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 2912 by tim, Fri Jun 30 02:45:29 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 bodies
71 > \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 227 | Line 214 | possible states. Each possible state of the system cor
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 395 | Line 341 | distribution,
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,
350 > as uniform over the whole integral. Thus, the number of phase points
351 > 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.
# Line 412 | Line 358 | With the help of stationary assumption
358   \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
359   \frac{d}{{dt}}(\delta v) = 0.
360   \end{equation}
361 < With the help of stationary assumption
362 < (\ref{introEquation:stationary}), we obtain the principle of the
363 < \emph{conservation of extension in phase space},
361 > With the help of the stationary assumption
362 > (Eq.~\ref{introEquation:stationary}), we obtain the principle of
363 > \emph{conservation of volume in phase space},
364   \begin{equation}
365   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
366   ...dq_f dp_1 } ..dp_f  = 0.
367   \label{introEquation:volumePreserving}
368   \end{equation}
369  
370 < \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
370 > \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
371  
372 < Liouville's theorem can be expresses in a variety of different forms
372 > Liouville's theorem can be expressed in a variety of different forms
373   which are convenient within different contexts. For any two function
374   $F$ and $G$ of the coordinates and momenta of a system, the Poisson
375   bracket ${F, G}$ is defined as
# Line 434 | Line 380 | Substituting equations of motion in Hamiltonian formal
380   q_i }}} \right)}.
381   \label{introEquation:poissonBracket}
382   \end{equation}
383 < Substituting equations of motion in Hamiltonian formalism(
384 < \ref{introEquation:motionHamiltonianCoordinate} ,
385 < \ref{introEquation:motionHamiltonianMomentum} ) into
386 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
387 < theorem using Poisson bracket notion,
383 > Substituting equations of motion in Hamiltonian formalism
384 > (Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
385 > Eq.~\ref{introEquation:motionHamiltonianMomentum}) into
386 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
387 > Liouville's theorem using Poisson bracket notion,
388   \begin{equation}
389   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
390   {\rho ,H} \right\}.
# Line 457 | Line 403 | expressed as
403   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
404   \label{introEquation:liouvilleTheoremInOperator}
405   \end{equation}
406 <
406 > which can help define a propagator $\rho (t) = e^{-iLt} \rho (0)$.
407   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
408  
409   Various thermodynamic properties can be calculated from Molecular
410   Dynamics simulation. By comparing experimental values with the
411   calculated properties, one can determine the accuracy of the
412 < simulation and the quality of the underlying model. However, both of
413 < experiment and computer simulation are usually performed during a
412 > simulation and the quality of the underlying model. However, both
413 > experiments and computer simulations are usually performed during a
414   certain time interval and the measurements are averaged over a
415 < period of them which is different from the average behavior of
416 < many-body system in Statistical Mechanics. Fortunately, Ergodic
417 < Hypothesis is proposed to make a connection between time average and
418 < ensemble average. It states that time average and average over the
419 < statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
415 > period of time which is different from the average behavior of
416 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
417 > Hypothesis makes a connection between time average and the ensemble
418 > average. It states that the time average and average over the
419 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}:
420   \begin{equation}
421   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
422   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 479 | Line 425 | sufficiently long time (longer than relaxation time),
425   where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
426   physical quantity and $\rho (p(t), q(t))$ is the equilibrium
427   distribution function. If an observation is averaged over a
428 < sufficiently long time (longer than relaxation time), all accessible
429 < microstates in phase space are assumed to be equally probed, giving
430 < a properly weighted statistical average. This allows the researcher
431 < freedom of choice when deciding how best to measure a given
432 < observable. In case an ensemble averaged approach sounds most
433 < reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
428 > sufficiently long time (longer than the relaxation time), all
429 > accessible microstates in phase space are assumed to be equally
430 > probed, giving a properly weighted statistical average. This allows
431 > the researcher freedom of choice when deciding how best to measure a
432 > given observable. In case an ensemble averaged approach sounds most
433 > reasonable, the Monte Carlo methods\cite{Metropolis1949} can be
434   utilized. Or if the system lends itself to a time averaging
435   approach, the Molecular Dynamics techniques in
436   Sec.~\ref{introSection:molecularDynamics} will be the best
437   choice\cite{Frenkel1996}.
438  
439   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
440 < A variety of numerical integrators were proposed to simulate the
441 < motions. They usually begin with an initial conditionals and move
442 < the objects in the direction governed by the differential equations.
443 < However, most of them ignore the hidden physical law contained
444 < within the equations. Since 1990, geometric integrators, which
445 < preserve various phase-flow invariants such as symplectic structure,
446 < volume and time reversal symmetry, are developed to address this
447 < issue. The velocity verlet method, which happens to be a simple
448 < example of symplectic integrator, continues to gain its popularity
449 < in molecular dynamics community. This fact can be partly explained
450 < by its geometric nature.
440 > A variety of numerical integrators have been proposed to simulate
441 > the motions of atoms in MD simulation. They usually begin with
442 > initial conditionals and move the objects in the direction governed
443 > by the differential equations. However, most of them ignore the
444 > hidden physical laws contained within the equations. Since 1990,
445 > geometric integrators, which preserve various phase-flow invariants
446 > such as symplectic structure, volume and time reversal symmetry,
447 > were developed to address this issue\cite{Dullweber1997,
448 > McLachlan1998, Leimkuhler1999}. The velocity Verlet method, which
449 > happens to be a simple example of symplectic integrator, continues
450 > to gain popularity in the molecular dynamics community. This fact
451 > can be partly explained by its geometric nature.
452  
453 < \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
454 < A \emph{manifold} is an abstract mathematical space. It locally
455 < looks like Euclidean space, but when viewed globally, it may have
456 < more complicate structure. A good example of manifold is the surface
457 < of Earth. It seems to be flat locally, but it is round if viewed as
458 < a whole. A \emph{differentiable manifold} (also known as
459 < \emph{smooth manifold}) is a manifold with an open cover in which
460 < the covering neighborhoods are all smoothly isomorphic to one
461 < another. In other words,it is possible to apply calculus on
515 < \emph{differentiable manifold}. A \emph{symplectic manifold} is
516 < defined as a pair $(M, \omega)$ which consisting of a
453 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
454 > A \emph{manifold} is an abstract mathematical space. It looks
455 > locally like Euclidean space, but when viewed globally, it may have
456 > more complicated structure. A good example of manifold is the
457 > surface of Earth. It seems to be flat locally, but it is round if
458 > viewed as a whole. A \emph{differentiable manifold} (also known as
459 > \emph{smooth manifold}) is a manifold on which it is possible to
460 > apply calculus\cite{Hirsch1997}. A \emph{symplectic manifold} is
461 > defined as a pair $(M, \omega)$ which consists of a
462   \emph{differentiable manifold} $M$ and a close, non-degenerated,
463   bilinear symplectic form, $\omega$. A symplectic form on a vector
464   space $V$ is a function $\omega(x, y)$ which satisfies
465   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
466   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
467 < $\omega(x, x) = 0$. Cross product operation in vector field is an
468 < example of symplectic form.
467 > $\omega(x, x) = 0$\cite{McDuff1998}. The cross product operation in
468 > vector field is an example of symplectic form. One of the
469 > motivations to study \emph{symplectic manifolds} in Hamiltonian
470 > Mechanics is that a symplectic manifold can represent all possible
471 > configurations of the system and the phase space of the system can
472 > be described by it's cotangent bundle\cite{Jost2002}. Every
473 > symplectic manifold is even dimensional. For instance, in Hamilton
474 > equations, coordinate and momentum always appear in pairs.
475  
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
476   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
477  
478 < For a ordinary differential system defined as
478 > For an ordinary differential system defined as
479   \begin{equation}
480   \dot x = f(x)
481   \end{equation}
482 < where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
482 > where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
483 > $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
484 > function and $J$ is the skew-symmetric matrix
485   \begin{equation}
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}
486   J = \left( {\begin{array}{*{20}c}
487     0 & I  \\
488     { - I} & 0  \\
# Line 560 | Line 492 | system can be rewritten as,
492   where $I$ is an identity matrix. Using this notation, Hamiltonian
493   system can be rewritten as,
494   \begin{equation}
495 < \frac{d}{{dt}}x = J\nabla _x H(x)
495 > \frac{d}{{dt}}x = J\nabla _x H(x).
496   \label{introEquation:compactHamiltonian}
497   \end{equation}In this case, $f$ is
498 < called a \emph{Hamiltonian vector field}.
499 <
568 < Another generalization of Hamiltonian dynamics is Poisson Dynamics,
498 > called a \emph{Hamiltonian vector field}. Another generalization of
499 > Hamiltonian dynamics is Poisson Dynamics\cite{Olver1986},
500   \begin{equation}
501   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
502   \end{equation}
503   The most obvious change being that matrix $J$ now depends on $x$.
573 The free rigid body is an example of Poisson system (actually a
574 Lie-Poisson system) with Hamiltonian function of angular kinetic
575 energy.
576 \begin{equation}
577 J(\pi ) = \left( {\begin{array}{*{20}c}
578   0 & {\pi _3 } & { - \pi _2 }  \\
579   { - \pi _3 } & 0 & {\pi _1 }  \\
580   {\pi _2 } & { - \pi _1 } & 0  \\
581 \end{array}} \right)
582 \end{equation}
504  
505 < \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}
505 > \subsection{\label{introSection:exactFlow}Exact Propagator}
506  
507 < \subsection{\label{introSection:exactFlow}Exact Flow}
508 <
591 < Let $x(t)$ be the exact solution of the ODE system,
507 > Let $x(t)$ be the exact solution of the ODE
508 > system,
509   \begin{equation}
510 < \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
511 < \end{equation}
512 < The exact flow(solution) $\varphi_\tau$ is defined by
513 < \[
514 < x(t+\tau) =\varphi_\tau(x(t))
510 > \frac{{dx}}{{dt}} = f(x), \label{introEquation:ODE}
511 > \end{equation} we can
512 > define its exact propagator $\varphi_\tau$:
513 > \[ x(t+\tau)
514 > =\varphi_\tau(x(t))
515   \]
516   where $\tau$ is a fixed time step and $\varphi$ is a map from phase
517 < space to itself. The flow has the continuous group property,
517 > space to itself. The propagator has the continuous group property,
518   \begin{equation}
519   \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
520   + \tau _2 } .
# Line 606 | Line 523 | Therefore, the exact flow is self-adjoint,
523   \begin{equation}
524   \varphi _\tau   \circ \varphi _{ - \tau }  = I
525   \end{equation}
526 < Therefore, the exact flow is self-adjoint,
526 > Therefore, the exact propagator is self-adjoint,
527   \begin{equation}
528   \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
529   \end{equation}
530 < The exact flow can also be written in terms of the of an operator,
530 > The exact propagator can also be written in terms of operator,
531   \begin{equation}
532   \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
533   }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
534   \label{introEquation:exponentialOperator}
535   \end{equation}
536 <
537 < In most cases, it is not easy to find the exact flow $\varphi_\tau$.
538 < Instead, we use a approximate map, $\psi_\tau$, which is usually
539 < called integrator. The order of an integrator $\psi_\tau$ is $p$, if
540 < the Taylor series of $\psi_\tau$ agree to order $p$,
536 > In most cases, it is not easy to find the exact propagator
537 > $\varphi_\tau$. Instead, we use an approximate map, $\psi_\tau$,
538 > which is usually called an integrator. The order of an integrator
539 > $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
540 > order $p$,
541   \begin{equation}
542 < \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
542 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
543   \end{equation}
544  
545   \subsection{\label{introSection:geometricProperties}Geometric Properties}
546  
547 < The hidden geometric properties of ODE and its flow play important
548 < roles in numerical studies. Many of them can be found in systems
549 < which occur naturally in applications.
550 <
551 < Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
635 < a \emph{symplectic} flow if it satisfies,
547 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
548 > ODE and its propagator play important roles in numerical studies.
549 > Many of them can be found in systems which occur naturally in
550 > applications. Let $\varphi$ be the propagator of Hamiltonian vector
551 > field, $\varphi$ is a \emph{symplectic} propagator if it satisfies,
552   \begin{equation}
553   {\varphi '}^T J \varphi ' = J.
554   \end{equation}
555   According to Liouville's theorem, the symplectic volume is invariant
556 < under a Hamiltonian flow, which is the basis for classical
557 < statistical mechanics. Furthermore, the flow of a Hamiltonian vector
558 < field on a symplectic manifold can be shown to be a
556 > under a Hamiltonian propagator, which is the basis for classical
557 > statistical mechanics. Furthermore, the propagator of a Hamiltonian
558 > vector field on a symplectic manifold can be shown to be a
559   symplectomorphism. As to the Poisson system,
560   \begin{equation}
561   {\varphi '}^T J \varphi ' = J \circ \varphi
562   \end{equation}
563 < is the property must be preserved by the integrator.
564 <
565 < It is possible to construct a \emph{volume-preserving} flow for a
566 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
567 < \det d\varphi  = 1$. One can show easily that a symplectic flow will
568 < be volume-preserving.
569 <
654 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
655 < will result in a new system,
563 > is the property that must be preserved by the integrator. It is
564 > possible to construct a \emph{volume-preserving} propagator for a
565 > source free ODE ($ \nabla \cdot f = 0 $), if the propagator
566 > satisfies $ \det d\varphi  = 1$. One can show easily that a
567 > symplectic propagator will be volume-preserving. Changing the
568 > variables $y = h(x)$ in an ODE (Eq.~\ref{introEquation:ODE}) will
569 > result in a new system,
570   \[
571   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
572   \]
573   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
574 < In other words, the flow of this vector field is reversible if and
575 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
576 <
577 < A \emph{first integral}, or conserved quantity of a general
578 < differential function is a function $ G:R^{2d}  \to R^d $ which is
665 < constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
574 > In other words, the propagator of this vector field is reversible if
575 > and only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
576 > conserved quantity of a general differential function is a function
577 > $ G:R^{2d}  \to R^d $ which is constant for all solutions of the ODE
578 > $\frac{{dx}}{{dt}} = f(x)$ ,
579   \[
580   \frac{{dG(x(t))}}{{dt}} = 0.
581   \]
582 < Using chain rule, one may obtain,
582 > Using the chain rule, one may obtain,
583   \[
584 < \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
584 > \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \cdot \nabla G,
585   \]
586 < which is the condition for conserving \emph{first integral}. For a
587 < canonical Hamiltonian system, the time evolution of an arbitrary
588 < smooth function $G$ is given by,
589 < \begin{equation}
590 < \begin{array}{c}
591 < \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}
586 > which is the condition for conserved quantities. For a canonical
587 > Hamiltonian system, the time evolution of an arbitrary smooth
588 > function $G$ is given by,
589 > \begin{eqnarray}
590 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \notag\\
591 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)).
592   \label{introEquation:firstIntegral1}
593 < \end{equation}
594 < Using poisson bracket notion, Equation
595 < \ref{introEquation:firstIntegral1} can be rewritten as
593 > \end{eqnarray}
594 > Using poisson bracket notion, Eq.~\ref{introEquation:firstIntegral1}
595 > can be rewritten as
596   \[
597   \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
598   \]
599 < Therefore, the sufficient condition for $G$ to be the \emph{first
600 < integral} of a Hamiltonian system is
601 < \[
602 < \left\{ {G,H} \right\} = 0.
603 < \]
604 < As well known, the Hamiltonian (or energy) H of a Hamiltonian system
605 < is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
695 < 0$.
599 > Therefore, the sufficient condition for $G$ to be a conserved
600 > quantity of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$ As
601 > is well known, the Hamiltonian (or energy) H of a Hamiltonian system
602 > is a conserved quantity, which is due to the fact $\{ H,H\}  = 0$.
603 > When designing any numerical methods, one should always try to
604 > preserve the structural properties of the original ODE and its
605 > propagator.
606  
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
607   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
608   A lot of well established and very effective numerical methods have
609 < been successful precisely because of their symplecticities even
609 > been successful precisely because of their symplectic nature even
610   though this fact was not recognized when they were first
611 < constructed. The most famous example is leapfrog methods in
612 < molecular dynamics. In general, symplectic integrators can be
611 > constructed. The most famous example is the Verlet-leapfrog method
612 > in molecular dynamics. In general, symplectic integrators can be
613   constructed using one of four different methods.
614   \begin{enumerate}
615   \item Generating functions
# Line 711 | Line 617 | constructed using one of four different methods.
617   \item Runge-Kutta methods
618   \item Splitting methods
619   \end{enumerate}
620 + Generating functions\cite{Channell1990} tend to lead to methods
621 + which are cumbersome and difficult to use. In dissipative systems,
622 + variational methods can capture the decay of energy
623 + accurately\cite{Kane2000}. Since they are geometrically unstable
624 + against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
625 + methods are not suitable for Hamiltonian system. Recently, various
626 + high-order explicit Runge-Kutta methods \cite{Owren1992,Chen2003}
627 + have been developed to overcome this instability. However, due to
628 + computational penalty involved in implementing the Runge-Kutta
629 + methods, they have not attracted much attention from the Molecular
630 + Dynamics community. Instead, splitting methods have been widely
631 + accepted since they exploit natural decompositions of the
632 + system\cite{Tuckerman1992, McLachlan1998}.
633  
634 < 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}.
634 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
635  
728 \subsubsection{\label{introSection:splittingMethod}Splitting Method}
729
636   The main idea behind splitting methods is to decompose the discrete
637 < $\varphi_h$ as a composition of simpler flows,
637 > $\varphi_h$ as a composition of simpler propagators,
638   \begin{equation}
639   \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
640   \varphi _{h_n }
641   \label{introEquation:FlowDecomposition}
642   \end{equation}
643 < where each of the sub-flow is chosen such that each represent a
644 < simpler integration of the system.
645 <
740 < Suppose that a Hamiltonian system takes the form,
643 > where each of the sub-propagator is chosen such that each represent
644 > a simpler integration of the system. Suppose that a Hamiltonian
645 > system takes the form,
646   \[
647   H = H_1 + H_2.
648   \]
649   Here, $H_1$ and $H_2$ may represent different physical processes of
650   the system. For instance, they may relate to kinetic and potential
651   energy respectively, which is a natural decomposition of the
652 < problem. If $H_1$ and $H_2$ can be integrated using exact flows
653 < $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
654 < order is then given by the Lie-Trotter formula
652 > problem. If $H_1$ and $H_2$ can be integrated using exact
653 > propagators $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a
654 > simple first order expression is then given by the Lie-Trotter
655 > formula
656   \begin{equation}
657   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
658   \label{introEquation:firstOrderSplitting}
# Line 755 | Line 661 | It is easy to show that any composition of symplectic
661   continuous $\varphi _i$ over a time $h$. By definition, as
662   $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
663   must follow that each operator $\varphi_i(t)$ is a symplectic map.
664 < It is easy to show that any composition of symplectic flows yields a
665 < symplectic map,
664 > It is easy to show that any composition of symplectic propagators
665 > yields a symplectic map,
666   \begin{equation}
667   (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
668   '\phi ' = \phi '^T J\phi ' = J,
# Line 764 | Line 670 | splitting in this context automatically generates a sy
670   \end{equation}
671   where $\phi$ and $\psi$ both are symplectic maps. Thus operator
672   splitting in this context automatically generates a symplectic map.
673 <
674 < The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
675 < introduces local errors proportional to $h^2$, while Strang
676 < splitting gives a second-order decomposition,
673 > The Lie-Trotter
674 > splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
675 > local errors proportional to $h^2$, while the Strang splitting gives
676 > a second-order decomposition,
677   \begin{equation}
678   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
679   _{1,h/2} , \label{introEquation:secondOrderSplitting}
680   \end{equation}
681 < which has a local error proportional to $h^3$. Sprang splitting's
682 < popularity in molecular simulation community attribute to its
683 < symmetric property,
681 > which has a local error proportional to $h^3$. The Strang
682 > splitting's popularity in molecular simulation community attribute
683 > to its symmetric property,
684   \begin{equation}
685   \varphi _h^{ - 1} = \varphi _{ - h}.
686   \label{introEquation:timeReversible}
687   \end{equation}
688  
689 < \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
689 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
690   The classical equation for a system consisting of interacting
691   particles can be written in Hamiltonian form,
692   \[
693   H = T + V
694   \]
695   where $T$ is the kinetic energy and $V$ is the potential energy.
696 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
696 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
697   obtains the following:
698   \begin{align}
699   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 800 | Line 706 | symplectic(\ref{introEquation:SymplecticFlowCompositio
706   \end{align}
707   where $F(t)$ is the force at time $t$. This integration scheme is
708   known as \emph{velocity verlet} which is
709 < symplectic(\ref{introEquation:SymplecticFlowComposition}),
710 < time-reversible(\ref{introEquation:timeReversible}) and
711 < volume-preserving (\ref{introEquation:volumePreserving}). These
709 > symplectic(Eq.~\ref{introEquation:SymplecticFlowComposition}),
710 > time-reversible(Eq.~\ref{introEquation:timeReversible}) and
711 > volume-preserving (Eq.~\ref{introEquation:volumePreserving}). These
712   geometric properties attribute to its long-time stability and its
713   popularity in the community. However, the most commonly used
714   velocity verlet integration scheme is written as below,
# Line 814 | Line 720 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
720      \label{introEquation:Lp9b}\\%
721   %
722   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
723 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
723 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
724   \end{align}
725   From the preceding splitting, one can see that the integration of
726   the equations of motion would follow:
# Line 823 | Line 729 | the equations of motion would follow:
729  
730   \item Use the half step velocities to move positions one whole step, $\Delta t$.
731  
732 < \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
732 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
733  
734   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
735   \end{enumerate}
736 <
737 < Simply switching the order of splitting and composing, a new
738 < integrator, the \emph{position verlet} integrator, can be generated,
736 > By simply switching the order of the propagators in the splitting
737 > and composing a new integrator, the \emph{position verlet}
738 > integrator, can be generated,
739   \begin{align}
740   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
741   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 837 | Line 743 | q(\Delta t)} \right]. %
743   %
744   q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
745   q(\Delta t)} \right]. %
746 < \label{introEquation:positionVerlet1}
746 > \label{introEquation:positionVerlet2}
747   \end{align}
748  
749 < \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
749 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
750  
751 < Baker-Campbell-Hausdorff formula can be used to determine the local
752 < error of splitting method in terms of commutator of the
753 < operators(\ref{introEquation:exponentialOperator}) associated with
754 < the sub-flow. For operators $hX$ and $hY$ which are associate to
755 < $\varphi_1(t)$ and $\varphi_2(t$ respectively , we have
751 > The Baker-Campbell-Hausdorff formula can be used to determine the
752 > local error of a splitting method in terms of the commutator of the
753 > operators(Eq.~\ref{introEquation:exponentialOperator}) associated with
754 > the sub-propagator. For operators $hX$ and $hY$ which are associated
755 > with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
756   \begin{equation}
757   \exp (hX + hY) = \exp (hZ)
758   \end{equation}
# Line 855 | Line 761 | Here, $[X,Y]$ is the commutators of operator $X$ and $
761   hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
762   {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
763   \end{equation}
764 < Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
764 > Here, $[X,Y]$ is the commutator of operator $X$ and $Y$ given by
765   \[
766   [X,Y] = XY - YX .
767   \]
768 < Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
769 < can obtain
768 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
769 > to the Strang splitting, we can obtain
770   \begin{eqnarray*}
771 < \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
772 < [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
773 < & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
774 < \ldots )
771 > \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 \\
772 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
773 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots
774 >                                   ).
775   \end{eqnarray*}
776 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
777 < error of Spring splitting is proportional to $h^3$. The same
778 < procedure can be applied to general splitting,  of the form
776 > Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local
777 > error of Strang splitting is proportional to $h^3$. The same
778 > procedure can be applied to a general splitting of the form
779   \begin{equation}
780   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
781   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
782   \end{equation}
783 < Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
784 < order method. Yoshida proposed an elegant way to compose higher
785 < order methods based on symmetric splitting. Given a symmetric second
786 < order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
787 < method can be constructed by composing,
783 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
784 > order methods. Yoshida proposed an elegant way to compose higher
785 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
786 > a symmetric second order base method $ \varphi _h^{(2)} $, a
787 > fourth-order symmetric method can be constructed by composing,
788   \[
789   \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
790   h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
# Line 888 | Line 794 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
794   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
795   \begin{equation}
796   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
797 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
797 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
798   \end{equation}
799 < , if the weights are chosen as
799 > if the weights are chosen as
800   \[
801   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
802   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 898 | Line 804 | As a special discipline of molecular modeling, Molecul
804  
805   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
806  
807 < As a special discipline of molecular modeling, Molecular dynamics
808 < has proven to be a powerful tool for studying the functions of
809 < biological systems, providing structural, thermodynamic and
810 < dynamical information.
807 > As one of the principal tools of molecular modeling, Molecular
808 > dynamics has proven to be a powerful tool for studying the functions
809 > of biological systems, providing structural, thermodynamic and
810 > dynamical information. The basic idea of molecular dynamics is that
811 > macroscopic properties are related to microscopic behavior and
812 > microscopic behavior can be calculated from the trajectories in
813 > simulations. For instance, instantaneous temperature of a
814 > Hamiltonian system of $N$ particles can be measured by
815 > \[
816 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
817 > \]
818 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
819 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
820 > the Boltzman constant.
821  
822 < \subsection{\label{introSec:mdInit}Initialization}
822 > A typical molecular dynamics run consists of three essential steps:
823 > \begin{enumerate}
824 >  \item Initialization
825 >    \begin{enumerate}
826 >    \item Preliminary preparation
827 >    \item Minimization
828 >    \item Heating
829 >    \item Equilibration
830 >    \end{enumerate}
831 >  \item Production
832 >  \item Analysis
833 > \end{enumerate}
834 > These three individual steps will be covered in the following
835 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
836 > initialization of a simulation. Sec.~\ref{introSection:production}
837 > discusses issues of production runs.
838 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
839 > analysis of trajectories.
840  
841 < \subsection{\label{introSec:forceEvaluation}Force Evaluation}
841 > \subsection{\label{introSec:initialSystemSettings}Initialization}
842  
843 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
843 > \subsubsection{\textbf{Preliminary preparation}}
844  
845 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
845 > When selecting the starting structure of a molecule for molecular
846 > simulation, one may retrieve its Cartesian coordinates from public
847 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
848 > thousands of crystal structures of molecules are discovered every
849 > year, many more remain unknown due to the difficulties of
850 > purification and crystallization. Even for molecules with known
851 > structures, some important information is missing. For example, a
852 > missing hydrogen atom which acts as donor in hydrogen bonding must
853 > be added. Moreover, in order to include electrostatic interactions,
854 > one may need to specify the partial charges for individual atoms.
855 > Under some circumstances, we may even need to prepare the system in
856 > a special configuration. For instance, when studying transport
857 > phenomenon in membrane systems, we may prepare the lipids in a
858 > bilayer structure instead of placing lipids randomly in solvent,
859 > since we are not interested in the slow self-aggregation process.
860  
861 < 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}}.
861 > \subsubsection{\textbf{Minimization}}
862  
863 < It is very important to develop stable and efficient methods to
864 < integrate the equations of motion of orientational degrees of
865 < freedom. Euler angles are the nature choice to describe the
866 < rotational degrees of freedom. However, due to its singularity, the
867 < numerical integration of corresponding equations of motion is very
868 < inefficient and inaccurate. Although an alternative integrator using
869 < different sets of Euler angles can overcome this difficulty\cite{},
870 < the computational penalty and the lost of angular momentum
871 < conservation still remain. A singularity free representation
872 < utilizing quaternions was developed by Evans in 1977. Unfortunately,
873 < this approach suffer from the nonseparable Hamiltonian resulted from
874 < quaternion representation, which prevents the symplectic algorithm
875 < to be utilized. Another different approach is to apply holonomic
876 < constraints to the atoms belonging to the rigid body. Each atom
877 < moves independently under the normal forces deriving from potential
878 < energy and constraint forces which are used to guarantee the
879 < rigidness. However, due to their iterative nature, SHAKE and Rattle
880 < algorithm converge very slowly when the number of constraint
881 < increases.
863 > It is quite possible that some of molecules in the system from
864 > preliminary preparation may be overlapping with each other. This
865 > close proximity leads to high initial potential energy which
866 > consequently jeopardizes any molecular dynamics simulations. To
867 > remove these steric overlaps, one typically performs energy
868 > minimization to find a more reasonable conformation. Several energy
869 > minimization methods have been developed to exploit the energy
870 > surface and to locate the local minimum. While converging slowly
871 > near the minimum, steepest descent method is extremely robust when
872 > systems are strongly anharmonic. Thus, it is often used to refine
873 > structures from crystallographic data. Relying on the Hessian,
874 > advanced methods like Newton-Raphson converge rapidly to a local
875 > minimum, but become unstable if the energy surface is far from
876 > quadratic. Another factor that must be taken into account, when
877 > choosing energy minimization method, is the size of the system.
878 > Steepest descent and conjugate gradient can deal with models of any
879 > size. Because of the limits on computer memory to store the hessian
880 > matrix and the computing power needed to diagonalize these matrices,
881 > most Newton-Raphson methods can not be used with very large systems.
882  
883 < 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{}.
883 > \subsubsection{\textbf{Heating}}
884  
885 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
885 > Typically, heating is performed by assigning random velocities
886 > according to a Maxwell-Boltzman distribution for a desired
887 > temperature. Beginning at a lower temperature and gradually
888 > increasing the temperature by assigning larger random velocities, we
889 > end up setting the temperature of the system to a final temperature
890 > at which the simulation will be conducted. In heating phase, we
891 > should also keep the system from drifting or rotating as a whole. To
892 > do this, the net linear momentum and angular momentum of the system
893 > is shifted to zero after each resampling from the Maxwell -Boltzman
894 > distribution.
895  
896 < \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
896 > \subsubsection{\textbf{Equilibration}}
897  
898 + The purpose of equilibration is to allow the system to evolve
899 + spontaneously for a period of time and reach equilibrium. The
900 + procedure is continued until various statistical properties, such as
901 + temperature, pressure, energy, volume and other structural
902 + properties \textit{etc}, become independent of time. Strictly
903 + speaking, minimization and heating are not necessary, provided the
904 + equilibration process is long enough. However, these steps can serve
905 + as a mean to arrive at an equilibrated structure in an effective
906 + way.
907 +
908 + \subsection{\label{introSection:production}Production}
909 +
910 + The production run is the most important step of the simulation, in
911 + which the equilibrated structure is used as a starting point and the
912 + motions of the molecules are collected for later analysis. In order
913 + to capture the macroscopic properties of the system, the molecular
914 + dynamics simulation must be performed by sampling correctly and
915 + efficiently from the relevant thermodynamic ensemble.
916 +
917 + The most expensive part of a molecular dynamics simulation is the
918 + calculation of non-bonded forces, such as van der Waals force and
919 + Coulombic forces \textit{etc}. For a system of $N$ particles, the
920 + complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
921 + which makes large simulations prohibitive in the absence of any
922 + algorithmic tricks. A natural approach to avoid system size issues
923 + is to represent the bulk behavior by a finite number of the
924 + particles. However, this approach will suffer from surface effects
925 + at the edges of the simulation. To offset this, \textit{Periodic
926 + boundary conditions} (see Fig.~\ref{introFig:pbc}) were developed to
927 + simulate bulk properties with a relatively small number of
928 + particles. In this method, the simulation box is replicated
929 + throughout space to form an infinite lattice. During the simulation,
930 + when a particle moves in the primary cell, its image in other cells
931 + move in exactly the same direction with exactly the same
932 + orientation. Thus, as a particle leaves the primary cell, one of its
933 + images will enter through the opposite face.
934 + \begin{figure}
935 + \centering
936 + \includegraphics[width=\linewidth]{pbc.eps}
937 + \caption[An illustration of periodic boundary conditions]{A 2-D
938 + illustration of periodic boundary conditions. As one particle leaves
939 + the left of the simulation box, an image of it enters the right.}
940 + \label{introFig:pbc}
941 + \end{figure}
942 +
943 + %cutoff and minimum image convention
944 + Another important technique to improve the efficiency of force
945 + evaluation is to apply spherical cutoffs where particles farther
946 + than a predetermined distance are not included in the calculation
947 + \cite{Frenkel1996}. The use of a cutoff radius will cause a
948 + discontinuity in the potential energy curve. Fortunately, one can
949 + shift a simple radial potential to ensure the potential curve go
950 + smoothly to zero at the cutoff radius. The cutoff strategy works
951 + well for Lennard-Jones interaction because of its short range
952 + nature. However, simply truncating the electrostatic interaction
953 + with the use of cutoffs has been shown to lead to severe artifacts
954 + in simulations. The Ewald summation, in which the slowly decaying
955 + Coulomb potential is transformed into direct and reciprocal sums
956 + with rapid and absolute convergence, has proved to minimize the
957 + periodicity artifacts in liquid simulations. Taking the advantages
958 + of the fast Fourier transform (FFT) for calculating discrete Fourier
959 + transforms, the particle mesh-based
960 + methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
961 + $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
962 + \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
963 + which treats Coulombic interactions exactly at short range, and
964 + approximate the potential at long range through multipolar
965 + expansion. In spite of their wide acceptance at the molecular
966 + simulation community, these two methods are difficult to implement
967 + correctly and efficiently. Instead, we use a damped and
968 + charge-neutralized Coulomb potential method developed by Wolf and
969 + his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
970 + particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
971   \begin{equation}
972 < H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
973 < V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
974 < \label{introEquation:RBHamiltonian}
972 > V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
973 > r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
974 > R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
975 > r_{ij})}{r_{ij}}\right\}, \label{introEquation:shiftedCoulomb}
976   \end{equation}
977 < Here, $q$ and $Q$  are the position and rotation matrix for the
978 < rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
979 < $J$, a diagonal matrix, is defined by
980 < \[
981 < I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
982 < \]
983 < where $I_{ii}$ is the diagonal element of the inertia tensor. This
984 < constrained Hamiltonian equation subjects to a holonomic constraint,
985 < \begin{equation}
986 < Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
973 < \end{equation}
974 < which is used to ensure rotation matrix's orthogonality.
975 < Differentiating \ref{introEquation:orthogonalConstraint} and using
976 < Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
977 < \begin{equation}
978 < Q^t PJ^{ - 1}  + J^{ - 1} P^t Q = 0 . \\
979 < \label{introEquation:RBFirstOrderConstraint}
980 < \end{equation}
977 > where $\alpha$ is the convergence parameter. Due to the lack of
978 > inherent periodicity and rapid convergence,this method is extremely
979 > efficient and easy to implement.
980 > \begin{figure}
981 > \centering
982 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
983 > \caption[An illustration of shifted Coulomb potential]{An
984 > illustration of shifted Coulomb potential.}
985 > \label{introFigure:shiftedCoulomb}
986 > \end{figure}
987  
988 < Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
983 < \ref{introEquation:motionHamiltonianMomentum}), one can write down
984 < the equations of motion,
985 < \[
986 < \begin{array}{c}
987 < \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
988 < \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
989 < \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
990 < \frac{{dP}}{{dt}} =  - \nabla _q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
991 < \end{array}
992 < \]
988 > %multiple time step
989  
990 + \subsection{\label{introSection:Analysis} Analysis}
991  
992 + Recently, advanced visualization techniques have been applied to
993 + monitor the motions of molecules. Although the dynamics of the
994 + system can be described qualitatively from animation, quantitative
995 + trajectory analysis is more useful. According to the principles of
996 + Statistical Mechanics in
997 + Sec.~\ref{introSection:statisticalMechanics}, one can compute
998 + thermodynamic properties, analyze fluctuations of structural
999 + parameters, and investigate time-dependent processes of the molecule
1000 + from the trajectories.
1001 +
1002 + \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1003 +
1004 + Thermodynamic properties, which can be expressed in terms of some
1005 + function of the coordinates and momenta of all particles in the
1006 + system, can be directly computed from molecular dynamics. The usual
1007 + way to measure the pressure is based on virial theorem of Clausius
1008 + which states that the virial is equal to $-3Nk_BT$. For a system
1009 + with forces between particles, the total virial, $W$, contains the
1010 + contribution from external pressure and interaction between the
1011 + particles:
1012   \[
1013 < M = \left\{ {(Q,P):Q^T Q = 1,Q^t PJ^{ - 1}  + J^{ - 1} P^t Q = 0}
1014 < \right\} .
1013 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1014 > f_{ij} } } \right\rangle
1015   \]
1016 + where $f_{ij}$ is the force between particle $i$ and $j$ at a
1017 + distance $r_{ij}$. Thus, the expression for the pressure is given
1018 + by:
1019 + \begin{equation}
1020 + P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1021 + < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1022 + \end{equation}
1023  
1024 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
1024 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1025  
1026 < \subsection{\label{introSection:symplecticDiscretizationRB}Symplectic Discretization of Euler Equations}
1026 > Structural Properties of a simple fluid can be described by a set of
1027 > distribution functions. Among these functions,the \emph{pair
1028 > distribution function}, also known as \emph{radial distribution
1029 > function}, is of most fundamental importance to liquid theory.
1030 > Experimentally, pair distribution functions can be gathered by
1031 > Fourier transforming raw data from a series of neutron diffraction
1032 > experiments and integrating over the surface factor
1033 > \cite{Powles1973}. The experimental results can serve as a criterion
1034 > to justify the correctness of a liquid model. Moreover, various
1035 > equilibrium thermodynamic and structural properties can also be
1036 > expressed in terms of the radial distribution function
1037 > \cite{Allen1987}. The pair distribution functions $g(r)$ gives the
1038 > probability that a particle $i$ will be located at a distance $r$
1039 > from a another particle $j$ in the system
1040 > \begin{equation}
1041 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1042 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1043 > (r)}{\rho}.
1044 > \end{equation}
1045 > Note that the delta function can be replaced by a histogram in
1046 > computer simulation. Peaks in $g(r)$ represent solvent shells, and
1047 > the height of these peaks gradually decreases to 1 as the liquid of
1048 > large distance approaches the bulk density.
1049  
1050  
1051 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1051 > \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1052 > Properties}}
1053  
1054 < \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
1055 <
1056 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
1010 <
1054 > Time-dependent properties are usually calculated using \emph{time
1055 > correlation functions}, which correlate random variables $A$ and $B$
1056 > at two different times,
1057   \begin{equation}
1058 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1059 < \label{introEquation:bathGLE}
1058 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1059 > \label{introEquation:timeCorrelationFunction}
1060   \end{equation}
1061 < where $H_B$ is harmonic bath Hamiltonian,
1061 > If $A$ and $B$ refer to same variable, this kind of correlation
1062 > functions are called \emph{autocorrelation functions}. One example
1063 > of auto correlation function is the velocity auto-correlation
1064 > function which is directly related to transport properties of
1065 > molecular liquids:
1066   \[
1067 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1068 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1067 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1068 > \right\rangle } dt
1069   \]
1070 < and $\Delta U$ is bilinear system-bath coupling,
1070 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1071 > function, which is averaged over time origins and over all the
1072 > atoms, the dipole autocorrelation functions is calculated for the
1073 > entire system. The dipole autocorrelation function is given by:
1074   \[
1075 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1075 > c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1076 > \right\rangle
1077   \]
1078 < Completing the square,
1078 > Here $u_{tot}$ is the net dipole of the entire system and is given
1079 > by
1080   \[
1081 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
1027 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1028 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1029 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
1030 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1081 > u_{tot} (t) = \sum\limits_i {u_i (t)}.
1082   \]
1083 < and putting it back into Eq.~\ref{introEquation:bathGLE},
1083 > In principle, many time correlation functions can be related to
1084 > Fourier transforms of the infrared, Raman, and inelastic neutron
1085 > scattering spectra of molecular liquids. In practice, one can
1086 > extract the IR spectrum from the intensity of the molecular dipole
1087 > fluctuation at each frequency using the following relationship:
1088   \[
1089 < H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1090 < {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1036 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1037 < w_\alpha ^2 }}x} \right)^2 } \right\}}
1089 > \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1090 > i2\pi vt} dt}.
1091   \]
1039 where
1040 \[
1041 W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1042 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1043 \]
1044 Since the first two terms of the new Hamiltonian depend only on the
1045 system coordinates, we can get the equations of motion for
1046 Generalized Langevin Dynamics by Hamilton's equations
1047 \ref{introEquation:motionHamiltonianCoordinate,
1048 introEquation:motionHamiltonianMomentum},
1049 \begin{align}
1050 \dot p &=  - \frac{{\partial H}}{{\partial x}}
1051       &= m\ddot x
1052       &= - \frac{{\partial W(x)}}{{\partial x}} - \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)}
1053 \label{introEquation:Lp5}
1054 \end{align}
1055 , and
1056 \begin{align}
1057 \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
1058                &= m\ddot x_\alpha
1059                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
1060 \end{align}
1092  
1093 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1093 > \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1094  
1095 < \[
1096 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1097 < \]
1095 > Rigid bodies are frequently involved in the modeling of different
1096 > areas, from engineering, physics, to chemistry. For example,
1097 > missiles and vehicles are usually modeled by rigid bodies.  The
1098 > movement of the objects in 3D gaming engines or other physics
1099 > simulators is governed by rigid body dynamics. In molecular
1100 > simulations, rigid bodies are used to simplify protein-protein
1101 > docking studies\cite{Gray2003}.
1102  
1103 < \[
1104 < L(x + y) = L(x) + L(y)
1105 < \]
1106 <
1107 < \[
1108 < L(ax) = aL(x)
1109 < \]
1110 <
1111 < \[
1112 < L(\dot x) = pL(x) - px(0)
1113 < \]
1114 <
1115 < \[
1116 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1117 < \]
1103 > It is very important to develop stable and efficient methods to
1104 > integrate the equations of motion for orientational degrees of
1105 > freedom. Euler angles are the natural choice to describe the
1106 > rotational degrees of freedom. However, due to $\frac {1}{sin
1107 > \theta}$ singularities, the numerical integration of corresponding
1108 > equations of these motion is very inefficient and inaccurate.
1109 > Although an alternative integrator using multiple sets of Euler
1110 > angles can overcome this difficulty\cite{Barojas1973}, the
1111 > computational penalty and the loss of angular momentum conservation
1112 > still remain. A singularity-free representation utilizing
1113 > quaternions was developed by Evans in 1977\cite{Evans1977}.
1114 > Unfortunately, this approach used a nonseparable Hamiltonian
1115 > resulting from the quaternion representation, which prevented the
1116 > symplectic algorithm from being utilized. Another different approach
1117 > is to apply holonomic constraints to the atoms belonging to the
1118 > rigid body. Each atom moves independently under the normal forces
1119 > deriving from potential energy and constraint forces which are used
1120 > to guarantee the rigidness. However, due to their iterative nature,
1121 > the SHAKE and Rattle algorithms also converge very slowly when the
1122 > number of constraints increases\cite{Ryckaert1977, Andersen1983}.
1123  
1124 + A break-through in geometric literature suggests that, in order to
1125 + develop a long-term integration scheme, one should preserve the
1126 + symplectic structure of the propagator. By introducing a conjugate
1127 + momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1128 + equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1129 + proposed to evolve the Hamiltonian system in a constraint manifold
1130 + by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1131 + An alternative method using the quaternion representation was
1132 + developed by Omelyan\cite{Omelyan1998}. However, both of these
1133 + methods are iterative and inefficient. In this section, we descibe a
1134 + symplectic Lie-Poisson integrator for rigid bodies developed by
1135 + Dullweber and his coworkers\cite{Dullweber1997} in depth.
1136 +
1137 + \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1138 + The motion of a rigid body is Hamiltonian with the Hamiltonian
1139 + function
1140 + \begin{equation}
1141 + H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1142 + V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1143 + \label{introEquation:RBHamiltonian}
1144 + \end{equation}
1145 + Here, $q$ and $Q$  are the position vector and rotation matrix for
1146 + the rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ ,
1147 + and $J$, a diagonal matrix, is defined by
1148   \[
1149 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1149 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1150   \]
1151 <
1152 < Some relatively important transformation,
1151 > where $I_{ii}$ is the diagonal element of the inertia tensor. This
1152 > constrained Hamiltonian equation is subjected to a holonomic
1153 > constraint,
1154 > \begin{equation}
1155 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1156 > \end{equation}
1157 > which is used to ensure the rotation matrix's unitarity. Using
1158 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} and Eq.~
1159 > \ref{introEquation:motionHamiltonianMomentum}, one can write down
1160 > the equations of motion,
1161 > \begin{eqnarray}
1162 > \frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\
1163 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q), \label{introEquation:RBMotionMomentum}\\
1164 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1},  \label{introEquation:RBMotionRotation}\\
1165 > \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1166 > \end{eqnarray}
1167 > Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and
1168 > using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1169 > \begin{equation}
1170 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1171 > \label{introEquation:RBFirstOrderConstraint}
1172 > \end{equation}
1173 > In general, there are two ways to satisfy the holonomic constraints.
1174 > We can use a constraint force provided by a Lagrange multiplier on
1175 > the normal manifold to keep the motion on the constraint space. Or
1176 > we can simply evolve the system on the constraint manifold. These
1177 > two methods have been proved to be equivalent. The holonomic
1178 > constraint and equations of motions define a constraint manifold for
1179 > rigid bodies
1180   \[
1181 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1181 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1182 > \right\}.
1183   \]
1184 <
1184 > Unfortunately, this constraint manifold is not $T^* SO(3)$ which is
1185 > a symplectic manifold on Lie rotation group $SO(3)$. However, it
1186 > turns out that under symplectic transformation, the cotangent space
1187 > and the phase space are diffeomorphic. By introducing
1188   \[
1189 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1189 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1190   \]
1191 <
1191 > the mechanical system subjected to a holonomic constraint manifold $M$
1192 > can be re-formulated as a Hamiltonian system on the cotangent space
1193   \[
1194 < L(1) = \frac{1}{p}
1194 > T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1195 > 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1196   \]
1197 <
1198 < First, the bath coordinates,
1197 > For a body fixed vector $X_i$ with respect to the center of mass of
1198 > the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1199 > given as
1200 > \begin{equation}
1201 > X_i^{lab} = Q X_i + q.
1202 > \end{equation}
1203 > Therefore, potential energy $V(q,Q)$ is defined by
1204   \[
1205 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
1104 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
1105 < }}L(x)
1205 > V(q,Q) = V(Q X_0 + q).
1206   \]
1207 + Hence, the force and torque are given by
1208   \[
1209 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1109 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1209 > \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1210   \]
1211 < Then, the system coordinates,
1212 < \begin{align}
1213 < mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1214 < \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1215 < }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1216 < (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1217 < }}\omega _\alpha ^2 L(x)} \right\}}
1218 < %
1219 < &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1220 < \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1221 < - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1222 < - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1223 < \end{align}
1224 < Then, the inverse transform,
1211 > and
1212 > \[
1213 > \nabla _Q V(q,Q) = F(q,Q)X_i^t
1214 > \]
1215 > respectively. As a common choice to describe the rotation dynamics
1216 > of the rigid body, the angular momentum on the body fixed frame $\Pi
1217 > = Q^t P$ is introduced to rewrite the equations of motion,
1218 > \begin{equation}
1219 > \begin{array}{l}
1220 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda,  \\
1221 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1},  \\
1222 > \end{array}
1223 > \label{introEqaution:RBMotionPI}
1224 > \end{equation}
1225 > as well as holonomic constraints $\Pi J^{ - 1}  + J^{ - 1} \Pi ^t  =
1226 > 0$ and $Q^T Q = 1$. For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a
1227 > matrix $\hat v \in so(3)^ \star$, the hat-map isomorphism,
1228 > \begin{equation}
1229 > v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1230 > {\begin{array}{*{20}c}
1231 >   0 & { - v_3 } & {v_2 }  \\
1232 >   {v_3 } & 0 & { - v_1 }  \\
1233 >   { - v_2 } & {v_1 } & 0  \\
1234 > \end{array}} \right),
1235 > \label{introEquation:hatmapIsomorphism}
1236 > \end{equation}
1237 > will let us associate the matrix products with traditional vector
1238 > operations
1239 > \[
1240 > \hat vu = v \times u.
1241 > \]
1242 > Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew
1243 > matrix,
1244 > \begin{eqnarray}
1245 > (\dot \Pi  - \dot \Pi ^T )&= &(\Pi  - \Pi ^T )(J^{ - 1} \Pi  + \Pi J^{ - 1} ) \notag \\
1246 > & & + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  -
1247 > (\Lambda  - \Lambda ^T ). \label{introEquation:skewMatrixPI}
1248 > \end{eqnarray}
1249 > Since $\Lambda$ is symmetric, the last term of
1250 > Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1251 > Lagrange multiplier $\Lambda$ is absent from the equations of
1252 > motion. This unique property eliminates the requirement of
1253 > iterations which can not be avoided in other methods\cite{Kol1997,
1254 > Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1255 > equation of motion for angular momentum in the body frame
1256 > \begin{equation}
1257 > \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1258 > F_i (r,Q)} \right) \times X_i }.
1259 > \label{introEquation:bodyAngularMotion}
1260 > \end{equation}
1261 > In the same manner, the equation of motion for rotation matrix is
1262 > given by
1263 > \[
1264 > \dot Q = Qskew(I^{ - 1} \pi ).
1265 > \]
1266  
1267 < \begin{align}
1268 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1128 < \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1129 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1130 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1131 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1132 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1133 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1134 < %
1135 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1136 < {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1137 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1138 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1139 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1140 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1141 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1142 < (\omega _\alpha  t)} \right\}}
1143 < \end{align}
1267 > \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1268 > Lie-Poisson Integrator for Free Rigid Bodies}
1269  
1270 + If there are no external forces exerted on the rigid body, the only
1271 + contribution to the rotational motion is from the kinetic energy
1272 + (the first term of \ref{introEquation:bodyAngularMotion}). The free
1273 + rigid body is an example of a Lie-Poisson system with Hamiltonian
1274 + function
1275   \begin{equation}
1276 < m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1277 < (t)\dot x(t - \tau )d\tau }  + R(t)
1148 < \label{introEuqation:GeneralizedLangevinDynamics}
1276 > T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1277 > \label{introEquation:rotationalKineticRB}
1278   \end{equation}
1279 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1280 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1281 < \[
1282 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1283 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1279 > where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1280 > Lie-Poisson structure matrix,
1281 > \begin{equation}
1282 > J(\pi ) = \left( {\begin{array}{*{20}c}
1283 >   0 & {\pi _3 } & { - \pi _2 }  \\
1284 >   { - \pi _3 } & 0 & {\pi _1 }  \\
1285 >   {\pi _2 } & { - \pi _1 } & 0  \\
1286 > \end{array}} \right).
1287 > \end{equation}
1288 > Thus, the dynamics of free rigid body is governed by
1289 > \begin{equation}
1290 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi ).
1291 > \end{equation}
1292 > One may notice that each $T_i^r$ in
1293 > Eq.~\ref{introEquation:rotationalKineticRB} can be solved exactly.
1294 > For instance, the equations of motion due to $T_1^r$ are given by
1295 > \begin{equation}
1296 > \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1297 > \label{introEqaution:RBMotionSingleTerm}
1298 > \end{equation}
1299 > with
1300 > \[ R_1  = \left( {\begin{array}{*{20}c}
1301 >   0 & 0 & 0  \\
1302 >   0 & 0 & {\pi _1 }  \\
1303 >   0 & { - \pi _1 } & 0  \\
1304 > \end{array}} \right).
1305   \]
1306 < For an infinite harmonic bath, we can use the spectral density and
1307 < an integral over frequencies.
1306 > The solutions of Eq.~\ref{introEqaution:RBMotionSingleTerm} is
1307 > \[
1308 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1309 > Q(0)e^{\Delta tR_1 }
1310 > \]
1311 > with
1312 > \[
1313 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1314 >   0 & 0 & 0  \\
1315 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1316 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1317 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1318 > \]
1319 > To reduce the cost of computing expensive functions in $e^{\Delta
1320 > tR_1 }$, we can use the Cayley transformation to obtain a
1321 > single-aixs propagator,
1322 > \begin{eqnarray*}
1323 > e^{\Delta tR_1 }  & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta
1324 > tR_1 ) \\
1325 > %
1326 > & \approx & \left( \begin{array}{ccc}
1327 > 1 & 0 & 0 \\
1328 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1329 > \theta^2 / 4} \\
1330 > 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1331 > \theta^2 / 4}
1332 > \end{array}
1333 > \right).
1334 > \end{eqnarray*}
1335 > The propagators for $T_2^r$ and $T_3^r$ can be found in the same
1336 > manner. In order to construct a second-order symplectic method, we
1337 > split the angular kinetic Hamiltonian function into five terms
1338 > \[
1339 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1340 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1341 > (\pi _1 ).
1342 > \]
1343 > By concatenating the propagators corresponding to these five terms,
1344 > we can obtain an symplectic integrator,
1345 > \[
1346 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1347 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1348 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1349 > _1 }.
1350 > \]
1351 > The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1352 > $F(\pi )$ and $G(\pi )$ is defined by
1353 > \[
1354 > \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1355 > ).
1356 > \]
1357 > If the Poisson bracket of a function $F$ with an arbitrary smooth
1358 > function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1359 > conserved quantity in Poisson system. We can easily verify that the
1360 > norm of the angular momentum, $\parallel \pi
1361 > \parallel$, is a \emph{Casimir}\cite{McLachlan1993}. Let$ F(\pi ) = S(\frac{{\parallel
1362 > \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1363 > then by the chain rule
1364 > \[
1365 > \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1366 > }}{2})\pi.
1367 > \]
1368 > Thus, $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel
1369 > \pi
1370 > \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1371 > Lie-Poisson integrator is found to be both extremely efficient and
1372 > stable. These properties can be explained by the fact the small
1373 > angle approximation is used and the norm of the angular momentum is
1374 > conserved.
1375  
1376 + \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1377 + Splitting for Rigid Body}
1378 +
1379 + The Hamiltonian of rigid body can be separated in terms of kinetic
1380 + energy and potential energy, $H = T(p,\pi ) + V(q,Q)$. The equations
1381 + of motion corresponding to potential energy and kinetic energy are
1382 + listed in Table~\ref{introTable:rbEquations}
1383 + \begin{table}
1384 + \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1385 + \label{introTable:rbEquations}
1386 + \begin{center}
1387 + \begin{tabular}{|l|l|}
1388 +  \hline
1389 +  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1390 +  Potential & Kinetic \\
1391 +  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1392 +  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1393 +  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1394 +  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1395 +  \hline
1396 + \end{tabular}
1397 + \end{center}
1398 + \end{table}
1399 + A second-order symplectic method is now obtained by the composition
1400 + of the position and velocity propagators,
1401   \[
1402 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1403 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1162 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1163 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1402 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1403 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1404   \]
1405 < The random forces depend only on initial conditions.
1405 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1406 > sub-propagators which corresponding to force and torque
1407 > respectively,
1408 > \[
1409 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1410 > _{\Delta t/2,\tau }.
1411 > \]
1412 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1413 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1414 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1415 > kinetic energy can be separated to translational kinetic term, $T^t
1416 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1417 > \begin{equation}
1418 > T(p,\pi ) =T^t (p) + T^r (\pi ).
1419 > \end{equation}
1420 > where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1421 > defined by Eq.~\ref{introEquation:rotationalKineticRB}. Therefore,
1422 > the corresponding propagators are given by
1423 > \[
1424 > \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1425 > _{\Delta t,T^r }.
1426 > \]
1427 > Finally, we obtain the overall symplectic propagators for freely
1428 > moving rigid bodies
1429 > \begin{eqnarray}
1430 > \varphi _{\Delta t}  &=& \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \notag\\
1431 >  & & \circ \varphi _{\Delta t,T^t }  \circ \varphi _{\Delta t/2,\pi _1 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi _1 }  \notag\\
1432 >  & & \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .
1433 > \label{introEquation:overallRBFlowMaps}
1434 > \end{eqnarray}
1435  
1436 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1437 < So we can define a new set of coordinates,
1436 > \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1437 > As an alternative to newtonian dynamics, Langevin dynamics, which
1438 > mimics a simple heat bath with stochastic and dissipative forces,
1439 > has been applied in a variety of studies. This section will review
1440 > the theory of Langevin dynamics. A brief derivation of generalized
1441 > Langevin equation will be given first. Following that, we will
1442 > discuss the physical meaning of the terms appearing in the equation
1443 > as well as the calculation of friction tensor from hydrodynamics
1444 > theory.
1445 >
1446 > \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1447 >
1448 > A harmonic bath model, in which an effective set of harmonic
1449 > oscillators are used to mimic the effect of a linearly responding
1450 > environment, has been widely used in quantum chemistry and
1451 > statistical mechanics. One of the successful applications of
1452 > Harmonic bath model is the derivation of the Generalized Langevin
1453 > Dynamics (GLE). Lets consider a system, in which the degree of
1454 > freedom $x$ is assumed to couple to the bath linearly, giving a
1455 > Hamiltonian of the form
1456 > \begin{equation}
1457 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1458 > \label{introEquation:bathGLE}.
1459 > \end{equation}
1460 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1461 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1462   \[
1463 < q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1464 < ^2 }}x(0)
1463 > H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1464 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1465 > \right\}}
1466   \]
1467 < This makes
1467 > where the index $\alpha$ runs over all the bath degrees of freedom,
1468 > $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1469 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1470 > coupling,
1471   \[
1472 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1472 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1473   \]
1474 < And since the $q$ coordinates are harmonic oscillators,
1474 > where $g_\alpha$ are the coupling constants between the bath
1475 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1476 > Introducing
1477   \[
1478 < \begin{array}{l}
1479 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1480 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1478 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1479 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1480 > \]
1481 > and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as
1482 > \[
1483 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1484 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1485 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1486 > w_\alpha ^2 }}x} \right)^2 } \right\}}.
1487 > \]
1488 > Since the first two terms of the new Hamiltonian depend only on the
1489 > system coordinates, we can get the equations of motion for
1490 > Generalized Langevin Dynamics by Hamilton's equations,
1491 > \begin{equation}
1492 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1493 > \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1494 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1495 > \label{introEquation:coorMotionGLE}
1496 > \end{equation}
1497 > and
1498 > \begin{equation}
1499 > m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1500 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1501 > \label{introEquation:bathMotionGLE}
1502 > \end{equation}
1503 > In order to derive an equation for $x$, the dynamics of the bath
1504 > variables $x_\alpha$ must be solved exactly first. As an integral
1505 > transform which is particularly useful in solving linear ordinary
1506 > differential equations,the Laplace transform is the appropriate tool
1507 > to solve this problem. The basic idea is to transform the difficult
1508 > differential equations into simple algebra problems which can be
1509 > solved easily. Then, by applying the inverse Laplace transform, we
1510 > can retrieve the solutions of the original problems. Let $f(t)$ be a
1511 > function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$
1512 > is a new function defined as
1513 > \[
1514 > L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1515 > \]
1516 > where  $p$ is real and  $L$ is called the Laplace Transform
1517 > Operator. Below are some important properties of Laplace transform
1518 > \begin{eqnarray*}
1519 > L(x + y)  & = & L(x) + L(y) \\
1520 > L(ax)     & = & aL(x) \\
1521 > L(\dot x) & = & pL(x) - px(0) \\
1522 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1523 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1524 > \end{eqnarray*}
1525 > Applying the Laplace transform to the bath coordinates, we obtain
1526 > \begin{eqnarray*}
1527 > 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), \\
1528 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}. \\
1529 > \end{eqnarray*}
1530 > In the same way, the system coordinates become
1531 > \begin{eqnarray*}
1532 > mL(\ddot x) & = &
1533 >  - \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\}}  \\
1534 >  & & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}.
1535 > \end{eqnarray*}
1536 > With the help of some relatively important inverse Laplace
1537 > transformations:
1538 > \[
1539 > \begin{array}{c}
1540 > L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1541 > L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1542 > L(1) = \frac{1}{p} \\
1543   \end{array}
1544   \]
1545 <
1546 < \begin{align}
1547 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1548 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1549 < (t)q_\beta  (0)} \right\rangle } }
1545 > we obtain
1546 > \begin{eqnarray*}
1547 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1548 > \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1549 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1550 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1551 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1552 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1553 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1554 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}\\
1555   %
1556 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1557 < \right\rangle \cos (\omega _\alpha  t)}
1558 < %
1559 < &= kT\xi (t)
1560 < \end{align}
1561 <
1556 > & = & -
1557 > \frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha
1558 > = 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha
1559 > ^2 }}} \right)\cos (\omega _\alpha
1560 > t)\dot x(t - \tau )d} \tau }  \\
1561 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1562 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1563 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1564 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1565 > \end{eqnarray*}
1566 > Introducing a \emph{dynamic friction kernel}
1567   \begin{equation}
1568 < \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1569 < \label{introEquation:secondFluctuationDissipation}
1568 > \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1569 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1570 > \label{introEquation:dynamicFrictionKernelDefinition}
1571   \end{equation}
1572 + and \emph{a random force}
1573 + \begin{equation}
1574 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1575 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1576 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1577 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1578 + \label{introEquation:randomForceDefinition}
1579 + \end{equation}
1580 + the equation of motion can be rewritten as
1581 + \begin{equation}
1582 + m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1583 + (t)\dot x(t - \tau )d\tau }  + R(t)
1584 + \label{introEuqation:GeneralizedLangevinDynamics}
1585 + \end{equation}
1586 + which is known as the \emph{generalized Langevin equation}.
1587  
1588 < \section{\label{introSection:hydroynamics}Hydrodynamics}
1588 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1589  
1590 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
1591 < \subsection{\label{introSection:analyticalApproach}Analytical
1592 < Approach}
1590 > One may notice that $R(t)$ depends only on initial conditions, which
1591 > implies it is completely deterministic within the context of a
1592 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1593 > uncorrelated to $x$ and $\dot x$,$\left\langle {x(t)R(t)}
1594 > \right\rangle  = 0, \left\langle {\dot x(t)R(t)} \right\rangle  =
1595 > 0.$ This property is what we expect from a truly random process. As
1596 > long as the model chosen for $R(t)$ was a gaussian distribution in
1597 > general, the stochastic nature of the GLE still remains.
1598 > %dynamic friction kernel
1599 > The convolution integral
1600 > \[
1601 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1602 > \]
1603 > depends on the entire history of the evolution of $x$, which implies
1604 > that the bath retains memory of previous motions. In other words,
1605 > the bath requires a finite time to respond to change in the motion
1606 > of the system. For a sluggish bath which responds slowly to changes
1607 > in the system coordinate, we may regard $\xi(t)$ as a constant
1608 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1609 > \[
1610 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1611 > \]
1612 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1613 > \[
1614 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1615 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1616 > \]
1617 > which can be used to describe the effect of dynamic caging in
1618 > viscous solvents. The other extreme is the bath that responds
1619 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1620 > taken as a $delta$ function in time:
1621 > \[
1622 > \xi (t) = 2\xi _0 \delta (t)
1623 > \]
1624 > Hence, the convolution integral becomes
1625 > \[
1626 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1627 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1628 > \]
1629 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1630 > \begin{equation}
1631 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1632 > x(t) + R(t) \label{introEquation:LangevinEquation}
1633 > \end{equation}
1634 > which is known as the Langevin equation. The static friction
1635 > coefficient $\xi _0$ can either be calculated from spectral density
1636 > or be determined by Stokes' law for regular shaped particles. A
1637 > brief review on calculating friction tensors for arbitrary shaped
1638 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1639  
1640 < \subsection{\label{introSection:approximationApproach}Approximation
1208 < Approach}
1640 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1641  
1642 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1643 < Body}
1644 <
1645 < \section{\label{introSection:correlationFunctions}Correlation Functions}
1642 > Defining a new set of coordinates
1643 > \[
1644 > q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1645 > ^2 }}x(0),
1646 > \]
1647 > we can rewrite $R(T)$ as
1648 > \[
1649 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1650 > \]
1651 > And since the $q$ coordinates are harmonic oscillators,
1652 > \begin{eqnarray*}
1653 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1654 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1655 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1656 > \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 } }  \\
1657 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1658 >  & = &kT\xi (t)
1659 > \end{eqnarray*}
1660 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1661 > \begin{equation}
1662 > \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1663 > \label{introEquation:secondFluctuationDissipation},
1664 > \end{equation}
1665 > which acts as a constraint on the possible ways in which one can
1666 > model the random force and friction kernel.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines