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

Comparing trunk/tengDissertation/Introduction.tex (file contents):
Revision 2725 by tim, Fri Apr 21 05:45:14 2006 UTC vs.
Revision 2793 by tim, Mon Jun 5 21:24:52 2006 UTC

# Line 93 | Line 93 | the kinetic, $K$, and potential energies, $U$ \cite{to
93   The actual trajectory, along which a dynamical system may move from
94   one point to another within a specified time, is derived by finding
95   the path which minimizes the time integral of the difference between
96 < the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
96 > the kinetic, $K$, and potential energies, $U$ \cite{Tolman1979}.
97   \begin{equation}
98   \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
99   \label{introEquation:halmitonianPrinciple1}
# Line 189 | Line 189 | known as the canonical equations of motions \cite{Gold
189   Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
190   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
191   equation of motion. Due to their symmetrical formula, they are also
192 < known as the canonical equations of motions \cite{Goldstein01}.
192 > known as the canonical equations of motions \cite{Goldstein2001}.
193  
194   An important difference between Lagrangian approach and the
195   Hamiltonian approach is that the Lagrangian is considered to be a
# Line 200 | Line 200 | equations\cite{Marion90}.
200   appropriate for application to statistical mechanics and quantum
201   mechanics, since it treats the coordinate and its time derivative as
202   independent variables and it only works with 1st-order differential
203 < equations\cite{Marion90}.
203 > equations\cite{Marion1990}.
204  
205   In Newtonian Mechanics, a system described by conservative forces
206   conserves the total energy \ref{introEquation:energyConservation}.
# Line 470 | Line 470 | statistical ensemble are identical \cite{Frenkel1996,
470   many-body system in Statistical Mechanics. Fortunately, Ergodic
471   Hypothesis is proposed to make a connection between time average and
472   ensemble average. It states that time average and average over the
473 < statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
473 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
474   \begin{equation}
475   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
476   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 484 | Line 484 | reasonable, the Monte Carlo techniques\cite{metropolis
484   a properly weighted statistical average. This allows the researcher
485   freedom of choice when deciding how best to measure a given
486   observable. In case an ensemble averaged approach sounds most
487 < reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
487 > reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
488   utilized. Or if the system lends itself to a time averaging
489   approach, the Molecular Dynamics techniques in
490   Sec.~\ref{introSection:molecularDynamics} will be the best
# Line 498 | Line 498 | issue. The velocity verlet method, which happens to be
498   within the equations. Since 1990, geometric integrators, which
499   preserve various phase-flow invariants such as symplectic structure,
500   volume and time reversal symmetry, are developed to address this
501 < issue. The velocity verlet method, which happens to be a simple
502 < example of symplectic integrator, continues to gain its popularity
503 < in molecular dynamics community. This fact can be partly explained
504 < by its geometric nature.
501 > issue\cite{Dullweber1997, McLachlan1998, Leimkuhler1999}. The
502 > velocity verlet method, which happens to be a simple example of
503 > symplectic integrator, continues to gain its popularity in molecular
504 > dynamics community. This fact can be partly explained by its
505 > geometric nature.
506  
507   \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
508   A \emph{manifold} is an abstract mathematical space. It locally
# Line 565 | Line 566 | Another generalization of Hamiltonian dynamics is Pois
566   \end{equation}In this case, $f$ is
567   called a \emph{Hamiltonian vector field}.
568  
569 < Another generalization of Hamiltonian dynamics is Poisson Dynamics,
569 > Another generalization of Hamiltonian dynamics is Poisson
570 > Dynamics\cite{Olver1986},
571   \begin{equation}
572   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
573   \end{equation}
# Line 612 | Line 614 | The hidden geometric properties of ODE and its flow pl
614  
615   \subsection{\label{introSection:geometricProperties}Geometric Properties}
616  
617 < The hidden geometric properties of ODE and its flow play important
618 < roles in numerical studies. Many of them can be found in systems
619 < which occur naturally in applications.
617 > The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
618 > and its flow play important roles in numerical studies. Many of them
619 > can be found in systems which occur naturally in applications.
620  
621   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
622   a \emph{symplectic} flow if it satisfies,
# Line 658 | Line 660 | smooth function $G$ is given by,
660   which is the condition for conserving \emph{first integral}. For a
661   canonical Hamiltonian system, the time evolution of an arbitrary
662   smooth function $G$ is given by,
663 < \begin{equation}
664 < \begin{array}{c}
665 < \frac{{dG(x(t))}}{{dt}} = [\nabla _x G(x(t))]^T \dot x(t) \\
666 <  = [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
665 < \end{array}
663 >
664 > \begin{eqnarray}
665 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
666 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
667   \label{introEquation:firstIntegral1}
668 < \end{equation}
668 > \end{eqnarray}
669 >
670 >
671   Using poisson bracket notion, Equation
672   \ref{introEquation:firstIntegral1} can be rewritten as
673   \[
# Line 679 | Line 682 | is a \emph{first integral}, which is due to the fact $
682   is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
683   0$.
684  
685 <
683 < When designing any numerical methods, one should always try to
685 > When designing any numerical methods, one should always try to
686   preserve the structural properties of the original ODE and its flow.
687  
688   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
# Line 697 | Line 699 | Generating function tends to lead to methods which are
699   \item Splitting methods
700   \end{enumerate}
701  
702 < Generating function tends to lead to methods which are cumbersome
703 < and difficult to use. In dissipative systems, variational methods
704 < can capture the decay of energy accurately. Since their
705 < geometrically unstable nature against non-Hamiltonian perturbations,
706 < ordinary implicit Runge-Kutta methods are not suitable for
707 < Hamiltonian system. Recently, various high-order explicit
708 < Runge--Kutta methods have been developed to overcome this
702 > Generating function\cite{Channell1990} tends to lead to methods
703 > which are cumbersome and difficult to use. In dissipative systems,
704 > variational methods can capture the decay of energy
705 > accurately\cite{Kane2000}. Since their geometrically unstable nature
706 > against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
707 > methods are not suitable for Hamiltonian system. Recently, various
708 > high-order explicit Runge-Kutta methods
709 > \cite{Owren1992,Chen2003}have been developed to overcome this
710   instability. However, due to computational penalty involved in
711   implementing the Runge-Kutta methods, they do not attract too much
712   attention from Molecular Dynamics community. Instead, splitting have
713   been widely accepted since they exploit natural decompositions of
714 < the system\cite{Tuckerman92}.
714 > the system\cite{Tuckerman1992, McLachlan1998}.
715  
716   \subsubsection{\label{introSection:splittingMethod}Splitting Method}
717  
# Line 831 | Line 834 | $\varphi_1(t)$ and $\varphi_2(t$ respectively , we hav
834   error of splitting method in terms of commutator of the
835   operators(\ref{introEquation:exponentialOperator}) associated with
836   the sub-flow. For operators $hX$ and $hY$ which are associate to
837 < $\varphi_1(t)$ and $\varphi_2(t$ respectively , we have
837 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
838   \begin{equation}
839   \exp (hX + hY) = \exp (hZ)
840   \end{equation}
# Line 844 | Line 847 | Applying Baker-Campbell-Hausdorff formula to Sprang sp
847   \[
848   [X,Y] = XY - YX .
849   \]
850 < Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
851 < can obtain
850 > Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
851 > Sprang splitting, we can obtain
852   \begin{eqnarray*}
853 < \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
854 < [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
855 < & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
853 < \ldots )
853 > \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2 [X,Y]/4 + h^2 [Y,X]/4 \\
854 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
855 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
856   \end{eqnarray*}
857   Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
858   error of Spring splitting is proportional to $h^3$. The same
# Line 859 | Line 861 | Careful choice of coefficient $a_1 ,\ldot , b_m$ will
861   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
862   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
863   \end{equation}
864 < Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
864 > Careful choice of coefficient $a_1 \ldot b_m$ will lead to higher
865   order method. Yoshida proposed an elegant way to compose higher
866 < order methods based on symmetric splitting. Given a symmetric second
867 < order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
868 < method can be constructed by composing,
866 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
867 > a symmetric second order base method $ \varphi _h^{(2)} $, a
868 > fourth-order symmetric method can be constructed by composing,
869   \[
870   \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
871   h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
# Line 984 | Line 986 | Production run is the most important steps of the simu
986  
987   \subsection{\label{introSection:production}Production}
988  
989 < Production run is the most important steps of the simulation, in
989 > Production run is the most important step of the simulation, in
990   which the equilibrated structure is used as a starting point and the
991   motions of the molecules are collected for later analysis. In order
992   to capture the macroscopic properties of the system, the molecular
# Line 1000 | Line 1002 | approach will suffer from the surface effect. To offse
1002   A natural approach to avoid system size issue is to represent the
1003   bulk behavior by a finite number of the particles. However, this
1004   approach will suffer from the surface effect. To offset this,
1005 < \textit{Periodic boundary condition} is developed to simulate bulk
1006 < properties with a relatively small number of particles. In this
1007 < method, the simulation box is replicated throughout space to form an
1008 < infinite lattice. During the simulation, when a particle moves in
1009 < the primary cell, its image in other cells move in exactly the same
1010 < direction with exactly the same orientation. Thus, as a particle
1011 < leaves the primary cell, one of its images will enter through the
1012 < opposite face.
1013 < %\begin{figure}
1014 < %\centering
1015 < %\includegraphics[width=\linewidth]{pbcFig.eps}
1016 < %\caption[An illustration of periodic boundary conditions]{A 2-D
1017 < %illustration of periodic boundary conditions. As one particle leaves
1018 < %the right of the simulation box, an image of it enters the left.}
1019 < %\label{introFig:pbc}
1020 < %\end{figure}
1005 > \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
1006 > is developed to simulate bulk properties with a relatively small
1007 > number of particles. In this method, the simulation box is
1008 > replicated throughout space to form an infinite lattice. During the
1009 > simulation, when a particle moves in the primary cell, its image in
1010 > other cells move in exactly the same direction with exactly the same
1011 > orientation. Thus, as a particle leaves the primary cell, one of its
1012 > images will enter through the opposite face.
1013 > \begin{figure}
1014 > \centering
1015 > \includegraphics[width=\linewidth]{pbc.eps}
1016 > \caption[An illustration of periodic boundary conditions]{A 2-D
1017 > illustration of periodic boundary conditions. As one particle leaves
1018 > the left of the simulation box, an image of it enters the right.}
1019 > \label{introFig:pbc}
1020 > \end{figure}
1021  
1022   %cutoff and minimum image convention
1023   Another important technique to improve the efficiency of force
1024   evaluation is to apply cutoff where particles farther than a
1025   predetermined distance, are not included in the calculation
1026   \cite{Frenkel1996}. The use of a cutoff radius will cause a
1027 < discontinuity in the potential energy curve
1028 < (Fig.~\ref{introFig:shiftPot}). Fortunately, one can shift the
1029 < potential to ensure the potential curve go smoothly to zero at the
1030 < cutoff radius. Cutoff strategy works pretty well for Lennard-Jones
1031 < interaction because of its short range nature. However, simply
1032 < truncating the electrostatic interaction with the use of cutoff has
1033 < been shown to lead to severe artifacts in simulations. Ewald
1034 < summation, in which the slowly conditionally convergent Coulomb
1035 < potential is transformed into direct and reciprocal sums with rapid
1036 < and absolute convergence, has proved to minimize the periodicity
1037 < artifacts in liquid simulations. Taking the advantages of the fast
1038 < Fourier transform (FFT) for calculating discrete Fourier transforms,
1039 < the particle mesh-based methods are accelerated from $O(N^{3/2})$ to
1040 < $O(N logN)$. An alternative approach is \emph{fast multipole
1041 < method}, which treats Coulombic interaction exactly at short range,
1042 < and approximate the potential at long range through multipolar
1043 < expansion. In spite of their wide acceptances at the molecular
1044 < simulation community, these two methods are hard to be implemented
1045 < correctly and efficiently. Instead, we use a damped and
1046 < charge-neutralized Coulomb potential method developed by Wolf and
1047 < his coworkers. The shifted Coulomb potential for particle $i$ and
1048 < particle $j$ at distance $r_{rj}$ is given by:
1027 > discontinuity in the potential energy curve. Fortunately, one can
1028 > shift the potential to ensure the potential curve go smoothly to
1029 > zero at the cutoff radius. Cutoff strategy works pretty well for
1030 > Lennard-Jones interaction because of its short range nature.
1031 > However, simply truncating the electrostatic interaction with the
1032 > use of cutoff has been shown to lead to severe artifacts in
1033 > simulations. Ewald summation, in which the slowly conditionally
1034 > convergent Coulomb potential is transformed into direct and
1035 > reciprocal sums with rapid and absolute convergence, has proved to
1036 > minimize the periodicity artifacts in liquid simulations. Taking the
1037 > advantages of the fast Fourier transform (FFT) for calculating
1038 > discrete Fourier transforms, the particle mesh-based
1039 > methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1040 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1041 > multipole method}\cite{Greengard1987, Greengard1994}, which treats
1042 > Coulombic interaction exactly at short range, and approximate the
1043 > potential at long range through multipolar expansion. In spite of
1044 > their wide acceptances at the molecular simulation community, these
1045 > two methods are hard to be implemented correctly and efficiently.
1046 > Instead, we use a damped and charge-neutralized Coulomb potential
1047 > method developed by Wolf and his coworkers\cite{Wolf1999}. The
1048 > shifted Coulomb potential for particle $i$ and particle $j$ at
1049 > distance $r_{rj}$ is given by:
1050   \begin{equation}
1051   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1052   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
# Line 1053 | Line 1056 | efficient and easy to implement.
1056   where $\alpha$ is the convergence parameter. Due to the lack of
1057   inherent periodicity and rapid convergence,this method is extremely
1058   efficient and easy to implement.
1059 < %\begin{figure}
1060 < %\centering
1061 < %\includegraphics[width=\linewidth]{pbcFig.eps}
1062 < %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
1063 < %\label{introFigure:shiftedCoulomb}
1064 < %\end{figure}
1059 > \begin{figure}
1060 > \centering
1061 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1062 > \caption[An illustration of shifted Coulomb potential]{An
1063 > illustration of shifted Coulomb potential.}
1064 > \label{introFigure:shiftedCoulomb}
1065 > \end{figure}
1066  
1067   %multiple time step
1068  
# Line 1104 | Line 1108 | integrating over the surface factor \cite{Powles73}. T
1108   function}, is of most fundamental importance to liquid-state theory.
1109   Pair distribution function can be gathered by Fourier transforming
1110   raw data from a series of neutron diffraction experiments and
1111 < integrating over the surface factor \cite{Powles73}. The experiment
1112 < result can serve as a criterion to justify the correctness of the
1113 < theory. Moreover, various equilibrium thermodynamic and structural
1114 < properties can also be expressed in terms of radial distribution
1115 < function \cite{allen87:csl}.
1111 > integrating over the surface factor \cite{Powles1973}. The
1112 > experiment result can serve as a criterion to justify the
1113 > correctness of the theory. Moreover, various equilibrium
1114 > thermodynamic and structural properties can also be expressed in
1115 > terms of radial distribution function \cite{Allen1987}.
1116  
1117   A pair distribution functions $g(r)$ gives the probability that a
1118   particle $i$ will be located at a distance $r$ from a another
# Line 1186 | Line 1190 | protein-protein docking study{\cite{Gray03}}.
1190   movement of the objects in 3D gaming engine or other physics
1191   simulator is governed by the rigid body dynamics. In molecular
1192   simulation, rigid body is used to simplify the model in
1193 < protein-protein docking study{\cite{Gray03}}.
1193 > protein-protein docking study\cite{Gray2003}.
1194  
1195   It is very important to develop stable and efficient methods to
1196   integrate the equations of motion of orientational degrees of
# Line 1194 | Line 1198 | different sets of Euler angles can overcome this diffi
1198   rotational degrees of freedom. However, due to its singularity, the
1199   numerical integration of corresponding equations of motion is very
1200   inefficient and inaccurate. Although an alternative integrator using
1201 < different sets of Euler angles can overcome this difficulty\cite{},
1202 < the computational penalty and the lost of angular momentum
1203 < conservation still remain. A singularity free representation
1204 < utilizing quaternions was developed by Evans in 1977. Unfortunately,
1205 < this approach suffer from the nonseparable Hamiltonian resulted from
1206 < quaternion representation, which prevents the symplectic algorithm
1207 < to be utilized. Another different approach is to apply holonomic
1208 < constraints to the atoms belonging to the rigid body. Each atom
1209 < moves independently under the normal forces deriving from potential
1210 < energy and constraint forces which are used to guarantee the
1211 < rigidness. However, due to their iterative nature, SHAKE and Rattle
1212 < algorithm converge very slowly when the number of constraint
1213 < increases.
1201 > different sets of Euler angles can overcome this
1202 > difficulty\cite{Barojas1973}, the computational penalty and the lost
1203 > of angular momentum conservation still remain. A singularity free
1204 > representation utilizing quaternions was developed by Evans in
1205 > 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1206 > nonseparable Hamiltonian resulted from quaternion representation,
1207 > which prevents the symplectic algorithm to be utilized. Another
1208 > different approach is to apply holonomic constraints to the atoms
1209 > belonging to the rigid body. Each atom moves independently under the
1210 > normal forces deriving from potential energy and constraint forces
1211 > which are used to guarantee the rigidness. However, due to their
1212 > iterative nature, SHAKE and Rattle algorithm converge very slowly
1213 > when the number of constraint increases\cite{Ryckaert1977,
1214 > Andersen1983}.
1215  
1216   The break through in geometric literature suggests that, in order to
1217   develop a long-term integration scheme, one should preserve the
1218   symplectic structure of the flow. Introducing conjugate momentum to
1219   rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1220 < symplectic integrator, RSHAKE, was proposed to evolve the
1221 < Hamiltonian system in a constraint manifold by iteratively
1220 > symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1221 > the Hamiltonian system in a constraint manifold by iteratively
1222   satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1223 < method using quaternion representation was developed by Omelyan.
1224 < However, both of these methods are iterative and inefficient. In
1225 < this section, we will present a symplectic Lie-Poisson integrator
1226 < for rigid body developed by Dullweber and his
1227 < coworkers\cite{Dullweber1997} in depth.
1223 > method using quaternion representation was developed by
1224 > Omelyan\cite{Omelyan1998}. However, both of these methods are
1225 > iterative and inefficient. In this section, we will present a
1226 > symplectic Lie-Poisson integrator for rigid body developed by
1227 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1228  
1229   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1230   The motion of the rigid body is Hamiltonian with the Hamiltonian
# Line 1238 | Line 1243 | Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1243   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1244   constrained Hamiltonian equation subjects to a holonomic constraint,
1245   \begin{equation}
1246 < Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1246 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1247   \end{equation}
1248   which is used to ensure rotation matrix's orthogonality.
1249   Differentiating \ref{introEquation:orthogonalConstraint} and using
# Line 1263 | Line 1268 | simply evolve the system in constraint manifold. The t
1268   In general, there are two ways to satisfy the holonomic constraints.
1269   We can use constraint force provided by lagrange multiplier on the
1270   normal manifold to keep the motion on constraint space. Or we can
1271 < simply evolve the system in constraint manifold. The two method are
1272 < proved to be equivalent. The holonomic constraint and equations of
1273 < motions define a constraint manifold for rigid body
1271 > simply evolve the system in constraint manifold. These two methods
1272 > are proved to be equivalent. The holonomic constraint and equations
1273 > of motions define a constraint manifold for rigid body
1274   \[
1275   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1276   \right\}.
# Line 1352 | Line 1357 | not be avoided in other methods\cite{}.
1357   \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1358   multiplier $\Lambda$ is absent from the equations of motion. This
1359   unique property eliminate the requirement of iterations which can
1360 < not be avoided in other methods\cite{}.
1360 > not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1361  
1362   Applying hat-map isomorphism, we obtain the equation of motion for
1363   angular momentum on body frame
# Line 1478 | Line 1483 | kinetic energy are listed in the below table,
1483   \]
1484   The equations of motion corresponding to potential energy and
1485   kinetic energy are listed in the below table,
1486 + \begin{table}
1487 + \caption{Equations of motion due to Potential and Kinetic Energies}
1488   \begin{center}
1489   \begin{tabular}{|l|l|}
1490    \hline
# Line 1490 | Line 1497 | A second-order symplectic method is now obtained by th
1497    \hline
1498   \end{tabular}
1499   \end{center}
1500 < A second-order symplectic method is now obtained by the composition
1501 < of the flow maps,
1500 > \end{table}
1501 > A second-order symplectic method is now obtained by the
1502 > composition of the flow maps,
1503   \[
1504   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1505   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
# Line 1616 | Line 1624 | Operator. Below are some important properties of Lapla
1624   \]
1625   where  $p$ is real and  $L$ is called the Laplace Transform
1626   Operator. Below are some important properties of Laplace transform
1619 \begin{equation}
1620 \begin{array}{c}
1621 L(x + y) = L(x) + L(y) \\
1622 L(ax) = aL(x) \\
1623 L(\dot x) = pL(x) - px(0) \\
1624 L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1625 L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1626 \end{array}
1627 \end{equation}
1627  
1628 + \begin{eqnarray*}
1629 + L(x + y)  & = & L(x) + L(y) \\
1630 + L(ax)     & = & aL(x) \\
1631 + L(\dot x) & = & pL(x) - px(0) \\
1632 + L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1633 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1634 + \end{eqnarray*}
1635 +
1636 +
1637   Applying Laplace transform to the bath coordinates, we obtain
1638 < \[
1639 < \begin{array}{c}
1640 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1641 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1642 < \end{array}
1635 < \]
1638 > \begin{eqnarray*}
1639 > p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) & = & - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1640 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1641 > \end{eqnarray*}
1642 >
1643   By the same way, the system coordinates become
1644 < \[
1645 < \begin{array}{c}
1646 < mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1647 <  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1641 < \end{array}
1642 < \]
1644 > \begin{eqnarray*}
1645 > mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1646 >  & & \mbox{} - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1647 > \end{eqnarray*}
1648  
1649   With the help of some relatively important inverse Laplace
1650   transformations:
# Line 1651 | Line 1656 | transformations:
1656   \end{array}
1657   \]
1658   , we obtain
1659 < \begin{align}
1660 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1659 > \[
1660 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1661   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1662   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1663   _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1664   - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1665   (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1666   _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1667 < %
1668 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1667 > \]
1668 > \[
1669 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1670   {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1671   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1672   t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
# Line 1668 | Line 1674 | t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  =
1674   \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1675   \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1676   (\omega _\alpha  t)} \right\}}
1677 < \end{align}
1677 > \]
1678  
1679   Introducing a \emph{dynamic friction kernel}
1680   \begin{equation}
# Line 1763 | Line 1769 | And since the $q$ coordinates are harmonic oscillators
1769   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1770   \]
1771   And since the $q$ coordinates are harmonic oscillators,
1772 < \[
1773 < \begin{array}{c}
1774 < \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1775 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1776 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1777 < \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1778 <  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1779 <  = kT\xi (t) \\
1780 < \end{array}
1781 < \]
1772 >
1773 > \begin{eqnarray*}
1774 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1775 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1776 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1777 > \left\langle {R(t)R(0)} \right\rangle & = & \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1778 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1779 >  & = &kT\xi (t) \\
1780 > \end{eqnarray*}
1781 >
1782   Thus, we recover the \emph{second fluctuation dissipation theorem}
1783   \begin{equation}
1784   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
# Line 1787 | Line 1793 | Equation, \zeta can be taken as a scalar. In general,
1793   when the system become more and more complicate. Instead, various
1794   approaches based on hydrodynamics have been developed to calculate
1795   the friction coefficients. The friction effect is isotropic in
1796 < Equation, \zeta can be taken as a scalar. In general, friction
1797 < tensor \Xi is a $6\times 6$ matrix given by
1796 > Equation, $\zeta$ can be taken as a scalar. In general, friction
1797 > tensor $\Xi$ is a $6\times 6$ matrix given by
1798   \[
1799   \Xi  = \left( {\begin{array}{*{20}c}
1800     {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
# Line 1844 | Line 1850 | coordinates by
1850   hydrodynamics theory, because their properties can be calculated
1851   exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1852   also called a triaxial ellipsoid, which is given in Cartesian
1853 < coordinates by
1853 > coordinates by\cite{Perrin1934, Perrin1936}
1854   \[
1855   \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1856   }} = 1
# Line 1888 | Line 1894 | unique\cite{Wegener79} as well as the intrinsic coupli
1894   hydrodynamic properties of rigid bodies. However, since the mapping
1895   from all possible ellipsoidal space, $r$-space, to all possible
1896   combination of rotational diffusion coefficients, $D$-space is not
1897 < unique\cite{Wegener79} as well as the intrinsic coupling between
1898 < translational and rotational motion of rigid body\cite{}, general
1899 < ellipsoid is not always suitable for modeling arbitrarily shaped
1900 < rigid molecule. A number of studies have been devoted to determine
1901 < the friction tensor for irregularly shaped rigid bodies using more
1902 < advanced method\cite{} where the molecule of interest was modeled by
1903 < combinations of spheres(beads)\cite{} and the hydrodynamics
1904 < properties of the molecule can be calculated using the hydrodynamic
1905 < interaction tensor. Let us consider a rigid assembly of $N$ beads
1906 < immersed in a continuous medium. Due to hydrodynamics interaction,
1907 < the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1908 < unperturbed velocity $v_i$,
1897 > unique\cite{Wegener1979} as well as the intrinsic coupling between
1898 > translational and rotational motion of rigid body, general ellipsoid
1899 > is not always suitable for modeling arbitrarily shaped rigid
1900 > molecule. A number of studies have been devoted to determine the
1901 > friction tensor for irregularly shaped rigid bodies using more
1902 > advanced method where the molecule of interest was modeled by
1903 > combinations of spheres(beads)\cite{Carrasco1999} and the
1904 > hydrodynamics properties of the molecule can be calculated using the
1905 > hydrodynamic interaction tensor. Let us consider a rigid assembly of
1906 > $N$ beads immersed in a continuous medium. Due to hydrodynamics
1907 > interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1908 > than its unperturbed velocity $v_i$,
1909   \[
1910   v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1911   \]
# Line 1920 | Line 1926 | introduced by Rotne and Prager\cite{} and improved by
1926   \end{equation}
1927   Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1928   A second order expression for element of different size was
1929 < introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1930 < la Torre and Bloomfield,
1929 > introduced by Rotne and Prager\cite{Rotne1969} and improved by
1930 > Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1931   \begin{equation}
1932   T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1933   \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
# Line 2022 | Line 2028 | the position of center of resistance,
2028   Using Equations \ref{introEquation:definitionCR} and
2029   \ref{introEquation:resistanceTensorTransformation}, one can locate
2030   the position of center of resistance,
2031 < \[
2032 < \left( \begin{array}{l}
2031 > \begin{eqnarray*}
2032 > \left( \begin{array}{l}
2033   x_{OR}  \\
2034   y_{OR}  \\
2035   z_{OR}  \\
2036 < \end{array} \right) = \left( {\begin{array}{*{20}c}
2036 > \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2037     {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2038     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2039     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2040 < \end{array}} \right)^{ - 1} \left( \begin{array}{l}
2040 > \end{array}} \right)^{ - 1}  \\
2041 >  & & \left( \begin{array}{l}
2042   (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2043   (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2044   (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2045 < \end{array} \right).
2046 < \]
2045 > \end{array} \right) \\
2046 > \end{eqnarray*}
2047 >
2048 >
2049 >
2050   where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2051   joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines