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 2791 by tim, Mon Jun 5 21:21:27 2006 UTC vs.
Revision 2905 by tim, Wed Jun 28 19:09:25 2006 UTC

# Line 6 | Line 6 | behind classical mechanics. Firstly, One can determine
6   Closely related to Classical Mechanics, Molecular Dynamics
7   simulations are carried out by integrating the equations of motion
8   for a given system of particles. There are three fundamental ideas
9 < behind classical mechanics. Firstly, One can determine the state of
9 > behind classical mechanics. Firstly, one can determine the state of
10   a mechanical system at any time of interest; Secondly, all the
11   mechanical properties of the system at that time can be determined
12   by combining the knowledge of the properties of the system with the
# Line 17 | Line 17 | Newton¡¯s first law defines a class of inertial frames
17   \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18   The discovery of Newton's three laws of mechanics which govern the
19   motion of particles is the foundation of the classical mechanics.
20 < Newton¡¯s first law defines a class of inertial frames. Inertial
20 > Newton's first law defines a class of inertial frames. Inertial
21   frames are reference frames where a particle not interacting with
22   other bodies will move with constant speed in the same direction.
23 < With respect to inertial frames Newton¡¯s second law has the form
23 > With respect to inertial frames, Newton's second law has the form
24   \begin{equation}
25 < F = \frac {dp}{dt} = \frac {mv}{dt}
25 > F = \frac {dp}{dt} = \frac {mdv}{dt}
26   \label{introEquation:newtonSecondLaw}
27   \end{equation}
28   A point mass interacting with other bodies moves with the
29   acceleration along the direction of the force acting on it. Let
30   $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
31   $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
32 < Newton¡¯s third law states that
32 > Newton's third law states that
33   \begin{equation}
34 < F_{ij} = -F_{ji}
34 > F_{ij} = -F_{ji}.
35   \label{introEquation:newtonThirdLaw}
36   \end{equation}
37
37   Conservation laws of Newtonian Mechanics play very important roles
38   in solving mechanics problems. The linear momentum of a particle is
39   conserved if it is free or it experiences no force. The second
# Line 46 | Line 45 | N \equiv r \times F \label{introEquation:torqueDefinit
45   \end{equation}
46   The torque $\tau$ with respect to the same origin is defined to be
47   \begin{equation}
48 < N \equiv r \times F \label{introEquation:torqueDefinition}
48 > \tau \equiv r \times F \label{introEquation:torqueDefinition}
49   \end{equation}
50   Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
51   \[
# Line 59 | Line 58 | thus,
58   \]
59   thus,
60   \begin{equation}
61 < \dot L = r \times \dot p = N
61 > \dot L = r \times \dot p = \tau
62   \end{equation}
63   If there are no external torques acting on a body, the angular
64   momentum of it is conserved. The last conservation theorem state
65 < that if all forces are conservative, Energy
66 < \begin{equation}E = T + V \label{introEquation:energyConservation}
65 > that if all forces are conservative, energy is conserved,
66 > \begin{equation}E = T + V. \label{introEquation:energyConservation}
67   \end{equation}
68 < is conserved. All of these conserved quantities are
69 < important factors to determine the quality of numerical integration
70 < scheme for rigid body \cite{Dullweber1997}.
68 > All of these conserved quantities are important factors to determine
69 > the quality of numerical integration schemes for rigid bodies
70 > \cite{Dullweber1997}.
71  
72   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
73  
74 < Newtonian Mechanics suffers from two important limitations: it
75 < describes their motion in special cartesian coordinate systems.
76 < Another limitation of Newtonian mechanics becomes obvious when we
77 < try to describe systems with large numbers of particles. It becomes
78 < very difficult to predict the properties of the system by carrying
79 < out calculations involving the each individual interaction between
80 < all the particles, even if we know all of the details of the
82 < interaction. In order to overcome some of the practical difficulties
83 < which arise in attempts to apply Newton's equation to complex
84 < system, alternative procedures may be developed.
74 > Newtonian Mechanics suffers from a important limitation: motions can
75 > only be described in cartesian coordinate systems which make it
76 > impossible to predict analytically the properties of the system even
77 > if we know all of the details of the interaction. In order to
78 > overcome some of the practical difficulties which arise in attempts
79 > to apply Newton's equation to complex system, approximate numerical
80 > procedures may be developed.
81  
82 < \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
83 < Principle}
82 > \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
83 > Principle}}
84  
85   Hamilton introduced the dynamical principle upon which it is
86 < possible to base all of mechanics and, indeed, most of classical
87 < physics. Hamilton's Principle may be stated as follow,
88 <
89 < The actual trajectory, along which a dynamical system may move from
90 < one point to another within a specified time, is derived by finding
91 < the path which minimizes the time integral of the difference between
96 < the kinetic, $K$, and potential energies, $U$ \cite{Tolman1979}.
86 > possible to base all of mechanics and most of classical physics.
87 > Hamilton's Principle may be stated as follows: the actual
88 > trajectory, along which a dynamical system may move from one point
89 > to another within a specified time, is derived by finding the path
90 > which minimizes the time integral of the difference between the
91 > kinetic $K$, and potential energies $U$,
92   \begin{equation}
93 < \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
93 > \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0}.
94   \label{introEquation:halmitonianPrinciple1}
95   \end{equation}
101
96   For simple mechanical systems, where the forces acting on the
97 < different part are derivable from a potential and the velocities are
98 < small compared with that of light, the Lagrangian function $L$ can
99 < be define as the difference between the kinetic energy of the system
106 < and its potential energy,
97 > different parts are derivable from a potential, the Lagrangian
98 > function $L$ can be defined as the difference between the kinetic
99 > energy of the system and its potential energy,
100   \begin{equation}
101 < L \equiv K - U = L(q_i ,\dot q_i ) ,
101 > L \equiv K - U = L(q_i ,\dot q_i ).
102   \label{introEquation:lagrangianDef}
103   \end{equation}
104 < then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
104 > Thus, Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
105   \begin{equation}
106 < \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
106 > \delta \int_{t_1 }^{t_2 } {L dt = 0} .
107   \label{introEquation:halmitonianPrinciple2}
108   \end{equation}
109  
110 < \subsubsection{\label{introSection:equationOfMotionLagrangian}The
111 < Equations of Motion in Lagrangian Mechanics}
110 > \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
111 > Equations of Motion in Lagrangian Mechanics}}
112  
113 < For a holonomic system of $f$ degrees of freedom, the equations of
114 < motion in the Lagrangian form is
113 > For a system of $f$ degrees of freedom, the equations of motion in
114 > the Lagrangian form is
115   \begin{equation}
116   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
117   \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
# Line 132 | Line 125 | independent of generalized velocities, the generalized
125   Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
126   introduced by William Rowan Hamilton in 1833 as a re-formulation of
127   classical mechanics. If the potential energy of a system is
128 < independent of generalized velocities, the generalized momenta can
136 < be defined as
128 > independent of velocities, the momenta can be defined as
129   \begin{equation}
130   p_i = \frac{\partial L}{\partial \dot q_i}
131   \label{introEquation:generalizedMomenta}
# Line 143 | Line 135 | p_i  = \frac{{\partial L}}{{\partial q_i }}
135   p_i  = \frac{{\partial L}}{{\partial q_i }}
136   \label{introEquation:generalizedMomentaDot}
137   \end{equation}
146
138   With the help of the generalized momenta, we may now define a new
139   quantity $H$ by the equation
140   \begin{equation}
# Line 151 | Line 142 | $L$ is the Lagrangian function for the system.
142   \label{introEquation:hamiltonianDefByLagrangian}
143   \end{equation}
144   where $ \dot q_1  \ldots \dot q_f $ are generalized velocities and
145 < $L$ is the Lagrangian function for the system.
146 <
156 < Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
157 < one can obtain
145 > $L$ is the Lagrangian function for the system. Differentiating
146 > Eq.~\ref{introEquation:hamiltonianDefByLagrangian}, one can obtain
147   \begin{equation}
148   dH = \sum\limits_k {\left( {p_k d\dot q_k  + \dot q_k dp_k  -
149   \frac{{\partial L}}{{\partial q_k }}dq_k  - \frac{{\partial
150   L}}{{\partial \dot q_k }}d\dot q_k } \right)}  - \frac{{\partial
151 < L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
151 > L}}{{\partial t}}dt . \label{introEquation:diffHamiltonian1}
152   \end{equation}
153 < Making use of  Eq.~\ref{introEquation:generalizedMomenta}, the
154 < second and fourth terms in the parentheses cancel. Therefore,
153 > Making use of Eq.~\ref{introEquation:generalizedMomenta}, the second
154 > and fourth terms in the parentheses cancel. Therefore,
155   Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
156   \begin{equation}
157   dH = \sum\limits_k {\left( {\dot q_k dp_k  - \dot p_k dq_k }
158 < \right)}  - \frac{{\partial L}}{{\partial t}}dt
158 > \right)}  - \frac{{\partial L}}{{\partial t}}dt .
159   \label{introEquation:diffHamiltonian2}
160   \end{equation}
161   By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
162   find
163   \begin{equation}
164 < \frac{{\partial H}}{{\partial p_k }} = q_k
164 > \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
165   \label{introEquation:motionHamiltonianCoordinate}
166   \end{equation}
167   \begin{equation}
168 < \frac{{\partial H}}{{\partial q_k }} =  - p_k
168 > \frac{{\partial H}}{{\partial q_k }} =  - \dot {p_k}
169   \label{introEquation:motionHamiltonianMomentum}
170   \end{equation}
171   and
# Line 185 | Line 174 | t}}
174   t}}
175   \label{introEquation:motionHamiltonianTime}
176   \end{equation}
177 <
189 < Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
177 > where Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
178   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
179   equation of motion. Due to their symmetrical formula, they are also
180   known as the canonical equations of motions \cite{Goldstein2001}.
181  
182   An important difference between Lagrangian approach and the
183   Hamiltonian approach is that the Lagrangian is considered to be a
184 < function of the generalized velocities $\dot q_i$ and the
185 < generalized coordinates $q_i$, while the Hamiltonian is considered
186 < to be a function of the generalized momenta $p_i$ and the conjugate
187 < generalized coordinate $q_i$. Hamiltonian Mechanics is more
188 < appropriate for application to statistical mechanics and quantum
189 < mechanics, since it treats the coordinate and its time derivative as
190 < independent variables and it only works with 1st-order differential
203 < equations\cite{Marion1990}.
204 <
184 > function of the generalized velocities $\dot q_i$ and coordinates
185 > $q_i$, while the Hamiltonian is considered to be a function of the
186 > generalized momenta $p_i$ and the conjugate coordinates $q_i$.
187 > Hamiltonian Mechanics is more appropriate for application to
188 > statistical mechanics and quantum mechanics, since it treats the
189 > coordinate and its time derivative as independent variables and it
190 > only works with 1st-order differential equations\cite{Marion1990}.
191   In Newtonian Mechanics, a system described by conservative forces
192 < conserves the total energy \ref{introEquation:energyConservation}.
193 < It follows that Hamilton's equations of motion conserve the total
194 < Hamiltonian.
192 > conserves the total energy
193 > (Eq.~\ref{introEquation:energyConservation}). It follows that
194 > Hamilton's equations of motion conserve the total Hamiltonian
195   \begin{equation}
196   \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
197   H}}{{\partial q_i }}\dot q_i  + \frac{{\partial H}}{{\partial p_i
198   }}\dot p_i } \right)}  = \sum\limits_i {\left( {\frac{{\partial
199   H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
200   \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
201 < q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
201 > q_i }}} \right) = 0}. \label{introEquation:conserveHalmitonian}
202   \end{equation}
203  
204   \section{\label{introSection:statisticalMechanics}Statistical
# Line 230 | Line 216 | momentum variables. Consider a dynamic system in a car
216   possible states. Each possible state of the system corresponds to
217   one unique point in the phase space. For mechanical systems, the
218   phase space usually consists of all possible values of position and
219 < momentum variables. Consider a dynamic system in a cartesian space,
220 < where each of the $6f$ coordinates and momenta is assigned to one of
221 < $6f$ mutually orthogonal axes, the phase space of this system is a
222 < $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
223 < \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
219 > momentum variables. Consider a dynamic system of $f$ particles in a
220 > cartesian space, where each of the $6f$ coordinates and momenta is
221 > assigned to one of $6f$ mutually orthogonal axes, the phase space of
222 > this system is a $6f$ dimensional space. A point, $x =
223 > (\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
224 > \over q} _1 , \ldots
225 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
226 > \over q} _f
227 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
228 > \over p} _1  \ldots
229 > ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
230 > \over p} _f )$ , with a unique set of values of $6f$ coordinates and
231   momenta is a phase space vector.
232 + %%%fix me
233  
234 < A microscopic state or microstate of a classical system is
241 < specification of the complete phase space vector of a system at any
242 < instant in time. An ensemble is defined as a collection of systems
243 < sharing one or more macroscopic characteristics but each being in a
244 < unique microstate. The complete ensemble is specified by giving all
245 < systems or microstates consistent with the common macroscopic
246 < characteristics of the ensemble. Although the state of each
247 < individual system in the ensemble could be precisely described at
248 < any instance in time by a suitable phase space vector, when using
249 < ensembles for statistical purposes, there is no need to maintain
250 < distinctions between individual systems, since the numbers of
251 < systems at any time in the different states which correspond to
252 < different regions of the phase space are more interesting. Moreover,
253 < in the point of view of statistical mechanics, one would prefer to
254 < use ensembles containing a large enough population of separate
255 < members so that the numbers of systems in such different states can
256 < be regarded as changing continuously as we traverse different
257 < regions of the phase space. The condition of an ensemble at any time
234 > In statistical mechanics, the condition of an ensemble at any time
235   can be regarded as appropriately specified by the density $\rho$
236   with which representative points are distributed over the phase
237 < space. The density of distribution for an ensemble with $f$ degrees
238 < of freedom is defined as,
237 > space. The density distribution for an ensemble with $f$ degrees of
238 > freedom is defined as,
239   \begin{equation}
240   \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
241   \label{introEquation:densityDistribution}
242   \end{equation}
243   Governed by the principles of mechanics, the phase points change
244 < their value which would change the density at any time at phase
245 < space. Hence, the density of distribution is also to be taken as a
246 < function of the time.
247 <
271 < The number of systems $\delta N$ at time $t$ can be determined by,
244 > their locations which would change the density at any time at phase
245 > space. Hence, the density distribution is also to be taken as a
246 > function of the time. The number of systems $\delta N$ at time $t$
247 > can be determined by,
248   \begin{equation}
249   \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
250   \label{introEquation:deltaN}
251   \end{equation}
252 < Assuming a large enough population of systems are exploited, we can
253 < sufficiently approximate $\delta N$ without introducing
254 < discontinuity when we go from one region in the phase space to
255 < another. By integrating over the whole phase space,
252 > Assuming a large enough population of systems, we can sufficiently
253 > approximate $\delta N$ without introducing discontinuity when we go
254 > from one region in the phase space to another. By integrating over
255 > the whole phase space,
256   \begin{equation}
257   N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
258   \label{introEquation:totalNumberSystem}
# Line 288 | Line 264 | With the help of Equation(\ref{introEquation:unitProba
264   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
265   \label{introEquation:unitProbability}
266   \end{equation}
267 < With the help of Equation(\ref{introEquation:unitProbability}) and
268 < the knowledge of the system, it is possible to calculate the average
267 > With the help of Eq.~\ref{introEquation:unitProbability} and the
268 > knowledge of the system, it is possible to calculate the average
269   value of any desired quantity which depends on the coordinates and
270   momenta of the system. Even when the dynamics of the real system is
271   complex, or stochastic, or even discontinuous, the average
272 < properties of the ensemble of possibilities as a whole may still
273 < remain well defined. For a classical system in thermal equilibrium
274 < with its environment, the ensemble average of a mechanical quantity,
275 < $\langle A(q , p) \rangle_t$, takes the form of an integral over the
276 < phase space of the system,
272 > properties of the ensemble of possibilities as a whole remaining
273 > well defined. For a classical system in thermal equilibrium with its
274 > environment, the ensemble average of a mechanical quantity, $\langle
275 > A(q , p) \rangle_t$, takes the form of an integral over the phase
276 > space of the system,
277   \begin{equation}
278   \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
279   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
280 < (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
280 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
281   \label{introEquation:ensembelAverage}
282   \end{equation}
283  
284   There are several different types of ensembles with different
285   statistical characteristics. As a function of macroscopic
286 < parameters, such as temperature \textit{etc}, partition function can
287 < be used to describe the statistical properties of a system in
288 < thermodynamic equilibrium.
289 <
290 < 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,
286 > parameters, such as temperature \textit{etc}, the partition function
287 > can be used to describe the statistical properties of a system in
288 > thermodynamic equilibrium. As an ensemble of systems, each of which
289 > is known to be thermally isolated and conserve energy, the
290 > Microcanonical ensemble (NVE) has a partition function like,
291   \begin{equation}
292 < \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
292 > \Omega (N,V,E) = e^{\beta TS}. \label{introEquation:NVEPartition}
293   \end{equation}
294 < A canonical ensemble(NVT)is an ensemble of systems, each of which
294 > A canonical ensemble (NVT) is an ensemble of systems, each of which
295   can share its energy with a large heat reservoir. The distribution
296   of the total energy amongst the possible dynamical states is given
297   by the partition function,
298   \begin{equation}
299 < \Omega (N,V,T) = e^{ - \beta A}
299 > \Omega (N,V,T) = e^{ - \beta A}.
300   \label{introEquation:NVTPartition}
301   \end{equation}
302   Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
303 < TS$. Since most experiment are carried out under constant pressure
304 < condition, isothermal-isobaric ensemble(NPT) play a very important
305 < role in molecular simulation. The isothermal-isobaric ensemble allow
306 < the system to exchange energy with a heat bath of temperature $T$
307 < and to change the volume as well. Its partition function is given as
303 > TS$. Since most experiments are carried out under constant pressure
304 > condition, the isothermal-isobaric ensemble (NPT) plays a very
305 > important role in molecular simulations. The isothermal-isobaric
306 > ensemble allow the system to exchange energy with a heat bath of
307 > temperature $T$ and to change the volume as well. Its partition
308 > function is given as
309   \begin{equation}
310   \Delta (N,P,T) =  - e^{\beta G}.
311   \label{introEquation:NPTPartition}
# Line 339 | Line 314 | The Liouville's theorem is the foundation on which sta
314  
315   \subsection{\label{introSection:liouville}Liouville's theorem}
316  
317 < The Liouville's theorem is the foundation on which statistical
318 < mechanics rests. It describes the time evolution of phase space
317 > Liouville's theorem is the foundation on which statistical mechanics
318 > rests. It describes the time evolution of the phase space
319   distribution function. In order to calculate the rate of change of
320 < $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
321 < consider the two faces perpendicular to the $q_1$ axis, which are
322 < located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
323 < leaving the opposite face is given by the expression,
320 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
321 > the two faces perpendicular to the $q_1$ axis, which are located at
322 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
323 > opposite face is given by the expression,
324   \begin{equation}
325   \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
326   \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
# Line 369 | Line 344 | divining $ \delta q_1  \ldots \delta q_f \delta p_1  \
344   + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
345   \end{equation}
346   which cancels the first terms of the right hand side. Furthermore,
347 < divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
347 > dividing $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
348   p_f $ in both sides, we can write out Liouville's theorem in a
349   simple form,
350   \begin{equation}
# Line 378 | Line 353 | simple form,
353   \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
354   \label{introEquation:liouvilleTheorem}
355   \end{equation}
381
356   Liouville's theorem states that the distribution function is
357   constant along any trajectory in phase space. In classical
358 < statistical mechanics, since the number of particles in the system
359 < is huge, we may be able to believe the system is stationary,
358 > statistical mechanics, since the number of members in an ensemble is
359 > huge and constant, we can assume the local density has no reason
360 > (other than classical mechanics) to change,
361   \begin{equation}
362   \frac{{\partial \rho }}{{\partial t}} = 0.
363   \label{introEquation:stationary}
# Line 395 | Line 370 | distribution,
370   \label{introEquation:densityAndHamiltonian}
371   \end{equation}
372  
373 < \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
373 > \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
374   Lets consider a region in the phase space,
375   \begin{equation}
376   \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
377   \end{equation}
378   If this region is small enough, the density $\rho$ can be regarded
379 < as uniform over the whole phase space. Thus, the number of phase
380 < points inside this region is given by,
379 > as uniform over the whole integral. Thus, the number of phase points
380 > inside this region is given by,
381   \begin{equation}
382   \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
383   dp_1 } ..dp_f.
# Line 414 | Line 389 | With the help of stationary assumption
389   \end{equation}
390   With the help of stationary assumption
391   (\ref{introEquation:stationary}), we obtain the principle of the
392 < \emph{conservation of extension in phase space},
392 > \emph{conservation of volume in phase space},
393   \begin{equation}
394   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
395   ...dq_f dp_1 } ..dp_f  = 0.
396   \label{introEquation:volumePreserving}
397   \end{equation}
398  
399 < \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
399 > \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
400  
401   Liouville's theorem can be expresses in a variety of different forms
402   which are convenient within different contexts. For any two function
# Line 434 | Line 409 | Substituting equations of motion in Hamiltonian formal
409   q_i }}} \right)}.
410   \label{introEquation:poissonBracket}
411   \end{equation}
412 < Substituting equations of motion in Hamiltonian formalism(
413 < \ref{introEquation:motionHamiltonianCoordinate} ,
414 < \ref{introEquation:motionHamiltonianMomentum} ) into
415 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
416 < theorem using Poisson bracket notion,
412 > Substituting equations of motion in Hamiltonian formalism
413 > (Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
414 > Eq.~\ref{introEquation:motionHamiltonianMomentum}) into
415 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
416 > Liouville's theorem using Poisson bracket notion,
417   \begin{equation}
418   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
419   {\rho ,H} \right\}.
# Line 463 | Line 438 | simulation and the quality of the underlying model. Ho
438   Various thermodynamic properties can be calculated from Molecular
439   Dynamics simulation. By comparing experimental values with the
440   calculated properties, one can determine the accuracy of the
441 < simulation and the quality of the underlying model. However, both of
442 < experiment and computer simulation are usually performed during a
441 > simulation and the quality of the underlying model. However, both
442 > experiments and computer simulations are usually performed during a
443   certain time interval and the measurements are averaged over a
444   period of them which is different from the average behavior of
445 < many-body system in Statistical Mechanics. Fortunately, Ergodic
446 < Hypothesis is proposed to make a connection between time average and
447 < ensemble average. It states that time average and average over the
448 < statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
445 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
446 > Hypothesis makes a connection between time average and the ensemble
447 > average. It states that the time average and average over the
448 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}:
449   \begin{equation}
450   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
451   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 484 | Line 459 | reasonable, the Monte Carlo techniques\cite{Metropolis
459   a properly weighted statistical average. This allows the researcher
460   freedom of choice when deciding how best to measure a given
461   observable. In case an ensemble averaged approach sounds most
462 < reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
462 > reasonable, the Monte Carlo methods\cite{Metropolis1949} can be
463   utilized. Or if the system lends itself to a time averaging
464   approach, the Molecular Dynamics techniques in
465   Sec.~\ref{introSection:molecularDynamics} will be the best
466   choice\cite{Frenkel1996}.
467  
468   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
469 < A variety of numerical integrators were proposed to simulate the
470 < motions. They usually begin with an initial conditionals and move
471 < the objects in the direction governed by the differential equations.
472 < However, most of them ignore the hidden physical law contained
473 < within the equations. Since 1990, geometric integrators, which
474 < preserve various phase-flow invariants such as symplectic structure,
475 < volume and time reversal symmetry, are developed to address this
476 < issue\cite{Dullweber1997, McLachlan1998, Leimkuhler1999}. The
477 < velocity verlet method, which happens to be a simple example of
478 < symplectic integrator, continues to gain its popularity in molecular
479 < dynamics community. This fact can be partly explained by its
480 < geometric nature.
469 > A variety of numerical integrators have been proposed to simulate
470 > the motions of atoms in MD simulation. They usually begin with
471 > initial conditionals and move the objects in the direction governed
472 > by the differential equations. However, most of them ignore the
473 > hidden physical laws contained within the equations. Since 1990,
474 > geometric integrators, which preserve various phase-flow invariants
475 > such as symplectic structure, volume and time reversal symmetry, are
476 > developed to address this issue\cite{Dullweber1997, McLachlan1998,
477 > Leimkuhler1999}. The velocity Verlet method, which happens to be a
478 > simple example of symplectic integrator, continues to gain
479 > popularity in the molecular dynamics community. This fact can be
480 > partly explained by its geometric nature.
481  
482 < \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
483 < A \emph{manifold} is an abstract mathematical space. It locally
484 < looks like Euclidean space, but when viewed globally, it may have
485 < more complicate structure. A good example of manifold is the surface
486 < of Earth. It seems to be flat locally, but it is round if viewed as
487 < a whole. A \emph{differentiable manifold} (also known as
488 < \emph{smooth manifold}) is a manifold with an open cover in which
489 < the covering neighborhoods are all smoothly isomorphic to one
490 < another. In other words,it is possible to apply calculus on
516 < \emph{differentiable manifold}. A \emph{symplectic manifold} is
517 < defined as a pair $(M, \omega)$ which consisting of a
482 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
483 > A \emph{manifold} is an abstract mathematical space. It looks
484 > locally like Euclidean space, but when viewed globally, it may have
485 > more complicated structure. A good example of manifold is the
486 > surface of Earth. It seems to be flat locally, but it is round if
487 > viewed as a whole. A \emph{differentiable manifold} (also known as
488 > \emph{smooth manifold}) is a manifold on which it is possible to
489 > apply calculus on \emph{differentiable manifold}. A \emph{symplectic
490 > manifold} is defined as a pair $(M, \omega)$ which consists of a
491   \emph{differentiable manifold} $M$ and a close, non-degenerated,
492   bilinear symplectic form, $\omega$. A symplectic form on a vector
493   space $V$ is a function $\omega(x, y)$ which satisfies
494   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
495   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
496 < $\omega(x, x) = 0$. Cross product operation in vector field is an
497 < example of symplectic form.
496 > $\omega(x, x) = 0$. The cross product operation in vector field is
497 > an example of symplectic form. One of the motivations to study
498 > \emph{symplectic manifolds} in Hamiltonian Mechanics is that a
499 > symplectic manifold can represent all possible configurations of the
500 > system and the phase space of the system can be described by it's
501 > cotangent bundle. Every symplectic manifold is even dimensional. For
502 > instance, in Hamilton equations, coordinate and momentum always
503 > appear in pairs.
504  
526 One of the motivations to study \emph{symplectic manifold} in
527 Hamiltonian Mechanics is that a symplectic manifold can represent
528 all possible configurations of the system and the phase space of the
529 system can be described by it's cotangent bundle. Every symplectic
530 manifold is even dimensional. For instance, in Hamilton equations,
531 coordinate and momentum always appear in pairs.
532
533 Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
534 \[
535 f : M \rightarrow N
536 \]
537 is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
538 the \emph{pullback} of $\eta$ under f is equal to $\omega$.
539 Canonical transformation is an example of symplectomorphism in
540 classical mechanics.
541
505   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
506  
507 < For a ordinary differential system defined as
507 > For an ordinary differential system defined as
508   \begin{equation}
509   \dot x = f(x)
510   \end{equation}
511 < where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
511 > where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
512 > $f(r) = J\nabla _x H(r)$. Here, $H = H (q, p)$ is Hamiltonian
513 > function and $J$ is the skew-symmetric matrix
514   \begin{equation}
550 f(r) = J\nabla _x H(r).
551 \end{equation}
552 $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
553 matrix
554 \begin{equation}
515   J = \left( {\begin{array}{*{20}c}
516     0 & I  \\
517     { - I} & 0  \\
# Line 561 | Line 521 | system can be rewritten as,
521   where $I$ is an identity matrix. Using this notation, Hamiltonian
522   system can be rewritten as,
523   \begin{equation}
524 < \frac{d}{{dt}}x = J\nabla _x H(x)
524 > \frac{d}{{dt}}x = J\nabla _x H(x).
525   \label{introEquation:compactHamiltonian}
526   \end{equation}In this case, $f$ is
527 < called a \emph{Hamiltonian vector field}.
528 <
569 < Another generalization of Hamiltonian dynamics is Poisson
570 < Dynamics\cite{Olver1986},
527 > called a \emph{Hamiltonian vector field}. Another generalization of
528 > Hamiltonian dynamics is Poisson Dynamics\cite{Olver1986},
529   \begin{equation}
530   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
531   \end{equation}
# Line 575 | Line 533 | Let $x(t)$ be the exact solution of the ODE system,
533  
534   \subsection{\label{introSection:exactFlow}Exact Flow}
535  
536 < Let $x(t)$ be the exact solution of the ODE system,
537 < \begin{equation}
538 < \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
539 < \end{equation}
540 < The exact flow(solution) $\varphi_\tau$ is defined by
583 < \[
584 < x(t+\tau) =\varphi_\tau(x(t))
536 > Let $x(t)$ be the exact solution of the ODE
537 > system,$\frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}$, we can
538 > define its exact flow(solution) $\varphi_\tau$
539 > \[ x(t+\tau)
540 > =\varphi_\tau(x(t))
541   \]
542   where $\tau$ is a fixed time step and $\varphi$ is a map from phase
543   space to itself. The flow has the continuous group property,
# Line 603 | Line 559 | The exact flow can also be written in terms of the of
559   }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
560   \label{introEquation:exponentialOperator}
561   \end{equation}
606
562   In most cases, it is not easy to find the exact flow $\varphi_\tau$.
563 < Instead, we use a approximate map, $\psi_\tau$, which is usually
563 > Instead, we use an approximate map, $\psi_\tau$, which is usually
564   called integrator. The order of an integrator $\psi_\tau$ is $p$, if
565   the Taylor series of $\psi_\tau$ agree to order $p$,
566   \begin{equation}
567 < \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
567 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
568   \end{equation}
569  
570   \subsection{\label{introSection:geometricProperties}Geometric Properties}
571  
572 < The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
573 < and its flow play important roles in numerical studies. Many of them
574 < can be found in systems which occur naturally in applications.
620 <
572 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
573 > ODE and its flow play important roles in numerical studies. Many of
574 > them can be found in systems which occur naturally in applications.
575   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
576   a \emph{symplectic} flow if it satisfies,
577   \begin{equation}
# Line 631 | Line 585 | is the property must be preserved by the integrator.
585   \begin{equation}
586   {\varphi '}^T J \varphi ' = J \circ \varphi
587   \end{equation}
588 < is the property must be preserved by the integrator.
589 <
590 < It is possible to construct a \emph{volume-preserving} flow for a
591 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
592 < \det d\varphi  = 1$. One can show easily that a symplectic flow will
593 < be volume-preserving.
640 <
641 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
642 < will result in a new system,
588 > is the property that must be preserved by the integrator. It is
589 > possible to construct a \emph{volume-preserving} flow for a source
590 > free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $ \det
591 > d\varphi  = 1$. One can show easily that a symplectic flow will be
592 > volume-preserving. Changing the variables $y = h(x)$ in an ODE
593 > (Eq.~\ref{introEquation:ODE}) will result in a new system,
594   \[
595   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
596   \]
597   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
598   In other words, the flow of this vector field is reversible if and
599 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
600 <
650 < A \emph{first integral}, or conserved quantity of a general
599 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
600 > \emph{first integral}, or conserved quantity of a general
601   differential function is a function $ G:R^{2d}  \to R^d $ which is
602   constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
603   \[
# Line 655 | Line 605 | Using chain rule, one may obtain,
605   \]
606   Using chain rule, one may obtain,
607   \[
608 < \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
608 > \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \dot \nabla G,
609   \]
610   which is the condition for conserving \emph{first integral}. For a
611   canonical Hamiltonian system, the time evolution of an arbitrary
612   smooth function $G$ is given by,
663
613   \begin{eqnarray}
614 < \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
615 <                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
614 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \notag\\
615 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)).
616   \label{introEquation:firstIntegral1}
617   \end{eqnarray}
618 <
619 <
671 < Using poisson bracket notion, Equation
672 < \ref{introEquation:firstIntegral1} can be rewritten as
618 > Using poisson bracket notion, Eq.~\ref{introEquation:firstIntegral1}
619 > can be rewritten as
620   \[
621   \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
622   \]
623   Therefore, the sufficient condition for $G$ to be the \emph{first
624 < integral} of a Hamiltonian system is
678 < \[
679 < \left\{ {G,H} \right\} = 0.
680 < \]
624 > integral} of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$
625   As well known, the Hamiltonian (or energy) H of a Hamiltonian system
626   is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
627 < 0$.
684 <
685 < When designing any numerical methods, one should always try to
627 > 0$. When designing any numerical methods, one should always try to
628   preserve the structural properties of the original ODE and its flow.
629  
630   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
631   A lot of well established and very effective numerical methods have
632   been successful precisely because of their symplecticities even
633   though this fact was not recognized when they were first
634 < constructed. The most famous example is leapfrog methods in
635 < molecular dynamics. In general, symplectic integrators can be
634 > constructed. The most famous example is the Verlet-leapfrog method
635 > in molecular dynamics. In general, symplectic integrators can be
636   constructed using one of four different methods.
637   \begin{enumerate}
638   \item Generating functions
# Line 698 | Line 640 | constructed using one of four different methods.
640   \item Runge-Kutta methods
641   \item Splitting methods
642   \end{enumerate}
701
643   Generating function\cite{Channell1990} tends to lead to methods
644   which are cumbersome and difficult to use. In dissipative systems,
645   variational methods can capture the decay of energy
646   accurately\cite{Kane2000}. Since their geometrically unstable nature
647   against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
648   methods are not suitable for Hamiltonian system. Recently, various
649 < high-order explicit Runge-Kutta methods
650 < \cite{Owren1992,Chen2003}have been developed to overcome this
651 < instability. However, due to computational penalty involved in
652 < implementing the Runge-Kutta methods, they do not attract too much
653 < attention from Molecular Dynamics community. Instead, splitting have
654 < been widely accepted since they exploit natural decompositions of
655 < the system\cite{Tuckerman1992, McLachlan1998}.
649 > high-order explicit Runge-Kutta methods \cite{Owren1992,Chen2003}
650 > have been developed to overcome this instability. However, due to
651 > computational penalty involved in implementing the Runge-Kutta
652 > methods, they have not attracted much attention from the Molecular
653 > Dynamics community. Instead, splitting methods have been widely
654 > accepted since they exploit natural decompositions of the
655 > system\cite{Tuckerman1992, McLachlan1998}.
656  
657 < \subsubsection{\label{introSection:splittingMethod}Splitting Method}
657 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
658  
659   The main idea behind splitting methods is to decompose the discrete
660   $\varphi_h$ as a composition of simpler flows,
# Line 723 | Line 664 | simpler integration of the system.
664   \label{introEquation:FlowDecomposition}
665   \end{equation}
666   where each of the sub-flow is chosen such that each represent a
667 < simpler integration of the system.
668 <
728 < Suppose that a Hamiltonian system takes the form,
667 > simpler integration of the system. Suppose that a Hamiltonian system
668 > takes the form,
669   \[
670   H = H_1 + H_2.
671   \]
# Line 734 | Line 674 | order is then given by the Lie-Trotter formula
674   energy respectively, which is a natural decomposition of the
675   problem. If $H_1$ and $H_2$ can be integrated using exact flows
676   $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
677 < order is then given by the Lie-Trotter formula
677 > order expression is then given by the Lie-Trotter formula
678   \begin{equation}
679   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
680   \label{introEquation:firstOrderSplitting}
# Line 753 | Line 693 | The Lie-Trotter splitting(\ref{introEquation:firstOrde
693   where $\phi$ and $\psi$ both are symplectic maps. Thus operator
694   splitting in this context automatically generates a symplectic map.
695  
696 < The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
697 < introduces local errors proportional to $h^2$, while Strang
698 < splitting gives a second-order decomposition,
696 > The Lie-Trotter
697 > splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
698 > local errors proportional to $h^2$, while Strang splitting gives a
699 > second-order decomposition,
700   \begin{equation}
701   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
702   _{1,h/2} , \label{introEquation:secondOrderSplitting}
703   \end{equation}
704 < which has a local error proportional to $h^3$. Sprang splitting's
705 < popularity in molecular simulation community attribute to its
706 < symmetric property,
704 > which has a local error proportional to $h^3$. The Sprang
705 > splitting's popularity in molecular simulation community attribute
706 > to its symmetric property,
707   \begin{equation}
708   \varphi _h^{ - 1} = \varphi _{ - h}.
709   \label{introEquation:timeReversible}
710   \end{equation}
711  
712 < \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
712 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
713   The classical equation for a system consisting of interacting
714   particles can be written in Hamiltonian form,
715   \[
716   H = T + V
717   \]
718   where $T$ is the kinetic energy and $V$ is the potential energy.
719 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
719 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
720   obtains the following:
721   \begin{align}
722   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 802 | Line 743 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
743      \label{introEquation:Lp9b}\\%
744   %
745   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
746 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
746 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
747   \end{align}
748   From the preceding splitting, one can see that the integration of
749   the equations of motion would follow:
# Line 811 | Line 752 | the equations of motion would follow:
752  
753   \item Use the half step velocities to move positions one whole step, $\Delta t$.
754  
755 < \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
755 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
756  
757   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
758   \end{enumerate}
759 <
760 < Simply switching the order of splitting and composing, a new
761 < integrator, the \emph{position verlet} integrator, can be generated,
759 > By simply switching the order of the propagators in the splitting
760 > and composing a new integrator, the \emph{position verlet}
761 > integrator, can be generated,
762   \begin{align}
763   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
764   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 828 | Line 769 | q(\Delta t)} \right]. %
769   \label{introEquation:positionVerlet2}
770   \end{align}
771  
772 < \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
772 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
773  
774 < Baker-Campbell-Hausdorff formula can be used to determine the local
775 < error of splitting method in terms of commutator of the
774 > The Baker-Campbell-Hausdorff formula can be used to determine the
775 > local error of splitting method in terms of the commutator of the
776   operators(\ref{introEquation:exponentialOperator}) associated with
777 < the sub-flow. For operators $hX$ and $hY$ which are associate to
777 > the sub-flow. For operators $hX$ and $hY$ which are associated with
778   $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
779   \begin{equation}
780   \exp (hX + hY) = \exp (hZ)
# Line 847 | Line 788 | Applying Baker-Campbell-Hausdorff formula\cite{Varadar
788   \[
789   [X,Y] = XY - YX .
790   \]
791 < Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
792 < Sprang splitting, we can obtain
791 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
792 > to the Sprang splitting, we can obtain
793   \begin{eqnarray*}
794   \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 \\
795                                     &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
796 <                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
796 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots
797 >                                   ).
798   \end{eqnarray*}
799 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
799 > Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local
800   error of Spring splitting is proportional to $h^3$. The same
801 < procedure can be applied to general splitting,  of the form
801 > procedure can be applied to a general splitting of the form
802   \begin{equation}
803   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
804   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
805   \end{equation}
806 < Careful choice of coefficient $a_1 \ldot b_m$ will lead to higher
807 < order method. Yoshida proposed an elegant way to compose higher
806 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
807 > order methods. Yoshida proposed an elegant way to compose higher
808   order methods based on symmetric splitting\cite{Yoshida1990}. Given
809   a symmetric second order base method $ \varphi _h^{(2)} $, a
810   fourth-order symmetric method can be constructed by composing,
# Line 875 | Line 817 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
817   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
818   \begin{equation}
819   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
820 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
820 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
821   \end{equation}
822 < , if the weights are chosen as
822 > if the weights are chosen as
823   \[
824   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
825   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 914 | Line 856 | initialization of a simulation. Sec.~\ref{introSec:pro
856   \end{enumerate}
857   These three individual steps will be covered in the following
858   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
859 < initialization of a simulation. Sec.~\ref{introSec:production} will
860 < discusses issues in production run. Sec.~\ref{introSection:Analysis}
861 < provides the theoretical tools for trajectory analysis.
859 > initialization of a simulation. Sec.~\ref{introSection:production}
860 > will discusse issues in production run.
861 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
862 > trajectory analysis.
863  
864   \subsection{\label{introSec:initialSystemSettings}Initialization}
865  
866 < \subsubsection{Preliminary preparation}
866 > \subsubsection{\textbf{Preliminary preparation}}
867  
868   When selecting the starting structure of a molecule for molecular
869   simulation, one may retrieve its Cartesian coordinates from public
870   databases, such as RCSB Protein Data Bank \textit{etc}. Although
871   thousands of crystal structures of molecules are discovered every
872   year, many more remain unknown due to the difficulties of
873 < purification and crystallization. Even for the molecule with known
874 < structure, some important information is missing. For example, the
873 > purification and crystallization. Even for molecules with known
874 > structure, some important information is missing. For example, a
875   missing hydrogen atom which acts as donor in hydrogen bonding must
876   be added. Moreover, in order to include electrostatic interaction,
877   one may need to specify the partial charges for individual atoms.
878   Under some circumstances, we may even need to prepare the system in
879 < a special setup. For instance, when studying transport phenomenon in
880 < membrane system, we may prepare the lipids in bilayer structure
881 < instead of placing lipids randomly in solvent, since we are not
882 < interested in self-aggregation and it takes a long time to happen.
879 > a special configuration. For instance, when studying transport
880 > phenomenon in membrane systems, we may prepare the lipids in a
881 > bilayer structure instead of placing lipids randomly in solvent,
882 > since we are not interested in the slow self-aggregation process.
883  
884 < \subsubsection{Minimization}
884 > \subsubsection{\textbf{Minimization}}
885  
886   It is quite possible that some of molecules in the system from
887 < preliminary preparation may be overlapped with each other. This
888 < close proximity leads to high potential energy which consequently
889 < jeopardizes any molecular dynamics simulations. To remove these
890 < steric overlaps, one typically performs energy minimization to find
891 < a more reasonable conformation. Several energy minimization methods
892 < have been developed to exploit the energy surface and to locate the
893 < local minimum. While converging slowly near the minimum, steepest
894 < descent method is extremely robust when systems are far from
895 < harmonic. Thus, it is often used to refine structure from
896 < crystallographic data. Relied on the gradient or hessian, advanced
897 < methods like conjugate gradient and Newton-Raphson converge rapidly
898 < to a local minimum, while become unstable if the energy surface is
899 < far from quadratic. Another factor must be taken into account, when
887 > preliminary preparation may be overlapping with each other. This
888 > close proximity leads to high initial potential energy which
889 > consequently jeopardizes any molecular dynamics simulations. To
890 > remove these steric overlaps, one typically performs energy
891 > minimization to find a more reasonable conformation. Several energy
892 > minimization methods have been developed to exploit the energy
893 > surface and to locate the local minimum. While converging slowly
894 > near the minimum, steepest descent method is extremely robust when
895 > systems are strongly anharmonic. Thus, it is often used to refine
896 > structure from crystallographic data. Relied on the gradient or
897 > hessian, advanced methods like Newton-Raphson converge rapidly to a
898 > local minimum, but become unstable if the energy surface is far from
899 > quadratic. Another factor that must be taken into account, when
900   choosing energy minimization method, is the size of the system.
901   Steepest descent and conjugate gradient can deal with models of any
902 < size. Because of the limit of computation power to calculate hessian
903 < matrix and insufficient storage capacity to store them, most
904 < Newton-Raphson methods can not be used with very large models.
902 > size. Because of the limits on computer memory to store the hessian
903 > matrix and the computing power needed to diagonalized these
904 > matrices, most Newton-Raphson methods can not be used with very
905 > large systems.
906  
907 < \subsubsection{Heating}
907 > \subsubsection{\textbf{Heating}}
908  
909   Typically, Heating is performed by assigning random velocities
910 < according to a Gaussian distribution for a temperature. Beginning at
911 < a lower temperature and gradually increasing the temperature by
912 < assigning greater random velocities, we end up with setting the
913 < temperature of the system to a final temperature at which the
914 < simulation will be conducted. In heating phase, we should also keep
915 < the system from drifting or rotating as a whole. Equivalently, the
916 < net linear momentum and angular momentum of the system should be
917 < shifted to zero.
910 > according to a Maxwell-Boltzman distribution for a desired
911 > temperature. Beginning at a lower temperature and gradually
912 > increasing the temperature by assigning larger random velocities, we
913 > end up with setting the temperature of the system to a final
914 > temperature at which the simulation will be conducted. In heating
915 > phase, we should also keep the system from drifting or rotating as a
916 > whole. To do this, the net linear momentum and angular momentum of
917 > the system is shifted to zero after each resampling from the Maxwell
918 > -Boltzman distribution.
919  
920 < \subsubsection{Equilibration}
920 > \subsubsection{\textbf{Equilibration}}
921  
922   The purpose of equilibration is to allow the system to evolve
923   spontaneously for a period of time and reach equilibrium. The
# Line 986 | Line 931 | Production run is the most important step of the simul
931  
932   \subsection{\label{introSection:production}Production}
933  
934 < Production run is the most important step of the simulation, in
934 > The production run is the most important step of the simulation, in
935   which the equilibrated structure is used as a starting point and the
936   motions of the molecules are collected for later analysis. In order
937   to capture the macroscopic properties of the system, the molecular
938 < dynamics simulation must be performed in correct and efficient way.
938 > dynamics simulation must be performed by sampling correctly and
939 > efficiently from the relevant thermodynamic ensemble.
940  
941   The most expensive part of a molecular dynamics simulation is the
942   calculation of non-bonded forces, such as van der Waals force and
943   Coulombic forces \textit{etc}. For a system of $N$ particles, the
944   complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
945   which making large simulations prohibitive in the absence of any
946 < computation saving techniques.
947 <
948 < A natural approach to avoid system size issue is to represent the
949 < bulk behavior by a finite number of the particles. However, this
950 < approach will suffer from the surface effect. To offset this,
1005 < \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
946 > algorithmic tricks. A natural approach to avoid system size issues
947 > is to represent the bulk behavior by a finite number of the
948 > particles. However, this approach will suffer from the surface
949 > effect at the edges of the simulation. To offset this,
950 > \textit{Periodic boundary conditions} (see Fig.~\ref{introFig:pbc})
951   is developed to simulate bulk properties with a relatively small
952   number of particles. In this method, the simulation box is
953   replicated throughout space to form an infinite lattice. During the
# Line 1021 | Line 966 | evaluation is to apply cutoff where particles farther
966  
967   %cutoff and minimum image convention
968   Another important technique to improve the efficiency of force
969 < evaluation is to apply cutoff where particles farther than a
970 < predetermined distance, are not included in the calculation
969 > evaluation is to apply spherical cutoff where particles farther than
970 > a predetermined distance are not included in the calculation
971   \cite{Frenkel1996}. The use of a cutoff radius will cause a
972   discontinuity in the potential energy curve. Fortunately, one can
973 < shift the potential to ensure the potential curve go smoothly to
974 < zero at the cutoff radius. Cutoff strategy works pretty well for
975 < Lennard-Jones interaction because of its short range nature.
976 < However, simply truncating the electrostatic interaction with the
977 < use of cutoff has been shown to lead to severe artifacts in
978 < simulations. Ewald summation, in which the slowly conditionally
979 < convergent Coulomb potential is transformed into direct and
980 < reciprocal sums with rapid and absolute convergence, has proved to
981 < minimize the periodicity artifacts in liquid simulations. Taking the
982 < advantages of the fast Fourier transform (FFT) for calculating
983 < discrete Fourier transforms, the particle mesh-based
973 > shift simple radial potential to ensure the potential curve go
974 > smoothly to zero at the cutoff radius. The cutoff strategy works
975 > well for Lennard-Jones interaction because of its short range
976 > nature. However, simply truncating the electrostatic interaction
977 > with the use of cutoffs has been shown to lead to severe artifacts
978 > in simulations. The Ewald summation, in which the slowly decaying
979 > Coulomb potential is transformed into direct and reciprocal sums
980 > with rapid and absolute convergence, has proved to minimize the
981 > periodicity artifacts in liquid simulations. Taking the advantages
982 > of the fast Fourier transform (FFT) for calculating discrete Fourier
983 > transforms, the particle mesh-based
984   methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
985 < $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
986 < multipole method}\cite{Greengard1987, Greengard1994}, which treats
987 < Coulombic interaction exactly at short range, and approximate the
988 < potential at long range through multipolar expansion. In spite of
989 < their wide acceptances at the molecular simulation community, these
990 < two methods are hard to be implemented correctly and efficiently.
991 < Instead, we use a damped and charge-neutralized Coulomb potential
992 < method developed by Wolf and his coworkers\cite{Wolf1999}. The
993 < shifted Coulomb potential for particle $i$ and particle $j$ at
994 < distance $r_{rj}$ is given by:
985 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
986 > \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
987 > which treats Coulombic interactions exactly at short range, and
988 > approximate the potential at long range through multipolar
989 > expansion. In spite of their wide acceptance at the molecular
990 > simulation community, these two methods are difficult to implement
991 > correctly and efficiently. Instead, we use a damped and
992 > charge-neutralized Coulomb potential method developed by Wolf and
993 > his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
994 > particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
995   \begin{equation}
996   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
997   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
# Line 1068 | Line 1013 | Recently, advanced visualization technique are widely
1013  
1014   \subsection{\label{introSection:Analysis} Analysis}
1015  
1016 < Recently, advanced visualization technique are widely applied to
1016 > Recently, advanced visualization technique have become applied to
1017   monitor the motions of molecules. Although the dynamics of the
1018   system can be described qualitatively from animation, quantitative
1019 < trajectory analysis are more appreciable. According to the
1020 < principles of Statistical Mechanics,
1019 > trajectory analysis are more useful. According to the principles of
1020 > Statistical Mechanics in
1021   Sec.~\ref{introSection:statisticalMechanics}, one can compute
1022 < thermodynamics properties, analyze fluctuations of structural
1022 > thermodynamic properties, analyze fluctuations of structural
1023   parameters, and investigate time-dependent processes of the molecule
1024   from the trajectories.
1025  
1026 < \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1026 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1027  
1028 < Thermodynamics properties, which can be expressed in terms of some
1028 > Thermodynamic properties, which can be expressed in terms of some
1029   function of the coordinates and momenta of all particles in the
1030   system, can be directly computed from molecular dynamics. The usual
1031   way to measure the pressure is based on virial theorem of Clausius
# Line 1100 | Line 1045 | P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\
1045   < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1046   \end{equation}
1047  
1048 < \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1048 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1049  
1050   Structural Properties of a simple fluid can be described by a set of
1051 < distribution functions. Among these functions,\emph{pair
1051 > distribution functions. Among these functions,the \emph{pair
1052   distribution function}, also known as \emph{radial distribution
1053 < function}, is of most fundamental importance to liquid-state theory.
1054 < Pair distribution function can be gathered by Fourier transforming
1055 < raw data from a series of neutron diffraction experiments and
1056 < integrating over the surface factor \cite{Powles1973}. The
1057 < experiment result can serve as a criterion to justify the
1058 < correctness of the theory. Moreover, various equilibrium
1059 < thermodynamic and structural properties can also be expressed in
1060 < terms of radial distribution function \cite{Allen1987}.
1061 <
1117 < A pair distribution functions $g(r)$ gives the probability that a
1053 > function}, is of most fundamental importance to liquid theory.
1054 > Experimentally, pair distribution function can be gathered by
1055 > Fourier transforming raw data from a series of neutron diffraction
1056 > experiments and integrating over the surface factor
1057 > \cite{Powles1973}. The experimental results can serve as a criterion
1058 > to justify the correctness of a liquid model. Moreover, various
1059 > equilibrium thermodynamic and structural properties can also be
1060 > expressed in terms of radial distribution function \cite{Allen1987}.
1061 > The pair distribution functions $g(r)$ gives the probability that a
1062   particle $i$ will be located at a distance $r$ from a another
1063   particle $j$ in the system
1064 < \[
1064 > \begin{equation}
1065   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1066 < \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1067 < \]
1066 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1067 > (r)}{\rho}.
1068 > \end{equation}
1069   Note that the delta function can be replaced by a histogram in
1070 < computer simulation. Figure
1071 < \ref{introFigure:pairDistributionFunction} shows a typical pair
1072 < distribution function for the liquid argon system. The occurrence of
1128 < several peaks in the plot of $g(r)$ suggests that it is more likely
1129 < to find particles at certain radial values than at others. This is a
1130 < result of the attractive interaction at such distances. Because of
1131 < the strong repulsive forces at short distance, the probability of
1132 < locating particles at distances less than about 2.5{\AA} from each
1133 < other is essentially zero.
1070 > computer simulation. Peaks in $g(r)$ represent solvent shells, and
1071 > the height of these peaks gradually decreases to 1 as the liquid of
1072 > large distance approaches the bulk density.
1073  
1135 %\begin{figure}
1136 %\centering
1137 %\includegraphics[width=\linewidth]{pdf.eps}
1138 %\caption[Pair distribution function for the liquid argon
1139 %]{Pair distribution function for the liquid argon}
1140 %\label{introFigure:pairDistributionFunction}
1141 %\end{figure}
1074  
1075 < \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1076 < Properties}
1075 > \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1076 > Properties}}
1077  
1078   Time-dependent properties are usually calculated using \emph{time
1079 < correlation function}, which correlates random variables $A$ and $B$
1080 < at two different time
1079 > correlation functions}, which correlate random variables $A$ and $B$
1080 > at two different times,
1081   \begin{equation}
1082   C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1083   \label{introEquation:timeCorrelationFunction}
1084   \end{equation}
1085   If $A$ and $B$ refer to same variable, this kind of correlation
1086 < function is called \emph{auto correlation function}. One example of
1087 < auto correlation function is velocity auto-correlation function
1088 < which is directly related to transport properties of molecular
1089 < liquids:
1086 > function is called an \emph{autocorrelation function}. One example
1087 > of an auto correlation function is the velocity auto-correlation
1088 > function which is directly related to transport properties of
1089 > molecular liquids:
1090   \[
1091   D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1092   \right\rangle } dt
1093   \]
1094 < where $D$ is diffusion constant. Unlike velocity autocorrelation
1095 < function which is averaging over time origins and over all the
1096 < atoms, dipole autocorrelation are calculated for the entire system.
1097 < The dipole autocorrelation function is given by:
1094 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1095 > function, which is averaging over time origins and over all the
1096 > atoms, the dipole autocorrelation functions are calculated for the
1097 > entire system. The dipole autocorrelation function is given by:
1098   \[
1099   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1100   \right\rangle
# Line 1170 | Line 1102 | u_{tot} (t) = \sum\limits_i {u_i (t)}
1102   Here $u_{tot}$ is the net dipole of the entire system and is given
1103   by
1104   \[
1105 < u_{tot} (t) = \sum\limits_i {u_i (t)}
1105 > u_{tot} (t) = \sum\limits_i {u_i (t)}.
1106   \]
1107   In principle, many time correlation functions can be related with
1108   Fourier transforms of the infrared, Raman, and inelastic neutron
# Line 1179 | Line 1111 | i2\pi vt} dt}
1111   each frequency using the following relationship:
1112   \[
1113   \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1114 < i2\pi vt} dt}
1114 > i2\pi vt} dt}.
1115   \]
1116  
1117   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
# Line 1188 | Line 1120 | simulator is governed by the rigid body dynamics. In m
1120   areas, from engineering, physics, to chemistry. For example,
1121   missiles and vehicle are usually modeled by rigid bodies.  The
1122   movement of the objects in 3D gaming engine or other physics
1123 < simulator is governed by the rigid body dynamics. In molecular
1124 < simulation, rigid body is used to simplify the model in
1125 < protein-protein docking study\cite{Gray2003}.
1123 > simulator is governed by rigid body dynamics. In molecular
1124 > simulations, rigid bodies are used to simplify protein-protein
1125 > docking studies\cite{Gray2003}.
1126  
1127   It is very important to develop stable and efficient methods to
1128 < integrate the equations of motion of orientational degrees of
1129 < freedom. Euler angles are the nature choice to describe the
1130 < rotational degrees of freedom. However, due to its singularity, the
1131 < numerical integration of corresponding equations of motion is very
1132 < inefficient and inaccurate. Although an alternative integrator using
1133 < different sets of Euler angles can overcome this
1134 < difficulty\cite{Barojas1973}, the computational penalty and the lost
1135 < of angular momentum conservation still remain. A singularity free
1136 < representation utilizing quaternions was developed by Evans in
1137 < 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1138 < nonseparable Hamiltonian resulted from quaternion representation,
1139 < which prevents the symplectic algorithm to be utilized. Another
1140 < different approach is to apply holonomic constraints to the atoms
1141 < belonging to the rigid body. Each atom moves independently under the
1142 < normal forces deriving from potential energy and constraint forces
1143 < which are used to guarantee the rigidness. However, due to their
1144 < iterative nature, SHAKE and Rattle algorithm converge very slowly
1145 < when the number of constraint increases\cite{Ryckaert1977,
1146 < Andersen1983}.
1128 > integrate the equations of motion for orientational degrees of
1129 > freedom. Euler angles are the natural choice to describe the
1130 > rotational degrees of freedom. However, due to $\frac {1}{sin
1131 > \theta}$ singularities, the numerical integration of corresponding
1132 > equations of motion is very inefficient and inaccurate. Although an
1133 > alternative integrator using multiple sets of Euler angles can
1134 > overcome this difficulty\cite{Barojas1973}, the computational
1135 > penalty and the loss of angular momentum conservation still remain.
1136 > A singularity-free representation utilizing quaternions was
1137 > developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1138 > approach uses a nonseparable Hamiltonian resulting from the
1139 > quaternion representation, which prevents the symplectic algorithm
1140 > to be utilized. Another different approach is to apply holonomic
1141 > constraints to the atoms belonging to the rigid body. Each atom
1142 > moves independently under the normal forces deriving from potential
1143 > energy and constraint forces which are used to guarantee the
1144 > rigidness. However, due to their iterative nature, the SHAKE and
1145 > Rattle algorithms also converge very slowly when the number of
1146 > constraints increases\cite{Ryckaert1977, Andersen1983}.
1147  
1148 < The break through in geometric literature suggests that, in order to
1148 > A break-through in geometric literature suggests that, in order to
1149   develop a long-term integration scheme, one should preserve the
1150 < symplectic structure of the flow. Introducing conjugate momentum to
1151 < rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1152 < symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1153 < the Hamiltonian system in a constraint manifold by iteratively
1154 < satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1155 < method using quaternion representation was developed by
1156 < Omelyan\cite{Omelyan1998}. However, both of these methods are
1157 < iterative and inefficient. In this section, we will present a
1150 > symplectic structure of the flow. By introducing a conjugate
1151 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1152 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1153 > proposed to evolve the Hamiltonian system in a constraint manifold
1154 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1155 > An alternative method using the quaternion representation was
1156 > developed by Omelyan\cite{Omelyan1998}. However, both of these
1157 > methods are iterative and inefficient. In this section, we descibe a
1158   symplectic Lie-Poisson integrator for rigid body developed by
1159   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1160  
1161 < \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1162 < The motion of the rigid body is Hamiltonian with the Hamiltonian
1161 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1162 > The motion of a rigid body is Hamiltonian with the Hamiltonian
1163   function
1164   \begin{equation}
1165   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
# Line 1241 | Line 1173 | constrained Hamiltonian equation subjects to a holonom
1173   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1174   \]
1175   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1176 < constrained Hamiltonian equation subjects to a holonomic constraint,
1176 > constrained Hamiltonian equation is subjected to a holonomic
1177 > constraint,
1178   \begin{equation}
1179   Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1180   \end{equation}
1181 < which is used to ensure rotation matrix's orthogonality.
1182 < Differentiating \ref{introEquation:orthogonalConstraint} and using
1183 < Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1181 > which is used to ensure rotation matrix's unitarity. Differentiating
1182 > Eq.~\ref{introEquation:orthogonalConstraint} and using
1183 > Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1184   \begin{equation}
1185   Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1186   \label{introEquation:RBFirstOrderConstraint}
1187   \end{equation}
1255
1188   Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1189   \ref{introEquation:motionHamiltonianMomentum}), one can write down
1190   the equations of motion,
1191 < \[
1192 < \begin{array}{c}
1193 < \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1194 < \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1195 < \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1196 < \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1265 < \end{array}
1266 < \]
1267 <
1191 > \begin{eqnarray}
1192 > \frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\
1193 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q), \label{introEquation:RBMotionMomentum}\\
1194 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1},  \label{introEquation:RBMotionRotation}\\
1195 > \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1196 > \end{eqnarray}
1197   In general, there are two ways to satisfy the holonomic constraints.
1198 < We can use constraint force provided by lagrange multiplier on the
1199 < normal manifold to keep the motion on constraint space. Or we can
1200 < simply evolve the system in constraint manifold. These two methods
1201 < are proved to be equivalent. The holonomic constraint and equations
1202 < of motions define a constraint manifold for rigid body
1198 > We can use a constraint force provided by a Lagrange multiplier on
1199 > the normal manifold to keep the motion on constraint space. Or we
1200 > can simply evolve the system on the constraint manifold. These two
1201 > methods have been proved to be equivalent. The holonomic constraint
1202 > and equations of motions define a constraint manifold for rigid
1203 > bodies
1204   \[
1205   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1206   \right\}.
1207   \]
1278
1208   Unfortunately, this constraint manifold is not the cotangent bundle
1209 < $T_{\star}SO(3)$. However, it turns out that under symplectic
1209 > $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1210 > rotation group $SO(3)$. However, it turns out that under symplectic
1211   transformation, the cotangent space and the phase space are
1212 < diffeomorphic. Introducing
1212 > diffeomorphic. By introducing
1213   \[
1214   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1215   \]
# Line 1289 | Line 1219 | T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \t
1219   T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1220   1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1221   \]
1292
1222   For a body fixed vector $X_i$ with respect to the center of mass of
1223   the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1224   given as
# Line 1308 | Line 1237 | respectively.
1237   \[
1238   \nabla _Q V(q,Q) = F(q,Q)X_i^t
1239   \]
1240 < respectively.
1241 <
1242 < As a common choice to describe the rotation dynamics of the rigid
1314 < body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1315 < rewrite the equations of motion,
1240 > respectively. As a common choice to describe the rotation dynamics
1241 > of the rigid body, the angular momentum on the body fixed frame $\Pi
1242 > = Q^t P$ is introduced to rewrite the equations of motion,
1243   \begin{equation}
1244   \begin{array}{l}
1245 < \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1246 < \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1245 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda,  \\
1246 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1},  \\
1247   \end{array}
1248   \label{introEqaution:RBMotionPI}
1249   \end{equation}
1250 < , as well as holonomic constraints,
1251 < \[
1252 < \begin{array}{l}
1326 < \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1327 < Q^T Q = 1 \\
1328 < \end{array}
1329 < \]
1330 <
1331 < For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1332 < so(3)^ \star$, the hat-map isomorphism,
1250 > as well as holonomic constraints $\Pi J^{ - 1}  + J^{ - 1} \Pi ^t  =
1251 > 0$ and $Q^T Q = 1$. For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a
1252 > matrix $\hat v \in so(3)^ \star$, the hat-map isomorphism,
1253   \begin{equation}
1254   v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1255   {\begin{array}{*{20}c}
# Line 1342 | Line 1262 | operations
1262   will let us associate the matrix products with traditional vector
1263   operations
1264   \[
1265 < \hat vu = v \times u
1265 > \hat vu = v \times u.
1266   \]
1267 <
1348 < Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1267 > Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew
1268   matrix,
1269 < \begin{equation}
1270 < (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1271 < ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1272 < - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1273 < (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1274 < \end{equation}
1275 < Since $\Lambda$ is symmetric, the last term of Equation
1276 < \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1277 < multiplier $\Lambda$ is absent from the equations of motion. This
1278 < unique property eliminate the requirement of iterations which can
1279 < not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1280 <
1362 < Applying hat-map isomorphism, we obtain the equation of motion for
1363 < angular momentum on body frame
1269 > \begin{eqnarray}
1270 > (\dot \Pi  - \dot \Pi ^T )&= &(\Pi  - \Pi ^T )(J^{ - 1} \Pi  + \Pi J^{ - 1} ) \notag \\
1271 > & & + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  -
1272 > (\Lambda  - \Lambda ^T ). \label{introEquation:skewMatrixPI}
1273 > \end{eqnarray}
1274 > Since $\Lambda$ is symmetric, the last term of
1275 > Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1276 > Lagrange multiplier $\Lambda$ is absent from the equations of
1277 > motion. This unique property eliminates the requirement of
1278 > iterations which can not be avoided in other methods\cite{Kol1997,
1279 > Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1280 > equation of motion for angular momentum on body frame
1281   \begin{equation}
1282   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1283   F_i (r,Q)} \right) \times X_i }.
# Line 1369 | Line 1286 | given by
1286   In the same manner, the equation of motion for rotation matrix is
1287   given by
1288   \[
1289 < \dot Q = Qskew(I^{ - 1} \pi )
1289 > \dot Q = Qskew(I^{ - 1} \pi ).
1290   \]
1291  
1292   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1293   Lie-Poisson Integrator for Free Rigid Body}
1294  
1295 < If there is not external forces exerted on the rigid body, the only
1296 < contribution to the rotational is from the kinetic potential (the
1297 < first term of \ref{ introEquation:bodyAngularMotion}). The free
1298 < rigid body is an example of Lie-Poisson system with Hamiltonian
1295 > If there are no external forces exerted on the rigid body, the only
1296 > contribution to the rotational motion is from the kinetic energy
1297 > (the first term of \ref{introEquation:bodyAngularMotion}). The free
1298 > rigid body is an example of a Lie-Poisson system with Hamiltonian
1299   function
1300   \begin{equation}
1301   T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
# Line 1391 | Line 1308 | J(\pi ) = \left( {\begin{array}{*{20}c}
1308     0 & {\pi _3 } & { - \pi _2 }  \\
1309     { - \pi _3 } & 0 & {\pi _1 }  \\
1310     {\pi _2 } & { - \pi _1 } & 0  \\
1311 < \end{array}} \right)
1311 > \end{array}} \right).
1312   \end{equation}
1313   Thus, the dynamics of free rigid body is governed by
1314   \begin{equation}
1315 < \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1315 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi ).
1316   \end{equation}
1317 <
1318 < One may notice that each $T_i^r$ in Equation
1319 < \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1403 < instance, the equations of motion due to $T_1^r$ are given by
1317 > One may notice that each $T_i^r$ in
1318 > Eq.~\ref{introEquation:rotationalKineticRB} can be solved exactly.
1319 > For instance, the equations of motion due to $T_1^r$ are given by
1320   \begin{equation}
1321   \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1322   \label{introEqaution:RBMotionSingleTerm}
1323   \end{equation}
1324 < where
1324 > with
1325   \[ R_1  = \left( {\begin{array}{*{20}c}
1326     0 & 0 & 0  \\
1327     0 & 0 & {\pi _1 }  \\
1328     0 & { - \pi _1 } & 0  \\
1329   \end{array}} \right).
1330   \]
1331 < The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1331 > The solutions of Eq.~\ref{introEqaution:RBMotionSingleTerm} is
1332   \[
1333   \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1334   Q(0)e^{\Delta tR_1 }
# Line 1426 | Line 1342 | tR_1 }$, we can use Cayley transformation,
1342   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1343   \]
1344   To reduce the cost of computing expensive functions in $e^{\Delta
1345 < tR_1 }$, we can use Cayley transformation,
1345 > tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1346 > propagator,
1347   \[
1348   e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1349 < )
1349 > ).
1350   \]
1351   The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1352 < manner.
1353 <
1437 < In order to construct a second-order symplectic method, we split the
1438 < angular kinetic Hamiltonian function can into five terms
1352 > manner. In order to construct a second-order symplectic method, we
1353 > split the angular kinetic Hamiltonian function into five terms
1354   \[
1355   T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1356   ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1357 < (\pi _1 )
1358 < \].
1359 < Concatenating flows corresponding to these five terms, we can obtain
1360 < an symplectic integrator,
1357 > (\pi _1 ).
1358 > \]
1359 > By concatenating the propagators corresponding to these five terms,
1360 > we can obtain an symplectic integrator,
1361   \[
1362   \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1363   \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1364   \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1365   _1 }.
1366   \]
1452
1367   The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1368   $F(\pi )$ and $G(\pi )$ is defined by
1369   \[
1370   \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1371 < )
1371 > ).
1372   \]
1373   If the Poisson bracket of a function $F$ with an arbitrary smooth
1374   function $G$ is zero, $F$ is a \emph{Casimir}, which is the
# Line 1465 | Line 1379 | then by the chain rule
1379   then by the chain rule
1380   \[
1381   \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1382 < }}{2})\pi
1382 > }}{2})\pi.
1383   \]
1384 < Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1384 > Thus, $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel
1385 > \pi
1386   \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1387 < Lie-Poisson integrator is found to be extremely efficient and stable
1388 < which can be explained by the fact the small angle approximation is
1389 < used and the norm of the angular momentum is conserved.
1387 > Lie-Poisson integrator is found to be both extremely efficient and
1388 > stable. These properties can be explained by the fact the small
1389 > angle approximation is used and the norm of the angular momentum is
1390 > conserved.
1391  
1392   \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1393   Splitting for Rigid Body}
1394  
1395   The Hamiltonian of rigid body can be separated in terms of kinetic
1396 < energy and potential energy,
1397 < \[
1398 < H = T(p,\pi ) + V(q,Q)
1483 < \]
1484 < The equations of motion corresponding to potential energy and
1485 < kinetic energy are listed in the below table,
1396 > energy and potential energy,$H = T(p,\pi ) + V(q,Q)$. The equations
1397 > of motion corresponding to potential energy and kinetic energy are
1398 > listed in the below table,
1399   \begin{table}
1400 < \caption{Equations of motion due to Potential and Kinetic Energies}
1400 > \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1401   \begin{center}
1402   \begin{tabular}{|l|l|}
1403    \hline
# Line 1498 | Line 1411 | A second-order symplectic method is now obtained by th
1411   \end{tabular}
1412   \end{center}
1413   \end{table}
1414 < A second-order symplectic method is now obtained by the
1415 < composition of the flow maps,
1414 > A second-order symplectic method is now obtained by the composition
1415 > of the position and velocity propagators,
1416   \[
1417   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1418   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1419   \]
1420   Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1421 < sub-flows which corresponding to force and torque respectively,
1421 > sub-propagators which corresponding to force and torque
1422 > respectively,
1423   \[
1424   \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1425   _{\Delta t/2,\tau }.
1426   \]
1427   Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1428 < $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1429 < order inside $\varphi _{\Delta t/2,V}$ does not matter.
1430 <
1431 < Furthermore, kinetic potential can be separated to translational
1518 < kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1428 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1429 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1430 > kinetic energy can be separated to translational kinetic term, $T^t
1431 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1432   \begin{equation}
1433   T(p,\pi ) =T^t (p) + T^r (\pi ).
1434   \end{equation}
1435   where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1436 < defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1437 < corresponding flow maps are given by
1436 > defined by Eq.~\ref{introEquation:rotationalKineticRB}. Therefore,
1437 > the corresponding propagators are given by
1438   \[
1439   \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1440   _{\Delta t,T^r }.
1441   \]
1442 < Finally, we obtain the overall symplectic flow maps for free moving
1443 < rigid body
1444 < \begin{equation}
1445 < \begin{array}{c}
1446 < \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1447 <  \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 }  \\
1535 <  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1536 < \end{array}
1442 > Finally, we obtain the overall symplectic propagators for freely
1443 > moving rigid bodies
1444 > \begin{eqnarray}
1445 > \varphi _{\Delta t}  &=& \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \notag\\
1446 >  & & \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\\
1447 >  & & \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1448   \label{introEquation:overallRBFlowMaps}
1449 < \end{equation}
1449 > \end{eqnarray}
1450  
1451   \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1452   As an alternative to newtonian dynamics, Langevin dynamics, which
1453   mimics a simple heat bath with stochastic and dissipative forces,
1454   has been applied in a variety of studies. This section will review
1455 < the theory of Langevin dynamics simulation. A brief derivation of
1456 < generalized Langevin equation will be given first. Follow that, we
1457 < will discuss the physical meaning of the terms appearing in the
1458 < equation as well as the calculation of friction tensor from
1459 < hydrodynamics theory.
1455 > the theory of Langevin dynamics. A brief derivation of generalized
1456 > Langevin equation will be given first. Following that, we will
1457 > discuss the physical meaning of the terms appearing in the equation
1458 > as well as the calculation of friction tensor from hydrodynamics
1459 > theory.
1460  
1461   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1462  
1463 < Harmonic bath model, in which an effective set of harmonic
1463 > A harmonic bath model, in which an effective set of harmonic
1464   oscillators are used to mimic the effect of a linearly responding
1465   environment, has been widely used in quantum chemistry and
1466   statistical mechanics. One of the successful applications of
1467 < Harmonic bath model is the derivation of Deriving Generalized
1468 < Langevin Dynamics. Lets consider a system, in which the degree of
1467 > Harmonic bath model is the derivation of the Generalized Langevin
1468 > Dynamics (GLE). Lets consider a system, in which the degree of
1469   freedom $x$ is assumed to couple to the bath linearly, giving a
1470   Hamiltonian of the form
1471   \begin{equation}
1472   H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1473   \label{introEquation:bathGLE}.
1474   \end{equation}
1475 < Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1476 < with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1475 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1476 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1477   \[
1478   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1479   }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
# Line 1570 | Line 1481 | the harmonic bath masses, and $\Delta U$ is bilinear s
1481   \]
1482   where the index $\alpha$ runs over all the bath degrees of freedom,
1483   $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1484 < the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1484 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1485   coupling,
1486   \[
1487   \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1488   \]
1489 < where $g_\alpha$ are the coupling constants between the bath and the
1490 < coordinate $x$. Introducing
1489 > where $g_\alpha$ are the coupling constants between the bath
1490 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1491 > Introducing
1492   \[
1493   W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1494   }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1495 < \] and combining the last two terms in Equation
1496 < \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1585 < Hamiltonian as
1495 > \]
1496 > and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as
1497   \[
1498   H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1499   {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1500   w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1501 < w_\alpha ^2 }}x} \right)^2 } \right\}}
1501 > w_\alpha ^2 }}x} \right)^2 } \right\}}.
1502   \]
1503   Since the first two terms of the new Hamiltonian depend only on the
1504   system coordinates, we can get the equations of motion for
1505 < Generalized Langevin Dynamics by Hamilton's equations
1595 < \ref{introEquation:motionHamiltonianCoordinate,
1596 < introEquation:motionHamiltonianMomentum},
1505 > Generalized Langevin Dynamics by Hamilton's equations,
1506   \begin{equation}
1507   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1508   \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
# Line 1606 | Line 1515 | m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x
1515   \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1516   \label{introEquation:bathMotionGLE}
1517   \end{equation}
1609
1518   In order to derive an equation for $x$, the dynamics of the bath
1519   variables $x_\alpha$ must be solved exactly first. As an integral
1520   transform which is particularly useful in solving linear ordinary
1521 < differential equations, Laplace transform is the appropriate tool to
1522 < solve this problem. The basic idea is to transform the difficult
1521 > differential equations,the Laplace transform is the appropriate tool
1522 > to solve this problem. The basic idea is to transform the difficult
1523   differential equations into simple algebra problems which can be
1524 < solved easily. Then applying inverse Laplace transform, also known
1525 < as the Bromwich integral, we can retrieve the solutions of the
1526 < original problems.
1527 <
1620 < Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1621 < transform of f(t) is a new function defined as
1524 > solved easily. Then, by applying the inverse Laplace transform, also
1525 > known as the Bromwich integral, we can retrieve the solutions of the
1526 > original problems. Let $f(t)$ be a function defined on $ [0,\infty )
1527 > $, the Laplace transform of $f(t)$ is a new function defined as
1528   \[
1529   L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1530   \]
1531   where  $p$ is real and  $L$ is called the Laplace Transform
1532   Operator. Below are some important properties of Laplace transform
1627
1533   \begin{eqnarray*}
1534   L(x + y)  & = & L(x) + L(y) \\
1535   L(ax)     & = & aL(x) \\
# Line 1632 | Line 1537 | Operator. Below are some important properties of Lapla
1537   L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1538   L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1539   \end{eqnarray*}
1540 <
1636 <
1637 < Applying Laplace transform to the bath coordinates, we obtain
1540 > Applying the Laplace transform to the bath coordinates, we obtain
1541   \begin{eqnarray*}
1542 < 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) \\
1543 < L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1542 > 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), \\
1543 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}. \\
1544   \end{eqnarray*}
1642
1545   By the same way, the system coordinates become
1546   \begin{eqnarray*}
1547 < mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1548 <  & & \mbox{} - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1547 > mL(\ddot x) & = &
1548 >  - \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\}}  \\
1549 >  & & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}.
1550   \end{eqnarray*}
1648
1551   With the help of some relatively important inverse Laplace
1552   transformations:
1553   \[
# Line 1655 | Line 1557 | transformations:
1557   L(1) = \frac{1}{p} \\
1558   \end{array}
1559   \]
1560 < , we obtain
1560 > we obtain
1561   \begin{eqnarray*}
1562 < m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} -
1562 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1563   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1564   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1565 < _\alpha  t)\dot x(t - \tau )d\tau  \\
1566 < & &\mbox{} - \left[ {g_\alpha  x_\alpha (0) - \frac{{g_\alpha
1567 < }}{{m_\alpha \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) -
1568 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1569 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}} \\
1570 < %
1571 < & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1565 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1566 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1567 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1568 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1569 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1570 > \end{eqnarray*}
1571 > \begin{eqnarray*}
1572 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1573   {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1574   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1575 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1576 < {\left[ {g_\alpha  x_\alpha  (0) \\
1577 < & & \mbox{} - \frac{{g_\alpha  }}{{m_\alpha \omega _\alpha  }}}
1575 > t)\dot x(t - \tau )d} \tau }  \\
1576 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1577 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1578   \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1579   (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1580   \end{eqnarray*}
# Line 1697 | Line 1600 | which is known as the \emph{generalized Langevin equat
1600   \end{equation}
1601   which is known as the \emph{generalized Langevin equation}.
1602  
1603 < \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1603 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1604  
1605   One may notice that $R(t)$ depends only on initial conditions, which
1606   implies it is completely deterministic within the context of a
1607   harmonic bath. However, it is easy to verify that $R(t)$ is totally
1608 < uncorrelated to $x$ and $\dot x$,
1609 < \[
1610 < \begin{array}{l}
1611 < \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1612 < \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1710 < \end{array}
1711 < \]
1712 < This property is what we expect from a truly random process. As long
1713 < as the model, which is gaussian distribution in general, chosen for
1714 < $R(t)$ is a truly random process, the stochastic nature of the GLE
1715 < still remains.
1716 <
1608 > uncorrelated to $x$ and $\dot x$,$\left\langle {x(t)R(t)}
1609 > \right\rangle  = 0, \left\langle {\dot x(t)R(t)} \right\rangle  =
1610 > 0.$ This property is what we expect from a truly random process. As
1611 > long as the model chosen for $R(t)$ was a gaussian distribution in
1612 > general, the stochastic nature of the GLE still remains.
1613   %dynamic friction kernel
1614   The convolution integral
1615   \[
# Line 1728 | Line 1624 | and Equation \ref{introEuqation:GeneralizedLangevinDyn
1624   \[
1625   \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1626   \]
1627 < and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1627 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1628   \[
1629   m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1630   \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1631   \]
1632 < which can be used to describe dynamic caging effect. The other
1633 < extreme is the bath that responds infinitely quickly to motions in
1634 < the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1635 < time:
1632 > which can be used to describe the effect of dynamic caging in
1633 > viscous solvents. The other extreme is the bath that responds
1634 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1635 > taken as a $delta$ function in time:
1636   \[
1637   \xi (t) = 2\xi _0 \delta (t)
1638   \]
# Line 1745 | Line 1641 | and Equation \ref{introEuqation:GeneralizedLangevinDyn
1641   \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1642   {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1643   \]
1644 < and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1644 > and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1645   \begin{equation}
1646   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1647   x(t) + R(t) \label{introEquation:LangevinEquation}
1648   \end{equation}
1649   which is known as the Langevin equation. The static friction
1650   coefficient $\xi _0$ can either be calculated from spectral density
1651 < or be determined by Stokes' law for regular shaped particles.A
1651 > or be determined by Stokes' law for regular shaped particles. A
1652   briefly review on calculating friction tensor for arbitrary shaped
1653   particles is given in Sec.~\ref{introSection:frictionTensor}.
1654  
1655 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1655 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1656  
1657 < Defining a new set of coordinates,
1657 > Defining a new set of coordinates
1658   \[
1659   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1660 < ^2 }}x(0)
1661 < \],
1660 > ^2 }}x(0),
1661 > \]
1662   we can rewrite $R(T)$ as
1663   \[
1664   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1665   \]
1666   And since the $q$ coordinates are harmonic oscillators,
1771
1667   \begin{eqnarray*}
1668   \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1669   \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
# Line 1777 | Line 1672 | And since the $q$ coordinates are harmonic oscillators
1672    & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1673    & = &kT\xi (t) \\
1674   \end{eqnarray*}
1780
1675   Thus, we recover the \emph{second fluctuation dissipation theorem}
1676   \begin{equation}
1677   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1678 < \label{introEquation:secondFluctuationDissipation}.
1678 > \label{introEquation:secondFluctuationDissipation},
1679   \end{equation}
1680 < In effect, it acts as a constraint on the possible ways in which one
1681 < can model the random force and friction kernel.
1788 <
1789 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
1790 < Theoretically, the friction kernel can be determined using velocity
1791 < autocorrelation function. However, this approach become impractical
1792 < when the system become more and more complicate. Instead, various
1793 < approaches based on hydrodynamics have been developed to calculate
1794 < the friction coefficients. The friction effect is isotropic in
1795 < Equation, $\zeta$ can be taken as a scalar. In general, friction
1796 < tensor $\Xi$ is a $6\times 6$ matrix given by
1797 < \[
1798 < \Xi  = \left( {\begin{array}{*{20}c}
1799 <   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1800 <   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1801 < \end{array}} \right).
1802 < \]
1803 < Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1804 < tensor and rotational resistance (friction) tensor respectively,
1805 < while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1806 < {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1807 < particle moves in a fluid, it may experience friction force or
1808 < torque along the opposite direction of the velocity or angular
1809 < velocity,
1810 < \[
1811 < \left( \begin{array}{l}
1812 < F_R  \\
1813 < \tau _R  \\
1814 < \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1815 <   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1816 <   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1817 < \end{array}} \right)\left( \begin{array}{l}
1818 < v \\
1819 < w \\
1820 < \end{array} \right)
1821 < \]
1822 < where $F_r$ is the friction force and $\tau _R$ is the friction
1823 < toque.
1824 <
1825 < \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1826 <
1827 < For a spherical particle, the translational and rotational friction
1828 < constant can be calculated from Stoke's law,
1829 < \[
1830 < \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1831 <   {6\pi \eta R} & 0 & 0  \\
1832 <   0 & {6\pi \eta R} & 0  \\
1833 <   0 & 0 & {6\pi \eta R}  \\
1834 < \end{array}} \right)
1835 < \]
1836 < and
1837 < \[
1838 < \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1839 <   {8\pi \eta R^3 } & 0 & 0  \\
1840 <   0 & {8\pi \eta R^3 } & 0  \\
1841 <   0 & 0 & {8\pi \eta R^3 }  \\
1842 < \end{array}} \right)
1843 < \]
1844 < where $\eta$ is the viscosity of the solvent and $R$ is the
1845 < hydrodynamics radius.
1846 <
1847 < Other non-spherical shape, such as cylinder and ellipsoid
1848 < \textit{etc}, are widely used as reference for developing new
1849 < hydrodynamics theory, because their properties can be calculated
1850 < exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1851 < also called a triaxial ellipsoid, which is given in Cartesian
1852 < coordinates by\cite{Perrin1934, Perrin1936}
1853 < \[
1854 < \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1855 < }} = 1
1856 < \]
1857 < where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1858 < due to the complexity of the elliptic integral, only the ellipsoid
1859 < with the restriction of two axes having to be equal, \textit{i.e.}
1860 < prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1861 < exactly. Introducing an elliptic integral parameter $S$ for prolate,
1862 < \[
1863 < S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1864 < } }}{b},
1865 < \]
1866 < and oblate,
1867 < \[
1868 < S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1869 < }}{a}
1870 < \],
1871 < one can write down the translational and rotational resistance
1872 < tensors
1873 < \[
1874 < \begin{array}{l}
1875 < \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1876 < \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1877 < \end{array},
1878 < \]
1879 < and
1880 < \[
1881 < \begin{array}{l}
1882 < \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1883 < \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1884 < \end{array}.
1885 < \]
1886 <
1887 < \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1888 <
1889 < Unlike spherical and other regular shaped molecules, there is not
1890 < analytical solution for friction tensor of any arbitrary shaped
1891 < rigid molecules. The ellipsoid of revolution model and general
1892 < triaxial ellipsoid model have been used to approximate the
1893 < hydrodynamic properties of rigid bodies. However, since the mapping
1894 < from all possible ellipsoidal space, $r$-space, to all possible
1895 < combination of rotational diffusion coefficients, $D$-space is not
1896 < unique\cite{Wegener1979} as well as the intrinsic coupling between
1897 < translational and rotational motion of rigid body, general ellipsoid
1898 < is not always suitable for modeling arbitrarily shaped rigid
1899 < molecule. A number of studies have been devoted to determine the
1900 < friction tensor for irregularly shaped rigid bodies using more
1901 < advanced method where the molecule of interest was modeled by
1902 < combinations of spheres(beads)\cite{Carrasco1999} and the
1903 < hydrodynamics properties of the molecule can be calculated using the
1904 < hydrodynamic interaction tensor. Let us consider a rigid assembly of
1905 < $N$ beads immersed in a continuous medium. Due to hydrodynamics
1906 < interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1907 < than its unperturbed velocity $v_i$,
1908 < \[
1909 < v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1910 < \]
1911 < where $F_i$ is the frictional force, and $T_{ij}$ is the
1912 < hydrodynamic interaction tensor. The friction force of $i$th bead is
1913 < proportional to its ``net'' velocity
1914 < \begin{equation}
1915 < F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1916 < \label{introEquation:tensorExpression}
1917 < \end{equation}
1918 < This equation is the basis for deriving the hydrodynamic tensor. In
1919 < 1930, Oseen and Burgers gave a simple solution to Equation
1920 < \ref{introEquation:tensorExpression}
1921 < \begin{equation}
1922 < T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1923 < R_{ij}^T }}{{R_{ij}^2 }}} \right).
1924 < \label{introEquation:oseenTensor}
1925 < \end{equation}
1926 < Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1927 < A second order expression for element of different size was
1928 < introduced by Rotne and Prager\cite{Rotne1969} and improved by
1929 < Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1930 < \begin{equation}
1931 < T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1932 < \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1933 < _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1934 < \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1935 < \label{introEquation:RPTensorNonOverlapped}
1936 < \end{equation}
1937 < Both of the Equation \ref{introEquation:oseenTensor} and Equation
1938 < \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1939 < \ge \sigma _i  + \sigma _j$. An alternative expression for
1940 < overlapping beads with the same radius, $\sigma$, is given by
1941 < \begin{equation}
1942 < T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1943 < \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1944 < \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1945 < \label{introEquation:RPTensorOverlapped}
1946 < \end{equation}
1947 <
1948 < To calculate the resistance tensor at an arbitrary origin $O$, we
1949 < construct a $3N \times 3N$ matrix consisting of $N \times N$
1950 < $B_{ij}$ blocks
1951 < \begin{equation}
1952 < B = \left( {\begin{array}{*{20}c}
1953 <   {B_{11} } &  \ldots  & {B_{1N} }  \\
1954 <    \vdots  &  \ddots  &  \vdots   \\
1955 <   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1956 < \end{array}} \right),
1957 < \end{equation}
1958 < where $B_{ij}$ is given by
1959 < \[
1960 < B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1961 < )T_{ij}
1962 < \]
1963 < where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1964 < $B$, we obtain
1965 <
1966 < \[
1967 < C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1968 <   {C_{11} } &  \ldots  & {C_{1N} }  \\
1969 <    \vdots  &  \ddots  &  \vdots   \\
1970 <   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1971 < \end{array}} \right)
1972 < \]
1973 < , which can be partitioned into $N \times N$ $3 \times 3$ block
1974 < $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1975 < \[
1976 < U_i  = \left( {\begin{array}{*{20}c}
1977 <   0 & { - z_i } & {y_i }  \\
1978 <   {z_i } & 0 & { - x_i }  \\
1979 <   { - y_i } & {x_i } & 0  \\
1980 < \end{array}} \right)
1981 < \]
1982 < where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1983 < bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1984 < arbitrary origin $O$ can be written as
1985 < \begin{equation}
1986 < \begin{array}{l}
1987 < \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1988 < \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1989 < \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1990 < \end{array}
1991 < \label{introEquation:ResistanceTensorArbitraryOrigin}
1992 < \end{equation}
1993 <
1994 < The resistance tensor depends on the origin to which they refer. The
1995 < proper location for applying friction force is the center of
1996 < resistance (reaction), at which the trace of rotational resistance
1997 < tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1998 < resistance is defined as an unique point of the rigid body at which
1999 < the translation-rotation coupling tensor are symmetric,
2000 < \begin{equation}
2001 < \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
2002 < \label{introEquation:definitionCR}
2003 < \end{equation}
2004 < Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2005 < we can easily find out that the translational resistance tensor is
2006 < origin independent, while the rotational resistance tensor and
2007 < translation-rotation coupling resistance tensor depend on the
2008 < origin. Given resistance tensor at an arbitrary origin $O$, and a
2009 < vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2010 < obtain the resistance tensor at $P$ by
2011 < \begin{equation}
2012 < \begin{array}{l}
2013 < \Xi _P^{tt}  = \Xi _O^{tt}  \\
2014 < \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2015 < \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2016 < \end{array}
2017 < \label{introEquation:resistanceTensorTransformation}
2018 < \end{equation}
2019 < where
2020 < \[
2021 < U_{OP}  = \left( {\begin{array}{*{20}c}
2022 <   0 & { - z_{OP} } & {y_{OP} }  \\
2023 <   {z_i } & 0 & { - x_{OP} }  \\
2024 <   { - y_{OP} } & {x_{OP} } & 0  \\
2025 < \end{array}} \right)
2026 < \]
2027 < Using Equations \ref{introEquation:definitionCR} and
2028 < \ref{introEquation:resistanceTensorTransformation}, one can locate
2029 < the position of center of resistance,
2030 < \begin{eqnarray*}
2031 < \left( \begin{array}{l}
2032 < x_{OR}  \\
2033 < y_{OR}  \\
2034 < z_{OR}  \\
2035 < \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2036 <   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2037 <   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2038 <   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2039 < \end{array}} \right)^{ - 1}  \\
2040 <  & & \left( \begin{array}{l}
2041 < (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2042 < (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2043 < (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2044 < \end{array} \right) \\
2045 < \end{eqnarray*}
2046 <
2047 <
2048 <
2049 < where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2050 < joining center of resistance $R$ and origin $O$.
1680 > which acts as a constraint on the possible ways in which one can
1681 > model the random force and friction kernel.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines