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 2697 by tim, Fri Apr 7 05:03:54 2006 UTC vs.
Revision 2699 by tim, Mon Apr 10 05:35:55 2006 UTC

# Line 212 | Line 212 | q_i }}} \right) = 0}
212   }}\dot p_i } \right)}  = \sum\limits_i {\left( {\frac{{\partial
213   H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
214   \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
215 < q_i }}} \right) = 0}
216 < \label{introEquation:conserveHalmitonian}
217 < \end{equation}
218 <
219 < When studying Hamiltonian system, it is more convenient to use
220 < notation
221 < \begin{equation}
222 < r = r(q,p)^T
223 < \end{equation}
224 < and to introduce a $2n \times 2n$ canonical structure matrix $J$,
225 < \begin{equation}
226 < J = \left( {\begin{array}{*{20}c}
227 <   0 & I  \\
228 <   { - I} & 0  \\
229 < \end{array}} \right)
230 < \label{introEquation:canonicalMatrix}
231 < \end{equation}
232 < where $I$ is a $n \times n$ identity matrix and $J$ is a
233 < skew-symmetric matrix ($ J^T  =  - J $). Thus, Hamiltonian system
234 < can be rewritten as,
235 < \begin{equation}
236 < \frac{d}{{dt}}r = J\nabla _r H(r)
237 < \label{introEquation:compactHamiltonian}
215 > q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
216   \end{equation}
217  
218   \section{\label{introSection:statisticalMechanics}Statistical
# Line 324 | Line 302 | classical mechanics. According to Liouville's theorem,
302   is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
303   the \emph{pullback} of $\eta$ under f is equal to $\omega$.
304   Canonical transformation is an example of symplectomorphism in
305 < classical mechanics. According to Liouville's theorem, the
328 < Hamiltonian \emph{flow} or \emph{symplectomorphism} generated by the
329 < Hamiltonian vector filed preserves the volume form on the phase
330 < space, which is the basis of classical statistical mechanics.
305 > classical mechanics.
306  
307 < \subsection{\label{introSection:exactFlow}The Exact Flow of ODE}
307 > \subsection{\label{introSection:ODE}Ordinary Differential Equations}
308  
309 < \subsection{\label{introSection:hamiltonianSplitting}Hamiltonian Splitting}
309 > For a ordinary differential system defined as
310 > \begin{equation}
311 > \dot x = f(x)
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).
316 > \end{equation}
317 > $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
318 > matrix
319 > \begin{equation}
320 > J = \left( {\begin{array}{*{20}c}
321 >   0 & I  \\
322 >   { - I} & 0  \\
323 > \end{array}} \right)
324 > \label{introEquation:canonicalMatrix}
325 > \end{equation}
326 > where $I$ is an identity matrix. Using this notation, Hamiltonian
327 > system can be rewritten as,
328 > \begin{equation}
329 > \frac{d}{{dt}}x = J\nabla _x H(x)
330 > \label{introEquation:compactHamiltonian}
331 > \end{equation}In this case, $f$ is
332 > called a \emph{Hamiltonian vector field}.
333 >
334 > Another generalization of Hamiltonian dynamics is Poisson Dynamics,
335 > \begin{equation}
336 > \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
337 > \end{equation}
338 > The most obvious change being that matrix $J$ now depends on $x$.
339 > The free rigid body is an example of Poisson system (actually a
340 > Lie-Poisson system) with Hamiltonian function of angular kinetic
341 > energy.
342 > \begin{equation}
343 > J(\pi ) = \left( {\begin{array}{*{20}c}
344 >   0 & {\pi _3 } & { - \pi _2 }  \\
345 >   { - \pi _3 } & 0 & {\pi _1 }  \\
346 >   {\pi _2 } & { - \pi _1 } & 0  \\
347 > \end{array}} \right)
348 > \end{equation}
349 >
350 > \begin{equation}
351 > H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
352 > }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
353 > \end{equation}
354 >
355 > \subsection{\label{introSection:geometricProperties}Geometric Properties}
356 > Let $x(t)$ be the exact solution of the ODE system,
357 > \begin{equation}
358 > \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
359 > \end{equation}
360 > The exact flow(solution) $\varphi_\tau$ is defined by
361 > \[
362 > x(t+\tau) =\varphi_\tau(x(t))
363 > \]
364 > where $\tau$ is a fixed time step and $\varphi$ is a map from phase
365 > space to itself. In most cases, it is not easy to find the exact
366 > flow $\varphi_\tau$. Instead, we use a approximate map, $\psi_\tau$,
367 > which is usually called integrator. The order of an integrator
368 > $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
369 > order $p$,
370 > \begin{equation}
371 > \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
372 > \end{equation}
373 >
374 > The hidden geometric properties of ODE and its flow play important
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 > '\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. 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 > '\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($
390 > \nabla \cdot f = 0 $) ODE, if the flow satisfies $ \det d\varphi  =
391 > 1$. Changing the variables $y = h(x)$ in a
392 > ODE\ref{introEquation:ODE} will result in a new system,
393 > \[
394 > \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
395 > \]
396 > The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
397 > In other words, the flow of this vector field is reversible if and
398 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. When
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: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