ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Langevin.tex
Revision: 2880
Committed: Thu Jun 22 22:19:02 2006 UTC (18 years ago) by tim
Content type: application/x-tex
File size: 27156 byte(s)
Log Message:
finish conclusion; some format checking correction

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 an excellent alternative to newtonian dynamics, Langevin
8 dynamics, which mimics a simple heat bath with stochastic and
9 dissipative forces, has been applied in a variety of studies. The
10 stochastic treatment of the solvent enables us to carry out
11 substantially longer time simulation. Implicit solvent Langevin
12 dynamics simulation of met-enkephalin not only outperforms explicit
13 solvent simulation on computation efficiency, but also agrees very
14 well with explicit solvent simulation on dynamics
15 properties\cite{Shen2002}. Recently, applying Langevin dynamics with
16 UNRES model, Liow and his coworkers suggest that protein folding
17 pathways can be possibly exploited within a reasonable amount of
18 time\cite{Liwo2005}. The stochastic nature of the Langevin dynamics
19 also enhances the sampling of the system and increases the
20 probability of crossing energy barrier\cite{Banerjee2004, Cui2003}.
21 Combining Langevin dynamics with Kramers's theory, Klimov and
22 Thirumalai identified the free-energy barrier by studying the
23 viscosity dependence of the protein folding rates\cite{Klimov1997}.
24 In order to account for solvent induced interactions missing from
25 implicit solvent model, Kaya incorporated desolvation free energy
26 barrier into implicit coarse-grained solvent model in protein
27 folding/unfolding study and discovered a higher free energy barrier
28 between the native and denatured states. Because of its stability
29 against noise, Langevin dynamics is very suitable for studying
30 remagnetization processes in various
31 systems\cite{Garcia-Palacios1998,Berkov2002,Denisov2003}. For
32 instance, the oscillation power spectrum of nanoparticles from
33 Langevin dynamics simulation has the same peak frequencies for
34 different wave vectors,which recovers the property of magnetic
35 excitations in small finite structures\cite{Berkov2005a}. In an
36 attempt to reduce the computational cost of simulation, multiple
37 time stepping (MTS) methods have been introduced and have been of
38 great interest to macromolecule and protein
39 community\cite{Tuckerman1992}. Relying on the observation that
40 forces between distant atoms generally demonstrate slower
41 fluctuations than forces between close atoms, MTS method are
42 generally implemented by evaluating the slowly fluctuating forces
43 less frequently than the fast ones. Unfortunately, nonlinear
44 instability resulting from increasing timestep in MTS simulation
45 have became a critical obstruction preventing the long time
46 simulation. Due to the coupling to the heat bath, Langevin dynamics
47 has been shown to be able to damp out the resonance artifact more
48 efficiently\cite{Sandu1999}.
49
50 %review langevin/browninan dynamics for arbitrarily shaped rigid body
51 Combining Langevin or Brownian dynamics with rigid body dynamics,
52 one can study the slow processes in biomolecular systems. Modeling
53 the DNA as a chain of rigid spheres beads, which subject to harmonic
54 potentials as well as excluded volume potentials, Mielke and his
55 coworkers discover rapid superhelical stress generations from the
56 stochastic simulation of twin supercoiling DNA with response to
57 induced torques\cite{Mielke2004}. Membrane fusion is another key
58 biological process which controls a variety of physiological
59 functions, such as release of neurotransmitters \textit{etc}. A
60 typical fusion event happens on the time scale of millisecond, which
61 is impracticable to study using all atomistic model with newtonian
62 mechanics. With the help of coarse-grained rigid body model and
63 stochastic dynamics, the fusion pathways were exploited by many
64 researchers\cite{Noguchi2001,Noguchi2002,Shillcock2005}. Due to the
65 difficulty of numerical integration of anisotropy rotation, most of
66 the rigid body models are simply modeled by sphere, cylinder,
67 ellipsoid or other regular shapes in stochastic simulations. In an
68 effort to account for the diffusion anisotropy of the arbitrary
69 particles, Fernandes and de la Torre improved the original Brownian
70 dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
71 incorporating a generalized $6\times6$ diffusion tensor and
72 introducing a simple rotation evolution scheme consisting of three
73 consecutive rotations\cite{Fernandes2002}. Unfortunately, unexpected
74 error and bias are introduced into the system due to the arbitrary
75 order of applying the noncommuting rotation
76 operators\cite{Beard2003}. Based on the observation the momentum
77 relaxation time is much less than the time step, one may ignore the
78 inertia in Brownian dynamics. However, assumption of the zero
79 average acceleration is not always true for cooperative motion which
80 is common in protein motion. An inertial Brownian dynamics (IBD) was
81 proposed to address this issue by adding an inertial correction
82 term\cite{Beard2001}. As a complement to IBD which has a lower bound
83 in time step because of the inertial relaxation time, long-time-step
84 inertial dynamics (LTID) can be used to investigate the inertial
85 behavior of the polymer segments in low friction
86 regime\cite{Beard2001}. LTID can also deal with the rotational
87 dynamics for nonskew bodies without translation-rotation coupling by
88 separating the translation and rotation motion and taking advantage
89 of the analytical solution of hydrodynamics properties. However,
90 typical nonskew bodies like cylinder and ellipsoid are inadequate to
91 represent most complex macromolecule assemblies. These intricate
92 molecules have been represented by a set of beads and their
93 hydrodynamics properties can be calculated using variant
94 hydrodynamic interaction tensors.
95
96 The goal of the present work is to develop a Langevin dynamics
97 algorithm for arbitrary rigid particles by integrating the accurate
98 estimation of friction tensor from hydrodynamics theory into the
99 sophisticated rigid body dynamics.
100
101 \section{Computational Methods{\label{methodSec}}}
102
103 \subsection{\label{introSection:frictionTensor}Friction Tensor}
104 Theoretically, the friction kernel can be determined using the
105 velocity autocorrelation function. However, this approach become
106 impractical when the system become more and more complicate.
107 Instead, various approaches based on hydrodynamics have been
108 developed to calculate the friction coefficients. In general,
109 friction tensor $\Xi$ is a $6\times 6$ matrix given by
110 \[
111 \Xi = \left( {\begin{array}{*{20}c}
112 {\Xi _{}^{tt} } & {\Xi _{}^{rt} } \\
113 {\Xi _{}^{tr} } & {\Xi _{}^{rr} } \\
114 \end{array}} \right).
115 \]
116 Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
117 tensor and rotational resistance (friction) tensor respectively,
118 while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
119 {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
120 particle moves in a fluid, it may experience friction force or
121 torque along the opposite direction of the velocity or angular
122 velocity,
123 \[
124 \left( \begin{array}{l}
125 F_R \\
126 \tau _R \\
127 \end{array} \right) = - \left( {\begin{array}{*{20}c}
128 {\Xi ^{tt} } & {\Xi ^{rt} } \\
129 {\Xi ^{tr} } & {\Xi ^{rr} } \\
130 \end{array}} \right)\left( \begin{array}{l}
131 v \\
132 w \\
133 \end{array} \right)
134 \]
135 where $F_r$ is the friction force and $\tau _R$ is the friction
136 toque.
137
138 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
139
140 For a spherical particle with slip boundary conditions, the
141 translational and rotational friction constant can be calculated
142 from Stoke's law,
143 \[
144 \Xi ^{tt} = \left( {\begin{array}{*{20}c}
145 {6\pi \eta R} & 0 & 0 \\
146 0 & {6\pi \eta R} & 0 \\
147 0 & 0 & {6\pi \eta R} \\
148 \end{array}} \right)
149 \]
150 and
151 \[
152 \Xi ^{rr} = \left( {\begin{array}{*{20}c}
153 {8\pi \eta R^3 } & 0 & 0 \\
154 0 & {8\pi \eta R^3 } & 0 \\
155 0 & 0 & {8\pi \eta R^3 } \\
156 \end{array}} \right)
157 \]
158 where $\eta$ is the viscosity of the solvent and $R$ is the
159 hydrodynamics radius.
160
161 Other non-spherical shape, such as cylinder and ellipsoid
162 \textit{etc}, are widely used as reference for developing new
163 hydrodynamics theory, because their properties can be calculated
164 exactly. In 1936, Perrin extended Stokes's law to general
165 ellipsoids, also called a triaxial ellipsoid, which is given in
166 Cartesian coordinates by\cite{Perrin1934, Perrin1936}
167 \[
168 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
169 }} = 1
170 \]
171 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
172 due to the complexity of the elliptic integral, only the ellipsoid
173 with the restriction of two axes having to be equal, \textit{i.e.}
174 prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
175 exactly. Introducing an elliptic integral parameter $S$ for prolate,
176 \[
177 S = \frac{2}{{\sqrt {a^2 - b^2 } }}\ln \frac{{a + \sqrt {a^2 - b^2
178 } }}{b},
179 \]
180 and oblate,
181 \[
182 S = \frac{2}{{\sqrt {b^2 - a^2 } }}arctg\frac{{\sqrt {b^2 - a^2 }
183 }}{a}
184 \],
185 one can write down the translational and rotational resistance
186 tensors
187 \[
188 \begin{array}{l}
189 \Xi _a^{tt} = 16\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - b^2 )S - 2a}} \\
190 \Xi _b^{tt} = \Xi _c^{tt} = 32\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - 3b^2 )S + 2a}} \\
191 \end{array},
192 \]
193 and
194 \[
195 \begin{array}{l}
196 \Xi _a^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^2 - b^2 )b^2 }}{{2a - b^2 S}} \\
197 \Xi _b^{rr} = \Xi _c^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^4 - b^4 )}}{{(2a^2 - b^2 )S - 2a}} \\
198 \end{array}.
199 \]
200
201 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
202
203 Unlike spherical and other simply shaped molecules, there is no
204 analytical solution for the friction tensor for arbitrarily shaped
205 rigid molecules. The ellipsoid of revolution model and general
206 triaxial ellipsoid model have been used to approximate the
207 hydrodynamic properties of rigid bodies. However, since the mapping
208 from all possible ellipsoidal spaces, $r$-space, to all possible
209 combination of rotational diffusion coefficients, $D$-space, is not
210 unique\cite{Wegener1979} as well as the intrinsic coupling between
211 translational and rotational motion of rigid body, general
212 ellipsoids are not always suitable for modeling arbitrarily shaped
213 rigid molecule. A number of studies have been devoted to determine
214 the friction tensor for irregularly shaped rigid bodies using more
215 advanced method where the molecule of interest was modeled by
216 combinations of spheres(beads)\cite{Carrasco1999} and the
217 hydrodynamics properties of the molecule can be calculated using the
218 hydrodynamic interaction tensor. Let us consider a rigid assembly of
219 $N$ beads immersed in a continuous medium. Due to hydrodynamics
220 interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
221 than its unperturbed velocity $v_i$,
222 \[
223 v'_i = v_i - \sum\limits_{j \ne i} {T_{ij} F_j }
224 \]
225 where $F_i$ is the frictional force, and $T_{ij}$ is the
226 hydrodynamic interaction tensor. The friction force of $i$th bead is
227 proportional to its ``net'' velocity
228 \begin{equation}
229 F_i = \zeta _i v_i - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
230 \label{introEquation:tensorExpression}
231 \end{equation}
232 This equation is the basis for deriving the hydrodynamic tensor. In
233 1930, Oseen and Burgers gave a simple solution to Equation
234 \ref{introEquation:tensorExpression}
235 \begin{equation}
236 T_{ij} = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
237 R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
238 \end{equation}
239 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
240 A second order expression for element of different size was
241 introduced by Rotne and Prager\cite{Rotne1969} and improved by
242 Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
243 \begin{equation}
244 T_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
245 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
246 _i^2 + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
247 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
248 \label{introEquation:RPTensorNonOverlapped}
249 \end{equation}
250 Both of the Equation \ref{introEquation:oseenTensor} and Equation
251 \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
252 \ge \sigma _i + \sigma _j$. An alternative expression for
253 overlapping beads with the same radius, $\sigma$, is given by
254 \begin{equation}
255 T_{ij} = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
256 \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
257 \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
258 \label{introEquation:RPTensorOverlapped}
259 \end{equation}
260
261 To calculate the resistance tensor at an arbitrary origin $O$, we
262 construct a $3N \times 3N$ matrix consisting of $N \times N$
263 $B_{ij}$ blocks
264 \begin{equation}
265 B = \left( {\begin{array}{*{20}c}
266 {B_{11} } & \ldots & {B_{1N} } \\
267 \vdots & \ddots & \vdots \\
268 {B_{N1} } & \cdots & {B_{NN} } \\
269 \end{array}} \right),
270 \end{equation}
271 where $B_{ij}$ is given by
272 \[
273 B_{ij} = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
274 )T_{ij}
275 \]
276 where $\delta _{ij}$ is Kronecker delta function. Inverting the $B$
277 matrix, we obtain
278
279 \[
280 C = B^{ - 1} = \left( {\begin{array}{*{20}c}
281 {C_{11} } & \ldots & {C_{1N} } \\
282 \vdots & \ddots & \vdots \\
283 {C_{N1} } & \cdots & {C_{NN} } \\
284 \end{array}} \right)
285 \]
286 , which can be partitioned into $N \times N$ $3 \times 3$ block
287 $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
288 \[
289 U_i = \left( {\begin{array}{*{20}c}
290 0 & { - z_i } & {y_i } \\
291 {z_i } & 0 & { - x_i } \\
292 { - y_i } & {x_i } & 0 \\
293 \end{array}} \right)
294 \]
295 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
296 bead $i$ and origin $O$. Hence, the elements of resistance tensor at
297 arbitrary origin $O$ can be written as
298 \begin{equation}
299 \begin{array}{l}
300 \Xi _{}^{tt} = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
301 \Xi _{}^{tr} = \Xi _{}^{rt} = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
302 \Xi _{}^{rr} = - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j \\
303 \end{array}
304 \label{introEquation:ResistanceTensorArbitraryOrigin}
305 \end{equation}
306
307 The resistance tensor depends on the origin to which they refer. The
308 proper location for applying friction force is the center of
309 resistance (or center of reaction), at which the trace of rotational
310 resistance tensor, $ \Xi ^{rr}$ reaches a minimum value.
311 Mathematically, the center of resistance is defined as an unique
312 point of the rigid body at which the translation-rotation coupling
313 tensor are symmetric,
314 \begin{equation}
315 \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
316 \label{introEquation:definitionCR}
317 \end{equation}
318 From Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
319 we can easily find out that the translational resistance tensor is
320 origin independent, while the rotational resistance tensor and
321 translation-rotation coupling resistance tensor depend on the
322 origin. Given the resistance tensor at an arbitrary origin $O$, and
323 a vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
324 obtain the resistance tensor at $P$ by
325 \begin{equation}
326 \begin{array}{l}
327 \Xi _P^{tt} = \Xi _O^{tt} \\
328 \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
329 \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{{tr} ^{^T }} \\
330 \end{array}
331 \label{introEquation:resistanceTensorTransformation}
332 \end{equation}
333 where
334 \[
335 U_{OP} = \left( {\begin{array}{*{20}c}
336 0 & { - z_{OP} } & {y_{OP} } \\
337 {z_i } & 0 & { - x_{OP} } \\
338 { - y_{OP} } & {x_{OP} } & 0 \\
339 \end{array}} \right)
340 \]
341 Using Equations \ref{introEquation:definitionCR} and
342 \ref{introEquation:resistanceTensorTransformation}, one can locate
343 the position of center of resistance,
344 \begin{eqnarray*}
345 \left( \begin{array}{l}
346 x_{OR} \\
347 y_{OR} \\
348 z_{OR} \\
349 \end{array} \right) & = &\left( {\begin{array}{*{20}c}
350 {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
351 { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
352 { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
353 \end{array}} \right)^{ - 1} \\
354 & & \left( \begin{array}{l}
355 (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
356 (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
357 (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
358 \end{array} \right) \\
359 \end{eqnarray*}
360
361 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
362 joining center of resistance $R$ and origin $O$.
363
364 \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
365
366 Consider a Langevin equation of motions in generalized coordinates
367 \begin{equation}
368 M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)} + F_{r,i} (t)
369 \label{LDGeneralizedForm}
370 \end{equation}
371 where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
372 and moment of inertial) matrix and $V_i$ is a generalized velocity,
373 $V_i = V_i(v_i,\omega _i)$. The right side of Eq.~
374 (\ref{LDGeneralizedForm}) consists of three generalized forces in
375 lab-fixed frame, systematic force $F_{s,i}$, dissipative force
376 $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
377 system in Newtownian mechanics typically refers to lab-fixed frame,
378 it is also convenient to handle the rotation of rigid body in
379 body-fixed frame. Thus the friction and random forces are calculated
380 in body-fixed frame and converted back to lab-fixed frame by:
381 \[
382 \begin{array}{l}
383 F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
384 F_{r,i}^l (t) = A^T F_{r,i}^b (t) \\
385 \end{array}.
386 \]
387 Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
388 the body-fixed velocity at center of resistance $v_{R,i}^b$ and
389 angular velocity $\omega _i$,
390 \begin{equation}
391 F_{r,i}^b (t) = \left( \begin{array}{l}
392 f_{r,i}^b (t) \\
393 \tau _{r,i}^b (t) \\
394 \end{array} \right) = - \left( {\begin{array}{*{20}c}
395 {\Xi _{R,t} } & {\Xi _{R,c}^T } \\
396 {\Xi _{R,c} } & {\Xi _{R,r} } \\
397 \end{array}} \right)\left( \begin{array}{l}
398 v_{R,i}^b (t) \\
399 \omega _i (t) \\
400 \end{array} \right),
401 \end{equation}
402 while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
403 with zero mean and variance
404 \begin{equation}
405 \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle =
406 \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle =
407 2k_B T\Xi _R \delta (t - t'). \label{randomForce}
408 \end{equation}
409
410 The equation of motion for $v_i$ can be written as
411 \begin{equation}
412 m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
413 f_{r,i}^l (t)
414 \end{equation}
415 Since the frictional force is applied at the center of resistance
416 which generally does not coincide with the center of mass, an extra
417 torque is exerted at the center of mass. Thus, the net body-fixed
418 frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
419 given by
420 \begin{equation}
421 \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
422 \end{equation}
423 where $r_{MR}$ is the vector from the center of mass to the center
424 of the resistance. Instead of integrating angular velocity in
425 lab-fixed frame, we consider the equation of motion of angular
426 momentum in body-fixed frame
427 \begin{equation}
428 \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
429 + \tau _{r,i}^b(t)
430 \end{equation}
431
432 Embedding the friction terms into force and torque, one can
433 integrate the langevin equations of motion for rigid body of
434 arbitrary shape in a velocity-Verlet style 2-part algorithm, where
435 $h= \delta t$:
436
437 {\tt moveA:}
438 \begin{align*}
439 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
440 + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
441 %
442 {\bf r}(t + h) &\leftarrow {\bf r}(t)
443 + h {\bf v}\left(t + h / 2 \right), \\
444 %
445 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
446 + \frac{h}{2} {\bf \tau}^b(t), \\
447 %
448 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
449 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
450 \end{align*}
451
452 In this context, the $\mathrm{rotate}$ function is the reversible
453 product of the three body-fixed rotations,
454 \begin{equation}
455 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
456 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
457 / 2) \cdot \mathsf{G}_x(a_x /2),
458 \end{equation}
459 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
460 rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
461 angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
462 axis $\alpha$,
463 \begin{equation}
464 \mathsf{G}_\alpha( \theta ) = \left\{
465 \begin{array}{lcl}
466 \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
467 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
468 j}(0).
469 \end{array}
470 \right.
471 \end{equation}
472 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
473 rotation matrix. For example, in the small-angle limit, the
474 rotation matrix around the body-fixed x-axis can be approximated as
475 \begin{equation}
476 \mathsf{R}_x(\theta) \approx \left(
477 \begin{array}{ccc}
478 1 & 0 & 0 \\
479 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
480 \theta^2 / 4} \\
481 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
482 \theta^2 / 4}
483 \end{array}
484 \right).
485 \end{equation}
486 All other rotations follow in a straightforward manner.
487
488 After the first part of the propagation, the forces and body-fixed
489 torques are calculated at the new positions and orientations
490
491 {\tt doForces:}
492 \begin{align*}
493 {\bf f}(t + h) &\leftarrow
494 - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
495 %
496 {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
497 \times \frac{\partial V}{\partial {\bf u}}, \\
498 %
499 {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
500 \cdot {\bf \tau}^s(t + h).
501 \end{align*}
502
503 {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
504 $\mathsf{A}$ is calculated in {\tt moveA}. Once the forces and
505 torques have been obtained at the new time step, the velocities can
506 be advanced to the same time value.
507
508 {\tt moveB:}
509 \begin{align*}
510 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
511 \right)
512 + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
513 %
514 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
515 \right)
516 + \frac{h}{2} {\bf \tau}^b(t + h) .
517 \end{align*}
518
519 \section{Results and Discussion}
520
521 The Langevin algorithm described in previous section has been
522 implemented in {\sc oopse}\cite{Meineke2005} and applied to the
523 studies of kinetics and thermodynamic properties in several systems.
524
525 \subsection{Temperature Control}
526
527 As shown in Eq.~\ref{randomForce}, random collisions associated with
528 the solvent's thermal motions is controlled by the external
529 temperature. The capability to maintain the temperature of the whole
530 system was usually used to measure the stability and efficiency of
531 the algorithm. In order to verify the stability of this new
532 algorithm, a series of simulations are performed on system
533 consisiting of 256 SSD water molecules with different viscosities.
534 Initial configuration for the simulations is taken from a 1ns NVT
535 simulation with a cubic box of 19.7166~\AA. All simulation are
536 carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
537 with reference temperature at 300~K. Average temperature as a
538 function of $\eta$ is shown in Table \ref{langevin:viscosity} where
539 the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
540 1$ poise. The better temperature control at higher viscosity can be
541 explained by the finite size effect and relative slow relaxation
542 rate at lower viscosity regime.
543 \begin{table}
544 \caption{Average temperatures from Langevin dynamics simulations of
545 SSD water molecules with reference temperature at 300~K.}
546 \label{langevin:viscosity}
547 \begin{center}
548 \begin{tabular}{|l|l|l|}
549 \hline
550 $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
551 1 & 300.47 & 10.99 \\
552 0.1 & 301.19 & 11.136 \\
553 0.01 & 303.04 & 11.796 \\
554 \hline
555 \end{tabular}
556 \end{center}
557 \end{table}
558
559 Another set of calculation were performed to study the efficiency of
560 temperature control using different temperature coupling schemes.
561 The starting configuration is cooled to 173~K and evolved using NVE,
562 NVT, and Langevin dynamic with time step of 2 fs.
563 Fig.~\ref{langevin:temperature} shows the heating curve obtained as
564 the systems reach equilibrium. The orange curve in
565 Fig.~\ref{langevin:temperature} represents the simulation using
566 Nos\'e-Hoover temperature scaling scheme with thermostat of 5 ps
567 which gives reasonable tight coupling, while the blue one from
568 Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
569 scaling to the desire temperature. In extremely lower friction
570 regime (when $ \eta \approx 0$), Langevin dynamics becomes normal
571 NVE (see green curve in Fig.~\ref{langevin:temperature}) which loses
572 the temperature control ability.
573
574 \begin{figure}
575 \centering
576 \includegraphics[width=\linewidth]{temperature.eps}
577 \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
578 temperature fluctuation versus time.} \label{langevin:temperature}
579 \end{figure}
580
581 \subsection{Langevin Dynamics of Banana Shaped Molecule}
582
583 In order to verify that Langevin dynamics can mimic the dynamics of
584 the systems absent of explicit solvents, we carried out two sets of
585 simulations and compare their dynamic properties.
586
587 \subsubsection{Simulations Contain One Banana Shaped Molecule}
588
589 Fig.~\ref{langevin:oneBanana} shows a snapshot of the simulation
590 made of 256 pentane molecules and one banana shaped molecule at
591 273~K. It has an equivalent implicit solvent system containing only
592 one banana shaped molecule with viscosity of 0.289 center poise. To
593 calculate the hydrodynamic properties of the banana shaped molecule,
594 we create a rough shell model (see Fig.~\ref{langevin:roughShell}),
595 in which the banana shaped molecule is represented as a ``shell''
596 made of 2266 small identical beads with size of 0.3 $\AA$ on the
597 surface. Applying the procedure described in
598 Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
599 identified the center of resistance at $(0, 0.7482, -0.1988)$, as
600 well as the resistance tensor,
601 \[
602 \left( {\begin{array}{*{20}c}
603 0.9261 & 1.310e-14 & -7.292e-15&5.067e-14&0.08585&0.2057\\
604 3.968e-14& 0.9270&-0.007063& 0.08585&6.764e-14&4.846e-14\\
605 -6.561e-16&-0.007063&0.7494&0.2057&4.846e-14&1.5036e-14\\
606 5.067e-14&0.0858&0.2057& 58.64& 8.563e-13&-8.5736\\
607 0.08585&6.764e-14&4.846e-14&1.555e-12&48.30&3.219&\\
608 0.2057&4.846e-14&1.5036e-14&-3.904e-13&3.219&10.7373\\
609 \end{array}} \right).
610 \]
611
612 \begin{figure}
613 \centering
614 \includegraphics[width=\linewidth]{roughShell.eps}
615 \caption[Rough shell model for banana shaped molecule]{Rough shell
616 model for banana shaped molecule.} \label{langevin:roughShell}
617 \end{figure}
618
619 \begin{figure}
620 \centering
621 \includegraphics[width=\linewidth]{oneBanana.eps}
622 \caption[Snapshot from Simulation of One Banana Shaped Molecules and
623 256 Pentane Molecules]{Snapshot from simulation of one Banana shaped
624 molecules and 256 pentane molecules.} \label{langevin:oneBanana}
625 \end{figure}
626
627 \subsubsection{Simulations Contain Two Banana Shaped Molecules}
628
629 \begin{figure}
630 \centering
631 \includegraphics[width=\linewidth]{twoBanana.eps}
632 \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
633 256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
634 molecules and 256 pentane molecules.} \label{langevin:twoBanana}
635 \end{figure}
636
637 \section{Conclusions}