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 2786 by tim, Sun Jun 4 20:18:07 2006 UTC vs.
Revision 2789 by tim, Mon Jun 5 21:00:46 2006 UTC

# Line 498 | Line 498 | issue\cite{}. The velocity verlet method, which happen
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\cite{}. The velocity verlet method, which happens to be a
502 < simple example of symplectic integrator, continues to gain its
503 < popularity in molecular dynamics community. This fact can be partly
504 < explained 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{Tuckerman1992}.
714 > the system\cite{Tuckerman1992, McLachlan1998}.
715  
716   \subsubsection{\label{introSection:splittingMethod}Splitting Method}
717  
# 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 [X,Y]/4 + h^2 [Y,X]/4 \\
854                                     &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
# Line 860 | Line 863 | order methods based on symmetric splitting. Given a sy
863   \end{equation}
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 983 | 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 999 | 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
# Line 1032 | Line 1035 | discrete Fourier transforms, the particle mesh-based m
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 methods are
1039 < accelerated from $O(N^{3/2})$ to $O(N logN)$. An alternative
1040 < approach is \emph{fast multipole method}, which treats Coulombic
1041 < interaction exactly at short range, and approximate the potential at
1042 < long range through multipolar expansion. In spite of their wide
1043 < acceptances at the molecular simulation community, these two methods
1044 < are hard to be implemented correctly and efficiently. Instead, we
1045 < use a damped and charge-neutralized Coulomb potential method
1046 < developed by Wolf and his coworkers. The shifted Coulomb potential
1047 < for particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
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 1051 | 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 1184 | Line 1190 | protein-protein docking study{\cite{Gray2003}}.
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{Gray2003}}.
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 1192 | 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 1350 | 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 1617 | 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
1627 < \begin{equation}
1628 < \begin{array}{c}
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{array}
1628 < \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 < \[
1639 < \begin{array}{c}
1640 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1641 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1642 < \end{array}
1636 < \]
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 < \[
1645 < \begin{array}{c}
1646 < mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1647 <  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1642 < \end{array}
1643 < \]
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 1652 | Line 1656 | transformations:
1656   \end{array}
1657   \]
1658   , we obtain
1659 < \begin{align}
1660 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1659 > \begin{eqnarray*}
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)
# Line 1661 | Line 1665 | _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
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
1668 > & = & \mbox{} - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1669   {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1670   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1671   t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
# Line 1669 | Line 1673 | t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  =
1673   \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1674   \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1675   (\omega _\alpha  t)} \right\}}
1676 < \end{align}
1676 > \end{eqnarray*}
1677  
1678   Introducing a \emph{dynamic friction kernel}
1679   \begin{equation}
# Line 1764 | Line 1768 | And since the $q$ coordinates are harmonic oscillators
1768   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1769   \]
1770   And since the $q$ coordinates are harmonic oscillators,
1771 < \[
1772 < \begin{array}{c}
1773 < \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1774 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1775 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1776 < \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1777 <  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1778 <  = kT\xi (t) \\
1779 < \end{array}
1780 < \]
1771 >
1772 > \begin{eqnarray*}
1773 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1774 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1775 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1776 > \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 } }  \\
1777 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1778 >  & = &kT\xi (t) \\
1779 > \end{eqnarray*}
1780 >
1781   Thus, we recover the \emph{second fluctuation dissipation theorem}
1782   \begin{equation}
1783   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
# Line 1845 | Line 1849 | coordinates by
1849   hydrodynamics theory, because their properties can be calculated
1850   exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1851   also called a triaxial ellipsoid, which is given in Cartesian
1852 < coordinates by
1852 > coordinates by\cite{Perrin1934, Perrin1936}
1853   \[
1854   \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1855   }} = 1
# Line 1890 | Line 1894 | translational and rotational motion of rigid body\cite
1894   from all possible ellipsoidal space, $r$-space, to all possible
1895   combination of rotational diffusion coefficients, $D$-space is not
1896   unique\cite{Wegener1979} as well as the intrinsic coupling between
1897 < translational and rotational motion of rigid body\cite{}, general
1898 < ellipsoid is not always suitable for modeling arbitrarily shaped
1899 < rigid molecule. A number of studies have been devoted to determine
1900 < the friction tensor for irregularly shaped rigid bodies using more
1901 < advanced method\cite{} where the molecule of interest was modeled by
1902 < combinations of spheres(beads)\cite{} and the hydrodynamics
1903 < properties of the molecule can be calculated using the hydrodynamic
1904 < interaction tensor. Let us consider a rigid assembly of $N$ beads
1905 < immersed in a continuous medium. Due to hydrodynamics interaction,
1906 < the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1907 < unperturbed velocity $v_i$,
1897 > translational and rotational motion of rigid body, general ellipsoid
1898 > is not always suitable for modeling arbitrarily shaped rigid
1899 > molecule. A number of studies have been devoted to determine the
1900 > friction tensor for irregularly shaped rigid bodies using more
1901 > advanced method where the molecule of interest was modeled by
1902 > combinations of spheres(beads)\cite{Carrasco1999} and the
1903 > hydrodynamics properties of the molecule can be calculated using the
1904 > hydrodynamic interaction tensor. Let us consider a rigid assembly of
1905 > $N$ beads immersed in a continuous medium. Due to hydrodynamics
1906 > interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1907 > than its unperturbed velocity $v_i$,
1908   \[
1909   v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1910   \]
# Line 1921 | Line 1925 | introduced by Rotne and Prager\cite{} and improved by
1925   \end{equation}
1926   Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1927   A second order expression for element of different size was
1928 < introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1929 < la Torre and Bloomfield,
1928 > introduced by Rotne and Prager\cite{Rotne1969} and improved by
1929 > Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1930   \begin{equation}
1931   T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1932   \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
# Line 2038 | Line 2042 | where $x_OR$, $y_OR$, $z_OR$ are the components of the
2042   (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2043   \end{array} \right).
2044   \]
2045 +
2046 +
2047 + \begin{eqnarray*}
2048 + \left( \begin{array}{l}
2049 + x_{OR}  \\
2050 + y_{OR}  \\
2051 + z_{OR}  \\
2052 + \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2053 +   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2054 +   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2055 +   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2056 + \end{array}} \right)^{ - 1}  \\
2057 +  & & \left( \begin{array}{l}
2058 + (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2059 + (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2060 + (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2061 + \end{array} \right) \\
2062 + \end{eqnarray*}
2063 +
2064 +
2065 +
2066   where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2067   joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines