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 2699 by tim, Mon Apr 10 05:35:55 2006 UTC vs.
Revision 2700 by tim, Tue Apr 11 03:38:09 2006 UTC

# Line 117 | Line 117 | for a holonomic system of $f$ degrees of freedom, the
117   \subsubsection{\label{introSection:equationOfMotionLagrangian}The
118   Equations of Motion in Lagrangian Mechanics}
119  
120 < for a holonomic system of $f$ degrees of freedom, the equations of
120 > For a holonomic system of $f$ degrees of freedom, the equations of
121   motion in the Lagrangian form is
122   \begin{equation}
123   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
# Line 221 | Line 221 | Statistical Mechanics concepts presented in this disse
221   The thermodynamic behaviors and properties of Molecular Dynamics
222   simulation are governed by the principle of Statistical Mechanics.
223   The following section will give a brief introduction to some of the
224 < Statistical Mechanics concepts presented in this dissertation.
224 > Statistical Mechanics concepts and theorem presented in this
225 > dissertation.
226  
227 < \subsection{\label{introSection:ensemble}Ensemble and Phase Space}
227 > \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
228 >
229 > Mathematically, phase space is the space which represents all
230 > possible states. Each possible state of the system corresponds to
231 > one unique point in the phase space. For mechanical systems, the
232 > phase space usually consists of all possible values of position and
233 > momentum variables. Consider a dynamic system in a cartesian space,
234 > where each of the $6f$ coordinates and momenta is assigned to one of
235 > $6f$ mutually orthogonal axes, the phase space of this system is a
236 > $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
237 > \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
238 > momenta is a phase space vector.
239 >
240 > A microscopic state or microstate of a classical system is
241 > specification of the complete phase space vector of a system at any
242 > instant in time. An ensemble is defined as a collection of systems
243 > sharing one or more macroscopic characteristics but each being in a
244 > unique microstate. The complete ensemble is specified by giving all
245 > systems or microstates consistent with the common macroscopic
246 > characteristics of the ensemble. Although the state of each
247 > individual system in the ensemble could be precisely described at
248 > any instance in time by a suitable phase space vector, when using
249 > ensembles for statistical purposes, there is no need to maintain
250 > distinctions between individual systems, since the numbers of
251 > systems at any time in the different states which correspond to
252 > different regions of the phase space are more interesting. Moreover,
253 > in the point of view of statistical mechanics, one would prefer to
254 > use ensembles containing a large enough population of separate
255 > members so that the numbers of systems in such different states can
256 > be regarded as changing continuously as we traverse different
257 > regions of the phase space. The condition of an ensemble at any time
258 > can be regarded as appropriately specified by the density $\rho$
259 > with which representative points are distributed over the phase
260 > space. The density of distribution for an ensemble with $f$ degrees
261 > of freedom is defined as,
262 > \begin{equation}
263 > \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
264 > \label{introEquation:densityDistribution}
265 > \end{equation}
266 > Governed by the principles of mechanics, the phase points change
267 > their value which would change the density at any time at phase
268 > space. Hence, the density of distribution is also to be taken as a
269 > function of the time.
270 >
271 > The number of systems $\delta N$ at time $t$ can be determined by,
272 > \begin{equation}
273 > \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
274 > \label{introEquation:deltaN}
275 > \end{equation}
276 > Assuming a large enough population of systems are exploited, we can
277 > sufficiently approximate $\delta N$ without introducing
278 > discontinuity when we go from one region in the phase space to
279 > another. By integrating over the whole phase space,
280 > \begin{equation}
281 > N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
282 > \label{introEquation:totalNumberSystem}
283 > \end{equation}
284 > gives us an expression for the total number of the systems. Hence,
285 > the probability per unit in the phase space can be obtained by,
286 > \begin{equation}
287 > \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
288 > {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
289 > \label{introEquation:unitProbability}
290 > \end{equation}
291 > With the help of Equation(\ref{introEquation:unitProbability}) and
292 > the knowledge of the system, it is possible to calculate the average
293 > value of any desired quantity which depends on the coordinates and
294 > momenta of the system. Even when the dynamics of the real system is
295 > complex, or stochastic, or even discontinuous, the average
296 > properties of the ensemble of possibilities as a whole may still
297 > remain well defined. For a classical system in thermal equilibrium
298 > with its environment, the ensemble average of a mechanical quantity,
299 > $\langle A(q , p) \rangle_t$, takes the form of an integral over the
300 > phase space of the system,
301 > \begin{equation}
302 > \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
303 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
304 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
305 > \label{introEquation:ensembelAverage}
306 > \end{equation}
307 >
308 > There are several different types of ensembles with different
309 > statistical characteristics. As a function of macroscopic
310 > parameters, such as temperature \textit{etc}, partition function can
311 > be used to describe the statistical properties of a system in
312 > thermodynamic equilibrium.
313 >
314 > As an ensemble of systems, each of which is known to be thermally
315 > isolated and conserve energy, Microcanonical ensemble(NVE) has a
316 > partition function like,
317 > \begin{equation}
318 > \Omega (N,V,E) = e^{\beta TS}
319 > \label{introEqaution:NVEPartition}.
320 > \end{equation}
321 > A canonical ensemble(NVT)is an ensemble of systems, each of which
322 > can share its energy with a large heat reservoir. The distribution
323 > of the total energy amongst the possible dynamical states is given
324 > by the partition function,
325 > \begin{equation}
326 > \Omega (N,V,T) = e^{ - \beta A}
327 > \label{introEquation:NVTPartition}
328 > \end{equation}
329 > Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
330 > TS$. Since most experiment are carried out under constant pressure
331 > condition, isothermal-isobaric ensemble(NPT) play a very important
332 > role in molecular simulation. The isothermal-isobaric ensemble allow
333 > the system to exchange energy with a heat bath of temperature $T$
334 > and to change the volume as well. Its partition function is given as
335 > \begin{equation}
336 > \Delta (N,P,T) =  - e^{\beta G}.
337 > \label{introEquation:NPTPartition}
338 > \end{equation}
339 > Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
340 >
341 > \subsection{\label{introSection:liouville}Liouville's theorem}
342 >
343 > The Liouville's theorem is the foundation on which statistical
344 > mechanics rests. It describes the time evolution of phase space
345 > distribution function. In order to calculate the rate of change of
346 > $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
347 > consider the two faces perpendicular to the $q_1$ axis, which are
348 > located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
349 > leaving the opposite face is given by the expression,
350 > \begin{equation}
351 > \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
352 > \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
353 > }}\delta q_1 } \right)\delta q_2  \ldots \delta q_f \delta p_1
354 > \ldots \delta p_f .
355 > \end{equation}
356 > Summing all over the phase space, we obtain
357 > \begin{equation}
358 > \frac{{d(\delta N)}}{{dt}} =  - \sum\limits_{i = 1}^f {\left[ {\rho
359 > \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
360 > \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
361 > {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  + \frac{{\partial
362 > \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
363 > \ldots \delta q_f \delta p_1  \ldots \delta p_f .
364 > \end{equation}
365 > Differentiating the equations of motion in Hamiltonian formalism
366 > (\ref{introEquation:motionHamiltonianCoordinate},
367 > \ref{introEquation:motionHamiltonianMomentum}), we can show,
368 > \begin{equation}
369 > \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
370 > + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
371 > \end{equation}
372 > which cancels the first terms of the right hand side. Furthermore,
373 > divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
374 > p_f $ in both sides, we can write out Liouville's theorem in a
375 > simple form,
376 > \begin{equation}
377 > \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
378 > {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  +
379 > \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
380 > \label{introEquation:liouvilleTheorem}
381 > \end{equation}
382  
383 + Liouville's theorem states that the distribution function is
384 + constant along any trajectory in phase space. In classical
385 + statistical mechanics, since the number of particles in the system
386 + is huge, we may be able to believe the system is stationary,
387 + \begin{equation}
388 + \frac{{\partial \rho }}{{\partial t}} = 0.
389 + \label{introEquation:stationary}
390 + \end{equation}
391 + In such stationary system, the density of distribution $\rho$ can be
392 + connected to the Hamiltonian $H$ through Maxwell-Boltzmann
393 + distribution,
394 + \begin{equation}
395 + \rho  \propto e^{ - \beta H}
396 + \label{introEquation:densityAndHamiltonian}
397 + \end{equation}
398 +
399 + Liouville's theorem can be expresses in a variety of different forms
400 + which are convenient within different contexts. For any two function
401 + $F$ and $G$ of the coordinates and momenta of a system, the Poisson
402 + bracket ${F, G}$ is defined as
403 + \begin{equation}
404 + \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
405 + F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
406 + \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
407 + q_i }}} \right)}.
408 + \label{introEquation:poissonBracket}
409 + \end{equation}
410 + Substituting equations of motion in Hamiltonian formalism(
411 + \ref{introEquation:motionHamiltonianCoordinate} ,
412 + \ref{introEquation:motionHamiltonianMomentum} ) into
413 + (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
414 + theorem using Poisson bracket notion,
415 + \begin{equation}
416 + \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
417 + {\rho ,H} \right\}.
418 + \label{introEquation:liouvilleTheromInPoissin}
419 + \end{equation}
420 + Moreover, the Liouville operator is defined as
421 + \begin{equation}
422 + iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
423 + p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
424 + H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
425 + \label{introEquation:liouvilleOperator}
426 + \end{equation}
427 + In terms of Liouville operator, Liouville's equation can also be
428 + expressed as
429 + \begin{equation}
430 + \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
431 + \label{introEquation:liouvilleTheoremInOperator}
432 + \end{equation}
433 +
434 +
435   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
436  
437   Various thermodynamic properties can be calculated from Molecular
# Line 239 | Line 446 | statistical ensemble are identical \cite{Frenkel1996,
446   ensemble average. It states that time average and average over the
447   statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
448   \begin{equation}
449 < \langle A \rangle_t = \mathop {\lim }\limits_{t \to \infty }
450 < \frac{1}{t}\int\limits_0^t {A(p(t),q(t))dt = \int\limits_\Gamma
451 < {A(p(t),q(t))} } \rho (p(t), q(t)) dpdq
449 > \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
450 > \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
451 > {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
452   \end{equation}
453 < where $\langle A \rangle_t$ is an equilibrium value of a physical
454 < quantity and $\rho (p(t), q(t))$ is the equilibrium distribution
455 < function. If an observation is averaged over a sufficiently long
456 < time (longer than relaxation time), all accessible microstates in
457 < phase space are assumed to be equally probed, giving a properly
458 < weighted statistical average. This allows the researcher freedom of
459 < choice when deciding how best to measure a given observable. In case
460 < an ensemble averaged approach sounds most reasonable, the Monte
461 < Carlo techniques\cite{metropolis:1949} can be utilized. Or if the
462 < system lends itself to a time averaging approach, the Molecular
463 < Dynamics techniques in Sec.~\ref{introSection:molecularDynamics}
464 < will be the best choice\cite{Frenkel1996}.
453 > where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
454 > physical quantity and $\rho (p(t), q(t))$ is the equilibrium
455 > distribution function. If an observation is averaged over a
456 > sufficiently long time (longer than relaxation time), all accessible
457 > microstates in phase space are assumed to be equally probed, giving
458 > a properly weighted statistical average. This allows the researcher
459 > freedom of choice when deciding how best to measure a given
460 > observable. In case an ensemble averaged approach sounds most
461 > reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
462 > utilized. Or if the system lends itself to a time averaging
463 > approach, the Molecular Dynamics techniques in
464 > Sec.~\ref{introSection:molecularDynamics} will be the best
465 > choice\cite{Frenkel1996}.
466  
467   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
468   A variety of numerical integrators were proposed to simulate the
# Line 442 | Line 650 | Suppose that a Hamiltonian system has a form with $H =
650   \label{introEquation:SymplecticFlowComposition}
651   \end{equation}
652   Suppose that a Hamiltonian system has a form with $H = T + V$
445
446
653  
654   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
655  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines