ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
Revision: 2700
Committed: Tue Apr 11 03:38:09 2006 UTC (18 years, 2 months ago) by tim
Content type: application/x-tex
File size: 37269 byte(s)
Log Message:
adding Liouville's theorem section and Phase Space and Ensemble section

File Contents

# User Rev Content
1 tim 2685 \chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND}
2    
3 tim 2693 \section{\label{introSection:classicalMechanics}Classical
4     Mechanics}
5 tim 2685
6 tim 2692 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.
16 tim 2685
17 tim 2693 \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18 tim 2694 The discovery of Newton's three laws of mechanics which govern the
19     motion of particles is the foundation of the classical mechanics.
20     Newton¡¯s first law defines a class of inertial frames. Inertial
21     frames are reference frames where a particle not interacting with
22     other bodies will move with constant speed in the same direction.
23     With respect to inertial frames Newton¡¯s second law has the form
24     \begin{equation}
25     F = \frac {dp}{dt} = \frac {mv}{dt}
26     \label{introEquation:newtonSecondLaw}
27     \end{equation}
28     A point mass interacting with other bodies moves with the
29     acceleration along the direction of the force acting on it. Let
30     $F_ij$ be the force that particle $i$ exerts on particle $j$, and
31     $F_ji$ be the force that particle $j$ exerts on particle $i$.
32     Newton¡¯s third law states that
33     \begin{equation}
34     F_ij = -F_ji
35     \label{introEquation:newtonThirdLaw}
36     \end{equation}
37 tim 2692
38 tim 2694 Conservation laws of Newtonian Mechanics play very important roles
39     in solving mechanics problems. The linear momentum of a particle is
40     conserved if it is free or it experiences no force. The second
41     conservation theorem concerns the angular momentum of a particle.
42     The angular momentum $L$ of a particle with respect to an origin
43     from which $r$ is measured is defined to be
44     \begin{equation}
45     L \equiv r \times p \label{introEquation:angularMomentumDefinition}
46     \end{equation}
47     The torque $\tau$ with respect to the same origin is defined to be
48     \begin{equation}
49     N \equiv r \times F \label{introEquation:torqueDefinition}
50     \end{equation}
51     Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52     \[
53     \dot L = \frac{d}{{dt}}(r \times p) = (\dot r \times p) + (r \times
54     \dot p)
55     \]
56     since
57     \[
58     \dot r \times p = \dot r \times mv = m\dot r \times \dot r \equiv 0
59     \]
60     thus,
61     \begin{equation}
62     \dot L = r \times \dot p = N
63     \end{equation}
64     If there are no external torques acting on a body, the angular
65     momentum of it is conserved. The last conservation theorem state
66 tim 2696 that if all forces are conservative, Energy
67     \begin{equation}E = T + V \label{introEquation:energyConservation}
68     \end{equation}
69     is conserved. All of these conserved quantities are
70     important factors to determine the quality of numerical integration
71     scheme for rigid body \cite{Dullweber1997}.
72 tim 2694
73 tim 2693 \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74 tim 2692
75     Newtonian Mechanics suffers from two important limitations: it
76     describes their motion in special cartesian coordinate systems.
77     Another limitation of Newtonian mechanics becomes obvious when we
78     try to describe systems with large numbers of particles. It becomes
79     very difficult to predict the properties of the system by carrying
80     out calculations involving the each individual interaction between
81     all the particles, even if we know all of the details of the
82     interaction. In order to overcome some of the practical difficulties
83     which arise in attempts to apply Newton's equation to complex
84     system, alternative procedures may be developed.
85    
86 tim 2694 \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
87 tim 2692 Principle}
88    
89     Hamilton introduced the dynamical principle upon which it is
90     possible to base all of mechanics and, indeed, most of classical
91     physics. Hamilton's Principle may be stated as follow,
92    
93     The actual trajectory, along which a dynamical system may move from
94     one point to another within a specified time, is derived by finding
95     the path which minimizes the time integral of the difference between
96 tim 2694 the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
97 tim 2692 \begin{equation}
98     \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
99 tim 2693 \label{introEquation:halmitonianPrinciple1}
100 tim 2692 \end{equation}
101    
102     For simple mechanical systems, where the forces acting on the
103     different part are derivable from a potential and the velocities are
104     small compared with that of light, the Lagrangian function $L$ can
105     be define as the difference between the kinetic energy of the system
106     and its potential energy,
107     \begin{equation}
108     L \equiv K - U = L(q_i ,\dot q_i ) ,
109     \label{introEquation:lagrangianDef}
110     \end{equation}
111     then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
112     \begin{equation}
113 tim 2693 \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
114     \label{introEquation:halmitonianPrinciple2}
115 tim 2692 \end{equation}
116    
117 tim 2694 \subsubsection{\label{introSection:equationOfMotionLagrangian}The
118 tim 2692 Equations of Motion in Lagrangian Mechanics}
119    
120 tim 2700 For a holonomic system of $f$ degrees of freedom, the equations of
121 tim 2692 motion in the Lagrangian form is
122     \begin{equation}
123     \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
124     \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
125 tim 2693 \label{introEquation:eqMotionLagrangian}
126 tim 2692 \end{equation}
127     where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
128     generalized velocity.
129    
130 tim 2693 \subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
131 tim 2692
132     Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
133     introduced by William Rowan Hamilton in 1833 as a re-formulation of
134     classical mechanics. If the potential energy of a system is
135     independent of generalized velocities, the generalized momenta can
136     be defined as
137     \begin{equation}
138     p_i = \frac{\partial L}{\partial \dot q_i}
139     \label{introEquation:generalizedMomenta}
140     \end{equation}
141 tim 2693 The Lagrange equations of motion are then expressed by
142 tim 2692 \begin{equation}
143 tim 2693 p_i = \frac{{\partial L}}{{\partial q_i }}
144     \label{introEquation:generalizedMomentaDot}
145     \end{equation}
146    
147     With the help of the generalized momenta, we may now define a new
148     quantity $H$ by the equation
149     \begin{equation}
150     H = \sum\limits_k {p_k \dot q_k } - L ,
151 tim 2692 \label{introEquation:hamiltonianDefByLagrangian}
152     \end{equation}
153     where $ \dot q_1 \ldots \dot q_f $ are generalized velocities and
154     $L$ is the Lagrangian function for the system.
155    
156 tim 2693 Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
157     one can obtain
158     \begin{equation}
159     dH = \sum\limits_k {\left( {p_k d\dot q_k + \dot q_k dp_k -
160     \frac{{\partial L}}{{\partial q_k }}dq_k - \frac{{\partial
161     L}}{{\partial \dot q_k }}d\dot q_k } \right)} - \frac{{\partial
162     L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
163     \end{equation}
164     Making use of Eq.~\ref{introEquation:generalizedMomenta}, the
165     second and fourth terms in the parentheses cancel. Therefore,
166     Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
167     \begin{equation}
168     dH = \sum\limits_k {\left( {\dot q_k dp_k - \dot p_k dq_k }
169     \right)} - \frac{{\partial L}}{{\partial t}}dt
170     \label{introEquation:diffHamiltonian2}
171     \end{equation}
172     By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
173     find
174     \begin{equation}
175     \frac{{\partial H}}{{\partial p_k }} = q_k
176     \label{introEquation:motionHamiltonianCoordinate}
177     \end{equation}
178     \begin{equation}
179     \frac{{\partial H}}{{\partial q_k }} = - p_k
180     \label{introEquation:motionHamiltonianMomentum}
181     \end{equation}
182     and
183     \begin{equation}
184     \frac{{\partial H}}{{\partial t}} = - \frac{{\partial L}}{{\partial
185     t}}
186     \label{introEquation:motionHamiltonianTime}
187     \end{equation}
188    
189     Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
190     Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
191     equation of motion. Due to their symmetrical formula, they are also
192 tim 2694 known as the canonical equations of motions \cite{Goldstein01}.
193 tim 2693
194 tim 2692 An important difference between Lagrangian approach and the
195     Hamiltonian approach is that the Lagrangian is considered to be a
196     function of the generalized velocities $\dot q_i$ and the
197     generalized coordinates $q_i$, while the Hamiltonian is considered
198     to be a function of the generalized momenta $p_i$ and the conjugate
199     generalized coordinate $q_i$. Hamiltonian Mechanics is more
200     appropriate for application to statistical mechanics and quantum
201     mechanics, since it treats the coordinate and its time derivative as
202     independent variables and it only works with 1st-order differential
203 tim 2694 equations\cite{Marion90}.
204 tim 2692
205 tim 2696 In Newtonian Mechanics, a system described by conservative forces
206     conserves the total energy \ref{introEquation:energyConservation}.
207     It follows that Hamilton's equations of motion conserve the total
208     Hamiltonian.
209     \begin{equation}
210     \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
211     H}}{{\partial q_i }}\dot q_i + \frac{{\partial H}}{{\partial p_i
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 tim 2698 q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
216 tim 2696 \end{equation}
217    
218 tim 2693 \section{\label{introSection:statisticalMechanics}Statistical
219     Mechanics}
220 tim 2692
221 tim 2694 The thermodynamic behaviors and properties of Molecular Dynamics
222 tim 2692 simulation are governed by the principle of Statistical Mechanics.
223     The following section will give a brief introduction to some of the
224 tim 2700 Statistical Mechanics concepts and theorem presented in this
225     dissertation.
226 tim 2692
227 tim 2700 \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
228 tim 2692
229 tim 2700 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 tim 2693 \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
436 tim 2692
437 tim 2695 Various thermodynamic properties can be calculated from Molecular
438     Dynamics simulation. By comparing experimental values with the
439     calculated properties, one can determine the accuracy of the
440     simulation and the quality of the underlying model. However, both of
441     experiment and computer simulation are usually performed during a
442     certain time interval and the measurements are averaged over a
443     period of them which is different from the average behavior of
444     many-body system in Statistical Mechanics. Fortunately, Ergodic
445     Hypothesis is proposed to make a connection between time average and
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 tim 2700 \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 tim 2695 \end{equation}
453 tim 2700 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 tim 2694
467 tim 2697 \section{\label{introSection:geometricIntegratos}Geometric Integrators}
468     A variety of numerical integrators were proposed to simulate the
469     motions. They usually begin with an initial conditionals and move
470     the objects in the direction governed by the differential equations.
471     However, most of them ignore the hidden physical law contained
472     within the equations. Since 1990, geometric integrators, which
473     preserve various phase-flow invariants such as symplectic structure,
474     volume and time reversal symmetry, are developed to address this
475     issue. The velocity verlet method, which happens to be a simple
476     example of symplectic integrator, continues to gain its popularity
477     in molecular dynamics community. This fact can be partly explained
478     by its geometric nature.
479    
480     \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
481     A \emph{manifold} is an abstract mathematical space. It locally
482     looks like Euclidean space, but when viewed globally, it may have
483     more complicate structure. A good example of manifold is the surface
484     of Earth. It seems to be flat locally, but it is round if viewed as
485     a whole. A \emph{differentiable manifold} (also known as
486     \emph{smooth manifold}) is a manifold with an open cover in which
487     the covering neighborhoods are all smoothly isomorphic to one
488     another. In other words,it is possible to apply calculus on
489     \emph{differentiable manifold}. A \emph{symplectic manifold} is
490     defined as a pair $(M, \omega)$ which consisting of a
491     \emph{differentiable manifold} $M$ and a close, non-degenerated,
492     bilinear symplectic form, $\omega$. A symplectic form on a vector
493     space $V$ is a function $\omega(x, y)$ which satisfies
494     $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
495     \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
496     $\omega(x, x) = 0$. Cross product operation in vector field is an
497     example of symplectic form.
498    
499     One of the motivations to study \emph{symplectic manifold} in
500     Hamiltonian Mechanics is that a symplectic manifold can represent
501     all possible configurations of the system and the phase space of the
502     system can be described by it's cotangent bundle. Every symplectic
503     manifold is even dimensional. For instance, in Hamilton equations,
504     coordinate and momentum always appear in pairs.
505    
506     Let $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
507     \[
508     f : M \rightarrow N
509     \]
510     is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
511     the \emph{pullback} of $\eta$ under f is equal to $\omega$.
512     Canonical transformation is an example of symplectomorphism in
513 tim 2698 classical mechanics.
514 tim 2697
515 tim 2698 \subsection{\label{introSection:ODE}Ordinary Differential Equations}
516 tim 2697
517 tim 2698 For a ordinary differential system defined as
518     \begin{equation}
519     \dot x = f(x)
520     \end{equation}
521     where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
522     \begin{equation}
523 tim 2699 f(r) = J\nabla _x H(r).
524 tim 2698 \end{equation}
525     $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
526     matrix
527     \begin{equation}
528     J = \left( {\begin{array}{*{20}c}
529     0 & I \\
530     { - I} & 0 \\
531     \end{array}} \right)
532     \label{introEquation:canonicalMatrix}
533     \end{equation}
534     where $I$ is an identity matrix. Using this notation, Hamiltonian
535     system can be rewritten as,
536     \begin{equation}
537     \frac{d}{{dt}}x = J\nabla _x H(x)
538     \label{introEquation:compactHamiltonian}
539     \end{equation}In this case, $f$ is
540     called a \emph{Hamiltonian vector field}.
541 tim 2697
542 tim 2698 Another generalization of Hamiltonian dynamics is Poisson Dynamics,
543     \begin{equation}
544     \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
545     \end{equation}
546     The most obvious change being that matrix $J$ now depends on $x$.
547     The free rigid body is an example of Poisson system (actually a
548     Lie-Poisson system) with Hamiltonian function of angular kinetic
549     energy.
550     \begin{equation}
551     J(\pi ) = \left( {\begin{array}{*{20}c}
552     0 & {\pi _3 } & { - \pi _2 } \\
553     { - \pi _3 } & 0 & {\pi _1 } \\
554     {\pi _2 } & { - \pi _1 } & 0 \\
555     \end{array}} \right)
556     \end{equation}
557    
558     \begin{equation}
559     H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
560     }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
561     \end{equation}
562    
563     \subsection{\label{introSection:geometricProperties}Geometric Properties}
564     Let $x(t)$ be the exact solution of the ODE system,
565     \begin{equation}
566     \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
567     \end{equation}
568     The exact flow(solution) $\varphi_\tau$ is defined by
569     \[
570     x(t+\tau) =\varphi_\tau(x(t))
571     \]
572     where $\tau$ is a fixed time step and $\varphi$ is a map from phase
573     space to itself. In most cases, it is not easy to find the exact
574     flow $\varphi_\tau$. Instead, we use a approximate map, $\psi_\tau$,
575     which is usually called integrator. The order of an integrator
576     $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
577     order $p$,
578     \begin{equation}
579     \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
580     \end{equation}
581    
582     The hidden geometric properties of ODE and its flow play important
583 tim 2699 roles in numerical studies. Let $\varphi$ be the flow of Hamiltonian
584     vector field, $\varphi$ is a \emph{symplectic} flow if it satisfies,
585 tim 2698 \begin{equation}
586 tim 2699 '\varphi^T J '\varphi = J.
587 tim 2698 \end{equation}
588     According to Liouville's theorem, the symplectic volume is invariant
589     under a Hamiltonian flow, which is the basis for classical
590 tim 2699 statistical mechanics. Furthermore, the flow of a Hamiltonian vector
591     field on a symplectic manifold can be shown to be a
592     symplectomorphism. As to the Poisson system,
593 tim 2698 \begin{equation}
594 tim 2699 '\varphi ^T J '\varphi = J \circ \varphi
595 tim 2698 \end{equation}
596     is the property must be preserved by the integrator. It is possible
597     to construct a \emph{volume-preserving} flow for a source free($
598     \nabla \cdot f = 0 $) ODE, if the flow satisfies $ \det d\varphi =
599     1$. Changing the variables $y = h(x)$ in a
600     ODE\ref{introEquation:ODE} will result in a new system,
601     \[
602     \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
603     \]
604     The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
605     In other words, the flow of this vector field is reversible if and
606     only if $ h \circ \varphi ^{ - 1} = \varphi \circ h $. When
607     designing any numerical methods, one should always try to preserve
608     the structural properties of the original ODE and its flow.
609    
610 tim 2699 \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
611     A lot of well established and very effective numerical methods have
612     been successful precisely because of their symplecticities even
613     though this fact was not recognized when they were first
614     constructed. The most famous example is leapfrog methods in
615     molecular dynamics. In general, symplectic integrators can be
616     constructed using one of four different methods.
617     \begin{enumerate}
618     \item Generating functions
619     \item Variational methods
620     \item Runge-Kutta methods
621     \item Splitting methods
622     \end{enumerate}
623 tim 2698
624 tim 2699 Generating function tends to lead to methods which are cumbersome
625     and difficult to use\cite{}. In dissipative systems, variational
626     methods can capture the decay of energy accurately\cite{}. Since
627     their geometrically unstable nature against non-Hamiltonian
628     perturbations, ordinary implicit Runge-Kutta methods are not
629     suitable for Hamiltonian system. Recently, various high-order
630     explicit Runge--Kutta methods have been developed to overcome this
631     instability \cite{}. However, due to computational penalty involved
632     in implementing the Runge-Kutta methods, they do not attract too
633     much attention from Molecular Dynamics community. Instead, splitting
634     have been widely accepted since they exploit natural decompositions
635     of the system\cite{Tuckerman92}. The main idea behind splitting
636     methods is to decompose the discrete $\varphi_h$ as a composition of
637     simpler flows,
638     \begin{equation}
639     \varphi _h = \varphi _{h_1 } \circ \varphi _{h_2 } \ldots \circ
640     \varphi _{h_n }
641     \label{introEquation:FlowDecomposition}
642     \end{equation}
643     where each of the sub-flow is chosen such that each represent a
644     simpler integration of the system. Let $\phi$ and $\psi$ both be
645     symplectic maps, it is easy to show that any composition of
646     symplectic flows yields a symplectic map,
647     \begin{equation}
648     (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
649     '\phi ' = \phi '^T J\phi ' = J.
650     \label{introEquation:SymplecticFlowComposition}
651     \end{equation}
652     Suppose that a Hamiltonian system has a form with $H = T + V$
653    
654 tim 2694 \section{\label{introSection:molecularDynamics}Molecular Dynamics}
655    
656     As a special discipline of molecular modeling, Molecular dynamics
657     has proven to be a powerful tool for studying the functions of
658     biological systems, providing structural, thermodynamic and
659     dynamical information.
660    
661     \subsection{\label{introSec:mdInit}Initialization}
662    
663     \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
664    
665 tim 2693 \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
666 tim 2692
667 tim 2694 A rigid body is a body in which the distance between any two given
668     points of a rigid body remains constant regardless of external
669     forces exerted on it. A rigid body therefore conserves its shape
670     during its motion.
671    
672     Applications of dynamics of rigid bodies.
673    
674 tim 2695 \subsection{\label{introSection:lieAlgebra}Lie Algebra}
675 tim 2694
676 tim 2695 \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
677    
678     \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
679    
680 tim 2694 %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
681    
682 tim 2693 \section{\label{introSection:correlationFunctions}Correlation Functions}
683 tim 2692
684 tim 2685 \section{\label{introSection:langevinDynamics}Langevin Dynamics}
685    
686 tim 2696 \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
687    
688 tim 2692 \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
689 tim 2685
690 tim 2696 \begin{equation}
691     H = \frac{{p^2 }}{{2m}} + U(x) + H_B + \Delta U(x,x_1 , \ldots x_N)
692     \label{introEquation:bathGLE}
693     \end{equation}
694     where $H_B$ is harmonic bath Hamiltonian,
695     \[
696     H_B =\sum\limits_{\alpha = 1}^N {\left\{ {\frac{{p_\alpha ^2
697     }}{{2m_\alpha }} + \frac{1}{2}m_\alpha w_\alpha ^2 } \right\}}
698     \]
699     and $\Delta U$ is bilinear system-bath coupling,
700     \[
701     \Delta U = - \sum\limits_{\alpha = 1}^N {g_\alpha x_\alpha x}
702     \]
703     Completing the square,
704     \[
705     H_B + \Delta U = \sum\limits_{\alpha = 1}^N {\left\{
706     {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
707     w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
708     w_\alpha ^2 }}x} \right)^2 } \right\}} - \sum\limits_{\alpha =
709     1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha w_\alpha ^2 }}} x^2
710     \]
711     and putting it back into Eq.~\ref{introEquation:bathGLE},
712     \[
713     H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha = 1}^N
714     {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
715     w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
716     w_\alpha ^2 }}x} \right)^2 } \right\}}
717     \]
718     where
719     \[
720     W(x) = U(x) - \sum\limits_{\alpha = 1}^N {\frac{{g_\alpha ^2
721     }}{{2m_\alpha w_\alpha ^2 }}} x^2
722     \]
723     Since the first two terms of the new Hamiltonian depend only on the
724     system coordinates, we can get the equations of motion for
725     Generalized Langevin Dynamics by Hamilton's equations
726     \ref{introEquation:motionHamiltonianCoordinate,
727     introEquation:motionHamiltonianMomentum},
728     \begin{align}
729     \dot p &= - \frac{{\partial H}}{{\partial x}}
730     &= m\ddot x
731     &= - \frac{{\partial W(x)}}{{\partial x}} - \sum\limits_{\alpha = 1}^N {g_\alpha \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right)}
732     \label{introEq:Lp5}
733     \end{align}
734     , and
735     \begin{align}
736     \dot p_\alpha &= - \frac{{\partial H}}{{\partial x_\alpha }}
737     &= m\ddot x_\alpha
738     &= \- m_\alpha w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha}}{{m_\alpha w_\alpha ^2 }}x} \right)
739     \end{align}
740    
741     \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
742    
743     \[
744     L(x) = \int_0^\infty {x(t)e^{ - pt} dt}
745     \]
746    
747     \[
748     L(x + y) = L(x) + L(y)
749     \]
750    
751     \[
752     L(ax) = aL(x)
753     \]
754    
755     \[
756     L(\dot x) = pL(x) - px(0)
757     \]
758    
759     \[
760     L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
761     \]
762    
763     \[
764     L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
765     \]
766    
767     Some relatively important transformation,
768     \[
769     L(\cos at) = \frac{p}{{p^2 + a^2 }}
770     \]
771    
772     \[
773     L(\sin at) = \frac{a}{{p^2 + a^2 }}
774     \]
775    
776     \[
777     L(1) = \frac{1}{p}
778     \]
779    
780     First, the bath coordinates,
781     \[
782     p^2 L(x_\alpha ) - px_\alpha (0) - \dot x_\alpha (0) = - \omega
783     _\alpha ^2 L(x_\alpha ) + \frac{{g_\alpha }}{{\omega _\alpha
784     }}L(x)
785     \]
786     \[
787     L(x_\alpha ) = \frac{{\frac{{g_\alpha }}{{\omega _\alpha }}L(x) +
788     px_\alpha (0) + \dot x_\alpha (0)}}{{p^2 + \omega _\alpha ^2 }}
789     \]
790     Then, the system coordinates,
791     \begin{align}
792     mL(\ddot x) &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
793     \sum\limits_{\alpha = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
794     }}{{\omega _\alpha }}L(x) + px_\alpha (0) + \dot x_\alpha
795     (0)}}{{p^2 + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
796     }}\omega _\alpha ^2 L(x)} \right\}}
797     %
798     &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
799     \sum\limits_{\alpha = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}\frac{p}{{p^2 + \omega _\alpha ^2 }}pL(x)
800     - \frac{p}{{p^2 + \omega _\alpha ^2 }}g_\alpha x_\alpha (0)
801     - \frac{1}{{p^2 + \omega _\alpha ^2 }}g_\alpha \dot x_\alpha (0)} \right\}}
802     \end{align}
803     Then, the inverse transform,
804    
805     \begin{align}
806     m\ddot x &= - \frac{{\partial W(x)}}{{\partial x}} -
807     \sum\limits_{\alpha = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
808     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
809     _\alpha t)\dot x(t - \tau )d\tau - \left[ {g_\alpha x_\alpha (0)
810     - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}} \right]\cos
811     (\omega _\alpha t) - \frac{{g_\alpha \dot x_\alpha (0)}}{{\omega
812     _\alpha }}\sin (\omega _\alpha t)} } \right\}}
813     %
814     &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
815     {\sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
816     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
817     t)\dot x(t - \tau )d} \tau } + \sum\limits_{\alpha = 1}^N {\left\{
818     {\left[ {g_\alpha x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha
819     \omega _\alpha }}} \right]\cos (\omega _\alpha t) +
820     \frac{{g_\alpha \dot x_\alpha (0)}}{{\omega _\alpha }}\sin
821     (\omega _\alpha t)} \right\}}
822     \end{align}
823    
824     \begin{equation}
825     m\ddot x = - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
826     (t)\dot x(t - \tau )d\tau } + R(t)
827     \label{introEuqation:GeneralizedLangevinDynamics}
828     \end{equation}
829     %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
830     %$W$ is the potential of mean force. $W(x) = - kT\ln p(x)$
831     \[
832     \xi (t) = \sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
833     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha t)}
834     \]
835     For an infinite harmonic bath, we can use the spectral density and
836     an integral over frequencies.
837    
838     \[
839     R(t) = \sum\limits_{\alpha = 1}^N {\left( {g_\alpha x_\alpha (0)
840     - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}x(0)}
841     \right)\cos (\omega _\alpha t)} + \frac{{\dot x_\alpha
842     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)
843     \]
844     The random forces depend only on initial conditions.
845    
846     \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
847     So we can define a new set of coordinates,
848     \[
849     q_\alpha (t) = x_\alpha (t) - \frac{1}{{m_\alpha \omega _\alpha
850     ^2 }}x(0)
851     \]
852     This makes
853     \[
854     R(t) = \sum\limits_{\alpha = 1}^N {g_\alpha q_\alpha (t)}
855     \]
856     And since the $q$ coordinates are harmonic oscillators,
857     \[
858     \begin{array}{l}
859     \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t) \\
860     \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle = \delta _{\alpha \beta } \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle \\
861     \end{array}
862     \]
863    
864     \begin{align}
865     \left\langle {R(t)R(0)} \right\rangle &= \sum\limits_\alpha
866     {\sum\limits_\beta {g_\alpha g_\beta \left\langle {q_\alpha
867     (t)q_\beta (0)} \right\rangle } }
868     %
869     &= \sum\limits_\alpha {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
870     \right\rangle \cos (\omega _\alpha t)}
871     %
872     &= kT\xi (t)
873     \end{align}
874    
875     \begin{equation}
876     \xi (t) = \left\langle {R(t)R(0)} \right\rangle
877     \label{introEquation:secondFluctuationDissipation}
878     \end{equation}
879    
880     \section{\label{introSection:hydroynamics}Hydrodynamics}
881    
882     \subsection{\label{introSection:frictionTensor} Friction Tensor}
883     \subsection{\label{introSection:analyticalApproach}Analytical
884     Approach}
885    
886     \subsection{\label{introSection:approximationApproach}Approximation
887     Approach}
888    
889     \subsection{\label{introSection:centersRigidBody}Centers of Rigid
890     Body}