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 2801 by tim, Tue Jun 6 14:56:36 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 858 | Line 861 | Careful choice of coefficient $a_1 \ldot b_m$ will lea
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 $a_1 \ldots b_m$ 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 911 | Line 914 | initialization of a simulation. Sec.~\ref{introSec:pro
914   \end{enumerate}
915   These three individual steps will be covered in the following
916   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
917 < initialization of a simulation. Sec.~\ref{introSec:production} will
918 < discusses issues in production run. Sec.~\ref{introSection:Analysis}
919 < provides the theoretical tools for trajectory analysis.
917 > initialization of a simulation. Sec.~\ref{introSection:production}
918 > will discusses issues in production run.
919 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
920 > trajectory analysis.
921  
922   \subsection{\label{introSec:initialSystemSettings}Initialization}
923  
# Line 983 | Line 987 | Production run is the most important steps of the simu
987  
988   \subsection{\label{introSection:production}Production}
989  
990 < Production run is the most important steps of the simulation, in
990 > Production run is the most important step of the simulation, in
991   which the equilibrated structure is used as a starting point and the
992   motions of the molecules are collected for later analysis. In order
993   to capture the macroscopic properties of the system, the molecular
# Line 999 | Line 1003 | approach will suffer from the surface effect. To offse
1003   A natural approach to avoid system size issue is to represent the
1004   bulk behavior by a finite number of the particles. However, this
1005   approach will suffer from the surface effect. To offset this,
1006 < \textit{Periodic boundary condition} is developed to simulate bulk
1007 < properties with a relatively small number of particles. In this
1008 < method, the simulation box is replicated throughout space to form an
1009 < infinite lattice. During the simulation, when a particle moves in
1010 < the primary cell, its image in other cells move in exactly the same
1011 < direction with exactly the same orientation. Thus, as a particle
1012 < leaves the primary cell, one of its images will enter through the
1013 < opposite face.
1014 < %\begin{figure}
1015 < %\centering
1016 < %\includegraphics[width=\linewidth]{pbcFig.eps}
1017 < %\caption[An illustration of periodic boundary conditions]{A 2-D
1018 < %illustration of periodic boundary conditions. As one particle leaves
1019 < %the right of the simulation box, an image of it enters the left.}
1020 < %\label{introFig:pbc}
1021 < %\end{figure}
1006 > \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
1007 > is developed to simulate bulk properties with a relatively small
1008 > number of particles. In this method, the simulation box is
1009 > replicated throughout space to form an infinite lattice. During the
1010 > simulation, when a particle moves in the primary cell, its image in
1011 > other cells move in exactly the same direction with exactly the same
1012 > orientation. Thus, as a particle leaves the primary cell, one of its
1013 > images will enter through the opposite face.
1014 > \begin{figure}
1015 > \centering
1016 > \includegraphics[width=\linewidth]{pbc.eps}
1017 > \caption[An illustration of periodic boundary conditions]{A 2-D
1018 > illustration of periodic boundary conditions. As one particle leaves
1019 > the left of the simulation box, an image of it enters the right.}
1020 > \label{introFig:pbc}
1021 > \end{figure}
1022  
1023   %cutoff and minimum image convention
1024   Another important technique to improve the efficiency of force
# Line 1032 | Line 1036 | discrete Fourier transforms, the particle mesh-based m
1036   reciprocal sums with rapid and absolute convergence, has proved to
1037   minimize the periodicity artifacts in liquid simulations. Taking the
1038   advantages of the fast Fourier transform (FFT) for calculating
1039 < discrete Fourier transforms, the particle mesh-based methods are
1040 < accelerated from $O(N^{3/2})$ to $O(N logN)$. An alternative
1041 < approach is \emph{fast multipole method}, which treats Coulombic
1042 < interaction exactly at short range, and approximate the potential at
1043 < long range through multipolar expansion. In spite of their wide
1044 < acceptances at the molecular simulation community, these two methods
1045 < are hard to be implemented correctly and efficiently. Instead, we
1046 < use a damped and charge-neutralized Coulomb potential method
1047 < developed by Wolf and his coworkers. The shifted Coulomb potential
1048 < for particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1039 > discrete Fourier transforms, the particle mesh-based
1040 > methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1041 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1042 > multipole method}\cite{Greengard1987, Greengard1994}, which treats
1043 > Coulombic interaction exactly at short range, and approximate the
1044 > potential at long range through multipolar expansion. In spite of
1045 > their wide acceptances at the molecular simulation community, these
1046 > two methods are hard to be implemented correctly and efficiently.
1047 > Instead, we use a damped and charge-neutralized Coulomb potential
1048 > method developed by Wolf and his coworkers\cite{Wolf1999}. The
1049 > shifted Coulomb potential for particle $i$ and particle $j$ at
1050 > distance $r_{rj}$ is given by:
1051   \begin{equation}
1052   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1053   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
# Line 1051 | Line 1057 | efficient and easy to implement.
1057   where $\alpha$ is the convergence parameter. Due to the lack of
1058   inherent periodicity and rapid convergence,this method is extremely
1059   efficient and easy to implement.
1060 < %\begin{figure}
1061 < %\centering
1062 < %\includegraphics[width=\linewidth]{pbcFig.eps}
1063 < %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
1064 < %\label{introFigure:shiftedCoulomb}
1065 < %\end{figure}
1060 > \begin{figure}
1061 > \centering
1062 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1063 > \caption[An illustration of shifted Coulomb potential]{An
1064 > illustration of shifted Coulomb potential.}
1065 > \label{introFigure:shiftedCoulomb}
1066 > \end{figure}
1067  
1068   %multiple time step
1069  
# Line 1184 | Line 1191 | protein-protein docking study{\cite{Gray2003}}.
1191   movement of the objects in 3D gaming engine or other physics
1192   simulator is governed by the rigid body dynamics. In molecular
1193   simulation, rigid body is used to simplify the model in
1194 < protein-protein docking study{\cite{Gray2003}}.
1194 > protein-protein docking study\cite{Gray2003}.
1195  
1196   It is very important to develop stable and efficient methods to
1197   integrate the equations of motion of orientational degrees of
# Line 1192 | Line 1199 | different sets of Euler angles can overcome this diffi
1199   rotational degrees of freedom. However, due to its singularity, the
1200   numerical integration of corresponding equations of motion is very
1201   inefficient and inaccurate. Although an alternative integrator using
1202 < different sets of Euler angles can overcome this difficulty\cite{},
1203 < the computational penalty and the lost of angular momentum
1204 < conservation still remain. A singularity free representation
1205 < utilizing quaternions was developed by Evans in 1977. Unfortunately,
1206 < this approach suffer from the nonseparable Hamiltonian resulted from
1207 < quaternion representation, which prevents the symplectic algorithm
1208 < to be utilized. Another different approach is to apply holonomic
1209 < constraints to the atoms belonging to the rigid body. Each atom
1210 < moves independently under the normal forces deriving from potential
1211 < energy and constraint forces which are used to guarantee the
1212 < rigidness. However, due to their iterative nature, SHAKE and Rattle
1213 < algorithm converge very slowly when the number of constraint
1214 < increases.
1202 > different sets of Euler angles can overcome this
1203 > difficulty\cite{Barojas1973}, the computational penalty and the lost
1204 > of angular momentum conservation still remain. A singularity free
1205 > representation utilizing quaternions was developed by Evans in
1206 > 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1207 > nonseparable Hamiltonian resulted from quaternion representation,
1208 > which prevents the symplectic algorithm to be utilized. Another
1209 > different approach is to apply holonomic constraints to the atoms
1210 > belonging to the rigid body. Each atom moves independently under the
1211 > normal forces deriving from potential energy and constraint forces
1212 > which are used to guarantee the rigidness. However, due to their
1213 > iterative nature, SHAKE and Rattle algorithm converge very slowly
1214 > when the number of constraint increases\cite{Ryckaert1977,
1215 > Andersen1983}.
1216  
1217   The break through in geometric literature suggests that, in order to
1218   develop a long-term integration scheme, one should preserve the
1219   symplectic structure of the flow. Introducing conjugate momentum to
1220   rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1221 < symplectic integrator, RSHAKE, was proposed to evolve the
1222 < Hamiltonian system in a constraint manifold by iteratively
1221 > symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1222 > the Hamiltonian system in a constraint manifold by iteratively
1223   satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1224 < method using quaternion representation was developed by Omelyan.
1225 < However, both of these methods are iterative and inefficient. In
1226 < this section, we will present a symplectic Lie-Poisson integrator
1227 < for rigid body developed by Dullweber and his
1228 < coworkers\cite{Dullweber1997} in depth.
1224 > method using quaternion representation was developed by
1225 > Omelyan\cite{Omelyan1998}. However, both of these methods are
1226 > iterative and inefficient. In this section, we will present a
1227 > symplectic Lie-Poisson integrator for rigid body developed by
1228 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1229  
1230   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1231   The motion of the rigid body is Hamiltonian with the Hamiltonian
# Line 1249 | Line 1257 | the equations of motion,
1257   Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1258   \ref{introEquation:motionHamiltonianMomentum}), one can write down
1259   the equations of motion,
1252 \[
1253 \begin{array}{c}
1254 \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1255 \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1256 \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1257 \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1258 \end{array}
1259 \]
1260  
1261 + \begin{eqnarray}
1262 + \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1263 + \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1264 + \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1265 + \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1266 + \end{eqnarray}
1267 +
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
# Line 1337 | Line 1344 | operations
1344   \[
1345   \hat vu = v \times u
1346   \]
1340
1347   Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1348   matrix,
1349   \begin{equation}
1350 < (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1350 > (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ {\bullet  ^T}
1351   ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1352   - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1353   (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
# Line 1350 | Line 1356 | not be avoided in other methods\cite{}.
1356   \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1357   multiplier $\Lambda$ is absent from the equations of motion. This
1358   unique property eliminate the requirement of iterations which can
1359 < not be avoided in other methods\cite{}.
1359 > not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1360  
1361   Applying hat-map isomorphism, we obtain the equation of motion for
1362   angular momentum on body frame
# Line 1370 | Line 1376 | first term of \ref{ introEquation:bodyAngularMotion}).
1376  
1377   If there is not external forces exerted on the rigid body, the only
1378   contribution to the rotational is from the kinetic potential (the
1379 < first term of \ref{ introEquation:bodyAngularMotion}). The free
1380 < rigid body is an example of Lie-Poisson system with Hamiltonian
1375 < function
1379 > first term of \ref{introEquation:bodyAngularMotion}). The free rigid
1380 > body is an example of Lie-Poisson system with Hamiltonian function
1381   \begin{equation}
1382   T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1383   \label{introEquation:rotationalKineticRB}
# Line 1617 | Line 1622 | Operator. Below are some important properties of Lapla
1622   \]
1623   where  $p$ is real and  $L$ is called the Laplace Transform
1624   Operator. Below are some important properties of Laplace transform
1625 < \begin{equation}
1626 < \begin{array}{c}
1627 < L(x + y) = L(x) + L(y) \\
1628 < L(ax) = aL(x) \\
1629 < L(\dot x) = pL(x) - px(0) \\
1630 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1631 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1632 < \end{array}
1628 < \end{equation}
1625 >
1626 > \begin{eqnarray*}
1627 > L(x + y)  & = & L(x) + L(y) \\
1628 > L(ax)     & = & aL(x) \\
1629 > L(\dot x) & = & pL(x) - px(0) \\
1630 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1631 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1632 > \end{eqnarray*}
1633  
1634 +
1635   Applying Laplace transform to the bath coordinates, we obtain
1636 < \[
1637 < \begin{array}{c}
1638 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1639 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1640 < \end{array}
1636 < \]
1636 > \begin{eqnarray*}
1637 > 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) \\
1638 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1639 > \end{eqnarray*}
1640 >
1641   By the same way, the system coordinates become
1642 < \[
1643 < \begin{array}{c}
1644 < mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1645 <  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1642 < \end{array}
1643 < \]
1642 > \begin{eqnarray*}
1643 > mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1644 >  & & \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\}}  \\
1645 > \end{eqnarray*}
1646  
1647   With the help of some relatively important inverse Laplace
1648   transformations:
# Line 1652 | Line 1654 | transformations:
1654   \end{array}
1655   \]
1656   , we obtain
1657 < \begin{align}
1658 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1657 > \begin{eqnarray*}
1658 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1659   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1660   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1661 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1662 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1663 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1664 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1665 < %
1666 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1661 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1662 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1663 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1664 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1665 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1666 > \end{eqnarray*}
1667 > \begin{eqnarray*}
1668 > m\ddot x & = & - \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\{
1672 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\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}
1673 <
1671 > t)\dot x(t - \tau )d} \tau }  \\
1672 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1673 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1674 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1675 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1676 > \end{eqnarray*}
1677   Introducing a \emph{dynamic friction kernel}
1678   \begin{equation}
1679   \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
# Line 1764 | Line 1767 | And since the $q$ coordinates are harmonic oscillators
1767   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1768   \]
1769   And since the $q$ coordinates are harmonic oscillators,
1770 < \[
1771 < \begin{array}{c}
1772 < \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1773 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1774 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1775 < \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1776 <  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1777 <  = kT\xi (t) \\
1778 < \end{array}
1779 < \]
1770 >
1771 > \begin{eqnarray*}
1772 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1773 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1774 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1775 > \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 } }  \\
1776 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1777 >  & = &kT\xi (t) \\
1778 > \end{eqnarray*}
1779 >
1780   Thus, we recover the \emph{second fluctuation dissipation theorem}
1781   \begin{equation}
1782   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
# Line 1845 | Line 1848 | coordinates by
1848   hydrodynamics theory, because their properties can be calculated
1849   exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1850   also called a triaxial ellipsoid, which is given in Cartesian
1851 < coordinates by
1851 > coordinates by\cite{Perrin1934, Perrin1936}
1852   \[
1853   \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1854   }} = 1
# Line 1890 | Line 1893 | translational and rotational motion of rigid body\cite
1893   from all possible ellipsoidal space, $r$-space, to all possible
1894   combination of rotational diffusion coefficients, $D$-space is not
1895   unique\cite{Wegener1979} as well as the intrinsic coupling between
1896 < translational and rotational motion of rigid body\cite{}, general
1897 < ellipsoid is not always suitable for modeling arbitrarily shaped
1898 < rigid molecule. A number of studies have been devoted to determine
1899 < the friction tensor for irregularly shaped rigid bodies using more
1900 < advanced method\cite{} where the molecule of interest was modeled by
1901 < combinations of spheres(beads)\cite{} and the hydrodynamics
1902 < properties of the molecule can be calculated using the hydrodynamic
1903 < interaction tensor. Let us consider a rigid assembly of $N$ beads
1904 < immersed in a continuous medium. Due to hydrodynamics interaction,
1905 < the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1906 < unperturbed velocity $v_i$,
1896 > translational and rotational motion of rigid body, general ellipsoid
1897 > is not always suitable for modeling arbitrarily shaped rigid
1898 > molecule. A number of studies have been devoted to determine the
1899 > friction tensor for irregularly shaped rigid bodies using more
1900 > advanced method where the molecule of interest was modeled by
1901 > combinations of spheres(beads)\cite{Carrasco1999} and the
1902 > hydrodynamics properties of the molecule can be calculated using the
1903 > hydrodynamic interaction tensor. Let us consider a rigid assembly of
1904 > $N$ beads immersed in a continuous medium. Due to hydrodynamics
1905 > interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1906 > than its unperturbed velocity $v_i$,
1907   \[
1908   v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1909   \]
# Line 1921 | Line 1924 | introduced by Rotne and Prager\cite{} and improved by
1924   \end{equation}
1925   Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1926   A second order expression for element of different size was
1927 < introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1928 < la Torre and Bloomfield,
1927 > introduced by Rotne and Prager\cite{Rotne1969} and improved by
1928 > Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1929   \begin{equation}
1930   T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1931   \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
# Line 2008 | Line 2011 | obtain the resistance tensor at $P$ by
2011   \begin{array}{l}
2012   \Xi _P^{tt}  = \Xi _O^{tt}  \\
2013   \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2014 < \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2014 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{{tr} ^{^T }}  \\
2015   \end{array}
2016   \label{introEquation:resistanceTensorTransformation}
2017   \end{equation}
# Line 2023 | Line 2026 | the position of center of resistance,
2026   Using Equations \ref{introEquation:definitionCR} and
2027   \ref{introEquation:resistanceTensorTransformation}, one can locate
2028   the position of center of resistance,
2029 < \[
2030 < \left( \begin{array}{l}
2029 > \begin{eqnarray*}
2030 > \left( \begin{array}{l}
2031   x_{OR}  \\
2032   y_{OR}  \\
2033   z_{OR}  \\
2034 < \end{array} \right) = \left( {\begin{array}{*{20}c}
2034 > \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2035     {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2036     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2037     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2038 < \end{array}} \right)^{ - 1} \left( \begin{array}{l}
2038 > \end{array}} \right)^{ - 1}  \\
2039 >  & & \left( \begin{array}{l}
2040   (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2041   (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2042   (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2043 < \end{array} \right).
2044 < \]
2043 > \end{array} \right) \\
2044 > \end{eqnarray*}
2045 >
2046 >
2047 >
2048   where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2049   joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines