ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
Revision: 2746
Committed: Wed May 3 16:44:46 2006 UTC (18 years, 4 months ago) by tim
Content type: application/x-tex
File size: 22599 byte(s)
Log Message:
initial draft

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     \author{Teng Lin and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
28     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     As an excellent alternative to newtonian dynamics, Langevin
53     dynamics, which mimics a simple heat bath with stochastic and
54     dissipative forces, has been applied in a variety of studies. The
55     stochastic treatment of the solvent enables us to carry out
56     substantially longer time simulation. Implicit solvent Langevin
57     dynamics simulation of met-enkephalin not only outperforms explicit
58     solvent simulation on computation efficiency, but also agrees very
59     well with explicit solvent simulation on dynamics
60     properties\cite{Shen2002}. Recently, applying Langevin dynamics with
61     UNRES model, Liow and his coworkers suggest that protein folding
62     pathways can be possibly exploited within a reasonable amount of
63     time\cite{Liwo2005}. The stochastic nature of the Langevin dynamics
64     also enhances the sampling of the system and increases the
65     probability of crossing energy barrier\cite{Banerjee2004, Cui2003}.
66     Combining Langevin dynamics with Kramers's theory, Klimov and
67     Thirumalai identified the free-energy barrier by studying the
68     viscosity dependence of the protein folding rates\cite{Klimov1997}.
69     In order to account for solvent induced interactions missing from
70     implicit solvent model, Kaya incorporated desolvation free energy
71     barrier into implicit coarse-grained solvent model in protein
72     folding/unfolding study and discovered a higher free energy barrier
73     between the native and denatured states. Because of its stability
74     against noise, Langevin dynamics is very suitable for studying
75     remagnetization processes in various
76     systems\cite{Garcia-Palacios1998,Berkov2002,Denisov2003}. For
77     instance, the oscillation power spectrum of nanoparticles from
78     Langevin dynamics simulation has the same peak frequencies for
79     different wave vectors,which recovers the property of magnetic
80     excitations in small finite structures\cite{Berkov2005a}. In an
81     attempt to reduce the computational cost of simulation, multiple
82     time stepping (MTS) methods have been introduced and have been of
83     great interest to macromolecule and protein
84     community\cite{Tuckerman1992}. Relying on the observation that
85     forces between distant atoms generally demonstrate slower
86     fluctuations than forces between close atoms, MTS method are
87     generally implemented by evaluating the slowly fluctuating forces
88     less frequently than the fast ones. Unfortunately, nonlinear
89     instability resulting from increasing timestep in MTS simulation
90     have became a critical obstruction preventing the long time
91     simulation. Due to the coupling to the heat bath, Langevin dynamics
92     has been shown to be able to damp out the resonance artifact more
93     efficiently\cite{Sandu1999}.
94    
95     %review rigid body dynamics
96     Rigid bodies are frequently involved in the modeling of different
97     areas, from engineering, physics, to chemistry. For example,
98     missiles and vehicle are usually modeled by rigid bodies. The
99     movement of the objects in 3D gaming engine or other physics
100     simulator is governed by the rigid body dynamics. In molecular
101     simulation, rigid body is used to simplify the model in
102     protein-protein docking study{\cite{Gray2003}}.
103    
104     It is very important to develop stable and efficient methods to
105     integrate the equations of motion of orientational degrees of
106     freedom. Euler angles are the nature choice to describe the
107     rotational degrees of freedom. However, due to its singularity, the
108     numerical integration of corresponding equations of motion is very
109     inefficient and inaccurate. Although an alternative integrator using
110     different sets of Euler angles can overcome this difficulty\cite{},
111     the computational penalty and the lost of angular momentum
112     conservation still remain. In 1977, a singularity free
113     representation utilizing quaternions was developed by
114     Evans\cite{Evans1977}. Unfortunately, this approach suffer from the
115     nonseparable Hamiltonian resulted from quaternion representation,
116     which prevents the symplectic algorithm to be utilized. Another
117     different approach is to apply holonomic constraints to the atoms
118     belonging to the rigid body\cite{}. Each atom moves independently
119     under the normal forces deriving from potential energy and
120     constraint forces which are used to guarantee the rigidness.
121     However, due to their iterative nature, SHAKE and Rattle algorithm
122     converge very slowly when the number of constraint increases.
123    
124     The break through in geometric literature suggests that, in order to
125     develop a long-term integration scheme, one should preserve the
126     geometric structure of the flow. Matubayasi and Nakahara developed a
127     time-reversible integrator for rigid bodies in quaternion
128     representation. Although it is not symplectic, this integrator still
129     demonstrates a better long-time energy conservation than traditional
130     methods because of the time-reversible nature. Extending
131     Trotter-Suzuki to general system with a flat phase space, Miller and
132     his colleagues devised an novel symplectic, time-reversible and
133     volume-preserving integrator in quaternion representation. However,
134     all of the integrators in quaternion representation suffer from the
135     computational penalty of constructing a rotation matrix from
136     quaternions to evolve coordinates and velocities at every time step.
137     An alternative integration scheme utilizing rotation matrix directly
138     is RSHAKE , in which a conjugate momentum to rotation matrix is
139     introduced to re-formulate the Hamiltonian's equation and the
140     Hamiltonian is evolved in a constraint manifold by iteratively
141     satisfying the orthogonality constraint. However, RSHAKE is
142     inefficient because of the iterative procedure. An extremely
143     efficient integration scheme in rotation matrix representation,
144     which also preserves the same structural properties of the
145     Hamiltonian flow as Miller's integrator, is proposed by Dullweber,
146     Leimkuhler and McLachlan (DLM)\cite{Dullweber1997}.
147    
148     %review langevin/browninan dynamics for arbitrarily shaped rigid body
149     Combining Langevin or Brownian dynamics with rigid body dynamics,
150     one can study the slow processes in biomolecular systems. Modeling
151     the DNA as a chain of rigid spheres beads, which subject to harmonic
152     potentials as well as excluded volume potentials, Mielke and his
153     coworkers discover rapid superhelical stress generations from the
154     stochastic simulation of twin supercoiling DNA with response to
155     induced torques\cite{Mielke2004}. Membrane fusion is another key
156     biological process which controls a variety of physiological
157     functions, such as release of neurotransmitters \textit{etc}. A
158     typical fusion event happens on the time scale of millisecond, which
159     is impracticable to study using all atomistic model with newtonian
160     mechanics. With the help of coarse-grained rigid body model and
161     stochastic dynamics, the fusion pathways were exploited by many
162     researchers\cite{Noguchi2001,Noguchi2002,Shillcock2005}. Due to the
163     difficulty of numerical integration of anisotropy rotation, most of
164     the rigid body models are simply modeled by sphere, cylinder,
165     ellipsoid or other regular shapes in stochastic simulations. In an
166     effort to account for the diffusion anisotropy of the arbitrary
167     particles, Fernandes and de la Torre improved the original Brownian
168     dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
169     incorporating a generalized $6\times6$ diffusion tensor and
170     introducing a simple rotation evolution scheme consisting of three
171     consecutive rotations\cite{Fernandes2002}. Unfortunately, unexpected
172     error and bias are introduced into the system due to the arbitrary
173     order of applying the noncommuting rotation
174     operators\cite{Beard2003}. Based on the observation the momentum
175     relaxation time is much less than the time step, one may ignore the
176     inertia in Brownian dynamics. However, assumption of the zero
177     average acceleration is not always true for cooperative motion which
178     is common in protein motion. An inertial Brownian dynamics (IBD) was
179     proposed to address this issue by adding an inertial correction
180     term\cite{Beard2001}. As a complement to IBD which has a lower bound
181     in time step because of the inertial relaxation time, long-time-step
182     inertial dynamics (LTID) can be used to investigate the inertial
183     behavior of the polymer segments in low friction
184     regime\cite{Beard2001}. LTID can also deal with the rotational
185     dynamics for nonskew bodies without translation-rotation coupling by
186     separating the translation and rotation motion and taking advantage
187     of the analytical solution of hydrodynamics properties. However,
188     typical nonskew bodies like cylinder and ellipsoid are inadequate to
189     represent most complex macromolecule assemblies. These intricate
190     molecules have been represented by a set of beads and their
191     hydrodynamics properties can be calculated using variant
192     hydrodynamic interaction tensors.
193    
194     The goal of the present work is to develop a Langevin dynamics
195     algorithm for arbitrary rigid particles by integrating the accurate
196     estimation of friction tensor from hydrodynamics theory into the
197     sophisticated rigid body dynamics.
198    
199     \section{Method{\label{methodSec}}}
200    
201     \subsection{Friction Tensor}
202    
203     For an arbitrary rigid body moves in a fluid, it may experience
204     friction force $f_r$ or friction torque $\tau _r$ along the opposite
205     direction of the velocity $v$ or angular velocity $\omega$ at
206     arbitrary origin $P$,
207     \begin{equation}
208     \left( \begin{array}{l}
209     f_r \\
210     \tau _r \\
211     \end{array} \right) = - \left( {\begin{array}{*{20}c}
212     {\Xi _{P,t} } & {\Xi _{P,c}^T } \\
213     {\Xi _{P,c} } & {\Xi _{P,r} } \\
214     \end{array}} \right)\left( \begin{array}{l}
215     \nu \\
216     \omega \\
217     \end{array} \right)
218     \end{equation}
219     where $\Xi _{P,t}t$ is the translation friction tensor, $\Xi _{P,r}$
220     is the rotational friction tensor and $\Xi _{P,c}$ is the
221     translation-rotation coupling tensor. The procedure of calculating
222     friction tensor using hydrodynamic tensor and comparison between
223     bead model and shell model were elaborated by Carrasco \textit{et
224     al}\cite{Carrasco1999}. An important property of the friction tensor
225     is that the translational friction tensor is independent of origin
226     while the rotational and coupling are sensitive to the choice of the
227     origin \cite{Brenner1967}, which can be described by
228     \begin{equation}
229     \begin{array}{c}
230     \Xi _{P,t} = \Xi _{O,t} = \Xi _t \\
231     \Xi _{P,c} = \Xi _{O,c} - r_{OP} \times \Xi _t \\
232     \Xi _{P,r} = \Xi _{O,r} - r_{OP} \times \Xi _t \times r_{OP} + \Xi _{O,c} \times r_{OP} - r_{OP} \times \Xi _{O,c}^T \\
233     \end{array}
234     \end{equation}
235     Where $O$ is another origin and $r_{OP}$ is the vector joining $O$
236     and $P$. It is also worthy of mention that both of translational and
237     rotational frictional tensors are always symmetric. In contrast,
238     coupling tensor is only symmetric at center of reaction:
239     \begin{equation}
240     \Xi _{R,c} = \Xi _{R,c}^T
241     \end{equation}
242     The proper location for applying friction force is the center of
243     reaction, at which the trace of rotational resistance tensor reaches
244     minimum.
245    
246     \subsection{Rigid body dynamics}
247    
248     The Hamiltonian of rigid body can be separated in terms of potential
249     energy $V(r,A)$ and kinetic energy $T(p,\pi)$,
250     \[
251     H = V(r,A) + T(v,\pi )
252     \]
253     A second-order symplectic method is now obtained by the composition
254     of the flow maps,
255     \[
256     \varphi _{\Delta t} = \varphi _{\Delta t/2,V} \circ \varphi
257     _{\Delta t,T} \circ \varphi _{\Delta t/2,V}.
258     \]
259     Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
260     sub-flows which corresponding to force and torque respectively,
261     \[
262     \varphi _{\Delta t/2,V} = \varphi _{\Delta t/2,F} \circ \varphi
263     _{\Delta t/2,\tau }.
264     \]
265     Since the associated operators of $\varphi _{\Delta t/2,F} $ and
266     $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
267     order inside $\varphi _{\Delta t/2,V}$ does not matter.
268    
269     Furthermore, kinetic potential can be separated to translational
270     kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
271     \begin{equation}
272     T(p,\pi ) =T^t (p) + T^r (\pi ).
273     \end{equation}
274     where $ T^t (p) = \frac{1}{2}v^T m v $ and $T^r (\pi )$ is defined
275     by \ref{introEquation:rotationalKineticRB}. Therefore, the
276     corresponding flow maps are given by
277     \[
278     \varphi _{\Delta t,T} = \varphi _{\Delta t,T^t } \circ \varphi
279     _{\Delta t,T^r }.
280     \]
281     The free rigid body is an example of Lie-Poisson system with
282     Hamiltonian function
283     \begin{equation}
284     T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
285     \label{introEquation:rotationalKineticRB}
286     \end{equation}
287     where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
288     Lie-Poisson structure matrix,
289     \begin{equation}
290     J(\pi ) = \left( {\begin{array}{*{20}c}
291     0 & {\pi _3 } & { - \pi _2 } \\
292     { - \pi _3 } & 0 & {\pi _1 } \\
293     {\pi _2 } & { - \pi _1 } & 0 \\
294     \end{array}} \right)
295     \end{equation}
296     Thus, the dynamics of free rigid body is governed by
297     \begin{equation}
298     \frac{d}{{dt}}\pi = J(\pi )\nabla _\pi T^r (\pi )
299     \end{equation}
300     One may notice that each $T_i^r$ in Equation
301     \ref{introEquation:rotationalKineticRB} can be solved exactly. For
302     instance, the equations of motion due to $T_1^r$ are given by
303     \begin{equation}
304     \frac{d}{{dt}}\pi = R_1 \pi ,\frac{d}{{dt}}A = AR_1
305     \label{introEqaution:RBMotionSingleTerm}
306     \end{equation}
307     where
308     \[ R_1 = \left( {\begin{array}{*{20}c}
309     0 & 0 & 0 \\
310     0 & 0 & {\pi _1 } \\
311     0 & { - \pi _1 } & 0 \\
312     \end{array}} \right).
313     \]
314     The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
315     \[
316     \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),A(\Delta t) =
317     A(0)e^{\Delta tR_1 }
318     \]
319     with
320     \[
321     e^{\Delta tR_1 } = \left( {\begin{array}{*{20}c}
322     0 & 0 & 0 \\
323     0 & {\cos \theta _1 } & {\sin \theta _1 } \\
324     0 & { - \sin \theta _1 } & {\cos \theta _1 } \\
325     \end{array}} \right),\theta _1 = \frac{{\pi _1 }}{{I_1 }}\Delta t.
326     \]
327     To reduce the cost of computing expensive functions in $e^{\Delta
328     tR_1 }$, we can use Cayley transformation,
329     \[
330     e^{\Delta tR_1 } \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
331     )
332     \]
333     The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
334     manner.
335    
336     In order to construct a second-order symplectic method, we split the
337     angular kinetic Hamiltonian function into five terms
338     \[
339     T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
340     ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
341     (\pi _1 )
342     \].
343     Concatenating flows corresponding to these five terms, we can obtain
344     the flow map for free rigid body,
345     \[
346     \varphi _{\Delta t,T^r } = \varphi _{\Delta t/2,\pi _1 } \circ
347     \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 }
348     \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi
349     _1 }.
350     \]
351    
352     The equations of motion corresponding to potential energy and
353     kinetic energy are listed in the below table,
354     \begin{center}
355     \begin{tabular}{|l|l|}
356     \hline
357     % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
358     Potential & Kinetic \\
359     $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
360     $\frac{d}{{dt}}p = - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
361     $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
362     $ \frac{d}{{dt}}\pi = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi = \pi \times I^{ - 1} \pi$\\
363     \hline
364     \end{tabular}
365     \end{center}
366    
367     Finally, we obtain the overall symplectic flow maps for free moving
368     rigid body
369     \begin{align*}
370     \varphi _{\Delta t} = &\varphi _{\Delta t/2,F} \circ \varphi _{\Delta t/2,\tau } \circ \\
371     &\varphi _{\Delta t,T^t } \circ \varphi _{\Delta t/2,\pi _1 } \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 } \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi _1 } \circ \\
372     &\varphi _{\Delta t/2,\tau } \circ \varphi _{\Delta t/2,F} .\\
373     \label{introEquation:overallRBFlowMaps}
374     \end{align*}
375    
376     \subsection{Langevin dynamics for rigid particles of arbitrary shape}
377    
378     Consider a Langevin equation of motions in generalized coordinates
379     \begin{equation}
380     M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)} + F_{r,i} (t)
381     \label{LDGeneralizedForm}
382     \end{equation}
383     where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
384     and moment of inertial) matrix and $V_i$ is a generalized velocity,
385     $V_i = V_i(v_i,\omega _i)$. The right side of Eq.
386     (\ref{LDGeneralizedForm}) consists of three generalized forces in
387     lab-fixed frame, systematic force $F_{s,i}$, dissipative force
388     $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
389     system in Newtownian mechanics typically refers to lab-fixed frame,
390     it is also convenient to handle the rotation of rigid body in
391     body-fixed frame. Thus the friction and random forces are calculated
392     in body-fixed frame and converted back to lab-fixed frame by:
393     \[
394     \begin{array}{l}
395     F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
396     F_{r,i}^l (t) = A^T F_{r,i}^b (t) \\
397     \end{array}.
398     \]
399     Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
400     the body-fixed velocity at center of resistance $v_{R,i}^b$ and
401     angular velocity $\omega _i$,
402     \begin{equation}
403     F_{r,i}^b (t) = \left( \begin{array}{l}
404     f_{r,i}^b (t) \\
405     \tau _{r,i}^b (t) \\
406     \end{array} \right) = - \left( {\begin{array}{*{20}c}
407     {\Xi _{R,t} } & {\Xi _{R,c}^T } \\
408     {\Xi _{R,c} } & {\Xi _{R,r} } \\
409     \end{array}} \right)\left( \begin{array}{l}
410     v_{R,i}^b (t) \\
411     \omega _i (t) \\
412     \end{array} \right),
413     \end{equation}
414     while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
415     with zero mean and variance
416     \begin{equation}
417     \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle =
418     \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle =
419     2k_B T\Xi _R \delta (t - t').
420     \end{equation}
421     The equation of motion for $v_i$ can be written as
422     \begin{equation}
423     m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
424     f_{r,i}^l (t)
425     \end{equation}
426     Since the frictional force is applied at the center of resistance
427     which generally does not coincide with the center of mass, an extra
428     torque is exerted at the center of mass. Thus, the net body-fixed
429     frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
430     given by
431     \begin{equation}
432     \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
433     \end{equation}
434     where $r_{MR}$ is the vector from the center of mass to the center
435     of the resistance. Instead of integrating angular velocity in
436     lab-fixed frame, we consider the equation of motion of angular
437     momentum in body-fixed frame
438     \begin{equation}
439     \dot \pi _i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b
440     (t) + \tau _{r,i}^b(t)
441     \end{equation}
442    
443     Embedding the friction terms into force and torque, one can
444     integrate the langevin equations of motion for rigid body of
445     arbitrary shape in a velocity-Verlet style 2-part algorithm, where
446     $h= \delta t$:
447    
448     {\tt part one:}
449     \begin{align*}
450     v_i (t + h/2) &\leftarrow v_i (t) + \frac{{hf_{t,i}^l (t)}}{{2m_i }} \\
451     \pi _i (t + h/2) &\leftarrow \pi _i (t) + \frac{{h\tau _{t,i}^b (t)}}{2} \\
452     r_i (t + h) &\leftarrow r_i (t) + hv_i (t + h/2) \\
453     A_i (t + h) &\leftarrow rotate\left( {h\pi _i (t + h/2)I^{ - 1} } \right) \\
454     \end{align*}
455     In this context, the $\mathrm{rotate}$ function is the reversible
456     product of five consecutive body-fixed rotations,
457     \begin{equation}
458     \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
459     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
460     / 2) \cdot \mathsf{G}_x(a_x /2),
461     \end{equation}
462     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
463     rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
464     angular momentum ($\pi$) by an angle $\theta$ around body-fixed axis
465     $\alpha$,
466     \begin{equation}
467     \mathsf{G}_\alpha( \theta ) = \left\{
468     \begin{array}{lcl}
469     \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
470     {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
471     j}(0).
472     \end{array}
473     \right.
474     \end{equation}
475     $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
476     rotation matrix. For example, in the small-angle limit, the
477     rotation matrix around the body-fixed x-axis can be approximated as
478     \begin{equation}
479     \mathsf{R}_x(\theta) \approx \left(
480     \begin{array}{ccc}
481     1 & 0 & 0 \\
482     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
483     \theta^2 / 4} \\
484     0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
485     \theta^2 / 4}
486     \end{array}
487     \right).
488     \end{equation}
489     All other rotations follow in a straightforward manner.
490    
491     After the first part of the propagation, the friction and random
492     forces are generated at the center of resistance in body-fixed frame
493     and converted back into lab-fixed frame
494     \[
495     f_{t,i}^l (t + h) = - \left( {\frac{{\partial V}}{{\partial r_i }}}
496     \right)_{r_i (t + h)} + A_i^T (t + h)[F_{f,i}^b (t + h) + F_{r,i}^b
497     (t + h)],
498     \]
499     while the system torque in lab-fixed frame is transformed into
500     body-fixed frame,
501     \[
502     \tau _{t,i}^b (t + h) = A\tau _{s,i}^l (t) + \tau _{n,i}^b (t) +
503     \tau _{r,i}^b (t).
504     \]
505     Once the forces and torques have been obtained at the new time step,
506     the velocities can be advanced to the same time value.
507    
508     {\tt part two:}
509     \begin{align*}
510     v_i (t) &\leftarrow v_i (t + h/2) + \frac{{hf_{t,i}^l (t + h)}}{{2m_i }} \\
511     \pi _i (t) &\leftarrow \pi _i (t + h/2) + \frac{{h\tau _{t,i}^b (t + h)}}{2} \\
512     \end{align*}
513    
514     \section{Results and discussion}
515    
516     \subsection{}
517    
518     \subsection{Lipid bilayer}
519    
520     \subsection{Liquid crystal}
521    
522     \section{Conclusions}
523    
524     \section{Acknowledgments}
525     Support for this project was provided by the National Science
526     Foundation under grant CHE-0134881. T.L. also acknowledges the
527     financial support from center of applied mathematics at University
528     of Notre Dame.
529     \newpage
530    
531     \bibliographystyle{jcp2}
532     \bibliography{langevin}
533    
534     \end{document}