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 2695 by tim, Thu Apr 6 04:07:57 2006 UTC vs.
Revision 2844 by tim, Fri Jun 9 19:06:37 2006 UTC

# Line 6 | Line 6 | behind classical mechanics. Firstly, One can determine
6   Closely related to Classical Mechanics, Molecular Dynamics
7   simulations are carried out by integrating the equations of motion
8   for a given system of particles. There are three fundamental ideas
9 < behind classical mechanics. Firstly, One can determine the state of
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
# Line 17 | Line 17 | Newton¡¯s first law defines a class of inertial frames
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
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
23 > With respect to inertial frames, Newton's second law has the form
24   \begin{equation}
25 < F = \frac {dp}{dt} = \frac {mv}{dt}
25 > F = \frac {dp}{dt} = \frac {mdv}{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
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
34 > F_{ij} = -F_{ji}
35   \label{introEquation:newtonThirdLaw}
36   \end{equation}
37  
# Line 46 | Line 46 | N \equiv r \times F \label{introEquation:torqueDefinit
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}
49 > \tau \equiv r \times F \label{introEquation:torqueDefinition}
50   \end{equation}
51   Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52   \[
# Line 59 | Line 59 | thus,
59   \]
60   thus,
61   \begin{equation}
62 < \dot L = r \times \dot p = N
62 > \dot L = r \times \dot p = \tau
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 $E = T + V$ is
67 < conserved. All of these conserved quantities are important factors
68 < to determine the quality of numerical integration scheme for rigid
69 < body \cite{Dullweber1997}.
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 > schemes for rigid bodies \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
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
80 < interaction. In order to overcome some of the practical difficulties
81 < which arise in attempts to apply Newton's equation to complex
82 < system, alternative procedures may be developed.
75 > Newtonian Mechanics suffers from two important limitations: motions
76 > can only be described in cartesian coordinate systems. Moreover, It
77 > become impossible to predict analytically the properties of the
78 > system even if we know all of the details of the interaction. In
79 > order to overcome some of the practical difficulties which arise in
80 > attempts to apply Newton's equation to complex system, approximate
81 > numerical procedures may be developed.
82  
83 < \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
84 < Principle}
83 > \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
84 > Principle}}
85  
86   Hamilton introduced the dynamical principle upon which it is
87 < possible to base all of mechanics and, indeed, most of classical
88 < physics. Hamilton's Principle may be stated as follow,
87 > possible to base all of mechanics and most of classical physics.
88 > Hamilton's Principle may be stated as follows,
89  
90   The actual trajectory, along which a dynamical system may move from
91   one point to another within a specified time, is derived by finding
92   the path which minimizes the time integral of the difference between
93 < the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
93 > the kinetic, $K$, and potential energies, $U$.
94   \begin{equation}
95   \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
96   \label{introEquation:halmitonianPrinciple1}
97   \end{equation}
98  
99   For simple mechanical systems, where the forces acting on the
100 < different part are derivable from a potential and the velocities are
101 < small compared with that of light, the Lagrangian function $L$ can
102 < be define as the difference between the kinetic energy of the system
104 < and its potential energy,
100 > different parts are derivable from a potential, the Lagrangian
101 > function $L$ can be defined as the difference between the kinetic
102 > energy of the system and its potential energy,
103   \begin{equation}
104   L \equiv K - U = L(q_i ,\dot q_i ) ,
105   \label{introEquation:lagrangianDef}
# Line 112 | Line 110 | then Eq.~\ref{introEquation:halmitonianPrinciple1} bec
110   \label{introEquation:halmitonianPrinciple2}
111   \end{equation}
112  
113 < \subsubsection{\label{introSection:equationOfMotionLagrangian}The
114 < Equations of Motion in Lagrangian Mechanics}
113 > \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
114 > Equations of Motion in Lagrangian Mechanics}}
115  
116 < for a holonomic system of $f$ degrees of freedom, the equations of
116 > For a holonomic system of $f$ degrees of freedom, the equations of
117   motion in the Lagrangian form is
118   \begin{equation}
119   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
# Line 130 | Line 128 | independent of generalized velocities, the generalized
128   Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
129   introduced by William Rowan Hamilton in 1833 as a re-formulation of
130   classical mechanics. If the potential energy of a system is
131 < independent of generalized velocities, the generalized momenta can
134 < be defined as
131 > independent of velocities, the momenta can be defined as
132   \begin{equation}
133   p_i = \frac{\partial L}{\partial \dot q_i}
134   \label{introEquation:generalizedMomenta}
# Line 170 | Line 167 | find
167   By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
168   find
169   \begin{equation}
170 < \frac{{\partial H}}{{\partial p_k }} = q_k
170 > \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
171   \label{introEquation:motionHamiltonianCoordinate}
172   \end{equation}
173   \begin{equation}
174 < \frac{{\partial H}}{{\partial q_k }} =  - p_k
174 > \frac{{\partial H}}{{\partial q_k }} =  - \dot {p_k}
175   \label{introEquation:motionHamiltonianMomentum}
176   \end{equation}
177   and
# Line 187 | Line 184 | known as the canonical equations of motions \cite{Gold
184   Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
185   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
186   equation of motion. Due to their symmetrical formula, they are also
187 < known as the canonical equations of motions \cite{Goldstein01}.
187 > known as the canonical equations of motions \cite{Goldstein2001}.
188  
189   An important difference between Lagrangian approach and the
190   Hamiltonian approach is that the Lagrangian is considered to be a
191 < function of the generalized velocities $\dot q_i$ and the
192 < generalized coordinates $q_i$, while the Hamiltonian is considered
193 < to be a function of the generalized momenta $p_i$ and the conjugate
194 < generalized coordinate $q_i$. Hamiltonian Mechanics is more
195 < appropriate for application to statistical mechanics and quantum
196 < mechanics, since it treats the coordinate and its time derivative as
197 < independent variables and it only works with 1st-order differential
201 < equations\cite{Marion90}.
191 > function of the generalized velocities $\dot q_i$ and coordinates
192 > $q_i$, while the Hamiltonian is considered to be a function of the
193 > generalized momenta $p_i$ and the conjugate coordinates $q_i$.
194 > Hamiltonian Mechanics is more appropriate for application to
195 > statistical mechanics and quantum mechanics, since it treats the
196 > coordinate and its time derivative as independent variables and it
197 > only works with 1st-order differential equations\cite{Marion1990}.
198  
199 < When studying Hamiltonian system, it is more convenient to use
200 < notation
199 > In Newtonian Mechanics, a system described by conservative forces
200 > conserves the total energy \ref{introEquation:energyConservation}.
201 > It follows that Hamilton's equations of motion conserve the total
202 > Hamiltonian.
203   \begin{equation}
204 < r = r(q,p)^T
204 > \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
205 > H}}{{\partial q_i }}\dot q_i  + \frac{{\partial H}}{{\partial p_i
206 > }}\dot p_i } \right)}  = \sum\limits_i {\left( {\frac{{\partial
207 > H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
208 > \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
209 > q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
210   \end{equation}
208 and to introduce a $2n \times 2n$ canonical structure matrix $J$,
209 \begin{equation}
210 J = \left( {\begin{array}{*{20}c}
211   0 & I  \\
212   { - I} & 0  \\
213 \end{array}} \right)
214 \label{introEquation:canonicalMatrix}
215 \end{equation}
216 where $I$ is a $n \times n$ identity matrix and $J$ is a
217 skew-symmetric matrix ($ J^T  =  - J $). Thus, Hamiltonian system
218 can be rewritten as,
219 \begin{equation}
220 \frac{d}{{dt}}r = J\nabla _r H(r)
221 \label{introEquation:compactHamiltonian}
222 \end{equation}
223
224 %\subsection{\label{introSection:canonicalTransformation}Canonical
225 %Transformation}
226
227 \section{\label{introSection:geometricIntegratos}Geometric Integrators}
228
229 \subsection{\label{introSection:symplecticMaps}Symplectic Maps and Methods}
230
231 \subsection{\label{Construction of Symplectic Methods}}
211  
212   \section{\label{introSection:statisticalMechanics}Statistical
213   Mechanics}
# Line 236 | Line 215 | Statistical Mechanics concepts presented in this disse
215   The thermodynamic behaviors and properties of Molecular Dynamics
216   simulation are governed by the principle of Statistical Mechanics.
217   The following section will give a brief introduction to some of the
218 < Statistical Mechanics concepts presented in this dissertation.
218 > Statistical Mechanics concepts and theorem presented in this
219 > dissertation.
220 >
221 > \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
222 >
223 > Mathematically, phase space is the space which represents all
224 > possible states. Each possible state of the system corresponds to
225 > one unique point in the phase space. For mechanical systems, the
226 > phase space usually consists of all possible values of position and
227 > momentum variables. Consider a dynamic system of $f$ particles in a
228 > cartesian space, where each of the $6f$ coordinates and momenta is
229 > assigned to one of $6f$ mutually orthogonal axes, the phase space of
230 > this system is a $6f$ dimensional space. A point, $x = (q_1 , \ldots
231 > ,q_f ,p_1 , \ldots ,p_f )$, with a unique set of values of $6f$
232 > coordinates and momenta is a phase space vector.
233 >
234 > A microscopic state or microstate of a classical system is
235 > specification of the complete phase space vector of a system at any
236 > instant in time. An ensemble is defined as a collection of systems
237 > sharing one or more macroscopic characteristics but each being in a
238 > unique microstate. The complete ensemble is specified by giving all
239 > systems or microstates consistent with the common macroscopic
240 > characteristics of the ensemble. Although the state of each
241 > individual system in the ensemble could be precisely described at
242 > any instance in time by a suitable phase space vector, when using
243 > ensembles for statistical purposes, there is no need to maintain
244 > distinctions between individual systems, since the numbers of
245 > systems at any time in the different states which correspond to
246 > different regions of the phase space are more interesting. Moreover,
247 > in the point of view of statistical mechanics, one would prefer to
248 > use ensembles containing a large enough population of separate
249 > members so that the numbers of systems in such different states can
250 > be regarded as changing continuously as we traverse different
251 > regions of the phase space. The condition of an ensemble at any time
252 > can be regarded as appropriately specified by the density $\rho$
253 > with which representative points are distributed over the phase
254 > space. The density distribution for an ensemble with $f$ degrees of
255 > freedom is defined as,
256 > \begin{equation}
257 > \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
258 > \label{introEquation:densityDistribution}
259 > \end{equation}
260 > Governed by the principles of mechanics, the phase points change
261 > their locations which would change the density at any time at phase
262 > space. Hence, the density distribution is also to be taken as a
263 > function of the time.
264 >
265 > The number of systems $\delta N$ at time $t$ can be determined by,
266 > \begin{equation}
267 > \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
268 > \label{introEquation:deltaN}
269 > \end{equation}
270 > Assuming a large enough population of systems, we can sufficiently
271 > approximate $\delta N$ without introducing discontinuity when we go
272 > from one region in the phase space to another. By integrating over
273 > the whole phase space,
274 > \begin{equation}
275 > N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
276 > \label{introEquation:totalNumberSystem}
277 > \end{equation}
278 > gives us an expression for the total number of the systems. Hence,
279 > the probability per unit in the phase space can be obtained by,
280 > \begin{equation}
281 > \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
282 > {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
283 > \label{introEquation:unitProbability}
284 > \end{equation}
285 > With the help of Equation(\ref{introEquation:unitProbability}) and
286 > the knowledge of the system, it is possible to calculate the average
287 > value of any desired quantity which depends on the coordinates and
288 > momenta of the system. Even when the dynamics of the real system is
289 > complex, or stochastic, or even discontinuous, the average
290 > properties of the ensemble of possibilities as a whole remaining
291 > well defined. For a classical system in thermal equilibrium with its
292 > environment, the ensemble average of a mechanical quantity, $\langle
293 > A(q , p) \rangle_t$, takes the form of an integral over the phase
294 > space of the system,
295 > \begin{equation}
296 > \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
297 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
298 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
299 > \label{introEquation:ensembelAverage}
300 > \end{equation}
301 >
302 > There are several different types of ensembles with different
303 > statistical characteristics. As a function of macroscopic
304 > parameters, such as temperature \textit{etc}, the partition function
305 > can be used to describe the statistical properties of a system in
306 > thermodynamic equilibrium.
307 >
308 > As an ensemble of systems, each of which is known to be thermally
309 > isolated and conserve energy, the Microcanonical ensemble(NVE) has a
310 > partition function like,
311 > \begin{equation}
312 > \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
313 > \end{equation}
314 > A canonical ensemble(NVT)is an ensemble of systems, each of which
315 > can share its energy with a large heat reservoir. The distribution
316 > of the total energy amongst the possible dynamical states is given
317 > by the partition function,
318 > \begin{equation}
319 > \Omega (N,V,T) = e^{ - \beta A}
320 > \label{introEquation:NVTPartition}
321 > \end{equation}
322 > Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
323 > TS$. Since most experiments are carried out under constant pressure
324 > condition, the isothermal-isobaric ensemble(NPT) plays a very
325 > important role in molecular simulations. The isothermal-isobaric
326 > ensemble allow the system to exchange energy with a heat bath of
327 > temperature $T$ and to change the volume as well. Its partition
328 > function is given as
329 > \begin{equation}
330 > \Delta (N,P,T) =  - e^{\beta G}.
331 > \label{introEquation:NPTPartition}
332 > \end{equation}
333 > Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
334 >
335 > \subsection{\label{introSection:liouville}Liouville's theorem}
336 >
337 > Liouville's theorem is the foundation on which statistical mechanics
338 > rests. It describes the time evolution of the phase space
339 > distribution function. In order to calculate the rate of change of
340 > $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
341 > consider the two faces perpendicular to the $q_1$ axis, which are
342 > located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
343 > leaving the opposite face is given by the expression,
344 > \begin{equation}
345 > \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
346 > \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
347 > }}\delta q_1 } \right)\delta q_2  \ldots \delta q_f \delta p_1
348 > \ldots \delta p_f .
349 > \end{equation}
350 > Summing all over the phase space, we obtain
351 > \begin{equation}
352 > \frac{{d(\delta N)}}{{dt}} =  - \sum\limits_{i = 1}^f {\left[ {\rho
353 > \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
354 > \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
355 > {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  + \frac{{\partial
356 > \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
357 > \ldots \delta q_f \delta p_1  \ldots \delta p_f .
358 > \end{equation}
359 > Differentiating the equations of motion in Hamiltonian formalism
360 > (\ref{introEquation:motionHamiltonianCoordinate},
361 > \ref{introEquation:motionHamiltonianMomentum}), we can show,
362 > \begin{equation}
363 > \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
364 > + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
365 > \end{equation}
366 > which cancels the first terms of the right hand side. Furthermore,
367 > dividing $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
368 > p_f $ in both sides, we can write out Liouville's theorem in a
369 > simple form,
370 > \begin{equation}
371 > \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
372 > {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  +
373 > \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
374 > \label{introEquation:liouvilleTheorem}
375 > \end{equation}
376  
377 < \subsection{\label{introSection::ensemble}Ensemble and Phase Space}
377 > Liouville's theorem states that the distribution function is
378 > constant along any trajectory in phase space. In classical
379 > statistical mechanics, since the number of particles in the system
380 > is huge, we may be able to believe the system is stationary,
381 > \begin{equation}
382 > \frac{{\partial \rho }}{{\partial t}} = 0.
383 > \label{introEquation:stationary}
384 > \end{equation}
385 > In such stationary system, the density of distribution $\rho$ can be
386 > connected to the Hamiltonian $H$ through Maxwell-Boltzmann
387 > distribution,
388 > \begin{equation}
389 > \rho  \propto e^{ - \beta H}
390 > \label{introEquation:densityAndHamiltonian}
391 > \end{equation}
392  
393 + \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
394 + Lets consider a region in the phase space,
395 + \begin{equation}
396 + \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
397 + \end{equation}
398 + If this region is small enough, the density $\rho$ can be regarded
399 + as uniform over the whole integral. Thus, the number of phase points
400 + inside this region is given by,
401 + \begin{equation}
402 + \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
403 + dp_1 } ..dp_f.
404 + \end{equation}
405 +
406 + \begin{equation}
407 + \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
408 + \frac{d}{{dt}}(\delta v) = 0.
409 + \end{equation}
410 + With the help of stationary assumption
411 + (\ref{introEquation:stationary}), we obtain the principle of the
412 + \emph{conservation of volume in phase space},
413 + \begin{equation}
414 + \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
415 + ...dq_f dp_1 } ..dp_f  = 0.
416 + \label{introEquation:volumePreserving}
417 + \end{equation}
418 +
419 + \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
420 +
421 + Liouville's theorem can be expresses in a variety of different forms
422 + which are convenient within different contexts. For any two function
423 + $F$ and $G$ of the coordinates and momenta of a system, the Poisson
424 + bracket ${F, G}$ is defined as
425 + \begin{equation}
426 + \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
427 + F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
428 + \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
429 + q_i }}} \right)}.
430 + \label{introEquation:poissonBracket}
431 + \end{equation}
432 + Substituting equations of motion in Hamiltonian formalism(
433 + \ref{introEquation:motionHamiltonianCoordinate} ,
434 + \ref{introEquation:motionHamiltonianMomentum} ) into
435 + (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
436 + theorem using Poisson bracket notion,
437 + \begin{equation}
438 + \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
439 + {\rho ,H} \right\}.
440 + \label{introEquation:liouvilleTheromInPoissin}
441 + \end{equation}
442 + Moreover, the Liouville operator is defined as
443 + \begin{equation}
444 + iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
445 + p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
446 + H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
447 + \label{introEquation:liouvilleOperator}
448 + \end{equation}
449 + In terms of Liouville operator, Liouville's equation can also be
450 + expressed as
451 + \begin{equation}
452 + \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
453 + \label{introEquation:liouvilleTheoremInOperator}
454 + \end{equation}
455 +
456   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
457  
458   Various thermodynamic properties can be calculated from Molecular
459   Dynamics simulation. By comparing experimental values with the
460   calculated properties, one can determine the accuracy of the
461 < simulation and the quality of the underlying model. However, both of
462 < experiment and computer simulation are usually performed during a
461 > simulation and the quality of the underlying model. However, both
462 > experiments and computer simulations are usually performed during a
463   certain time interval and the measurements are averaged over a
464   period of them which is different from the average behavior of
465 < many-body system in Statistical Mechanics. Fortunately, Ergodic
466 < Hypothesis is proposed to make a connection between time average and
467 < ensemble average. It states that time average and average over the
468 < statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
465 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
466 > Hypothesis makes a connection between time average and the ensemble
467 > average. It states that the time average and average over the
468 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
469   \begin{equation}
470 < \langle A \rangle_t = \mathop {\lim }\limits_{t \to \infty }
471 < \frac{1}{t}\int\limits_0^t {A(p(t),q(t))dt = \int\limits_\Gamma
472 < {A(p(t),q(t))} } \rho (p(t), q(t)) dpdq
470 > \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
471 > \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
472 > {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
473   \end{equation}
474 < where $\langle A \rangle_t$ is an equilibrium value of a physical
475 < quantity and $\rho (p(t), q(t))$ is the equilibrium distribution
476 < function. If an observation is averaged over a sufficiently long
477 < time (longer than relaxation time), all accessible microstates in
478 < phase space are assumed to be equally probed, giving a properly
479 < weighted statistical average. This allows the researcher freedom of
480 < choice when deciding how best to measure a given observable. In case
481 < an ensemble averaged approach sounds most reasonable, the Monte
482 < Carlo techniques\cite{metropolis:1949} can be utilized. Or if the
483 < system lends itself to a time averaging approach, the Molecular
484 < Dynamics techniques in Sec.~\ref{introSection:molecularDynamics}
485 < will be the best choice.
474 > where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
475 > physical quantity and $\rho (p(t), q(t))$ is the equilibrium
476 > distribution function. If an observation is averaged over a
477 > sufficiently long time (longer than relaxation time), all accessible
478 > microstates in phase space are assumed to be equally probed, giving
479 > a properly weighted statistical average. This allows the researcher
480 > freedom of choice when deciding how best to measure a given
481 > observable. In case an ensemble averaged approach sounds most
482 > reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
483 > utilized. Or if the system lends itself to a time averaging
484 > approach, the Molecular Dynamics techniques in
485 > Sec.~\ref{introSection:molecularDynamics} will be the best
486 > choice\cite{Frenkel1996}.
487  
488 < \section{\label{introSection:molecularDynamics}Molecular Dynamics}
488 > \section{\label{introSection:geometricIntegratos}Geometric Integrators}
489 > A variety of numerical integrators have been proposed to simulate
490 > the motions of atoms in MD simulation. They usually begin with
491 > initial conditionals and move the objects in the direction governed
492 > by the differential equations. However, most of them ignore the
493 > hidden physical laws contained within the equations. Since 1990,
494 > geometric integrators, which preserve various phase-flow invariants
495 > such as symplectic structure, volume and time reversal symmetry, are
496 > developed to address this issue\cite{Dullweber1997, McLachlan1998,
497 > Leimkuhler1999}. The velocity verlet method, which happens to be a
498 > simple example of symplectic integrator, continues to gain
499 > popularity in the molecular dynamics community. This fact can be
500 > partly explained by its geometric nature.
501  
502 < As a special discipline of molecular modeling, Molecular dynamics
503 < has proven to be a powerful tool for studying the functions of
504 < biological systems, providing structural, thermodynamic and
505 < dynamical information.
502 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
503 > A \emph{manifold} is an abstract mathematical space. It looks
504 > locally like Euclidean space, but when viewed globally, it may have
505 > more complicated structure. A good example of manifold is the
506 > surface of Earth. It seems to be flat locally, but it is round if
507 > viewed as a whole. A \emph{differentiable manifold} (also known as
508 > \emph{smooth manifold}) is a manifold on which it is possible to
509 > apply calculus on \emph{differentiable manifold}. A \emph{symplectic
510 > manifold} is defined as a pair $(M, \omega)$ which consists of a
511 > \emph{differentiable manifold} $M$ and a close, non-degenerated,
512 > bilinear symplectic form, $\omega$. A symplectic form on a vector
513 > space $V$ is a function $\omega(x, y)$ which satisfies
514 > $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
515 > \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
516 > $\omega(x, x) = 0$. The cross product operation in vector field is
517 > an example of symplectic form.
518  
519 < \subsection{\label{introSec:mdInit}Initialization}
519 > One of the motivations to study \emph{symplectic manifolds} in
520 > Hamiltonian Mechanics is that a symplectic manifold can represent
521 > all possible configurations of the system and the phase space of the
522 > system can be described by it's cotangent bundle. Every symplectic
523 > manifold is even dimensional. For instance, in Hamilton equations,
524 > coordinate and momentum always appear in pairs.
525  
526 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
526 > \subsection{\label{introSection:ODE}Ordinary Differential Equations}
527  
528 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
528 > For an ordinary differential system defined as
529 > \begin{equation}
530 > \dot x = f(x)
531 > \end{equation}
532 > where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
533 > \begin{equation}
534 > f(r) = J\nabla _x H(r).
535 > \end{equation}
536 > $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
537 > matrix
538 > \begin{equation}
539 > J = \left( {\begin{array}{*{20}c}
540 >   0 & I  \\
541 >   { - I} & 0  \\
542 > \end{array}} \right)
543 > \label{introEquation:canonicalMatrix}
544 > \end{equation}
545 > where $I$ is an identity matrix. Using this notation, Hamiltonian
546 > system can be rewritten as,
547 > \begin{equation}
548 > \frac{d}{{dt}}x = J\nabla _x H(x)
549 > \label{introEquation:compactHamiltonian}
550 > \end{equation}In this case, $f$ is
551 > called a \emph{Hamiltonian vector field}.
552  
553 < A rigid body is a body in which the distance between any two given
554 < points of a rigid body remains constant regardless of external
555 < forces exerted on it. A rigid body therefore conserves its shape
556 < during its motion.
553 > Another generalization of Hamiltonian dynamics is Poisson
554 > Dynamics\cite{Olver1986},
555 > \begin{equation}
556 > \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
557 > \end{equation}
558 > The most obvious change being that matrix $J$ now depends on $x$.
559  
560 < Applications of dynamics of rigid bodies.
560 > \subsection{\label{introSection:exactFlow}Exact Flow}
561  
562 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
563 <
564 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
565 <
566 < \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
567 <
568 < %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
569 <
570 < \section{\label{introSection:correlationFunctions}Correlation Functions}
562 > Let $x(t)$ be the exact solution of the ODE system,
563 > \begin{equation}
564 > \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
565 > \end{equation}
566 > The exact flow(solution) $\varphi_\tau$ is defined by
567 > \[
568 > x(t+\tau) =\varphi_\tau(x(t))
569 > \]
570 > where $\tau$ is a fixed time step and $\varphi$ is a map from phase
571 > space to itself. The flow has the continuous group property,
572 > \begin{equation}
573 > \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
574 > + \tau _2 } .
575 > \end{equation}
576 > In particular,
577 > \begin{equation}
578 > \varphi _\tau   \circ \varphi _{ - \tau }  = I
579 > \end{equation}
580 > Therefore, the exact flow is self-adjoint,
581 > \begin{equation}
582 > \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
583 > \end{equation}
584 > The exact flow can also be written in terms of the of an operator,
585 > \begin{equation}
586 > \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
587 > }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
588 > \label{introEquation:exponentialOperator}
589 > \end{equation}
590  
591 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
591 > In most cases, it is not easy to find the exact flow $\varphi_\tau$.
592 > Instead, we use a approximate map, $\psi_\tau$, which is usually
593 > called integrator. The order of an integrator $\psi_\tau$ is $p$, if
594 > the Taylor series of $\psi_\tau$ agree to order $p$,
595 > \begin{equation}
596 > \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
597 > \end{equation}
598  
599 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
599 > \subsection{\label{introSection:geometricProperties}Geometric Properties}
600  
601 < \subsection{\label{introSection:hydroynamics}Hydrodynamics}
601 > The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
602 > and its flow play important roles in numerical studies. Many of them
603 > can be found in systems which occur naturally in applications.
604 >
605 > Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
606 > a \emph{symplectic} flow if it satisfies,
607 > \begin{equation}
608 > {\varphi '}^T J \varphi ' = J.
609 > \end{equation}
610 > According to Liouville's theorem, the symplectic volume is invariant
611 > under a Hamiltonian flow, which is the basis for classical
612 > statistical mechanics. Furthermore, the flow of a Hamiltonian vector
613 > field on a symplectic manifold can be shown to be a
614 > symplectomorphism. As to the Poisson system,
615 > \begin{equation}
616 > {\varphi '}^T J \varphi ' = J \circ \varphi
617 > \end{equation}
618 > is the property must be preserved by the integrator.
619 >
620 > It is possible to construct a \emph{volume-preserving} flow for a
621 > source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
622 > \det d\varphi  = 1$. One can show easily that a symplectic flow will
623 > be volume-preserving.
624 >
625 > Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
626 > will result in a new system,
627 > \[
628 > \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
629 > \]
630 > The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
631 > In other words, the flow of this vector field is reversible if and
632 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
633 >
634 > A \emph{first integral}, or conserved quantity of a general
635 > differential function is a function $ G:R^{2d}  \to R^d $ which is
636 > constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
637 > \[
638 > \frac{{dG(x(t))}}{{dt}} = 0.
639 > \]
640 > Using chain rule, one may obtain,
641 > \[
642 > \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
643 > \]
644 > which is the condition for conserving \emph{first integral}. For a
645 > canonical Hamiltonian system, the time evolution of an arbitrary
646 > smooth function $G$ is given by,
647 >
648 > \begin{eqnarray}
649 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
650 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
651 > \label{introEquation:firstIntegral1}
652 > \end{eqnarray}
653 >
654 >
655 > Using poisson bracket notion, Equation
656 > \ref{introEquation:firstIntegral1} can be rewritten as
657 > \[
658 > \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
659 > \]
660 > Therefore, the sufficient condition for $G$ to be the \emph{first
661 > integral} of a Hamiltonian system is
662 > \[
663 > \left\{ {G,H} \right\} = 0.
664 > \]
665 > As well known, the Hamiltonian (or energy) H of a Hamiltonian system
666 > is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
667 > 0$.
668 >
669 > When designing any numerical methods, one should always try to
670 > preserve the structural properties of the original ODE and its flow.
671 >
672 > \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
673 > A lot of well established and very effective numerical methods have
674 > been successful precisely because of their symplecticities even
675 > though this fact was not recognized when they were first
676 > constructed. The most famous example is the Verlet-leapfrog methods
677 > in molecular dynamics. In general, symplectic integrators can be
678 > constructed using one of four different methods.
679 > \begin{enumerate}
680 > \item Generating functions
681 > \item Variational methods
682 > \item Runge-Kutta methods
683 > \item Splitting methods
684 > \end{enumerate}
685 >
686 > Generating function\cite{Channell1990} tends to lead to methods
687 > which are cumbersome and difficult to use. In dissipative systems,
688 > variational methods can capture the decay of energy
689 > accurately\cite{Kane2000}. Since their geometrically unstable nature
690 > against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
691 > methods are not suitable for Hamiltonian system. Recently, various
692 > high-order explicit Runge-Kutta methods
693 > \cite{Owren1992,Chen2003}have been developed to overcome this
694 > instability. However, due to computational penalty involved in
695 > implementing the Runge-Kutta methods, they have not attracted much
696 > attention from the Molecular Dynamics community. Instead, splitting
697 > methods have been widely accepted since they exploit natural
698 > decompositions of the system\cite{Tuckerman1992, McLachlan1998}.
699 >
700 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
701 >
702 > The main idea behind splitting methods is to decompose the discrete
703 > $\varphi_h$ as a composition of simpler flows,
704 > \begin{equation}
705 > \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
706 > \varphi _{h_n }
707 > \label{introEquation:FlowDecomposition}
708 > \end{equation}
709 > where each of the sub-flow is chosen such that each represent a
710 > simpler integration of the system.
711 >
712 > Suppose that a Hamiltonian system takes the form,
713 > \[
714 > H = H_1 + H_2.
715 > \]
716 > Here, $H_1$ and $H_2$ may represent different physical processes of
717 > the system. For instance, they may relate to kinetic and potential
718 > energy respectively, which is a natural decomposition of the
719 > problem. If $H_1$ and $H_2$ can be integrated using exact flows
720 > $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
721 > order expression is then given by the Lie-Trotter formula
722 > \begin{equation}
723 > \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
724 > \label{introEquation:firstOrderSplitting}
725 > \end{equation}
726 > where $\varphi _h$ is the result of applying the corresponding
727 > continuous $\varphi _i$ over a time $h$. By definition, as
728 > $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
729 > must follow that each operator $\varphi_i(t)$ is a symplectic map.
730 > It is easy to show that any composition of symplectic flows yields a
731 > symplectic map,
732 > \begin{equation}
733 > (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
734 > '\phi ' = \phi '^T J\phi ' = J,
735 > \label{introEquation:SymplecticFlowComposition}
736 > \end{equation}
737 > where $\phi$ and $\psi$ both are symplectic maps. Thus operator
738 > splitting in this context automatically generates a symplectic map.
739 >
740 > The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
741 > introduces local errors proportional to $h^2$, while Strang
742 > splitting gives a second-order decomposition,
743 > \begin{equation}
744 > \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
745 > _{1,h/2} , \label{introEquation:secondOrderSplitting}
746 > \end{equation}
747 > which has a local error proportional to $h^3$. The Sprang
748 > splitting's popularity in molecular simulation community attribute
749 > to its symmetric property,
750 > \begin{equation}
751 > \varphi _h^{ - 1} = \varphi _{ - h}.
752 > \label{introEquation:timeReversible}
753 > \end{equation},appendixFig:architecture
754 >
755 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Example of Splitting Method}}
756 > The classical equation for a system consisting of interacting
757 > particles can be written in Hamiltonian form,
758 > \[
759 > H = T + V
760 > \]
761 > where $T$ is the kinetic energy and $V$ is the potential energy.
762 > Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
763 > obtains the following:
764 > \begin{align}
765 > q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
766 >    \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
767 > \label{introEquation:Lp10a} \\%
768 > %
769 > \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
770 >    \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
771 > \label{introEquation:Lp10b}
772 > \end{align}
773 > where $F(t)$ is the force at time $t$. This integration scheme is
774 > known as \emph{velocity verlet} which is
775 > symplectic(\ref{introEquation:SymplecticFlowComposition}),
776 > time-reversible(\ref{introEquation:timeReversible}) and
777 > volume-preserving (\ref{introEquation:volumePreserving}). These
778 > geometric properties attribute to its long-time stability and its
779 > popularity in the community. However, the most commonly used
780 > velocity verlet integration scheme is written as below,
781 > \begin{align}
782 > \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
783 >    \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
784 > %
785 > q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
786 >    \label{introEquation:Lp9b}\\%
787 > %
788 > \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
789 >    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
790 > \end{align}
791 > From the preceding splitting, one can see that the integration of
792 > the equations of motion would follow:
793 > \begin{enumerate}
794 > \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
795 >
796 > \item Use the half step velocities to move positions one whole step, $\Delta t$.
797 >
798 > \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
799 >
800 > \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
801 > \end{enumerate}
802 >
803 > Simply switching the order of splitting and composing, a new
804 > integrator, the \emph{position verlet} integrator, can be generated,
805 > \begin{align}
806 > \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
807 > \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
808 > \label{introEquation:positionVerlet1} \\%
809 > %
810 > q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
811 > q(\Delta t)} \right]. %
812 > \label{introEquation:positionVerlet2}
813 > \end{align}
814 >
815 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
816 >
817 > Baker-Campbell-Hausdorff formula can be used to determine the local
818 > error of splitting method in terms of commutator of the
819 > operators(\ref{introEquation:exponentialOperator}) associated with
820 > the sub-flow. For operators $hX$ and $hY$ which are associate to
821 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
822 > \begin{equation}
823 > \exp (hX + hY) = \exp (hZ)
824 > \end{equation}
825 > where
826 > \begin{equation}
827 > hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
828 > {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
829 > \end{equation}
830 > Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
831 > \[
832 > [X,Y] = XY - YX .
833 > \]
834 > Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
835 > Sprang splitting, we can obtain
836 > \begin{eqnarray*}
837 > \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2 [X,Y]/4 + h^2 [Y,X]/4 \\
838 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
839 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
840 > \end{eqnarray*}
841 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
842 > error of Spring splitting is proportional to $h^3$. The same
843 > procedure can be applied to general splitting,  of the form
844 > \begin{equation}
845 > \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
846 > 1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
847 > \end{equation}
848 > Careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
849 > order method. Yoshida proposed an elegant way to compose higher
850 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
851 > a symmetric second order base method $ \varphi _h^{(2)} $, a
852 > fourth-order symmetric method can be constructed by composing,
853 > \[
854 > \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
855 > h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
856 > \]
857 > where $ \alpha  =  - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
858 > = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
859 > integrator $ \varphi _h^{(2n + 2)}$ can be composed by
860 > \begin{equation}
861 > \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
862 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
863 > \end{equation}
864 > , if the weights are chosen as
865 > \[
866 > \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
867 > \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
868 > \]
869 >
870 > \section{\label{introSection:molecularDynamics}Molecular Dynamics}
871 >
872 > As one of the principal tools of molecular modeling, Molecular
873 > dynamics has proven to be a powerful tool for studying the functions
874 > of biological systems, providing structural, thermodynamic and
875 > dynamical information. The basic idea of molecular dynamics is that
876 > macroscopic properties are related to microscopic behavior and
877 > microscopic behavior can be calculated from the trajectories in
878 > simulations. For instance, instantaneous temperature of an
879 > Hamiltonian system of $N$ particle can be measured by
880 > \[
881 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
882 > \]
883 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
884 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
885 > the boltzman constant.
886 >
887 > A typical molecular dynamics run consists of three essential steps:
888 > \begin{enumerate}
889 >  \item Initialization
890 >    \begin{enumerate}
891 >    \item Preliminary preparation
892 >    \item Minimization
893 >    \item Heating
894 >    \item Equilibration
895 >    \end{enumerate}
896 >  \item Production
897 >  \item Analysis
898 > \end{enumerate}
899 > These three individual steps will be covered in the following
900 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
901 > initialization of a simulation. Sec.~\ref{introSection:production}
902 > will discusses issues in production run.
903 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
904 > trajectory analysis.
905 >
906 > \subsection{\label{introSec:initialSystemSettings}Initialization}
907 >
908 > \subsubsection{\textbf{Preliminary preparation}}
909 >
910 > When selecting the starting structure of a molecule for molecular
911 > simulation, one may retrieve its Cartesian coordinates from public
912 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
913 > thousands of crystal structures of molecules are discovered every
914 > year, many more remain unknown due to the difficulties of
915 > purification and crystallization. Even for the molecule with known
916 > structure, some important information is missing. For example, the
917 > missing hydrogen atom which acts as donor in hydrogen bonding must
918 > be added. Moreover, in order to include electrostatic interaction,
919 > one may need to specify the partial charges for individual atoms.
920 > Under some circumstances, we may even need to prepare the system in
921 > a special setup. For instance, when studying transport phenomenon in
922 > membrane system, we may prepare the lipids in bilayer structure
923 > instead of placing lipids randomly in solvent, since we are not
924 > interested in self-aggregation and it takes a long time to happen.
925 >
926 > \subsubsection{\textbf{Minimization}}
927 >
928 > It is quite possible that some of molecules in the system from
929 > preliminary preparation may be overlapped with each other. This
930 > close proximity leads to high potential energy which consequently
931 > jeopardizes any molecular dynamics simulations. To remove these
932 > steric overlaps, one typically performs energy minimization to find
933 > a more reasonable conformation. Several energy minimization methods
934 > have been developed to exploit the energy surface and to locate the
935 > local minimum. While converging slowly near the minimum, steepest
936 > descent method is extremely robust when systems are far from
937 > harmonic. Thus, it is often used to refine structure from
938 > crystallographic data. Relied on the gradient or hessian, advanced
939 > methods like conjugate gradient and Newton-Raphson converge rapidly
940 > to a local minimum, while become unstable if the energy surface is
941 > far from quadratic. Another factor must be taken into account, when
942 > choosing energy minimization method, is the size of the system.
943 > Steepest descent and conjugate gradient can deal with models of any
944 > size. Because of the limit of computation power to calculate hessian
945 > matrix and insufficient storage capacity to store them, most
946 > Newton-Raphson methods can not be used with very large models.
947 >
948 > \subsubsection{\textbf{Heating}}
949 >
950 > Typically, Heating is performed by assigning random velocities
951 > according to a Gaussian distribution for a temperature. Beginning at
952 > a lower temperature and gradually increasing the temperature by
953 > assigning greater random velocities, we end up with setting the
954 > temperature of the system to a final temperature at which the
955 > simulation will be conducted. In heating phase, we should also keep
956 > the system from drifting or rotating as a whole. Equivalently, the
957 > net linear momentum and angular momentum of the system should be
958 > shifted to zero.
959 >
960 > \subsubsection{\textbf{Equilibration}}
961 >
962 > The purpose of equilibration is to allow the system to evolve
963 > spontaneously for a period of time and reach equilibrium. The
964 > procedure is continued until various statistical properties, such as
965 > temperature, pressure, energy, volume and other structural
966 > properties \textit{etc}, become independent of time. Strictly
967 > speaking, minimization and heating are not necessary, provided the
968 > equilibration process is long enough. However, these steps can serve
969 > as a means to arrive at an equilibrated structure in an effective
970 > way.
971 >
972 > \subsection{\label{introSection:production}Production}
973 >
974 > Production run is the most important step of the simulation, in
975 > which the equilibrated structure is used as a starting point and the
976 > motions of the molecules are collected for later analysis. In order
977 > to capture the macroscopic properties of the system, the molecular
978 > dynamics simulation must be performed in correct and efficient way.
979 >
980 > The most expensive part of a molecular dynamics simulation is the
981 > calculation of non-bonded forces, such as van der Waals force and
982 > Coulombic forces \textit{etc}. For a system of $N$ particles, the
983 > complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
984 > which making large simulations prohibitive in the absence of any
985 > computation saving techniques.
986 >
987 > A natural approach to avoid system size issue is to represent the
988 > bulk behavior by a finite number of the particles. However, this
989 > approach will suffer from the surface effect. To offset this,
990 > \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
991 > is developed to simulate bulk properties with a relatively small
992 > number of particles. In this method, the simulation box is
993 > replicated throughout space to form an infinite lattice. During the
994 > simulation, when a particle moves in the primary cell, its image in
995 > other cells move in exactly the same direction with exactly the same
996 > orientation. Thus, as a particle leaves the primary cell, one of its
997 > images will enter through the opposite face.
998 > \begin{figure}
999 > \centering
1000 > \includegraphics[width=\linewidth]{pbc.eps}
1001 > \caption[An illustration of periodic boundary conditions]{A 2-D
1002 > illustration of periodic boundary conditions. As one particle leaves
1003 > the left of the simulation box, an image of it enters the right.}
1004 > \label{introFig:pbc}
1005 > \end{figure}
1006 >
1007 > %cutoff and minimum image convention
1008 > Another important technique to improve the efficiency of force
1009 > evaluation is to apply cutoff where particles farther than a
1010 > predetermined distance, are not included in the calculation
1011 > \cite{Frenkel1996}. The use of a cutoff radius will cause a
1012 > discontinuity in the potential energy curve. Fortunately, one can
1013 > shift the potential to ensure the potential curve go smoothly to
1014 > zero at the cutoff radius. Cutoff strategy works pretty well for
1015 > Lennard-Jones interaction because of its short range nature.
1016 > However, simply truncating the electrostatic interaction with the
1017 > use of cutoff has been shown to lead to severe artifacts in
1018 > simulations. Ewald summation, in which the slowly conditionally
1019 > convergent Coulomb potential is transformed into direct and
1020 > reciprocal sums with rapid and absolute convergence, has proved to
1021 > minimize the periodicity artifacts in liquid simulations. Taking the
1022 > advantages of the fast Fourier transform (FFT) for calculating
1023 > discrete Fourier transforms, the particle mesh-based
1024 > methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1025 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1026 > multipole method}\cite{Greengard1987, Greengard1994}, which treats
1027 > Coulombic interaction exactly at short range, and approximate the
1028 > potential at long range through multipolar expansion. In spite of
1029 > their wide acceptances at the molecular simulation community, these
1030 > two methods are hard to be implemented correctly and efficiently.
1031 > Instead, we use a damped and charge-neutralized Coulomb potential
1032 > method developed by Wolf and his coworkers\cite{Wolf1999}. The
1033 > shifted Coulomb potential for particle $i$ and particle $j$ at
1034 > distance $r_{rj}$ is given by:
1035 > \begin{equation}
1036 > V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1037 > r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1038 > R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1039 > r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1040 > \end{equation}
1041 > where $\alpha$ is the convergence parameter. Due to the lack of
1042 > inherent periodicity and rapid convergence,this method is extremely
1043 > efficient and easy to implement.
1044 > \begin{figure}
1045 > \centering
1046 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1047 > \caption[An illustration of shifted Coulomb potential]{An
1048 > illustration of shifted Coulomb potential.}
1049 > \label{introFigure:shiftedCoulomb}
1050 > \end{figure}
1051 >
1052 > %multiple time step
1053 >
1054 > \subsection{\label{introSection:Analysis} Analysis}
1055 >
1056 > Recently, advanced visualization technique are widely applied to
1057 > monitor the motions of molecules. Although the dynamics of the
1058 > system can be described qualitatively from animation, quantitative
1059 > trajectory analysis are more appreciable. According to the
1060 > principles of Statistical Mechanics,
1061 > Sec.~\ref{introSection:statisticalMechanics}, one can compute
1062 > thermodynamics properties, analyze fluctuations of structural
1063 > parameters, and investigate time-dependent processes of the molecule
1064 > from the trajectories.
1065 >
1066 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamics Properties}}
1067 >
1068 > Thermodynamics properties, which can be expressed in terms of some
1069 > function of the coordinates and momenta of all particles in the
1070 > system, can be directly computed from molecular dynamics. The usual
1071 > way to measure the pressure is based on virial theorem of Clausius
1072 > which states that the virial is equal to $-3Nk_BT$. For a system
1073 > with forces between particles, the total virial, $W$, contains the
1074 > contribution from external pressure and interaction between the
1075 > particles:
1076 > \[
1077 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1078 > f_{ij} } } \right\rangle
1079 > \]
1080 > where $f_{ij}$ is the force between particle $i$ and $j$ at a
1081 > distance $r_{ij}$. Thus, the expression for the pressure is given
1082 > by:
1083 > \begin{equation}
1084 > P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1085 > < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1086 > \end{equation}
1087 >
1088 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1089 >
1090 > Structural Properties of a simple fluid can be described by a set of
1091 > distribution functions. Among these functions,\emph{pair
1092 > distribution function}, also known as \emph{radial distribution
1093 > function}, is of most fundamental importance to liquid-state theory.
1094 > Pair distribution function can be gathered by Fourier transforming
1095 > raw data from a series of neutron diffraction experiments and
1096 > integrating over the surface factor \cite{Powles1973}. The
1097 > experiment result can serve as a criterion to justify the
1098 > correctness of the theory. Moreover, various equilibrium
1099 > thermodynamic and structural properties can also be expressed in
1100 > terms of radial distribution function \cite{Allen1987}.
1101 >
1102 > A pair distribution functions $g(r)$ gives the probability that a
1103 > particle $i$ will be located at a distance $r$ from a another
1104 > particle $j$ in the system
1105 > \[
1106 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1107 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1108 > \]
1109 > Note that the delta function can be replaced by a histogram in
1110 > computer simulation. Figure
1111 > \ref{introFigure:pairDistributionFunction} shows a typical pair
1112 > distribution function for the liquid argon system. The occurrence of
1113 > several peaks in the plot of $g(r)$ suggests that it is more likely
1114 > to find particles at certain radial values than at others. This is a
1115 > result of the attractive interaction at such distances. Because of
1116 > the strong repulsive forces at short distance, the probability of
1117 > locating particles at distances less than about 2.5{\AA} from each
1118 > other is essentially zero.
1119 >
1120 > %\begin{figure}
1121 > %\centering
1122 > %\includegraphics[width=\linewidth]{pdf.eps}
1123 > %\caption[Pair distribution function for the liquid argon
1124 > %]{Pair distribution function for the liquid argon}
1125 > %\label{introFigure:pairDistributionFunction}
1126 > %\end{figure}
1127 >
1128 > \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1129 > Properties}}
1130 >
1131 > Time-dependent properties are usually calculated using \emph{time
1132 > correlation function}, which correlates random variables $A$ and $B$
1133 > at two different time
1134 > \begin{equation}
1135 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1136 > \label{introEquation:timeCorrelationFunction}
1137 > \end{equation}
1138 > If $A$ and $B$ refer to same variable, this kind of correlation
1139 > function is called \emph{auto correlation function}. One example of
1140 > auto correlation function is velocity auto-correlation function
1141 > which is directly related to transport properties of molecular
1142 > liquids:
1143 > \[
1144 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1145 > \right\rangle } dt
1146 > \]
1147 > where $D$ is diffusion constant. Unlike velocity autocorrelation
1148 > function which is averaging over time origins and over all the
1149 > atoms, dipole autocorrelation are calculated for the entire system.
1150 > The dipole autocorrelation function is given by:
1151 > \[
1152 > c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1153 > \right\rangle
1154 > \]
1155 > Here $u_{tot}$ is the net dipole of the entire system and is given
1156 > by
1157 > \[
1158 > u_{tot} (t) = \sum\limits_i {u_i (t)}
1159 > \]
1160 > In principle, many time correlation functions can be related with
1161 > Fourier transforms of the infrared, Raman, and inelastic neutron
1162 > scattering spectra of molecular liquids. In practice, one can
1163 > extract the IR spectrum from the intensity of dipole fluctuation at
1164 > each frequency using the following relationship:
1165 > \[
1166 > \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1167 > i2\pi vt} dt}
1168 > \]
1169 >
1170 > \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1171 >
1172 > Rigid bodies are frequently involved in the modeling of different
1173 > areas, from engineering, physics, to chemistry. For example,
1174 > missiles and vehicle are usually modeled by rigid bodies.  The
1175 > movement of the objects in 3D gaming engine or other physics
1176 > simulator is governed by the rigid body dynamics. In molecular
1177 > simulation, rigid body is used to simplify the model in
1178 > protein-protein docking study\cite{Gray2003}.
1179 >
1180 > It is very important to develop stable and efficient methods to
1181 > integrate the equations of motion of orientational degrees of
1182 > freedom. Euler angles are the nature choice to describe the
1183 > rotational degrees of freedom. However, due to its singularity, the
1184 > numerical integration of corresponding equations of motion is very
1185 > inefficient and inaccurate. Although an alternative integrator using
1186 > different sets of Euler angles can overcome this
1187 > difficulty\cite{Barojas1973}, the computational penalty and the lost
1188 > of angular momentum conservation still remain. A singularity free
1189 > representation utilizing quaternions was developed by Evans in
1190 > 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1191 > nonseparable Hamiltonian resulted from quaternion representation,
1192 > which prevents the symplectic algorithm to be utilized. Another
1193 > different approach is to apply holonomic constraints to the atoms
1194 > belonging to the rigid body. Each atom moves independently under the
1195 > normal forces deriving from potential energy and constraint forces
1196 > which are used to guarantee the rigidness. However, due to their
1197 > iterative nature, SHAKE and Rattle algorithm converge very slowly
1198 > when the number of constraint increases\cite{Ryckaert1977,
1199 > Andersen1983}.
1200 >
1201 > The break through in geometric literature suggests that, in order to
1202 > develop a long-term integration scheme, one should preserve the
1203 > symplectic structure of the flow. Introducing conjugate momentum to
1204 > rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1205 > symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1206 > the Hamiltonian system in a constraint manifold by iteratively
1207 > satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1208 > method using quaternion representation was developed by
1209 > Omelyan\cite{Omelyan1998}. However, both of these methods are
1210 > iterative and inefficient. In this section, we will present a
1211 > symplectic Lie-Poisson integrator for rigid body developed by
1212 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1213 >
1214 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1215 > The motion of the rigid body is Hamiltonian with the Hamiltonian
1216 > function
1217 > \begin{equation}
1218 > H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1219 > V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1220 > \label{introEquation:RBHamiltonian}
1221 > \end{equation}
1222 > Here, $q$ and $Q$  are the position and rotation matrix for the
1223 > rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1224 > $J$, a diagonal matrix, is defined by
1225 > \[
1226 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1227 > \]
1228 > where $I_{ii}$ is the diagonal element of the inertia tensor. This
1229 > constrained Hamiltonian equation subjects to a holonomic constraint,
1230 > \begin{equation}
1231 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1232 > \end{equation}
1233 > which is used to ensure rotation matrix's orthogonality.
1234 > Differentiating \ref{introEquation:orthogonalConstraint} and using
1235 > Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1236 > \begin{equation}
1237 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1238 > \label{introEquation:RBFirstOrderConstraint}
1239 > \end{equation}
1240 >
1241 > Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1242 > \ref{introEquation:motionHamiltonianMomentum}), one can write down
1243 > the equations of motion,
1244 >
1245 > \begin{eqnarray}
1246 > \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1247 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1248 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1249 > \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1250 > \end{eqnarray}
1251 >
1252 > In general, there are two ways to satisfy the holonomic constraints.
1253 > We can use constraint force provided by lagrange multiplier on the
1254 > normal manifold to keep the motion on constraint space. Or we can
1255 > simply evolve the system in constraint manifold. These two methods
1256 > are proved to be equivalent. The holonomic constraint and equations
1257 > of motions define a constraint manifold for rigid body
1258 > \[
1259 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1260 > \right\}.
1261 > \]
1262 >
1263 > Unfortunately, this constraint manifold is not the cotangent bundle
1264 > $T_{\star}SO(3)$. However, it turns out that under symplectic
1265 > transformation, the cotangent space and the phase space are
1266 > diffeomorphic. Introducing
1267 > \[
1268 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1269 > \]
1270 > the mechanical system subject to a holonomic constraint manifold $M$
1271 > can be re-formulated as a Hamiltonian system on the cotangent space
1272 > \[
1273 > T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1274 > 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1275 > \]
1276 >
1277 > For a body fixed vector $X_i$ with respect to the center of mass of
1278 > the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1279 > given as
1280 > \begin{equation}
1281 > X_i^{lab} = Q X_i + q.
1282 > \end{equation}
1283 > Therefore, potential energy $V(q,Q)$ is defined by
1284 > \[
1285 > V(q,Q) = V(Q X_0 + q).
1286 > \]
1287 > Hence, the force and torque are given by
1288 > \[
1289 > \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1290 > \]
1291 > and
1292 > \[
1293 > \nabla _Q V(q,Q) = F(q,Q)X_i^t
1294 > \]
1295 > respectively.
1296 >
1297 > As a common choice to describe the rotation dynamics of the rigid
1298 > body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1299 > rewrite the equations of motion,
1300 > \begin{equation}
1301 > \begin{array}{l}
1302 > \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1303 > \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1304 > \end{array}
1305 > \label{introEqaution:RBMotionPI}
1306 > \end{equation}
1307 > , as well as holonomic constraints,
1308 > \[
1309 > \begin{array}{l}
1310 > \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1311 > Q^T Q = 1 \\
1312 > \end{array}
1313 > \]
1314 >
1315 > For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1316 > so(3)^ \star$, the hat-map isomorphism,
1317 > \begin{equation}
1318 > v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1319 > {\begin{array}{*{20}c}
1320 >   0 & { - v_3 } & {v_2 }  \\
1321 >   {v_3 } & 0 & { - v_1 }  \\
1322 >   { - v_2 } & {v_1 } & 0  \\
1323 > \end{array}} \right),
1324 > \label{introEquation:hatmapIsomorphism}
1325 > \end{equation}
1326 > will let us associate the matrix products with traditional vector
1327 > operations
1328 > \[
1329 > \hat vu = v \times u
1330 > \]
1331 > Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1332 > matrix,
1333 > \begin{equation}
1334 > (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ {\bullet  ^T}
1335 > ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1336 > - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1337 > (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1338 > \end{equation}
1339 > Since $\Lambda$ is symmetric, the last term of Equation
1340 > \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1341 > multiplier $\Lambda$ is absent from the equations of motion. This
1342 > unique property eliminate the requirement of iterations which can
1343 > not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1344 >
1345 > Applying hat-map isomorphism, we obtain the equation of motion for
1346 > angular momentum on body frame
1347 > \begin{equation}
1348 > \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1349 > F_i (r,Q)} \right) \times X_i }.
1350 > \label{introEquation:bodyAngularMotion}
1351 > \end{equation}
1352 > In the same manner, the equation of motion for rotation matrix is
1353 > given by
1354 > \[
1355 > \dot Q = Qskew(I^{ - 1} \pi )
1356 > \]
1357 >
1358 > \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1359 > Lie-Poisson Integrator for Free Rigid Body}
1360 >
1361 > If there is not external forces exerted on the rigid body, the only
1362 > contribution to the rotational is from the kinetic potential (the
1363 > first term of \ref{introEquation:bodyAngularMotion}). The free rigid
1364 > body is an example of Lie-Poisson system with Hamiltonian function
1365 > \begin{equation}
1366 > T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1367 > \label{introEquation:rotationalKineticRB}
1368 > \end{equation}
1369 > where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1370 > Lie-Poisson structure matrix,
1371 > \begin{equation}
1372 > J(\pi ) = \left( {\begin{array}{*{20}c}
1373 >   0 & {\pi _3 } & { - \pi _2 }  \\
1374 >   { - \pi _3 } & 0 & {\pi _1 }  \\
1375 >   {\pi _2 } & { - \pi _1 } & 0  \\
1376 > \end{array}} \right)
1377 > \end{equation}
1378 > Thus, the dynamics of free rigid body is governed by
1379 > \begin{equation}
1380 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1381 > \end{equation}
1382 >
1383 > One may notice that each $T_i^r$ in Equation
1384 > \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1385 > instance, the equations of motion due to $T_1^r$ are given by
1386 > \begin{equation}
1387 > \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1388 > \label{introEqaution:RBMotionSingleTerm}
1389 > \end{equation}
1390 > where
1391 > \[ R_1  = \left( {\begin{array}{*{20}c}
1392 >   0 & 0 & 0  \\
1393 >   0 & 0 & {\pi _1 }  \\
1394 >   0 & { - \pi _1 } & 0  \\
1395 > \end{array}} \right).
1396 > \]
1397 > The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1398 > \[
1399 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1400 > Q(0)e^{\Delta tR_1 }
1401 > \]
1402 > with
1403 > \[
1404 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1405 >   0 & 0 & 0  \\
1406 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1407 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1408 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1409 > \]
1410 > To reduce the cost of computing expensive functions in $e^{\Delta
1411 > tR_1 }$, we can use Cayley transformation,
1412 > \[
1413 > e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1414 > )
1415 > \]
1416 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1417 > manner.
1418 >
1419 > In order to construct a second-order symplectic method, we split the
1420 > angular kinetic Hamiltonian function can into five terms
1421 > \[
1422 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1423 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1424 > (\pi _1 )
1425 > \].
1426 > Concatenating flows corresponding to these five terms, we can obtain
1427 > an symplectic integrator,
1428 > \[
1429 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1430 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1431 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1432 > _1 }.
1433 > \]
1434 >
1435 > The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1436 > $F(\pi )$ and $G(\pi )$ is defined by
1437 > \[
1438 > \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1439 > )
1440 > \]
1441 > If the Poisson bracket of a function $F$ with an arbitrary smooth
1442 > function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1443 > conserved quantity in Poisson system. We can easily verify that the
1444 > norm of the angular momentum, $\parallel \pi
1445 > \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1446 > \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1447 > then by the chain rule
1448 > \[
1449 > \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1450 > }}{2})\pi
1451 > \]
1452 > Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1453 > \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1454 > Lie-Poisson integrator is found to be extremely efficient and stable
1455 > which can be explained by the fact the small angle approximation is
1456 > used and the norm of the angular momentum is conserved.
1457 >
1458 > \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1459 > Splitting for Rigid Body}
1460 >
1461 > The Hamiltonian of rigid body can be separated in terms of kinetic
1462 > energy and potential energy,
1463 > \[
1464 > H = T(p,\pi ) + V(q,Q)
1465 > \]
1466 > The equations of motion corresponding to potential energy and
1467 > kinetic energy are listed in the below table,
1468 > \begin{table}
1469 > \caption{Equations of motion due to Potential and Kinetic Energies}
1470 > \begin{center}
1471 > \begin{tabular}{|l|l|}
1472 >  \hline
1473 >  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1474 >  Potential & Kinetic \\
1475 >  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1476 >  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1477 >  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1478 >  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1479 >  \hline
1480 > \end{tabular}
1481 > \end{center}
1482 > \end{table}
1483 > A second-order symplectic method is now obtained by the
1484 > composition of the flow maps,
1485 > \[
1486 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1487 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1488 > \]
1489 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1490 > sub-flows which corresponding to force and torque respectively,
1491 > \[
1492 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1493 > _{\Delta t/2,\tau }.
1494 > \]
1495 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1496 > $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1497 > order inside $\varphi _{\Delta t/2,V}$ does not matter.
1498 >
1499 > Furthermore, kinetic potential can be separated to translational
1500 > kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1501 > \begin{equation}
1502 > T(p,\pi ) =T^t (p) + T^r (\pi ).
1503 > \end{equation}
1504 > where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1505 > defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1506 > corresponding flow maps are given by
1507 > \[
1508 > \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1509 > _{\Delta t,T^r }.
1510 > \]
1511 > Finally, we obtain the overall symplectic flow maps for free moving
1512 > rigid body
1513 > \begin{equation}
1514 > \begin{array}{c}
1515 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1516 >  \circ \varphi _{\Delta t,T^t }  \circ \varphi _{\Delta t/2,\pi _1 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi _1 }  \\
1517 >  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1518 > \end{array}
1519 > \label{introEquation:overallRBFlowMaps}
1520 > \end{equation}
1521 >
1522 > \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1523 > As an alternative to newtonian dynamics, Langevin dynamics, which
1524 > mimics a simple heat bath with stochastic and dissipative forces,
1525 > has been applied in a variety of studies. This section will review
1526 > the theory of Langevin dynamics simulation. A brief derivation of
1527 > generalized Langevin equation will be given first. Follow that, we
1528 > will discuss the physical meaning of the terms appearing in the
1529 > equation as well as the calculation of friction tensor from
1530 > hydrodynamics theory.
1531 >
1532 > \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1533 >
1534 > Harmonic bath model, in which an effective set of harmonic
1535 > oscillators are used to mimic the effect of a linearly responding
1536 > environment, has been widely used in quantum chemistry and
1537 > statistical mechanics. One of the successful applications of
1538 > Harmonic bath model is the derivation of Deriving Generalized
1539 > Langevin Dynamics. Lets consider a system, in which the degree of
1540 > freedom $x$ is assumed to couple to the bath linearly, giving a
1541 > Hamiltonian of the form
1542 > \begin{equation}
1543 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1544 > \label{introEquation:bathGLE}.
1545 > \end{equation}
1546 > Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1547 > with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1548 > \[
1549 > H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1550 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1551 > \right\}}
1552 > \]
1553 > where the index $\alpha$ runs over all the bath degrees of freedom,
1554 > $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1555 > the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1556 > coupling,
1557 > \[
1558 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1559 > \]
1560 > where $g_\alpha$ are the coupling constants between the bath and the
1561 > coordinate $x$. Introducing
1562 > \[
1563 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1564 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1565 > \] and combining the last two terms in Equation
1566 > \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1567 > Hamiltonian as
1568 > \[
1569 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1570 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1571 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1572 > w_\alpha ^2 }}x} \right)^2 } \right\}}
1573 > \]
1574 > Since the first two terms of the new Hamiltonian depend only on the
1575 > system coordinates, we can get the equations of motion for
1576 > Generalized Langevin Dynamics by Hamilton's equations
1577 > \ref{introEquation:motionHamiltonianCoordinate,
1578 > introEquation:motionHamiltonianMomentum},
1579 > \begin{equation}
1580 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1581 > \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1582 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1583 > \label{introEquation:coorMotionGLE}
1584 > \end{equation}
1585 > and
1586 > \begin{equation}
1587 > m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1588 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1589 > \label{introEquation:bathMotionGLE}
1590 > \end{equation}
1591 >
1592 > In order to derive an equation for $x$, the dynamics of the bath
1593 > variables $x_\alpha$ must be solved exactly first. As an integral
1594 > transform which is particularly useful in solving linear ordinary
1595 > differential equations, Laplace transform is the appropriate tool to
1596 > solve this problem. The basic idea is to transform the difficult
1597 > differential equations into simple algebra problems which can be
1598 > solved easily. Then applying inverse Laplace transform, also known
1599 > as the Bromwich integral, we can retrieve the solutions of the
1600 > original problems.
1601 >
1602 > Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1603 > transform of f(t) is a new function defined as
1604 > \[
1605 > L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1606 > \]
1607 > where  $p$ is real and  $L$ is called the Laplace Transform
1608 > Operator. Below are some important properties of Laplace transform
1609 >
1610 > \begin{eqnarray*}
1611 > L(x + y)  & = & L(x) + L(y) \\
1612 > L(ax)     & = & aL(x) \\
1613 > L(\dot x) & = & pL(x) - px(0) \\
1614 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1615 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1616 > \end{eqnarray*}
1617 >
1618 >
1619 > Applying Laplace transform to the bath coordinates, we obtain
1620 > \begin{eqnarray*}
1621 > p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) & = & - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1622 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1623 > \end{eqnarray*}
1624 >
1625 > By the same way, the system coordinates become
1626 > \begin{eqnarray*}
1627 > mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1628 >  & & \mbox{} - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1629 > \end{eqnarray*}
1630 >
1631 > With the help of some relatively important inverse Laplace
1632 > transformations:
1633 > \[
1634 > \begin{array}{c}
1635 > L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1636 > L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1637 > L(1) = \frac{1}{p} \\
1638 > \end{array}
1639 > \]
1640 > , we obtain
1641 > \begin{eqnarray*}
1642 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1643 > \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1644 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1645 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1646 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1647 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1648 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1649 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1650 > \end{eqnarray*}
1651 > \begin{eqnarray*}
1652 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1653 > {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1654 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1655 > t)\dot x(t - \tau )d} \tau }  \\
1656 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1657 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1658 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1659 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1660 > \end{eqnarray*}
1661 > Introducing a \emph{dynamic friction kernel}
1662 > \begin{equation}
1663 > \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1664 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1665 > \label{introEquation:dynamicFrictionKernelDefinition}
1666 > \end{equation}
1667 > and \emph{a random force}
1668 > \begin{equation}
1669 > R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1670 > - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1671 > \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1672 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1673 > \label{introEquation:randomForceDefinition}
1674 > \end{equation}
1675 > the equation of motion can be rewritten as
1676 > \begin{equation}
1677 > m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1678 > (t)\dot x(t - \tau )d\tau }  + R(t)
1679 > \label{introEuqation:GeneralizedLangevinDynamics}
1680 > \end{equation}
1681 > which is known as the \emph{generalized Langevin equation}.
1682 >
1683 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1684 >
1685 > One may notice that $R(t)$ depends only on initial conditions, which
1686 > implies it is completely deterministic within the context of a
1687 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1688 > uncorrelated to $x$ and $\dot x$,
1689 > \[
1690 > \begin{array}{l}
1691 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1692 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1693 > \end{array}
1694 > \]
1695 > This property is what we expect from a truly random process. As long
1696 > as the model, which is gaussian distribution in general, chosen for
1697 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1698 > still remains.
1699 >
1700 > %dynamic friction kernel
1701 > The convolution integral
1702 > \[
1703 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1704 > \]
1705 > depends on the entire history of the evolution of $x$, which implies
1706 > that the bath retains memory of previous motions. In other words,
1707 > the bath requires a finite time to respond to change in the motion
1708 > of the system. For a sluggish bath which responds slowly to changes
1709 > in the system coordinate, we may regard $\xi(t)$ as a constant
1710 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1711 > \[
1712 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1713 > \]
1714 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1715 > \[
1716 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1717 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1718 > \]
1719 > which can be used to describe dynamic caging effect. The other
1720 > extreme is the bath that responds infinitely quickly to motions in
1721 > the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1722 > time:
1723 > \[
1724 > \xi (t) = 2\xi _0 \delta (t)
1725 > \]
1726 > Hence, the convolution integral becomes
1727 > \[
1728 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1729 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1730 > \]
1731 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1732 > \begin{equation}
1733 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1734 > x(t) + R(t) \label{introEquation:LangevinEquation}
1735 > \end{equation}
1736 > which is known as the Langevin equation. The static friction
1737 > coefficient $\xi _0$ can either be calculated from spectral density
1738 > or be determined by Stokes' law for regular shaped particles.A
1739 > briefly review on calculating friction tensor for arbitrary shaped
1740 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1741 >
1742 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1743 >
1744 > Defining a new set of coordinates,
1745 > \[
1746 > q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1747 > ^2 }}x(0)
1748 > \],
1749 > we can rewrite $R(T)$ as
1750 > \[
1751 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1752 > \]
1753 > And since the $q$ coordinates are harmonic oscillators,
1754 >
1755 > \begin{eqnarray*}
1756 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1757 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1758 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1759 > \left\langle {R(t)R(0)} \right\rangle & = & \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1760 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1761 >  & = &kT\xi (t) \\
1762 > \end{eqnarray*}
1763 >
1764 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1765 > \begin{equation}
1766 > \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1767 > \label{introEquation:secondFluctuationDissipation}.
1768 > \end{equation}
1769 > In effect, it acts as a constraint on the possible ways in which one
1770 > can model the random force and friction kernel.
1771 >
1772 > \subsection{\label{introSection:frictionTensor} Friction Tensor}
1773 > Theoretically, the friction kernel can be determined using velocity
1774 > autocorrelation function. However, this approach become impractical
1775 > when the system become more and more complicate. Instead, various
1776 > approaches based on hydrodynamics have been developed to calculate
1777 > the friction coefficients. The friction effect is isotropic in
1778 > Equation, $\zeta$ can be taken as a scalar. In general, friction
1779 > tensor $\Xi$ is a $6\times 6$ matrix given by
1780 > \[
1781 > \Xi  = \left( {\begin{array}{*{20}c}
1782 >   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1783 >   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1784 > \end{array}} \right).
1785 > \]
1786 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1787 > tensor and rotational resistance (friction) tensor respectively,
1788 > while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1789 > {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1790 > particle moves in a fluid, it may experience friction force or
1791 > torque along the opposite direction of the velocity or angular
1792 > velocity,
1793 > \[
1794 > \left( \begin{array}{l}
1795 > F_R  \\
1796 > \tau _R  \\
1797 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1798 >   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1799 >   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1800 > \end{array}} \right)\left( \begin{array}{l}
1801 > v \\
1802 > w \\
1803 > \end{array} \right)
1804 > \]
1805 > where $F_r$ is the friction force and $\tau _R$ is the friction
1806 > toque.
1807 >
1808 > \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
1809 >
1810 > For a spherical particle, the translational and rotational friction
1811 > constant can be calculated from Stoke's law,
1812 > \[
1813 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1814 >   {6\pi \eta R} & 0 & 0  \\
1815 >   0 & {6\pi \eta R} & 0  \\
1816 >   0 & 0 & {6\pi \eta R}  \\
1817 > \end{array}} \right)
1818 > \]
1819 > and
1820 > \[
1821 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1822 >   {8\pi \eta R^3 } & 0 & 0  \\
1823 >   0 & {8\pi \eta R^3 } & 0  \\
1824 >   0 & 0 & {8\pi \eta R^3 }  \\
1825 > \end{array}} \right)
1826 > \]
1827 > where $\eta$ is the viscosity of the solvent and $R$ is the
1828 > hydrodynamics radius.
1829 >
1830 > Other non-spherical shape, such as cylinder and ellipsoid
1831 > \textit{etc}, are widely used as reference for developing new
1832 > hydrodynamics theory, because their properties can be calculated
1833 > exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1834 > also called a triaxial ellipsoid, which is given in Cartesian
1835 > coordinates by\cite{Perrin1934, Perrin1936}
1836 > \[
1837 > \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1838 > }} = 1
1839 > \]
1840 > where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1841 > due to the complexity of the elliptic integral, only the ellipsoid
1842 > with the restriction of two axes having to be equal, \textit{i.e.}
1843 > prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1844 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
1845 > \[
1846 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1847 > } }}{b},
1848 > \]
1849 > and oblate,
1850 > \[
1851 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1852 > }}{a}
1853 > \],
1854 > one can write down the translational and rotational resistance
1855 > tensors
1856 > \[
1857 > \begin{array}{l}
1858 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1859 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1860 > \end{array},
1861 > \]
1862 > and
1863 > \[
1864 > \begin{array}{l}
1865 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1866 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1867 > \end{array}.
1868 > \]
1869 >
1870 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shape}}
1871 >
1872 > Unlike spherical and other regular shaped molecules, there is not
1873 > analytical solution for friction tensor of any arbitrary shaped
1874 > rigid molecules. The ellipsoid of revolution model and general
1875 > triaxial ellipsoid model have been used to approximate the
1876 > hydrodynamic properties of rigid bodies. However, since the mapping
1877 > from all possible ellipsoidal space, $r$-space, to all possible
1878 > combination of rotational diffusion coefficients, $D$-space is not
1879 > unique\cite{Wegener1979} as well as the intrinsic coupling between
1880 > translational and rotational motion of rigid body, general ellipsoid
1881 > is not always suitable for modeling arbitrarily shaped rigid
1882 > molecule. A number of studies have been devoted to determine the
1883 > friction tensor for irregularly shaped rigid bodies using more
1884 > advanced method where the molecule of interest was modeled by
1885 > combinations of spheres(beads)\cite{Carrasco1999} and the
1886 > hydrodynamics properties of the molecule can be calculated using the
1887 > hydrodynamic interaction tensor. Let us consider a rigid assembly of
1888 > $N$ beads immersed in a continuous medium. Due to hydrodynamics
1889 > interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1890 > than its unperturbed velocity $v_i$,
1891 > \[
1892 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1893 > \]
1894 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1895 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1896 > proportional to its ``net'' velocity
1897 > \begin{equation}
1898 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1899 > \label{introEquation:tensorExpression}
1900 > \end{equation}
1901 > This equation is the basis for deriving the hydrodynamic tensor. In
1902 > 1930, Oseen and Burgers gave a simple solution to Equation
1903 > \ref{introEquation:tensorExpression}
1904 > \begin{equation}
1905 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1906 > R_{ij}^T }}{{R_{ij}^2 }}} \right).
1907 > \label{introEquation:oseenTensor}
1908 > \end{equation}
1909 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1910 > A second order expression for element of different size was
1911 > introduced by Rotne and Prager\cite{Rotne1969} and improved by
1912 > Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1913 > \begin{equation}
1914 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1915 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1916 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1917 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1918 > \label{introEquation:RPTensorNonOverlapped}
1919 > \end{equation}
1920 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1921 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1922 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1923 > overlapping beads with the same radius, $\sigma$, is given by
1924 > \begin{equation}
1925 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1926 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1927 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1928 > \label{introEquation:RPTensorOverlapped}
1929 > \end{equation}
1930 >
1931 > To calculate the resistance tensor at an arbitrary origin $O$, we
1932 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1933 > $B_{ij}$ blocks
1934 > \begin{equation}
1935 > B = \left( {\begin{array}{*{20}c}
1936 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1937 >    \vdots  &  \ddots  &  \vdots   \\
1938 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1939 > \end{array}} \right),
1940 > \end{equation}
1941 > where $B_{ij}$ is given by
1942 > \[
1943 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1944 > )T_{ij}
1945 > \]
1946 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1947 > $B$, we obtain
1948 >
1949 > \[
1950 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1951 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1952 >    \vdots  &  \ddots  &  \vdots   \\
1953 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1954 > \end{array}} \right)
1955 > \]
1956 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1957 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1958 > \[
1959 > U_i  = \left( {\begin{array}{*{20}c}
1960 >   0 & { - z_i } & {y_i }  \\
1961 >   {z_i } & 0 & { - x_i }  \\
1962 >   { - y_i } & {x_i } & 0  \\
1963 > \end{array}} \right)
1964 > \]
1965 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1966 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1967 > arbitrary origin $O$ can be written as
1968 > \begin{equation}
1969 > \begin{array}{l}
1970 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1971 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1972 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1973 > \end{array}
1974 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1975 > \end{equation}
1976 >
1977 > The resistance tensor depends on the origin to which they refer. The
1978 > proper location for applying friction force is the center of
1979 > resistance (reaction), at which the trace of rotational resistance
1980 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1981 > resistance is defined as an unique point of the rigid body at which
1982 > the translation-rotation coupling tensor are symmetric,
1983 > \begin{equation}
1984 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1985 > \label{introEquation:definitionCR}
1986 > \end{equation}
1987 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
1988 > we can easily find out that the translational resistance tensor is
1989 > origin independent, while the rotational resistance tensor and
1990 > translation-rotation coupling resistance tensor depend on the
1991 > origin. Given resistance tensor at an arbitrary origin $O$, and a
1992 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
1993 > obtain the resistance tensor at $P$ by
1994 > \begin{equation}
1995 > \begin{array}{l}
1996 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
1997 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
1998 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{{tr} ^{^T }}  \\
1999 > \end{array}
2000 > \label{introEquation:resistanceTensorTransformation}
2001 > \end{equation}
2002 > where
2003 > \[
2004 > U_{OP}  = \left( {\begin{array}{*{20}c}
2005 >   0 & { - z_{OP} } & {y_{OP} }  \\
2006 >   {z_i } & 0 & { - x_{OP} }  \\
2007 >   { - y_{OP} } & {x_{OP} } & 0  \\
2008 > \end{array}} \right)
2009 > \]
2010 > Using Equations \ref{introEquation:definitionCR} and
2011 > \ref{introEquation:resistanceTensorTransformation}, one can locate
2012 > the position of center of resistance,
2013 > \begin{eqnarray*}
2014 > \left( \begin{array}{l}
2015 > x_{OR}  \\
2016 > y_{OR}  \\
2017 > z_{OR}  \\
2018 > \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2019 >   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2020 >   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2021 >   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2022 > \end{array}} \right)^{ - 1}  \\
2023 >  & & \left( \begin{array}{l}
2024 > (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2025 > (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2026 > (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2027 > \end{array} \right) \\
2028 > \end{eqnarray*}
2029 >
2030 >
2031 >
2032 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2033 > joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines