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 3299 by gezelter, Mon Jan 7 21:30:15 2008 UTC vs.
Revision 3340 by gezelter, Thu Jan 31 22:13:55 2008 UTC

# Line 2 | Line 2
2   %\documentclass[aps,prb,preprint]{revtex4}
3   \documentclass[11pt]{article}
4   \usepackage{endfloat}
5 < \usepackage{amsmath,bm}
5 > \usepackage{amsmath}
6   \usepackage{amssymb}
7   \usepackage{times}
8   \usepackage{mathptmx}
# 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' 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{HuseyinKaya07012005}
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 Stokes' law,
89 > \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  
97 < It is very important to develop stable and efficient methods to
98 < integrate the equations of motion for orientational degrees of
99 < freedom. Euler angles are the natural choice to describe the
100 < rotational degrees of freedom. However, due to $\frac {1}{sin
101 < \theta}$ singularities, the numerical integration of corresponding
102 < equations of these motion is very inefficient and inaccurate.
103 < Although an alternative integrator using multiple sets of Euler
104 < angles can overcome this difficulty\cite{Barojas1973}, the
97 > 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 > 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 > motion.\cite{FIXMAN:1986lr,Ramachandran1996}
110 >
111 > 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 > \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 > 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 >
149 > Euler angles are a natural choice to describe the rotational degrees
150 > of freedom.  However, due to $\frac{1}{\sin \theta}$ singularities, the
151 > 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 < still remain. A singularity-free representation utilizing
156 < quaternions was developed by Evans in 1977.\cite{Evans1977}
157 < Unfortunately, this approach used a nonseparable Hamiltonian
158 < resulting from the quaternion representation, which prevented the
159 < symplectic algorithm from being utilized. Another different approach
103 < 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}
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  
161 < A break-through in geometric literature suggests that, in order to
162 < develop a long-term integration scheme, one should preserve the
163 < symplectic structure of the propagator. By introducing a conjugate
164 < momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
165 < equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
166 < proposed to evolve the Hamiltonian system in a constraint manifold
167 < by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
117 < An alternative method using the quaternion representation was
118 < developed by Omelyan.\cite{Omelyan1998} However, both of these
119 < methods are iterative and inefficient. In this section, we descibe a
120 < symplectic Lie-Poisson integrator for rigid bodies developed by
121 < Dullweber and his coworkers\cite{Dullweber1997} in depth.
161 > 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  
169 < %review langevin/browninan dynamics for arbitrarily shaped rigid body
170 < Combining Langevin or Brownian dynamics with rigid body dynamics,
171 < one can study slow processes in biomolecular systems. Modeling DNA
172 < as a chain of rigid beads, which are subject to harmonic potentials
173 < as well as excluded volume potentials, Mielke and his coworkers
174 < discovered rapid superhelical stress generations from the stochastic
175 < simulation of twin supercoiling DNA with response to induced
176 < torques.\cite{Mielke2004} Membrane fusion is another key biological
177 < process which controls a variety of physiological functions, such as
178 < release of neurotransmitters \textit{etc}. A typical fusion event
179 < happens on the time scale of a millisecond, which is impractical to
180 < study using atomistic models with newtonian mechanics. With the help
181 < of coarse-grained rigid body model and stochastic dynamics, the
136 < fusion pathways were explored by many
137 < researchers.\cite{Noguchi2001,Noguchi2002,Shillcock2005} Due to the
138 < difficulty of numerical integration of anisotropic rotation, most of
139 < the rigid body models are simply modeled using spheres, cylinders,
140 < ellipsoids or other regular shapes in stochastic simulations. In an
141 < effort to account for the diffusion anisotropy of arbitrary
142 < particles, Fernandes and de la Torre improved the original Brownian
143 < dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
144 < incorporating a generalized $6\times6$ diffusion tensor and
145 < introducing a simple rotation evolution scheme consisting of three
146 < consecutive rotations.\cite{Fernandes2002} Unfortunately, unexpected
147 < errors and biases are introduced into the system due to the
148 < arbitrary order of applying the noncommuting rotation
149 < operators.\cite{Beard2003} Based on the observation the momentum
150 < relaxation time is much less than the time step, one may ignore the
151 < inertia in Brownian dynamics. However, the assumption of zero
152 < average acceleration is not always true for cooperative motion which
153 < is common in protein motion. An inertial Brownian dynamics (IBD) was
154 < proposed to address this issue by adding an inertial correction
155 < term.\cite{Beard2000} As a complement to IBD which has a lower bound
156 < in time step because of the inertial relaxation time, long-time-step
157 < inertial dynamics (LTID) can be used to investigate the inertial
158 < behavior of the polymer segments in low friction
159 < regime.\cite{Beard2000} LTID can also deal with the rotational
160 < dynamics for nonskew bodies without translation-rotation coupling by
161 < separating the translation and rotation motion and taking advantage
162 < of the analytical solution of hydrodynamics properties. However,
163 < typical nonskew bodies like cylinders and ellipsoids are inadequate
164 < 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.
169 > 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  
183   The goal of the present work is to develop a Langevin dynamics
184   algorithm for arbitrary-shaped rigid particles by integrating the
185 < accurate estimation of friction tensor from hydrodynamics theory
186 < into the sophisticated rigid body dynamics algorithms.
185 > accurate estimation of friction tensor from hydrodynamics theory into
186 > a symplectic rigid body dynamics propagator.  In the sections below,
187 > we review some of the theory of hydrodynamic tensors developed
188 > 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  
194 < \section{Computational Methods{\label{methodSec}}}
195 <
176 < \subsection{\label{introSection:frictionTensor}Friction Tensor}
177 < Theoretically, the friction kernel can be determined using the
194 > \subsection{\label{introSection:frictionTensor}The Friction Tensor}
195 > Theoretically, a complete friction kernel can be determined using the
196   velocity autocorrelation function. However, this approach becomes
197 < impractical when the system becomes more and more complicated.
198 < Instead, various approaches based on hydrodynamics have been
199 < developed to calculate the friction coefficients. In general, the
200 < friction tensor $\Xi$ is a $6\times 6$ matrix given by
201 < \[
202 < \Xi  = \left( {\begin{array}{*{20}c}
203 <   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
204 <   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
205 < \end{array}} \right).
206 < \]
207 < Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are $3 \times 3$
208 < translational friction tensor and rotational resistance (friction)
209 < tensor respectively, while ${\Xi^{tr} }$ is translation-rotation
210 < coupling tensor and $ {\Xi^{rt} }$ is rotation-translation coupling
211 < tensor. When a particle moves in a fluid, it may experience friction
212 < force or torque along the opposite direction of the velocity or
213 < angular velocity,
214 < \[
197 > impractical when the solute becomes complex. Instead, various
198 > approaches based on hydrodynamics have been developed to calculate the
199 > friction coefficients. In general, the friction tensor $\Xi$ is a
200 > $6\times 6$ matrix given by
201 > \begin{equation}
202 > \Xi  = \left( \begin{array}{*{20}c}
203 >   \Xi^{tt} & \Xi^{rt}  \\
204 >   \Xi^{tr} & \Xi^{rr}  \\
205 > \end{array} \right).
206 > \end{equation}
207 > Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
208 > rotational resistance (friction) tensors respectively, while
209 > $\Xi^{tr}$ is translation-rotation coupling tensor and $\Xi^{rt}$ is
210 > rotation-translation coupling tensor. When a particle moves in a
211 > fluid, it may experience friction force ($\mathbf{f}_f$) and torque
212 > ($\mathbf{\tau}_f$) in opposition to the directions of the velocity
213 > ($\mathbf{v}$) and body-fixed angular velocity ($\mathbf{\omega}$),
214 > \begin{equation}
215   \left( \begin{array}{l}
216 < F_R  \\
217 < \tau _R  \\
218 < \end{array} \right) =  - \left( {\begin{array}{*{20}c}
219 <   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
220 <   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
221 < \end{array}} \right)\left( \begin{array}{l}
222 < v \\
223 < w \\
224 < \end{array} \right)
225 < \]
208 < where $F_r$ is the friction force and $\tau _R$ is the friction
209 < torque.
216 > \mathbf{f}_f  \\
217 > \mathbf{\tau}_f  \\
218 > \end{array} \right) =  - \left( \begin{array}{*{20}c}
219 >   \Xi^{tt} & \Xi^{rt}  \\
220 >   \Xi^{tr} & \Xi^{rr}  \\
221 > \end{array} \right)\left( \begin{array}{l}
222 > \mathbf{v} \\
223 > \mathbf{\omega} \\
224 > \end{array} \right).
225 > \end{equation}
226  
227   \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
228 <
229 < For a spherical particle with slip boundary conditions, the
230 < translational and rotational friction constant can be calculated
231 < from Stoke's law,
232 < \[
217 < \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
228 > For a spherical particle under ``stick'' boundary conditions, the
229 > translational and rotational friction tensors can be calculated from
230 > Stokes' law,
231 > \begin{equation}
232 > \Xi^{tt}  = \left( \begin{array}{*{20}c}
233     {6\pi \eta R} & 0 & 0  \\
234     0 & {6\pi \eta R} & 0  \\
235     0 & 0 & {6\pi \eta R}  \\
236 < \end{array}} \right)
237 < \]
236 > \end{array} \right)
237 > \end{equation}
238   and
239 < \[
240 < \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
239 > \begin{equation}
240 > \Xi^{rr}  = \left( \begin{array}{*{20}c}
241     {8\pi \eta R^3 } & 0 & 0  \\
242     0 & {8\pi \eta R^3 } & 0  \\
243     0 & 0 & {8\pi \eta R^3 }  \\
244 < \end{array}} \right)
245 < \]
244 > \end{array} \right)
245 > \end{equation}
246   where $\eta$ is the viscosity of the solvent and $R$ is the
247   hydrodynamic radius.
248  
249   Other non-spherical shapes, such as cylinders and ellipsoids, are
250 < widely used as references for developing new hydrodynamics theory,
250 > widely used as references for developing new hydrodynamics theories,
251   because their properties can be calculated exactly. In 1936, Perrin
252 < extended Stokes's law to general ellipsoids, also called a triaxial
253 < ellipsoid, which is given in Cartesian coordinates
254 < by\cite{Perrin1934, Perrin1936}
255 < \[
256 < \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
257 < }} = 1
258 < \]
259 < where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
260 < due to the complexity of the elliptic integral, only the ellipsoid
261 < with the restriction of two axes being equal, \textit{i.e.}
262 < prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
263 < exactly. Introducing an elliptic integral parameter $S$ for prolate
264 < ellipsoids :
265 < \[
266 < S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
267 < } }}{b},
268 < \]
269 < and oblate ellipsoids:
270 < \[
271 < S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
272 < }}{a},
258 < \]
259 < one can write down the translational and rotational resistance
260 < tensors
261 < \begin{eqnarray*}
262 < \Xi _a^{tt}  & = & 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}}. \\
263 < \Xi _b^{tt}  & = & \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S +
264 < 2a}},
265 < \end{eqnarray*}
266 < and
267 < \begin{eqnarray*}
268 < \Xi _a^{rr} & = & \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}}, \\
269 < \Xi _b^{rr} & = & \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}}.
252 > extended Stokes' law to general ellipsoids which are given in
253 > Cartesian coordinates by~\cite{Perrin1934,Perrin1936}
254 > \begin{equation}
255 > \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1.
256 > \end{equation}
257 > Here, the semi-axes are of lengths $a$, $b$, and $c$. Due to the
258 > complexity of the elliptic integral, only uniaxial ellipsoids, either
259 > prolate ($a \ge b = c$) or oblate ($a < b = c$), can be solved
260 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
261 > \begin{equation}
262 > S = \frac{2}{\sqrt{a^2  - b^2}} \ln \frac{a + \sqrt{a^2  - b^2}}{b},
263 > \end{equation}
264 > and oblate,
265 > \begin{equation}
266 > S = \frac{2}{\sqrt {b^2  - a^2 }} \arctan \frac{\sqrt {b^2  - a^2}}{a},
267 > \end{equation}
268 > ellipsoids, one can write down the translational and rotational
269 > resistance tensors:
270 > \begin{eqnarray*}
271 > \Xi_a^{tt}  & = & 16\pi \eta \frac{a^2  - b^2}{(2a^2  - b^2 )S - 2a}. \\
272 > \Xi_b^{tt} =  \Xi_c^{tt} & = & 32\pi \eta \frac{a^2  - b^2 }{(2a^2 - 3b^2 )S + 2a},
273   \end{eqnarray*}
274 + for oblate, and
275 + \begin{eqnarray*}
276 + \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2  - b^2 )b^2}{2a - b^2 S}, \\
277 + \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4  - b^4)}{(2a^2  - b^2 )S - 2a}
278 + \end{eqnarray*}
279 + for prolate ellipsoids. For both spherical and ellipsoidal particles,
280 + the translation-rotation and rotation-translation coupling tensors are
281 + zero.
282  
283   \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
273
284   Unlike spherical and other simply shaped molecules, there is no
285   analytical solution for the friction tensor for arbitrarily shaped
286   rigid molecules. The ellipsoid of revolution model and general
287   triaxial ellipsoid model have been used to approximate the
288 < hydrodynamic properties of rigid bodies. However, since the mapping
289 < from all possible ellipsoidal spaces, $r$-space, to all possible
290 < combination of rotational diffusion coefficients, $D$-space, is not
291 < unique\cite{Wegener1979} as well as the intrinsic coupling between
292 < translational and rotational motion of rigid bodies, general
293 < ellipsoids are not always suitable for modeling arbitrarily shaped
294 < rigid molecules. A number of studies have been devoted to
288 > hydrodynamic properties of rigid bodies. However, the mapping from all
289 > possible ellipsoidal spaces, $r$-space, to all possible combination of
290 > rotational diffusion coefficients, $D$-space, is not
291 > unique.\cite{Wegener1979} Additionally, because there is intrinsic
292 > coupling between translational and rotational motion of rigid bodies,
293 > general ellipsoids are not always suitable for modeling arbitrarily
294 > shaped rigid molecules.  A number of studies have been devoted to
295   determining the friction tensor for irregularly shaped rigid bodies
296 < using more advanced methods where the molecule of interest was
297 < modeled by a combinations of spheres\cite{Carrasco1999} and the
298 < hydrodynamics properties of the molecule can be calculated using the
299 < hydrodynamic interaction tensor. Let us consider a rigid assembly of
300 < $N$ beads immersed in a continuous medium. Due to hydrodynamic
301 < interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
302 < than its unperturbed velocity $v_i$,
303 < \[
296 > using more advanced methods where the molecule of interest was modeled
297 > by a combinations of spheres\cite{Carrasco1999} and the hydrodynamics
298 > properties of the molecule can be calculated using the hydrodynamic
299 > interaction tensor.
300 >
301 > Consider a rigid assembly of $N$ beads immersed in a continuous
302 > medium. Due to hydrodynamic interaction, the ``net'' velocity of $i$th
303 > bead, $v'_i$ is different than its unperturbed velocity $v_i$,
304 > \begin{equation}
305   v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
306 < \]
307 < where $F_i$ is the frictional force, and $T_{ij}$ is the
308 < hydrodynamic interaction tensor. The friction force of $i$th bead is
309 < proportional to its ``net'' velocity
306 > \end{equation}
307 > where $F_i$ is the frictional force, and $T_{ij}$ is the hydrodynamic
308 > interaction tensor. The frictional force on the $i^\mathrm{th}$ bead
309 > is proportional to its ``net'' velocity
310   \begin{equation}
311   F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
312   \label{introEquation:tensorExpression}
# Line 332 | Line 343 | B = \left( {\begin{array}{*{20}c}
343   construct a $3N \times 3N$ matrix consisting of $N \times N$
344   $B_{ij}$ blocks
345   \begin{equation}
346 < B = \left( {\begin{array}{*{20}c}
347 <   {B_{11} } &  \ldots  & {B_{1N} }  \\
346 > B = \left( \begin{array}{*{20}c}
347 >   B_{11} &  \ldots  & B_{1N}   \\
348      \vdots  &  \ddots  &  \vdots   \\
349 <   {B_{N1} } &  \cdots  & {B_{NN} }  \\
350 < \end{array}} \right),
349 >   B_{N1} &  \cdots  & B_{NN} \\
350 > \end{array} \right),
351   \end{equation}
352   where $B_{ij}$ is given by
353 < \[
353 > \begin{equation}
354   B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
355   )T_{ij}
356 < \]
356 > \end{equation}
357   where $\delta _{ij}$ is the Kronecker delta function. Inverting the
358   $B$ matrix, we obtain
359   \[
# Line 365 | Line 376 | arbitrary origin $O$ can be written as
376   bead $i$ and origin $O$, the elements of resistance tensor at
377   arbitrary origin $O$ can be written as
378   \begin{eqnarray}
379 + \label{introEquation:ResistanceTensorArbitraryOrigin}
380   \Xi _{}^{tt}  & = & \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
381   \Xi _{}^{tr}  & = & \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
382 < \Xi _{}^{rr}  & = &  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
383 < \label{introEquation:ResistanceTensorArbitraryOrigin}
382 > \Xi _{}^{rr}  & = &  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } }
383 > U_j  + 6 \eta V {\bf I}. \notag
384   \end{eqnarray}
385 + The final term in the expression for $\Xi^{rr}$ is correction that
386 + accounts for errors in the rotational motion of certain kinds of bead
387 + models. The additive correction uses the solvent viscosity ($\eta$)
388 + as well as the total volume of the beads that contribute to the
389 + hydrodynamic model,
390 + \begin{equation}
391 + V = \frac{4 \pi}{3} \sum_{i=1}^{N} \sigma_i^3,
392 + \end{equation}
393 + where $\sigma_i$ is the radius of bead $i$.  This correction term was
394 + rigorously tested and compared with the analytical results for
395 + two-sphere and ellipsoidal systems by Garcia de la Torre and
396 + Rodes.\cite{Torre:1983lr}
397 +
398 +
399   The resistance tensor depends on the origin to which they refer. The
400   proper location for applying the friction force is the center of
401   resistance (or center of reaction), at which the trace of rotational
# Line 400 | Line 426 | U_{OP}  = \left( {\begin{array}{*{20}c}
426   \[
427   U_{OP}  = \left( {\begin{array}{*{20}c}
428     0 & { - z_{OP} } & {y_{OP} }  \\
429 <   {z_i } & 0 & { - x_{OP} }  \\
429 >   {z_{OP} } & 0 & { - x_{OP} }  \\
430     { - y_{OP} } & {x_{OP} } & 0  \\
431   \end{array}} \right)
432   \]
# Line 412 | Line 438 | locate the position of center of resistance,
438   x_{OR}  \\
439   y_{OR}  \\
440   z_{OR}  \\
441 < \end{array} \right) & = &\left( {\begin{array}{*{20}c}
441 > \end{array} \right) & = &\left( \begin{array}{*{20}c}
442     {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
443     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
444     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
445 < \end{array}} \right)^{ - 1}  \\
445 > \end{array} \right)^{ - 1}  \\
446    & & \left( \begin{array}{l}
447   (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
448   (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
449   (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
450   \end{array} \right) \\
451   \end{eqnarray*}
452 < where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
452 > where $x_{OR}$, $y_{OR}$, $z_{OR}$ are the components of the vector
453   joining center of resistance $R$ and origin $O$.
454  
429 \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
455  
456 + \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
457 +
458   Consider the Langevin equations of motion in generalized coordinates
459   \begin{equation}
460 < M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i} (t)
460 > \mathbf{M} \dot{\mathbf{V}}(t) = \mathbf{F}_{s}(t) +
461 > \mathbf{F}_{f}(t)  + \mathbf{F}_{r}(t)
462   \label{LDGeneralizedForm}
463   \end{equation}
464 < where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
465 < and moment of inertial) matrix and $V_i$ is a generalized velocity,
466 < $V_i = V_i(v_i,\omega _i)$. The right side of
467 < Eq.~\ref{LDGeneralizedForm} consists of three generalized forces in
468 < lab-fixed frame, systematic force $F_{s,i}$, dissipative force
469 < $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
470 < system in Newtownian mechanics typically refers to lab-fixed frame,
471 < it is also convenient to handle the rotation of rigid body in
472 < body-fixed frame. Thus the friction and random forces are calculated
473 < in body-fixed frame and converted back to lab-fixed frame by:
474 < \[
475 < \begin{array}{l}
476 < F_{f,i}^l (t) = Q^T F_{f,i}^b (t), \\
449 < F_{r,i}^l (t) = Q^T F_{r,i}^b (t). \\
450 < \end{array}
451 < \]
452 < Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
453 < the body-fixed velocity at center of resistance $v_{R,i}^b$ and
454 < angular velocity $\omega _i$
464 > where $\mathbf{M}$ is a $6 \times 6$ diagonal mass matrix (which
465 > includes the mass of the rigid body as well as the moments of inertia
466 > in the body-fixed frame) and $\mathbf{V}$ is a generalized velocity,
467 > $\mathbf{V} =
468 > \left\{\mathbf{v},\mathbf{\omega}\right\}$. The right side of
469 > Eq.~\ref{LDGeneralizedForm} consists of three generalized forces: a
470 > system force $\mathbf{F}_{s}$, a frictional or dissipative force
471 > $\mathbf{F}_{f}$ and stochastic force $\mathbf{F}_{r}$. While the
472 > evolution of the system in Newtonian mechanics is typically done in
473 > the lab-fixed frame, it is convenient to handle the dynamics of rigid
474 > bodies in the body-fixed frame. Thus the friction and random forces
475 > are calculated in body-fixed frame and may be converted back to
476 > lab-fixed frame using the rigid body's rotation matrix ($Q$):
477   \begin{equation}
478 < F_{r,i}^b (t) = \left( \begin{array}{l}
479 < f_{r,i}^b (t) \\
480 < \tau _{r,i}^b (t) \\
481 < \end{array} \right) =  - \left( {\begin{array}{*{20}c}
482 <   {\Xi _{R,t} } & {\Xi _{R,c}^T }  \\
483 <   {\Xi _{R,c} } & {\Xi _{R,r} }  \\
484 < \end{array}} \right)\left( \begin{array}{l}
485 < v_{R,i}^b (t) \\
486 < \omega _i (t) \\
478 > \mathbf{F}_{f,r} =
479 > \left( \begin{array}{c}
480 > \mathbf{f}_{f,r} \\
481 > \mathbf{\tau}_{f,r}
482 > \end{array} \right)
483 > =
484 > \left( \begin{array}{c}
485 > Q^{T} \mathbf{f}^{b}_{f,r} \\
486 > Q^{T} \mathbf{\tau}^{b}_{f,r}
487 > \end{array} \right)
488 > \end{equation}
489 > The body-fixed friction force, $\mathbf{F}_{f}^b$, is proportional to
490 > the velocity at the center of resistance $\mathbf{v}_{R}^b$ (in the
491 > body-fixed frame) and the angular velocity $\mathbf{\omega}$
492 > \begin{equation}
493 > \mathbf{F}_{f}^b (t) = \left( \begin{array}{l}
494 > \mathbf{f}_{f}^b (t) \\
495 > \mathbf{\tau}_{f}^b (t) \\
496 > \end{array} \right) =  - \left( \begin{array}{*{20}c}
497 >   \Xi_{R,t} & \Xi_{R,c}^T  \\
498 >   \Xi_{R,c} & \Xi_{R,r}    \\
499 > \end{array} \right)\left( \begin{array}{l}
500 > \mathbf{v}_{R}^b (t) \\
501 > \mathbf{\omega} (t) \\
502   \end{array} \right),
503   \end{equation}
504 < while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
505 < with zero mean and variance
504 > while the random force, $\mathbf{F}_{r}$, is a Gaussian stochastic
505 > variable with zero mean and variance
506   \begin{equation}
507 < \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle  =
508 < \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle  =
509 < 2k_B T\Xi _R \delta (t - t'). \label{randomForce}
507 > \left\langle {\mathbf{F}_{r}(t) (\mathbf{F}_{r}(t'))^T } \right\rangle  =
508 > \left\langle {\mathbf{F}_{r}^b (t) (\mathbf{F}_{r}^b (t'))^T } \right\rangle  =
509 > 2 k_B T \Xi_R \delta(t - t'). \label{randomForce}
510   \end{equation}
511 < The equation of motion for $v_i$ can be written as
511 > $\Xi_R$ is the $6\times6$ resistance tensor at the center of
512 > resistance.  Once this tensor is known for a given rigid body,
513 > obtaining a stochastic vector that has the properties in
514 > Eq. (\ref{eq:randomForce}) can be done efficiently by carrying out a
515 > one-time Cholesky decomposition to obtain the square root matrix of
516 > the resistance tensor $\Xi_R = \mathbf{S} \mathbf{S}^{T}$, where
517 > $\mathbf{S}$ is a lower triangular matrix.\cite{SchlickBook} A vector
518 > with the statistics required for the random force can then be obtained
519 > by multiplying $\mathbf{S}$ onto a 6-vector $Z$ which has elements
520 > chosen from a Gaussian distribution, such that:
521   \begin{equation}
522 < m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
523 < f_{r,i}^l (t)
522 > \langle Z_i \rangle = 0, \hspace{1in} \langle Z_i \cdot Z_j \rangle = \frac{2 k_B
523 > T}{\delta t} \delta_{ij}.
524   \end{equation}
525 + The random force, $F_{r}^{b} = \mathbf{S} Z$, can be shown to have the
526 + correct ohmic
527 +
528 +
529 + Each
530 + time a random force vector is needed, a gaussian random vector is
531 + generated and then the square root matrix is multiplied onto this
532 + vector.
533 +
534 + The equation of motion for $\mathbf{v}$ can be written as
535 + \begin{equation}
536 + m \dot{\mathbf{v}} (t) =  \mathbf{f}_{s}^l (t) + \mathbf{f}_{f}^l (t) +
537 + \mathbf{f}_{r}^l (t)
538 + \end{equation}
539   Since the frictional force is applied at the center of resistance
540   which generally does not coincide with the center of mass, an extra
541   torque is exerted at the center of mass. Thus, the net body-fixed
542 < frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
542 > frictional torque at the center of mass, $\tau_{f}^b (t)$, is
543   given by
544   \begin{equation}
545 < \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
545 > \tau_{f}^b \leftarrow \tau_{f}^b + \mathbf{r}_{MR} \times \mathbf{f}_{f}^b
546   \end{equation}
547   where $r_{MR}$ is the vector from the center of mass to the center
548   of the resistance. Instead of integrating the angular velocity in
549   lab-fixed frame, we consider the equation of angular momentum in
550   body-fixed frame
551   \begin{equation}
552 < \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
493 < + \tau _{r,i}^b(t)
552 > \dot j(t) = \tau_{s}^b (t) + \tau_{f}^b (t) + \tau_{r}^b(t)
553   \end{equation}
554 < Embedding the friction terms into force and torque, one can
555 < integrate the langevin equations of motion for rigid body of
556 < arbitrary shape in a velocity-Verlet style 2-part algorithm, where
498 < $h= \delta t$:
554 > Embedding the friction terms into force and torque, one can integrate
555 > the Langevin equations of motion for rigid body of arbitrary shape in
556 > a velocity-Verlet style 2-part algorithm, where $h= \delta t$:
557  
558   {\tt moveA:}
559   \begin{align*}
# Line 574 | Line 632 | the velocities can be advanced to the same time value.
632      + \frac{h}{2} {\bf \tau}^b(t + h) .
633   \end{align*}
634  
635 < \section{Results and Discussion}
635 > \section{Validating the Method\label{sec:validating}}
636 > In order to validate our Langevin integrator for arbitrarily-shaped
637 > rigid bodies, we implemented the algorithm in {\sc
638 > oopse}\cite{Meineke2005} and  compared the results of this algorithm
639 > with the known
640 > hydrodynamic limiting behavior for a few model systems, and to
641 > microcanonical molecular dynamics simulations for some more
642 > complicated bodies. The model systems and their analytical behavior
643 > (if known) are summarized below. Parameters for the primary particles
644 > comprising our model systems are given in table \ref{tab:parameters},
645 > and a sketch of the arrangement of these primary particles into the
646 > model rigid bodies is shown in figure \ref{fig:models}. In table
647 > \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
648 > ellipsoidal (Gay-Berne) particles.  For spherical particles, the value
649 > of the Lennard-Jones $\sigma$ parameter is the particle diameter
650 > ($d$).  Gay-Berne ellipsoids have an energy scaling parameter,
651 > $\epsilon^s$, which describes the well depth for two identical
652 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a
653 > well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
654 > describes the ratio between the well depths in the {\it end-to-end}
655 > and side-by-side configurations.  For spheres, $\epsilon^r \equiv 1$.
656 > Moments of inertia are also required to describe the motion of primary
657 > particles with orientational degrees of freedom.
658  
659 < The Langevin algorithm described in previous section has been
660 < 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}
659 > \begin{table*}
660 > \begin{minipage}{\linewidth}
661   \begin{center}
662 < \begin{tabular}{lll}
663 <  \hline
664 <  $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
665 <  \hline
666 <  1    & 300.47 & 10.99 \\
667 <  0.1  & 301.19 & 11.136 \\
668 <  0.01 & 303.04 & 11.796 \\
669 <  \hline
662 > \caption{Parameters for the primary particles in use by the rigid body
663 > models in figure \ref{fig:models}.}
664 > \begin{tabular}{lrcccccccc}
665 > \hline
666 > & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
667 > & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
668 > $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
669 > Sphere   & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75  & 802.75  \\
670 > Ellipsoid & & 4.6 & 13.8  & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
671 > Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1   & 190 & 802.75 & 802.75 & 802.75 \\
672 > Banana  &(3 identical ellipsoids)& 4.2 & 11.2  & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
673 > Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
674 >       & Ellipsoidal Tail & 4.6 & 13.8  & 0.8   & 0.2 & 760 & 45000 & 45000 & 9000 \\
675 > Solvent &  & 4.7 & $= d$ & 0.8 & 1   & 72.06 & & & \\
676 > \hline
677   \end{tabular}
678 + \label{tab:parameters}
679   \end{center}
680 < \end{table}
680 > \end{minipage}
681 > \end{table*}
682  
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
683   \begin{figure}
684   \centering
685 < \includegraphics[width=\linewidth]{temperature.pdf}
686 < \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
687 < temperature fluctuation versus time.} \label{langevin:temperature}
685 > \includegraphics[width=3in]{sketch}
686 > \caption[Sketch of the model systems]{A sketch of the model systems
687 > used in evaluating the behavior of the rigid body Langevin
688 > integrator.} \label{fig:models}
689   \end{figure}
690  
691 < \subsection{Comparisons with Analytic and MD simulation results}
691 > \subsection{Simulation Methodology}
692 > We performed reference microcanonical simulations with explicit
693 > solvents for each of the different model system.  In each case there
694 > was one solute model and 1929 solvent molecules present in the
695 > simulation box.  All simulations were equilibrated using a
696 > constant-pressure and temperature integrator with target values of 300
697 > K for the temperature and 1 atm for pressure.  Following this stage,
698 > further equilibration and sampling was done in a microcanonical
699 > ensemble.  Since the model bodies are typically quite massive, we were
700 > able to use a time step of 25 fs.
701  
702 < In order to validate our langevin simulation procedure for
703 < arbitrarily-shaped rigid bodies, we compared the results of this
704 < procedure with the known hydrodynamic limiting behavior for a few
705 < model systems, and to microcanonical molecular dynamics simulations
706 < for some more complicated bodies.  The model systems and their
707 < analytical behavior (if known) are summarized below.
708 <
709 < \subsubsection{Spherical particles}
702 > The model systems studied used both Lennard-Jones spheres as well as
703 > uniaxial Gay-Berne ellipoids. In its original form, the Gay-Berne
704 > potential was a single site model for the interactions of rigid
705 > ellipsoidal molecules.\cite{Gay81} It can be thought of as a
706 > modification of the Gaussian overlap model originally described by
707 > Berne and Pechukas.\cite{Berne72} The potential is constructed in the
708 > familiar form of the Lennard-Jones function using
709 > orientation-dependent $\sigma$ and $\epsilon$ parameters,
710 > \begin{equation*}
711 > V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
712 > r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
713 > {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u
714 > }_i},
715 > {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
716 > -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
717 > {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
718 > \label{eq:gb}
719 > \end{equation*}
720  
721 + The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
722 + \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
723 + \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
724 + are dependent on the relative orientations of the two ellipsoids (${\bf
725 + \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
726 + inter-ellipsoid separation (${\bf \hat{r}}_{ij}$).  The shape and
727 + attractiveness of each ellipsoid is governed by a relatively small set
728 + of parameters: $l$ and $d$ describe the length and width of each
729 + uniaxial ellipsoid, while $\epsilon^s$, which describes the well depth
730 + for two identical ellipsoids in a {\it side-by-side} configuration.
731 + Additionally, a well depth aspect ratio, $\epsilon^r = \epsilon^e /
732 + \epsilon^s$, describes the ratio between the well depths in the {\it
733 + end-to-end} and side-by-side configurations.  Details of the potential
734 + are given elsewhere,\cite{Luckhurst90,Golubkov06,SunGezelter08} and an
735 + excellent overview of the computational methods that can be used to
736 + efficiently compute forces and torques for this potential can be found
737 + in Ref. \citen{Golubkov06}
738 +
739 + For the interaction between nonequivalent uniaxial ellipsoids (or
740 + between spheres and ellipsoids), the spheres are treated as ellipsoids
741 + with an aspect ratio of 1 ($d = l$) and with an well depth ratio
742 + ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$).  The form of the
743 + Gay-Berne potential we are using was generalized by Cleaver {\it et
744 + al.} and is appropriate for dissimilar uniaxial
745 + ellipsoids.\cite{Cleaver96}
746 +
747 + A switching function was applied to all potentials to smoothly turn
748 + off the interactions between a range of $22$ and $25$ \AA.  The
749 + switching function was the standard (cubic) function,
750 + \begin{equation}
751 + s(r) =
752 +        \begin{cases}
753 +        1 & \text{if $r \le r_{\text{sw}}$},\\
754 +        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
755 +        {(r_{\text{cut}} - r_{\text{sw}})^3}
756 +        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
757 +        0 & \text{if $r > r_{\text{cut}}$.}
758 +        \end{cases}
759 + \label{eq:switchingFunc}
760 + \end{equation}
761 +
762 + To measure shear viscosities from our microcanonical simulations, we
763 + used the Einstein form of the pressure correlation function,\cite{hess:209}
764 + \begin{equation}
765 + \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
766 + \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \right\rangle_{t_0}.
767 + \label{eq:shear}
768 + \end{equation}
769 + A similar form exists for the bulk viscosity
770 + \begin{equation}
771 + \kappa = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
772 + \int_{t_0}^{t_0 + t}
773 + \left(P\left(t'\right)-\left\langle P \right\rangle \right)dt'
774 + \right)^2 \right\rangle_{t_0}.
775 + \end{equation}
776 + Alternatively, the shear viscosity can also be calculated using a
777 + Green-Kubo formula with the off-diagonal pressure tensor correlation function,
778 + \begin{equation}
779 + \eta = \frac{V}{k_B T} \int_0^{\infty} \left\langle P_{xz}(t_0) P_{xz}(t_0
780 + + t) \right\rangle_{t_0} dt,
781 + \end{equation}
782 + although this method converges extremely slowly and is not practical
783 + for obtaining viscosities from molecular dynamics simulations.
784 +
785 + The Langevin dynamics for the different model systems were performed
786 + at the same temperature as the average temperature of the
787 + microcanonical simulations and with a solvent viscosity taken from
788 + Eq. (\ref{eq:shear}) applied to these simulations.  We used 1024
789 + independent solute simulations to obtain statistics on our Langevin
790 + integrator.
791 +
792 + \subsection{Analysis}
793 +
794 + The quantities of interest when comparing the Langevin integrator to
795 + analytic hydrodynamic equations and to molecular dynamics simulations
796 + are typically translational diffusion constants and orientational
797 + relaxation times.  Translational diffusion constants for point
798 + particles are computed easily from the long-time slope of the
799 + mean-square displacement,
800 + \begin{equation}
801 + D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \left\langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \right\rangle,
802 + \end{equation}
803 + of the solute molecules.  For models in which the translational
804 + diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
805 + (i.e. any non-spherically-symmetric rigid body), it is possible to
806 + compute the diffusive behavior for motion parallel to each body-fixed
807 + axis by projecting the displacement of the particle onto the
808 + body-fixed reference frame at $t=0$.  With an isotropic solvent, as we
809 + have used in this study, there are differences between the three
810 + diffusion constants, but these must converge to the same value at
811 + longer times.  Translational diffusion constants for the different
812 + shaped models are shown in table \ref{tab:translation}.
813 +
814 + In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
815 + diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
816 + {\it around} a particular body-fixed axis and {\it not} the diffusion
817 + of a vector pointing along the axis.  However, these eigenvalues can
818 + be combined to find 5 characteristic rotational relaxation
819 + times,\cite{PhysRev.119.53,Berne90}
820 + \begin{eqnarray}
821 + 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
822 + 1 / \tau_2 & = & 6 D_r - 2 \Delta \\
823 + 1 / \tau_3 & = & 3 (D_r + D_1)  \\
824 + 1 / \tau_4 & = & 3 (D_r + D_2) \\
825 + 1 / \tau_5 & = & 3 (D_r + D_3)
826 + \end{eqnarray}
827 + where
828 + \begin{equation}
829 + D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
830 + \end{equation}
831 + and
832 + \begin{equation}
833 + \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
834 + \end{equation}
835 + Each of these characteristic times can be used to predict the decay of
836 + part of the rotational correlation function when $\ell = 2$,
837 + \begin{equation}
838 + C_2(t)  =  \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
839 + \end{equation}
840 + This is the same as the $F^2_{0,0}(t)$ correlation function that
841 + appears in Ref. \citen{Berne90}.  The amplitudes of the two decay
842 + terms are expressed in terms of three dimensionless functions of the
843 + eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
844 + 2\Delta)$, and $N = 2 \sqrt{\Delta b}$.  Similar expressions can be
845 + obtained for other angular momentum correlation
846 + functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
847 + studied, only one of the amplitudes of the two decay terms was
848 + non-zero, so it was possible to derive a single relaxation time for
849 + each of the hydrodynamic tensors. In many cases, these characteristic
850 + times are averaged and reported in the literature as a single relaxation
851 + time,\cite{Garcia-de-la-Torre:1997qy}
852 + \begin{equation}
853 + 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
854 + \end{equation}
855 + although for the cases reported here, this averaging is not necessary
856 + and only one of the five relaxation times is relevant.
857 +
858 + To test the Langevin integrator's behavior for rotational relaxation,
859 + we have compared the analytical orientational relaxation times (if
860 + they are known) with the general result from the diffusion tensor and
861 + with the results from both the explicitly solvated molecular dynamics
862 + and Langevin simulations.  Relaxation times from simulations (both
863 + microcanonical and Langevin), were computed using Legendre polynomial
864 + correlation functions for a unit vector (${\bf u}$) fixed along one or
865 + more of the body-fixed axes of the model.
866 + \begin{equation}
867 + C_{\ell}(t)  =  \left\langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
868 + u}_{i}(0) \right) \right\rangle
869 + \end{equation}
870 + For simulations in the high-friction limit, orientational correlation
871 + times can then be obtained from exponential fits of this function, or by
872 + integrating,
873 + \begin{equation}
874 + \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
875 + \end{equation}
876 + In lower-friction solvents, the Legendre correlation functions often
877 + exhibit non-exponential decay, and may not be characterized by a
878 + single decay constant.
879 +
880 + In table \ref{tab:rotation} we show the characteristic rotational
881 + relaxation times (based on the diffusion tensor) for each of the model
882 + systems compared with the values obtained via microcanonical and Langevin
883 + simulations.
884 +
885 + \subsection{Spherical particles}
886   Our model system for spherical particles was a Lennard-Jones sphere of
887   diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
888   4.7 \AA).  The well depth ($\epsilon$) for both particles was set to
889 < an arbitrary value of 0.8 kcal/mol.  Parameters for our model systems
654 < are given in table \ref{tab:parameters}, and a sketch of these model
655 < rigid bodies is shown in figure \ref{fig:sketch}.
889 > an arbitrary value of 0.8 kcal/mol.  
890  
891   The Stokes-Einstein behavior of large spherical particles in
892   hydrodynamic flows is well known, giving translational friction
893   coefficients of $6 \pi \eta R$ (stick boundary conditions) and
894 < rotational friction coefficients of $8 \pi \eta R^3$.  Recently, Reid
895 < and Skinner have computed the behavior of spherical tag particles in
896 < molecular dynamics simulations, and have shown that {\it slip}
897 < boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
894 > rotational friction coefficients of $8 \pi \eta R^3$.  Recently,
895 > Schmidt and Skinner have computed the behavior of spherical tag
896 > particles in molecular dynamics simulations, and have shown that {\it
897 > slip} boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
898   appropriate for molecule-sized spheres embedded in a sea of spherical
899 < solvent particles.\cite{ReidAndSkinner}
899 > solvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
900  
901   Our simulation results show similar behavior to the behavior observed
902 < by Reid and Skinner.  The diffusion constant obtained from our
902 > by Schmidt and Skinner.  The diffusion constant obtained from our
903   microcanonical molecular dynamics simulations lies between the slip
904   and stick boundary condition results obtained via Stokes-Einstein
905   behavior.  Since the Langevin integrator assumes Stokes-Einstein stick
# Line 674 | Line 908 | conditions.
908   agreement with the hydrodynamic results for spherical particles.  One
909   avenue for improvement of the method would be to compute elements of
910   $\Xi_{tt}$ assuming behavior intermediate between the two boundary
911 < conditions.
911 > conditions.
912  
913 < In these simulations, our spherical particles were structureless, so
914 < there is no way to obtain rotational correlation times for this model
915 < system.
913 > In the explicit solvent simulations, both our solute and solvent
914 > particles were structureless, exerting no torques upon each other.
915 > Therefore, there are not rotational correlation times available for
916 > this model system.
917  
918 < \subsubsection{Ellipsoids}
919 < For uniaxial ellipsoids ($a > b = c$) , Perrin's formulae for both
918 > \subsection{Ellipsoids}
919 > For uniaxial ellipsoids ($a > b = c$), Perrin's formulae for both
920   translational and rotational diffusion of each of the body-fixed axes
921   can be combined to give a single translational diffusion
922 < constant,\cite{PecoraBerne}
922 > constant,\cite{Berne90}
923   \begin{equation}
924   D = \frac{k_B T}{6 \pi \eta a} G(\rho),
925   \label{Dperrin}
# Line 705 | Line 940 | boundary conditions most closely resembled the simulat
940   Again, there is some uncertainty about the correct boundary conditions
941   to use for molecular-scale ellipsoids in a sea of similarly-sized
942   solvent particles.  Ravichandran and Bagchi found that {\it slip}
943 < boundary conditions most closely resembled the simulation results, in
944 < agreement with earlier work of Tang and Evans.\cite{}
943 > boundary conditions most closely resembled the simulation
944 > results,\cite{Ravichandran:1999fk} in agreement with earlier work of
945 > Tang and Evans.\cite{TANG:1993lr}
946  
947 < As in the case of our spherical model system, the Langevin integrator
948 < reproduces almost exactly the behavior of the Perrin formulae (which
949 < is unsurprising given that the Perrin formulae were used to derive the
947 > Even though there are analytic resistance tensors for ellipsoids, we
948 > constructed a rough-shell model using 2135 beads (each with a diameter
949 > of 0.25 \AA) to approximate the shape of the model ellipsoid.  We
950 > compared the Langevin dynamics from both the simple ellipsoidal
951 > resistance tensor and the rough shell approximation with
952 > microcanonical simulations and the predictions of Perrin.  As in the
953 > case of our spherical model system, the Langevin integrator reproduces
954 > almost exactly the behavior of the Perrin formulae (which is
955 > unsurprising given that the Perrin formulae were used to derive the
956   drag and random forces applied to the ellipsoid).  We obtain
957   translational diffusion constants and rotational correlation times
958   that are within a few percent of the analytic values for both the
959   exact treatment of the diffusion tensor as well as the rough-shell
960   model for the ellipsoid.
961  
962 < The agreement with the translational diffusion constants from the
963 < microcanonical simulations is quite good, although the rotational
964 < correlation times are as much as a factor of 2 different from the
965 < predictions of the Perrin hydrodynamic model.  
962 > The translational diffusion constants from the microcanonical simulations
963 > agree well with the predictions of the Perrin model, although the rotational
964 > correlation times are a factor of 2 shorter than expected from hydrodynamic
965 > theory.  One explanation for the slower rotation
966 > of explicitly-solvated ellipsoids is the possibility that solute-solvent
967 > collisions happen at both ends of the solute whenever the principal
968 > axis of the ellipsoid is turning. In the upper portion of figure
969 > \ref{fig:explanation} we sketch a physical picture of this explanation.
970 > Since our Langevin integrator is providing nearly quantitative agreement with
971 > the Perrin model, it also predicts orientational diffusion for ellipsoids that
972 > exceed explicitly solvated correlation times by a factor of two.
973  
974 < \subsubsection{Rigid dumbells}
975 <
976 < Perhaps the only composite rigid body for which analytic expressions
977 < for the hydrodynamic tensor are available is the two-sphere dumbell
978 < model.  This model consists of two non-overlapping spheres held by a
979 < rigid bond connecting their centers. There are competing expressions
980 < for the 6x6 resistance tensor for this
981 < model. Equation (\ref{introEquation:oseenTensor}) above gives the original
982 < Oseen tensor, while the second order expression introduced by Rotne
983 < and Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la Torre and
735 < Bloomfield,\cite{Torre1977} is given above as
974 > \subsection{Rigid dumbbells}
975 > Perhaps the only {\it composite} rigid body for which analytic
976 > expressions for the hydrodynamic tensor are available is the
977 > two-sphere dumbbell model.  This model consists of two non-overlapping
978 > spheres held by a rigid bond connecting their centers. There are
979 > competing expressions for the 6x6 resistance tensor for this
980 > model. Equation (\ref{introEquation:oseenTensor}) above gives the
981 > original Oseen tensor, while the second order expression introduced by
982 > Rotne and Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la
983 > Torre and Bloomfield,\cite{Torre1977} is given above as
984   Eq. (\ref{introEquation:RPTensorNonOverlapped}).  In our case, we use
985   a model dumbbell in which the two spheres are identical Lennard-Jones
986   particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
987 < a distance of 6.65 \AA\ ??.  
987 > a distance of 6.532 \AA.
988  
989   The theoretical values for the translational diffusion constant of the
990   dumbbell are calculated from the work of Stimson and Jeffery, who
991   studied the motion of this system in a flow parallel to the
992 < inter-sphere axis,\cite{StimsonJeffery26} and Davis, who studied the
993 < motion in a flow perpendicular to the inter-sphere axis.\cite{Davis69}
992 > inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
993 > motion in a flow {\it perpendicular} to the inter-sphere
994 > axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
995 > orientational} correlation times for this model system (other than
996 > those derived from the 6 x 6 tensors mentioned above).
997  
998 < How did we do?  Does Analytic reproduce MD?  Does LD reproduce
999 < Analytic or MD?
998 > The bead model for this model system comprises the two large spheres
999 > by themselves, while the rough shell approximation used 3368 separate
1000 > beads (each with a diameter of 0.25 \AA) to approximate the shape of
1001 > the rigid body.  The hydrodynamics tensors computed from both the bead
1002 > and rough shell models are remarkably similar.  Computing the initial
1003 > hydrodynamic tensor for a rough shell model can be quite expensive (in
1004 > this case it requires inverting a 10104 x 10104 matrix), while the
1005 > bead model is typically easy to compute (in this case requiring
1006 > inversion of a 6 x 6 matrix).  
1007  
1008 < \subsubsection{Ellipsoidal-composite banana-shaped molecules}
1008 > \begin{figure}
1009 > \centering
1010 > \includegraphics[width=2in]{RoughShell}
1011 > \caption[Model rigid bodies and their rough shell approximations]{The
1012 > model rigid bodies (left column) used to test this algorithm and their
1013 > rough-shell approximations (right-column) that were used to compute
1014 > the hydrodynamic tensors.  The top two models (ellipsoid and dumbbell)
1015 > have analytic solutions and were used to test the rough shell
1016 > approximation.  The lower two models (banana and lipid) were compared
1017 > with explicitly-solvated molecular dynamics simulations. }
1018 > \label{fig:roughShell}
1019 > \end{figure}
1020  
1021 < Banana-shaped rigid bodies composed of composites of Gay-Berne
1022 < ellipsoids have been used by Orlandi {\it et al.} to observe
1023 < mesophases in coarse-grained models bent-core liquid crystalline
1024 < molecules.\cite{OrlandiZannoni06}  We have used the overlapping
1021 >
1022 > Once the hydrodynamic tensor has been computed, there is no additional
1023 > penalty for carrying out a Langevin simulation with either of the two
1024 > different hydrodynamics models.  Our naive expectation is that since
1025 > the rigid body's surface is roughened under the various shell models,
1026 > the diffusion constants will be even farther from the ``slip''
1027 > boundary conditions than observed for the bead model (which uses a
1028 > Stokes-Einstein model to arrive at the hydrodynamic tensor).  For the
1029 > dumbbell, this prediction is correct although all of the Langevin
1030 > diffusion constants are within 6\% of the diffusion constant predicted
1031 > from the fully solvated system.
1032 >
1033 > For rotational motion, Langevin integration (and the hydrodynamic tensor)
1034 > yields rotational correlation times that are substantially shorter than those
1035 > obtained from explicitly-solvated simulations.  It is likely that this is due
1036 > to the large size of the explicit solvent spheres, a feature that prevents
1037 > the solvent from coming in contact with a substantial fraction of the surface
1038 > area of the dumbbell.  Therefore, the explicit solvent only provides drag
1039 > over a substantially reduced surface area of this model, while the
1040 > hydrodynamic theories utilize the entire surface area for estimating
1041 > rotational diffusion.  A sketch of the free volume available in the explicit
1042 > solvent simulations is shown in figure \ref{fig:explanation}.
1043 >
1044 >
1045 > \begin{figure}
1046 > \centering
1047 > \includegraphics[width=6in]{explanation}
1048 > \caption[Explanations of the differences between orientational
1049 > correlation times for explicitly-solvated models and hydrodynamics
1050 > predictions]{Explanations of the differences between orientational
1051 > correlation times for explicitly-solvated models and hydrodynamic
1052 > predictions.   For the ellipsoids (upper figures), rotation of the
1053 > principal axis can involve correlated collisions at both sides of the
1054 > solute.  In the rigid dumbbell model (lower figures), the large size
1055 > of the explicit solvent spheres prevents them from coming in contact
1056 > with a substantial fraction of the surface area of the dumbbell.
1057 > Therefore, the explicit solvent only provides drag over a
1058 > substantially reduced surface area of this model, where the
1059 > hydrodynamic theories utilize the entire surface area for estimating
1060 > rotational diffusion.
1061 > } \label{fig:explanation}
1062 > \end{figure}
1063 >
1064 >
1065 >
1066 > \subsection{Composite banana-shaped molecules}
1067 > Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids have
1068 > been used by Orlandi {\it et al.} to observe mesophases in
1069 > coarse-grained models for bent-core liquid crystalline
1070 > molecules.\cite{Orlandi:2006fk} We have used the same overlapping
1071   ellipsoids as a way to test the behavior of our algorithm for a
1072   structure of some interest to the materials science community,
1073   although since we are interested in capturing only the hydrodynamic
1074 < behavior of this model, we leave out the dipolar interactions of the
1075 < original Orlandi model.
1076 <
1077 < \subsubsection{Composite sphero-ellipsoids}
1074 > behavior of this model, we have left out the dipolar interactions of
1075 > the original Orlandi model.
1076 >
1077 > A reference system composed of a single banana rigid body embedded in a
1078 > sea of 1929 solvent particles was created and run under standard
1079 > (microcanonical) molecular dynamics.  The resulting viscosity of this
1080 > mixture was 0.298 centipoise (as estimated using Eq. (\ref{eq:shear})).
1081 > To calculate the hydrodynamic properties of the banana rigid body model,
1082 > we created a rough shell (see Fig.~\ref{fig:roughShell}), in which
1083 > the banana is represented as a ``shell'' made of 3321 identical beads
1084 > (0.25 \AA\  in diameter) distributed on the surface.  Applying the
1085 > procedure described in Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1086 > identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0 \AA), as
1087 > well as the resistance tensor,
1088 > \begin{equation*}
1089 > \Xi =
1090 > \left( {\begin{array}{*{20}c}
1091 > 0.9261 & 0 & 0&0&0.08585&0.2057\\
1092 > 0& 0.9270&-0.007063& 0.08585&0&0\\
1093 > 0&-0.007063&0.7494&0.2057&0&0\\
1094 > 0&0.0858&0.2057& 58.64& 0&0\\0.08585&0&0&0&48.30&3.219&\\0.2057&0&0&0&3.219&10.7373\\\end{array}} \right),
1095 > \end{equation*}
1096 > where the units for translational, translation-rotation coupling and
1097 > rotational tensors are (kcal fs / mol \AA$^2$), (kcal fs / mol \AA\ rad),
1098 > and (kcal fs / mol rad$^2$), respectively.
1099  
1100 + The Langevin rigid-body integrator (and the hydrodynamic diffusion tensor)
1101 + are essentially quantitative for translational diffusion of this model.  
1102 + Orientational correlation times under the Langevin rigid-body integrator
1103 + are within 11\% of the values obtained from explicit solvent, but these
1104 + models also exhibit some solvent inaccessible surface area in the
1105 + explicitly-solvated case.  
1106 +
1107 + \subsection{Composite sphero-ellipsoids}
1108   Spherical heads perched on the ends of Gay-Berne ellipsoids have been
1109 < used recently as models for lipid molecules.\cite{SunGezelter08,AytonVoth??}
1109 > used recently as models for lipid
1110 > molecules.\cite{SunGezelter08,Ayton01}
1111 > MORE DETAILS
1112  
1113 + A reference system composed of a single lipid rigid body embedded in a
1114 + sea of 1929 solvent particles was created and run under standard
1115 + (microcanonical) molecular dynamics.  The resulting viscosity of this
1116 + mixture was 0.349 centipoise (as estimated using
1117 + Eq. (\ref{eq:shear})).  To calculate the hydrodynamic properties of
1118 + the lipid rigid body model, we created a rough shell (see
1119 + Fig.~\ref{fig:roughShell}), in which the lipid is represented as a
1120 + ``shell'' made of 3550 identical beads (0.25 \AA\ in diameter)
1121 + distributed on the surface.  Applying the procedure described in
1122 + Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1123 + identified the center of resistance, ${\bf r} = $(0 \AA, 0 \AA, 1.46
1124 + \AA).
1125  
768 We performed several NVE
769 simulations with explicit solvents for different shaped
770 molecules. There are one solute molecule and 1929 solvent molecules in
771 NVE simulation. The parameters are shown in table
772 \ref{tab:parameters}. The force field between spheres is standard
773 Lennard-Jones, and ellipsoids interact with other ellipsoids and
774 spheres with generalized Gay-Berne potential. All simulations are
775 carried out at 300 K and 1 Atm. The time step is 25 ns, and a
776 switching function was applied to all potentials to smoothly turn off
777 the interactions between a range of $22$ and $25$ \AA.  The switching
778 function was the standard (cubic) function,
779 \begin{equation}
780 s(r) =
781        \begin{cases}
782        1 & \text{if $r \le r_{\text{sw}}$},\\
783        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
784        {(r_{\text{cut}} - r_{\text{sw}})^3}
785        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
786        0 & \text{if $r > r_{\text{cut}}$.}
787        \end{cases}
788 \label{eq:switchingFunc}
789 \end{equation}
790 We have computed translational diffusion constants for lipid molecules
791 from the mean-square displacement,
792 \begin{equation}
793 D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
794 \end{equation}
795 of the solute molecules. Translational diffusion constants for the
796 different shaped molecules are shown in table
797 \ref{tab:translation}.  We have also computed orientational correlation
798 times for different shaped molecules from fits of the second-order
799 Legendre polynomial correlation function,
800 \begin{equation}
801 C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
802 \mu}_{i}(0) \right)
803 \end{equation}
804 the results are shown in table \ref{tab:rotation}. We used einstein
805 format of the pressure correlation function,
806 \begin{equation}
807 C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
808 \mu}_{i}(0) \right)
809 \end{equation}
810 to estimate the viscosity of the systems from NVE simulations. The
811 viscosity can also be calculated by Green-Kubo pressure correlaton
812 function,
813 \begin{equation}
814 C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
815 \mu}_{i}(0) \right)
816 \end{equation}
817 However, this method converges slowly, and the statistics are not good
818 enough to give us a very accurate value. The langevin dynamics
819 simulations for different shaped molecules are performed at the same
820 conditions as the NVE simulations with viscosity estimated from NVE
821 simulations. To get better statistics, 1024 non-interacting solute
822 molecules are put into one simulation box for each langevin
823 simulation, this is equal to 1024 simulations for single solute
824 systems. The diffusion constants and rotation relaxation times for
825 different shaped molecules are shown in table \ref{tab:translation}
826 and \ref{tab:rotation} to compare to the results calculated from NVE
827 simulations. The theoretical values for sphere is calculated from the
828 Stokes-Einstein law, the theoretical values for ellipsoid is
829 calculated from Perrin's fomula,  The exact method is
830 applied to the langevin dynamics simulations for sphere and ellipsoid,
831 the bead model is applied to the simulation for dumbbell molecule, and
832 the rough shell model is applied to ellipsoid, dumbbell, banana and
833 lipid molecules. The results from all the langevin dynamics
834 simulations, including exact, bead model and rough shell, match the
835 theoretical values perfectly for all different shaped molecules. This
836 indicates that our simulation package for langevin dynamics is working
837 well. The approxiate methods ( bead model and rough shell model) are
838 accurate enough for the current simulations. The goal of the langevin
839 dynamics theory is to replace the explicit solvents by the friction
840 forces. We compared the dynamic properties of different shaped
841 molecules in langevin dynamics simulations with that in NVE
842 simulations. The results are reasonable close. Overall, the
843 translational diffusion constants calculated from langevin dynamics
844 simulations are very close to the values from the NVE simulation. For
845 sphere and lipid molecules, the diffusion constants are a little bit
846 off from the NVE simulation results. One possible reason is that the
847 calculation of the viscosity is very difficult to be accurate. Another
848 possible reason is that although we save very frequently during the
849 NVE simulations and run pretty long time simulations, there is only
850 one solute molecule in the system which makes the calculation for the
851 diffusion constant difficult. The sphere molecule behaves as a free
852 rotor in the solvent, so there is no rotation relaxation time
853 calculated from NVE simulations. The rotation relaxation time is not
854 very close to the NVE simulations results. The banana and lipid
855 molecules match the NVE simulations results pretty well. The mismatch
856 between langevin dynamics and NVE simulation for ellipsoid is possibly
857 caused by the slip boundary condition. For dumbbell, the mismatch is
858 caused by the size of the solvent molecule is pretty large compared to
859 dumbbell molecule in NVE simulations.
1126  
1127 + \subsection{Summary}
1128   According to our simulations, the langevin dynamics is a reliable
1129   theory to apply to replace the explicit solvents, especially for the
1130   translation properties. For large molecules, the rotation properties
1131   are also mimiced reasonablly well.
1132  
1133 < \begin{table*}
1134 < \begin{minipage}{\linewidth}
1135 < \begin{center}
1136 < \caption{}
1137 < \begin{tabular}{llccccccc}
1138 < \hline
1139 <  & & Sphere & Ellipsoid & Dumbbell(2 spheres) & Banana(3 ellpsoids) &
1140 < Lipid(head) & lipid(tail) & Solvent \\
1141 < \hline
1142 < $d$ (\AA) & & 6.5 & 4.6  & 6.5 &  4.2 & 6.5 & 4.6 & 4.7 \\
1143 < $l$ (\AA) & & $= d$ & 13.8 & $=d$ & 11.2 & $=d$ & 13.8 & 4.7 \\
1144 < $\epsilon^s$ (kcal/mol) & & 0.8 & 0.8 & 0.8 & 0.8 & 0.185 & 0.8 & 0.8 \\
1145 < $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 & 0.2 & 1 & 0.2 & 1 \\
1146 < $m$ (amu) & & 190 & 200 & 190 & 240 & 196 & 760 & 72.06 \\
1147 < %$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
1148 < %\multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
1149 < %\multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
883 < %\multicolumn{2}c{$I_{zz}$} &  0 &    9000 & N/A \\
884 < %$\mu$ (Debye) & & varied & 0 & 0 \\
885 < \end{tabular}
886 < \label{tab:parameters}
887 < \end{center}
888 < \end{minipage}
889 < \end{table*}
1133 > \begin{figure}
1134 > \centering
1135 > \includegraphics[width=\linewidth]{graph}
1136 > \caption[Mean squared displacements and orientational
1137 > correlation functions for each of the model rigid bodies.]{The
1138 > mean-squared displacements ($\langle r^2(t) \rangle$) and
1139 > orientational correlation functions ($C_2(t)$) for each of the model
1140 > rigid bodies studied.  The circles are the results for microcanonical
1141 > simulations with explicit solvent molecules, while the other data sets
1142 > are results for Langevin dynamics using the different hydrodynamic
1143 > tensor approximations.  The Perrin model for the ellipsoids is
1144 > considered the ``exact'' hydrodynamic behavior (this can also be said
1145 > for the translational motion of the dumbbell operating under the bead
1146 > model). In most cases, the various hydrodynamics models reproduce
1147 > each other quantitatively.}
1148 > \label{fig:results}
1149 > \end{figure}
1150  
1151   \begin{table*}
1152   \begin{minipage}{\linewidth}
1153   \begin{center}
1154 < \caption{}
1155 < \begin{tabular}{lccccc}
1154 > \caption{Translational diffusion constants (D) for the model systems
1155 > calculated using microcanonical simulations (with explicit solvent),
1156 > theoretical predictions, and Langevin simulations (with implicit solvent).
1157 > Analytical solutions for the exactly-solved hydrodynamics models are
1158 > from Refs. \citen{Einstein05} (sphere), \citen{Perrin1934} and \citen{Perrin1936}
1159 > (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1160 > (dumbbell). The other model systems have no known analytic solution.
1161 > All  diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1162 > $10^{-4}$ \AA$^2$  / fs). }
1163 > \begin{tabular}{lccccccc}
1164   \hline
1165 < & & & & &Translation \\
1166 < \hline
1167 < & NVE &  & Theoretical & Langevin & \\
900 < \hline
901 < & $\eta$ & D & D & method & D \\
1165 > & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1166 > \cline{2-3} \cline{5-7}
1167 > model & $\eta$ (centipoise)  & D & & Analytical & method & Hydrodynamics & simulation \\
1168   \hline
1169 < sphere & 3.480159e-03 & 1.643135e-04 & 1.942779e-04 & exact & 1.982283e-04 \\
1170 < ellipsoid & 2.551262e-03 & 2.437492e-04 & 2.335756e-04 & exact & 2.374905e-04 \\
1171 < & 2.551262e-03  & 2.437492e-04 & 2.335756e-04 & rough shell & 2.284088e-04 \\
1172 < dumbell & 2.41276e-03  & 2.129432e-04 & 2.090239e-04 & bead model & 2.148098e-04 \\
1173 < & 2.41276e-03 & 2.129432e-04 & 2.090239e-04 & rough shell & 2.013219e-04 \\
1174 < banana & 2.9846e-03 & 1.527819e-04 &  & rough shell & 1.54807e-04 \\
1175 < lipid & 3.488661e-03 & 0.9562979e-04 &  & rough shell & 1.320987e-04 \\
1169 > sphere    & 0.279  & 3.06 & & 2.42 & exact       & 2.42 & 2.33 \\
1170 > ellipsoid & 0.255  & 2.44 & & 2.34 & exact       & 2.34 & 2.37 \\
1171 >          & 0.255  & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1172 > dumbbell  & 0.308  & 2.06 & & 1.64 & bead model  & 1.65 & 1.62 \\
1173 >          & 0.308  & 2.06 & & 1.64 & rough shell & 1.59 & 1.62 \\
1174 > banana    & 0.298  & 1.53 & &      & rough shell & 1.56 & 1.55 \\
1175 > lipid     & 0.349  & 0.96 & &      & rough shell & 1.33 & 1.32 \\
1176   \end{tabular}
1177   \label{tab:translation}
1178   \end{center}
# Line 916 | Line 1182 | lipid & 3.488661e-03 & 0.9562979e-04 &  & rough shell
1182   \begin{table*}
1183   \begin{minipage}{\linewidth}
1184   \begin{center}
1185 < \caption{}
1186 < \begin{tabular}{lccccc}
1185 > \caption{Orientational relaxation times ($\tau$) for the model systems using
1186 > microcanonical simulation (with explicit solvent), theoretical
1187 > predictions, and Langevin simulations (with implicit solvent). All
1188 > relaxation times are for the rotational correlation function with
1189 > $\ell = 2$ and are reported in units of ps.  The ellipsoidal model has
1190 > an exact solution for the orientational correlation time due to
1191 > Perrin, but the other model systems have no known analytic solution.}
1192 > \begin{tabular}{lccccccc}
1193   \hline
1194 < & & & & &Rotation \\
1195 < \hline
1196 < & NVE &  & Theoretical & Langevin & \\
925 < \hline
926 < & $\eta$ & $\tau_0$ & $\tau_0$ & method & $\tau_0$ \\
1194 > & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1195 > \cline{2-3} \cline{5-7}
1196 > model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic  & simulation \\
1197   \hline
1198 < sphere & 3.480159e-03 &  & 1.208178e+04 & exact & 1.20628e+04 \\
1199 < ellipsoid & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & exact & 2.21507e+04 \\
1200 < & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & rough shell & 2.21714e+04 \\
1201 < dumbell & 2.41276e-03 & 1.42974e+04 &  & bead model & 7.12435e+04 \\
1202 < & 2.41276e-03 & 1.42974e+04 &  & rough shell & 7.04765e+04 \\
1203 < banana & 2.9846e-03 & 6.38323e+04 &  & rough shell & 7.0945e+04 \\
1204 < lipid & 3.488661e-03 & 7.79595e+04 &  & rough shell & 7.78886e+04 \\
1198 > sphere    & 0.279  &      & & 9.69 & exact       & 9.69 & 9.64 \\
1199 > ellipsoid & 0.255  & 46.7 & & 22.0 & exact       & 22.0 & 22.2 \\
1200 >          & 0.255  & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1201 > dumbbell  & 0.308  & 14.1 & &      & bead model  & 50.0 & 50.1 \\
1202 >          & 0.308  & 14.1 & &      & rough shell & 41.5 & 41.3 \\
1203 > banana    & 0.298  & 63.8 & &      & rough shell & 70.9 & 70.9 \\
1204 > lipid     & 0.349  & 78.0 & &      & rough shell & 76.9 & 77.9 \\
1205 > \hline
1206   \end{tabular}
1207   \label{tab:rotation}
1208   \end{center}
1209   \end{minipage}
1210   \end{table*}
1211  
1212 < Langevin dynamics simulations are applied to study the formation of
1213 < the ripple phase of lipid membranes. The initial configuration is
1212 > \section{Application: A rigid-body lipid bilayer}
1213 >
1214 > The Langevin dynamics integrator was applied to study the formation of
1215 > corrugated structures emerging from simulations of the coarse grained
1216 > lipid molecular models presented above.  The initial configuration is
1217   taken from our molecular dynamics studies on lipid bilayers with
1218 < lennard-Jones sphere solvents. The solvent molecules are excluded from
1219 < the system, the experimental value of water viscosity is applied to
1220 < mimic the heat bath. Fig. XXX is the snapshot of the stable
1221 < configuration of the system, the ripple structure stayed stable after
1222 < 100 ns run. The efficiency of the simulation is increased by one order
1218 > lennard-Jones sphere solvents. The solvent molecules were excluded
1219 > from the system, and the experimental value for the viscosity of water
1220 > at 20C ($\eta = 1.00$ cp) was used to mimic the hydrodynamic effects
1221 > of the solvent.  The absence of explicit solvent molecules and the
1222 > stability of the integrator allowed us to take timesteps of 50 fs.  A
1223 > total simulation run time of 100 ns was sampled.
1224 > Fig. \ref{fig:bilayer} shows the configuration of the system after 100
1225 > ns, and the ripple structure remains stable during the entire
1226 > trajectory.  Compared with using explicit bead-model solvent
1227 > molecules, the efficiency of the simulation has increased by an order
1228   of magnitude.
1229  
951 \subsection{Langevin Dynamics of Banana Shaped Molecules}
952
953 In order to verify that Langevin dynamics can mimic the dynamics of
954 the systems absent of explicit solvents, we carried out two sets of
955 simulations and compare their dynamic properties.
956 Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
957 made of 256 pentane molecules and two banana shaped molecules at
958 273~K. It has an equivalent implicit solvent system containing only
959 two banana shaped molecules with viscosity of 0.289 center poise. To
960 calculate the hydrodynamic properties of the banana shaped molecule,
961 we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
962 in which the banana shaped molecule is represented as a ``shell''
963 made of 2266 small identical beads with size of 0.3 \AA on the
964 surface. Applying the procedure described in
965 Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
966 identified the center of resistance at (0 $\rm{\AA}$, 0.7482 $\rm{\AA}$,
967 -0.1988 $\rm{\AA}$), as well as the resistance tensor,
968 \[
969 \left( {\begin{array}{*{20}c}
970 0.9261 & 0 & 0&0&0.08585&0.2057\\
971 0& 0.9270&-0.007063& 0.08585&0&0\\
972 0&-0.007063&0.7494&0.2057&0&0\\
973 0&0.0858&0.2057& 58.64& 0&0\\
974 0.08585&0&0&0&48.30&3.219&\\
975 0.2057&0&0&0&3.219&10.7373\\
976 \end{array}} \right).
977 \]
978 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.
979 Curves of the velocity auto-correlation functions in
980 Fig.~\ref{langevin:vacf} were shown to match each other very well.
981 However, because of the stochastic nature, simulation using Langevin
982 dynamics was shown to decay slightly faster than MD. In order to
983 study the rotational motion of the molecules, we also calculated the
984 auto-correlation function of the principle axis of the second GB
985 particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
986 probably due to the reason that we used the experimental viscosity directly instead of calculating bulk viscosity from simulation.
987
1230   \begin{figure}
1231   \centering
1232 < \includegraphics[width=\linewidth]{roughShell.pdf}
1233 < \caption[Rough shell model for banana shaped molecule]{Rough shell
1234 < model for banana shaped molecule.} \label{langevin:roughShell}
1232 > \includegraphics[width=\linewidth]{bilayer}
1233 > \caption[Snapshot of a bilayer of rigid-body models for lipids]{A
1234 > snapshot of a bilayer composed of rigid-body models for lipid
1235 > molecules evolving using the Langevin integrator described in this
1236 > work.} \label{fig:bilayer}
1237   \end{figure}
1238  
995 \begin{figure}
996 \centering
997 \includegraphics[width=\linewidth]{twoBanana.pdf}
998 \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
999 256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
1000 molecules and 256 pentane molecules.} \label{langevin:twoBanana}
1001 \end{figure}
1002
1003 \begin{figure}
1004 \centering
1005 \includegraphics[width=\linewidth]{vacf.pdf}
1006 \caption[Plots of Velocity Auto-correlation Functions]{Velocity
1007 auto-correlation functions of NVE (explicit solvent) in blue and
1008 Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
1009 \end{figure}
1010
1011 \begin{figure}
1012 \centering
1013 \includegraphics[width=\linewidth]{uacf.pdf}
1014 \caption[Auto-correlation functions of the principle axis of the
1015 middle GB particle]{Auto-correlation functions of the principle axis
1016 of the middle GB particle of NVE (blue) and Langevin dynamics
1017 (red).} \label{langevin:uacf}
1018 \end{figure}
1019
1239   \section{Conclusions}
1240  
1241   We have presented a new Langevin algorithm by incorporating the
1242   hydrodynamics properties of arbitrary shaped molecules into an
1243 < advanced symplectic integration scheme. The temperature control
1244 < ability of this algorithm was demonstrated by a set of simulations
1245 < with different viscosities. It was also shown to have significant
1246 < advantage of producing rapid thermal equilibration over
1028 < Nos\'{e}-Hoover method. Further studies in systems involving banana
1029 < shaped molecules illustrated that the dynamic properties could be
1030 < preserved by using this new algorithm as an implicit solvent model.
1243 > advanced symplectic integration scheme. Further studies in systems
1244 > involving banana shaped molecules illustrated that the dynamic
1245 > properties could be preserved by using this new algorithm as an
1246 > implicit solvent model.
1247  
1248  
1249   \section{Acknowledgments}
# Line 1037 | Line 1253 | of Notre Dame.
1253   of Notre Dame.
1254   \newpage
1255  
1256 < \bibliographystyle{jcp2}
1256 > \bibliographystyle{jcp}
1257   \bibliography{langevin}
1258  
1259   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines