ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3299
Committed: Mon Jan 7 21:30:15 2008 UTC (16 years, 7 months ago) by gezelter
Content type: application/x-tex
File size: 46342 byte(s)
Log Message:
working on model systems section

File Contents

# User Rev Content
1 tim 2746 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2     %\documentclass[aps,prb,preprint]{revtex4}
3     \documentclass[11pt]{article}
4     \usepackage{endfloat}
5     \usepackage{amsmath,bm}
6     \usepackage{amssymb}
7     \usepackage{times}
8     \usepackage{mathptmx}
9     \usepackage{setspace}
10     \usepackage{tabularx}
11     \usepackage{graphicx}
12     \usepackage{booktabs}
13     \usepackage{bibentry}
14     \usepackage{mathrsfs}
15     \usepackage[ref]{overcite}
16     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18     9.0in \textwidth 6.5in \brokenpenalty=10000
19     \renewcommand{\baselinestretch}{1.2}
20 gezelter 3299 \renewcommand\citemid{\ } % no comma in optional referenc note
21 tim 2746
22     \begin{document}
23    
24 gezelter 3205 \title{An algorithm for performing Langevin dynamics on rigid bodies of arbitrary shape }
25 tim 2746
26 gezelter 3299 \author{Xiuquan Sun, Teng Lin and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
27 tim 2746 gezelter@nd.edu} \\
28     Department of Chemistry and Biochemistry\\
29     University of Notre Dame\\
30     Notre Dame, Indiana 46556}
31    
32     \date{\today}
33    
34     \maketitle \doublespacing
35    
36     \begin{abstract}
37    
38     \end{abstract}
39    
40     \newpage
41    
42     %\narrowtext
43    
44     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45     % BODY OF TEXT
46     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
47    
48     \section{Introduction}
49    
50     %applications of langevin dynamics
51 tim 2999 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
75 tim 2746 instance, the oscillation power spectrum of nanoparticles from
76     Langevin dynamics simulation has the same peak frequencies for
77 tim 2999 different wave vectors, which recovers the property of magnetic
78     excitations in small finite structures.\cite{Berkov2005a}
79 tim 2746
80     %review rigid body dynamics
81     Rigid bodies are frequently involved in the modeling of different
82     areas, from engineering, physics, to chemistry. For example,
83     missiles and vehicle are usually modeled by rigid bodies. The
84     movement of the objects in 3D gaming engine or other physics
85     simulator is governed by the rigid body dynamics. In molecular
86     simulation, rigid body is used to simplify the model in
87     protein-protein docking study{\cite{Gray2003}}.
88    
89     It is very important to develop stable and efficient methods to
90 tim 2999 integrate the equations of motion for orientational degrees of
91     freedom. Euler angles are the natural choice to describe the
92     rotational degrees of freedom. However, due to $\frac {1}{sin
93     \theta}$ singularities, the numerical integration of corresponding
94     equations of these motion is very inefficient and inaccurate.
95     Although an alternative integrator using multiple sets of Euler
96     angles can overcome this difficulty\cite{Barojas1973}, the
97     computational penalty and the loss of angular momentum conservation
98     still remain. A singularity-free representation utilizing
99     quaternions was developed by Evans in 1977.\cite{Evans1977}
100     Unfortunately, this approach used a nonseparable Hamiltonian
101     resulting from the quaternion representation, which prevented the
102     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}
109 tim 2746
110 tim 2999 A break-through in geometric literature suggests that, in order to
111 tim 2746 develop a long-term integration scheme, one should preserve the
112 tim 2999 symplectic structure of the propagator. By introducing a conjugate
113     momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
114     equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
115     proposed to evolve the Hamiltonian system in a constraint manifold
116     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.
122 tim 2746
123     %review langevin/browninan dynamics for arbitrarily shaped rigid body
124     Combining Langevin or Brownian dynamics with rigid body dynamics,
125 tim 2999 one can study slow processes in biomolecular systems. Modeling DNA
126     as a chain of rigid beads, which are subject to harmonic potentials
127     as well as excluded volume potentials, Mielke and his coworkers
128     discovered rapid superhelical stress generations from the stochastic
129     simulation of twin supercoiling DNA with response to induced
130     torques.\cite{Mielke2004} Membrane fusion is another key biological
131     process which controls a variety of physiological functions, such as
132     release of neurotransmitters \textit{etc}. A typical fusion event
133     happens on the time scale of a millisecond, which is impractical to
134     study using atomistic models with newtonian mechanics. With the help
135     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 tim 2746 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 tim 2999 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 tim 2746 relaxation time is much less than the time step, one may ignore the
151 tim 2999 inertia in Brownian dynamics. However, the assumption of zero
152 tim 2746 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 tim 2999 term.\cite{Beard2000} As a complement to IBD which has a lower bound
156 tim 2746 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 tim 2999 regime.\cite{Beard2000} LTID can also deal with the rotational
160 tim 2746 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 tim 2999 typical nonskew bodies like cylinders and ellipsoids are inadequate
164     to represent most complex macromolecule assemblies. These intricate
165 tim 2746 molecules have been represented by a set of beads and their
166 tim 2999 hydrodynamic properties can be calculated using variants on the
167     standard hydrodynamic interaction tensors.
168 tim 2746
169     The goal of the present work is to develop a Langevin dynamics
170 tim 2999 algorithm for arbitrary-shaped rigid particles by integrating the
171     accurate estimation of friction tensor from hydrodynamics theory
172     into the sophisticated rigid body dynamics algorithms.
173 tim 2746
174 tim 2999 \section{Computational Methods{\label{methodSec}}}
175 tim 2746
176 tim 2999 \subsection{\label{introSection:frictionTensor}Friction Tensor}
177     Theoretically, the friction kernel can be determined using the
178     velocity autocorrelation function. However, this approach becomes
179     impractical when the system becomes more and more complicated.
180     Instead, various approaches based on hydrodynamics have been
181     developed to calculate the friction coefficients. In general, the
182     friction tensor $\Xi$ is a $6\times 6$ matrix given by
183     \[
184     \Xi = \left( {\begin{array}{*{20}c}
185     {\Xi _{}^{tt} } & {\Xi _{}^{rt} } \\
186     {\Xi _{}^{tr} } & {\Xi _{}^{rr} } \\
187     \end{array}} \right).
188     \]
189     Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are $3 \times 3$
190     translational friction tensor and rotational resistance (friction)
191     tensor respectively, while ${\Xi^{tr} }$ is translation-rotation
192     coupling tensor and $ {\Xi^{rt} }$ is rotation-translation coupling
193     tensor. When a particle moves in a fluid, it may experience friction
194     force or torque along the opposite direction of the velocity or
195     angular velocity,
196     \[
197 tim 2746 \left( \begin{array}{l}
198 tim 2999 F_R \\
199     \tau _R \\
200 tim 2746 \end{array} \right) = - \left( {\begin{array}{*{20}c}
201 tim 2999 {\Xi ^{tt} } & {\Xi ^{rt} } \\
202     {\Xi ^{tr} } & {\Xi ^{rr} } \\
203 tim 2746 \end{array}} \right)\left( \begin{array}{l}
204 tim 2999 v \\
205     w \\
206 tim 2746 \end{array} \right)
207 tim 2999 \]
208     where $F_r$ is the friction force and $\tau _R$ is the friction
209     torque.
210 tim 2746
211 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
212 tim 2746
213 tim 2999 For a spherical particle with slip boundary conditions, the
214     translational and rotational friction constant can be calculated
215     from Stoke's law,
216 tim 2746 \[
217 tim 2999 \Xi ^{tt} = \left( {\begin{array}{*{20}c}
218     {6\pi \eta R} & 0 & 0 \\
219     0 & {6\pi \eta R} & 0 \\
220     0 & 0 & {6\pi \eta R} \\
221     \end{array}} \right)
222 tim 2746 \]
223 tim 2999 and
224 tim 2746 \[
225 tim 2999 \Xi ^{rr} = \left( {\begin{array}{*{20}c}
226     {8\pi \eta R^3 } & 0 & 0 \\
227     0 & {8\pi \eta R^3 } & 0 \\
228     0 & 0 & {8\pi \eta R^3 } \\
229     \end{array}} \right)
230 tim 2746 \]
231 tim 2999 where $\eta$ is the viscosity of the solvent and $R$ is the
232     hydrodynamic radius.
233    
234     Other non-spherical shapes, such as cylinders and ellipsoids, are
235     widely used as references for developing new hydrodynamics theory,
236     because their properties can be calculated exactly. In 1936, Perrin
237     extended Stokes's law to general ellipsoids, also called a triaxial
238     ellipsoid, which is given in Cartesian coordinates
239     by\cite{Perrin1934, Perrin1936}
240 tim 2746 \[
241 tim 2999 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
242     }} = 1
243 tim 2746 \]
244 tim 2999 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
245     due to the complexity of the elliptic integral, only the ellipsoid
246     with the restriction of two axes being equal, \textit{i.e.}
247     prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
248     exactly. Introducing an elliptic integral parameter $S$ for prolate
249     ellipsoids :
250     \[
251     S = \frac{2}{{\sqrt {a^2 - b^2 } }}\ln \frac{{a + \sqrt {a^2 - b^2
252     } }}{b},
253     \]
254     and oblate ellipsoids:
255     \[
256     S = \frac{2}{{\sqrt {b^2 - a^2 } }}arctg\frac{{\sqrt {b^2 - a^2 }
257     }}{a},
258     \]
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}}.
270     \end{eqnarray*}
271 tim 2746
272 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
273    
274     Unlike spherical and other simply shaped molecules, there is no
275     analytical solution for the friction tensor for arbitrarily shaped
276     rigid molecules. The ellipsoid of revolution model and general
277     triaxial ellipsoid model have been used to approximate the
278     hydrodynamic properties of rigid bodies. However, since the mapping
279     from all possible ellipsoidal spaces, $r$-space, to all possible
280     combination of rotational diffusion coefficients, $D$-space, is not
281     unique\cite{Wegener1979} as well as the intrinsic coupling between
282     translational and rotational motion of rigid bodies, general
283     ellipsoids are not always suitable for modeling arbitrarily shaped
284     rigid molecules. A number of studies have been devoted to
285     determining the friction tensor for irregularly shaped rigid bodies
286     using more advanced methods where the molecule of interest was
287     modeled by a combinations of spheres\cite{Carrasco1999} and the
288     hydrodynamics properties of the molecule can be calculated using the
289     hydrodynamic interaction tensor. Let us consider a rigid assembly of
290     $N$ beads immersed in a continuous medium. Due to hydrodynamic
291     interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
292     than its unperturbed velocity $v_i$,
293 tim 2746 \[
294 tim 2999 v'_i = v_i - \sum\limits_{j \ne i} {T_{ij} F_j }
295 tim 2746 \]
296 tim 2999 where $F_i$ is the frictional force, and $T_{ij}$ is the
297     hydrodynamic interaction tensor. The friction force of $i$th bead is
298     proportional to its ``net'' velocity
299 tim 2746 \begin{equation}
300 tim 2999 F_i = \zeta _i v_i - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
301     \label{introEquation:tensorExpression}
302 tim 2746 \end{equation}
303 tim 2999 This equation is the basis for deriving the hydrodynamic tensor. In
304     1930, Oseen and Burgers gave a simple solution to
305     Eq.~\ref{introEquation:tensorExpression}
306 tim 2746 \begin{equation}
307 tim 2999 T_{ij} = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
308     R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
309 tim 2746 \end{equation}
310 tim 2999 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
311     A second order expression for element of different size was
312     introduced by Rotne and Prager\cite{Rotne1969} and improved by
313     Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
314 tim 2746 \begin{equation}
315 tim 2999 T_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
316     \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
317     _i^2 + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
318     \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
319     \label{introEquation:RPTensorNonOverlapped}
320 tim 2746 \end{equation}
321 tim 2999 Both of the Eq.~\ref{introEquation:oseenTensor} and
322     Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption
323     $R_{ij} \ge \sigma _i + \sigma _j$. An alternative expression for
324     overlapping beads with the same radius, $\sigma$, is given by
325 tim 2746 \begin{equation}
326 tim 2999 T_{ij} = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
327     \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
328     \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
329     \label{introEquation:RPTensorOverlapped}
330 tim 2746 \end{equation}
331 tim 2999 To calculate the resistance tensor at an arbitrary origin $O$, we
332     construct a $3N \times 3N$ matrix consisting of $N \times N$
333     $B_{ij}$ blocks
334     \begin{equation}
335     B = \left( {\begin{array}{*{20}c}
336     {B_{11} } & \ldots & {B_{1N} } \\
337     \vdots & \ddots & \vdots \\
338     {B_{N1} } & \cdots & {B_{NN} } \\
339     \end{array}} \right),
340     \end{equation}
341     where $B_{ij}$ is given by
342 tim 2746 \[
343 tim 2999 B_{ij} = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
344     )T_{ij}
345 tim 2746 \]
346 tim 2999 where $\delta _{ij}$ is the Kronecker delta function. Inverting the
347     $B$ matrix, we obtain
348 tim 2746 \[
349 tim 2999 C = B^{ - 1} = \left( {\begin{array}{*{20}c}
350     {C_{11} } & \ldots & {C_{1N} } \\
351     \vdots & \ddots & \vdots \\
352     {C_{N1} } & \cdots & {C_{NN} } \\
353     \end{array}} \right),
354 tim 2746 \]
355 tim 2999 which can be partitioned into $N \times N$ $3 \times 3$ block
356     $C_{ij}$. With the help of $C_{ij}$ and the skew matrix $U_i$
357 tim 2746 \[
358 tim 2999 U_i = \left( {\begin{array}{*{20}c}
359     0 & { - z_i } & {y_i } \\
360     {z_i } & 0 & { - x_i } \\
361     { - y_i } & {x_i } & 0 \\
362     \end{array}} \right)
363 tim 2746 \]
364 tim 2999 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
365     bead $i$ and origin $O$, the elements of resistance tensor at
366     arbitrary origin $O$ can be written as
367     \begin{eqnarray}
368     \Xi _{}^{tt} & = & \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
369     \Xi _{}^{tr} & = & \Xi _{}^{rt} = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
370     \Xi _{}^{rr} & = & - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
371     \label{introEquation:ResistanceTensorArbitraryOrigin}
372     \end{eqnarray}
373     The resistance tensor depends on the origin to which they refer. The
374     proper location for applying the friction force is the center of
375     resistance (or center of reaction), at which the trace of rotational
376     resistance tensor, $ \Xi ^{rr}$ reaches a minimum value.
377     Mathematically, the center of resistance is defined as an unique
378     point of the rigid body at which the translation-rotation coupling
379     tensors are symmetric,
380     \begin{equation}
381     \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
382     \label{introEquation:definitionCR}
383     \end{equation}
384     From Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
385     we can easily derive that the translational resistance tensor is
386     origin independent, while the rotational resistance tensor and
387     translation-rotation coupling resistance tensor depend on the
388     origin. Given the resistance tensor at an arbitrary origin $O$, and
389     a vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
390     obtain the resistance tensor at $P$ by
391     \begin{equation}
392     \begin{array}{l}
393     \Xi _P^{tt} = \Xi _O^{tt} \\
394     \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
395     \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{{tr} ^{^T }} \\
396     \end{array}
397     \label{introEquation:resistanceTensorTransformation}
398     \end{equation}
399     where
400 tim 2746 \[
401 tim 2999 U_{OP} = \left( {\begin{array}{*{20}c}
402     0 & { - z_{OP} } & {y_{OP} } \\
403     {z_i } & 0 & { - x_{OP} } \\
404     { - y_{OP} } & {x_{OP} } & 0 \\
405     \end{array}} \right)
406 tim 2746 \]
407 tim 2999 Using Eq.~\ref{introEquation:definitionCR} and
408     Eq.~\ref{introEquation:resistanceTensorTransformation}, one can
409     locate the position of center of resistance,
410     \begin{eqnarray*}
411     \left( \begin{array}{l}
412     x_{OR} \\
413     y_{OR} \\
414     z_{OR} \\
415     \end{array} \right) & = &\left( {\begin{array}{*{20}c}
416     {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
417     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
418     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
419     \end{array}} \right)^{ - 1} \\
420     & & \left( \begin{array}{l}
421     (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
422     (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
423     (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
424     \end{array} \right) \\
425     \end{eqnarray*}
426     where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
427     joining center of resistance $R$ and origin $O$.
428 tim 2746
429 tim 2999 \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
430 tim 2746
431 tim 2999 Consider the Langevin equations of motion in generalized coordinates
432 tim 2746 \begin{equation}
433     M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)} + F_{r,i} (t)
434     \label{LDGeneralizedForm}
435     \end{equation}
436     where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
437     and moment of inertial) matrix and $V_i$ is a generalized velocity,
438 tim 2999 $V_i = V_i(v_i,\omega _i)$. The right side of
439     Eq.~\ref{LDGeneralizedForm} consists of three generalized forces in
440 tim 2746 lab-fixed frame, systematic force $F_{s,i}$, dissipative force
441     $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
442     system in Newtownian mechanics typically refers to lab-fixed frame,
443     it is also convenient to handle the rotation of rigid body in
444     body-fixed frame. Thus the friction and random forces are calculated
445     in body-fixed frame and converted back to lab-fixed frame by:
446     \[
447     \begin{array}{l}
448 tim 2999 F_{f,i}^l (t) = Q^T F_{f,i}^b (t), \\
449     F_{r,i}^l (t) = Q^T F_{r,i}^b (t). \\
450     \end{array}
451 tim 2746 \]
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 tim 2999 angular velocity $\omega _i$
455 tim 2746 \begin{equation}
456     F_{r,i}^b (t) = \left( \begin{array}{l}
457     f_{r,i}^b (t) \\
458     \tau _{r,i}^b (t) \\
459     \end{array} \right) = - \left( {\begin{array}{*{20}c}
460     {\Xi _{R,t} } & {\Xi _{R,c}^T } \\
461     {\Xi _{R,c} } & {\Xi _{R,r} } \\
462     \end{array}} \right)\left( \begin{array}{l}
463     v_{R,i}^b (t) \\
464     \omega _i (t) \\
465     \end{array} \right),
466     \end{equation}
467     while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
468     with zero mean and variance
469     \begin{equation}
470     \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle =
471     \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle =
472 tim 2999 2k_B T\Xi _R \delta (t - t'). \label{randomForce}
473 tim 2746 \end{equation}
474     The equation of motion for $v_i$ can be written as
475     \begin{equation}
476     m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
477     f_{r,i}^l (t)
478     \end{equation}
479     Since the frictional force is applied at the center of resistance
480     which generally does not coincide with the center of mass, an extra
481     torque is exerted at the center of mass. Thus, the net body-fixed
482     frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
483     given by
484     \begin{equation}
485     \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
486     \end{equation}
487     where $r_{MR}$ is the vector from the center of mass to the center
488 tim 2999 of the resistance. Instead of integrating the angular velocity in
489     lab-fixed frame, we consider the equation of angular momentum in
490     body-fixed frame
491 tim 2746 \begin{equation}
492 tim 2999 \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
493     + \tau _{r,i}^b(t)
494 tim 2746 \end{equation}
495     Embedding the friction terms into force and torque, one can
496     integrate the langevin equations of motion for rigid body of
497     arbitrary shape in a velocity-Verlet style 2-part algorithm, where
498     $h= \delta t$:
499    
500 tim 2999 {\tt moveA:}
501 tim 2746 \begin{align*}
502 tim 2999 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
503     + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
504     %
505     {\bf r}(t + h) &\leftarrow {\bf r}(t)
506     + h {\bf v}\left(t + h / 2 \right), \\
507     %
508     {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
509     + \frac{h}{2} {\bf \tau}^b(t), \\
510     %
511     \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
512     (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
513 tim 2746 \end{align*}
514     In this context, the $\mathrm{rotate}$ function is the reversible
515 tim 2999 product of the three body-fixed rotations,
516 tim 2746 \begin{equation}
517     \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
518     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
519     / 2) \cdot \mathsf{G}_x(a_x /2),
520     \end{equation}
521     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
522 tim 2999 rotates both the rotation matrix ($\mathsf{Q}$) and the body-fixed
523     angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
524     axis $\alpha$,
525 tim 2746 \begin{equation}
526     \mathsf{G}_\alpha( \theta ) = \left\{
527     \begin{array}{lcl}
528 tim 2999 \mathsf{Q}(t) & \leftarrow & \mathsf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
529 tim 2746 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
530     j}(0).
531     \end{array}
532     \right.
533     \end{equation}
534     $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
535     rotation matrix. For example, in the small-angle limit, the
536     rotation matrix around the body-fixed x-axis can be approximated as
537     \begin{equation}
538     \mathsf{R}_x(\theta) \approx \left(
539     \begin{array}{ccc}
540     1 & 0 & 0 \\
541     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
542     \theta^2 / 4} \\
543     0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
544     \theta^2 / 4}
545     \end{array}
546     \right).
547     \end{equation}
548 tim 2999 All other rotations follow in a straightforward manner. After the
549     first part of the propagation, the forces and body-fixed torques are
550     calculated at the new positions and orientations
551 tim 2746
552 tim 2999 {\tt doForces:}
553     \begin{align*}
554     {\bf f}(t + h) &\leftarrow
555     - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
556     %
557     {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
558     \times \frac{\partial V}{\partial {\bf u}}, \\
559     %
560     {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{Q}(t + h)
561     \cdot {\bf \tau}^s(t + h).
562     \end{align*}
563 tim 2746 Once the forces and torques have been obtained at the new time step,
564     the velocities can be advanced to the same time value.
565    
566 tim 2999 {\tt moveB:}
567 tim 2746 \begin{align*}
568 tim 2999 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
569     \right)
570     + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
571     %
572     {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
573     \right)
574     + \frac{h}{2} {\bf \tau}^b(t + h) .
575 tim 2746 \end{align*}
576    
577 tim 2999 \section{Results and Discussion}
578 tim 2746
579 tim 2999 The Langevin algorithm described in previous section has been
580     implemented in {\sc oopse}\cite{Meineke2005} and applied to studies
581     of the static and dynamic properties in several systems.
582 tim 2746
583 tim 2999 \subsection{Temperature Control}
584 tim 2746
585 tim 2999 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}
605     \begin{center}
606     \begin{tabular}{lll}
607     \hline
608     $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
609     \hline
610     1 & 300.47 & 10.99 \\
611     0.1 & 301.19 & 11.136 \\
612     0.01 & 303.04 & 11.796 \\
613     \hline
614     \end{tabular}
615     \end{center}
616     \end{table}
617 tim 2746
618 tim 2999 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$, Langevin dynamics becomes normal
629     NVE (see orange curve in Fig.~\ref{langevin:temperature}) which
630     loses the temperature control ability.
631    
632     \begin{figure}
633     \centering
634 gezelter 3000 \includegraphics[width=\linewidth]{temperature.pdf}
635 tim 2999 \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
636     temperature fluctuation versus time.} \label{langevin:temperature}
637     \end{figure}
638    
639 gezelter 3299 \subsection{Comparisons with Analytic and MD simulation results}
640 xsun 3298
641 gezelter 3299 In order to validate our langevin simulation procedure for
642     arbitrarily-shaped rigid bodies, we compared the results of this
643     procedure with the known hydrodynamic limiting behavior for a few
644     model systems, and to microcanonical molecular dynamics simulations
645     for some more complicated bodies. The model systems and their
646     analytical behavior (if known) are summarized below.
647    
648     \subsubsection{Spherical particles}
649    
650     Our model system for spherical particles was a Lennard-Jones sphere of
651     diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
652     4.7 \AA). The well depth ($\epsilon$) for both particles was set to
653     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}.
656    
657     The Stokes-Einstein behavior of large spherical particles in
658     hydrodynamic flows is well known, giving translational friction
659     coefficients of $6 \pi \eta R$ (stick boundary conditions) and
660     rotational friction coefficients of $8 \pi \eta R^3$. Recently, Reid
661     and Skinner have computed the behavior of spherical tag particles in
662     molecular dynamics simulations, and have shown that {\it slip}
663     boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
664     appropriate for molecule-sized spheres embedded in a sea of spherical
665     solvent particles.\cite{ReidAndSkinner}
666    
667     Our simulation results show similar behavior to the behavior observed
668     by Reid and Skinner. The diffusion constant obtained from our
669     microcanonical molecular dynamics simulations lies between the slip
670     and stick boundary condition results obtained via Stokes-Einstein
671     behavior. Since the Langevin integrator assumes Stokes-Einstein stick
672     boundary conditions in calculating the drag and random forces for
673     spherical particles, our Langevin routine obtains nearly quantitative
674     agreement with the hydrodynamic results for spherical particles. One
675     avenue for improvement of the method would be to compute elements of
676     $\Xi_{tt}$ assuming behavior intermediate between the two boundary
677     conditions.
678    
679     In these simulations, our spherical particles were structureless, so
680     there is no way to obtain rotational correlation times for this model
681     system.
682    
683     \subsubsection{Ellipsoids}
684     For uniaxial ellipsoids ($a > b = c$) , Perrin's formulae for both
685     translational and rotational diffusion of each of the body-fixed axes
686     can be combined to give a single translational diffusion
687     constant,\cite{PecoraBerne}
688     \begin{equation}
689     D = \frac{k_B T}{6 \pi \eta a} G(\rho),
690     \label{Dperrin}
691     \end{equation}
692     as well as a single rotational diffusion coefficient,
693     \begin{equation}
694     \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - \rho^2)
695     G(\rho) - 1}{1 - \rho^4} \right\}.
696     \label{ThetaPerrin}
697     \end{equation}
698     In these expressions, $G(\rho)$ is a function of the axial ratio
699     ($\rho = b / a$), which for prolate ellipsoids, is
700     \begin{equation}
701     G(\rho) = (1- \rho^2)^{-1/2} \ln \left\{ \frac{1 + (1 -
702     \rho^2)^{1/2}}{\rho} \right\}
703     \label{GPerrin}
704     \end{equation}
705     Again, there is some uncertainty about the correct boundary conditions
706     to use for molecular-scale ellipsoids in a sea of similarly-sized
707     solvent particles. Ravichandran and Bagchi found that {\it slip}
708     boundary conditions most closely resembled the simulation results, in
709     agreement with earlier work of Tang and Evans.\cite{}
710    
711     As in the case of our spherical model system, the Langevin integrator
712     reproduces almost exactly the behavior of the Perrin formulae (which
713     is unsurprising given that the Perrin formulae were used to derive the
714     drag and random forces applied to the ellipsoid). We obtain
715     translational diffusion constants and rotational correlation times
716     that are within a few percent of the analytic values for both the
717     exact treatment of the diffusion tensor as well as the rough-shell
718     model for the ellipsoid.
719    
720     The agreement with the translational diffusion constants from the
721     microcanonical simulations is quite good, although the rotational
722     correlation times are as much as a factor of 2 different from the
723     predictions of the Perrin hydrodynamic model.
724    
725     \subsubsection{Rigid dumbells}
726    
727     Perhaps the only composite rigid body for which analytic expressions
728     for the hydrodynamic tensor are available is the two-sphere dumbell
729     model. This model consists of two non-overlapping spheres held by a
730     rigid bond connecting their centers. There are competing expressions
731     for the 6x6 resistance tensor for this
732     model. Equation (\ref{introEquation:oseenTensor}) above gives the original
733     Oseen tensor, while the second order expression introduced by Rotne
734     and Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la Torre and
735     Bloomfield,\cite{Torre1977} is given above as
736     Eq. (\ref{introEquation:RPTensorNonOverlapped}). In our case, we use
737     a model dumbbell in which the two spheres are identical Lennard-Jones
738     particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
739     a distance of 6.65 \AA\ ??.
740    
741     The theoretical values for the translational diffusion constant of the
742     dumbbell are calculated from the work of Stimson and Jeffery, who
743     studied the motion of this system in a flow parallel to the
744     inter-sphere axis,\cite{StimsonJeffery26} and Davis, who studied the
745     motion in a flow perpendicular to the inter-sphere axis.\cite{Davis69}
746    
747     How did we do? Does Analytic reproduce MD? Does LD reproduce
748     Analytic or MD?
749    
750     \subsubsection{Ellipsoidal-composite banana-shaped molecules}
751    
752     Banana-shaped rigid bodies composed of composites of Gay-Berne
753     ellipsoids have been used by Orlandi {\it et al.} to observe
754     mesophases in coarse-grained models bent-core liquid crystalline
755     molecules.\cite{OrlandiZannoni06} We have used the overlapping
756     ellipsoids as a way to test the behavior of our algorithm for a
757     structure of some interest to the materials science community,
758     although since we are interested in capturing only the hydrodynamic
759     behavior of this model, we leave out the dipolar interactions of the
760     original Orlandi model.
761    
762     \subsubsection{Composite sphero-ellipsoids}
763    
764     Spherical heads perched on the ends of Gay-Berne ellipsoids have been
765     used recently as models for lipid molecules.\cite{SunGezelter08,AytonVoth??}
766    
767    
768     We performed several NVE
769 xsun 3298 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 gezelter 3299 calculated from Perrin's fomula, The exact method is
830 xsun 3298 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.
860    
861     According to our simulations, the langevin dynamics is a reliable
862     theory to apply to replace the explicit solvents, especially for the
863     translation properties. For large molecules, the rotation properties
864     are also mimiced reasonablly well.
865    
866     \begin{table*}
867     \begin{minipage}{\linewidth}
868     \begin{center}
869     \caption{}
870     \begin{tabular}{llccccccc}
871     \hline
872     & & Sphere & Ellipsoid & Dumbbell(2 spheres) & Banana(3 ellpsoids) &
873     Lipid(head) & lipid(tail) & Solvent \\
874     \hline
875     $d$ (\AA) & & 6.5 & 4.6 & 6.5 & 4.2 & 6.5 & 4.6 & 4.7 \\
876     $l$ (\AA) & & $= d$ & 13.8 & $=d$ & 11.2 & $=d$ & 13.8 & 4.7 \\
877     $\epsilon^s$ (kcal/mol) & & 0.8 & 0.8 & 0.8 & 0.8 & 0.185 & 0.8 & 0.8 \\
878     $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 & 0.2 & 1 & 0.2 & 1 \\
879     $m$ (amu) & & 190 & 200 & 190 & 240 & 196 & 760 & 72.06 \\
880     %$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
881     %\multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
882     %\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*}
890    
891     \begin{table*}
892     \begin{minipage}{\linewidth}
893     \begin{center}
894     \caption{}
895     \begin{tabular}{lccccc}
896     \hline
897     & & & & &Translation \\
898     \hline
899     & NVE & & Theoretical & Langevin & \\
900     \hline
901     & $\eta$ & D & D & method & D \\
902     \hline
903     sphere & 3.480159e-03 & 1.643135e-04 & 1.942779e-04 & exact & 1.982283e-04 \\
904     ellipsoid & 2.551262e-03 & 2.437492e-04 & 2.335756e-04 & exact & 2.374905e-04 \\
905     & 2.551262e-03 & 2.437492e-04 & 2.335756e-04 & rough shell & 2.284088e-04 \\
906     dumbell & 2.41276e-03 & 2.129432e-04 & 2.090239e-04 & bead model & 2.148098e-04 \\
907     & 2.41276e-03 & 2.129432e-04 & 2.090239e-04 & rough shell & 2.013219e-04 \\
908     banana & 2.9846e-03 & 1.527819e-04 & & rough shell & 1.54807e-04 \\
909     lipid & 3.488661e-03 & 0.9562979e-04 & & rough shell & 1.320987e-04 \\
910     \end{tabular}
911     \label{tab:translation}
912     \end{center}
913     \end{minipage}
914     \end{table*}
915    
916     \begin{table*}
917     \begin{minipage}{\linewidth}
918     \begin{center}
919     \caption{}
920     \begin{tabular}{lccccc}
921     \hline
922     & & & & &Rotation \\
923     \hline
924     & NVE & & Theoretical & Langevin & \\
925     \hline
926     & $\eta$ & $\tau_0$ & $\tau_0$ & method & $\tau_0$ \\
927     \hline
928     sphere & 3.480159e-03 & & 1.208178e+04 & exact & 1.20628e+04 \\
929     ellipsoid & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & exact & 2.21507e+04 \\
930     & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & rough shell & 2.21714e+04 \\
931     dumbell & 2.41276e-03 & 1.42974e+04 & & bead model & 7.12435e+04 \\
932     & 2.41276e-03 & 1.42974e+04 & & rough shell & 7.04765e+04 \\
933     banana & 2.9846e-03 & 6.38323e+04 & & rough shell & 7.0945e+04 \\
934     lipid & 3.488661e-03 & 7.79595e+04 & & rough shell & 7.78886e+04 \\
935     \end{tabular}
936     \label{tab:rotation}
937     \end{center}
938     \end{minipage}
939     \end{table*}
940    
941     Langevin dynamics simulations are applied to study the formation of
942     the ripple phase of lipid membranes. The initial configuration is
943     taken from our molecular dynamics studies on lipid bilayers with
944     lennard-Jones sphere solvents. The solvent molecules are excluded from
945     the system, the experimental value of water viscosity is applied to
946     mimic the heat bath. Fig. XXX is the snapshot of the stable
947     configuration of the system, the ripple structure stayed stable after
948     100 ns run. The efficiency of the simulation is increased by one order
949     of magnitude.
950    
951 tim 2999 \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    
988     \begin{figure}
989     \centering
990 gezelter 3000 \includegraphics[width=\linewidth]{roughShell.pdf}
991 tim 2999 \caption[Rough shell model for banana shaped molecule]{Rough shell
992     model for banana shaped molecule.} \label{langevin:roughShell}
993     \end{figure}
994    
995     \begin{figure}
996     \centering
997 gezelter 3000 \includegraphics[width=\linewidth]{twoBanana.pdf}
998 tim 2999 \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 gezelter 3000 \includegraphics[width=\linewidth]{vacf.pdf}
1006 tim 2999 \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 gezelter 3000 \includegraphics[width=\linewidth]{uacf.pdf}
1014 tim 2999 \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    
1020 tim 2746 \section{Conclusions}
1021    
1022 tim 2999 We have presented a new Langevin algorithm by incorporating the
1023     hydrodynamics properties of arbitrary shaped molecules into an
1024     advanced symplectic integration scheme. The temperature control
1025     ability of this algorithm was demonstrated by a set of simulations
1026     with different viscosities. It was also shown to have significant
1027     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.
1031    
1032    
1033 tim 2746 \section{Acknowledgments}
1034     Support for this project was provided by the National Science
1035     Foundation under grant CHE-0134881. T.L. also acknowledges the
1036     financial support from center of applied mathematics at University
1037     of Notre Dame.
1038     \newpage
1039    
1040     \bibliographystyle{jcp2}
1041     \bibliography{langevin}
1042    
1043     \end{document}