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 2719 by tim, Tue Apr 18 22:38:19 2006 UTC vs.
Revision 2776 by tim, Thu May 25 21:32:14 2006 UTC

# Line 831 | Line 831 | $\varphi_1(t)$ and $\varphi_2(t$ respectively , we hav
831   error of splitting method in terms of commutator of the
832   operators(\ref{introEquation:exponentialOperator}) associated with
833   the sub-flow. For operators $hX$ and $hY$ which are associate to
834 < $\varphi_1(t)$ and $\varphi_2(t$ respectively , we have
834 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
835   \begin{equation}
836   \exp (hX + hY) = \exp (hZ)
837   \end{equation}
# Line 846 | Line 846 | can obtain
846   \]
847   Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
848   can obtain
849 < \begin{eqnarray*}
849 > \begin{equation}
850   \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
851   [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
852   & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
853   \ldots )
854 < \end{eqnarray*}
854 > \end{equation}
855   Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
856   error of Spring splitting is proportional to $h^3$. The same
857   procedure can be applied to general splitting,  of the form
# Line 883 | Line 883 | As a special discipline of molecular modeling, Molecul
883  
884   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
885  
886 < As a special discipline of molecular modeling, Molecular dynamics
887 < has proven to be a powerful tool for studying the functions of
888 < biological systems, providing structural, thermodynamic and
889 < dynamical information.
890 <
891 < One of the principal tools for modeling proteins, nucleic acids and
892 < their complexes. Stability of proteins Folding of proteins.
893 < Molecular recognition by:proteins, DNA, RNA, lipids, hormones STP,
894 < etc. Enzyme reactions Rational design of biologically active
895 < molecules (drug design) Small and large-scale conformational
896 < changes. determination and construction of 3D structures (homology,
897 < Xray diffraction, NMR) Dynamic processes such as ion transport in
898 < biological systems.
899 <
900 < Macroscopic properties are related to microscopic behavior.
886 > As one of the principal tools of molecular modeling, Molecular
887 > dynamics has proven to be a powerful tool for studying the functions
888 > of biological systems, providing structural, thermodynamic and
889 > dynamical information. The basic idea of molecular dynamics is that
890 > macroscopic properties are related to microscopic behavior and
891 > microscopic behavior can be calculated from the trajectories in
892 > simulations. For instance, instantaneous temperature of an
893 > Hamiltonian system of $N$ particle can be measured by
894 > \[
895 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
896 > \]
897 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
898 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
899 > the boltzman constant.
900  
901 < Time dependent (and independent) microscopic behavior of a molecule
902 < can be calculated by molecular dynamics simulations.
903 <
904 < \subsection{\label{introSec:mdInit}Initialization}
905 <
906 < \subsection{\label{introSec:forceEvaluation}Force Evaluation}
901 > A typical molecular dynamics run consists of three essential steps:
902 > \begin{enumerate}
903 >  \item Initialization
904 >    \begin{enumerate}
905 >    \item Preliminary preparation
906 >    \item Minimization
907 >    \item Heating
908 >    \item Equilibration
909 >    \end{enumerate}
910 >  \item Production
911 >  \item Analysis
912 > \end{enumerate}
913 > These three individual steps will be covered in the following
914 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
915 > initialization of a simulation. Sec.~\ref{introSec:production} will
916 > discusses issues in production run. Sec.~\ref{introSection:Analysis}
917 > provides the theoretical tools for trajectory analysis.
918  
919 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
919 > \subsection{\label{introSec:initialSystemSettings}Initialization}
920 >
921 > \subsubsection{Preliminary preparation}
922 >
923 > When selecting the starting structure of a molecule for molecular
924 > simulation, one may retrieve its Cartesian coordinates from public
925 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
926 > thousands of crystal structures of molecules are discovered every
927 > year, many more remain unknown due to the difficulties of
928 > purification and crystallization. Even for the molecule with known
929 > structure, some important information is missing. For example, the
930 > missing hydrogen atom which acts as donor in hydrogen bonding must
931 > be added. Moreover, in order to include electrostatic interaction,
932 > one may need to specify the partial charges for individual atoms.
933 > Under some circumstances, we may even need to prepare the system in
934 > a special setup. For instance, when studying transport phenomenon in
935 > membrane system, we may prepare the lipids in bilayer structure
936 > instead of placing lipids randomly in solvent, since we are not
937 > interested in self-aggregation and it takes a long time to happen.
938 >
939 > \subsubsection{Minimization}
940 >
941 > It is quite possible that some of molecules in the system from
942 > preliminary preparation may be overlapped with each other. This
943 > close proximity leads to high potential energy which consequently
944 > jeopardizes any molecular dynamics simulations. To remove these
945 > steric overlaps, one typically performs energy minimization to find
946 > a more reasonable conformation. Several energy minimization methods
947 > have been developed to exploit the energy surface and to locate the
948 > local minimum. While converging slowly near the minimum, steepest
949 > descent method is extremely robust when systems are far from
950 > harmonic. Thus, it is often used to refine structure from
951 > crystallographic data. Relied on the gradient or hessian, advanced
952 > methods like conjugate gradient and Newton-Raphson converge rapidly
953 > to a local minimum, while become unstable if the energy surface is
954 > far from quadratic. Another factor must be taken into account, when
955 > choosing energy minimization method, is the size of the system.
956 > Steepest descent and conjugate gradient can deal with models of any
957 > size. Because of the limit of computation power to calculate hessian
958 > matrix and insufficient storage capacity to store them, most
959 > Newton-Raphson methods can not be used with very large models.
960  
961 + \subsubsection{Heating}
962 +
963 + Typically, Heating is performed by assigning random velocities
964 + according to a Gaussian distribution for a temperature. Beginning at
965 + a lower temperature and gradually increasing the temperature by
966 + assigning greater random velocities, we end up with setting the
967 + temperature of the system to a final temperature at which the
968 + simulation will be conducted. In heating phase, we should also keep
969 + the system from drifting or rotating as a whole. Equivalently, the
970 + net linear momentum and angular momentum of the system should be
971 + shifted to zero.
972 +
973 + \subsubsection{Equilibration}
974 +
975 + The purpose of equilibration is to allow the system to evolve
976 + spontaneously for a period of time and reach equilibrium. The
977 + procedure is continued until various statistical properties, such as
978 + temperature, pressure, energy, volume and other structural
979 + properties \textit{etc}, become independent of time. Strictly
980 + speaking, minimization and heating are not necessary, provided the
981 + equilibration process is long enough. However, these steps can serve
982 + as a means to arrive at an equilibrated structure in an effective
983 + way.
984 +
985 + \subsection{\label{introSection:production}Production}
986 +
987 + Production run is the most important steps of the simulation, in
988 + which the equilibrated structure is used as a starting point and the
989 + motions of the molecules are collected for later analysis. In order
990 + to capture the macroscopic properties of the system, the molecular
991 + dynamics simulation must be performed in correct and efficient way.
992 +
993 + The most expensive part of a molecular dynamics simulation is the
994 + calculation of non-bonded forces, such as van der Waals force and
995 + Coulombic forces \textit{etc}. For a system of $N$ particles, the
996 + complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
997 + which making large simulations prohibitive in the absence of any
998 + computation saving techniques.
999 +
1000 + A natural approach to avoid system size issue is to represent the
1001 + bulk behavior by a finite number of the particles. However, this
1002 + approach will suffer from the surface effect. To offset this,
1003 + \textit{Periodic boundary condition} is developed to simulate bulk
1004 + properties with a relatively small number of particles. In this
1005 + method, the simulation box is replicated throughout space to form an
1006 + infinite lattice. During the simulation, when a particle moves in
1007 + the primary cell, its image in other cells move in exactly the same
1008 + direction with exactly the same orientation. Thus, as a particle
1009 + leaves the primary cell, one of its images will enter through the
1010 + opposite face.
1011 + %\begin{figure}
1012 + %\centering
1013 + %\includegraphics[width=\linewidth]{pbcFig.eps}
1014 + %\caption[An illustration of periodic boundary conditions]{A 2-D
1015 + %illustration of periodic boundary conditions. As one particle leaves
1016 + %the right of the simulation box, an image of it enters the left.}
1017 + %\label{introFig:pbc}
1018 + %\end{figure}
1019 +
1020 + %cutoff and minimum image convention
1021 + Another important technique to improve the efficiency of force
1022 + evaluation is to apply cutoff where particles farther than a
1023 + predetermined distance, are not included in the calculation
1024 + \cite{Frenkel1996}. The use of a cutoff radius will cause a
1025 + discontinuity in the potential energy curve. Fortunately, one can
1026 + shift the potential to ensure the potential curve go smoothly to
1027 + zero at the cutoff radius. Cutoff strategy works pretty well for
1028 + Lennard-Jones interaction because of its short range nature.
1029 + However, simply truncating the electrostatic interaction with the
1030 + use of cutoff has been shown to lead to severe artifacts in
1031 + simulations. Ewald summation, in which the slowly conditionally
1032 + convergent Coulomb potential is transformed into direct and
1033 + reciprocal sums with rapid and absolute convergence, has proved to
1034 + minimize the periodicity artifacts in liquid simulations. Taking the
1035 + advantages of the fast Fourier transform (FFT) for calculating
1036 + discrete Fourier transforms, the particle mesh-based methods are
1037 + accelerated from $O(N^{3/2})$ to $O(N logN)$. An alternative
1038 + approach is \emph{fast multipole method}, which treats Coulombic
1039 + interaction exactly at short range, and approximate the potential at
1040 + long range through multipolar expansion. In spite of their wide
1041 + acceptances at the molecular simulation community, these two methods
1042 + are hard to be implemented correctly and efficiently. Instead, we
1043 + use a damped and charge-neutralized Coulomb potential method
1044 + developed by Wolf and his coworkers. The shifted Coulomb potential
1045 + for particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1046 + \begin{equation}
1047 + V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1048 + r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1049 + R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1050 + r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1051 + \end{equation}
1052 + where $\alpha$ is the convergence parameter. Due to the lack of
1053 + inherent periodicity and rapid convergence,this method is extremely
1054 + efficient and easy to implement.
1055 + %\begin{figure}
1056 + %\centering
1057 + %\includegraphics[width=\linewidth]{pbcFig.eps}
1058 + %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
1059 + %\label{introFigure:shiftedCoulomb}
1060 + %\end{figure}
1061 +
1062 + %multiple time step
1063 +
1064 + \subsection{\label{introSection:Analysis} Analysis}
1065 +
1066 + Recently, advanced visualization technique are widely applied to
1067 + monitor the motions of molecules. Although the dynamics of the
1068 + system can be described qualitatively from animation, quantitative
1069 + trajectory analysis are more appreciable. According to the
1070 + principles of Statistical Mechanics,
1071 + Sec.~\ref{introSection:statisticalMechanics}, one can compute
1072 + thermodynamics properties, analyze fluctuations of structural
1073 + parameters, and investigate time-dependent processes of the molecule
1074 + from the trajectories.
1075 +
1076 + \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1077 +
1078 + Thermodynamics properties, which can be expressed in terms of some
1079 + function of the coordinates and momenta of all particles in the
1080 + system, can be directly computed from molecular dynamics. The usual
1081 + way to measure the pressure is based on virial theorem of Clausius
1082 + which states that the virial is equal to $-3Nk_BT$. For a system
1083 + with forces between particles, the total virial, $W$, contains the
1084 + contribution from external pressure and interaction between the
1085 + particles:
1086 + \[
1087 + W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1088 + f_{ij} } } \right\rangle
1089 + \]
1090 + where $f_{ij}$ is the force between particle $i$ and $j$ at a
1091 + distance $r_{ij}$. Thus, the expression for the pressure is given
1092 + by:
1093 + \begin{equation}
1094 + P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1095 + < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1096 + \end{equation}
1097 +
1098 + \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1099 +
1100 + Structural Properties of a simple fluid can be described by a set of
1101 + distribution functions. Among these functions,\emph{pair
1102 + distribution function}, also known as \emph{radial distribution
1103 + function}, is of most fundamental importance to liquid-state theory.
1104 + Pair distribution function can be gathered by Fourier transforming
1105 + raw data from a series of neutron diffraction experiments and
1106 + integrating over the surface factor \cite{Powles73}. The experiment
1107 + result can serve as a criterion to justify the correctness of the
1108 + theory. Moreover, various equilibrium thermodynamic and structural
1109 + properties can also be expressed in terms of radial distribution
1110 + function \cite{allen87:csl}.
1111 +
1112 + A pair distribution functions $g(r)$ gives the probability that a
1113 + particle $i$ will be located at a distance $r$ from a another
1114 + particle $j$ in the system
1115 + \[
1116 + g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1117 + \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1118 + \]
1119 + Note that the delta function can be replaced by a histogram in
1120 + computer simulation. Figure
1121 + \ref{introFigure:pairDistributionFunction} shows a typical pair
1122 + distribution function for the liquid argon system. The occurrence of
1123 + several peaks in the plot of $g(r)$ suggests that it is more likely
1124 + to find particles at certain radial values than at others. This is a
1125 + result of the attractive interaction at such distances. Because of
1126 + the strong repulsive forces at short distance, the probability of
1127 + locating particles at distances less than about 2.5{\AA} from each
1128 + other is essentially zero.
1129 +
1130 + %\begin{figure}
1131 + %\centering
1132 + %\includegraphics[width=\linewidth]{pdf.eps}
1133 + %\caption[Pair distribution function for the liquid argon
1134 + %]{Pair distribution function for the liquid argon}
1135 + %\label{introFigure:pairDistributionFunction}
1136 + %\end{figure}
1137 +
1138 + \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1139 + Properties}
1140 +
1141 + Time-dependent properties are usually calculated using \emph{time
1142 + correlation function}, which correlates random variables $A$ and $B$
1143 + at two different time
1144 + \begin{equation}
1145 + C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1146 + \label{introEquation:timeCorrelationFunction}
1147 + \end{equation}
1148 + If $A$ and $B$ refer to same variable, this kind of correlation
1149 + function is called \emph{auto correlation function}. One example of
1150 + auto correlation function is velocity auto-correlation function
1151 + which is directly related to transport properties of molecular
1152 + liquids:
1153 + \[
1154 + D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1155 + \right\rangle } dt
1156 + \]
1157 + where $D$ is diffusion constant. Unlike velocity autocorrelation
1158 + function which is averaging over time origins and over all the
1159 + atoms, dipole autocorrelation are calculated for the entire system.
1160 + The dipole autocorrelation function is given by:
1161 + \[
1162 + c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1163 + \right\rangle
1164 + \]
1165 + Here $u_{tot}$ is the net dipole of the entire system and is given
1166 + by
1167 + \[
1168 + u_{tot} (t) = \sum\limits_i {u_i (t)}
1169 + \]
1170 + In principle, many time correlation functions can be related with
1171 + Fourier transforms of the infrared, Raman, and inelastic neutron
1172 + scattering spectra of molecular liquids. In practice, one can
1173 + extract the IR spectrum from the intensity of dipole fluctuation at
1174 + each frequency using the following relationship:
1175 + \[
1176 + \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1177 + i2\pi vt} dt}
1178 + \]
1179 +
1180   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1181  
1182   Rigid bodies are frequently involved in the modeling of different
# Line 968 | Line 1237 | Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1237   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1238   constrained Hamiltonian equation subjects to a holonomic constraint,
1239   \begin{equation}
1240 < Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1240 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1241   \end{equation}
1242   which is used to ensure rotation matrix's orthogonality.
1243   Differentiating \ref{introEquation:orthogonalConstraint} and using
# Line 993 | Line 1262 | simply evolve the system in constraint manifold. The t
1262   In general, there are two ways to satisfy the holonomic constraints.
1263   We can use constraint force provided by lagrange multiplier on the
1264   normal manifold to keep the motion on constraint space. Or we can
1265 < simply evolve the system in constraint manifold. The two method are
1266 < proved to be equivalent. The holonomic constraint and equations of
1267 < motions define a constraint manifold for rigid body
1265 > simply evolve the system in constraint manifold. These two methods
1266 > are proved to be equivalent. The holonomic constraint and equations
1267 > of motions define a constraint manifold for rigid body
1268   \[
1269   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1270   \right\}.
# Line 1156 | Line 1425 | e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1
1425   e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1426   )
1427   \]
1428 <
1160 < The flow maps for $T_2^r$ and $T_2^r$ can be found in the same
1428 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1429   manner.
1430  
1431   In order to construct a second-order symplectic method, we split the
# Line 1209 | Line 1477 | kinetic energy are listed in the below table,
1477   \]
1478   The equations of motion corresponding to potential energy and
1479   kinetic energy are listed in the below table,
1480 + \begin{table}
1481 + \caption{Equations of motion due to Potential and Kinetic Energies}
1482   \begin{center}
1483   \begin{tabular}{|l|l|}
1484    \hline
# Line 1221 | Line 1491 | A second-order symplectic method is now obtained by th
1491    \hline
1492   \end{tabular}
1493   \end{center}
1494 < A second-order symplectic method is now obtained by the composition
1495 < of the flow maps,
1494 > \end{table}
1495 > A second-order symplectic method is now obtained by the
1496 > composition of the flow maps,
1497   \[
1498   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1499   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
# Line 1480 | Line 1751 | particles is given in section \ref{introSection:fricti
1751   coefficient $\xi _0$ can either be calculated from spectral density
1752   or be determined by Stokes' law for regular shaped particles.A
1753   briefly review on calculating friction tensor for arbitrary shaped
1754 < particles is given in section \ref{introSection:frictionTensor}.
1754 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1755  
1756   \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1757  
# Line 1518 | Line 1789 | Equation, \zeta can be taken as a scalar. In general,
1789   when the system become more and more complicate. Instead, various
1790   approaches based on hydrodynamics have been developed to calculate
1791   the friction coefficients. The friction effect is isotropic in
1792 < Equation, \zeta can be taken as a scalar. In general, friction
1793 < tensor \Xi is a $6\times 6$ matrix given by
1792 > Equation, $\zeta$ can be taken as a scalar. In general, friction
1793 > tensor $\Xi$ is a $6\times 6$ matrix given by
1794   \[
1795   \Xi  = \left( {\begin{array}{*{20}c}
1796     {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
# Line 1770 | Line 2041 | joining center of resistance $R$ and origin $O$.
2041   \]
2042   where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2043   joining center of resistance $R$ and origin $O$.
1773
1774 %\section{\label{introSection:correlationFunctions}Correlation Functions}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines