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

Comparing trunk/tengDissertation/Langevin.tex (file contents):
Revision 2908 by tim, Wed Jun 28 19:09:25 2006 UTC vs.
Revision 2909 by tim, Thu Jun 29 23:00:35 2006 UTC

# Line 4 | Line 4 | As an excellent alternative to newtonian dynamics, Lan
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{Palacios1998,Berkov2002,Denisov2003}. For instance, the
32 < oscillation power spectrum of nanoparticles from Langevin dynamics
33 < simulation has the same peak frequencies for different wave
34 < vectors,which recovers the property of magnetic excitations in small
35 < finite structures\cite{Berkov2005a}. In an attempt to reduce the
36 < computational cost of simulation, multiple time stepping (MTS)
37 < methods have been introduced and have been of great interest to
38 < macromolecule and protein community\cite{Tuckerman1992}. Relying on
39 < the observation that forces between distant atoms generally
40 < demonstrate slower fluctuations than forces between close atoms, MTS
41 < method are generally implemented by evaluating the slowly
42 < fluctuating forces less frequently than the fast ones.
43 < Unfortunately, nonlinear instability resulting from increasing
44 < timestep in MTS simulation have became a critical obstruction
45 < preventing the long time simulation. Due to the coupling to the heat
46 < bath, Langevin dynamics has been shown to be able to damp out the
47 < resonance artifact more efficiently\cite{Sandu1999}.
7 > As alternative to Newtonian dynamics, Langevin dynamics, which
8 > mimics a simple heat bath with stochastic and dissipative forces,
9 > has been applied in a variety of studies. The stochastic treatment
10 > of the solvent enables us to carry out substantially longer time
11 > simulations. Implicit solvent Langevin dynamics simulations of
12 > met-enkephalin not only outperform explicit solvent simulations for
13 > computational efficiency, but also agrees very well with explicit
14 > solvent simulations for dynamical properties\cite{Shen2002}.
15 > Recently, applying Langevin dynamics with the UNRES model, Liow and
16 > his coworkers suggest that protein folding pathways can be possibly
17 > explored within a reasonable amount of time\cite{Liwo2005}. The
18 > stochastic nature of the Langevin dynamics also enhances the
19 > sampling of the system and increases the probability of crossing
20 > energy barriers\cite{Banerjee2004, Cui2003}. Combining Langevin
21 > dynamics with Kramers's theory, Klimov and Thirumalai identified
22 > free-energy barriers by studying the viscosity dependence of the
23 > protein folding rates\cite{Klimov1997}. In order to account for
24 > solvent induced interactions missing from implicit solvent model,
25 > Kaya incorporated desolvation free energy barrier into implicit
26 > coarse-grained solvent model in protein folding/unfolding studies
27 > and discovered a higher free energy barrier between the native and
28 > denatured states. Because of its stability against noise, Langevin
29 > dynamics is very suitable for studying remagnetization processes in
30 > various systems\cite{Palacios1998,Berkov2002,Denisov2003}. For
31 > instance, the oscillation power spectrum of nanoparticles from
32 > Langevin dynamics simulation has the same peak frequencies for
33 > different wave vectors, which recovers the property of magnetic
34 > excitations in small finite structures\cite{Berkov2005a}.
35  
36   %review langevin/browninan dynamics for arbitrarily shaped rigid body
37   Combining Langevin or Brownian dynamics with rigid body dynamics,
38 < one can study the slow processes in biomolecular systems. Modeling
39 < the DNA as a chain of rigid spheres beads, which subject to harmonic
40 < potentials as well as excluded volume potentials, Mielke and his
41 < coworkers discover rapid superhelical stress generations from the
42 < stochastic simulation of twin supercoiling DNA with response to
43 < induced torques\cite{Mielke2004}. Membrane fusion is another key
44 < biological process which controls a variety of physiological
45 < functions, such as release of neurotransmitters \textit{etc}. A
46 < typical fusion event happens on the time scale of millisecond, which
47 < is impracticable to study using all atomistic model with newtonian
48 < mechanics. With the help of coarse-grained rigid body model and
49 < stochastic dynamics, the fusion pathways were exploited by many
38 > one can study slow processes in biomolecular systems. Modeling DNA
39 > as a chain of rigid beads, which are subject to harmonic potentials
40 > as well as excluded volume potentials, Mielke and his coworkers
41 > discovered rapid superhelical stress generations from the stochastic
42 > simulation of twin supercoiling DNA with response to induced
43 > torques\cite{Mielke2004}. Membrane fusion is another key biological
44 > process which controls a variety of physiological functions, such as
45 > release of neurotransmitters \textit{etc}. A typical fusion event
46 > happens on the time scale of millisecond, which is impractical to
47 > study using atomistic models with newtonian mechanics. With the help
48 > of coarse-grained rigid body model and stochastic dynamics, the
49 > fusion pathways were explored by many
50   researchers\cite{Noguchi2001,Noguchi2002,Shillcock2005}. Due to the
51 < difficulty of numerical integration of anisotropy rotation, most of
52 < the rigid body models are simply modeled by sphere, cylinder,
53 < ellipsoid or other regular shapes in stochastic simulations. In an
54 < effort to account for the diffusion anisotropy of the arbitrary
51 > difficulty of numerical integration of anisotropic rotation, most of
52 > the rigid body models are simply modeled using spheres, cylinders,
53 > ellipsoids or other regular shapes in stochastic simulations. In an
54 > effort to account for the diffusion anisotropy of arbitrary
55   particles, Fernandes and de la Torre improved the original Brownian
56   dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
57   incorporating a generalized $6\times6$ diffusion tensor and
58   introducing a simple rotation evolution scheme consisting of three
59   consecutive rotations\cite{Fernandes2002}. Unfortunately, unexpected
60 < error and bias are introduced into the system due to the arbitrary
61 < order of applying the noncommuting rotation
60 > errors and biases are introduced into the system due to the
61 > arbitrary order of applying the noncommuting rotation
62   operators\cite{Beard2003}. Based on the observation the momentum
63   relaxation time is much less than the time step, one may ignore the
64 < inertia in Brownian dynamics. However, assumption of the zero
64 > inertia in Brownian dynamics. However, the assumption of zero
65   average acceleration is not always true for cooperative motion which
66   is common in protein motion. An inertial Brownian dynamics (IBD) was
67   proposed to address this issue by adding an inertial correction
# Line 86 | Line 73 | typical nonskew bodies like cylinder and ellipsoid are
73   dynamics for nonskew bodies without translation-rotation coupling by
74   separating the translation and rotation motion and taking advantage
75   of the analytical solution of hydrodynamics properties. However,
76 < typical nonskew bodies like cylinder and ellipsoid are inadequate to
77 < represent most complex macromolecule assemblies. These intricate
76 > typical nonskew bodies like cylinders and ellipsoids are inadequate
77 > to represent most complex macromolecule assemblies. These intricate
78   molecules have been represented by a set of beads and their
79 < hydrodynamics properties can be calculated using variant
80 < hydrodynamic interaction tensors.
79 > hydrodynamic properties can be calculated using variants on the
80 > standard hydrodynamic interaction tensors.
81  
82   The goal of the present work is to develop a Langevin dynamics
83 < algorithm for arbitrary rigid particles by integrating the accurate
84 < estimation of friction tensor from hydrodynamics theory into the
85 < sophisticated rigid body dynamics.
83 > algorithm for arbitrary-shaped rigid particles by integrating the
84 > accurate estimation of friction tensor from hydrodynamics theory
85 > into the sophisticated rigid body dynamics algorithms.
86  
87   \section{Computational Methods{\label{methodSec}}}
88  
89   \subsection{\label{introSection:frictionTensor}Friction Tensor}
90   Theoretically, the friction kernel can be determined using the
91 < velocity autocorrelation function. However, this approach become
92 < impractical when the system become more and more complicate.
91 > velocity autocorrelation function. However, this approach becomes
92 > impractical when the system becomes more and more complicated.
93   Instead, various approaches based on hydrodynamics have been
94 < developed to calculate the friction coefficients. In general,
94 > developed to calculate the friction coefficients. In general, the
95   friction tensor $\Xi$ is a $6\times 6$ matrix given by
96   \[
97   \Xi  = \left( {\begin{array}{*{20}c}
# Line 112 | Line 99 | Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translatio
99     {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
100   \end{array}} \right).
101   \]
102 < Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
103 < tensor and rotational resistance (friction) tensor respectively,
104 < while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
105 < {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
106 < particle moves in a fluid, it may experience friction force or
107 < torque along the opposite direction of the velocity or angular
108 < velocity,
102 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are $3 \times 3$
103 > translational friction tensor and rotational resistance (friction)
104 > tensor respectively, while ${\Xi^{tr} }$ is translation-rotation
105 > coupling tensor and $ {\Xi^{rt} }$ is rotation-translation coupling
106 > tensor. When a particle moves in a fluid, it may experience friction
107 > force or torque along the opposite direction of the velocity or
108 > angular velocity,
109   \[
110   \left( \begin{array}{l}
111   F_R  \\
# Line 132 | Line 119 | toque.
119   \end{array} \right)
120   \]
121   where $F_r$ is the friction force and $\tau _R$ is the friction
122 < toque.
122 > torque.
123  
124 < \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
124 > \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
125  
126   For a spherical particle with slip boundary conditions, the
127   translational and rotational friction constant can be calculated
# Line 155 | Line 142 | hydrodynamics radius.
142   \end{array}} \right)
143   \]
144   where $\eta$ is the viscosity of the solvent and $R$ is the
145 < hydrodynamics radius.
145 > hydrodynamic radius.
146  
147 < Other non-spherical shape, such as cylinder and ellipsoid
148 < \textit{etc}, are widely used as reference for developing new
149 < hydrodynamics theory, because their properties can be calculated
150 < exactly. In 1936, Perrin extended Stokes's law to general
151 < ellipsoids, also called a triaxial ellipsoid, which is given in
152 < Cartesian coordinates by\cite{Perrin1934, Perrin1936}
147 > Other non-spherical shapes, such as cylinders and ellipsoids, are
148 > widely used as references for developing new hydrodynamics theory,
149 > because their properties can be calculated exactly. In 1936, Perrin
150 > extended Stokes's law to general ellipsoids, also called a triaxial
151 > ellipsoid, which is given in Cartesian coordinates
152 > by\cite{Perrin1934, Perrin1936}
153   \[
154   \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
155   }} = 1
156   \]
157   where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
158   due to the complexity of the elliptic integral, only the ellipsoid
159 < with the restriction of two axes having to be equal, \textit{i.e.}
159 > with the restriction of two axes being equal, \textit{i.e.}
160   prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
161   exactly. Introducing an elliptic integral parameter $S$ for prolate
162 < :
162 > ellipsoids :
163   \[
164   S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
165   } }}{b},
166   \]
167 < and oblate :
167 > and oblate ellipsoids:
168   \[
169   S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
170 < }}{a}.
170 > }}{a},
171   \]
172   one can write down the translational and rotational resistance
173   tensors
174 < \[
188 < \begin{array}{l}
174 > \begin{eqnarray*}
175   \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}}. \\
176   \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S +
177   2a}},
178 < \end{array}
193 < \]
178 > \end{eqnarray*}
179   and
180 < \[
196 < \begin{array}{l}
180 > \begin{eqnarray*}
181   \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}}, \\
182 < \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
183 < \end{array}.
200 < \]
182 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}}.
183 > \end{eqnarray}
184  
185   \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
186  
# Line 209 | Line 192 | translational and rotational motion of rigid body, gen
192   from all possible ellipsoidal spaces, $r$-space, to all possible
193   combination of rotational diffusion coefficients, $D$-space, is not
194   unique\cite{Wegener1979} as well as the intrinsic coupling between
195 < translational and rotational motion of rigid body, general
195 > translational and rotational motion of rigid bodies, general
196   ellipsoids are not always suitable for modeling arbitrarily shaped
197 < rigid molecule. A number of studies have been devoted to determine
198 < the friction tensor for irregularly shaped rigid bodies using more
199 < advanced method where the molecule of interest was modeled by
200 < combinations of spheres(beads)\cite{Carrasco1999} and the
197 > rigid molecules. A number of studies have been devoted to
198 > determining the friction tensor for irregularly shaped rigid bodies
199 > using more advanced methods where the molecule of interest was
200 > modeled by a combinations of spheres\cite{Carrasco1999} and the
201   hydrodynamics properties of the molecule can be calculated using the
202   hydrodynamic interaction tensor. Let us consider a rigid assembly of
203 < $N$ beads immersed in a continuous medium. Due to hydrodynamics
203 > $N$ beads immersed in a continuous medium. Due to hydrodynamic
204   interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
205   than its unperturbed velocity $v_i$,
206   \[
# Line 273 | Line 256 | where $\delta _{ij}$ is Kronecker delta function. Inve
256   B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
257   )T_{ij}
258   \]
259 < where $\delta _{ij}$ is Kronecker delta function. Inverting the $B$
260 < matrix, we obtain
259 > where $\delta _{ij}$ is the Kronecker delta function. Inverting the
260 > $B$ matrix, we obtain
261   \[
262   C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
263     {C_{11} } &  \ldots  & {C_{1N} }  \\
# Line 283 | Line 266 | $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U
266   \end{array}} \right),
267   \]
268   which can be partitioned into $N \times N$ $3 \times 3$ block
269 < $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
269 > $C_{ij}$. With the help of $C_{ij}$ and the skew matrix $U_i$
270   \[
271   U_i  = \left( {\begin{array}{*{20}c}
272     0 & { - z_i } & {y_i }  \\
# Line 292 | Line 275 | bead $i$ and origin $O$. Hence, the elements of resist
275   \end{array}} \right)
276   \]
277   where $x_i$, $y_i$, $z_i$ are the components of the vector joining
278 < bead $i$ and origin $O$. Hence, the elements of resistance tensor at
278 > bead $i$ and origin $O$, the elements of resistance tensor at
279   arbitrary origin $O$ can be written as
280   \begin{eqnarray}
281   \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
# Line 300 | Line 283 | arbitrary origin $O$ can be written as
283   \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
284   \label{introEquation:ResistanceTensorArbitraryOrigin}
285   \end{eqnarray}
303
286   The resistance tensor depends on the origin to which they refer. The
287 < proper location for applying friction force is the center of
287 > proper location for applying the friction force is the center of
288   resistance (or center of reaction), at which the trace of rotational
289   resistance tensor, $ \Xi ^{rr}$ reaches a minimum value.
290   Mathematically, the center of resistance is defined as an unique
291   point of the rigid body at which the translation-rotation coupling
292 < tensor are symmetric,
292 > tensors are symmetric,
293   \begin{equation}
294   \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
295   \label{introEquation:definitionCR}
296   \end{equation}
297   From Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
298 < we can easily find out that the translational resistance tensor is
298 > we can easily derive that the translational resistance tensor is
299   origin independent, while the rotational resistance tensor and
300   translation-rotation coupling resistance tensor depend on the
301   origin. Given the resistance tensor at an arbitrary origin $O$, and
# Line 359 | Line 341 | Consider a Langevin equation of motions in generalized
341  
342   \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
343  
344 < Consider a Langevin equation of motions in generalized coordinates
344 > Consider the Langevin equations of motion in generalized coordinates
345   \begin{equation}
346   M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i} (t)
347   \label{LDGeneralizedForm}
# Line 376 | Line 358 | in body-fixed frame and converted back to lab-fixed fr
358   in body-fixed frame and converted back to lab-fixed frame by:
359   \[
360   \begin{array}{l}
361 < F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
362 < F_{r,i}^l (t) = A^T F_{r,i}^b (t). \\
361 > F_{f,i}^l (t) = Q^T F_{f,i}^b (t), \\
362 > F_{r,i}^l (t) = Q^T F_{r,i}^b (t). \\
363   \end{array}
364   \]
365   Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
# Line 416 | Line 398 | of the resistance. Instead of integrating angular velo
398   \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
399   \end{equation}
400   where $r_{MR}$ is the vector from the center of mass to the center
401 < of the resistance. Instead of integrating angular velocity in
402 < lab-fixed frame, we consider the equation of motion of angular
403 < momentum in body-fixed frame
401 > of the resistance. Instead of integrating the angular velocity in
402 > lab-fixed frame, we consider the equation of angular momentum in
403 > body-fixed frame
404   \begin{equation}
405   \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
406   + \tau _{r,i}^b(t)
# Line 439 | Line 421 | $h= \delta t$:
421   {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
422      + \frac{h}{2} {\bf \tau}^b(t), \\
423   %
424 < \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
424 > \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
425      (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
426   \end{align*}
427   In this context, the $\mathrm{rotate}$ function is the reversible
# Line 450 | Line 432 | rotates both the rotation matrix ($\mathsf{A}$) and th
432   / 2) \cdot \mathsf{G}_x(a_x /2),
433   \end{equation}
434   where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
435 < rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
435 > rotates both the rotation matrix ($\mathsf{Q}$) and the body-fixed
436   angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
437   axis $\alpha$,
438   \begin{equation}
439   \mathsf{G}_\alpha( \theta ) = \left\{
440   \begin{array}{lcl}
441 < \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
441 > \mathsf{Q}(t) & \leftarrow & \mathsf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
442   {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
443   j}(0).
444   \end{array}
# Line 488 | Line 470 | calculated at the new positions and orientations
470   {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
471      \times \frac{\partial V}{\partial {\bf u}}, \\
472   %
473 < {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
473 > {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{Q}(t + h)
474      \cdot {\bf \tau}^s(t + h).
475   \end{align*}
476   Once the forces and torques have been obtained at the new time step,
# Line 508 | Line 490 | implemented in {\sc oopse}\cite{Meineke2005} and appli
490   \section{Results and Discussion}
491  
492   The Langevin algorithm described in previous section has been
493 < implemented in {\sc oopse}\cite{Meineke2005} and applied to the
494 < studies of kinetics and thermodynamic properties in several systems.
493 > implemented in {\sc oopse}\cite{Meineke2005} and applied to studies
494 > of the static and dynamic properties in several systems.
495  
496   \subsection{Temperature Control}
497  
# Line 520 | Line 502 | Initial configuration for the simulations is taken fro
502   the algorithm. In order to verify the stability of this new
503   algorithm, a series of simulations are performed on system
504   consisiting of 256 SSD water molecules with different viscosities.
505 < Initial configuration for the simulations is taken from a 1ns NVT
506 < simulation with a cubic box of 19.7166~\AA. All simulation are
505 > The initial configuration for the simulations is taken from a 1ns
506 > NVT simulation with a cubic box of 19.7166~\AA. All simulation are
507   carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
508 < with reference temperature at 300~K. Average temperature as a
508 > with reference temperature at 300~K. The average temperature as a
509   function of $\eta$ is shown in Table \ref{langevin:viscosity} where
510   the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
511   1$ poise. The better temperature control at higher viscosity can be
# Line 558 | Line 540 | NVE (see green curve in Fig.~\ref{langevin:temperature
540   Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
541   scaling to the desire temperature. In extremely lower friction
542   regime (when $ \eta  \approx 0$), Langevin dynamics becomes normal
543 < NVE (see green curve in Fig.~\ref{langevin:temperature}) which loses
544 < the temperature control ability.
543 > NVE (see orange curve in Fig.~\ref{langevin:temperature}) which
544 > loses the temperature control ability.
545  
546   \begin{figure}
547   \centering
# Line 578 | Line 560 | we create a rough shell model (see Fig.~\ref{langevin:
560   273~K. It has an equivalent implicit solvent system containing only
561   two banana shaped molecules with viscosity of 0.289 center poise. To
562   calculate the hydrodynamic properties of the banana shaped molecule,
563 < we create a rough shell model (see Fig.~\ref{langevin:roughShell}),
563 > we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
564   in which the banana shaped molecule is represented as a ``shell''
565   made of 2266 small identical beads with size of 0.3 \AA on the
566   surface. Applying the procedure described in
567   Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
568 < identified the center of resistance at $(0, 0.7482, -0.1988)$, as
569 < well as the resistance tensor,
568 > identified the center of resistance at $(0\AA, 0.7482\AA,
569 > -0.1988\AA)$, as well as the resistance tensor,
570   \[
571   \left( {\begin{array}{*{20}c}
572   0.9261 & 0 & 0&0&0.08585&0.2057\\
# Line 606 | Line 588 | Curves of velocity auto-correlation functions in
588   %\end{array}} \right).
589   %\]
590  
591 < Curves of velocity auto-correlation functions in
591 > Curves of the velocity auto-correlation functions in
592   Fig.~\ref{langevin:vacf} were shown to match each other very well.
593   However, because of the stochastic nature, simulation using Langevin
594 < dynamics was shown to decay slightly fast. In order to study the
595 < rotational motion of the molecules, we also calculated the auto-
596 < correlation function of the principle axis of the second GB
597 < particle, $u$.
594 > dynamics was shown to decay slightly faster than MD. In order to
595 > study the rotational motion of the molecules, we also calculated the
596 > auto- correlation function of the principle axis of the second GB
597 > particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
598 > probably due to the reason that the viscosity using in the
599 > simulations only partially preserved the dynamics of the system.
600  
601   \begin{figure}
602   \centering
# Line 633 | Line 617 | auto-correlation functions in NVE (blue) and Langevin
617   \centering
618   \includegraphics[width=\linewidth]{vacf.eps}
619   \caption[Plots of Velocity Auto-correlation Functions]{Velocity
620 < auto-correlation functions in NVE (blue) and Langevin dynamics
621 < (red).} \label{langevin:vacf}
620 > auto-correlation functions of NVE (explicit solvent) in blue) and
621 > Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
622   \end{figure}
623  
624   \begin{figure}
# Line 642 | Line 626 | of the middle GB particle in NVE (blue) and Langevin d
626   \includegraphics[width=\linewidth]{uacf.eps}
627   \caption[Auto-correlation functions of the principle axis of the
628   middle GB particle]{Auto-correlation functions of the principle axis
629 < of the middle GB particle in NVE (blue) and Langevin dynamics
630 < (red).} \label{langevin:twoBanana}
629 > of the middle GB particle of NVE (blue) and Langevin dynamics
630 > (red).} \label{langevin:uacf}
631   \end{figure}
632  
633   \section{Conclusions}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines