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 2905 by tim, Wed Jun 28 19:09:25 2006 UTC vs.
Revision 2907 by tim, Thu Jun 29 16:57:37 2006 UTC

# Line 3 | Line 3 | Closely related to Classical Mechanics, Molecular Dyna
3   \section{\label{introSection:classicalMechanics}Classical
4   Mechanics}
5  
6 < Closely related to Classical Mechanics, Molecular Dynamics
7 < simulations are carried out by integrating the equations of motion
8 < for a given system of particles. There are three fundamental ideas
9 < behind classical mechanics. Firstly, one can determine the state of
10 < a mechanical system at any time of interest; Secondly, all the
11 < mechanical properties of the system at that time can be determined
12 < by combining the knowledge of the properties of the system with the
13 < specification of this state; Finally, the specification of the state
14 < when further combine with the laws of mechanics will also be
15 < sufficient to predict the future behavior of the system.
6 > Using equations of motion derived from Classical Mechanics,
7 > Molecular Dynamics simulations are carried out by integrating the
8 > equations of motion for a given system of particles. There are three
9 > fundamental ideas behind classical mechanics. Firstly, one can
10 > determine the state of a mechanical system at any time of interest;
11 > Secondly, all the mechanical properties of the system at that time
12 > can be determined by combining the knowledge of the properties of
13 > the system with the specification of this state; Finally, the
14 > specification of the state when further combined with the laws of
15 > mechanics will also be sufficient to predict the future behavior of
16 > the system.
17  
18   \subsection{\label{introSection:newtonian}Newtonian Mechanics}
19   The discovery of Newton's three laws of mechanics which govern the
# Line 71 | Line 72 | Newtonian Mechanics suffers from a important limitatio
72  
73   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
75 < Newtonian Mechanics suffers from a important limitation: motions can
75 > Newtonian Mechanics suffers from an important limitation: motion can
76   only be described in cartesian coordinate systems which make it
77   impossible to predict analytically the properties of the system even
78   if we know all of the details of the interaction. In order to
79   overcome some of the practical difficulties which arise in attempts
80 < to apply Newton's equation to complex system, approximate numerical
80 > to apply Newton's equation to complex systems, approximate numerical
81   procedures may be developed.
82  
83   \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
# Line 84 | Line 85 | Hamilton's Principle may be stated as follows: the act
85  
86   Hamilton introduced the dynamical principle upon which it is
87   possible to base all of mechanics and most of classical physics.
88 < Hamilton's Principle may be stated as follows: the actual
89 < trajectory, along which a dynamical system may move from one point
90 < to another within a specified time, is derived by finding the path
91 < which minimizes the time integral of the difference between the
92 < kinetic $K$, and potential energies $U$,
88 > Hamilton's Principle may be stated as follows: the trajectory, along
89 > which a dynamical system may move from one point to another within a
90 > specified time, is derived by finding the path which minimizes the
91 > time integral of the difference between the kinetic $K$, and
92 > potential energies $U$,
93   \begin{equation}
94   \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0}.
95   \label{introEquation:halmitonianPrinciple1}
# Line 213 | Line 214 | possible states. Each possible state of the system cor
214   \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
215  
216   Mathematically, phase space is the space which represents all
217 < possible states. Each possible state of the system corresponds to
218 < one unique point in the phase space. For mechanical systems, the
219 < phase space usually consists of all possible values of position and
220 < momentum variables. Consider a dynamic system of $f$ particles in a
221 < cartesian space, where each of the $6f$ coordinates and momenta is
222 < assigned to one of $6f$ mutually orthogonal axes, the phase space of
223 < this system is a $6f$ dimensional space. A point, $x =
217 > possible states of a system. Each possible state of the system
218 > corresponds to one unique point in the phase space. For mechanical
219 > systems, the phase space usually consists of all possible values of
220 > position and momentum variables. Consider a dynamic system of $f$
221 > particles in a cartesian space, where each of the $6f$ coordinates
222 > and momenta is assigned to one of $6f$ mutually orthogonal axes, the
223 > phase space of this system is a $6f$ dimensional space. A point, $x
224 > =
225   (\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
226   \over q} _1 , \ldots
227   ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
# Line 241 | Line 243 | their locations which would change the density at any
243   \label{introEquation:densityDistribution}
244   \end{equation}
245   Governed by the principles of mechanics, the phase points change
246 < their locations which would change the density at any time at phase
246 > their locations which changes the density at any time at phase
247   space. Hence, the density distribution is also to be taken as a
248   function of the time. The number of systems $\delta N$ at time $t$
249   can be determined by,
# Line 249 | Line 251 | Assuming a large enough population of systems, we can
251   \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
252   \label{introEquation:deltaN}
253   \end{equation}
254 < Assuming a large enough population of systems, we can sufficiently
254 > Assuming enough copies of the systems, we can sufficiently
255   approximate $\delta N$ without introducing discontinuity when we go
256   from one region in the phase space to another. By integrating over
257   the whole phase space,
# Line 257 | Line 259 | gives us an expression for the total number of the sys
259   N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
260   \label{introEquation:totalNumberSystem}
261   \end{equation}
262 < gives us an expression for the total number of the systems. Hence,
263 < the probability per unit in the phase space can be obtained by,
262 > gives us an expression for the total number of copies. Hence, the
263 > probability per unit volume in the phase space can be obtained by,
264   \begin{equation}
265   \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
266   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
# Line 267 | Line 269 | momenta of the system. Even when the dynamics of the r
269   With the help of Eq.~\ref{introEquation:unitProbability} and the
270   knowledge of the system, it is possible to calculate the average
271   value of any desired quantity which depends on the coordinates and
272 < momenta of the system. Even when the dynamics of the real system is
272 > momenta of the system. Even when the dynamics of the real system are
273   complex, or stochastic, or even discontinuous, the average
274 < properties of the ensemble of possibilities as a whole remaining
275 < well defined. For a classical system in thermal equilibrium with its
274 > properties of the ensemble of possibilities as a whole remain well
275 > defined. For a classical system in thermal equilibrium with its
276   environment, the ensemble average of a mechanical quantity, $\langle
277   A(q , p) \rangle_t$, takes the form of an integral over the phase
278   space of the system,
# Line 355 | Line 357 | statistical mechanics, since the number of members in
357   \end{equation}
358   Liouville's theorem states that the distribution function is
359   constant along any trajectory in phase space. In classical
360 < statistical mechanics, since the number of members in an ensemble is
361 < huge and constant, we can assume the local density has no reason
362 < (other than classical mechanics) to change,
360 > statistical mechanics, since the number of system copies in an
361 > ensemble is huge and constant, we can assume the local density has
362 > no reason (other than classical mechanics) to change,
363   \begin{equation}
364   \frac{{\partial \rho }}{{\partial t}} = 0.
365   \label{introEquation:stationary}
# Line 387 | Line 389 | With the help of stationary assumption
389   \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
390   \frac{d}{{dt}}(\delta v) = 0.
391   \end{equation}
392 < With the help of stationary assumption
393 < (\ref{introEquation:stationary}), we obtain the principle of the
392 > With the help of the stationary assumption
393 > (Eq.~\ref{introEquation:stationary}), we obtain the principle of
394   \emph{conservation of volume in phase space},
395   \begin{equation}
396   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
# Line 398 | Line 400 | Liouville's theorem can be expresses in a variety of d
400  
401   \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
402  
403 < Liouville's theorem can be expresses in a variety of different forms
403 > Liouville's theorem can be expressed in a variety of different forms
404   which are convenient within different contexts. For any two function
405   $F$ and $G$ of the coordinates and momenta of a system, the Poisson
406   bracket ${F, G}$ is defined as
# Line 432 | Line 434 | expressed as
434   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
435   \label{introEquation:liouvilleTheoremInOperator}
436   \end{equation}
437 <
437 > which can help define a propagator $\rho (t) = e^{-iLt} \rho (0)$.
438   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
439  
440   Various thermodynamic properties can be calculated from Molecular
# Line 441 | Line 443 | period of them which is different from the average beh
443   simulation and the quality of the underlying model. However, both
444   experiments and computer simulations are usually performed during a
445   certain time interval and the measurements are averaged over a
446 < period of them which is different from the average behavior of
446 > period of time which is different from the average behavior of
447   many-body system in Statistical Mechanics. Fortunately, the Ergodic
448   Hypothesis makes a connection between time average and the ensemble
449   average. It states that the time average and average over the
# Line 454 | Line 456 | sufficiently long time (longer than relaxation time),
456   where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
457   physical quantity and $\rho (p(t), q(t))$ is the equilibrium
458   distribution function. If an observation is averaged over a
459 < sufficiently long time (longer than relaxation time), all accessible
460 < microstates in phase space are assumed to be equally probed, giving
461 < a properly weighted statistical average. This allows the researcher
462 < freedom of choice when deciding how best to measure a given
463 < observable. In case an ensemble averaged approach sounds most
459 > sufficiently long time (longer than the relaxation time), all
460 > accessible microstates in phase space are assumed to be equally
461 > probed, giving a properly weighted statistical average. This allows
462 > the researcher freedom of choice when deciding how best to measure a
463 > given observable. In case an ensemble averaged approach sounds most
464   reasonable, the Monte Carlo methods\cite{Metropolis1949} can be
465   utilized. Or if the system lends itself to a time averaging
466   approach, the Molecular Dynamics techniques in
# Line 472 | Line 474 | such as symplectic structure, volume and time reversal
474   by the differential equations. However, most of them ignore the
475   hidden physical laws contained within the equations. Since 1990,
476   geometric integrators, which preserve various phase-flow invariants
477 < such as symplectic structure, volume and time reversal symmetry, are
478 < developed to address this issue\cite{Dullweber1997, McLachlan1998,
479 < Leimkuhler1999}. The velocity Verlet method, which happens to be a
480 < simple example of symplectic integrator, continues to gain
481 < popularity in the molecular dynamics community. This fact can be
482 < partly explained by its geometric nature.
477 > such as symplectic structure, volume and time reversal symmetry,
478 > were developed to address this issue\cite{Dullweber1997,
479 > McLachlan1998, Leimkuhler1999}. The velocity Verlet method, which
480 > happens to be a simple example of symplectic integrator, continues
481 > to gain popularity in the molecular dynamics community. This fact
482 > can be partly explained by its geometric nature.
483  
484   \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
485   A \emph{manifold} is an abstract mathematical space. It looks
# Line 486 | Line 488 | apply calculus on \emph{differentiable manifold}. A \e
488   surface of Earth. It seems to be flat locally, but it is round if
489   viewed as a whole. A \emph{differentiable manifold} (also known as
490   \emph{smooth manifold}) is a manifold on which it is possible to
491 < apply calculus on \emph{differentiable manifold}. A \emph{symplectic
492 < manifold} is defined as a pair $(M, \omega)$ which consists of a
491 > apply calculus\cite{Hirsch1997}. A \emph{symplectic manifold} is
492 > defined as a pair $(M, \omega)$ which consists of a
493   \emph{differentiable manifold} $M$ and a close, non-degenerated,
494   bilinear symplectic form, $\omega$. A symplectic form on a vector
495   space $V$ is a function $\omega(x, y)$ which satisfies
496   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
497   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
498 < $\omega(x, x) = 0$. The cross product operation in vector field is
499 < an example of symplectic form. One of the motivations to study
500 < \emph{symplectic manifolds} in Hamiltonian Mechanics is that a
501 < symplectic manifold can represent all possible configurations of the
502 < system and the phase space of the system can be described by it's
503 < cotangent bundle. Every symplectic manifold is even dimensional. For
504 < instance, in Hamilton equations, coordinate and momentum always
505 < appear in pairs.
498 > $\omega(x, x) = 0$\cite{McDuff1998}. The cross product operation in
499 > vector field is an example of symplectic form. One of the
500 > motivations to study \emph{symplectic manifolds} in Hamiltonian
501 > Mechanics is that a symplectic manifold can represent all possible
502 > configurations of the system and the phase space of the system can
503 > be described by it's cotangent bundle\cite{Jost2002}. Every
504 > symplectic manifold is even dimensional. For instance, in Hamilton
505 > equations, coordinate and momentum always appear in pairs.
506  
507   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
508  
# Line 509 | Line 511 | $f(r) = J\nabla _x H(r)$. Here, $H = H (q, p)$ is Hami
511   \dot x = f(x)
512   \end{equation}
513   where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
514 < $f(r) = J\nabla _x H(r)$. Here, $H = H (q, p)$ is Hamiltonian
514 > $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
515   function and $J$ is the skew-symmetric matrix
516   \begin{equation}
517   J = \left( {\begin{array}{*{20}c}
# Line 531 | Line 533 | The most obvious change being that matrix $J$ now depe
533   \end{equation}
534   The most obvious change being that matrix $J$ now depends on $x$.
535  
536 < \subsection{\label{introSection:exactFlow}Exact Flow}
536 > \subsection{\label{introSection:exactFlow}Exact Propagator}
537  
538   Let $x(t)$ be the exact solution of the ODE
539   system,$\frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}$, we can
540 < define its exact flow(solution) $\varphi_\tau$
540 > define its exact propagator(solution) $\varphi_\tau$
541   \[ x(t+\tau)
542   =\varphi_\tau(x(t))
543   \]
544   where $\tau$ is a fixed time step and $\varphi$ is a map from phase
545 < space to itself. The flow has the continuous group property,
545 > space to itself. The propagator has the continuous group property,
546   \begin{equation}
547   \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
548   + \tau _2 } .
# Line 549 | Line 551 | Therefore, the exact flow is self-adjoint,
551   \begin{equation}
552   \varphi _\tau   \circ \varphi _{ - \tau }  = I
553   \end{equation}
554 < Therefore, the exact flow is self-adjoint,
554 > Therefore, the exact propagator is self-adjoint,
555   \begin{equation}
556   \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
557   \end{equation}
558 < The exact flow can also be written in terms of the of an operator,
558 > The exact propagator can also be written in terms of operator,
559   \begin{equation}
560   \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
561   }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
562   \label{introEquation:exponentialOperator}
563   \end{equation}
564 < In most cases, it is not easy to find the exact flow $\varphi_\tau$.
565 < Instead, we use an approximate map, $\psi_\tau$, which is usually
566 < called integrator. The order of an integrator $\psi_\tau$ is $p$, if
567 < the Taylor series of $\psi_\tau$ agree to order $p$,
564 > In most cases, it is not easy to find the exact propagator
565 > $\varphi_\tau$. Instead, we use an approximate map, $\psi_\tau$,
566 > which is usually called an integrator. The order of an integrator
567 > $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
568 > order $p$,
569   \begin{equation}
570   \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
571   \end{equation}
# Line 570 | Line 573 | ODE and its flow play important roles in numerical stu
573   \subsection{\label{introSection:geometricProperties}Geometric Properties}
574  
575   The hidden geometric properties\cite{Budd1999, Marsden1998} of an
576 < ODE and its flow play important roles in numerical studies. Many of
577 < them can be found in systems which occur naturally in applications.
578 < Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
579 < a \emph{symplectic} flow if it satisfies,
576 > ODE and its propagator play important roles in numerical studies.
577 > Many of them can be found in systems which occur naturally in
578 > applications. Let $\varphi$ be the propagator of Hamiltonian vector
579 > field, $\varphi$ is a \emph{symplectic} propagator if it satisfies,
580   \begin{equation}
581   {\varphi '}^T J \varphi ' = J.
582   \end{equation}
583   According to Liouville's theorem, the symplectic volume is invariant
584 < under a Hamiltonian flow, which is the basis for classical
585 < statistical mechanics. Furthermore, the flow of a Hamiltonian vector
586 < field on a symplectic manifold can be shown to be a
584 > under a Hamiltonian propagator, which is the basis for classical
585 > statistical mechanics. Furthermore, the propagator of a Hamiltonian
586 > vector field on a symplectic manifold can be shown to be a
587   symplectomorphism. As to the Poisson system,
588   \begin{equation}
589   {\varphi '}^T J \varphi ' = J \circ \varphi
590   \end{equation}
591   is the property that must be preserved by the integrator. It is
592 < possible to construct a \emph{volume-preserving} flow for a source
593 < free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $ \det
594 < d\varphi  = 1$. One can show easily that a symplectic flow will be
595 < volume-preserving. Changing the variables $y = h(x)$ in an ODE
596 < (Eq.~\ref{introEquation:ODE}) will result in a new system,
592 > possible to construct a \emph{volume-preserving} propagator for a
593 > source free ODE ($ \nabla \cdot f = 0 $), if the propagator
594 > satisfies $ \det d\varphi  = 1$. One can show easily that a
595 > symplectic propagator will be volume-preserving. Changing the
596 > variables $y = h(x)$ in an ODE (Eq.~\ref{introEquation:ODE}) will
597 > result in a new system,
598   \[
599   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
600   \]
601   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
602 < In other words, the flow of this vector field is reversible if and
603 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
604 < \emph{first integral}, or conserved quantity of a general
605 < differential function is a function $ G:R^{2d}  \to R^d $ which is
606 < constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
602 > In other words, the propagator of this vector field is reversible if
603 > and only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
604 > conserved quantity of a general differential function is a function
605 > $ G:R^{2d}  \to R^d $ which is constant for all solutions of the ODE
606 > $\frac{{dx}}{{dt}} = f(x)$ ,
607   \[
608   \frac{{dG(x(t))}}{{dt}} = 0.
609   \]
610 < Using chain rule, one may obtain,
610 > Using the chain rule, one may obtain,
611   \[
612   \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \dot \nabla G,
613   \]
614 < which is the condition for conserving \emph{first integral}. For a
615 < canonical Hamiltonian system, the time evolution of an arbitrary
616 < smooth function $G$ is given by,
614 > which is the condition for conserved quantities. For a canonical
615 > Hamiltonian system, the time evolution of an arbitrary smooth
616 > function $G$ is given by,
617   \begin{eqnarray}
618   \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \notag\\
619                          & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)).
# Line 620 | Line 624 | Therefore, the sufficient condition for $G$ to be the
624   \[
625   \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
626   \]
627 < Therefore, the sufficient condition for $G$ to be the \emph{first
628 < integral} of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$
629 < As well known, the Hamiltonian (or energy) H of a Hamiltonian system
630 < is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
631 < 0$. When designing any numerical methods, one should always try to
632 < preserve the structural properties of the original ODE and its flow.
627 > Therefore, the sufficient condition for $G$ to be a conserved
628 > quantity of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$ As
629 > is well known, the Hamiltonian (or energy) H of a Hamiltonian system
630 > is a conserved quantity, which is due to the fact $\{ H,H\}  = 0$.
631 > When designing any numerical methods, one should always try to
632 > preserve the structural properties of the original ODE and its
633 > propagator.
634  
635   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
636   A lot of well established and very effective numerical methods have
637 < been successful precisely because of their symplecticities even
637 > been successful precisely because of their symplectic nature even
638   though this fact was not recognized when they were first
639   constructed. The most famous example is the Verlet-leapfrog method
640   in molecular dynamics. In general, symplectic integrators can be
# Line 640 | Line 645 | Generating function\cite{Channell1990} tends to lead t
645   \item Runge-Kutta methods
646   \item Splitting methods
647   \end{enumerate}
648 < Generating function\cite{Channell1990} tends to lead to methods
648 > Generating functions\cite{Channell1990} tend to lead to methods
649   which are cumbersome and difficult to use. In dissipative systems,
650   variational methods can capture the decay of energy
651 < accurately\cite{Kane2000}. Since their geometrically unstable nature
651 > accurately\cite{Kane2000}. Since they are geometrically unstable
652   against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
653   methods are not suitable for Hamiltonian system. Recently, various
654   high-order explicit Runge-Kutta methods \cite{Owren1992,Chen2003}
# Line 657 | Line 662 | $\varphi_h$ as a composition of simpler flows,
662   \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
663  
664   The main idea behind splitting methods is to decompose the discrete
665 < $\varphi_h$ as a composition of simpler flows,
665 > $\varphi_h$ as a composition of simpler propagators,
666   \begin{equation}
667   \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
668   \varphi _{h_n }
669   \label{introEquation:FlowDecomposition}
670   \end{equation}
671 < where each of the sub-flow is chosen such that each represent a
672 < simpler integration of the system. Suppose that a Hamiltonian system
673 < takes the form,
671 > where each of the sub-propagator is chosen such that each represent
672 > a simpler integration of the system. Suppose that a Hamiltonian
673 > system takes the form,
674   \[
675   H = H_1 + H_2.
676   \]
677   Here, $H_1$ and $H_2$ may represent different physical processes of
678   the system. For instance, they may relate to kinetic and potential
679   energy respectively, which is a natural decomposition of the
680 < problem. If $H_1$ and $H_2$ can be integrated using exact flows
681 < $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
682 < order expression is then given by the Lie-Trotter formula
680 > problem. If $H_1$ and $H_2$ can be integrated using exact
681 > propagators $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a
682 > simple first order expression is then given by the Lie-Trotter
683 > formula
684   \begin{equation}
685   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
686   \label{introEquation:firstOrderSplitting}
# Line 683 | Line 689 | It is easy to show that any composition of symplectic
689   continuous $\varphi _i$ over a time $h$. By definition, as
690   $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
691   must follow that each operator $\varphi_i(t)$ is a symplectic map.
692 < It is easy to show that any composition of symplectic flows yields a
693 < symplectic map,
692 > It is easy to show that any composition of symplectic propagators
693 > yields a symplectic map,
694   \begin{equation}
695   (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
696   '\phi ' = \phi '^T J\phi ' = J,
# Line 692 | Line 698 | splitting in this context automatically generates a sy
698   \end{equation}
699   where $\phi$ and $\psi$ both are symplectic maps. Thus operator
700   splitting in this context automatically generates a symplectic map.
695
701   The Lie-Trotter
702   splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
703 < local errors proportional to $h^2$, while Strang splitting gives a
704 < second-order decomposition,
703 > local errors proportional to $h^2$, while the Strang splitting gives
704 > a second-order decomposition,
705   \begin{equation}
706   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
707   _{1,h/2} , \label{introEquation:secondOrderSplitting}
708   \end{equation}
709 < which has a local error proportional to $h^3$. The Sprang
709 > which has a local error proportional to $h^3$. The Strang
710   splitting's popularity in molecular simulation community attribute
711   to its symmetric property,
712   \begin{equation}
# Line 772 | Line 777 | local error of splitting method in terms of the commut
777   \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
778  
779   The Baker-Campbell-Hausdorff formula can be used to determine the
780 < local error of splitting method in terms of the commutator of the
780 > local error of a splitting method in terms of the commutator of the
781   operators(\ref{introEquation:exponentialOperator}) associated with
782 < the sub-flow. For operators $hX$ and $hY$ which are associated with
783 < $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
782 > the sub-propagator. For operators $hX$ and $hY$ which are associated
783 > with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
784   \begin{equation}
785   \exp (hX + hY) = \exp (hZ)
786   \end{equation}
# Line 784 | Line 789 | Here, $[X,Y]$ is the commutators of operator $X$ and $
789   hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
790   {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
791   \end{equation}
792 < Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
792 > Here, $[X,Y]$ is the commutator of operator $X$ and $Y$ given by
793   \[
794   [X,Y] = XY - YX .
795   \]
796   Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
797 < to the Sprang splitting, we can obtain
797 > to the Strang splitting, we can obtain
798   \begin{eqnarray*}
799   \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 \\
800                                     &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
# Line 797 | Line 802 | error of Spring splitting is proportional to $h^3$. Th
802                                     ).
803   \end{eqnarray*}
804   Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local
805 < error of Spring splitting is proportional to $h^3$. The same
805 > error of Strang splitting is proportional to $h^3$. The same
806   procedure can be applied to a general splitting of the form
807   \begin{equation}
808   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
# Line 833 | Line 838 | simulations. For instance, instantaneous temperature o
838   dynamical information. The basic idea of molecular dynamics is that
839   macroscopic properties are related to microscopic behavior and
840   microscopic behavior can be calculated from the trajectories in
841 < simulations. For instance, instantaneous temperature of an
842 < Hamiltonian system of $N$ particle can be measured by
841 > simulations. For instance, instantaneous temperature of a
842 > Hamiltonian system of $N$ particles can be measured by
843   \[
844   T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
845   \]
846   where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
847   respectively, $f$ is the number of degrees of freedom, and $k_B$ is
848 < the boltzman constant.
848 > the Boltzman constant.
849  
850   A typical molecular dynamics run consists of three essential steps:
851   \begin{enumerate}
# Line 857 | Line 862 | will discusse issues in production run.
862   These three individual steps will be covered in the following
863   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
864   initialization of a simulation. Sec.~\ref{introSection:production}
865 < will discusse issues in production run.
865 > will discuss issues of production runs.
866   Sec.~\ref{introSection:Analysis} provides the theoretical tools for
867 < trajectory analysis.
867 > analysis of trajectories.
868  
869   \subsection{\label{introSec:initialSystemSettings}Initialization}
870  
# Line 871 | Line 876 | structure, some important information is missing. For
876   thousands of crystal structures of molecules are discovered every
877   year, many more remain unknown due to the difficulties of
878   purification and crystallization. Even for molecules with known
879 < structure, some important information is missing. For example, a
879 > structures, some important information is missing. For example, a
880   missing hydrogen atom which acts as donor in hydrogen bonding must
881 < be added. Moreover, in order to include electrostatic interaction,
881 > be added. Moreover, in order to include electrostatic interactions,
882   one may need to specify the partial charges for individual atoms.
883   Under some circumstances, we may even need to prepare the system in
884   a special configuration. For instance, when studying transport
# Line 893 | Line 898 | structure from crystallographic data. Relied on the gr
898   surface and to locate the local minimum. While converging slowly
899   near the minimum, steepest descent method is extremely robust when
900   systems are strongly anharmonic. Thus, it is often used to refine
901 < structure from crystallographic data. Relied on the gradient or
902 < hessian, advanced methods like Newton-Raphson converge rapidly to a
903 < local minimum, but become unstable if the energy surface is far from
901 > structures from crystallographic data. Relying on the Hessian,
902 > advanced methods like Newton-Raphson converge rapidly to a local
903 > minimum, but become unstable if the energy surface is far from
904   quadratic. Another factor that must be taken into account, when
905   choosing energy minimization method, is the size of the system.
906   Steepest descent and conjugate gradient can deal with models of any
907   size. Because of the limits on computer memory to store the hessian
908 < matrix and the computing power needed to diagonalized these
909 < matrices, most Newton-Raphson methods can not be used with very
905 < large systems.
908 > matrix and the computing power needed to diagonalize these matrices,
909 > most Newton-Raphson methods can not be used with very large systems.
910  
911   \subsubsection{\textbf{Heating}}
912  
913 < Typically, Heating is performed by assigning random velocities
913 > Typically, heating is performed by assigning random velocities
914   according to a Maxwell-Boltzman distribution for a desired
915   temperature. Beginning at a lower temperature and gradually
916   increasing the temperature by assigning larger random velocities, we
917 < end up with setting the temperature of the system to a final
918 < temperature at which the simulation will be conducted. In heating
919 < phase, we should also keep the system from drifting or rotating as a
920 < whole. To do this, the net linear momentum and angular momentum of
921 < the system is shifted to zero after each resampling from the Maxwell
922 < -Boltzman distribution.
917 > end up setting the temperature of the system to a final temperature
918 > at which the simulation will be conducted. In heating phase, we
919 > should also keep the system from drifting or rotating as a whole. To
920 > do this, the net linear momentum and angular momentum of the system
921 > is shifted to zero after each resampling from the Maxwell -Boltzman
922 > distribution.
923  
924   \subsubsection{\textbf{Equilibration}}
925  
# Line 942 | Line 946 | which making large simulations prohibitive in the abse
946   calculation of non-bonded forces, such as van der Waals force and
947   Coulombic forces \textit{etc}. For a system of $N$ particles, the
948   complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
949 < which making large simulations prohibitive in the absence of any
949 > which makes large simulations prohibitive in the absence of any
950   algorithmic tricks. A natural approach to avoid system size issues
951   is to represent the bulk behavior by a finite number of the
952 < particles. However, this approach will suffer from the surface
953 < effect at the edges of the simulation. To offset this,
954 < \textit{Periodic boundary conditions} (see Fig.~\ref{introFig:pbc})
955 < is developed to simulate bulk properties with a relatively small
956 < number of particles. In this method, the simulation box is
957 < replicated throughout space to form an infinite lattice. During the
958 < simulation, when a particle moves in the primary cell, its image in
959 < other cells move in exactly the same direction with exactly the same
952 > particles. However, this approach will suffer from surface effects
953 > at the edges of the simulation. To offset this, \textit{Periodic
954 > boundary conditions} (see Fig.~\ref{introFig:pbc}) were developed to
955 > simulate bulk properties with a relatively small number of
956 > particles. In this method, the simulation box is replicated
957 > throughout space to form an infinite lattice. During the simulation,
958 > when a particle moves in the primary cell, its image in other cells
959 > move in exactly the same direction with exactly the same
960   orientation. Thus, as a particle leaves the primary cell, one of its
961   images will enter through the opposite face.
962   \begin{figure}
# Line 966 | Line 970 | evaluation is to apply spherical cutoff where particle
970  
971   %cutoff and minimum image convention
972   Another important technique to improve the efficiency of force
973 < evaluation is to apply spherical cutoff where particles farther than
974 < a predetermined distance are not included in the calculation
973 > evaluation is to apply spherical cutoffs where particles farther
974 > than a predetermined distance are not included in the calculation
975   \cite{Frenkel1996}. The use of a cutoff radius will cause a
976   discontinuity in the potential energy curve. Fortunately, one can
977 < shift simple radial potential to ensure the potential curve go
977 > shift a simple radial potential to ensure the potential curve go
978   smoothly to zero at the cutoff radius. The cutoff strategy works
979   well for Lennard-Jones interaction because of its short range
980   nature. However, simply truncating the electrostatic interaction
# Line 1016 | Line 1020 | trajectory analysis are more useful. According to the
1020   Recently, advanced visualization technique have become applied to
1021   monitor the motions of molecules. Although the dynamics of the
1022   system can be described qualitatively from animation, quantitative
1023 < trajectory analysis are more useful. According to the principles of
1023 > trajectory analysis is more useful. According to the principles of
1024   Statistical Mechanics in
1025   Sec.~\ref{introSection:statisticalMechanics}, one can compute
1026   thermodynamic properties, analyze fluctuations of structural
# Line 1051 | Line 1055 | Experimentally, pair distribution function can be gath
1055   distribution functions. Among these functions,the \emph{pair
1056   distribution function}, also known as \emph{radial distribution
1057   function}, is of most fundamental importance to liquid theory.
1058 < Experimentally, pair distribution function can be gathered by
1058 > Experimentally, pair distribution functions can be gathered by
1059   Fourier transforming raw data from a series of neutron diffraction
1060   experiments and integrating over the surface factor
1061   \cite{Powles1973}. The experimental results can serve as a criterion
1062   to justify the correctness of a liquid model. Moreover, various
1063   equilibrium thermodynamic and structural properties can also be
1064 < expressed in terms of radial distribution function \cite{Allen1987}.
1065 < The pair distribution functions $g(r)$ gives the probability that a
1066 < particle $i$ will be located at a distance $r$ from a another
1067 < particle $j$ in the system
1064 > expressed in terms of the radial distribution function
1065 > \cite{Allen1987}. The pair distribution functions $g(r)$ gives the
1066 > probability that a particle $i$ will be located at a distance $r$
1067 > from a another particle $j$ in the system
1068   \begin{equation}
1069   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1070   \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
# Line 1092 | Line 1096 | function, which is averaging over time origins and ove
1096   \right\rangle } dt
1097   \]
1098   where $D$ is diffusion constant. Unlike the velocity autocorrelation
1099 < function, which is averaging over time origins and over all the
1100 < atoms, the dipole autocorrelation functions are calculated for the
1099 > function, which is averaged over time origins and over all the
1100 > atoms, the dipole autocorrelation functions is calculated for the
1101   entire system. The dipole autocorrelation function is given by:
1102   \[
1103   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
# Line 1104 | Line 1108 | In principle, many time correlation functions can be r
1108   \[
1109   u_{tot} (t) = \sum\limits_i {u_i (t)}.
1110   \]
1111 < In principle, many time correlation functions can be related with
1111 > In principle, many time correlation functions can be related to
1112   Fourier transforms of the infrared, Raman, and inelastic neutron
1113   scattering spectra of molecular liquids. In practice, one can
1114 < extract the IR spectrum from the intensity of dipole fluctuation at
1115 < each frequency using the following relationship:
1114 > extract the IR spectrum from the intensity of the molecular dipole
1115 > fluctuation at each frequency using the following relationship:
1116   \[
1117   \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1118   i2\pi vt} dt}.
# Line 1118 | Line 1122 | missiles and vehicle are usually modeled by rigid bodi
1122  
1123   Rigid bodies are frequently involved in the modeling of different
1124   areas, from engineering, physics, to chemistry. For example,
1125 < missiles and vehicle are usually modeled by rigid bodies.  The
1126 < movement of the objects in 3D gaming engine or other physics
1127 < simulator is governed by rigid body dynamics. In molecular
1125 > missiles and vehicles are usually modeled by rigid bodies.  The
1126 > movement of the objects in 3D gaming engines or other physics
1127 > simulators is governed by rigid body dynamics. In molecular
1128   simulations, rigid bodies are used to simplify protein-protein
1129   docking studies\cite{Gray2003}.
1130  
# Line 1129 | Line 1133 | equations of motion is very inefficient and inaccurate
1133   freedom. Euler angles are the natural choice to describe the
1134   rotational degrees of freedom. However, due to $\frac {1}{sin
1135   \theta}$ singularities, the numerical integration of corresponding
1136 < equations of motion is very inefficient and inaccurate. Although an
1137 < alternative integrator using multiple sets of Euler angles can
1138 < overcome this difficulty\cite{Barojas1973}, the computational
1139 < penalty and the loss of angular momentum conservation still remain.
1140 < A singularity-free representation utilizing quaternions was
1141 < developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1142 < approach uses a nonseparable Hamiltonian resulting from the
1143 < quaternion representation, which prevents the symplectic algorithm
1144 < to be utilized. Another different approach is to apply holonomic
1145 < constraints to the atoms belonging to the rigid body. Each atom
1146 < moves independently under the normal forces deriving from potential
1147 < energy and constraint forces which are used to guarantee the
1148 < rigidness. However, due to their iterative nature, the SHAKE and
1149 < Rattle algorithms also converge very slowly when the number of
1150 < constraints increases\cite{Ryckaert1977, Andersen1983}.
1136 > equations of these motion is very inefficient and inaccurate.
1137 > Although an alternative integrator using multiple sets of Euler
1138 > angles can overcome this difficulty\cite{Barojas1973}, the
1139 > computational penalty and the loss of angular momentum conservation
1140 > still remain. A singularity-free representation utilizing
1141 > quaternions was developed by Evans in 1977\cite{Evans1977}.
1142 > Unfortunately, this approach uses a nonseparable Hamiltonian
1143 > resulting from the quaternion representation, which prevents the
1144 > symplectic algorithm from being utilized. Another different approach
1145 > is to apply holonomic constraints to the atoms belonging to the
1146 > rigid body. Each atom moves independently under the normal forces
1147 > deriving from potential energy and constraint forces which are used
1148 > to guarantee the rigidness. However, due to their iterative nature,
1149 > the SHAKE and Rattle algorithms also converge very slowly when the
1150 > number of constraints increases\cite{Ryckaert1977, Andersen1983}.
1151  
1152   A break-through in geometric literature suggests that, in order to
1153   develop a long-term integration scheme, one should preserve the
1154 < symplectic structure of the flow. By introducing a conjugate
1154 > symplectic structure of the propagator. By introducing a conjugate
1155   momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1156   equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1157   proposed to evolve the Hamiltonian system in a constraint manifold
# Line 1155 | Line 1159 | symplectic Lie-Poisson integrator for rigid body devel
1159   An alternative method using the quaternion representation was
1160   developed by Omelyan\cite{Omelyan1998}. However, both of these
1161   methods are iterative and inefficient. In this section, we descibe a
1162 < symplectic Lie-Poisson integrator for rigid body developed by
1162 > symplectic Lie-Poisson integrator for rigid bodies developed by
1163   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1164  
1165   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
# Line 1348 | Line 1352 | The flow maps for $T_2^r$ and $T_3^r$ can be found in
1352   e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1353   ).
1354   \]
1355 < The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1355 > The propagator maps for $T_2^r$ and $T_3^r$ can be found in the same
1356   manner. In order to construct a second-order symplectic method, we
1357   split the angular kinetic Hamiltonian function into five terms
1358   \[

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines