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 2706 by tim, Thu Apr 13 04:47:47 2006 UTC vs.
Revision 2950 by tim, Tue Jul 18 16:42:49 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
20   motion of particles is the foundation of the classical mechanics.
21 < Newton¡¯s first law defines a class of inertial frames. Inertial
21 > Newton's first law defines a class of inertial frames. Inertial
22   frames are reference frames where a particle not interacting with
23   other bodies will move with constant speed in the same direction.
24 < With respect to inertial frames Newton¡¯s second law has the form
24 > With respect to inertial frames, Newton's second law has the form
25   \begin{equation}
26 < F = \frac {dp}{dt} = \frac {mv}{dt}
26 > F = \frac {dp}{dt} = \frac {mdv}{dt}
27   \label{introEquation:newtonSecondLaw}
28   \end{equation}
29   A point mass interacting with other bodies moves with the
30   acceleration along the direction of the force acting on it. Let
31   $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
32   $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
33 < Newton¡¯s third law states that
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 46 | Line 46 | N \equiv r \times F \label{introEquation:torqueDefinit
46   \end{equation}
47   The torque $\tau$ with respect to the same origin is defined to be
48   \begin{equation}
49 < N \equiv r \times F \label{introEquation:torqueDefinition}
49 > \tau \equiv r \times F \label{introEquation:torqueDefinition}
50   \end{equation}
51   Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52   \[
# Line 59 | Line 59 | thus,
59   \]
60   thus,
61   \begin{equation}
62 < \dot L = r \times \dot p = N
62 > \dot L = r \times \dot p = \tau
63   \end{equation}
64   If there are no external torques acting on a body, the angular
65   momentum of it is conserved. The last conservation theorem state
66 < that if all forces are conservative, Energy
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 < scheme for rigid body \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: it
76 < describes their motion in special cartesian coordinate systems.
77 < Another limitation of Newtonian mechanics becomes obvious when we
78 < try to describe systems with large numbers of particles. It becomes
79 < very difficult to predict the properties of the system by carrying
80 < out calculations involving the each individual interaction between
81 < all the particles, even if we know all of the details of the
82 < interaction. In order to overcome some of the practical difficulties
83 < which arise in attempts to apply Newton's equation to complex
84 < system, alternative procedures may be developed.
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}Hamilton's
84 < Principle}
83 > \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
84 > Principle}}
85  
86   Hamilton introduced the dynamical principle upon which it is
87 < possible to base all of mechanics and, indeed, most of classical
88 < physics. Hamilton's Principle may be stated as follow,
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
96 < the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
87 > possible to base all of mechanics and most of classical physics.
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}
101
97   For simple mechanical systems, where the forces acting on the
98 < different part are derivable from a potential and the velocities are
99 < small compared with that of light, the Lagrangian function $L$ can
100 < be define as the difference between the kinetic energy of the system
106 < and its potential energy,
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  
111 < \subsubsection{\label{introSection:equationOfMotionLagrangian}The
112 < Equations of Motion in Lagrangian Mechanics}
111 > \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
112 > Equations of Motion in Lagrangian Mechanics}}
113  
114 < For a holonomic system of $f$ degrees of freedom, the equations of
115 < motion in the Lagrangian form is
114 > For a system of $f$ degrees of freedom, the equations of motion in
115 > the Lagrangian form is
116   \begin{equation}
117   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
118   \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
# Line 132 | Line 126 | independent of generalized velocities, the generalized
126   Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
127   introduced by William Rowan Hamilton in 1833 as a re-formulation of
128   classical mechanics. If the potential energy of a system is
129 < independent of generalized velocities, the generalized momenta can
136 < be defined as
129 > independent of velocities, the momenta can be defined as
130   \begin{equation}
131   p_i = \frac{\partial L}{\partial \dot q_i}
132   \label{introEquation:generalizedMomenta}
# Line 143 | 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}
146
139   With the help of the generalized momenta, we may now define a new
140   quantity $H$ by the equation
141   \begin{equation}
# Line 151 | 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 <
156 < Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
157 < 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
163   find
164   \begin{equation}
165 < \frac{{\partial H}}{{\partial p_k }} = q_k
165 > \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
166   \label{introEquation:motionHamiltonianCoordinate}
167   \end{equation}
168   \begin{equation}
169 < \frac{{\partial H}}{{\partial q_k }} =  - p_k
169 > \frac{{\partial H}}{{\partial q_k }} =  - \dot {p_k}
170   \label{introEquation:motionHamiltonianMomentum}
171   \end{equation}
172   and
# Line 185 | Line 175 | t}}
175   t}}
176   \label{introEquation:motionHamiltonianTime}
177   \end{equation}
178 <
189 < 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{Goldstein01}.
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
185 < function of the generalized velocities $\dot q_i$ and the
186 < generalized coordinates $q_i$, while the Hamiltonian is considered
187 < to be a function of the generalized momenta $p_i$ and the conjugate
188 < generalized coordinate $q_i$. Hamiltonian Mechanics is more
189 < appropriate for application to statistical mechanics and quantum
190 < mechanics, since it treats the coordinate and its time derivative as
191 < independent variables and it only works with 1st-order differential
203 < equations\cite{Marion90}.
204 <
185 > function of the generalized velocities $\dot q_i$ and coordinates
186 > $q_i$, while the Hamiltonian is considered to be a function of the
187 > generalized momenta $p_i$ and the conjugate coordinates $q_i$.
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}
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 221 | 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 in a cartesian space,
221 < where each of the $6f$ coordinates and momenta is assigned to one of
222 < $6f$ mutually orthogonal axes, the phase space of this system is a
223 < $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
224 < \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
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 < A microscopic state or microstate of a classical system is
241 < specification of the complete phase space vector of a system at any
242 < instant in time. An ensemble is defined as a collection of systems
243 < sharing one or more macroscopic characteristics but each being in a
244 < unique microstate. The complete ensemble is specified by giving all
245 < systems or microstates consistent with the common macroscopic
246 < characteristics of the ensemble. Although the state of each
247 < individual system in the ensemble could be precisely described at
248 < any instance in time by a suitable phase space vector, when using
249 < ensembles for statistical purposes, there is no need to maintain
250 < distinctions between individual systems, since the numbers of
251 < systems at any time in the different states which correspond to
252 < different regions of the phase space are more interesting. Moreover,
253 < in the point of view of statistical mechanics, one would prefer to
254 < use ensembles containing a large enough population of separate
255 < members so that the numbers of systems in such different states can
256 < be regarded as changing continuously as we traverse different
257 < regions of the phase space. The condition of an ensemble at any time
236 > In statistical mechanics, the condition of an ensemble at any time
237   can be regarded as appropriately specified by the density $\rho$
238   with which representative points are distributed over the phase
239 < space. The density of distribution for an ensemble with $f$ degrees
240 < of freedom is defined as,
239 > space. The density distribution for an ensemble with $f$ degrees of
240 > freedom is defined as,
241   \begin{equation}
242   \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
243   \label{introEquation:densityDistribution}
244   \end{equation}
245   Governed by the principles of mechanics, the phase points change
246 < their value which would change the density at any time at phase
247 < space. Hence, the density of distribution is also to be taken as a
248 < function of the time.
249 <
271 < The number of systems $\delta N$ at time $t$ can be determined by,
246 > their locations which changes the density at any time at phase
247 > space. Hence, the density distribution is also to be taken as a
248 > function of the time. The number of systems $\delta N$ at time $t$
249 > can be determined by,
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 are exploited, we can
255 < sufficiently approximate $\delta N$ without introducing
256 < discontinuity when we go from one region in the phase space to
257 < another. By integrating over the whole phase space,
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,
258   \begin{equation}
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 }}.
267   \label{introEquation:unitProbability}
268   \end{equation}
269 < With the help of Equation(\ref{introEquation:unitProbability}) and
270 < the knowledge of the system, it is possible to calculate the average
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 may still
275 < remain well defined. For a classical system in thermal equilibrium
276 < with its environment, the ensemble average of a mechanical quantity,
277 < $\langle A(q , p) \rangle_t$, takes the form of an integral over the
278 < phase space of the system,
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  
308 There are several different types of ensembles with different
309 statistical characteristics. As a function of macroscopic
310 parameters, such as temperature \textit{etc}, partition function can
311 be used to describe the statistical properties of a system in
312 thermodynamic equilibrium.
313
314 As an ensemble of systems, each of which is known to be thermally
315 isolated and conserve energy, Microcanonical ensemble(NVE) has a
316 partition function like,
317 \begin{equation}
318 \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
319 \end{equation}
320 A canonical ensemble(NVT)is an ensemble of systems, each of which
321 can share its energy with a large heat reservoir. The distribution
322 of the total energy amongst the possible dynamical states is given
323 by the partition function,
324 \begin{equation}
325 \Omega (N,V,T) = e^{ - \beta A}
326 \label{introEquation:NVTPartition}
327 \end{equation}
328 Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
329 TS$. Since most experiment are carried out under constant pressure
330 condition, isothermal-isobaric ensemble(NPT) play a very important
331 role in molecular simulation. The isothermal-isobaric ensemble allow
332 the system to exchange energy with a heat bath of temperature $T$
333 and to change the volume as well. Its partition function is given as
334 \begin{equation}
335 \Delta (N,P,T) =  - e^{\beta G}.
336 \label{introEquation:NPTPartition}
337 \end{equation}
338 Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
339
286   \subsection{\label{introSection:liouville}Liouville's theorem}
287  
288 < The Liouville's theorem is the foundation on which statistical
289 < mechanics rests. It describes the time evolution of phase space
288 > Liouville's theorem is the foundation on which statistical mechanics
289 > rests. It describes the time evolution of the phase space
290   distribution function. In order to calculate the rate of change of
291 < $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
292 < consider the two faces perpendicular to the $q_1$ axis, which are
293 < located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
294 < leaving the opposite face is given by the expression,
291 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
292 > the two faces perpendicular to the $q_1$ axis, which are located at
293 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
294 > opposite face is given by the expression,
295   \begin{equation}
296   \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
297   \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
# Line 369 | Line 315 | divining $ \delta q_1  \ldots \delta q_f \delta p_1  \
315   + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
316   \end{equation}
317   which cancels the first terms of the right hand side. Furthermore,
318 < divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
318 > dividing $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
319   p_f $ in both sides, we can write out Liouville's theorem in a
320   simple form,
321   \begin{equation}
# Line 378 | Line 324 | simple form,
324   \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
325   \label{introEquation:liouvilleTheorem}
326   \end{equation}
381
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 particles in the system
330 < is huge, we may be able to believe the system is stationary,
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 391 | Line 337 | distribution,
337   connected to the Hamiltonian $H$ through Maxwell-Boltzmann
338   distribution,
339   \begin{equation}
340 < \rho  \propto e^{ - \beta H}
340 > \rho  \propto e^{ - \beta H}.
341   \label{introEquation:densityAndHamiltonian}
342   \end{equation}
343  
344 < \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
344 > \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
345   Lets consider a region in the phase space,
346   \begin{equation}
347   \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
348   \end{equation}
349   If this region is small enough, the density $\rho$ can be regarded
350 < as uniform over the whole phase space. Thus, the number of phase
351 < points inside this region is given by,
352 < \begin{equation}
353 < \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
354 < dp_1 } ..dp_f.
409 < \end{equation}
410 <
411 < \begin{equation}
412 < \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
350 > as uniform over the whole integral. Thus, the number of phase points
351 > inside this region is given by,
352 > \begin{eqnarray}
353 > \delta N &=& \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f,\\
354 > \frac{{d(\delta N)}}{{dt}} &=& \frac{{d\rho }}{{dt}}\delta v + \rho
355   \frac{d}{{dt}}(\delta v) = 0.
356 < \end{equation}
357 < With the help of stationary assumption
358 < (\ref{introEquation:stationary}), we obtain the principle of the
359 < \emph{conservation of extension in phase space},
356 > \end{eqnarray}
357 > With the help of the stationary assumption
358 > (Eq.~\ref{introEquation:stationary}), we obtain the principle of
359 > \emph{conservation of volume in phase space},
360   \begin{equation}
361   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
362   ...dq_f dp_1 } ..dp_f  = 0.
363   \label{introEquation:volumePreserving}
364   \end{equation}
365  
366 < \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
366 > \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
367  
368 < Liouville's theorem can be expresses in a variety of different forms
368 > Liouville's theorem can be expressed in a variety of different forms
369   which are convenient within different contexts. For any two function
370   $F$ and $G$ of the coordinates and momenta of a system, the Poisson
371 < bracket ${F, G}$ is defined as
371 > bracket $\{F,G\}$ is defined as
372   \begin{equation}
373   \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
374   F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
# Line 434 | Line 376 | Substituting equations of motion in Hamiltonian formal
376   q_i }}} \right)}.
377   \label{introEquation:poissonBracket}
378   \end{equation}
379 < Substituting equations of motion in Hamiltonian formalism(
380 < \ref{introEquation:motionHamiltonianCoordinate} ,
381 < \ref{introEquation:motionHamiltonianMomentum} ) into
382 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
383 < theorem using Poisson bracket notion,
379 > Substituting equations of motion in Hamiltonian formalism
380 > (Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
381 > Eq.~\ref{introEquation:motionHamiltonianMomentum}) into
382 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
383 > Liouville's theorem using Poisson bracket notion,
384   \begin{equation}
385   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
386   {\rho ,H} \right\}.
# Line 457 | Line 399 | expressed as
399   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
400   \label{introEquation:liouvilleTheoremInOperator}
401   \end{equation}
402 <
402 > which can help define a propagator $\rho (t) = e^{-iLt} \rho (0)$.
403   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
404  
405   Various thermodynamic properties can be calculated from Molecular
406   Dynamics simulation. By comparing experimental values with the
407   calculated properties, one can determine the accuracy of the
408 < simulation and the quality of the underlying model. However, both of
409 < experiment and computer simulation are usually performed during a
408 > simulation and the quality of the underlying model. However, both
409 > experiments and computer simulations are usually performed during a
410   certain time interval and the measurements are averaged over a
411 < period of them which is different from the average behavior of
412 < many-body system in Statistical Mechanics. Fortunately, Ergodic
413 < Hypothesis is proposed to make a connection between time average and
414 < ensemble average. It states that time average and average over the
415 < statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
411 > period of time which is different from the average behavior of
412 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
413 > Hypothesis makes a connection between time average and the ensemble
414 > average. It states that the time average and average over the
415 > statistical ensemble are identical:\cite{Frenkel1996, Leach2001}
416   \begin{equation}
417   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
418   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 479 | Line 421 | sufficiently long time (longer than relaxation time),
421   where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
422   physical quantity and $\rho (p(t), q(t))$ is the equilibrium
423   distribution function. If an observation is averaged over a
424 < sufficiently long time (longer than relaxation time), all accessible
425 < microstates in phase space are assumed to be equally probed, giving
426 < a properly weighted statistical average. This allows the researcher
427 < freedom of choice when deciding how best to measure a given
428 < observable. In case an ensemble averaged approach sounds most
429 < reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
424 > sufficiently long time (longer than the relaxation time), all
425 > accessible microstates in phase space are assumed to be equally
426 > probed, giving a properly weighted statistical average. This allows
427 > the researcher freedom of choice when deciding how best to measure a
428 > given observable. In case an ensemble averaged approach sounds most
429 > reasonable, the Monte Carlo methods\cite{Metropolis1949} can be
430   utilized. Or if the system lends itself to a time averaging
431   approach, the Molecular Dynamics techniques in
432   Sec.~\ref{introSection:molecularDynamics} will be the best
433 < choice\cite{Frenkel1996}.
433 > choice.\cite{Frenkel1996}
434  
435   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
436 < A variety of numerical integrators were proposed to simulate the
437 < motions. They usually begin with an initial conditionals and move
438 < the objects in the direction governed by the differential equations.
439 < However, most of them ignore the hidden physical law contained
440 < within the equations. Since 1990, geometric integrators, which
441 < preserve various phase-flow invariants such as symplectic structure,
442 < volume and time reversal symmetry, are developed to address this
443 < issue. The velocity verlet method, which happens to be a simple
444 < example of symplectic integrator, continues to gain its popularity
445 < in molecular dynamics community. This fact can be partly explained
446 < by its geometric nature.
436 > A variety of numerical integrators have been proposed to simulate
437 > the motions of atoms in MD simulation. They usually begin with
438 > initial conditions and move the objects in the direction governed by
439 > the differential equations. However, most of them ignore the hidden
440 > physical laws contained within the equations. Since 1990, geometric
441 > integrators, which preserve various phase-flow invariants such as
442 > symplectic structure, volume and time reversal symmetry, were
443 > developed to address this issue.\cite{Dullweber1997, McLachlan1998,
444 > Leimkuhler1999} The velocity Verlet method, which happens to be a
445 > simple example of symplectic integrator, continues to gain
446 > popularity in the molecular dynamics community. This fact can be
447 > partly explained by its geometric nature.
448  
449 < \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
450 < A \emph{manifold} is an abstract mathematical space. It locally
451 < looks like Euclidean space, but when viewed globally, it may have
452 < more complicate structure. A good example of manifold is the surface
453 < of Earth. It seems to be flat locally, but it is round if viewed as
454 < a whole. A \emph{differentiable manifold} (also known as
455 < \emph{smooth manifold}) is a manifold with an open cover in which
456 < the covering neighborhoods are all smoothly isomorphic to one
457 < another. In other words,it is possible to apply calculus on
458 < \emph{differentiable manifold}. A \emph{symplectic manifold} is
516 < defined as a pair $(M, \omega)$ which consisting of a
517 < \emph{differentiable manifold} $M$ and a close, non-degenerated,
449 > \subsection{\label{introSection:symplecticManifold}Manifolds and Bundles}
450 > A \emph{manifold} is an abstract mathematical space. It looks
451 > locally like Euclidean space, but when viewed globally, it may have
452 > more complicated structure. A good example of manifold is the
453 > surface of Earth. It seems to be flat locally, but it is round if
454 > viewed as a whole. A \emph{differentiable manifold} (also known as
455 > \emph{smooth manifold}) is a manifold on which it is possible to
456 > apply calculus.\cite{Hirsch1997} A \emph{symplectic manifold} is
457 > defined as a pair $(M, \omega)$ which consists of a
458 > \emph{differentiable manifold} $M$ and a close, non-degenerate,
459   bilinear symplectic form, $\omega$. A symplectic form on a vector
460   space $V$ is a function $\omega(x, y)$ which satisfies
461   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
462   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
463 < $\omega(x, x) = 0$. Cross product operation in vector field is an
464 < example of symplectic form.
463 > $\omega(x, x) = 0$.\cite{McDuff1998} The cross product operation in
464 > vector field is an example of symplectic form.
465 > Given vector spaces $V$ and $W$ over same field $F$, $f: V \to W$ is a linear transformation if
466 > \begin{eqnarray*}
467 > f(x+y) & = & f(x) + f(y) \\
468 > f(ax) & = & af(x)
469 > \end{eqnarray*}
470 > are always satisfied for any two vectors $x$ and $y$ in $V$ and any scalar $a$ in $F$. One can define the dual vector space $V^*$ of $V$ if any two built-in linear transformations $\phi$ and $\psi$ in $V^*$ satisfy the following definition of addition and scalar multiplication:
471 > \begin{eqnarray*}
472 > (\phi+\psi)(x) & = & \phi(x)+\psi(x) \\
473 > (a\phi)(x) & = & a \phi(x)
474 > \end{eqnarray*}
475 > for all $a$ in $F$ and $x$ in $V$. For a manifold $M$, one can define a tangent vector of a tangent space $TM_q$ at every point $q$
476 > \begin{equation}
477 > \dot q = \mathop {\lim }\limits_{t \to 0} \frac{{\phi (t) - \phi (0)}}{t}
478 > \end{equation}
479 > where $\phi(0)=q$ and $\phi(t) \in M$. One may also define a cotangent space $T^*M_q$ as the dual space of the tangent space $TM_q$. The tangent space and the cotangent space are isomorphic to each other, since they are both real vector spaces with same dimension.
480 > The union of tangent spaces at every point of $M$ is called the tangent bundle of $M$ and is denoted by $TM$, while cotangent bundle $T^*M$ is defined as the union of the cotangent spaces to $M$.\cite{Jost2002} For a Hamiltonian system with configuration manifold $V$, the $(q,\dot q)$ phase space is the tangent bundle of the configuration manifold $V$, while the cotangent bundle is represented by $(q,p)$.
481  
525 One of the motivations to study \emph{symplectic manifold} in
526 Hamiltonian Mechanics is that a symplectic manifold can represent
527 all possible configurations of the system and the phase space of the
528 system can be described by it's cotangent bundle. Every symplectic
529 manifold is even dimensional. For instance, in Hamilton equations,
530 coordinate and momentum always appear in pairs.
531
532 Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
533 \[
534 f : M \rightarrow N
535 \]
536 is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
537 the \emph{pullback} of $\eta$ under f is equal to $\omega$.
538 Canonical transformation is an example of symplectomorphism in
539 classical mechanics.
540
482   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
483  
484 < For a ordinary differential system defined as
484 > For an ordinary differential system defined as
485   \begin{equation}
486   \dot x = f(x)
487   \end{equation}
488 < where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
488 > where $x = x(q,p)$, this system is a canonical Hamiltonian, if
489 > $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
490 > function and $J$ is the skew-symmetric matrix
491   \begin{equation}
549 f(r) = J\nabla _x H(r).
550 \end{equation}
551 $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
552 matrix
553 \begin{equation}
492   J = \left( {\begin{array}{*{20}c}
493     0 & I  \\
494     { - I} & 0  \\
# Line 560 | Line 498 | system can be rewritten as,
498   where $I$ is an identity matrix. Using this notation, Hamiltonian
499   system can be rewritten as,
500   \begin{equation}
501 < \frac{d}{{dt}}x = J\nabla _x H(x)
501 > \frac{d}{{dt}}x = J\nabla _x H(x).
502   \label{introEquation:compactHamiltonian}
503   \end{equation}In this case, $f$ is
504 < called a \emph{Hamiltonian vector field}.
505 <
568 < Another generalization of Hamiltonian dynamics is Poisson Dynamics,
504 > called a \emph{Hamiltonian vector field}. Another generalization of
505 > Hamiltonian dynamics is Poisson Dynamics,\cite{Olver1986}
506   \begin{equation}
507   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
508   \end{equation}
509 < The most obvious change being that matrix $J$ now depends on $x$.
510 < The free rigid body is an example of Poisson system (actually a
574 < Lie-Poisson system) with Hamiltonian function of angular kinetic
575 < energy.
576 < \begin{equation}
577 < J(\pi ) = \left( {\begin{array}{*{20}c}
578 <   0 & {\pi _3 } & { - \pi _2 }  \\
579 <   { - \pi _3 } & 0 & {\pi _1 }  \\
580 <   {\pi _2 } & { - \pi _1 } & 0  \\
581 < \end{array}} \right)
582 < \end{equation}
509 > where the most obvious change being that matrix $J$ now depends on
510 > $x$.
511  
512 < \begin{equation}
585 < H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
586 < }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
587 < \end{equation}
512 > \subsection{\label{introSection:exactFlow}Exact Propagator}
513  
514 < \subsection{\label{introSection:exactFlow}Exact Flow}
515 <
591 < Let $x(t)$ be the exact solution of the ODE system,
514 > Let $x(t)$ be the exact solution of the ODE
515 > system,
516   \begin{equation}
517 < \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
518 < \end{equation}
519 < The exact flow(solution) $\varphi_\tau$ is defined by
520 < \[
521 < x(t+\tau) =\varphi_\tau(x(t))
517 > \frac{{dx}}{{dt}} = f(x), \label{introEquation:ODE}
518 > \end{equation} we can
519 > define its exact propagator $\varphi_\tau$:
520 > \[ x(t+\tau)
521 > =\varphi_\tau(x(t))
522   \]
523   where $\tau$ is a fixed time step and $\varphi$ is a map from phase
524 < space to itself. The flow has the continuous group property,
524 > space to itself. The propagator has the continuous group property,
525   \begin{equation}
526   \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
527   + \tau _2 } .
# Line 606 | Line 530 | Therefore, the exact flow is self-adjoint,
530   \begin{equation}
531   \varphi _\tau   \circ \varphi _{ - \tau }  = I
532   \end{equation}
533 < Therefore, the exact flow is self-adjoint,
533 > Therefore, the exact propagator is self-adjoint,
534   \begin{equation}
535   \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
536   \end{equation}
537 < The exact flow can also be written in terms of the of an operator,
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 < \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
616 < }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
617 < \label{introEquation:exponentialOperator}
543 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
544   \end{equation}
545  
620 In most cases, it is not easy to find the exact flow $\varphi_\tau$.
621 Instead, we use a approximate map, $\psi_\tau$, which is usually
622 called integrator. The order of an integrator $\psi_\tau$ is $p$, if
623 the Taylor series of $\psi_\tau$ agree to order $p$,
624 \begin{equation}
625 \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
626 \end{equation}
627
546   \subsection{\label{introSection:geometricProperties}Geometric Properties}
547  
548 < The hidden geometric properties of ODE and its flow play important
549 < roles in numerical studies. Many of them can be found in systems
550 < which occur naturally in applications.
551 <
552 < Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
635 < a \emph{symplectic} flow if it satisfies,
548 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
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 must be preserved by the integrator.
565 <
566 < It is possible to construct a \emph{volume-preserving} flow for a
567 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
568 < \det d\varphi  = 1$. One can show easily that a symplectic flow will
569 < be volume-preserving.
570 <
654 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
655 < 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
665 < 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,
590 < \begin{equation}
591 < \begin{array}{c}
592 < \frac{{dG(x(t))}}{{dt}} = [\nabla _x G(x(t))]^T \dot x(t) \\
679 <  = [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
680 < \end{array}
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) \notag\\
592 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)).
593   \label{introEquation:firstIntegral1}
594 < \end{equation}
595 < Using poisson bracket notion, Equation
596 < \ref{introEquation:firstIntegral1} can be rewritten as
594 > \end{eqnarray}
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 < \[
603 < \left\{ {G,H} \right\} = 0.
604 < \]
605 < As well known, the Hamiltonian (or energy) H of a Hamiltonian system
606 < is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
695 < 0$.
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
606 > propagator.
607  
697
698 When designing any numerical methods, one should always try to
699 preserve the structural properties of the original ODE and its flow.
700
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 leapfrog methods in
613 < molecular dynamics. In general, symplectic integrators can be
612 > constructed. The most famous example is the Verlet-leapfrog method
613 > in molecular dynamics. In general, symplectic integrators can be
614   constructed using one of four different methods.
615   \begin{enumerate}
616   \item Generating functions
# Line 711 | Line 618 | constructed using one of four different methods.
618   \item Runge-Kutta methods
619   \item Splitting methods
620   \end{enumerate}
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 they are geometrically unstable
625 + against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
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 < Generating function tends to lead to methods which are cumbersome
716 < and difficult to use. In dissipative systems, variational methods
717 < can capture the decay of energy accurately. Since their
718 < geometrically unstable nature against non-Hamiltonian perturbations,
719 < ordinary implicit Runge-Kutta methods are not suitable for
720 < Hamiltonian system. Recently, various high-order explicit
721 < Runge--Kutta methods have been developed to overcome this
722 < instability. However, due to computational penalty involved in
723 < implementing the Runge-Kutta methods, they do not attract too much
724 < attention from Molecular Dynamics community. Instead, splitting have
725 < been widely accepted since they exploit natural decompositions of
726 < the system\cite{Tuckerman92}.
636 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
637  
728 \subsubsection{\label{introSection:splittingMethod}Splitting Method}
729
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 <
740 < 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 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 755 | 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 764 | 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 $h^3$. Sprang splitting's
684 < popularity in molecular simulation community attribute to its
685 < symmetric property,
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}
687   \varphi _h^{ - 1} = \varphi _{ - h}.
688   \label{introEquation:timeReversible}
689   \end{equation}
690  
691 < \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
691 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
692   The classical equation for a system consisting of interacting
693   particles can be written in Hamiltonian form,
694   \[
695   H = T + V
696   \]
697   where $T$ is the kinetic energy and $V$ is the potential energy.
698 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
698 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
699   obtains the following:
700   \begin{align}
701   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 800 | 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 814 | Line 722 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
722      \label{introEquation:Lp9b}\\%
723   %
724   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
725 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
725 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
726   \end{align}
727   From the preceding splitting, one can see that the integration of
728   the equations of motion would follow:
# Line 823 | 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, $\mathbf{r}(\Delta t)$, 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}
738 <
739 < Simply switching the order of splitting and composing, a new
740 < integrator, the \emph{position verlet} integrator, can be generated,
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,
741   \begin{align}
742   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
743   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 837 | Line 745 | q(\Delta t)} \right]. %
745   %
746   q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
747   q(\Delta t)} \right]. %
748 < \label{introEquation:positionVerlet1}
748 > \label{introEquation:positionVerlet2}
749   \end{align}
750  
751 < \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
751 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
752  
753 < Baker-Campbell-Hausdorff formula can be used to determine the local
754 < error of splitting method in terms of commutator of the
755 < operators(\ref{introEquation:exponentialOperator}) associated with
756 < the sub-flow. For operators $hX$ and $hY$ which are associate to
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
756 > operators associated
757 > with the sub-propagator. For operators $hX$ and $hY$ which are
758 > associated with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we
759 > have
760   \begin{equation}
761   \exp (hX + hY) = \exp (hZ)
762   \end{equation}
# Line 855 | Line 765 | Here, $[X,Y]$ is the commutators of operator $X$ and $
765   hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
766   {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
767   \end{equation}
768 < Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
768 > Here, $[X,Y]$ is the commutator of operator $X$ and $Y$ given by
769   \[
770   [X,Y] = XY - YX .
771   \]
772 < Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
773 < can obtain
772 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
773 > to the Strang splitting, we can obtain
774   \begin{eqnarray*}
775 < \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
776 < [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
777 < & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
778 < \ldots )
775 > \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 \\
776 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
777 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots
778 >                                   ).
779   \end{eqnarray*}
780 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
781 < error of Spring splitting is proportional to $h^3$. The same
782 < procedure can be applied to general splitting,  of the form
780 > Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local
781 > error of Strang splitting is proportional to $h^3$. The same
782 > procedure can be applied to a general splitting of the form
783   \begin{equation}
784   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
785   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
786   \end{equation}
787 < Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
788 < order method. Yoshida proposed an elegant way to compose higher
789 < order methods based on symmetric splitting. Given a symmetric second
790 < order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
791 < method can be constructed by composing,
787 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
788 > order methods. Yoshida proposed an elegant way to compose higher
789 > order methods based on symmetric splitting.\cite{Yoshida1990} Given
790 > a symmetric second order base method $ \varphi _h^{(2)} $, a
791 > fourth-order symmetric method can be constructed by composing,
792   \[
793   \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
794   h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
# Line 888 | Line 798 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
798   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
799   \begin{equation}
800   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
801 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
801 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
802   \end{equation}
803 < , if the weights are chosen as
803 > if the weights are chosen as
804   \[
805   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
806   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 898 | Line 808 | As a special discipline of molecular modeling, Molecul
808  
809   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
810  
811 < As a special discipline of molecular modeling, Molecular dynamics
812 < has proven to be a powerful tool for studying the functions of
813 < biological systems, providing structural, thermodynamic and
814 < dynamical information.
811 > As one of the principal tools of molecular modeling, Molecular
812 > dynamics has proven to be a powerful tool for studying the functions
813 > of biological systems, providing structural, thermodynamic and
814 > dynamical information. The basic idea of molecular dynamics is that
815 > macroscopic properties are related to microscopic behavior and
816 > microscopic behavior can be calculated from the trajectories in
817 > simulations. For instance, instantaneous temperature of a
818 > Hamiltonian system of $N$ particles can be measured by
819 > \[
820 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
821 > \]
822 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
823 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
824 > the Boltzman constant.
825  
826 < \subsection{\label{introSec:mdInit}Initialization}
826 > A typical molecular dynamics run consists of three essential steps:
827 > \begin{enumerate}
828 >  \item Initialization
829 >    \begin{enumerate}
830 >    \item Preliminary preparation
831 >    \item Minimization
832 >    \item Heating
833 >    \item Equilibration
834 >    \end{enumerate}
835 >  \item Production
836 >  \item Analysis
837 > \end{enumerate}
838 > These three individual steps will be covered in the following
839 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
840 > initialization of a simulation. Sec.~\ref{introSection:production}
841 > discusses issues of production runs.
842 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
843 > analysis of trajectories.
844  
845 < \subsection{\label{introSec:forceEvaluation}Force Evaluation}
845 > \subsection{\label{introSec:initialSystemSettings}Initialization}
846  
847 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
847 > \subsubsection{\textbf{Preliminary preparation}}
848  
849 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
849 > When selecting the starting structure of a molecule for molecular
850 > simulation, one may retrieve its Cartesian coordinates from public
851 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
852 > thousands of crystal structures of molecules are discovered every
853 > year, many more remain unknown due to the difficulties of
854 > purification and crystallization. Even for molecules with known
855 > structures, some important information is missing. For example, a
856 > missing hydrogen atom which acts as donor in hydrogen bonding must
857 > be added. Moreover, in order to include electrostatic interactions,
858 > one may need to specify the partial charges for individual atoms.
859 > Under some circumstances, we may even need to prepare the system in
860 > a special configuration. For instance, when studying transport
861 > phenomenon in membrane systems, we may prepare the lipids in a
862 > bilayer structure instead of placing lipids randomly in solvent,
863 > since we are not interested in the slow self-aggregation process.
864  
865 < Rigid bodies are frequently involved in the modeling of different
915 < areas, from engineering, physics, to chemistry. For example,
916 < missiles and vehicle are usually modeled by rigid bodies.  The
917 < movement of the objects in 3D gaming engine or other physics
918 < simulator is governed by the rigid body dynamics. In molecular
919 < simulation, rigid body is used to simplify the model in
920 < protein-protein docking study{\cite{Gray03}}.
865 > \subsubsection{\textbf{Minimization}}
866  
867 < It is very important to develop stable and efficient methods to
868 < integrate the equations of motion of orientational degrees of
869 < freedom. Euler angles are the nature choice to describe the
870 < rotational degrees of freedom. However, due to its singularity, the
871 < numerical integration of corresponding equations of motion is very
872 < inefficient and inaccurate. Although an alternative integrator using
873 < different sets of Euler angles can overcome this difficulty\cite{},
874 < the computational penalty and the lost of angular momentum
875 < conservation still remain. A singularity free representation
876 < utilizing quaternions was developed by Evans in 1977. Unfortunately,
877 < this approach suffer from the nonseparable Hamiltonian resulted from
878 < quaternion representation, which prevents the symplectic algorithm
879 < to be utilized. Another different approach is to apply holonomic
880 < constraints to the atoms belonging to the rigid body. Each atom
881 < moves independently under the normal forces deriving from potential
882 < energy and constraint forces which are used to guarantee the
883 < rigidness. However, due to their iterative nature, SHAKE and Rattle
884 < algorithm converge very slowly when the number of constraint
885 < increases.
867 > It is quite possible that some of molecules in the system from
868 > preliminary preparation may be overlapping with each other. This
869 > close proximity leads to high initial potential energy which
870 > consequently jeopardizes any molecular dynamics simulations. To
871 > remove these steric overlaps, one typically performs energy
872 > minimization to find a more reasonable conformation. Several energy
873 > minimization methods have been developed to exploit the energy
874 > surface and to locate the local minimum. While converging slowly
875 > near the minimum, the steepest descent method is extremely robust when
876 > systems are strongly anharmonic. Thus, it is often used to refine
877 > structures from crystallographic data. Relying on the Hessian,
878 > advanced methods like Newton-Raphson converge rapidly to a local
879 > minimum, but become unstable if the energy surface is far from
880 > quadratic. Another factor that must be taken into account, when
881 > choosing energy minimization method, is the size of the system.
882 > Steepest descent and conjugate gradient can deal with models of any
883 > size. Because of the limits on computer memory to store the hessian
884 > matrix and the computing power needed to diagonalize these matrices,
885 > most Newton-Raphson methods can not be used with very large systems.
886  
887 < The break through in geometric literature suggests that, in order to
943 < develop a long-term integration scheme, one should preserve the
944 < symplectic structure of the flow. Introducing conjugate momentum to
945 < rotation matrix $A$ and re-formulating Hamiltonian's equation, a
946 < symplectic integrator, RSHAKE, was proposed to evolve the
947 < Hamiltonian system in a constraint manifold by iteratively
948 < satisfying the orthogonality constraint $A_t A = 1$. An alternative
949 < method using quaternion representation was developed by Omelyan.
950 < However, both of these methods are iterative and inefficient. In
951 < this section, we will present a symplectic Lie-Poisson integrator
952 < for rigid body developed by Dullweber and his coworkers\cite{}.
887 > \subsubsection{\textbf{Heating}}
888  
889 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
889 > Typically, heating is performed by assigning random velocities
890 > according to a Maxwell-Boltzman distribution for a desired
891 > temperature. Beginning at a lower temperature and gradually
892 > increasing the temperature by assigning larger random velocities, we
893 > end up setting the temperature of the system to a final temperature
894 > at which the simulation will be conducted. In the heating phase, we
895 > should also keep the system from drifting or rotating as a whole. To
896 > do this, the net linear momentum and angular momentum of the system
897 > is shifted to zero after each resampling from the Maxwell -Boltzman
898 > distribution.
899  
900 < \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
900 > \subsubsection{\textbf{Equilibration}}
901 >
902 > The purpose of equilibration is to allow the system to evolve
903 > spontaneously for a period of time and reach equilibrium. The
904 > procedure is continued until various statistical properties, such as
905 > temperature, pressure, energy, volume and other structural
906 > properties \textit{etc}, become independent of time. Strictly
907 > speaking, minimization and heating are not necessary, provided the
908 > equilibration process is long enough. However, these steps can serve
909 > as a mean to arrive at an equilibrated structure in an effective
910 > way.
911 >
912 > \subsection{\label{introSection:production}Production}
913 >
914 > The production run is the most important step of the simulation, in
915 > which the equilibrated structure is used as a starting point and the
916 > motions of the molecules are collected for later analysis. In order
917 > to capture the macroscopic properties of the system, the molecular
918 > dynamics simulation must be performed by sampling correctly and
919 > efficiently from the relevant thermodynamic ensemble.
920 >
921 > The most expensive part of a molecular dynamics simulation is the
922 > calculation of non-bonded forces, such as van der Waals force and
923 > Coulombic forces \textit{etc}. For a system of $N$ particles, the
924 > complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
925 > which makes large simulations prohibitive in the absence of any
926 > algorithmic tricks. A natural approach to avoid system size issues
927 > is to represent the bulk behavior by a finite number of the
928 > particles. However, this approach will suffer from surface effects
929 > at the edges of the simulation. To offset this, \textit{Periodic
930 > boundary conditions} (see Fig.~\ref{introFig:pbc}) were developed to
931 > simulate bulk properties with a relatively small number of
932 > particles. In this method, the simulation box is replicated
933 > throughout space to form an infinite lattice. During the simulation,
934 > when a particle moves in the primary cell, its image in other cells
935 > move in exactly the same direction with exactly the same
936 > orientation. Thus, as a particle leaves the primary cell, one of its
937 > images will enter through the opposite face.
938 > \begin{figure}
939 > \centering
940 > \includegraphics[width=\linewidth]{pbc.eps}
941 > \caption[An illustration of periodic boundary conditions]{A 2-D
942 > illustration of periodic boundary conditions. As one particle leaves
943 > the left of the simulation box, an image of it enters the right.}
944 > \label{introFig:pbc}
945 > \end{figure}
946  
947 + %cutoff and minimum image convention
948 + Another important technique to improve the efficiency of force
949 + evaluation is to apply spherical cutoffs where particles farther
950 + than a predetermined distance are not included in the
951 + calculation.\cite{Frenkel1996} The use of a cutoff radius will cause
952 + a discontinuity in the potential energy curve. Fortunately, one can
953 + shift a simple radial potential to ensure the potential curve go
954 + smoothly to zero at the cutoff radius. The cutoff strategy works
955 + well for Lennard-Jones interaction because of its short range
956 + nature. However, simply truncating the electrostatic interaction
957 + with the use of cutoffs has been shown to lead to severe artifacts
958 + in simulations. The Ewald summation, in which the slowly decaying
959 + Coulomb potential is transformed into direct and reciprocal sums
960 + with rapid and absolute convergence, has proved to minimize the
961 + periodicity artifacts in liquid simulations. Taking advantage of
962 + fast Fourier transform (FFT) techniques for calculating discrete
963 + Fourier transforms, the particle mesh-based
964 + methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
965 + $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
966 + \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
967 + which treats Coulombic interactions exactly at short range, and
968 + approximate the potential at long range through multipolar
969 + expansion. In spite of their wide acceptance at the molecular
970 + simulation community, these two methods are difficult to implement
971 + correctly and efficiently. Instead, we use a damped and
972 + charge-neutralized Coulomb potential method developed by Wolf and
973 + his coworkers.\cite{Wolf1999} The shifted Coulomb potential for
974 + particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
975   \begin{equation}
976 < H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
977 < V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
978 < \label{introEquation:RBHamiltonian}
976 > V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
977 > r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
978 > R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
979 > r_{ij})}{r_{ij}}\right\}, \label{introEquation:shiftedCoulomb}
980   \end{equation}
981 < Here, $q$ and $Q$  are the position and rotation matrix for the
982 < rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
983 < $J$, a diagonal matrix, is defined by
981 > where $\alpha$ is the convergence parameter. Due to the lack of
982 > inherent periodicity and rapid convergence,this method is extremely
983 > efficient and easy to implement.
984 > \begin{figure}
985 > \centering
986 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
987 > \caption[An illustration of shifted Coulomb potential]{An
988 > illustration of shifted Coulomb potential.}
989 > \label{introFigure:shiftedCoulomb}
990 > \end{figure}
991 >
992 > \subsection{\label{introSection:Analysis} Analysis}
993 >
994 > According to the principles of
995 > Statistical Mechanics in
996 > Sec.~\ref{introSection:statisticalMechanics}, one can compute
997 > thermodynamic properties, analyze fluctuations of structural
998 > parameters, and investigate time-dependent processes of the molecule
999 > from the trajectories.
1000 >
1001 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1002 >
1003 > Thermodynamic properties, which can be expressed in terms of some
1004 > function of the coordinates and momenta of all particles in the
1005 > system, can be directly computed from molecular dynamics. The usual
1006 > way to measure the pressure is based on virial theorem of Clausius
1007 > which states that the virial is equal to $-3Nk_BT$. For a system
1008 > with forces between particles, the total virial, $W$, contains the
1009 > contribution from external pressure and interaction between the
1010 > particles:
1011   \[
1012 < I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1012 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1013 > f_{ij} } } \right\rangle
1014   \]
1015 < where $I_{ii}$ is the diagonal element of the inertia tensor. This
1016 < constrained Hamiltonian equation subjects to a holonomic constraint,
1015 > where $f_{ij}$ is the force between particle $i$ and $j$ at a
1016 > distance $r_{ij}$. Thus, the expression for the pressure is given
1017 > by:
1018   \begin{equation}
1019 < Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1019 > P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1020 > < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1021   \end{equation}
1022 < which is used to ensure rotation matrix's orthogonality.
1023 < Differentiating \ref{introEquation:orthogonalConstraint} and using
1024 < Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1022 >
1023 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1024 >
1025 > Structural Properties of a simple fluid can be described by a set of
1026 > distribution functions. Among these functions,the \emph{pair
1027 > distribution function}, also known as \emph{radial distribution
1028 > function}, is of most fundamental importance to liquid theory.
1029 > Experimentally, pair distribution functions can be gathered by
1030 > Fourier transforming raw data from a series of neutron diffraction
1031 > experiments and integrating over the surface
1032 > factor.\cite{Powles1973} The experimental results can serve as a
1033 > criterion to justify the correctness of a liquid model. Moreover,
1034 > various equilibrium thermodynamic and structural properties can also
1035 > be expressed in terms of the radial distribution
1036 > function.\cite{Allen1987} The pair distribution functions $g(r)$
1037 > gives the probability that a particle $i$ will be located at a
1038 > distance $r$ from a another particle $j$ in the system
1039   \begin{equation}
1040 < Q^t PJ^{ - 1}  + J^{ - 1} P^t Q = 0 . \\
1041 < \label{introEquation:RBFirstOrderConstraint}
1040 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1041 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1042 > (r)}{\rho}.
1043   \end{equation}
1044 + Note that the delta function can be replaced by a histogram in
1045 + computer simulation. Peaks in $g(r)$ represent solvent shells, and
1046 + the height of these peaks gradually decreases to 1 as the liquid of
1047 + large distance approaches the bulk density.
1048  
982 Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
983 \ref{introEquation:motionHamiltonianMomentum}), one can write down
984 the equations of motion,
985 \[
986 \begin{array}{c}
987 \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
988 \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
989 \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
990 \frac{{dP}}{{dt}} =  - \nabla _q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
991 \end{array}
992 \]
1049  
1050 + \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1051 + Properties}}
1052  
1053 < \[
1054 < M = \left\{ {(Q,P):Q^T Q = 1,Q^t PJ^{ - 1}  + J^{ - 1} P^t Q = 0}
1055 < \right\} .
1056 < \]
1053 > Time-dependent properties are usually calculated using \emph{time
1054 > correlation functions}, which correlate random variables $A$ and $B$
1055 > at two different times,
1056 > \begin{equation}
1057 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1058 > \label{introEquation:timeCorrelationFunction}
1059 > \end{equation}
1060 > If $A$ and $B$ refer to same variable, this kind of correlation
1061 > functions are called \emph{autocorrelation functions}. One typical example is the velocity autocorrelation
1062 > function which is directly related to transport properties of
1063 > molecular liquids:
1064 > \begin{equation}
1065 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1066 > \right\rangle } dt
1067 > \end{equation}
1068 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1069 > function, which is averaged over time origins and over all the
1070 > atoms, the dipole autocorrelation functions is calculated for the
1071 > entire system. The dipole autocorrelation function is given by:
1072 > \begin{equation}
1073 > c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1074 > \right\rangle
1075 > \end{equation}
1076 > Here $u_{tot}$ is the net dipole of the entire system and is given
1077 > by
1078 > \begin{equation}
1079 > u_{tot} (t) = \sum\limits_i {u_i (t)}.
1080 > \end{equation}
1081 > In principle, many time correlation functions can be related to
1082 > Fourier transforms of the infrared, Raman, and inelastic neutron
1083 > scattering spectra of molecular liquids. In practice, one can
1084 > extract the IR spectrum from the intensity of the molecular dipole
1085 > fluctuation at each frequency using the following relationship:
1086 > \begin{equation}
1087 > \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1088 > i2\pi vt} dt}.
1089 > \end{equation}
1090  
1091 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
1091 > \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1092  
1093 < \subsection{\label{introSection:symplecticDiscretizationRB}Symplectic Discretization of Euler Equations}
1093 > Rigid bodies are frequently involved in the modeling of different
1094 > areas, including engineering, physics and chemistry. For example,
1095 > missiles and vehicles are usually modeled by rigid bodies.  The
1096 > movement of the objects in 3D gaming engines or other physics
1097 > simulators is governed by rigid body dynamics. In molecular
1098 > simulations, rigid bodies are used to simplify protein-protein
1099 > docking studies.\cite{Gray2003}
1100  
1101 + It is very important to develop stable and efficient methods to
1102 + integrate the equations of motion for orientational degrees of
1103 + freedom. Euler angles are the natural choice to describe the
1104 + rotational degrees of freedom. However, due to $\frac {1}{sin
1105 + \theta}$ singularities, the numerical integration of corresponding
1106 + equations of these motion is very inefficient and inaccurate.
1107 + Although an alternative integrator using multiple sets of Euler
1108 + angles can overcome this difficulty\cite{Barojas1973}, the
1109 + computational penalty and the loss of angular momentum conservation
1110 + still remain. A singularity-free representation utilizing
1111 + quaternions was developed by Evans in 1977.\cite{Evans1977}
1112 + Unfortunately, this approach used a nonseparable Hamiltonian
1113 + resulting from the quaternion representation, which prevented the
1114 + symplectic algorithm from being utilized. Another different approach
1115 + is to apply holonomic constraints to the atoms belonging to the
1116 + rigid body. Each atom moves independently under the normal forces
1117 + deriving from potential energy and constraint forces which are used
1118 + to guarantee the rigidness. However, due to their iterative nature,
1119 + the SHAKE and Rattle algorithms also converge very slowly when the
1120 + number of constraints increases.\cite{Ryckaert1977, Andersen1983}
1121  
1122 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1122 > A break-through in geometric literature suggests that, in order to
1123 > develop a long-term integration scheme, one should preserve the
1124 > symplectic structure of the propagator. By introducing a conjugate
1125 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1126 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1127 > proposed to evolve the Hamiltonian system in a constraint manifold
1128 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1129 > An alternative method using the quaternion representation was
1130 > developed by Omelyan.\cite{Omelyan1998} However, both of these
1131 > methods are iterative and inefficient. In this section, we descibe a
1132 > symplectic Lie-Poisson integrator for rigid bodies developed by
1133 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1134  
1135 < \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
1136 <
1009 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
1010 <
1135 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1136 > The Hamiltonian of a rigid body is given by
1137   \begin{equation}
1138 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1139 < \label{introEquation:bathGLE}
1138 > H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1139 > V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1140 > \label{introEquation:RBHamiltonian}
1141   \end{equation}
1142 < where $H_B$ is harmonic bath Hamiltonian,
1142 > Here, $q$ and $Q$  are the position vector and rotation matrix for
1143 > the rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ ,
1144 > and $J$, a diagonal matrix, is defined by
1145   \[
1146 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1018 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1146 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1147   \]
1148 < and $\Delta U$ is bilinear system-bath coupling,
1149 < \[
1150 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1151 < \]
1152 < Completing the square,
1148 > where $I_{ii}$ is the diagonal element of the inertia tensor. This
1149 > constrained Hamiltonian equation is subjected to a holonomic
1150 > constraint,
1151 > \begin{equation}
1152 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1153 > \end{equation}
1154 > which is used to ensure the rotation matrix's unitarity. Using
1155 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} and Eq.~
1156 > \ref{introEquation:motionHamiltonianMomentum}, one can write down
1157 > the equations of motion,
1158 > \begin{eqnarray}
1159 > \frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\
1160 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q), \label{introEquation:RBMotionMomentum}\\
1161 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1},  \label{introEquation:RBMotionRotation}\\
1162 > \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1163 > \end{eqnarray}
1164 > Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and
1165 > using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1166 > \begin{equation}
1167 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1168 > \label{introEquation:RBFirstOrderConstraint}
1169 > \end{equation}
1170 > In general, there are two ways to satisfy the holonomic constraints.
1171 > We can use a constraint force provided by a Lagrange multiplier on
1172 > the normal manifold to keep the motion on the constraint space. Or
1173 > we can simply evolve the system on the constraint manifold. These
1174 > two methods have been proved to be equivalent. The holonomic
1175 > constraint and equations of motions define a constraint manifold for
1176 > rigid bodies
1177   \[
1178 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
1179 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1028 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1029 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
1030 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1178 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1179 > \right\}.
1180   \]
1181 < and putting it back into Eq.~\ref{introEquation:bathGLE},
1181 > Unfortunately, this constraint manifold is not $T^* SO(3)$ which is
1182 > a symplectic manifold on Lie rotation group $SO(3)$. However, it
1183 > turns out that under symplectic transformation, the cotangent space
1184 > and the phase space are diffeomorphic. By introducing
1185   \[
1186 < H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1035 < {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1036 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1037 < w_\alpha ^2 }}x} \right)^2 } \right\}}
1186 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1187   \]
1188 < where
1188 > the mechanical system subjected to a holonomic constraint manifold $M$
1189 > can be re-formulated as a Hamiltonian system on the cotangent space
1190   \[
1191 < W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1192 < }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1191 > T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1192 > 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1193   \]
1194 < Since the first two terms of the new Hamiltonian depend only on the
1195 < system coordinates, we can get the equations of motion for
1196 < Generalized Langevin Dynamics by Hamilton's equations
1197 < \ref{introEquation:motionHamiltonianCoordinate,
1198 < introEquation:motionHamiltonianMomentum},
1199 < \begin{align}
1200 < \dot p &=  - \frac{{\partial H}}{{\partial x}}
1051 <       &= m\ddot x
1052 <       &= - \frac{{\partial W(x)}}{{\partial x}} - \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)}
1053 < \label{introEquation:Lp5}
1054 < \end{align}
1055 < , and
1056 < \begin{align}
1057 < \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
1058 <                &= m\ddot x_\alpha
1059 <                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
1060 < \end{align}
1061 <
1062 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1063 <
1194 > For a body fixed vector $X_i$ with respect to the center of mass of
1195 > the rigid body, its corresponding lab fixed vector $X_i^{lab}$  is
1196 > given as
1197 > \begin{equation}
1198 > X_i^{lab} = Q X_i + q.
1199 > \end{equation}
1200 > Therefore, potential energy $V(q,Q)$ is defined by
1201   \[
1202 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1202 > V(q,Q) = V(Q X_0 + q).
1203   \]
1204 <
1204 > Hence, the force and torque are given by
1205   \[
1206 < L(x + y) = L(x) + L(y)
1206 > \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1207   \]
1208 <
1208 > and
1209   \[
1210 < L(ax) = aL(x)
1210 > \nabla _Q V(q,Q) = F(q,Q)X_i^t
1211   \]
1212 <
1212 > respectively. As a common choice to describe the rotation dynamics
1213 > of the rigid body, the angular momentum on the body fixed frame $\Pi
1214 > = Q^t P$ is introduced to rewrite the equations of motion,
1215 > \begin{equation}
1216 > \begin{array}{l}
1217 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda,  \\
1218 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1},  \\
1219 > \end{array}
1220 > \label{introEqaution:RBMotionPI}
1221 > \end{equation}
1222 > as well as holonomic constraints $\Pi J^{ - 1}  + J^{ - 1} \Pi ^t  =
1223 > 0$ and $Q^T Q = 1$. For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a
1224 > matrix $\hat v \in so(3)^ \star$, the hat-map isomorphism,
1225 > \begin{equation}
1226 > v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1227 > {\begin{array}{*{20}c}
1228 >   0 & { - v_3 } & {v_2 }  \\
1229 >   {v_3 } & 0 & { - v_1 }  \\
1230 >   { - v_2 } & {v_1 } & 0  \\
1231 > \end{array}} \right),
1232 > \label{introEquation:hatmapIsomorphism}
1233 > \end{equation}
1234 > will let us associate the matrix products with traditional vector
1235 > operations
1236   \[
1237 < L(\dot x) = pL(x) - px(0)
1237 > \hat vu = v \times u.
1238   \]
1239 <
1239 > Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew
1240 > matrix,
1241 > \begin{eqnarray}
1242 > (\dot \Pi  - \dot \Pi ^T )&= &(\Pi  - \Pi ^T )(J^{ - 1} \Pi  + \Pi J^{ - 1} ) \notag \\
1243 > & & + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  -
1244 > (\Lambda  - \Lambda ^T ). \label{introEquation:skewMatrixPI}
1245 > \end{eqnarray}
1246 > Since $\Lambda$ is symmetric, the last term of
1247 > Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1248 > Lagrange multiplier $\Lambda$ is absent from the equations of
1249 > motion. This unique property eliminates the requirement of
1250 > iterations which can not be avoided in other methods.\cite{Kol1997,
1251 > Omelyan1998} Applying the hat-map isomorphism, we obtain the
1252 > equation of motion for angular momentum
1253 > \begin{equation}
1254 > \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1255 > F_i (r,Q)} \right) \times X_i }.
1256 > \label{introEquation:bodyAngularMotion}
1257 > \end{equation}
1258 > In the same manner, the equation of motion for rotation matrix is
1259 > given by
1260   \[
1261 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1261 > \dot Q = Qskew(I^{ - 1} \pi ).
1262   \]
1263  
1264 < \[
1265 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1086 < \]
1264 > \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1265 > Lie-Poisson Integrator for Free Rigid Bodies}
1266  
1267 < Some relatively important transformation,
1268 < \[
1269 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1267 > If there are no external forces exerted on the rigid body, the only
1268 > contribution to the rotational motion is from the kinetic energy
1269 > (the first term of \ref{introEquation:bodyAngularMotion}). The free
1270 > rigid body is an example of a Lie-Poisson system with Hamiltonian
1271 > function
1272 > \begin{equation}
1273 > T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1274 > \label{introEquation:rotationalKineticRB}
1275 > \end{equation}
1276 > where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1277 > Lie-Poisson structure matrix,
1278 > \begin{equation}
1279 > J(\pi ) = \left( {\begin{array}{*{20}c}
1280 >   0 & {\pi _3 } & { - \pi _2 }  \\
1281 >   { - \pi _3 } & 0 & {\pi _1 }  \\
1282 >   {\pi _2 } & { - \pi _1 } & 0  \\
1283 > \end{array}} \right).
1284 > \end{equation}
1285 > Thus, the dynamics of free rigid body is governed by
1286 > \begin{equation}
1287 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi ).
1288 > \end{equation}
1289 > One may notice that each $T_i^r$ in
1290 > Eq.~\ref{introEquation:rotationalKineticRB} can be solved exactly.
1291 > For instance, the equations of motion due to $T_1^r$ are given by
1292 > \begin{equation}
1293 > \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1294 > \label{introEqaution:RBMotionSingleTerm}
1295 > \end{equation}
1296 > with
1297 > \[ R_1  = \left( {\begin{array}{*{20}c}
1298 >   0 & 0 & 0  \\
1299 >   0 & 0 & {\pi _1 }  \\
1300 >   0 & { - \pi _1 } & 0  \\
1301 > \end{array}} \right).
1302   \]
1303 <
1303 > The solutions of Eq.~\ref{introEqaution:RBMotionSingleTerm} is
1304   \[
1305 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1305 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1306 > Q(0)e^{\Delta tR_1 }
1307   \]
1308 <
1308 > with
1309   \[
1310 < L(1) = \frac{1}{p}
1310 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1311 >   0 & 0 & 0  \\
1312 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1313 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1314 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1315 > \]
1316 > To reduce the cost of computing expensive functions in $e^{\Delta
1317 > tR_1 }$, we can use the Cayley transformation to obtain a
1318 > single-aixs propagator,
1319 > \begin{eqnarray*}
1320 > e^{\Delta tR_1 }  & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta
1321 > tR_1 ) \\
1322 > %
1323 > & \approx & \left( \begin{array}{ccc}
1324 > 1 & 0 & 0 \\
1325 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1326 > \theta^2 / 4} \\
1327 > 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1328 > \theta^2 / 4}
1329 > \end{array}
1330 > \right).
1331 > \end{eqnarray*}
1332 > The propagators for $T_2^r$ and $T_3^r$ can be found in the same
1333 > manner. In order to construct a second-order symplectic method, we
1334 > split the angular kinetic Hamiltonian function into five terms
1335 > \[
1336 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1337 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1338 > (\pi _1 ).
1339   \]
1340 + By concatenating the propagators corresponding to these five terms,
1341 + we can obtain an symplectic integrator,
1342 + \[
1343 + \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1344 + \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1345 + \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1346 + _1 }.
1347 + \]
1348 + The non-canonical Lie-Poisson bracket $\{F, G\}$ of two functions $F(\pi )$ and $G(\pi )$ is defined by
1349 + \[
1350 + \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1351 + ).
1352 + \]
1353 + If the Poisson bracket of a function $F$ with an arbitrary smooth
1354 + function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1355 + conserved quantity in Poisson system. We can easily verify that the
1356 + norm of the angular momentum, $\parallel \pi
1357 + \parallel$, is a \emph{Casimir}.\cite{McLachlan1993} Let $F(\pi ) = S(\frac{{\parallel
1358 + \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1359 + then by the chain rule
1360 + \[
1361 + \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1362 + }}{2})\pi.
1363 + \]
1364 + Thus, $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel
1365 + \pi
1366 + \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1367 + Lie-Poisson integrator is found to be both extremely efficient and
1368 + stable. These properties can be explained by the fact the small
1369 + angle approximation is used and the norm of the angular momentum is
1370 + conserved.
1371  
1372 < First, the bath coordinates,
1372 > \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1373 > Splitting for Rigid Body}
1374 >
1375 > The Hamiltonian of rigid body can be separated in terms of kinetic
1376 > energy and potential energy, $H = T(p,\pi ) + V(q,Q)$. The equations
1377 > of motion corresponding to potential energy and kinetic energy are
1378 > listed in Table~\ref{introTable:rbEquations}.
1379 > \begin{table}
1380 > \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1381 > \label{introTable:rbEquations}
1382 > \begin{center}
1383 > \begin{tabular}{|l|l|}
1384 >  \hline
1385 >  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1386 >  Potential & Kinetic \\
1387 >  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1388 >  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1389 >  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1390 >  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1391 >  \hline
1392 > \end{tabular}
1393 > \end{center}
1394 > \end{table}
1395 > A second-order symplectic method is now obtained by the composition
1396 > of the position and velocity propagators,
1397   \[
1398 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
1399 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
1105 < }}L(x)
1398 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1399 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1400   \]
1401 + Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1402 + sub-propagators which corresponding to force and torque
1403 + respectively,
1404   \[
1405 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1406 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1405 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1406 > _{\Delta t/2,\tau }.
1407   \]
1408 < Then, the system coordinates,
1409 < \begin{align}
1410 < mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1411 < \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1412 < }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1413 < (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1414 < }}\omega _\alpha ^2 L(x)} \right\}}
1415 < %
1416 < &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1417 < \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1418 < - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1419 < - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1420 < \end{align}
1421 < Then, the inverse transform,
1408 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1409 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1410 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1411 > kinetic energy can be separated to translational kinetic term, $T^t
1412 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1413 > \begin{equation}
1414 > T(p,\pi ) =T^t (p) + T^r (\pi ).
1415 > \end{equation}
1416 > where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1417 > defined by Eq.~\ref{introEquation:rotationalKineticRB}. Therefore,
1418 > the corresponding propagators are given by
1419 > \[
1420 > \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1421 > _{\Delta t,T^r }.
1422 > \]
1423 > Finally, we obtain the overall symplectic propagators for freely
1424 > moving rigid bodies
1425 > \begin{eqnarray}
1426 > \varphi _{\Delta t}  &=& \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \notag\\
1427 >  & & \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\\
1428 >  & & \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .
1429 > \label{introEquation:overallRBFlowMaps}
1430 > \end{eqnarray}
1431  
1432 < \begin{align}
1433 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1432 > \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1433 > As an alternative to newtonian dynamics, Langevin dynamics, which
1434 > mimics a simple heat bath with stochastic and dissipative forces,
1435 > has been applied in a variety of studies. This section will review
1436 > the theory of Langevin dynamics. A brief derivation of the generalized
1437 > Langevin equation will be given first. Following that, we will
1438 > discuss the physical meaning of the terms appearing in the equation.
1439 >
1440 > \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1441 >
1442 > A harmonic bath model, in which an effective set of harmonic
1443 > oscillators are used to mimic the effect of a linearly responding
1444 > environment, has been widely used in quantum chemistry and
1445 > statistical mechanics. One of the successful applications of
1446 > Harmonic bath model is the derivation of the Generalized Langevin
1447 > Dynamics (GLE). Consider a system, in which the degree of
1448 > freedom $x$ is assumed to couple to the bath linearly, giving a
1449 > Hamiltonian of the form
1450 > \begin{equation}
1451 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1452 > \label{introEquation:bathGLE}.
1453 > \end{equation}
1454 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1455 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1456 > \[
1457 > H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1458 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  x_\alpha ^2 }
1459 > \right\}}
1460 > \]
1461 > where the index $\alpha$ runs over all the bath degrees of freedom,
1462 > $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1463 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1464 > coupling,
1465 > \[
1466 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1467 > \]
1468 > where $g_\alpha$ are the coupling constants between the bath
1469 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1470 > Introducing
1471 > \[
1472 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1473 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1474 > \]
1475 > and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as
1476 > \[
1477 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1478 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1479 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1480 > w_\alpha ^2 }}x} \right)^2 } \right\}}.
1481 > \]
1482 > Since the first two terms of the new Hamiltonian depend only on the
1483 > system coordinates, we can get the equations of motion for
1484 > Generalized Langevin Dynamics by Hamilton's equations,
1485 > \begin{equation}
1486 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1487 > \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1488 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1489 > \label{introEquation:coorMotionGLE}
1490 > \end{equation}
1491 > and
1492 > \begin{equation}
1493 > m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1494 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1495 > \label{introEquation:bathMotionGLE}
1496 > \end{equation}
1497 > In order to derive an equation for $x$, the dynamics of the bath
1498 > variables $x_\alpha$ must be solved exactly first. As an integral
1499 > transform which is particularly useful in solving linear ordinary
1500 > differential equations,the Laplace transform is the appropriate tool
1501 > to solve this problem. The basic idea is to transform the difficult
1502 > differential equations into simple algebra problems which can be
1503 > solved easily. Then, by applying the inverse Laplace transform, we
1504 > can retrieve the solutions of the original problems. Let $f(t)$ be a
1505 > function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$
1506 > is a new function defined as
1507 > \[
1508 > L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1509 > \]
1510 > where  $p$ is real and  $L$ is called the Laplace Transform
1511 > Operator. Below are some important properties of the Laplace transform
1512 > \begin{eqnarray*}
1513 > L(x + y)  & = & L(x) + L(y) \\
1514 > L(ax)     & = & aL(x) \\
1515 > L(\dot x) & = & pL(x) - px(0) \\
1516 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1517 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1518 > \end{eqnarray*}
1519 > Applying the Laplace transform to the bath coordinates, we obtain
1520 > \begin{eqnarray*}
1521 > 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), \\
1522 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}. \\
1523 > \end{eqnarray*}
1524 > In the same way, the system coordinates become
1525 > \begin{eqnarray*}
1526 > mL(\ddot x) & = &
1527 >  - \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\}}  \\
1528 >  & & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}.
1529 > \end{eqnarray*}
1530 > With the help of some relatively important inverse Laplace
1531 > transformations:
1532 > \[
1533 > \begin{array}{c}
1534 > L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1535 > L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1536 > L(1) = \frac{1}{p} \\
1537 > \end{array}
1538 > \]
1539 > we obtain
1540 > \begin{eqnarray*}
1541 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1542   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1543   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1544 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1545 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1546 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1547 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1544 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1545 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1546 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1547 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1548 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}\\
1549   %
1550 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1551 < {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1552 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1553 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1554 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1555 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1556 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1557 < (\omega _\alpha  t)} \right\}}
1558 < \end{align}
1559 <
1550 > & = & -
1551 > \frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha
1552 > = 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha
1553 > ^2 }}} \right)\cos (\omega _\alpha
1554 > t)\dot x(t - \tau )d} \tau }  \\
1555 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1556 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1557 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1558 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1559 > \end{eqnarray*}
1560 > Introducing a \emph{dynamic friction kernel}
1561   \begin{equation}
1562 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1563 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1564 + \label{introEquation:dynamicFrictionKernelDefinition}
1565 + \end{equation}
1566 + and \emph{a random force}
1567 + \begin{equation}
1568 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1569 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1570 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1571 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1572 + \label{introEquation:randomForceDefinition}
1573 + \end{equation}
1574 + the equation of motion can be rewritten as
1575 + \begin{equation}
1576   m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1577   (t)\dot x(t - \tau )d\tau }  + R(t)
1578   \label{introEuqation:GeneralizedLangevinDynamics}
1579   \end{equation}
1580 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1581 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1580 > which is known as the \emph{generalized Langevin equation} (GLE).
1581 >
1582 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1583 >
1584 > One may notice that $R(t)$ depends only on initial conditions, which
1585 > implies it is completely deterministic within the context of a
1586 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1587 > uncorrelated to $x$ and $\dot x$, $\left\langle {x(t)R(t)}
1588 > \right\rangle  = 0, \left\langle {\dot x(t)R(t)} \right\rangle  =
1589 > 0.$ This property is what we expect from a truly random process. As
1590 > long as the model chosen for $R(t)$ was a gaussian distribution in
1591 > general, the stochastic nature of the GLE still remains.
1592 > %dynamic friction kernel
1593 > The convolution integral
1594   \[
1595 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1154 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1595 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1596   \]
1597 < For an infinite harmonic bath, we can use the spectral density and
1598 < an integral over frequencies.
1599 <
1597 > depends on the entire history of the evolution of $x$, which implies
1598 > that the bath retains memory of previous motions. In other words,
1599 > the bath requires a finite time to respond to change in the motion
1600 > of the system. For a sluggish bath which responds slowly to changes
1601 > in the system coordinate, we may regard $\xi(t)$ as a constant
1602 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1603   \[
1604 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1161 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1162 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1163 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1604 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1605   \]
1606 < The random forces depend only on initial conditions.
1166 <
1167 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1168 < So we can define a new set of coordinates,
1606 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1607   \[
1608 < q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1609 < ^2 }}x(0)
1608 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1609 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1610   \]
1611 < This makes
1611 > which can be used to describe the effect of dynamic caging in
1612 > viscous solvents. The other extreme is the bath that responds
1613 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1614 > taken as a $delta$ function in time:
1615   \[
1616 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1616 > \xi (t) = 2\xi _0 \delta (t).
1617   \]
1618 < And since the $q$ coordinates are harmonic oscillators,
1618 > Hence, the convolution integral becomes
1619   \[
1620 < \begin{array}{l}
1621 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1181 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1182 < \end{array}
1620 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1621 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1622   \]
1623 + and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes the
1624 + Langevin equation
1625 + \begin{equation}
1626 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1627 + x(t) + R(t) \label{introEquation:LangevinEquation}.
1628 + \end{equation}
1629 + The static friction coefficient $\xi _0$ can either be calculated
1630 + from spectral density or be determined by Stokes' law for regular
1631 + shaped particles. A brief review on calculating friction tensors for
1632 + arbitrary shaped particles is given in
1633 + Sec.~\ref{introSection:frictionTensor}.
1634  
1635 < \begin{align}
1186 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1187 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1188 < (t)q_\beta  (0)} \right\rangle } }
1189 < %
1190 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1191 < \right\rangle \cos (\omega _\alpha  t)}
1192 < %
1193 < &= kT\xi (t)
1194 < \end{align}
1635 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1636  
1637 + Defining a new set of coordinates
1638 + \[
1639 + q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1640 + ^2 }}x(0),
1641 + \]
1642 + we can rewrite $R(t)$ as
1643 + \[
1644 + R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1645 + \]
1646 + And since the $q$ coordinates are harmonic oscillators,
1647 + \begin{eqnarray*}
1648 + \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1649 + \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1650 + \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1651 + \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 } }  \\
1652 +  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1653 +  & = &kT\xi (t)
1654 + \end{eqnarray*}
1655 + Thus, we recover the \emph{second fluctuation dissipation theorem}
1656   \begin{equation}
1657   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1658 < \label{introEquation:secondFluctuationDissipation}
1658 > \label{introEquation:secondFluctuationDissipation},
1659   \end{equation}
1660 <
1661 < \section{\label{introSection:hydroynamics}Hydrodynamics}
1202 <
1203 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
1204 < \subsection{\label{introSection:analyticalApproach}Analytical
1205 < Approach}
1206 <
1207 < \subsection{\label{introSection:approximationApproach}Approximation
1208 < Approach}
1209 <
1210 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1211 < Body}
1212 <
1213 < \section{\label{introSection:correlationFunctions}Correlation Functions}
1660 > which acts as a constraint on the possible ways in which one can
1661 > model the random force and friction kernel.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines