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 2938 by tim, Mon Jul 17 15:28:44 2006 UTC vs.
Revision 2941 by tim, Mon Jul 17 20:01:05 2006 UTC

# Line 67 | Line 67 | the quality of numerical integration schemes for rigid
67   \begin{equation}E = T + V. \label{introEquation:energyConservation}
68   \end{equation}
69   All of these conserved quantities are important factors to determine
70 < the quality of numerical integration schemes for rigid bodies
71 < \cite{Dullweber1997}.
70 > the quality of numerical integration schemes for rigid
71 > bodies.\cite{Dullweber1997}
72  
73   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
# Line 178 | Line 178 | known as the canonical equations of motions \cite{Gold
178   where Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
179   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
180   equation of motion. Due to their symmetrical formula, they are also
181 < known as the canonical equations of motions \cite{Goldstein2001}.
181 > known as the canonical equations of motions.\cite{Goldstein2001}
182  
183   An important difference between Lagrangian approach and the
184   Hamiltonian approach is that the Lagrangian is considered to be a
# Line 188 | Line 188 | only works with 1st-order differential equations\cite{
188   Hamiltonian Mechanics is more appropriate for application to
189   statistical mechanics and quantum mechanics, since it treats the
190   coordinate and its time derivative as independent variables and it
191 < only works with 1st-order differential equations\cite{Marion1990}.
191 > only works with 1st-order differential equations.\cite{Marion1990}
192   In Newtonian Mechanics, a system described by conservative forces
193   conserves the total energy
194   (Eq.~\ref{introEquation:energyConservation}). It follows that
# Line 416 | Line 416 | statistical ensemble are identical \cite{Frenkel1996,
416   many-body system in Statistical Mechanics. Fortunately, the Ergodic
417   Hypothesis makes a connection between time average and the ensemble
418   average. It states that the time average and average over the
419 < statistical ensemble are identical \cite{Frenkel1996, Leach2001}:
419 > statistical ensemble are identical:\cite{Frenkel1996, Leach2001}
420   \begin{equation}
421   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
422   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 434 | Line 434 | choice\cite{Frenkel1996}.
434   utilized. Or if the system lends itself to a time averaging
435   approach, the Molecular Dynamics techniques in
436   Sec.~\ref{introSection:molecularDynamics} will be the best
437 < choice\cite{Frenkel1996}.
437 > choice.\cite{Frenkel1996}
438  
439   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
440   A variety of numerical integrators have been proposed to simulate
441   the motions of atoms in MD simulation. They usually begin with
442 < initial conditions and move the objects in the direction governed
443 < by the differential equations. However, most of them ignore the
444 < hidden physical laws contained within the equations. Since 1990,
445 < geometric integrators, which preserve various phase-flow invariants
446 < such as symplectic structure, volume and time reversal symmetry,
447 < were developed to address this issue\cite{Dullweber1997,
448 < McLachlan1998, Leimkuhler1999}. The velocity Verlet method, which
449 < happens to be a simple example of symplectic integrator, continues
450 < to gain popularity in the molecular dynamics community. This fact
451 < can be partly explained by its geometric nature.
442 > initial conditions and move the objects in the direction governed by
443 > the differential equations. However, most of them ignore the hidden
444 > physical laws contained within the equations. Since 1990, geometric
445 > integrators, which preserve various phase-flow invariants such as
446 > symplectic structure, volume and time reversal symmetry, were
447 > developed to address this issue.\cite{Dullweber1997, McLachlan1998,
448 > Leimkuhler1999} The velocity Verlet method, which happens to be a
449 > simple example of symplectic integrator, continues to gain
450 > popularity in the molecular dynamics community. This fact can be
451 > partly explained by its geometric nature.
452  
453   \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
454   A \emph{manifold} is an abstract mathematical space. It looks
# Line 457 | Line 457 | apply calculus\cite{Hirsch1997}. A \emph{symplectic ma
457   surface of Earth. It seems to be flat locally, but it is round if
458   viewed as a whole. A \emph{differentiable manifold} (also known as
459   \emph{smooth manifold}) is a manifold on which it is possible to
460 < apply calculus\cite{Hirsch1997}. A \emph{symplectic manifold} is
460 > apply calculus.\cite{Hirsch1997} A \emph{symplectic manifold} is
461   defined as a pair $(M, \omega)$ which consists of a
462   \emph{differentiable manifold} $M$ and a close, non-degenerate,
463   bilinear symplectic form, $\omega$. A symplectic form on a vector
464   space $V$ is a function $\omega(x, y)$ which satisfies
465   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
466   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
467 < $\omega(x, x) = 0$\cite{McDuff1998}. The cross product operation in
467 > $\omega(x, x) = 0$.\cite{McDuff1998} The cross product operation in
468   vector field is an example of symplectic form. One of the
469   motivations to study \emph{symplectic manifolds} in Hamiltonian
470   Mechanics is that a symplectic manifold can represent all possible
471   configurations of the system and the phase space of the system can
472 < be described by it's cotangent bundle\cite{Jost2002}. Every
472 > be described by it's cotangent bundle.\cite{Jost2002} Every
473   symplectic manifold is even dimensional. For instance, in Hamilton
474   equations, coordinate and momentum always appear in pairs.
475  
# Line 496 | Line 496 | Hamiltonian dynamics is Poisson Dynamics\cite{Olver198
496   \label{introEquation:compactHamiltonian}
497   \end{equation}In this case, $f$ is
498   called a \emph{Hamiltonian vector field}. Another generalization of
499 < Hamiltonian dynamics is Poisson Dynamics\cite{Olver1986},
499 > Hamiltonian dynamics is Poisson Dynamics,\cite{Olver1986}
500   \begin{equation}
501   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
502   \end{equation}
503 < The most obvious change being that matrix $J$ now depends on $x$.
503 > where the most obvious change being that matrix $J$ now depends on
504 > $x$.
505  
506   \subsection{\label{introSection:exactFlow}Exact Propagator}
507  
# Line 620 | Line 621 | accurately\cite{Kane2000}. Since they are geometricall
621   Generating functions\cite{Channell1990} tend to lead to methods
622   which are cumbersome and difficult to use. In dissipative systems,
623   variational methods can capture the decay of energy
624 < accurately\cite{Kane2000}. Since they are geometrically unstable
624 > accurately.\cite{Kane2000} Since they are geometrically unstable
625   against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
626   methods are not suitable for Hamiltonian system. Recently, various
627   high-order explicit Runge-Kutta methods \cite{Owren1992,Chen2003}
# Line 629 | Line 630 | system\cite{Tuckerman1992, McLachlan1998}.
630   methods, they have not attracted much attention from the Molecular
631   Dynamics community. Instead, splitting methods have been widely
632   accepted since they exploit natural decompositions of the
633 < system\cite{Tuckerman1992, McLachlan1998}.
633 > system.\cite{McLachlan1998, Tuckerman1992}
634  
635   \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
636  
# Line 673 | Line 674 | a second-order decomposition,
674   The Lie-Trotter
675   splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
676   local errors proportional to $h^2$, while the Strang splitting gives
677 < a second-order decomposition,
677 > a second-order decomposition,\cite{Strang1968}
678   \begin{equation}
679   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
680   _{1,h/2} , \label{introEquation:secondOrderSplitting}
# Line 748 | Line 749 | The Baker-Campbell-Hausdorff formula can be used to de
749  
750   \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
751  
752 < The Baker-Campbell-Hausdorff formula can be used to determine the
753 < local error of a splitting method in terms of the commutator of the
754 < operators(Eq.~\ref{introEquation:exponentialOperator}) associated with
755 < the sub-propagator. For operators $hX$ and $hY$ which are associated
756 < with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
752 > The Baker-Campbell-Hausdorff formula\cite{Gilmore1974} can be used
753 > to determine the local error of a splitting method in terms of the
754 > commutator of the
755 > operators(Eq.~\ref{introEquation:exponentialOperator}) associated
756 > with the sub-propagator. For operators $hX$ and $hY$ which are
757 > associated with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we
758 > have
759   \begin{equation}
760   \exp (hX + hY) = \exp (hZ)
761   \end{equation}
# Line 782 | Line 785 | order methods based on symmetric splitting\cite{Yoshid
785   \end{equation}
786   A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
787   order methods. Yoshida proposed an elegant way to compose higher
788 < order methods based on symmetric splitting\cite{Yoshida1990}. Given
788 > order methods based on symmetric splitting.\cite{Yoshida1990} Given
789   a symmetric second order base method $ \varphi _h^{(2)} $, a
790   fourth-order symmetric method can be constructed by composing,
791   \[
# Line 943 | Line 946 | than a predetermined distance are not included in the
946   %cutoff and minimum image convention
947   Another important technique to improve the efficiency of force
948   evaluation is to apply spherical cutoffs where particles farther
949 < than a predetermined distance are not included in the calculation
950 < \cite{Frenkel1996}. The use of a cutoff radius will cause a
951 < discontinuity in the potential energy curve. Fortunately, one can
949 > than a predetermined distance are not included in the
950 > calculation.\cite{Frenkel1996} The use of a cutoff radius will cause
951 > a discontinuity in the potential energy curve. Fortunately, one can
952   shift a simple radial potential to ensure the potential curve go
953   smoothly to zero at the cutoff radius. The cutoff strategy works
954   well for Lennard-Jones interaction because of its short range
# Line 954 | Line 957 | periodicity artifacts in liquid simulations. Taking ad
957   in simulations. The Ewald summation, in which the slowly decaying
958   Coulomb potential is transformed into direct and reciprocal sums
959   with rapid and absolute convergence, has proved to minimize the
960 < periodicity artifacts in liquid simulations. Taking advantage
961 < of fast Fourier transform (FFT) techniques for calculating discrete Fourier
962 < transforms, the particle mesh-based
960 > periodicity artifacts in liquid simulations. Taking advantage of
961 > fast Fourier transform (FFT) techniques for calculating discrete
962 > Fourier transforms, the particle mesh-based
963   methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
964   $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
965   \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
# Line 966 | Line 969 | his coworkers\cite{Wolf1999}. The shifted Coulomb pote
969   simulation community, these two methods are difficult to implement
970   correctly and efficiently. Instead, we use a damped and
971   charge-neutralized Coulomb potential method developed by Wolf and
972 < his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
972 > his coworkers.\cite{Wolf1999} The shifted Coulomb potential for
973   particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
974   \begin{equation}
975   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
# Line 1029 | Line 1032 | experiments and integrating over the surface factor
1032   function}, is of most fundamental importance to liquid theory.
1033   Experimentally, pair distribution functions can be gathered by
1034   Fourier transforming raw data from a series of neutron diffraction
1035 < experiments and integrating over the surface factor
1036 < \cite{Powles1973}. The experimental results can serve as a criterion
1037 < to justify the correctness of a liquid model. Moreover, various
1038 < equilibrium thermodynamic and structural properties can also be
1039 < expressed in terms of the radial distribution function
1040 < \cite{Allen1987}. The pair distribution functions $g(r)$ gives the
1041 < probability that a particle $i$ will be located at a distance $r$
1042 < from a another particle $j$ in the system
1035 > experiments and integrating over the surface
1036 > factor.\cite{Powles1973} The experimental results can serve as a
1037 > criterion to justify the correctness of a liquid model. Moreover,
1038 > various equilibrium thermodynamic and structural properties can also
1039 > be expressed in terms of the radial distribution
1040 > function.\cite{Allen1987} The pair distribution functions $g(r)$
1041 > gives the probability that a particle $i$ will be located at a
1042 > distance $r$ from a another particle $j$ in the system
1043   \begin{equation}
1044   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1045   \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
# Line 1097 | Line 1100 | docking studies\cite{Gray2003}.
1100   movement of the objects in 3D gaming engines or other physics
1101   simulators is governed by rigid body dynamics. In molecular
1102   simulations, rigid bodies are used to simplify protein-protein
1103 < docking studies\cite{Gray2003}.
1103 > docking studies.\cite{Gray2003}
1104  
1105   It is very important to develop stable and efficient methods to
1106   integrate the equations of motion for orientational degrees of
# Line 1109 | Line 1112 | quaternions was developed by Evans in 1977\cite{Evans1
1112   angles can overcome this difficulty\cite{Barojas1973}, the
1113   computational penalty and the loss of angular momentum conservation
1114   still remain. A singularity-free representation utilizing
1115 < quaternions was developed by Evans in 1977\cite{Evans1977}.
1115 > quaternions was developed by Evans in 1977.\cite{Evans1977}
1116   Unfortunately, this approach used a nonseparable Hamiltonian
1117   resulting from the quaternion representation, which prevented the
1118   symplectic algorithm from being utilized. Another different approach
# Line 1118 | Line 1121 | number of constraints increases\cite{Ryckaert1977, And
1121   deriving from potential energy and constraint forces which are used
1122   to guarantee the rigidness. However, due to their iterative nature,
1123   the SHAKE and Rattle algorithms also converge very slowly when the
1124 < number of constraints increases\cite{Ryckaert1977, Andersen1983}.
1124 > number of constraints increases.\cite{Ryckaert1977, Andersen1983}
1125  
1126   A break-through in geometric literature suggests that, in order to
1127   develop a long-term integration scheme, one should preserve the
# Line 1128 | Line 1131 | developed by Omelyan\cite{Omelyan1998}. However, both
1131   proposed to evolve the Hamiltonian system in a constraint manifold
1132   by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1133   An alternative method using the quaternion representation was
1134 < developed by Omelyan\cite{Omelyan1998}. However, both of these
1134 > developed by Omelyan.\cite{Omelyan1998} However, both of these
1135   methods are iterative and inefficient. In this section, we descibe a
1136   symplectic Lie-Poisson integrator for rigid bodies developed by
1137   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1138  
1139   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1140 < The Hamiltonian of a rigid body is given by
1140 > The Hamiltonian of a rigid body is given by
1141   \begin{equation}
1142   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1143   V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
# Line 1248 | Line 1251 | iterations which can not be avoided in other methods\c
1251   Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1252   Lagrange multiplier $\Lambda$ is absent from the equations of
1253   motion. This unique property eliminates the requirement of
1254 < iterations which can not be avoided in other methods\cite{Kol1997,
1255 < Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1254 > iterations which can not be avoided in other methods.\cite{Kol1997,
1255 > Omelyan1998} Applying the hat-map isomorphism, we obtain the
1256   equation of motion for angular momentum in the body frame
1257   \begin{equation}
1258   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
# Line 1355 | Line 1358 | norm of the angular momentum, $\parallel \pi
1358   function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1359   conserved quantity in Poisson system. We can easily verify that the
1360   norm of the angular momentum, $\parallel \pi
1361 < \parallel$, is a \emph{Casimir}\cite{McLachlan1993}. Let $F(\pi ) = S(\frac{{\parallel
1361 > \parallel$, is a \emph{Casimir}.\cite{McLachlan1993} Let $F(\pi ) = S(\frac{{\parallel
1362   \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1363   then by the chain rule
1364   \[
# Line 1376 | Line 1379 | listed in Table~\ref{introTable:rbEquations}.
1379   The Hamiltonian of rigid body can be separated in terms of kinetic
1380   energy and potential energy, $H = T(p,\pi ) + V(q,Q)$. The equations
1381   of motion corresponding to potential energy and kinetic energy are
1382 < listed in Table~\ref{introTable:rbEquations}.
1382 > listed in Table~\ref{introTable:rbEquations}.
1383   \begin{table}
1384   \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1385   \label{introTable:rbEquations}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines