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 2698 by tim, Fri Apr 7 22:05:48 2006 UTC vs.
Revision 2699 by tim, Mon Apr 10 05:35:55 2006 UTC

# Line 312 | Line 312 | f(r) = J\nabla _x H(r)
312   \end{equation}
313   where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
314   \begin{equation}
315 < f(r) = J\nabla _x H(r)
315 > f(r) = J\nabla _x H(r).
316   \end{equation}
317   $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
318   matrix
# Line 372 | Line 372 | roles in numerical studies. The flow of a Hamiltonian
372   \end{equation}
373  
374   The hidden geometric properties of ODE and its flow play important
375 < roles in numerical studies. The flow of a Hamiltonian vector field
376 < on a symplectic manifold is a symplectomorphism. Let $\varphi$ be
377 < the flow of Hamiltonian vector field, $\varphi$ is a
378 < \emph{symplectic} flow if it satisfies,
375 > roles in numerical studies. Let $\varphi$ be the flow of Hamiltonian
376 > vector field, $\varphi$ is a \emph{symplectic} flow if it satisfies,
377   \begin{equation}
378 < d \varphi^T J d \varphi = J.
378 > '\varphi^T J '\varphi = J.
379   \end{equation}
380   According to Liouville's theorem, the symplectic volume is invariant
381   under a Hamiltonian flow, which is the basis for classical
382 < statistical mechanics. As to the Poisson system,
382 > statistical mechanics. Furthermore, the flow of a Hamiltonian vector
383 > field on a symplectic manifold can be shown to be a
384 > symplectomorphism. As to the Poisson system,
385   \begin{equation}
386 < d\varphi ^T Jd\varphi  = J \circ \varphi
386 > '\varphi ^T J '\varphi  = J \circ \varphi
387   \end{equation}
388   is the property must be preserved by the integrator. It is possible
389   to construct a \emph{volume-preserving} flow for a source free($
# Line 399 | Line 399 | the structural properties of the original ODE and its
399   designing any numerical methods, one should always try to preserve
400   the structural properties of the original ODE and its flow.
401  
402 < \subsection{\label{introSection:splittingAndComposition}Splitting and Composition Methods}
402 > \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
403 > A lot of well established and very effective numerical methods have
404 > been successful precisely because of their symplecticities even
405 > though this fact was not recognized when they were first
406 > constructed. The most famous example is leapfrog methods in
407 > molecular dynamics. In general, symplectic integrators can be
408 > constructed using one of four different methods.
409 > \begin{enumerate}
410 > \item Generating functions
411 > \item Variational methods
412 > \item Runge-Kutta methods
413 > \item Splitting methods
414 > \end{enumerate}
415 >
416 > Generating function tends to lead to methods which are cumbersome
417 > and difficult to use\cite{}. In dissipative systems, variational
418 > methods can capture the decay of energy accurately\cite{}. Since
419 > their geometrically unstable nature against non-Hamiltonian
420 > perturbations, ordinary implicit Runge-Kutta methods are not
421 > suitable for Hamiltonian system. Recently, various high-order
422 > explicit Runge--Kutta methods have been developed to overcome this
423 > instability \cite{}. However, due to computational penalty involved
424 > in implementing the Runge-Kutta methods, they do not attract too
425 > much attention from Molecular Dynamics community. Instead, splitting
426 > have been widely accepted since they exploit natural decompositions
427 > of the system\cite{Tuckerman92}. The main idea behind splitting
428 > methods is to decompose the discrete $\varphi_h$ as a composition of
429 > simpler flows,
430 > \begin{equation}
431 > \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
432 > \varphi _{h_n }
433 > \label{introEquation:FlowDecomposition}
434 > \end{equation}
435 > where each of the sub-flow is chosen such that each represent a
436 > simpler integration of the system. Let $\phi$ and $\psi$ both be
437 > symplectic maps, it is easy to show that any composition of
438 > symplectic flows yields a symplectic map,
439 > \begin{equation}
440 > (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
441 > '\phi ' = \phi '^T J\phi ' = J.
442 > \label{introEquation:SymplecticFlowComposition}
443 > \end{equation}
444 > Suppose that a Hamiltonian system has a form with $H = T + V$
445  
446 +
447 +
448   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
449  
450   As a special discipline of molecular modeling, Molecular dynamics

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines