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 2792 by tim, Mon Jun 5 21:23:04 2006 UTC vs.
Revision 2898 by tim, Tue Jun 27 03:22:42 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
# Line 68 | Line 67 | scheme for rigid body \cite{Dullweber1997}.
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}.
70 > schemes for rigid bodies \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 two important limitations: motions
75 > can only be described in cartesian coordinate systems. Moreover, it
76 > becomes impossible to predict analytically the properties of the
77 > system even if we know all of the details of the interaction. In
78 > order to overcome some of the practical difficulties which arise in
79 > attempts to apply Newton's equation to complex system, approximate
80 > numerical 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} ,
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 ) ,
102   \label{introEquation:lagrangianDef}
# Line 114 | Line 107 | then Eq.~\ref{introEquation:halmitonianPrinciple1} bec
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
# Line 172 | Line 161 | find
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}
188
177   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
# Line 193 | Line 181 | function of the generalized velocities $\dot q_i$ and
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}.
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  
192   In Newtonian Mechanics, a system described by conservative forces
193   conserves the total energy \ref{introEquation:energyConservation}.
# Line 230 | Line 217 | momentum variables. Consider a dynamic system in a car
217   possible states. Each possible state of the system corresponds to
218   one unique point in the phase space. For mechanical systems, the
219   phase space usually consists of all possible values of position and
220 < momentum variables. Consider a dynamic system in a cartesian space,
221 < where each of the $6f$ coordinates and momenta is assigned to one of
222 < $6f$ mutually orthogonal axes, the phase space of this system is a
223 < $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
224 < \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
225 < momenta is a phase space vector.
220 > momentum variables. Consider a dynamic system of $f$ particles in a
221 > cartesian space, where each of the $6f$ coordinates and momenta is
222 > assigned to one of $6f$ mutually orthogonal axes, the phase space of
223 > this system is a $6f$ dimensional space. A point, $x = (\rightarrow
224 > q_1 , \ldots ,\rightarrow q_f ,\rightarrow p_1 , \ldots ,\rightarrow
225 > p_f )$, with a unique set of values of $6f$ coordinates and momenta
226 > is a phase space vector.
227 > %%%fix me
228  
229 < 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
229 > In statistical mechanics, the condition of an ensemble at any time
230   can be regarded as appropriately specified by the density $\rho$
231   with which representative points are distributed over the phase
232 < space. The density of distribution for an ensemble with $f$ degrees
233 < of freedom is defined as,
232 > space. The density distribution for an ensemble with $f$ degrees of
233 > freedom is defined as,
234   \begin{equation}
235   \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
236   \label{introEquation:densityDistribution}
237   \end{equation}
238   Governed by the principles of mechanics, the phase points change
239 < their value which would change the density at any time at phase
240 < space. Hence, the density of distribution is also to be taken as a
239 > their locations which would change the density at any time at phase
240 > space. Hence, the density distribution is also to be taken as a
241   function of the time.
242  
243   The number of systems $\delta N$ at time $t$ can be determined by,
# Line 273 | Line 245 | Assuming a large enough population of systems are expl
245   \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
246   \label{introEquation:deltaN}
247   \end{equation}
248 < Assuming a large enough population of systems are exploited, we can
249 < sufficiently approximate $\delta N$ without introducing
250 < discontinuity when we go from one region in the phase space to
251 < another. By integrating over the whole phase space,
248 > Assuming a large enough population of systems, we can sufficiently
249 > approximate $\delta N$ without introducing discontinuity when we go
250 > from one region in the phase space to another. By integrating over
251 > the whole phase space,
252   \begin{equation}
253   N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
254   \label{introEquation:totalNumberSystem}
# Line 288 | Line 260 | With the help of Equation(\ref{introEquation:unitProba
260   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
261   \label{introEquation:unitProbability}
262   \end{equation}
263 < With the help of Equation(\ref{introEquation:unitProbability}) and
264 < the knowledge of the system, it is possible to calculate the average
263 > With the help of Eq.~\ref{introEquation:unitProbability} and the
264 > knowledge of the system, it is possible to calculate the average
265   value of any desired quantity which depends on the coordinates and
266   momenta of the system. Even when the dynamics of the real system is
267   complex, or stochastic, or even discontinuous, the average
268 < properties of the ensemble of possibilities as a whole may still
269 < remain well defined. For a classical system in thermal equilibrium
270 < with its environment, the ensemble average of a mechanical quantity,
271 < $\langle A(q , p) \rangle_t$, takes the form of an integral over the
272 < phase space of the system,
268 > properties of the ensemble of possibilities as a whole remaining
269 > well defined. For a classical system in thermal equilibrium with its
270 > environment, the ensemble average of a mechanical quantity, $\langle
271 > A(q , p) \rangle_t$, takes the form of an integral over the phase
272 > space of the system,
273   \begin{equation}
274   \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
275   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
# Line 307 | Line 279 | parameters, such as temperature \textit{etc}, partitio
279  
280   There are several different types of ensembles with different
281   statistical characteristics. As a function of macroscopic
282 < parameters, such as temperature \textit{etc}, partition function can
283 < be used to describe the statistical properties of a system in
284 < thermodynamic equilibrium.
285 <
286 < 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,
282 > parameters, such as temperature \textit{etc}, the partition function
283 > can be used to describe the statistical properties of a system in
284 > thermodynamic equilibrium. As an ensemble of systems, each of which
285 > is known to be thermally isolated and conserve energy, the
286 > Microcanonical ensemble (NVE) has a partition function like,
287   \begin{equation}
288   \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
289   \end{equation}
290 < A canonical ensemble(NVT)is an ensemble of systems, each of which
290 > A canonical ensemble (NVT)is an ensemble of systems, each of which
291   can share its energy with a large heat reservoir. The distribution
292   of the total energy amongst the possible dynamical states is given
293   by the partition function,
# Line 326 | Line 296 | TS$. Since most experiment are carried out under const
296   \label{introEquation:NVTPartition}
297   \end{equation}
298   Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
299 < TS$. Since most experiment are carried out under constant pressure
300 < condition, isothermal-isobaric ensemble(NPT) play a very important
301 < role in molecular simulation. The isothermal-isobaric ensemble allow
302 < the system to exchange energy with a heat bath of temperature $T$
303 < and to change the volume as well. Its partition function is given as
299 > TS$. Since most experiments are carried out under constant pressure
300 > condition, the isothermal-isobaric ensemble (NPT) plays a very
301 > important role in molecular simulations. The isothermal-isobaric
302 > ensemble allow the system to exchange energy with a heat bath of
303 > temperature $T$ and to change the volume as well. Its partition
304 > function is given as
305   \begin{equation}
306   \Delta (N,P,T) =  - e^{\beta G}.
307   \label{introEquation:NPTPartition}
# Line 339 | Line 310 | The Liouville's theorem is the foundation on which sta
310  
311   \subsection{\label{introSection:liouville}Liouville's theorem}
312  
313 < The Liouville's theorem is the foundation on which statistical
314 < mechanics rests. It describes the time evolution of phase space
313 > Liouville's theorem is the foundation on which statistical mechanics
314 > rests. It describes the time evolution of the phase space
315   distribution function. In order to calculate the rate of change of
316 < $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
317 < consider the two faces perpendicular to the $q_1$ axis, which are
318 < located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
319 < leaving the opposite face is given by the expression,
316 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
317 > the two faces perpendicular to the $q_1$ axis, which are located at
318 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
319 > opposite face is given by the expression,
320   \begin{equation}
321   \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
322   \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
# Line 369 | Line 340 | divining $ \delta q_1  \ldots \delta q_f \delta p_1  \
340   + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
341   \end{equation}
342   which cancels the first terms of the right hand side. Furthermore,
343 < divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
343 > dividing $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
344   p_f $ in both sides, we can write out Liouville's theorem in a
345   simple form,
346   \begin{equation}
# Line 381 | Line 352 | statistical mechanics, since the number of particles i
352  
353   Liouville's theorem states that the distribution function is
354   constant along any trajectory in phase space. In classical
355 < statistical mechanics, since the number of particles in the system
356 < is huge, we may be able to believe the system is stationary,
355 > statistical mechanics, since the number of members in an ensemble is
356 > huge and constant, we can assume the local density has no reason
357 > (other than classical mechanics) to change,
358   \begin{equation}
359   \frac{{\partial \rho }}{{\partial t}} = 0.
360   \label{introEquation:stationary}
# Line 395 | Line 367 | distribution,
367   \label{introEquation:densityAndHamiltonian}
368   \end{equation}
369  
370 < \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
370 > \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
371   Lets consider a region in the phase space,
372   \begin{equation}
373   \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
374   \end{equation}
375   If this region is small enough, the density $\rho$ can be regarded
376 < as uniform over the whole phase space. Thus, the number of phase
377 < points inside this region is given by,
376 > as uniform over the whole integral. Thus, the number of phase points
377 > inside this region is given by,
378   \begin{equation}
379   \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
380   dp_1 } ..dp_f.
# Line 414 | Line 386 | With the help of stationary assumption
386   \end{equation}
387   With the help of stationary assumption
388   (\ref{introEquation:stationary}), we obtain the principle of the
389 < \emph{conservation of extension in phase space},
389 > \emph{conservation of volume in phase space},
390   \begin{equation}
391   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
392   ...dq_f dp_1 } ..dp_f  = 0.
393   \label{introEquation:volumePreserving}
394   \end{equation}
395  
396 < \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
396 > \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
397  
398   Liouville's theorem can be expresses in a variety of different forms
399   which are convenient within different contexts. For any two function
# Line 435 | Line 407 | Substituting equations of motion in Hamiltonian formal
407   \label{introEquation:poissonBracket}
408   \end{equation}
409   Substituting equations of motion in Hamiltonian formalism(
410 < \ref{introEquation:motionHamiltonianCoordinate} ,
411 < \ref{introEquation:motionHamiltonianMomentum} ) into
412 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
413 < theorem using Poisson bracket notion,
410 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
411 > Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
412 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
413 > Liouville's theorem using Poisson bracket notion,
414   \begin{equation}
415   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
416   {\rho ,H} \right\}.
# Line 463 | Line 435 | simulation and the quality of the underlying model. Ho
435   Various thermodynamic properties can be calculated from Molecular
436   Dynamics simulation. By comparing experimental values with the
437   calculated properties, one can determine the accuracy of the
438 < simulation and the quality of the underlying model. However, both of
439 < experiment and computer simulation are usually performed during a
438 > simulation and the quality of the underlying model. However, both
439 > experiments and computer simulations are usually performed during a
440   certain time interval and the measurements are averaged over a
441   period of them which is different from the average behavior of
442 < many-body system in Statistical Mechanics. Fortunately, Ergodic
443 < Hypothesis is proposed to make a connection between time average and
444 < ensemble average. It states that time average and average over the
442 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
443 > Hypothesis makes a connection between time average and the ensemble
444 > average. It states that the time average and average over the
445   statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
446   \begin{equation}
447   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
# Line 491 | Line 463 | A variety of numerical integrators were proposed to si
463   choice\cite{Frenkel1996}.
464  
465   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
466 < A variety of numerical integrators were proposed to simulate the
467 < motions. They usually begin with an initial conditionals and move
468 < the objects in the direction governed by the differential equations.
469 < However, most of them ignore the hidden physical law contained
470 < within the equations. Since 1990, geometric integrators, which
471 < preserve various phase-flow invariants such as symplectic structure,
472 < volume and time reversal symmetry, are developed to address this
473 < issue\cite{Dullweber1997, McLachlan1998, Leimkuhler1999}. The
474 < velocity verlet method, which happens to be a simple example of
475 < symplectic integrator, continues to gain its popularity in molecular
476 < dynamics community. This fact can be partly explained by its
477 < geometric nature.
466 > A variety of numerical integrators have been proposed to simulate
467 > the motions of atoms in MD simulation. They usually begin with
468 > initial conditionals and move the objects in the direction governed
469 > by the differential equations. However, most of them ignore the
470 > hidden physical laws contained within the equations. Since 1990,
471 > geometric integrators, which preserve various phase-flow invariants
472 > such as symplectic structure, volume and time reversal symmetry, are
473 > developed to address this issue\cite{Dullweber1997, McLachlan1998,
474 > Leimkuhler1999}. The velocity Verlet method, which happens to be a
475 > simple example of symplectic integrator, continues to gain
476 > popularity in the molecular dynamics community. This fact can be
477 > partly explained by its geometric nature.
478  
479 < \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
480 < A \emph{manifold} is an abstract mathematical space. It locally
481 < looks like Euclidean space, but when viewed globally, it may have
482 < more complicate structure. A good example of manifold is the surface
483 < of Earth. It seems to be flat locally, but it is round if viewed as
484 < a whole. A \emph{differentiable manifold} (also known as
485 < \emph{smooth manifold}) is a manifold with an open cover in which
486 < the covering neighborhoods are all smoothly isomorphic to one
487 < 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
479 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
480 > A \emph{manifold} is an abstract mathematical space. It looks
481 > locally like Euclidean space, but when viewed globally, it may have
482 > more complicated structure. A good example of manifold is the
483 > surface of Earth. It seems to be flat locally, but it is round if
484 > viewed as a whole. A \emph{differentiable manifold} (also known as
485 > \emph{smooth manifold}) is a manifold on which it is possible to
486 > apply calculus on \emph{differentiable manifold}. A \emph{symplectic
487 > manifold} is defined as a pair $(M, \omega)$ which consists of a
488   \emph{differentiable manifold} $M$ and a close, non-degenerated,
489   bilinear symplectic form, $\omega$. A symplectic form on a vector
490   space $V$ is a function $\omega(x, y)$ which satisfies
491   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
492   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
493 < $\omega(x, x) = 0$. Cross product operation in vector field is an
494 < example of symplectic form.
493 > $\omega(x, x) = 0$. The cross product operation in vector field is
494 > an example of symplectic form.
495  
496 < One of the motivations to study \emph{symplectic manifold} in
496 > One of the motivations to study \emph{symplectic manifolds} in
497   Hamiltonian Mechanics is that a symplectic manifold can represent
498   all possible configurations of the system and the phase space of the
499   system can be described by it's cotangent bundle. Every symplectic
500   manifold is even dimensional. For instance, in Hamilton equations,
501   coordinate and momentum always appear in pairs.
502  
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
503   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
504  
505 < For a ordinary differential system defined as
505 > For an ordinary differential system defined as
506   \begin{equation}
507   \dot x = f(x)
508   \end{equation}
509 < where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
509 > where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
510   \begin{equation}
511   f(r) = J\nabla _x H(r).
512   \end{equation}
# Line 605 | Line 566 | Instead, we use a approximate map, $\psi_\tau$, which
566   \end{equation}
567  
568   In most cases, it is not easy to find the exact flow $\varphi_\tau$.
569 < Instead, we use a approximate map, $\psi_\tau$, which is usually
569 > Instead, we use an approximate map, $\psi_\tau$, which is usually
570   called integrator. The order of an integrator $\psi_\tau$ is $p$, if
571   the Taylor series of $\psi_\tau$ agree to order $p$,
572   \begin{equation}
573 < \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
573 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
574   \end{equation}
575  
576   \subsection{\label{introSection:geometricProperties}Geometric Properties}
577  
578 < The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
579 < and its flow play important roles in numerical studies. Many of them
580 < can be found in systems which occur naturally in applications.
620 <
578 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
579 > ODE and its flow play important roles in numerical studies. Many of
580 > them can be found in systems which occur naturally in applications.
581   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
582   a \emph{symplectic} flow if it satisfies,
583   \begin{equation}
# Line 631 | Line 591 | is the property must be preserved by the integrator.
591   \begin{equation}
592   {\varphi '}^T J \varphi ' = J \circ \varphi
593   \end{equation}
594 < is the property must be preserved by the integrator.
595 <
596 < It is possible to construct a \emph{volume-preserving} flow for a
597 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
598 < \det d\varphi  = 1$. One can show easily that a symplectic flow will
599 < be volume-preserving.
640 <
641 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
642 < will result in a new system,
594 > is the property that must be preserved by the integrator. It is
595 > possible to construct a \emph{volume-preserving} flow for a source
596 > free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $ \det
597 > d\varphi  = 1$. One can show easily that a symplectic flow will be
598 > volume-preserving. Changing the variables $y = h(x)$ in an ODE
599 > (Eq.~\ref{introEquation:ODE}) will result in a new system,
600   \[
601   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
602   \]
603   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
604   In other words, the flow of this vector field is reversible if and
605 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
606 <
650 < A \emph{first integral}, or conserved quantity of a general
605 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
606 > \emph{first integral}, or conserved quantity of a general
607   differential function is a function $ G:R^{2d}  \to R^d $ which is
608   constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
609   \[
# Line 660 | Line 616 | smooth function $G$ is given by,
616   which is the condition for conserving \emph{first integral}. For a
617   canonical Hamiltonian system, the time evolution of an arbitrary
618   smooth function $G$ is given by,
663
619   \begin{eqnarray}
620   \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
621                          & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
622   \label{introEquation:firstIntegral1}
623   \end{eqnarray}
669
670
624   Using poisson bracket notion, Equation
625   \ref{introEquation:firstIntegral1} can be rewritten as
626   \[
# Line 680 | Line 633 | is a \emph{first integral}, which is due to the fact $
633   \]
634   As well known, the Hamiltonian (or energy) H of a Hamiltonian system
635   is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
636 < 0$.
684 <
685 < When designing any numerical methods, one should always try to
636 > 0$. When designing any numerical methods, one should always try to
637   preserve the structural properties of the original ODE and its flow.
638  
639   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
640   A lot of well established and very effective numerical methods have
641   been successful precisely because of their symplecticities even
642   though this fact was not recognized when they were first
643 < constructed. The most famous example is leapfrog methods in
644 < molecular dynamics. In general, symplectic integrators can be
643 > constructed. The most famous example is the Verlet-leapfrog method
644 > in molecular dynamics. In general, symplectic integrators can be
645   constructed using one of four different methods.
646   \begin{enumerate}
647   \item Generating functions
# Line 708 | Line 659 | implementing the Runge-Kutta methods, they do not attr
659   high-order explicit Runge-Kutta methods
660   \cite{Owren1992,Chen2003}have been developed to overcome this
661   instability. However, due to computational penalty involved in
662 < implementing the Runge-Kutta methods, they do not attract too much
663 < attention from Molecular Dynamics community. Instead, splitting have
664 < been widely accepted since they exploit natural decompositions of
665 < the system\cite{Tuckerman1992, McLachlan1998}.
662 > implementing the Runge-Kutta methods, they have not attracted much
663 > attention from the Molecular Dynamics community. Instead, splitting
664 > methods have been widely accepted since they exploit natural
665 > decompositions of the system\cite{Tuckerman1992, McLachlan1998}.
666  
667 < \subsubsection{\label{introSection:splittingMethod}Splitting Method}
667 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
668  
669   The main idea behind splitting methods is to decompose the discrete
670   $\varphi_h$ as a composition of simpler flows,
# Line 723 | Line 674 | simpler integration of the system.
674   \label{introEquation:FlowDecomposition}
675   \end{equation}
676   where each of the sub-flow is chosen such that each represent a
677 < simpler integration of the system.
678 <
728 < Suppose that a Hamiltonian system takes the form,
677 > simpler integration of the system. Suppose that a Hamiltonian system
678 > takes the form,
679   \[
680   H = H_1 + H_2.
681   \]
# Line 734 | Line 684 | order is then given by the Lie-Trotter formula
684   energy respectively, which is a natural decomposition of the
685   problem. If $H_1$ and $H_2$ can be integrated using exact flows
686   $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
687 < order is then given by the Lie-Trotter formula
687 > order expression is then given by the Lie-Trotter formula
688   \begin{equation}
689   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
690   \label{introEquation:firstOrderSplitting}
# Line 760 | Line 710 | which has a local error proportional to $h^3$. Sprang
710   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
711   _{1,h/2} , \label{introEquation:secondOrderSplitting}
712   \end{equation}
713 < which has a local error proportional to $h^3$. Sprang splitting's
714 < popularity in molecular simulation community attribute to its
715 < symmetric property,
713 > which has a local error proportional to $h^3$. The Sprang
714 > splitting's popularity in molecular simulation community attribute
715 > to its symmetric property,
716   \begin{equation}
717   \varphi _h^{ - 1} = \varphi _{ - h}.
718   \label{introEquation:timeReversible}
719   \end{equation}
720  
721 < \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
721 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
722   The classical equation for a system consisting of interacting
723   particles can be written in Hamiltonian form,
724   \[
725   H = T + V
726   \]
727   where $T$ is the kinetic energy and $V$ is the potential energy.
728 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
728 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
729   obtains the following:
730   \begin{align}
731   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 802 | Line 752 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
752      \label{introEquation:Lp9b}\\%
753   %
754   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
755 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
755 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
756   \end{align}
757   From the preceding splitting, one can see that the integration of
758   the equations of motion would follow:
# Line 811 | Line 761 | the equations of motion would follow:
761  
762   \item Use the half step velocities to move positions one whole step, $\Delta t$.
763  
764 < \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
764 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
765  
766   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
767   \end{enumerate}
768  
769 < Simply switching the order of splitting and composing, a new
770 < integrator, the \emph{position verlet} integrator, can be generated,
769 > By simply switching the order of the propagators in the splitting
770 > and composing a new integrator, the \emph{position verlet}
771 > integrator, can be generated,
772   \begin{align}
773   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
774   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 828 | Line 779 | q(\Delta t)} \right]. %
779   \label{introEquation:positionVerlet2}
780   \end{align}
781  
782 < \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
782 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
783  
784 < Baker-Campbell-Hausdorff formula can be used to determine the local
785 < error of splitting method in terms of commutator of the
784 > The Baker-Campbell-Hausdorff formula can be used to determine the
785 > local error of splitting method in terms of the commutator of the
786   operators(\ref{introEquation:exponentialOperator}) associated with
787 < the sub-flow. For operators $hX$ and $hY$ which are associate to
787 > the sub-flow. For operators $hX$ and $hY$ which are associated with
788   $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
789   \begin{equation}
790   \exp (hX + hY) = \exp (hZ)
# Line 847 | Line 798 | Applying Baker-Campbell-Hausdorff formula\cite{Varadar
798   \[
799   [X,Y] = XY - YX .
800   \]
801 < Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
802 < Sprang splitting, we can obtain
801 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
802 > to the Sprang splitting, we can obtain
803   \begin{eqnarray*}
804   \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2 [X,Y]/4 + h^2 [Y,X]/4 \\
805                                     &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
806                                     &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
807   \end{eqnarray*}
808 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
808 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
809   error of Spring splitting is proportional to $h^3$. The same
810 < procedure can be applied to general splitting,  of the form
810 > procedure can be applied to a general splitting,  of the form
811   \begin{equation}
812   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
813   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
814   \end{equation}
815 < Careful choice of coefficient $a_1 \ldot b_m$ will lead to higher
816 < order method. Yoshida proposed an elegant way to compose higher
815 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
816 > order methods. Yoshida proposed an elegant way to compose higher
817   order methods based on symmetric splitting\cite{Yoshida1990}. Given
818   a symmetric second order base method $ \varphi _h^{(2)} $, a
819   fourth-order symmetric method can be constructed by composing,
# Line 875 | Line 826 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
826   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
827   \begin{equation}
828   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
829 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
829 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
830   \end{equation}
831 < , if the weights are chosen as
831 > if the weights are chosen as
832   \[
833   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
834   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 914 | Line 865 | initialization of a simulation. Sec.~\ref{introSec:pro
865   \end{enumerate}
866   These three individual steps will be covered in the following
867   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
868 < initialization of a simulation. Sec.~\ref{introSec:production} will
869 < discusses issues in production run. Sec.~\ref{introSection:Analysis}
870 < provides the theoretical tools for trajectory analysis.
868 > initialization of a simulation. Sec.~\ref{introSection:production}
869 > will discusse issues in production run.
870 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
871 > trajectory analysis.
872  
873   \subsection{\label{introSec:initialSystemSettings}Initialization}
874  
875 < \subsubsection{Preliminary preparation}
875 > \subsubsection{\textbf{Preliminary preparation}}
876  
877   When selecting the starting structure of a molecule for molecular
878   simulation, one may retrieve its Cartesian coordinates from public
879   databases, such as RCSB Protein Data Bank \textit{etc}. Although
880   thousands of crystal structures of molecules are discovered every
881   year, many more remain unknown due to the difficulties of
882 < purification and crystallization. Even for the molecule with known
883 < structure, some important information is missing. For example, the
882 > purification and crystallization. Even for molecules with known
883 > structure, some important information is missing. For example, a
884   missing hydrogen atom which acts as donor in hydrogen bonding must
885   be added. Moreover, in order to include electrostatic interaction,
886   one may need to specify the partial charges for individual atoms.
887   Under some circumstances, we may even need to prepare the system in
888 < a special setup. For instance, when studying transport phenomenon in
889 < membrane system, we may prepare the lipids in bilayer structure
890 < instead of placing lipids randomly in solvent, since we are not
891 < interested in self-aggregation and it takes a long time to happen.
888 > a special configuration. For instance, when studying transport
889 > phenomenon in membrane systems, we may prepare the lipids in a
890 > bilayer structure instead of placing lipids randomly in solvent,
891 > since we are not interested in the slow self-aggregation process.
892  
893 < \subsubsection{Minimization}
893 > \subsubsection{\textbf{Minimization}}
894  
895   It is quite possible that some of molecules in the system from
896 < preliminary preparation may be overlapped with each other. This
897 < close proximity leads to high potential energy which consequently
898 < jeopardizes any molecular dynamics simulations. To remove these
899 < steric overlaps, one typically performs energy minimization to find
900 < a more reasonable conformation. Several energy minimization methods
901 < have been developed to exploit the energy surface and to locate the
902 < local minimum. While converging slowly near the minimum, steepest
903 < descent method is extremely robust when systems are far from
904 < harmonic. Thus, it is often used to refine structure from
905 < crystallographic data. Relied on the gradient or hessian, advanced
906 < methods like conjugate gradient and Newton-Raphson converge rapidly
907 < to a local minimum, while become unstable if the energy surface is
908 < far from quadratic. Another factor must be taken into account, when
896 > preliminary preparation may be overlapping with each other. This
897 > close proximity leads to high initial potential energy which
898 > consequently jeopardizes any molecular dynamics simulations. To
899 > remove these steric overlaps, one typically performs energy
900 > minimization to find a more reasonable conformation. Several energy
901 > minimization methods have been developed to exploit the energy
902 > surface and to locate the local minimum. While converging slowly
903 > near the minimum, steepest descent method is extremely robust when
904 > systems are strongly anharmonic. Thus, it is often used to refine
905 > structure from crystallographic data. Relied on the gradient or
906 > hessian, advanced methods like Newton-Raphson converge rapidly to a
907 > local minimum, but become unstable if the energy surface is far from
908 > quadratic. Another factor that must be taken into account, when
909   choosing energy minimization method, is the size of the system.
910   Steepest descent and conjugate gradient can deal with models of any
911 < size. Because of the limit of computation power to calculate hessian
912 < matrix and insufficient storage capacity to store them, most
913 < Newton-Raphson methods can not be used with very large models.
911 > size. Because of the limits on computer memory to store the hessian
912 > matrix and the computing power needed to diagonalized these
913 > matrices, most Newton-Raphson methods can not be used with very
914 > large systems.
915  
916 < \subsubsection{Heating}
916 > \subsubsection{\textbf{Heating}}
917  
918   Typically, Heating is performed by assigning random velocities
919 < according to a Gaussian distribution for a temperature. Beginning at
920 < a lower temperature and gradually increasing the temperature by
921 < assigning greater random velocities, we end up with setting the
922 < temperature of the system to a final temperature at which the
923 < simulation will be conducted. In heating phase, we should also keep
924 < the system from drifting or rotating as a whole. Equivalently, the
925 < net linear momentum and angular momentum of the system should be
926 < shifted to zero.
919 > according to a Maxwell-Boltzman distribution for a desired
920 > temperature. Beginning at a lower temperature and gradually
921 > increasing the temperature by assigning larger random velocities, we
922 > end up with setting the temperature of the system to a final
923 > temperature at which the simulation will be conducted. In heating
924 > phase, we should also keep the system from drifting or rotating as a
925 > whole. To do this, the net linear momentum and angular momentum of
926 > the system is shifted to zero after each resampling from the Maxwell
927 > -Boltzman distribution.
928  
929 < \subsubsection{Equilibration}
929 > \subsubsection{\textbf{Equilibration}}
930  
931   The purpose of equilibration is to allow the system to evolve
932   spontaneously for a period of time and reach equilibrium. The
# Line 986 | Line 940 | Production run is the most important step of the simul
940  
941   \subsection{\label{introSection:production}Production}
942  
943 < Production run is the most important step of the simulation, in
943 > The production run is the most important step of the simulation, in
944   which the equilibrated structure is used as a starting point and the
945   motions of the molecules are collected for later analysis. In order
946   to capture the macroscopic properties of the system, the molecular
947 < dynamics simulation must be performed in correct and efficient way.
947 > dynamics simulation must be performed by sampling correctly and
948 > efficiently from the relevant thermodynamic ensemble.
949  
950   The most expensive part of a molecular dynamics simulation is the
951   calculation of non-bonded forces, such as van der Waals force and
952   Coulombic forces \textit{etc}. For a system of $N$ particles, the
953   complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
954   which making large simulations prohibitive in the absence of any
955 < computation saving techniques.
955 > algorithmic tricks.
956  
957 < A natural approach to avoid system size issue is to represent the
957 > A natural approach to avoid system size issues is to represent the
958   bulk behavior by a finite number of the particles. However, this
959 < approach will suffer from the surface effect. To offset this,
960 < \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
961 < is developed to simulate bulk properties with a relatively small
962 < number of particles. In this method, the simulation box is
963 < replicated throughout space to form an infinite lattice. During the
964 < simulation, when a particle moves in the primary cell, its image in
965 < other cells move in exactly the same direction with exactly the same
966 < orientation. Thus, as a particle leaves the primary cell, one of its
967 < images will enter through the opposite face.
959 > approach will suffer from the surface effect at the edges of the
960 > simulation. To offset this, \textit{Periodic boundary conditions}
961 > (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
962 > properties with a relatively small number of particles. In this
963 > method, the simulation box is replicated throughout space to form an
964 > infinite lattice. During the simulation, when a particle moves in
965 > the primary cell, its image in other cells move in exactly the same
966 > direction with exactly the same orientation. Thus, as a particle
967 > leaves the primary cell, one of its images will enter through the
968 > opposite face.
969   \begin{figure}
970   \centering
971   \includegraphics[width=\linewidth]{pbc.eps}
# Line 1021 | Line 977 | evaluation is to apply cutoff where particles farther
977  
978   %cutoff and minimum image convention
979   Another important technique to improve the efficiency of force
980 < evaluation is to apply cutoff where particles farther than a
981 < predetermined distance, are not included in the calculation
980 > evaluation is to apply spherical cutoff where particles farther than
981 > a predetermined distance are not included in the calculation
982   \cite{Frenkel1996}. The use of a cutoff radius will cause a
983   discontinuity in the potential energy curve. Fortunately, one can
984 < shift the potential to ensure the potential curve go smoothly to
985 < zero at the cutoff radius. Cutoff strategy works pretty well for
986 < Lennard-Jones interaction because of its short range nature.
987 < However, simply truncating the electrostatic interaction with the
988 < use of cutoff has been shown to lead to severe artifacts in
989 < simulations. Ewald summation, in which the slowly conditionally
990 < convergent Coulomb potential is transformed into direct and
991 < reciprocal sums with rapid and absolute convergence, has proved to
992 < minimize the periodicity artifacts in liquid simulations. Taking the
993 < advantages of the fast Fourier transform (FFT) for calculating
994 < discrete Fourier transforms, the particle mesh-based
984 > shift simple radial potential to ensure the potential curve go
985 > smoothly to zero at the cutoff radius. The cutoff strategy works
986 > well for Lennard-Jones interaction because of its short range
987 > nature. However, simply truncating the electrostatic interaction
988 > with the use of cutoffs has been shown to lead to severe artifacts
989 > in simulations. The Ewald summation, in which the slowly decaying
990 > Coulomb potential is transformed into direct and reciprocal sums
991 > with rapid and absolute convergence, has proved to minimize the
992 > periodicity artifacts in liquid simulations. Taking the advantages
993 > of the fast Fourier transform (FFT) for calculating discrete Fourier
994 > transforms, the particle mesh-based
995   methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
996 < $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
997 < multipole method}\cite{Greengard1987, Greengard1994}, which treats
998 < Coulombic interaction exactly at short range, and approximate the
999 < potential at long range through multipolar expansion. In spite of
1000 < their wide acceptances at the molecular simulation community, these
1001 < two methods are hard to be implemented correctly and efficiently.
1002 < Instead, we use a damped and charge-neutralized Coulomb potential
1003 < method developed by Wolf and his coworkers\cite{Wolf1999}. The
1004 < shifted Coulomb potential for particle $i$ and particle $j$ at
1005 < distance $r_{rj}$ is given by:
996 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
997 > \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
998 > which treats Coulombic interactions exactly at short range, and
999 > approximate the potential at long range through multipolar
1000 > expansion. In spite of their wide acceptance at the molecular
1001 > simulation community, these two methods are difficult to implement
1002 > correctly and efficiently. Instead, we use a damped and
1003 > charge-neutralized Coulomb potential method developed by Wolf and
1004 > his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
1005 > particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1006   \begin{equation}
1007   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1008   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
# Line 1068 | Line 1024 | Recently, advanced visualization technique are widely
1024  
1025   \subsection{\label{introSection:Analysis} Analysis}
1026  
1027 < Recently, advanced visualization technique are widely applied to
1027 > Recently, advanced visualization technique have become applied to
1028   monitor the motions of molecules. Although the dynamics of the
1029   system can be described qualitatively from animation, quantitative
1030 < trajectory analysis are more appreciable. According to the
1031 < principles of Statistical Mechanics,
1032 < Sec.~\ref{introSection:statisticalMechanics}, one can compute
1033 < thermodynamics properties, analyze fluctuations of structural
1034 < parameters, and investigate time-dependent processes of the molecule
1079 < from the trajectories.
1030 > trajectory analysis are more useful. According to the principles of
1031 > Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
1032 > one can compute thermodynamic properties, analyze fluctuations of
1033 > structural parameters, and investigate time-dependent processes of
1034 > the molecule from the trajectories.
1035  
1036 < \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1036 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1037  
1038 < Thermodynamics properties, which can be expressed in terms of some
1038 > Thermodynamic properties, which can be expressed in terms of some
1039   function of the coordinates and momenta of all particles in the
1040   system, can be directly computed from molecular dynamics. The usual
1041   way to measure the pressure is based on virial theorem of Clausius
# Line 1100 | Line 1055 | P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\
1055   < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1056   \end{equation}
1057  
1058 < \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1058 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1059  
1060   Structural Properties of a simple fluid can be described by a set of
1061 < distribution functions. Among these functions,\emph{pair
1061 > distribution functions. Among these functions,the \emph{pair
1062   distribution function}, also known as \emph{radial distribution
1063 < function}, is of most fundamental importance to liquid-state theory.
1064 < Pair distribution function can be gathered by Fourier transforming
1065 < raw data from a series of neutron diffraction experiments and
1066 < integrating over the surface factor \cite{Powles1973}. The
1067 < experiment result can serve as a criterion to justify the
1068 < correctness of the theory. Moreover, various equilibrium
1069 < thermodynamic and structural properties can also be expressed in
1070 < terms of radial distribution function \cite{Allen1987}.
1063 > function}, is of most fundamental importance to liquid theory.
1064 > Experimentally, pair distribution function can be gathered by
1065 > Fourier transforming raw data from a series of neutron diffraction
1066 > experiments and integrating over the surface factor
1067 > \cite{Powles1973}. The experimental results can serve as a criterion
1068 > to justify the correctness of a liquid model. Moreover, various
1069 > equilibrium thermodynamic and structural properties can also be
1070 > expressed in terms of radial distribution function \cite{Allen1987}.
1071  
1072 < A pair distribution functions $g(r)$ gives the probability that a
1072 > The pair distribution functions $g(r)$ gives the probability that a
1073   particle $i$ will be located at a distance $r$ from a another
1074   particle $j$ in the system
1075   \[
1076   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1077 < \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1077 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1078 > (r)}{\rho}.
1079   \]
1080   Note that the delta function can be replaced by a histogram in
1081 < computer simulation. Figure
1082 < \ref{introFigure:pairDistributionFunction} shows a typical pair
1083 < 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.
1081 > computer simulation. Peaks in $g(r)$ represent solvent shells, and
1082 > the height of these peaks gradually decreases to 1 as the liquid of
1083 > large distance approaches the bulk density.
1084  
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}
1085  
1086 < \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1087 < Properties}
1086 > \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1087 > Properties}}
1088  
1089   Time-dependent properties are usually calculated using \emph{time
1090 < correlation function}, which correlates random variables $A$ and $B$
1091 < at two different time
1090 > correlation functions}, which correlate random variables $A$ and $B$
1091 > at two different times,
1092   \begin{equation}
1093   C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1094   \label{introEquation:timeCorrelationFunction}
1095   \end{equation}
1096   If $A$ and $B$ refer to same variable, this kind of correlation
1097 < function is called \emph{auto correlation function}. One example of
1098 < auto correlation function is velocity auto-correlation function
1099 < which is directly related to transport properties of molecular
1100 < liquids:
1097 > function is called an \emph{autocorrelation function}. One example
1098 > of an auto correlation function is the velocity auto-correlation
1099 > function which is directly related to transport properties of
1100 > molecular liquids:
1101   \[
1102   D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1103   \right\rangle } dt
1104   \]
1105 < where $D$ is diffusion constant. Unlike velocity autocorrelation
1106 < function which is averaging over time origins and over all the
1107 < atoms, dipole autocorrelation are calculated for the entire system.
1108 < The dipole autocorrelation function is given by:
1105 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1106 > function, which is averaging over time origins and over all the
1107 > atoms, the dipole autocorrelation functions are calculated for the
1108 > entire system. The dipole autocorrelation function is given by:
1109   \[
1110   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1111   \right\rangle
# Line 1188 | Line 1131 | simulator is governed by the rigid body dynamics. In m
1131   areas, from engineering, physics, to chemistry. For example,
1132   missiles and vehicle are usually modeled by rigid bodies.  The
1133   movement of the objects in 3D gaming engine or other physics
1134 < simulator is governed by the rigid body dynamics. In molecular
1135 < simulation, rigid body is used to simplify the model in
1136 < protein-protein docking study\cite{Gray2003}.
1134 > simulator is governed by rigid body dynamics. In molecular
1135 > simulations, rigid bodies are used to simplify protein-protein
1136 > docking studies\cite{Gray2003}.
1137  
1138   It is very important to develop stable and efficient methods to
1139 < integrate the equations of motion of orientational degrees of
1140 < freedom. Euler angles are the nature choice to describe the
1141 < rotational degrees of freedom. However, due to its singularity, the
1142 < numerical integration of corresponding equations of motion is very
1143 < inefficient and inaccurate. Although an alternative integrator using
1144 < different sets of Euler angles can overcome this
1145 < difficulty\cite{Barojas1973}, the computational penalty and the lost
1146 < of angular momentum conservation still remain. A singularity free
1147 < representation utilizing quaternions was developed by Evans in
1148 < 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1149 < nonseparable Hamiltonian resulted from quaternion representation,
1150 < which prevents the symplectic algorithm to be utilized. Another
1151 < different approach is to apply holonomic constraints to the atoms
1152 < belonging to the rigid body. Each atom moves independently under the
1153 < normal forces deriving from potential energy and constraint forces
1154 < which are used to guarantee the rigidness. However, due to their
1155 < iterative nature, SHAKE and Rattle algorithm converge very slowly
1156 < when the number of constraint increases\cite{Ryckaert1977,
1157 < Andersen1983}.
1139 > integrate the equations of motion for orientational degrees of
1140 > freedom. Euler angles are the natural choice to describe the
1141 > rotational degrees of freedom. However, due to $\frac {1}{sin
1142 > \theta}$ singularities, the numerical integration of corresponding
1143 > equations of motion is very inefficient and inaccurate. Although an
1144 > alternative integrator using multiple sets of Euler angles can
1145 > overcome this difficulty\cite{Barojas1973}, the computational
1146 > penalty and the loss of angular momentum conservation still remain.
1147 > A singularity-free representation utilizing quaternions was
1148 > developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1149 > approach uses a nonseparable Hamiltonian resulting from the
1150 > quaternion representation, which prevents the symplectic algorithm
1151 > to be utilized. Another different approach is to apply holonomic
1152 > constraints to the atoms belonging to the rigid body. Each atom
1153 > moves independently under the normal forces deriving from potential
1154 > energy and constraint forces which are used to guarantee the
1155 > rigidness. However, due to their iterative nature, the SHAKE and
1156 > Rattle algorithms also converge very slowly when the number of
1157 > constraints increases\cite{Ryckaert1977, Andersen1983}.
1158  
1159 < The break through in geometric literature suggests that, in order to
1159 > A break-through in geometric literature suggests that, in order to
1160   develop a long-term integration scheme, one should preserve the
1161 < symplectic structure of the flow. Introducing conjugate momentum to
1162 < rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1163 < symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1164 < the Hamiltonian system in a constraint manifold by iteratively
1165 < satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1166 < method using quaternion representation was developed by
1167 < Omelyan\cite{Omelyan1998}. However, both of these methods are
1168 < iterative and inefficient. In this section, we will present a
1161 > symplectic structure of the flow. By introducing a conjugate
1162 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1163 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1164 > proposed to evolve the Hamiltonian system in a constraint manifold
1165 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1166 > An alternative method using the quaternion representation was
1167 > developed by Omelyan\cite{Omelyan1998}. However, both of these
1168 > methods are iterative and inefficient. In this section, we descibe a
1169   symplectic Lie-Poisson integrator for rigid body developed by
1170   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1171  
1172 < \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1173 < The motion of the rigid body is Hamiltonian with the Hamiltonian
1172 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1173 > The motion of a rigid body is Hamiltonian with the Hamiltonian
1174   function
1175   \begin{equation}
1176   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
# Line 1241 | Line 1184 | constrained Hamiltonian equation subjects to a holonom
1184   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1185   \]
1186   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1187 < constrained Hamiltonian equation subjects to a holonomic constraint,
1187 > constrained Hamiltonian equation is subjected to a holonomic
1188 > constraint,
1189   \begin{equation}
1190   Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1191   \end{equation}
1192 < which is used to ensure rotation matrix's orthogonality.
1193 < Differentiating \ref{introEquation:orthogonalConstraint} and using
1194 < Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1192 > which is used to ensure rotation matrix's unitarity. Differentiating
1193 > \ref{introEquation:orthogonalConstraint} and using Equation
1194 > \ref{introEquation:RBMotionMomentum}, one may obtain,
1195   \begin{equation}
1196   Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1197   \label{introEquation:RBFirstOrderConstraint}
# Line 1256 | Line 1200 | the equations of motion,
1200   Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1201   \ref{introEquation:motionHamiltonianMomentum}), one can write down
1202   the equations of motion,
1259 \[
1260 \begin{array}{c}
1261 \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1262 \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1263 \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1264 \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1265 \end{array}
1266 \]
1203  
1204 + \begin{eqnarray}
1205 + \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1206 + \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1207 + \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1208 + \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1209 + \end{eqnarray}
1210 +
1211   In general, there are two ways to satisfy the holonomic constraints.
1212 < We can use constraint force provided by lagrange multiplier on the
1213 < normal manifold to keep the motion on constraint space. Or we can
1214 < simply evolve the system in constraint manifold. These two methods
1215 < are proved to be equivalent. The holonomic constraint and equations
1216 < of motions define a constraint manifold for rigid body
1212 > We can use a constraint force provided by a Lagrange multiplier on
1213 > the normal manifold to keep the motion on constraint space. Or we
1214 > can simply evolve the system on the constraint manifold. These two
1215 > methods have been proved to be equivalent. The holonomic constraint
1216 > and equations of motions define a constraint manifold for rigid
1217 > bodies
1218   \[
1219   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1220   \right\}.
1221   \]
1222  
1223   Unfortunately, this constraint manifold is not the cotangent bundle
1224 < $T_{\star}SO(3)$. However, it turns out that under symplectic
1224 > $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1225 > rotation group $SO(3)$. However, it turns out that under symplectic
1226   transformation, the cotangent space and the phase space are
1227 < diffeomorphic. Introducing
1227 > diffeomorphic. By introducing
1228   \[
1229   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1230   \]
# Line 1311 | Line 1256 | body, angular momentum on body frame $\Pi  = Q^t P$ is
1256   respectively.
1257  
1258   As a common choice to describe the rotation dynamics of the rigid
1259 < body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1260 < rewrite the equations of motion,
1259 > body, the angular momentum on the body fixed frame $\Pi  = Q^t P$ is
1260 > introduced to rewrite the equations of motion,
1261   \begin{equation}
1262   \begin{array}{l}
1263 < \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1264 < \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1263 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1264 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1}  \\
1265   \end{array}
1266   \label{introEqaution:RBMotionPI}
1267   \end{equation}
# Line 1344 | Line 1289 | operations
1289   \[
1290   \hat vu = v \times u
1291   \]
1347
1292   Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1293   matrix,
1294 < \begin{equation}
1295 < (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1296 < ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1297 < - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1298 < (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1299 < \end{equation}
1294 >
1295 > \begin{eqnarray*}
1296 > (\dot \Pi  - \dot \Pi ^T ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{
1297 > }}(J^{ - 1} \Pi  + \Pi J^{ - 1} ) + \sum\limits_i {[Q^T F_i
1298 > (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  - (\Lambda  - \Lambda ^T ).
1299 > \label{introEquation:skewMatrixPI}
1300 > \end{eqnarray*}
1301 >
1302   Since $\Lambda$ is symmetric, the last term of Equation
1303   \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1304   multiplier $\Lambda$ is absent from the equations of motion. This
1305 < unique property eliminate the requirement of iterations which can
1305 > unique property eliminates the requirement of iterations which can
1306   not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1307  
1308 < Applying hat-map isomorphism, we obtain the equation of motion for
1309 < angular momentum on body frame
1308 > Applying the hat-map isomorphism, we obtain the equation of motion
1309 > for angular momentum on body frame
1310   \begin{equation}
1311   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1312   F_i (r,Q)} \right) \times X_i }.
# Line 1375 | Line 1321 | If there is not external forces exerted on the rigid b
1321   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1322   Lie-Poisson Integrator for Free Rigid Body}
1323  
1324 < If there is not external forces exerted on the rigid body, the only
1325 < contribution to the rotational is from the kinetic potential (the
1326 < first term of \ref{ introEquation:bodyAngularMotion}). The free
1327 < rigid body is an example of Lie-Poisson system with Hamiltonian
1324 > If there are no external forces exerted on the rigid body, the only
1325 > contribution to the rotational motion is from the kinetic energy
1326 > (the first term of \ref{introEquation:bodyAngularMotion}). The free
1327 > rigid body is an example of a Lie-Poisson system with Hamiltonian
1328   function
1329   \begin{equation}
1330   T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
# Line 1426 | Line 1372 | tR_1 }$, we can use Cayley transformation,
1372   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1373   \]
1374   To reduce the cost of computing expensive functions in $e^{\Delta
1375 < tR_1 }$, we can use Cayley transformation,
1375 > tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1376 > propagator,
1377   \[
1378   e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1379   )
1380   \]
1381   The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1382 < manner.
1383 <
1437 < In order to construct a second-order symplectic method, we split the
1438 < angular kinetic Hamiltonian function can into five terms
1382 > manner. In order to construct a second-order symplectic method, we
1383 > split the angular kinetic Hamiltonian function can into five terms
1384   \[
1385   T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1386   ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1387 < (\pi _1 )
1388 < \].
1389 < Concatenating flows corresponding to these five terms, we can obtain
1390 < an symplectic integrator,
1387 > (\pi _1 ).
1388 > \]
1389 > By concatenating the propagators corresponding to these five terms,
1390 > we can obtain an symplectic integrator,
1391   \[
1392   \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1393   \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
# Line 1469 | Line 1414 | Lie-Poisson integrator is found to be extremely effici
1414   \]
1415   Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1416   \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1417 < Lie-Poisson integrator is found to be extremely efficient and stable
1418 < which can be explained by the fact the small angle approximation is
1419 < used and the norm of the angular momentum is conserved.
1417 > Lie-Poisson integrator is found to be both extremely efficient and
1418 > stable. These properties can be explained by the fact the small
1419 > angle approximation is used and the norm of the angular momentum is
1420 > conserved.
1421  
1422   \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1423   Splitting for Rigid Body}
# Line 1484 | Line 1430 | kinetic energy are listed in the below table,
1430   The equations of motion corresponding to potential energy and
1431   kinetic energy are listed in the below table,
1432   \begin{table}
1433 < \caption{Equations of motion due to Potential and Kinetic Energies}
1433 > \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1434   \begin{center}
1435   \begin{tabular}{|l|l|}
1436    \hline
# Line 1498 | Line 1444 | A second-order symplectic method is now obtained by th
1444   \end{tabular}
1445   \end{center}
1446   \end{table}
1447 < A second-order symplectic method is now obtained by the
1448 < composition of the flow maps,
1447 > A second-order symplectic method is now obtained by the composition
1448 > of the position and velocity propagators,
1449   \[
1450   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1451   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1452   \]
1453   Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1454 < sub-flows which corresponding to force and torque respectively,
1454 > sub-propagators which corresponding to force and torque
1455 > respectively,
1456   \[
1457   \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1458   _{\Delta t/2,\tau }.
1459   \]
1460   Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1461 < $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1462 < order inside $\varphi _{\Delta t/2,V}$ does not matter.
1463 <
1464 < Furthermore, kinetic potential can be separated to translational
1518 < kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1461 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1462 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1463 > kinetic energy can be separated to translational kinetic term, $T^t
1464 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1465   \begin{equation}
1466   T(p,\pi ) =T^t (p) + T^r (\pi ).
1467   \end{equation}
1468   where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1469   defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1470 < corresponding flow maps are given by
1470 > corresponding propagators are given by
1471   \[
1472   \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1473   _{\Delta t,T^r }.
1474   \]
1475 < Finally, we obtain the overall symplectic flow maps for free moving
1476 < rigid body
1475 > Finally, we obtain the overall symplectic propagators for freely
1476 > moving rigid bodies
1477   \begin{equation}
1478   \begin{array}{c}
1479   \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
# Line 1541 | Line 1487 | the theory of Langevin dynamics simulation. A brief de
1487   As an alternative to newtonian dynamics, Langevin dynamics, which
1488   mimics a simple heat bath with stochastic and dissipative forces,
1489   has been applied in a variety of studies. This section will review
1490 < the theory of Langevin dynamics simulation. A brief derivation of
1491 < generalized Langevin equation will be given first. Follow that, we
1492 < will discuss the physical meaning of the terms appearing in the
1493 < equation as well as the calculation of friction tensor from
1494 < hydrodynamics theory.
1490 > the theory of Langevin dynamics. A brief derivation of generalized
1491 > Langevin equation will be given first. Following that, we will
1492 > discuss the physical meaning of the terms appearing in the equation
1493 > as well as the calculation of friction tensor from hydrodynamics
1494 > theory.
1495  
1496   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1497  
1498 < Harmonic bath model, in which an effective set of harmonic
1498 > A harmonic bath model, in which an effective set of harmonic
1499   oscillators are used to mimic the effect of a linearly responding
1500   environment, has been widely used in quantum chemistry and
1501   statistical mechanics. One of the successful applications of
1502 < Harmonic bath model is the derivation of Deriving Generalized
1503 < Langevin Dynamics. Lets consider a system, in which the degree of
1502 > Harmonic bath model is the derivation of the Generalized Langevin
1503 > Dynamics (GLE). Lets consider a system, in which the degree of
1504   freedom $x$ is assumed to couple to the bath linearly, giving a
1505   Hamiltonian of the form
1506   \begin{equation}
1507   H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1508   \label{introEquation:bathGLE}.
1509   \end{equation}
1510 < Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1511 < with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1510 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1511 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1512   \[
1513   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1514   }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
# Line 1570 | Line 1516 | the harmonic bath masses, and $\Delta U$ is bilinear s
1516   \]
1517   where the index $\alpha$ runs over all the bath degrees of freedom,
1518   $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1519 < the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1519 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1520   coupling,
1521   \[
1522   \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1523   \]
1524 < where $g_\alpha$ are the coupling constants between the bath and the
1525 < coordinate $x$. Introducing
1524 > where $g_\alpha$ are the coupling constants between the bath
1525 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1526 > Introducing
1527   \[
1528   W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1529   }}{{2m_\alpha  w_\alpha ^2 }}} x^2
# Line 1591 | Line 1538 | Generalized Langevin Dynamics by Hamilton's equations
1538   \]
1539   Since the first two terms of the new Hamiltonian depend only on the
1540   system coordinates, we can get the equations of motion for
1541 < Generalized Langevin Dynamics by Hamilton's equations
1595 < \ref{introEquation:motionHamiltonianCoordinate,
1596 < introEquation:motionHamiltonianMomentum},
1541 > Generalized Langevin Dynamics by Hamilton's equations,
1542   \begin{equation}
1543   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1544   \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
# Line 1610 | Line 1555 | differential equations, Laplace transform is the appro
1555   In order to derive an equation for $x$, the dynamics of the bath
1556   variables $x_\alpha$ must be solved exactly first. As an integral
1557   transform which is particularly useful in solving linear ordinary
1558 < differential equations, Laplace transform is the appropriate tool to
1559 < solve this problem. The basic idea is to transform the difficult
1558 > differential equations,the Laplace transform is the appropriate tool
1559 > to solve this problem. The basic idea is to transform the difficult
1560   differential equations into simple algebra problems which can be
1561 < solved easily. Then applying inverse Laplace transform, also known
1562 < as the Bromwich integral, we can retrieve the solutions of the
1561 > solved easily. Then, by applying the inverse Laplace transform, also
1562 > known as the Bromwich integral, we can retrieve the solutions of the
1563   original problems.
1564  
1565   Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
# Line 1634 | Line 1579 | Applying Laplace transform to the bath coordinates, we
1579   \end{eqnarray*}
1580  
1581  
1582 < Applying Laplace transform to the bath coordinates, we obtain
1582 > Applying the Laplace transform to the bath coordinates, we obtain
1583   \begin{eqnarray*}
1584   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) \\
1585   L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
# Line 1657 | Line 1602 | m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}}
1602   \]
1603   , we obtain
1604   \begin{eqnarray*}
1605 < m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} -
1605 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1606   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1607   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1608 < _\alpha  t)\dot x(t - \tau )d\tau  \\
1609 < & & - \left[ {g_\alpha  x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha
1610 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) -
1611 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1612 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}} \\
1613 < %
1614 < & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1608 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1609 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1610 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1611 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1612 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1613 > \end{eqnarray*}
1614 > \begin{eqnarray*}
1615 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1616   {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1617   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1618 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1619 < {\left[ {g_\alpha  x_\alpha  (0) \\
1620 < & & - \frac{{g_\alpha  }}{{m_\alpha \omega _\alpha  }}} \right]\cos
1621 < (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha (0)}}{{\omega
1622 < _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1618 > t)\dot x(t - \tau )d} \tau }  \\
1619 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1620 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1621 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1622 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1623   \end{eqnarray*}
1624   Introducing a \emph{dynamic friction kernel}
1625   \begin{equation}
# Line 1697 | Line 1643 | which is known as the \emph{generalized Langevin equat
1643   \end{equation}
1644   which is known as the \emph{generalized Langevin equation}.
1645  
1646 < \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1646 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1647  
1648   One may notice that $R(t)$ depends only on initial conditions, which
1649   implies it is completely deterministic within the context of a
# Line 1710 | Line 1656 | as the model, which is gaussian distribution in genera
1656   \end{array}
1657   \]
1658   This property is what we expect from a truly random process. As long
1659 < as the model, which is gaussian distribution in general, chosen for
1660 < $R(t)$ is a truly random process, the stochastic nature of the GLE
1715 < still remains.
1659 > as the model chosen for $R(t)$ was a gaussian distribution in
1660 > general, the stochastic nature of the GLE still remains.
1661  
1662   %dynamic friction kernel
1663   The convolution integral
# Line 1733 | Line 1678 | which can be used to describe dynamic caging effect. T
1678   m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1679   \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1680   \]
1681 < which can be used to describe dynamic caging effect. The other
1682 < extreme is the bath that responds infinitely quickly to motions in
1683 < the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1684 < time:
1681 > which can be used to describe the effect of dynamic caging in
1682 > viscous solvents. The other extreme is the bath that responds
1683 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1684 > taken as a $delta$ function in time:
1685   \[
1686   \xi (t) = 2\xi _0 \delta (t)
1687   \]
# Line 1752 | Line 1697 | or be determined by Stokes' law for regular shaped par
1697   \end{equation}
1698   which is known as the Langevin equation. The static friction
1699   coefficient $\xi _0$ can either be calculated from spectral density
1700 < or be determined by Stokes' law for regular shaped particles.A
1700 > or be determined by Stokes' law for regular shaped particles. A
1701   briefly review on calculating friction tensor for arbitrary shaped
1702   particles is given in Sec.~\ref{introSection:frictionTensor}.
1703  
1704 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1704 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1705  
1706   Defining a new set of coordinates,
1707   \[
# Line 1785 | Line 1730 | can model the random force and friction kernel.
1730   \end{equation}
1731   In effect, it acts as a constraint on the possible ways in which one
1732   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 $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines