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 2895 by tim, Tue Jun 27 02:42:30 2006 UTC vs.
Revision 2945 by tim, Tue Jul 18 14:19:49 2006 UTC

# Line 3 | Line 3 | Closely related to Classical Mechanics, Molecular Dyna
3   \section{\label{introSection:classicalMechanics}Classical
4   Mechanics}
5  
6 < Closely related to Classical Mechanics, Molecular Dynamics
7 < simulations are carried out by integrating the equations of motion
8 < for a given system of particles. There are three fundamental ideas
9 < behind classical mechanics. Firstly, one can determine the state of
10 < a mechanical system at any time of interest; Secondly, all the
11 < mechanical properties of the system at that time can be determined
12 < by combining the knowledge of the properties of the system with the
13 < specification of this state; Finally, the specification of the state
14 < when further combine with the laws of mechanics will also be
15 < sufficient to predict the future behavior of the system.
6 > Using equations of motion derived from Classical Mechanics,
7 > Molecular Dynamics simulations are carried out by integrating the
8 > equations of motion for a given system of particles. There are three
9 > fundamental ideas behind classical mechanics. Firstly, one can
10 > determine the state of a mechanical system at any time of interest;
11 > Secondly, all the mechanical properties of the system at that time
12 > can be determined by combining the knowledge of the properties of
13 > the system with the specification of this state; Finally, the
14 > specification of the state when further combined with the laws of
15 > mechanics will also be sufficient to predict the future behavior of
16 > the system.
17  
18   \subsection{\label{introSection:newtonian}Newtonian Mechanics}
19   The discovery of Newton's three laws of mechanics which govern the
# Line 31 | Line 32 | F_{ij} = -F_{ji}
32   $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
33   Newton's third law states that
34   \begin{equation}
35 < F_{ij} = -F_{ji}
35 > F_{ij} = -F_{ji}.
36   \label{introEquation:newtonThirdLaw}
37   \end{equation}
37
38   Conservation laws of Newtonian Mechanics play very important roles
39   in solving mechanics problems. The linear momentum of a particle is
40   conserved if it is free or it experiences no force. The second
# Line 63 | Line 63 | that if all forces are conservative, Energy
63   \end{equation}
64   If there are no external torques acting on a body, the angular
65   momentum of it is conserved. The last conservation theorem state
66 < that if all forces are conservative, Energy
67 < \begin{equation}E = T + V \label{introEquation:energyConservation}
66 > that if all forces are conservative, energy is conserved,
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}.
69 > All of these conserved quantities are important factors to determine
70 > the quality of numerical integration schemes for rigid
71 > bodies.\cite{Dullweber1997}
72  
73   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
75 < Newtonian Mechanics suffers from two important limitations: motions
76 < can only be described in cartesian coordinate systems. Moreover, it
77 < becomes 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.
75 > Newtonian Mechanics suffers from an important limitation: motion can
76 > only be described in cartesian coordinate systems which make it
77 > impossible to predict analytically the properties of the system even
78 > if we know all of the details of the interaction. In order to
79 > overcome some of the practical difficulties which arise in attempts
80 > to apply Newton's equation to complex systems, approximate numerical
81 > procedures may be developed.
82  
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 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, $U$.
88 > Hamilton's Principle may be stated as follows: the trajectory, along
89 > which a dynamical system may move from one point to another within a
90 > specified time, is derived by finding the path which minimizes the
91 > time integral of the difference between the kinetic $K$, and
92 > potential energies $U$,
93   \begin{equation}
94 < \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
94 > \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0}.
95   \label{introEquation:halmitonianPrinciple1}
96   \end{equation}
98
97   For simple mechanical systems, where the forces acting on the
98   different parts are derivable from a potential, the Lagrangian
99   function $L$ can be defined as the difference between the kinetic
100   energy of the system and its potential energy,
101   \begin{equation}
102 < L \equiv K - U = L(q_i ,\dot q_i ) ,
102 > L \equiv K - U = L(q_i ,\dot q_i ).
103   \label{introEquation:lagrangianDef}
104   \end{equation}
105 < then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
105 > Thus, Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
106   \begin{equation}
107 < \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
107 > \delta \int_{t_1 }^{t_2 } {L dt = 0} .
108   \label{introEquation:halmitonianPrinciple2}
109   \end{equation}
110  
# Line 138 | Line 136 | p_i  = \frac{{\partial L}}{{\partial q_i }}
136   p_i  = \frac{{\partial L}}{{\partial q_i }}
137   \label{introEquation:generalizedMomentaDot}
138   \end{equation}
141
139   With the help of the generalized momenta, we may now define a new
140   quantity $H$ by the equation
141   \begin{equation}
# Line 146 | Line 143 | $L$ is the Lagrangian function for the system.
143   \label{introEquation:hamiltonianDefByLagrangian}
144   \end{equation}
145   where $ \dot q_1  \ldots \dot q_f $ are generalized velocities and
146 < $L$ is the Lagrangian function for the system.
147 <
151 < Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
152 < one can obtain
146 > $L$ is the Lagrangian function for the system. Differentiating
147 > Eq.~\ref{introEquation:hamiltonianDefByLagrangian}, one can obtain
148   \begin{equation}
149   dH = \sum\limits_k {\left( {p_k d\dot q_k  + \dot q_k dp_k  -
150   \frac{{\partial L}}{{\partial q_k }}dq_k  - \frac{{\partial
151   L}}{{\partial \dot q_k }}d\dot q_k } \right)}  - \frac{{\partial
152 < L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
152 > L}}{{\partial t}}dt . \label{introEquation:diffHamiltonian1}
153   \end{equation}
154 < Making use of  Eq.~\ref{introEquation:generalizedMomenta}, the
155 < second and fourth terms in the parentheses cancel. Therefore,
154 > Making use of Eq.~\ref{introEquation:generalizedMomenta}, the second
155 > and fourth terms in the parentheses cancel. Therefore,
156   Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
157   \begin{equation}
158   dH = \sum\limits_k {\left( {\dot q_k dp_k  - \dot p_k dq_k }
159 < \right)}  - \frac{{\partial L}}{{\partial t}}dt
159 > \right)}  - \frac{{\partial L}}{{\partial t}}dt .
160   \label{introEquation:diffHamiltonian2}
161   \end{equation}
162   By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
# Line 180 | Line 175 | t}}
175   t}}
176   \label{introEquation:motionHamiltonianTime}
177   \end{equation}
178 <
184 < Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
178 > where Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
179   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
180   equation of motion. Due to their symmetrical formula, they are also
181 < known as the canonical equations of motions \cite{Goldstein2001}.
181 > known as the canonical equations of motions.\cite{Goldstein2001}
182  
183   An important difference between Lagrangian approach and the
184   Hamiltonian approach is that the Lagrangian is considered to be a
# Line 194 | Line 188 | only works with 1st-order differential equations\cite{
188   Hamiltonian Mechanics is more appropriate for application to
189   statistical mechanics and quantum mechanics, since it treats the
190   coordinate and its time derivative as independent variables and it
191 < only works with 1st-order differential equations\cite{Marion1990}.
198 <
191 > only works with 1st-order differential equations.\cite{Marion1990}
192   In Newtonian Mechanics, a system described by conservative forces
193 < conserves the total energy \ref{introEquation:energyConservation}.
194 < It follows that Hamilton's equations of motion conserve the total
195 < Hamiltonian.
193 > conserves the total energy
194 > (Eq.~\ref{introEquation:energyConservation}). It follows that
195 > Hamilton's equations of motion conserve the total Hamiltonian
196   \begin{equation}
197   \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
198   H}}{{\partial q_i }}\dot q_i  + \frac{{\partial H}}{{\partial p_i
199   }}\dot p_i } \right)}  = \sum\limits_i {\left( {\frac{{\partial
200   H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
201   \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
202 < q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
202 > q_i }}} \right) = 0}. \label{introEquation:conserveHalmitonian}
203   \end{equation}
204  
205   \section{\label{introSection:statisticalMechanics}Statistical
# Line 215 | Line 208 | Statistical Mechanics concepts and theorem presented i
208   The thermodynamic behaviors and properties of Molecular Dynamics
209   simulation are governed by the principle of Statistical Mechanics.
210   The following section will give a brief introduction to some of the
211 < Statistical Mechanics concepts and theorem presented in this
211 > Statistical Mechanics concepts and theorems presented in this
212   dissertation.
213  
214   \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
215  
216   Mathematically, phase space is the space which represents all
217 < possible states. Each possible state of the system corresponds to
218 < one unique point in the phase space. For mechanical systems, the
219 < phase space usually consists of all possible values of position and
220 < momentum variables. Consider a dynamic system of $f$ particles in a
221 < cartesian space, where each of the $6f$ coordinates and momenta is
222 < assigned to one of $6f$ mutually orthogonal axes, the phase space of
223 < this system is a $6f$ dimensional space. A point, $x = (\rightarrow
224 < q_1 , \ldots ,\rightarrow q_f ,\rightarrow p_1 , \ldots ,\rightarrow
225 < p_f )$, with a unique set of values of $6f$ coordinates and momenta
226 < is a phase space vector.
217 > possible states of a system. Each possible state of the system
218 > corresponds to one unique point in the phase space. For mechanical
219 > systems, the phase space usually consists of all possible values of
220 > position and momentum variables. Consider a dynamic system of $f$
221 > particles in a cartesian space, where each of the $6f$ coordinates
222 > and momenta is assigned to one of $6f$ mutually orthogonal axes, the
223 > phase space of this system is a $6f$ dimensional space. A point, $x
224 > =
225 > (\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
226 > \over q} _1 , \ldots
227 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
228 > \over q} _f
229 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
230 > \over p} _1  \ldots
231 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
232 > \over p} _f )$ , with a unique set of values of $6f$ coordinates and
233 > momenta is a phase space vector.
234   %%%fix me
235  
236   In statistical mechanics, the condition of an ensemble at any time
# Line 243 | Line 243 | their locations which would change the density at any
243   \label{introEquation:densityDistribution}
244   \end{equation}
245   Governed by the principles of mechanics, the phase points change
246 < their locations which would change the density at any time at phase
246 > their locations which changes the density at any time at phase
247   space. Hence, the density distribution is also to be taken as a
248 < function of the time.
249 <
250 < The number of systems $\delta N$ at time $t$ can be determined by,
248 > function of the time. The number of systems $\delta N$ at time $t$
249 > can be determined by,
250   \begin{equation}
251   \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
252   \label{introEquation:deltaN}
253   \end{equation}
254 < Assuming a large enough population of systems, we can sufficiently
254 > Assuming enough copies of the systems, we can sufficiently
255   approximate $\delta N$ without introducing discontinuity when we go
256   from one region in the phase space to another. By integrating over
257   the whole phase space,
# Line 260 | Line 259 | gives us an expression for the total number of the sys
259   N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
260   \label{introEquation:totalNumberSystem}
261   \end{equation}
262 < gives us an expression for the total number of the systems. Hence,
263 < the probability per unit in the phase space can be obtained by,
262 > gives us an expression for the total number of copies. Hence, the
263 > probability per unit volume in the phase space can be obtained by,
264   \begin{equation}
265   \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
266   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
# Line 270 | Line 269 | momenta of the system. Even when the dynamics of the r
269   With the help of Eq.~\ref{introEquation:unitProbability} and the
270   knowledge of the system, it is possible to calculate the average
271   value of any desired quantity which depends on the coordinates and
272 < momenta of the system. Even when the dynamics of the real system is
272 > momenta of the system. Even when the dynamics of the real system are
273   complex, or stochastic, or even discontinuous, the average
274 < properties of the ensemble of possibilities as a whole remaining
275 < well defined. For a classical system in thermal equilibrium with its
274 > properties of the ensemble of possibilities as a whole remain well
275 > defined. For a classical system in thermal equilibrium with its
276   environment, the ensemble average of a mechanical quantity, $\langle
277   A(q , p) \rangle_t$, takes the form of an integral over the phase
278   space of the system,
279   \begin{equation}
280   \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
281   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
282 < (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
282 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
283   \label{introEquation:ensembelAverage}
284   \end{equation}
285  
287 There are several different types of ensembles with different
288 statistical characteristics. As a function of macroscopic
289 parameters, such as temperature \textit{etc}, the partition function
290 can be used to describe the statistical properties of a system in
291 thermodynamic equilibrium.
292
293 As an ensemble of systems, each of which is known to be thermally
294 isolated and conserve energy, the Microcanonical ensemble (NVE) has
295 a partition function like,
296 \begin{equation}
297 \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
298 \end{equation}
299 A canonical ensemble (NVT)is an ensemble of systems, each of which
300 can share its energy with a large heat reservoir. The distribution
301 of the total energy amongst the possible dynamical states is given
302 by the partition function,
303 \begin{equation}
304 \Omega (N,V,T) = e^{ - \beta A}
305 \label{introEquation:NVTPartition}
306 \end{equation}
307 Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
308 TS$. Since most experiments are carried out under constant pressure
309 condition, the isothermal-isobaric ensemble (NPT) plays a very
310 important role in molecular simulations. The isothermal-isobaric
311 ensemble allow the system to exchange energy with a heat bath of
312 temperature $T$ and to change the volume as well. Its partition
313 function is given as
314 \begin{equation}
315 \Delta (N,P,T) =  - e^{\beta G}.
316 \label{introEquation:NPTPartition}
317 \end{equation}
318 Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
319
286   \subsection{\label{introSection:liouville}Liouville's theorem}
287  
288   Liouville's theorem is the foundation on which statistical mechanics
# Line 358 | Line 324 | simple form,
324   \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
325   \label{introEquation:liouvilleTheorem}
326   \end{equation}
361
327   Liouville's theorem states that the distribution function is
328   constant along any trajectory in phase space. In classical
329 < statistical mechanics, since the number of members in an ensemble is
330 < huge and constant, we can assume the local density has no reason
331 < (other than classical mechanics) to change,
329 > statistical mechanics, since the number of system copies in an
330 > ensemble is huge and constant, we can assume the local density has
331 > no reason (other than classical mechanics) to change,
332   \begin{equation}
333   \frac{{\partial \rho }}{{\partial t}} = 0.
334   \label{introEquation:stationary}
# Line 393 | Line 358 | With the help of stationary assumption
358   \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
359   \frac{d}{{dt}}(\delta v) = 0.
360   \end{equation}
361 < With the help of stationary assumption
362 < (\ref{introEquation:stationary}), we obtain the principle of the
361 > With the help of the stationary assumption
362 > (Eq.~\ref{introEquation:stationary}), we obtain the principle of
363   \emph{conservation of volume in phase space},
364   \begin{equation}
365   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
# Line 404 | Line 369 | Liouville's theorem can be expresses in a variety of d
369  
370   \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
371  
372 < Liouville's theorem can be expresses in a variety of different forms
372 > Liouville's theorem can be expressed in a variety of different forms
373   which are convenient within different contexts. For any two function
374   $F$ and $G$ of the coordinates and momenta of a system, the Poisson
375 < bracket ${F, G}$ is defined as
375 > bracket $\{F,G\}$ is defined as
376   \begin{equation}
377   \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
378   F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
# Line 415 | Line 380 | Substituting equations of motion in Hamiltonian formal
380   q_i }}} \right)}.
381   \label{introEquation:poissonBracket}
382   \end{equation}
383 < Substituting equations of motion in Hamiltonian formalism(
384 < Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
385 < Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
383 > Substituting equations of motion in Hamiltonian formalism
384 > (Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
385 > Eq.~\ref{introEquation:motionHamiltonianMomentum}) into
386   (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
387   Liouville's theorem using Poisson bracket notion,
388   \begin{equation}
# Line 438 | Line 403 | expressed as
403   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
404   \label{introEquation:liouvilleTheoremInOperator}
405   \end{equation}
406 <
406 > which can help define a propagator $\rho (t) = e^{-iLt} \rho (0)$.
407   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
408  
409   Various thermodynamic properties can be calculated from Molecular
# Line 447 | Line 412 | period of them which is different from the average beh
412   simulation and the quality of the underlying model. However, both
413   experiments and computer simulations are usually performed during a
414   certain time interval and the measurements are averaged over a
415 < period of them which is different from the average behavior of
415 > period of time which is different from the average behavior of
416   many-body system in Statistical Mechanics. Fortunately, the Ergodic
417   Hypothesis makes a connection between time average and the ensemble
418   average. It states that the time average and average over the
419 < statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
419 > statistical ensemble are identical:\cite{Frenkel1996, Leach2001}
420   \begin{equation}
421   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
422   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 460 | Line 425 | sufficiently long time (longer than relaxation time),
425   where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
426   physical quantity and $\rho (p(t), q(t))$ is the equilibrium
427   distribution function. If an observation is averaged over a
428 < sufficiently long time (longer than relaxation time), all accessible
429 < microstates in phase space are assumed to be equally probed, giving
430 < a properly weighted statistical average. This allows the researcher
431 < freedom of choice when deciding how best to measure a given
432 < observable. In case an ensemble averaged approach sounds most
433 < reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
428 > sufficiently long time (longer than the relaxation time), all
429 > accessible microstates in phase space are assumed to be equally
430 > probed, giving a properly weighted statistical average. This allows
431 > the researcher freedom of choice when deciding how best to measure a
432 > given observable. In case an ensemble averaged approach sounds most
433 > reasonable, the Monte Carlo methods\cite{Metropolis1949} can be
434   utilized. Or if the system lends itself to a time averaging
435   approach, the Molecular Dynamics techniques in
436   Sec.~\ref{introSection:molecularDynamics} will be the best
437 < choice\cite{Frenkel1996}.
437 > choice.\cite{Frenkel1996}
438  
439   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
440   A variety of numerical integrators have been proposed to simulate
441   the motions of atoms in MD simulation. They usually begin with
442 < initial conditionals and move the objects in the direction governed
443 < by the differential equations. However, most of them ignore the
444 < hidden physical laws contained within the equations. Since 1990,
445 < geometric integrators, which preserve various phase-flow invariants
446 < such as symplectic structure, volume and time reversal symmetry, are
447 < developed to address this issue\cite{Dullweber1997, McLachlan1998,
448 < Leimkuhler1999}. The velocity Verlet method, which happens to be a
442 > initial conditions and move the objects in the direction governed by
443 > the differential equations. However, most of them ignore the hidden
444 > physical laws contained within the equations. Since 1990, geometric
445 > integrators, which preserve various phase-flow invariants such as
446 > symplectic structure, volume and time reversal symmetry, were
447 > developed to address this issue.\cite{Dullweber1997, McLachlan1998,
448 > Leimkuhler1999} The velocity Verlet method, which happens to be a
449   simple example of symplectic integrator, continues to gain
450   popularity in the molecular dynamics community. This fact can be
451   partly explained by its geometric nature.
# Line 492 | Line 457 | apply calculus on \emph{differentiable manifold}. A \e
457   surface of Earth. It seems to be flat locally, but it is round if
458   viewed as a whole. A \emph{differentiable manifold} (also known as
459   \emph{smooth manifold}) is a manifold on which it is possible to
460 < apply calculus on \emph{differentiable manifold}. A \emph{symplectic
461 < manifold} is defined as a pair $(M, \omega)$ which consists of a
462 < \emph{differentiable manifold} $M$ and a close, non-degenerated,
460 > apply calculus.\cite{Hirsch1997} A \emph{symplectic manifold} is
461 > defined as a pair $(M, \omega)$ which consists of a
462 > \emph{differentiable manifold} $M$ and a close, non-degenerate,
463   bilinear symplectic form, $\omega$. A symplectic form on a vector
464   space $V$ is a function $\omega(x, y)$ which satisfies
465   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
466   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
467 < $\omega(x, x) = 0$. The cross product operation in vector field is
468 < an example of symplectic form.
467 > $\omega(x, x) = 0$.\cite{McDuff1998} The cross product operation in
468 > vector field is an example of symplectic form. One of the
469 > motivations to study \emph{symplectic manifolds} in Hamiltonian
470 > Mechanics is that a symplectic manifold can represent all possible
471 > configurations of the system and the phase space of the system can
472 > be described by it's cotangent bundle.\cite{Jost2002} Every
473 > symplectic manifold is even dimensional. For instance, in Hamilton
474 > equations, coordinate and momentum always appear in pairs.
475  
505 One of the motivations to study \emph{symplectic manifolds} in
506 Hamiltonian Mechanics is that a symplectic manifold can represent
507 all possible configurations of the system and the phase space of the
508 system can be described by it's cotangent bundle. Every symplectic
509 manifold is even dimensional. For instance, in Hamilton equations,
510 coordinate and momentum always appear in pairs.
511
476   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
477  
478   For an ordinary differential system defined as
479   \begin{equation}
480   \dot x = f(x)
481   \end{equation}
482 < where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
483 < \begin{equation}
484 < f(r) = J\nabla _x H(r).
521 < \end{equation}
522 < $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
523 < matrix
482 > where $x = x(q,p)$, this system is a canonical Hamiltonian, if
483 > $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
484 > function and $J$ is the skew-symmetric matrix
485   \begin{equation}
486   J = \left( {\begin{array}{*{20}c}
487     0 & I  \\
# Line 531 | Line 492 | system can be rewritten as,
492   where $I$ is an identity matrix. Using this notation, Hamiltonian
493   system can be rewritten as,
494   \begin{equation}
495 < \frac{d}{{dt}}x = J\nabla _x H(x)
495 > \frac{d}{{dt}}x = J\nabla _x H(x).
496   \label{introEquation:compactHamiltonian}
497   \end{equation}In this case, $f$ is
498 < called a \emph{Hamiltonian vector field}.
499 <
539 < Another generalization of Hamiltonian dynamics is Poisson
540 < Dynamics\cite{Olver1986},
498 > called a \emph{Hamiltonian vector field}. Another generalization of
499 > Hamiltonian dynamics is Poisson Dynamics,\cite{Olver1986}
500   \begin{equation}
501   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
502   \end{equation}
503 < The most obvious change being that matrix $J$ now depends on $x$.
503 > where the most obvious change being that matrix $J$ now depends on
504 > $x$.
505  
506 < \subsection{\label{introSection:exactFlow}Exact Flow}
506 > \subsection{\label{introSection:exactFlow}Exact Propagator}
507  
508 < Let $x(t)$ be the exact solution of the ODE system,
508 > Let $x(t)$ be the exact solution of the ODE
509 > system,
510   \begin{equation}
511 < \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
512 < \end{equation}
513 < The exact flow(solution) $\varphi_\tau$ is defined by
514 < \[
515 < x(t+\tau) =\varphi_\tau(x(t))
511 > \frac{{dx}}{{dt}} = f(x), \label{introEquation:ODE}
512 > \end{equation} we can
513 > define its exact propagator $\varphi_\tau$:
514 > \[ x(t+\tau)
515 > =\varphi_\tau(x(t))
516   \]
517   where $\tau$ is a fixed time step and $\varphi$ is a map from phase
518 < space to itself. The flow has the continuous group property,
518 > space to itself. The propagator has the continuous group property,
519   \begin{equation}
520   \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
521   + \tau _2 } .
# Line 563 | Line 524 | Therefore, the exact flow is self-adjoint,
524   \begin{equation}
525   \varphi _\tau   \circ \varphi _{ - \tau }  = I
526   \end{equation}
527 < Therefore, the exact flow is self-adjoint,
527 > Therefore, the exact propagator is self-adjoint,
528   \begin{equation}
529   \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
530   \end{equation}
531 < The exact flow can also be written in terms of the of an operator,
531 > The exact propagator can also be written as an operator,
532   \begin{equation}
533   \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
534   }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
535   \label{introEquation:exponentialOperator}
536   \end{equation}
537 <
538 < In most cases, it is not easy to find the exact flow $\varphi_\tau$.
539 < Instead, we use an approximate map, $\psi_\tau$, which is usually
540 < called integrator. The order of an integrator $\psi_\tau$ is $p$, if
541 < the Taylor series of $\psi_\tau$ agree to order $p$,
537 > In most cases, it is not easy to find the exact propagator
538 > $\varphi_\tau$. Instead, we use an approximate map, $\psi_\tau$,
539 > which is usually called an integrator. The order of an integrator
540 > $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
541 > order $p$,
542   \begin{equation}
543   \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
544   \end{equation}
# Line 585 | Line 546 | ODE and its flow play important roles in numerical stu
546   \subsection{\label{introSection:geometricProperties}Geometric Properties}
547  
548   The hidden geometric properties\cite{Budd1999, Marsden1998} of an
549 < ODE and its flow play important roles in numerical studies. Many of
550 < them can be found in systems which occur naturally in applications.
551 <
552 < Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
592 < a \emph{symplectic} flow if it satisfies,
549 > ODE and its propagator play important roles in numerical studies.
550 > Many of them can be found in systems which occur naturally in
551 > applications. Let $\varphi$ be the propagator of Hamiltonian vector
552 > field, $\varphi$ is a \emph{symplectic} propagator if it satisfies,
553   \begin{equation}
554   {\varphi '}^T J \varphi ' = J.
555   \end{equation}
556   According to Liouville's theorem, the symplectic volume is invariant
557 < under a Hamiltonian flow, which is the basis for classical
558 < statistical mechanics. Furthermore, the flow of a Hamiltonian vector
559 < field on a symplectic manifold can be shown to be a
557 > under a Hamiltonian propagator, which is the basis for classical
558 > statistical mechanics. Furthermore, the propagator of a Hamiltonian
559 > vector field on a symplectic manifold can be shown to be a
560   symplectomorphism. As to the Poisson system,
561   \begin{equation}
562   {\varphi '}^T J \varphi ' = J \circ \varphi
563   \end{equation}
564 < is the property that must be preserved by the integrator.
565 <
566 < It is possible to construct a \emph{volume-preserving} flow for a
567 < source free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $
568 < \det d\varphi  = 1$. One can show easily that a symplectic flow will
569 < be volume-preserving.
570 <
611 < Changing the variables $y = h(x)$ in an ODE
612 < (Eq.~\ref{introEquation:ODE}) will result in a new system,
564 > is the property that must be preserved by the integrator. It is
565 > possible to construct a \emph{volume-preserving} propagator for a
566 > source free ODE ($ \nabla \cdot f = 0 $), if the propagator
567 > satisfies $ \det d\varphi  = 1$. One can show easily that a
568 > symplectic propagator will be volume-preserving. Changing the
569 > variables $y = h(x)$ in an ODE (Eq.~\ref{introEquation:ODE}) will
570 > result in a new system,
571   \[
572   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
573   \]
574   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
575 < In other words, the flow of this vector field is reversible if and
576 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
577 <
578 < A \emph{first integral}, or conserved quantity of a general
579 < differential function is a function $ G:R^{2d}  \to R^d $ which is
622 < constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
575 > In other words, the propagator of this vector field is reversible if
576 > and only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
577 > conserved quantity of a general differential function is a function
578 > $ G:R^{2d}  \to R^d $ which is constant for all solutions of the ODE
579 > $\frac{{dx}}{{dt}} = f(x)$ ,
580   \[
581   \frac{{dG(x(t))}}{{dt}} = 0.
582   \]
583 < Using chain rule, one may obtain,
583 > Using the chain rule, one may obtain,
584   \[
585 < \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
585 > \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \cdot \nabla G,
586   \]
587 < which is the condition for conserving \emph{first integral}. For a
588 < canonical Hamiltonian system, the time evolution of an arbitrary
589 < smooth function $G$ is given by,
633 <
587 > which is the condition for conserved quantities. For a canonical
588 > Hamiltonian system, the time evolution of an arbitrary smooth
589 > function $G$ is given by,
590   \begin{eqnarray}
591 < \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
592 <                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
591 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \notag\\
592 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)).
593   \label{introEquation:firstIntegral1}
594   \end{eqnarray}
595 <
596 <
641 < Using poisson bracket notion, Equation
642 < \ref{introEquation:firstIntegral1} can be rewritten as
595 > Using poisson bracket notion, Eq.~\ref{introEquation:firstIntegral1}
596 > can be rewritten as
597   \[
598   \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
599   \]
600 < Therefore, the sufficient condition for $G$ to be the \emph{first
601 < integral} of a Hamiltonian system is
602 < \[
603 < \left\{ {G,H} \right\} = 0.
650 < \]
651 < As well known, the Hamiltonian (or energy) H of a Hamiltonian system
652 < is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
653 < 0$.
654 <
600 > Therefore, the sufficient condition for $G$ to be a conserved
601 > quantity of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$ As
602 > is well known, the Hamiltonian (or energy) H of a Hamiltonian system
603 > is a conserved quantity, which is due to the fact $\{ H,H\}  = 0$.
604   When designing any numerical methods, one should always try to
605 < preserve the structural properties of the original ODE and its flow.
605 > preserve the structural properties of the original ODE and its
606 > propagator.
607  
608   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
609   A lot of well established and very effective numerical methods have
610 < been successful precisely because of their symplecticities even
610 > been successful precisely because of their symplectic nature even
611   though this fact was not recognized when they were first
612   constructed. The most famous example is the Verlet-leapfrog method
613   in molecular dynamics. In general, symplectic integrators can be
# Line 668 | Line 618 | constructed using one of four different methods.
618   \item Runge-Kutta methods
619   \item Splitting methods
620   \end{enumerate}
621 <
672 < Generating function\cite{Channell1990} tends to lead to methods
621 > Generating functions\cite{Channell1990} tend to lead to methods
622   which are cumbersome and difficult to use. In dissipative systems,
623   variational methods can capture the decay of energy
624 < accurately\cite{Kane2000}. Since their geometrically unstable nature
624 > accurately.\cite{Kane2000} Since they are geometrically unstable
625   against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
626 < methods are not suitable for Hamiltonian system. Recently, various
627 < high-order explicit Runge-Kutta methods
628 < \cite{Owren1992,Chen2003}have been developed to overcome this
629 < instability. However, due to computational penalty involved in
630 < implementing the Runge-Kutta methods, they have not attracted much
631 < attention from the Molecular Dynamics community. Instead, splitting
632 < methods have been widely accepted since they exploit natural
633 < decompositions of the system\cite{Tuckerman1992, McLachlan1998}.
626 > methods are not suitable for Hamiltonian
627 > system.\cite{Cartwright1992} Recently, various high-order explicit
628 > Runge-Kutta methods \cite{Owren1992,Chen2003} have been developed to
629 > overcome this instability. However, due to computational penalty
630 > involved in implementing the Runge-Kutta methods, they have not
631 > attracted much attention from the Molecular Dynamics community.
632 > Instead, splitting methods have been widely accepted since they
633 > exploit natural decompositions of the system.\cite{McLachlan1998,
634 > Tuckerman1992}
635  
636   \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
637  
638   The main idea behind splitting methods is to decompose the discrete
639 < $\varphi_h$ as a composition of simpler flows,
639 > $\varphi_h$ as a composition of simpler propagators,
640   \begin{equation}
641   \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
642   \varphi _{h_n }
643   \label{introEquation:FlowDecomposition}
644   \end{equation}
645 < where each of the sub-flow is chosen such that each represent a
646 < simpler integration of the system.
647 <
698 < Suppose that a Hamiltonian system takes the form,
645 > where each of the sub-propagator is chosen such that each represent
646 > a simpler integration of the system. Suppose that a Hamiltonian
647 > system takes the form,
648   \[
649   H = H_1 + H_2.
650   \]
651   Here, $H_1$ and $H_2$ may represent different physical processes of
652   the system. For instance, they may relate to kinetic and potential
653   energy respectively, which is a natural decomposition of the
654 < problem. If $H_1$ and $H_2$ can be integrated using exact flows
655 < $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
656 < order expression is then given by the Lie-Trotter formula
654 > problem. If $H_1$ and $H_2$ can be integrated using exact
655 > propagators $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a
656 > simple first order expression is then given by the Lie-Trotter
657 > formula\cite{Trotter1959}
658   \begin{equation}
659   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
660   \label{introEquation:firstOrderSplitting}
# Line 713 | Line 663 | It is easy to show that any composition of symplectic
663   continuous $\varphi _i$ over a time $h$. By definition, as
664   $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
665   must follow that each operator $\varphi_i(t)$ is a symplectic map.
666 < It is easy to show that any composition of symplectic flows yields a
667 < symplectic map,
666 > It is easy to show that any composition of symplectic propagators
667 > yields a symplectic map,
668   \begin{equation}
669   (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
670   '\phi ' = \phi '^T J\phi ' = J,
# Line 722 | Line 672 | splitting in this context automatically generates a sy
672   \end{equation}
673   where $\phi$ and $\psi$ both are symplectic maps. Thus operator
674   splitting in this context automatically generates a symplectic map.
675 <
676 < The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
677 < introduces local errors proportional to $h^2$, while Strang
678 < splitting gives a second-order decomposition,
675 > The Lie-Trotter
676 > splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
677 > local errors proportional to $h^2$, while the Strang splitting gives
678 > a second-order decomposition,\cite{Strang1968}
679   \begin{equation}
680   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
681   _{1,h/2} , \label{introEquation:secondOrderSplitting}
682   \end{equation}
683 < which has a local error proportional to $h^3$. The Sprang
683 > which has a local error proportional to $h^3$. The Strang
684   splitting's popularity in molecular simulation community attribute
685   to its symmetric property,
686   \begin{equation}
# Line 758 | Line 708 | symplectic(\ref{introEquation:SymplecticFlowCompositio
708   \end{align}
709   where $F(t)$ is the force at time $t$. This integration scheme is
710   known as \emph{velocity verlet} which is
711 < symplectic(\ref{introEquation:SymplecticFlowComposition}),
712 < time-reversible(\ref{introEquation:timeReversible}) and
713 < volume-preserving (\ref{introEquation:volumePreserving}). These
711 > symplectic(Eq.~\ref{introEquation:SymplecticFlowComposition}),
712 > time-reversible(Eq.~\ref{introEquation:timeReversible}) and
713 > volume-preserving (Eq.~\ref{introEquation:volumePreserving}). These
714   geometric properties attribute to its long-time stability and its
715   popularity in the community. However, the most commonly used
716   velocity verlet integration scheme is written as below,
# Line 781 | Line 731 | the equations of motion would follow:
731  
732   \item Use the half step velocities to move positions one whole step, $\Delta t$.
733  
734 < \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
734 > \item Evaluate the forces at the new positions, $q(\Delta t)$, and use the new forces to complete the velocity move.
735  
736   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
737   \end{enumerate}
788
738   By simply switching the order of the propagators in the splitting
739   and composing a new integrator, the \emph{position verlet}
740   integrator, can be generated,
# Line 801 | Line 750 | The Baker-Campbell-Hausdorff formula can be used to de
750  
751   \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
752  
753 < The Baker-Campbell-Hausdorff formula can be used to determine the
754 < local error of splitting method in terms of the commutator of the
755 < operators(\ref{introEquation:exponentialOperator}) associated with
756 < the sub-flow. For operators $hX$ and $hY$ which are associated with
757 < $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
753 > The Baker-Campbell-Hausdorff formula\cite{Gilmore1974} can be used
754 > to determine the local error of a splitting method in terms of the
755 > commutator of the operators associated with the sub-propagator. For
756 > operators $hX$ and $hY$ which are associated with $\varphi_1(t)$ and
757 > $\varphi_2(t)$ respectively , we have
758   \begin{equation}
759   \exp (hX + hY) = \exp (hZ)
760   \end{equation}
# Line 814 | Line 763 | Here, $[X,Y]$ is the commutators of operator $X$ and $
763   hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
764   {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
765   \end{equation}
766 < Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
766 > Here, $[X,Y]$ is the commutator of operator $X$ and $Y$ given by
767   \[
768   [X,Y] = XY - YX .
769   \]
770   Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
771 < to the Sprang splitting, we can obtain
771 > to the Strang splitting, we can obtain
772   \begin{eqnarray*}
773   \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 \\
774                                     &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
775 <                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
775 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots
776 >                                   ).
777   \end{eqnarray*}
778 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
779 < error of Spring splitting is proportional to $h^3$. The same
780 < procedure can be applied to a general splitting,  of the form
778 > Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local
779 > error of Strang splitting is proportional to $h^3$. The same
780 > procedure can be applied to a general splitting of the form
781   \begin{equation}
782   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
783   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
784   \end{equation}
785   A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
786   order methods. Yoshida proposed an elegant way to compose higher
787 < order methods based on symmetric splitting\cite{Yoshida1990}. Given
787 > order methods based on symmetric splitting.\cite{Yoshida1990} Given
788   a symmetric second order base method $ \varphi _h^{(2)} $, a
789   fourth-order symmetric method can be constructed by composing,
790   \[
# Line 862 | Line 812 | simulations. For instance, instantaneous temperature o
812   dynamical information. The basic idea of molecular dynamics is that
813   macroscopic properties are related to microscopic behavior and
814   microscopic behavior can be calculated from the trajectories in
815 < simulations. For instance, instantaneous temperature of an
816 < Hamiltonian system of $N$ particle can be measured by
815 > simulations. For instance, instantaneous temperature of a
816 > Hamiltonian system of $N$ particles can be measured by
817   \[
818   T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
819   \]
820   where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
821   respectively, $f$ is the number of degrees of freedom, and $k_B$ is
822 < the boltzman constant.
822 > the Boltzman constant.
823  
824   A typical molecular dynamics run consists of three essential steps:
825   \begin{enumerate}
# Line 886 | Line 836 | will discusse issues in production run.
836   These three individual steps will be covered in the following
837   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
838   initialization of a simulation. Sec.~\ref{introSection:production}
839 < will discusse issues in production run.
839 > discusses issues of production runs.
840   Sec.~\ref{introSection:Analysis} provides the theoretical tools for
841 < trajectory analysis.
841 > analysis of trajectories.
842  
843   \subsection{\label{introSec:initialSystemSettings}Initialization}
844  
# Line 900 | Line 850 | structure, some important information is missing. For
850   thousands of crystal structures of molecules are discovered every
851   year, many more remain unknown due to the difficulties of
852   purification and crystallization. Even for molecules with known
853 < structure, some important information is missing. For example, a
853 > structures, some important information is missing. For example, a
854   missing hydrogen atom which acts as donor in hydrogen bonding must
855 < be added. Moreover, in order to include electrostatic interaction,
855 > be added. Moreover, in order to include electrostatic interactions,
856   one may need to specify the partial charges for individual atoms.
857   Under some circumstances, we may even need to prepare the system in
858   a special configuration. For instance, when studying transport
# Line 920 | Line 870 | near the minimum, steepest descent method is extremely
870   minimization to find a more reasonable conformation. Several energy
871   minimization methods have been developed to exploit the energy
872   surface and to locate the local minimum. While converging slowly
873 < near the minimum, steepest descent method is extremely robust when
873 > near the minimum, the steepest descent method is extremely robust when
874   systems are strongly anharmonic. Thus, it is often used to refine
875 < structure from crystallographic data. Relied on the gradient or
876 < hessian, advanced methods like Newton-Raphson converge rapidly to a
877 < local minimum, but become unstable if the energy surface is far from
875 > structures from crystallographic data. Relying on the Hessian,
876 > advanced methods like Newton-Raphson converge rapidly to a local
877 > minimum, but become unstable if the energy surface is far from
878   quadratic. Another factor that must be taken into account, when
879   choosing energy minimization method, is the size of the system.
880   Steepest descent and conjugate gradient can deal with models of any
881   size. Because of the limits on computer memory to store the hessian
882 < matrix and the computing power needed to diagonalized these
883 < matrices, most Newton-Raphson methods can not be used with very
934 < large systems.
882 > matrix and the computing power needed to diagonalize these matrices,
883 > most Newton-Raphson methods can not be used with very large systems.
884  
885   \subsubsection{\textbf{Heating}}
886  
887 < Typically, Heating is performed by assigning random velocities
887 > Typically, heating is performed by assigning random velocities
888   according to a Maxwell-Boltzman distribution for a desired
889   temperature. Beginning at a lower temperature and gradually
890   increasing the temperature by assigning larger random velocities, we
891 < end up with setting the temperature of the system to a final
892 < temperature at which the simulation will be conducted. In heating
893 < phase, we should also keep the system from drifting or rotating as a
894 < whole. To do this, the net linear momentum and angular momentum of
895 < the system is shifted to zero after each resampling from the Maxwell
896 < -Boltzman distribution.
891 > end up setting the temperature of the system to a final temperature
892 > at which the simulation will be conducted. In the heating phase, we
893 > should also keep the system from drifting or rotating as a whole. To
894 > do this, the net linear momentum and angular momentum of the system
895 > is shifted to zero after each resampling from the Maxwell -Boltzman
896 > distribution.
897  
898   \subsubsection{\textbf{Equilibration}}
899  
# Line 955 | Line 904 | as a means to arrive at an equilibrated structure in a
904   properties \textit{etc}, become independent of time. Strictly
905   speaking, minimization and heating are not necessary, provided the
906   equilibration process is long enough. However, these steps can serve
907 < as a means to arrive at an equilibrated structure in an effective
907 > as a mean to arrive at an equilibrated structure in an effective
908   way.
909  
910   \subsection{\label{introSection:production}Production}
# Line 971 | Line 920 | which making large simulations prohibitive in the abse
920   calculation of non-bonded forces, such as van der Waals force and
921   Coulombic forces \textit{etc}. For a system of $N$ particles, the
922   complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
923 < which making large simulations prohibitive in the absence of any
924 < algorithmic tricks.
925 <
926 < A natural approach to avoid system size issues is to represent the
927 < bulk behavior by a finite number of the particles. However, this
928 < approach will suffer from the surface effect at the edges of the
929 < simulation. To offset this, \textit{Periodic boundary conditions}
930 < (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
931 < properties with a relatively small number of particles. In this
932 < method, the simulation box is replicated throughout space to form an
933 < infinite lattice. During the simulation, when a particle moves in
934 < the primary cell, its image in other cells move in exactly the same
935 < direction with exactly the same orientation. Thus, as a particle
987 < leaves the primary cell, one of its images will enter through the
988 < opposite face.
923 > which makes large simulations prohibitive in the absence of any
924 > algorithmic tricks. A natural approach to avoid system size issues
925 > is to represent the bulk behavior by a finite number of the
926 > particles. However, this approach will suffer from surface effects
927 > at the edges of the simulation. To offset this, \textit{Periodic
928 > boundary conditions} (see Fig.~\ref{introFig:pbc}) were developed to
929 > simulate bulk properties with a relatively small number of
930 > particles. In this method, the simulation box is replicated
931 > throughout space to form an infinite lattice. During the simulation,
932 > when a particle moves in the primary cell, its image in other cells
933 > move in exactly the same direction with exactly the same
934 > orientation. Thus, as a particle leaves the primary cell, one of its
935 > images will enter through the opposite face.
936   \begin{figure}
937   \centering
938   \includegraphics[width=\linewidth]{pbc.eps}
# Line 997 | Line 944 | evaluation is to apply spherical cutoff where particle
944  
945   %cutoff and minimum image convention
946   Another important technique to improve the efficiency of force
947 < evaluation is to apply spherical cutoff where particles farther than
948 < a predetermined distance are not included in the calculation
949 < \cite{Frenkel1996}. The use of a cutoff radius will cause a
950 < discontinuity in the potential energy curve. Fortunately, one can
951 < shift simple radial potential to ensure the potential curve go
947 > evaluation is to apply spherical cutoffs where particles farther
948 > than a predetermined distance are not included in the
949 > calculation.\cite{Frenkel1996} The use of a cutoff radius will cause
950 > a discontinuity in the potential energy curve. Fortunately, one can
951 > shift a simple radial potential to ensure the potential curve go
952   smoothly to zero at the cutoff radius. The cutoff strategy works
953   well for Lennard-Jones interaction because of its short range
954   nature. However, simply truncating the electrostatic interaction
# Line 1009 | Line 956 | periodicity artifacts in liquid simulations. Taking th
956   in simulations. The Ewald summation, in which the slowly decaying
957   Coulomb potential is transformed into direct and reciprocal sums
958   with rapid and absolute convergence, has proved to minimize the
959 < periodicity artifacts in liquid simulations. Taking the advantages
960 < of the fast Fourier transform (FFT) for calculating discrete Fourier
961 < transforms, the particle mesh-based
959 > periodicity artifacts in liquid simulations. Taking advantage of
960 > fast Fourier transform (FFT) techniques for calculating discrete
961 > Fourier transforms, the particle mesh-based
962   methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
963   $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
964   \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
# Line 1021 | Line 968 | his coworkers\cite{Wolf1999}. The shifted Coulomb pote
968   simulation community, these two methods are difficult to implement
969   correctly and efficiently. Instead, we use a damped and
970   charge-neutralized Coulomb potential method developed by Wolf and
971 < his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
971 > his coworkers.\cite{Wolf1999} The shifted Coulomb potential for
972   particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
973   \begin{equation}
974   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
975   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
976   R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
977 < r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
977 > r_{ij})}{r_{ij}}\right\}, \label{introEquation:shiftedCoulomb}
978   \end{equation}
979   where $\alpha$ is the convergence parameter. Due to the lack of
980   inherent periodicity and rapid convergence,this method is extremely
# Line 1044 | Line 991 | Recently, advanced visualization technique have become
991  
992   \subsection{\label{introSection:Analysis} Analysis}
993  
994 < Recently, advanced visualization technique have become applied to
994 > Recently, advanced visualization techniques have been applied to
995   monitor the motions of molecules. Although the dynamics of the
996   system can be described qualitatively from animation, quantitative
997 < trajectory analysis are more useful. According to the principles of
998 < Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
999 < one can compute thermodynamic properties, analyze fluctuations of
1000 < structural parameters, and investigate time-dependent processes of
1001 < the molecule from the trajectories.
997 > trajectory analysis is more useful. According to the principles of
998 > Statistical Mechanics in
999 > Sec.~\ref{introSection:statisticalMechanics}, one can compute
1000 > thermodynamic properties, analyze fluctuations of structural
1001 > parameters, and investigate time-dependent processes of the molecule
1002 > from the trajectories.
1003  
1004   \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1005  
# Line 1081 | Line 1029 | Experimentally, pair distribution function can be gath
1029   distribution functions. Among these functions,the \emph{pair
1030   distribution function}, also known as \emph{radial distribution
1031   function}, is of most fundamental importance to liquid theory.
1032 < Experimentally, pair distribution function can be gathered by
1032 > Experimentally, pair distribution functions can be gathered by
1033   Fourier transforming raw data from a series of neutron diffraction
1034 < experiments and integrating over the surface factor
1035 < \cite{Powles1973}. The experimental results can serve as a criterion
1036 < to justify the correctness of a liquid model. Moreover, various
1037 < equilibrium thermodynamic and structural properties can also be
1038 < expressed in terms of radial distribution function \cite{Allen1987}.
1039 <
1040 < The pair distribution functions $g(r)$ gives the probability that a
1041 < particle $i$ will be located at a distance $r$ from a another
1042 < particle $j$ in the system
1095 < \[
1034 > experiments and integrating over the surface
1035 > factor.\cite{Powles1973} The experimental results can serve as a
1036 > criterion to justify the correctness of a liquid model. Moreover,
1037 > various equilibrium thermodynamic and structural properties can also
1038 > be expressed in terms of the radial distribution
1039 > function.\cite{Allen1987} The pair distribution functions $g(r)$
1040 > gives the probability that a particle $i$ will be located at a
1041 > distance $r$ from a another particle $j$ in the system
1042 > \begin{equation}
1043   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1044   \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1045   (r)}{\rho}.
1046 < \]
1046 > \end{equation}
1047   Note that the delta function can be replaced by a histogram in
1048   computer simulation. Peaks in $g(r)$ represent solvent shells, and
1049   the height of these peaks gradually decreases to 1 as the liquid of
# Line 1114 | Line 1061 | function is called an \emph{autocorrelation function}.
1061   \label{introEquation:timeCorrelationFunction}
1062   \end{equation}
1063   If $A$ and $B$ refer to same variable, this kind of correlation
1064 < function is called an \emph{autocorrelation function}. One example
1118 < of an auto correlation function is the velocity auto-correlation
1064 > functions are called \emph{autocorrelation functions}. One typical example is the velocity autocorrelation
1065   function which is directly related to transport properties of
1066   molecular liquids:
1067 < \[
1067 > \begin{equation}
1068   D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1069   \right\rangle } dt
1070 < \]
1070 > \end{equation}
1071   where $D$ is diffusion constant. Unlike the velocity autocorrelation
1072 < function, which is averaging over time origins and over all the
1073 < atoms, the dipole autocorrelation functions are calculated for the
1072 > function, which is averaged over time origins and over all the
1073 > atoms, the dipole autocorrelation functions is calculated for the
1074   entire system. The dipole autocorrelation function is given by:
1075 < \[
1075 > \begin{equation}
1076   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1077   \right\rangle
1078 < \]
1078 > \end{equation}
1079   Here $u_{tot}$ is the net dipole of the entire system and is given
1080   by
1081 < \[
1082 < u_{tot} (t) = \sum\limits_i {u_i (t)}
1083 < \]
1084 < In principle, many time correlation functions can be related with
1081 > \begin{equation}
1082 > u_{tot} (t) = \sum\limits_i {u_i (t)}.
1083 > \end{equation}
1084 > In principle, many time correlation functions can be related to
1085   Fourier transforms of the infrared, Raman, and inelastic neutron
1086   scattering spectra of molecular liquids. In practice, one can
1087 < extract the IR spectrum from the intensity of dipole fluctuation at
1088 < each frequency using the following relationship:
1089 < \[
1087 > extract the IR spectrum from the intensity of the molecular dipole
1088 > fluctuation at each frequency using the following relationship:
1089 > \begin{equation}
1090   \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1091 < i2\pi vt} dt}
1092 < \]
1091 > i2\pi vt} dt}.
1092 > \end{equation}
1093  
1094   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1095  
1096   Rigid bodies are frequently involved in the modeling of different
1097 < areas, from engineering, physics, to chemistry. For example,
1098 < missiles and vehicle are usually modeled by rigid bodies.  The
1099 < movement of the objects in 3D gaming engine or other physics
1100 < simulator is governed by rigid body dynamics. In molecular
1097 > areas, including engineering, physics and chemistry. For example,
1098 > missiles and vehicles are usually modeled by rigid bodies.  The
1099 > movement of the objects in 3D gaming engines or other physics
1100 > simulators is governed by rigid body dynamics. In molecular
1101   simulations, rigid bodies are used to simplify protein-protein
1102 < docking studies\cite{Gray2003}.
1102 > docking studies.\cite{Gray2003}
1103  
1104   It is very important to develop stable and efficient methods to
1105   integrate the equations of motion for orientational degrees of
1106   freedom. Euler angles are the natural choice to describe the
1107   rotational degrees of freedom. However, due to $\frac {1}{sin
1108   \theta}$ singularities, the numerical integration of corresponding
1109 < equations of motion is very inefficient and inaccurate. Although an
1110 < alternative integrator using multiple sets of Euler angles can
1111 < overcome this difficulty\cite{Barojas1973}, the computational
1112 < penalty and the loss of angular momentum conservation still remain.
1113 < A singularity-free representation utilizing quaternions was
1114 < developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1115 < approach uses a nonseparable Hamiltonian resulting from the
1116 < quaternion representation, which prevents the symplectic algorithm
1117 < to be utilized. Another different approach is to apply holonomic
1118 < constraints to the atoms belonging to the rigid body. Each atom
1119 < moves independently under the normal forces deriving from potential
1120 < energy and constraint forces which are used to guarantee the
1121 < rigidness. However, due to their iterative nature, the SHAKE and
1122 < Rattle algorithms also converge very slowly when the number of
1123 < constraints increases\cite{Ryckaert1977, Andersen1983}.
1109 > equations of these motion is very inefficient and inaccurate.
1110 > Although an alternative integrator using multiple sets of Euler
1111 > angles can overcome this difficulty\cite{Barojas1973}, the
1112 > computational penalty and the loss of angular momentum conservation
1113 > still remain. A singularity-free representation utilizing
1114 > quaternions was developed by Evans in 1977.\cite{Evans1977}
1115 > Unfortunately, this approach used a nonseparable Hamiltonian
1116 > resulting from the quaternion representation, which prevented the
1117 > symplectic algorithm from being utilized. Another different approach
1118 > is to apply holonomic constraints to the atoms belonging to the
1119 > rigid body. Each atom moves independently under the normal forces
1120 > deriving from potential energy and constraint forces which are used
1121 > to guarantee the rigidness. However, due to their iterative nature,
1122 > the SHAKE and Rattle algorithms also converge very slowly when the
1123 > number of constraints increases.\cite{Ryckaert1977, Andersen1983}
1124  
1125   A break-through in geometric literature suggests that, in order to
1126   develop a long-term integration scheme, one should preserve the
1127 < symplectic structure of the flow. By introducing a conjugate
1127 > symplectic structure of the propagator. By introducing a conjugate
1128   momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1129   equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1130   proposed to evolve the Hamiltonian system in a constraint manifold
1131   by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1132   An alternative method using the quaternion representation was
1133 < developed by Omelyan\cite{Omelyan1998}. However, both of these
1133 > developed by Omelyan.\cite{Omelyan1998} However, both of these
1134   methods are iterative and inefficient. In this section, we descibe a
1135 < symplectic Lie-Poisson integrator for rigid body developed by
1135 > symplectic Lie-Poisson integrator for rigid bodies developed by
1136   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1137  
1138   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1139 < The motion of a rigid body is Hamiltonian with the Hamiltonian
1194 < function
1139 > The Hamiltonian of a rigid body is given by
1140   \begin{equation}
1141   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1142   V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1143   \label{introEquation:RBHamiltonian}
1144   \end{equation}
1145 < Here, $q$ and $Q$  are the position and rotation matrix for the
1146 < rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1147 < $J$, a diagonal matrix, is defined by
1145 > Here, $q$ and $Q$  are the position vector and rotation matrix for
1146 > the rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ ,
1147 > and $J$, a diagonal matrix, is defined by
1148   \[
1149   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1150   \]
# Line 1209 | Line 1154 | which is used to ensure rotation matrix's unitarity. D
1154   \begin{equation}
1155   Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1156   \end{equation}
1157 < which is used to ensure rotation matrix's unitarity. Differentiating
1158 < \ref{introEquation:orthogonalConstraint} and using Equation
1159 < \ref{introEquation:RBMotionMomentum}, one may obtain,
1215 < \begin{equation}
1216 < Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1217 < \label{introEquation:RBFirstOrderConstraint}
1218 < \end{equation}
1219 <
1220 < Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1221 < \ref{introEquation:motionHamiltonianMomentum}), one can write down
1157 > which is used to ensure the rotation matrix's unitarity. Using
1158 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} and Eq.~
1159 > \ref{introEquation:motionHamiltonianMomentum}, one can write down
1160   the equations of motion,
1223
1161   \begin{eqnarray}
1162 < \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1163 < \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1164 < \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1162 > \frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\
1163 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q), \label{introEquation:RBMotionMomentum}\\
1164 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1},  \label{introEquation:RBMotionRotation}\\
1165   \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1166   \end{eqnarray}
1167 <
1167 > Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and
1168 > using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1169 > \begin{equation}
1170 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1171 > \label{introEquation:RBFirstOrderConstraint}
1172 > \end{equation}
1173   In general, there are two ways to satisfy the holonomic constraints.
1174   We can use a constraint force provided by a Lagrange multiplier on
1175 < the normal manifold to keep the motion on constraint space. Or we
1176 < can simply evolve the system on the constraint manifold. These two
1177 < methods have been proved to be equivalent. The holonomic constraint
1178 < and equations of motions define a constraint manifold for rigid
1179 < bodies
1175 > the normal manifold to keep the motion on the constraint space. Or
1176 > we can simply evolve the system on the constraint manifold. These
1177 > two methods have been proved to be equivalent. The holonomic
1178 > constraint and equations of motions define a constraint manifold for
1179 > rigid bodies
1180   \[
1181   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1182   \right\}.
1183   \]
1184 <
1185 < Unfortunately, this constraint manifold is not the cotangent bundle
1186 < $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1187 < rotation group $SO(3)$. However, it turns out that under symplectic
1246 < transformation, the cotangent space and the phase space are
1247 < diffeomorphic. By introducing
1184 > Unfortunately, this constraint manifold is not $T^* SO(3)$ which is
1185 > a symplectic manifold on Lie rotation group $SO(3)$. However, it
1186 > turns out that under symplectic transformation, the cotangent space
1187 > and the phase space are diffeomorphic. By introducing
1188   \[
1189   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1190   \]
1191 < the mechanical system subject to a holonomic constraint manifold $M$
1191 > the mechanical system subjected to a holonomic constraint manifold $M$
1192   can be re-formulated as a Hamiltonian system on the cotangent space
1193   \[
1194   T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1195   1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1196   \]
1257
1197   For a body fixed vector $X_i$ with respect to the center of mass of
1198   the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1199   given as
# Line 1273 | Line 1212 | respectively.
1212   \[
1213   \nabla _Q V(q,Q) = F(q,Q)X_i^t
1214   \]
1215 < respectively.
1216 <
1217 < As a common choice to describe the rotation dynamics of the rigid
1279 < body, the angular momentum on the body fixed frame $\Pi  = Q^t P$ is
1280 < introduced to rewrite the equations of motion,
1215 > respectively. As a common choice to describe the rotation dynamics
1216 > of the rigid body, the angular momentum on the body fixed frame $\Pi
1217 > = Q^t P$ is introduced to rewrite the equations of motion,
1218   \begin{equation}
1219   \begin{array}{l}
1220 < \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1221 < \dot Q  = Q\Pi {\rm{ }}J^{ - 1}  \\
1220 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda,  \\
1221 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1},  \\
1222   \end{array}
1223   \label{introEqaution:RBMotionPI}
1224   \end{equation}
1225 < , as well as holonomic constraints,
1226 < \[
1227 < \begin{array}{l}
1291 < \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1292 < Q^T Q = 1 \\
1293 < \end{array}
1294 < \]
1295 <
1296 < For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1297 < so(3)^ \star$, the hat-map isomorphism,
1225 > as well as holonomic constraints $\Pi J^{ - 1}  + J^{ - 1} \Pi ^t  =
1226 > 0$ and $Q^T Q = 1$. For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a
1227 > matrix $\hat v \in so(3)^ \star$, the hat-map isomorphism,
1228   \begin{equation}
1229   v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1230   {\begin{array}{*{20}c}
# Line 1307 | Line 1237 | operations
1237   will let us associate the matrix products with traditional vector
1238   operations
1239   \[
1240 < \hat vu = v \times u
1240 > \hat vu = v \times u.
1241   \]
1242 < Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1242 > Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew
1243   matrix,
1244 <
1245 < \begin{eqnarray*}
1246 < (\dot \Pi  - \dot \Pi ^T ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{
1247 < }}(J^{ - 1} \Pi  + \Pi J^{ - 1} ) + \sum\limits_i {[Q^T F_i
1248 < (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  - (\Lambda  - \Lambda ^T ).
1249 < \label{introEquation:skewMatrixPI}
1250 < \end{eqnarray*}
1251 <
1252 < Since $\Lambda$ is symmetric, the last term of Equation
1253 < \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1254 < multiplier $\Lambda$ is absent from the equations of motion. This
1255 < unique property eliminates the requirement of iterations which can
1326 < not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1327 <
1328 < Applying the hat-map isomorphism, we obtain the equation of motion
1329 < for angular momentum on body frame
1244 > \begin{eqnarray}
1245 > (\dot \Pi  - \dot \Pi ^T )&= &(\Pi  - \Pi ^T )(J^{ - 1} \Pi  + \Pi J^{ - 1} ) \notag \\
1246 > & & + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  -
1247 > (\Lambda  - \Lambda ^T ). \label{introEquation:skewMatrixPI}
1248 > \end{eqnarray}
1249 > Since $\Lambda$ is symmetric, the last term of
1250 > Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1251 > Lagrange multiplier $\Lambda$ is absent from the equations of
1252 > motion. This unique property eliminates the requirement of
1253 > iterations which can not be avoided in other methods.\cite{Kol1997,
1254 > Omelyan1998} Applying the hat-map isomorphism, we obtain the
1255 > equation of motion for angular momentum in the body frame
1256   \begin{equation}
1257   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1258   F_i (r,Q)} \right) \times X_i }.
# Line 1335 | Line 1261 | given by
1261   In the same manner, the equation of motion for rotation matrix is
1262   given by
1263   \[
1264 < \dot Q = Qskew(I^{ - 1} \pi )
1264 > \dot Q = Qskew(I^{ - 1} \pi ).
1265   \]
1266  
1267   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1268 < Lie-Poisson Integrator for Free Rigid Body}
1268 > Lie-Poisson Integrator for Free Rigid Bodies}
1269  
1270   If there are no external forces exerted on the rigid body, the only
1271   contribution to the rotational motion is from the kinetic energy
# Line 1357 | Line 1283 | J(\pi ) = \left( {\begin{array}{*{20}c}
1283     0 & {\pi _3 } & { - \pi _2 }  \\
1284     { - \pi _3 } & 0 & {\pi _1 }  \\
1285     {\pi _2 } & { - \pi _1 } & 0  \\
1286 < \end{array}} \right)
1286 > \end{array}} \right).
1287   \end{equation}
1288   Thus, the dynamics of free rigid body is governed by
1289   \begin{equation}
1290 < \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1290 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi ).
1291   \end{equation}
1292 <
1293 < One may notice that each $T_i^r$ in Equation
1294 < \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1369 < instance, the equations of motion due to $T_1^r$ are given by
1292 > One may notice that each $T_i^r$ in
1293 > Eq.~\ref{introEquation:rotationalKineticRB} can be solved exactly.
1294 > For instance, the equations of motion due to $T_1^r$ are given by
1295   \begin{equation}
1296   \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1297   \label{introEqaution:RBMotionSingleTerm}
1298   \end{equation}
1299 < where
1299 > with
1300   \[ R_1  = \left( {\begin{array}{*{20}c}
1301     0 & 0 & 0  \\
1302     0 & 0 & {\pi _1 }  \\
1303     0 & { - \pi _1 } & 0  \\
1304   \end{array}} \right).
1305   \]
1306 < The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1306 > The solutions of Eq.~\ref{introEqaution:RBMotionSingleTerm} is
1307   \[
1308   \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1309   Q(0)e^{\Delta tR_1 }
# Line 1392 | Line 1317 | tR_1 }$, we can use Cayley transformation to obtain a
1317   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1318   \]
1319   To reduce the cost of computing expensive functions in $e^{\Delta
1320 < tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1321 < propagator,
1322 < \[
1323 < e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1324 < )
1325 < \]
1326 < The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1320 > tR_1 }$, we can use the Cayley transformation to obtain a
1321 > single-aixs propagator,
1322 > \begin{eqnarray*}
1323 > e^{\Delta tR_1 }  & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta
1324 > tR_1 ) \\
1325 > %
1326 > & \approx & \left( \begin{array}{ccc}
1327 > 1 & 0 & 0 \\
1328 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1329 > \theta^2 / 4} \\
1330 > 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1331 > \theta^2 / 4}
1332 > \end{array}
1333 > \right).
1334 > \end{eqnarray*}
1335 > The propagators for $T_2^r$ and $T_3^r$ can be found in the same
1336   manner. In order to construct a second-order symplectic method, we
1337 < split the angular kinetic Hamiltonian function can into five terms
1337 > split the angular kinetic Hamiltonian function into five terms
1338   \[
1339   T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1340   ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
# Line 1414 | Line 1348 | _1 }.
1348   \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1349   _1 }.
1350   \]
1351 <
1418 < The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1419 < $F(\pi )$ and $G(\pi )$ is defined by
1351 > The non-canonical Lie-Poisson bracket $\{F, G\}$ of two functions $F(\pi )$ and $G(\pi )$ is defined by
1352   \[
1353   \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1354 < )
1354 > ).
1355   \]
1356   If the Poisson bracket of a function $F$ with an arbitrary smooth
1357   function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1358   conserved quantity in Poisson system. We can easily verify that the
1359   norm of the angular momentum, $\parallel \pi
1360 < \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1360 > \parallel$, is a \emph{Casimir}.\cite{McLachlan1993} Let $F(\pi ) = S(\frac{{\parallel
1361   \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1362   then by the chain rule
1363   \[
1364   \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1365 < }}{2})\pi
1365 > }}{2})\pi.
1366   \]
1367 < Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1367 > Thus, $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel
1368 > \pi
1369   \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1370   Lie-Poisson integrator is found to be both extremely efficient and
1371   stable. These properties can be explained by the fact the small
# Line 1443 | Line 1376 | energy and potential energy,
1376   Splitting for Rigid Body}
1377  
1378   The Hamiltonian of rigid body can be separated in terms of kinetic
1379 < energy and potential energy,
1380 < \[
1381 < H = T(p,\pi ) + V(q,Q)
1449 < \]
1450 < The equations of motion corresponding to potential energy and
1451 < kinetic energy are listed in the below table,
1379 > energy and potential energy, $H = T(p,\pi ) + V(q,Q)$. The equations
1380 > of motion corresponding to potential energy and kinetic energy are
1381 > listed in Table~\ref{introTable:rbEquations}.
1382   \begin{table}
1383   \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1384 + \label{introTable:rbEquations}
1385   \begin{center}
1386   \begin{tabular}{|l|l|}
1387    \hline
# Line 1486 | Line 1417 | defined by \ref{introEquation:rotationalKineticRB}. Th
1417   T(p,\pi ) =T^t (p) + T^r (\pi ).
1418   \end{equation}
1419   where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1420 < defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1421 < corresponding propagators are given by
1420 > defined by Eq.~\ref{introEquation:rotationalKineticRB}. Therefore,
1421 > the corresponding propagators are given by
1422   \[
1423   \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1424   _{\Delta t,T^r }.
1425   \]
1426   Finally, we obtain the overall symplectic propagators for freely
1427   moving rigid bodies
1428 < \begin{equation}
1429 < \begin{array}{c}
1430 < \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1431 <  \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 }  \\
1501 <  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1502 < \end{array}
1428 > \begin{eqnarray}
1429 > \varphi _{\Delta t}  &=& \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \notag\\
1430 >  & & \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 }  \notag\\
1431 >  & & \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .
1432   \label{introEquation:overallRBFlowMaps}
1433 < \end{equation}
1433 > \end{eqnarray}
1434  
1435   \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1436   As an alternative to newtonian dynamics, Langevin dynamics, which
1437   mimics a simple heat bath with stochastic and dissipative forces,
1438   has been applied in a variety of studies. This section will review
1439 < the theory of Langevin dynamics. A brief derivation of generalized
1439 > the theory of Langevin dynamics. A brief derivation of the generalized
1440   Langevin equation will be given first. Following that, we will
1441 < discuss the physical meaning of the terms appearing in the equation
1513 < as well as the calculation of friction tensor from hydrodynamics
1514 < theory.
1441 > discuss the physical meaning of the terms appearing in the equation.
1442  
1443   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1444  
# Line 1520 | Line 1447 | Dynamics (GLE). Lets consider a system, in which the d
1447   environment, has been widely used in quantum chemistry and
1448   statistical mechanics. One of the successful applications of
1449   Harmonic bath model is the derivation of the Generalized Langevin
1450 < Dynamics (GLE). Lets consider a system, in which the degree of
1450 > Dynamics (GLE). Consider a system, in which the degree of
1451   freedom $x$ is assumed to couple to the bath linearly, giving a
1452   Hamiltonian of the form
1453   \begin{equation}
# Line 1531 | Line 1458 | H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_
1458   with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1459   \[
1460   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1461 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1461 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  x_\alpha ^2 }
1462   \right\}}
1463   \]
1464   where the index $\alpha$ runs over all the bath degrees of freedom,
# Line 1547 | Line 1474 | W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\a
1474   \[
1475   W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1476   }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1477 < \] and combining the last two terms in Equation
1478 < \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1552 < Hamiltonian as
1477 > \]
1478 > and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as
1479   \[
1480   H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1481   {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1482   w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1483 < w_\alpha ^2 }}x} \right)^2 } \right\}}
1483 > w_\alpha ^2 }}x} \right)^2 } \right\}}.
1484   \]
1485   Since the first two terms of the new Hamiltonian depend only on the
1486   system coordinates, we can get the equations of motion for
# Line 1571 | Line 1497 | m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x
1497   \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1498   \label{introEquation:bathMotionGLE}
1499   \end{equation}
1574
1500   In order to derive an equation for $x$, the dynamics of the bath
1501   variables $x_\alpha$ must be solved exactly first. As an integral
1502   transform which is particularly useful in solving linear ordinary
1503   differential equations,the Laplace transform is the appropriate tool
1504   to solve this problem. The basic idea is to transform the difficult
1505   differential equations into simple algebra problems which can be
1506 < solved easily. Then, by applying the inverse Laplace transform, also
1507 < known as the Bromwich integral, we can retrieve the solutions of the
1508 < original problems.
1509 <
1585 < Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1586 < transform of f(t) is a new function defined as
1506 > solved easily. Then, by applying the inverse Laplace transform, we
1507 > can retrieve the solutions of the original problems. Let $f(t)$ be a
1508 > function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$
1509 > is a new function defined as
1510   \[
1511   L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1512   \]
1513   where  $p$ is real and  $L$ is called the Laplace Transform
1514 < Operator. Below are some important properties of Laplace transform
1592 <
1514 > Operator. Below are some important properties of the Laplace transform
1515   \begin{eqnarray*}
1516   L(x + y)  & = & L(x) + L(y) \\
1517   L(ax)     & = & aL(x) \\
# Line 1597 | Line 1519 | Operator. Below are some important properties of Lapla
1519   L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1520   L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1521   \end{eqnarray*}
1600
1601
1522   Applying the Laplace transform to the bath coordinates, we obtain
1523   \begin{eqnarray*}
1524 < 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) \\
1525 < L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1524 > 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), \\
1525 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}. \\
1526   \end{eqnarray*}
1527 <
1608 < By the same way, the system coordinates become
1527 > In the same way, the system coordinates become
1528   \begin{eqnarray*}
1529 < mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1530 <  & & \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\}}  \\
1529 > mL(\ddot x) & = &
1530 >  - \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\}}  \\
1531 >  & & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}.
1532   \end{eqnarray*}
1613
1533   With the help of some relatively important inverse Laplace
1534   transformations:
1535   \[
# Line 1620 | Line 1539 | transformations:
1539   L(1) = \frac{1}{p} \\
1540   \end{array}
1541   \]
1542 < , we obtain
1542 > we obtain
1543   \begin{eqnarray*}
1544   m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1545   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
# Line 1629 | Line 1548 | x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _
1548   & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1549   x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1550   \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1551 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1552 < \end{eqnarray*}
1553 < \begin{eqnarray*}
1554 < m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1555 < {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1556 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1551 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}\\
1552 > %
1553 > & = & -
1554 > \frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha
1555 > = 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha
1556 > ^2 }}} \right)\cos (\omega _\alpha
1557   t)\dot x(t - \tau )d} \tau }  \\
1558   & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1559   x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
# Line 1661 | Line 1580 | which is known as the \emph{generalized Langevin equat
1580   (t)\dot x(t - \tau )d\tau }  + R(t)
1581   \label{introEuqation:GeneralizedLangevinDynamics}
1582   \end{equation}
1583 < which is known as the \emph{generalized Langevin equation}.
1583 > which is known as the \emph{generalized Langevin equation} (GLE).
1584  
1585   \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1586  
1587   One may notice that $R(t)$ depends only on initial conditions, which
1588   implies it is completely deterministic within the context of a
1589   harmonic bath. However, it is easy to verify that $R(t)$ is totally
1590 < uncorrelated to $x$ and $\dot x$,
1591 < \[
1592 < \begin{array}{l}
1593 < \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1675 < \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1676 < \end{array}
1677 < \]
1678 < This property is what we expect from a truly random process. As long
1679 < as the model chosen for $R(t)$ was a gaussian distribution in
1590 > uncorrelated to $x$ and $\dot x$, $\left\langle {x(t)R(t)}
1591 > \right\rangle  = 0, \left\langle {\dot x(t)R(t)} \right\rangle  =
1592 > 0.$ This property is what we expect from a truly random process. As
1593 > long as the model chosen for $R(t)$ was a gaussian distribution in
1594   general, the stochastic nature of the GLE still remains.
1681
1595   %dynamic friction kernel
1596   The convolution integral
1597   \[
# Line 1693 | Line 1606 | and Equation \ref{introEuqation:GeneralizedLangevinDyn
1606   \[
1607   \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1608   \]
1609 < and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1609 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1610   \[
1611   m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1612   \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
# Line 1703 | Line 1616 | taken as a $delta$ function in time:
1616   infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1617   taken as a $delta$ function in time:
1618   \[
1619 < \xi (t) = 2\xi _0 \delta (t)
1619 > \xi (t) = 2\xi _0 \delta (t).
1620   \]
1621   Hence, the convolution integral becomes
1622   \[
1623   \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1624   {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1625   \]
1626 < and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1626 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1627   \begin{equation}
1628   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1629   x(t) + R(t) \label{introEquation:LangevinEquation}
# Line 1718 | Line 1631 | briefly review on calculating friction tensor for arbi
1631   which is known as the Langevin equation. The static friction
1632   coefficient $\xi _0$ can either be calculated from spectral density
1633   or be determined by Stokes' law for regular shaped particles. A
1634 < briefly review on calculating friction tensor for arbitrary shaped
1634 > brief review on calculating friction tensors for arbitrary shaped
1635   particles is given in Sec.~\ref{introSection:frictionTensor}.
1636  
1637   \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1638  
1639 < Defining a new set of coordinates,
1639 > Defining a new set of coordinates
1640   \[
1641   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1642 < ^2 }}x(0)
1643 < \],
1644 < we can rewrite $R(T)$ as
1642 > ^2 }}x(0),
1643 > \]
1644 > we can rewrite $R(t)$ as
1645   \[
1646   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1647   \]
1648   And since the $q$ coordinates are harmonic oscillators,
1736
1649   \begin{eqnarray*}
1650   \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1651   \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1652   \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1653   \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 } }  \\
1654    & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1655 <  & = &kT\xi (t) \\
1655 >  & = &kT\xi (t)
1656   \end{eqnarray*}
1745
1657   Thus, we recover the \emph{second fluctuation dissipation theorem}
1658   \begin{equation}
1659   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1660 < \label{introEquation:secondFluctuationDissipation}.
1660 > \label{introEquation:secondFluctuationDissipation},
1661   \end{equation}
1662 < In effect, it acts as a constraint on the possible ways in which one
1663 < can model the random force and friction kernel.
1662 > which acts as a constraint on the possible ways in which one can
1663 > model the random force and friction kernel.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines