ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Langevin.tex
Revision: 2851
Committed: Sun Jun 11 02:06:01 2006 UTC (18 years, 1 month ago) by tim
Content type: application/x-tex
File size: 25179 byte(s)
Log Message:
seperate Langevin chapter

File Contents

# Content
1
2 \chapter{\label{chapt:methodology}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 rigid body dynamics
51 Rigid bodies are frequently involved in the modeling of different
52 areas, from engineering, physics, to chemistry. For example,
53 missiles and vehicle are usually modeled by rigid bodies. The
54 movement of the objects in 3D gaming engine or other physics
55 simulator is governed by the rigid body dynamics. In molecular
56 simulation, rigid body is used to simplify the model in
57 protein-protein docking study{\cite{Gray2003}}.
58
59 It is very important to develop stable and efficient methods to
60 integrate the equations of motion of orientational degrees of
61 freedom. Euler angles are the nature choice to describe the
62 rotational degrees of freedom. However, due to its singularity, the
63 numerical integration of corresponding equations of motion is very
64 inefficient and inaccurate. Although an alternative integrator using
65 different sets of Euler angles can overcome this difficulty\cite{},
66 the computational penalty and the lost of angular momentum
67 conservation still remain. In 1977, a singularity free
68 representation utilizing quaternions was developed by
69 Evans\cite{Evans1977}. Unfortunately, this approach suffer from the
70 nonseparable Hamiltonian resulted from quaternion representation,
71 which prevents the symplectic algorithm to be utilized. Another
72 different approach is to apply holonomic constraints to the atoms
73 belonging to the rigid body\cite{}. Each atom moves independently
74 under the normal forces deriving from potential energy and
75 constraint forces which are used to guarantee the rigidness.
76 However, due to their iterative nature, SHAKE and Rattle algorithm
77 converge very slowly when the number of constraint increases.
78
79 The break through in geometric literature suggests that, in order to
80 develop a long-term integration scheme, one should preserve the
81 geometric structure of the flow. Matubayasi and Nakahara developed a
82 time-reversible integrator for rigid bodies in quaternion
83 representation. Although it is not symplectic, this integrator still
84 demonstrates a better long-time energy conservation than traditional
85 methods because of the time-reversible nature. Extending
86 Trotter-Suzuki to general system with a flat phase space, Miller and
87 his colleagues devised an novel symplectic, time-reversible and
88 volume-preserving integrator in quaternion representation. However,
89 all of the integrators in quaternion representation suffer from the
90 computational penalty of constructing a rotation matrix from
91 quaternions to evolve coordinates and velocities at every time step.
92 An alternative integration scheme utilizing rotation matrix directly
93 is RSHAKE , in which a conjugate momentum to rotation matrix is
94 introduced to re-formulate the Hamiltonian's equation and the
95 Hamiltonian is evolved in a constraint manifold by iteratively
96 satisfying the orthogonality constraint. However, RSHAKE is
97 inefficient because of the iterative procedure. An extremely
98 efficient integration scheme in rotation matrix representation,
99 which also preserves the same structural properties of the
100 Hamiltonian flow as Miller's integrator, is proposed by Dullweber,
101 Leimkuhler and McLachlan (DLM).
102
103 %review langevin/browninan dynamics for arbitrarily shaped rigid body
104 Combining Langevin or Brownian dynamics with rigid body dynamics,
105 one can study the slow processes in biomolecular systems. Modeling
106 the DNA as a chain of rigid spheres beads, which subject to harmonic
107 potentials as well as excluded volume potentials, Mielke and his
108 coworkers discover rapid superhelical stress generations from the
109 stochastic simulation of twin supercoiling DNA with response to
110 induced torques\cite{Mielke2004}. Membrane fusion is another key
111 biological process which controls a variety of physiological
112 functions, such as release of neurotransmitters \textit{etc}. A
113 typical fusion event happens on the time scale of millisecond, which
114 is impracticable to study using all atomistic model with newtonian
115 mechanics. With the help of coarse-grained rigid body model and
116 stochastic dynamics, the fusion pathways were exploited by many
117 researchers\cite{Noguchi2001,Noguchi2002,Shillcock2005}. Due to the
118 difficulty of numerical integration of anisotropy rotation, most of
119 the rigid body models are simply modeled by sphere, cylinder,
120 ellipsoid or other regular shapes in stochastic simulations. In an
121 effort to account for the diffusion anisotropy of the arbitrary
122 particles, Fernandes and de la Torre improved the original Brownian
123 dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
124 incorporating a generalized $6\times6$ diffusion tensor and
125 introducing a simple rotation evolution scheme consisting of three
126 consecutive rotations\cite{Fernandes2002}. Unfortunately, unexpected
127 error and bias are introduced into the system due to the arbitrary
128 order of applying the noncommuting rotation
129 operators\cite{Beard2003}. Based on the observation the momentum
130 relaxation time is much less than the time step, one may ignore the
131 inertia in Brownian dynamics. However, assumption of the zero
132 average acceleration is not always true for cooperative motion which
133 is common in protein motion. An inertial Brownian dynamics (IBD) was
134 proposed to address this issue by adding an inertial correction
135 term\cite{Beard2001}. As a complement to IBD which has a lower bound
136 in time step because of the inertial relaxation time, long-time-step
137 inertial dynamics (LTID) can be used to investigate the inertial
138 behavior of the polymer segments in low friction
139 regime\cite{Beard2001}. LTID can also deal with the rotational
140 dynamics for nonskew bodies without translation-rotation coupling by
141 separating the translation and rotation motion and taking advantage
142 of the analytical solution of hydrodynamics properties. However,
143 typical nonskew bodies like cylinder and ellipsoid are inadequate to
144 represent most complex macromolecule assemblies. These intricate
145 molecules have been represented by a set of beads and their
146 hydrodynamics properties can be calculated using variant
147 hydrodynamic interaction tensors.
148
149 The goal of the present work is to develop a Langevin dynamics
150 algorithm for arbitrary rigid particles by integrating the accurate
151 estimation of friction tensor from hydrodynamics theory into the
152 sophisticated rigid body dynamics.
153
154 \section{Method{\label{methodSec}}}
155
156 \subsection{\label{introSection:frictionTensor} Friction Tensor}
157 Theoretically, the friction kernel can be determined using velocity
158 autocorrelation function. However, this approach become impractical
159 when the system become more and more complicate. Instead, various
160 approaches based on hydrodynamics have been developed to calculate
161 the friction coefficients. The friction effect is isotropic in
162 Equation, $\zeta$ can be taken as a scalar. In general, friction
163 tensor $\Xi$ is a $6\times 6$ matrix given by
164 \[
165 \Xi = \left( {\begin{array}{*{20}c}
166 {\Xi _{}^{tt} } & {\Xi _{}^{rt} } \\
167 {\Xi _{}^{tr} } & {\Xi _{}^{rr} } \\
168 \end{array}} \right).
169 \]
170 Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
171 tensor and rotational resistance (friction) tensor respectively,
172 while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
173 {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
174 particle moves in a fluid, it may experience friction force or
175 torque along the opposite direction of the velocity or angular
176 velocity,
177 \[
178 \left( \begin{array}{l}
179 F_R \\
180 \tau _R \\
181 \end{array} \right) = - \left( {\begin{array}{*{20}c}
182 {\Xi ^{tt} } & {\Xi ^{rt} } \\
183 {\Xi ^{tr} } & {\Xi ^{rr} } \\
184 \end{array}} \right)\left( \begin{array}{l}
185 v \\
186 w \\
187 \end{array} \right)
188 \]
189 where $F_r$ is the friction force and $\tau _R$ is the friction
190 toque.
191
192 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
193
194 For a spherical particle, the translational and rotational friction
195 constant can be calculated from Stoke's law,
196 \[
197 \Xi ^{tt} = \left( {\begin{array}{*{20}c}
198 {6\pi \eta R} & 0 & 0 \\
199 0 & {6\pi \eta R} & 0 \\
200 0 & 0 & {6\pi \eta R} \\
201 \end{array}} \right)
202 \]
203 and
204 \[
205 \Xi ^{rr} = \left( {\begin{array}{*{20}c}
206 {8\pi \eta R^3 } & 0 & 0 \\
207 0 & {8\pi \eta R^3 } & 0 \\
208 0 & 0 & {8\pi \eta R^3 } \\
209 \end{array}} \right)
210 \]
211 where $\eta$ is the viscosity of the solvent and $R$ is the
212 hydrodynamics radius.
213
214 Other non-spherical shape, such as cylinder and ellipsoid
215 \textit{etc}, are widely used as reference for developing new
216 hydrodynamics theory, because their properties can be calculated
217 exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
218 also called a triaxial ellipsoid, which is given in Cartesian
219 coordinates by\cite{Perrin1934, Perrin1936}
220 \[
221 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
222 }} = 1
223 \]
224 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
225 due to the complexity of the elliptic integral, only the ellipsoid
226 with the restriction of two axes having to be equal, \textit{i.e.}
227 prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
228 exactly. Introducing an elliptic integral parameter $S$ for prolate,
229 \[
230 S = \frac{2}{{\sqrt {a^2 - b^2 } }}\ln \frac{{a + \sqrt {a^2 - b^2
231 } }}{b},
232 \]
233 and oblate,
234 \[
235 S = \frac{2}{{\sqrt {b^2 - a^2 } }}arctg\frac{{\sqrt {b^2 - a^2 }
236 }}{a}
237 \],
238 one can write down the translational and rotational resistance
239 tensors
240 \[
241 \begin{array}{l}
242 \Xi _a^{tt} = 16\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - b^2 )S - 2a}} \\
243 \Xi _b^{tt} = \Xi _c^{tt} = 32\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - 3b^2 )S + 2a}} \\
244 \end{array},
245 \]
246 and
247 \[
248 \begin{array}{l}
249 \Xi _a^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^2 - b^2 )b^2 }}{{2a - b^2 S}} \\
250 \Xi _b^{rr} = \Xi _c^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^4 - b^4 )}}{{(2a^2 - b^2 )S - 2a}} \\
251 \end{array}.
252 \]
253
254 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shape}}
255
256 Unlike spherical and other regular shaped molecules, there is not
257 analytical solution for friction tensor of any arbitrary shaped
258 rigid molecules. The ellipsoid of revolution model and general
259 triaxial ellipsoid model have been used to approximate the
260 hydrodynamic properties of rigid bodies. However, since the mapping
261 from all possible ellipsoidal space, $r$-space, to all possible
262 combination of rotational diffusion coefficients, $D$-space is not
263 unique\cite{Wegener1979} as well as the intrinsic coupling between
264 translational and rotational motion of rigid body, general ellipsoid
265 is not always suitable for modeling arbitrarily shaped rigid
266 molecule. A number of studies have been devoted to determine the
267 friction tensor for irregularly shaped rigid bodies using more
268 advanced method where the molecule of interest was modeled by
269 combinations of spheres(beads)\cite{Carrasco1999} and the
270 hydrodynamics properties of the molecule can be calculated using the
271 hydrodynamic interaction tensor. Let us consider a rigid assembly of
272 $N$ beads immersed in a continuous medium. Due to hydrodynamics
273 interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
274 than its unperturbed velocity $v_i$,
275 \[
276 v'_i = v_i - \sum\limits_{j \ne i} {T_{ij} F_j }
277 \]
278 where $F_i$ is the frictional force, and $T_{ij}$ is the
279 hydrodynamic interaction tensor. The friction force of $i$th bead is
280 proportional to its ``net'' velocity
281 \begin{equation}
282 F_i = \zeta _i v_i - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
283 \label{introEquation:tensorExpression}
284 \end{equation}
285 This equation is the basis for deriving the hydrodynamic tensor. In
286 1930, Oseen and Burgers gave a simple solution to Equation
287 \ref{introEquation:tensorExpression}
288 \begin{equation}
289 T_{ij} = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
290 R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
291 \end{equation}
292 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
293 A second order expression for element of different size was
294 introduced by Rotne and Prager\cite{Rotne1969} and improved by
295 Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
296 \begin{equation}
297 T_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
298 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
299 _i^2 + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
300 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
301 \label{introEquation:RPTensorNonOverlapped}
302 \end{equation}
303 Both of the Equation \ref{introEquation:oseenTensor} and Equation
304 \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
305 \ge \sigma _i + \sigma _j$. An alternative expression for
306 overlapping beads with the same radius, $\sigma$, is given by
307 \begin{equation}
308 T_{ij} = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
309 \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
310 \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
311 \label{introEquation:RPTensorOverlapped}
312 \end{equation}
313
314 To calculate the resistance tensor at an arbitrary origin $O$, we
315 construct a $3N \times 3N$ matrix consisting of $N \times N$
316 $B_{ij}$ blocks
317 \begin{equation}
318 B = \left( {\begin{array}{*{20}c}
319 {B_{11} } & \ldots & {B_{1N} } \\
320 \vdots & \ddots & \vdots \\
321 {B_{N1} } & \cdots & {B_{NN} } \\
322 \end{array}} \right),
323 \end{equation}
324 where $B_{ij}$ is given by
325 \[
326 B_{ij} = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
327 )T_{ij}
328 \]
329 where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
330 $B$, we obtain
331
332 \[
333 C = B^{ - 1} = \left( {\begin{array}{*{20}c}
334 {C_{11} } & \ldots & {C_{1N} } \\
335 \vdots & \ddots & \vdots \\
336 {C_{N1} } & \cdots & {C_{NN} } \\
337 \end{array}} \right)
338 \]
339 , which can be partitioned into $N \times N$ $3 \times 3$ block
340 $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
341 \[
342 U_i = \left( {\begin{array}{*{20}c}
343 0 & { - z_i } & {y_i } \\
344 {z_i } & 0 & { - x_i } \\
345 { - y_i } & {x_i } & 0 \\
346 \end{array}} \right)
347 \]
348 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
349 bead $i$ and origin $O$. Hence, the elements of resistance tensor at
350 arbitrary origin $O$ can be written as
351 \begin{equation}
352 \begin{array}{l}
353 \Xi _{}^{tt} = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
354 \Xi _{}^{tr} = \Xi _{}^{rt} = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
355 \Xi _{}^{rr} = - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j \\
356 \end{array}
357 \label{introEquation:ResistanceTensorArbitraryOrigin}
358 \end{equation}
359
360 The resistance tensor depends on the origin to which they refer. The
361 proper location for applying friction force is the center of
362 resistance (reaction), at which the trace of rotational resistance
363 tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
364 resistance is defined as an unique point of the rigid body at which
365 the translation-rotation coupling tensor are symmetric,
366 \begin{equation}
367 \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
368 \label{introEquation:definitionCR}
369 \end{equation}
370 Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
371 we can easily find out that the translational resistance tensor is
372 origin independent, while the rotational resistance tensor and
373 translation-rotation coupling resistance tensor depend on the
374 origin. Given resistance tensor at an arbitrary origin $O$, and a
375 vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
376 obtain the resistance tensor at $P$ by
377 \begin{equation}
378 \begin{array}{l}
379 \Xi _P^{tt} = \Xi _O^{tt} \\
380 \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
381 \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{{tr} ^{^T }} \\
382 \end{array}
383 \label{introEquation:resistanceTensorTransformation}
384 \end{equation}
385 where
386 \[
387 U_{OP} = \left( {\begin{array}{*{20}c}
388 0 & { - z_{OP} } & {y_{OP} } \\
389 {z_i } & 0 & { - x_{OP} } \\
390 { - y_{OP} } & {x_{OP} } & 0 \\
391 \end{array}} \right)
392 \]
393 Using Equations \ref{introEquation:definitionCR} and
394 \ref{introEquation:resistanceTensorTransformation}, one can locate
395 the position of center of resistance,
396 \begin{eqnarray*}
397 \left( \begin{array}{l}
398 x_{OR} \\
399 y_{OR} \\
400 z_{OR} \\
401 \end{array} \right) & = &\left( {\begin{array}{*{20}c}
402 {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
403 { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
404 { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
405 \end{array}} \right)^{ - 1} \\
406 & & \left( \begin{array}{l}
407 (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
408 (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
409 (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
410 \end{array} \right) \\
411 \end{eqnarray*}
412
413 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
414 joining center of resistance $R$ and origin $O$.
415
416 \subsection{Langevin dynamics for rigid particles of arbitrary shape}
417
418 Consider a Langevin equation of motions in generalized coordinates
419 \begin{equation}
420 M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)} + F_{r,i} (t)
421 \label{LDGeneralizedForm}
422 \end{equation}
423 where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
424 and moment of inertial) matrix and $V_i$ is a generalized velocity,
425 $V_i = V_i(v_i,\omega _i)$. The right side of Eq.
426 (\ref{LDGeneralizedForm}) consists of three generalized forces in
427 lab-fixed frame, systematic force $F_{s,i}$, dissipative force
428 $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
429 system in Newtownian mechanics typically refers to lab-fixed frame,
430 it is also convenient to handle the rotation of rigid body in
431 body-fixed frame. Thus the friction and random forces are calculated
432 in body-fixed frame and converted back to lab-fixed frame by:
433 \[
434 \begin{array}{l}
435 F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
436 F_{r,i}^l (t) = A^T F_{r,i}^b (t) \\
437 \end{array}.
438 \]
439 Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
440 the body-fixed velocity at center of resistance $v_{R,i}^b$ and
441 angular velocity $\omega _i$,
442 \begin{equation}
443 F_{r,i}^b (t) = \left( \begin{array}{l}
444 f_{r,i}^b (t) \\
445 \tau _{r,i}^b (t) \\
446 \end{array} \right) = - \left( {\begin{array}{*{20}c}
447 {\Xi _{R,t} } & {\Xi _{R,c}^T } \\
448 {\Xi _{R,c} } & {\Xi _{R,r} } \\
449 \end{array}} \right)\left( \begin{array}{l}
450 v_{R,i}^b (t) \\
451 \omega _i (t) \\
452 \end{array} \right),
453 \end{equation}
454 while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
455 with zero mean and variance
456 \begin{equation}
457 \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle =
458 \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle =
459 2k_B T\Xi _R \delta (t - t').
460 \end{equation}
461
462 The equation of motion for $v_i$ can be written as
463 \begin{equation}
464 m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
465 f_{r,i}^l (t)
466 \end{equation}
467 Since the frictional force is applied at the center of resistance
468 which generally does not coincide with the center of mass, an extra
469 torque is exerted at the center of mass. Thus, the net body-fixed
470 frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
471 given by
472 \begin{equation}
473 \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
474 \end{equation}
475 where $r_{MR}$ is the vector from the center of mass to the center
476 of the resistance. Instead of integrating angular velocity in
477 lab-fixed frame, we consider the equation of motion of angular
478 momentum in body-fixed frame
479 \begin{equation}
480 \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
481 + \tau _{r,i}^b(t)
482 \end{equation}
483
484 Embedding the friction terms into force and torque, one can
485 integrate the langevin equations of motion for rigid body of
486 arbitrary shape in a velocity-Verlet style 2-part algorithm, where
487 $h= \delta t$:
488
489 {\tt moveA:}
490 \begin{align*}
491 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
492 + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
493 %
494 {\bf r}(t + h) &\leftarrow {\bf r}(t)
495 + h {\bf v}\left(t + h / 2 \right), \\
496 %
497 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
498 + \frac{h}{2} {\bf \tau}^b(t), \\
499 %
500 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
501 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
502 \end{align*}
503
504 In this context, the $\mathrm{rotate}$ function is the reversible
505 product of the three body-fixed rotations,
506 \begin{equation}
507 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
508 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
509 / 2) \cdot \mathsf{G}_x(a_x /2),
510 \end{equation}
511 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
512 rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
513 angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
514 axis $\alpha$,
515 \begin{equation}
516 \mathsf{G}_\alpha( \theta ) = \left\{
517 \begin{array}{lcl}
518 \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
519 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
520 j}(0).
521 \end{array}
522 \right.
523 \end{equation}
524 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
525 rotation matrix. For example, in the small-angle limit, the
526 rotation matrix around the body-fixed x-axis can be approximated as
527 \begin{equation}
528 \mathsf{R}_x(\theta) \approx \left(
529 \begin{array}{ccc}
530 1 & 0 & 0 \\
531 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
532 \theta^2 / 4} \\
533 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
534 \theta^2 / 4}
535 \end{array}
536 \right).
537 \end{equation}
538 All other rotations follow in a straightforward manner.
539
540 After the first part of the propagation, the forces and body-fixed
541 torques are calculated at the new positions and orientations
542
543 {\tt doForces:}
544 \begin{align*}
545 {\bf f}(t + h) &\leftarrow
546 - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
547 %
548 {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
549 \times \frac{\partial V}{\partial {\bf u}}, \\
550 %
551 {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
552 \cdot {\bf \tau}^s(t + h).
553 \end{align*}
554
555 {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
556 $\mathsf{A}$ is calculated in {\tt moveA}. Once the forces and
557 torques have been obtained at the new time step, the velocities can
558 be advanced to the same time value.
559
560 {\tt moveB:}
561 \begin{align*}
562 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
563 \right)
564 + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
565 %
566 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
567 \right)
568 + \frac{h}{2} {\bf \tau}^b(t + h) .
569 \end{align*}
570
571 \section{Results and discussion}
572
573 \section{Conclusions}