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 2692 by tim, Tue Apr 4 21:32:58 2006 UTC vs.
Revision 2706 by tim, Thu Apr 13 04:47:47 2006 UTC

# Line 1 | Line 1
1   \chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND}
2  
3 < \section{\label{introSection:molecularDynamics}Molecular Dynamics}
3 > \section{\label{introSection:classicalMechanics}Classical
4 > Mechanics}
5  
5 As a special discipline of molecular modeling, Molecular dynamics
6 has proven to be a powerful tool for studying the functions of
7 biological systems, providing structural, thermodynamic and
8 dynamical information.
9
10 \subsection{\label{introSection:classicalMechanics}Classical Mechanics}
11
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
# Line 20 | Line 14 | sufficient to predict the future behavior of the syste
14   when further combine with the laws of mechanics will also be
15   sufficient to predict the future behavior of the system.
16  
17 < \subsubsection{\label{introSection:newtonian}Newtonian Mechanics}
17 > \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18 > 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  
38 < \subsubsection{\label{introSection:lagrangian}Lagrangian Mechanics}
38 > 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 > 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  
73 + \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74 +
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
# Line 35 | Line 83 | system, alternative procedures may be developed.
83   which arise in attempts to apply Newton's equation to complex
84   system, alternative procedures may be developed.
85  
86 < \subsubsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
86 > \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
87   Principle}
88  
89   Hamilton introduced the dynamical principle upon which it is
# Line 45 | Line 93 | the kinetic, $K$, and potential energies, $U$.
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 < the kinetic, $K$, and potential energies, $U$.
96 > the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
97   \begin{equation}
98   \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
99 < \lable{introEquation:halmitonianPrinciple1}
99 > \label{introEquation:halmitonianPrinciple1}
100   \end{equation}
101  
102   For simple mechanical systems, where the forces acting on the
# Line 62 | Line 110 | then Eq.~\ref{introEquation:halmitonianPrinciple1} bec
110   \end{equation}
111   then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
112   \begin{equation}
113 < \delta \int_{t_1 }^{t_2 } {K dt = 0} ,
114 < \lable{introEquation:halmitonianPrinciple2}
113 > \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
114 > \label{introEquation:halmitonianPrinciple2}
115   \end{equation}
116  
117 < \subsubsubsection{\label{introSection:equationOfMotionLagrangian}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 }} -
124   \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
125 < \lable{introEquation:eqMotionLagrangian}
125 > \label{introEquation:eqMotionLagrangian}
126   \end{equation}
127   where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
128   generalized velocity.
129  
130 < \subsubsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
130 > \subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
131  
132   Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
133   introduced by William Rowan Hamilton in 1833 as a re-formulation of
# Line 90 | Line 138 | With the help of these momenta, we may now define a ne
138   p_i = \frac{\partial L}{\partial \dot q_i}
139   \label{introEquation:generalizedMomenta}
140   \end{equation}
141 < With the help of these momenta, we may now define a new quantity $H$
94 < by the equation
141 > The Lagrange equations of motion are then expressed by
142   \begin{equation}
143 < H = p_1 \dot q_1  +  \ldots  + p_f \dot q_f  - L,
143 > 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   \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 + 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 + known as the canonical equations of motions \cite{Goldstein01}.
193 +
194   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
# Line 108 | Line 200 | equations.
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 < equations.
203 > equations\cite{Marion90}.
204  
205 + 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 + q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
216 + \end{equation}
217  
218 < \subsubsection{\label{introSection:canonicalTransformation}Canonical Transformation}
218 > \section{\label{introSection:statisticalMechanics}Statistical
219 > Mechanics}
220  
221 < \subsection{\label{introSection:statisticalMechanics}Statistical Mechanics}
117 <
118 < The thermodynamic behaviors and properties  of Molecular Dynamics
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 < \subsubsection{\label{introSection::ensemble}Ensemble}
227 > \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
228  
229 < \subsubsection{\label{introSection:ergodic}The Ergodic Hypothesis}
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 < \subsection{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
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 < \subsection{\label{introSection:correlationFunctions}Correlation Functions}
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} \label{introEquation:NVEPartition}.
319 > \end{equation}
320 > A canonical ensemble(NVT)is an ensemble of systems, each of which
321 > can share its energy with a large heat reservoir. The distribution
322 > of the total energy amongst the possible dynamical states is given
323 > by the partition function,
324 > \begin{equation}
325 > \Omega (N,V,T) = e^{ - \beta A}
326 > \label{introEquation:NVTPartition}
327 > \end{equation}
328 > Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
329 > TS$. Since most experiment are carried out under constant pressure
330 > condition, isothermal-isobaric ensemble(NPT) play a very important
331 > role in molecular simulation. The isothermal-isobaric ensemble allow
332 > the system to exchange energy with a heat bath of temperature $T$
333 > and to change the volume as well. Its partition function is given as
334 > \begin{equation}
335 > \Delta (N,P,T) =  - e^{\beta G}.
336 > \label{introEquation:NPTPartition}
337 > \end{equation}
338 > Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
339 >
340 > \subsection{\label{introSection:liouville}Liouville's theorem}
341 >
342 > The Liouville's theorem is the foundation on which statistical
343 > mechanics rests. It describes the time evolution of phase space
344 > distribution function. In order to calculate the rate of change of
345 > $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
346 > consider the two faces perpendicular to the $q_1$ axis, which are
347 > located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
348 > leaving the opposite face is given by the expression,
349 > \begin{equation}
350 > \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
351 > \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
352 > }}\delta q_1 } \right)\delta q_2  \ldots \delta q_f \delta p_1
353 > \ldots \delta p_f .
354 > \end{equation}
355 > Summing all over the phase space, we obtain
356 > \begin{equation}
357 > \frac{{d(\delta N)}}{{dt}} =  - \sum\limits_{i = 1}^f {\left[ {\rho
358 > \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
359 > \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
360 > {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  + \frac{{\partial
361 > \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
362 > \ldots \delta q_f \delta p_1  \ldots \delta p_f .
363 > \end{equation}
364 > Differentiating the equations of motion in Hamiltonian formalism
365 > (\ref{introEquation:motionHamiltonianCoordinate},
366 > \ref{introEquation:motionHamiltonianMomentum}), we can show,
367 > \begin{equation}
368 > \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
369 > + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
370 > \end{equation}
371 > which cancels the first terms of the right hand side. Furthermore,
372 > divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
373 > p_f $ in both sides, we can write out Liouville's theorem in a
374 > simple form,
375 > \begin{equation}
376 > \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
377 > {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  +
378 > \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
379 > \label{introEquation:liouvilleTheorem}
380 > \end{equation}
381 >
382 > Liouville's theorem states that the distribution function is
383 > constant along any trajectory in phase space. In classical
384 > statistical mechanics, since the number of particles in the system
385 > is huge, we may be able to believe the system is stationary,
386 > \begin{equation}
387 > \frac{{\partial \rho }}{{\partial t}} = 0.
388 > \label{introEquation:stationary}
389 > \end{equation}
390 > In such stationary system, the density of distribution $\rho$ can be
391 > connected to the Hamiltonian $H$ through Maxwell-Boltzmann
392 > distribution,
393 > \begin{equation}
394 > \rho  \propto e^{ - \beta H}
395 > \label{introEquation:densityAndHamiltonian}
396 > \end{equation}
397 >
398 > \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
399 > Lets consider a region in the phase space,
400 > \begin{equation}
401 > \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
402 > \end{equation}
403 > If this region is small enough, the density $\rho$ can be regarded
404 > as uniform over the whole phase space. Thus, the number of phase
405 > points inside this region is given by,
406 > \begin{equation}
407 > \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
408 > dp_1 } ..dp_f.
409 > \end{equation}
410 >
411 > \begin{equation}
412 > \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
413 > \frac{d}{{dt}}(\delta v) = 0.
414 > \end{equation}
415 > With the help of stationary assumption
416 > (\ref{introEquation:stationary}), we obtain the principle of the
417 > \emph{conservation of extension in phase space},
418 > \begin{equation}
419 > \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
420 > ...dq_f dp_1 } ..dp_f  = 0.
421 > \label{introEquation:volumePreserving}
422 > \end{equation}
423 >
424 > \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
425 >
426 > Liouville's theorem can be expresses in a variety of different forms
427 > which are convenient within different contexts. For any two function
428 > $F$ and $G$ of the coordinates and momenta of a system, the Poisson
429 > bracket ${F, G}$ is defined as
430 > \begin{equation}
431 > \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
432 > F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
433 > \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
434 > q_i }}} \right)}.
435 > \label{introEquation:poissonBracket}
436 > \end{equation}
437 > Substituting equations of motion in Hamiltonian formalism(
438 > \ref{introEquation:motionHamiltonianCoordinate} ,
439 > \ref{introEquation:motionHamiltonianMomentum} ) into
440 > (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
441 > theorem using Poisson bracket notion,
442 > \begin{equation}
443 > \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
444 > {\rho ,H} \right\}.
445 > \label{introEquation:liouvilleTheromInPoissin}
446 > \end{equation}
447 > Moreover, the Liouville operator is defined as
448 > \begin{equation}
449 > iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
450 > p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
451 > H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
452 > \label{introEquation:liouvilleOperator}
453 > \end{equation}
454 > In terms of Liouville operator, Liouville's equation can also be
455 > expressed as
456 > \begin{equation}
457 > \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
458 > \label{introEquation:liouvilleTheoremInOperator}
459 > \end{equation}
460 >
461 > \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
462 >
463 > Various thermodynamic properties can be calculated from Molecular
464 > Dynamics simulation. By comparing experimental values with the
465 > calculated properties, one can determine the accuracy of the
466 > simulation and the quality of the underlying model. However, both of
467 > experiment and computer simulation are usually performed during a
468 > certain time interval and the measurements are averaged over a
469 > period of them which is different from the average behavior of
470 > many-body system in Statistical Mechanics. Fortunately, Ergodic
471 > Hypothesis is proposed to make a connection between time average and
472 > ensemble average. It states that time average and average over the
473 > statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
474 > \begin{equation}
475 > \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
476 > \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
477 > {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
478 > \end{equation}
479 > where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
480 > physical quantity and $\rho (p(t), q(t))$ is the equilibrium
481 > distribution function. If an observation is averaged over a
482 > sufficiently long time (longer than relaxation time), all accessible
483 > microstates in phase space are assumed to be equally probed, giving
484 > a properly weighted statistical average. This allows the researcher
485 > freedom of choice when deciding how best to measure a given
486 > observable. In case an ensemble averaged approach sounds most
487 > reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
488 > utilized. Or if the system lends itself to a time averaging
489 > approach, the Molecular Dynamics techniques in
490 > Sec.~\ref{introSection:molecularDynamics} will be the best
491 > choice\cite{Frenkel1996}.
492 >
493 > \section{\label{introSection:geometricIntegratos}Geometric Integrators}
494 > A variety of numerical integrators were proposed to simulate the
495 > motions. They usually begin with an initial conditionals and move
496 > the objects in the direction governed by the differential equations.
497 > However, most of them ignore the hidden physical law contained
498 > within the equations. Since 1990, geometric integrators, which
499 > preserve various phase-flow invariants such as symplectic structure,
500 > volume and time reversal symmetry, are developed to address this
501 > issue. The velocity verlet method, which happens to be a simple
502 > example of symplectic integrator, continues to gain its popularity
503 > in molecular dynamics community. This fact can be partly explained
504 > by its geometric nature.
505 >
506 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
507 > A \emph{manifold} is an abstract mathematical space. It locally
508 > looks like Euclidean space, but when viewed globally, it may have
509 > more complicate structure. A good example of manifold is the surface
510 > of Earth. It seems to be flat locally, but it is round if viewed as
511 > a whole. A \emph{differentiable manifold} (also known as
512 > \emph{smooth manifold}) is a manifold with an open cover in which
513 > the covering neighborhoods are all smoothly isomorphic to one
514 > another. In other words,it is possible to apply calculus on
515 > \emph{differentiable manifold}. A \emph{symplectic manifold} is
516 > defined as a pair $(M, \omega)$ which consisting of a
517 > \emph{differentiable manifold} $M$ and a close, non-degenerated,
518 > bilinear symplectic form, $\omega$. A symplectic form on a vector
519 > space $V$ is a function $\omega(x, y)$ which satisfies
520 > $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
521 > \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
522 > $\omega(x, x) = 0$. Cross product operation in vector field is an
523 > example of symplectic form.
524 >
525 > One of the motivations to study \emph{symplectic manifold} in
526 > Hamiltonian Mechanics is that a symplectic manifold can represent
527 > all possible configurations of the system and the phase space of the
528 > system can be described by it's cotangent bundle. Every symplectic
529 > manifold is even dimensional. For instance, in Hamilton equations,
530 > coordinate and momentum always appear in pairs.
531 >
532 > Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
533 > \[
534 > f : M \rightarrow N
535 > \]
536 > is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
537 > the \emph{pullback} of $\eta$ under f is equal to $\omega$.
538 > Canonical transformation is an example of symplectomorphism in
539 > classical mechanics.
540 >
541 > \subsection{\label{introSection:ODE}Ordinary Differential Equations}
542 >
543 > For a ordinary differential system defined as
544 > \begin{equation}
545 > \dot x = f(x)
546 > \end{equation}
547 > where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
548 > \begin{equation}
549 > f(r) = J\nabla _x H(r).
550 > \end{equation}
551 > $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
552 > matrix
553 > \begin{equation}
554 > J = \left( {\begin{array}{*{20}c}
555 >   0 & I  \\
556 >   { - I} & 0  \\
557 > \end{array}} \right)
558 > \label{introEquation:canonicalMatrix}
559 > \end{equation}
560 > where $I$ is an identity matrix. Using this notation, Hamiltonian
561 > system can be rewritten as,
562 > \begin{equation}
563 > \frac{d}{{dt}}x = J\nabla _x H(x)
564 > \label{introEquation:compactHamiltonian}
565 > \end{equation}In this case, $f$ is
566 > called a \emph{Hamiltonian vector field}.
567 >
568 > Another generalization of Hamiltonian dynamics is Poisson Dynamics,
569 > \begin{equation}
570 > \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
571 > \end{equation}
572 > The most obvious change being that matrix $J$ now depends on $x$.
573 > The free rigid body is an example of Poisson system (actually a
574 > Lie-Poisson system) with Hamiltonian function of angular kinetic
575 > energy.
576 > \begin{equation}
577 > J(\pi ) = \left( {\begin{array}{*{20}c}
578 >   0 & {\pi _3 } & { - \pi _2 }  \\
579 >   { - \pi _3 } & 0 & {\pi _1 }  \\
580 >   {\pi _2 } & { - \pi _1 } & 0  \\
581 > \end{array}} \right)
582 > \end{equation}
583 >
584 > \begin{equation}
585 > H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
586 > }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
587 > \end{equation}
588 >
589 > \subsection{\label{introSection:exactFlow}Exact Flow}
590 >
591 > Let $x(t)$ be the exact solution of the ODE system,
592 > \begin{equation}
593 > \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
594 > \end{equation}
595 > The exact flow(solution) $\varphi_\tau$ is defined by
596 > \[
597 > x(t+\tau) =\varphi_\tau(x(t))
598 > \]
599 > where $\tau$ is a fixed time step and $\varphi$ is a map from phase
600 > space to itself. The flow has the continuous group property,
601 > \begin{equation}
602 > \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
603 > + \tau _2 } .
604 > \end{equation}
605 > In particular,
606 > \begin{equation}
607 > \varphi _\tau   \circ \varphi _{ - \tau }  = I
608 > \end{equation}
609 > Therefore, the exact flow is self-adjoint,
610 > \begin{equation}
611 > \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
612 > \end{equation}
613 > The exact flow can also be written in terms of the of an operator,
614 > \begin{equation}
615 > \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
616 > }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
617 > \label{introEquation:exponentialOperator}
618 > \end{equation}
619 >
620 > In most cases, it is not easy to find the exact flow $\varphi_\tau$.
621 > Instead, we use a approximate map, $\psi_\tau$, which is usually
622 > called integrator. The order of an integrator $\psi_\tau$ is $p$, if
623 > the Taylor series of $\psi_\tau$ agree to order $p$,
624 > \begin{equation}
625 > \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
626 > \end{equation}
627 >
628 > \subsection{\label{introSection:geometricProperties}Geometric Properties}
629 >
630 > The hidden geometric properties of ODE and its flow play important
631 > roles in numerical studies. Many of them can be found in systems
632 > which occur naturally in applications.
633 >
634 > Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
635 > a \emph{symplectic} flow if it satisfies,
636 > \begin{equation}
637 > {\varphi '}^T J \varphi ' = J.
638 > \end{equation}
639 > According to Liouville's theorem, the symplectic volume is invariant
640 > under a Hamiltonian flow, which is the basis for classical
641 > statistical mechanics. Furthermore, the flow of a Hamiltonian vector
642 > field on a symplectic manifold can be shown to be a
643 > symplectomorphism. As to the Poisson system,
644 > \begin{equation}
645 > {\varphi '}^T J \varphi ' = J \circ \varphi
646 > \end{equation}
647 > is the property must be preserved by the integrator.
648  
649 + It is possible to construct a \emph{volume-preserving} flow for a
650 + source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
651 + \det d\varphi  = 1$. One can show easily that a symplectic flow will
652 + be volume-preserving.
653 +
654 + Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
655 + will result in a new system,
656 + \[
657 + \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
658 + \]
659 + The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
660 + In other words, the flow of this vector field is reversible if and
661 + only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
662 +
663 + A \emph{first integral}, or conserved quantity of a general
664 + differential function is a function $ G:R^{2d}  \to R^d $ which is
665 + constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
666 + \[
667 + \frac{{dG(x(t))}}{{dt}} = 0.
668 + \]
669 + Using chain rule, one may obtain,
670 + \[
671 + \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
672 + \]
673 + which is the condition for conserving \emph{first integral}. For a
674 + canonical Hamiltonian system, the time evolution of an arbitrary
675 + smooth function $G$ is given by,
676 + \begin{equation}
677 + \begin{array}{c}
678 + \frac{{dG(x(t))}}{{dt}} = [\nabla _x G(x(t))]^T \dot x(t) \\
679 +  = [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
680 + \end{array}
681 + \label{introEquation:firstIntegral1}
682 + \end{equation}
683 + Using poisson bracket notion, Equation
684 + \ref{introEquation:firstIntegral1} can be rewritten as
685 + \[
686 + \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
687 + \]
688 + Therefore, the sufficient condition for $G$ to be the \emph{first
689 + integral} of a Hamiltonian system is
690 + \[
691 + \left\{ {G,H} \right\} = 0.
692 + \]
693 + As well known, the Hamiltonian (or energy) H of a Hamiltonian system
694 + is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
695 + 0$.
696 +
697 +
698 + When designing any numerical methods, one should always try to
699 + preserve the structural properties of the original ODE and its flow.
700 +
701 + \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
702 + A lot of well established and very effective numerical methods have
703 + been successful precisely because of their symplecticities even
704 + though this fact was not recognized when they were first
705 + constructed. The most famous example is leapfrog methods in
706 + molecular dynamics. In general, symplectic integrators can be
707 + constructed using one of four different methods.
708 + \begin{enumerate}
709 + \item Generating functions
710 + \item Variational methods
711 + \item Runge-Kutta methods
712 + \item Splitting methods
713 + \end{enumerate}
714 +
715 + Generating function tends to lead to methods which are cumbersome
716 + and difficult to use. In dissipative systems, variational methods
717 + can capture the decay of energy accurately. Since their
718 + geometrically unstable nature against non-Hamiltonian perturbations,
719 + ordinary implicit Runge-Kutta methods are not suitable for
720 + Hamiltonian system. Recently, various high-order explicit
721 + Runge--Kutta methods have been developed to overcome this
722 + instability. However, due to computational penalty involved in
723 + implementing the Runge-Kutta methods, they do not attract too much
724 + attention from Molecular Dynamics community. Instead, splitting have
725 + been widely accepted since they exploit natural decompositions of
726 + the system\cite{Tuckerman92}.
727 +
728 + \subsubsection{\label{introSection:splittingMethod}Splitting Method}
729 +
730 + The main idea behind splitting methods is to decompose the discrete
731 + $\varphi_h$ as a composition of simpler flows,
732 + \begin{equation}
733 + \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
734 + \varphi _{h_n }
735 + \label{introEquation:FlowDecomposition}
736 + \end{equation}
737 + where each of the sub-flow is chosen such that each represent a
738 + simpler integration of the system.
739 +
740 + Suppose that a Hamiltonian system takes the form,
741 + \[
742 + H = H_1 + H_2.
743 + \]
744 + Here, $H_1$ and $H_2$ may represent different physical processes of
745 + the system. For instance, they may relate to kinetic and potential
746 + energy respectively, which is a natural decomposition of the
747 + problem. If $H_1$ and $H_2$ can be integrated using exact flows
748 + $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
749 + order is then given by the Lie-Trotter formula
750 + \begin{equation}
751 + \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
752 + \label{introEquation:firstOrderSplitting}
753 + \end{equation}
754 + where $\varphi _h$ is the result of applying the corresponding
755 + continuous $\varphi _i$ over a time $h$. By definition, as
756 + $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
757 + must follow that each operator $\varphi_i(t)$ is a symplectic map.
758 + It is easy to show that any composition of symplectic flows yields a
759 + symplectic map,
760 + \begin{equation}
761 + (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
762 + '\phi ' = \phi '^T J\phi ' = J,
763 + \label{introEquation:SymplecticFlowComposition}
764 + \end{equation}
765 + where $\phi$ and $\psi$ both are symplectic maps. Thus operator
766 + splitting in this context automatically generates a symplectic map.
767 +
768 + The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
769 + introduces local errors proportional to $h^2$, while Strang
770 + splitting gives a second-order decomposition,
771 + \begin{equation}
772 + \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
773 + _{1,h/2} , \label{introEquation:secondOrderSplitting}
774 + \end{equation}
775 + which has a local error proportional to $h^3$. Sprang splitting's
776 + popularity in molecular simulation community attribute to its
777 + symmetric property,
778 + \begin{equation}
779 + \varphi _h^{ - 1} = \varphi _{ - h}.
780 + \label{introEquation:timeReversible}
781 + \end{equation}
782 +
783 + \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
784 + The classical equation for a system consisting of interacting
785 + particles can be written in Hamiltonian form,
786 + \[
787 + H = T + V
788 + \]
789 + where $T$ is the kinetic energy and $V$ is the potential energy.
790 + Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
791 + obtains the following:
792 + \begin{align}
793 + q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
794 +    \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
795 + \label{introEquation:Lp10a} \\%
796 + %
797 + \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
798 +    \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
799 + \label{introEquation:Lp10b}
800 + \end{align}
801 + where $F(t)$ is the force at time $t$. This integration scheme is
802 + known as \emph{velocity verlet} which is
803 + symplectic(\ref{introEquation:SymplecticFlowComposition}),
804 + time-reversible(\ref{introEquation:timeReversible}) and
805 + volume-preserving (\ref{introEquation:volumePreserving}). These
806 + geometric properties attribute to its long-time stability and its
807 + popularity in the community. However, the most commonly used
808 + velocity verlet integration scheme is written as below,
809 + \begin{align}
810 + \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
811 +    \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
812 + %
813 + q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
814 +    \label{introEquation:Lp9b}\\%
815 + %
816 + \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
817 +    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
818 + \end{align}
819 + From the preceding splitting, one can see that the integration of
820 + the equations of motion would follow:
821 + \begin{enumerate}
822 + \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
823 +
824 + \item Use the half step velocities to move positions one whole step, $\Delta t$.
825 +
826 + \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
827 +
828 + \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
829 + \end{enumerate}
830 +
831 + Simply switching the order of splitting and composing, a new
832 + integrator, the \emph{position verlet} integrator, can be generated,
833 + \begin{align}
834 + \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
835 + \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
836 + \label{introEquation:positionVerlet1} \\%
837 + %
838 + q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
839 + q(\Delta t)} \right]. %
840 + \label{introEquation:positionVerlet1}
841 + \end{align}
842 +
843 + \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
844 +
845 + Baker-Campbell-Hausdorff formula can be used to determine the local
846 + error of splitting method in terms of commutator of the
847 + operators(\ref{introEquation:exponentialOperator}) associated with
848 + the sub-flow. For operators $hX$ and $hY$ which are associate to
849 + $\varphi_1(t)$ and $\varphi_2(t$ respectively , we have
850 + \begin{equation}
851 + \exp (hX + hY) = \exp (hZ)
852 + \end{equation}
853 + where
854 + \begin{equation}
855 + hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
856 + {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
857 + \end{equation}
858 + Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
859 + \[
860 + [X,Y] = XY - YX .
861 + \]
862 + Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
863 + can obtain
864 + \begin{eqnarray*}
865 + \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
866 + [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
867 + & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
868 + \ldots )
869 + \end{eqnarray*}
870 + Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
871 + error of Spring splitting is proportional to $h^3$. The same
872 + procedure can be applied to general splitting,  of the form
873 + \begin{equation}
874 + \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
875 + 1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
876 + \end{equation}
877 + Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
878 + order method. Yoshida proposed an elegant way to compose higher
879 + order methods based on symmetric splitting. Given a symmetric second
880 + order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
881 + method can be constructed by composing,
882 + \[
883 + \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
884 + h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
885 + \]
886 + where $ \alpha  =  - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
887 + = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
888 + integrator $ \varphi _h^{(2n + 2)}$ can be composed by
889 + \begin{equation}
890 + \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
891 + _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
892 + \end{equation}
893 + , if the weights are chosen as
894 + \[
895 + \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
896 + \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
897 + \]
898 +
899 + \section{\label{introSection:molecularDynamics}Molecular Dynamics}
900 +
901 + As a special discipline of molecular modeling, Molecular dynamics
902 + has proven to be a powerful tool for studying the functions of
903 + biological systems, providing structural, thermodynamic and
904 + dynamical information.
905 +
906 + \subsection{\label{introSec:mdInit}Initialization}
907 +
908 + \subsection{\label{introSec:forceEvaluation}Force Evaluation}
909 +
910 + \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
911 +
912 + \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
913 +
914 + Rigid bodies are frequently involved in the modeling of different
915 + areas, from engineering, physics, to chemistry. For example,
916 + missiles and vehicle are usually modeled by rigid bodies.  The
917 + movement of the objects in 3D gaming engine or other physics
918 + simulator is governed by the rigid body dynamics. In molecular
919 + simulation, rigid body is used to simplify the model in
920 + protein-protein docking study{\cite{Gray03}}.
921 +
922 + It is very important to develop stable and efficient methods to
923 + integrate the equations of motion of orientational degrees of
924 + freedom. Euler angles are the nature choice to describe the
925 + rotational degrees of freedom. However, due to its singularity, the
926 + numerical integration of corresponding equations of motion is very
927 + inefficient and inaccurate. Although an alternative integrator using
928 + different sets of Euler angles can overcome this difficulty\cite{},
929 + the computational penalty and the lost of angular momentum
930 + conservation still remain. A singularity free representation
931 + utilizing quaternions was developed by Evans in 1977. Unfortunately,
932 + this approach suffer from the nonseparable Hamiltonian resulted from
933 + quaternion representation, which prevents the symplectic algorithm
934 + to be utilized. Another different approach is to apply holonomic
935 + constraints to the atoms belonging to the rigid body. Each atom
936 + moves independently under the normal forces deriving from potential
937 + energy and constraint forces which are used to guarantee the
938 + rigidness. However, due to their iterative nature, SHAKE and Rattle
939 + algorithm converge very slowly when the number of constraint
940 + increases.
941 +
942 + The break through in geometric literature suggests that, in order to
943 + develop a long-term integration scheme, one should preserve the
944 + symplectic structure of the flow. Introducing conjugate momentum to
945 + rotation matrix $A$ and re-formulating Hamiltonian's equation, a
946 + symplectic integrator, RSHAKE, was proposed to evolve the
947 + Hamiltonian system in a constraint manifold by iteratively
948 + satisfying the orthogonality constraint $A_t A = 1$. An alternative
949 + method using quaternion representation was developed by Omelyan.
950 + However, both of these methods are iterative and inefficient. In
951 + this section, we will present a symplectic Lie-Poisson integrator
952 + for rigid body developed by Dullweber and his coworkers\cite{}.
953 +
954 + \subsection{\label{introSection:lieAlgebra}Lie Algebra}
955 +
956 + \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
957 +
958 + \begin{equation}
959 + H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
960 + V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
961 + \label{introEquation:RBHamiltonian}
962 + \end{equation}
963 + Here, $q$ and $Q$  are the position and rotation matrix for the
964 + rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
965 + $J$, a diagonal matrix, is defined by
966 + \[
967 + I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
968 + \]
969 + where $I_{ii}$ is the diagonal element of the inertia tensor. This
970 + constrained Hamiltonian equation subjects to a holonomic constraint,
971 + \begin{equation}
972 + Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
973 + \end{equation}
974 + which is used to ensure rotation matrix's orthogonality.
975 + Differentiating \ref{introEquation:orthogonalConstraint} and using
976 + Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
977 + \begin{equation}
978 + Q^t PJ^{ - 1}  + J^{ - 1} P^t Q = 0 . \\
979 + \label{introEquation:RBFirstOrderConstraint}
980 + \end{equation}
981 +
982 + Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
983 + \ref{introEquation:motionHamiltonianMomentum}), one can write down
984 + the equations of motion,
985 + \[
986 + \begin{array}{c}
987 + \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
988 + \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
989 + \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
990 + \frac{{dP}}{{dt}} =  - \nabla _q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
991 + \end{array}
992 + \]
993 +
994 +
995 + \[
996 + M = \left\{ {(Q,P):Q^T Q = 1,Q^t PJ^{ - 1}  + J^{ - 1} P^t Q = 0}
997 + \right\} .
998 + \]
999 +
1000 + \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
1001 +
1002 + \subsection{\label{introSection:symplecticDiscretizationRB}Symplectic Discretization of Euler Equations}
1003 +
1004 +
1005   \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1006  
1007 + \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
1008 +
1009   \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
1010  
1011 < \subsection{\label{introSection:hydroynamics}Hydrodynamics}
1011 > \begin{equation}
1012 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1013 > \label{introEquation:bathGLE}
1014 > \end{equation}
1015 > where $H_B$ is harmonic bath Hamiltonian,
1016 > \[
1017 > H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1018 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1019 > \]
1020 > and $\Delta U$ is bilinear system-bath coupling,
1021 > \[
1022 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1023 > \]
1024 > Completing the square,
1025 > \[
1026 > H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
1027 > {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1028 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1029 > w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
1030 > 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1031 > \]
1032 > and putting it back into Eq.~\ref{introEquation:bathGLE},
1033 > \[
1034 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1035 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1036 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1037 > w_\alpha ^2 }}x} \right)^2 } \right\}}
1038 > \]
1039 > where
1040 > \[
1041 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1042 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1043 > \]
1044 > Since the first two terms of the new Hamiltonian depend only on the
1045 > system coordinates, we can get the equations of motion for
1046 > Generalized Langevin Dynamics by Hamilton's equations
1047 > \ref{introEquation:motionHamiltonianCoordinate,
1048 > introEquation:motionHamiltonianMomentum},
1049 > \begin{align}
1050 > \dot p &=  - \frac{{\partial H}}{{\partial x}}
1051 >       &= m\ddot x
1052 >       &= - \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)}
1053 > \label{introEquation:Lp5}
1054 > \end{align}
1055 > , and
1056 > \begin{align}
1057 > \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
1058 >                &= m\ddot x_\alpha
1059 >                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
1060 > \end{align}
1061 >
1062 > \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1063 >
1064 > \[
1065 > L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1066 > \]
1067 >
1068 > \[
1069 > L(x + y) = L(x) + L(y)
1070 > \]
1071 >
1072 > \[
1073 > L(ax) = aL(x)
1074 > \]
1075 >
1076 > \[
1077 > L(\dot x) = pL(x) - px(0)
1078 > \]
1079 >
1080 > \[
1081 > L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1082 > \]
1083 >
1084 > \[
1085 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1086 > \]
1087 >
1088 > Some relatively important transformation,
1089 > \[
1090 > L(\cos at) = \frac{p}{{p^2  + a^2 }}
1091 > \]
1092 >
1093 > \[
1094 > L(\sin at) = \frac{a}{{p^2  + a^2 }}
1095 > \]
1096 >
1097 > \[
1098 > L(1) = \frac{1}{p}
1099 > \]
1100 >
1101 > First, the bath coordinates,
1102 > \[
1103 > p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
1104 > _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
1105 > }}L(x)
1106 > \]
1107 > \[
1108 > L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1109 > px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1110 > \]
1111 > Then, the system coordinates,
1112 > \begin{align}
1113 > mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1114 > \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1115 > }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1116 > (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1117 > }}\omega _\alpha ^2 L(x)} \right\}}
1118 > %
1119 > &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1120 > \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1121 > - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1122 > - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1123 > \end{align}
1124 > Then, the inverse transform,
1125 >
1126 > \begin{align}
1127 > m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1128 > \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1129 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1130 > _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1131 > - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1132 > (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1133 > _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1134 > %
1135 > &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1136 > {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1137 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1138 > t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1139 > {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1140 > \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1141 > \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1142 > (\omega _\alpha  t)} \right\}}
1143 > \end{align}
1144 >
1145 > \begin{equation}
1146 > m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1147 > (t)\dot x(t - \tau )d\tau }  + R(t)
1148 > \label{introEuqation:GeneralizedLangevinDynamics}
1149 > \end{equation}
1150 > %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1151 > %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1152 > \[
1153 > \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1154 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1155 > \]
1156 > For an infinite harmonic bath, we can use the spectral density and
1157 > an integral over frequencies.
1158 >
1159 > \[
1160 > R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1161 > - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1162 > \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1163 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1164 > \]
1165 > The random forces depend only on initial conditions.
1166 >
1167 > \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1168 > So we can define a new set of coordinates,
1169 > \[
1170 > q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1171 > ^2 }}x(0)
1172 > \]
1173 > This makes
1174 > \[
1175 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1176 > \]
1177 > And since the $q$ coordinates are harmonic oscillators,
1178 > \[
1179 > \begin{array}{l}
1180 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1181 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1182 > \end{array}
1183 > \]
1184 >
1185 > \begin{align}
1186 > \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1187 > {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1188 > (t)q_\beta  (0)} \right\rangle } }
1189 > %
1190 > &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1191 > \right\rangle \cos (\omega _\alpha  t)}
1192 > %
1193 > &= kT\xi (t)
1194 > \end{align}
1195 >
1196 > \begin{equation}
1197 > \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1198 > \label{introEquation:secondFluctuationDissipation}
1199 > \end{equation}
1200 >
1201 > \section{\label{introSection:hydroynamics}Hydrodynamics}
1202 >
1203 > \subsection{\label{introSection:frictionTensor} Friction Tensor}
1204 > \subsection{\label{introSection:analyticalApproach}Analytical
1205 > Approach}
1206 >
1207 > \subsection{\label{introSection:approximationApproach}Approximation
1208 > Approach}
1209 >
1210 > \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1211 > Body}
1212 >
1213 > \section{\label{introSection:correlationFunctions}Correlation Functions}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines