ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Langevin.tex
Revision: 2858
Committed: Tue Jun 13 02:08:23 2006 UTC (18 years ago) by tim
Content type: application/x-tex
File size: 22583 byte(s)
Log Message:
adding some images

File Contents

# User Rev Content
1 tim 2851
2 tim 2854 \chapter{\label{chapt:methodology}LANGEVIN DYNAMICS FOR RIGID BODIES OF ARBITRARY SHAPE}
3 tim 2851
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 tim 2858 \section{Computational methods{\label{methodSec}}}
102 tim 2851
103 tim 2853 \subsection{\label{introSection:frictionTensor}Friction Tensor}
104 tim 2851 Theoretically, the friction kernel can be determined using velocity
105     autocorrelation function. However, this approach become impractical
106     when the system become more and more complicate. Instead, various
107     approaches based on hydrodynamics have been developed to calculate
108     the friction coefficients. The friction effect is isotropic in
109     Equation, $\zeta$ can be taken as a scalar. In general, friction
110     tensor $\Xi$ is a $6\times 6$ matrix given by
111     \[
112     \Xi = \left( {\begin{array}{*{20}c}
113     {\Xi _{}^{tt} } & {\Xi _{}^{rt} } \\
114     {\Xi _{}^{tr} } & {\Xi _{}^{rr} } \\
115     \end{array}} \right).
116     \]
117     Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
118     tensor and rotational resistance (friction) tensor respectively,
119     while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
120     {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
121     particle moves in a fluid, it may experience friction force or
122     torque along the opposite direction of the velocity or angular
123     velocity,
124     \[
125     \left( \begin{array}{l}
126     F_R \\
127     \tau _R \\
128     \end{array} \right) = - \left( {\begin{array}{*{20}c}
129     {\Xi ^{tt} } & {\Xi ^{rt} } \\
130     {\Xi ^{tr} } & {\Xi ^{rr} } \\
131     \end{array}} \right)\left( \begin{array}{l}
132     v \\
133     w \\
134     \end{array} \right)
135     \]
136     where $F_r$ is the friction force and $\tau _R$ is the friction
137     toque.
138    
139     \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
140    
141     For a spherical particle, the translational and rotational friction
142     constant can be calculated 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 ellipsoid,
165     also called a triaxial ellipsoid, which is given in Cartesian
166     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 Shape}}
202    
203     Unlike spherical and other regular shaped molecules, there is not
204     analytical solution for friction tensor of any arbitrary 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 space, $r$-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 ellipsoid
212     is not always suitable for modeling arbitrarily shaped rigid
213     molecule. A number of studies have been devoted to determine the
214     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 matrix
277     $B$, 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 (reaction), at which the trace of rotational resistance
310     tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
311     resistance is defined as an unique point of the rigid body at which
312     the translation-rotation coupling tensor are symmetric,
313     \begin{equation}
314     \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
315     \label{introEquation:definitionCR}
316     \end{equation}
317     Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
318     we can easily find out that the translational resistance tensor is
319     origin independent, while the rotational resistance tensor and
320     translation-rotation coupling resistance tensor depend on the
321     origin. Given resistance tensor at an arbitrary origin $O$, and a
322     vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
323     obtain the resistance tensor at $P$ by
324     \begin{equation}
325     \begin{array}{l}
326     \Xi _P^{tt} = \Xi _O^{tt} \\
327     \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
328     \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{{tr} ^{^T }} \\
329     \end{array}
330     \label{introEquation:resistanceTensorTransformation}
331     \end{equation}
332     where
333     \[
334     U_{OP} = \left( {\begin{array}{*{20}c}
335     0 & { - z_{OP} } & {y_{OP} } \\
336     {z_i } & 0 & { - x_{OP} } \\
337     { - y_{OP} } & {x_{OP} } & 0 \\
338     \end{array}} \right)
339     \]
340     Using Equations \ref{introEquation:definitionCR} and
341     \ref{introEquation:resistanceTensorTransformation}, one can locate
342     the position of center of resistance,
343     \begin{eqnarray*}
344     \left( \begin{array}{l}
345     x_{OR} \\
346     y_{OR} \\
347     z_{OR} \\
348     \end{array} \right) & = &\left( {\begin{array}{*{20}c}
349     {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
350     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
351     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
352     \end{array}} \right)^{ - 1} \\
353     & & \left( \begin{array}{l}
354     (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
355     (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
356     (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
357     \end{array} \right) \\
358     \end{eqnarray*}
359    
360     where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
361     joining center of resistance $R$ and origin $O$.
362    
363 tim 2858 \subsection{Langevin dynamics for rigid particles of arbitrary shape\label{LDRB}}
364 tim 2851
365     Consider a Langevin equation of motions in generalized coordinates
366     \begin{equation}
367     M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)} + F_{r,i} (t)
368     \label{LDGeneralizedForm}
369     \end{equation}
370     where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
371     and moment of inertial) matrix and $V_i$ is a generalized velocity,
372 tim 2857 $V_i = V_i(v_i,\omega _i)$. The right side of Eq.~
373 tim 2851 (\ref{LDGeneralizedForm}) consists of three generalized forces in
374     lab-fixed frame, systematic force $F_{s,i}$, dissipative force
375     $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
376     system in Newtownian mechanics typically refers to lab-fixed frame,
377     it is also convenient to handle the rotation of rigid body in
378     body-fixed frame. Thus the friction and random forces are calculated
379     in body-fixed frame and converted back to lab-fixed frame by:
380     \[
381     \begin{array}{l}
382     F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
383     F_{r,i}^l (t) = A^T F_{r,i}^b (t) \\
384     \end{array}.
385     \]
386     Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
387     the body-fixed velocity at center of resistance $v_{R,i}^b$ and
388     angular velocity $\omega _i$,
389     \begin{equation}
390     F_{r,i}^b (t) = \left( \begin{array}{l}
391     f_{r,i}^b (t) \\
392     \tau _{r,i}^b (t) \\
393     \end{array} \right) = - \left( {\begin{array}{*{20}c}
394     {\Xi _{R,t} } & {\Xi _{R,c}^T } \\
395     {\Xi _{R,c} } & {\Xi _{R,r} } \\
396     \end{array}} \right)\left( \begin{array}{l}
397     v_{R,i}^b (t) \\
398     \omega _i (t) \\
399     \end{array} \right),
400     \end{equation}
401     while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
402     with zero mean and variance
403     \begin{equation}
404     \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle =
405     \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle =
406     2k_B T\Xi _R \delta (t - t').
407     \end{equation}
408    
409     The equation of motion for $v_i$ can be written as
410     \begin{equation}
411     m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
412     f_{r,i}^l (t)
413     \end{equation}
414     Since the frictional force is applied at the center of resistance
415     which generally does not coincide with the center of mass, an extra
416     torque is exerted at the center of mass. Thus, the net body-fixed
417     frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
418     given by
419     \begin{equation}
420     \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
421     \end{equation}
422     where $r_{MR}$ is the vector from the center of mass to the center
423     of the resistance. Instead of integrating angular velocity in
424     lab-fixed frame, we consider the equation of motion of angular
425     momentum in body-fixed frame
426     \begin{equation}
427     \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
428     + \tau _{r,i}^b(t)
429     \end{equation}
430    
431     Embedding the friction terms into force and torque, one can
432     integrate the langevin equations of motion for rigid body of
433     arbitrary shape in a velocity-Verlet style 2-part algorithm, where
434     $h= \delta t$:
435    
436     {\tt moveA:}
437     \begin{align*}
438     {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
439     + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
440     %
441     {\bf r}(t + h) &\leftarrow {\bf r}(t)
442     + h {\bf v}\left(t + h / 2 \right), \\
443     %
444     {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
445     + \frac{h}{2} {\bf \tau}^b(t), \\
446     %
447     \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
448     (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
449     \end{align*}
450    
451     In this context, the $\mathrm{rotate}$ function is the reversible
452     product of the three body-fixed rotations,
453     \begin{equation}
454     \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
455     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
456     / 2) \cdot \mathsf{G}_x(a_x /2),
457     \end{equation}
458     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
459     rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
460     angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
461     axis $\alpha$,
462     \begin{equation}
463     \mathsf{G}_\alpha( \theta ) = \left\{
464     \begin{array}{lcl}
465     \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
466     {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
467     j}(0).
468     \end{array}
469     \right.
470     \end{equation}
471     $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
472     rotation matrix. For example, in the small-angle limit, the
473     rotation matrix around the body-fixed x-axis can be approximated as
474     \begin{equation}
475     \mathsf{R}_x(\theta) \approx \left(
476     \begin{array}{ccc}
477     1 & 0 & 0 \\
478     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
479     \theta^2 / 4} \\
480     0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
481     \theta^2 / 4}
482     \end{array}
483     \right).
484     \end{equation}
485     All other rotations follow in a straightforward manner.
486    
487     After the first part of the propagation, the forces and body-fixed
488     torques are calculated at the new positions and orientations
489    
490     {\tt doForces:}
491     \begin{align*}
492     {\bf f}(t + h) &\leftarrow
493     - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
494     %
495     {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
496     \times \frac{\partial V}{\partial {\bf u}}, \\
497     %
498     {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
499     \cdot {\bf \tau}^s(t + h).
500     \end{align*}
501    
502     {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
503     $\mathsf{A}$ is calculated in {\tt moveA}. Once the forces and
504     torques have been obtained at the new time step, the velocities can
505     be advanced to the same time value.
506    
507     {\tt moveB:}
508     \begin{align*}
509     {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
510     \right)
511     + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
512     %
513     {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
514     \right)
515     + \frac{h}{2} {\bf \tau}^b(t + h) .
516     \end{align*}
517    
518     \section{Results and discussion}
519    
520 tim 2858 The Langevin algorithm described in Sec.~\ref{LDRB} has been
521     implemented in {\sc oopse}\cite{Meineke2005} and applied to several
522     test systems.
523    
524     \subsection{Langevin dynamics of}
525    
526 tim 2857 \begin{figure}
527     \centering
528 tim 2858 \includegraphics[width=\linewidth]{temperature.eps}
529     \caption[]{.} \label{langevin:temperature}
530 tim 2857 \end{figure}
531    
532 tim 2858 \subsection{LD of banana-shaped molecule}
533    
534 tim 2857 \begin{figure}
535     \centering
536 tim 2858 \includegraphics[width=\linewidth]{one_banana.eps}
537     \caption[]{.} \label{langevin:banana}
538 tim 2857 \end{figure}
539    
540 tim 2851 \section{Conclusions}