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 2844 by tim, Fri Jun 9 19:06:37 2006 UTC vs.
Revision 2895 by tim, Tue Jun 27 02:42:30 2006 UTC

# Line 73 | Line 73 | can only be described in cartesian coordinate systems.
73   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
75   Newtonian Mechanics suffers from two important limitations: motions
76 < can only be described in cartesian coordinate systems. Moreover, It
77 < become impossible to predict analytically the properties of the
76 > can only be described in cartesian coordinate systems. Moreover, it
77 > becomes impossible to predict analytically the properties of the
78   system even if we know all of the details of the interaction. In
79   order to overcome some of the practical difficulties which arise in
80   attempts to apply Newton's equation to complex system, approximate
# Line 113 | Line 113 | For a holonomic system of $f$ degrees of freedom, the
113   \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
114   Equations of Motion in Lagrangian Mechanics}}
115  
116 < For a holonomic system of $f$ degrees of freedom, the equations of
117 < motion in the Lagrangian form is
116 > For a system of $f$ degrees of freedom, the equations of motion in
117 > the Lagrangian form is
118   \begin{equation}
119   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
120   \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
# Line 227 | Line 227 | this system is a $6f$ dimensional space. A point, $x =
227   momentum variables. Consider a dynamic system of $f$ particles in a
228   cartesian space, where each of the $6f$ coordinates and momenta is
229   assigned to one of $6f$ mutually orthogonal axes, the phase space of
230 < this system is a $6f$ dimensional space. A point, $x = (q_1 , \ldots
231 < ,q_f ,p_1 , \ldots ,p_f )$, with a unique set of values of $6f$
232 < coordinates and momenta is a phase space vector.
230 > this system is a $6f$ dimensional space. A point, $x = (\rightarrow
231 > q_1 , \ldots ,\rightarrow q_f ,\rightarrow p_1 , \ldots ,\rightarrow
232 > p_f )$, with a unique set of values of $6f$ coordinates and momenta
233 > is a phase space vector.
234 > %%%fix me
235  
236 < A microscopic state or microstate of a classical system is
235 < specification of the complete phase space vector of a system at any
236 < instant in time. An ensemble is defined as a collection of systems
237 < sharing one or more macroscopic characteristics but each being in a
238 < unique microstate. The complete ensemble is specified by giving all
239 < systems or microstates consistent with the common macroscopic
240 < characteristics of the ensemble. Although the state of each
241 < individual system in the ensemble could be precisely described at
242 < any instance in time by a suitable phase space vector, when using
243 < ensembles for statistical purposes, there is no need to maintain
244 < distinctions between individual systems, since the numbers of
245 < systems at any time in the different states which correspond to
246 < different regions of the phase space are more interesting. Moreover,
247 < in the point of view of statistical mechanics, one would prefer to
248 < use ensembles containing a large enough population of separate
249 < members so that the numbers of systems in such different states can
250 < be regarded as changing continuously as we traverse different
251 < regions of the phase space. The condition of an ensemble at any time
236 > In statistical mechanics, the condition of an ensemble at any time
237   can be regarded as appropriately specified by the density $\rho$
238   with which representative points are distributed over the phase
239   space. The density distribution for an ensemble with $f$ degrees of
# Line 282 | Line 267 | With the help of Equation(\ref{introEquation:unitProba
267   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
268   \label{introEquation:unitProbability}
269   \end{equation}
270 < With the help of Equation(\ref{introEquation:unitProbability}) and
271 < the knowledge of the system, it is possible to calculate the average
270 > With the help of Eq.~\ref{introEquation:unitProbability} and the
271 > knowledge of the system, it is possible to calculate the average
272   value of any desired quantity which depends on the coordinates and
273   momenta of the system. Even when the dynamics of the real system is
274   complex, or stochastic, or even discontinuous, the average
# Line 306 | Line 291 | isolated and conserve energy, the Microcanonical ensem
291   thermodynamic equilibrium.
292  
293   As an ensemble of systems, each of which is known to be thermally
294 < isolated and conserve energy, the Microcanonical ensemble(NVE) has a
295 < partition function like,
294 > isolated and conserve energy, the Microcanonical ensemble (NVE) has
295 > a partition function like,
296   \begin{equation}
297   \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
298   \end{equation}
299 < A canonical ensemble(NVT)is an ensemble of systems, each of which
299 > A canonical ensemble (NVT)is an ensemble of systems, each of which
300   can share its energy with a large heat reservoir. The distribution
301   of the total energy amongst the possible dynamical states is given
302   by the partition function,
# Line 321 | Line 306 | condition, the isothermal-isobaric ensemble(NPT) plays
306   \end{equation}
307   Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
308   TS$. Since most experiments are carried out under constant pressure
309 < condition, the isothermal-isobaric ensemble(NPT) plays a very
309 > condition, the isothermal-isobaric ensemble (NPT) plays a very
310   important role in molecular simulations. The isothermal-isobaric
311   ensemble allow the system to exchange energy with a heat bath of
312   temperature $T$ and to change the volume as well. Its partition
# Line 337 | Line 322 | $\rho$, we begin from Equation(\ref{introEquation:delt
322   Liouville's theorem is the foundation on which statistical mechanics
323   rests. It describes the time evolution of the phase space
324   distribution function. In order to calculate the rate of change of
325 < $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
326 < consider the two faces perpendicular to the $q_1$ axis, which are
327 < located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
328 < leaving the opposite face is given by the expression,
325 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
326 > the two faces perpendicular to the $q_1$ axis, which are located at
327 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
328 > opposite face is given by the expression,
329   \begin{equation}
330   \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
331   \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
# Line 376 | Line 361 | statistical mechanics, since the number of particles i
361  
362   Liouville's theorem states that the distribution function is
363   constant along any trajectory in phase space. In classical
364 < statistical mechanics, since the number of particles in the system
365 < is huge, we may be able to believe the system is stationary,
364 > statistical mechanics, since the number of members in an ensemble is
365 > huge and constant, we can assume the local density has no reason
366 > (other than classical mechanics) to change,
367   \begin{equation}
368   \frac{{\partial \rho }}{{\partial t}} = 0.
369   \label{introEquation:stationary}
# Line 430 | Line 416 | Substituting equations of motion in Hamiltonian formal
416   \label{introEquation:poissonBracket}
417   \end{equation}
418   Substituting equations of motion in Hamiltonian formalism(
419 < \ref{introEquation:motionHamiltonianCoordinate} ,
420 < \ref{introEquation:motionHamiltonianMomentum} ) into
421 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
422 < theorem using Poisson bracket notion,
419 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
420 > Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
421 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
422 > Liouville's theorem using Poisson bracket notion,
423   \begin{equation}
424   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
425   {\rho ,H} \right\}.
# Line 494 | Line 480 | Leimkuhler1999}. The velocity verlet method, which hap
480   geometric integrators, which preserve various phase-flow invariants
481   such as symplectic structure, volume and time reversal symmetry, are
482   developed to address this issue\cite{Dullweber1997, McLachlan1998,
483 < Leimkuhler1999}. The velocity verlet method, which happens to be a
483 > Leimkuhler1999}. The velocity Verlet method, which happens to be a
484   simple example of symplectic integrator, continues to gain
485   popularity in the molecular dynamics community. This fact can be
486   partly explained by its geometric nature.
# Line 589 | Line 575 | Instead, we use a approximate map, $\psi_\tau$, which
575   \end{equation}
576  
577   In most cases, it is not easy to find the exact flow $\varphi_\tau$.
578 < Instead, we use a approximate map, $\psi_\tau$, which is usually
578 > Instead, we use an approximate map, $\psi_\tau$, which is usually
579   called integrator. The order of an integrator $\psi_\tau$ is $p$, if
580   the Taylor series of $\psi_\tau$ agree to order $p$,
581   \begin{equation}
582 < \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
582 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
583   \end{equation}
584  
585   \subsection{\label{introSection:geometricProperties}Geometric Properties}
586  
587 < The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
588 < and its flow play important roles in numerical studies. Many of them
589 < can be found in systems which occur naturally in applications.
587 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
588 > ODE and its flow play important roles in numerical studies. Many of
589 > them can be found in systems which occur naturally in applications.
590  
591   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
592   a \emph{symplectic} flow if it satisfies,
# Line 615 | Line 601 | is the property must be preserved by the integrator.
601   \begin{equation}
602   {\varphi '}^T J \varphi ' = J \circ \varphi
603   \end{equation}
604 < is the property must be preserved by the integrator.
604 > is the property that must be preserved by the integrator.
605  
606   It is possible to construct a \emph{volume-preserving} flow for a
607 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
607 > source free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $
608   \det d\varphi  = 1$. One can show easily that a symplectic flow will
609   be volume-preserving.
610  
611 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
612 < will result in a new system,
611 > Changing the variables $y = h(x)$ in an ODE
612 > (Eq.~\ref{introEquation:ODE}) will result in a new system,
613   \[
614   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
615   \]
# Line 673 | Line 659 | constructed. The most famous example is the Verlet-lea
659   A lot of well established and very effective numerical methods have
660   been successful precisely because of their symplecticities even
661   though this fact was not recognized when they were first
662 < constructed. The most famous example is the Verlet-leapfrog methods
662 > constructed. The most famous example is the Verlet-leapfrog method
663   in molecular dynamics. In general, symplectic integrators can be
664   constructed using one of four different methods.
665   \begin{enumerate}
# Line 750 | Line 736 | to its symmetric property,
736   \begin{equation}
737   \varphi _h^{ - 1} = \varphi _{ - h}.
738   \label{introEquation:timeReversible}
739 < \end{equation},appendixFig:architecture
739 > \end{equation}
740  
741 < \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Example of Splitting Method}}
741 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
742   The classical equation for a system consisting of interacting
743   particles can be written in Hamiltonian form,
744   \[
745   H = T + V
746   \]
747   where $T$ is the kinetic energy and $V$ is the potential energy.
748 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
748 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
749   obtains the following:
750   \begin{align}
751   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 786 | Line 772 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
772      \label{introEquation:Lp9b}\\%
773   %
774   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
775 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
775 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
776   \end{align}
777   From the preceding splitting, one can see that the integration of
778   the equations of motion would follow:
# Line 795 | Line 781 | the equations of motion would follow:
781  
782   \item Use the half step velocities to move positions one whole step, $\Delta t$.
783  
784 < \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
784 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
785  
786   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
787   \end{enumerate}
788  
789 < Simply switching the order of splitting and composing, a new
790 < integrator, the \emph{position verlet} integrator, can be generated,
789 > By simply switching the order of the propagators in the splitting
790 > and composing a new integrator, the \emph{position verlet}
791 > integrator, can be generated,
792   \begin{align}
793   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
794   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 814 | Line 801 | Baker-Campbell-Hausdorff formula can be used to determ
801  
802   \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
803  
804 < Baker-Campbell-Hausdorff formula can be used to determine the local
805 < error of splitting method in terms of commutator of the
804 > The Baker-Campbell-Hausdorff formula can be used to determine the
805 > local error of splitting method in terms of the commutator of the
806   operators(\ref{introEquation:exponentialOperator}) associated with
807 < the sub-flow. For operators $hX$ and $hY$ which are associate to
807 > the sub-flow. For operators $hX$ and $hY$ which are associated with
808   $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
809   \begin{equation}
810   \exp (hX + hY) = \exp (hZ)
# Line 831 | Line 818 | Applying Baker-Campbell-Hausdorff formula\cite{Varadar
818   \[
819   [X,Y] = XY - YX .
820   \]
821 < Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
822 < Sprang splitting, we can obtain
821 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
822 > to the Sprang splitting, we can obtain
823   \begin{eqnarray*}
824   \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 \\
825                                     &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
826                                     &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
827   \end{eqnarray*}
828 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
828 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
829   error of Spring splitting is proportional to $h^3$. The same
830 < procedure can be applied to general splitting,  of the form
830 > procedure can be applied to a general splitting,  of the form
831   \begin{equation}
832   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
833   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
834   \end{equation}
835 < Careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
836 < order method. Yoshida proposed an elegant way to compose higher
835 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
836 > order methods. Yoshida proposed an elegant way to compose higher
837   order methods based on symmetric splitting\cite{Yoshida1990}. Given
838   a symmetric second order base method $ \varphi _h^{(2)} $, a
839   fourth-order symmetric method can be constructed by composing,
# Line 859 | Line 846 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
846   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
847   \begin{equation}
848   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
849 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
849 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
850   \end{equation}
851 < , if the weights are chosen as
851 > if the weights are chosen as
852   \[
853   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
854   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 899 | Line 886 | will discusses issues in production run.
886   These three individual steps will be covered in the following
887   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
888   initialization of a simulation. Sec.~\ref{introSection:production}
889 < will discusses issues in production run.
889 > will discusse issues in production run.
890   Sec.~\ref{introSection:Analysis} provides the theoretical tools for
891   trajectory analysis.
892  
# Line 912 | Line 899 | purification and crystallization. Even for the molecul
899   databases, such as RCSB Protein Data Bank \textit{etc}. Although
900   thousands of crystal structures of molecules are discovered every
901   year, many more remain unknown due to the difficulties of
902 < purification and crystallization. Even for the molecule with known
903 < structure, some important information is missing. For example, the
902 > purification and crystallization. Even for molecules with known
903 > structure, some important information is missing. For example, a
904   missing hydrogen atom which acts as donor in hydrogen bonding must
905   be added. Moreover, in order to include electrostatic interaction,
906   one may need to specify the partial charges for individual atoms.
907   Under some circumstances, we may even need to prepare the system in
908 < a special setup. For instance, when studying transport phenomenon in
909 < membrane system, we may prepare the lipids in bilayer structure
910 < instead of placing lipids randomly in solvent, since we are not
911 < interested in self-aggregation and it takes a long time to happen.
908 > a special configuration. For instance, when studying transport
909 > phenomenon in membrane systems, we may prepare the lipids in a
910 > bilayer structure instead of placing lipids randomly in solvent,
911 > since we are not interested in the slow self-aggregation process.
912  
913   \subsubsection{\textbf{Minimization}}
914  
915   It is quite possible that some of molecules in the system from
916 < preliminary preparation may be overlapped with each other. This
917 < close proximity leads to high potential energy which consequently
918 < jeopardizes any molecular dynamics simulations. To remove these
919 < steric overlaps, one typically performs energy minimization to find
920 < a more reasonable conformation. Several energy minimization methods
921 < have been developed to exploit the energy surface and to locate the
922 < local minimum. While converging slowly near the minimum, steepest
923 < descent method is extremely robust when systems are far from
924 < harmonic. Thus, it is often used to refine structure from
925 < crystallographic data. Relied on the gradient or hessian, advanced
926 < methods like conjugate gradient and Newton-Raphson converge rapidly
927 < to a local minimum, while become unstable if the energy surface is
928 < far from quadratic. Another factor must be taken into account, when
916 > preliminary preparation may be overlapping with each other. This
917 > close proximity leads to high initial potential energy which
918 > consequently jeopardizes any molecular dynamics simulations. To
919 > remove these steric overlaps, one typically performs energy
920 > minimization to find a more reasonable conformation. Several energy
921 > minimization methods have been developed to exploit the energy
922 > surface and to locate the local minimum. While converging slowly
923 > near the minimum, steepest descent method is extremely robust when
924 > systems are strongly anharmonic. Thus, it is often used to refine
925 > structure from crystallographic data. Relied on the gradient or
926 > hessian, advanced methods like Newton-Raphson converge rapidly to a
927 > local minimum, but become unstable if the energy surface is far from
928 > quadratic. Another factor that must be taken into account, when
929   choosing energy minimization method, is the size of the system.
930   Steepest descent and conjugate gradient can deal with models of any
931 < size. Because of the limit of computation power to calculate hessian
932 < matrix and insufficient storage capacity to store them, most
933 < Newton-Raphson methods can not be used with very large models.
931 > size. Because of the limits on computer memory to store the hessian
932 > matrix and the computing power needed to diagonalized these
933 > matrices, most Newton-Raphson methods can not be used with very
934 > large systems.
935  
936   \subsubsection{\textbf{Heating}}
937  
938   Typically, Heating is performed by assigning random velocities
939 < according to a Gaussian distribution for a temperature. Beginning at
940 < a lower temperature and gradually increasing the temperature by
941 < assigning greater random velocities, we end up with setting the
942 < temperature of the system to a final temperature at which the
943 < simulation will be conducted. In heating phase, we should also keep
944 < the system from drifting or rotating as a whole. Equivalently, the
945 < net linear momentum and angular momentum of the system should be
946 < shifted to zero.
939 > according to a Maxwell-Boltzman distribution for a desired
940 > temperature. Beginning at a lower temperature and gradually
941 > increasing the temperature by assigning larger random velocities, we
942 > end up with setting the temperature of the system to a final
943 > temperature at which the simulation will be conducted. In heating
944 > phase, we should also keep the system from drifting or rotating as a
945 > whole. To do this, the net linear momentum and angular momentum of
946 > the system is shifted to zero after each resampling from the Maxwell
947 > -Boltzman distribution.
948  
949   \subsubsection{\textbf{Equilibration}}
950  
# Line 971 | Line 960 | Production run is the most important step of the simul
960  
961   \subsection{\label{introSection:production}Production}
962  
963 < Production run is the most important step of the simulation, in
963 > The production run is the most important step of the simulation, in
964   which the equilibrated structure is used as a starting point and the
965   motions of the molecules are collected for later analysis. In order
966   to capture the macroscopic properties of the system, the molecular
967 < dynamics simulation must be performed in correct and efficient way.
967 > dynamics simulation must be performed by sampling correctly and
968 > efficiently from the relevant thermodynamic ensemble.
969  
970   The most expensive part of a molecular dynamics simulation is the
971   calculation of non-bonded forces, such as van der Waals force and
972   Coulombic forces \textit{etc}. For a system of $N$ particles, the
973   complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
974   which making large simulations prohibitive in the absence of any
975 < computation saving techniques.
975 > algorithmic tricks.
976  
977 < A natural approach to avoid system size issue is to represent the
977 > A natural approach to avoid system size issues is to represent the
978   bulk behavior by a finite number of the particles. However, this
979 < approach will suffer from the surface effect. To offset this,
980 < \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
981 < is developed to simulate bulk properties with a relatively small
982 < number of particles. In this method, the simulation box is
983 < replicated throughout space to form an infinite lattice. During the
984 < simulation, when a particle moves in the primary cell, its image in
985 < other cells move in exactly the same direction with exactly the same
986 < orientation. Thus, as a particle leaves the primary cell, one of its
987 < images will enter through the opposite face.
979 > approach will suffer from the surface effect at the edges of the
980 > simulation. To offset this, \textit{Periodic boundary conditions}
981 > (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
982 > properties with a relatively small number of particles. In this
983 > method, the simulation box is replicated throughout space to form an
984 > infinite lattice. During the simulation, when a particle moves in
985 > the primary cell, its image in other cells move in exactly the same
986 > direction with exactly the same orientation. Thus, as a particle
987 > leaves the primary cell, one of its images will enter through the
988 > opposite face.
989   \begin{figure}
990   \centering
991   \includegraphics[width=\linewidth]{pbc.eps}
# Line 1006 | Line 997 | evaluation is to apply cutoff where particles farther
997  
998   %cutoff and minimum image convention
999   Another important technique to improve the efficiency of force
1000 < evaluation is to apply cutoff where particles farther than a
1001 < predetermined distance, are not included in the calculation
1000 > evaluation is to apply spherical cutoff where particles farther than
1001 > a predetermined distance are not included in the calculation
1002   \cite{Frenkel1996}. The use of a cutoff radius will cause a
1003   discontinuity in the potential energy curve. Fortunately, one can
1004 < shift the potential to ensure the potential curve go smoothly to
1005 < zero at the cutoff radius. Cutoff strategy works pretty well for
1006 < Lennard-Jones interaction because of its short range nature.
1007 < However, simply truncating the electrostatic interaction with the
1008 < use of cutoff has been shown to lead to severe artifacts in
1009 < simulations. Ewald summation, in which the slowly conditionally
1010 < convergent Coulomb potential is transformed into direct and
1011 < reciprocal sums with rapid and absolute convergence, has proved to
1012 < minimize the periodicity artifacts in liquid simulations. Taking the
1013 < advantages of the fast Fourier transform (FFT) for calculating
1014 < discrete Fourier transforms, the particle mesh-based
1004 > shift simple radial potential to ensure the potential curve go
1005 > smoothly to zero at the cutoff radius. The cutoff strategy works
1006 > well for Lennard-Jones interaction because of its short range
1007 > nature. However, simply truncating the electrostatic interaction
1008 > with the use of cutoffs has been shown to lead to severe artifacts
1009 > in simulations. The Ewald summation, in which the slowly decaying
1010 > Coulomb potential is transformed into direct and reciprocal sums
1011 > with rapid and absolute convergence, has proved to minimize the
1012 > periodicity artifacts in liquid simulations. Taking the advantages
1013 > of the fast Fourier transform (FFT) for calculating discrete Fourier
1014 > transforms, the particle mesh-based
1015   methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1016 < $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1017 < multipole method}\cite{Greengard1987, Greengard1994}, which treats
1018 < Coulombic interaction exactly at short range, and approximate the
1019 < potential at long range through multipolar expansion. In spite of
1020 < their wide acceptances at the molecular simulation community, these
1021 < two methods are hard to be implemented correctly and efficiently.
1022 < Instead, we use a damped and charge-neutralized Coulomb potential
1023 < method developed by Wolf and his coworkers\cite{Wolf1999}. The
1024 < shifted Coulomb potential for particle $i$ and particle $j$ at
1025 < distance $r_{rj}$ is given by:
1016 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
1017 > \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
1018 > which treats Coulombic interactions exactly at short range, and
1019 > approximate the potential at long range through multipolar
1020 > expansion. In spite of their wide acceptance at the molecular
1021 > simulation community, these two methods are difficult to implement
1022 > correctly and efficiently. Instead, we use a damped and
1023 > charge-neutralized Coulomb potential method developed by Wolf and
1024 > his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
1025 > particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1026   \begin{equation}
1027   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1028   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
# Line 1053 | Line 1044 | Recently, advanced visualization technique are widely
1044  
1045   \subsection{\label{introSection:Analysis} Analysis}
1046  
1047 < Recently, advanced visualization technique are widely applied to
1047 > Recently, advanced visualization technique have become applied to
1048   monitor the motions of molecules. Although the dynamics of the
1049   system can be described qualitatively from animation, quantitative
1050 < trajectory analysis are more appreciable. According to the
1051 < principles of Statistical Mechanics,
1052 < Sec.~\ref{introSection:statisticalMechanics}, one can compute
1053 < thermodynamics properties, analyze fluctuations of structural
1054 < parameters, and investigate time-dependent processes of the molecule
1064 < from the trajectories.
1050 > trajectory analysis are more useful. According to the principles of
1051 > Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
1052 > one can compute thermodynamic properties, analyze fluctuations of
1053 > structural parameters, and investigate time-dependent processes of
1054 > the molecule from the trajectories.
1055  
1056 < \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamics Properties}}
1056 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1057  
1058 < Thermodynamics properties, which can be expressed in terms of some
1058 > Thermodynamic properties, which can be expressed in terms of some
1059   function of the coordinates and momenta of all particles in the
1060   system, can be directly computed from molecular dynamics. The usual
1061   way to measure the pressure is based on virial theorem of Clausius
# Line 1088 | Line 1078 | distribution functions. Among these functions,\emph{pa
1078   \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1079  
1080   Structural Properties of a simple fluid can be described by a set of
1081 < distribution functions. Among these functions,\emph{pair
1081 > distribution functions. Among these functions,the \emph{pair
1082   distribution function}, also known as \emph{radial distribution
1083 < function}, is of most fundamental importance to liquid-state theory.
1084 < Pair distribution function can be gathered by Fourier transforming
1085 < raw data from a series of neutron diffraction experiments and
1086 < integrating over the surface factor \cite{Powles1973}. The
1087 < experiment result can serve as a criterion to justify the
1088 < correctness of the theory. Moreover, various equilibrium
1089 < thermodynamic and structural properties can also be expressed in
1090 < terms of radial distribution function \cite{Allen1987}.
1083 > function}, is of most fundamental importance to liquid theory.
1084 > Experimentally, pair distribution function can be gathered by
1085 > Fourier transforming raw data from a series of neutron diffraction
1086 > experiments and integrating over the surface factor
1087 > \cite{Powles1973}. The experimental results can serve as a criterion
1088 > to justify the correctness of a liquid model. Moreover, various
1089 > equilibrium thermodynamic and structural properties can also be
1090 > expressed in terms of radial distribution function \cite{Allen1987}.
1091  
1092 < A pair distribution functions $g(r)$ gives the probability that a
1092 > The pair distribution functions $g(r)$ gives the probability that a
1093   particle $i$ will be located at a distance $r$ from a another
1094   particle $j$ in the system
1095   \[
1096   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1097 < \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1097 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1098 > (r)}{\rho}.
1099   \]
1100   Note that the delta function can be replaced by a histogram in
1101 < computer simulation. Figure
1102 < \ref{introFigure:pairDistributionFunction} shows a typical pair
1103 < distribution function for the liquid argon system. The occurrence of
1113 < several peaks in the plot of $g(r)$ suggests that it is more likely
1114 < to find particles at certain radial values than at others. This is a
1115 < result of the attractive interaction at such distances. Because of
1116 < the strong repulsive forces at short distance, the probability of
1117 < locating particles at distances less than about 2.5{\AA} from each
1118 < other is essentially zero.
1101 > computer simulation. Peaks in $g(r)$ represent solvent shells, and
1102 > the height of these peaks gradually decreases to 1 as the liquid of
1103 > large distance approaches the bulk density.
1104  
1120 %\begin{figure}
1121 %\centering
1122 %\includegraphics[width=\linewidth]{pdf.eps}
1123 %\caption[Pair distribution function for the liquid argon
1124 %]{Pair distribution function for the liquid argon}
1125 %\label{introFigure:pairDistributionFunction}
1126 %\end{figure}
1105  
1106   \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1107   Properties}}
1108  
1109   Time-dependent properties are usually calculated using \emph{time
1110 < correlation function}, which correlates random variables $A$ and $B$
1111 < at two different time
1110 > correlation functions}, which correlate random variables $A$ and $B$
1111 > at two different times,
1112   \begin{equation}
1113   C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1114   \label{introEquation:timeCorrelationFunction}
1115   \end{equation}
1116   If $A$ and $B$ refer to same variable, this kind of correlation
1117 < function is called \emph{auto correlation function}. One example of
1118 < auto correlation function is velocity auto-correlation function
1119 < which is directly related to transport properties of molecular
1120 < liquids:
1117 > function is called an \emph{autocorrelation function}. One example
1118 > of an auto correlation function is the velocity auto-correlation
1119 > function which is directly related to transport properties of
1120 > molecular liquids:
1121   \[
1122   D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1123   \right\rangle } dt
1124   \]
1125 < where $D$ is diffusion constant. Unlike velocity autocorrelation
1126 < function which is averaging over time origins and over all the
1127 < atoms, dipole autocorrelation are calculated for the entire system.
1128 < The dipole autocorrelation function is given by:
1125 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1126 > function, which is averaging over time origins and over all the
1127 > atoms, the dipole autocorrelation functions are calculated for the
1128 > entire system. The dipole autocorrelation function is given by:
1129   \[
1130   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1131   \right\rangle
# Line 1173 | Line 1151 | simulator is governed by the rigid body dynamics. In m
1151   areas, from engineering, physics, to chemistry. For example,
1152   missiles and vehicle are usually modeled by rigid bodies.  The
1153   movement of the objects in 3D gaming engine or other physics
1154 < simulator is governed by the rigid body dynamics. In molecular
1155 < simulation, rigid body is used to simplify the model in
1156 < protein-protein docking study\cite{Gray2003}.
1154 > simulator is governed by rigid body dynamics. In molecular
1155 > simulations, rigid bodies are used to simplify protein-protein
1156 > docking studies\cite{Gray2003}.
1157  
1158   It is very important to develop stable and efficient methods to
1159 < integrate the equations of motion of orientational degrees of
1160 < freedom. Euler angles are the nature choice to describe the
1161 < rotational degrees of freedom. However, due to its singularity, the
1162 < numerical integration of corresponding equations of motion is very
1163 < inefficient and inaccurate. Although an alternative integrator using
1164 < different sets of Euler angles can overcome this
1165 < difficulty\cite{Barojas1973}, the computational penalty and the lost
1166 < of angular momentum conservation still remain. A singularity free
1167 < representation utilizing quaternions was developed by Evans in
1168 < 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1169 < nonseparable Hamiltonian resulted from quaternion representation,
1170 < which prevents the symplectic algorithm to be utilized. Another
1171 < different approach is to apply holonomic constraints to the atoms
1172 < belonging to the rigid body. Each atom moves independently under the
1173 < normal forces deriving from potential energy and constraint forces
1174 < which are used to guarantee the rigidness. However, due to their
1175 < iterative nature, SHAKE and Rattle algorithm converge very slowly
1176 < when the number of constraint increases\cite{Ryckaert1977,
1177 < Andersen1983}.
1159 > integrate the equations of motion for orientational degrees of
1160 > freedom. Euler angles are the natural choice to describe the
1161 > rotational degrees of freedom. However, due to $\frac {1}{sin
1162 > \theta}$ singularities, the numerical integration of corresponding
1163 > equations of motion is very inefficient and inaccurate. Although an
1164 > alternative integrator using multiple sets of Euler angles can
1165 > overcome this difficulty\cite{Barojas1973}, the computational
1166 > penalty and the loss of angular momentum conservation still remain.
1167 > A singularity-free representation utilizing quaternions was
1168 > developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1169 > approach uses a nonseparable Hamiltonian resulting from the
1170 > quaternion representation, which prevents the symplectic algorithm
1171 > to be utilized. Another different approach is to apply holonomic
1172 > constraints to the atoms belonging to the rigid body. Each atom
1173 > moves independently under the normal forces deriving from potential
1174 > energy and constraint forces which are used to guarantee the
1175 > rigidness. However, due to their iterative nature, the SHAKE and
1176 > Rattle algorithms also converge very slowly when the number of
1177 > constraints increases\cite{Ryckaert1977, Andersen1983}.
1178  
1179 < The break through in geometric literature suggests that, in order to
1179 > A break-through in geometric literature suggests that, in order to
1180   develop a long-term integration scheme, one should preserve the
1181 < symplectic structure of the flow. Introducing conjugate momentum to
1182 < rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1183 < symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1184 < the Hamiltonian system in a constraint manifold by iteratively
1185 < satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1186 < method using quaternion representation was developed by
1187 < Omelyan\cite{Omelyan1998}. However, both of these methods are
1188 < iterative and inefficient. In this section, we will present a
1181 > symplectic structure of the flow. By introducing a conjugate
1182 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1183 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1184 > proposed to evolve the Hamiltonian system in a constraint manifold
1185 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1186 > An alternative method using the quaternion representation was
1187 > developed by Omelyan\cite{Omelyan1998}. However, both of these
1188 > methods are iterative and inefficient. In this section, we descibe a
1189   symplectic Lie-Poisson integrator for rigid body developed by
1190   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1191  
1192 < \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1193 < The motion of the rigid body is Hamiltonian with the Hamiltonian
1192 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1193 > The motion of a rigid body is Hamiltonian with the Hamiltonian
1194   function
1195   \begin{equation}
1196   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
# Line 1226 | Line 1204 | constrained Hamiltonian equation subjects to a holonom
1204   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1205   \]
1206   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1207 < constrained Hamiltonian equation subjects to a holonomic constraint,
1207 > constrained Hamiltonian equation is subjected to a holonomic
1208 > constraint,
1209   \begin{equation}
1210   Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1211   \end{equation}
1212 < which is used to ensure rotation matrix's orthogonality.
1213 < Differentiating \ref{introEquation:orthogonalConstraint} and using
1214 < Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1212 > which is used to ensure rotation matrix's unitarity. Differentiating
1213 > \ref{introEquation:orthogonalConstraint} and using Equation
1214 > \ref{introEquation:RBMotionMomentum}, one may obtain,
1215   \begin{equation}
1216   Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1217   \label{introEquation:RBFirstOrderConstraint}
# Line 1250 | Line 1229 | We can use constraint force provided by lagrange multi
1229   \end{eqnarray}
1230  
1231   In general, there are two ways to satisfy the holonomic constraints.
1232 < We can use constraint force provided by lagrange multiplier on the
1233 < normal manifold to keep the motion on constraint space. Or we can
1234 < simply evolve the system in constraint manifold. These two methods
1235 < are proved to be equivalent. The holonomic constraint and equations
1236 < of motions define a constraint manifold for rigid body
1232 > We can use a constraint force provided by a Lagrange multiplier on
1233 > the normal manifold to keep the motion on constraint space. Or we
1234 > can simply evolve the system on the constraint manifold. These two
1235 > methods have been proved to be equivalent. The holonomic constraint
1236 > and equations of motions define a constraint manifold for rigid
1237 > bodies
1238   \[
1239   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1240   \right\}.
1241   \]
1242  
1243   Unfortunately, this constraint manifold is not the cotangent bundle
1244 < $T_{\star}SO(3)$. However, it turns out that under symplectic
1244 > $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1245 > rotation group $SO(3)$. However, it turns out that under symplectic
1246   transformation, the cotangent space and the phase space are
1247 < diffeomorphic. Introducing
1247 > diffeomorphic. By introducing
1248   \[
1249   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1250   \]
# Line 1295 | Line 1276 | body, angular momentum on body frame $\Pi  = Q^t P$ is
1276   respectively.
1277  
1278   As a common choice to describe the rotation dynamics of the rigid
1279 < body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1280 < rewrite the equations of motion,
1279 > body, the angular momentum on the body fixed frame $\Pi  = Q^t P$ is
1280 > introduced to rewrite the equations of motion,
1281   \begin{equation}
1282   \begin{array}{l}
1283 < \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1284 < \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1283 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1284 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1}  \\
1285   \end{array}
1286   \label{introEqaution:RBMotionPI}
1287   \end{equation}
# Line 1330 | Line 1311 | matrix,
1311   \]
1312   Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1313   matrix,
1314 < \begin{equation}
1315 < (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ {\bullet  ^T}
1316 < ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1317 < - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1318 < (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1319 < \end{equation}
1314 >
1315 > \begin{eqnarray*}
1316 > (\dot \Pi  - \dot \Pi ^T ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{
1317 > }}(J^{ - 1} \Pi  + \Pi J^{ - 1} ) + \sum\limits_i {[Q^T F_i
1318 > (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  - (\Lambda  - \Lambda ^T ).
1319 > \label{introEquation:skewMatrixPI}
1320 > \end{eqnarray*}
1321 >
1322   Since $\Lambda$ is symmetric, the last term of Equation
1323   \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1324   multiplier $\Lambda$ is absent from the equations of motion. This
1325 < unique property eliminate the requirement of iterations which can
1325 > unique property eliminates the requirement of iterations which can
1326   not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1327  
1328 < Applying hat-map isomorphism, we obtain the equation of motion for
1329 < angular momentum on body frame
1328 > Applying the hat-map isomorphism, we obtain the equation of motion
1329 > for angular momentum on body frame
1330   \begin{equation}
1331   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1332   F_i (r,Q)} \right) \times X_i }.
# Line 1358 | Line 1341 | If there is not external forces exerted on the rigid b
1341   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1342   Lie-Poisson Integrator for Free Rigid Body}
1343  
1344 < If there is not external forces exerted on the rigid body, the only
1345 < contribution to the rotational is from the kinetic potential (the
1346 < first term of \ref{introEquation:bodyAngularMotion}). The free rigid
1347 < body is an example of Lie-Poisson system with Hamiltonian function
1344 > If there are no external forces exerted on the rigid body, the only
1345 > contribution to the rotational motion is from the kinetic energy
1346 > (the first term of \ref{introEquation:bodyAngularMotion}). The free
1347 > rigid body is an example of a Lie-Poisson system with Hamiltonian
1348 > function
1349   \begin{equation}
1350   T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1351   \label{introEquation:rotationalKineticRB}
# Line 1408 | Line 1392 | tR_1 }$, we can use Cayley transformation,
1392   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1393   \]
1394   To reduce the cost of computing expensive functions in $e^{\Delta
1395 < tR_1 }$, we can use Cayley transformation,
1395 > tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1396 > propagator,
1397   \[
1398   e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1399   )
1400   \]
1401   The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1402 < manner.
1403 <
1419 < In order to construct a second-order symplectic method, we split the
1420 < angular kinetic Hamiltonian function can into five terms
1402 > manner. In order to construct a second-order symplectic method, we
1403 > split the angular kinetic Hamiltonian function can into five terms
1404   \[
1405   T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1406   ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1407 < (\pi _1 )
1408 < \].
1409 < Concatenating flows corresponding to these five terms, we can obtain
1410 < an symplectic integrator,
1407 > (\pi _1 ).
1408 > \]
1409 > By concatenating the propagators corresponding to these five terms,
1410 > we can obtain an symplectic integrator,
1411   \[
1412   \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1413   \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
# Line 1451 | Line 1434 | Lie-Poisson integrator is found to be extremely effici
1434   \]
1435   Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1436   \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1437 < Lie-Poisson integrator is found to be extremely efficient and stable
1438 < which can be explained by the fact the small angle approximation is
1439 < used and the norm of the angular momentum is conserved.
1437 > Lie-Poisson integrator is found to be both extremely efficient and
1438 > stable. These properties can be explained by the fact the small
1439 > angle approximation is used and the norm of the angular momentum is
1440 > conserved.
1441  
1442   \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1443   Splitting for Rigid Body}
# Line 1466 | Line 1450 | kinetic energy are listed in the below table,
1450   The equations of motion corresponding to potential energy and
1451   kinetic energy are listed in the below table,
1452   \begin{table}
1453 < \caption{Equations of motion due to Potential and Kinetic Energies}
1453 > \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1454   \begin{center}
1455   \begin{tabular}{|l|l|}
1456    \hline
# Line 1480 | Line 1464 | A second-order symplectic method is now obtained by th
1464   \end{tabular}
1465   \end{center}
1466   \end{table}
1467 < A second-order symplectic method is now obtained by the
1468 < composition of the flow maps,
1467 > A second-order symplectic method is now obtained by the composition
1468 > of the position and velocity propagators,
1469   \[
1470   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1471   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1472   \]
1473   Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1474 < sub-flows which corresponding to force and torque respectively,
1474 > sub-propagators which corresponding to force and torque
1475 > respectively,
1476   \[
1477   \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1478   _{\Delta t/2,\tau }.
1479   \]
1480   Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1481 < $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1482 < order inside $\varphi _{\Delta t/2,V}$ does not matter.
1483 <
1484 < Furthermore, kinetic potential can be separated to translational
1500 < kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1481 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1482 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1483 > kinetic energy can be separated to translational kinetic term, $T^t
1484 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1485   \begin{equation}
1486   T(p,\pi ) =T^t (p) + T^r (\pi ).
1487   \end{equation}
1488   where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1489   defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1490 < corresponding flow maps are given by
1490 > corresponding propagators are given by
1491   \[
1492   \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1493   _{\Delta t,T^r }.
1494   \]
1495 < Finally, we obtain the overall symplectic flow maps for free moving
1496 < rigid body
1495 > Finally, we obtain the overall symplectic propagators for freely
1496 > moving rigid bodies
1497   \begin{equation}
1498   \begin{array}{c}
1499   \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
# Line 1523 | Line 1507 | the theory of Langevin dynamics simulation. A brief de
1507   As an alternative to newtonian dynamics, Langevin dynamics, which
1508   mimics a simple heat bath with stochastic and dissipative forces,
1509   has been applied in a variety of studies. This section will review
1510 < the theory of Langevin dynamics simulation. A brief derivation of
1511 < generalized Langevin equation will be given first. Follow that, we
1512 < will discuss the physical meaning of the terms appearing in the
1513 < equation as well as the calculation of friction tensor from
1514 < hydrodynamics theory.
1510 > the theory of Langevin dynamics. A brief derivation of generalized
1511 > Langevin equation will be given first. Following that, we will
1512 > discuss the physical meaning of the terms appearing in the equation
1513 > as well as the calculation of friction tensor from hydrodynamics
1514 > theory.
1515  
1516   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1517  
1518 < Harmonic bath model, in which an effective set of harmonic
1518 > A harmonic bath model, in which an effective set of harmonic
1519   oscillators are used to mimic the effect of a linearly responding
1520   environment, has been widely used in quantum chemistry and
1521   statistical mechanics. One of the successful applications of
1522 < Harmonic bath model is the derivation of Deriving Generalized
1523 < Langevin Dynamics. Lets consider a system, in which the degree of
1522 > Harmonic bath model is the derivation of the Generalized Langevin
1523 > Dynamics (GLE). Lets consider a system, in which the degree of
1524   freedom $x$ is assumed to couple to the bath linearly, giving a
1525   Hamiltonian of the form
1526   \begin{equation}
1527   H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1528   \label{introEquation:bathGLE}.
1529   \end{equation}
1530 < Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1531 < with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1530 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1531 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1532   \[
1533   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1534   }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
# Line 1552 | Line 1536 | the harmonic bath masses, and $\Delta U$ is bilinear s
1536   \]
1537   where the index $\alpha$ runs over all the bath degrees of freedom,
1538   $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1539 < the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1539 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1540   coupling,
1541   \[
1542   \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1543   \]
1544 < where $g_\alpha$ are the coupling constants between the bath and the
1545 < coordinate $x$. Introducing
1544 > where $g_\alpha$ are the coupling constants between the bath
1545 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1546 > Introducing
1547   \[
1548   W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1549   }}{{2m_\alpha  w_\alpha ^2 }}} x^2
# Line 1573 | Line 1558 | Generalized Langevin Dynamics by Hamilton's equations
1558   \]
1559   Since the first two terms of the new Hamiltonian depend only on the
1560   system coordinates, we can get the equations of motion for
1561 < Generalized Langevin Dynamics by Hamilton's equations
1577 < \ref{introEquation:motionHamiltonianCoordinate,
1578 < introEquation:motionHamiltonianMomentum},
1561 > Generalized Langevin Dynamics by Hamilton's equations,
1562   \begin{equation}
1563   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1564   \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
# Line 1592 | Line 1575 | differential equations, Laplace transform is the appro
1575   In order to derive an equation for $x$, the dynamics of the bath
1576   variables $x_\alpha$ must be solved exactly first. As an integral
1577   transform which is particularly useful in solving linear ordinary
1578 < differential equations, Laplace transform is the appropriate tool to
1579 < solve this problem. The basic idea is to transform the difficult
1578 > differential equations,the Laplace transform is the appropriate tool
1579 > to solve this problem. The basic idea is to transform the difficult
1580   differential equations into simple algebra problems which can be
1581 < solved easily. Then applying inverse Laplace transform, also known
1582 < as the Bromwich integral, we can retrieve the solutions of the
1581 > solved easily. Then, by applying the inverse Laplace transform, also
1582 > known as the Bromwich integral, we can retrieve the solutions of the
1583   original problems.
1584  
1585   Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
# Line 1616 | Line 1599 | Applying Laplace transform to the bath coordinates, we
1599   \end{eqnarray*}
1600  
1601  
1602 < Applying Laplace transform to the bath coordinates, we obtain
1602 > Applying the Laplace transform to the bath coordinates, we obtain
1603   \begin{eqnarray*}
1604   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) \\
1605   L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
# Line 1693 | Line 1676 | as the model, which is gaussian distribution in genera
1676   \end{array}
1677   \]
1678   This property is what we expect from a truly random process. As long
1679 < as the model, which is gaussian distribution in general, chosen for
1680 < $R(t)$ is a truly random process, the stochastic nature of the GLE
1698 < still remains.
1679 > as the model chosen for $R(t)$ was a gaussian distribution in
1680 > general, the stochastic nature of the GLE still remains.
1681  
1682   %dynamic friction kernel
1683   The convolution integral
# Line 1716 | Line 1698 | which can be used to describe dynamic caging effect. T
1698   m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1699   \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1700   \]
1701 < which can be used to describe dynamic caging effect. The other
1702 < extreme is the bath that responds infinitely quickly to motions in
1703 < the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1704 < time:
1701 > which can be used to describe the effect of dynamic caging in
1702 > viscous solvents. The other extreme is the bath that responds
1703 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1704 > taken as a $delta$ function in time:
1705   \[
1706   \xi (t) = 2\xi _0 \delta (t)
1707   \]
# Line 1735 | Line 1717 | or be determined by Stokes' law for regular shaped par
1717   \end{equation}
1718   which is known as the Langevin equation. The static friction
1719   coefficient $\xi _0$ can either be calculated from spectral density
1720 < or be determined by Stokes' law for regular shaped particles.A
1720 > or be determined by Stokes' law for regular shaped particles. A
1721   briefly review on calculating friction tensor for arbitrary shaped
1722   particles is given in Sec.~\ref{introSection:frictionTensor}.
1723  
# Line 1768 | Line 1750 | can model the random force and friction kernel.
1750   \end{equation}
1751   In effect, it acts as a constraint on the possible ways in which one
1752   can model the random force and friction kernel.
1771
1772 \subsection{\label{introSection:frictionTensor} Friction Tensor}
1773 Theoretically, the friction kernel can be determined using velocity
1774 autocorrelation function. However, this approach become impractical
1775 when the system become more and more complicate. Instead, various
1776 approaches based on hydrodynamics have been developed to calculate
1777 the friction coefficients. The friction effect is isotropic in
1778 Equation, $\zeta$ can be taken as a scalar. In general, friction
1779 tensor $\Xi$ is a $6\times 6$ matrix given by
1780 \[
1781 \Xi  = \left( {\begin{array}{*{20}c}
1782   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1783   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1784 \end{array}} \right).
1785 \]
1786 Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1787 tensor and rotational resistance (friction) tensor respectively,
1788 while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1789 {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1790 particle moves in a fluid, it may experience friction force or
1791 torque along the opposite direction of the velocity or angular
1792 velocity,
1793 \[
1794 \left( \begin{array}{l}
1795 F_R  \\
1796 \tau _R  \\
1797 \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1798   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1799   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1800 \end{array}} \right)\left( \begin{array}{l}
1801 v \\
1802 w \\
1803 \end{array} \right)
1804 \]
1805 where $F_r$ is the friction force and $\tau _R$ is the friction
1806 toque.
1807
1808 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
1809
1810 For a spherical particle, the translational and rotational friction
1811 constant can be calculated from Stoke's law,
1812 \[
1813 \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1814   {6\pi \eta R} & 0 & 0  \\
1815   0 & {6\pi \eta R} & 0  \\
1816   0 & 0 & {6\pi \eta R}  \\
1817 \end{array}} \right)
1818 \]
1819 and
1820 \[
1821 \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1822   {8\pi \eta R^3 } & 0 & 0  \\
1823   0 & {8\pi \eta R^3 } & 0  \\
1824   0 & 0 & {8\pi \eta R^3 }  \\
1825 \end{array}} \right)
1826 \]
1827 where $\eta$ is the viscosity of the solvent and $R$ is the
1828 hydrodynamics radius.
1829
1830 Other non-spherical shape, such as cylinder and ellipsoid
1831 \textit{etc}, are widely used as reference for developing new
1832 hydrodynamics theory, because their properties can be calculated
1833 exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1834 also called a triaxial ellipsoid, which is given in Cartesian
1835 coordinates by\cite{Perrin1934, Perrin1936}
1836 \[
1837 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1838 }} = 1
1839 \]
1840 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1841 due to the complexity of the elliptic integral, only the ellipsoid
1842 with the restriction of two axes having to be equal, \textit{i.e.}
1843 prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1844 exactly. Introducing an elliptic integral parameter $S$ for prolate,
1845 \[
1846 S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1847 } }}{b},
1848 \]
1849 and oblate,
1850 \[
1851 S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1852 }}{a}
1853 \],
1854 one can write down the translational and rotational resistance
1855 tensors
1856 \[
1857 \begin{array}{l}
1858 \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1859 \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1860 \end{array},
1861 \]
1862 and
1863 \[
1864 \begin{array}{l}
1865 \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1866 \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1867 \end{array}.
1868 \]
1869
1870 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shape}}
1871
1872 Unlike spherical and other regular shaped molecules, there is not
1873 analytical solution for friction tensor of any arbitrary shaped
1874 rigid molecules. The ellipsoid of revolution model and general
1875 triaxial ellipsoid model have been used to approximate the
1876 hydrodynamic properties of rigid bodies. However, since the mapping
1877 from all possible ellipsoidal space, $r$-space, to all possible
1878 combination of rotational diffusion coefficients, $D$-space is not
1879 unique\cite{Wegener1979} as well as the intrinsic coupling between
1880 translational and rotational motion of rigid body, general ellipsoid
1881 is not always suitable for modeling arbitrarily shaped rigid
1882 molecule. A number of studies have been devoted to determine the
1883 friction tensor for irregularly shaped rigid bodies using more
1884 advanced method where the molecule of interest was modeled by
1885 combinations of spheres(beads)\cite{Carrasco1999} and the
1886 hydrodynamics properties of the molecule can be calculated using the
1887 hydrodynamic interaction tensor. Let us consider a rigid assembly of
1888 $N$ beads immersed in a continuous medium. Due to hydrodynamics
1889 interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1890 than its unperturbed velocity $v_i$,
1891 \[
1892 v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1893 \]
1894 where $F_i$ is the frictional force, and $T_{ij}$ is the
1895 hydrodynamic interaction tensor. The friction force of $i$th bead is
1896 proportional to its ``net'' velocity
1897 \begin{equation}
1898 F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1899 \label{introEquation:tensorExpression}
1900 \end{equation}
1901 This equation is the basis for deriving the hydrodynamic tensor. In
1902 1930, Oseen and Burgers gave a simple solution to Equation
1903 \ref{introEquation:tensorExpression}
1904 \begin{equation}
1905 T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1906 R_{ij}^T }}{{R_{ij}^2 }}} \right).
1907 \label{introEquation:oseenTensor}
1908 \end{equation}
1909 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1910 A second order expression for element of different size was
1911 introduced by Rotne and Prager\cite{Rotne1969} and improved by
1912 Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1913 \begin{equation}
1914 T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1915 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1916 _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1917 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1918 \label{introEquation:RPTensorNonOverlapped}
1919 \end{equation}
1920 Both of the Equation \ref{introEquation:oseenTensor} and Equation
1921 \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1922 \ge \sigma _i  + \sigma _j$. An alternative expression for
1923 overlapping beads with the same radius, $\sigma$, is given by
1924 \begin{equation}
1925 T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1926 \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1927 \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1928 \label{introEquation:RPTensorOverlapped}
1929 \end{equation}
1930
1931 To calculate the resistance tensor at an arbitrary origin $O$, we
1932 construct a $3N \times 3N$ matrix consisting of $N \times N$
1933 $B_{ij}$ blocks
1934 \begin{equation}
1935 B = \left( {\begin{array}{*{20}c}
1936   {B_{11} } &  \ldots  & {B_{1N} }  \\
1937    \vdots  &  \ddots  &  \vdots   \\
1938   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1939 \end{array}} \right),
1940 \end{equation}
1941 where $B_{ij}$ is given by
1942 \[
1943 B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1944 )T_{ij}
1945 \]
1946 where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1947 $B$, we obtain
1948
1949 \[
1950 C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1951   {C_{11} } &  \ldots  & {C_{1N} }  \\
1952    \vdots  &  \ddots  &  \vdots   \\
1953   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1954 \end{array}} \right)
1955 \]
1956 , which can be partitioned into $N \times N$ $3 \times 3$ block
1957 $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1958 \[
1959 U_i  = \left( {\begin{array}{*{20}c}
1960   0 & { - z_i } & {y_i }  \\
1961   {z_i } & 0 & { - x_i }  \\
1962   { - y_i } & {x_i } & 0  \\
1963 \end{array}} \right)
1964 \]
1965 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1966 bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1967 arbitrary origin $O$ can be written as
1968 \begin{equation}
1969 \begin{array}{l}
1970 \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1971 \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1972 \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1973 \end{array}
1974 \label{introEquation:ResistanceTensorArbitraryOrigin}
1975 \end{equation}
1976
1977 The resistance tensor depends on the origin to which they refer. The
1978 proper location for applying friction force is the center of
1979 resistance (reaction), at which the trace of rotational resistance
1980 tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1981 resistance is defined as an unique point of the rigid body at which
1982 the translation-rotation coupling tensor are symmetric,
1983 \begin{equation}
1984 \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1985 \label{introEquation:definitionCR}
1986 \end{equation}
1987 Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
1988 we can easily find out that the translational resistance tensor is
1989 origin independent, while the rotational resistance tensor and
1990 translation-rotation coupling resistance tensor depend on the
1991 origin. Given resistance tensor at an arbitrary origin $O$, and a
1992 vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
1993 obtain the resistance tensor at $P$ by
1994 \begin{equation}
1995 \begin{array}{l}
1996 \Xi _P^{tt}  = \Xi _O^{tt}  \\
1997 \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
1998 \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{{tr} ^{^T }}  \\
1999 \end{array}
2000 \label{introEquation:resistanceTensorTransformation}
2001 \end{equation}
2002 where
2003 \[
2004 U_{OP}  = \left( {\begin{array}{*{20}c}
2005   0 & { - z_{OP} } & {y_{OP} }  \\
2006   {z_i } & 0 & { - x_{OP} }  \\
2007   { - y_{OP} } & {x_{OP} } & 0  \\
2008 \end{array}} \right)
2009 \]
2010 Using Equations \ref{introEquation:definitionCR} and
2011 \ref{introEquation:resistanceTensorTransformation}, one can locate
2012 the position of center of resistance,
2013 \begin{eqnarray*}
2014 \left( \begin{array}{l}
2015 x_{OR}  \\
2016 y_{OR}  \\
2017 z_{OR}  \\
2018 \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2019   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2020   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2021   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2022 \end{array}} \right)^{ - 1}  \\
2023  & & \left( \begin{array}{l}
2024 (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2025 (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2026 (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2027 \end{array} \right) \\
2028 \end{eqnarray*}
2029
2030
2031
2032 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2033 joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines