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 2912 by tim, Fri Jun 30 02:45:29 2006 UTC vs.
Revision 2944 by tim, Tue Jul 18 05:47:33 2006 UTC

# Line 67 | Line 67 | All of these conserved quantities are important factor
67   \begin{equation}E = T + V. \label{introEquation:energyConservation}
68   \end{equation}
69   All of these conserved quantities are important factors to determine
70 < the quality of numerical integration schemes for rigid bodies
71 < \cite{Dullweber1997}.
70 > the quality of numerical integration schemes for rigid
71 > bodies.\cite{Dullweber1997}
72  
73   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
# Line 178 | Line 178 | equation of motion. Due to their symmetrical formula,
178   where Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
179   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
180   equation of motion. Due to their symmetrical formula, they are also
181 < known as the canonical equations of motions \cite{Goldstein2001}.
181 > known as the canonical equations of motions.\cite{Goldstein2001}
182  
183   An important difference between Lagrangian approach and the
184   Hamiltonian approach is that the Lagrangian is considered to be a
# Line 188 | Line 188 | coordinate and its time derivative as independent vari
188   Hamiltonian Mechanics is more appropriate for application to
189   statistical mechanics and quantum mechanics, since it treats the
190   coordinate and its time derivative as independent variables and it
191 < only works with 1st-order differential equations\cite{Marion1990}.
191 > only works with 1st-order differential equations.\cite{Marion1990}
192   In Newtonian Mechanics, a system described by conservative forces
193   conserves the total energy
194   (Eq.~\ref{introEquation:energyConservation}). It follows that
# 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 372 | 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 416 | Line 416 | average. It states that the time average and average o
416   many-body system in Statistical Mechanics. Fortunately, the Ergodic
417   Hypothesis makes a connection between time average and the ensemble
418   average. It states that the time average and average over the
419 < statistical ensemble are identical \cite{Frenkel1996, Leach2001}:
419 > statistical ensemble are identical:\cite{Frenkel1996, Leach2001}
420   \begin{equation}
421   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
422   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 434 | Line 434 | Sec.~\ref{introSection:molecularDynamics} will be the
434   utilized. Or if the system lends itself to a time averaging
435   approach, the Molecular Dynamics techniques in
436   Sec.~\ref{introSection:molecularDynamics} will be the best
437 < choice\cite{Frenkel1996}.
437 > choice.\cite{Frenkel1996}
438  
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
443 < by the differential equations. However, most of them ignore the
444 < hidden physical laws contained within the equations. Since 1990,
445 < geometric integrators, which preserve various phase-flow invariants
446 < such as symplectic structure, volume and time reversal symmetry,
447 < were developed to address this issue\cite{Dullweber1997,
448 < McLachlan1998, Leimkuhler1999}. The velocity Verlet method, which
449 < happens to be a simple example of symplectic integrator, continues
450 < to gain popularity in the molecular dynamics community. This fact
451 < can be partly explained by its geometric nature.
442 > initial conditions and move the objects in the direction governed by
443 > the differential equations. However, most of them ignore the hidden
444 > physical laws contained within the equations. Since 1990, geometric
445 > integrators, which preserve various phase-flow invariants such as
446 > symplectic structure, volume and time reversal symmetry, were
447 > developed to address this issue.\cite{Dullweber1997, McLachlan1998,
448 > Leimkuhler1999} The velocity Verlet method, which happens to be a
449 > simple example of symplectic integrator, continues to gain
450 > popularity in the molecular dynamics community. This fact can be
451 > partly explained by its geometric nature.
452  
453 < \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
453 > \subsection{\label{introSection:symplecticManifold}Manifolds and Bundles}
454   A \emph{manifold} is an abstract mathematical space. It looks
455   locally like Euclidean space, but when viewed globally, it may have
456   more complicated structure. A good example of manifold is the
457   surface of Earth. It seems to be flat locally, but it is round if
458   viewed as a whole. A \emph{differentiable manifold} (also known as
459   \emph{smooth manifold}) is a manifold on which it is possible to
460 < apply calculus\cite{Hirsch1997}. A \emph{symplectic manifold} is
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   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
466   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
467 < $\omega(x, x) = 0$\cite{McDuff1998}. The cross product operation in
468 < vector field is an example of symplectic form. One of the
469 < motivations to study \emph{symplectic manifolds} in Hamiltonian
470 < Mechanics is that a symplectic manifold can represent all possible
471 < configurations of the system and the phase space of the system can
472 < be described by it's cotangent bundle\cite{Jost2002}. Every
473 < symplectic manifold is even dimensional. For instance, in Hamilton
474 < equations, coordinate and momentum always appear in pairs.
467 > $\omega(x, x) = 0$.\cite{McDuff1998} The cross product operation in
468 > vector field is an example of symplectic form.
469 > Given vector spaces $V$ and $W$ over same field $F$, $f: V \to W$ is a linear transformation if
470 > \begin{eqnarray*}
471 > f(x+y) & = & f(x) + f(y) \\
472 > f(ax) & = & af(x)      
473 > \end{eqnarray*}
474 > are always satisfied for any two vectors $x$ and $y$ in $V$ and any scalar $a$ in $F$. One can define the dual vector space $V^*$ of $V$ if any two built-in linear transformations $\phi$ and $\psi$ in $V^*$ satisfy the following definition of addition and scalar multiplication:
475 > \begin{eqnarray*}
476 > (\phi+\psi)(x) & = & \phi(x)+\psi(x) \\
477 > (a\phi)(x) & = & a \phi(x)
478 > \end{eqnarray*}
479 > for all $a$ in $F$ and $x$ in $V$. For a manifold $M$, one can define a tangent vector of a tangent space $TM_q$ at every point $q$
480 > \begin{equation}
481 > \dot q = \mathop {\lim }\limits_{t \to 0} \frac{{\phi (t) - \phi (0)}}{t}      
482 > \end{equation}
483 > where $\phi(0)=q$ and $\phi(t) \in M$. One may also define a cotangent space $T^*M_q$ as the dual space of the tangent space $TM_q$. The tangent space and the cotangent space are isomorphic to each other, since they are both real vector spaces with same dimension.
484 > The union of tangent spaces at every point of $M$ is called the tangent bundle of $M$ and is denoted by $TM$, while cotangent bundle $T^*M$ is defined as the union of the cotangent spaces to $M$.\cite{Jost2002} For a Hamiltonian system with configuration manifold $V$, the $(q,\dot q)$ phase space is the tangent bundle of the configuration manifold $V$, while the cotangent bundle is represented by $(q,p)$.
485  
486   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
487  
# Line 479 | Line 489 | For an ordinary differential system defined as
489   \begin{equation}
490   \dot x = f(x)
491   \end{equation}
492 < where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
492 > where $x = x(q,p)$, this system is a canonical Hamiltonian, if
493   $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
494   function and $J$ is the skew-symmetric matrix
495   \begin{equation}
# Line 496 | Line 506 | called a \emph{Hamiltonian vector field}. Another gene
506   \label{introEquation:compactHamiltonian}
507   \end{equation}In this case, $f$ is
508   called a \emph{Hamiltonian vector field}. Another generalization of
509 < Hamiltonian dynamics is Poisson Dynamics\cite{Olver1986},
509 > Hamiltonian dynamics is Poisson Dynamics,\cite{Olver1986}
510   \begin{equation}
511   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
512   \end{equation}
513 < The most obvious change being that matrix $J$ now depends on $x$.
513 > where the most obvious change being that matrix $J$ now depends on
514 > $x$.
515  
516   \subsection{\label{introSection:exactFlow}Exact Propagator}
517  
# Line 527 | Line 538 | Therefore, the exact propagator is self-adjoint,
538   \begin{equation}
539   \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
540   \end{equation}
530 The exact propagator can also be written in terms of 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).
534 \label{introEquation:exponentialOperator}
535 \end{equation}
541   In most cases, it is not easy to find the exact propagator
542   $\varphi_\tau$. Instead, we use an approximate map, $\psi_\tau$,
543   which is usually called an integrator. The order of an integrator
# Line 620 | Line 625 | variational methods can capture the decay of energy
625   Generating functions\cite{Channell1990} tend to lead to methods
626   which are cumbersome and difficult to use. In dissipative systems,
627   variational methods can capture the decay of energy
628 < accurately\cite{Kane2000}. Since they are geometrically unstable
628 > accurately.\cite{Kane2000} Since they are geometrically unstable
629   against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
630 < methods are not suitable for Hamiltonian system. Recently, various
631 < high-order explicit Runge-Kutta methods \cite{Owren1992,Chen2003}
632 < have been developed to overcome this instability. However, due to
633 < computational penalty involved in implementing the Runge-Kutta
634 < methods, they have not attracted much attention from the Molecular
635 < Dynamics community. Instead, splitting methods have been widely
636 < accepted since they exploit natural decompositions of the
637 < system\cite{Tuckerman1992, McLachlan1998}.
630 > methods are not suitable for Hamiltonian
631 > system.\cite{Cartwright1992} Recently, various high-order explicit
632 > Runge-Kutta methods \cite{Owren1992,Chen2003} have been developed to
633 > overcome this instability. However, due to computational penalty
634 > involved in implementing the Runge-Kutta methods, they have not
635 > attracted much attention from the Molecular Dynamics community.
636 > Instead, splitting methods have been widely accepted since they
637 > exploit natural decompositions of the system.\cite{McLachlan1998,
638 > Tuckerman1992}
639  
640   \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
641  
# Line 652 | Line 658 | simple first order expression is then given by the Lie
658   problem. If $H_1$ and $H_2$ can be integrated using exact
659   propagators $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a
660   simple first order expression is then given by the Lie-Trotter
661 < formula
661 > formula\cite{Trotter1959}
662   \begin{equation}
663   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
664   \label{introEquation:firstOrderSplitting}
# Line 673 | Line 679 | local errors proportional to $h^2$, while the Strang s
679   The Lie-Trotter
680   splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
681   local errors proportional to $h^2$, while the Strang splitting gives
682 < a second-order decomposition,
682 > a second-order decomposition,\cite{Strang1968}
683   \begin{equation}
684   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
685   _{1,h/2} , \label{introEquation:secondOrderSplitting}
# Line 729 | Line 735 | the equations of motion would follow:
735  
736   \item Use the half step velocities to move positions one whole step, $\Delta t$.
737  
738 < \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
738 > \item Evaluate the forces at the new positions, $q(\Delta t)$, and use the new forces to complete the velocity move.
739  
740   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
741   \end{enumerate}
# Line 748 | Line 754 | q(\Delta t)} \right]. %
754  
755   \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
756  
757 < The Baker-Campbell-Hausdorff formula can be used to determine the
758 < local error of a splitting method in terms of the commutator of the
759 < operators(Eq.~\ref{introEquation:exponentialOperator}) associated with
760 < the sub-propagator. For operators $hX$ and $hY$ which are associated
761 < with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
757 > The Baker-Campbell-Hausdorff formula\cite{Gilmore1974} can be used
758 > to determine the local error of a splitting method in terms of the
759 > commutator of the
760 > operators(Eq.~\ref{introEquation:exponentialOperator}) associated
761 > with the sub-propagator. For operators $hX$ and $hY$ which are
762 > associated with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we
763 > have
764   \begin{equation}
765   \exp (hX + hY) = \exp (hZ)
766   \end{equation}
# Line 782 | Line 790 | order methods. Yoshida proposed an elegant way to comp
790   \end{equation}
791   A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
792   order methods. Yoshida proposed an elegant way to compose higher
793 < order methods based on symmetric splitting\cite{Yoshida1990}. Given
793 > order methods based on symmetric splitting.\cite{Yoshida1990} Given
794   a symmetric second order base method $ \varphi _h^{(2)} $, a
795   fourth-order symmetric method can be constructed by composing,
796   \[
# Line 868 | Line 876 | surface and to locate the local minimum. While converg
876   minimization to find a more reasonable conformation. Several energy
877   minimization methods have been developed to exploit the energy
878   surface and to locate the local minimum. While converging slowly
879 < near the minimum, steepest descent method is extremely robust when
879 > near the minimum, the steepest descent method is extremely robust when
880   systems are strongly anharmonic. Thus, it is often used to refine
881   structures from crystallographic data. Relying on the Hessian,
882   advanced methods like Newton-Raphson converge rapidly to a local
# Line 887 | Line 895 | end up setting the temperature of the system to a fina
895   temperature. Beginning at a lower temperature and gradually
896   increasing the temperature by assigning larger random velocities, we
897   end up setting the temperature of the system to a final temperature
898 < at which the simulation will be conducted. In heating phase, we
898 > at which the simulation will be conducted. In the heating phase, we
899   should also keep the system from drifting or rotating as a whole. To
900   do this, the net linear momentum and angular momentum of the system
901   is shifted to zero after each resampling from the Maxwell -Boltzman
# Line 943 | Line 951 | evaluation is to apply spherical cutoffs where particl
951   %cutoff and minimum image convention
952   Another important technique to improve the efficiency of force
953   evaluation is to apply spherical cutoffs where particles farther
954 < than a predetermined distance are not included in the calculation
955 < \cite{Frenkel1996}. The use of a cutoff radius will cause a
956 < discontinuity in the potential energy curve. Fortunately, one can
954 > than a predetermined distance are not included in the
955 > calculation.\cite{Frenkel1996} The use of a cutoff radius will cause
956 > a discontinuity in the potential energy curve. Fortunately, one can
957   shift a simple radial potential to ensure the potential curve go
958   smoothly to zero at the cutoff radius. The cutoff strategy works
959   well for Lennard-Jones interaction because of its short range
# Line 954 | Line 962 | with rapid and absolute convergence, has proved to min
962   in simulations. The Ewald summation, in which the slowly decaying
963   Coulomb potential is transformed into direct and reciprocal sums
964   with rapid and absolute convergence, has proved to minimize the
965 < periodicity artifacts in liquid simulations. Taking the advantages
966 < of the fast Fourier transform (FFT) for calculating discrete Fourier
967 < transforms, the particle mesh-based
965 > periodicity artifacts in liquid simulations. Taking advantage of
966 > fast Fourier transform (FFT) techniques for calculating discrete
967 > Fourier transforms, the particle mesh-based
968   methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
969   $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
970   \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
# Line 966 | Line 974 | charge-neutralized Coulomb potential method developed
974   simulation community, these two methods are difficult to implement
975   correctly and efficiently. Instead, we use a damped and
976   charge-neutralized Coulomb potential method developed by Wolf and
977 < his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
977 > his coworkers.\cite{Wolf1999} The shifted Coulomb potential for
978   particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
979   \begin{equation}
980   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
# Line 1029 | Line 1037 | Fourier transforming raw data from a series of neutron
1037   function}, is of most fundamental importance to liquid theory.
1038   Experimentally, pair distribution functions can be gathered by
1039   Fourier transforming raw data from a series of neutron diffraction
1040 < experiments and integrating over the surface factor
1041 < \cite{Powles1973}. The experimental results can serve as a criterion
1042 < to justify the correctness of a liquid model. Moreover, various
1043 < equilibrium thermodynamic and structural properties can also be
1044 < expressed in terms of the radial distribution function
1045 < \cite{Allen1987}. The pair distribution functions $g(r)$ gives the
1046 < probability that a particle $i$ will be located at a distance $r$
1047 < from a another particle $j$ in the system
1040 > experiments and integrating over the surface
1041 > factor.\cite{Powles1973} The experimental results can serve as a
1042 > criterion to justify the correctness of a liquid model. Moreover,
1043 > various equilibrium thermodynamic and structural properties can also
1044 > be expressed in terms of the radial distribution
1045 > function.\cite{Allen1987} The pair distribution functions $g(r)$
1046 > gives the probability that a particle $i$ will be located at a
1047 > distance $r$ from a another particle $j$ in the system
1048   \begin{equation}
1049   g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1050   \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
# Line 1059 | Line 1067 | If $A$ and $B$ refer to same variable, this kind of co
1067   \label{introEquation:timeCorrelationFunction}
1068   \end{equation}
1069   If $A$ and $B$ refer to same variable, this kind of correlation
1070 < functions are called \emph{autocorrelation functions}. One example
1063 < of auto correlation function is the velocity auto-correlation
1070 > functions are called \emph{autocorrelation functions}. One typical example is the velocity autocorrelation
1071   function which is directly related to transport properties of
1072   molecular liquids:
1073 < \[
1073 > \begin{equation}
1074   D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1075   \right\rangle } dt
1076 < \]
1076 > \end{equation}
1077   where $D$ is diffusion constant. Unlike the velocity autocorrelation
1078   function, which is averaged over time origins and over all the
1079   atoms, the dipole autocorrelation functions is calculated for the
1080   entire system. The dipole autocorrelation function is given by:
1081 < \[
1081 > \begin{equation}
1082   c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1083   \right\rangle
1084 < \]
1084 > \end{equation}
1085   Here $u_{tot}$ is the net dipole of the entire system and is given
1086   by
1087 < \[
1087 > \begin{equation}
1088   u_{tot} (t) = \sum\limits_i {u_i (t)}.
1089 < \]
1089 > \end{equation}
1090   In principle, many time correlation functions can be related to
1091   Fourier transforms of the infrared, Raman, and inelastic neutron
1092   scattering spectra of molecular liquids. In practice, one can
1093   extract the IR spectrum from the intensity of the molecular dipole
1094   fluctuation at each frequency using the following relationship:
1095 < \[
1095 > \begin{equation}
1096   \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1097   i2\pi vt} dt}.
1098 < \]
1098 > \end{equation}
1099  
1100   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1101  
1102   Rigid bodies are frequently involved in the modeling of different
1103 < areas, from engineering, physics, to chemistry. For example,
1103 > areas, including engineering, physics and chemistry. For example,
1104   missiles and vehicles are usually modeled by rigid bodies.  The
1105   movement of the objects in 3D gaming engines or other physics
1106   simulators is governed by rigid body dynamics. In molecular
1107   simulations, rigid bodies are used to simplify protein-protein
1108 < docking studies\cite{Gray2003}.
1108 > docking studies.\cite{Gray2003}
1109  
1110   It is very important to develop stable and efficient methods to
1111   integrate the equations of motion for orientational degrees of
# Line 1110 | Line 1117 | still remain. A singularity-free representation utiliz
1117   angles can overcome this difficulty\cite{Barojas1973}, the
1118   computational penalty and the loss of angular momentum conservation
1119   still remain. A singularity-free representation utilizing
1120 < quaternions was developed by Evans in 1977\cite{Evans1977}.
1120 > quaternions was developed by Evans in 1977.\cite{Evans1977}
1121   Unfortunately, this approach used a nonseparable Hamiltonian
1122   resulting from the quaternion representation, which prevented the
1123   symplectic algorithm from being utilized. Another different approach
# Line 1119 | Line 1126 | the SHAKE and Rattle algorithms also converge very slo
1126   deriving from potential energy and constraint forces which are used
1127   to guarantee the rigidness. However, due to their iterative nature,
1128   the SHAKE and Rattle algorithms also converge very slowly when the
1129 < number of constraints increases\cite{Ryckaert1977, Andersen1983}.
1129 > number of constraints increases.\cite{Ryckaert1977, Andersen1983}
1130  
1131   A break-through in geometric literature suggests that, in order to
1132   develop a long-term integration scheme, one should preserve the
# Line 1129 | Line 1136 | An alternative method using the quaternion representat
1136   proposed to evolve the Hamiltonian system in a constraint manifold
1137   by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1138   An alternative method using the quaternion representation was
1139 < developed by Omelyan\cite{Omelyan1998}. However, both of these
1139 > developed by Omelyan.\cite{Omelyan1998} However, both of these
1140   methods are iterative and inefficient. In this section, we descibe a
1141   symplectic Lie-Poisson integrator for rigid bodies developed by
1142   Dullweber and his coworkers\cite{Dullweber1997} in depth.
1143  
1144   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1145 < The motion of a rigid body is Hamiltonian with the Hamiltonian
1139 < function
1145 > The Hamiltonian of a rigid body is given by
1146   \begin{equation}
1147   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1148   V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
# Line 1195 | Line 1201 | For a body fixed vector $X_i$ with respect to the cent
1201   1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1202   \]
1203   For a body fixed vector $X_i$ with respect to the center of mass of
1204 < the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1204 > the rigid body, its corresponding lab fixed vector $X_i^{lab}$  is
1205   given as
1206   \begin{equation}
1207   X_i^{lab} = Q X_i + q.
# Line 1250 | Line 1256 | motion. This unique property eliminates the requiremen
1256   Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1257   Lagrange multiplier $\Lambda$ is absent from the equations of
1258   motion. This unique property eliminates the requirement of
1259 < iterations which can not be avoided in other methods\cite{Kol1997,
1260 < Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1259 > iterations which can not be avoided in other methods.\cite{Kol1997,
1260 > Omelyan1998} Applying the hat-map isomorphism, we obtain the
1261   equation of motion for angular momentum in the body frame
1262   \begin{equation}
1263   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
# Line 1348 | Line 1354 | _1 }.
1354   \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1355   _1 }.
1356   \]
1357 < The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1352 < $F(\pi )$ and $G(\pi )$ is defined by
1357 > The non-canonical Lie-Poisson bracket $\{F, G\}$ of two functions $F(\pi )$ and $G(\pi )$ is defined by
1358   \[
1359   \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1360   ).
# Line 1358 | Line 1363 | norm of the angular momentum, $\parallel \pi
1363   function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1364   conserved quantity in Poisson system. We can easily verify that the
1365   norm of the angular momentum, $\parallel \pi
1366 < \parallel$, is a \emph{Casimir}\cite{McLachlan1993}. Let$ F(\pi ) = S(\frac{{\parallel
1366 > \parallel$, is a \emph{Casimir}.\cite{McLachlan1993} Let $F(\pi ) = S(\frac{{\parallel
1367   \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1368   then by the chain rule
1369   \[
# Line 1379 | Line 1384 | of motion corresponding to potential energy and kineti
1384   The Hamiltonian of rigid body can be separated in terms of kinetic
1385   energy and potential energy, $H = T(p,\pi ) + V(q,Q)$. The equations
1386   of motion corresponding to potential energy and kinetic energy are
1387 < listed in Table~\ref{introTable:rbEquations}
1387 > listed in Table~\ref{introTable:rbEquations}.
1388   \begin{table}
1389   \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1390   \label{introTable:rbEquations}
# Line 1437 | Line 1442 | has been applied in a variety of studies. This section
1442   As an alternative to newtonian dynamics, Langevin dynamics, which
1443   mimics a simple heat bath with stochastic and dissipative forces,
1444   has been applied in a variety of studies. This section will review
1445 < the theory of Langevin dynamics. A brief derivation of generalized
1445 > the theory of Langevin dynamics. A brief derivation of the generalized
1446   Langevin equation will be given first. Following that, we will
1447 < discuss the physical meaning of the terms appearing in the equation
1443 < as well as the calculation of friction tensor from hydrodynamics
1444 < theory.
1447 > discuss the physical meaning of the terms appearing in the equation.
1448  
1449   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1450  
# Line 1450 | Line 1453 | Harmonic bath model is the derivation of the Generaliz
1453   environment, has been widely used in quantum chemistry and
1454   statistical mechanics. One of the successful applications of
1455   Harmonic bath model is the derivation of the Generalized Langevin
1456 < Dynamics (GLE). Lets consider a system, in which the degree of
1456 > Dynamics (GLE). Consider a system, in which the degree of
1457   freedom $x$ is assumed to couple to the bath linearly, giving a
1458   Hamiltonian of the form
1459   \begin{equation}
# Line 1461 | Line 1464 | H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_
1464   with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1465   \[
1466   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1467 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1467 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  x_\alpha ^2 }
1468   \right\}}
1469   \]
1470   where the index $\alpha$ runs over all the bath degrees of freedom,
# Line 1514 | Line 1517 | where  $p$ is real and  $L$ is called the Laplace Tran
1517   L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1518   \]
1519   where  $p$ is real and  $L$ is called the Laplace Transform
1520 < Operator. Below are some important properties of Laplace transform
1520 > Operator. Below are some important properties of the Laplace transform
1521   \begin{eqnarray*}
1522   L(x + y)  & = & L(x) + L(y) \\
1523   L(ax)     & = & aL(x) \\
# Line 1583 | Line 1586 | m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int
1586   (t)\dot x(t - \tau )d\tau }  + R(t)
1587   \label{introEuqation:GeneralizedLangevinDynamics}
1588   \end{equation}
1589 < which is known as the \emph{generalized Langevin equation}.
1589 > which is known as the \emph{generalized Langevin equation} (GLE).
1590  
1591   \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1592  
1593   One may notice that $R(t)$ depends only on initial conditions, which
1594   implies it is completely deterministic within the context of a
1595   harmonic bath. However, it is easy to verify that $R(t)$ is totally
1596 < uncorrelated to $x$ and $\dot x$,$\left\langle {x(t)R(t)}
1596 > uncorrelated to $x$ and $\dot x$, $\left\langle {x(t)R(t)}
1597   \right\rangle  = 0, \left\langle {\dot x(t)R(t)} \right\rangle  =
1598   0.$ This property is what we expect from a truly random process. As
1599   long as the model chosen for $R(t)$ was a gaussian distribution in
# Line 1619 | Line 1622 | taken as a $delta$ function in time:
1622   infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1623   taken as a $delta$ function in time:
1624   \[
1625 < \xi (t) = 2\xi _0 \delta (t)
1625 > \xi (t) = 2\xi _0 \delta (t).
1626   \]
1627   Hence, the convolution integral becomes
1628   \[
# Line 1644 | Line 1647 | q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \o
1647   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1648   ^2 }}x(0),
1649   \]
1650 < we can rewrite $R(T)$ as
1650 > we can rewrite $R(t)$ as
1651   \[
1652   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1653   \]

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines