ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3308
Committed: Fri Jan 11 17:04:12 2008 UTC (16 years, 7 months ago) by gezelter
Content type: application/x-tex
File size: 52862 byte(s)
Log Message:
updates

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 gezelter 3305 \section{Results}
578 gezelter 3302 In order to validate our Langevin integrator for arbitrarily-shaped
579 gezelter 3305 rigid bodies, we implemented the algorithm in {\sc
580     oopse}\cite{Meineke2005} and compared the results of this algorithm
581     with the known
582 gezelter 3302 hydrodynamic limiting behavior for a few model systems, and to
583     microcanonical molecular dynamics simulations for some more
584     complicated bodies. The model systems and their analytical behavior
585     (if known) are summarized below. Parameters for the primary particles
586     comprising our model systems are given in table \ref{tab:parameters},
587     and a sketch of the arrangement of these primary particles into the
588 gezelter 3305 model rigid bodies is shown in figure \ref{fig:models}. In table
589     \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
590     ellipsoidal (Gay-Berne) particles. For spherical particles, the value
591     of the Lennard-Jones $\sigma$ parameter is the particle diameter
592     ($d$). Gay-Berne ellipsoids have an energy scaling parameter,
593     $\epsilon^s$, which describes the well depth for two identical
594     ellipsoids in a {\it side-by-side} configuration. Additionally, a
595     well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
596     describes the ratio between the well depths in the {\it end-to-end}
597     and side-by-side configurations. For spheres, $\epsilon^r \equiv 1$.
598     Moments of inertia are also required to describe the motion of primary
599     particles with orientational degrees of freedom.
600 gezelter 3299
601 gezelter 3302 \begin{table*}
602     \begin{minipage}{\linewidth}
603     \begin{center}
604     \caption{Parameters for the primary particles in use by the rigid body
605     models in figure \ref{fig:models}.}
606     \begin{tabular}{lrcccccccc}
607     \hline
608     & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
609     & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
610     $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
611 gezelter 3308 Sphere & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
612 gezelter 3302 Ellipsoid & & 4.6 & 13.8 & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
613 gezelter 3308 Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75 & 802.75 \\
614 gezelter 3302 Banana &(3 identical ellipsoids)& 4.2 & 11.2 & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
615     Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
616     & Ellipsoidal Tail & 4.6 & 13.8 & 0.8 & 0.2 & 760 & 45000 & 45000 & 9000 \\
617     Solvent & & 4.7 & $= d$ & 0.8 & 1 & 72.06 & & & \\
618     \hline
619     \end{tabular}
620     \label{tab:parameters}
621     \end{center}
622     \end{minipage}
623     \end{table*}
624    
625 gezelter 3305 \begin{figure}
626     \centering
627     \includegraphics[width=3in]{sketch}
628     \caption[Sketch of the model systems]{A sketch of the model systems
629     used in evaluating the behavior of the rigid body Langevin
630     integrator.} \label{fig:models}
631     \end{figure}
632    
633 gezelter 3302 \subsection{Simulation Methodology}
634    
635     We performed reference microcanonical simulations with explicit
636     solvents for each of the different model system. In each case there
637     was one solute model and 1929 solvent molecules present in the
638     simulation box. All simulations were equilibrated using a
639     constant-pressure and temperature integrator with target values of 300
640     K for the temperature and 1 atm for pressure. Following this stage,
641     further equilibration and sampling was done in a microcanonical
642 gezelter 3305 ensemble. Since the model bodies are typically quite massive, we were
643     able to use a time step of 25 fs. A switching function was applied to
644 gezelter 3302 all potentials to smoothly turn off the interactions between a range
645     of $22$ and $25$ \AA. The switching function was the standard (cubic)
646     function,
647     \begin{equation}
648     s(r) =
649     \begin{cases}
650     1 & \text{if $r \le r_{\text{sw}}$},\\
651     \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
652     {(r_{\text{cut}} - r_{\text{sw}})^3}
653     & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
654     0 & \text{if $r > r_{\text{cut}}$.}
655     \end{cases}
656     \label{eq:switchingFunc}
657     \end{equation}
658     To measure shear viscosities from our microcanonical simulations, we
659     used the Einstein form of the pressure correlation function,\cite{hess:209}
660     \begin{equation}
661     \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \langle \left(
662     \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \rangle_{t_0}.
663     \label{eq:shear}
664     \end{equation}
665     A similar form exists for the bulk viscosity
666     \begin{equation}
667     \kappa = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \langle \left(
668     \int_{t_0}^{t_0 + t}
669     \left(P\left(t'\right)-\langle P \rangle \right)dt'
670     \right)^2 \rangle_{t_0}.
671     \end{equation}
672     Alternatively, the shear viscosity can also be calculated using a
673     Green-Kubo formula with the off-diagonal pressure tensor correlation function,
674     \begin{equation}
675     \eta = \frac{V}{k_B T} \int_0^{\infty} \langle P_{xz}(t_0) P_{xz}(t_0
676     + t) \rangle_{t_0} dt,
677     \end{equation}
678     although this method converges extremely slowly and is not practical
679     for obtaining viscosities from molecular dynamics simulations.
680    
681     The Langevin dynamics for the different model systems were performed
682     at the same temperature as the average temperature of the
683     microcanonical simulations and with a solvent viscosity taken from
684 gezelter 3305 Eq. (\ref{eq:shear}) applied to these simulations. We used 1024
685     independent solute simulations to obtain statistics on our Langevin
686     integrator.
687 gezelter 3302
688     \subsection{Analysis}
689    
690     The quantities of interest when comparing the Langevin integrator to
691     analytic hydrodynamic equations and to molecular dynamics simulations
692     are typically translational diffusion constants and orientational
693     relaxation times. Translational diffusion constants for point
694     particles are computed easily from the long-time slope of the
695     mean-square displacement,
696     \begin{equation}
697     D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
698     \end{equation}
699     of the solute molecules. For models in which the translational
700 gezelter 3305 diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
701     (i.e. any non-spherically-symmetric rigid body), it is possible to
702     compute the diffusive behavior for motion parallel to each body-fixed
703     axis by projecting the displacement of the particle onto the
704     body-fixed reference frame at $t=0$. With an isotropic solvent, as we
705     have used in this study, there are differences between the three
706 gezelter 3302 diffusion constants, but these must converge to the same value at
707     longer times. Translational diffusion constants for the different
708 gezelter 3305 shaped models are shown in table \ref{tab:translation}.
709 gezelter 3302
710 gezelter 3305 In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
711 gezelter 3302 diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
712     {\it around} a particular body-fixed axis and {\it not} the diffusion
713     of a vector pointing along the axis. However, these eigenvalues can
714     be combined to find 5 characteristic rotational relaxation
715 gezelter 3305 times,\cite{PhysRev.119.53,Berne90}
716 gezelter 3302 \begin{eqnarray}
717 gezelter 3305 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
718     1 / \tau_2 & = & 6 D_r - 2 \Delta \\
719     1 / \tau_3 & = & 3 (D_r + D_1) \\
720     1 / \tau_4 & = & 3 (D_r + D_2) \\
721     1 / \tau_5 & = & 3 (D_r + D_3)
722 gezelter 3302 \end{eqnarray}
723     where
724     \begin{equation}
725     D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
726     \end{equation}
727     and
728     \begin{equation}
729 gezelter 3305 \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
730 gezelter 3302 \end{equation}
731 gezelter 3305 Each of these characteristic times can be used to predict the decay of
732     part of the rotational correlation function when $\ell = 2$,
733 gezelter 3302 \begin{equation}
734 gezelter 3305 C_2(t) = \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
735 gezelter 3302 \end{equation}
736 gezelter 3305 This is the same as the $F^2_{0,0}(t)$ correlation function that
737     appears in Ref. \citen{Berne90}. The amplitudes of the two decay
738     terms are expressed in terms of three dimensionless functions of the
739     eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
740     2\Delta)$, and $N = 2 \sqrt{\Delta b}$. Similar expressions can be
741     obtained for other angular momentum correlation
742     functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
743     studied, only one of the amplitudes of the two decay terms was
744     non-zero, so it was possible to derive a single relaxation time for
745     each of the hydrodynamic tensors. In many cases, these characteristic
746     times are averaged and reported in the literature as a single relaxation
747     time,\cite{Garcia-de-la-Torre:1997qy}
748 gezelter 3302 \begin{equation}
749 gezelter 3305 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
750     \end{equation}
751     although for the cases reported here, this averaging is not necessary
752     and only one of the five relaxation times is relevant.
753    
754     To test the Langevin integrator's behavior for rotational relaxation,
755     we have compared the analytical orientational relaxation times (if
756     they are known) with the general result from the diffusion tensor and
757     with the results from both the explicitly solvated molecular dynamics
758     and Langevin simulations. Relaxation times from simulations (both
759     microcanonical and Langevin), were computed using Legendre polynomial
760     correlation functions for a unit vector (${\bf u}$) fixed along one or
761     more of the body-fixed axes of the model.
762     \begin{equation}
763 gezelter 3302 C_{\ell}(t) = \langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
764     u}_{i}(0) \right)
765     \end{equation}
766     For simulations in the high-friction limit, orientational correlation
767     times can then be obtained from exponential fits of this function, or by
768     integrating,
769     \begin{equation}
770 gezelter 3305 \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
771 gezelter 3302 \end{equation}
772 gezelter 3305 In lower-friction solvents, the Legendre correlation functions often
773     exhibit non-exponential decay, and may not be characterized by a
774     single decay constant.
775 gezelter 3302
776     In table \ref{tab:rotation} we show the characteristic rotational
777     relaxation times (based on the diffusion tensor) for each of the model
778     systems compared with the values obtained via microcanonical and Langevin
779     simulations.
780    
781 gezelter 3305 \subsection{Spherical particles}
782 gezelter 3299 Our model system for spherical particles was a Lennard-Jones sphere of
783     diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
784     4.7 \AA). The well depth ($\epsilon$) for both particles was set to
785 gezelter 3302 an arbitrary value of 0.8 kcal/mol.
786 gezelter 3299
787     The Stokes-Einstein behavior of large spherical particles in
788     hydrodynamic flows is well known, giving translational friction
789     coefficients of $6 \pi \eta R$ (stick boundary conditions) and
790 gezelter 3302 rotational friction coefficients of $8 \pi \eta R^3$. Recently,
791     Schmidt and Skinner have computed the behavior of spherical tag
792     particles in molecular dynamics simulations, and have shown that {\it
793     slip} boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
794 gezelter 3299 appropriate for molecule-sized spheres embedded in a sea of spherical
795 gezelter 3305 qsolvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
796 gezelter 3299
797     Our simulation results show similar behavior to the behavior observed
798 gezelter 3302 by Schmidt and Skinner. The diffusion constant obtained from our
799 gezelter 3299 microcanonical molecular dynamics simulations lies between the slip
800     and stick boundary condition results obtained via Stokes-Einstein
801     behavior. Since the Langevin integrator assumes Stokes-Einstein stick
802     boundary conditions in calculating the drag and random forces for
803     spherical particles, our Langevin routine obtains nearly quantitative
804     agreement with the hydrodynamic results for spherical particles. One
805     avenue for improvement of the method would be to compute elements of
806     $\Xi_{tt}$ assuming behavior intermediate between the two boundary
807 gezelter 3302 conditions.
808 gezelter 3299
809     In these simulations, our spherical particles were structureless, so
810     there is no way to obtain rotational correlation times for this model
811     system.
812    
813     \subsubsection{Ellipsoids}
814     For uniaxial ellipsoids ($a > b = c$) , Perrin's formulae for both
815     translational and rotational diffusion of each of the body-fixed axes
816     can be combined to give a single translational diffusion
817 gezelter 3302 constant,\cite{Berne90}
818 gezelter 3299 \begin{equation}
819     D = \frac{k_B T}{6 \pi \eta a} G(\rho),
820     \label{Dperrin}
821     \end{equation}
822     as well as a single rotational diffusion coefficient,
823     \begin{equation}
824     \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - \rho^2)
825     G(\rho) - 1}{1 - \rho^4} \right\}.
826     \label{ThetaPerrin}
827     \end{equation}
828     In these expressions, $G(\rho)$ is a function of the axial ratio
829     ($\rho = b / a$), which for prolate ellipsoids, is
830     \begin{equation}
831     G(\rho) = (1- \rho^2)^{-1/2} \ln \left\{ \frac{1 + (1 -
832     \rho^2)^{1/2}}{\rho} \right\}
833     \label{GPerrin}
834     \end{equation}
835     Again, there is some uncertainty about the correct boundary conditions
836     to use for molecular-scale ellipsoids in a sea of similarly-sized
837     solvent particles. Ravichandran and Bagchi found that {\it slip}
838 gezelter 3302 boundary conditions most closely resembled the simulation
839     results,\cite{Ravichandran:1999fk} in agreement with earlier work of
840     Tang and Evans.\cite{TANG:1993lr}
841 gezelter 3299
842 gezelter 3305 Even though there are analytic resistance tensors for ellipsoids, we
843     constructed a rough-shell model using 2135 beads (each with a diameter
844     of 0.25 \AA) to approximate the shape of the modle ellipsoid. We
845     compared the Langevin dynamics from both the simple ellipsoidal
846     resistance tensor and the rough shell approximation with
847     microcanonical simulations and the predictions of Perrin. As in the
848     case of our spherical model system, the Langevin integrator reproduces
849     almost exactly the behavior of the Perrin formulae (which is
850     unsurprising given that the Perrin formulae were used to derive the
851 gezelter 3299 drag and random forces applied to the ellipsoid). We obtain
852     translational diffusion constants and rotational correlation times
853     that are within a few percent of the analytic values for both the
854     exact treatment of the diffusion tensor as well as the rough-shell
855     model for the ellipsoid.
856    
857 gezelter 3308 The translational diffusion constants from the microcanonical simulations
858     agree well with the predictions of the Perrin model, although the rotational
859     correlation times are a factor of 2 shorter than expected from hydrodynamic
860     theory. One explanation for the slower rotation
861     of explicitly-solvated ellipsoids is the possibility that solute-solvent
862     collisions happen at both ends of the solute whenever the principal
863     axis of the ellipsoid is turning. In the upper portion of figure
864     \ref{fig:explanation} we sketch a physical picture of this explanation.
865     Since our Langevin integrator is providing nearly quantitative agreement with
866     the Perrin model, it also predicts orientational diffusion for ellipsoids that
867     exceed explicitly solvated correlation times by a factor of two.
868 gezelter 3299
869 gezelter 3308 \begin{figure}
870     \centering
871     \includegraphics[width=6in]{explanation}
872     \caption[Explanations of the differences between orientational correlation times for explicitly-solvated models and hydrodynamics predictions]{Explanations of the differences between orientational correlation times for explicitly-solvated models and hydrodynamic predictions. For the ellipsoids (upper figures), rotation of the principal axis can involve correlated collisions at both sides of the solute. In the rigid dumbbell model (lower figures), the large size of the explicit solvent spheres prevents them from coming in contact with a substantial fraction of the surface area of the dumbbell. Therefore, the explicit solvent only provides drag over a substantially reduced surface area of this model, where the hydrodynamic theories utilize the entire surface area for estimating rotational diffusion.
873     } \label{fig:explanation}
874     \end{figure}
875    
876 gezelter 3302 \subsubsection{Rigid dumbbells}
877     Perhaps the only {\it composite} rigid body for which analytic
878     expressions for the hydrodynamic tensor are available is the
879     two-sphere dumbbell model. This model consists of two non-overlapping
880     spheres held by a rigid bond connecting their centers. There are
881     competing expressions for the 6x6 resistance tensor for this
882     model. Equation (\ref{introEquation:oseenTensor}) above gives the
883     original Oseen tensor, while the second order expression introduced by
884     Rotne and Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la
885     Torre and Bloomfield,\cite{Torre1977} is given above as
886 gezelter 3299 Eq. (\ref{introEquation:RPTensorNonOverlapped}). In our case, we use
887     a model dumbbell in which the two spheres are identical Lennard-Jones
888     particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
889 gezelter 3302 a distance of 6.532 \AA.
890 gezelter 3299
891     The theoretical values for the translational diffusion constant of the
892     dumbbell are calculated from the work of Stimson and Jeffery, who
893     studied the motion of this system in a flow parallel to the
894 gezelter 3302 inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
895     motion in a flow {\it perpendicular} to the inter-sphere
896     axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
897     orientational} correlation times for this model system (other than
898 gezelter 3305 those derived from the 6 x 6 tensors mentioned above).
899 gezelter 3299
900 gezelter 3305 The bead model for this model system comprises the two large spheres
901     by themselves, while the rough shell approximation used 3368 separate
902     beads (each with a diameter of 0.25 \AA) to approximate the shape of
903     the rigid body. The hydrodynamics tensors computed from both the bead
904     and rough shell models are remarkably similar. Computing the initial
905     hydrodynamic tensor for a rough shell model can be quite expensive (in
906     this case it requires inverting a 10104 x 10104 matrix), while the
907     bead model is typically easy to compute (in this case requiring
908 gezelter 3308 inversion of a 6 x 6 matrix).
909 gezelter 3305
910 gezelter 3308 \begin{figure}
911     \centering
912     \includegraphics[width=3in]{RoughShell}
913     \caption[Model rigid bodies and their rough shell approximations]{The
914     model rigid bodies (left column) used to test this algorithm and their
915     rough-shell approximations (right-column) that were used to compute
916     the hydrodynamic tensors. The top two models (ellipsoid and dumbbell)
917     have analytic solutions and were used to test the rough shell
918     approximation. The lower two models (banana and lipid) were compared
919     with explicitly-solvated molecular dynamics simulations. }
920     \label{fig:roughShell}
921     \end{figure}
922    
923    
924 gezelter 3305 Once the hydrodynamic tensor has been computed, there is no additional
925     penalty for carrying out a Langevin simulation with either of the two
926     different hydrodynamics models. Our naive expectation is that since
927     the rigid body's surface is roughened under the various shell models,
928     the diffusion constants will be even farther from the ``slip''
929     boundary conditions than observed for the bead model (which uses a
930     Stokes-Einstein model to arrive at the hydrodynamic tensor). For the
931     dumbbell, this prediction is correct although all of the Langevin
932     diffusion constants are within 6\% of the diffusion constant predicted
933     from the fully solvated system.
934    
935 gezelter 3308 For rotational motion, Langevin integration (and the hydrodynamic tensor)
936     yields rotational correlation times that are substantially shorter than those
937     obtained from explicitly-solvated simulations. It is likely that this is due
938     to the large size of the explicit solvent spheres, a feature that prevents
939     the solvent from coming in contact with a substantial fraction of the surface
940     area of the dumbbell. Therefore, the explicit solvent only provides drag
941     over a substantially reduced surface area of this model, while the
942     hydrodynamic theories utilize the entire surface area for estimating
943     rotational diffusion. A sketch of the free volume available in the explicit
944     solvent simulations is shown in figure \ref{fig:explanation}.
945 gezelter 3305
946 gezelter 3299 \subsubsection{Ellipsoidal-composite banana-shaped molecules}
947 gezelter 3308 Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids
948     have been used by Orlandi {\it et al.} to observe
949 gezelter 3299 mesophases in coarse-grained models bent-core liquid crystalline
950 gezelter 3308 molecules.\cite{Orlandi:2006fk} We have used the same overlapping
951 gezelter 3299 ellipsoids as a way to test the behavior of our algorithm for a
952     structure of some interest to the materials science community,
953     although since we are interested in capturing only the hydrodynamic
954 gezelter 3308 behavior of this model, we have left out the dipolar interactions of the
955 gezelter 3299 original Orlandi model.
956 gezelter 3308
957     A reference system composed of a single banana rigid body embedded in a
958     sea of 1929 solvent particles was created and run under standard
959     (microcanonical) molecular dynamics. The resulting viscosity of this
960     mixture was 0.298 centipoise (as estimated using Eq. (\ref{eq:shear})).
961     To calculate the hydrodynamic properties of the banana rigid body model,
962     we created a rough shell (see Fig.~\ref{langevin:roughShell}), in which
963     the banana is represented as a ``shell'' made of 3321 identical beads
964     (0.25 \AA in diameter) distributed on the surface. Applying the
965     procedure described in Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
966     identified the center of resistance at (0 $\rm{\AA}$, 0.81 $\rm{\AA}$, 0 $\rm{\AA}$), as well as the resistance tensor,
967     \[
968     \left( {\begin{array}{*{20}c}
969     0.9261 & 0 & 0&0&0.08585&0.2057\\
970     0& 0.9270&-0.007063& 0.08585&0&0\\
971     0&-0.007063&0.7494&0.2057&0&0\\
972     0&0.0858&0.2057& 58.64& 0&0\\0.08585&0&0&0&48.30&3.219&\\0.2057&0&0&0&3.219&10.7373\\\end{array}} \right).
973     \]
974     where the units for translational, translation-rotation coupling and rotational
975     tensors are $\frac{kcal \cdot fs}{mol \cdot \rm{\AA}^2}$, $\frac{kcal \cdot fs}{
976     mol \cdot \rm{\AA} \cdot rad}$ and $\frac{kcal \cdot fs}{mol \cdot rad^2}$ respe
977     ctively.
978 gezelter 3299
979 gezelter 3308 The Langevin rigid-body integrator (and the hydrodynamic diffusion tensor)
980     are essentially quantitative for translational diffusion of this model.
981     Orientational correlation times under the Langevin rigid-body integrator
982     are within 11\% of the values obtained from explicit solvent, but these
983     models also exhibit some solvent inaccessible surface area in the
984     explicitly-solvated case.
985    
986 gezelter 3299 \subsubsection{Composite sphero-ellipsoids}
987     Spherical heads perched on the ends of Gay-Berne ellipsoids have been
988 gezelter 3302 used recently as models for lipid molecules.\cite{SunGezelter08,Ayton01}
989 gezelter 3299
990 gezelter 3302 The diffusion constants and rotation relaxation times for
991 xsun 3298 different shaped molecules are shown in table \ref{tab:translation}
992     and \ref{tab:rotation} to compare to the results calculated from NVE
993     simulations. The theoretical values for sphere is calculated from the
994     Stokes-Einstein law, the theoretical values for ellipsoid is
995 gezelter 3299 calculated from Perrin's fomula, The exact method is
996 xsun 3298 applied to the langevin dynamics simulations for sphere and ellipsoid,
997     the bead model is applied to the simulation for dumbbell molecule, and
998     the rough shell model is applied to ellipsoid, dumbbell, banana and
999     lipid molecules. The results from all the langevin dynamics
1000     simulations, including exact, bead model and rough shell, match the
1001     theoretical values perfectly for all different shaped molecules. This
1002     indicates that our simulation package for langevin dynamics is working
1003     well. The approxiate methods ( bead model and rough shell model) are
1004     accurate enough for the current simulations. The goal of the langevin
1005     dynamics theory is to replace the explicit solvents by the friction
1006     forces. We compared the dynamic properties of different shaped
1007     molecules in langevin dynamics simulations with that in NVE
1008     simulations. The results are reasonable close. Overall, the
1009     translational diffusion constants calculated from langevin dynamics
1010     simulations are very close to the values from the NVE simulation. For
1011     sphere and lipid molecules, the diffusion constants are a little bit
1012     off from the NVE simulation results. One possible reason is that the
1013     calculation of the viscosity is very difficult to be accurate. Another
1014     possible reason is that although we save very frequently during the
1015     NVE simulations and run pretty long time simulations, there is only
1016     one solute molecule in the system which makes the calculation for the
1017     diffusion constant difficult. The sphere molecule behaves as a free
1018     rotor in the solvent, so there is no rotation relaxation time
1019     calculated from NVE simulations. The rotation relaxation time is not
1020     very close to the NVE simulations results. The banana and lipid
1021     molecules match the NVE simulations results pretty well. The mismatch
1022     between langevin dynamics and NVE simulation for ellipsoid is possibly
1023     caused by the slip boundary condition. For dumbbell, the mismatch is
1024     caused by the size of the solvent molecule is pretty large compared to
1025     dumbbell molecule in NVE simulations.
1026    
1027     According to our simulations, the langevin dynamics is a reliable
1028     theory to apply to replace the explicit solvents, especially for the
1029     translation properties. For large molecules, the rotation properties
1030     are also mimiced reasonablly well.
1031    
1032     \begin{table*}
1033     \begin{minipage}{\linewidth}
1034     \begin{center}
1035 gezelter 3305 \caption{Translational diffusion constants (D) for the model systems
1036     calculated using microcanonical simulations (with explicit solvent),
1037     theoretical predictions, and Langevin simulations (with implicit solvent).
1038     Analytical solutions for the exactly-solved hydrodynamics models are
1039     from Refs. \citen{Einstein05} (sphere), \citen{Perrin1934} and \citen{Perrin1936}
1040     (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1041     (dumbbell). The other model systems have no known analytic solution.
1042     All diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1043     $10^{-4}$ \AA$^2$ / fs). }
1044     \begin{tabular}{lccccccc}
1045 xsun 3298 \hline
1046 gezelter 3305 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1047     \cline{2-3} \cline{5-7}
1048     model & $\eta$ (centipoise) & D & & Analytical & method & Hydrodynamics & simulation \\
1049 xsun 3298 \hline
1050 gezelter 3308 sphere & 0.348 & 1.64 & & 1.94 & exact & 1.94 & 1.98 \\
1051 gezelter 3305 ellipsoid & 0.255 & 2.44 & & 2.34 & exact & 2.34 & 2.37 \\
1052     & 0.255 & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1053 gezelter 3308 dumbbell & 0.241 & 2.13 & & 2.09 & bead model & 2.10 & 2.15 \\
1054     & 0.241 & 2.13 & & 2.09 & rough shell & 2.03 & 2.01 \\
1055 gezelter 3305 banana & 0.298 & 1.53 & & & rough shell & 1.56 & 1.55 \\
1056     lipid & 0.349 & 0.96 & & & rough shell & 1.33 & 1.32 \\
1057 xsun 3298 \end{tabular}
1058     \label{tab:translation}
1059     \end{center}
1060     \end{minipage}
1061     \end{table*}
1062    
1063     \begin{table*}
1064     \begin{minipage}{\linewidth}
1065     \begin{center}
1066 gezelter 3305 \caption{Orientational relaxation times ($\tau$) for the model systems using
1067     microcanonical simulation (with explicit solvent), theoretical
1068     predictions, and Langevin simulations (with implicit solvent). All
1069     relaxation times are for the rotational correlation function with
1070     $\ell = 2$ and are reported in units of ps. The ellipsoidal model has
1071     an exact solution for the orientational correlation time due to
1072     Perrin, but the other model systems have no known analytic solution.}
1073     \begin{tabular}{lccccccc}
1074 xsun 3298 \hline
1075 gezelter 3305 & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1076     \cline{2-3} \cline{5-7}
1077     model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic & simulation \\
1078 xsun 3298 \hline
1079 gezelter 3305 ellipsoid & 0.255 & 46.7 & & 22.0 & exact & 22.0 & 22.2 \\
1080     & 0.255 & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1081 gezelter 3308 dumbbell & 0.241 & 14.3 & & & bead model & 39.2 & 71.2 \\
1082     & 0.241 & 14.3 & & & rough shell & 32.6 & 70.5 \\
1083 gezelter 3305 banana & 0.298 & 63.8 & & & rough shell & 70.9 & 70.9 \\
1084     lipid & 0.349 & 78.0 & & & rough shell & 76.9 & 77.9 \\
1085     \hline
1086 xsun 3298 \end{tabular}
1087     \label{tab:rotation}
1088     \end{center}
1089     \end{minipage}
1090     \end{table*}
1091    
1092     Langevin dynamics simulations are applied to study the formation of
1093     the ripple phase of lipid membranes. The initial configuration is
1094     taken from our molecular dynamics studies on lipid bilayers with
1095     lennard-Jones sphere solvents. The solvent molecules are excluded from
1096     the system, the experimental value of water viscosity is applied to
1097     mimic the heat bath. Fig. XXX is the snapshot of the stable
1098     configuration of the system, the ripple structure stayed stable after
1099     100 ns run. The efficiency of the simulation is increased by one order
1100     of magnitude.
1101    
1102 tim 2746 \section{Conclusions}
1103    
1104 tim 2999 We have presented a new Langevin algorithm by incorporating the
1105     hydrodynamics properties of arbitrary shaped molecules into an
1106 gezelter 3308 advanced symplectic integration scheme. Further studies in systems
1107     involving banana shaped molecules illustrated that the dynamic
1108     properties could be preserved by using this new algorithm as an
1109     implicit solvent model.
1110 tim 2999
1111    
1112 tim 2746 \section{Acknowledgments}
1113     Support for this project was provided by the National Science
1114     Foundation under grant CHE-0134881. T.L. also acknowledges the
1115     financial support from center of applied mathematics at University
1116     of Notre Dame.
1117     \newpage
1118    
1119 gezelter 3305 \bibliographystyle{jcp}
1120 tim 2746 \bibliography{langevin}
1121    
1122     \end{document}