ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 2999
Committed: Mon Sep 4 15:05:46 2006 UTC (18 years ago) by tim
Content type: application/x-tex
File size: 31743 byte(s)
Log Message:
new draft of the Langevin paper

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{epsf}
8     \usepackage{times}
9     \usepackage{mathptmx}
10     \usepackage{setspace}
11     \usepackage{tabularx}
12     \usepackage{graphicx}
13     \usepackage{booktabs}
14     \usepackage{bibentry}
15     \usepackage{mathrsfs}
16     \usepackage[ref]{overcite}
17     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
18     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
19     9.0in \textwidth 6.5in \brokenpenalty=10000
20     \renewcommand{\baselinestretch}{1.2}
21     \renewcommand\citemid{\ } % no comma in optional reference note
22    
23     \begin{document}
24    
25     \title{Langevin Dynamics for Rigid Body of Arbitrary Shape }
26    
27 tim 2999 \author{Teng Lin, Xiuquan Sun and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
28 tim 2746 gezelter@nd.edu} \\
29     Department of Chemistry and Biochemistry\\
30     University of Notre Dame\\
31     Notre Dame, Indiana 46556}
32    
33     \date{\today}
34    
35     \maketitle \doublespacing
36    
37     \begin{abstract}
38    
39     \end{abstract}
40    
41     \newpage
42    
43     %\narrowtext
44    
45     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
46     % BODY OF TEXT
47     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
48    
49     \section{Introduction}
50    
51     %applications of langevin dynamics
52 tim 2999 As alternative to Newtonian dynamics, Langevin dynamics, which
53     mimics a simple heat bath with stochastic and dissipative forces,
54     has been applied in a variety of studies. The stochastic treatment
55     of the solvent enables us to carry out substantially longer time
56     simulations. Implicit solvent Langevin dynamics simulations of
57     met-enkephalin not only outperform explicit solvent simulations for
58     computational efficiency, but also agrees very well with explicit
59     solvent simulations for dynamical properties.\cite{Shen2002}
60     Recently, applying Langevin dynamics with the UNRES model, Liow and
61     his coworkers suggest that protein folding pathways can be possibly
62     explored within a reasonable amount of time.\cite{Liwo2005} The
63     stochastic nature of the Langevin dynamics also enhances the
64     sampling of the system and increases the probability of crossing
65     energy barriers.\cite{Banerjee2004, Cui2003} Combining Langevin
66     dynamics with Kramers's theory, Klimov and Thirumalai identified
67     free-energy barriers by studying the viscosity dependence of the
68     protein folding rates.\cite{Klimov1997} In order to account for
69     solvent induced interactions missing from implicit solvent model,
70     Kaya incorporated desolvation free energy barrier into implicit
71     coarse-grained solvent model in protein folding/unfolding studies
72     and discovered a higher free energy barrier between the native and
73     denatured states. Because of its stability against noise, Langevin
74     dynamics is very suitable for studying remagnetization processes in
75     various systems.\cite{Palacios1998,Berkov2002,Denisov2003} For
76 tim 2746 instance, the oscillation power spectrum of nanoparticles from
77     Langevin dynamics simulation has the same peak frequencies for
78 tim 2999 different wave vectors, which recovers the property of magnetic
79     excitations in small finite structures.\cite{Berkov2005a}
80 tim 2746
81     %review rigid body dynamics
82     Rigid bodies are frequently involved in the modeling of different
83     areas, from engineering, physics, to chemistry. For example,
84     missiles and vehicle are usually modeled by rigid bodies. The
85     movement of the objects in 3D gaming engine or other physics
86     simulator is governed by the rigid body dynamics. In molecular
87     simulation, rigid body is used to simplify the model in
88     protein-protein docking study{\cite{Gray2003}}.
89    
90     It is very important to develop stable and efficient methods to
91 tim 2999 integrate the equations of motion for orientational degrees of
92     freedom. Euler angles are the natural choice to describe the
93     rotational degrees of freedom. However, due to $\frac {1}{sin
94     \theta}$ singularities, the numerical integration of corresponding
95     equations of these motion is very inefficient and inaccurate.
96     Although an alternative integrator using multiple sets of Euler
97     angles can overcome this difficulty\cite{Barojas1973}, the
98     computational penalty and the loss of angular momentum conservation
99     still remain. A singularity-free representation utilizing
100     quaternions was developed by Evans in 1977.\cite{Evans1977}
101     Unfortunately, this approach used a nonseparable Hamiltonian
102     resulting from the quaternion representation, which prevented the
103     symplectic algorithm from being utilized. Another different approach
104     is to apply holonomic constraints to the atoms belonging to the
105     rigid body. Each atom moves independently under the normal forces
106     deriving from potential energy and constraint forces which are used
107     to guarantee the rigidness. However, due to their iterative nature,
108     the SHAKE and Rattle algorithms also converge very slowly when the
109     number of constraints increases.\cite{Ryckaert1977, Andersen1983}
110 tim 2746
111 tim 2999 A break-through in geometric literature suggests that, in order to
112 tim 2746 develop a long-term integration scheme, one should preserve the
113 tim 2999 symplectic structure of the propagator. By introducing a conjugate
114     momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
115     equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
116     proposed to evolve the Hamiltonian system in a constraint manifold
117     by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
118     An alternative method using the quaternion representation was
119     developed by Omelyan.\cite{Omelyan1998} However, both of these
120     methods are iterative and inefficient. In this section, we descibe a
121     symplectic Lie-Poisson integrator for rigid bodies developed by
122     Dullweber and his coworkers\cite{Dullweber1997} in depth.
123 tim 2746
124     %review langevin/browninan dynamics for arbitrarily shaped rigid body
125     Combining Langevin or Brownian dynamics with rigid body dynamics,
126 tim 2999 one can study slow processes in biomolecular systems. Modeling DNA
127     as a chain of rigid beads, which are subject to harmonic potentials
128     as well as excluded volume potentials, Mielke and his coworkers
129     discovered rapid superhelical stress generations from the stochastic
130     simulation of twin supercoiling DNA with response to induced
131     torques.\cite{Mielke2004} Membrane fusion is another key biological
132     process which controls a variety of physiological functions, such as
133     release of neurotransmitters \textit{etc}. A typical fusion event
134     happens on the time scale of a millisecond, which is impractical to
135     study using atomistic models with newtonian mechanics. With the help
136     of coarse-grained rigid body model and stochastic dynamics, the
137     fusion pathways were explored by many
138     researchers.\cite{Noguchi2001,Noguchi2002,Shillcock2005} Due to the
139     difficulty of numerical integration of anisotropic rotation, most of
140     the rigid body models are simply modeled using spheres, cylinders,
141     ellipsoids or other regular shapes in stochastic simulations. In an
142     effort to account for the diffusion anisotropy of arbitrary
143 tim 2746 particles, Fernandes and de la Torre improved the original Brownian
144     dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
145     incorporating a generalized $6\times6$ diffusion tensor and
146     introducing a simple rotation evolution scheme consisting of three
147 tim 2999 consecutive rotations.\cite{Fernandes2002} Unfortunately, unexpected
148     errors and biases are introduced into the system due to the
149     arbitrary order of applying the noncommuting rotation
150     operators.\cite{Beard2003} Based on the observation the momentum
151 tim 2746 relaxation time is much less than the time step, one may ignore the
152 tim 2999 inertia in Brownian dynamics. However, the assumption of zero
153 tim 2746 average acceleration is not always true for cooperative motion which
154     is common in protein motion. An inertial Brownian dynamics (IBD) was
155     proposed to address this issue by adding an inertial correction
156 tim 2999 term.\cite{Beard2000} As a complement to IBD which has a lower bound
157 tim 2746 in time step because of the inertial relaxation time, long-time-step
158     inertial dynamics (LTID) can be used to investigate the inertial
159     behavior of the polymer segments in low friction
160 tim 2999 regime.\cite{Beard2000} LTID can also deal with the rotational
161 tim 2746 dynamics for nonskew bodies without translation-rotation coupling by
162     separating the translation and rotation motion and taking advantage
163     of the analytical solution of hydrodynamics properties. However,
164 tim 2999 typical nonskew bodies like cylinders and ellipsoids are inadequate
165     to represent most complex macromolecule assemblies. These intricate
166 tim 2746 molecules have been represented by a set of beads and their
167 tim 2999 hydrodynamic properties can be calculated using variants on the
168     standard hydrodynamic interaction tensors.
169 tim 2746
170     The goal of the present work is to develop a Langevin dynamics
171 tim 2999 algorithm for arbitrary-shaped rigid particles by integrating the
172     accurate estimation of friction tensor from hydrodynamics theory
173     into the sophisticated rigid body dynamics algorithms.
174 tim 2746
175 tim 2999 \section{Computational Methods{\label{methodSec}}}
176 tim 2746
177 tim 2999 \subsection{\label{introSection:frictionTensor}Friction Tensor}
178     Theoretically, the friction kernel can be determined using the
179     velocity autocorrelation function. However, this approach becomes
180     impractical when the system becomes more and more complicated.
181     Instead, various approaches based on hydrodynamics have been
182     developed to calculate the friction coefficients. In general, the
183     friction tensor $\Xi$ is a $6\times 6$ matrix given by
184     \[
185     \Xi = \left( {\begin{array}{*{20}c}
186     {\Xi _{}^{tt} } & {\Xi _{}^{rt} } \\
187     {\Xi _{}^{tr} } & {\Xi _{}^{rr} } \\
188     \end{array}} \right).
189     \]
190     Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are $3 \times 3$
191     translational friction tensor and rotational resistance (friction)
192     tensor respectively, while ${\Xi^{tr} }$ is translation-rotation
193     coupling tensor and $ {\Xi^{rt} }$ is rotation-translation coupling
194     tensor. When a particle moves in a fluid, it may experience friction
195     force or torque along the opposite direction of the velocity or
196     angular velocity,
197     \[
198 tim 2746 \left( \begin{array}{l}
199 tim 2999 F_R \\
200     \tau _R \\
201 tim 2746 \end{array} \right) = - \left( {\begin{array}{*{20}c}
202 tim 2999 {\Xi ^{tt} } & {\Xi ^{rt} } \\
203     {\Xi ^{tr} } & {\Xi ^{rr} } \\
204 tim 2746 \end{array}} \right)\left( \begin{array}{l}
205 tim 2999 v \\
206     w \\
207 tim 2746 \end{array} \right)
208 tim 2999 \]
209     where $F_r$ is the friction force and $\tau _R$ is the friction
210     torque.
211 tim 2746
212 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
213 tim 2746
214 tim 2999 For a spherical particle with slip boundary conditions, the
215     translational and rotational friction constant can be calculated
216     from Stoke's law,
217 tim 2746 \[
218 tim 2999 \Xi ^{tt} = \left( {\begin{array}{*{20}c}
219     {6\pi \eta R} & 0 & 0 \\
220     0 & {6\pi \eta R} & 0 \\
221     0 & 0 & {6\pi \eta R} \\
222     \end{array}} \right)
223 tim 2746 \]
224 tim 2999 and
225 tim 2746 \[
226 tim 2999 \Xi ^{rr} = \left( {\begin{array}{*{20}c}
227     {8\pi \eta R^3 } & 0 & 0 \\
228     0 & {8\pi \eta R^3 } & 0 \\
229     0 & 0 & {8\pi \eta R^3 } \\
230     \end{array}} \right)
231 tim 2746 \]
232 tim 2999 where $\eta$ is the viscosity of the solvent and $R$ is the
233     hydrodynamic radius.
234    
235     Other non-spherical shapes, such as cylinders and ellipsoids, are
236     widely used as references for developing new hydrodynamics theory,
237     because their properties can be calculated exactly. In 1936, Perrin
238     extended Stokes's law to general ellipsoids, also called a triaxial
239     ellipsoid, which is given in Cartesian coordinates
240     by\cite{Perrin1934, Perrin1936}
241 tim 2746 \[
242 tim 2999 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
243     }} = 1
244 tim 2746 \]
245 tim 2999 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
246     due to the complexity of the elliptic integral, only the ellipsoid
247     with the restriction of two axes being equal, \textit{i.e.}
248     prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
249     exactly. Introducing an elliptic integral parameter $S$ for prolate
250     ellipsoids :
251     \[
252     S = \frac{2}{{\sqrt {a^2 - b^2 } }}\ln \frac{{a + \sqrt {a^2 - b^2
253     } }}{b},
254     \]
255     and oblate ellipsoids:
256     \[
257     S = \frac{2}{{\sqrt {b^2 - a^2 } }}arctg\frac{{\sqrt {b^2 - a^2 }
258     }}{a},
259     \]
260     one can write down the translational and rotational resistance
261     tensors
262     \begin{eqnarray*}
263     \Xi _a^{tt} & = & 16\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - b^2 )S - 2a}}. \\
264     \Xi _b^{tt} & = & \Xi _c^{tt} = 32\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - 3b^2 )S +
265     2a}},
266     \end{eqnarray*}
267     and
268     \begin{eqnarray*}
269     \Xi _a^{rr} & = & \frac{{32\pi }}{3}\eta \frac{{(a^2 - b^2 )b^2 }}{{2a - b^2 S}}, \\
270     \Xi _b^{rr} & = & \Xi _c^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^4 - b^4 )}}{{(2a^2 - b^2 )S - 2a}}.
271     \end{eqnarray*}
272 tim 2746
273 tim 2999 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
274    
275     Unlike spherical and other simply shaped molecules, there is no
276     analytical solution for the friction tensor for arbitrarily shaped
277     rigid molecules. The ellipsoid of revolution model and general
278     triaxial ellipsoid model have been used to approximate the
279     hydrodynamic properties of rigid bodies. However, since the mapping
280     from all possible ellipsoidal spaces, $r$-space, to all possible
281     combination of rotational diffusion coefficients, $D$-space, is not
282     unique\cite{Wegener1979} as well as the intrinsic coupling between
283     translational and rotational motion of rigid bodies, general
284     ellipsoids are not always suitable for modeling arbitrarily shaped
285     rigid molecules. A number of studies have been devoted to
286     determining the friction tensor for irregularly shaped rigid bodies
287     using more advanced methods where the molecule of interest was
288     modeled by a combinations of spheres\cite{Carrasco1999} and the
289     hydrodynamics properties of the molecule can be calculated using the
290     hydrodynamic interaction tensor. Let us consider a rigid assembly of
291     $N$ beads immersed in a continuous medium. Due to hydrodynamic
292     interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
293     than its unperturbed velocity $v_i$,
294 tim 2746 \[
295 tim 2999 v'_i = v_i - \sum\limits_{j \ne i} {T_{ij} F_j }
296 tim 2746 \]
297 tim 2999 where $F_i$ is the frictional force, and $T_{ij}$ is the
298     hydrodynamic interaction tensor. The friction force of $i$th bead is
299     proportional to its ``net'' velocity
300 tim 2746 \begin{equation}
301 tim 2999 F_i = \zeta _i v_i - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
302     \label{introEquation:tensorExpression}
303 tim 2746 \end{equation}
304 tim 2999 This equation is the basis for deriving the hydrodynamic tensor. In
305     1930, Oseen and Burgers gave a simple solution to
306     Eq.~\ref{introEquation:tensorExpression}
307 tim 2746 \begin{equation}
308 tim 2999 T_{ij} = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
309     R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
310 tim 2746 \end{equation}
311 tim 2999 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
312     A second order expression for element of different size was
313     introduced by Rotne and Prager\cite{Rotne1969} and improved by
314     Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
315 tim 2746 \begin{equation}
316 tim 2999 T_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
317     \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
318     _i^2 + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
319     \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
320     \label{introEquation:RPTensorNonOverlapped}
321 tim 2746 \end{equation}
322 tim 2999 Both of the Eq.~\ref{introEquation:oseenTensor} and
323     Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption
324     $R_{ij} \ge \sigma _i + \sigma _j$. An alternative expression for
325     overlapping beads with the same radius, $\sigma$, is given by
326 tim 2746 \begin{equation}
327 tim 2999 T_{ij} = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
328     \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
329     \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
330     \label{introEquation:RPTensorOverlapped}
331 tim 2746 \end{equation}
332 tim 2999 To calculate the resistance tensor at an arbitrary origin $O$, we
333     construct a $3N \times 3N$ matrix consisting of $N \times N$
334     $B_{ij}$ blocks
335     \begin{equation}
336     B = \left( {\begin{array}{*{20}c}
337     {B_{11} } & \ldots & {B_{1N} } \\
338     \vdots & \ddots & \vdots \\
339     {B_{N1} } & \cdots & {B_{NN} } \\
340     \end{array}} \right),
341     \end{equation}
342     where $B_{ij}$ is given by
343 tim 2746 \[
344 tim 2999 B_{ij} = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
345     )T_{ij}
346 tim 2746 \]
347 tim 2999 where $\delta _{ij}$ is the Kronecker delta function. Inverting the
348     $B$ matrix, we obtain
349 tim 2746 \[
350 tim 2999 C = B^{ - 1} = \left( {\begin{array}{*{20}c}
351     {C_{11} } & \ldots & {C_{1N} } \\
352     \vdots & \ddots & \vdots \\
353     {C_{N1} } & \cdots & {C_{NN} } \\
354     \end{array}} \right),
355 tim 2746 \]
356 tim 2999 which can be partitioned into $N \times N$ $3 \times 3$ block
357     $C_{ij}$. With the help of $C_{ij}$ and the skew matrix $U_i$
358 tim 2746 \[
359 tim 2999 U_i = \left( {\begin{array}{*{20}c}
360     0 & { - z_i } & {y_i } \\
361     {z_i } & 0 & { - x_i } \\
362     { - y_i } & {x_i } & 0 \\
363     \end{array}} \right)
364 tim 2746 \]
365 tim 2999 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
366     bead $i$ and origin $O$, the elements of resistance tensor at
367     arbitrary origin $O$ can be written as
368     \begin{eqnarray}
369     \Xi _{}^{tt} & = & \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
370     \Xi _{}^{tr} & = & \Xi _{}^{rt} = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
371     \Xi _{}^{rr} & = & - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
372     \label{introEquation:ResistanceTensorArbitraryOrigin}
373     \end{eqnarray}
374     The resistance tensor depends on the origin to which they refer. The
375     proper location for applying the friction force is the center of
376     resistance (or center of reaction), at which the trace of rotational
377     resistance tensor, $ \Xi ^{rr}$ reaches a minimum value.
378     Mathematically, the center of resistance is defined as an unique
379     point of the rigid body at which the translation-rotation coupling
380     tensors are symmetric,
381     \begin{equation}
382     \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
383     \label{introEquation:definitionCR}
384     \end{equation}
385     From Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
386     we can easily derive that the translational resistance tensor is
387     origin independent, while the rotational resistance tensor and
388     translation-rotation coupling resistance tensor depend on the
389     origin. Given the resistance tensor at an arbitrary origin $O$, and
390     a vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
391     obtain the resistance tensor at $P$ by
392     \begin{equation}
393     \begin{array}{l}
394     \Xi _P^{tt} = \Xi _O^{tt} \\
395     \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
396     \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{{tr} ^{^T }} \\
397     \end{array}
398     \label{introEquation:resistanceTensorTransformation}
399     \end{equation}
400     where
401 tim 2746 \[
402 tim 2999 U_{OP} = \left( {\begin{array}{*{20}c}
403     0 & { - z_{OP} } & {y_{OP} } \\
404     {z_i } & 0 & { - x_{OP} } \\
405     { - y_{OP} } & {x_{OP} } & 0 \\
406     \end{array}} \right)
407 tim 2746 \]
408 tim 2999 Using Eq.~\ref{introEquation:definitionCR} and
409     Eq.~\ref{introEquation:resistanceTensorTransformation}, one can
410     locate the position of center of resistance,
411     \begin{eqnarray*}
412     \left( \begin{array}{l}
413     x_{OR} \\
414     y_{OR} \\
415     z_{OR} \\
416     \end{array} \right) & = &\left( {\begin{array}{*{20}c}
417     {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
418     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
419     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
420     \end{array}} \right)^{ - 1} \\
421     & & \left( \begin{array}{l}
422     (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
423     (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
424     (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
425     \end{array} \right) \\
426     \end{eqnarray*}
427     where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
428     joining center of resistance $R$ and origin $O$.
429 tim 2746
430 tim 2999 \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
431 tim 2746
432 tim 2999 Consider the Langevin equations of motion in generalized coordinates
433 tim 2746 \begin{equation}
434     M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)} + F_{r,i} (t)
435     \label{LDGeneralizedForm}
436     \end{equation}
437     where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
438     and moment of inertial) matrix and $V_i$ is a generalized velocity,
439 tim 2999 $V_i = V_i(v_i,\omega _i)$. The right side of
440     Eq.~\ref{LDGeneralizedForm} consists of three generalized forces in
441 tim 2746 lab-fixed frame, systematic force $F_{s,i}$, dissipative force
442     $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
443     system in Newtownian mechanics typically refers to lab-fixed frame,
444     it is also convenient to handle the rotation of rigid body in
445     body-fixed frame. Thus the friction and random forces are calculated
446     in body-fixed frame and converted back to lab-fixed frame by:
447     \[
448     \begin{array}{l}
449 tim 2999 F_{f,i}^l (t) = Q^T F_{f,i}^b (t), \\
450     F_{r,i}^l (t) = Q^T F_{r,i}^b (t). \\
451     \end{array}
452 tim 2746 \]
453     Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
454     the body-fixed velocity at center of resistance $v_{R,i}^b$ and
455 tim 2999 angular velocity $\omega _i$
456 tim 2746 \begin{equation}
457     F_{r,i}^b (t) = \left( \begin{array}{l}
458     f_{r,i}^b (t) \\
459     \tau _{r,i}^b (t) \\
460     \end{array} \right) = - \left( {\begin{array}{*{20}c}
461     {\Xi _{R,t} } & {\Xi _{R,c}^T } \\
462     {\Xi _{R,c} } & {\Xi _{R,r} } \\
463     \end{array}} \right)\left( \begin{array}{l}
464     v_{R,i}^b (t) \\
465     \omega _i (t) \\
466     \end{array} \right),
467     \end{equation}
468     while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
469     with zero mean and variance
470     \begin{equation}
471     \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle =
472     \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle =
473 tim 2999 2k_B T\Xi _R \delta (t - t'). \label{randomForce}
474 tim 2746 \end{equation}
475     The equation of motion for $v_i$ can be written as
476     \begin{equation}
477     m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
478     f_{r,i}^l (t)
479     \end{equation}
480     Since the frictional force is applied at the center of resistance
481     which generally does not coincide with the center of mass, an extra
482     torque is exerted at the center of mass. Thus, the net body-fixed
483     frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
484     given by
485     \begin{equation}
486     \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
487     \end{equation}
488     where $r_{MR}$ is the vector from the center of mass to the center
489 tim 2999 of the resistance. Instead of integrating the angular velocity in
490     lab-fixed frame, we consider the equation of angular momentum in
491     body-fixed frame
492 tim 2746 \begin{equation}
493 tim 2999 \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
494     + \tau _{r,i}^b(t)
495 tim 2746 \end{equation}
496     Embedding the friction terms into force and torque, one can
497     integrate the langevin equations of motion for rigid body of
498     arbitrary shape in a velocity-Verlet style 2-part algorithm, where
499     $h= \delta t$:
500    
501 tim 2999 {\tt moveA:}
502 tim 2746 \begin{align*}
503 tim 2999 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
504     + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
505     %
506     {\bf r}(t + h) &\leftarrow {\bf r}(t)
507     + h {\bf v}\left(t + h / 2 \right), \\
508     %
509     {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
510     + \frac{h}{2} {\bf \tau}^b(t), \\
511     %
512     \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
513     (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
514 tim 2746 \end{align*}
515     In this context, the $\mathrm{rotate}$ function is the reversible
516 tim 2999 product of the three body-fixed rotations,
517 tim 2746 \begin{equation}
518     \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
519     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
520     / 2) \cdot \mathsf{G}_x(a_x /2),
521     \end{equation}
522     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
523 tim 2999 rotates both the rotation matrix ($\mathsf{Q}$) and the body-fixed
524     angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
525     axis $\alpha$,
526 tim 2746 \begin{equation}
527     \mathsf{G}_\alpha( \theta ) = \left\{
528     \begin{array}{lcl}
529 tim 2999 \mathsf{Q}(t) & \leftarrow & \mathsf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
530 tim 2746 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
531     j}(0).
532     \end{array}
533     \right.
534     \end{equation}
535     $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
536     rotation matrix. For example, in the small-angle limit, the
537     rotation matrix around the body-fixed x-axis can be approximated as
538     \begin{equation}
539     \mathsf{R}_x(\theta) \approx \left(
540     \begin{array}{ccc}
541     1 & 0 & 0 \\
542     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
543     \theta^2 / 4} \\
544     0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
545     \theta^2 / 4}
546     \end{array}
547     \right).
548     \end{equation}
549 tim 2999 All other rotations follow in a straightforward manner. After the
550     first part of the propagation, the forces and body-fixed torques are
551     calculated at the new positions and orientations
552 tim 2746
553 tim 2999 {\tt doForces:}
554     \begin{align*}
555     {\bf f}(t + h) &\leftarrow
556     - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
557     %
558     {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
559     \times \frac{\partial V}{\partial {\bf u}}, \\
560     %
561     {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{Q}(t + h)
562     \cdot {\bf \tau}^s(t + h).
563     \end{align*}
564 tim 2746 Once the forces and torques have been obtained at the new time step,
565     the velocities can be advanced to the same time value.
566    
567 tim 2999 {\tt moveB:}
568 tim 2746 \begin{align*}
569 tim 2999 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
570     \right)
571     + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
572     %
573     {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
574     \right)
575     + \frac{h}{2} {\bf \tau}^b(t + h) .
576 tim 2746 \end{align*}
577    
578 tim 2999 \section{Results and Discussion}
579 tim 2746
580 tim 2999 The Langevin algorithm described in previous section has been
581     implemented in {\sc oopse}\cite{Meineke2005} and applied to studies
582     of the static and dynamic properties in several systems.
583 tim 2746
584 tim 2999 \subsection{Temperature Control}
585 tim 2746
586 tim 2999 As shown in Eq.~\ref{randomForce}, random collisions associated with
587     the solvent's thermal motions is controlled by the external
588     temperature. The capability to maintain the temperature of the whole
589     system was usually used to measure the stability and efficiency of
590     the algorithm. In order to verify the stability of this new
591     algorithm, a series of simulations are performed on system
592     consisiting of 256 SSD water molecules with different viscosities.
593     The initial configuration for the simulations is taken from a 1ns
594     NVT simulation with a cubic box of 19.7166~\AA. All simulation are
595     carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
596     with reference temperature at 300~K. The average temperature as a
597     function of $\eta$ is shown in Table \ref{langevin:viscosity} where
598     the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
599     1$ poise. The better temperature control at higher viscosity can be
600     explained by the finite size effect and relative slow relaxation
601     rate at lower viscosity regime.
602     \begin{table}
603     \caption{AVERAGE TEMPERATURES FROM LANGEVIN DYNAMICS SIMULATIONS OF
604     SSD WATER MOLECULES WITH REFERENCE TEMPERATURE AT 300~K.}
605     \label{langevin:viscosity}
606     \begin{center}
607     \begin{tabular}{lll}
608     \hline
609     $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
610     \hline
611     1 & 300.47 & 10.99 \\
612     0.1 & 301.19 & 11.136 \\
613     0.01 & 303.04 & 11.796 \\
614     \hline
615     \end{tabular}
616     \end{center}
617     \end{table}
618 tim 2746
619 tim 2999 Another set of calculations were performed to study the efficiency of
620     temperature control using different temperature coupling schemes.
621     The starting configuration is cooled to 173~K and evolved using NVE,
622     NVT, and Langevin dynamic with time step of 2 fs.
623     Fig.~\ref{langevin:temperature} shows the heating curve obtained as
624     the systems reach equilibrium. The orange curve in
625     Fig.~\ref{langevin:temperature} represents the simulation using
626     Nos\'e-Hoover temperature scaling scheme with thermostat of 5 ps
627     which gives reasonable tight coupling, while the blue one from
628     Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
629     scaling to the desire temperature. When $ \eta = 0$, Langevin dynamics becomes normal
630     NVE (see orange curve in Fig.~\ref{langevin:temperature}) which
631     loses the temperature control ability.
632    
633     \begin{figure}
634     \centering
635     \includegraphics[width=\linewidth]{temperature.eps}
636     \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
637     temperature fluctuation versus time.} \label{langevin:temperature}
638     \end{figure}
639    
640     \subsection{Langevin Dynamics of Banana Shaped Molecules}
641    
642     In order to verify that Langevin dynamics can mimic the dynamics of
643     the systems absent of explicit solvents, we carried out two sets of
644     simulations and compare their dynamic properties.
645     Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
646     made of 256 pentane molecules and two banana shaped molecules at
647     273~K. It has an equivalent implicit solvent system containing only
648     two banana shaped molecules with viscosity of 0.289 center poise. To
649     calculate the hydrodynamic properties of the banana shaped molecule,
650     we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
651     in which the banana shaped molecule is represented as a ``shell''
652     made of 2266 small identical beads with size of 0.3 \AA on the
653     surface. Applying the procedure described in
654     Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
655     identified the center of resistance at (0 $\rm{\AA}$, 0.7482 $\rm{\AA}$,
656     -0.1988 $\rm{\AA}$), as well as the resistance tensor,
657     \[
658     \left( {\begin{array}{*{20}c}
659     0.9261 & 0 & 0&0&0.08585&0.2057\\
660     0& 0.9270&-0.007063& 0.08585&0&0\\
661     0&-0.007063&0.7494&0.2057&0&0\\
662     0&0.0858&0.2057& 58.64& 0&0\\
663     0.08585&0&0&0&48.30&3.219&\\
664     0.2057&0&0&0&3.219&10.7373\\
665     \end{array}} \right).
666     \]
667     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.
668     Curves of the velocity auto-correlation functions in
669     Fig.~\ref{langevin:vacf} were shown to match each other very well.
670     However, because of the stochastic nature, simulation using Langevin
671     dynamics was shown to decay slightly faster than MD. In order to
672     study the rotational motion of the molecules, we also calculated the
673     auto-correlation function of the principle axis of the second GB
674     particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
675     probably due to the reason that we used the experimental viscosity directly instead of calculating bulk viscosity from simulation.
676    
677     \begin{figure}
678     \centering
679     \includegraphics[width=\linewidth]{roughShell.eps}
680     \caption[Rough shell model for banana shaped molecule]{Rough shell
681     model for banana shaped molecule.} \label{langevin:roughShell}
682     \end{figure}
683    
684     \begin{figure}
685     \centering
686     \includegraphics[width=\linewidth]{twoBanana.eps}
687     \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
688     256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
689     molecules and 256 pentane molecules.} \label{langevin:twoBanana}
690     \end{figure}
691    
692     \begin{figure}
693     \centering
694     \includegraphics[width=\linewidth]{vacf.eps}
695     \caption[Plots of Velocity Auto-correlation Functions]{Velocity
696     auto-correlation functions of NVE (explicit solvent) in blue and
697     Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
698     \end{figure}
699    
700     \begin{figure}
701     \centering
702     \includegraphics[width=\linewidth]{uacf.eps}
703     \caption[Auto-correlation functions of the principle axis of the
704     middle GB particle]{Auto-correlation functions of the principle axis
705     of the middle GB particle of NVE (blue) and Langevin dynamics
706     (red).} \label{langevin:uacf}
707     \end{figure}
708    
709 tim 2746 \section{Conclusions}
710    
711 tim 2999 We have presented a new Langevin algorithm by incorporating the
712     hydrodynamics properties of arbitrary shaped molecules into an
713     advanced symplectic integration scheme. The temperature control
714     ability of this algorithm was demonstrated by a set of simulations
715     with different viscosities. It was also shown to have significant
716     advantage of producing rapid thermal equilibration over
717     Nos\'{e}-Hoover method. Further studies in systems involving banana
718     shaped molecules illustrated that the dynamic properties could be
719     preserved by using this new algorithm as an implicit solvent model.
720    
721    
722 tim 2746 \section{Acknowledgments}
723     Support for this project was provided by the National Science
724     Foundation under grant CHE-0134881. T.L. also acknowledges the
725     financial support from center of applied mathematics at University
726     of Notre Dame.
727     \newpage
728    
729     \bibliographystyle{jcp2}
730     \bibliography{langevin}
731    
732     \end{document}