ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.tex
(Generate patch)

Comparing trunk/langevin/langevin.tex (file contents):
Revision 3332 by xsun, Fri Jan 18 22:07:35 2008 UTC vs.
Revision 3333 by gezelter, Thu Jan 24 14:16:07 2008 UTC

# Line 68 | Line 68 | Kramers's theory, Klimov and Thirumalai identified fre
68   The stochastic nature of Langevin dynamics also enhances the sampling
69   of the system and increases the probability of crossing energy
70   barriers.\cite{Cui2003,Banerjee2004} Combining Langevin dynamics with
71 < Kramers's theory, Klimov and Thirumalai identified free-energy
71 > Kramers' theory, Klimov and Thirumalai identified free-energy
72   barriers by studying the viscosity dependence of the protein folding
73   rates.\cite{Klimov1997} In order to account for solvent induced
74   interactions missing from the implicit solvent model, Kaya
# Line 85 | Line 85 | individual atoms are taken from the Stokes-Einstein hy
85   finite structures.\cite{Berkov2005a}]
86  
87   In typical LD simulations, the friction and random forces on
88 < individual atoms are taken from the Stokes-Einstein hydrodynamic
89 < approximation,
88 > individual atoms are taken from Stokes' law,
89   \begin{eqnarray}
90   m \dot{v}(t) & = & -\nabla U(x) - \xi m v(t) + R(t) \\
91   \langle R(t) \rangle & = & 0 \\
# Line 95 | Line 94 | The use of rigid substructures,\cite{???}
94   where $\xi \approx 6 \pi \eta a$.  Here $\eta$ is the viscosity of the
95   implicit solvent, and $a$ is the hydrodynamic radius of the atom.
96  
97 < The use of rigid substructures,\cite{???}
98 < coarse-graining,\cite{Ayton,Sun,Zannoni} and ellipsoidal
99 < representations of protein side chains~\cite{Schulten} has made the
100 < use of the Stokes-Einstein approximation problematic.  A rigid
101 < substructure moves as a single unit with orientational as well as
102 < translational degrees of freedom.  This requires a more general
97 > The use of rigid substructures,\cite{Chun:2000fj}
98 > coarse-graining,\cite{Ayton01,Golubkov06,Orlandi:2006fk,SunGezelter08}
99 > and ellipsoidal representations of protein side chains~\cite{Fogolari:1996lr}
100 > has made the use of the Stokes-Einstein approximation problematic.  A
101 > rigid substructure moves as a single unit with orientational as well
102 > as translational degrees of freedom.  This requires a more general
103   treatment of the hydrodynamics than the spherical approximation
104   provides.  The atoms involved in a rigid or coarse-grained structure
105   should properly have solvent-mediated interactions with each
106   other. The theory of interactions {\it between} bodies moving through
107   a fluid has been developed over the past century and has been applied
108   to simulations of Brownian
109 < motion.\cite{MarshallNewton,GarciaDeLaTorre} There a need to have a
111 < more thorough treatment of hydrodynamics included in the library of
112 < methods available for performing Langevin simulations.
109 > motion.\cite{FIXMAN:1986lr,Ramachandran1996}
110  
111 + In order to account for the diffusion anisotropy of arbitrarily-shaped
112 + particles, Fernandes and Garc\'{i}a de la Torre improved the original
113 + Brownian dynamics simulation algorithm~\cite{Ermak1978,Allison1991} by
114 + incorporating a generalized $6\times6$ diffusion tensor and
115 + introducing a rotational evolution scheme consisting of three
116 + consecutive rotations.\cite{Fernandes2002} Unfortunately, biases are
117 + introduced into the system due to the arbitrary order of applying the
118 + noncommuting rotation operators.\cite{Beard2003} Based on the
119 + observation the momentum relaxation time is much less than the time
120 + step, one may ignore the inertia in Brownian dynamics.  However, the
121 + assumption of zero average acceleration is not always true for
122 + cooperative motion which is common in proteins. An inertial Brownian
123 + dynamics (IBD) was proposed to address this issue by adding an
124 + inertial correction term.\cite{Beard2000} As a complement to IBD which
125 + has a lower bound in time step because of the inertial relaxation
126 + time, long-time-step inertial dynamics (LTID) can be used to
127 + investigate the inertial behavior of linked polymer segments in a low
128 + friction regime.\cite{Beard2000} LTID can also deal with the
129 + rotational dynamics for nonskew bodies without translation-rotation
130 + coupling by separating the translation and rotation motion and taking
131 + advantage of the analytical solution of hydrodynamics
132 + properties. However, typical nonskew bodies like cylinders and
133 + ellipsoids are inadequate to represent most complex macromolecular
134 + assemblies.  There is therefore a need for incorporating the
135 + hydrodynamics of complex (and potentially skew) rigid bodies in the
136 + library of methods available for performing Langevin simulations.
137 +
138   \subsection{Rigid Body Dynamics}
139   Rigid bodies are frequently involved in the modeling of large
140   collections of particles that move as a single unit.  In molecular
141   simulations, rigid bodies have been used to simplify protein-protein
142 < docking,\cite{Gray2003} and lipid bilayer simulations.\cite{Sun2008}
143 < Many of the water models in common use are also rigid-body
144 < models,\cite{TIPs,SPC/E} although they are typically evolved using
145 < constraints rather than rigid body equations of motion.
142 > docking,\cite{Gray2003} and lipid bilayer
143 > simulations.\cite{SunGezelter08} Many of the water models in common
144 > use are also rigid-body
145 > models,\cite{Jorgensen83,Berendsen81,Berendsen87} although they are
146 > typically evolved using constraints rather than rigid body equations
147 > of motion.
148  
149 < Euler angles are a natural choice to describe the rotational
150 < degrees of freedom.  However, due to $1 \over \sin \theta$
151 < singularities, the numerical integration of corresponding equations of
152 < these motion can become inaccurate (and inefficient).  Although an
153 < alternative integrator using multiple sets of Euler angles can
154 < overcome this problem,\cite{Barojas1973} the computational penalty and
155 < the loss of angular momentum conservation remain. A singularity-free
156 < representation utilizing quaternions was developed by Evans in
157 < 1977.\cite{Evans1977} Unfortunately, this approach uses a nonseparable
158 < Hamiltonian resulting from the quaternion representation, which
159 < prevented symplectic algorithms from being utilized until very
134 < recently.\cite{Miller2002} Another approach is the application of
135 < holonomic constraints to the atoms belonging to the rigid body. Each
136 < atom moves independently under the normal forces deriving from
137 < potential energy and constraint forces which are used to guarantee the
138 < rigidness. However, due to their iterative nature, the SHAKE and
139 < Rattle algorithms also converge very slowly when the number of
140 < constraints increases.\cite{Ryckaert1977,Andersen1983}
149 > Euler angles are a natural choice to describe the rotational degrees
150 > of freedom.  However, due to $1 \over \sin \theta$ singularities, the
151 > numerical integration of corresponding equations of these motion can
152 > become inaccurate (and inefficient).  Although the use of multiple
153 > sets of Euler angles can overcome this problem,\cite{Barojas1973} the
154 > computational penalty and the loss of angular momentum conservation
155 > remain.  A singularity-free representation utilizing quaternions was
156 > developed by Evans in 1977.\cite{Evans1977} The Evans quaternion
157 > approach uses a nonseparable Hamiltonian, and this has prevented
158 > symplectic algorithms from being utilized until very
159 > recently.\cite{Miller2002}
160  
161 < A breakthrough in geometric literature suggests that, in order to
162 < develop a long-term integration scheme, one should preserve the
163 < symplectic structure of the propagator. By introducing a conjugate
164 < momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
165 < equation, a symplectic integrator, RSHAKE,\cite{Kol1997} was proposed
166 < to evolve the Hamiltonian system in a constraint manifold by
167 < iteratively satisfying the orthogonality constraint $Q^T Q = 1$.  An
149 < alternative method using the quaternion representation was developed
150 < by Omelyan.\cite{Omelyan1998} However, both of these methods are
151 < iterative and suffer from some related inefficiencies. A symplectic
152 < Lie-Poisson integrator for rigid bodies developed by Dullweber {\it et
153 < al.}\cite{Dullweber1997} gets around most of the limitations mentioned
154 < above and has become the basis for our Langevin integrator.
161 > Another approach is the application of holonomic constraints to the
162 > atoms belonging to the rigid body.  Each atom moves independently
163 > under the normal forces deriving from potential energy and constraints
164 > are used to guarantee rigidity. However, due to their iterative
165 > nature, the SHAKE and RATTLE algorithms converge very slowly when the
166 > number of constraints (and the number of particles that belong to the
167 > rigid body) increases.\cite{Ryckaert1977,Andersen1983}
168  
169 + In order to develop a stable and efficient integration scheme that
170 + preserves most constants of the motion, symplectic propagators are
171 + necessary.  By introducing a conjugate momentum to the rotation matrix
172 + $Q$ and re-formulating Hamilton's equations, a symplectic
173 + orientational integrator, RSHAKE,\cite{Kol1997} was proposed to evolve
174 + rigid bodies on a constraint manifold by iteratively satisfying the
175 + orthogonality constraint $Q^T Q = 1$.  An alternative method using the
176 + quaternion representation was developed by Omelyan.\cite{Omelyan1998}
177 + However, both of these methods are iterative and suffer from some
178 + related inefficiencies. A symplectic Lie-Poisson integrator for rigid
179 + bodies developed by Dullweber {\it et al.}\cite{Dullweber1997} removes
180 + most of the limitations mentioned above and is therefore the basis for
181 + our Langevin integrator.
182  
157 \subsection{The Hydrodynamic tensor and Brownian dynamics}
158 Combining Brownian dynamics with rigid substructures, one can study
159 slow processes in biomolecular systems.  Modeling DNA as a chain of
160 beads which are subject to harmonic potentials as well as excluded
161 volume potentials, Mielke and his coworkers discovered rapid
162 superhelical stress generations from the stochastic simulation of twin
163 supercoiling DNA with response to induced torques.\cite{Mielke2004}
164 Membrane fusion is another key biological process which controls a
165 variety of physiological functions, such as release of
166 neurotransmitters \textit{etc}. A typical fusion event happens on the
167 time scale of a millisecond, which is impractical to study using
168 atomistic models with newtonian mechanics. With the help of
169 coarse-grained rigid body model and stochastic dynamics, the fusion
170 pathways were explored by Noguchi and others.\cite{Noguchi2001,Noguchi2002,Shillcock2005}
171
172 Due to the difficulty of numerically integrating anisotropic
173 rotational motion, most of the coarse-grained rigid body models are
174 treated as spheres, cylinders, ellipsoids or other regular shapes in
175 stochastic simulations.  In an effort to account for the diffusion
176 anisotropy of arbitrarily-shaped particles, Fernandes and Garc\'{i}a
177 de la Torre improved the original Brownian dynamics simulation
178 algorithm~\cite{Ermak1978,Allison1991} by incorporating a generalized
179 $6\times6$ diffusion tensor and introducing a rotational evolution
180 scheme consisting of three consecutive rotations.\cite{Fernandes2002}
181 Unfortunately, biases are introduced into the system due to the
182 arbitrary order of applying the noncommuting rotation
183 operators.\cite{Beard2003} Based on the observation the momentum
184 relaxation time is much less than the time step, one may ignore the
185 inertia in Brownian dynamics.  However, the assumption of zero average
186 acceleration is not always true for cooperative motion which is common
187 in proteins. An inertial Brownian dynamics (IBD) was proposed to
188 address this issue by adding an inertial correction
189 term.\cite{Beard2000} As a complement to IBD which has a lower bound
190 in time step because of the inertial relaxation time, long-time-step
191 inertial dynamics (LTID) can be used to investigate the inertial
192 behavior of the polymer segments in low friction
193 regime.\cite{Beard2000} LTID can also deal with the rotational
194 dynamics for nonskew bodies without translation-rotation coupling by
195 separating the translation and rotation motion and taking advantage of
196 the analytical solution of hydrodynamics properties. However, typical
197 nonskew bodies like cylinders and ellipsoids are inadequate to
198 represent most complex macromolecular assemblies.
199
183   The goal of the present work is to develop a Langevin dynamics
184   algorithm for arbitrary-shaped rigid particles by integrating the
185   accurate estimation of friction tensor from hydrodynamics theory into
186   a symplectic rigid body dynamics propagator.  In the sections below,
187 < we review some of the theory of hydrodynamic tensors developed for
188 < Brownian simulations of rigid multi-particle systems, we then present
189 < our integration method for a set of generalized Langevin equations of
190 < motion, and we compare the behavior of the new Langevin integrator to
191 < dynamical quantities obtained via explicit solvent molecular dynamics.
187 > we review some of the theory of hydrodynamic tensors developed
188 > primarily for Brownian simulations of multi-particle systems, we then
189 > present our integration method for a set of generalized Langevin
190 > equations of motion, and we compare the behavior of the new Langevin
191 > integrator to dynamical quantities obtained via explicit solvent
192 > molecular dynamics.
193  
194   \subsection{\label{introSection:frictionTensor}The Friction Tensor}
195   Theoretically, a complete friction kernel can be determined using the
196   velocity autocorrelation function. However, this approach becomes
197 < impractical when the solute becomes complex.  Instead, various
197 > impractical when the solute becomes complex. Instead, various
198   approaches based on hydrodynamics have been developed to calculate the
199   friction coefficients. In general, the friction tensor $\Xi$ is a
200   $6\times 6$ matrix given by
201   \begin{equation}
202 < \Xi  = \left( {\begin{array}{*{20}c}
203 <   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
204 <   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
205 < \end{array}} \right).
202 > \Xi  = \left( \begin{array}{*{20}c}
203 >   \Xi^{tt} & \Xi^{rt}  \\
204 >   \Xi^{tr} & \Xi^{rr}  \\
205 > \end{array} \right).
206   \end{equation}
207   Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
208   rotational resistance (friction) tensors respectively, while
# Line 231 | Line 215 | fluid, it may experience friction force ($\mathbf{F}_f
215   \left( \begin{array}{l}
216   \mathbf{F}_f  \\
217   \mathbf{\tau}_f  \\
218 < \end{array} \right) =  - \left( {\begin{array}{*{20}c}
219 <   \Xi ^{tt} & \Xi ^{rt}  \\
220 <   \Xi ^{tr} & \Xi ^{rr}  \\
221 < \end{array}} \right)\left( \begin{array}{l}
218 > \end{array} \right) =  - \left( \begin{array}{*{20}c}
219 >   \Xi^{tt} & \Xi^{rt}  \\
220 >   \Xi^{tr} & \Xi^{rr}  \\
221 > \end{array} \right)\left( \begin{array}{l}
222   \mathbf{v} \\
223   \mathbf{\omega} \\
224   \end{array} \right).
# Line 243 | Line 227 | Stoke's law,
227   \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
228   For a spherical particle under ``stick'' boundary conditions, the
229   translational and rotational friction tensors can be calculated from
230 < Stoke's law,
230 > Stokes' law,
231   \begin{equation}
232 < \Xi^{tt}  = \left( {\begin{array}{*{20}c}
232 > \Xi^{tt}  = \left( \begin{array}{*{20}c}
233     {6\pi \eta R} & 0 & 0  \\
234     0 & {6\pi \eta R} & 0  \\
235     0 & 0 & {6\pi \eta R}  \\
236 < \end{array}} \right)
236 > \end{array} \right)
237   \end{equation}
238   and
239   \begin{equation}
240 < \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
240 > \Xi^{rr}  = \left( \begin{array}{*{20}c}
241     {8\pi \eta R^3 } & 0 & 0  \\
242     0 & {8\pi \eta R^3 } & 0  \\
243     0 & 0 & {8\pi \eta R^3 }  \\
244 < \end{array}} \right)
244 > \end{array} \right)
245   \end{equation}
246   where $\eta$ is the viscosity of the solvent and $R$ is the
247   hydrodynamic radius.
# Line 265 | Line 249 | extended Stokes's law to general ellipsoids, also call
249   Other non-spherical shapes, such as cylinders and ellipsoids, are
250   widely used as references for developing new hydrodynamics theories,
251   because their properties can be calculated exactly. In 1936, Perrin
252 < extended Stokes's law to general ellipsoids, also called a triaxial
253 < ellipsoid, which is given in Cartesian coordinates
270 < by\cite{Perrin1934,Perrin1936}
252 > extended Stokes' law to general ellipsoids which are given in
253 > Cartesian coordinates by~\cite{Perrin1934,Perrin1936}
254   \begin{equation}
255 < \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1
255 > \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1.
256   \end{equation}
257 < where the semi-axes are of lengths $a$, $b$, and $c$. Due to the
258 < complexity of the elliptic integral, only uniaxial ellipsoids,
259 < {\it i.e.} prolate ($ a \ge b = c$) and oblate ($ a < b = c $), can
260 < be solved exactly. Introducing an elliptic integral parameter $S$ for
278 < prolate ellipsoids :
257 > Here, the semi-axes are of lengths $a$, $b$, and $c$. Due to the
258 > complexity of the elliptic integral, only uniaxial ellipsoids, either
259 > prolate ($a \ge b = c$) or oblate ($a < b = c$), can be solved
260 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
261   \begin{equation}
262   S = \frac{2}{\sqrt{a^2  - b^2}} \ln \frac{a + \sqrt{a^2  - b^2}}{b},
263   \end{equation}
264 < and oblate ellipsoids:
264 > and oblate,
265   \begin{equation}
266   S = \frac{2}{\sqrt {b^2  - a^2 }} \arctan \frac{\sqrt {b^2  - a^2}}{a},
267   \end{equation}
268 < one can write down the translational and rotational resistance
269 < tensors for oblate,
268 > ellipsoids, one can write down the translational and rotational
269 > resistance tensors:
270   \begin{eqnarray*}
271   \Xi_a^{tt}  & = & 16\pi \eta \frac{a^2  - b^2}{(2a^2  - b^2 )S - 2a}. \\
272   \Xi_b^{tt} =  \Xi_c^{tt} & = & 32\pi \eta \frac{a^2  - b^2 }{(2a^2 - 3b^2 )S + 2a},
273   \end{eqnarray*}
274 < and prolate,
274 > for oblate, and
275   \begin{eqnarray*}
276   \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2  - b^2 )b^2}{2a - b^2 S}, \\
277   \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4  - b^4)}{(2a^2  - b^2 )S - 2a}
278   \end{eqnarray*}
279 < ellipsoids. For both spherical and ellipsoidal particles, the
280 < translation-rotation and rotation-translation coupling tensors are
279 > for prolate ellipsoids. For both spherical and ellipsoidal particles,
280 > the translation-rotation and rotation-translation coupling tensors are
281   zero.
282  
283   \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
302
284   Unlike spherical and other simply shaped molecules, there is no
285   analytical solution for the friction tensor for arbitrarily shaped
286   rigid molecules. The ellipsoid of revolution model and general
# Line 315 | Line 296 | interaction tensor. Let us consider a rigid assembly o
296   using more advanced methods where the molecule of interest was modeled
297   by a combinations of spheres\cite{Carrasco1999} and the hydrodynamics
298   properties of the molecule can be calculated using the hydrodynamic
299 < interaction tensor. Let us consider a rigid assembly of $N$ beads
300 < immersed in a continuous medium. Due to hydrodynamic interaction, the
301 < ``net'' velocity of $i$th bead, $v'_i$ is different than its
302 < unperturbed velocity $v_i$,
303 < \[
299 > interaction tensor.
300 >
301 > Consider a rigid assembly of $N$ beads immersed in a continuous
302 > medium. Due to hydrodynamic interaction, the ``net'' velocity of $i$th
303 > bead, $v'_i$ is different than its unperturbed velocity $v_i$,
304 > \begin{equation}
305   v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
306 < \]
307 < where $F_i$ is the frictional force, and $T_{ij}$ is the
308 < hydrodynamic interaction tensor. The friction force of $i$th bead is
309 < proportional to its ``net'' velocity
306 > \end{equation}
307 > where $F_i$ is the frictional force, and $T_{ij}$ is the hydrodynamic
308 > interaction tensor. The frictional force on the $i^\mathrm{th}$ bead
309 > is proportional to its ``net'' velocity
310   \begin{equation}
311   F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
312   \label{introEquation:tensorExpression}
# Line 361 | Line 343 | B = \left( {\begin{array}{*{20}c}
343   construct a $3N \times 3N$ matrix consisting of $N \times N$
344   $B_{ij}$ blocks
345   \begin{equation}
346 < B = \left( {\begin{array}{*{20}c}
347 <   {B_{11} } &  \ldots  & {B_{1N} }  \\
346 > B = \left( \begin{array}{*{20}c}
347 >   B_{11} &  \ldots  & B_{1N}   \\
348      \vdots  &  \ddots  &  \vdots   \\
349 <   {B_{N1} } &  \cdots  & {B_{NN} }  \\
350 < \end{array}} \right),
349 >   B_{N1} &  \cdots  & B_{NN} \\
350 > \end{array} \right),
351   \end{equation}
352   where $B_{ij}$ is given by
353 < \[
353 > \begin{equation}
354   B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
355   )T_{ij}
356 < \]
356 > \end{equation}
357   where $\delta _{ij}$ is the Kronecker delta function. Inverting the
358   $B$ matrix, we obtain
359   \[
# Line 456 | Line 438 | locate the position of center of resistance,
438   x_{OR}  \\
439   y_{OR}  \\
440   z_{OR}  \\
441 < \end{array} \right) & = &\left( {\begin{array}{*{20}c}
441 > \end{array} \right) & = &\left( \begin{array}{*{20}c}
442     {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
443     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
444     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
445 < \end{array}} \right)^{ - 1}  \\
445 > \end{array} \right)^{ - 1}  \\
446    & & \left( \begin{array}{l}
447   (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
448   (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
# Line 474 | Line 456 | M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i}
456   \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
457   Consider the Langevin equations of motion in generalized coordinates
458   \begin{equation}
459 < M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i} (t)
459 > \mathbf{M}_i \dot \mathbf{V}_i(t) = \mathbf{F}_{s,i}(t) + \mathbf{F}_{f,i}(t)  + \mathbf{R}_{i}(t)
460   \label{LDGeneralizedForm}
461   \end{equation}
462 < where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
463 < and moment of inertial) matrix and $V_i$ is a generalized velocity,
464 < $V_i = V_i(v_i,\omega _i)$. The right side of
465 < Eq.~\ref{LDGeneralizedForm} consists of three generalized forces in
466 < lab-fixed frame, systematic force $F_{s,i}$, dissipative force
467 < $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
468 < system in Newtownian mechanics typically refers to lab-fixed frame,
469 < it is also convenient to handle the rotation of rigid body in
470 < body-fixed frame. Thus the friction and random forces are calculated
471 < in body-fixed frame and converted back to lab-fixed frame by:
472 < \[
462 > where $\mathbf{M}_i$ is a $6\times6$ diagonal mass matrix (which
463 > includes the rigid body mass and moments of inertia) and $\mathbf{V}_i$ is a
464 > generalized velocity, $\mathbf{V}_i =
465 > \left\{\mathbf{v}_i,\mathbf{\omega}_i \right\}$. The right side of
466 > Eq.~\ref{LDGeneralizedForm} consists of three generalized forces: a
467 > system force $\mathbf{F}_{s,i}$, a frictional or dissipative force
468 > $\mathbf{F}_{f,i}$ and stochastic force $\mathbf{R}_{i}$. While the
469 > evolution of the system in Newtownian mechanics is typically done in the
470 > lab-fixed frame, it is convenient to handle the rotation of rigid
471 > bodies in the body-fixed frame. Thus the friction and random forces are
472 > calculated in body-fixed frame and converted back to lab-fixed frame
473 > using the rigid body's rotation matrix ($Q_i$):
474 > \begin{equation}
475   \begin{array}{l}
476 < F_{f,i}^l (t) = Q^T F_{f,i}^b (t), \\
477 < F_{r,i}^l (t) = Q^T F_{r,i}^b (t). \\
476 > \mathbf{F}_{f,i}(t) = Q_{i}^{T} \mathbf{F}_{f,i}^b (t), \\
477 > \mathbf{R}_{i}(t) = Q_{i}^{T} \mathbf{R}_{i}^b (t). \\
478   \end{array}
479 < \]
480 < Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
481 < the body-fixed velocity at center of resistance $v_{R,i}^b$ and
482 < angular velocity $\omega _i$
479 > \end{equation}
480 > Here, the body-fixed friction force $\mathbf{F}_{f,i}^b$ is proportional to
481 > the body-fixed velocity at the center of resistance $\mathbf{v}_{R,i}^b$ and
482 > angular velocity $\mathbf{\omega}_i$
483   \begin{equation}
484 < F_{r,i}^b (t) = \left( \begin{array}{l}
485 < f_{r,i}^b (t) \\
486 < \tau _{r,i}^b (t) \\
487 < \end{array} \right) =  - \left( {\begin{array}{*{20}c}
488 <   {\Xi _{R,t} } & {\Xi _{R,c}^T }  \\
489 <   {\Xi _{R,c} } & {\Xi _{R,r} }  \\
490 < \end{array}} \right)\left( \begin{array}{l}
491 < v_{R,i}^b (t) \\
492 < \omega _i (t) \\
484 > \mathbf{F}_{f,i}^b (t) = \left( \begin{array}{l}
485 > \mathbf{f}_{f,i}^b (t) \\
486 > \mathbf{\tau}_{f,i}^b (t) \\
487 > \end{array} \right) =  - \left( \begin{array}{*{20}c}
488 >   \Xi_{R,t} & \Xi_{R,c}^T  \\
489 >   \Xi_{R,c} & \Xi_{R,r}    \\
490 > \end{array} \right)\left( \begin{array}{l}
491 > \mathbf{v}_{R,i}^b (t) \\
492 > \mathbf{\omega}_i (t) \\
493   \end{array} \right),
494   \end{equation}
495 < while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
495 > while the random force $\mathbf{R}_{i}^l$ is a Gaussian stochastic variable
496   with zero mean and variance
497   \begin{equation}
498 < \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle  =
499 < \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle  =
500 < 2k_B T\Xi _R \delta (t - t'). \label{randomForce}
498 > \left\langle {\mathbf{R}_{i}^l (t) (\mathbf{R}_{i}^l (t'))^T } \right\rangle  =
499 > \left\langle {\mathbf{R}_{i}^b (t) (\mathbf{R}_{i}^b (t'))^T } \right\rangle  =
500 > 2 k_B T \Xi_R \delta(t - t'). \label{randomForce}
501   \end{equation}
502 < The equation of motion for $v_i$ can be written as
502 > Once the $6\times6$ resistance tensor at the center of resistance
503 > ($\Xi_R$) is known, obtaining a stochastic vector that has the
504 > properties in Eq. (\ref{eq:randomForce}) can be done efficiently by
505 > carrying out a one-time Cholesky decomposition to obtain the square
506 > root matrix of $\Xi_R$.\cite{SchlickBook} Each time a random force
507 > vector is needed, a gaussian random vector is generated and then the
508 > square root matrix is multiplied onto this vector.
509 >
510 > The equation of motion for $\mathbf{v}_i$ can be written as
511   \begin{equation}
512 < m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
513 < f_{r,i}^l (t)
512 > m\dot \mathbf{v}_i (t) =  \mathbf{f}_{s,i} (t) + \mathbf{f}_{f,i}^l (t) +
513 > \mathbf{R}_{i}^l (t)
514   \end{equation}
515   Since the frictional force is applied at the center of resistance
516   which generally does not coincide with the center of mass, an extra
517   torque is exerted at the center of mass. Thus, the net body-fixed
518 < frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
518 > frictional torque at the center of mass, $\tau_{f,i}^b (t)$, is
519   given by
520   \begin{equation}
521 < \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
521 > \tau_{f,i}^b \leftarrow \tau_{f,i}^b + \mathbf{r}_{MR} \times \mathbf{f}_{r,i}^b
522   \end{equation}
523   where $r_{MR}$ is the vector from the center of mass to the center
524   of the resistance. Instead of integrating the angular velocity in
525   lab-fixed frame, we consider the equation of angular momentum in
526   body-fixed frame
527   \begin{equation}
528 < \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
537 < + \tau _{r,i}^b(t)
528 > \dot j_i (t) = \tau_{s,i} (t) + \tau_{f,i}^b (t) + \mathbf{R}_{i}^b(t)
529   \end{equation}
530 < Embedding the friction terms into force and torque, one can
531 < integrate the langevin equations of motion for rigid body of
532 < arbitrary shape in a velocity-Verlet style 2-part algorithm, where
542 < $h= \delta t$:
530 > Embedding the friction terms into force and torque, one can integrate
531 > the Langevin equations of motion for rigid body of arbitrary shape in
532 > a velocity-Verlet style 2-part algorithm, where $h= \delta t$:
533  
534   {\tt moveA:}
535   \begin{align*}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines