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 3315 by gezelter, Thu Jan 17 22:45:16 2008 UTC vs.
Revision 3316 by gezelter, Fri Jan 18 20:43:53 2008 UTC

# Line 48 | Line 48 | As alternative to Newtonian dynamics, Langevin dynamic
48   \section{Introduction}
49  
50   %applications of langevin dynamics
51 < As alternative to Newtonian dynamics, Langevin dynamics, which
52 < mimics a simple heat bath with stochastic and dissipative forces,
53 < has been applied in a variety of studies. The stochastic treatment
54 < of the solvent enables us to carry out substantially longer time
55 < simulations. Implicit solvent Langevin dynamics simulations of
56 < met-enkephalin not only outperform explicit solvent simulations for
57 < computational efficiency, but also agrees very well with explicit
58 < solvent simulations for dynamical properties.\cite{Shen2002}
59 < Recently, applying Langevin dynamics with the UNRES model, Liow and
60 < his coworkers suggest that protein folding pathways can be possibly
61 < explored within a reasonable amount of time.\cite{Liwo2005} The
62 < stochastic nature of the Langevin dynamics also enhances the
63 < sampling of the system and increases the probability of crossing
64 < energy barriers.\cite{Banerjee2004, Cui2003} Combining Langevin
65 < dynamics with Kramers's theory, Klimov and Thirumalai identified
66 < free-energy barriers by studying the viscosity dependence of the
67 < protein folding rates.\cite{Klimov1997} In order to account for
68 < solvent induced interactions missing from implicit solvent model,
69 < Kaya incorporated desolvation free energy barrier into implicit
70 < coarse-grained solvent model in protein folding/unfolding studies
71 < and discovered a higher free energy barrier between the native and
72 < denatured states. Because of its stability against noise, Langevin
73 < dynamics is very suitable for studying remagnetization processes in
74 < various systems.\cite{Palacios1998,Berkov2002,Denisov2003} For
51 > Langevin dynamics, which mimics a simple heat bath with stochastic and
52 > dissipative forces, has been applied in a variety of situations as an
53 > alternative to molecular dynamics with explicit solvent molecules.
54 > The stochastic treatment of the solvent allows the use of simulations
55 > with substantially longer time and length scales.  In general, the
56 > dynamic and structural properties obtained from Langevin simulations
57 > agree quite well with similar properties obtained from explicit
58 > solvent simulations.
59 >
60 > Recent examples of the usefulness of Langevin simulations include a
61 > study of met-enkephalin in which Langevin simulations predicted
62 > dynamical properties that were largely in agreement with explicit
63 > solvent simulations.\cite{Shen2002} By applying Langevin dynamics with
64 > the UNRES model, Liow and his coworkers suggest that protein folding
65 > pathways can be explored within a reasonable amount of
66 > time.\cite{Liwo2005}
67 >
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
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
75 > incorporated a desolvation free energy barrier into protein
76 > folding/unfolding studies and discovered a higher free energy barrier
77 > between the native and denatured states.\cite{XXX}
78 >
79 > Because of its stability against noise, Langevin dynamics has also
80 > proven useful for studying remagnetization processes in various
81 > systems.\cite{Palacios1998,Berkov2002,Denisov2003} [Check: For
82   instance, the oscillation power spectrum of nanoparticles from
83 < Langevin dynamics simulation has the same peak frequencies for
84 < different wave vectors, which recovers the property of magnetic
85 < excitations in small finite structures.\cite{Berkov2005a}
83 > Langevin dynamics has the same peak frequencies for different wave
84 > vectors, which recovers the property of magnetic excitations in small
85 > finite structures.\cite{Berkov2005a}]
86  
87 < %review rigid body dynamics
88 < Rigid bodies are frequently involved in the modeling of different
89 < areas, from engineering, physics, to chemistry. For example,
90 < missiles and vehicle are usually modeled by rigid bodies.  The
91 < movement of the objects in 3D gaming engine or other physics
92 < simulator is governed by the rigid body dynamics. In molecular
93 < simulation, rigid body is used to simplify the model in
94 < protein-protein docking study{\cite{Gray2003}}.
87 > In typical LD simulations, the friction and random forces on
88 > individual atoms are taken from the Stokes-Einstein hydrodynamic
89 > approximation,
90 > \begin{eqnarray}
91 > m \dot{v}(t) & = & -\nabla U(x) - \xi m v(t) + R(t) \\
92 > \langle R(t) \rangle & = & 0 \\
93 > \langle R(t) R(t') \rangle & = & 2 k_B T \xi m \delta(t - t')
94 > \end{eqnarray}
95 > where $\xi \approx 6 \pi \eta a$.  Here $\eta$ is the viscosity of the
96 > implicit solvent, and $a$ is the hydrodynamic radius of the atom.
97  
98 < It is very important to develop stable and efficient methods to
99 < integrate the equations of motion for orientational degrees of
100 < freedom. Euler angles are the natural choice to describe the
101 < rotational degrees of freedom. However, due to $\frac {1}{sin
102 < \theta}$ singularities, the numerical integration of corresponding
103 < equations of these motion is very inefficient and inaccurate.
104 < Although an alternative integrator using multiple sets of Euler
105 < angles can overcome this difficulty\cite{Barojas1973}, the
106 < computational penalty and the loss of angular momentum conservation
107 < still remain. A singularity-free representation utilizing
108 < quaternions was developed by Evans in 1977.\cite{Evans1977}
109 < Unfortunately, this approach used a nonseparable Hamiltonian
110 < resulting from the quaternion representation, which prevented the
111 < symplectic algorithm from being utilized. Another different approach
112 < is to apply holonomic constraints to the atoms belonging to the
104 < rigid body. Each atom moves independently under the normal forces
105 < deriving from potential energy and constraint forces which are used
106 < to guarantee the rigidness. However, due to their iterative nature,
107 < the SHAKE and Rattle algorithms also converge very slowly when the
108 < number of constraints increases.\cite{Ryckaert1977, Andersen1983}
98 > The use of rigid substructures,\cite{???}
99 > coarse-graining,\cite{Ayton,Sun,Zannoni} and ellipsoidal
100 > representations of protein side chains~\cite{Schulten} has made the
101 > use of the Stokes-Einstein approximation problematic.  A rigid
102 > substructure moves as a single unit with orientational as well as
103 > translational degrees of freedom.  This requires a more general
104 > treatment of the hydrodynamics than the spherical approximation
105 > provides.  The atoms involved in a rigid or coarse-grained structure
106 > should properly have solvent-mediated interactions with each
107 > other. The theory of interactions {\it between} bodies moving through
108 > a fluid has been developed over the past century and has been applied
109 > to simulations of Brownian
110 > 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.
113  
114 < A break-through in geometric literature suggests that, in order to
114 > \subsection{Rigid Body Dynamics}
115 > Rigid bodies are frequently involved in the modeling of large
116 > collections of particles that move as a single unit.  In molecular
117 > simulations, rigid bodies have been used to simplify protein-protein
118 > docking,\cite{Gray2003} and lipid bilayer simulations.\cite{Sun2008}
119 > Many of the water models in common use are also rigid-body
120 > models,\cite{TIPs,SPC/E} although they are typically evolved using
121 > constraints rather than rigid body equations of motion.
122 >
123 > Euler angles are a natural choice to describe the rotational
124 > degrees of freedom.  However, due to $1 \over \sin \theta$
125 > singularities, the numerical integration of corresponding equations of
126 > these motion can become inaccurate (and inefficient).  Although an
127 > alternative integrator using multiple sets of Euler angles can
128 > overcome this problem,\cite{Barojas1973} the computational penalty and
129 > the loss of angular momentum conservation remain. A singularity-free
130 > representation utilizing quaternions was developed by Evans in
131 > 1977.\cite{Evans1977} Unfortunately, this approach uses a nonseparable
132 > Hamiltonian resulting from the quaternion representation, which
133 > 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}
141 >
142 > A breakthrough in geometric literature suggests that, in order to
143   develop a long-term integration scheme, one should preserve the
144   symplectic structure of the propagator. By introducing a conjugate
145   momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
146 < equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
147 < proposed to evolve the Hamiltonian system in a constraint manifold
148 < by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
149 < An alternative method using the quaternion representation was
150 < developed by Omelyan.\cite{Omelyan1998} However, both of these
151 < methods are iterative and inefficient. In this section, we descibe a
152 < symplectic Lie-Poisson integrator for rigid bodies developed by
153 < Dullweber and his coworkers\cite{Dullweber1997} in depth.
146 > equation, a symplectic integrator, RSHAKE,\cite{Kol1997} was proposed
147 > to evolve the Hamiltonian system in a constraint manifold by
148 > 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.
155  
156 < %review langevin/browninan dynamics for arbitrarily shaped rigid body
157 < Combining Langevin or Brownian dynamics with rigid body dynamics,
158 < one can study slow processes in biomolecular systems. Modeling DNA
159 < as a chain of rigid beads, which are subject to harmonic potentials
160 < as well as excluded volume potentials, Mielke and his coworkers
161 < discovered rapid superhelical stress generations from the stochastic
162 < simulation of twin supercoiling DNA with response to induced
163 < torques.\cite{Mielke2004} Membrane fusion is another key biological
164 < process which controls a variety of physiological functions, such as
165 < release of neurotransmitters \textit{etc}. A typical fusion event
166 < happens on the time scale of a millisecond, which is impractical to
167 < study using atomistic models with newtonian mechanics. With the help
168 < of coarse-grained rigid body model and stochastic dynamics, the
169 < fusion pathways were explored by many
170 < researchers.\cite{Noguchi2001,Noguchi2002,Shillcock2005} Due to the
171 < difficulty of numerical integration of anisotropic rotation, most of
172 < the rigid body models are simply modeled using spheres, cylinders,
173 < ellipsoids or other regular shapes in stochastic simulations. In an
174 < effort to account for the diffusion anisotropy of arbitrary
175 < particles, Fernandes and de la Torre improved the original Brownian
176 < dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
177 < incorporating a generalized $6\times6$ diffusion tensor and
178 < introducing a simple rotation evolution scheme consisting of three
179 < consecutive rotations.\cite{Fernandes2002} Unfortunately, unexpected
180 < errors and biases are introduced into the system due to the
156 >
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
186 < average acceleration is not always true for cooperative motion which
187 < is common in protein motion. An inertial Brownian dynamics (IBD) was
188 < proposed to address this issue by adding an inertial correction
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
196 < of the analytical solution of hydrodynamics properties. However,
197 < typical nonskew bodies like cylinders and ellipsoids are inadequate
198 < to represent most complex macromolecule assemblies. These intricate
165 < molecules have been represented by a set of beads and their
166 < hydrodynamic properties can be calculated using variants on the
167 < standard hydrodynamic interaction tensors.
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  
200   The goal of the present work is to develop a Langevin dynamics
201   algorithm for arbitrary-shaped rigid particles by integrating the
202 < accurate estimation of friction tensor from hydrodynamics theory
203 < into the sophisticated rigid body dynamics algorithms.
202 > accurate estimation of friction tensor from hydrodynamics theory into
203 > a symplectic rigid body dynamics propagator.  In the sections below,
204 > we review some of the theory of hydrodynamic tensors developed for
205 > Brownian simulations of rigid multi-particle systems, we then present
206 > our integration method for a set of generalized Langevin equations of
207 > motion, and we compare the behavior of the new Langevin integrator to
208 > dynamical quantities obtained via explicit solvent molecular dynamics.
209  
210 < \subsection{\label{introSection:frictionTensor}Friction Tensor}
211 < Theoretically, the friction kernel can be determined using the
210 > \subsection{\label{introSection:frictionTensor}The Friction Tensor}
211 > Theoretically, a complete friction kernel can be determined using the
212   velocity autocorrelation function. However, this approach becomes
213 < impractical when the system becomes more and more complicated.
214 < Instead, various approaches based on hydrodynamics have been
215 < developed to calculate the friction coefficients. In general, the
216 < friction tensor $\Xi$ is a $6\times 6$ matrix given by
217 < \[
213 > impractical when the solute becomes complex.  Instead, various
214 > approaches based on hydrodynamics have been developed to calculate the
215 > friction coefficients. In general, the friction tensor $\Xi$ is a
216 > $6\times 6$ matrix given by
217 > \begin{equation}
218   \Xi  = \left( {\begin{array}{*{20}c}
219     {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
220     {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
221   \end{array}} \right).
222 < \]
223 < Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are $3 \times 3$
224 < translational friction tensor and rotational resistance (friction)
225 < tensor respectively, while ${\Xi^{tr} }$ is translation-rotation
226 < coupling tensor and $ {\Xi^{rt} }$ is rotation-translation coupling
227 < tensor. When a particle moves in a fluid, it may experience friction
228 < force or torque along the opposite direction of the velocity or
229 < angular velocity,
230 < \[
222 > \end{equation}
223 > Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
224 > rotational resistance (friction) tensors respectively, while
225 > $\Xi^{tr}$ is translation-rotation coupling tensor and $\Xi^{rt}$ is
226 > rotation-translation coupling tensor. When a particle moves in a
227 > fluid, it may experience friction force ($\mathbf{F}_f$) and torque
228 > ($\mathbf{\tau}_f$) in opposition to the directions of the velocity
229 > ($\mathbf{v}$) and body-fixed angular velocity ($\mathbf{\omega}$),
230 > \begin{equation}
231   \left( \begin{array}{l}
232 < F_R  \\
233 < \tau _R  \\
232 > \mathbf{F}_f  \\
233 > \mathbf{\tau}_f  \\
234   \end{array} \right) =  - \left( {\begin{array}{*{20}c}
235 <   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
236 <   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
235 >   \Xi ^{tt} & \Xi ^{rt}  \\
236 >   \Xi ^{tr} & \Xi ^{rr}  \\
237   \end{array}} \right)\left( \begin{array}{l}
238 < v \\
239 < w \\
240 < \end{array} \right)
241 < \]
206 < where $F_r$ is the friction force and $\tau _R$ is the friction
207 < torque.
238 > \mathbf{v} \\
239 > \mathbf{\omega} \\
240 > \end{array} \right).
241 > \end{equation}
242  
243   \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
244 <
245 < For a spherical particle with slip boundary conditions, the
246 < translational and rotational friction constant can be calculated
247 < from Stoke's law,
248 < \[
215 < \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
244 > For a spherical particle under ``stick'' boundary conditions, the
245 > translational and rotational friction tensors can be calculated from
246 > Stoke's law,
247 > \begin{equation}
248 > \Xi^{tt}  = \left( {\begin{array}{*{20}c}
249     {6\pi \eta R} & 0 & 0  \\
250     0 & {6\pi \eta R} & 0  \\
251     0 & 0 & {6\pi \eta R}  \\
252   \end{array}} \right)
253 < \]
253 > \end{equation}
254   and
255 < \[
255 > \begin{equation}
256   \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
257     {8\pi \eta R^3 } & 0 & 0  \\
258     0 & {8\pi \eta R^3 } & 0  \\
259     0 & 0 & {8\pi \eta R^3 }  \\
260   \end{array}} \right)
261 < \]
261 > \end{equation}
262   where $\eta$ is the viscosity of the solvent and $R$ is the
263   hydrodynamic radius.
264  
265   Other non-spherical shapes, such as cylinders and ellipsoids, are
266 < widely used as references for developing new hydrodynamics theory,
266 > widely used as references for developing new hydrodynamics theories,
267   because their properties can be calculated exactly. In 1936, Perrin
268   extended Stokes's law to general ellipsoids, also called a triaxial
269   ellipsoid, which is given in Cartesian coordinates
270 < by\cite{Perrin1934, Perrin1936}
271 < \[
272 < \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
273 < }} = 1
274 < \]
275 < where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
276 < due to the complexity of the elliptic integral, only the ellipsoid
277 < with the restriction of two axes being equal, \textit{i.e.}
278 < prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
279 < exactly. Introducing an elliptic integral parameter $S$ for prolate
280 < ellipsoids :
281 < \[
249 < S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
250 < } }}{b},
251 < \]
270 > by\cite{Perrin1934,Perrin1936}
271 > \begin{equation}
272 > \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1
273 > \end{equation}
274 > where the semi-axes are of lengths $a$, $b$, and $c$. Due to the
275 > complexity of the elliptic integral, only uniaxial ellipsoids,
276 > {\it i.e.} prolate ($ a \ge b = c$) and oblate ($ a < b = c $), can
277 > be solved exactly. Introducing an elliptic integral parameter $S$ for
278 > prolate ellipsoids :
279 > \begin{equation}
280 > S = \frac{2}{\sqrt{a^2  - b^2}} \ln \frac{a + \sqrt{a^2  - b^2}}{b},
281 > \end{equation}
282   and oblate ellipsoids:
283 < \[
284 < S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
285 < }}{a},
256 < \]
283 > \begin{equation}
284 > S = \frac{2}{\sqrt {b^2  - a^2 }} \arctan \frac{\sqrt {b^2  - a^2}}{a},
285 > \end{equation}
286   one can write down the translational and rotational resistance
287 < tensors
287 > tensors for oblate,
288   \begin{eqnarray*}
289 < \Xi _a^{tt}  & = & 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}}. \\
290 < \Xi _b^{tt}  & = & \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S +
262 < 2a}},
289 > \Xi_a^{tt}  & = & 16\pi \eta \frac{a^2  - b^2}{(2a^2  - b^2 )S - 2a}. \\
290 > \Xi_b^{tt} =  \Xi_c^{tt} & = & 32\pi \eta \frac{a^2  - b^2 }{(2a^2 - 3b^2 )S + 2a},
291   \end{eqnarray*}
292 < and
292 > and prolate,
293   \begin{eqnarray*}
294 < \Xi _a^{rr} & = & \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}}, \\
295 < \Xi _b^{rr} & = & \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}}.
294 > \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2  - b^2 )b^2}{2a - b^2 S}, \\
295 > \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4  - b^4)}{(2a^2  - b^2 )S - 2a}
296   \end{eqnarray*}
297 + ellipsoids. For both spherical and ellipsoidal particles, the
298 + translation-rotation and rotation-translation coupling tensors are
299 + zero.
300  
301   \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
302  
# Line 273 | Line 304 | hydrodynamic properties of rigid bodies. However, sinc
304   analytical solution for the friction tensor for arbitrarily shaped
305   rigid molecules. The ellipsoid of revolution model and general
306   triaxial ellipsoid model have been used to approximate the
307 < hydrodynamic properties of rigid bodies. However, since the mapping
308 < from all possible ellipsoidal spaces, $r$-space, to all possible
309 < combination of rotational diffusion coefficients, $D$-space, is not
310 < unique\cite{Wegener1979} as well as the intrinsic coupling between
311 < translational and rotational motion of rigid bodies, general
312 < ellipsoids are not always suitable for modeling arbitrarily shaped
313 < rigid molecules. A number of studies have been devoted to
307 > hydrodynamic properties of rigid bodies. However, the mapping from all
308 > possible ellipsoidal spaces, $r$-space, to all possible combination of
309 > rotational diffusion coefficients, $D$-space, is not
310 > unique.\cite{Wegener1979} Additionally, because there is intrinsic
311 > coupling between translational and rotational motion of rigid bodies,
312 > general ellipsoids are not always suitable for modeling arbitrarily
313 > shaped rigid molecules.  A number of studies have been devoted to
314   determining the friction tensor for irregularly shaped rigid bodies
315 < using more advanced methods where the molecule of interest was
316 < modeled by a combinations of spheres\cite{Carrasco1999} and the
317 < hydrodynamics properties of the molecule can be calculated using the
318 < hydrodynamic interaction tensor. Let us consider a rigid assembly of
319 < $N$ beads immersed in a continuous medium. Due to hydrodynamic
320 < interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
321 < than its unperturbed velocity $v_i$,
315 > using more advanced methods where the molecule of interest was modeled
316 > by a combinations of spheres\cite{Carrasco1999} and the hydrodynamics
317 > properties of the molecule can be calculated using the hydrodynamic
318 > interaction tensor. Let us consider a rigid assembly of $N$ beads
319 > immersed in a continuous medium. Due to hydrodynamic interaction, the
320 > ``net'' velocity of $i$th bead, $v'_i$ is different than its
321 > unperturbed velocity $v_i$,
322   \[
323   v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
324   \]

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines