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 2705 by tim, Wed Apr 12 21:20:13 2006 UTC vs.
Revision 2904 by tim, Wed Jun 28 17:36:32 2006 UTC

# Line 6 | Line 6 | behind classical mechanics. Firstly, One can determine
6   Closely related to Classical Mechanics, Molecular Dynamics
7   simulations are carried out by integrating the equations of motion
8   for a given system of particles. There are three fundamental ideas
9 < behind classical mechanics. Firstly, One can determine the state of
9 > behind classical mechanics. Firstly, one can determine the state of
10   a mechanical system at any time of interest; Secondly, all the
11   mechanical properties of the system at that time can be determined
12   by combining the knowledge of the properties of the system with the
# Line 17 | Line 17 | Newton¡¯s first law defines a class of inertial frames
17   \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18   The discovery of Newton's three laws of mechanics which govern the
19   motion of particles is the foundation of the classical mechanics.
20 < Newton¡¯s first law defines a class of inertial frames. Inertial
20 > Newton's first law defines a class of inertial frames. Inertial
21   frames are reference frames where a particle not interacting with
22   other bodies will move with constant speed in the same direction.
23 < With respect to inertial frames Newton¡¯s second law has the form
23 > With respect to inertial frames, Newton's second law has the form
24   \begin{equation}
25 < F = \frac {dp}{dt} = \frac {mv}{dt}
25 > F = \frac {dp}{dt} = \frac {mdv}{dt}
26   \label{introEquation:newtonSecondLaw}
27   \end{equation}
28   A point mass interacting with other bodies moves with the
29   acceleration along the direction of the force acting on it. Let
30   $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
31   $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
32 < Newton¡¯s third law states that
32 > Newton's third law states that
33   \begin{equation}
34 < F_{ij} = -F_{ji}
34 > F_{ij} = -F_{ji}.
35   \label{introEquation:newtonThirdLaw}
36   \end{equation}
37
37   Conservation laws of Newtonian Mechanics play very important roles
38   in solving mechanics problems. The linear momentum of a particle is
39   conserved if it is free or it experiences no force. The second
# Line 46 | Line 45 | N \equiv r \times F \label{introEquation:torqueDefinit
45   \end{equation}
46   The torque $\tau$ with respect to the same origin is defined to be
47   \begin{equation}
48 < N \equiv r \times F \label{introEquation:torqueDefinition}
48 > \tau \equiv r \times F \label{introEquation:torqueDefinition}
49   \end{equation}
50   Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
51   \[
# Line 59 | Line 58 | thus,
58   \]
59   thus,
60   \begin{equation}
61 < \dot L = r \times \dot p = N
61 > \dot L = r \times \dot p = \tau
62   \end{equation}
63   If there are no external torques acting on a body, the angular
64   momentum of it is conserved. The last conservation theorem state
65 < that if all forces are conservative, Energy
66 < \begin{equation}E = T + V \label{introEquation:energyConservation}
65 > that if all forces are conservative, energy is conserved,
66 > \begin{equation}E = T + V. \label{introEquation:energyConservation}
67   \end{equation}
68 < is conserved. All of these conserved quantities are
69 < important factors to determine the quality of numerical integration
70 < scheme for rigid body \cite{Dullweber1997}.
68 > All of these conserved quantities are important factors to determine
69 > the quality of numerical integration schemes for rigid bodies
70 > \cite{Dullweber1997}.
71  
72   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
73  
74 < Newtonian Mechanics suffers from two important limitations: it
75 < describes their motion in special cartesian coordinate systems.
76 < Another limitation of Newtonian mechanics becomes obvious when we
77 < try to describe systems with large numbers of particles. It becomes
78 < very difficult to predict the properties of the system by carrying
79 < out calculations involving the each individual interaction between
80 < 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.
74 > Newtonian Mechanics suffers from a important limitation: motions can
75 > only be described in cartesian coordinate systems which make it
76 > impossible to predict analytically the properties of the system even
77 > if we know all of the details of the interaction. In order to
78 > overcome some of the practical difficulties which arise in attempts
79 > to apply Newton's equation to complex system, approximate numerical
80 > procedures may be developed.
81  
82 < \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
83 < Principle}
82 > \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
83 > Principle}}
84  
85   Hamilton introduced the dynamical principle upon which it is
86 < possible to base all of mechanics and, indeed, most of classical
87 < physics. Hamilton's Principle may be stated as follow,
88 <
89 < The actual trajectory, along which a dynamical system may move from
90 < one point to another within a specified time, is derived by finding
91 < the path which minimizes the time integral of the difference between
96 < the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
86 > possible to base all of mechanics and most of classical physics.
87 > Hamilton's Principle may be stated as follows: the actual
88 > trajectory, along which a dynamical system may move from one point
89 > to another within a specified time, is derived by finding the path
90 > which minimizes the time integral of the difference between the
91 > kinetic, $K$, and potential energies, $U$,
92   \begin{equation}
93 < \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
93 > \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0}.
94   \label{introEquation:halmitonianPrinciple1}
95   \end{equation}
101
96   For simple mechanical systems, where the forces acting on the
97 < different part are derivable from a potential and the velocities are
98 < small compared with that of light, the Lagrangian function $L$ can
99 < be define as the difference between the kinetic energy of the system
106 < and its potential energy,
97 > different parts are derivable from a potential, the Lagrangian
98 > function $L$ can be defined as the difference between the kinetic
99 > energy of the system and its potential energy,
100   \begin{equation}
101 < L \equiv K - U = L(q_i ,\dot q_i ) ,
101 > L \equiv K - U = L(q_i ,\dot q_i ).
102   \label{introEquation:lagrangianDef}
103   \end{equation}
104 < then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
104 > Thus, Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
105   \begin{equation}
106 < \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
106 > \delta \int_{t_1 }^{t_2 } {L dt = 0} .
107   \label{introEquation:halmitonianPrinciple2}
108   \end{equation}
109  
110 < \subsubsection{\label{introSection:equationOfMotionLagrangian}The
111 < Equations of Motion in Lagrangian Mechanics}
110 > \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
111 > Equations of Motion in Lagrangian Mechanics}}
112  
113 < For a holonomic system of $f$ degrees of freedom, the equations of
114 < motion in the Lagrangian form is
113 > For a system of $f$ degrees of freedom, the equations of motion in
114 > the Lagrangian form is
115   \begin{equation}
116   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
117   \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
# Line 132 | Line 125 | independent of generalized velocities, the generalized
125   Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
126   introduced by William Rowan Hamilton in 1833 as a re-formulation of
127   classical mechanics. If the potential energy of a system is
128 < independent of generalized velocities, the generalized momenta can
136 < be defined as
128 > independent of velocities, the momenta can be defined as
129   \begin{equation}
130   p_i = \frac{\partial L}{\partial \dot q_i}
131   \label{introEquation:generalizedMomenta}
# Line 143 | Line 135 | p_i  = \frac{{\partial L}}{{\partial q_i }}
135   p_i  = \frac{{\partial L}}{{\partial q_i }}
136   \label{introEquation:generalizedMomentaDot}
137   \end{equation}
146
138   With the help of the generalized momenta, we may now define a new
139   quantity $H$ by the equation
140   \begin{equation}
# Line 151 | Line 142 | $L$ is the Lagrangian function for the system.
142   \label{introEquation:hamiltonianDefByLagrangian}
143   \end{equation}
144   where $ \dot q_1  \ldots \dot q_f $ are generalized velocities and
145 < $L$ is the Lagrangian function for the system.
146 <
156 < Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
157 < one can obtain
145 > $L$ is the Lagrangian function for the system. Differentiating
146 > Eq.~\ref{introEquation:hamiltonianDefByLagrangian}, one can obtain
147   \begin{equation}
148   dH = \sum\limits_k {\left( {p_k d\dot q_k  + \dot q_k dp_k  -
149   \frac{{\partial L}}{{\partial q_k }}dq_k  - \frac{{\partial
150   L}}{{\partial \dot q_k }}d\dot q_k } \right)}  - \frac{{\partial
151 < L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
151 > L}}{{\partial t}}dt . \label{introEquation:diffHamiltonian1}
152   \end{equation}
153 < Making use of  Eq.~\ref{introEquation:generalizedMomenta}, the
154 < second and fourth terms in the parentheses cancel. Therefore,
153 > Making use of Eq.~\ref{introEquation:generalizedMomenta}, the second
154 > and fourth terms in the parentheses cancel. Therefore,
155   Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
156   \begin{equation}
157   dH = \sum\limits_k {\left( {\dot q_k dp_k  - \dot p_k dq_k }
158 < \right)}  - \frac{{\partial L}}{{\partial t}}dt
158 > \right)}  - \frac{{\partial L}}{{\partial t}}dt .
159   \label{introEquation:diffHamiltonian2}
160   \end{equation}
161   By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
162   find
163   \begin{equation}
164 < \frac{{\partial H}}{{\partial p_k }} = q_k
164 > \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
165   \label{introEquation:motionHamiltonianCoordinate}
166   \end{equation}
167   \begin{equation}
168 < \frac{{\partial H}}{{\partial q_k }} =  - p_k
168 > \frac{{\partial H}}{{\partial q_k }} =  - \dot {p_k}
169   \label{introEquation:motionHamiltonianMomentum}
170   \end{equation}
171   and
# Line 185 | Line 174 | t}}
174   t}}
175   \label{introEquation:motionHamiltonianTime}
176   \end{equation}
177 <
189 < Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
177 > where Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
178   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
179   equation of motion. Due to their symmetrical formula, they are also
180 < known as the canonical equations of motions \cite{Goldstein01}.
180 > known as the canonical equations of motions \cite{Goldstein2001}.
181  
182   An important difference between Lagrangian approach and the
183   Hamiltonian approach is that the Lagrangian is considered to be a
184 < function of the generalized velocities $\dot q_i$ and the
185 < generalized coordinates $q_i$, while the Hamiltonian is considered
186 < to be a function of the generalized momenta $p_i$ and the conjugate
187 < generalized coordinate $q_i$. Hamiltonian Mechanics is more
188 < appropriate for application to statistical mechanics and quantum
189 < mechanics, since it treats the coordinate and its time derivative as
190 < independent variables and it only works with 1st-order differential
203 < equations\cite{Marion90}.
204 <
184 > function of the generalized velocities $\dot q_i$ and coordinates
185 > $q_i$, while the Hamiltonian is considered to be a function of the
186 > generalized momenta $p_i$ and the conjugate coordinates $q_i$.
187 > Hamiltonian Mechanics is more appropriate for application to
188 > statistical mechanics and quantum mechanics, since it treats the
189 > coordinate and its time derivative as independent variables and it
190 > only works with 1st-order differential equations\cite{Marion1990}.
191   In Newtonian Mechanics, a system described by conservative forces
192 < conserves the total energy \ref{introEquation:energyConservation}.
193 < It follows that Hamilton's equations of motion conserve the total
194 < Hamiltonian.
192 > conserves the total energy
193 > (Eq.~\ref{introEquation:energyConservation}). It follows that
194 > Hamilton's equations of motion conserve the total Hamiltonian
195   \begin{equation}
196   \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
197   H}}{{\partial q_i }}\dot q_i  + \frac{{\partial H}}{{\partial p_i
198   }}\dot p_i } \right)}  = \sum\limits_i {\left( {\frac{{\partial
199   H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
200   \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
201 < q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
201 > q_i }}} \right) = 0}. \label{introEquation:conserveHalmitonian}
202   \end{equation}
203  
204   \section{\label{introSection:statisticalMechanics}Statistical
# Line 230 | Line 216 | momentum variables. Consider a dynamic system in a car
216   possible states. Each possible state of the system corresponds to
217   one unique point in the phase space. For mechanical systems, the
218   phase space usually consists of all possible values of position and
219 < momentum variables. Consider a dynamic system in a cartesian space,
220 < where each of the $6f$ coordinates and momenta is assigned to one of
221 < $6f$ mutually orthogonal axes, the phase space of this system is a
222 < $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
223 < \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
219 > momentum variables. Consider a dynamic system of $f$ particles in a
220 > cartesian space, where each of the $6f$ coordinates and momenta is
221 > assigned to one of $6f$ mutually orthogonal axes, the phase space of
222 > this system is a $6f$ dimensional space. A point, $x =
223 > (\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
224 > \over q} _1 , \ldots
225 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
226 > \over q} _f
227 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
228 > \over p} _1  \ldots
229 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
230 > \over p} _f )$ , with a unique set of values of $6f$ coordinates and
231   momenta is a phase space vector.
232 + %%%fix me
233  
234 < 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
234 > In statistical mechanics, the condition of an ensemble at any time
235   can be regarded as appropriately specified by the density $\rho$
236   with which representative points are distributed over the phase
237 < space. The density of distribution for an ensemble with $f$ degrees
238 < of freedom is defined as,
237 > space. The density distribution for an ensemble with $f$ degrees of
238 > freedom is defined as,
239   \begin{equation}
240   \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
241   \label{introEquation:densityDistribution}
242   \end{equation}
243   Governed by the principles of mechanics, the phase points change
244 < their value which would change the density at any time at phase
245 < space. Hence, the density of distribution is also to be taken as a
244 > their locations which would change the density at any time at phase
245 > space. Hence, the density distribution is also to be taken as a
246   function of the time.
247  
248   The number of systems $\delta N$ at time $t$ can be determined by,
# Line 273 | Line 250 | Assuming a large enough population of systems are expl
250   \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
251   \label{introEquation:deltaN}
252   \end{equation}
253 < Assuming a large enough population of systems are exploited, we can
254 < sufficiently approximate $\delta N$ without introducing
255 < discontinuity when we go from one region in the phase space to
256 < another. By integrating over the whole phase space,
253 > Assuming a large enough population of systems, we can sufficiently
254 > approximate $\delta N$ without introducing discontinuity when we go
255 > from one region in the phase space to another. By integrating over
256 > the whole phase space,
257   \begin{equation}
258   N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
259   \label{introEquation:totalNumberSystem}
# Line 288 | Line 265 | With the help of Equation(\ref{introEquation:unitProba
265   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
266   \label{introEquation:unitProbability}
267   \end{equation}
268 < With the help of Equation(\ref{introEquation:unitProbability}) and
269 < the knowledge of the system, it is possible to calculate the average
268 > With the help of Eq.~\ref{introEquation:unitProbability} and the
269 > knowledge of the system, it is possible to calculate the average
270   value of any desired quantity which depends on the coordinates and
271   momenta of the system. Even when the dynamics of the real system is
272   complex, or stochastic, or even discontinuous, the average
273 < properties of the ensemble of possibilities as a whole may still
274 < remain well defined. For a classical system in thermal equilibrium
275 < with its environment, the ensemble average of a mechanical quantity,
276 < $\langle A(q , p) \rangle_t$, takes the form of an integral over the
277 < phase space of the system,
273 > properties of the ensemble of possibilities as a whole remaining
274 > well defined. For a classical system in thermal equilibrium with its
275 > environment, the ensemble average of a mechanical quantity, $\langle
276 > A(q , p) \rangle_t$, takes the form of an integral over the phase
277 > space of the system,
278   \begin{equation}
279   \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
280   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
# Line 307 | Line 284 | parameters, such as temperature \textit{etc}, partitio
284  
285   There are several different types of ensembles with different
286   statistical characteristics. As a function of macroscopic
287 < parameters, such as temperature \textit{etc}, partition function can
288 < be used to describe the statistical properties of a system in
289 < thermodynamic equilibrium.
290 <
291 < 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,
287 > parameters, such as temperature \textit{etc}, the partition function
288 > can be used to describe the statistical properties of a system in
289 > thermodynamic equilibrium. As an ensemble of systems, each of which
290 > is known to be thermally isolated and conserve energy, the
291 > Microcanonical ensemble (NVE) has a partition function like,
292   \begin{equation}
293 < \Omega (N,V,E) = e^{\beta TS}
319 < \label{introEqaution:NVEPartition}.
293 > \Omega (N,V,E) = e^{\beta TS}. \label{introEquation:NVEPartition}
294   \end{equation}
295 < A canonical ensemble(NVT)is an ensemble of systems, each of which
295 > A canonical ensemble (NVT)is an ensemble of systems, each of which
296   can share its energy with a large heat reservoir. The distribution
297   of the total energy amongst the possible dynamical states is given
298   by the partition function,
299   \begin{equation}
300 < \Omega (N,V,T) = e^{ - \beta A}
300 > \Omega (N,V,T) = e^{ - \beta A}.
301   \label{introEquation:NVTPartition}
302   \end{equation}
303   Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
304 < TS$. Since most experiment are carried out under constant pressure
305 < condition, isothermal-isobaric ensemble(NPT) play a very important
306 < role in molecular simulation. The isothermal-isobaric ensemble allow
307 < the system to exchange energy with a heat bath of temperature $T$
308 < and to change the volume as well. Its partition function is given as
304 > TS$. Since most experiments are carried out under constant pressure
305 > condition, the isothermal-isobaric ensemble (NPT) plays a very
306 > important role in molecular simulations. The isothermal-isobaric
307 > ensemble allow the system to exchange energy with a heat bath of
308 > temperature $T$ and to change the volume as well. Its partition
309 > function is given as
310   \begin{equation}
311   \Delta (N,P,T) =  - e^{\beta G}.
312   \label{introEquation:NPTPartition}
# Line 340 | Line 315 | The Liouville's theorem is the foundation on which sta
315  
316   \subsection{\label{introSection:liouville}Liouville's theorem}
317  
318 < The Liouville's theorem is the foundation on which statistical
319 < mechanics rests. It describes the time evolution of phase space
318 > Liouville's theorem is the foundation on which statistical mechanics
319 > rests. It describes the time evolution of the phase space
320   distribution function. In order to calculate the rate of change of
321 < $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
322 < consider the two faces perpendicular to the $q_1$ axis, which are
323 < located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
324 < leaving the opposite face is given by the expression,
321 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
322 > the two faces perpendicular to the $q_1$ axis, which are located at
323 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
324 > opposite face is given by the expression,
325   \begin{equation}
326   \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
327   \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
# Line 370 | Line 345 | divining $ \delta q_1  \ldots \delta q_f \delta p_1  \
345   + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
346   \end{equation}
347   which cancels the first terms of the right hand side. Furthermore,
348 < divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
348 > dividing $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
349   p_f $ in both sides, we can write out Liouville's theorem in a
350   simple form,
351   \begin{equation}
# Line 379 | Line 354 | simple form,
354   \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
355   \label{introEquation:liouvilleTheorem}
356   \end{equation}
382
357   Liouville's theorem states that the distribution function is
358   constant along any trajectory in phase space. In classical
359 < statistical mechanics, since the number of particles in the system
360 < is huge, we may be able to believe the system is stationary,
359 > statistical mechanics, since the number of members in an ensemble is
360 > huge and constant, we can assume the local density has no reason
361 > (other than classical mechanics) to change,
362   \begin{equation}
363   \frac{{\partial \rho }}{{\partial t}} = 0.
364   \label{introEquation:stationary}
# Line 396 | Line 371 | distribution,
371   \label{introEquation:densityAndHamiltonian}
372   \end{equation}
373  
374 < \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
374 > \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
375   Lets consider a region in the phase space,
376   \begin{equation}
377   \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
378   \end{equation}
379   If this region is small enough, the density $\rho$ can be regarded
380 < as uniform over the whole phase space. Thus, the number of phase
381 < points inside this region is given by,
380 > as uniform over the whole integral. Thus, the number of phase points
381 > inside this region is given by,
382   \begin{equation}
383   \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
384   dp_1 } ..dp_f.
# Line 415 | Line 390 | With the help of stationary assumption
390   \end{equation}
391   With the help of stationary assumption
392   (\ref{introEquation:stationary}), we obtain the principle of the
393 < \emph{conservation of extension in phase space},
393 > \emph{conservation of volume in phase space},
394   \begin{equation}
395   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
396   ...dq_f dp_1 } ..dp_f  = 0.
397   \label{introEquation:volumePreserving}
398   \end{equation}
399  
400 < \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
400 > \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
401  
402   Liouville's theorem can be expresses in a variety of different forms
403   which are convenient within different contexts. For any two function
# Line 436 | Line 411 | Substituting equations of motion in Hamiltonian formal
411   \label{introEquation:poissonBracket}
412   \end{equation}
413   Substituting equations of motion in Hamiltonian formalism(
414 < \ref{introEquation:motionHamiltonianCoordinate} ,
415 < \ref{introEquation:motionHamiltonianMomentum} ) into
416 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
417 < theorem using Poisson bracket notion,
414 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
415 > Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
416 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
417 > Liouville's theorem using Poisson bracket notion,
418   \begin{equation}
419   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
420   {\rho ,H} \right\}.
# Line 464 | Line 439 | simulation and the quality of the underlying model. Ho
439   Various thermodynamic properties can be calculated from Molecular
440   Dynamics simulation. By comparing experimental values with the
441   calculated properties, one can determine the accuracy of the
442 < simulation and the quality of the underlying model. However, both of
443 < experiment and computer simulation are usually performed during a
442 > simulation and the quality of the underlying model. However, both
443 > experiments and computer simulations are usually performed during a
444   certain time interval and the measurements are averaged over a
445   period of them which is different from the average behavior of
446 < many-body system in Statistical Mechanics. Fortunately, Ergodic
447 < Hypothesis is proposed to make a connection between time average and
448 < ensemble average. It states that time average and average over the
449 < statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
446 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
447 > Hypothesis makes a connection between time average and the ensemble
448 > average. It states that the time average and average over the
449 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
450   \begin{equation}
451   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
452   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 485 | Line 460 | reasonable, the Monte Carlo techniques\cite{metropolis
460   a properly weighted statistical average. This allows the researcher
461   freedom of choice when deciding how best to measure a given
462   observable. In case an ensemble averaged approach sounds most
463 < reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
463 > reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
464   utilized. Or if the system lends itself to a time averaging
465   approach, the Molecular Dynamics techniques in
466   Sec.~\ref{introSection:molecularDynamics} will be the best
467   choice\cite{Frenkel1996}.
468  
469   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
470 < A variety of numerical integrators were proposed to simulate the
471 < motions. They usually begin with an initial conditionals and move
472 < the objects in the direction governed by the differential equations.
473 < However, most of them ignore the hidden physical law contained
474 < within the equations. Since 1990, geometric integrators, which
475 < preserve various phase-flow invariants such as symplectic structure,
476 < volume and time reversal symmetry, are developed to address this
477 < issue. The velocity verlet method, which happens to be a simple
478 < example of symplectic integrator, continues to gain its popularity
479 < in molecular dynamics community. This fact can be partly explained
480 < by its geometric nature.
470 > A variety of numerical integrators have been proposed to simulate
471 > the motions of atoms in MD simulation. They usually begin with
472 > initial conditionals and move the objects in the direction governed
473 > by the differential equations. However, most of them ignore the
474 > hidden physical laws contained within the equations. Since 1990,
475 > geometric integrators, which preserve various phase-flow invariants
476 > such as symplectic structure, volume and time reversal symmetry, are
477 > developed to address this issue\cite{Dullweber1997, McLachlan1998,
478 > Leimkuhler1999}. The velocity Verlet method, which happens to be a
479 > simple example of symplectic integrator, continues to gain
480 > popularity in the molecular dynamics community. This fact can be
481 > partly explained by its geometric nature.
482  
483 < \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
484 < A \emph{manifold} is an abstract mathematical space. It locally
485 < looks like Euclidean space, but when viewed globally, it may have
486 < more complicate structure. A good example of manifold is the surface
487 < of Earth. It seems to be flat locally, but it is round if viewed as
488 < a whole. A \emph{differentiable manifold} (also known as
489 < \emph{smooth manifold}) is a manifold with an open cover in which
490 < the covering neighborhoods are all smoothly isomorphic to one
491 < another. In other words,it is possible to apply calculus on
516 < \emph{differentiable manifold}. A \emph{symplectic manifold} is
517 < defined as a pair $(M, \omega)$ which consisting of a
483 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
484 > A \emph{manifold} is an abstract mathematical space. It looks
485 > locally like Euclidean space, but when viewed globally, it may have
486 > more complicated structure. A good example of manifold is the
487 > surface of Earth. It seems to be flat locally, but it is round if
488 > viewed as a whole. A \emph{differentiable manifold} (also known as
489 > \emph{smooth manifold}) is a manifold on which it is possible to
490 > apply calculus on \emph{differentiable manifold}. A \emph{symplectic
491 > manifold} is defined as a pair $(M, \omega)$ which consists of a
492   \emph{differentiable manifold} $M$ and a close, non-degenerated,
493   bilinear symplectic form, $\omega$. A symplectic form on a vector
494   space $V$ is a function $\omega(x, y)$ which satisfies
495   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
496   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
497 < $\omega(x, x) = 0$. Cross product operation in vector field is an
498 < example of symplectic form.
497 > $\omega(x, x) = 0$. The cross product operation in vector field is
498 > an example of symplectic form. One of the motivations to study
499 > \emph{symplectic manifolds} in Hamiltonian Mechanics is that a
500 > symplectic manifold can represent all possible configurations of the
501 > system and the phase space of the system can be described by it's
502 > cotangent bundle. Every symplectic manifold is even dimensional. For
503 > instance, in Hamilton equations, coordinate and momentum always
504 > appear in pairs.
505  
526 One of the motivations to study \emph{symplectic manifold} in
527 Hamiltonian Mechanics is that a symplectic manifold can represent
528 all possible configurations of the system and the phase space of the
529 system can be described by it's cotangent bundle. Every symplectic
530 manifold is even dimensional. For instance, in Hamilton equations,
531 coordinate and momentum always appear in pairs.
532
533 Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
534 \[
535 f : M \rightarrow N
536 \]
537 is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
538 the \emph{pullback} of $\eta$ under f is equal to $\omega$.
539 Canonical transformation is an example of symplectomorphism in
540 classical mechanics.
541
506   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
507  
508 < For a ordinary differential system defined as
508 > For an ordinary differential system defined as
509   \begin{equation}
510   \dot x = f(x)
511   \end{equation}
512 < where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
512 > where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
513   \begin{equation}
514   f(r) = J\nabla _x H(r).
515   \end{equation}
# Line 564 | Line 528 | called a \emph{Hamiltonian vector field}.
528   \frac{d}{{dt}}x = J\nabla _x H(x)
529   \label{introEquation:compactHamiltonian}
530   \end{equation}In this case, $f$ is
531 < called a \emph{Hamiltonian vector field}.
532 <
569 < Another generalization of Hamiltonian dynamics is Poisson Dynamics,
531 > called a \emph{Hamiltonian vector field}. Another generalization of
532 > Hamiltonian dynamics is Poisson Dynamics\cite{Olver1986},
533   \begin{equation}
534   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
535   \end{equation}
536   The most obvious change being that matrix $J$ now depends on $x$.
574 The free rigid body is an example of Poisson system (actually a
575 Lie-Poisson system) with Hamiltonian function of angular kinetic
576 energy.
577 \begin{equation}
578 J(\pi ) = \left( {\begin{array}{*{20}c}
579   0 & {\pi _3 } & { - \pi _2 }  \\
580   { - \pi _3 } & 0 & {\pi _1 }  \\
581   {\pi _2 } & { - \pi _1 } & 0  \\
582 \end{array}} \right)
583 \end{equation}
537  
585 \begin{equation}
586 H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
587 }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
588 \end{equation}
589
538   \subsection{\label{introSection:exactFlow}Exact Flow}
539  
540   Let $x(t)$ be the exact solution of the ODE system,
# Line 619 | Line 567 | Instead, we use a approximate map, $\psi_\tau$, which
567   \end{equation}
568  
569   In most cases, it is not easy to find the exact flow $\varphi_\tau$.
570 < Instead, we use a approximate map, $\psi_\tau$, which is usually
570 > Instead, we use an approximate map, $\psi_\tau$, which is usually
571   called integrator. The order of an integrator $\psi_\tau$ is $p$, if
572   the Taylor series of $\psi_\tau$ agree to order $p$,
573   \begin{equation}
574 < \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
574 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
575   \end{equation}
576  
577   \subsection{\label{introSection:geometricProperties}Geometric Properties}
578  
579 < The hidden geometric properties of ODE and its flow play important
580 < roles in numerical studies. Many of them can be found in systems
581 < which occur naturally in applications.
634 <
579 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
580 > ODE and its flow play important roles in numerical studies. Many of
581 > them can be found in systems which occur naturally in applications.
582   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
583   a \emph{symplectic} flow if it satisfies,
584   \begin{equation}
# Line 645 | Line 592 | is the property must be preserved by the integrator.
592   \begin{equation}
593   {\varphi '}^T J \varphi ' = J \circ \varphi
594   \end{equation}
595 < is the property must be preserved by the integrator.
596 <
597 < It is possible to construct a \emph{volume-preserving} flow for a
598 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
599 < \det d\varphi  = 1$. One can show easily that a symplectic flow will
600 < be volume-preserving.
654 <
655 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
656 < will result in a new system,
595 > is the property that must be preserved by the integrator. It is
596 > possible to construct a \emph{volume-preserving} flow for a source
597 > free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $ \det
598 > d\varphi  = 1$. One can show easily that a symplectic flow will be
599 > volume-preserving. Changing the variables $y = h(x)$ in an ODE
600 > (Eq.~\ref{introEquation:ODE}) will result in a new system,
601   \[
602   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
603   \]
604   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
605   In other words, the flow of this vector field is reversible if and
606 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
607 <
664 < A \emph{first integral}, or conserved quantity of a general
606 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
607 > \emph{first integral}, or conserved quantity of a general
608   differential function is a function $ G:R^{2d}  \to R^d $ which is
609   constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
610   \[
# Line 674 | Line 617 | smooth function $G$ is given by,
617   which is the condition for conserving \emph{first integral}. For a
618   canonical Hamiltonian system, the time evolution of an arbitrary
619   smooth function $G$ is given by,
620 < \begin{equation}
621 < \begin{array}{c}
622 < \frac{{dG(x(t))}}{{dt}} = [\nabla _x G(x(t))]^T \dot x(t) \\
680 <  = [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
681 < \end{array}
620 > \begin{eqnarray}
621 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
622 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
623   \label{introEquation:firstIntegral1}
624 < \end{equation}
624 > \end{eqnarray}
625   Using poisson bracket notion, Equation
626   \ref{introEquation:firstIntegral1} can be rewritten as
627   \[
# Line 693 | Line 634 | is a \emph{first integral}, which is due to the fact $
634   \]
635   As well known, the Hamiltonian (or energy) H of a Hamiltonian system
636   is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
637 < 0$.
697 <
698 <
699 < When designing any numerical methods, one should always try to
637 > 0$. When designing any numerical methods, one should always try to
638   preserve the structural properties of the original ODE and its flow.
639  
640   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
641   A lot of well established and very effective numerical methods have
642   been successful precisely because of their symplecticities even
643   though this fact was not recognized when they were first
644 < constructed. The most famous example is leapfrog methods in
645 < molecular dynamics. In general, symplectic integrators can be
644 > constructed. The most famous example is the Verlet-leapfrog method
645 > in molecular dynamics. In general, symplectic integrators can be
646   constructed using one of four different methods.
647   \begin{enumerate}
648   \item Generating functions
# Line 713 | Line 651 | Generating function tends to lead to methods which are
651   \item Splitting methods
652   \end{enumerate}
653  
654 < Generating function tends to lead to methods which are cumbersome
655 < and difficult to use. In dissipative systems, variational methods
656 < can capture the decay of energy accurately. Since their
657 < geometrically unstable nature against non-Hamiltonian perturbations,
658 < ordinary implicit Runge-Kutta methods are not suitable for
659 < Hamiltonian system. Recently, various high-order explicit
660 < Runge--Kutta methods have been developed to overcome this
654 > Generating function\cite{Channell1990} tends to lead to methods
655 > which are cumbersome and difficult to use. In dissipative systems,
656 > variational methods can capture the decay of energy
657 > accurately\cite{Kane2000}. Since their geometrically unstable nature
658 > against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
659 > methods are not suitable for Hamiltonian system. Recently, various
660 > high-order explicit Runge-Kutta methods
661 > \cite{Owren1992,Chen2003}have been developed to overcome this
662   instability. However, due to computational penalty involved in
663 < implementing the Runge-Kutta methods, they do not attract too much
664 < attention from Molecular Dynamics community. Instead, splitting have
665 < been widely accepted since they exploit natural decompositions of
666 < the system\cite{Tuckerman92}.
663 > implementing the Runge-Kutta methods, they have not attracted much
664 > attention from the Molecular Dynamics community. Instead, splitting
665 > methods have been widely accepted since they exploit natural
666 > decompositions of the system\cite{Tuckerman1992, McLachlan1998}.
667  
668 < \subsubsection{\label{introSection:splittingMethod}Splitting Method}
668 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
669  
670   The main idea behind splitting methods is to decompose the discrete
671   $\varphi_h$ as a composition of simpler flows,
# Line 736 | Line 675 | simpler integration of the system.
675   \label{introEquation:FlowDecomposition}
676   \end{equation}
677   where each of the sub-flow is chosen such that each represent a
678 < simpler integration of the system.
679 <
741 < Suppose that a Hamiltonian system takes the form,
678 > simpler integration of the system. Suppose that a Hamiltonian system
679 > takes the form,
680   \[
681   H = H_1 + H_2.
682   \]
# Line 747 | Line 685 | order is then given by the Lie-Trotter formula
685   energy respectively, which is a natural decomposition of the
686   problem. If $H_1$ and $H_2$ can be integrated using exact flows
687   $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
688 < order is then given by the Lie-Trotter formula
688 > order expression is then given by the Lie-Trotter formula
689   \begin{equation}
690   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
691   \label{introEquation:firstOrderSplitting}
# Line 771 | Line 709 | _{1,h/2} ,
709   splitting gives a second-order decomposition,
710   \begin{equation}
711   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
712 < _{1,h/2} ,
775 < \label{introEqaution:secondOrderSplitting}
712 > _{1,h/2} , \label{introEquation:secondOrderSplitting}
713   \end{equation}
714 < which has a local error proportional to $h^3$. Sprang splitting's
715 < popularity in molecular simulation community attribute to its
716 < symmetric property,
714 > which has a local error proportional to $h^3$. The Sprang
715 > splitting's popularity in molecular simulation community attribute
716 > to its symmetric property,
717   \begin{equation}
718   \varphi _h^{ - 1} = \varphi _{ - h}.
719   \label{introEquation:timeReversible}
720   \end{equation}
721  
722 < \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
722 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
723   The classical equation for a system consisting of interacting
724   particles can be written in Hamiltonian form,
725   \[
726   H = T + V
727   \]
728   where $T$ is the kinetic energy and $V$ is the potential energy.
729 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
729 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
730   obtains the following:
731   \begin{align}
732   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 816 | Line 753 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
753      \label{introEquation:Lp9b}\\%
754   %
755   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
756 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
756 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
757   \end{align}
758   From the preceding splitting, one can see that the integration of
759   the equations of motion would follow:
# Line 825 | Line 762 | the equations of motion would follow:
762  
763   \item Use the half step velocities to move positions one whole step, $\Delta t$.
764  
765 < \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
765 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
766  
767   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
768   \end{enumerate}
769 <
770 < Simply switching the order of splitting and composing, a new
771 < integrator, the \emph{position verlet} integrator, can be generated,
769 > By simply switching the order of the propagators in the splitting
770 > and composing a new integrator, the \emph{position verlet}
771 > integrator, can be generated,
772   \begin{align}
773   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
774   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 839 | Line 776 | q(\Delta t)} \right]. %
776   %
777   q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
778   q(\Delta t)} \right]. %
779 < \label{introEquation:positionVerlet1}
779 > \label{introEquation:positionVerlet2}
780   \end{align}
781  
782 < \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
782 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
783  
784 < Baker-Campbell-Hausdorff formula can be used to determine the local
785 < error of splitting method in terms of commutator of the
784 > The Baker-Campbell-Hausdorff formula can be used to determine the
785 > local error of splitting method in terms of the commutator of the
786   operators(\ref{introEquation:exponentialOperator}) associated with
787 < the sub-flow. For operators $hX$ and $hY$ which are associate to
788 < $\varphi_1(t)$ and $\varphi_2(t$ respectively , we have
787 > the sub-flow. For operators $hX$ and $hY$ which are associated with
788 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
789   \begin{equation}
790   \exp (hX + hY) = \exp (hZ)
791   \end{equation}
# Line 861 | Line 798 | Applying Baker-Campbell-Hausdorff formula to Sprang sp
798   \[
799   [X,Y] = XY - YX .
800   \]
801 < Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
802 < can obtain
801 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
802 > to the Sprang splitting, we can obtain
803   \begin{eqnarray*}
804 < \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
805 < [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
806 < & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
807 < \ldots )
808 < \end{eqnarray*}
872 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
804 > \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 \\
805 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
806 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
807 > \end{eqnarray*}
808 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
809   error of Spring splitting is proportional to $h^3$. The same
810 < procedure can be applied to general splitting,  of the form
810 > procedure can be applied to a general splitting,  of the form
811   \begin{equation}
812   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
813   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
814   \end{equation}
815 < Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
816 < order method. Yoshida proposed an elegant way to compose higher
817 < order methods based on symmetric splitting. Given a symmetric second
818 < order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
819 < method can be constructed by composing,
815 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
816 > order methods. Yoshida proposed an elegant way to compose higher
817 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
818 > a symmetric second order base method $ \varphi _h^{(2)} $, a
819 > fourth-order symmetric method can be constructed by composing,
820   \[
821   \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
822   h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
# Line 890 | Line 826 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
826   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
827   \begin{equation}
828   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
829 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
829 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
830   \end{equation}
831 < , if the weights are chosen as
831 > if the weights are chosen as
832   \[
833   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
834   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 900 | Line 836 | As a special discipline of molecular modeling, Molecul
836  
837   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
838  
839 < As a special discipline of molecular modeling, Molecular dynamics
840 < has proven to be a powerful tool for studying the functions of
841 < biological systems, providing structural, thermodynamic and
842 < dynamical information.
839 > As one of the principal tools of molecular modeling, Molecular
840 > dynamics has proven to be a powerful tool for studying the functions
841 > of biological systems, providing structural, thermodynamic and
842 > dynamical information. The basic idea of molecular dynamics is that
843 > macroscopic properties are related to microscopic behavior and
844 > microscopic behavior can be calculated from the trajectories in
845 > simulations. For instance, instantaneous temperature of an
846 > Hamiltonian system of $N$ particle can be measured by
847 > \[
848 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
849 > \]
850 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
851 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
852 > the boltzman constant.
853  
854 < \subsection{\label{introSec:mdInit}Initialization}
854 > A typical molecular dynamics run consists of three essential steps:
855 > \begin{enumerate}
856 >  \item Initialization
857 >    \begin{enumerate}
858 >    \item Preliminary preparation
859 >    \item Minimization
860 >    \item Heating
861 >    \item Equilibration
862 >    \end{enumerate}
863 >  \item Production
864 >  \item Analysis
865 > \end{enumerate}
866 > These three individual steps will be covered in the following
867 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
868 > initialization of a simulation. Sec.~\ref{introSection:production}
869 > will discusse issues in production run.
870 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
871 > trajectory analysis.
872  
873 < \subsection{\label{introSec:forceEvaluation}Force Evaluation}
873 > \subsection{\label{introSec:initialSystemSettings}Initialization}
874  
875 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
875 > \subsubsection{\textbf{Preliminary preparation}}
876 >
877 > When selecting the starting structure of a molecule for molecular
878 > simulation, one may retrieve its Cartesian coordinates from public
879 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
880 > thousands of crystal structures of molecules are discovered every
881 > year, many more remain unknown due to the difficulties of
882 > purification and crystallization. Even for molecules with known
883 > structure, some important information is missing. For example, a
884 > missing hydrogen atom which acts as donor in hydrogen bonding must
885 > be added. Moreover, in order to include electrostatic interaction,
886 > one may need to specify the partial charges for individual atoms.
887 > Under some circumstances, we may even need to prepare the system in
888 > a special configuration. For instance, when studying transport
889 > phenomenon in membrane systems, we may prepare the lipids in a
890 > bilayer structure instead of placing lipids randomly in solvent,
891 > since we are not interested in the slow self-aggregation process.
892 >
893 > \subsubsection{\textbf{Minimization}}
894 >
895 > It is quite possible that some of molecules in the system from
896 > preliminary preparation may be overlapping with each other. This
897 > close proximity leads to high initial potential energy which
898 > consequently jeopardizes any molecular dynamics simulations. To
899 > remove these steric overlaps, one typically performs energy
900 > minimization to find a more reasonable conformation. Several energy
901 > minimization methods have been developed to exploit the energy
902 > surface and to locate the local minimum. While converging slowly
903 > near the minimum, steepest descent method is extremely robust when
904 > systems are strongly anharmonic. Thus, it is often used to refine
905 > structure from crystallographic data. Relied on the gradient or
906 > hessian, advanced methods like Newton-Raphson converge rapidly to a
907 > local minimum, but become unstable if the energy surface is far from
908 > quadratic. Another factor that must be taken into account, when
909 > choosing energy minimization method, is the size of the system.
910 > Steepest descent and conjugate gradient can deal with models of any
911 > size. Because of the limits on computer memory to store the hessian
912 > matrix and the computing power needed to diagonalized these
913 > matrices, most Newton-Raphson methods can not be used with very
914 > large systems.
915 >
916 > \subsubsection{\textbf{Heating}}
917 >
918 > Typically, Heating is performed by assigning random velocities
919 > according to a Maxwell-Boltzman distribution for a desired
920 > temperature. Beginning at a lower temperature and gradually
921 > increasing the temperature by assigning larger random velocities, we
922 > end up with setting the temperature of the system to a final
923 > temperature at which the simulation will be conducted. In heating
924 > phase, we should also keep the system from drifting or rotating as a
925 > whole. To do this, the net linear momentum and angular momentum of
926 > the system is shifted to zero after each resampling from the Maxwell
927 > -Boltzman distribution.
928 >
929 > \subsubsection{\textbf{Equilibration}}
930 >
931 > The purpose of equilibration is to allow the system to evolve
932 > spontaneously for a period of time and reach equilibrium. The
933 > procedure is continued until various statistical properties, such as
934 > temperature, pressure, energy, volume and other structural
935 > properties \textit{etc}, become independent of time. Strictly
936 > speaking, minimization and heating are not necessary, provided the
937 > equilibration process is long enough. However, these steps can serve
938 > as a means to arrive at an equilibrated structure in an effective
939 > way.
940 >
941 > \subsection{\label{introSection:production}Production}
942 >
943 > The production run is the most important step of the simulation, in
944 > which the equilibrated structure is used as a starting point and the
945 > motions of the molecules are collected for later analysis. In order
946 > to capture the macroscopic properties of the system, the molecular
947 > dynamics simulation must be performed by sampling correctly and
948 > efficiently from the relevant thermodynamic ensemble.
949 >
950 > The most expensive part of a molecular dynamics simulation is the
951 > calculation of non-bonded forces, such as van der Waals force and
952 > Coulombic forces \textit{etc}. For a system of $N$ particles, the
953 > complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
954 > which making large simulations prohibitive in the absence of any
955 > algorithmic tricks.
956 >
957 > A natural approach to avoid system size issues is to represent the
958 > bulk behavior by a finite number of the particles. However, this
959 > approach will suffer from the surface effect at the edges of the
960 > simulation. To offset this, \textit{Periodic boundary conditions}
961 > (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
962 > properties with a relatively small number of particles. In this
963 > method, the simulation box is replicated throughout space to form an
964 > infinite lattice. During the simulation, when a particle moves in
965 > the primary cell, its image in other cells move in exactly the same
966 > direction with exactly the same orientation. Thus, as a particle
967 > leaves the primary cell, one of its images will enter through the
968 > opposite face.
969 > \begin{figure}
970 > \centering
971 > \includegraphics[width=\linewidth]{pbc.eps}
972 > \caption[An illustration of periodic boundary conditions]{A 2-D
973 > illustration of periodic boundary conditions. As one particle leaves
974 > the left of the simulation box, an image of it enters the right.}
975 > \label{introFig:pbc}
976 > \end{figure}
977 >
978 > %cutoff and minimum image convention
979 > Another important technique to improve the efficiency of force
980 > evaluation is to apply spherical cutoff where particles farther than
981 > a predetermined distance are not included in the calculation
982 > \cite{Frenkel1996}. The use of a cutoff radius will cause a
983 > discontinuity in the potential energy curve. Fortunately, one can
984 > shift simple radial potential to ensure the potential curve go
985 > smoothly to zero at the cutoff radius. The cutoff strategy works
986 > well for Lennard-Jones interaction because of its short range
987 > nature. However, simply truncating the electrostatic interaction
988 > with the use of cutoffs has been shown to lead to severe artifacts
989 > in simulations. The Ewald summation, in which the slowly decaying
990 > Coulomb potential is transformed into direct and reciprocal sums
991 > with rapid and absolute convergence, has proved to minimize the
992 > periodicity artifacts in liquid simulations. Taking the advantages
993 > of the fast Fourier transform (FFT) for calculating discrete Fourier
994 > transforms, the particle mesh-based
995 > methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
996 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
997 > \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
998 > which treats Coulombic interactions exactly at short range, and
999 > approximate the potential at long range through multipolar
1000 > expansion. In spite of their wide acceptance at the molecular
1001 > simulation community, these two methods are difficult to implement
1002 > correctly and efficiently. Instead, we use a damped and
1003 > charge-neutralized Coulomb potential method developed by Wolf and
1004 > his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
1005 > particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1006 > \begin{equation}
1007 > V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1008 > r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1009 > R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1010 > r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1011 > \end{equation}
1012 > where $\alpha$ is the convergence parameter. Due to the lack of
1013 > inherent periodicity and rapid convergence,this method is extremely
1014 > efficient and easy to implement.
1015 > \begin{figure}
1016 > \centering
1017 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1018 > \caption[An illustration of shifted Coulomb potential]{An
1019 > illustration of shifted Coulomb potential.}
1020 > \label{introFigure:shiftedCoulomb}
1021 > \end{figure}
1022 >
1023 > %multiple time step
1024 >
1025 > \subsection{\label{introSection:Analysis} Analysis}
1026 >
1027 > Recently, advanced visualization technique have become applied to
1028 > monitor the motions of molecules. Although the dynamics of the
1029 > system can be described qualitatively from animation, quantitative
1030 > trajectory analysis are more useful. According to the principles of
1031 > Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
1032 > one can compute thermodynamic properties, analyze fluctuations of
1033 > structural parameters, and investigate time-dependent processes of
1034 > the molecule from the trajectories.
1035 >
1036 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1037 >
1038 > Thermodynamic properties, which can be expressed in terms of some
1039 > function of the coordinates and momenta of all particles in the
1040 > system, can be directly computed from molecular dynamics. The usual
1041 > way to measure the pressure is based on virial theorem of Clausius
1042 > which states that the virial is equal to $-3Nk_BT$. For a system
1043 > with forces between particles, the total virial, $W$, contains the
1044 > contribution from external pressure and interaction between the
1045 > particles:
1046 > \[
1047 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1048 > f_{ij} } } \right\rangle
1049 > \]
1050 > where $f_{ij}$ is the force between particle $i$ and $j$ at a
1051 > distance $r_{ij}$. Thus, the expression for the pressure is given
1052 > by:
1053 > \begin{equation}
1054 > P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1055 > < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1056 > \end{equation}
1057 >
1058 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1059 >
1060 > Structural Properties of a simple fluid can be described by a set of
1061 > distribution functions. Among these functions,the \emph{pair
1062 > distribution function}, also known as \emph{radial distribution
1063 > function}, is of most fundamental importance to liquid theory.
1064 > Experimentally, pair distribution function can be gathered by
1065 > Fourier transforming raw data from a series of neutron diffraction
1066 > experiments and integrating over the surface factor
1067 > \cite{Powles1973}. The experimental results can serve as a criterion
1068 > to justify the correctness of a liquid model. Moreover, various
1069 > equilibrium thermodynamic and structural properties can also be
1070 > expressed in terms of radial distribution function \cite{Allen1987}.
1071 > The pair distribution functions $g(r)$ gives the probability that a
1072 > particle $i$ will be located at a distance $r$ from a another
1073 > particle $j$ in the system
1074 > \[
1075 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1076 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1077 > (r)}{\rho}.
1078 > \]
1079 > Note that the delta function can be replaced by a histogram in
1080 > computer simulation. Peaks in $g(r)$ represent solvent shells, and
1081 > the height of these peaks gradually decreases to 1 as the liquid of
1082 > large distance approaches the bulk density.
1083 >
1084 >
1085 > \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1086 > Properties}}
1087 >
1088 > Time-dependent properties are usually calculated using \emph{time
1089 > correlation functions}, which correlate random variables $A$ and $B$
1090 > at two different times,
1091 > \begin{equation}
1092 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1093 > \label{introEquation:timeCorrelationFunction}
1094 > \end{equation}
1095 > If $A$ and $B$ refer to same variable, this kind of correlation
1096 > function is called an \emph{autocorrelation function}. One example
1097 > of an auto correlation function is the velocity auto-correlation
1098 > function which is directly related to transport properties of
1099 > molecular liquids:
1100 > \[
1101 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1102 > \right\rangle } dt
1103 > \]
1104 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1105 > function, which is averaging over time origins and over all the
1106 > atoms, the dipole autocorrelation functions are calculated for the
1107 > entire system. The dipole autocorrelation function is given by:
1108 > \[
1109 > c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1110 > \right\rangle
1111 > \]
1112 > Here $u_{tot}$ is the net dipole of the entire system and is given
1113 > by
1114 > \[
1115 > u_{tot} (t) = \sum\limits_i {u_i (t)}
1116 > \]
1117 > In principle, many time correlation functions can be related with
1118 > Fourier transforms of the infrared, Raman, and inelastic neutron
1119 > scattering spectra of molecular liquids. In practice, one can
1120 > extract the IR spectrum from the intensity of dipole fluctuation at
1121 > each frequency using the following relationship:
1122 > \[
1123 > \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1124 > i2\pi vt} dt}
1125 > \]
1126  
1127   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1128  
# Line 917 | Line 1130 | simulator is governed by the rigid body dynamics. In m
1130   areas, from engineering, physics, to chemistry. For example,
1131   missiles and vehicle are usually modeled by rigid bodies.  The
1132   movement of the objects in 3D gaming engine or other physics
1133 < simulator is governed by the rigid body dynamics. In molecular
1134 < simulation, rigid body is used to simplify the model in
1135 < protein-protein docking study{\cite{Gray03}}.
1133 > simulator is governed by rigid body dynamics. In molecular
1134 > simulations, rigid bodies are used to simplify protein-protein
1135 > docking studies\cite{Gray2003}.
1136  
1137   It is very important to develop stable and efficient methods to
1138 < integrate the equations of motion of orientational degrees of
1139 < freedom. Euler angles are the nature choice to describe the
1140 < rotational degrees of freedom. However, due to its singularity, the
1141 < numerical integration of corresponding equations of motion is very
1142 < inefficient and inaccurate. Although an alternative integrator using
1143 < different sets of Euler angles can overcome this difficulty\cite{},
1144 < the computational penalty and the lost of angular momentum
1145 < conservation still remain. A singularity free representation
1146 < utilizing quaternions was developed by Evans in 1977. Unfortunately,
1147 < this approach suffer from the nonseparable Hamiltonian resulted from
1138 > integrate the equations of motion for orientational degrees of
1139 > freedom. Euler angles are the natural choice to describe the
1140 > rotational degrees of freedom. However, due to $\frac {1}{sin
1141 > \theta}$ singularities, the numerical integration of corresponding
1142 > equations of motion is very inefficient and inaccurate. Although an
1143 > alternative integrator using multiple sets of Euler angles can
1144 > overcome this difficulty\cite{Barojas1973}, the computational
1145 > penalty and the loss of angular momentum conservation still remain.
1146 > A singularity-free representation utilizing quaternions was
1147 > developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1148 > approach uses a nonseparable Hamiltonian resulting from the
1149   quaternion representation, which prevents the symplectic algorithm
1150   to be utilized. Another different approach is to apply holonomic
1151   constraints to the atoms belonging to the rigid body. Each atom
1152   moves independently under the normal forces deriving from potential
1153   energy and constraint forces which are used to guarantee the
1154 < rigidness. However, due to their iterative nature, SHAKE and Rattle
1155 < algorithm converge very slowly when the number of constraint
1156 < increases.
1154 > rigidness. However, due to their iterative nature, the SHAKE and
1155 > Rattle algorithms also converge very slowly when the number of
1156 > constraints increases\cite{Ryckaert1977, Andersen1983}.
1157  
1158 < The break through in geometric literature suggests that, in order to
1158 > A break-through in geometric literature suggests that, in order to
1159   develop a long-term integration scheme, one should preserve the
1160 < symplectic structure of the flow. Introducing conjugate momentum to
1161 < rotation matrix $A$ and re-formulating Hamiltonian's equation, a
1162 < symplectic integrator, RSHAKE, was proposed to evolve the
1163 < Hamiltonian system in a constraint manifold by iteratively
1164 < satisfying the orthogonality constraint $A_t A = 1$. An alternative
1165 < method using quaternion representation was developed by Omelyan.
1166 < However, both of these methods are iterative and inefficient. In
1167 < this section, we will present a symplectic Lie-Poisson integrator
1168 < for rigid body developed by Dullweber and his coworkers\cite{}.
1160 > symplectic structure of the flow. By introducing a conjugate
1161 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1162 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1163 > proposed to evolve the Hamiltonian system in a constraint manifold
1164 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1165 > An alternative method using the quaternion representation was
1166 > developed by Omelyan\cite{Omelyan1998}. However, both of these
1167 > methods are iterative and inefficient. In this section, we descibe a
1168 > symplectic Lie-Poisson integrator for rigid body developed by
1169 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1170  
1171 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
1172 <
1173 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
959 <
960 < \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
961 <
962 < \section{\label{introSection:correlationFunctions}Correlation Functions}
963 <
964 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
965 <
966 < \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
967 <
968 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
969 <
1171 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1172 > The motion of a rigid body is Hamiltonian with the Hamiltonian
1173 > function
1174   \begin{equation}
1175 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1176 < \label{introEquation:bathGLE}
1175 > H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1176 > V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1177 > \label{introEquation:RBHamiltonian}
1178   \end{equation}
1179 < where $H_B$ is harmonic bath Hamiltonian,
1179 > Here, $q$ and $Q$  are the position and rotation matrix for the
1180 > rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1181 > $J$, a diagonal matrix, is defined by
1182   \[
1183 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
977 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1183 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1184   \]
1185 < and $\Delta U$ is bilinear system-bath coupling,
1185 > where $I_{ii}$ is the diagonal element of the inertia tensor. This
1186 > constrained Hamiltonian equation is subjected to a holonomic
1187 > constraint,
1188 > \begin{equation}
1189 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1190 > \end{equation}
1191 > which is used to ensure rotation matrix's unitarity. Differentiating
1192 > \ref{introEquation:orthogonalConstraint} and using Equation
1193 > \ref{introEquation:RBMotionMomentum}, one may obtain,
1194 > \begin{equation}
1195 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1196 > \label{introEquation:RBFirstOrderConstraint}
1197 > \end{equation}
1198 > Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1199 > \ref{introEquation:motionHamiltonianMomentum}), one can write down
1200 > the equations of motion,
1201 > \begin{eqnarray}
1202 > \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1203 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1204 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1205 > \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1206 > \end{eqnarray}
1207 > In general, there are two ways to satisfy the holonomic constraints.
1208 > We can use a constraint force provided by a Lagrange multiplier on
1209 > the normal manifold to keep the motion on constraint space. Or we
1210 > can simply evolve the system on the constraint manifold. These two
1211 > methods have been proved to be equivalent. The holonomic constraint
1212 > and equations of motions define a constraint manifold for rigid
1213 > bodies
1214   \[
1215 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1215 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1216 > \right\}.
1217   \]
1218 < Completing the square,
1218 > Unfortunately, this constraint manifold is not the cotangent bundle
1219 > $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1220 > rotation group $SO(3)$. However, it turns out that under symplectic
1221 > transformation, the cotangent space and the phase space are
1222 > diffeomorphic. By introducing
1223   \[
1224 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
986 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
987 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
988 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
989 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1224 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1225   \]
1226 < and putting it back into Eq.~\ref{introEquation:bathGLE},
1226 > the mechanical system subject to a holonomic constraint manifold $M$
1227 > can be re-formulated as a Hamiltonian system on the cotangent space
1228   \[
1229 < H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1230 < {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
995 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
996 < w_\alpha ^2 }}x} \right)^2 } \right\}}
1229 > T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1230 > 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1231   \]
1232 < where
1232 > For a body fixed vector $X_i$ with respect to the center of mass of
1233 > the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1234 > given as
1235 > \begin{equation}
1236 > X_i^{lab} = Q X_i + q.
1237 > \end{equation}
1238 > Therefore, potential energy $V(q,Q)$ is defined by
1239   \[
1240 < W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1001 < }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1240 > V(q,Q) = V(Q X_0 + q).
1241   \]
1242 < Since the first two terms of the new Hamiltonian depend only on the
1004 < system coordinates, we can get the equations of motion for
1005 < Generalized Langevin Dynamics by Hamilton's equations
1006 < \ref{introEquation:motionHamiltonianCoordinate,
1007 < introEquation:motionHamiltonianMomentum},
1008 < \begin{align}
1009 < \dot p &=  - \frac{{\partial H}}{{\partial x}}
1010 <       &= m\ddot x
1011 <       &= - \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)}
1012 < \label{introEquation:Lp5}
1013 < \end{align}
1014 < , and
1015 < \begin{align}
1016 < \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
1017 <                &= m\ddot x_\alpha
1018 <                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
1019 < \end{align}
1020 <
1021 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1022 <
1242 > Hence, the force and torque are given by
1243   \[
1244 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1244 > \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1245   \]
1246 <
1246 > and
1247   \[
1248 < L(x + y) = L(x) + L(y)
1248 > \nabla _Q V(q,Q) = F(q,Q)X_i^t
1249   \]
1250 <
1250 > respectively. As a common choice to describe the rotation dynamics
1251 > of the rigid body, the angular momentum on the body fixed frame $\Pi
1252 > = Q^t P$ is introduced to rewrite the equations of motion,
1253 > \begin{equation}
1254 > \begin{array}{l}
1255 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda,  \\
1256 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1},  \\
1257 > \end{array}
1258 > \label{introEqaution:RBMotionPI}
1259 > \end{equation}
1260 > as well as holonomic constraints,
1261   \[
1262 < L(ax) = aL(x)
1262 > \begin{array}{l}
1263 > \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0, \\
1264 > Q^T Q = 1 .\\
1265 > \end{array}
1266   \]
1267 <
1267 > For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1268 > so(3)^ \star$, the hat-map isomorphism,
1269 > \begin{equation}
1270 > v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1271 > {\begin{array}{*{20}c}
1272 >   0 & { - v_3 } & {v_2 }  \\
1273 >   {v_3 } & 0 & { - v_1 }  \\
1274 >   { - v_2 } & {v_1 } & 0  \\
1275 > \end{array}} \right),
1276 > \label{introEquation:hatmapIsomorphism}
1277 > \end{equation}
1278 > will let us associate the matrix products with traditional vector
1279 > operations
1280   \[
1281 < L(\dot x) = pL(x) - px(0)
1281 > \hat vu = v \times u.
1282   \]
1283 <
1283 > Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew
1284 > matrix,
1285 > \begin{eqnarray}
1286 > (\dot \Pi  - \dot \Pi ^T ){\rm{ }} &= &{\rm{ }}(\Pi  - \Pi ^T ){\rm{
1287 > }}(J^{ - 1} \Pi  + \Pi J^{ - 1} ) \notag \\
1288 > + \sum\limits_i {[Q^T F_i
1289 > (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  - (\Lambda  - \Lambda ^T ).
1290 > \label{introEquation:skewMatrixPI}
1291 > \end{eqnarray}
1292 > Since $\Lambda$ is symmetric, the last term of
1293 > Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1294 > Lagrange multiplier $\Lambda$ is absent from the equations of
1295 > motion. This unique property eliminates the requirement of
1296 > iterations which can not be avoided in other methods\cite{Kol1997,
1297 > Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1298 > equation of motion for angular momentum on body frame
1299 > \begin{equation}
1300 > \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1301 > F_i (r,Q)} \right) \times X_i }.
1302 > \label{introEquation:bodyAngularMotion}
1303 > \end{equation}
1304 > In the same manner, the equation of motion for rotation matrix is
1305 > given by
1306   \[
1307 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1307 > \dot Q = Qskew(I^{ - 1} \pi ).
1308   \]
1309  
1310 + \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1311 + Lie-Poisson Integrator for Free Rigid Body}
1312 +
1313 + If there are no external forces exerted on the rigid body, the only
1314 + contribution to the rotational motion is from the kinetic energy
1315 + (the first term of \ref{introEquation:bodyAngularMotion}). The free
1316 + rigid body is an example of a Lie-Poisson system with Hamiltonian
1317 + function
1318 + \begin{equation}
1319 + T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1320 + \label{introEquation:rotationalKineticRB}
1321 + \end{equation}
1322 + where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1323 + Lie-Poisson structure matrix,
1324 + \begin{equation}
1325 + J(\pi ) = \left( {\begin{array}{*{20}c}
1326 +   0 & {\pi _3 } & { - \pi _2 }  \\
1327 +   { - \pi _3 } & 0 & {\pi _1 }  \\
1328 +   {\pi _2 } & { - \pi _1 } & 0  \\
1329 + \end{array}} \right).
1330 + \end{equation}
1331 + Thus, the dynamics of free rigid body is governed by
1332 + \begin{equation}
1333 + \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi ).
1334 + \end{equation}
1335 + One may notice that each $T_i^r$ in Equation
1336 + \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1337 + instance, the equations of motion due to $T_1^r$ are given by
1338 + \begin{equation}
1339 + \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1340 + \label{introEqaution:RBMotionSingleTerm}
1341 + \end{equation}
1342 + where
1343 + \[ R_1  = \left( {\begin{array}{*{20}c}
1344 +   0 & 0 & 0  \\
1345 +   0 & 0 & {\pi _1 }  \\
1346 +   0 & { - \pi _1 } & 0  \\
1347 + \end{array}} \right).
1348 + \]
1349 + The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1350   \[
1351 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1351 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1352 > Q(0)e^{\Delta tR_1 }
1353   \]
1354 <
1047 < Some relatively important transformation,
1354 > with
1355   \[
1356 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1356 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1357 >   0 & 0 & 0  \\
1358 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1359 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1360 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1361   \]
1362 <
1362 > To reduce the cost of computing expensive functions in $e^{\Delta
1363 > tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1364 > propagator,
1365   \[
1366 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1366 > e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1367 > ).
1368   \]
1369 <
1369 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1370 > manner. In order to construct a second-order symplectic method, we
1371 > split the angular kinetic Hamiltonian function can into five terms
1372   \[
1373 < L(1) = \frac{1}{p}
1373 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1374 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1375 > (\pi _1 ).
1376   \]
1377 <
1378 < First, the bath coordinates,
1377 > By concatenating the propagators corresponding to these five terms,
1378 > we can obtain an symplectic integrator,
1379   \[
1380 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
1381 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
1382 < }}L(x)
1380 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1381 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1382 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1383 > _1 }.
1384   \]
1385 + The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1386 + $F(\pi )$ and $G(\pi )$ is defined by
1387   \[
1388 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1389 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1388 > \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1389 > ).
1390   \]
1391 < Then, the system coordinates,
1392 < \begin{align}
1393 < mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1394 < \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1395 < }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1396 < (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1397 < }}\omega _\alpha ^2 L(x)} \right\}}
1398 < %
1399 < &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1400 < \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1401 < - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1402 < - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1403 < \end{align}
1404 < Then, the inverse transform,
1391 > If the Poisson bracket of a function $F$ with an arbitrary smooth
1392 > function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1393 > conserved quantity in Poisson system. We can easily verify that the
1394 > norm of the angular momentum, $\parallel \pi
1395 > \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1396 > \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1397 > then by the chain rule
1398 > \[
1399 > \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1400 > }}{2})\pi.
1401 > \]
1402 > Thus, $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel
1403 > \pi
1404 > \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1405 > Lie-Poisson integrator is found to be both extremely efficient and
1406 > stable. These properties can be explained by the fact the small
1407 > angle approximation is used and the norm of the angular momentum is
1408 > conserved.
1409  
1410 < \begin{align}
1411 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1410 > \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1411 > Splitting for Rigid Body}
1412 >
1413 > The Hamiltonian of rigid body can be separated in terms of kinetic
1414 > energy and potential energy,
1415 > \[
1416 > H = T(p,\pi ) + V(q,Q).
1417 > \]
1418 > The equations of motion corresponding to potential energy and
1419 > kinetic energy are listed in the below table,
1420 > \begin{table}
1421 > \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1422 > \begin{center}
1423 > \begin{tabular}{|l|l|}
1424 >  \hline
1425 >  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1426 >  Potential & Kinetic \\
1427 >  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1428 >  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1429 >  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1430 >  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1431 >  \hline
1432 > \end{tabular}
1433 > \end{center}
1434 > \end{table}
1435 > A second-order symplectic method is now obtained by the composition
1436 > of the position and velocity propagators,
1437 > \[
1438 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1439 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1440 > \]
1441 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1442 > sub-propagators which corresponding to force and torque
1443 > respectively,
1444 > \[
1445 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1446 > _{\Delta t/2,\tau }.
1447 > \]
1448 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1449 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1450 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1451 > kinetic energy can be separated to translational kinetic term, $T^t
1452 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1453 > \begin{equation}
1454 > T(p,\pi ) =T^t (p) + T^r (\pi ).
1455 > \end{equation}
1456 > where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1457 > defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1458 > corresponding propagators are given by
1459 > \[
1460 > \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1461 > _{\Delta t,T^r }.
1462 > \]
1463 > Finally, we obtain the overall symplectic propagators for freely
1464 > moving rigid bodies
1465 > \begin{eqnarray*}
1466 > \varphi _{\Delta t}  &=& \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1467 >  & & \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 }  \\
1468 >  & & \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1469 > \label{introEquation:overallRBFlowMaps}
1470 > \end{eqnarray*}
1471 >
1472 > \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1473 > As an alternative to newtonian dynamics, Langevin dynamics, which
1474 > mimics a simple heat bath with stochastic and dissipative forces,
1475 > has been applied in a variety of studies. This section will review
1476 > the theory of Langevin dynamics. A brief derivation of generalized
1477 > Langevin equation will be given first. Following that, we will
1478 > discuss the physical meaning of the terms appearing in the equation
1479 > as well as the calculation of friction tensor from hydrodynamics
1480 > theory.
1481 >
1482 > \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1483 >
1484 > A harmonic bath model, in which an effective set of harmonic
1485 > oscillators are used to mimic the effect of a linearly responding
1486 > environment, has been widely used in quantum chemistry and
1487 > statistical mechanics. One of the successful applications of
1488 > Harmonic bath model is the derivation of the Generalized Langevin
1489 > Dynamics (GLE). Lets consider a system, in which the degree of
1490 > freedom $x$ is assumed to couple to the bath linearly, giving a
1491 > Hamiltonian of the form
1492 > \begin{equation}
1493 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1494 > \label{introEquation:bathGLE}.
1495 > \end{equation}
1496 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1497 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1498 > \[
1499 > H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1500 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1501 > \right\}}
1502 > \]
1503 > where the index $\alpha$ runs over all the bath degrees of freedom,
1504 > $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1505 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1506 > coupling,
1507 > \[
1508 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1509 > \]
1510 > where $g_\alpha$ are the coupling constants between the bath
1511 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1512 > Introducing
1513 > \[
1514 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1515 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1516 > \]
1517 > and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as
1518 > \[
1519 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1520 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1521 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1522 > w_\alpha ^2 }}x} \right)^2 } \right\}}.
1523 > \]
1524 > Since the first two terms of the new Hamiltonian depend only on the
1525 > system coordinates, we can get the equations of motion for
1526 > Generalized Langevin Dynamics by Hamilton's equations,
1527 > \begin{equation}
1528 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1529 > \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1530 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1531 > \label{introEquation:coorMotionGLE}
1532 > \end{equation}
1533 > and
1534 > \begin{equation}
1535 > m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1536 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1537 > \label{introEquation:bathMotionGLE}
1538 > \end{equation}
1539 > In order to derive an equation for $x$, the dynamics of the bath
1540 > variables $x_\alpha$ must be solved exactly first. As an integral
1541 > transform which is particularly useful in solving linear ordinary
1542 > differential equations,the Laplace transform is the appropriate tool
1543 > to solve this problem. The basic idea is to transform the difficult
1544 > differential equations into simple algebra problems which can be
1545 > solved easily. Then, by applying the inverse Laplace transform, also
1546 > known as the Bromwich integral, we can retrieve the solutions of the
1547 > original problems. Let $f(t)$ be a function defined on $ [0,\infty )
1548 > $. The Laplace transform of f(t) is a new function defined as
1549 > \[
1550 > L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1551 > \]
1552 > where  $p$ is real and  $L$ is called the Laplace Transform
1553 > Operator. Below are some important properties of Laplace transform
1554 > \begin{eqnarray*}
1555 > L(x + y)  & = & L(x) + L(y) \\
1556 > L(ax)     & = & aL(x) \\
1557 > L(\dot x) & = & pL(x) - px(0) \\
1558 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1559 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1560 > \end{eqnarray*}
1561 > Applying the Laplace transform to the bath coordinates, we obtain
1562 > \begin{eqnarray*}
1563 > 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) \\
1564 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1565 > \end{eqnarray*}
1566 > By the same way, the system coordinates become
1567 > \begin{eqnarray*}
1568 > mL(\ddot x) & = &
1569 >  - \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\}}  \\
1570 >  & & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}
1571 > \end{eqnarray*}
1572 > With the help of some relatively important inverse Laplace
1573 > transformations:
1574 > \[
1575 > \begin{array}{c}
1576 > L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1577 > L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1578 > L(1) = \frac{1}{p} \\
1579 > \end{array}
1580 > \]
1581 > we obtain
1582 > \begin{eqnarray*}
1583 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1584   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1585   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1586 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1587 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1588 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1589 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1590 < %
1591 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1586 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1587 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1588 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1589 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1590 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1591 > \end{eqnarray*}
1592 > \begin{eqnarray*}
1593 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1594   {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1595   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1596 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1597 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1598 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1599 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1600 < (\omega _\alpha  t)} \right\}}
1601 < \end{align}
1602 <
1596 > t)\dot x(t - \tau )d} \tau }  \\
1597 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1598 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1599 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1600 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1601 > \end{eqnarray*}
1602 > Introducing a \emph{dynamic friction kernel}
1603   \begin{equation}
1604 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1605 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1606 + \label{introEquation:dynamicFrictionKernelDefinition}
1607 + \end{equation}
1608 + and \emph{a random force}
1609 + \begin{equation}
1610 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1611 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1612 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1613 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1614 + \label{introEquation:randomForceDefinition}
1615 + \end{equation}
1616 + the equation of motion can be rewritten as
1617 + \begin{equation}
1618   m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1619   (t)\dot x(t - \tau )d\tau }  + R(t)
1620   \label{introEuqation:GeneralizedLangevinDynamics}
1621   \end{equation}
1622 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1623 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1622 > which is known as the \emph{generalized Langevin equation}.
1623 >
1624 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1625 >
1626 > One may notice that $R(t)$ depends only on initial conditions, which
1627 > implies it is completely deterministic within the context of a
1628 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1629 > uncorrelated to $x$ and $\dot x$,
1630   \[
1631 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1632 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1631 > \begin{array}{l}
1632 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1633 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1634 > \end{array}
1635   \]
1636 < For an infinite harmonic bath, we can use the spectral density and
1637 < an integral over frequencies.
1636 > This property is what we expect from a truly random process. As long
1637 > as the model chosen for $R(t)$ was a gaussian distribution in
1638 > general, the stochastic nature of the GLE still remains.
1639  
1640 + %dynamic friction kernel
1641 + The convolution integral
1642   \[
1643 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1120 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1121 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1122 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1643 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1644   \]
1645 < The random forces depend only on initial conditions.
1646 <
1647 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1648 < So we can define a new set of coordinates,
1645 > depends on the entire history of the evolution of $x$, which implies
1646 > that the bath retains memory of previous motions. In other words,
1647 > the bath requires a finite time to respond to change in the motion
1648 > of the system. For a sluggish bath which responds slowly to changes
1649 > in the system coordinate, we may regard $\xi(t)$ as a constant
1650 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1651   \[
1652 < q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1130 < ^2 }}x(0)
1652 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1653   \]
1654 < This makes
1654 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1655   \[
1656 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1656 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1657 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1658   \]
1659 < And since the $q$ coordinates are harmonic oscillators,
1659 > which can be used to describe the effect of dynamic caging in
1660 > viscous solvents. The other extreme is the bath that responds
1661 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1662 > taken as a $delta$ function in time:
1663   \[
1664 < \begin{array}{l}
1139 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1140 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1141 < \end{array}
1664 > \xi (t) = 2\xi _0 \delta (t)
1665   \]
1666 + Hence, the convolution integral becomes
1667 + \[
1668 + \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1669 + {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1670 + \]
1671 + and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1672 + \begin{equation}
1673 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1674 + x(t) + R(t) \label{introEquation:LangevinEquation}
1675 + \end{equation}
1676 + which is known as the Langevin equation. The static friction
1677 + coefficient $\xi _0$ can either be calculated from spectral density
1678 + or be determined by Stokes' law for regular shaped particles. A
1679 + briefly review on calculating friction tensor for arbitrary shaped
1680 + particles is given in Sec.~\ref{introSection:frictionTensor}.
1681  
1682 < \begin{align}
1145 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1146 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1147 < (t)q_\beta  (0)} \right\rangle } }
1148 < %
1149 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1150 < \right\rangle \cos (\omega _\alpha  t)}
1151 < %
1152 < &= kT\xi (t)
1153 < \end{align}
1682 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1683  
1684 + Defining a new set of coordinates,
1685 + \[
1686 + q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1687 + ^2 }}x(0)
1688 + \],
1689 + we can rewrite $R(T)$ as
1690 + \[
1691 + R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1692 + \]
1693 + And since the $q$ coordinates are harmonic oscillators,
1694 + \begin{eqnarray*}
1695 + \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1696 + \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1697 + \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1698 + \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 } }  \\
1699 +  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1700 +  & = &kT\xi (t) \\
1701 + \end{eqnarray*}
1702 + Thus, we recover the \emph{second fluctuation dissipation theorem}
1703   \begin{equation}
1704   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1705 < \label{introEquation:secondFluctuationDissipation}
1705 > \label{introEquation:secondFluctuationDissipation}.
1706   \end{equation}
1707 <
1708 < \section{\label{introSection:hydroynamics}Hydrodynamics}
1161 <
1162 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
1163 < \subsection{\label{introSection:analyticalApproach}Analytical
1164 < Approach}
1165 <
1166 < \subsection{\label{introSection:approximationApproach}Approximation
1167 < Approach}
1168 <
1169 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1170 < Body}
1707 > In effect, it acts as a constraint on the possible ways in which one
1708 > can model the random force and friction kernel.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines