + Since our Langevin integrator is providing nearly quantitative agreement with
957 + the Perrin model, it also predicts orientational diffusion for ellipsoids that
958 + exceed explicitly solvated correlation times by a factor of two.
959 +
960 + \subsection{Rigid dumbbells}
961 + Perhaps the only {\it composite} rigid body for which analytic
962 + expressions for the hydrodynamic tensor are available is the
963 + two-sphere dumbbell model.  This model consists of two non-overlapping
964 + spheres held by a rigid bond connecting their centers. There are
965 + competing expressions for the 6x6 resistance tensor for this
966 + model. Equation (\ref{introEquation:oseenTensor}) above gives the
967 + original Oseen tensor, while the second order expression introduced by
968 + Rotne and Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la
969 + Torre and Bloomfield,\cite{Torre1977} is given above as
970 + Eq. (\ref{introEquation:RPTensorNonOverlapped}).  In our case, we use
971 + a model dumbbell in which the two spheres are identical Lennard-Jones
972 + particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
973 + a distance of 6.532 \AA.
974 +
975 + The theoretical values for the translational diffusion constant of the
976 + dumbbell are calculated from the work of Stimson and Jeffery, who
977 + studied the motion of this system in a flow parallel to the
978 + inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
979 + motion in a flow {\it perpendicular} to the inter-sphere
980 + axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
981 + orientational} correlation times for this model system (other than
982 + those derived from the 6 x 6 tensors mentioned above).
983 +
984 + The bead model for this model system comprises the two large spheres
985 + by themselves, while the rough shell approximation used 3368 separate
986 + beads (each with a diameter of 0.25 \AA) to approximate the shape of
987 + the rigid body.  The hydrodynamics tensors computed from both the bead
988 + and rough shell models are remarkably similar.  Computing the initial
989 + hydrodynamic tensor for a rough shell model can be quite expensive (in
990 + this case it requires inverting a 10104 x 10104 matrix), while the
991 + bead model is typically easy to compute (in this case requiring
992 + inversion of a 6 x 6 matrix).  
993 +
994 + \begin{figure}
995 + \centering
996 + \includegraphics[width=2in]{RoughShell}
997 + \caption[Model rigid bodies and their rough shell approximations]{The
998 + model rigid bodies (left column) used to test this algorithm and their
999 + rough-shell approximations (right-column) that were used to compute
1000 + the hydrodynamic tensors.  The top two models (ellipsoid and dumbbell)
1001 + have analytic solutions and were used to test the rough shell
1002 + approximation.  The lower two models (banana and lipid) were compared
1003 + with explicitly-solvated molecular dynamics simulations. }
1004 + \label{fig:roughShell}
1005 + \end{figure}
1006 +
1007 +
1008 + Once the hydrodynamic tensor has been computed, there is no additional
1009 + penalty for carrying out a Langevin simulation with either of the two
1010 + different hydrodynamics models.  Our naive expectation is that since
1011 + the rigid body's surface is roughened under the various shell models,
1012 + the diffusion constants will be even farther from the ``slip''
1013 + boundary conditions than observed for the bead model (which uses a
1014 + Stokes-Einstein model to arrive at the hydrodynamic tensor).  For the
1015 + dumbbell, this prediction is correct although all of the Langevin
1016 + diffusion constants are within 6\% of the diffusion constant predicted
1017 + from the fully solvated system.
1018 +
1019 + For rotational motion, Langevin integration (and the hydrodynamic tensor)
1020 + yields rotational correlation times that are substantially shorter than those
1021 + obtained from explicitly-solvated simulations.  It is likely that this is due
1022 + to the large size of the explicit solvent spheres, a feature that prevents
1023 + the solvent from coming in contact with a substantial fraction of the surface
1024 + area of the dumbbell.  Therefore, the explicit solvent only provides drag
1025 + over a substantially reduced surface area of this model, while the
1026 + hydrodynamic theories utilize the entire surface area for estimating
1027 + rotational diffusion.  A sketch of the free volume available in the explicit
1028 + solvent simulations is shown in figure \ref{fig:explanation}.
1029 +
1030 +
1031 + \begin{figure}
1032 + \centering
1033 + \includegraphics[width=6in]{explanation}
1034 + \caption[Explanations of the differences between orientational
1035 + correlation times for explicitly-solvated models and hydrodynamics
1036 + predictions]{Explanations of the differences between orientational
1037 + correlation times for explicitly-solvated models and hydrodynamic
1038 + predictions.   For the ellipsoids (upper figures), rotation of the
1039 + principal axis can involve correlated collisions at both sides of the
1040 + solute.  In the rigid dumbbell model (lower figures), the large size
1041 + of the explicit solvent spheres prevents them from coming in contact
1042 + with a substantial fraction of the surface area of the dumbbell.
1043 + Therefore, the explicit solvent only provides drag over a
1044 + substantially reduced surface area of this model, where the
1045 + hydrodynamic theories utilize the entire surface area for estimating
1046 + rotational diffusion.
1047 + } \label{fig:explanation}
1048 + \end{figure}
1049 +
1050 +
1051 +
1052 + \subsection{Composite banana-shaped molecules}
1053 + Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids have
1054 + been used by Orlandi {\it et al.} to observe mesophases in
1055 + coarse-grained models for bent-core liquid crystalline
1056 + molecules.\cite{Orlandi:2006fk} We have used the same overlapping
1057 + ellipsoids as a way to test the behavior of our algorithm for a
1058 + structure of some interest to the materials science community,
1059 + although since we are interested in capturing only the hydrodynamic
1060 + behavior of this model, we have left out the dipolar interactions of
1061 + the original Orlandi model.
1062 +
1063 + A reference system composed of a single banana rigid body embedded in a
1064 + sea of 1929 solvent particles was created and run under standard
1065 + (microcanonical) molecular dynamics.  The resulting viscosity of this
1066 + mixture was 0.298 centipoise (as estimated using Eq. (\ref{eq:shear})).
1067 + To calculate the hydrodynamic properties of the banana rigid body model,
1068 + we created a rough shell (see Fig.~\ref{fig:roughShell}), in which
1069 + the banana is represented as a ``shell'' made of 3321 identical beads
1070 + (0.25 \AA\  in diameter) distributed on the surface.  Applying the
1071 + procedure described in Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1072 + identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0 \AA), as
1073 + well as the resistance tensor,
1074 + \begin{equation*}
1075 + \Xi =
1076 + \left( {\begin{array}{*{20}c}
1077 + 0.9261 & 0 & 0&0&0.08585&0.2057\\
1078 + 0& 0.9270&-0.007063& 0.08585&0&0\\
1079 + 0&-0.007063&0.7494&0.2057&0&0\\
1080 + 0&0.0858&0.2057& 58.64& 0&0\\0.08585&0&0&0&48.30&3.219&\\0.2057&0&0&0&3.219&10.7373\\\end{array}} \right),
1081 + \end{equation*}
1082 + where the units for translational, translation-rotation coupling and
1083 + rotational tensors are (kcal fs / mol \AA$^2$), (kcal fs / mol \AA\ rad),
1084 + and (kcal fs / mol rad$^2$), respectively.
1085 +
1086 + The Langevin rigid-body integrator (and the hydrodynamic diffusion tensor)
1087 + are essentially quantitative for translational diffusion of this model.  
1088 + Orientational correlation times under the Langevin rigid-body integrator
1089 + are within 11\% of the values obtained from explicit solvent, but these
1090 + models also exhibit some solvent inaccessible surface area in the
1091 + explicitly-solvated case.  
1092 +
1093 + \subsection{Composite sphero-ellipsoids}
1094 + Spherical heads perched on the ends of Gay-Berne ellipsoids have been
1095 + used recently as models for lipid
1096 + molecules.\cite{SunGezelter08,Ayton01}
1097 + MORE DETAILS
1098 +
1099 + A reference system composed of a single lipid rigid body embedded in a
1100 + sea of 1929 solvent particles was created and run under standard
1101 + (microcanonical) molecular dynamics.  The resulting viscosity of this
1102 + mixture was 0.349 centipoise (as estimated using
1103 + Eq. (\ref{eq:shear})).  To calculate the hydrodynamic properties of
1104 + the lipid rigid body model, we created a rough shell (see
1105 + Fig.~\ref{fig:roughShell}), in which the lipid is represented as a
1106 + ``shell'' made of 3550 identical beads (0.25 \AA\ in diameter)
1107 + distributed on the surface.  Applying the procedure described in
1108 + Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1109 + identified the center of resistance, ${\bf r} = $(0 \AA, 0 \AA, 1.46
1110 + \AA).
1111 +
1112 +
1113 + \subsection{Summary}
1114   According to our simulations, the langevin dynamics is a reliable
1115   theory to apply to replace the explicit solvents, especially for the
1116   translation properties. For large molecules, the rotation properties
1117   are also mimiced reasonablly well.
1118  
1119 < \begin{table*}
1120 < \begin{minipage}{\linewidth}
1121 < \begin{center}
1122 < \caption{}
1123 < \begin{tabular}{llccccccc}
1124 < \hline
1125 <  & & Sphere & Ellipsoid & Dumbbell(2 spheres) & Banana(3 ellpsoids) &
1126 < Lipid(head) & lipid(tail) & Solvent \\
1127 < \hline
1128 < $d$ (\AA) & & 6.5 & 4.6  & 6.5 &  4.2 & 6.5 & 4.6 & 4.7 \\
1129 < $l$ (\AA) & & $= d$ & 13.8 & $=d$ & 11.2 & $=d$ & 13.8 & 4.7 \\
1130 < $\epsilon^s$ (kcal/mol) & & 0.8 & 0.8 & 0.8 & 0.8 & 0.185 & 0.8 & 0.8 \\
1131 < $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 & 0.2 & 1 & 0.2 & 1 \\
1132 < $m$ (amu) & & 190 & 200 & 190 & 240 & 196 & 760 & 72.06 \\
1133 < %$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
1134 < %\multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
1135 < %\multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
757 < %\multicolumn{2}c{$I_{zz}$} &  0 &    9000 & N/A \\
758 < %$\mu$ (Debye) & & varied & 0 & 0 \\
759 < \end{tabular}
760 < \label{tab:parameters}
761 < \end{center}
762 < \end{minipage}
763 < \end{table*}
1119 > \begin{figure}
1120 > \centering
1121 > \includegraphics[width=\linewidth]{graph}
1122 > \caption[Mean squared displacements and orientational
1123 > correlation functions for each of the model rigid bodies.]{The
1124 > mean-squared displacements ($\langle r^2(t) \rangle$) and
1125 > orientational correlation functions ($C_2(t)$) for each of the model
1126 > rigid bodies studied.  The circles are the results for microcanonical
1127 > simulations with explicit solvent molecules, while the other data sets
1128 > are results for Langevin dynamics using the different hydrodynamic
1129 > tensor approximations.  The Perrin model for the ellipsoids is
1130 > considered the ``exact'' hydrodynamic behavior (this can also be said
1131 > for the translational motion of the dumbbell operating under the bead
1132 > model). In most cases, the various hydrodynamics models reproduce
1133 > each other quantitatively.}
1134 > \label{fig:results}
1135 > \end{figure}
1136  
1137   \begin{table*}
1138   \begin{minipage}{\linewidth}
1139   \begin{center}
1140 < \caption{}
1141 < \begin{tabular}{lccccc}
1140 > \caption{Translational diffusion constants (D) for the model systems
1141 > calculated using microcanonical simulations (with explicit solvent),
1142 > theoretical predictions, and Langevin simulations (with implicit solvent).
1143 > Analytical solutions for the exactly-solved hydrodynamics models are
1144 > from Refs. \citen{Einstein05} (sphere), \citen{Perrin1934} and \citen{Perrin1936}
1145 > (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1146 > (dumbbell). The other model systems have no known analytic solution.
1147 > All  diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1148 > $10^{-4}$ \AA$^2$  / fs). }
1149 > \begin{tabular}{lccccccc}
1150   \hline
1151 < & & & & &Translation \\
1152 < \hline
1153 < & NVE &  & Theoretical & Langevin & \\
774 < \hline
775 < & $\eta$ & D & D & method & D \\
1151 > & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1152 > \cline{2-3} \cline{5-7}
1153 > model & $\eta$ (centipoise)  & D & & Analytical & method & Hydrodynamics & simulation \\
1154   \hline
1155 < sphere & 3.480159e-03 & 1.643135e-04 & 1.942779e-04 & exact & 1.982283e-04 \\
1156 < ellipsoid & 2.551262e-03 & 2.437492e-04 & 2.335756e-04 & exact & 2.374905e-04 \\
1157 < & 2.551262e-03  & 2.437492e-04 & 2.335756e-04 & rough shell & 2.284088e-04 \\
1158 < dumbell & 2.41276e-03  & 2.129432e-04 & 2.090239e-04 & bead model & 2.148098e-04 \\
1159 < & 2.41276e-03 & 2.129432e-04 & 2.090239e-04 & rough shell & 2.013219e-04 \\
1160 < banana & 2.9846e-03 & 1.527819e-04 &  & rough shell & 1.54807e-04 \\
1161 < lipid & 3.488661e-03 & 0.9562979e-04 &  & rough shell & 1.320987e-04 \\
1155 > sphere    & 0.279  & 3.06 & & 2.42 & exact       & 2.42 & 2.33 \\
1156 > ellipsoid & 0.255  & 2.44 & & 2.34 & exact       & 2.34 & 2.37 \\
1157 >          & 0.255  & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1158 > dumbbell  & 0.308  & 2.06 & & 1.64 & bead model  & 1.65 & 1.62 \\
1159 >          & 0.308  & 2.06 & & 1.64 & rough shell & 1.59 & 1.62 \\
1160 > banana    & 0.298  & 1.53 & &      & rough shell & 1.56 & 1.55 \\
1161 > lipid     & 0.349  & 0.96 & &      & rough shell & 1.33 & 1.32 \\
1162   \end{tabular}
1163   \label{tab:translation}
1164   \end{center}
# Line 790 | Line 1168 | lipid & 3.488661e-03 & 0.9562979e-04 &  & rough shell
1168   \begin{table*}
1169   \begin{minipage}{\linewidth}
1170   \begin{center}
1171 < \caption{}
1172 < \begin{tabular}{lccccc}
1171 > \caption{Orientational relaxation times ($\tau$) for the model systems using
1172 > microcanonical simulation (with explicit solvent), theoretical
1173 > predictions, and Langevin simulations (with implicit solvent). All
1174 > relaxation times are for the rotational correlation function with
1175 > $\ell = 2$ and are reported in units of ps.  The ellipsoidal model has
1176 > an exact solution for the orientational correlation time due to
1177 > Perrin, but the other model systems have no known analytic solution.}
1178 > \begin{tabular}{lccccccc}
1179   \hline
1180 < & & & & &Rotation \\
1181 < \hline
1182 < & NVE &  & Theoretical & Langevin & \\
799 < \hline
800 < & $\eta$ & $\tau_0$ & $\tau_0$ & method & $\tau_0$ \\
1180 > & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1181 > \cline{2-3} \cline{5-7}
1182 > model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic  & simulation \\
1183   \hline
1184 < sphere & 3.480159e-03 &  & 1.208178e+04 & exact & 1.20628e+04 \\
1185 < ellipsoid & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & exact & 2.21507e+04 \\
1186 < & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & rough shell & 2.21714e+04 \\
1187 < dumbell & 2.41276e-03 & 1.42974e+04 &  & bead model & 7.12435e+04 \\
1188 < & 2.41276e-03 & 1.42974e+04 &  & rough shell & 7.04765e+04 \\
1189 < banana & 2.9846e-03 & 6.38323e+04 &  & rough shell & 7.0945e+04 \\
1190 < lipid & 3.488661e-03 & 7.79595e+04 &  & rough shell & 7.78886e+04 \\
1184 > sphere    & 0.279  &      & & 9.69 & exact       & 9.69 & 9.64 \\
1185 > ellipsoid & 0.255  & 46.7 & & 22.0 & exact       & 22.0 & 22.2 \\
1186 >          & 0.255  & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1187 > dumbbell  & 0.308  & 14.1 & &      & bead model  & 50.0 & 50.1 \\
1188 >          & 0.308  & 14.1 & &      & rough shell & 41.5 & 41.3 \\
1189 > banana    & 0.298  & 63.8 & &      & rough shell & 70.9 & 70.9 \\
1190 > lipid     & 0.349  & 78.0 & &      & rough shell & 76.9 & 77.9 \\
1191 > \hline
1192   \end{tabular}
1193   \label{tab:rotation}
1194   \end{center}
1195   \end{minipage}
1196   \end{table*}
1197  
1198 < Langevin dynamics simulations are applied to study the formation of
1199 < the ripple phase of lipid membranes. The initial configuration is
1198 > \section{Application: A rigid-body lipid bilayer}
1199 >
1200 > The Langevin dynamics integrator was applied to study the formation of
1201 > corrugated structures emerging from simulations of the coarse grained
1202 > lipid molecular models presented above.  The initial configuration is
1203   taken from our molecular dynamics studies on lipid bilayers with
1204 < lennard-Jones sphere solvents. The solvent molecules are excluded from
1205 < the system, the experimental value of water viscosity is applied to
1206 < mimic the heat bath. Fig. XXX is the snapshot of the stable
1207 < configuration of the system, the ripple structure stayed stable after
1208 < 100 ns run. The efficiency of the simulation is increased by one order
1204 > lennard-Jones sphere solvents. The solvent molecules were excluded
1205 > from the system, and the experimental value for the viscosity of water
1206 > at 20C ($\eta = 1.00$ cp) was used to mimic the hydrodynamic effects
1207 > of the solvent.  The absence of explicit solvent molecules and the
1208 > stability of the integrator allowed us to take timesteps of 50 fs.  A
1209 > total simulation run time of 100 ns was sampled.
1210 > Fig. \ref{fig:bilayer} shows the configuration of the system after 100
1211 > ns, and the ripple structure remains stable during the entire
1212 > trajectory.  Compared with using explicit bead-model solvent
1213 > molecules, the efficiency of the simulation has increased by an order
1214   of magnitude.
1215  
825 \subsection{Langevin Dynamics of Banana Shaped Molecules}
826
827 In order to verify that Langevin dynamics can mimic the dynamics of
828 the systems absent of explicit solvents, we carried out two sets of
829 simulations and compare their dynamic properties.
830 Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
831 made of 256 pentane molecules and two banana shaped molecules at
832 273~K. It has an equivalent implicit solvent system containing only
833 two banana shaped molecules with viscosity of 0.289 center poise. To
834 calculate the hydrodynamic properties of the banana shaped molecule,
835 we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
836 in which the banana shaped molecule is represented as a ``shell''
837 made of 2266 small identical beads with size of 0.3 \AA on the
838 surface. Applying the procedure described in
839 Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
840 identified the center of resistance at (0 $\rm{\AA}$, 0.7482 $\rm{\AA}$,
841 -0.1988 $\rm{\AA}$), as well as the resistance tensor,
842 \[
843 \left( {\begin{array}{*{20}c}
844 0.9261 & 0 & 0&0&0.08585&0.2057\\
845 0& 0.9270&-0.007063& 0.08585&0&0\\
846 0&-0.007063&0.7494&0.2057&0&0\\
847 0&0.0858&0.2057& 58.64& 0&0\\
848 0.08585&0&0&0&48.30&3.219&\\
849 0.2057&0&0&0&3.219&10.7373\\
850 \end{array}} \right).
851 \]
852 where the units for translational, translation-rotation coupling and rotational tensors are $\frac{kcal \cdot fs}{mol \cdot \rm{\AA}^2}$, $\frac{kcal \cdot fs}{mol \cdot \rm{\AA} \cdot rad}$ and $\frac{kcal \cdot fs}{mol \cdot rad^2}$ respectively.
853 Curves of the velocity auto-correlation functions in
854 Fig.~\ref{langevin:vacf} were shown to match each other very well.
855 However, because of the stochastic nature, simulation using Langevin
856 dynamics was shown to decay slightly faster than MD. In order to
857 study the rotational motion of the molecules, we also calculated the
858 auto-correlation function of the principle axis of the second GB
859 particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
860 probably due to the reason that we used the experimental viscosity directly instead of calculating bulk viscosity from simulation.
861
1216   \begin{figure}
1217   \centering
1218 < \includegraphics[width=\linewidth]{roughShell.pdf}
1219 < \caption[Rough shell model for banana shaped molecule]{Rough shell
1220 < model for banana shaped molecule.} \label{langevin:roughShell}
1218 > \includegraphics[width=\linewidth]{bilayer}
1219 > \caption[Snapshot of a bilayer of rigid-body models for lipids]{A
1220 > snapshot of a bilayer composed of rigid-body models for lipid
1221 > molecules evolving using the Langevin integrator described in this
1222 > work.} \label{fig:bilayer}
1223   \end{figure}
1224  
869 \begin{figure}
870 \centering
871 \includegraphics[width=\linewidth]{twoBanana.pdf}
872 \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
873 256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
874 molecules and 256 pentane molecules.} \label{langevin:twoBanana}
875 \end{figure}
876
877 \begin{figure}
878 \centering
879 \includegraphics[width=\linewidth]{vacf.pdf}
880 \caption[Plots of Velocity Auto-correlation Functions]{Velocity
881 auto-correlation functions of NVE (explicit solvent) in blue and
882 Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
883 \end{figure}
884
885 \begin{figure}
886 \centering
887 \includegraphics[width=\linewidth]{uacf.pdf}
888 \caption[Auto-correlation functions of the principle axis of the
889 middle GB particle]{Auto-correlation functions of the principle axis
890 of the middle GB particle of NVE (blue) and Langevin dynamics
891 (red).} \label{langevin:uacf}
892 \end{figure}
893
1225   \section{Conclusions}
1226  
1227   We have presented a new Langevin algorithm by incorporating the
1228   hydrodynamics properties of arbitrary shaped molecules into an
1229 < advanced symplectic integration scheme. The temperature control
1230 < ability of this algorithm was demonstrated by a set of simulations
1231 < with different viscosities. It was also shown to have significant
1232 < advantage of producing rapid thermal equilibration over
902 < Nos\'{e}-Hoover method. Further studies in systems involving banana
903 < shaped molecules illustrated that the dynamic properties could be
904 < preserved by using this new algorithm as an implicit solvent model.
1229 > advanced symplectic integration scheme. Further studies in systems
1230 > involving banana shaped molecules illustrated that the dynamic
1231 > properties could be preserved by using this new algorithm as an
1232 > implicit solvent model.
1233  
1234  
1235   \section{Acknowledgments}
# Line 911 | Line 1239 | of Notre Dame.
1239   of Notre Dame.
1240   \newpage
1241  
1242 < \bibliographystyle{jcp2}
1242 > \bibliographystyle{jcp}
1243   \bibliography{langevin}
1244  
1245   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines