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 2908 by tim, Thu Jun 29 18:21:09 2006 UTC vs.
Revision 2938 by tim, Mon Jul 17 15:28:44 2006 UTC

# Line 208 | Line 208 | The following section will give a brief introduction t
208   The thermodynamic behaviors and properties of Molecular Dynamics
209   simulation are governed by the principle of Statistical Mechanics.
210   The following section will give a brief introduction to some of the
211 < Statistical Mechanics concepts and theorem presented in this
211 > Statistical Mechanics concepts and theorems presented in this
212   dissertation.
213  
214   \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
# Line 281 | Line 281 | space of the system,
281   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
282   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
283   \label{introEquation:ensembelAverage}
284 \end{equation}
285
286 There are several different types of ensembles with different
287 statistical characteristics. As a function of macroscopic
288 parameters, such as temperature \textit{etc}, the partition function
289 can be used to describe the statistical properties of a system in
290 thermodynamic equilibrium. As an ensemble of systems, each of which
291 is known to be thermally isolated and conserve energy, the
292 Microcanonical ensemble (NVE) has a partition function like,
293 \begin{equation}
294 \Omega (N,V,E) = e^{\beta TS}. \label{introEquation:NVEPartition}
295 \end{equation}
296 A canonical ensemble (NVT) is an ensemble of systems, each of which
297 can share its energy with a large heat reservoir. The distribution
298 of the total energy amongst the possible dynamical states is given
299 by the partition function,
300 \begin{equation}
301 \Omega (N,V,T) = e^{ - \beta A}.
302 \label{introEquation:NVTPartition}
303 \end{equation}
304 Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
305 TS$. Since most experiments are carried out under constant pressure
306 condition, the isothermal-isobaric ensemble (NPT) plays a very
307 important role in molecular simulations. The isothermal-isobaric
308 ensemble allow the system to exchange energy with a heat bath of
309 temperature $T$ and to change the volume as well. Its partition
310 function is given as
311 \begin{equation}
312 \Delta (N,P,T) =  - e^{\beta G}.
313 \label{introEquation:NPTPartition}
284   \end{equation}
315 Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
285  
286   \subsection{\label{introSection:liouville}Liouville's theorem}
287  
# Line 403 | Line 372 | $F$ and $G$ of the coordinates and momenta of a system
372   Liouville's theorem can be expressed in a variety of different forms
373   which are convenient within different contexts. For any two function
374   $F$ and $G$ of the coordinates and momenta of a system, the Poisson
375 < bracket ${F, G}$ is defined as
375 > bracket $\{F,G\}$ is defined as
376   \begin{equation}
377   \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
378   F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
# Line 470 | Line 439 | the motions of atoms in MD simulation. They usually be
439   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
440   A variety of numerical integrators have been proposed to simulate
441   the motions of atoms in MD simulation. They usually begin with
442 < initial conditionals and move the objects in the direction governed
442 > initial conditions and move the objects in the direction governed
443   by the differential equations. However, most of them ignore the
444   hidden physical laws contained within the equations. Since 1990,
445   geometric integrators, which preserve various phase-flow invariants
# Line 490 | Line 459 | defined as a pair $(M, \omega)$ which consists of a
459   \emph{smooth manifold}) is a manifold on which it is possible to
460   apply calculus\cite{Hirsch1997}. A \emph{symplectic manifold} is
461   defined as a pair $(M, \omega)$ which consists of a
462 < \emph{differentiable manifold} $M$ and a close, non-degenerated,
462 > \emph{differentiable manifold} $M$ and a close, non-degenerate,
463   bilinear symplectic form, $\omega$. A symplectic form on a vector
464   space $V$ is a function $\omega(x, y)$ which satisfies
465   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
# Line 510 | Line 479 | For an ordinary differential system defined as
479   \begin{equation}
480   \dot x = f(x)
481   \end{equation}
482 < where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
482 > where $x = x(q,p)$, this system is a canonical Hamiltonian, if
483   $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
484   function and $J$ is the skew-symmetric matrix
485   \begin{equation}
# Line 536 | Line 505 | Let $x(t)$ be the exact solution of the ODE
505   \subsection{\label{introSection:exactFlow}Exact Propagator}
506  
507   Let $x(t)$ be the exact solution of the ODE
508 < system,$\frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}$, we can
509 < define its exact propagator(solution) $\varphi_\tau$
508 > system,
509 > \begin{equation}
510 > \frac{{dx}}{{dt}} = f(x), \label{introEquation:ODE}
511 > \end{equation} we can
512 > define its exact propagator $\varphi_\tau$:
513   \[ x(t+\tau)
514   =\varphi_\tau(x(t))
515   \]
# Line 555 | Line 527 | Therefore, the exact propagator is self-adjoint,
527   \begin{equation}
528   \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
529   \end{equation}
530 < The exact propagator can also be written in terms of operator,
530 > The exact propagator can also be written as an operator,
531   \begin{equation}
532   \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
533   }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
# Line 609 | Line 581 | Using the chain rule, one may obtain,
581   \]
582   Using the chain rule, one may obtain,
583   \[
584 < \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \dot \nabla G,
584 > \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \cdot \nabla G,
585   \]
586   which is the condition for conserved quantities. For a canonical
587   Hamiltonian system, the time evolution of an arbitrary smooth
# Line 734 | Line 706 | known as \emph{velocity verlet} which is
706   \end{align}
707   where $F(t)$ is the force at time $t$. This integration scheme is
708   known as \emph{velocity verlet} which is
709 < symplectic(\ref{introEquation:SymplecticFlowComposition}),
710 < time-reversible(\ref{introEquation:timeReversible}) and
711 < volume-preserving (\ref{introEquation:volumePreserving}). These
709 > symplectic(Eq.~\ref{introEquation:SymplecticFlowComposition}),
710 > time-reversible(Eq.~\ref{introEquation:timeReversible}) and
711 > volume-preserving (Eq.~\ref{introEquation:volumePreserving}). These
712   geometric properties attribute to its long-time stability and its
713   popularity in the community. However, the most commonly used
714   velocity verlet integration scheme is written as below,
# Line 757 | Line 729 | the equations of motion would follow:
729  
730   \item Use the half step velocities to move positions one whole step, $\Delta t$.
731  
732 < \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
732 > \item Evaluate the forces at the new positions, $q(\Delta t)$, and use the new forces to complete the velocity move.
733  
734   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
735   \end{enumerate}
# Line 778 | Line 750 | local error of a splitting method in terms of the comm
750  
751   The Baker-Campbell-Hausdorff formula can be used to determine the
752   local error of a splitting method in terms of the commutator of the
753 < operators(\ref{introEquation:exponentialOperator}) associated with
753 > operators(Eq.~\ref{introEquation:exponentialOperator}) associated with
754   the sub-propagator. For operators $hX$ and $hY$ which are associated
755   with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
756   \begin{equation}
# Line 862 | Line 834 | initialization of a simulation. Sec.~\ref{introSection
834   These three individual steps will be covered in the following
835   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
836   initialization of a simulation. Sec.~\ref{introSection:production}
837 < will discuss issues of production runs.
837 > discusses issues of production runs.
838   Sec.~\ref{introSection:Analysis} provides the theoretical tools for
839   analysis of trajectories.
840  
# Line 896 | Line 868 | surface and to locate the local minimum. While converg
868   minimization to find a more reasonable conformation. Several energy
869   minimization methods have been developed to exploit the energy
870   surface and to locate the local minimum. While converging slowly
871 < near the minimum, steepest descent method is extremely robust when
871 > near the minimum, the steepest descent method is extremely robust when
872   systems are strongly anharmonic. Thus, it is often used to refine
873   structures from crystallographic data. Relying on the Hessian,
874   advanced methods like Newton-Raphson converge rapidly to a local
# Line 915 | Line 887 | end up setting the temperature of the system to a fina
887   temperature. Beginning at a lower temperature and gradually
888   increasing the temperature by assigning larger random velocities, we
889   end up setting the temperature of the system to a final temperature
890 < at which the simulation will be conducted. In heating phase, we
890 > at which the simulation will be conducted. In the heating phase, we
891   should also keep the system from drifting or rotating as a whole. To
892   do this, the net linear momentum and angular momentum of the system
893   is shifted to zero after each resampling from the Maxwell -Boltzman
# Line 930 | Line 902 | equilibration process is long enough. However, these s
902   properties \textit{etc}, become independent of time. Strictly
903   speaking, minimization and heating are not necessary, provided the
904   equilibration process is long enough. However, these steps can serve
905 < as a means to arrive at an equilibrated structure in an effective
905 > as a mean to arrive at an equilibrated structure in an effective
906   way.
907  
908   \subsection{\label{introSection:production}Production}
# Line 982 | Line 954 | with rapid and absolute convergence, has proved to min
954   in simulations. The Ewald summation, in which the slowly decaying
955   Coulomb potential is transformed into direct and reciprocal sums
956   with rapid and absolute convergence, has proved to minimize the
957 < periodicity artifacts in liquid simulations. Taking the advantages
958 < of the fast Fourier transform (FFT) for calculating discrete Fourier
957 > periodicity artifacts in liquid simulations. Taking advantage
958 > of fast Fourier transform (FFT) techniques for calculating discrete Fourier
959   transforms, the particle mesh-based
960   methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
961   $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
# Line 1000 | Line 972 | R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
972   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
973   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
974   R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
975 < r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
975 > r_{ij})}{r_{ij}}\right\}, \label{introEquation:shiftedCoulomb}
976   \end{equation}
977   where $\alpha$ is the convergence parameter. Due to the lack of
978   inherent periodicity and rapid convergence,this method is extremely
# Line 1017 | Line 989 | illustration of shifted Coulomb potential.}
989  
990   \subsection{\label{introSection:Analysis} Analysis}
991  
992 < Recently, advanced visualization technique have become applied to
992 > Recently, advanced visualization techniques have been applied to
993   monitor the motions of molecules. Although the dynamics of the
994   system can be described qualitatively from animation, quantitative
995   trajectory analysis is more useful. According to the principles of
# Line 1087 | Line 1059 | If $A$ and $B$ refer to same variable, this kind of co
1059   \label{introEquation:timeCorrelationFunction}
1060   \end{equation}
1061   If $A$ and $B$ refer to same variable, this kind of correlation
1062 < function is called an \emph{autocorrelation function}. One example
1091 < of an auto correlation function is the velocity auto-correlation
1062 > functions are called \emph{autocorrelation functions}. One typical example is the velocity autocorrelation
1063   function which is directly related to transport properties of
1064   molecular liquids:
1065 < \[
1065 > \begin{equation}
1066   D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1067   \right\rangle } dt
1068 < \]
1068 > \end{equation}
1069   where $D$ is diffusion constant. Unlike the velocity autocorrelation
1070   function, which is averaged over time origins and over all the
1071   atoms, the dipole autocorrelation functions is calculated for the
1072   entire system. The dipole autocorrelation function is given by:
1073 < \[
1073 > \begin{equation}
1074   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1075   \right\rangle
1076 < \]
1076 > \end{equation}
1077   Here $u_{tot}$ is the net dipole of the entire system and is given
1078   by
1079 < \[
1079 > \begin{equation}
1080   u_{tot} (t) = \sum\limits_i {u_i (t)}.
1081 < \]
1081 > \end{equation}
1082   In principle, many time correlation functions can be related to
1083   Fourier transforms of the infrared, Raman, and inelastic neutron
1084   scattering spectra of molecular liquids. In practice, one can
1085   extract the IR spectrum from the intensity of the molecular dipole
1086   fluctuation at each frequency using the following relationship:
1087 < \[
1087 > \begin{equation}
1088   \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1089   i2\pi vt} dt}.
1090 < \]
1090 > \end{equation}
1091  
1092   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1093  
1094   Rigid bodies are frequently involved in the modeling of different
1095 < areas, from engineering, physics, to chemistry. For example,
1095 > areas, including engineering, physics and chemistry. For example,
1096   missiles and vehicles are usually modeled by rigid bodies.  The
1097   movement of the objects in 3D gaming engines or other physics
1098   simulators is governed by rigid body dynamics. In molecular
# Line 1139 | Line 1110 | quaternions was developed by Evans in 1977\cite{Evans1
1110   computational penalty and the loss of angular momentum conservation
1111   still remain. A singularity-free representation utilizing
1112   quaternions was developed by Evans in 1977\cite{Evans1977}.
1113 < Unfortunately, this approach uses a nonseparable Hamiltonian
1114 < resulting from the quaternion representation, which prevents the
1113 > Unfortunately, this approach used a nonseparable Hamiltonian
1114 > resulting from the quaternion representation, which prevented the
1115   symplectic algorithm from being utilized. Another different approach
1116   is to apply holonomic constraints to the atoms belonging to the
1117   rigid body. Each atom moves independently under the normal forces
# Line 1163 | Line 1134 | Dullweber and his coworkers\cite{Dullweber1997} in dep
1134   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1135  
1136   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1137 < The motion of a rigid body is Hamiltonian with the Hamiltonian
1167 < function
1137 > The Hamiltonian of a rigid body is given by
1138   \begin{equation}
1139   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1140   V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1141   \label{introEquation:RBHamiltonian}
1142   \end{equation}
1143 < Here, $q$ and $Q$  are the position and rotation matrix for the
1144 < rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1145 < $J$, a diagonal matrix, is defined by
1143 > Here, $q$ and $Q$  are the position vector and rotation matrix for
1144 > the rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ ,
1145 > and $J$, a diagonal matrix, is defined by
1146   \[
1147   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1148   \]
# Line 1182 | Line 1152 | Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1152   \begin{equation}
1153   Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1154   \end{equation}
1155 < which is used to ensure rotation matrix's unitarity. Differentiating
1156 < Eq.~\ref{introEquation:orthogonalConstraint} and using
1157 < Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1188 < \begin{equation}
1189 < Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1190 < \label{introEquation:RBFirstOrderConstraint}
1191 < \end{equation}
1192 < Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1193 < \ref{introEquation:motionHamiltonianMomentum}), one can write down
1155 > which is used to ensure the rotation matrix's unitarity. Using
1156 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} and Eq.~
1157 > \ref{introEquation:motionHamiltonianMomentum}, one can write down
1158   the equations of motion,
1159   \begin{eqnarray}
1160   \frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\
# Line 1198 | Line 1162 | the equations of motion,
1162   \frac{{dQ}}{{dt}} & = & PJ^{ - 1},  \label{introEquation:RBMotionRotation}\\
1163   \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1164   \end{eqnarray}
1165 + Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and
1166 + using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1167 + \begin{equation}
1168 + Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1169 + \label{introEquation:RBFirstOrderConstraint}
1170 + \end{equation}
1171   In general, there are two ways to satisfy the holonomic constraints.
1172   We can use a constraint force provided by a Lagrange multiplier on
1173 < the normal manifold to keep the motion on constraint space. Or we
1174 < can simply evolve the system on the constraint manifold. These two
1175 < methods have been proved to be equivalent. The holonomic constraint
1176 < and equations of motions define a constraint manifold for rigid
1177 < bodies
1173 > the normal manifold to keep the motion on the constraint space. Or
1174 > we can simply evolve the system on the constraint manifold. These
1175 > two methods have been proved to be equivalent. The holonomic
1176 > constraint and equations of motions define a constraint manifold for
1177 > rigid bodies
1178   \[
1179   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1180   \right\}.
1181   \]
1182 < Unfortunately, this constraint manifold is not the cotangent bundle
1183 < $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1184 < rotation group $SO(3)$. However, it turns out that under symplectic
1185 < transformation, the cotangent space and the phase space are
1216 < diffeomorphic. By introducing
1182 > Unfortunately, this constraint manifold is not $T^* SO(3)$ which is
1183 > a symplectic manifold on Lie rotation group $SO(3)$. However, it
1184 > turns out that under symplectic transformation, the cotangent space
1185 > and the phase space are diffeomorphic. By introducing
1186   \[
1187   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1188   \]
1189 < the mechanical system subject to a holonomic constraint manifold $M$
1189 > the mechanical system subjected to a holonomic constraint manifold $M$
1190   can be re-formulated as a Hamiltonian system on the cotangent space
1191   \[
1192   T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
# Line 1281 | Line 1250 | Omelyan1998}. Applying the hat-map isomorphism, we obt
1250   motion. This unique property eliminates the requirement of
1251   iterations which can not be avoided in other methods\cite{Kol1997,
1252   Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1253 < equation of motion for angular momentum on body frame
1253 > equation of motion for angular momentum in the body frame
1254   \begin{equation}
1255   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1256   F_i (r,Q)} \right) \times X_i }.
# Line 1294 | Line 1263 | given by
1263   \]
1264  
1265   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1266 < Lie-Poisson Integrator for Free Rigid Body}
1266 > Lie-Poisson Integrator for Free Rigid Bodies}
1267  
1268   If there are no external forces exerted on the rigid body, the only
1269   contribution to the rotational motion is from the kinetic energy
# Line 1346 | Line 1315 | To reduce the cost of computing expensive functions in
1315   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1316   \]
1317   To reduce the cost of computing expensive functions in $e^{\Delta
1318 < tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1319 < propagator,
1320 < \[
1321 < e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1322 < ).
1323 < \]
1324 < The propagator maps for $T_2^r$ and $T_3^r$ can be found in the same
1318 > tR_1 }$, we can use the Cayley transformation to obtain a
1319 > single-aixs propagator,
1320 > \begin{eqnarray*}
1321 > e^{\Delta tR_1 }  & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta
1322 > tR_1 ) \\
1323 > %
1324 > & \approx & \left( \begin{array}{ccc}
1325 > 1 & 0 & 0 \\
1326 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1327 > \theta^2 / 4} \\
1328 > 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1329 > \theta^2 / 4}
1330 > \end{array}
1331 > \right).
1332 > \end{eqnarray*}
1333 > The propagators for $T_2^r$ and $T_3^r$ can be found in the same
1334   manner. In order to construct a second-order symplectic method, we
1335   split the angular kinetic Hamiltonian function into five terms
1336   \[
# Line 1368 | Line 1346 | _1 }.
1346   \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1347   _1 }.
1348   \]
1349 < The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1372 < $F(\pi )$ and $G(\pi )$ is defined by
1349 > The non-canonical Lie-Poisson bracket $\{F, G\}$ of two functions $F(\pi )$ and $G(\pi )$ is defined by
1350   \[
1351   \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1352   ).
# Line 1378 | Line 1355 | norm of the angular momentum, $\parallel \pi
1355   function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1356   conserved quantity in Poisson system. We can easily verify that the
1357   norm of the angular momentum, $\parallel \pi
1358 < \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1358 > \parallel$, is a \emph{Casimir}\cite{McLachlan1993}. Let $F(\pi ) = S(\frac{{\parallel
1359   \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1360   then by the chain rule
1361   \[
# Line 1397 | Line 1374 | The Hamiltonian of rigid body can be separated in term
1374   Splitting for Rigid Body}
1375  
1376   The Hamiltonian of rigid body can be separated in terms of kinetic
1377 < energy and potential energy,$H = T(p,\pi ) + V(q,Q)$. The equations
1377 > energy and potential energy, $H = T(p,\pi ) + V(q,Q)$. The equations
1378   of motion corresponding to potential energy and kinetic energy are
1379 < listed in the below table,
1379 > listed in Table~\ref{introTable:rbEquations}.
1380   \begin{table}
1381   \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1382 + \label{introTable:rbEquations}
1383   \begin{center}
1384   \begin{tabular}{|l|l|}
1385    \hline
# Line 1456 | Line 1434 | has been applied in a variety of studies. This section
1434   As an alternative to newtonian dynamics, Langevin dynamics, which
1435   mimics a simple heat bath with stochastic and dissipative forces,
1436   has been applied in a variety of studies. This section will review
1437 < the theory of Langevin dynamics. A brief derivation of generalized
1437 > the theory of Langevin dynamics. A brief derivation of the generalized
1438   Langevin equation will be given first. Following that, we will
1439 < discuss the physical meaning of the terms appearing in the equation
1462 < as well as the calculation of friction tensor from hydrodynamics
1463 < theory.
1439 > discuss the physical meaning of the terms appearing in the equation.
1440  
1441   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1442  
# Line 1469 | Line 1445 | Harmonic bath model is the derivation of the Generaliz
1445   environment, has been widely used in quantum chemistry and
1446   statistical mechanics. One of the successful applications of
1447   Harmonic bath model is the derivation of the Generalized Langevin
1448 < Dynamics (GLE). Lets consider a system, in which the degree of
1448 > Dynamics (GLE). Consider a system, in which the degree of
1449   freedom $x$ is assumed to couple to the bath linearly, giving a
1450   Hamiltonian of the form
1451   \begin{equation}
# Line 1480 | Line 1456 | H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_
1456   with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1457   \[
1458   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1459 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1459 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  x_\alpha ^2 }
1460   \right\}}
1461   \]
1462   where the index $\alpha$ runs over all the bath degrees of freedom,
# Line 1525 | Line 1501 | differential equations into simple algebra problems wh
1501   differential equations,the Laplace transform is the appropriate tool
1502   to solve this problem. The basic idea is to transform the difficult
1503   differential equations into simple algebra problems which can be
1504 < solved easily. Then, by applying the inverse Laplace transform, also
1505 < known as the Bromwich integral, we can retrieve the solutions of the
1506 < original problems. Let $f(t)$ be a function defined on $ [0,\infty )
1507 < $, the Laplace transform of $f(t)$ is a new function defined as
1504 > solved easily. Then, by applying the inverse Laplace transform, we
1505 > can retrieve the solutions of the original problems. Let $f(t)$ be a
1506 > function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$
1507 > is a new function defined as
1508   \[
1509   L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1510   \]
1511   where  $p$ is real and  $L$ is called the Laplace Transform
1512 < Operator. Below are some important properties of Laplace transform
1512 > Operator. Below are some important properties of the Laplace transform
1513   \begin{eqnarray*}
1514   L(x + y)  & = & L(x) + L(y) \\
1515   L(ax)     & = & aL(x) \\
# Line 1546 | Line 1522 | L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega
1522   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), \\
1523   L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}. \\
1524   \end{eqnarray*}
1525 < By the same way, the system coordinates become
1525 > In the same way, the system coordinates become
1526   \begin{eqnarray*}
1527   mL(\ddot x) & = &
1528    - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
# Line 1570 | Line 1546 | x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _
1546   & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1547   x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1548   \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1549 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1550 < \end{eqnarray*}
1551 < \begin{eqnarray*}
1552 < m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1553 < {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1554 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1549 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}\\
1550 > %
1551 > & = & -
1552 > \frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha
1553 > = 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha
1554 > ^2 }}} \right)\cos (\omega _\alpha
1555   t)\dot x(t - \tau )d} \tau }  \\
1556   & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1557   x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
# Line 1602 | Line 1578 | m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int
1578   (t)\dot x(t - \tau )d\tau }  + R(t)
1579   \label{introEuqation:GeneralizedLangevinDynamics}
1580   \end{equation}
1581 < which is known as the \emph{generalized Langevin equation}.
1581 > which is known as the \emph{generalized Langevin equation} (GLE).
1582  
1583   \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1584  
1585   One may notice that $R(t)$ depends only on initial conditions, which
1586   implies it is completely deterministic within the context of a
1587   harmonic bath. However, it is easy to verify that $R(t)$ is totally
1588 < uncorrelated to $x$ and $\dot x$,$\left\langle {x(t)R(t)}
1588 > uncorrelated to $x$ and $\dot x$, $\left\langle {x(t)R(t)}
1589   \right\rangle  = 0, \left\langle {\dot x(t)R(t)} \right\rangle  =
1590   0.$ This property is what we expect from a truly random process. As
1591   long as the model chosen for $R(t)$ was a gaussian distribution in
# Line 1638 | Line 1614 | taken as a $delta$ function in time:
1614   infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1615   taken as a $delta$ function in time:
1616   \[
1617 < \xi (t) = 2\xi _0 \delta (t)
1617 > \xi (t) = 2\xi _0 \delta (t).
1618   \]
1619   Hence, the convolution integral becomes
1620   \[
# Line 1653 | Line 1629 | or be determined by Stokes' law for regular shaped par
1629   which is known as the Langevin equation. The static friction
1630   coefficient $\xi _0$ can either be calculated from spectral density
1631   or be determined by Stokes' law for regular shaped particles. A
1632 < briefly review on calculating friction tensor for arbitrary shaped
1632 > brief review on calculating friction tensors for arbitrary shaped
1633   particles is given in Sec.~\ref{introSection:frictionTensor}.
1634  
1635   \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
# Line 1663 | Line 1639 | q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \o
1639   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1640   ^2 }}x(0),
1641   \]
1642 < we can rewrite $R(T)$ as
1642 > we can rewrite $R(t)$ as
1643   \[
1644   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1645   \]

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines