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 2786 by tim, Sun Jun 4 20:18:07 2006 UTC vs.
Revision 2911 by tim, Thu Jun 29 23:56:11 2006 UTC

# Line 3 | Line 3 | Closely related to Classical Mechanics, Molecular Dyna
3   \section{\label{introSection:classicalMechanics}Classical
4   Mechanics}
5  
6 < Closely related to Classical Mechanics, Molecular Dynamics
7 < simulations are carried out by integrating the equations of motion
8 < for a given system of particles. There are three fundamental ideas
9 < behind classical mechanics. Firstly, One can determine the state of
10 < a mechanical system at any time of interest; Secondly, all the
11 < mechanical properties of the system at that time can be determined
12 < by combining the knowledge of the properties of the system with the
13 < specification of this state; Finally, the specification of the state
14 < when further combine with the laws of mechanics will also be
15 < sufficient to predict the future behavior of the system.
6 > Using equations of motion derived from Classical Mechanics,
7 > Molecular Dynamics simulations are carried out by integrating the
8 > equations of motion for a given system of particles. There are three
9 > fundamental ideas behind classical mechanics. Firstly, one can
10 > determine the state of a mechanical system at any time of interest;
11 > Secondly, all the mechanical properties of the system at that time
12 > can be determined by combining the knowledge of the properties of
13 > the system with the specification of this state; Finally, the
14 > specification of the state when further combined with the laws of
15 > mechanics will also be sufficient to predict the future behavior of
16 > the system.
17  
18   \subsection{\label{introSection:newtonian}Newtonian Mechanics}
19   The discovery of Newton's three laws of mechanics which govern the
20   motion of particles is the foundation of the classical mechanics.
21 < Newton¡¯s first law defines a class of inertial frames. Inertial
21 > Newton's first law defines a class of inertial frames. Inertial
22   frames are reference frames where a particle not interacting with
23   other bodies will move with constant speed in the same direction.
24 < With respect to inertial frames Newton¡¯s second law has the form
24 > With respect to inertial frames, Newton's second law has the form
25   \begin{equation}
26 < F = \frac {dp}{dt} = \frac {mv}{dt}
26 > F = \frac {dp}{dt} = \frac {mdv}{dt}
27   \label{introEquation:newtonSecondLaw}
28   \end{equation}
29   A point mass interacting with other bodies moves with the
30   acceleration along the direction of the force acting on it. Let
31   $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
32   $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
33 < Newton¡¯s third law states that
33 > Newton's third law states that
34   \begin{equation}
35 < F_{ij} = -F_{ji}
35 > F_{ij} = -F_{ji}.
36   \label{introEquation:newtonThirdLaw}
37   \end{equation}
37
38   Conservation laws of Newtonian Mechanics play very important roles
39   in solving mechanics problems. The linear momentum of a particle is
40   conserved if it is free or it experiences no force. The second
# Line 46 | Line 46 | N \equiv r \times F \label{introEquation:torqueDefinit
46   \end{equation}
47   The torque $\tau$ with respect to the same origin is defined to be
48   \begin{equation}
49 < N \equiv r \times F \label{introEquation:torqueDefinition}
49 > \tau \equiv r \times F \label{introEquation:torqueDefinition}
50   \end{equation}
51   Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52   \[
# Line 59 | Line 59 | thus,
59   \]
60   thus,
61   \begin{equation}
62 < \dot L = r \times \dot p = N
62 > \dot L = r \times \dot p = \tau
63   \end{equation}
64   If there are no external torques acting on a body, the angular
65   momentum of it is conserved. The last conservation theorem state
66 < that if all forces are conservative, Energy
67 < \begin{equation}E = T + V \label{introEquation:energyConservation}
66 > that if all forces are conservative, energy is conserved,
67 > \begin{equation}E = T + V. \label{introEquation:energyConservation}
68   \end{equation}
69 < is conserved. All of these conserved quantities are
70 < important factors to determine the quality of numerical integration
71 < scheme for rigid body \cite{Dullweber1997}.
69 > All of these conserved quantities are important factors to determine
70 > the quality of numerical integration schemes for rigid bodies
71 > \cite{Dullweber1997}.
72  
73   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
75 < Newtonian Mechanics suffers from two important limitations: it
76 < describes their motion in special cartesian coordinate systems.
77 < Another limitation of Newtonian mechanics becomes obvious when we
78 < try to describe systems with large numbers of particles. It becomes
79 < very difficult to predict the properties of the system by carrying
80 < out calculations involving the each individual interaction between
81 < all the particles, even if we know all of the details of the
82 < interaction. In order to overcome some of the practical difficulties
83 < which arise in attempts to apply Newton's equation to complex
84 < system, alternative procedures may be developed.
75 > Newtonian Mechanics suffers from an important limitation: motion can
76 > only be described in cartesian coordinate systems which make it
77 > impossible to predict analytically the properties of the system even
78 > if we know all of the details of the interaction. In order to
79 > overcome some of the practical difficulties which arise in attempts
80 > to apply Newton's equation to complex systems, approximate numerical
81 > procedures may be developed.
82  
83 < \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
84 < Principle}
83 > \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
84 > Principle}}
85  
86   Hamilton introduced the dynamical principle upon which it is
87 < possible to base all of mechanics and, indeed, most of classical
88 < physics. Hamilton's Principle may be stated as follow,
89 <
90 < The actual trajectory, along which a dynamical system may move from
91 < one point to another within a specified time, is derived by finding
92 < the path which minimizes the time integral of the difference between
96 < the kinetic, $K$, and potential energies, $U$ \cite{Tolman1979}.
87 > possible to base all of mechanics and most of classical physics.
88 > Hamilton's Principle may be stated as follows: the trajectory, along
89 > which a dynamical system may move from one point to another within a
90 > specified time, is derived by finding the path which minimizes the
91 > time integral of the difference between the kinetic $K$, and
92 > potential energies $U$,
93   \begin{equation}
94 < \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
94 > \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0}.
95   \label{introEquation:halmitonianPrinciple1}
96   \end{equation}
101
97   For simple mechanical systems, where the forces acting on the
98 < different part are derivable from a potential and the velocities are
99 < small compared with that of light, the Lagrangian function $L$ can
100 < be define as the difference between the kinetic energy of the system
106 < and its potential energy,
98 > different parts are derivable from a potential, the Lagrangian
99 > function $L$ can be defined as the difference between the kinetic
100 > energy of the system and its potential energy,
101   \begin{equation}
102 < L \equiv K - U = L(q_i ,\dot q_i ) ,
102 > L \equiv K - U = L(q_i ,\dot q_i ).
103   \label{introEquation:lagrangianDef}
104   \end{equation}
105 < then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
105 > Thus, Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
106   \begin{equation}
107 < \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
107 > \delta \int_{t_1 }^{t_2 } {L dt = 0} .
108   \label{introEquation:halmitonianPrinciple2}
109   \end{equation}
110  
111 < \subsubsection{\label{introSection:equationOfMotionLagrangian}The
112 < Equations of Motion in Lagrangian Mechanics}
111 > \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
112 > Equations of Motion in Lagrangian Mechanics}}
113  
114 < For a holonomic system of $f$ degrees of freedom, the equations of
115 < motion in the Lagrangian form is
114 > For a system of $f$ degrees of freedom, the equations of motion in
115 > the Lagrangian form is
116   \begin{equation}
117   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
118   \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
# Line 132 | Line 126 | independent of generalized velocities, the generalized
126   Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
127   introduced by William Rowan Hamilton in 1833 as a re-formulation of
128   classical mechanics. If the potential energy of a system is
129 < independent of generalized velocities, the generalized momenta can
136 < be defined as
129 > independent of velocities, the momenta can be defined as
130   \begin{equation}
131   p_i = \frac{\partial L}{\partial \dot q_i}
132   \label{introEquation:generalizedMomenta}
# Line 143 | Line 136 | p_i  = \frac{{\partial L}}{{\partial q_i }}
136   p_i  = \frac{{\partial L}}{{\partial q_i }}
137   \label{introEquation:generalizedMomentaDot}
138   \end{equation}
146
139   With the help of the generalized momenta, we may now define a new
140   quantity $H$ by the equation
141   \begin{equation}
# Line 151 | Line 143 | $L$ is the Lagrangian function for the system.
143   \label{introEquation:hamiltonianDefByLagrangian}
144   \end{equation}
145   where $ \dot q_1  \ldots \dot q_f $ are generalized velocities and
146 < $L$ is the Lagrangian function for the system.
147 <
156 < Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
157 < one can obtain
146 > $L$ is the Lagrangian function for the system. Differentiating
147 > Eq.~\ref{introEquation:hamiltonianDefByLagrangian}, one can obtain
148   \begin{equation}
149   dH = \sum\limits_k {\left( {p_k d\dot q_k  + \dot q_k dp_k  -
150   \frac{{\partial L}}{{\partial q_k }}dq_k  - \frac{{\partial
151   L}}{{\partial \dot q_k }}d\dot q_k } \right)}  - \frac{{\partial
152 < L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
152 > L}}{{\partial t}}dt . \label{introEquation:diffHamiltonian1}
153   \end{equation}
154 < Making use of  Eq.~\ref{introEquation:generalizedMomenta}, the
155 < second and fourth terms in the parentheses cancel. Therefore,
154 > Making use of Eq.~\ref{introEquation:generalizedMomenta}, the second
155 > and fourth terms in the parentheses cancel. Therefore,
156   Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
157   \begin{equation}
158   dH = \sum\limits_k {\left( {\dot q_k dp_k  - \dot p_k dq_k }
159 < \right)}  - \frac{{\partial L}}{{\partial t}}dt
159 > \right)}  - \frac{{\partial L}}{{\partial t}}dt .
160   \label{introEquation:diffHamiltonian2}
161   \end{equation}
162   By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
163   find
164   \begin{equation}
165 < \frac{{\partial H}}{{\partial p_k }} = q_k
165 > \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
166   \label{introEquation:motionHamiltonianCoordinate}
167   \end{equation}
168   \begin{equation}
169 < \frac{{\partial H}}{{\partial q_k }} =  - p_k
169 > \frac{{\partial H}}{{\partial q_k }} =  - \dot {p_k}
170   \label{introEquation:motionHamiltonianMomentum}
171   \end{equation}
172   and
# Line 185 | Line 175 | t}}
175   t}}
176   \label{introEquation:motionHamiltonianTime}
177   \end{equation}
178 <
189 < Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
178 > where Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
179   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
180   equation of motion. Due to their symmetrical formula, they are also
181   known as the canonical equations of motions \cite{Goldstein2001}.
182  
183   An important difference between Lagrangian approach and the
184   Hamiltonian approach is that the Lagrangian is considered to be a
185 < function of the generalized velocities $\dot q_i$ and the
186 < generalized coordinates $q_i$, while the Hamiltonian is considered
187 < to be a function of the generalized momenta $p_i$ and the conjugate
188 < generalized coordinate $q_i$. Hamiltonian Mechanics is more
189 < appropriate for application to statistical mechanics and quantum
190 < mechanics, since it treats the coordinate and its time derivative as
191 < independent variables and it only works with 1st-order differential
203 < equations\cite{Marion1990}.
204 <
185 > function of the generalized velocities $\dot q_i$ and coordinates
186 > $q_i$, while the Hamiltonian is considered to be a function of the
187 > generalized momenta $p_i$ and the conjugate coordinates $q_i$.
188 > Hamiltonian Mechanics is more appropriate for application to
189 > statistical mechanics and quantum mechanics, since it treats the
190 > coordinate and its time derivative as independent variables and it
191 > only works with 1st-order differential equations\cite{Marion1990}.
192   In Newtonian Mechanics, a system described by conservative forces
193 < conserves the total energy \ref{introEquation:energyConservation}.
194 < It follows that Hamilton's equations of motion conserve the total
195 < Hamiltonian.
193 > conserves the total energy
194 > (Eq.~\ref{introEquation:energyConservation}). It follows that
195 > Hamilton's equations of motion conserve the total Hamiltonian
196   \begin{equation}
197   \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
198   H}}{{\partial q_i }}\dot q_i  + \frac{{\partial H}}{{\partial p_i
199   }}\dot p_i } \right)}  = \sum\limits_i {\left( {\frac{{\partial
200   H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
201   \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
202 < q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
202 > q_i }}} \right) = 0}. \label{introEquation:conserveHalmitonian}
203   \end{equation}
204  
205   \section{\label{introSection:statisticalMechanics}Statistical
# Line 227 | Line 214 | possible states. Each possible state of the system cor
214   \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
215  
216   Mathematically, phase space is the space which represents all
217 < possible states. Each possible state of the system corresponds to
218 < one unique point in the phase space. For mechanical systems, the
219 < phase space usually consists of all possible values of position and
220 < momentum variables. Consider a dynamic system in a cartesian space,
221 < where each of the $6f$ coordinates and momenta is assigned to one of
222 < $6f$ mutually orthogonal axes, the phase space of this system is a
223 < $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
224 < \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
217 > possible states of a system. Each possible state of the system
218 > corresponds to one unique point in the phase space. For mechanical
219 > systems, the phase space usually consists of all possible values of
220 > position and momentum variables. Consider a dynamic system of $f$
221 > particles in a cartesian space, where each of the $6f$ coordinates
222 > and momenta is assigned to one of $6f$ mutually orthogonal axes, the
223 > phase space of this system is a $6f$ dimensional space. A point, $x
224 > =
225 > (\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
226 > \over q} _1 , \ldots
227 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
228 > \over q} _f
229 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
230 > \over p} _1  \ldots
231 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
232 > \over p} _f )$ , with a unique set of values of $6f$ coordinates and
233   momenta is a phase space vector.
234 + %%%fix me
235  
236 < A microscopic state or microstate of a classical system is
241 < specification of the complete phase space vector of a system at any
242 < instant in time. An ensemble is defined as a collection of systems
243 < sharing one or more macroscopic characteristics but each being in a
244 < unique microstate. The complete ensemble is specified by giving all
245 < systems or microstates consistent with the common macroscopic
246 < characteristics of the ensemble. Although the state of each
247 < individual system in the ensemble could be precisely described at
248 < any instance in time by a suitable phase space vector, when using
249 < ensembles for statistical purposes, there is no need to maintain
250 < distinctions between individual systems, since the numbers of
251 < systems at any time in the different states which correspond to
252 < different regions of the phase space are more interesting. Moreover,
253 < in the point of view of statistical mechanics, one would prefer to
254 < use ensembles containing a large enough population of separate
255 < members so that the numbers of systems in such different states can
256 < be regarded as changing continuously as we traverse different
257 < regions of the phase space. The condition of an ensemble at any time
236 > In statistical mechanics, the condition of an ensemble at any time
237   can be regarded as appropriately specified by the density $\rho$
238   with which representative points are distributed over the phase
239 < space. The density of distribution for an ensemble with $f$ degrees
240 < of freedom is defined as,
239 > space. The density distribution for an ensemble with $f$ degrees of
240 > freedom is defined as,
241   \begin{equation}
242   \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
243   \label{introEquation:densityDistribution}
244   \end{equation}
245   Governed by the principles of mechanics, the phase points change
246 < their value which would change the density at any time at phase
247 < space. Hence, the density of distribution is also to be taken as a
248 < function of the time.
249 <
271 < The number of systems $\delta N$ at time $t$ can be determined by,
246 > their locations which changes the density at any time at phase
247 > space. Hence, the density distribution is also to be taken as a
248 > function of the time. The number of systems $\delta N$ at time $t$
249 > can be determined by,
250   \begin{equation}
251   \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
252   \label{introEquation:deltaN}
253   \end{equation}
254 < Assuming a large enough population of systems are exploited, we can
255 < sufficiently approximate $\delta N$ without introducing
256 < discontinuity when we go from one region in the phase space to
257 < another. By integrating over the whole phase space,
254 > Assuming enough copies of the systems, we can sufficiently
255 > approximate $\delta N$ without introducing discontinuity when we go
256 > from one region in the phase space to another. By integrating over
257 > the whole phase space,
258   \begin{equation}
259   N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
260   \label{introEquation:totalNumberSystem}
261   \end{equation}
262 < gives us an expression for the total number of the systems. Hence,
263 < the probability per unit in the phase space can be obtained by,
262 > gives us an expression for the total number of copies. Hence, the
263 > probability per unit volume in the phase space can be obtained by,
264   \begin{equation}
265   \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
266   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
267   \label{introEquation:unitProbability}
268   \end{equation}
269 < With the help of Equation(\ref{introEquation:unitProbability}) and
270 < the knowledge of the system, it is possible to calculate the average
269 > With the help of Eq.~\ref{introEquation:unitProbability} and the
270 > knowledge of the system, it is possible to calculate the average
271   value of any desired quantity which depends on the coordinates and
272 < momenta of the system. Even when the dynamics of the real system is
272 > momenta of the system. Even when the dynamics of the real system are
273   complex, or stochastic, or even discontinuous, the average
274 < properties of the ensemble of possibilities as a whole may still
275 < remain well defined. For a classical system in thermal equilibrium
276 < with its environment, the ensemble average of a mechanical quantity,
277 < $\langle A(q , p) \rangle_t$, takes the form of an integral over the
278 < phase space of the system,
274 > properties of the ensemble of possibilities as a whole remain well
275 > defined. For a classical system in thermal equilibrium with its
276 > environment, the ensemble average of a mechanical quantity, $\langle
277 > A(q , p) \rangle_t$, takes the form of an integral over the phase
278 > space of the system,
279   \begin{equation}
280   \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
281   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
282 < (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
282 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
283   \label{introEquation:ensembelAverage}
284   \end{equation}
285  
308 There are several different types of ensembles with different
309 statistical characteristics. As a function of macroscopic
310 parameters, such as temperature \textit{etc}, partition function can
311 be used to describe the statistical properties of a system in
312 thermodynamic equilibrium.
313
314 As an ensemble of systems, each of which is known to be thermally
315 isolated and conserve energy, Microcanonical ensemble(NVE) has a
316 partition function like,
317 \begin{equation}
318 \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
319 \end{equation}
320 A canonical ensemble(NVT)is an ensemble of systems, each of which
321 can share its energy with a large heat reservoir. The distribution
322 of the total energy amongst the possible dynamical states is given
323 by the partition function,
324 \begin{equation}
325 \Omega (N,V,T) = e^{ - \beta A}
326 \label{introEquation:NVTPartition}
327 \end{equation}
328 Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
329 TS$. Since most experiment are carried out under constant pressure
330 condition, isothermal-isobaric ensemble(NPT) play a very important
331 role in molecular simulation. The isothermal-isobaric ensemble allow
332 the system to exchange energy with a heat bath of temperature $T$
333 and to change the volume as well. Its partition function is given as
334 \begin{equation}
335 \Delta (N,P,T) =  - e^{\beta G}.
336 \label{introEquation:NPTPartition}
337 \end{equation}
338 Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
339
286   \subsection{\label{introSection:liouville}Liouville's theorem}
287  
288 < The Liouville's theorem is the foundation on which statistical
289 < mechanics rests. It describes the time evolution of phase space
288 > Liouville's theorem is the foundation on which statistical mechanics
289 > rests. It describes the time evolution of the phase space
290   distribution function. In order to calculate the rate of change of
291 < $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
292 < consider the two faces perpendicular to the $q_1$ axis, which are
293 < located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
294 < leaving the opposite face is given by the expression,
291 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
292 > the two faces perpendicular to the $q_1$ axis, which are located at
293 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
294 > opposite face is given by the expression,
295   \begin{equation}
296   \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
297   \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
# Line 369 | Line 315 | divining $ \delta q_1  \ldots \delta q_f \delta p_1  \
315   + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
316   \end{equation}
317   which cancels the first terms of the right hand side. Furthermore,
318 < divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
318 > dividing $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
319   p_f $ in both sides, we can write out Liouville's theorem in a
320   simple form,
321   \begin{equation}
# Line 378 | Line 324 | simple form,
324   \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
325   \label{introEquation:liouvilleTheorem}
326   \end{equation}
381
327   Liouville's theorem states that the distribution function is
328   constant along any trajectory in phase space. In classical
329 < statistical mechanics, since the number of particles in the system
330 < is huge, we may be able to believe the system is stationary,
329 > statistical mechanics, since the number of system copies in an
330 > ensemble is huge and constant, we can assume the local density has
331 > no reason (other than classical mechanics) to change,
332   \begin{equation}
333   \frac{{\partial \rho }}{{\partial t}} = 0.
334   \label{introEquation:stationary}
# Line 395 | Line 341 | distribution,
341   \label{introEquation:densityAndHamiltonian}
342   \end{equation}
343  
344 < \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
344 > \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
345   Lets consider a region in the phase space,
346   \begin{equation}
347   \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
348   \end{equation}
349   If this region is small enough, the density $\rho$ can be regarded
350 < as uniform over the whole phase space. Thus, the number of phase
351 < points inside this region is given by,
350 > as uniform over the whole integral. Thus, the number of phase points
351 > inside this region is given by,
352   \begin{equation}
353   \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
354   dp_1 } ..dp_f.
# Line 412 | Line 358 | With the help of stationary assumption
358   \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
359   \frac{d}{{dt}}(\delta v) = 0.
360   \end{equation}
361 < With the help of stationary assumption
362 < (\ref{introEquation:stationary}), we obtain the principle of the
363 < \emph{conservation of extension in phase space},
361 > With the help of the stationary assumption
362 > (Eq.~\ref{introEquation:stationary}), we obtain the principle of
363 > \emph{conservation of volume in phase space},
364   \begin{equation}
365   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
366   ...dq_f dp_1 } ..dp_f  = 0.
367   \label{introEquation:volumePreserving}
368   \end{equation}
369  
370 < \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
370 > \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
371  
372 < Liouville's theorem can be expresses in a variety of different forms
372 > Liouville's theorem can be expressed in a variety of different forms
373   which are convenient within different contexts. For any two function
374   $F$ and $G$ of the coordinates and momenta of a system, the Poisson
375   bracket ${F, G}$ is defined as
# Line 434 | Line 380 | Substituting equations of motion in Hamiltonian formal
380   q_i }}} \right)}.
381   \label{introEquation:poissonBracket}
382   \end{equation}
383 < Substituting equations of motion in Hamiltonian formalism(
384 < \ref{introEquation:motionHamiltonianCoordinate} ,
385 < \ref{introEquation:motionHamiltonianMomentum} ) into
386 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
387 < theorem using Poisson bracket notion,
383 > Substituting equations of motion in Hamiltonian formalism
384 > (Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
385 > Eq.~\ref{introEquation:motionHamiltonianMomentum}) into
386 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
387 > Liouville's theorem using Poisson bracket notion,
388   \begin{equation}
389   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
390   {\rho ,H} \right\}.
# Line 457 | Line 403 | expressed as
403   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
404   \label{introEquation:liouvilleTheoremInOperator}
405   \end{equation}
406 <
406 > which can help define a propagator $\rho (t) = e^{-iLt} \rho (0)$.
407   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
408  
409   Various thermodynamic properties can be calculated from Molecular
410   Dynamics simulation. By comparing experimental values with the
411   calculated properties, one can determine the accuracy of the
412 < simulation and the quality of the underlying model. However, both of
413 < experiment and computer simulation are usually performed during a
412 > simulation and the quality of the underlying model. However, both
413 > experiments and computer simulations are usually performed during a
414   certain time interval and the measurements are averaged over a
415 < period of them which is different from the average behavior of
416 < many-body system in Statistical Mechanics. Fortunately, Ergodic
417 < Hypothesis is proposed to make a connection between time average and
418 < ensemble average. It states that time average and average over the
419 < statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
415 > period of time which is different from the average behavior of
416 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
417 > Hypothesis makes a connection between time average and the ensemble
418 > average. It states that the time average and average over the
419 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}:
420   \begin{equation}
421   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
422   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 479 | Line 425 | sufficiently long time (longer than relaxation time),
425   where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
426   physical quantity and $\rho (p(t), q(t))$ is the equilibrium
427   distribution function. If an observation is averaged over a
428 < sufficiently long time (longer than relaxation time), all accessible
429 < microstates in phase space are assumed to be equally probed, giving
430 < a properly weighted statistical average. This allows the researcher
431 < freedom of choice when deciding how best to measure a given
432 < observable. In case an ensemble averaged approach sounds most
433 < reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
428 > sufficiently long time (longer than the relaxation time), all
429 > accessible microstates in phase space are assumed to be equally
430 > probed, giving a properly weighted statistical average. This allows
431 > the researcher freedom of choice when deciding how best to measure a
432 > given observable. In case an ensemble averaged approach sounds most
433 > reasonable, the Monte Carlo methods\cite{Metropolis1949} can be
434   utilized. Or if the system lends itself to a time averaging
435   approach, the Molecular Dynamics techniques in
436   Sec.~\ref{introSection:molecularDynamics} will be the best
437   choice\cite{Frenkel1996}.
438  
439   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
440 < A variety of numerical integrators were proposed to simulate the
441 < motions. They usually begin with an initial conditionals and move
442 < the objects in the direction governed by the differential equations.
443 < However, most of them ignore the hidden physical law contained
444 < within the equations. Since 1990, geometric integrators, which
445 < preserve various phase-flow invariants such as symplectic structure,
446 < volume and time reversal symmetry, are developed to address this
447 < issue\cite{}. The velocity verlet method, which happens to be a
448 < simple example of symplectic integrator, continues to gain its
449 < popularity in molecular dynamics community. This fact can be partly
450 < explained by its geometric nature.
440 > A variety of numerical integrators have been proposed to simulate
441 > the motions of atoms in MD simulation. They usually begin with
442 > initial conditionals and move the objects in the direction governed
443 > by the differential equations. However, most of them ignore the
444 > hidden physical laws contained within the equations. Since 1990,
445 > geometric integrators, which preserve various phase-flow invariants
446 > such as symplectic structure, volume and time reversal symmetry,
447 > were developed to address this issue\cite{Dullweber1997,
448 > McLachlan1998, Leimkuhler1999}. The velocity Verlet method, which
449 > happens to be a simple example of symplectic integrator, continues
450 > to gain popularity in the molecular dynamics community. This fact
451 > can be partly explained by its geometric nature.
452  
453 < \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
454 < A \emph{manifold} is an abstract mathematical space. It locally
455 < looks like Euclidean space, but when viewed globally, it may have
456 < more complicate structure. A good example of manifold is the surface
457 < of Earth. It seems to be flat locally, but it is round if viewed as
458 < a whole. A \emph{differentiable manifold} (also known as
459 < \emph{smooth manifold}) is a manifold with an open cover in which
460 < the covering neighborhoods are all smoothly isomorphic to one
461 < another. In other words,it is possible to apply calculus on
515 < \emph{differentiable manifold}. A \emph{symplectic manifold} is
516 < defined as a pair $(M, \omega)$ which consisting of a
453 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
454 > A \emph{manifold} is an abstract mathematical space. It looks
455 > locally like Euclidean space, but when viewed globally, it may have
456 > more complicated structure. A good example of manifold is the
457 > surface of Earth. It seems to be flat locally, but it is round if
458 > viewed as a whole. A \emph{differentiable manifold} (also known as
459 > \emph{smooth manifold}) is a manifold on which it is possible to
460 > apply calculus\cite{Hirsch1997}. A \emph{symplectic manifold} is
461 > defined as a pair $(M, \omega)$ which consists of a
462   \emph{differentiable manifold} $M$ and a close, non-degenerated,
463   bilinear symplectic form, $\omega$. A symplectic form on a vector
464   space $V$ is a function $\omega(x, y)$ which satisfies
465   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
466   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
467 < $\omega(x, x) = 0$. Cross product operation in vector field is an
468 < example of symplectic form.
467 > $\omega(x, x) = 0$\cite{McDuff1998}. The cross product operation in
468 > vector field is an example of symplectic form. One of the
469 > motivations to study \emph{symplectic manifolds} in Hamiltonian
470 > Mechanics is that a symplectic manifold can represent all possible
471 > configurations of the system and the phase space of the system can
472 > be described by it's cotangent bundle\cite{Jost2002}. Every
473 > symplectic manifold is even dimensional. For instance, in Hamilton
474 > equations, coordinate and momentum always appear in pairs.
475  
525 One of the motivations to study \emph{symplectic manifold} in
526 Hamiltonian Mechanics is that a symplectic manifold can represent
527 all possible configurations of the system and the phase space of the
528 system can be described by it's cotangent bundle. Every symplectic
529 manifold is even dimensional. For instance, in Hamilton equations,
530 coordinate and momentum always appear in pairs.
531
532 Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
533 \[
534 f : M \rightarrow N
535 \]
536 is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
537 the \emph{pullback} of $\eta$ under f is equal to $\omega$.
538 Canonical transformation is an example of symplectomorphism in
539 classical mechanics.
540
476   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
477  
478 < For a ordinary differential system defined as
478 > For an ordinary differential system defined as
479   \begin{equation}
480   \dot x = f(x)
481   \end{equation}
482 < where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
482 > where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
483 > $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
484 > function and $J$ is the skew-symmetric matrix
485   \begin{equation}
549 f(r) = J\nabla _x H(r).
550 \end{equation}
551 $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
552 matrix
553 \begin{equation}
486   J = \left( {\begin{array}{*{20}c}
487     0 & I  \\
488     { - I} & 0  \\
# Line 560 | Line 492 | system can be rewritten as,
492   where $I$ is an identity matrix. Using this notation, Hamiltonian
493   system can be rewritten as,
494   \begin{equation}
495 < \frac{d}{{dt}}x = J\nabla _x H(x)
495 > \frac{d}{{dt}}x = J\nabla _x H(x).
496   \label{introEquation:compactHamiltonian}
497   \end{equation}In this case, $f$ is
498 < called a \emph{Hamiltonian vector field}.
499 <
568 < Another generalization of Hamiltonian dynamics is Poisson Dynamics,
498 > called a \emph{Hamiltonian vector field}. Another generalization of
499 > Hamiltonian dynamics is Poisson Dynamics\cite{Olver1986},
500   \begin{equation}
501   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
502   \end{equation}
503   The most obvious change being that matrix $J$ now depends on $x$.
504  
505 < \subsection{\label{introSection:exactFlow}Exact Flow}
505 > \subsection{\label{introSection:exactFlow}Exact Propagator}
506  
507 < Let $x(t)$ be the exact solution of the ODE system,
508 < \begin{equation}
509 < \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
510 < \end{equation}
511 < The exact flow(solution) $\varphi_\tau$ is defined by
581 < \[
582 < x(t+\tau) =\varphi_\tau(x(t))
507 > Let $x(t)$ be the exact solution of the ODE
508 > system,$\frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}$, we can
509 > define its exact propagator(solution) $\varphi_\tau$
510 > \[ x(t+\tau)
511 > =\varphi_\tau(x(t))
512   \]
513   where $\tau$ is a fixed time step and $\varphi$ is a map from phase
514 < space to itself. The flow has the continuous group property,
514 > space to itself. The propagator has the continuous group property,
515   \begin{equation}
516   \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
517   + \tau _2 } .
# Line 591 | Line 520 | Therefore, the exact flow is self-adjoint,
520   \begin{equation}
521   \varphi _\tau   \circ \varphi _{ - \tau }  = I
522   \end{equation}
523 < Therefore, the exact flow is self-adjoint,
523 > Therefore, the exact propagator is self-adjoint,
524   \begin{equation}
525   \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
526   \end{equation}
527 < The exact flow can also be written in terms of the of an operator,
527 > The exact propagator can also be written in terms of operator,
528   \begin{equation}
529   \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
530   }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
531   \label{introEquation:exponentialOperator}
532   \end{equation}
533 <
534 < In most cases, it is not easy to find the exact flow $\varphi_\tau$.
535 < Instead, we use a approximate map, $\psi_\tau$, which is usually
536 < called integrator. The order of an integrator $\psi_\tau$ is $p$, if
537 < the Taylor series of $\psi_\tau$ agree to order $p$,
533 > In most cases, it is not easy to find the exact propagator
534 > $\varphi_\tau$. Instead, we use an approximate map, $\psi_\tau$,
535 > which is usually called an integrator. The order of an integrator
536 > $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
537 > order $p$,
538   \begin{equation}
539 < \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
539 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
540   \end{equation}
541  
542   \subsection{\label{introSection:geometricProperties}Geometric Properties}
543  
544 < The hidden geometric properties of ODE and its flow play important
545 < roles in numerical studies. Many of them can be found in systems
546 < which occur naturally in applications.
547 <
548 < Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
620 < a \emph{symplectic} flow if it satisfies,
544 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
545 > ODE and its propagator play important roles in numerical studies.
546 > Many of them can be found in systems which occur naturally in
547 > applications. Let $\varphi$ be the propagator of Hamiltonian vector
548 > field, $\varphi$ is a \emph{symplectic} propagator if it satisfies,
549   \begin{equation}
550   {\varphi '}^T J \varphi ' = J.
551   \end{equation}
552   According to Liouville's theorem, the symplectic volume is invariant
553 < under a Hamiltonian flow, which is the basis for classical
554 < statistical mechanics. Furthermore, the flow of a Hamiltonian vector
555 < field on a symplectic manifold can be shown to be a
553 > under a Hamiltonian propagator, which is the basis for classical
554 > statistical mechanics. Furthermore, the propagator of a Hamiltonian
555 > vector field on a symplectic manifold can be shown to be a
556   symplectomorphism. As to the Poisson system,
557   \begin{equation}
558   {\varphi '}^T J \varphi ' = J \circ \varphi
559   \end{equation}
560 < is the property must be preserved by the integrator.
561 <
562 < It is possible to construct a \emph{volume-preserving} flow for a
563 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
564 < \det d\varphi  = 1$. One can show easily that a symplectic flow will
565 < be volume-preserving.
566 <
639 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
640 < will result in a new system,
560 > is the property that must be preserved by the integrator. It is
561 > possible to construct a \emph{volume-preserving} propagator for a
562 > source free ODE ($ \nabla \cdot f = 0 $), if the propagator
563 > satisfies $ \det d\varphi  = 1$. One can show easily that a
564 > symplectic propagator will be volume-preserving. Changing the
565 > variables $y = h(x)$ in an ODE (Eq.~\ref{introEquation:ODE}) will
566 > result in a new system,
567   \[
568   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
569   \]
570   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
571 < In other words, the flow of this vector field is reversible if and
572 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
573 <
574 < A \emph{first integral}, or conserved quantity of a general
575 < differential function is a function $ G:R^{2d}  \to R^d $ which is
650 < constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
571 > In other words, the propagator of this vector field is reversible if
572 > and only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
573 > conserved quantity of a general differential function is a function
574 > $ G:R^{2d}  \to R^d $ which is constant for all solutions of the ODE
575 > $\frac{{dx}}{{dt}} = f(x)$ ,
576   \[
577   \frac{{dG(x(t))}}{{dt}} = 0.
578   \]
579 < Using chain rule, one may obtain,
579 > Using the chain rule, one may obtain,
580   \[
581 < \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
581 > \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \dot \nabla G,
582   \]
583 < which is the condition for conserving \emph{first integral}. For a
584 < canonical Hamiltonian system, the time evolution of an arbitrary
585 < smooth function $G$ is given by,
586 < \begin{equation}
587 < \begin{array}{c}
588 < \frac{{dG(x(t))}}{{dt}} = [\nabla _x G(x(t))]^T \dot x(t) \\
664 <  = [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
665 < \end{array}
583 > which is the condition for conserved quantities. For a canonical
584 > Hamiltonian system, the time evolution of an arbitrary smooth
585 > function $G$ is given by,
586 > \begin{eqnarray}
587 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \notag\\
588 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)).
589   \label{introEquation:firstIntegral1}
590 < \end{equation}
591 < Using poisson bracket notion, Equation
592 < \ref{introEquation:firstIntegral1} can be rewritten as
590 > \end{eqnarray}
591 > Using poisson bracket notion, Eq.~\ref{introEquation:firstIntegral1}
592 > can be rewritten as
593   \[
594   \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
595   \]
596 < Therefore, the sufficient condition for $G$ to be the \emph{first
597 < integral} of a Hamiltonian system is
598 < \[
599 < \left\{ {G,H} \right\} = 0.
600 < \]
601 < As well known, the Hamiltonian (or energy) H of a Hamiltonian system
602 < is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
680 < 0$.
596 > Therefore, the sufficient condition for $G$ to be a conserved
597 > quantity of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$ As
598 > is well known, the Hamiltonian (or energy) H of a Hamiltonian system
599 > is a conserved quantity, which is due to the fact $\{ H,H\}  = 0$.
600 > When designing any numerical methods, one should always try to
601 > preserve the structural properties of the original ODE and its
602 > propagator.
603  
682
683 When designing any numerical methods, one should always try to
684 preserve the structural properties of the original ODE and its flow.
685
604   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
605   A lot of well established and very effective numerical methods have
606 < been successful precisely because of their symplecticities even
606 > been successful precisely because of their symplectic nature even
607   though this fact was not recognized when they were first
608 < constructed. The most famous example is leapfrog methods in
609 < molecular dynamics. In general, symplectic integrators can be
608 > constructed. The most famous example is the Verlet-leapfrog method
609 > in molecular dynamics. In general, symplectic integrators can be
610   constructed using one of four different methods.
611   \begin{enumerate}
612   \item Generating functions
# Line 696 | Line 614 | constructed using one of four different methods.
614   \item Runge-Kutta methods
615   \item Splitting methods
616   \end{enumerate}
617 <
618 < Generating function tends to lead to methods which are cumbersome
619 < and difficult to use. In dissipative systems, variational methods
620 < can capture the decay of energy accurately. Since their
621 < geometrically unstable nature against non-Hamiltonian perturbations,
622 < ordinary implicit Runge-Kutta methods are not suitable for
623 < Hamiltonian system. Recently, various high-order explicit
624 < Runge--Kutta methods have been developed to overcome this
625 < instability. However, due to computational penalty involved in
626 < implementing the Runge-Kutta methods, they do not attract too much
627 < attention from Molecular Dynamics community. Instead, splitting have
628 < been widely accepted since they exploit natural decompositions of
629 < the system\cite{Tuckerman1992}.
617 > Generating functions\cite{Channell1990} tend to lead to methods
618 > which are cumbersome and difficult to use. In dissipative systems,
619 > variational methods can capture the decay of energy
620 > accurately\cite{Kane2000}. Since they are geometrically unstable
621 > against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
622 > methods are not suitable for Hamiltonian system. Recently, various
623 > high-order explicit Runge-Kutta methods \cite{Owren1992,Chen2003}
624 > have been developed to overcome this instability. However, due to
625 > computational penalty involved in implementing the Runge-Kutta
626 > methods, they have not attracted much attention from the Molecular
627 > Dynamics community. Instead, splitting methods have been widely
628 > accepted since they exploit natural decompositions of the
629 > system\cite{Tuckerman1992, McLachlan1998}.
630  
631 < \subsubsection{\label{introSection:splittingMethod}Splitting Method}
631 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
632  
633   The main idea behind splitting methods is to decompose the discrete
634 < $\varphi_h$ as a composition of simpler flows,
634 > $\varphi_h$ as a composition of simpler propagators,
635   \begin{equation}
636   \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
637   \varphi _{h_n }
638   \label{introEquation:FlowDecomposition}
639   \end{equation}
640 < where each of the sub-flow is chosen such that each represent a
641 < simpler integration of the system.
642 <
725 < Suppose that a Hamiltonian system takes the form,
640 > where each of the sub-propagator is chosen such that each represent
641 > a simpler integration of the system. Suppose that a Hamiltonian
642 > system takes the form,
643   \[
644   H = H_1 + H_2.
645   \]
646   Here, $H_1$ and $H_2$ may represent different physical processes of
647   the system. For instance, they may relate to kinetic and potential
648   energy respectively, which is a natural decomposition of the
649 < problem. If $H_1$ and $H_2$ can be integrated using exact flows
650 < $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
651 < order is then given by the Lie-Trotter formula
649 > problem. If $H_1$ and $H_2$ can be integrated using exact
650 > propagators $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a
651 > simple first order expression is then given by the Lie-Trotter
652 > formula
653   \begin{equation}
654   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
655   \label{introEquation:firstOrderSplitting}
# Line 740 | Line 658 | It is easy to show that any composition of symplectic
658   continuous $\varphi _i$ over a time $h$. By definition, as
659   $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
660   must follow that each operator $\varphi_i(t)$ is a symplectic map.
661 < It is easy to show that any composition of symplectic flows yields a
662 < symplectic map,
661 > It is easy to show that any composition of symplectic propagators
662 > yields a symplectic map,
663   \begin{equation}
664   (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
665   '\phi ' = \phi '^T J\phi ' = J,
# Line 749 | Line 667 | splitting in this context automatically generates a sy
667   \end{equation}
668   where $\phi$ and $\psi$ both are symplectic maps. Thus operator
669   splitting in this context automatically generates a symplectic map.
670 <
671 < The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
672 < introduces local errors proportional to $h^2$, while Strang
673 < splitting gives a second-order decomposition,
670 > The Lie-Trotter
671 > splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
672 > local errors proportional to $h^2$, while the Strang splitting gives
673 > a second-order decomposition,
674   \begin{equation}
675   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
676   _{1,h/2} , \label{introEquation:secondOrderSplitting}
677   \end{equation}
678 < which has a local error proportional to $h^3$. Sprang splitting's
679 < popularity in molecular simulation community attribute to its
680 < symmetric property,
678 > which has a local error proportional to $h^3$. The Strang
679 > splitting's popularity in molecular simulation community attribute
680 > to its symmetric property,
681   \begin{equation}
682   \varphi _h^{ - 1} = \varphi _{ - h}.
683   \label{introEquation:timeReversible}
684   \end{equation}
685  
686 < \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
686 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
687   The classical equation for a system consisting of interacting
688   particles can be written in Hamiltonian form,
689   \[
690   H = T + V
691   \]
692   where $T$ is the kinetic energy and $V$ is the potential energy.
693 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
693 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
694   obtains the following:
695   \begin{align}
696   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 799 | Line 717 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
717      \label{introEquation:Lp9b}\\%
718   %
719   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
720 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
720 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
721   \end{align}
722   From the preceding splitting, one can see that the integration of
723   the equations of motion would follow:
# Line 808 | Line 726 | the equations of motion would follow:
726  
727   \item Use the half step velocities to move positions one whole step, $\Delta t$.
728  
729 < \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
729 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
730  
731   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
732   \end{enumerate}
733 <
734 < Simply switching the order of splitting and composing, a new
735 < integrator, the \emph{position verlet} integrator, can be generated,
733 > By simply switching the order of the propagators in the splitting
734 > and composing a new integrator, the \emph{position verlet}
735 > integrator, can be generated,
736   \begin{align}
737   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
738   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 825 | Line 743 | q(\Delta t)} \right]. %
743   \label{introEquation:positionVerlet2}
744   \end{align}
745  
746 < \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
746 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
747  
748 < Baker-Campbell-Hausdorff formula can be used to determine the local
749 < error of splitting method in terms of commutator of the
748 > The Baker-Campbell-Hausdorff formula can be used to determine the
749 > local error of a splitting method in terms of the commutator of the
750   operators(\ref{introEquation:exponentialOperator}) associated with
751 < the sub-flow. For operators $hX$ and $hY$ which are associate to
752 < $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
751 > the sub-propagator. For operators $hX$ and $hY$ which are associated
752 > with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
753   \begin{equation}
754   \exp (hX + hY) = \exp (hZ)
755   \end{equation}
# Line 840 | Line 758 | Here, $[X,Y]$ is the commutators of operator $X$ and $
758   hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
759   {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
760   \end{equation}
761 < Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
761 > Here, $[X,Y]$ is the commutator of operator $X$ and $Y$ given by
762   \[
763   [X,Y] = XY - YX .
764   \]
765 < Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
766 < can obtain
765 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
766 > to the Strang splitting, we can obtain
767   \begin{eqnarray*}
768   \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 \\
769                                     &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
770 <                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
770 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots
771 >                                   ).
772   \end{eqnarray*}
773 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
774 < error of Spring splitting is proportional to $h^3$. The same
775 < procedure can be applied to general splitting,  of the form
773 > Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local
774 > error of Strang splitting is proportional to $h^3$. The same
775 > procedure can be applied to a general splitting of the form
776   \begin{equation}
777   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
778   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
779   \end{equation}
780 < Careful choice of coefficient $a_1 \ldot b_m$ will lead to higher
781 < order method. Yoshida proposed an elegant way to compose higher
782 < order methods based on symmetric splitting. Given a symmetric second
783 < order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
784 < method can be constructed by composing,
780 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
781 > order methods. Yoshida proposed an elegant way to compose higher
782 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
783 > a symmetric second order base method $ \varphi _h^{(2)} $, a
784 > fourth-order symmetric method can be constructed by composing,
785   \[
786   \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
787   h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
# Line 872 | Line 791 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
791   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
792   \begin{equation}
793   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
794 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
794 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
795   \end{equation}
796 < , if the weights are chosen as
796 > if the weights are chosen as
797   \[
798   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
799   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 888 | Line 807 | simulations. For instance, instantaneous temperature o
807   dynamical information. The basic idea of molecular dynamics is that
808   macroscopic properties are related to microscopic behavior and
809   microscopic behavior can be calculated from the trajectories in
810 < simulations. For instance, instantaneous temperature of an
811 < Hamiltonian system of $N$ particle can be measured by
810 > simulations. For instance, instantaneous temperature of a
811 > Hamiltonian system of $N$ particles can be measured by
812   \[
813   T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
814   \]
815   where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
816   respectively, $f$ is the number of degrees of freedom, and $k_B$ is
817 < the boltzman constant.
817 > the Boltzman constant.
818  
819   A typical molecular dynamics run consists of three essential steps:
820   \begin{enumerate}
# Line 911 | Line 830 | initialization of a simulation. Sec.~\ref{introSec:pro
830   \end{enumerate}
831   These three individual steps will be covered in the following
832   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
833 < initialization of a simulation. Sec.~\ref{introSec:production} will
834 < discusses issues in production run. Sec.~\ref{introSection:Analysis}
835 < provides the theoretical tools for trajectory analysis.
833 > initialization of a simulation. Sec.~\ref{introSection:production}
834 > will discuss issues of production runs.
835 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
836 > analysis of trajectories.
837  
838   \subsection{\label{introSec:initialSystemSettings}Initialization}
839  
840 < \subsubsection{Preliminary preparation}
840 > \subsubsection{\textbf{Preliminary preparation}}
841  
842   When selecting the starting structure of a molecule for molecular
843   simulation, one may retrieve its Cartesian coordinates from public
844   databases, such as RCSB Protein Data Bank \textit{etc}. Although
845   thousands of crystal structures of molecules are discovered every
846   year, many more remain unknown due to the difficulties of
847 < purification and crystallization. Even for the molecule with known
848 < structure, some important information is missing. For example, the
847 > purification and crystallization. Even for molecules with known
848 > structures, some important information is missing. For example, a
849   missing hydrogen atom which acts as donor in hydrogen bonding must
850 < be added. Moreover, in order to include electrostatic interaction,
850 > be added. Moreover, in order to include electrostatic interactions,
851   one may need to specify the partial charges for individual atoms.
852   Under some circumstances, we may even need to prepare the system in
853 < a special setup. For instance, when studying transport phenomenon in
854 < membrane system, we may prepare the lipids in bilayer structure
855 < instead of placing lipids randomly in solvent, since we are not
856 < interested in self-aggregation and it takes a long time to happen.
853 > a special configuration. For instance, when studying transport
854 > phenomenon in membrane systems, we may prepare the lipids in a
855 > bilayer structure instead of placing lipids randomly in solvent,
856 > since we are not interested in the slow self-aggregation process.
857  
858 < \subsubsection{Minimization}
858 > \subsubsection{\textbf{Minimization}}
859  
860   It is quite possible that some of molecules in the system from
861 < preliminary preparation may be overlapped with each other. This
862 < close proximity leads to high potential energy which consequently
863 < jeopardizes any molecular dynamics simulations. To remove these
864 < steric overlaps, one typically performs energy minimization to find
865 < a more reasonable conformation. Several energy minimization methods
866 < have been developed to exploit the energy surface and to locate the
867 < local minimum. While converging slowly near the minimum, steepest
868 < descent method is extremely robust when systems are far from
869 < harmonic. Thus, it is often used to refine structure from
870 < crystallographic data. Relied on the gradient or hessian, advanced
871 < methods like conjugate gradient and Newton-Raphson converge rapidly
872 < to a local minimum, while become unstable if the energy surface is
873 < far from quadratic. Another factor must be taken into account, when
861 > preliminary preparation may be overlapping with each other. This
862 > close proximity leads to high initial potential energy which
863 > consequently jeopardizes any molecular dynamics simulations. To
864 > remove these steric overlaps, one typically performs energy
865 > minimization to find a more reasonable conformation. Several energy
866 > minimization methods have been developed to exploit the energy
867 > surface and to locate the local minimum. While converging slowly
868 > near the minimum, steepest descent method is extremely robust when
869 > systems are strongly anharmonic. Thus, it is often used to refine
870 > structures from crystallographic data. Relying on the Hessian,
871 > advanced methods like Newton-Raphson converge rapidly to a local
872 > minimum, but become unstable if the energy surface is far from
873 > quadratic. Another factor that must be taken into account, when
874   choosing energy minimization method, is the size of the system.
875   Steepest descent and conjugate gradient can deal with models of any
876 < size. Because of the limit of computation power to calculate hessian
877 < matrix and insufficient storage capacity to store them, most
878 < Newton-Raphson methods can not be used with very large models.
876 > size. Because of the limits on computer memory to store the hessian
877 > matrix and the computing power needed to diagonalize these matrices,
878 > most Newton-Raphson methods can not be used with very large systems.
879  
880 < \subsubsection{Heating}
880 > \subsubsection{\textbf{Heating}}
881  
882 < Typically, Heating is performed by assigning random velocities
883 < according to a Gaussian distribution for a temperature. Beginning at
884 < a lower temperature and gradually increasing the temperature by
885 < assigning greater random velocities, we end up with setting the
886 < temperature of the system to a final temperature at which the
887 < simulation will be conducted. In heating phase, we should also keep
888 < the system from drifting or rotating as a whole. Equivalently, the
889 < net linear momentum and angular momentum of the system should be
890 < shifted to zero.
882 > Typically, heating is performed by assigning random velocities
883 > according to a Maxwell-Boltzman distribution for a desired
884 > temperature. Beginning at a lower temperature and gradually
885 > increasing the temperature by assigning larger random velocities, we
886 > end up setting the temperature of the system to a final temperature
887 > at which the simulation will be conducted. In heating phase, we
888 > should also keep the system from drifting or rotating as a whole. To
889 > do this, the net linear momentum and angular momentum of the system
890 > is shifted to zero after each resampling from the Maxwell -Boltzman
891 > distribution.
892  
893 < \subsubsection{Equilibration}
893 > \subsubsection{\textbf{Equilibration}}
894  
895   The purpose of equilibration is to allow the system to evolve
896   spontaneously for a period of time and reach equilibrium. The
# Line 983 | Line 904 | Production run is the most important steps of the simu
904  
905   \subsection{\label{introSection:production}Production}
906  
907 < Production run is the most important steps of the simulation, in
907 > The production run is the most important step of the simulation, in
908   which the equilibrated structure is used as a starting point and the
909   motions of the molecules are collected for later analysis. In order
910   to capture the macroscopic properties of the system, the molecular
911 < dynamics simulation must be performed in correct and efficient way.
911 > dynamics simulation must be performed by sampling correctly and
912 > efficiently from the relevant thermodynamic ensemble.
913  
914   The most expensive part of a molecular dynamics simulation is the
915   calculation of non-bonded forces, such as van der Waals force and
916   Coulombic forces \textit{etc}. For a system of $N$ particles, the
917   complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
918 < which making large simulations prohibitive in the absence of any
919 < computation saving techniques.
918 > which makes large simulations prohibitive in the absence of any
919 > algorithmic tricks. A natural approach to avoid system size issues
920 > is to represent the bulk behavior by a finite number of the
921 > particles. However, this approach will suffer from surface effects
922 > at the edges of the simulation. To offset this, \textit{Periodic
923 > boundary conditions} (see Fig.~\ref{introFig:pbc}) were developed to
924 > simulate bulk properties with a relatively small number of
925 > particles. In this method, the simulation box is replicated
926 > throughout space to form an infinite lattice. During the simulation,
927 > when a particle moves in the primary cell, its image in other cells
928 > move in exactly the same direction with exactly the same
929 > orientation. Thus, as a particle leaves the primary cell, one of its
930 > images will enter through the opposite face.
931 > \begin{figure}
932 > \centering
933 > \includegraphics[width=\linewidth]{pbc.eps}
934 > \caption[An illustration of periodic boundary conditions]{A 2-D
935 > illustration of periodic boundary conditions. As one particle leaves
936 > the left of the simulation box, an image of it enters the right.}
937 > \label{introFig:pbc}
938 > \end{figure}
939  
999 A natural approach to avoid system size issue is to represent the
1000 bulk behavior by a finite number of the particles. However, this
1001 approach will suffer from the surface effect. To offset this,
1002 \textit{Periodic boundary condition} is developed to simulate bulk
1003 properties with a relatively small number of particles. In this
1004 method, the simulation box is replicated throughout space to form an
1005 infinite lattice. During the simulation, when a particle moves in
1006 the primary cell, its image in other cells move in exactly the same
1007 direction with exactly the same orientation. Thus, as a particle
1008 leaves the primary cell, one of its images will enter through the
1009 opposite face.
1010 %\begin{figure}
1011 %\centering
1012 %\includegraphics[width=\linewidth]{pbcFig.eps}
1013 %\caption[An illustration of periodic boundary conditions]{A 2-D
1014 %illustration of periodic boundary conditions. As one particle leaves
1015 %the right of the simulation box, an image of it enters the left.}
1016 %\label{introFig:pbc}
1017 %\end{figure}
1018
940   %cutoff and minimum image convention
941   Another important technique to improve the efficiency of force
942 < evaluation is to apply cutoff where particles farther than a
943 < predetermined distance, are not included in the calculation
942 > evaluation is to apply spherical cutoffs where particles farther
943 > than a predetermined distance are not included in the calculation
944   \cite{Frenkel1996}. The use of a cutoff radius will cause a
945   discontinuity in the potential energy curve. Fortunately, one can
946 < shift the potential to ensure the potential curve go smoothly to
947 < zero at the cutoff radius. Cutoff strategy works pretty well for
948 < Lennard-Jones interaction because of its short range nature.
949 < However, simply truncating the electrostatic interaction with the
950 < use of cutoff has been shown to lead to severe artifacts in
951 < simulations. Ewald summation, in which the slowly conditionally
952 < convergent Coulomb potential is transformed into direct and
953 < reciprocal sums with rapid and absolute convergence, has proved to
954 < minimize the periodicity artifacts in liquid simulations. Taking the
955 < advantages of the fast Fourier transform (FFT) for calculating
956 < discrete Fourier transforms, the particle mesh-based methods are
957 < accelerated from $O(N^{3/2})$ to $O(N logN)$. An alternative
958 < approach is \emph{fast multipole method}, which treats Coulombic
959 < interaction exactly at short range, and approximate the potential at
960 < long range through multipolar expansion. In spite of their wide
961 < acceptances at the molecular simulation community, these two methods
962 < are hard to be implemented correctly and efficiently. Instead, we
963 < use a damped and charge-neutralized Coulomb potential method
964 < developed by Wolf and his coworkers. The shifted Coulomb potential
965 < for particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
946 > shift a simple radial potential to ensure the potential curve go
947 > smoothly to zero at the cutoff radius. The cutoff strategy works
948 > well for Lennard-Jones interaction because of its short range
949 > nature. However, simply truncating the electrostatic interaction
950 > with the use of cutoffs has been shown to lead to severe artifacts
951 > in simulations. The Ewald summation, in which the slowly decaying
952 > Coulomb potential is transformed into direct and reciprocal sums
953 > with rapid and absolute convergence, has proved to minimize the
954 > periodicity artifacts in liquid simulations. Taking the advantages
955 > of the fast Fourier transform (FFT) for calculating discrete Fourier
956 > transforms, the particle mesh-based
957 > methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
958 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
959 > \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
960 > which treats Coulombic interactions exactly at short range, and
961 > approximate the potential at long range through multipolar
962 > expansion. In spite of their wide acceptance at the molecular
963 > simulation community, these two methods are difficult to implement
964 > correctly and efficiently. Instead, we use a damped and
965 > charge-neutralized Coulomb potential method developed by Wolf and
966 > his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
967 > particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
968   \begin{equation}
969   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
970   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
# Line 1051 | Line 974 | efficient and easy to implement.
974   where $\alpha$ is the convergence parameter. Due to the lack of
975   inherent periodicity and rapid convergence,this method is extremely
976   efficient and easy to implement.
977 < %\begin{figure}
978 < %\centering
979 < %\includegraphics[width=\linewidth]{pbcFig.eps}
980 < %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
981 < %\label{introFigure:shiftedCoulomb}
982 < %\end{figure}
977 > \begin{figure}
978 > \centering
979 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
980 > \caption[An illustration of shifted Coulomb potential]{An
981 > illustration of shifted Coulomb potential.}
982 > \label{introFigure:shiftedCoulomb}
983 > \end{figure}
984  
985   %multiple time step
986  
987   \subsection{\label{introSection:Analysis} Analysis}
988  
989 < Recently, advanced visualization technique are widely applied to
989 > Recently, advanced visualization technique have become applied to
990   monitor the motions of molecules. Although the dynamics of the
991   system can be described qualitatively from animation, quantitative
992 < trajectory analysis are more appreciable. According to the
993 < principles of Statistical Mechanics,
992 > trajectory analysis is more useful. According to the principles of
993 > Statistical Mechanics in
994   Sec.~\ref{introSection:statisticalMechanics}, one can compute
995 < thermodynamics properties, analyze fluctuations of structural
995 > thermodynamic properties, analyze fluctuations of structural
996   parameters, and investigate time-dependent processes of the molecule
997   from the trajectories.
998  
999 < \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
999 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1000  
1001 < Thermodynamics properties, which can be expressed in terms of some
1001 > Thermodynamic properties, which can be expressed in terms of some
1002   function of the coordinates and momenta of all particles in the
1003   system, can be directly computed from molecular dynamics. The usual
1004   way to measure the pressure is based on virial theorem of Clausius
# Line 1094 | Line 1018 | P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\
1018   < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1019   \end{equation}
1020  
1021 < \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1021 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1022  
1023   Structural Properties of a simple fluid can be described by a set of
1024 < distribution functions. Among these functions,\emph{pair
1024 > distribution functions. Among these functions,the \emph{pair
1025   distribution function}, also known as \emph{radial distribution
1026 < function}, is of most fundamental importance to liquid-state theory.
1027 < Pair distribution function can be gathered by Fourier transforming
1028 < raw data from a series of neutron diffraction experiments and
1029 < integrating over the surface factor \cite{Powles1973}. The
1030 < experiment result can serve as a criterion to justify the
1031 < correctness of the theory. Moreover, various equilibrium
1032 < thermodynamic and structural properties can also be expressed in
1033 < terms of radial distribution function \cite{Allen1987}.
1034 <
1035 < A pair distribution functions $g(r)$ gives the probability that a
1036 < particle $i$ will be located at a distance $r$ from a another
1037 < particle $j$ in the system
1114 < \[
1026 > function}, is of most fundamental importance to liquid theory.
1027 > Experimentally, pair distribution functions can be gathered by
1028 > Fourier transforming raw data from a series of neutron diffraction
1029 > experiments and integrating over the surface factor
1030 > \cite{Powles1973}. The experimental results can serve as a criterion
1031 > to justify the correctness of a liquid model. Moreover, various
1032 > equilibrium thermodynamic and structural properties can also be
1033 > expressed in terms of the radial distribution function
1034 > \cite{Allen1987}. The pair distribution functions $g(r)$ gives the
1035 > probability that a particle $i$ will be located at a distance $r$
1036 > from a another particle $j$ in the system
1037 > \begin{equation}
1038   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1039 < \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1040 < \]
1039 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1040 > (r)}{\rho}.
1041 > \end{equation}
1042   Note that the delta function can be replaced by a histogram in
1043 < computer simulation. Figure
1044 < \ref{introFigure:pairDistributionFunction} shows a typical pair
1045 < distribution function for the liquid argon system. The occurrence of
1122 < several peaks in the plot of $g(r)$ suggests that it is more likely
1123 < to find particles at certain radial values than at others. This is a
1124 < result of the attractive interaction at such distances. Because of
1125 < the strong repulsive forces at short distance, the probability of
1126 < locating particles at distances less than about 2.5{\AA} from each
1127 < other is essentially zero.
1043 > computer simulation. Peaks in $g(r)$ represent solvent shells, and
1044 > the height of these peaks gradually decreases to 1 as the liquid of
1045 > large distance approaches the bulk density.
1046  
1129 %\begin{figure}
1130 %\centering
1131 %\includegraphics[width=\linewidth]{pdf.eps}
1132 %\caption[Pair distribution function for the liquid argon
1133 %]{Pair distribution function for the liquid argon}
1134 %\label{introFigure:pairDistributionFunction}
1135 %\end{figure}
1047  
1048 < \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1049 < Properties}
1048 > \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1049 > Properties}}
1050  
1051   Time-dependent properties are usually calculated using \emph{time
1052 < correlation function}, which correlates random variables $A$ and $B$
1053 < at two different time
1052 > correlation functions}, which correlate random variables $A$ and $B$
1053 > at two different times,
1054   \begin{equation}
1055   C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1056   \label{introEquation:timeCorrelationFunction}
1057   \end{equation}
1058   If $A$ and $B$ refer to same variable, this kind of correlation
1059 < function is called \emph{auto correlation function}. One example of
1060 < auto correlation function is velocity auto-correlation function
1061 < which is directly related to transport properties of molecular
1062 < liquids:
1059 > function is called an \emph{autocorrelation function}. One example
1060 > of an auto correlation function is the velocity auto-correlation
1061 > function which is directly related to transport properties of
1062 > molecular liquids:
1063   \[
1064   D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1065   \right\rangle } dt
1066   \]
1067 < where $D$ is diffusion constant. Unlike velocity autocorrelation
1068 < function which is averaging over time origins and over all the
1069 < atoms, dipole autocorrelation are calculated for the entire system.
1070 < The dipole autocorrelation function is given by:
1067 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1068 > function, which is averaged over time origins and over all the
1069 > atoms, the dipole autocorrelation functions is calculated for the
1070 > entire system. The dipole autocorrelation function is given by:
1071   \[
1072   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1073   \right\rangle
# Line 1164 | Line 1075 | u_{tot} (t) = \sum\limits_i {u_i (t)}
1075   Here $u_{tot}$ is the net dipole of the entire system and is given
1076   by
1077   \[
1078 < u_{tot} (t) = \sum\limits_i {u_i (t)}
1078 > u_{tot} (t) = \sum\limits_i {u_i (t)}.
1079   \]
1080 < In principle, many time correlation functions can be related with
1080 > In principle, many time correlation functions can be related to
1081   Fourier transforms of the infrared, Raman, and inelastic neutron
1082   scattering spectra of molecular liquids. In practice, one can
1083 < extract the IR spectrum from the intensity of dipole fluctuation at
1084 < each frequency using the following relationship:
1083 > extract the IR spectrum from the intensity of the molecular dipole
1084 > fluctuation at each frequency using the following relationship:
1085   \[
1086   \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1087 < i2\pi vt} dt}
1087 > i2\pi vt} dt}.
1088   \]
1089  
1090   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1091  
1092   Rigid bodies are frequently involved in the modeling of different
1093   areas, from engineering, physics, to chemistry. For example,
1094 < missiles and vehicle are usually modeled by rigid bodies.  The
1095 < movement of the objects in 3D gaming engine or other physics
1096 < simulator is governed by the rigid body dynamics. In molecular
1097 < simulation, rigid body is used to simplify the model in
1098 < protein-protein docking study{\cite{Gray2003}}.
1094 > missiles and vehicles are usually modeled by rigid bodies.  The
1095 > movement of the objects in 3D gaming engines or other physics
1096 > simulators is governed by rigid body dynamics. In molecular
1097 > simulations, rigid bodies are used to simplify protein-protein
1098 > docking studies\cite{Gray2003}.
1099  
1100   It is very important to develop stable and efficient methods to
1101 < integrate the equations of motion of orientational degrees of
1102 < freedom. Euler angles are the nature choice to describe the
1103 < rotational degrees of freedom. However, due to its singularity, the
1104 < numerical integration of corresponding equations of motion is very
1105 < inefficient and inaccurate. Although an alternative integrator using
1106 < different sets of Euler angles can overcome this difficulty\cite{},
1107 < the computational penalty and the lost of angular momentum
1108 < conservation still remain. A singularity free representation
1109 < utilizing quaternions was developed by Evans in 1977. Unfortunately,
1110 < this approach suffer from the nonseparable Hamiltonian resulted from
1111 < quaternion representation, which prevents the symplectic algorithm
1112 < to be utilized. Another different approach is to apply holonomic
1113 < constraints to the atoms belonging to the rigid body. Each atom
1114 < moves independently under the normal forces deriving from potential
1115 < energy and constraint forces which are used to guarantee the
1116 < rigidness. However, due to their iterative nature, SHAKE and Rattle
1117 < algorithm converge very slowly when the number of constraint
1118 < increases.
1101 > integrate the equations of motion for orientational degrees of
1102 > freedom. Euler angles are the natural choice to describe the
1103 > rotational degrees of freedom. However, due to $\frac {1}{sin
1104 > \theta}$ singularities, the numerical integration of corresponding
1105 > equations of these motion is very inefficient and inaccurate.
1106 > Although an alternative integrator using multiple sets of Euler
1107 > angles can overcome this difficulty\cite{Barojas1973}, the
1108 > computational penalty and the loss of angular momentum conservation
1109 > still remain. A singularity-free representation utilizing
1110 > quaternions was developed by Evans in 1977\cite{Evans1977}.
1111 > Unfortunately, this approach uses a nonseparable Hamiltonian
1112 > resulting from the quaternion representation, which prevents the
1113 > symplectic algorithm from being utilized. Another different approach
1114 > is to apply holonomic constraints to the atoms belonging to the
1115 > rigid body. Each atom moves independently under the normal forces
1116 > deriving from potential energy and constraint forces which are used
1117 > to guarantee the rigidness. However, due to their iterative nature,
1118 > the SHAKE and Rattle algorithms also converge very slowly when the
1119 > number of constraints increases\cite{Ryckaert1977, Andersen1983}.
1120  
1121 < The break through in geometric literature suggests that, in order to
1121 > A break-through in geometric literature suggests that, in order to
1122   develop a long-term integration scheme, one should preserve the
1123 < symplectic structure of the flow. Introducing conjugate momentum to
1124 < rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1125 < symplectic integrator, RSHAKE, was proposed to evolve the
1126 < Hamiltonian system in a constraint manifold by iteratively
1127 < satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1128 < method using quaternion representation was developed by Omelyan.
1129 < However, both of these methods are iterative and inefficient. In
1130 < this section, we will present a symplectic Lie-Poisson integrator
1131 < for rigid body developed by Dullweber and his
1132 < coworkers\cite{Dullweber1997} in depth.
1123 > symplectic structure of the propagator. By introducing a conjugate
1124 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1125 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1126 > proposed to evolve the Hamiltonian system in a constraint manifold
1127 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1128 > An alternative method using the quaternion representation was
1129 > developed by Omelyan\cite{Omelyan1998}. However, both of these
1130 > methods are iterative and inefficient. In this section, we descibe a
1131 > symplectic Lie-Poisson integrator for rigid bodies developed by
1132 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1133  
1134 < \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1135 < The motion of the rigid body is Hamiltonian with the Hamiltonian
1134 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1135 > The motion of a rigid body is Hamiltonian with the Hamiltonian
1136   function
1137   \begin{equation}
1138   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1139   V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1140   \label{introEquation:RBHamiltonian}
1141   \end{equation}
1142 < Here, $q$ and $Q$  are the position and rotation matrix for the
1143 < rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1144 < $J$, a diagonal matrix, is defined by
1142 > Here, $q$ and $Q$  are the position vector and rotation matrix for
1143 > the rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ ,
1144 > and $J$, a diagonal matrix, is defined by
1145   \[
1146   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1147   \]
1148   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1149 < constrained Hamiltonian equation subjects to a holonomic constraint,
1149 > constrained Hamiltonian equation is subjected to a holonomic
1150 > constraint,
1151   \begin{equation}
1152   Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1153   \end{equation}
1154 < which is used to ensure rotation matrix's orthogonality.
1155 < Differentiating \ref{introEquation:orthogonalConstraint} and using
1156 < Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1154 > which is used to ensure the rotation matrix's unitarity. Using
1155 > Equation (\ref{introEquation:motionHamiltonianCoordinate},
1156 > \ref{introEquation:motionHamiltonianMomentum}), one can write down
1157 > the equations of motion,
1158 > \begin{eqnarray}
1159 > \frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\
1160 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q), \label{introEquation:RBMotionMomentum}\\
1161 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1},  \label{introEquation:RBMotionRotation}\\
1162 > \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1163 > \end{eqnarray}
1164 > Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and
1165 > using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1166   \begin{equation}
1167   Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1168   \label{introEquation:RBFirstOrderConstraint}
1169   \end{equation}
1248
1249 Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1250 \ref{introEquation:motionHamiltonianMomentum}), one can write down
1251 the equations of motion,
1252 \[
1253 \begin{array}{c}
1254 \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1255 \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1256 \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1257 \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1258 \end{array}
1259 \]
1260
1170   In general, there are two ways to satisfy the holonomic constraints.
1171 < We can use constraint force provided by lagrange multiplier on the
1172 < normal manifold to keep the motion on constraint space. Or we can
1173 < simply evolve the system in constraint manifold. These two methods
1174 < are proved to be equivalent. The holonomic constraint and equations
1175 < of motions define a constraint manifold for rigid body
1171 > We can use a constraint force provided by a Lagrange multiplier on
1172 > the normal manifold to keep the motion on the constraint space. Or
1173 > we can simply evolve the system on the constraint manifold. These
1174 > two methods have been proved to be equivalent. The holonomic
1175 > constraint and equations of motions define a constraint manifold for
1176 > rigid bodies
1177   \[
1178   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1179   \right\}.
1180   \]
1181 <
1182 < Unfortunately, this constraint manifold is not the cotangent bundle
1183 < $T_{\star}SO(3)$. However, it turns out that under symplectic
1184 < transformation, the cotangent space and the phase space are
1275 < diffeomorphic. Introducing
1181 > Unfortunately, this constraint manifold is not $T^* SO(3)$ which is
1182 > a symplectic manifold on Lie rotation group $SO(3)$. However, it
1183 > turns out that under symplectic transformation, the cotangent space
1184 > and the phase space are diffeomorphic. By introducing
1185   \[
1186   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1187   \]
# Line 1282 | Line 1191 | T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \t
1191   T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1192   1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1193   \]
1285
1194   For a body fixed vector $X_i$ with respect to the center of mass of
1195   the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1196   given as
# Line 1301 | Line 1209 | respectively.
1209   \[
1210   \nabla _Q V(q,Q) = F(q,Q)X_i^t
1211   \]
1212 < respectively.
1213 <
1214 < As a common choice to describe the rotation dynamics of the rigid
1307 < body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1308 < rewrite the equations of motion,
1212 > respectively. As a common choice to describe the rotation dynamics
1213 > of the rigid body, the angular momentum on the body fixed frame $\Pi
1214 > = Q^t P$ is introduced to rewrite the equations of motion,
1215   \begin{equation}
1216   \begin{array}{l}
1217 < \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1218 < \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1217 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda,  \\
1218 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1},  \\
1219   \end{array}
1220   \label{introEqaution:RBMotionPI}
1221   \end{equation}
1222 < , as well as holonomic constraints,
1223 < \[
1224 < \begin{array}{l}
1319 < \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1320 < Q^T Q = 1 \\
1321 < \end{array}
1322 < \]
1323 <
1324 < For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1325 < so(3)^ \star$, the hat-map isomorphism,
1222 > as well as holonomic constraints $\Pi J^{ - 1}  + J^{ - 1} \Pi ^t  =
1223 > 0$ and $Q^T Q = 1$. For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a
1224 > matrix $\hat v \in so(3)^ \star$, the hat-map isomorphism,
1225   \begin{equation}
1226   v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1227   {\begin{array}{*{20}c}
# Line 1335 | Line 1234 | operations
1234   will let us associate the matrix products with traditional vector
1235   operations
1236   \[
1237 < \hat vu = v \times u
1237 > \hat vu = v \times u.
1238   \]
1239 <
1341 < Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1239 > Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew
1240   matrix,
1241 + \begin{eqnarray}
1242 + (\dot \Pi  - \dot \Pi ^T )&= &(\Pi  - \Pi ^T )(J^{ - 1} \Pi  + \Pi J^{ - 1} ) \notag \\
1243 + & & + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  -
1244 + (\Lambda  - \Lambda ^T ). \label{introEquation:skewMatrixPI}
1245 + \end{eqnarray}
1246 + Since $\Lambda$ is symmetric, the last term of
1247 + Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1248 + Lagrange multiplier $\Lambda$ is absent from the equations of
1249 + motion. This unique property eliminates the requirement of
1250 + iterations which can not be avoided in other methods\cite{Kol1997,
1251 + Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1252 + equation of motion for angular momentum in the body frame
1253   \begin{equation}
1344 (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1345 ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1346 - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1347 (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1348 \end{equation}
1349 Since $\Lambda$ is symmetric, the last term of Equation
1350 \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1351 multiplier $\Lambda$ is absent from the equations of motion. This
1352 unique property eliminate the requirement of iterations which can
1353 not be avoided in other methods\cite{}.
1354
1355 Applying hat-map isomorphism, we obtain the equation of motion for
1356 angular momentum on body frame
1357 \begin{equation}
1254   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1255   F_i (r,Q)} \right) \times X_i }.
1256   \label{introEquation:bodyAngularMotion}
# Line 1362 | Line 1258 | given by
1258   In the same manner, the equation of motion for rotation matrix is
1259   given by
1260   \[
1261 < \dot Q = Qskew(I^{ - 1} \pi )
1261 > \dot Q = Qskew(I^{ - 1} \pi ).
1262   \]
1263  
1264   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1265 < Lie-Poisson Integrator for Free Rigid Body}
1265 > Lie-Poisson Integrator for Free Rigid Bodies}
1266  
1267 < If there is not external forces exerted on the rigid body, the only
1268 < contribution to the rotational is from the kinetic potential (the
1269 < first term of \ref{ introEquation:bodyAngularMotion}). The free
1270 < rigid body is an example of Lie-Poisson system with Hamiltonian
1267 > If there are no external forces exerted on the rigid body, the only
1268 > contribution to the rotational motion is from the kinetic energy
1269 > (the first term of \ref{introEquation:bodyAngularMotion}). The free
1270 > rigid body is an example of a Lie-Poisson system with Hamiltonian
1271   function
1272   \begin{equation}
1273   T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
# Line 1384 | Line 1280 | J(\pi ) = \left( {\begin{array}{*{20}c}
1280     0 & {\pi _3 } & { - \pi _2 }  \\
1281     { - \pi _3 } & 0 & {\pi _1 }  \\
1282     {\pi _2 } & { - \pi _1 } & 0  \\
1283 < \end{array}} \right)
1283 > \end{array}} \right).
1284   \end{equation}
1285   Thus, the dynamics of free rigid body is governed by
1286   \begin{equation}
1287 < \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1287 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi ).
1288   \end{equation}
1289 <
1290 < One may notice that each $T_i^r$ in Equation
1291 < \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1396 < instance, the equations of motion due to $T_1^r$ are given by
1289 > One may notice that each $T_i^r$ in
1290 > Eq.~\ref{introEquation:rotationalKineticRB} can be solved exactly.
1291 > For instance, the equations of motion due to $T_1^r$ are given by
1292   \begin{equation}
1293   \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1294   \label{introEqaution:RBMotionSingleTerm}
1295   \end{equation}
1296 < where
1296 > with
1297   \[ R_1  = \left( {\begin{array}{*{20}c}
1298     0 & 0 & 0  \\
1299     0 & 0 & {\pi _1 }  \\
1300     0 & { - \pi _1 } & 0  \\
1301   \end{array}} \right).
1302   \]
1303 < The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1303 > The solutions of Eq.~\ref{introEqaution:RBMotionSingleTerm} is
1304   \[
1305   \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1306   Q(0)e^{\Delta tR_1 }
# Line 1419 | Line 1314 | tR_1 }$, we can use Cayley transformation,
1314   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1315   \]
1316   To reduce the cost of computing expensive functions in $e^{\Delta
1317 < tR_1 }$, we can use Cayley transformation,
1317 > tR_1 }$, we can use the Cayley transformation to obtain a
1318 > single-aixs propagator,
1319 > \begin{eqnarray*}
1320 > e^{\Delta tR_1 }  & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta
1321 > tR_1 ) \\
1322 > %
1323 > & \approx & \left( \begin{array}{ccc}
1324 > 1 & 0 & 0 \\
1325 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1326 > \theta^2 / 4} \\
1327 > 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1328 > \theta^2 / 4}
1329 > \end{array}
1330 > \right).
1331 > \end{eqnarray*}
1332 > The propagators for $T_2^r$ and $T_3^r$ can be found in the same
1333 > manner. In order to construct a second-order symplectic method, we
1334 > split the angular kinetic Hamiltonian function into five terms
1335   \[
1424 e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1425 )
1426 \]
1427 The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1428 manner.
1429
1430 In order to construct a second-order symplectic method, we split the
1431 angular kinetic Hamiltonian function can into five terms
1432 \[
1336   T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1337   ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1338 < (\pi _1 )
1339 < \].
1340 < Concatenating flows corresponding to these five terms, we can obtain
1341 < an symplectic integrator,
1338 > (\pi _1 ).
1339 > \]
1340 > By concatenating the propagators corresponding to these five terms,
1341 > we can obtain an symplectic integrator,
1342   \[
1343   \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1344   \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1345   \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1346   _1 }.
1347   \]
1445
1348   The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1349   $F(\pi )$ and $G(\pi )$ is defined by
1350   \[
1351   \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1352 < )
1352 > ).
1353   \]
1354   If the Poisson bracket of a function $F$ with an arbitrary smooth
1355   function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1356   conserved quantity in Poisson system. We can easily verify that the
1357   norm of the angular momentum, $\parallel \pi
1358 < \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1358 > \parallel$, is a \emph{Casimir}\cite{McLachlan1993}. Let$ F(\pi ) = S(\frac{{\parallel
1359   \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1360   then by the chain rule
1361   \[
1362   \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1363 < }}{2})\pi
1363 > }}{2})\pi.
1364   \]
1365 < Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1365 > Thus, $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel
1366 > \pi
1367   \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1368 < Lie-Poisson integrator is found to be extremely efficient and stable
1369 < which can be explained by the fact the small angle approximation is
1370 < used and the norm of the angular momentum is conserved.
1368 > Lie-Poisson integrator is found to be both extremely efficient and
1369 > stable. These properties can be explained by the fact the small
1370 > angle approximation is used and the norm of the angular momentum is
1371 > conserved.
1372  
1373   \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1374   Splitting for Rigid Body}
1375  
1376   The Hamiltonian of rigid body can be separated in terms of kinetic
1377 < energy and potential energy,
1378 < \[
1379 < H = T(p,\pi ) + V(q,Q)
1476 < \]
1477 < The equations of motion corresponding to potential energy and
1478 < kinetic energy are listed in the below table,
1377 > energy and potential energy,$H = T(p,\pi ) + V(q,Q)$. The equations
1378 > of motion corresponding to potential energy and kinetic energy are
1379 > listed in the below table,
1380   \begin{table}
1381 < \caption{Equations of motion due to Potential and Kinetic Energies}
1381 > \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1382   \begin{center}
1383   \begin{tabular}{|l|l|}
1384    \hline
# Line 1491 | Line 1392 | A second-order symplectic method is now obtained by th
1392   \end{tabular}
1393   \end{center}
1394   \end{table}
1395 < A second-order symplectic method is now obtained by the
1396 < composition of the flow maps,
1395 > A second-order symplectic method is now obtained by the composition
1396 > of the position and velocity propagators,
1397   \[
1398   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1399   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1400   \]
1401   Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1402 < sub-flows which corresponding to force and torque respectively,
1402 > sub-propagators which corresponding to force and torque
1403 > respectively,
1404   \[
1405   \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1406   _{\Delta t/2,\tau }.
1407   \]
1408   Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1409 < $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1410 < order inside $\varphi _{\Delta t/2,V}$ does not matter.
1411 <
1412 < Furthermore, kinetic potential can be separated to translational
1511 < kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1409 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1410 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1411 > kinetic energy can be separated to translational kinetic term, $T^t
1412 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1413   \begin{equation}
1414   T(p,\pi ) =T^t (p) + T^r (\pi ).
1415   \end{equation}
1416   where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1417 < defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1418 < corresponding flow maps are given by
1417 > defined by Eq.~\ref{introEquation:rotationalKineticRB}. Therefore,
1418 > the corresponding propagators are given by
1419   \[
1420   \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1421   _{\Delta t,T^r }.
1422   \]
1423 < Finally, we obtain the overall symplectic flow maps for free moving
1424 < rigid body
1425 < \begin{equation}
1426 < \begin{array}{c}
1427 < \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1428 <  \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 }  \\
1528 <  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1529 < \end{array}
1423 > Finally, we obtain the overall symplectic propagators for freely
1424 > moving rigid bodies
1425 > \begin{eqnarray}
1426 > \varphi _{\Delta t}  &=& \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \notag\\
1427 >  & & \circ \varphi _{\Delta t,T^t }  \circ \varphi _{\Delta t/2,\pi _1 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi _1 }  \notag\\
1428 >  & & \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .
1429   \label{introEquation:overallRBFlowMaps}
1430 < \end{equation}
1430 > \end{eqnarray}
1431  
1432   \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1433   As an alternative to newtonian dynamics, Langevin dynamics, which
1434   mimics a simple heat bath with stochastic and dissipative forces,
1435   has been applied in a variety of studies. This section will review
1436 < the theory of Langevin dynamics simulation. A brief derivation of
1437 < generalized Langevin equation will be given first. Follow that, we
1438 < will discuss the physical meaning of the terms appearing in the
1439 < equation as well as the calculation of friction tensor from
1440 < hydrodynamics theory.
1436 > the theory of Langevin dynamics. A brief derivation of generalized
1437 > Langevin equation will be given first. Following that, we will
1438 > discuss the physical meaning of the terms appearing in the equation
1439 > as well as the calculation of friction tensor from hydrodynamics
1440 > theory.
1441  
1442   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1443  
1444 < Harmonic bath model, in which an effective set of harmonic
1444 > A harmonic bath model, in which an effective set of harmonic
1445   oscillators are used to mimic the effect of a linearly responding
1446   environment, has been widely used in quantum chemistry and
1447   statistical mechanics. One of the successful applications of
1448 < Harmonic bath model is the derivation of Deriving Generalized
1449 < Langevin Dynamics. Lets consider a system, in which the degree of
1448 > Harmonic bath model is the derivation of the Generalized Langevin
1449 > Dynamics (GLE). Lets consider a system, in which the degree of
1450   freedom $x$ is assumed to couple to the bath linearly, giving a
1451   Hamiltonian of the form
1452   \begin{equation}
1453   H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1454   \label{introEquation:bathGLE}.
1455   \end{equation}
1456 < Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1457 < with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1456 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1457 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1458   \[
1459   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1460   }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
# Line 1563 | Line 1462 | the harmonic bath masses, and $\Delta U$ is bilinear s
1462   \]
1463   where the index $\alpha$ runs over all the bath degrees of freedom,
1464   $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1465 < the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1465 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1466   coupling,
1467   \[
1468   \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1469   \]
1470 < where $g_\alpha$ are the coupling constants between the bath and the
1471 < coordinate $x$. Introducing
1470 > where $g_\alpha$ are the coupling constants between the bath
1471 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1472 > Introducing
1473   \[
1474   W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1475   }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1476 < \] and combining the last two terms in Equation
1477 < \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1578 < Hamiltonian as
1476 > \]
1477 > and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as
1478   \[
1479   H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1480   {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1481   w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1482 < w_\alpha ^2 }}x} \right)^2 } \right\}}
1482 > w_\alpha ^2 }}x} \right)^2 } \right\}}.
1483   \]
1484   Since the first two terms of the new Hamiltonian depend only on the
1485   system coordinates, we can get the equations of motion for
1486 < Generalized Langevin Dynamics by Hamilton's equations
1588 < \ref{introEquation:motionHamiltonianCoordinate,
1589 < introEquation:motionHamiltonianMomentum},
1486 > Generalized Langevin Dynamics by Hamilton's equations,
1487   \begin{equation}
1488   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1489   \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
# Line 1599 | Line 1496 | m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x
1496   \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1497   \label{introEquation:bathMotionGLE}
1498   \end{equation}
1602
1499   In order to derive an equation for $x$, the dynamics of the bath
1500   variables $x_\alpha$ must be solved exactly first. As an integral
1501   transform which is particularly useful in solving linear ordinary
1502 < differential equations, Laplace transform is the appropriate tool to
1503 < solve this problem. The basic idea is to transform the difficult
1502 > differential equations,the Laplace transform is the appropriate tool
1503 > to solve this problem. The basic idea is to transform the difficult
1504   differential equations into simple algebra problems which can be
1505 < solved easily. Then applying inverse Laplace transform, also known
1506 < as the Bromwich integral, we can retrieve the solutions of the
1507 < original problems.
1508 <
1613 < Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1614 < transform of f(t) is a new function defined as
1505 > solved easily. Then, by applying the inverse Laplace transform, we
1506 > can retrieve the solutions of the original problems. Let $f(t)$ be a
1507 > function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$
1508 > is a new function defined as
1509   \[
1510   L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1511   \]
1512   where  $p$ is real and  $L$ is called the Laplace Transform
1513   Operator. Below are some important properties of Laplace transform
1514 < \begin{equation}
1515 < \begin{array}{c}
1516 < L(x + y) = L(x) + L(y) \\
1517 < L(ax) = aL(x) \\
1518 < L(\dot x) = pL(x) - px(0) \\
1519 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1520 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1521 < \end{array}
1522 < \end{equation}
1523 <
1524 < Applying Laplace transform to the bath coordinates, we obtain
1525 < \[
1526 < \begin{array}{c}
1527 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1528 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1635 < \end{array}
1636 < \]
1637 < By the same way, the system coordinates become
1638 < \[
1639 < \begin{array}{c}
1640 < mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1514 > \begin{eqnarray*}
1515 > L(x + y)  & = & L(x) + L(y) \\
1516 > L(ax)     & = & aL(x) \\
1517 > L(\dot x) & = & pL(x) - px(0) \\
1518 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1519 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1520 > \end{eqnarray*}
1521 > Applying the Laplace transform to the bath coordinates, we obtain
1522 > \begin{eqnarray*}
1523 > p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) & = & - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x), \\
1524 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}. \\
1525 > \end{eqnarray*}
1526 > In the same way, the system coordinates become
1527 > \begin{eqnarray*}
1528 > mL(\ddot x) & = &
1529    - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1530 < \end{array}
1531 < \]
1644 <
1530 >  & & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}.
1531 > \end{eqnarray*}
1532   With the help of some relatively important inverse Laplace
1533   transformations:
1534   \[
# Line 1651 | Line 1538 | transformations:
1538   L(1) = \frac{1}{p} \\
1539   \end{array}
1540   \]
1541 < , we obtain
1542 < \begin{align}
1543 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1541 > we obtain
1542 > \begin{eqnarray*}
1543 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1544   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1545   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1546 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1547 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1548 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1549 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1546 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1547 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1548 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1549 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1550 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}\\
1551   %
1552 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1553 < {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1554 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1555 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1556 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1557 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1558 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1559 < (\omega _\alpha  t)} \right\}}
1560 < \end{align}
1561 <
1552 > & = & -
1553 > \frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha
1554 > = 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha
1555 > ^2 }}} \right)\cos (\omega _\alpha
1556 > t)\dot x(t - \tau )d} \tau }  \\
1557 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1558 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1559 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1560 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1561 > \end{eqnarray*}
1562   Introducing a \emph{dynamic friction kernel}
1563   \begin{equation}
1564   \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
# Line 1693 | Line 1581 | which is known as the \emph{generalized Langevin equat
1581   \end{equation}
1582   which is known as the \emph{generalized Langevin equation}.
1583  
1584 < \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1584 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1585  
1586   One may notice that $R(t)$ depends only on initial conditions, which
1587   implies it is completely deterministic within the context of a
1588   harmonic bath. However, it is easy to verify that $R(t)$ is totally
1589 < uncorrelated to $x$ and $\dot x$,
1590 < \[
1591 < \begin{array}{l}
1592 < \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1593 < \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1706 < \end{array}
1707 < \]
1708 < This property is what we expect from a truly random process. As long
1709 < as the model, which is gaussian distribution in general, chosen for
1710 < $R(t)$ is a truly random process, the stochastic nature of the GLE
1711 < still remains.
1712 <
1589 > uncorrelated to $x$ and $\dot x$,$\left\langle {x(t)R(t)}
1590 > \right\rangle  = 0, \left\langle {\dot x(t)R(t)} \right\rangle  =
1591 > 0.$ This property is what we expect from a truly random process. As
1592 > long as the model chosen for $R(t)$ was a gaussian distribution in
1593 > general, the stochastic nature of the GLE still remains.
1594   %dynamic friction kernel
1595   The convolution integral
1596   \[
# Line 1724 | Line 1605 | and Equation \ref{introEuqation:GeneralizedLangevinDyn
1605   \[
1606   \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1607   \]
1608 < and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1608 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1609   \[
1610   m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1611   \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1612   \]
1613 < which can be used to describe dynamic caging effect. The other
1614 < extreme is the bath that responds infinitely quickly to motions in
1615 < the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1616 < time:
1613 > which can be used to describe the effect of dynamic caging in
1614 > viscous solvents. The other extreme is the bath that responds
1615 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1616 > taken as a $delta$ function in time:
1617   \[
1618   \xi (t) = 2\xi _0 \delta (t)
1619   \]
# Line 1741 | Line 1622 | and Equation \ref{introEuqation:GeneralizedLangevinDyn
1622   \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1623   {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1624   \]
1625 < and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1625 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1626   \begin{equation}
1627   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1628   x(t) + R(t) \label{introEquation:LangevinEquation}
1629   \end{equation}
1630   which is known as the Langevin equation. The static friction
1631   coefficient $\xi _0$ can either be calculated from spectral density
1632 < or be determined by Stokes' law for regular shaped particles.A
1633 < briefly review on calculating friction tensor for arbitrary shaped
1632 > or be determined by Stokes' law for regular shaped particles. A
1633 > brief review on calculating friction tensors for arbitrary shaped
1634   particles is given in Sec.~\ref{introSection:frictionTensor}.
1635  
1636 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1636 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1637  
1638 < Defining a new set of coordinates,
1638 > Defining a new set of coordinates
1639   \[
1640   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1641 < ^2 }}x(0)
1642 < \],
1641 > ^2 }}x(0),
1642 > \]
1643   we can rewrite $R(T)$ as
1644   \[
1645   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1646   \]
1647   And since the $q$ coordinates are harmonic oscillators,
1648 < \[
1649 < \begin{array}{c}
1650 < \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1651 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1652 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1653 < \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1654 <  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1655 <  = kT\xi (t) \\
1775 < \end{array}
1776 < \]
1648 > \begin{eqnarray*}
1649 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1650 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1651 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1652 > \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 } }  \\
1653 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1654 >  & = &kT\xi (t)
1655 > \end{eqnarray*}
1656   Thus, we recover the \emph{second fluctuation dissipation theorem}
1657   \begin{equation}
1658   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1659 < \label{introEquation:secondFluctuationDissipation}.
1659 > \label{introEquation:secondFluctuationDissipation},
1660   \end{equation}
1661 < In effect, it acts as a constraint on the possible ways in which one
1662 < can model the random force and friction kernel.
1784 <
1785 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
1786 < Theoretically, the friction kernel can be determined using velocity
1787 < autocorrelation function. However, this approach become impractical
1788 < when the system become more and more complicate. Instead, various
1789 < approaches based on hydrodynamics have been developed to calculate
1790 < the friction coefficients. The friction effect is isotropic in
1791 < Equation, $\zeta$ can be taken as a scalar. In general, friction
1792 < tensor $\Xi$ is a $6\times 6$ matrix given by
1793 < \[
1794 < \Xi  = \left( {\begin{array}{*{20}c}
1795 <   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1796 <   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1797 < \end{array}} \right).
1798 < \]
1799 < Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1800 < tensor and rotational resistance (friction) tensor respectively,
1801 < while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1802 < {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1803 < particle moves in a fluid, it may experience friction force or
1804 < torque along the opposite direction of the velocity or angular
1805 < velocity,
1806 < \[
1807 < \left( \begin{array}{l}
1808 < F_R  \\
1809 < \tau _R  \\
1810 < \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1811 <   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1812 <   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1813 < \end{array}} \right)\left( \begin{array}{l}
1814 < v \\
1815 < w \\
1816 < \end{array} \right)
1817 < \]
1818 < where $F_r$ is the friction force and $\tau _R$ is the friction
1819 < toque.
1820 <
1821 < \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1822 <
1823 < For a spherical particle, the translational and rotational friction
1824 < constant can be calculated from Stoke's law,
1825 < \[
1826 < \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1827 <   {6\pi \eta R} & 0 & 0  \\
1828 <   0 & {6\pi \eta R} & 0  \\
1829 <   0 & 0 & {6\pi \eta R}  \\
1830 < \end{array}} \right)
1831 < \]
1832 < and
1833 < \[
1834 < \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1835 <   {8\pi \eta R^3 } & 0 & 0  \\
1836 <   0 & {8\pi \eta R^3 } & 0  \\
1837 <   0 & 0 & {8\pi \eta R^3 }  \\
1838 < \end{array}} \right)
1839 < \]
1840 < where $\eta$ is the viscosity of the solvent and $R$ is the
1841 < hydrodynamics radius.
1842 <
1843 < Other non-spherical shape, such as cylinder and ellipsoid
1844 < \textit{etc}, are widely used as reference for developing new
1845 < hydrodynamics theory, because their properties can be calculated
1846 < exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1847 < also called a triaxial ellipsoid, which is given in Cartesian
1848 < coordinates by
1849 < \[
1850 < \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1851 < }} = 1
1852 < \]
1853 < where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1854 < due to the complexity of the elliptic integral, only the ellipsoid
1855 < with the restriction of two axes having to be equal, \textit{i.e.}
1856 < prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1857 < exactly. Introducing an elliptic integral parameter $S$ for prolate,
1858 < \[
1859 < S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1860 < } }}{b},
1861 < \]
1862 < and oblate,
1863 < \[
1864 < S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1865 < }}{a}
1866 < \],
1867 < one can write down the translational and rotational resistance
1868 < tensors
1869 < \[
1870 < \begin{array}{l}
1871 < \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1872 < \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1873 < \end{array},
1874 < \]
1875 < and
1876 < \[
1877 < \begin{array}{l}
1878 < \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1879 < \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1880 < \end{array}.
1881 < \]
1882 <
1883 < \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1884 <
1885 < Unlike spherical and other regular shaped molecules, there is not
1886 < analytical solution for friction tensor of any arbitrary shaped
1887 < rigid molecules. The ellipsoid of revolution model and general
1888 < triaxial ellipsoid model have been used to approximate the
1889 < hydrodynamic properties of rigid bodies. However, since the mapping
1890 < from all possible ellipsoidal space, $r$-space, to all possible
1891 < combination of rotational diffusion coefficients, $D$-space is not
1892 < unique\cite{Wegener1979} as well as the intrinsic coupling between
1893 < translational and rotational motion of rigid body\cite{}, general
1894 < ellipsoid is not always suitable for modeling arbitrarily shaped
1895 < rigid molecule. A number of studies have been devoted to determine
1896 < the friction tensor for irregularly shaped rigid bodies using more
1897 < advanced method\cite{} where the molecule of interest was modeled by
1898 < combinations of spheres(beads)\cite{} and the hydrodynamics
1899 < properties of the molecule can be calculated using the hydrodynamic
1900 < interaction tensor. Let us consider a rigid assembly of $N$ beads
1901 < immersed in a continuous medium. Due to hydrodynamics interaction,
1902 < the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1903 < unperturbed velocity $v_i$,
1904 < \[
1905 < v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1906 < \]
1907 < where $F_i$ is the frictional force, and $T_{ij}$ is the
1908 < hydrodynamic interaction tensor. The friction force of $i$th bead is
1909 < proportional to its ``net'' velocity
1910 < \begin{equation}
1911 < F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1912 < \label{introEquation:tensorExpression}
1913 < \end{equation}
1914 < This equation is the basis for deriving the hydrodynamic tensor. In
1915 < 1930, Oseen and Burgers gave a simple solution to Equation
1916 < \ref{introEquation:tensorExpression}
1917 < \begin{equation}
1918 < T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1919 < R_{ij}^T }}{{R_{ij}^2 }}} \right).
1920 < \label{introEquation:oseenTensor}
1921 < \end{equation}
1922 < Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1923 < A second order expression for element of different size was
1924 < introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1925 < la Torre and Bloomfield,
1926 < \begin{equation}
1927 < T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1928 < \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1929 < _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1930 < \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1931 < \label{introEquation:RPTensorNonOverlapped}
1932 < \end{equation}
1933 < Both of the Equation \ref{introEquation:oseenTensor} and Equation
1934 < \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1935 < \ge \sigma _i  + \sigma _j$. An alternative expression for
1936 < overlapping beads with the same radius, $\sigma$, is given by
1937 < \begin{equation}
1938 < T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1939 < \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1940 < \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1941 < \label{introEquation:RPTensorOverlapped}
1942 < \end{equation}
1943 <
1944 < To calculate the resistance tensor at an arbitrary origin $O$, we
1945 < construct a $3N \times 3N$ matrix consisting of $N \times N$
1946 < $B_{ij}$ blocks
1947 < \begin{equation}
1948 < B = \left( {\begin{array}{*{20}c}
1949 <   {B_{11} } &  \ldots  & {B_{1N} }  \\
1950 <    \vdots  &  \ddots  &  \vdots   \\
1951 <   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1952 < \end{array}} \right),
1953 < \end{equation}
1954 < where $B_{ij}$ is given by
1955 < \[
1956 < B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1957 < )T_{ij}
1958 < \]
1959 < where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1960 < $B$, we obtain
1961 <
1962 < \[
1963 < C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1964 <   {C_{11} } &  \ldots  & {C_{1N} }  \\
1965 <    \vdots  &  \ddots  &  \vdots   \\
1966 <   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1967 < \end{array}} \right)
1968 < \]
1969 < , which can be partitioned into $N \times N$ $3 \times 3$ block
1970 < $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1971 < \[
1972 < U_i  = \left( {\begin{array}{*{20}c}
1973 <   0 & { - z_i } & {y_i }  \\
1974 <   {z_i } & 0 & { - x_i }  \\
1975 <   { - y_i } & {x_i } & 0  \\
1976 < \end{array}} \right)
1977 < \]
1978 < where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1979 < bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1980 < arbitrary origin $O$ can be written as
1981 < \begin{equation}
1982 < \begin{array}{l}
1983 < \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1984 < \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1985 < \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1986 < \end{array}
1987 < \label{introEquation:ResistanceTensorArbitraryOrigin}
1988 < \end{equation}
1989 <
1990 < The resistance tensor depends on the origin to which they refer. The
1991 < proper location for applying friction force is the center of
1992 < resistance (reaction), at which the trace of rotational resistance
1993 < tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1994 < resistance is defined as an unique point of the rigid body at which
1995 < the translation-rotation coupling tensor are symmetric,
1996 < \begin{equation}
1997 < \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1998 < \label{introEquation:definitionCR}
1999 < \end{equation}
2000 < Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2001 < we can easily find out that the translational resistance tensor is
2002 < origin independent, while the rotational resistance tensor and
2003 < translation-rotation coupling resistance tensor depend on the
2004 < origin. Given resistance tensor at an arbitrary origin $O$, and a
2005 < vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2006 < obtain the resistance tensor at $P$ by
2007 < \begin{equation}
2008 < \begin{array}{l}
2009 < \Xi _P^{tt}  = \Xi _O^{tt}  \\
2010 < \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2011 < \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2012 < \end{array}
2013 < \label{introEquation:resistanceTensorTransformation}
2014 < \end{equation}
2015 < where
2016 < \[
2017 < U_{OP}  = \left( {\begin{array}{*{20}c}
2018 <   0 & { - z_{OP} } & {y_{OP} }  \\
2019 <   {z_i } & 0 & { - x_{OP} }  \\
2020 <   { - y_{OP} } & {x_{OP} } & 0  \\
2021 < \end{array}} \right)
2022 < \]
2023 < Using Equations \ref{introEquation:definitionCR} and
2024 < \ref{introEquation:resistanceTensorTransformation}, one can locate
2025 < the position of center of resistance,
2026 < \[
2027 < \left( \begin{array}{l}
2028 < x_{OR}  \\
2029 < y_{OR}  \\
2030 < z_{OR}  \\
2031 < \end{array} \right) = \left( {\begin{array}{*{20}c}
2032 <   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2033 <   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2034 <   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2035 < \end{array}} \right)^{ - 1} \left( \begin{array}{l}
2036 < (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2037 < (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2038 < (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2039 < \end{array} \right).
2040 < \]
2041 < where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2042 < joining center of resistance $R$ and origin $O$.
1661 > which acts as a constraint on the possible ways in which one can
1662 > model the random force and friction kernel.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines