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 2851 by tim, Sun Jun 11 02:06:01 2006 UTC vs.
Revision 2905 by tim, Wed Jun 28 19:09:25 2006 UTC

# Line 1 | Line 1
1  
2 < \chapter{\label{chapt:methodology}Langevin Dynamics for Rigid Bodies of Arbitrary Shape}
2 > \chapter{\label{chapt:langevin}LANGEVIN DYNAMICS FOR RIGID BODIES OF ARBITRARY SHAPE}
3  
4   \section{Introduction}
5  
# Line 28 | Line 28 | systems\cite{Garcia-Palacios1998,Berkov2002,Denisov200
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}.
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}.
48  
50 %review rigid body dynamics
51 Rigid bodies are frequently involved in the modeling of different
52 areas, from engineering, physics, to chemistry. For example,
53 missiles and vehicle are usually modeled by rigid bodies.  The
54 movement of the objects in 3D gaming engine or other physics
55 simulator is governed by the rigid body dynamics. In molecular
56 simulation, rigid body is used to simplify the model in
57 protein-protein docking study{\cite{Gray2003}}.
58
59 It is very important to develop stable and efficient methods to
60 integrate the equations of motion of orientational degrees of
61 freedom. Euler angles are the nature choice to describe the
62 rotational degrees of freedom. However, due to its singularity, the
63 numerical integration of corresponding equations of motion is very
64 inefficient and inaccurate. Although an alternative integrator using
65 different sets of Euler angles can overcome this difficulty\cite{},
66 the computational penalty and the lost of angular momentum
67 conservation still remain. In 1977, a singularity free
68 representation utilizing quaternions was developed by
69 Evans\cite{Evans1977}. Unfortunately, this approach suffer from the
70 nonseparable Hamiltonian resulted from quaternion representation,
71 which prevents the symplectic algorithm to be utilized. Another
72 different approach is to apply holonomic constraints to the atoms
73 belonging to the rigid body\cite{}. Each atom moves independently
74 under the normal forces deriving from potential energy and
75 constraint forces which are used to guarantee the rigidness.
76 However, due to their iterative nature, SHAKE and Rattle algorithm
77 converge very slowly when the number of constraint increases.
78
79 The break through in geometric literature suggests that, in order to
80 develop a long-term integration scheme, one should preserve the
81 geometric structure of the flow. Matubayasi and Nakahara developed a
82 time-reversible integrator for rigid bodies in quaternion
83 representation. Although it is not symplectic, this integrator still
84 demonstrates a better long-time energy conservation than traditional
85 methods because of the time-reversible nature. Extending
86 Trotter-Suzuki to general system with a flat phase space, Miller and
87 his colleagues devised an novel symplectic, time-reversible and
88 volume-preserving integrator in quaternion representation. However,
89 all of the integrators in quaternion representation suffer from the
90 computational penalty of constructing a rotation matrix from
91 quaternions to evolve coordinates and velocities at every time step.
92 An alternative integration scheme utilizing rotation matrix directly
93 is RSHAKE , in which a conjugate momentum to rotation matrix is
94 introduced to re-formulate the Hamiltonian's equation and the
95 Hamiltonian is evolved in a constraint manifold by iteratively
96 satisfying the orthogonality constraint. However, RSHAKE is
97 inefficient because of the iterative procedure. An extremely
98 efficient integration scheme in rotation matrix representation,
99 which also preserves the same structural properties of the
100 Hamiltonian flow as Miller's integrator, is proposed by Dullweber,
101 Leimkuhler and McLachlan (DLM).
102
49   %review langevin/browninan dynamics for arbitrarily shaped rigid body
50   Combining Langevin or Brownian dynamics with rigid body dynamics,
51   one can study the slow processes in biomolecular systems. Modeling
# Line 132 | Line 78 | term\cite{Beard2001}. As a complement to IBD which has
78   average acceleration is not always true for cooperative motion which
79   is common in protein motion. An inertial Brownian dynamics (IBD) was
80   proposed to address this issue by adding an inertial correction
81 < term\cite{Beard2001}. As a complement to IBD which has a lower bound
81 > term\cite{Beard2000}. As a complement to IBD which has a lower bound
82   in time step because of the inertial relaxation time, long-time-step
83   inertial dynamics (LTID) can be used to investigate the inertial
84   behavior of the polymer segments in low friction
85 < regime\cite{Beard2001}. LTID can also deal with the rotational
85 > regime\cite{Beard2000}. LTID can also deal with the rotational
86   dynamics for nonskew bodies without translation-rotation coupling by
87   separating the translation and rotation motion and taking advantage
88   of the analytical solution of hydrodynamics properties. However,
# Line 151 | Line 97 | sophisticated rigid body dynamics.
97   estimation of friction tensor from hydrodynamics theory into the
98   sophisticated rigid body dynamics.
99  
100 < \section{Method{\label{methodSec}}}
100 > \section{Computational Methods{\label{methodSec}}}
101  
102 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
103 < Theoretically, the friction kernel can be determined using velocity
104 < autocorrelation function. However, this approach become impractical
105 < when the system become more and more complicate. Instead, various
106 < approaches based on hydrodynamics have been developed to calculate
107 < the friction coefficients. The friction effect is isotropic in
108 < Equation, $\zeta$ can be taken as a scalar. In general, friction
163 < tensor $\Xi$ is a $6\times 6$ matrix given by
102 > \subsection{\label{introSection:frictionTensor}Friction Tensor}
103 > Theoretically, the friction kernel can be determined using the
104 > velocity autocorrelation function. However, this approach become
105 > impractical when the system become more and more complicate.
106 > Instead, various approaches based on hydrodynamics have been
107 > developed to calculate the friction coefficients. In general,
108 > friction tensor $\Xi$ is a $6\times 6$ matrix given by
109   \[
110   \Xi  = \left( {\begin{array}{*{20}c}
111     {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
# Line 191 | Line 136 | For a spherical particle, the translational and rotati
136  
137   \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
138  
139 < For a spherical particle, the translational and rotational friction
140 < constant can be calculated from Stoke's law,
139 > For a spherical particle with slip boundary conditions, the
140 > translational and rotational friction constant can be calculated
141 > from Stoke's law,
142   \[
143   \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
144     {6\pi \eta R} & 0 & 0  \\
# Line 214 | Line 160 | exactly. In 1936, Perrin extended Stokes's law to gene
160   Other non-spherical shape, such as cylinder and ellipsoid
161   \textit{etc}, are widely used as reference for developing new
162   hydrodynamics theory, because their properties can be calculated
163 < exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
164 < also called a triaxial ellipsoid, which is given in Cartesian
165 < coordinates by\cite{Perrin1934, Perrin1936}
163 > exactly. In 1936, Perrin extended Stokes's law to general
164 > ellipsoids, also called a triaxial ellipsoid, which is given in
165 > Cartesian coordinates by\cite{Perrin1934, Perrin1936}
166   \[
167   \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
168   }} = 1
# Line 225 | Line 171 | exactly. Introducing an elliptic integral parameter $S
171   due to the complexity of the elliptic integral, only the ellipsoid
172   with the restriction of two axes having to be equal, \textit{i.e.}
173   prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
174 < exactly. Introducing an elliptic integral parameter $S$ for prolate,
174 > exactly. Introducing an elliptic integral parameter $S$ for prolate
175 > :
176   \[
177   S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
178   } }}{b},
179   \]
180 < and oblate,
180 > and oblate :
181   \[
182   S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
183 < }}{a}
184 < \],
183 > }}{a}.
184 > \]
185   one can write down the translational and rotational resistance
186   tensors
187   \[
188   \begin{array}{l}
189 < \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
190 < \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
191 < \end{array},
189 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}}. \\
190 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S +
191 > 2a}},
192 > \end{array}
193   \]
194   and
195   \[
196   \begin{array}{l}
197 < \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
197 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}}, \\
198   \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
199   \end{array}.
200   \]
201  
202 < \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shape}}
202 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
203  
204 < Unlike spherical and other regular shaped molecules, there is not
205 < analytical solution for friction tensor of any arbitrary shaped
204 > Unlike spherical and other simply shaped molecules, there is no
205 > analytical solution for the friction tensor for arbitrarily shaped
206   rigid molecules. The ellipsoid of revolution model and general
207   triaxial ellipsoid model have been used to approximate the
208   hydrodynamic properties of rigid bodies. However, since the mapping
209 < from all possible ellipsoidal space, $r$-space, to all possible
210 < combination of rotational diffusion coefficients, $D$-space is not
209 > from all possible ellipsoidal spaces, $r$-space, to all possible
210 > combination of rotational diffusion coefficients, $D$-space, is not
211   unique\cite{Wegener1979} as well as the intrinsic coupling between
212 < translational and rotational motion of rigid body, general ellipsoid
213 < is not always suitable for modeling arbitrarily shaped rigid
214 < molecule. A number of studies have been devoted to determine the
215 < friction tensor for irregularly shaped rigid bodies using more
212 > translational and rotational motion of rigid body, general
213 > ellipsoids are not always suitable for modeling arbitrarily shaped
214 > rigid molecule. A number of studies have been devoted to determine
215 > the friction tensor for irregularly shaped rigid bodies using more
216   advanced method where the molecule of interest was modeled by
217   combinations of spheres(beads)\cite{Carrasco1999} and the
218   hydrodynamics properties of the molecule can be calculated using the
# Line 283 | Line 231 | This equation is the basis for deriving the hydrodynam
231   \label{introEquation:tensorExpression}
232   \end{equation}
233   This equation is the basis for deriving the hydrodynamic tensor. In
234 < 1930, Oseen and Burgers gave a simple solution to Equation
235 < \ref{introEquation:tensorExpression}
234 > 1930, Oseen and Burgers gave a simple solution to
235 > Eq.~\ref{introEquation:tensorExpression}
236   \begin{equation}
237   T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
238   R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
# Line 300 | Line 248 | Both of the Equation \ref{introEquation:oseenTensor} a
248   \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
249   \label{introEquation:RPTensorNonOverlapped}
250   \end{equation}
251 < Both of the Equation \ref{introEquation:oseenTensor} and Equation
252 < \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
253 < \ge \sigma _i  + \sigma _j$. An alternative expression for
251 > Both of the Eq.~\ref{introEquation:oseenTensor} and
252 > Eq.~\ref{introEquation:RPTensorNonOverlapped} have an assumption
253 > $R_{ij} \ge \sigma _i  + \sigma _j$. An alternative expression for
254   overlapping beads with the same radius, $\sigma$, is given by
255   \begin{equation}
256   T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
# Line 310 | Line 258 | T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left(
258   \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
259   \label{introEquation:RPTensorOverlapped}
260   \end{equation}
313
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
# Line 326 | Line 273 | where $\delta _{ij}$ is Kronecker delta function. Inve
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
331 <
276 > where $\delta _{ij}$ is Kronecker delta function. Inverting the $B$
277 > matrix, we obtain
278   \[
279   C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
280     {C_{11} } &  \ldots  & {C_{1N} }  \\
281      \vdots  &  \ddots  &  \vdots   \\
282     {C_{N1} } &  \cdots  & {C_{NN} }  \\
283 < \end{array}} \right)
283 > \end{array}} \right),
284   \]
285 < , which can be partitioned into $N \times N$ $3 \times 3$ block
285 > which can be partitioned into $N \times N$ $3 \times 3$ block
286   $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
287   \[
288   U_i  = \left( {\begin{array}{*{20}c}
# Line 348 | Line 294 | arbitrary origin $O$ can be written as
294   where $x_i$, $y_i$, $z_i$ are the components of the vector joining
295   bead $i$ and origin $O$. Hence, the elements of resistance tensor at
296   arbitrary origin $O$ can be written as
297 < \begin{equation}
298 < \begin{array}{l}
353 < \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
297 > \begin{eqnarray}
298 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
299   \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
300 < \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
356 < \end{array}
300 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
301   \label{introEquation:ResistanceTensorArbitraryOrigin}
302 < \end{equation}
302 > \end{eqnarray}
303  
304   The resistance tensor depends on the origin to which they refer. The
305   proper location for applying friction force is the center of
306 < resistance (reaction), at which the trace of rotational resistance
307 < tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
308 < resistance is defined as an unique point of the rigid body at which
309 < the translation-rotation coupling tensor are symmetric,
306 > resistance (or center of reaction), at which the trace of rotational
307 > resistance tensor, $ \Xi ^{rr}$ reaches a minimum value.
308 > Mathematically, the center of resistance is defined as an unique
309 > point of the rigid body at which the translation-rotation coupling
310 > tensor are symmetric,
311   \begin{equation}
312   \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
313   \label{introEquation:definitionCR}
314   \end{equation}
315 < Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
315 > From Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
316   we can easily find out that the translational resistance tensor is
317   origin independent, while the rotational resistance tensor and
318   translation-rotation coupling resistance tensor depend on the
319 < origin. Given resistance tensor at an arbitrary origin $O$, and a
320 < vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
319 > origin. Given the resistance tensor at an arbitrary origin $O$, and
320 > a vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
321   obtain the resistance tensor at $P$ by
322   \begin{equation}
323   \begin{array}{l}
# Line 390 | Line 335 | Using Equations \ref{introEquation:definitionCR} and
335     { - y_{OP} } & {x_{OP} } & 0  \\
336   \end{array}} \right)
337   \]
338 < Using Equations \ref{introEquation:definitionCR} and
339 < \ref{introEquation:resistanceTensorTransformation}, one can locate
340 < the position of center of resistance,
338 > Using Eq.~\ref{introEquation:definitionCR} and
339 > Eq.~\ref{introEquation:resistanceTensorTransformation}, one can
340 > locate the position of center of resistance,
341   \begin{eqnarray*}
342   \left( \begin{array}{l}
343   x_{OR}  \\
# Line 409 | Line 354 | the position of center of resistance,
354   (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
355   \end{array} \right) \\
356   \end{eqnarray*}
412
357   where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
358   joining center of resistance $R$ and origin $O$.
359  
360 < \subsection{Langevin dynamics for rigid particles of arbitrary shape}
360 > \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
361  
362   Consider a Langevin equation of motions in generalized coordinates
363   \begin{equation}
# Line 422 | Line 366 | $V_i = V_i(v_i,\omega _i)$. The right side of Eq.
366   \end{equation}
367   where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
368   and moment of inertial) matrix and $V_i$ is a generalized velocity,
369 < $V_i = V_i(v_i,\omega _i)$. The right side of Eq.
370 < (\ref{LDGeneralizedForm}) consists of three generalized forces in
369 > $V_i = V_i(v_i,\omega _i)$. The right side of
370 > Eq.~\ref{LDGeneralizedForm} consists of three generalized forces in
371   lab-fixed frame, systematic force $F_{s,i}$, dissipative force
372   $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
373   system in Newtownian mechanics typically refers to lab-fixed frame,
# Line 433 | Line 377 | in body-fixed frame and converted back to lab-fixed fr
377   \[
378   \begin{array}{l}
379   F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
380 < F_{r,i}^l (t) = A^T F_{r,i}^b (t) \\
381 < \end{array}.
380 > F_{r,i}^l (t) = A^T F_{r,i}^b (t). \\
381 > \end{array}
382   \]
383   Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
384   the body-fixed velocity at center of resistance $v_{R,i}^b$ and
385 < angular velocity $\omega _i$,
385 > angular velocity $\omega _i$
386   \begin{equation}
387   F_{r,i}^b (t) = \left( \begin{array}{l}
388   f_{r,i}^b (t) \\
# Line 456 | Line 400 | with zero mean and variance
400   \begin{equation}
401   \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle  =
402   \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle  =
403 < 2k_B T\Xi _R \delta (t - t').
403 > 2k_B T\Xi _R \delta (t - t'). \label{randomForce}
404   \end{equation}
461
405   The equation of motion for $v_i$ can be written as
406   \begin{equation}
407   m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
# Line 480 | Line 423 | + \tau _{r,i}^b(t)
423   \dot j_i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b (t)
424   + \tau _{r,i}^b(t)
425   \end{equation}
483
426   Embedding the friction terms into force and torque, one can
427   integrate the langevin equations of motion for rigid body of
428   arbitrary shape in a velocity-Verlet style 2-part algorithm, where
# Line 500 | Line 442 | $h= \delta t$:
442   \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
443      (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
444   \end{align*}
503
445   In this context, the $\mathrm{rotate}$ function is the reversible
446   product of the three body-fixed rotations,
447   \begin{equation}
# Line 535 | Line 476 | All other rotations follow in a straightforward manner
476   \end{array}
477   \right).
478   \end{equation}
479 < All other rotations follow in a straightforward manner.
479 > All other rotations follow in a straightforward manner. After the
480 > first part of the propagation, the forces and body-fixed torques are
481 > calculated at the new positions and orientations
482  
540 After the first part of the propagation, the forces and body-fixed
541 torques are calculated at the new positions and orientations
542
483   {\tt doForces:}
484   \begin{align*}
485   {\bf f}(t + h) &\leftarrow
# Line 551 | Line 491 | torques are calculated at the new positions and orient
491   {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
492      \cdot {\bf \tau}^s(t + h).
493   \end{align*}
494 + Once the forces and torques have been obtained at the new time step,
495 + the velocities can be advanced to the same time value.
496  
555 {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
556 $\mathsf{A}$ is calculated in {\tt moveA}.  Once the forces and
557 torques have been obtained at the new time step, the velocities can
558 be advanced to the same time value.
559
497   {\tt moveB:}
498   \begin{align*}
499   {\bf v}\left(t + h \right)  &\leftarrow  {\bf v}\left(t + h / 2
# Line 568 | Line 505 | be advanced to the same time value.
505      + \frac{h}{2} {\bf \tau}^b(t + h) .
506   \end{align*}
507  
508 < \section{Results and discussion}
508 > \section{Results and Discussion}
509  
510 + The Langevin algorithm described in previous section has been
511 + implemented in {\sc oopse}\cite{Meineke2005} and applied to the
512 + studies of kinetics and thermodynamic properties in several systems.
513 +
514 + \subsection{Temperature Control}
515 +
516 + As shown in Eq.~\ref{randomForce}, random collisions associated with
517 + the solvent's thermal motions is controlled by the external
518 + temperature. The capability to maintain the temperature of the whole
519 + system was usually used to measure the stability and efficiency of
520 + the algorithm. In order to verify the stability of this new
521 + algorithm, a series of simulations are performed on system
522 + consisiting of 256 SSD water molecules with different viscosities.
523 + Initial configuration for the simulations is taken from a 1ns NVT
524 + simulation with a cubic box of 19.7166~\AA. All simulation are
525 + carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
526 + with reference temperature at 300~K. Average temperature as a
527 + function of $\eta$ is shown in Table \ref{langevin:viscosity} where
528 + the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
529 + 1$ poise. The better temperature control at higher viscosity can be
530 + explained by the finite size effect and relative slow relaxation
531 + rate at lower viscosity regime.
532 + \begin{table}
533 + \caption{AVERAGE TEMPERATURES FROM LANGEVIN DYNAMICS SIMULATIONS OF
534 + SSD WATER MOLECULES WITH REFERENCE TEMPERATURE AT 300~K.}
535 + \label{langevin:viscosity}
536 + \begin{center}
537 + \begin{tabular}{lll}
538 +  \hline
539 +  $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
540 +  \hline
541 +  1    & 300.47 & 10.99 \\
542 +  0.1  & 301.19 & 11.136 \\
543 +  0.01 & 303.04 & 11.796 \\
544 +  \hline
545 + \end{tabular}
546 + \end{center}
547 + \end{table}
548 +
549 + Another set of calculation were performed to study the efficiency of
550 + temperature control using different temperature coupling schemes.
551 + The starting configuration is cooled to 173~K and evolved using NVE,
552 + NVT, and Langevin dynamic with time step of 2 fs.
553 + Fig.~\ref{langevin:temperature} shows the heating curve obtained as
554 + the systems reach equilibrium. The orange curve in
555 + Fig.~\ref{langevin:temperature} represents the simulation using
556 + Nos\'e-Hoover temperature scaling scheme with thermostat of 5 ps
557 + which gives reasonable tight coupling, while the blue one from
558 + Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
559 + scaling to the desire temperature. In extremely lower friction
560 + regime (when $ \eta  \approx 0$), Langevin dynamics becomes normal
561 + NVE (see green curve in Fig.~\ref{langevin:temperature}) which loses
562 + the temperature control ability.
563 +
564 + \begin{figure}
565 + \centering
566 + \includegraphics[width=\linewidth]{temperature.eps}
567 + \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
568 + temperature fluctuation versus time.} \label{langevin:temperature}
569 + \end{figure}
570 +
571 + \subsection{Langevin Dynamics of Banana Shaped Molecules}
572 +
573 + In order to verify that Langevin dynamics can mimic the dynamics of
574 + the systems absent of explicit solvents, we carried out two sets of
575 + simulations and compare their dynamic properties.
576 + Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
577 + made of 256 pentane molecules and two banana shaped molecules at
578 + 273~K. It has an equivalent implicit solvent system containing only
579 + two banana shaped molecules with viscosity of 0.289 center poise. To
580 + calculate the hydrodynamic properties of the banana shaped molecule,
581 + we create a rough shell model (see Fig.~\ref{langevin:roughShell}),
582 + in which the banana shaped molecule is represented as a ``shell''
583 + made of 2266 small identical beads with size of 0.3 \AA on the
584 + surface. Applying the procedure described in
585 + Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
586 + identified the center of resistance at $(0, 0.7482, -0.1988)$, as
587 + well as the resistance tensor,
588 + \[
589 + \left( {\begin{array}{*{20}c}
590 + 0.9261 & 0 & 0&0&0.08585&0.2057\\
591 + 0& 0.9270&-0.007063& 0.08585&0&0\\
592 + 0&0.007063&0.7494&0.2057&0&0\\
593 + 0&0.0858&0.2057& 58.64& 0&-8.5736\\
594 + 0.08585&0&0&0&48.30&3.219&\\
595 + 0.2057&0&0&0&3.219&10.7373\\
596 + \end{array}} \right).
597 + \]
598 + %\[
599 + %\left( {\begin{array}{*{20}c}
600 + %0.9261 & 1.310e-14 & -7.292e-15&5.067e-14&0.08585&0.2057\\
601 + %3.968e-14& 0.9270&-0.007063& 0.08585&6.764e-14&4.846e-14\\
602 + %-6.561e-16&-0.007063&0.7494&0.2057&4.846e-14&1.5036e-14\\
603 + %5.067e-14&0.0858&0.2057& 58.64& 8.563e-13&-8.5736\\
604 + %0.08585&6.764e-14&4.846e-14&1.555e-12&48.30&3.219&\\
605 + %0.2057&4.846e-14&1.5036e-14&-3.904e-13&3.219&10.7373\\
606 + %\end{array}} \right).
607 + %\]
608 +
609 + Curves of velocity auto-correlation functions in
610 + Fig.~\ref{langevin:vacf} were shown to match each other very well.
611 + However, because of the stochastic nature, simulation using Langevin
612 + dynamics was shown to decay slightly fast. In order to study the
613 + rotational motion of the molecules, we also calculated the auto-
614 + correlation function of the principle axis of the second GB
615 + particle, $u$.
616 +
617 + \begin{figure}
618 + \centering
619 + \includegraphics[width=\linewidth]{roughShell.eps}
620 + \caption[Rough shell model for banana shaped molecule]{Rough shell
621 + model for banana shaped molecule.} \label{langevin:roughShell}
622 + \end{figure}
623 +
624 + \begin{figure}
625 + \centering
626 + \includegraphics[width=\linewidth]{twoBanana.eps}
627 + \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
628 + 256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
629 + molecules and 256 pentane molecules.} \label{langevin:twoBanana}
630 + \end{figure}
631 +
632 + \begin{figure}
633 + \centering
634 + \includegraphics[width=\linewidth]{vacf.eps}
635 + \caption[Plots of Velocity Auto-correlation Functions]{Velocity
636 + auto-correlation functions in NVE (blue) and Langevin dynamics
637 + (red).} \label{langevin:vacf}
638 + \end{figure}
639 +
640 + \begin{figure}
641 + \centering
642 + \includegraphics[width=\linewidth]{uacf.eps}
643 + \caption[Auto-correlation functions of the principle axis of the
644 + middle GB particle]{Auto-correlation functions of the principle axis
645 + of the middle GB particle in NVE (blue) and Langevin dynamics
646 + (red).} \label{langevin:twoBanana}
647 + \end{figure}
648 +
649   \section{Conclusions}
650 +
651 + We have presented a new Langevin algorithm by incorporating the
652 + hydrodynamics properties of arbitrary shaped molecules into an
653 + advanced symplectic integration scheme. The temperature control
654 + ability of this algorithm was demonstrated by a set of simulations
655 + with different viscosities. It was also shown to have significant
656 + advantage of producing rapid thermal equilibration over
657 + Nos\'{e}-Hoover method. Further studies in systems involving banana
658 + shaped molecules illustrated that the dynamic properties could be
659 + preserved by using this new algorithm as an implicit solvent model.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines