| 1 | xsun | 3336 | \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND} | 
| 2 | xsun | 3347 |  | 
| 3 |  |  | \section{Background on the Problem\label{In:sec:pro}} | 
| 4 | xsun | 3365 | Phospholipid molecules are the primary topic of this dissertation | 
| 5 |  |  | because of their critical role as the foundation of biological | 
| 6 |  |  | membranes. Lipids, when dispersed in water, self assemble into bilayer | 
| 7 |  |  | structures. The phase behavior of lipid bilayers have been explored | 
| 8 |  |  | experimentally~\cite{Cevc87}, however, a complete understanding of the | 
| 9 |  |  | mechanism for self-assembly has not been achieved. | 
| 10 | xsun | 3347 |  | 
| 11 |  |  | \subsection{Ripple Phase\label{In:ssec:ripple}} | 
| 12 | xsun | 3365 | The $P_{\beta'}$ {\it ripple phase} of lipid bilayers, named from the | 
| 13 | xsun | 3347 | periodic buckling of the membrane, is an intermediate phase which is | 
| 14 |  |  | developed either from heating the gel phase $L_{\beta'}$ or cooling | 
| 15 | xsun | 3365 | the fluid phase $L_\alpha$. Although the ripple phase has been | 
| 16 |  |  | observed in several | 
| 17 | xsun | 3347 | experiments~\cite{Sun96,Katsaras00,Copeland80,Meyer96,Kaasgaard03}, | 
| 18 | xsun | 3365 | the physical mechanism for the formation of the ripple phase has never been | 
| 19 | xsun | 3347 | explained and the microscopic structure of the ripple phase has never | 
| 20 |  |  | been elucidated by experiments. Computational simulation is a perfect | 
| 21 | xsun | 3365 | tool to study the microscopic properties for a system. However, the | 
| 22 |  |  | large length scale the ripple structure and the long time scale | 
| 23 | xsun | 3347 | of the formation of the ripples are crucial obstacles to performing | 
| 24 | xsun | 3365 | the actual work. The principal ideas explored in this dissertation are | 
| 25 |  |  | attempts to break the computational task up by | 
| 26 | xsun | 3347 | \begin{itemize} | 
| 27 | xsun | 3365 | \item Simplifying the lipid model. | 
| 28 |  |  | \item Improving algorithm for integrating the equations of motion. | 
| 29 | xsun | 3347 | \end{itemize} | 
| 30 | xsun | 3365 | In chapters~\ref{chap:mc} and~\ref{chap:md}, we use a simple point | 
| 31 |  |  | dipole model and a coarse-grained model to perform the Monte Carlo and | 
| 32 |  |  | Molecular Dynamics simulations respectively, and in | 
| 33 |  |  | chapter~\ref{chap:ld}, we develop a Langevin Dynamics algorithm which | 
| 34 |  |  | excludes the explicit solvent to improve the efficiency of the | 
| 35 |  |  | simulations. | 
| 36 | xsun | 3347 |  | 
| 37 |  |  | \subsection{Lattice Model\label{In:ssec:model}} | 
| 38 | xsun | 3365 | The gel-like characteristic of the ripple phase makes it feasible to | 
| 39 |  |  | apply a lattice model to study the system. It is claimed that the | 
| 40 |  |  | packing of the lipid molecules in ripple phase is | 
| 41 | xsun | 3347 | hexagonal~\cite{Cevc87}. The popular $2$ dimensional lattice models, | 
| 42 | xsun | 3365 | {\it i.e.}, the Ising, $X-Y$, and Heisenberg models, show {\it | 
| 43 |  |  | frustration} on triangular lattices. The Hamiltonian of the systems | 
| 44 |  |  | are given by | 
| 45 |  |  | \begin{equation} | 
| 46 |  |  | H = | 
| 47 |  |  | \begin{cases} | 
| 48 |  |  | -J \sum_n \sum_{n'} s_n s_n' & \text{Ising}, \\ | 
| 49 |  |  | -J \sum_n \sum_{n'} \vec s_n \cdot \vec s_{n'} & \text{$X-Y$ and | 
| 50 |  |  | Heisenberg}, | 
| 51 |  |  | \end{cases} | 
| 52 |  |  | \end{equation} | 
| 53 |  |  | where $J$ has non zero value only when spins $s_n$ ($\vec s_n$) and | 
| 54 | xsun | 3371 | $s_{n'}$ ($\vec s_{n'}$) are the nearest neighbours. When $J > 0$, the | 
| 55 |  |  | spins prefer aligned with each other, and if $J < 0$, the spins want | 
| 56 |  |  | to be anti-aligned. | 
| 57 |  |  |  | 
| 58 | xsun | 3347 | \begin{figure} | 
| 59 |  |  | \centering | 
| 60 | xsun | 3371 | \includegraphics[width=3in]{./figures/inFrustration.pdf} | 
| 61 |  |  | \caption{Frustration on triangular lattice, the spins and dipoles are | 
| 62 |  |  | represented by arrows. The multiple local minima of energy states | 
| 63 |  |  | induce the frustration for spins and dipoles picking the directions.} | 
| 64 | xsun | 3354 | \label{Infig:frustration} | 
| 65 | xsun | 3347 | \end{figure} | 
| 66 | xsun | 3371 | The spins in figure~\ref{Infig:frustration} shows an illustration of | 
| 67 |  |  | the frustration for $J < 0$ on a triangular lattice. There are | 
| 68 |  |  | multiple local minima energy states which are independent of the | 
| 69 |  |  | direction of the spin on top of the triangle, therefore infinite | 
| 70 |  |  | possibilities for the packing of spins induce what is known as | 
| 71 |  |  | ``complete regular frustration'' which leads to disordered low | 
| 72 |  |  | temperature phases. The similarity goes to the dipoles on a hexagonal | 
| 73 |  |  | lattice, which are shown by the dipoles in | 
| 74 |  |  | figure~\ref{Infig:frustration}. In this circumstance, the dipoles want | 
| 75 |  |  | to be aligned, however, due to the long wave fluctuation, at low | 
| 76 |  |  | temperature, the aligned state becomes unstable, vortex is formed and | 
| 77 |  |  | results in multiple local minima of energy states. The dipole on the | 
| 78 |  |  | center of the hexagonal lattice is frustrated. | 
| 79 | xsun | 3347 |  | 
| 80 |  |  | The lack of translational degree of freedom in lattice models prevents | 
| 81 | xsun | 3365 | their utilization in models for surface buckling which would | 
| 82 |  |  | correspond to ripple formation. In chapter~\ref{chap:mc}, a modified | 
| 83 |  |  | lattice model is introduced to tackle this specific situation. | 
| 84 | xsun | 3347 |  | 
| 85 |  |  | \section{Overview of Classical Statistical Mechanics\label{In:sec:SM}} | 
| 86 |  |  | Statistical mechanics provides a way to calculate the macroscopic | 
| 87 |  |  | properties of a system from the molecular interactions used in | 
| 88 |  |  | computational simulations. This section serves as a brief introduction | 
| 89 |  |  | to key concepts of classical statistical mechanics that we used in | 
| 90 |  |  | this dissertation. Tolman gives an excellent introduction to the | 
| 91 |  |  | principles of statistical mechanics~\cite{Tolman1979}. A large part of | 
| 92 |  |  | section~\ref{In:sec:SM} will follow Tolman's notation. | 
| 93 |  |  |  | 
| 94 |  |  | \subsection{Ensembles\label{In:ssec:ensemble}} | 
| 95 |  |  | In classical mechanics, the state of the system is completely | 
| 96 |  |  | described by the positions and momenta of all particles. If we have an | 
| 97 |  |  | $N$ particle system, there are $6N$ coordinates ($3N$ position $(q_1, | 
| 98 |  |  | q_2, \ldots, q_{3N})$ and $3N$ momenta $(p_1, p_2, \ldots, p_{3N})$) | 
| 99 |  |  | to define the instantaneous state of the system. Each single set of | 
| 100 |  |  | the $6N$ coordinates can be considered as a unique point in a $6N$ | 
| 101 |  |  | dimensional space where each perpendicular axis is one of | 
| 102 |  |  | $q_{i\alpha}$ or $p_{i\alpha}$ ($i$ is the particle and $\alpha$ is | 
| 103 |  |  | the spatial axis). This $6N$ dimensional space is known as phase | 
| 104 |  |  | space. The instantaneous state of the system is a single point in | 
| 105 |  |  | phase space. A trajectory is a connected path of points in phase | 
| 106 |  |  | space. An ensemble is a collection of systems described by the same | 
| 107 |  |  | macroscopic observables but which have microscopic state | 
| 108 |  |  | distributions. In phase space an ensemble is a collection of a set of | 
| 109 |  |  | representive points. A density distribution $\rho(q^N, p^N)$ of the | 
| 110 |  |  | representive points in phase space describes the condition of an | 
| 111 |  |  | ensemble of identical systems. Since this density may change with | 
| 112 |  |  | time, it is also a function of time. $\rho(q^N, p^N, t)$ describes the | 
| 113 |  |  | ensemble at a time $t$, and $\rho(q^N, p^N, t')$ describes the | 
| 114 |  |  | ensemble at a later time $t'$. For convenience, we will use $\rho$ | 
| 115 |  |  | instead of $\rho(q^N, p^N, t)$ in the following disccusion. If we | 
| 116 |  |  | normalize $\rho$ to unity, | 
| 117 |  |  | \begin{equation} | 
| 118 |  |  | 1 = \int d \vec q~^N \int d \vec p~^N \rho, | 
| 119 | xsun | 3354 | \label{Ineq:normalized} | 
| 120 | xsun | 3347 | \end{equation} | 
| 121 |  |  | then the value of $\rho$ gives the probability of finding the system | 
| 122 |  |  | in a unit volume in the phase space. | 
| 123 |  |  |  | 
| 124 |  |  | Liouville's theorem describes the change in density $\rho$ with | 
| 125 |  |  | time. The number of representive points at a given volume in the phase | 
| 126 |  |  | space at any instant can be written as: | 
| 127 |  |  | \begin{equation} | 
| 128 | xsun | 3354 | \label{Ineq:deltaN} | 
| 129 | xsun | 3347 | \delta N = \rho~\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N. | 
| 130 |  |  | \end{equation} | 
| 131 |  |  | To calculate the change in the number of representive points in this | 
| 132 |  |  | volume, let us consider a simple condition: the change in the number | 
| 133 |  |  | of representive points in $q_1$ axis. The rate of the number of the | 
| 134 |  |  | representive points entering the volume at $q_1$ per unit time is: | 
| 135 |  |  | \begin{equation} | 
| 136 | xsun | 3354 | \label{Ineq:deltaNatq1} | 
| 137 | xsun | 3347 | \rho~\dot q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N, | 
| 138 |  |  | \end{equation} | 
| 139 |  |  | and the rate of the number of representive points leaving the volume | 
| 140 |  |  | at another position $q_1 + \delta q_1$ is: | 
| 141 |  |  | \begin{equation} | 
| 142 | xsun | 3354 | \label{Ineq:deltaNatq1plusdeltaq1} | 
| 143 | xsun | 3347 | \left( \rho + \frac{\partial \rho}{\partial q_1} \delta q_1 \right)\left(\dot q_1 + | 
| 144 |  |  | \frac{\partial \dot q_1}{\partial q_1} \delta q_1 \right)\delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N. | 
| 145 |  |  | \end{equation} | 
| 146 |  |  | Here the higher order differentials are neglected. So the change of | 
| 147 |  |  | the number of the representive points is the difference of | 
| 148 | xsun | 3354 | eq.~\ref{Ineq:deltaNatq1} and eq.~\ref{Ineq:deltaNatq1plusdeltaq1}, which | 
| 149 | xsun | 3347 | gives us: | 
| 150 |  |  | \begin{equation} | 
| 151 | xsun | 3354 | \label{Ineq:deltaNatq1axis} | 
| 152 | xsun | 3347 | -\left(\rho \frac{\partial {\dot q_1}}{\partial q_1} + \frac{\partial {\rho}}{\partial q_1} \dot q_1 \right)\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N, | 
| 153 |  |  | \end{equation} | 
| 154 |  |  | where, higher order differetials are neglected. If we sum over all the | 
| 155 |  |  | axes in the phase space, we can get the change of the number of | 
| 156 |  |  | representive points in a given volume with time: | 
| 157 |  |  | \begin{equation} | 
| 158 | xsun | 3354 | \label{Ineq:deltaNatGivenVolume} | 
| 159 | xsun | 3347 | \frac{d(\delta N)}{dt} = -\sum_{i=1}^N \left[\rho \left(\frac{\partial | 
| 160 |  |  | {\dot q_i}}{\partial q_i} + \frac{\partial | 
| 161 |  |  | {\dot p_i}}{\partial p_i}\right) + \left( \frac{\partial {\rho}}{\partial | 
| 162 |  |  | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i\right)\right]\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N. | 
| 163 |  |  | \end{equation} | 
| 164 |  |  | From Hamilton's equation of motion, | 
| 165 |  |  | \begin{equation} | 
| 166 |  |  | \frac{\partial {\dot q_i}}{\partial q_i} = - \frac{\partial | 
| 167 |  |  | {\dot p_i}}{\partial p_i}, | 
| 168 | xsun | 3354 | \label{Ineq:canonicalFormOfEquationOfMotion} | 
| 169 | xsun | 3347 | \end{equation} | 
| 170 |  |  | this cancels out the first term on the right side of | 
| 171 | xsun | 3354 | eq.~\ref{Ineq:deltaNatGivenVolume}. If both sides of | 
| 172 |  |  | eq.~\ref{Ineq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2 | 
| 173 | xsun | 3347 | \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N$, then we | 
| 174 |  |  | can derive Liouville's theorem: | 
| 175 |  |  | \begin{equation} | 
| 176 |  |  | \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = -\sum_{i} \left( | 
| 177 |  |  | \frac{\partial {\rho}}{\partial | 
| 178 |  |  | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right). | 
| 179 | xsun | 3354 | \label{Ineq:simpleFormofLiouville} | 
| 180 | xsun | 3347 | \end{equation} | 
| 181 |  |  | This is the basis of statistical mechanics. If we move the right | 
| 182 | xsun | 3354 | side of equation~\ref{Ineq:simpleFormofLiouville} to the left, we | 
| 183 | xsun | 3347 | will obtain | 
| 184 |  |  | \begin{equation} | 
| 185 |  |  | \left( \frac{\partial \rho}{\partial t} \right)_{q, p} + \sum_{i} \left( | 
| 186 |  |  | \frac{\partial {\rho}}{\partial | 
| 187 |  |  | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right) | 
| 188 |  |  | = 0. | 
| 189 | xsun | 3354 | \label{Ineq:anotherFormofLiouville} | 
| 190 | xsun | 3347 | \end{equation} | 
| 191 |  |  | It is easy to note that the left side of | 
| 192 | xsun | 3354 | equation~\ref{Ineq:anotherFormofLiouville} is the total derivative of | 
| 193 | xsun | 3347 | $\rho$ with respect of $t$, which means | 
| 194 |  |  | \begin{equation} | 
| 195 |  |  | \frac{d \rho}{dt} = 0, | 
| 196 | xsun | 3354 | \label{Ineq:conservationofRho} | 
| 197 | xsun | 3347 | \end{equation} | 
| 198 |  |  | and the rate of density change is zero in the neighborhood of any | 
| 199 |  |  | selected moving representive points in the phase space. | 
| 200 |  |  |  | 
| 201 |  |  | The condition of the ensemble is determined by the density | 
| 202 |  |  | distribution. If we consider the density distribution as only a | 
| 203 |  |  | function of $q$ and $p$, which means the rate of change of the phase | 
| 204 |  |  | space density in the neighborhood of all representive points in the | 
| 205 |  |  | phase space is zero, | 
| 206 |  |  | \begin{equation} | 
| 207 |  |  | \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = 0. | 
| 208 | xsun | 3354 | \label{Ineq:statEquilibrium} | 
| 209 | xsun | 3347 | \end{equation} | 
| 210 |  |  | We may conclude the ensemble is in {\it statistical equilibrium}. An | 
| 211 |  |  | ensemble in statistical equilibrium often means the system is also in | 
| 212 |  |  | macroscopic equilibrium. If $\left( \frac{\partial \rho}{\partial t} | 
| 213 |  |  | \right)_{q, p} = 0$, then | 
| 214 |  |  | \begin{equation} | 
| 215 |  |  | \sum_{i} \left( | 
| 216 |  |  | \frac{\partial {\rho}}{\partial | 
| 217 |  |  | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right) | 
| 218 |  |  | = 0. | 
| 219 | xsun | 3354 | \label{Ineq:constantofMotion} | 
| 220 | xsun | 3347 | \end{equation} | 
| 221 |  |  | If $\rho$ is a function only of some constant of the motion, $\rho$ is | 
| 222 |  |  | independent of time. For a conservative system, the energy of the | 
| 223 |  |  | system is one of the constants of the motion. Here are several | 
| 224 |  |  | examples: when the density distribution is constant everywhere in the | 
| 225 |  |  | phase space, | 
| 226 |  |  | \begin{equation} | 
| 227 |  |  | \rho = \mathrm{const.} | 
| 228 | xsun | 3354 | \label{Ineq:uniformEnsemble} | 
| 229 | xsun | 3347 | \end{equation} | 
| 230 |  |  | the ensemble is called {\it uniform ensemble}.  Another useful | 
| 231 |  |  | ensemble is called {\it microcanonical ensemble}, for which: | 
| 232 |  |  | \begin{equation} | 
| 233 |  |  | \rho = \delta \left( H(q^N, p^N) - E \right) \frac{1}{\Sigma (N, V, E)} | 
| 234 | xsun | 3354 | \label{Ineq:microcanonicalEnsemble} | 
| 235 | xsun | 3347 | \end{equation} | 
| 236 |  |  | where $\Sigma(N, V, E)$ is a normalization constant parameterized by | 
| 237 |  |  | $N$, the total number of particles, $V$, the total physical volume and | 
| 238 |  |  | $E$, the total energy. The physical meaning of $\Sigma(N, V, E)$ is | 
| 239 |  |  | the phase space volume accessible to a microcanonical system with | 
| 240 |  |  | energy $E$ evolving under Hamilton's equations. $H(q^N, p^N)$ is the | 
| 241 |  |  | Hamiltonian of the system. The Gibbs entropy is defined as | 
| 242 |  |  | \begin{equation} | 
| 243 |  |  | S = - k_B \int d \vec q~^N \int d \vec p~^N \rho \ln [C^N \rho], | 
| 244 | xsun | 3354 | \label{Ineq:gibbsEntropy} | 
| 245 | xsun | 3347 | \end{equation} | 
| 246 |  |  | where $k_B$ is the Boltzmann constant and $C^N$ is a number which | 
| 247 |  |  | makes the argument of $\ln$ dimensionless, in this case, it is the | 
| 248 |  |  | total phase space volume of one state. The entropy in microcanonical | 
| 249 |  |  | ensemble is given by | 
| 250 |  |  | \begin{equation} | 
| 251 |  |  | S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right). | 
| 252 | xsun | 3354 | \label{Ineq:entropy} | 
| 253 | xsun | 3347 | \end{equation} | 
| 254 |  |  | If the density distribution $\rho$ is given by | 
| 255 |  |  | \begin{equation} | 
| 256 |  |  | \rho = \frac{1}{Z_N}e^{-H(q^N, p^N) / k_B T}, | 
| 257 | xsun | 3354 | \label{Ineq:canonicalEnsemble} | 
| 258 | xsun | 3347 | \end{equation} | 
| 259 |  |  | the ensemble is known as the {\it canonical ensemble}. Here, | 
| 260 |  |  | \begin{equation} | 
| 261 |  |  | Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N  e^{-H(q^N, p^N) / k_B T}, | 
| 262 | xsun | 3354 | \label{Ineq:partitionFunction} | 
| 263 | xsun | 3347 | \end{equation} | 
| 264 |  |  | which is also known as {\it partition function}. $\Gamma$ indicates | 
| 265 |  |  | that the integral is over all the phase space. In the canonical | 
| 266 |  |  | ensemble, $N$, the total number of particles, $V$, total volume, and | 
| 267 |  |  | $T$, the temperature are constants. The systems with the lowest | 
| 268 |  |  | energies hold the largest population. According to maximum principle, | 
| 269 |  |  | the thermodynamics maximizes the entropy $S$, | 
| 270 |  |  | \begin{equation} | 
| 271 |  |  | \begin{array}{ccc} | 
| 272 |  |  | \delta S = 0 & \mathrm{and} & \delta^2 S < 0. | 
| 273 |  |  | \end{array} | 
| 274 | xsun | 3354 | \label{Ineq:maximumPrinciple} | 
| 275 | xsun | 3347 | \end{equation} | 
| 276 | xsun | 3354 | From Eq.~\ref{Ineq:maximumPrinciple} and two constrains of the canonical | 
| 277 | xsun | 3347 | ensemble, {\it i.e.}, total probability and average energy conserved, | 
| 278 |  |  | the partition function is calculated as | 
| 279 |  |  | \begin{equation} | 
| 280 |  |  | Z_N = e^{-A/k_B T}, | 
| 281 | xsun | 3354 | \label{Ineq:partitionFunctionWithFreeEnergy} | 
| 282 | xsun | 3347 | \end{equation} | 
| 283 |  |  | where $A$ is the Helmholtz free energy. The significance of | 
| 284 | xsun | 3354 | Eq.~\ref{Ineq:entropy} and~\ref{Ineq:partitionFunctionWithFreeEnergy} is | 
| 285 | xsun | 3347 | that they serve as a connection between macroscopic properties of the | 
| 286 |  |  | system and the distribution of the microscopic states. | 
| 287 |  |  |  | 
| 288 |  |  | There is an implicit assumption that our arguments are based on so | 
| 289 |  |  | far. A representive point in the phase space is equally to be found in | 
| 290 |  |  | any same extent of the regions. In other words, all energetically | 
| 291 |  |  | accessible states are represented equally, the probabilities to find | 
| 292 |  |  | the system in any of the accessible states is equal. This is called | 
| 293 |  |  | equal a {\it priori} probabilities. | 
| 294 |  |  |  | 
| 295 |  |  | \subsection{Statistical Average\label{In:ssec:average}} | 
| 296 |  |  | Given the density distribution $\rho$ in the phase space, the average | 
| 297 |  |  | of any quantity ($F(q^N, p^N$)) which depends on the coordinates | 
| 298 |  |  | ($q^N$) and the momenta ($p^N$) for all the systems in the ensemble | 
| 299 |  |  | can be calculated based on the definition shown by | 
| 300 | xsun | 3354 | Eq.~\ref{Ineq:statAverage1} | 
| 301 | xsun | 3347 | \begin{equation} | 
| 302 |  |  | \langle F(q^N, p^N, t) \rangle = \frac{\int d \vec q~^N \int d \vec p~^N | 
| 303 |  |  | F(q^N, p^N, t) \rho}{\int d \vec q~^N \int d \vec p~^N \rho}. | 
| 304 | xsun | 3354 | \label{Ineq:statAverage1} | 
| 305 | xsun | 3347 | \end{equation} | 
| 306 |  |  | Since the density distribution $\rho$ is normalized to unity, the mean | 
| 307 |  |  | value of $F(q^N, p^N)$ is simplified to | 
| 308 |  |  | \begin{equation} | 
| 309 |  |  | \langle F(q^N, p^N, t) \rangle = \int d \vec q~^N \int d \vec p~^N F(q^N, | 
| 310 |  |  | p^N, t) \rho, | 
| 311 | xsun | 3354 | \label{Ineq:statAverage2} | 
| 312 | xsun | 3347 | \end{equation} | 
| 313 |  |  | called {\it ensemble average}. However, the quantity is often averaged | 
| 314 |  |  | for a finite time in real experiments, | 
| 315 |  |  | \begin{equation} | 
| 316 |  |  | \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty} | 
| 317 |  |  | \frac{1}{T} \int_{t_0}^{t_0+T} F(q^N, p^N, t) dt. | 
| 318 | xsun | 3354 | \label{Ineq:timeAverage1} | 
| 319 | xsun | 3347 | \end{equation} | 
| 320 |  |  | Usually this time average is independent of $t_0$ in statistical | 
| 321 | xsun | 3354 | mechanics, so Eq.~\ref{Ineq:timeAverage1} becomes | 
| 322 | xsun | 3347 | \begin{equation} | 
| 323 |  |  | \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty} | 
| 324 |  |  | \frac{1}{T} \int_{0}^{T} F(q^N, p^N, t) dt | 
| 325 | xsun | 3354 | \label{Ineq:timeAverage2} | 
| 326 | xsun | 3347 | \end{equation} | 
| 327 |  |  | for an infinite time interval. | 
| 328 |  |  |  | 
| 329 |  |  | {\it ergodic hypothesis}, an important hypothesis from the statistical | 
| 330 |  |  | mechanics point of view, states that the system will eventually pass | 
| 331 |  |  | arbitrarily close to any point that is energetically accessible in | 
| 332 |  |  | phase space. Mathematically, this leads to | 
| 333 |  |  | \begin{equation} | 
| 334 |  |  | \langle F(q^N, p^N, t) \rangle = \langle F(q^N, p^N) \rangle_t. | 
| 335 | xsun | 3354 | \label{Ineq:ergodicity} | 
| 336 | xsun | 3347 | \end{equation} | 
| 337 | xsun | 3354 | Eq.~\ref{Ineq:ergodicity} validates the Monte Carlo method which we will | 
| 338 | xsun | 3347 | discuss in section~\ref{In:ssec:mc}. An ensemble average of a quantity | 
| 339 |  |  | can be related to the time average measured in the experiments. | 
| 340 |  |  |  | 
| 341 |  |  | \subsection{Correlation Function\label{In:ssec:corr}} | 
| 342 |  |  | Thermodynamic properties can be computed by equillibrium statistical | 
| 343 |  |  | mechanics. On the other hand, {\it Time correlation function} is a | 
| 344 |  |  | powerful method to understand the evolution of a dynamic system in | 
| 345 |  |  | non-equillibrium statistical mechanics. Imagine a property $A(q^N, | 
| 346 |  |  | p^N, t)$ as a function of coordinates $q^N$ and momenta $p^N$ has an | 
| 347 |  |  | intial value at $t_0$, at a later time $t_0 + \tau$ this value is | 
| 348 |  |  | changed. If $\tau$ is very small, the change of the value is minor, | 
| 349 |  |  | and the later value of $A(q^N, p^N, t_0 + | 
| 350 |  |  | \tau)$ is correlated to its initial value. Howere, when $\tau$ is large, | 
| 351 |  |  | this correlation is lost. The correlation function is a measurement of | 
| 352 |  |  | this relationship and is defined by~\cite{Berne90} | 
| 353 |  |  | \begin{equation} | 
| 354 |  |  | C(t) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarrow \infty} | 
| 355 |  |  | \frac{1}{T} \int_{0}^{T} dt A(t) A(t + \tau). | 
| 356 | xsun | 3354 | \label{Ineq:autocorrelationFunction} | 
| 357 | xsun | 3347 | \end{equation} | 
| 358 | xsun | 3354 | Eq.~\ref{Ineq:autocorrelationFunction} is the correlation function of a | 
| 359 | xsun | 3347 | single variable, called {\it autocorrelation function}. The defination | 
| 360 |  |  | of the correlation function for two different variables is similar to | 
| 361 |  |  | that of autocorrelation function, which is | 
| 362 |  |  | \begin{equation} | 
| 363 |  |  | C(t) = \langle A(0)B(\tau) \rangle = \lim_{T \rightarrow \infty} | 
| 364 |  |  | \frac{1}{T} \int_{0}^{T} dt A(t) B(t + \tau), | 
| 365 | xsun | 3354 | \label{Ineq:crosscorrelationFunction} | 
| 366 | xsun | 3347 | \end{equation} | 
| 367 |  |  | and called {\it cross correlation function}. | 
| 368 |  |  |  | 
| 369 | xsun | 3354 | In section~\ref{In:ssec:average} we know from Eq.~\ref{Ineq:ergodicity} | 
| 370 | xsun | 3347 | the relationship between time average and ensemble average. We can put | 
| 371 |  |  | the correlation function in a classical mechanics form, | 
| 372 |  |  | \begin{equation} | 
| 373 |  |  | C(t) = \langle A(0)A(\tau) \rangle = \int d \vec q~^N \int d \vec p~^N A(t) A(t + \tau) \rho(q, p) | 
| 374 | xsun | 3354 | \label{Ineq:autocorrelationFunctionCM} | 
| 375 | xsun | 3347 | \end{equation} | 
| 376 |  |  | and | 
| 377 |  |  | \begin{equation} | 
| 378 |  |  | C(t) = \langle A(0)B(\tau) \rangle = \int d \vec q~^N \int d \vec p~^N A(t) B(t + \tau) | 
| 379 |  |  | \rho(q, p) | 
| 380 | xsun | 3354 | \label{Ineq:crosscorrelationFunctionCM} | 
| 381 | xsun | 3347 | \end{equation} | 
| 382 |  |  | as autocorrelation function and cross correlation function | 
| 383 |  |  | respectively. $\rho(q, p)$ is the density distribution at equillibrium | 
| 384 |  |  | in phase space. In many cases, the correlation function decay is a | 
| 385 |  |  | single exponential | 
| 386 |  |  | \begin{equation} | 
| 387 |  |  | C(t) \sim e^{-t / \tau_r}, | 
| 388 | xsun | 3354 | \label{Ineq:relaxation} | 
| 389 | xsun | 3347 | \end{equation} | 
| 390 |  |  | where $\tau_r$ is known as relaxation time which discribes the rate of | 
| 391 |  |  | the decay. | 
| 392 |  |  |  | 
| 393 |  |  | \section{Methodolody\label{In:sec:method}} | 
| 394 |  |  | The simulations performed in this dissertation are branched into two | 
| 395 |  |  | main catalog, Monte Carlo and Molecular Dynamics. There are two main | 
| 396 |  |  | difference between Monte Carlo and Molecular Dynamics simulations. One | 
| 397 |  |  | is that the Monte Carlo simulation is time independent, and Molecular | 
| 398 |  |  | Dynamics simulation is time involved. Another dissimilar is that the | 
| 399 |  |  | Monte Carlo is a stochastic process, the configuration of the system | 
| 400 |  |  | is not determinated by its past, however, using Moleuclar Dynamics, | 
| 401 |  |  | the system is propagated by Newton's equation of motion, the | 
| 402 |  |  | trajectory of the system evolved in the phase space is determined. A | 
| 403 |  |  | brief introduction of the two algorithms are given in | 
| 404 |  |  | section~\ref{In:ssec:mc} and~\ref{In:ssec:md}. An extension of the | 
| 405 |  |  | Molecular Dynamics, Langevin Dynamics, is introduced by | 
| 406 |  |  | section~\ref{In:ssec:ld}. | 
| 407 |  |  |  | 
| 408 |  |  | \subsection{Monte Carlo\label{In:ssec:mc}} | 
| 409 |  |  | Monte Carlo algorithm was first introduced by Metropolis {\it et | 
| 410 |  |  | al.}.~\cite{Metropolis53} Basic Monte Carlo algorithm is usually | 
| 411 |  |  | applied to the canonical ensemble, a Boltzmann-weighted ensemble, in | 
| 412 |  |  | which the $N$, the total number of particles, $V$, total volume, $T$, | 
| 413 |  |  | temperature are constants. The average energy is given by substituding | 
| 414 | xsun | 3354 | Eq.~\ref{Ineq:canonicalEnsemble} into Eq.~\ref{Ineq:statAverage2}, | 
| 415 | xsun | 3347 | \begin{equation} | 
| 416 |  |  | \langle E \rangle = \frac{1}{Z_N} \int d \vec q~^N \int d \vec p~^N E e^{-H(q^N, p^N) / k_B T}. | 
| 417 | xsun | 3354 | \label{Ineq:energyofCanonicalEnsemble} | 
| 418 | xsun | 3347 | \end{equation} | 
| 419 |  |  | So are the other properties of the system. The Hamiltonian is the | 
| 420 |  |  | summation of Kinetic energy $K(p^N)$ as a function of momenta and | 
| 421 |  |  | Potential energy $U(q^N)$ as a function of positions, | 
| 422 |  |  | \begin{equation} | 
| 423 |  |  | H(q^N, p^N) = K(p^N) + U(q^N). | 
| 424 | xsun | 3354 | \label{Ineq:hamiltonian} | 
| 425 | xsun | 3347 | \end{equation} | 
| 426 |  |  | If the property $A$ is only a function of position ($ A = A(q^N)$), | 
| 427 |  |  | the mean value of $A$ is reduced to | 
| 428 |  |  | \begin{equation} | 
| 429 |  |  | \langle A \rangle = \frac{\int d \vec q~^N \int d \vec p~^N A e^{-U(q^N) / k_B T}}{\int d \vec q~^N \int d \vec p~^N e^{-U(q^N) / k_B T}}, | 
| 430 | xsun | 3354 | \label{Ineq:configurationIntegral} | 
| 431 | xsun | 3347 | \end{equation} | 
| 432 |  |  | The kinetic energy $K(p^N)$ is factored out in | 
| 433 | xsun | 3354 | Eq.~\ref{Ineq:configurationIntegral}. $\langle A | 
| 434 | xsun | 3347 | \rangle$ is a configuration integral now, and the | 
| 435 | xsun | 3354 | Eq.~\ref{Ineq:configurationIntegral} is equivalent to | 
| 436 | xsun | 3347 | \begin{equation} | 
| 437 |  |  | \langle A \rangle = \int d \vec q~^N A \rho(q^N). | 
| 438 | xsun | 3354 | \label{Ineq:configurationAve} | 
| 439 | xsun | 3347 | \end{equation} | 
| 440 |  |  |  | 
| 441 |  |  | In a Monte Carlo simulation of canonical ensemble, the probability of | 
| 442 |  |  | the system being in a state $s$ is $\rho_s$, the change of this | 
| 443 |  |  | probability with time is given by | 
| 444 |  |  | \begin{equation} | 
| 445 |  |  | \frac{d \rho_s}{dt} = \sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ], | 
| 446 | xsun | 3354 | \label{Ineq:timeChangeofProb} | 
| 447 | xsun | 3347 | \end{equation} | 
| 448 |  |  | where $w_{ss'}$ is the tansition probability of going from state $s$ | 
| 449 |  |  | to state $s'$. Since $\rho_s$ is independent of time at equilibrium, | 
| 450 |  |  | \begin{equation} | 
| 451 |  |  | \frac{d \rho_{s}^{equilibrium}}{dt} = 0, | 
| 452 | xsun | 3354 | \label{Ineq:equiProb} | 
| 453 | xsun | 3347 | \end{equation} | 
| 454 |  |  | which means $\sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ]$ is $0$ | 
| 455 |  |  | for all $s'$. So | 
| 456 |  |  | \begin{equation} | 
| 457 |  |  | \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = \frac{w_{s's}}{w_{ss'}}. | 
| 458 | xsun | 3354 | \label{Ineq:relationshipofRhoandW} | 
| 459 | xsun | 3347 | \end{equation} | 
| 460 |  |  | If | 
| 461 |  |  | \begin{equation} | 
| 462 |  |  | \frac{w_{s's}}{w_{ss'}} = e^{-(U_s - U_{s'}) / k_B T}, | 
| 463 | xsun | 3354 | \label{Ineq:conditionforBoltzmannStatistics} | 
| 464 | xsun | 3347 | \end{equation} | 
| 465 |  |  | then | 
| 466 |  |  | \begin{equation} | 
| 467 |  |  | \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = e^{-(U_s - U_{s'}) / k_B T}. | 
| 468 | xsun | 3354 | \label{Ineq:satisfyofBoltzmannStatistics} | 
| 469 | xsun | 3347 | \end{equation} | 
| 470 | xsun | 3354 | Eq.~\ref{Ineq:satisfyofBoltzmannStatistics} implies that | 
| 471 | xsun | 3347 | $\rho^{equilibrium}$ satisfies Boltzmann statistics. An algorithm, | 
| 472 |  |  | shows how Monte Carlo simulation generates a transition probability | 
| 473 | xsun | 3354 | governed by \ref{Ineq:conditionforBoltzmannStatistics}, is schemed as | 
| 474 | xsun | 3347 | \begin{enumerate} | 
| 475 | xsun | 3354 | \item\label{Initm:oldEnergy} Choose an particle randomly, calculate the energy. | 
| 476 |  |  | \item\label{Initm:newEnergy} Make a random displacement for particle, | 
| 477 | xsun | 3347 | calculate the new energy. | 
| 478 |  |  | \begin{itemize} | 
| 479 | xsun | 3354 | \item Keep the new configuration and return to step~\ref{Initm:oldEnergy} if energy | 
| 480 | xsun | 3347 | goes down. | 
| 481 |  |  | \item Pick a random number between $[0,1]$ if energy goes up. | 
| 482 |  |  | \begin{itemize} | 
| 483 |  |  | \item Keep the new configuration and return to | 
| 484 | xsun | 3354 | step~\ref{Initm:oldEnergy} if the random number smaller than | 
| 485 | xsun | 3347 | $e^{-(U_{new} - U_{old})} / k_B T$. | 
| 486 |  |  | \item Keep the old configuration and return to | 
| 487 | xsun | 3354 | step~\ref{Initm:oldEnergy} if the random number larger than | 
| 488 | xsun | 3347 | $e^{-(U_{new} - U_{old})} / k_B T$. | 
| 489 |  |  | \end{itemize} | 
| 490 |  |  | \end{itemize} | 
| 491 | xsun | 3354 | \item\label{Initm:accumulateAvg} Accumulate the average after it converges. | 
| 492 | xsun | 3347 | \end{enumerate} | 
| 493 |  |  | It is important to notice that the old configuration has to be sampled | 
| 494 |  |  | again if it is kept. | 
| 495 |  |  |  | 
| 496 |  |  | \subsection{Molecular Dynamics\label{In:ssec:md}} | 
| 497 |  |  | Although some of properites of the system can be calculated from the | 
| 498 |  |  | ensemble average in Monte Carlo simulations, due to the nature of | 
| 499 |  |  | lacking in the time dependence, it is impossible to gain information | 
| 500 |  |  | of those dynamic properties from Monte Carlo simulations. Molecular | 
| 501 |  |  | Dynamics is a measurement of the evolution of the positions and | 
| 502 |  |  | momenta of the particles in the system. The evolution of the system | 
| 503 |  |  | obeys laws of classical mechanics, in most situations, there is no | 
| 504 |  |  | need for the count of the quantum effects. For a real experiment, the | 
| 505 |  |  | instantaneous positions and momenta of the particles in the system are | 
| 506 |  |  | neither important nor measurable, the observable quantities are | 
| 507 |  |  | usually a average value for a finite time interval. These quantities | 
| 508 |  |  | are expressed as a function of positions and momenta in Melecular | 
| 509 |  |  | Dynamics simulations. Like the thermal temperature of the system is | 
| 510 |  |  | defined by | 
| 511 |  |  | \begin{equation} | 
| 512 |  |  | \frac{1}{2} k_B T = \langle \frac{1}{2} m v_\alpha \rangle, | 
| 513 | xsun | 3354 | \label{Ineq:temperature} | 
| 514 | xsun | 3347 | \end{equation} | 
| 515 |  |  | here $m$ is the mass of the particle and $v_\alpha$ is the $\alpha$ | 
| 516 |  |  | component of the velocity of the particle. The right side of | 
| 517 | xsun | 3354 | Eq.~\ref{Ineq:temperature} is the average kinetic energy of the | 
| 518 | xsun | 3347 | system. A simple Molecular Dynamics simulation scheme | 
| 519 |  |  | is:~\cite{Frenkel1996} | 
| 520 |  |  | \begin{enumerate} | 
| 521 | xsun | 3354 | \item\label{Initm:initialize} Assign the initial positions and momenta | 
| 522 | xsun | 3347 | for the particles in the system. | 
| 523 | xsun | 3354 | \item\label{Initm:calcForce} Calculate the forces. | 
| 524 |  |  | \item\label{Initm:equationofMotion} Integrate the equation of motion. | 
| 525 | xsun | 3347 | \begin{itemize} | 
| 526 | xsun | 3354 | \item Return to step~\ref{Initm:calcForce} if the equillibrium is | 
| 527 | xsun | 3347 | not achieved. | 
| 528 | xsun | 3354 | \item Go to step~\ref{Initm:calcAvg} if the equillibrium is | 
| 529 | xsun | 3347 | achieved. | 
| 530 |  |  | \end{itemize} | 
| 531 | xsun | 3354 | \item\label{Initm:calcAvg} Compute the quantities we are interested in. | 
| 532 | xsun | 3347 | \end{enumerate} | 
| 533 |  |  | The initial positions of the particles are chosen as that there is no | 
| 534 |  |  | overlap for the particles. The initial velocities at first are | 
| 535 |  |  | distributed randomly to the particles, and then shifted to make the | 
| 536 |  |  | momentum of the system $0$, at last scaled to the desired temperature | 
| 537 | xsun | 3354 | of the simulation according Eq.~\ref{Ineq:temperature}. | 
| 538 | xsun | 3347 |  | 
| 539 | xsun | 3354 | The core of Molecular Dynamics simulations is step~\ref{Initm:calcForce} | 
| 540 |  |  | and~\ref{Initm:equationofMotion}. The calculation of the forces are | 
| 541 | xsun | 3347 | often involved numerous effort, this is the most time consuming step | 
| 542 |  |  | in the Molecular Dynamics scheme. The evaluation of the forces is | 
| 543 |  |  | followed by | 
| 544 |  |  | \begin{equation} | 
| 545 |  |  | f(q) = - \frac{\partial U(q)}{\partial q}, | 
| 546 | xsun | 3354 | \label{Ineq:force} | 
| 547 | xsun | 3347 | \end{equation} | 
| 548 |  |  | $U(q)$ is the potential of the system. Once the forces computed, are | 
| 549 |  |  | the positions and velocities updated by integrating Newton's equation | 
| 550 |  |  | of motion, | 
| 551 |  |  | \begin{equation} | 
| 552 |  |  | f(q) = \frac{dp}{dt} = \frac{m dv}{dt}. | 
| 553 | xsun | 3354 | \label{Ineq:newton} | 
| 554 | xsun | 3347 | \end{equation} | 
| 555 |  |  | Here is an example of integrating algorithms, Verlet algorithm, which | 
| 556 |  |  | is one of the best algorithms to integrate Newton's equation of | 
| 557 |  |  | motion. The Taylor expension of the position at time $t$ is | 
| 558 |  |  | \begin{equation} | 
| 559 |  |  | q(t+\Delta t)= q(t) + v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 + | 
| 560 |  |  | \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} + | 
| 561 |  |  | \mathcal{O}(\Delta t^4) | 
| 562 | xsun | 3354 | \label{Ineq:verletFuture} | 
| 563 | xsun | 3347 | \end{equation} | 
| 564 |  |  | for a later time $t+\Delta t$, and | 
| 565 |  |  | \begin{equation} | 
| 566 |  |  | q(t-\Delta t)= q(t) - v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 - | 
| 567 |  |  | \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} + | 
| 568 |  |  | \mathcal{O}(\Delta t^4) , | 
| 569 | xsun | 3354 | \label{Ineq:verletPrevious} | 
| 570 | xsun | 3347 | \end{equation} | 
| 571 |  |  | for a previous time $t-\Delta t$. The summation of the | 
| 572 | xsun | 3354 | Eq.~\ref{Ineq:verletFuture} and~\ref{Ineq:verletPrevious} gives | 
| 573 | xsun | 3347 | \begin{equation} | 
| 574 |  |  | q(t+\Delta t)+q(t-\Delta t) = | 
| 575 |  |  | 2q(t) + \frac{f(t)}{m}\Delta t^2 + \mathcal{O}(\Delta t^4), | 
| 576 | xsun | 3354 | \label{Ineq:verletSum} | 
| 577 | xsun | 3347 | \end{equation} | 
| 578 |  |  | so, the new position can be expressed as | 
| 579 |  |  | \begin{equation} | 
| 580 |  |  | q(t+\Delta t) \approx | 
| 581 |  |  | 2q(t) - q(t-\Delta t) + \frac{f(t)}{m}\Delta t^2. | 
| 582 | xsun | 3354 | \label{Ineq:newPosition} | 
| 583 | xsun | 3347 | \end{equation} | 
| 584 |  |  | The higher order of the $\Delta t$ is omitted. | 
| 585 |  |  |  | 
| 586 |  |  | Numerous technics and tricks are applied to Molecular Dynamics | 
| 587 |  |  | simulation to gain more efficiency or more accuracy. The simulation | 
| 588 |  |  | engine used in this dissertation for the Molecular Dynamics | 
| 589 |  |  | simulations is {\sc oopse}, more details of the algorithms and | 
| 590 |  |  | technics can be found in~\cite{Meineke2005}. | 
| 591 |  |  |  | 
| 592 |  |  | \subsection{Langevin Dynamics\label{In:ssec:ld}} | 
| 593 |  |  | In many cases, the properites of the solvent in a system, like the | 
| 594 |  |  | lipid-water system studied in this dissertation, are less important to | 
| 595 |  |  | the researchers. However, the major computational expense is spent on | 
| 596 |  |  | the solvent in the Molecular Dynamics simulations because of the large | 
| 597 |  |  | number of the solvent molecules compared to that of solute molecules, | 
| 598 |  |  | the ratio of the number of lipid molecules to the number of water | 
| 599 |  |  | molecules is $1:25$ in our lipid-water system. The efficiency of the | 
| 600 |  |  | Molecular Dynamics simulations is greatly reduced. | 
| 601 |  |  |  | 
| 602 |  |  | As an extension of the Molecular Dynamics simulations, the Langevin | 
| 603 |  |  | Dynamics seeks a way to avoid integrating equation of motion for | 
| 604 |  |  | solvent particles without losing the Brownian properites of solute | 
| 605 |  |  | particles. A common approximation is that the coupling of the solute | 
| 606 |  |  | and solvent is expressed as a set of harmonic oscillators. So the | 
| 607 |  |  | Hamiltonian of such a system is written as | 
| 608 |  |  | \begin{equation} | 
| 609 |  |  | H = \frac{p^2}{2m} + U(q) + H_B + \Delta U(q), | 
| 610 | xsun | 3354 | \label{Ineq:hamiltonianofCoupling} | 
| 611 | xsun | 3347 | \end{equation} | 
| 612 |  |  | where $H_B$ is the Hamiltonian of the bath which equals to | 
| 613 |  |  | \begin{equation} | 
| 614 |  |  | H_B = \sum_{\alpha = 1}^{N} \left\{ \frac{p_\alpha^2}{2m_\alpha} + | 
| 615 |  |  | \frac{1}{2} m_\alpha \omega_\alpha^2 q_\alpha^2\right\}, | 
| 616 | xsun | 3354 | \label{Ineq:hamiltonianofBath} | 
| 617 | xsun | 3347 | \end{equation} | 
| 618 |  |  | $\alpha$ is all the degree of freedoms of the bath, $\omega$ is the | 
| 619 |  |  | bath frequency, and $\Delta U(q)$ is the bilinear coupling given by | 
| 620 |  |  | \begin{equation} | 
| 621 |  |  | \Delta U = -\sum_{\alpha = 1}^{N} g_\alpha q_\alpha q, | 
| 622 | xsun | 3354 | \label{Ineq:systemBathCoupling} | 
| 623 | xsun | 3347 | \end{equation} | 
| 624 |  |  | where $g$ is the coupling constant. By solving the Hamilton's equation | 
| 625 |  |  | of motion, the {\it Generalized Langevin Equation} for this system is | 
| 626 |  |  | derived to | 
| 627 |  |  | \begin{equation} | 
| 628 |  |  | m \ddot q = -\frac{\partial W(q)}{\partial q} - \int_0^t \xi(t) \dot q(t-t')dt' + R(t), | 
| 629 | xsun | 3354 | \label{Ineq:gle} | 
| 630 | xsun | 3347 | \end{equation} | 
| 631 |  |  | with mean force, | 
| 632 |  |  | \begin{equation} | 
| 633 |  |  | W(q) = U(q) - \sum_{\alpha = 1}^N \frac{g_\alpha^2}{2m_\alpha | 
| 634 |  |  | \omega_\alpha^2}q^2, | 
| 635 | xsun | 3354 | \label{Ineq:meanForce} | 
| 636 | xsun | 3347 | \end{equation} | 
| 637 |  |  | being only a dependence of coordinates of the solute particles, {\it | 
| 638 |  |  | friction kernel}, | 
| 639 |  |  | \begin{equation} | 
| 640 |  |  | \xi(t) = \sum_{\alpha = 1}^N \frac{-g_\alpha}{m_\alpha | 
| 641 |  |  | \omega_\alpha} \cos(\omega_\alpha t), | 
| 642 | xsun | 3354 | \label{Ineq:xiforGLE} | 
| 643 | xsun | 3347 | \end{equation} | 
| 644 |  |  | and the random force, | 
| 645 |  |  | \begin{equation} | 
| 646 |  |  | R(t) = \sum_{\alpha = 1}^N \left( g_\alpha q_\alpha(0)-\frac{g_\alpha}{m_\alpha | 
| 647 |  |  | \omega_\alpha^2}q(0)\right) \cos(\omega_\alpha t) + \frac{\dot | 
| 648 |  |  | q_\alpha(0)}{\omega_\alpha} \sin(\omega_\alpha t), | 
| 649 | xsun | 3354 | \label{Ineq:randomForceforGLE} | 
| 650 | xsun | 3347 | \end{equation} | 
| 651 |  |  | as only a dependence of the initial conditions. The relationship of | 
| 652 |  |  | friction kernel $\xi(t)$ and random force $R(t)$ is given by | 
| 653 |  |  | \begin{equation} | 
| 654 |  |  | \xi(t) = \frac{1}{k_B T} \langle R(t)R(0) \rangle | 
| 655 | xsun | 3354 | \label{Ineq:relationshipofXiandR} | 
| 656 | xsun | 3347 | \end{equation} | 
| 657 |  |  | from their definitions. In Langevin limit, the friction is treated | 
| 658 |  |  | static, which means | 
| 659 |  |  | \begin{equation} | 
| 660 |  |  | \xi(t) = 2 \xi_0 \delta(t). | 
| 661 | xsun | 3354 | \label{Ineq:xiofStaticFriction} | 
| 662 | xsun | 3347 | \end{equation} | 
| 663 | xsun | 3354 | After substitude $\xi(t)$ into Eq.~\ref{Ineq:gle} with | 
| 664 |  |  | Eq.~\ref{Ineq:xiofStaticFriction}, {\it Langevin Equation} is extracted | 
| 665 | xsun | 3347 | to | 
| 666 |  |  | \begin{equation} | 
| 667 |  |  | m \ddot q = -\frac{\partial U(q)}{\partial q} - \xi \dot q(t) + R(t). | 
| 668 | xsun | 3354 | \label{Ineq:langevinEquation} | 
| 669 | xsun | 3347 | \end{equation} | 
| 670 |  |  | The applying of Langevin Equation to dynamic simulations is discussed | 
| 671 |  |  | in Ch.~\ref{chap:ld}. |