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 2881 by tim, Fri Jun 23 20:21:54 2006 UTC

# 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 231 | Line 231 | A microscopic state or microstate of a classical syste
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.
233  
234 + %%%fix me
235   A microscopic state or microstate of a classical system is
236   specification of the complete phase space vector of a system at any
237   instant in time. An ensemble is defined as a collection of systems
# Line 282 | Line 283 | With the help of Equation(\ref{introEquation:unitProba
283   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
284   \label{introEquation:unitProbability}
285   \end{equation}
286 < With the help of Equation(\ref{introEquation:unitProbability}) and
287 < the knowledge of the system, it is possible to calculate the average
286 > With the help of Eq.~\ref{introEquation:unitProbability} and the
287 > knowledge of the system, it is possible to calculate the average
288   value of any desired quantity which depends on the coordinates and
289   momenta of the system. Even when the dynamics of the real system is
290   complex, or stochastic, or even discontinuous, the average
# Line 306 | Line 307 | isolated and conserve energy, the Microcanonical ensem
307   thermodynamic equilibrium.
308  
309   As an ensemble of systems, each of which is known to be thermally
310 < isolated and conserve energy, the Microcanonical ensemble(NVE) has a
311 < partition function like,
310 > isolated and conserve energy, the Microcanonical ensemble (NVE) has
311 > a partition function like,
312   \begin{equation}
313   \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
314   \end{equation}
315 < A canonical ensemble(NVT)is an ensemble of systems, each of which
315 > A canonical ensemble (NVT)is an ensemble of systems, each of which
316   can share its energy with a large heat reservoir. The distribution
317   of the total energy amongst the possible dynamical states is given
318   by the partition function,
# Line 321 | Line 322 | condition, the isothermal-isobaric ensemble(NPT) plays
322   \end{equation}
323   Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
324   TS$. Since most experiments are carried out under constant pressure
325 < condition, the isothermal-isobaric ensemble(NPT) plays a very
325 > condition, the isothermal-isobaric ensemble (NPT) plays a very
326   important role in molecular simulations. The isothermal-isobaric
327   ensemble allow the system to exchange energy with a heat bath of
328   temperature $T$ and to change the volume as well. Its partition
# Line 337 | Line 338 | $\rho$, we begin from Equation(\ref{introEquation:delt
338   Liouville's theorem is the foundation on which statistical mechanics
339   rests. It describes the time evolution of the phase space
340   distribution function. In order to calculate the rate of change of
341 < $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
342 < consider the two faces perpendicular to the $q_1$ axis, which are
343 < located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
344 < leaving the opposite face is given by the expression,
341 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
342 > the two faces perpendicular to the $q_1$ axis, which are located at
343 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
344 > opposite face is given by the expression,
345   \begin{equation}
346   \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
347   \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
# Line 376 | Line 377 | statistical mechanics, since the number of particles i
377  
378   Liouville's theorem states that the distribution function is
379   constant along any trajectory in phase space. In classical
380 < statistical mechanics, since the number of particles in the system
381 < is huge, we may be able to believe the system is stationary,
380 > statistical mechanics, since the number of members in an ensemble is
381 > huge and constant, we can assume the local density has no reason
382 > (other than classical mechanics) to change,
383   \begin{equation}
384   \frac{{\partial \rho }}{{\partial t}} = 0.
385   \label{introEquation:stationary}
# Line 430 | Line 432 | Substituting equations of motion in Hamiltonian formal
432   \label{introEquation:poissonBracket}
433   \end{equation}
434   Substituting equations of motion in Hamiltonian formalism(
435 < \ref{introEquation:motionHamiltonianCoordinate} ,
436 < \ref{introEquation:motionHamiltonianMomentum} ) into
437 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
438 < theorem using Poisson bracket notion,
435 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
436 > Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
437 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
438 > Liouville's theorem using Poisson bracket notion,
439   \begin{equation}
440   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
441   {\rho ,H} \right\}.
# Line 494 | Line 496 | Leimkuhler1999}. The velocity verlet method, which hap
496   geometric integrators, which preserve various phase-flow invariants
497   such as symplectic structure, volume and time reversal symmetry, are
498   developed to address this issue\cite{Dullweber1997, McLachlan1998,
499 < Leimkuhler1999}. The velocity verlet method, which happens to be a
499 > Leimkuhler1999}. The velocity Verlet method, which happens to be a
500   simple example of symplectic integrator, continues to gain
501   popularity in the molecular dynamics community. This fact can be
502   partly explained by its geometric nature.
# Line 589 | Line 591 | Instead, we use a approximate map, $\psi_\tau$, which
591   \end{equation}
592  
593   In most cases, it is not easy to find the exact flow $\varphi_\tau$.
594 < Instead, we use a approximate map, $\psi_\tau$, which is usually
594 > Instead, we use an approximate map, $\psi_\tau$, which is usually
595   called integrator. The order of an integrator $\psi_\tau$ is $p$, if
596   the Taylor series of $\psi_\tau$ agree to order $p$,
597   \begin{equation}
598 < \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
598 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
599   \end{equation}
600  
601   \subsection{\label{introSection:geometricProperties}Geometric Properties}
602  
603 < The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
604 < and its flow play important roles in numerical studies. Many of them
605 < can be found in systems which occur naturally in applications.
603 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
604 > ODE and its flow play important roles in numerical studies. Many of
605 > them can be found in systems which occur naturally in applications.
606  
607   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
608   a \emph{symplectic} flow if it satisfies,
# Line 615 | Line 617 | is the property must be preserved by the integrator.
617   \begin{equation}
618   {\varphi '}^T J \varphi ' = J \circ \varphi
619   \end{equation}
620 < is the property must be preserved by the integrator.
620 > is the property that must be preserved by the integrator.
621  
622   It is possible to construct a \emph{volume-preserving} flow for a
623 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
623 > source free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $
624   \det d\varphi  = 1$. One can show easily that a symplectic flow will
625   be volume-preserving.
626  
627 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
628 < will result in a new system,
627 > Changing the variables $y = h(x)$ in an ODE
628 > (Eq.~\ref{introEquation:ODE}) will result in a new system,
629   \[
630   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
631   \]
# Line 673 | Line 675 | constructed. The most famous example is the Verlet-lea
675   A lot of well established and very effective numerical methods have
676   been successful precisely because of their symplecticities even
677   though this fact was not recognized when they were first
678 < constructed. The most famous example is the Verlet-leapfrog methods
678 > constructed. The most famous example is the Verlet-leapfrog method
679   in molecular dynamics. In general, symplectic integrators can be
680   constructed using one of four different methods.
681   \begin{enumerate}
# Line 752 | Line 754 | to its symmetric property,
754   \label{introEquation:timeReversible}
755   \end{equation},appendixFig:architecture
756  
757 < \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Example of Splitting Method}}
757 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
758   The classical equation for a system consisting of interacting
759   particles can be written in Hamiltonian form,
760   \[
761   H = T + V
762   \]
763   where $T$ is the kinetic energy and $V$ is the potential energy.
764 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
764 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
765   obtains the following:
766   \begin{align}
767   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 786 | Line 788 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
788      \label{introEquation:Lp9b}\\%
789   %
790   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
791 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
791 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
792   \end{align}
793   From the preceding splitting, one can see that the integration of
794   the equations of motion would follow:
# Line 795 | Line 797 | the equations of motion would follow:
797  
798   \item Use the half step velocities to move positions one whole step, $\Delta t$.
799  
800 < \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
800 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
801  
802   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
803   \end{enumerate}
804  
805 < Simply switching the order of splitting and composing, a new
806 < integrator, the \emph{position verlet} integrator, can be generated,
805 > By simply switching the order of the propagators in the splitting
806 > and composing a new integrator, the \emph{position verlet}
807 > integrator, can be generated,
808   \begin{align}
809   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
810   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 814 | Line 817 | Baker-Campbell-Hausdorff formula can be used to determ
817  
818   \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
819  
820 < Baker-Campbell-Hausdorff formula can be used to determine the local
821 < error of splitting method in terms of commutator of the
820 > The Baker-Campbell-Hausdorff formula can be used to determine the
821 > local error of splitting method in terms of the commutator of the
822   operators(\ref{introEquation:exponentialOperator}) associated with
823 < the sub-flow. For operators $hX$ and $hY$ which are associate to
823 > the sub-flow. For operators $hX$ and $hY$ which are associated with
824   $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
825   \begin{equation}
826   \exp (hX + hY) = \exp (hZ)
# Line 831 | Line 834 | Applying Baker-Campbell-Hausdorff formula\cite{Varadar
834   \[
835   [X,Y] = XY - YX .
836   \]
837 < Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
838 < Sprang splitting, we can obtain
837 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
838 > to the Sprang splitting, we can obtain
839   \begin{eqnarray*}
840   \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 \\
841                                     &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
842                                     &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
843   \end{eqnarray*}
844 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
844 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
845   error of Spring splitting is proportional to $h^3$. The same
846 < procedure can be applied to general splitting,  of the form
846 > procedure can be applied to a general splitting,  of the form
847   \begin{equation}
848   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
849   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
850   \end{equation}
851 < Careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
852 < order method. Yoshida proposed an elegant way to compose higher
851 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
852 > order methods. Yoshida proposed an elegant way to compose higher
853   order methods based on symmetric splitting\cite{Yoshida1990}. Given
854   a symmetric second order base method $ \varphi _h^{(2)} $, a
855   fourth-order symmetric method can be constructed by composing,
# Line 859 | Line 862 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
862   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
863   \begin{equation}
864   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
865 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
865 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
866   \end{equation}
867 < , if the weights are chosen as
867 > if the weights are chosen as
868   \[
869   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
870   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 899 | Line 902 | will discusses issues in production run.
902   These three individual steps will be covered in the following
903   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
904   initialization of a simulation. Sec.~\ref{introSection:production}
905 < will discusses issues in production run.
905 > will discusse issues in production run.
906   Sec.~\ref{introSection:Analysis} provides the theoretical tools for
907   trajectory analysis.
908  
# Line 912 | Line 915 | purification and crystallization. Even for the molecul
915   databases, such as RCSB Protein Data Bank \textit{etc}. Although
916   thousands of crystal structures of molecules are discovered every
917   year, many more remain unknown due to the difficulties of
918 < purification and crystallization. Even for the molecule with known
919 < structure, some important information is missing. For example, the
918 > purification and crystallization. Even for molecules with known
919 > structure, some important information is missing. For example, a
920   missing hydrogen atom which acts as donor in hydrogen bonding must
921   be added. Moreover, in order to include electrostatic interaction,
922   one may need to specify the partial charges for individual atoms.
923   Under some circumstances, we may even need to prepare the system in
924 < a special setup. For instance, when studying transport phenomenon in
925 < membrane system, we may prepare the lipids in bilayer structure
926 < instead of placing lipids randomly in solvent, since we are not
927 < interested in self-aggregation and it takes a long time to happen.
924 > a special configuration. For instance, when studying transport
925 > phenomenon in membrane systems, we may prepare the lipids in a
926 > bilayer structure instead of placing lipids randomly in solvent,
927 > since we are not interested in the slow self-aggregation process.
928  
929   \subsubsection{\textbf{Minimization}}
930  
931   It is quite possible that some of molecules in the system from
932 < preliminary preparation may be overlapped with each other. This
933 < close proximity leads to high potential energy which consequently
934 < jeopardizes any molecular dynamics simulations. To remove these
935 < steric overlaps, one typically performs energy minimization to find
936 < a more reasonable conformation. Several energy minimization methods
937 < have been developed to exploit the energy surface and to locate the
938 < local minimum. While converging slowly near the minimum, steepest
939 < descent method is extremely robust when systems are far from
940 < harmonic. Thus, it is often used to refine structure from
941 < crystallographic data. Relied on the gradient or hessian, advanced
942 < methods like conjugate gradient and Newton-Raphson converge rapidly
943 < to a local minimum, while become unstable if the energy surface is
944 < far from quadratic. Another factor must be taken into account, when
932 > preliminary preparation may be overlapping with each other. This
933 > close proximity leads to high initial potential energy which
934 > consequently jeopardizes any molecular dynamics simulations. To
935 > remove these steric overlaps, one typically performs energy
936 > minimization to find a more reasonable conformation. Several energy
937 > minimization methods have been developed to exploit the energy
938 > surface and to locate the local minimum. While converging slowly
939 > near the minimum, steepest descent method is extremely robust when
940 > systems are strongly anharmonic. Thus, it is often used to refine
941 > structure from crystallographic data. Relied on the gradient or
942 > hessian, advanced methods like Newton-Raphson converge rapidly to a
943 > local minimum, but become unstable if the energy surface is far from
944 > quadratic. Another factor that must be taken into account, when
945   choosing energy minimization method, is the size of the system.
946   Steepest descent and conjugate gradient can deal with models of any
947 < size. Because of the limit of computation power to calculate hessian
948 < matrix and insufficient storage capacity to store them, most
949 < Newton-Raphson methods can not be used with very large models.
947 > size. Because of the limits on computer memory to store the hessian
948 > matrix and the computing power needed to diagonalized these
949 > matrices, most Newton-Raphson methods can not be used with very
950 > large systems.
951  
952   \subsubsection{\textbf{Heating}}
953  
954   Typically, Heating is performed by assigning random velocities
955 < according to a Gaussian distribution for a temperature. Beginning at
956 < a lower temperature and gradually increasing the temperature by
957 < assigning greater random velocities, we end up with setting the
958 < temperature of the system to a final temperature at which the
959 < simulation will be conducted. In heating phase, we should also keep
960 < the system from drifting or rotating as a whole. Equivalently, the
961 < net linear momentum and angular momentum of the system should be
962 < shifted to zero.
955 > according to a Maxwell-Boltzman distribution for a desired
956 > temperature. Beginning at a lower temperature and gradually
957 > increasing the temperature by assigning larger random velocities, we
958 > end up with setting the temperature of the system to a final
959 > temperature at which the simulation will be conducted. In heating
960 > phase, we should also keep the system from drifting or rotating as a
961 > whole. To do this, the net linear momentum and angular momentum of
962 > the system is shifted to zero after each resampling from the Maxwell
963 > -Boltzman distribution.
964  
965   \subsubsection{\textbf{Equilibration}}
966  
# Line 971 | Line 976 | Production run is the most important step of the simul
976  
977   \subsection{\label{introSection:production}Production}
978  
979 < Production run is the most important step of the simulation, in
979 > The production run is the most important step of the simulation, in
980   which the equilibrated structure is used as a starting point and the
981   motions of the molecules are collected for later analysis. In order
982   to capture the macroscopic properties of the system, the molecular
983 < dynamics simulation must be performed in correct and efficient way.
983 > dynamics simulation must be performed by sampling correctly and
984 > efficiently from the relevant thermodynamic ensemble.
985  
986   The most expensive part of a molecular dynamics simulation is the
987   calculation of non-bonded forces, such as van der Waals force and
988   Coulombic forces \textit{etc}. For a system of $N$ particles, the
989   complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
990   which making large simulations prohibitive in the absence of any
991 < computation saving techniques.
991 > algorithmic tricks.
992  
993 < A natural approach to avoid system size issue is to represent the
993 > A natural approach to avoid system size issues is to represent the
994   bulk behavior by a finite number of the particles. However, this
995 < approach will suffer from the surface effect. To offset this,
996 < \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
997 < is developed to simulate bulk properties with a relatively small
998 < number of particles. In this method, the simulation box is
999 < replicated throughout space to form an infinite lattice. During the
1000 < simulation, when a particle moves in the primary cell, its image in
1001 < other cells move in exactly the same direction with exactly the same
1002 < orientation. Thus, as a particle leaves the primary cell, one of its
1003 < images will enter through the opposite face.
995 > approach will suffer from the surface effect at the edges of the
996 > simulation. To offset this, \textit{Periodic boundary conditions}
997 > (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
998 > properties with a relatively small number of particles. In this
999 > method, the simulation box is replicated throughout space to form an
1000 > infinite lattice. During the simulation, when a particle moves in
1001 > the primary cell, its image in other cells move in exactly the same
1002 > direction with exactly the same orientation. Thus, as a particle
1003 > leaves the primary cell, one of its images will enter through the
1004 > opposite face.
1005   \begin{figure}
1006   \centering
1007   \includegraphics[width=\linewidth]{pbc.eps}
# Line 1006 | Line 1013 | evaluation is to apply cutoff where particles farther
1013  
1014   %cutoff and minimum image convention
1015   Another important technique to improve the efficiency of force
1016 < evaluation is to apply cutoff where particles farther than a
1017 < predetermined distance, are not included in the calculation
1016 > evaluation is to apply spherical cutoff where particles farther than
1017 > a predetermined distance are not included in the calculation
1018   \cite{Frenkel1996}. The use of a cutoff radius will cause a
1019   discontinuity in the potential energy curve. Fortunately, one can
1020 < shift the potential to ensure the potential curve go smoothly to
1021 < zero at the cutoff radius. Cutoff strategy works pretty well for
1022 < Lennard-Jones interaction because of its short range nature.
1023 < However, simply truncating the electrostatic interaction with the
1024 < use of cutoff has been shown to lead to severe artifacts in
1025 < simulations. Ewald summation, in which the slowly conditionally
1026 < convergent Coulomb potential is transformed into direct and
1027 < reciprocal sums with rapid and absolute convergence, has proved to
1028 < minimize the periodicity artifacts in liquid simulations. Taking the
1029 < advantages of the fast Fourier transform (FFT) for calculating
1030 < discrete Fourier transforms, the particle mesh-based
1020 > shift simple radial potential to ensure the potential curve go
1021 > smoothly to zero at the cutoff radius. The cutoff strategy works
1022 > well for Lennard-Jones interaction because of its short range
1023 > nature. However, simply truncating the electrostatic interaction
1024 > with the use of cutoffs has been shown to lead to severe artifacts
1025 > in simulations. The Ewald summation, in which the slowly decaying
1026 > Coulomb potential is transformed into direct and reciprocal sums
1027 > with rapid and absolute convergence, has proved to minimize the
1028 > periodicity artifacts in liquid simulations. Taking the advantages
1029 > of the fast Fourier transform (FFT) for calculating discrete Fourier
1030 > transforms, the particle mesh-based
1031   methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1032 < $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1033 < multipole method}\cite{Greengard1987, Greengard1994}, which treats
1034 < Coulombic interaction exactly at short range, and approximate the
1035 < potential at long range through multipolar expansion. In spite of
1036 < their wide acceptances at the molecular simulation community, these
1037 < two methods are hard to be implemented correctly and efficiently.
1038 < Instead, we use a damped and charge-neutralized Coulomb potential
1039 < method developed by Wolf and his coworkers\cite{Wolf1999}. The
1040 < shifted Coulomb potential for particle $i$ and particle $j$ at
1041 < distance $r_{rj}$ is given by:
1032 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
1033 > \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
1034 > which treats Coulombic interactions exactly at short range, and
1035 > approximate the potential at long range through multipolar
1036 > expansion. In spite of their wide acceptance at the molecular
1037 > simulation community, these two methods are difficult to implement
1038 > correctly and efficiently. Instead, we use a damped and
1039 > charge-neutralized Coulomb potential method developed by Wolf and
1040 > his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
1041 > particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1042   \begin{equation}
1043   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1044   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
# Line 1053 | Line 1060 | Recently, advanced visualization technique are widely
1060  
1061   \subsection{\label{introSection:Analysis} Analysis}
1062  
1063 < Recently, advanced visualization technique are widely applied to
1063 > Recently, advanced visualization technique have become applied to
1064   monitor the motions of molecules. Although the dynamics of the
1065   system can be described qualitatively from animation, quantitative
1066 < trajectory analysis are more appreciable. According to the
1067 < principles of Statistical Mechanics,
1068 < Sec.~\ref{introSection:statisticalMechanics}, one can compute
1069 < thermodynamics properties, analyze fluctuations of structural
1070 < parameters, and investigate time-dependent processes of the molecule
1064 < from the trajectories.
1066 > trajectory analysis are more useful. According to the principles of
1067 > Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
1068 > one can compute thermodynamic properties, analyze fluctuations of
1069 > structural parameters, and investigate time-dependent processes of
1070 > the molecule from the trajectories.
1071  
1072 < \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamics Properties}}
1072 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1073  
1074 < Thermodynamics properties, which can be expressed in terms of some
1074 > Thermodynamic properties, which can be expressed in terms of some
1075   function of the coordinates and momenta of all particles in the
1076   system, can be directly computed from molecular dynamics. The usual
1077   way to measure the pressure is based on virial theorem of Clausius
# Line 1088 | Line 1094 | distribution functions. Among these functions,\emph{pa
1094   \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1095  
1096   Structural Properties of a simple fluid can be described by a set of
1097 < distribution functions. Among these functions,\emph{pair
1097 > distribution functions. Among these functions,the \emph{pair
1098   distribution function}, also known as \emph{radial distribution
1099 < function}, is of most fundamental importance to liquid-state theory.
1100 < Pair distribution function can be gathered by Fourier transforming
1101 < raw data from a series of neutron diffraction experiments and
1102 < integrating over the surface factor \cite{Powles1973}. The
1103 < experiment result can serve as a criterion to justify the
1104 < correctness of the theory. Moreover, various equilibrium
1105 < thermodynamic and structural properties can also be expressed in
1106 < terms of radial distribution function \cite{Allen1987}.
1099 > function}, is of most fundamental importance to liquid theory.
1100 > Experimentally, pair distribution function can be gathered by
1101 > Fourier transforming raw data from a series of neutron diffraction
1102 > experiments and integrating over the surface factor
1103 > \cite{Powles1973}. The experimental results can serve as a criterion
1104 > to justify the correctness of a liquid model. Moreover, various
1105 > equilibrium thermodynamic and structural properties can also be
1106 > expressed in terms of radial distribution function \cite{Allen1987}.
1107  
1108 < A pair distribution functions $g(r)$ gives the probability that a
1108 > The pair distribution functions $g(r)$ gives the probability that a
1109   particle $i$ will be located at a distance $r$ from a another
1110   particle $j$ in the system
1111   \[
1112   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1113 < \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1113 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1114 > (r)}{\rho}.
1115   \]
1116   Note that the delta function can be replaced by a histogram in
1117 < computer simulation. Figure
1118 < \ref{introFigure:pairDistributionFunction} shows a typical pair
1119 < 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.
1117 > computer simulation. Peaks in $g(r)$ represent solvent shells, and
1118 > the height of these peaks gradually decreases to 1 as the liquid of
1119 > large distance approaches the bulk density.
1120  
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}
1121  
1122   \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1123   Properties}}
1124  
1125   Time-dependent properties are usually calculated using \emph{time
1126 < correlation function}, which correlates random variables $A$ and $B$
1127 < at two different time
1126 > correlation functions}, which correlate random variables $A$ and $B$
1127 > at two different times,
1128   \begin{equation}
1129   C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1130   \label{introEquation:timeCorrelationFunction}
1131   \end{equation}
1132   If $A$ and $B$ refer to same variable, this kind of correlation
1133 < function is called \emph{auto correlation function}. One example of
1134 < auto correlation function is velocity auto-correlation function
1135 < which is directly related to transport properties of molecular
1136 < liquids:
1133 > function is called an \emph{autocorrelation function}. One example
1134 > of an auto correlation function is the velocity auto-correlation
1135 > function which is directly related to transport properties of
1136 > molecular liquids:
1137   \[
1138   D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1139   \right\rangle } dt
1140   \]
1141 < where $D$ is diffusion constant. Unlike velocity autocorrelation
1142 < function which is averaging over time origins and over all the
1143 < atoms, dipole autocorrelation are calculated for the entire system.
1144 < The dipole autocorrelation function is given by:
1141 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1142 > function, which is averaging over time origins and over all the
1143 > atoms, the dipole autocorrelation functions are calculated for the
1144 > entire system. The dipole autocorrelation function is given by:
1145   \[
1146   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1147   \right\rangle
# Line 1173 | Line 1167 | simulator is governed by the rigid body dynamics. In m
1167   areas, from engineering, physics, to chemistry. For example,
1168   missiles and vehicle are usually modeled by rigid bodies.  The
1169   movement of the objects in 3D gaming engine or other physics
1170 < simulator is governed by the rigid body dynamics. In molecular
1171 < simulation, rigid body is used to simplify the model in
1172 < protein-protein docking study\cite{Gray2003}.
1170 > simulator is governed by rigid body dynamics. In molecular
1171 > simulations, rigid bodies are used to simplify protein-protein
1172 > docking studies\cite{Gray2003}.
1173  
1174   It is very important to develop stable and efficient methods to
1175 < integrate the equations of motion of orientational degrees of
1176 < freedom. Euler angles are the nature choice to describe the
1177 < rotational degrees of freedom. However, due to its singularity, the
1178 < numerical integration of corresponding equations of motion is very
1179 < inefficient and inaccurate. Although an alternative integrator using
1180 < different sets of Euler angles can overcome this
1181 < difficulty\cite{Barojas1973}, the computational penalty and the lost
1182 < of angular momentum conservation still remain. A singularity free
1183 < representation utilizing quaternions was developed by Evans in
1184 < 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1185 < nonseparable Hamiltonian resulted from quaternion representation,
1186 < which prevents the symplectic algorithm to be utilized. Another
1187 < different approach is to apply holonomic constraints to the atoms
1188 < belonging to the rigid body. Each atom moves independently under the
1189 < normal forces deriving from potential energy and constraint forces
1190 < which are used to guarantee the rigidness. However, due to their
1191 < iterative nature, SHAKE and Rattle algorithm converge very slowly
1192 < when the number of constraint increases\cite{Ryckaert1977,
1193 < Andersen1983}.
1175 > integrate the equations of motion for orientational degrees of
1176 > freedom. Euler angles are the natural choice to describe the
1177 > rotational degrees of freedom. However, due to $\frac {1}{sin
1178 > \theta}$ singularities, the numerical integration of corresponding
1179 > equations of motion is very inefficient and inaccurate. Although an
1180 > alternative integrator using multiple sets of Euler angles can
1181 > overcome this difficulty\cite{Barojas1973}, the computational
1182 > penalty and the loss of angular momentum conservation still remain.
1183 > A singularity-free representation utilizing quaternions was
1184 > developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1185 > approach uses a nonseparable Hamiltonian resulting from the
1186 > quaternion representation, which prevents the symplectic algorithm
1187 > to be utilized. Another different approach is to apply holonomic
1188 > constraints to the atoms belonging to the rigid body. Each atom
1189 > moves independently under the normal forces deriving from potential
1190 > energy and constraint forces which are used to guarantee the
1191 > rigidness. However, due to their iterative nature, the SHAKE and
1192 > Rattle algorithms also converge very slowly when the number of
1193 > constraints increases\cite{Ryckaert1977, Andersen1983}.
1194  
1195 < The break through in geometric literature suggests that, in order to
1195 > A break-through in geometric literature suggests that, in order to
1196   develop a long-term integration scheme, one should preserve the
1197 < symplectic structure of the flow. Introducing conjugate momentum to
1198 < rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1199 < symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1200 < the Hamiltonian system in a constraint manifold by iteratively
1201 < satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1202 < method using quaternion representation was developed by
1203 < Omelyan\cite{Omelyan1998}. However, both of these methods are
1204 < iterative and inefficient. In this section, we will present a
1197 > symplectic structure of the flow. By introducing a conjugate
1198 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1199 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1200 > proposed to evolve the Hamiltonian system in a constraint manifold
1201 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1202 > An alternative method using the quaternion representation was
1203 > developed by Omelyan\cite{Omelyan1998}. However, both of these
1204 > methods are iterative and inefficient. In this section, we descibe a
1205   symplectic Lie-Poisson integrator for rigid body developed by
1206   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1207  
1208 < \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1209 < The motion of the rigid body is Hamiltonian with the Hamiltonian
1208 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1209 > The motion of a rigid body is Hamiltonian with the Hamiltonian
1210   function
1211   \begin{equation}
1212   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
# Line 1226 | Line 1220 | constrained Hamiltonian equation subjects to a holonom
1220   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1221   \]
1222   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1223 < constrained Hamiltonian equation subjects to a holonomic constraint,
1223 > constrained Hamiltonian equation is subjected to a holonomic
1224 > constraint,
1225   \begin{equation}
1226   Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1227   \end{equation}
1228 < which is used to ensure rotation matrix's orthogonality.
1229 < Differentiating \ref{introEquation:orthogonalConstraint} and using
1230 < Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1228 > which is used to ensure rotation matrix's unitarity. Differentiating
1229 > \ref{introEquation:orthogonalConstraint} and using Equation
1230 > \ref{introEquation:RBMotionMomentum}, one may obtain,
1231   \begin{equation}
1232   Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1233   \label{introEquation:RBFirstOrderConstraint}
# Line 1250 | Line 1245 | We can use constraint force provided by lagrange multi
1245   \end{eqnarray}
1246  
1247   In general, there are two ways to satisfy the holonomic constraints.
1248 < We can use constraint force provided by lagrange multiplier on the
1249 < normal manifold to keep the motion on constraint space. Or we can
1250 < simply evolve the system in constraint manifold. These two methods
1251 < are proved to be equivalent. The holonomic constraint and equations
1252 < of motions define a constraint manifold for rigid body
1248 > We can use a constraint force provided by a Lagrange multiplier on
1249 > the normal manifold to keep the motion on constraint space. Or we
1250 > can simply evolve the system on the constraint manifold. These two
1251 > methods have been proved to be equivalent. The holonomic constraint
1252 > and equations of motions define a constraint manifold for rigid
1253 > bodies
1254   \[
1255   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1256   \right\}.
# Line 1263 | Line 1259 | diffeomorphic. Introducing
1259   Unfortunately, this constraint manifold is not the cotangent bundle
1260   $T_{\star}SO(3)$. However, it turns out that under symplectic
1261   transformation, the cotangent space and the phase space are
1262 < diffeomorphic. Introducing
1262 > diffeomorphic. By introducing
1263   \[
1264   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1265   \]
# Line 1295 | Line 1291 | body, angular momentum on body frame $\Pi  = Q^t P$ is
1291   respectively.
1292  
1293   As a common choice to describe the rotation dynamics of the rigid
1294 < body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1295 < rewrite the equations of motion,
1294 > body, the angular momentum on the body fixed frame $\Pi  = Q^t P$ is
1295 > introduced to rewrite the equations of motion,
1296   \begin{equation}
1297   \begin{array}{l}
1298   \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
# Line 1339 | Line 1335 | unique property eliminate the requirement of iteration
1335   Since $\Lambda$ is symmetric, the last term of Equation
1336   \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1337   multiplier $\Lambda$ is absent from the equations of motion. This
1338 < unique property eliminate the requirement of iterations which can
1338 > unique property eliminates the requirement of iterations which can
1339   not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1340  
1341 < Applying hat-map isomorphism, we obtain the equation of motion for
1342 < angular momentum on body frame
1341 > Applying the hat-map isomorphism, we obtain the equation of motion
1342 > for angular momentum on body frame
1343   \begin{equation}
1344   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1345   F_i (r,Q)} \right) \times X_i }.
# Line 1358 | Line 1354 | If there is not external forces exerted on the rigid b
1354   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1355   Lie-Poisson Integrator for Free Rigid Body}
1356  
1357 < If there is not external forces exerted on the rigid body, the only
1358 < contribution to the rotational is from the kinetic potential (the
1359 < first term of \ref{introEquation:bodyAngularMotion}). The free rigid
1360 < body is an example of Lie-Poisson system with Hamiltonian function
1357 > If there are no external forces exerted on the rigid body, the only
1358 > contribution to the rotational motion is from the kinetic energy
1359 > (the first term of \ref{introEquation:bodyAngularMotion}). The free
1360 > rigid body is an example of a Lie-Poisson system with Hamiltonian
1361 > function
1362   \begin{equation}
1363   T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1364   \label{introEquation:rotationalKineticRB}
# Line 1408 | Line 1405 | tR_1 }$, we can use Cayley transformation,
1405   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1406   \]
1407   To reduce the cost of computing expensive functions in $e^{\Delta
1408 < tR_1 }$, we can use Cayley transformation,
1408 > tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1409 > propagator,
1410   \[
1411   e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1412   )
1413   \]
1414   The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1415 < manner.
1416 <
1419 < In order to construct a second-order symplectic method, we split the
1420 < angular kinetic Hamiltonian function can into five terms
1415 > manner. In order to construct a second-order symplectic method, we
1416 > split the angular kinetic Hamiltonian function can into five terms
1417   \[
1418   T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1419   ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1420 < (\pi _1 )
1421 < \].
1422 < Concatenating flows corresponding to these five terms, we can obtain
1423 < an symplectic integrator,
1420 > (\pi _1 ).
1421 > \]
1422 > By concatenating the propagators corresponding to these five terms,
1423 > we can obtain an symplectic integrator,
1424   \[
1425   \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1426   \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
# Line 1451 | Line 1447 | Lie-Poisson integrator is found to be extremely effici
1447   \]
1448   Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1449   \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1450 < Lie-Poisson integrator is found to be extremely efficient and stable
1451 < which can be explained by the fact the small angle approximation is
1452 < used and the norm of the angular momentum is conserved.
1450 > Lie-Poisson integrator is found to be both extremely efficient and
1451 > stable. These properties can be explained by the fact the small
1452 > angle approximation is used and the norm of the angular momentum is
1453 > conserved.
1454  
1455   \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1456   Splitting for Rigid Body}
# Line 1480 | Line 1477 | A second-order symplectic method is now obtained by th
1477   \end{tabular}
1478   \end{center}
1479   \end{table}
1480 < A second-order symplectic method is now obtained by the
1481 < composition of the flow maps,
1480 > A second-order symplectic method is now obtained by the composition
1481 > of the position and velocity propagators,
1482   \[
1483   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1484   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1485   \]
1486   Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1487 < sub-flows which corresponding to force and torque respectively,
1487 > sub-propagators which corresponding to force and torque
1488 > respectively,
1489   \[
1490   \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1491   _{\Delta t/2,\tau }.
1492   \]
1493   Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1494 < $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1495 < order inside $\varphi _{\Delta t/2,V}$ does not matter.
1496 <
1497 < Furthermore, kinetic potential can be separated to translational
1500 < kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1494 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1495 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1496 > kinetic energy can be separated to translational kinetic term, $T^t
1497 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1498   \begin{equation}
1499   T(p,\pi ) =T^t (p) + T^r (\pi ).
1500   \end{equation}
1501   where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1502   defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1503 < corresponding flow maps are given by
1503 > corresponding propagators are given by
1504   \[
1505   \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1506   _{\Delta t,T^r }.
1507   \]
1508 < Finally, we obtain the overall symplectic flow maps for free moving
1509 < rigid body
1508 > Finally, we obtain the overall symplectic propagators for freely
1509 > moving rigid bodies
1510   \begin{equation}
1511   \begin{array}{c}
1512   \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
# Line 1523 | Line 1520 | the theory of Langevin dynamics simulation. A brief de
1520   As an alternative to newtonian dynamics, Langevin dynamics, which
1521   mimics a simple heat bath with stochastic and dissipative forces,
1522   has been applied in a variety of studies. This section will review
1523 < the theory of Langevin dynamics simulation. A brief derivation of
1524 < generalized Langevin equation will be given first. Follow that, we
1525 < will discuss the physical meaning of the terms appearing in the
1526 < equation as well as the calculation of friction tensor from
1527 < hydrodynamics theory.
1523 > the theory of Langevin dynamics. A brief derivation of generalized
1524 > Langevin equation will be given first. Following that, we will
1525 > discuss the physical meaning of the terms appearing in the equation
1526 > as well as the calculation of friction tensor from hydrodynamics
1527 > theory.
1528  
1529   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1530  
1531 < Harmonic bath model, in which an effective set of harmonic
1531 > A harmonic bath model, in which an effective set of harmonic
1532   oscillators are used to mimic the effect of a linearly responding
1533   environment, has been widely used in quantum chemistry and
1534   statistical mechanics. One of the successful applications of
1535 < Harmonic bath model is the derivation of Deriving Generalized
1536 < Langevin Dynamics. Lets consider a system, in which the degree of
1535 > Harmonic bath model is the derivation of the Generalized Langevin
1536 > Dynamics (GLE). Lets consider a system, in which the degree of
1537   freedom $x$ is assumed to couple to the bath linearly, giving a
1538   Hamiltonian of the form
1539   \begin{equation}
1540   H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1541   \label{introEquation:bathGLE}.
1542   \end{equation}
1543 < Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1544 < with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1543 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1544 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1545   \[
1546   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1547   }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
# Line 1552 | Line 1549 | the harmonic bath masses, and $\Delta U$ is bilinear s
1549   \]
1550   where the index $\alpha$ runs over all the bath degrees of freedom,
1551   $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1552 < the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1552 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1553   coupling,
1554   \[
1555   \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1556   \]
1557 < where $g_\alpha$ are the coupling constants between the bath and the
1558 < coordinate $x$. Introducing
1557 > where $g_\alpha$ are the coupling constants between the bath
1558 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1559 > Introducing
1560   \[
1561   W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1562   }}{{2m_\alpha  w_\alpha ^2 }}} x^2
# Line 1573 | Line 1571 | Generalized Langevin Dynamics by Hamilton's equations
1571   \]
1572   Since the first two terms of the new Hamiltonian depend only on the
1573   system coordinates, we can get the equations of motion for
1574 < Generalized Langevin Dynamics by Hamilton's equations
1577 < \ref{introEquation:motionHamiltonianCoordinate,
1578 < introEquation:motionHamiltonianMomentum},
1574 > Generalized Langevin Dynamics by Hamilton's equations,
1575   \begin{equation}
1576   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1577   \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
# Line 1592 | Line 1588 | differential equations, Laplace transform is the appro
1588   In order to derive an equation for $x$, the dynamics of the bath
1589   variables $x_\alpha$ must be solved exactly first. As an integral
1590   transform which is particularly useful in solving linear ordinary
1591 < differential equations, Laplace transform is the appropriate tool to
1592 < solve this problem. The basic idea is to transform the difficult
1591 > differential equations,the Laplace transform is the appropriate tool
1592 > to solve this problem. The basic idea is to transform the difficult
1593   differential equations into simple algebra problems which can be
1594 < solved easily. Then applying inverse Laplace transform, also known
1595 < as the Bromwich integral, we can retrieve the solutions of the
1594 > solved easily. Then, by applying the inverse Laplace transform, also
1595 > known as the Bromwich integral, we can retrieve the solutions of the
1596   original problems.
1597  
1598   Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
# Line 1616 | Line 1612 | Applying Laplace transform to the bath coordinates, we
1612   \end{eqnarray*}
1613  
1614  
1615 < Applying Laplace transform to the bath coordinates, we obtain
1615 > Applying the Laplace transform to the bath coordinates, we obtain
1616   \begin{eqnarray*}
1617   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) \\
1618   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 1689 | as the model, which is gaussian distribution in genera
1689   \end{array}
1690   \]
1691   This property is what we expect from a truly random process. As long
1692 < as the model, which is gaussian distribution in general, chosen for
1693 < $R(t)$ is a truly random process, the stochastic nature of the GLE
1698 < still remains.
1692 > as the model chosen for $R(t)$ was a gaussian distribution in
1693 > general, the stochastic nature of the GLE still remains.
1694  
1695   %dynamic friction kernel
1696   The convolution integral
# Line 1716 | Line 1711 | which can be used to describe dynamic caging effect. T
1711   m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1712   \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1713   \]
1714 < which can be used to describe dynamic caging effect. The other
1715 < extreme is the bath that responds infinitely quickly to motions in
1716 < the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1717 < time:
1714 > which can be used to describe the effect of dynamic caging in
1715 > viscous solvents. The other extreme is the bath that responds
1716 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1717 > taken as a $delta$ function in time:
1718   \[
1719   \xi (t) = 2\xi _0 \delta (t)
1720   \]
# Line 1735 | Line 1730 | or be determined by Stokes' law for regular shaped par
1730   \end{equation}
1731   which is known as the Langevin equation. The static friction
1732   coefficient $\xi _0$ can either be calculated from spectral density
1733 < or be determined by Stokes' law for regular shaped particles.A
1733 > or be determined by Stokes' law for regular shaped particles. A
1734   briefly review on calculating friction tensor for arbitrary shaped
1735   particles is given in Sec.~\ref{introSection:frictionTensor}.
1736  
# Line 1768 | Line 1763 | can model the random force and friction kernel.
1763   \end{equation}
1764   In effect, it acts as a constraint on the possible ways in which one
1765   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