ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 3000
Committed: Mon Sep 4 16:36:51 2006 UTC (18 years ago) by gezelter
Content type: application/x-tex
File size: 31725 byte(s)
Log Message:
Converting to pdflatex, changing eps figures to pdf, adding Makefile
and applescript previewer

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     \renewcommand\citemid{\ } % no comma in optional reference note
21    
22     \begin{document}
23    
24     \title{Langevin Dynamics for Rigid Body of Arbitrary Shape }
25    
26 tim 2999 \author{Teng Lin, Xiuquan Sun 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     \subsection{Langevin Dynamics of Banana Shaped Molecules}
640    
641     In order to verify that Langevin dynamics can mimic the dynamics of
642     the systems absent of explicit solvents, we carried out two sets of
643     simulations and compare their dynamic properties.
644     Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
645     made of 256 pentane molecules and two banana shaped molecules at
646     273~K. It has an equivalent implicit solvent system containing only
647     two banana shaped molecules with viscosity of 0.289 center poise. To
648     calculate the hydrodynamic properties of the banana shaped molecule,
649     we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
650     in which the banana shaped molecule is represented as a ``shell''
651     made of 2266 small identical beads with size of 0.3 \AA on the
652     surface. Applying the procedure described in
653     Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
654     identified the center of resistance at (0 $\rm{\AA}$, 0.7482 $\rm{\AA}$,
655     -0.1988 $\rm{\AA}$), as well as the resistance tensor,
656     \[
657     \left( {\begin{array}{*{20}c}
658     0.9261 & 0 & 0&0&0.08585&0.2057\\
659     0& 0.9270&-0.007063& 0.08585&0&0\\
660     0&-0.007063&0.7494&0.2057&0&0\\
661     0&0.0858&0.2057& 58.64& 0&0\\
662     0.08585&0&0&0&48.30&3.219&\\
663     0.2057&0&0&0&3.219&10.7373\\
664     \end{array}} \right).
665     \]
666     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.
667     Curves of the velocity auto-correlation functions in
668     Fig.~\ref{langevin:vacf} were shown to match each other very well.
669     However, because of the stochastic nature, simulation using Langevin
670     dynamics was shown to decay slightly faster than MD. In order to
671     study the rotational motion of the molecules, we also calculated the
672     auto-correlation function of the principle axis of the second GB
673     particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
674     probably due to the reason that we used the experimental viscosity directly instead of calculating bulk viscosity from simulation.
675    
676     \begin{figure}
677     \centering
678 gezelter 3000 \includegraphics[width=\linewidth]{roughShell.pdf}
679 tim 2999 \caption[Rough shell model for banana shaped molecule]{Rough shell
680     model for banana shaped molecule.} \label{langevin:roughShell}
681     \end{figure}
682    
683     \begin{figure}
684     \centering
685 gezelter 3000 \includegraphics[width=\linewidth]{twoBanana.pdf}
686 tim 2999 \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
687     256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
688     molecules and 256 pentane molecules.} \label{langevin:twoBanana}
689     \end{figure}
690    
691     \begin{figure}
692     \centering
693 gezelter 3000 \includegraphics[width=\linewidth]{vacf.pdf}
694 tim 2999 \caption[Plots of Velocity Auto-correlation Functions]{Velocity
695     auto-correlation functions of NVE (explicit solvent) in blue and
696     Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
697     \end{figure}
698    
699     \begin{figure}
700     \centering
701 gezelter 3000 \includegraphics[width=\linewidth]{uacf.pdf}
702 tim 2999 \caption[Auto-correlation functions of the principle axis of the
703     middle GB particle]{Auto-correlation functions of the principle axis
704     of the middle GB particle of NVE (blue) and Langevin dynamics
705     (red).} \label{langevin:uacf}
706     \end{figure}
707    
708 tim 2746 \section{Conclusions}
709    
710 tim 2999 We have presented a new Langevin algorithm by incorporating the
711     hydrodynamics properties of arbitrary shaped molecules into an
712     advanced symplectic integration scheme. The temperature control
713     ability of this algorithm was demonstrated by a set of simulations
714     with different viscosities. It was also shown to have significant
715     advantage of producing rapid thermal equilibration over
716     Nos\'{e}-Hoover method. Further studies in systems involving banana
717     shaped molecules illustrated that the dynamic properties could be
718     preserved by using this new algorithm as an implicit solvent model.
719    
720    
721 tim 2746 \section{Acknowledgments}
722     Support for this project was provided by the National Science
723     Foundation under grant CHE-0134881. T.L. also acknowledges the
724     financial support from center of applied mathematics at University
725     of Notre Dame.
726     \newpage
727    
728     \bibliographystyle{jcp2}
729     \bibliography{langevin}
730    
731     \end{document}