ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Langevin.tex
Revision: 2941
Committed: Mon Jul 17 20:01:05 2006 UTC (17 years, 11 months ago) by tim
Content type: application/x-tex
File size: 27873 byte(s)
Log Message:
references corrections

File Contents

# Content
1
2 \chapter{\label{chapt:langevin}LANGEVIN DYNAMICS FOR RIGID BODIES OF ARBITRARY SHAPE}
3
4 \section{Introduction}
5
6 %applications of langevin dynamics
7 As alternative to Newtonian dynamics, Langevin dynamics, which
8 mimics a simple heat bath with stochastic and dissipative forces,
9 has been applied in a variety of studies. The stochastic treatment
10 of the solvent enables us to carry out substantially longer time
11 simulations. Implicit solvent Langevin dynamics simulations of
12 met-enkephalin not only outperform explicit solvent simulations for
13 computational efficiency, but also agrees very well with explicit
14 solvent simulations for dynamical properties.\cite{Shen2002}
15 Recently, applying Langevin dynamics with the UNRES model, Liow and
16 his coworkers suggest that protein folding pathways can be possibly
17 explored within a reasonable amount of time.\cite{Liwo2005} The
18 stochastic nature of the Langevin dynamics also enhances the
19 sampling of the system and increases the probability of crossing
20 energy barriers.\cite{Banerjee2004, Cui2003} Combining Langevin
21 dynamics with Kramers's theory, Klimov and Thirumalai identified
22 free-energy barriers by studying the viscosity dependence of the
23 protein folding rates.\cite{Klimov1997} In order to account for
24 solvent induced interactions missing from implicit solvent model,
25 Kaya incorporated desolvation free energy barrier into implicit
26 coarse-grained solvent model in protein folding/unfolding studies
27 and discovered a higher free energy barrier between the native and
28 denatured states. Because of its stability against noise, Langevin
29 dynamics is very suitable for studying remagnetization processes in
30 various systems.\cite{Palacios1998,Berkov2002,Denisov2003} For
31 instance, the oscillation power spectrum of nanoparticles from
32 Langevin dynamics simulation has the same peak frequencies for
33 different wave vectors, which recovers the property of magnetic
34 excitations in small finite structures.\cite{Berkov2005a}
35
36 %review langevin/browninan dynamics for arbitrarily shaped rigid body
37 Combining Langevin or Brownian dynamics with rigid body dynamics,
38 one can study slow processes in biomolecular systems. Modeling DNA
39 as a chain of rigid beads, which are subject to harmonic potentials
40 as well as excluded volume potentials, Mielke and his coworkers
41 discovered rapid superhelical stress generations from the stochastic
42 simulation of twin supercoiling DNA with response to induced
43 torques.\cite{Mielke2004} Membrane fusion is another key biological
44 process which controls a variety of physiological functions, such as
45 release of neurotransmitters \textit{etc}. A typical fusion event
46 happens on the time scale of a millisecond, which is impractical to
47 study using atomistic models with newtonian mechanics. With the help
48 of coarse-grained rigid body model and stochastic dynamics, the
49 fusion pathways were explored by many
50 researchers.\cite{Noguchi2001,Noguchi2002,Shillcock2005} Due to the
51 difficulty of numerical integration of anisotropic rotation, most of
52 the rigid body models are simply modeled using spheres, cylinders,
53 ellipsoids or other regular shapes in stochastic simulations. In an
54 effort to account for the diffusion anisotropy of arbitrary
55 particles, Fernandes and de la Torre improved the original Brownian
56 dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
57 incorporating a generalized $6\times6$ diffusion tensor and
58 introducing a simple rotation evolution scheme consisting of three
59 consecutive rotations.\cite{Fernandes2002} Unfortunately, unexpected
60 errors and biases are introduced into the system due to the
61 arbitrary order of applying the noncommuting rotation
62 operators.\cite{Beard2003} Based on the observation the momentum
63 relaxation time is much less than the time step, one may ignore the
64 inertia in Brownian dynamics. However, the assumption of zero
65 average acceleration is not always true for cooperative motion which
66 is common in protein motion. An inertial Brownian dynamics (IBD) was
67 proposed to address this issue by adding an inertial correction
68 term.\cite{Beard2000} As a complement to IBD which has a lower bound
69 in time step because of the inertial relaxation time, long-time-step
70 inertial dynamics (LTID) can be used to investigate the inertial
71 behavior of the polymer segments in low friction
72 regime.\cite{Beard2000} LTID can also deal with the rotational
73 dynamics for nonskew bodies without translation-rotation coupling by
74 separating the translation and rotation motion and taking advantage
75 of the analytical solution of hydrodynamics properties. However,
76 typical nonskew bodies like cylinders and ellipsoids are inadequate
77 to represent most complex macromolecule assemblies. These intricate
78 molecules have been represented by a set of beads and their
79 hydrodynamic properties can be calculated using variants on the
80 standard hydrodynamic interaction tensors.
81
82 The goal of the present work is to develop a Langevin dynamics
83 algorithm for arbitrary-shaped rigid particles by integrating the
84 accurate estimation of friction tensor from hydrodynamics theory
85 into the sophisticated rigid body dynamics algorithms.
86
87 \section{Computational Methods{\label{methodSec}}}
88
89 \subsection{\label{introSection:frictionTensor}Friction Tensor}
90 Theoretically, the friction kernel can be determined using the
91 velocity autocorrelation function. However, this approach becomes
92 impractical when the system becomes more and more complicated.
93 Instead, various approaches based on hydrodynamics have been
94 developed to calculate the friction coefficients. In general, the
95 friction tensor $\Xi$ is a $6\times 6$ matrix given by
96 \[
97 \Xi = \left( {\begin{array}{*{20}c}
98 {\Xi _{}^{tt} } & {\Xi _{}^{rt} } \\
99 {\Xi _{}^{tr} } & {\Xi _{}^{rr} } \\
100 \end{array}} \right).
101 \]
102 Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are $3 \times 3$
103 translational friction tensor and rotational resistance (friction)
104 tensor respectively, while ${\Xi^{tr} }$ is translation-rotation
105 coupling tensor and $ {\Xi^{rt} }$ is rotation-translation coupling
106 tensor. When a particle moves in a fluid, it may experience friction
107 force or torque along the opposite direction of the velocity or
108 angular velocity,
109 \[
110 \left( \begin{array}{l}
111 F_R \\
112 \tau _R \\
113 \end{array} \right) = - \left( {\begin{array}{*{20}c}
114 {\Xi ^{tt} } & {\Xi ^{rt} } \\
115 {\Xi ^{tr} } & {\Xi ^{rr} } \\
116 \end{array}} \right)\left( \begin{array}{l}
117 v \\
118 w \\
119 \end{array} \right)
120 \]
121 where $F_r$ is the friction force and $\tau _R$ is the friction
122 torque.
123
124 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
125
126 For a spherical particle with slip boundary conditions, the
127 translational and rotational friction constant can be calculated
128 from Stoke's law,
129 \[
130 \Xi ^{tt} = \left( {\begin{array}{*{20}c}
131 {6\pi \eta R} & 0 & 0 \\
132 0 & {6\pi \eta R} & 0 \\
133 0 & 0 & {6\pi \eta R} \\
134 \end{array}} \right)
135 \]
136 and
137 \[
138 \Xi ^{rr} = \left( {\begin{array}{*{20}c}
139 {8\pi \eta R^3 } & 0 & 0 \\
140 0 & {8\pi \eta R^3 } & 0 \\
141 0 & 0 & {8\pi \eta R^3 } \\
142 \end{array}} \right)
143 \]
144 where $\eta$ is the viscosity of the solvent and $R$ is the
145 hydrodynamic radius.
146
147 Other non-spherical shapes, such as cylinders and ellipsoids, are
148 widely used as references for developing new hydrodynamics theory,
149 because their properties can be calculated exactly. In 1936, Perrin
150 extended Stokes's law to general ellipsoids, also called a triaxial
151 ellipsoid, which is given in Cartesian coordinates
152 by\cite{Perrin1934, Perrin1936}
153 \[
154 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
155 }} = 1
156 \]
157 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
158 due to the complexity of the elliptic integral, only the ellipsoid
159 with the restriction of two axes being equal, \textit{i.e.}
160 prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
161 exactly. Introducing an elliptic integral parameter $S$ for prolate
162 ellipsoids :
163 \[
164 S = \frac{2}{{\sqrt {a^2 - b^2 } }}\ln \frac{{a + \sqrt {a^2 - b^2
165 } }}{b},
166 \]
167 and oblate ellipsoids:
168 \[
169 S = \frac{2}{{\sqrt {b^2 - a^2 } }}arctg\frac{{\sqrt {b^2 - a^2 }
170 }}{a},
171 \]
172 one can write down the translational and rotational resistance
173 tensors
174 \begin{eqnarray*}
175 \Xi _a^{tt} & = & 16\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - b^2 )S - 2a}}. \\
176 \Xi _b^{tt} & = & \Xi _c^{tt} = 32\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - 3b^2 )S +
177 2a}},
178 \end{eqnarray*}
179 and
180 \begin{eqnarray*}
181 \Xi _a^{rr} & = & \frac{{32\pi }}{3}\eta \frac{{(a^2 - b^2 )b^2 }}{{2a - b^2 S}}, \\
182 \Xi _b^{rr} & = & \Xi _c^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^4 - b^4 )}}{{(2a^2 - b^2 )S - 2a}}.
183 \end{eqnarray*}
184
185 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
186
187 Unlike spherical and other simply shaped molecules, there is no
188 analytical solution for the friction tensor for arbitrarily shaped
189 rigid molecules. The ellipsoid of revolution model and general
190 triaxial ellipsoid model have been used to approximate the
191 hydrodynamic properties of rigid bodies. However, since the mapping
192 from all possible ellipsoidal spaces, $r$-space, to all possible
193 combination of rotational diffusion coefficients, $D$-space, is not
194 unique\cite{Wegener1979} as well as the intrinsic coupling between
195 translational and rotational motion of rigid bodies, general
196 ellipsoids are not always suitable for modeling arbitrarily shaped
197 rigid molecules. A number of studies have been devoted to
198 determining the friction tensor for irregularly shaped rigid bodies
199 using more advanced methods where the molecule of interest was
200 modeled by a combinations of spheres\cite{Carrasco1999} and the
201 hydrodynamics properties of the molecule can be calculated using the
202 hydrodynamic interaction tensor. Let us consider a rigid assembly of
203 $N$ beads immersed in a continuous medium. Due to hydrodynamic
204 interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
205 than its unperturbed velocity $v_i$,
206 \[
207 v'_i = v_i - \sum\limits_{j \ne i} {T_{ij} F_j }
208 \]
209 where $F_i$ is the frictional force, and $T_{ij}$ is the
210 hydrodynamic interaction tensor. The friction force of $i$th bead is
211 proportional to its ``net'' velocity
212 \begin{equation}
213 F_i = \zeta _i v_i - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
214 \label{introEquation:tensorExpression}
215 \end{equation}
216 This equation is the basis for deriving the hydrodynamic tensor. In
217 1930, Oseen and Burgers gave a simple solution to
218 Eq.~\ref{introEquation:tensorExpression}
219 \begin{equation}
220 T_{ij} = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
221 R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
222 \end{equation}
223 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
224 A second order expression for element of different size was
225 introduced by Rotne and Prager\cite{Rotne1969} and improved by
226 Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
227 \begin{equation}
228 T_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
229 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
230 _i^2 + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
231 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
232 \label{introEquation:RPTensorNonOverlapped}
233 \end{equation}
234 Both of the Eq.~\ref{introEquation:oseenTensor} and
235 Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption
236 $R_{ij} \ge \sigma _i + \sigma _j$. An alternative expression for
237 overlapping beads with the same radius, $\sigma$, is given by
238 \begin{equation}
239 T_{ij} = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
240 \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
241 \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
242 \label{introEquation:RPTensorOverlapped}
243 \end{equation}
244 To calculate the resistance tensor at an arbitrary origin $O$, we
245 construct a $3N \times 3N$ matrix consisting of $N \times N$
246 $B_{ij}$ blocks
247 \begin{equation}
248 B = \left( {\begin{array}{*{20}c}
249 {B_{11} } & \ldots & {B_{1N} } \\
250 \vdots & \ddots & \vdots \\
251 {B_{N1} } & \cdots & {B_{NN} } \\
252 \end{array}} \right),
253 \end{equation}
254 where $B_{ij}$ is given by
255 \[
256 B_{ij} = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
257 )T_{ij}
258 \]
259 where $\delta _{ij}$ is the Kronecker delta function. Inverting the
260 $B$ matrix, we obtain
261 \[
262 C = B^{ - 1} = \left( {\begin{array}{*{20}c}
263 {C_{11} } & \ldots & {C_{1N} } \\
264 \vdots & \ddots & \vdots \\
265 {C_{N1} } & \cdots & {C_{NN} } \\
266 \end{array}} \right),
267 \]
268 which can be partitioned into $N \times N$ $3 \times 3$ block
269 $C_{ij}$. With the help of $C_{ij}$ and the skew matrix $U_i$
270 \[
271 U_i = \left( {\begin{array}{*{20}c}
272 0 & { - z_i } & {y_i } \\
273 {z_i } & 0 & { - x_i } \\
274 { - y_i } & {x_i } & 0 \\
275 \end{array}} \right)
276 \]
277 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
278 bead $i$ and origin $O$, the elements of resistance tensor at
279 arbitrary origin $O$ can be written as
280 \begin{eqnarray}
281 \Xi _{}^{tt} & = & \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
282 \Xi _{}^{tr} & = & \Xi _{}^{rt} = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
283 \Xi _{}^{rr} & = & - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
284 \label{introEquation:ResistanceTensorArbitraryOrigin}
285 \end{eqnarray}
286 The resistance tensor depends on the origin to which they refer. The
287 proper location for applying the friction force is the center of
288 resistance (or center of reaction), at which the trace of rotational
289 resistance tensor, $ \Xi ^{rr}$ reaches a minimum value.
290 Mathematically, the center of resistance is defined as an unique
291 point of the rigid body at which the translation-rotation coupling
292 tensors are symmetric,
293 \begin{equation}
294 \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
295 \label{introEquation:definitionCR}
296 \end{equation}
297 From Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
298 we can easily derive that the translational resistance tensor is
299 origin independent, while the rotational resistance tensor and
300 translation-rotation coupling resistance tensor depend on the
301 origin. Given the resistance tensor at an arbitrary origin $O$, and
302 a vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
303 obtain the resistance tensor at $P$ by
304 \begin{equation}
305 \begin{array}{l}
306 \Xi _P^{tt} = \Xi _O^{tt} \\
307 \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
308 \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{{tr} ^{^T }} \\
309 \end{array}
310 \label{introEquation:resistanceTensorTransformation}
311 \end{equation}
312 where
313 \[
314 U_{OP} = \left( {\begin{array}{*{20}c}
315 0 & { - z_{OP} } & {y_{OP} } \\
316 {z_i } & 0 & { - x_{OP} } \\
317 { - y_{OP} } & {x_{OP} } & 0 \\
318 \end{array}} \right)
319 \]
320 Using Eq.~\ref{introEquation:definitionCR} and
321 Eq.~\ref{introEquation:resistanceTensorTransformation}, one can
322 locate the position of center of resistance,
323 \begin{eqnarray*}
324 \left( \begin{array}{l}
325 x_{OR} \\
326 y_{OR} \\
327 z_{OR} \\
328 \end{array} \right) & = &\left( {\begin{array}{*{20}c}
329 {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
330 { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
331 { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
332 \end{array}} \right)^{ - 1} \\
333 & & \left( \begin{array}{l}
334 (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
335 (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
336 (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
337 \end{array} \right) \\
338 \end{eqnarray*}
339 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
340 joining center of resistance $R$ and origin $O$.
341
342 \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
343
344 Consider the Langevin equations of motion in generalized coordinates
345 \begin{equation}
346 M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)} + F_{r,i} (t)
347 \label{LDGeneralizedForm}
348 \end{equation}
349 where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
350 and moment of inertial) matrix and $V_i$ is a generalized velocity,
351 $V_i = V_i(v_i,\omega _i)$. The right side of
352 Eq.~\ref{LDGeneralizedForm} consists of three generalized forces in
353 lab-fixed frame, systematic force $F_{s,i}$, dissipative force
354 $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
355 system in Newtownian mechanics typically refers to lab-fixed frame,
356 it is also convenient to handle the rotation of rigid body in
357 body-fixed frame. Thus the friction and random forces are calculated
358 in body-fixed frame and converted back to lab-fixed frame by:
359 \[
360 \begin{array}{l}
361 F_{f,i}^l (t) = Q^T F_{f,i}^b (t), \\
362 F_{r,i}^l (t) = Q^T F_{r,i}^b (t). \\
363 \end{array}
364 \]
365 Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
366 the body-fixed velocity at center of resistance $v_{R,i}^b$ and
367 angular velocity $\omega _i$
368 \begin{equation}
369 F_{r,i}^b (t) = \left( \begin{array}{l}
370 f_{r,i}^b (t) \\
371 \tau _{r,i}^b (t) \\
372 \end{array} \right) = - \left( {\begin{array}{*{20}c}
373 {\Xi _{R,t} } & {\Xi _{R,c}^T } \\
374 {\Xi _{R,c} } & {\Xi _{R,r} } \\
375 \end{array}} \right)\left( \begin{array}{l}
376 v_{R,i}^b (t) \\
377 \omega _i (t) \\
378 \end{array} \right),
379 \end{equation}
380 while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
381 with zero mean and variance
382 \begin{equation}
383 \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle =
384 \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle =
385 2k_B T\Xi _R \delta (t - t'). \label{randomForce}
386 \end{equation}
387 The equation of motion for $v_i$ can be written as
388 \begin{equation}
389 m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
390 f_{r,i}^l (t)
391 \end{equation}
392 Since the frictional force is applied at the center of resistance
393 which generally does not coincide with the center of mass, an extra
394 torque is exerted at the center of mass. Thus, the net body-fixed
395 frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
396 given by
397 \begin{equation}
398 \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
399 \end{equation}
400 where $r_{MR}$ is the vector from the center of mass to the center
401 of the resistance. Instead of integrating the angular velocity in
402 lab-fixed frame, we consider the equation of angular momentum in
403 body-fixed frame
404 \begin{equation}
405 \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
406 + \tau _{r,i}^b(t)
407 \end{equation}
408 Embedding the friction terms into force and torque, one can
409 integrate the langevin equations of motion for rigid body of
410 arbitrary shape in a velocity-Verlet style 2-part algorithm, where
411 $h= \delta t$:
412
413 {\tt moveA:}
414 \begin{align*}
415 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
416 + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
417 %
418 {\bf r}(t + h) &\leftarrow {\bf r}(t)
419 + h {\bf v}\left(t + h / 2 \right), \\
420 %
421 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
422 + \frac{h}{2} {\bf \tau}^b(t), \\
423 %
424 \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
425 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
426 \end{align*}
427 In this context, the $\mathrm{rotate}$ function is the reversible
428 product of the three body-fixed rotations,
429 \begin{equation}
430 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
431 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
432 / 2) \cdot \mathsf{G}_x(a_x /2),
433 \end{equation}
434 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
435 rotates both the rotation matrix ($\mathsf{Q}$) and the body-fixed
436 angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
437 axis $\alpha$,
438 \begin{equation}
439 \mathsf{G}_\alpha( \theta ) = \left\{
440 \begin{array}{lcl}
441 \mathsf{Q}(t) & \leftarrow & \mathsf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
442 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
443 j}(0).
444 \end{array}
445 \right.
446 \end{equation}
447 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
448 rotation matrix. For example, in the small-angle limit, the
449 rotation matrix around the body-fixed x-axis can be approximated as
450 \begin{equation}
451 \mathsf{R}_x(\theta) \approx \left(
452 \begin{array}{ccc}
453 1 & 0 & 0 \\
454 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
455 \theta^2 / 4} \\
456 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
457 \theta^2 / 4}
458 \end{array}
459 \right).
460 \end{equation}
461 All other rotations follow in a straightforward manner. After the
462 first part of the propagation, the forces and body-fixed torques are
463 calculated at the new positions and orientations
464
465 {\tt doForces:}
466 \begin{align*}
467 {\bf f}(t + h) &\leftarrow
468 - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
469 %
470 {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
471 \times \frac{\partial V}{\partial {\bf u}}, \\
472 %
473 {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{Q}(t + h)
474 \cdot {\bf \tau}^s(t + h).
475 \end{align*}
476 Once the forces and torques have been obtained at the new time step,
477 the velocities can be advanced to the same time value.
478
479 {\tt moveB:}
480 \begin{align*}
481 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
482 \right)
483 + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
484 %
485 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
486 \right)
487 + \frac{h}{2} {\bf \tau}^b(t + h) .
488 \end{align*}
489
490 \section{Results and Discussion}
491
492 The Langevin algorithm described in previous section has been
493 implemented in {\sc oopse}\cite{Meineke2005} and applied to studies
494 of the static and dynamic properties in several systems.
495
496 \subsection{Temperature Control}
497
498 As shown in Eq.~\ref{randomForce}, random collisions associated with
499 the solvent's thermal motions is controlled by the external
500 temperature. The capability to maintain the temperature of the whole
501 system was usually used to measure the stability and efficiency of
502 the algorithm. In order to verify the stability of this new
503 algorithm, a series of simulations are performed on system
504 consisiting of 256 SSD water molecules with different viscosities.
505 The initial configuration for the simulations is taken from a 1ns
506 NVT simulation with a cubic box of 19.7166~\AA. All simulation are
507 carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
508 with reference temperature at 300~K. The average temperature as a
509 function of $\eta$ is shown in Table \ref{langevin:viscosity} where
510 the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
511 1$ poise. The better temperature control at higher viscosity can be
512 explained by the finite size effect and relative slow relaxation
513 rate at lower viscosity regime.
514 \begin{table}
515 \caption{AVERAGE TEMPERATURES FROM LANGEVIN DYNAMICS SIMULATIONS OF
516 SSD WATER MOLECULES WITH REFERENCE TEMPERATURE AT 300~K.}
517 \label{langevin:viscosity}
518 \begin{center}
519 \begin{tabular}{lll}
520 \hline
521 $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
522 \hline
523 1 & 300.47 & 10.99 \\
524 0.1 & 301.19 & 11.136 \\
525 0.01 & 303.04 & 11.796 \\
526 \hline
527 \end{tabular}
528 \end{center}
529 \end{table}
530
531 Another set of calculations were performed to study the efficiency of
532 temperature control using different temperature coupling schemes.
533 The starting configuration is cooled to 173~K and evolved using NVE,
534 NVT, and Langevin dynamic with time step of 2 fs.
535 Fig.~\ref{langevin:temperature} shows the heating curve obtained as
536 the systems reach equilibrium. The orange curve in
537 Fig.~\ref{langevin:temperature} represents the simulation using
538 Nos\'e-Hoover temperature scaling scheme with thermostat of 5 ps
539 which gives reasonable tight coupling, while the blue one from
540 Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
541 scaling to the desire temperature. When $ \eta = 0$, Langevin dynamics becomes normal
542 NVE (see orange curve in Fig.~\ref{langevin:temperature}) which
543 loses the temperature control ability.
544
545 \begin{figure}
546 \centering
547 \includegraphics[width=\linewidth]{temperature.eps}
548 \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
549 temperature fluctuation versus time.} \label{langevin:temperature}
550 \end{figure}
551
552 \subsection{Langevin Dynamics of Banana Shaped Molecules}
553
554 In order to verify that Langevin dynamics can mimic the dynamics of
555 the systems absent of explicit solvents, we carried out two sets of
556 simulations and compare their dynamic properties.
557 Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
558 made of 256 pentane molecules and two banana shaped molecules at
559 273~K. It has an equivalent implicit solvent system containing only
560 two banana shaped molecules with viscosity of 0.289 center poise. To
561 calculate the hydrodynamic properties of the banana shaped molecule,
562 we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
563 in which the banana shaped molecule is represented as a ``shell''
564 made of 2266 small identical beads with size of 0.3 \AA on the
565 surface. Applying the procedure described in
566 Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
567 identified the center of resistance at (0 $\rm{\AA}$, 0.7482 $\rm{\AA}$,
568 -0.1988 $\rm{\AA}$), as well as the resistance tensor,
569 \[
570 \left( {\begin{array}{*{20}c}
571 0.9261 & 0 & 0&0&0.08585&0.2057\\
572 0& 0.9270&-0.007063& 0.08585&0&0\\
573 0&-0.007063&0.7494&0.2057&0&0\\
574 0&0.0858&0.2057& 58.64& 0&0\\
575 0.08585&0&0&0&48.30&3.219&\\
576 0.2057&0&0&0&3.219&10.7373\\
577 \end{array}} \right).
578 \]
579 %\[
580 %\left( {\begin{array}{*{20}c}
581 %0.9261 & 1.310e-14 & -7.292e-15&5.067e-14&0.08585&0.2057\\
582 %3.968e-14& 0.9270&-0.007063& 0.08585&6.764e-14&4.846e-14\\
583 %-6.561e-16&-0.007063&0.7494&0.2057&4.846e-14&1.5036e-14\\
584 %5.067e-14&0.0858&0.2057& 58.64& 8.563e-13&-8.5736\\
585 %0.08585&6.764e-14&4.846e-14&1.555e-12&48.30&3.219&\\
586 %0.2057&4.846e-14&1.5036e-14&-3.904e-13&3.219&10.7373\\
587 %\end{array}} \right).
588 %\]
589
590 Curves of the velocity auto-correlation functions in
591 Fig.~\ref{langevin:vacf} were shown to match each other very well.
592 However, because of the stochastic nature, simulation using Langevin
593 dynamics was shown to decay slightly faster than MD. In order to
594 study the rotational motion of the molecules, we also calculated the
595 auto-correlation function of the principle axis of the second GB
596 particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
597 probably due to the reason that the viscosity using in the
598 simulations only partially preserved the dynamics of the system.
599
600 \begin{figure}
601 \centering
602 \includegraphics[width=\linewidth]{roughShell.eps}
603 \caption[Rough shell model for banana shaped molecule]{Rough shell
604 model for banana shaped molecule.} \label{langevin:roughShell}
605 \end{figure}
606
607 \begin{figure}
608 \centering
609 \includegraphics[width=\linewidth]{twoBanana.eps}
610 \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
611 256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
612 molecules and 256 pentane molecules.} \label{langevin:twoBanana}
613 \end{figure}
614
615 \begin{figure}
616 \centering
617 \includegraphics[width=\linewidth]{vacf.eps}
618 \caption[Plots of Velocity Auto-correlation Functions]{Velocity
619 auto-correlation functions of NVE (explicit solvent) in blue and
620 Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
621 \end{figure}
622
623 \begin{figure}
624 \centering
625 \includegraphics[width=\linewidth]{uacf.eps}
626 \caption[Auto-correlation functions of the principle axis of the
627 middle GB particle]{Auto-correlation functions of the principle axis
628 of the middle GB particle of NVE (blue) and Langevin dynamics
629 (red).} \label{langevin:uacf}
630 \end{figure}
631
632 \section{Conclusions}
633
634 We have presented a new Langevin algorithm by incorporating the
635 hydrodynamics properties of arbitrary shaped molecules into an
636 advanced symplectic integration scheme. The temperature control
637 ability of this algorithm was demonstrated by a set of simulations
638 with different viscosities. It was also shown to have significant
639 advantage of producing rapid thermal equilibration over
640 Nos\'{e}-Hoover method. Further studies in systems involving banana
641 shaped molecules illustrated that the dynamic properties could be
642 preserved by using this new algorithm as an implicit solvent model.