| 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 | xsun | 3375 | membranes. The chemical structure of phospholipids includes a head | 
| 7 |  |  | group with a large dipole moment which is due to the large charge | 
| 8 |  |  | separation between phosphate and amino alcohol, and a nonpolar tail | 
| 9 |  |  | that contains fatty acid chains. Depending on the specific alcohol | 
| 10 |  |  | which the phosphate and fatty acid chains are esterified to, the | 
| 11 |  |  | phospholipids are divided into glycerophospholipids and | 
| 12 |  |  | sphingophospholipids.~\cite{Cevc80} The chemical structures are shown | 
| 13 |  |  | in figure~\ref{Infig:lipid}. | 
| 14 | xsun | 3374 | \begin{figure} | 
| 15 |  |  | \centering | 
| 16 |  |  | \includegraphics[width=\linewidth]{./figures/inLipid.pdf} | 
| 17 | xsun | 3383 | \caption[The chemical structure of lipids]{The chemical structure of | 
| 18 |  |  | glycerophospholipids (left) and sphingophospholipids | 
| 19 |  |  | (right).\cite{Cevc80}} | 
| 20 | xsun | 3374 | \label{Infig:lipid} | 
| 21 |  |  | \end{figure} | 
| 22 | xsun | 3375 | Glycerophospholipids are the dominant phospholipids in biological | 
| 23 |  |  | membranes. The type of glycerophospholipid depends on the identity of | 
| 24 |  |  | the X group, and the chains. For example, if X is choline | 
| 25 |  |  | [(CH$_3$)$_3$N$^+$CH$_2$CH$_2$OH], the lipids are known as | 
| 26 |  |  | phosphatidylcholine (PC), or if X is ethanolamine | 
| 27 |  |  | [H$_3$N$^+$CH$_2$CH$_2$OH], the lipids are known as | 
| 28 |  |  | phosphatidyethanolamine (PE). Table~\ref{Intab:pc} listed a number | 
| 29 |  |  | types of phosphatidycholine with different fatty acids as the lipid | 
| 30 |  |  | chains. | 
| 31 | xsun | 3374 | \begin{table*} | 
| 32 |  |  | \begin{minipage}{\linewidth} | 
| 33 |  |  | \begin{center} | 
| 34 | xsun | 3383 | \caption{A NUMBER TYPES OF PHOSPHATIDYCHOLINE} | 
| 35 | xsun | 3374 | \begin{tabular}{lll} | 
| 36 |  |  | \hline | 
| 37 |  |  | & Fatty acid & Generic Name \\ | 
| 38 |  |  | \hline | 
| 39 |  |  | \textcolor{red}{DMPC} & Myristic: CH$_3$(CH$_2$)$_{12}$COOH & | 
| 40 |  |  | \textcolor{red}{D}i\textcolor{red}{M}yristoyl\textcolor{red}{P}hosphatidyl\textcolor{red}{C}holine \\ | 
| 41 |  |  | \textcolor{red}{DPPC} & Palmitic: CH$_3$(CH$_2$)$_{14}$COOH & \textcolor{red}{D}i\textcolor{red}{P}almtoyl\textcolor{red}{P}hosphatidyl\textcolor{red}{C}holine | 
| 42 |  |  | \\ | 
| 43 |  |  | \textcolor{red}{DSPC} & Stearic: CH$_3$(CH$_2$)$_{16}$COOH & \textcolor{red}{D}i\textcolor{red}{S}tearoyl\textcolor{red}{P}hosphatidyl\textcolor{red}{C}holine \\ | 
| 44 |  |  | \end{tabular} | 
| 45 |  |  | \label{Intab:pc} | 
| 46 |  |  | \end{center} | 
| 47 |  |  | \end{minipage} | 
| 48 |  |  | \end{table*} | 
| 49 | xsun | 3377 | When dispersed in water, lipids self assemble into a number of | 
| 50 | xsun | 3374 | topologically distinct bilayer structures. The phase behavior of lipid | 
| 51 |  |  | bilayers has been explored experimentally~\cite{Cevc80}, however, a | 
| 52 |  |  | complete understanding of the mechanism and driving forces behind the | 
| 53 |  |  | various phases has not been achieved. | 
| 54 | xsun | 3347 |  | 
| 55 | xsun | 3375 | \subsection{The Ripple Phase\label{In:ssec:ripple}} | 
| 56 | xsun | 3365 | The $P_{\beta'}$ {\it ripple phase} of lipid bilayers, named from the | 
| 57 | xsun | 3347 | periodic buckling of the membrane, is an intermediate phase which is | 
| 58 |  |  | developed either from heating the gel phase $L_{\beta'}$ or cooling | 
| 59 | xsun | 3375 | the fluid phase $L_\alpha$. A sketch of the phases is shown in | 
| 60 | xsun | 3374 | figure~\ref{Infig:phaseDiagram}.~\cite{Cevc80} | 
| 61 | xsun | 3372 | \begin{figure} | 
| 62 |  |  | \centering | 
| 63 |  |  | \includegraphics[width=\linewidth]{./figures/inPhaseDiagram.pdf} | 
| 64 | xsun | 3383 | \caption[Phases of PC lipid bilayers]{Phases of PC lipid | 
| 65 |  |  | bilayers. With increasing temperature, phosphotidylcholine (PC) | 
| 66 |  |  | bilayers can go through $L_{\beta'} \rightarrow P_{\beta'}$ (gel | 
| 67 |  |  | $\rightarrow$ ripple) and $P_{\beta'} \rightarrow L_\alpha$ (ripple | 
| 68 |  |  | $\rightarrow$ fluid) phase transitions.~\cite{Cevc80}} | 
| 69 | xsun | 3372 | \label{Infig:phaseDiagram} | 
| 70 |  |  | \end{figure} | 
| 71 | xsun | 3375 | Most structural information about the ripple phase has been obtained | 
| 72 |  |  | by X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture | 
| 73 | xsun | 3372 | electron microscopy (FFEM).~\cite{Copeland80,Meyer96} The X-ray | 
| 74 |  |  | diffraction work by Katsaras {\it et al.} showed that a rich phase | 
| 75 |  |  | diagram exhibiting both {\it asymmetric} and {\it symmetric} ripples | 
| 76 |  |  | is possible for lecithin bilayers.\cite{Katsaras00} Recently, | 
| 77 |  |  | Kaasgaard {\it et al.} used atomic force microscopy (AFM) to observe | 
| 78 |  |  | ripple phase morphology in bilayers supported on | 
| 79 |  |  | mica.~\cite{Kaasgaard03} | 
| 80 |  |  | \begin{figure} | 
| 81 |  |  | \centering | 
| 82 |  |  | \includegraphics[width=\linewidth]{./figures/inRipple.pdf} | 
| 83 | xsun | 3383 | \caption[Experimental observations of the riple phase]{Experimental | 
| 84 |  |  | observations of the riple phase. The top image is an electrostatic | 
| 85 |  |  | density map obtained by Sun {\it et al.} using X-ray | 
| 86 |  |  | diffraction~\cite{Sun96}.  The lower figures are the surface topology | 
| 87 |  |  | of various ripple domains in bilayers supported in mica. The AFM | 
| 88 |  |  | images were observed by Kaasgaard {\it et al.}.~\cite{Kaasgaard03}} | 
| 89 | xsun | 3372 | \label{Infig:ripple} | 
| 90 |  |  | \end{figure} | 
| 91 | xsun | 3375 | Figure~\ref{Infig:ripple} shows the ripple phase oberved by both X-ray | 
| 92 | xsun | 3372 | diffraction and AFM. The experimental results provide strong support | 
| 93 |  |  | for a 2-dimensional triangular packing lattice of the lipid molecules | 
| 94 |  |  | within the ripple phase.  This is a notable change from the observed | 
| 95 | xsun | 3374 | lipid packing within the gel phase,~\cite{Cevc80} although Tenchov | 
| 96 | xsun | 3375 | {\it et al.} have recently observed near-triangular packing in some | 
| 97 | xsun | 3372 | phosphatidylcholine (PC) gel phases.~\cite{Tenchov2001} However, the | 
| 98 |  |  | physical mechanism for the formation of the ripple phase has never | 
| 99 |  |  | been explained and the microscopic structure of the ripple phase has | 
| 100 |  |  | never been elucidated by experiments. Computational simulation is a | 
| 101 | xsun | 3383 | very good tool to study the microscopic properties for a | 
| 102 | xsun | 3375 | system. However, the large length scale of the ripple structures and | 
| 103 |  |  | the long time required for the formation of the ripples are crucial | 
| 104 |  |  | obstacles to performing the actual work. The principal ideas explored | 
| 105 |  |  | in this dissertation are attempts to break the computational task up | 
| 106 |  |  | by | 
| 107 | xsun | 3347 | \begin{itemize} | 
| 108 | xsun | 3365 | \item Simplifying the lipid model. | 
| 109 | xsun | 3375 | \item Improving the algorithm for integrating the equations of motion. | 
| 110 | xsun | 3347 | \end{itemize} | 
| 111 | xsun | 3365 | In chapters~\ref{chap:mc} and~\ref{chap:md}, we use a simple point | 
| 112 | xsun | 3375 | dipole spin model and a coarse-grained molecular scale model to | 
| 113 | xsun | 3372 | perform the Monte Carlo and Molecular Dynamics simulations | 
| 114 |  |  | respectively, and in chapter~\ref{chap:ld}, we develop a Langevin | 
| 115 |  |  | Dynamics algorithm which excludes the explicit solvent to improve the | 
| 116 |  |  | efficiency of the simulations. | 
| 117 | xsun | 3347 |  | 
| 118 | xsun | 3375 | \subsection{Lattice Models\label{In:ssec:model}} | 
| 119 |  |  | The gel-like characteristic (laterally immobile molecules) exhibited | 
| 120 | xsun | 3372 | by the ripple phase makes it feasible to apply a lattice model to | 
| 121 |  |  | study the system. The popular $2$ dimensional lattice models, {\it | 
| 122 |  |  | i.e.}, the Ising, $X-Y$, and Heisenberg models, show {\it frustration} | 
| 123 |  |  | on triangular lattices. The Hamiltonians of these systems are given by | 
| 124 | xsun | 3365 | \begin{equation} | 
| 125 |  |  | H = | 
| 126 |  |  | \begin{cases} | 
| 127 |  |  | -J \sum_n \sum_{n'} s_n s_n' & \text{Ising}, \\ | 
| 128 |  |  | -J \sum_n \sum_{n'} \vec s_n \cdot \vec s_{n'} & \text{$X-Y$ and | 
| 129 |  |  | Heisenberg}, | 
| 130 |  |  | \end{cases} | 
| 131 |  |  | \end{equation} | 
| 132 |  |  | where $J$ has non zero value only when spins $s_n$ ($\vec s_n$) and | 
| 133 | xsun | 3372 | $s_{n'}$ ($\vec s_{n'}$) are nearest neighbors. When $J > 0$, spins | 
| 134 |  |  | prefer an aligned structure, and if $J < 0$, spins prefer an | 
| 135 |  |  | anti-aligned structure. | 
| 136 | xsun | 3371 |  | 
| 137 | xsun | 3347 | \begin{figure} | 
| 138 |  |  | \centering | 
| 139 | xsun | 3371 | \includegraphics[width=3in]{./figures/inFrustration.pdf} | 
| 140 | xsun | 3383 | \caption[Frustration on triangular lattice]{Frustration on triangular | 
| 141 |  |  | lattice, the spins and dipoles are represented by arrows. The multiple | 
| 142 |  |  | local minima of energy states induce frustration for spins and dipoles | 
| 143 |  |  | resulting in disordered low-temperature phases.} | 
| 144 | xsun | 3354 | \label{Infig:frustration} | 
| 145 | xsun | 3347 | \end{figure} | 
| 146 | xsun | 3375 | The spins in figure~\ref{Infig:frustration} illustrate frustration for | 
| 147 |  |  | $J < 0$ on a triangular lattice. There are multiple local minima | 
| 148 |  |  | energy states which are independent of the direction of the spin on | 
| 149 |  |  | top of the triangle, therefore infinite possibilities for orienting | 
| 150 |  |  | large numbers spins. This induces what is known as ``complete regular | 
| 151 |  |  | frustration'' which leads to disordered low temperature phases. This | 
| 152 |  |  | behavior extends to dipoles on a triangular lattices, which are shown | 
| 153 |  |  | in the lower portion of figure~\ref{Infig:frustration}. In this case, | 
| 154 |  |  | dipole-aligned structures are energetically favorable, however, at low | 
| 155 |  |  | temperature, vortices are easily formed, and, this results in multiple | 
| 156 |  |  | local minima of energy states for a central dipole. The dipole on the | 
| 157 |  |  | center of the hexagonal lattice is therefore frustrated. | 
| 158 | xsun | 3347 |  | 
| 159 | xsun | 3372 | The lack of translational degrees of freedom in lattice models | 
| 160 |  |  | prevents their utilization in models for surface buckling. In | 
| 161 |  |  | chapter~\ref{chap:mc}, a modified lattice model is introduced to | 
| 162 |  |  | tackle this specific situation. | 
| 163 | xsun | 3347 |  | 
| 164 |  |  | \section{Overview of Classical Statistical Mechanics\label{In:sec:SM}} | 
| 165 |  |  | Statistical mechanics provides a way to calculate the macroscopic | 
| 166 |  |  | properties of a system from the molecular interactions used in | 
| 167 |  |  | computational simulations. This section serves as a brief introduction | 
| 168 |  |  | to key concepts of classical statistical mechanics that we used in | 
| 169 |  |  | this dissertation. Tolman gives an excellent introduction to the | 
| 170 | xsun | 3372 | principles of statistical mechanics.~\cite{Tolman1979} A large part of | 
| 171 | xsun | 3375 | section~\ref{In:sec:SM} follows Tolman's notation. | 
| 172 | xsun | 3347 |  | 
| 173 |  |  | \subsection{Ensembles\label{In:ssec:ensemble}} | 
| 174 |  |  | In classical mechanics, the state of the system is completely | 
| 175 |  |  | described by the positions and momenta of all particles. If we have an | 
| 176 | xsun | 3372 | $N$ particle system, there are $6N$ coordinates ($3N$ positions $(q_1, | 
| 177 | xsun | 3347 | q_2, \ldots, q_{3N})$ and $3N$ momenta $(p_1, p_2, \ldots, p_{3N})$) | 
| 178 |  |  | to define the instantaneous state of the system. Each single set of | 
| 179 |  |  | the $6N$ coordinates can be considered as a unique point in a $6N$ | 
| 180 |  |  | dimensional space where each perpendicular axis is one of | 
| 181 |  |  | $q_{i\alpha}$ or $p_{i\alpha}$ ($i$ is the particle and $\alpha$ is | 
| 182 |  |  | the spatial axis). This $6N$ dimensional space is known as phase | 
| 183 |  |  | space. The instantaneous state of the system is a single point in | 
| 184 |  |  | phase space. A trajectory is a connected path of points in phase | 
| 185 |  |  | space. An ensemble is a collection of systems described by the same | 
| 186 |  |  | macroscopic observables but which have microscopic state | 
| 187 |  |  | distributions. In phase space an ensemble is a collection of a set of | 
| 188 |  |  | representive points. A density distribution $\rho(q^N, p^N)$ of the | 
| 189 |  |  | representive points in phase space describes the condition of an | 
| 190 |  |  | ensemble of identical systems. Since this density may change with | 
| 191 |  |  | time, it is also a function of time. $\rho(q^N, p^N, t)$ describes the | 
| 192 |  |  | ensemble at a time $t$, and $\rho(q^N, p^N, t')$ describes the | 
| 193 |  |  | ensemble at a later time $t'$. For convenience, we will use $\rho$ | 
| 194 |  |  | instead of $\rho(q^N, p^N, t)$ in the following disccusion. If we | 
| 195 |  |  | normalize $\rho$ to unity, | 
| 196 |  |  | \begin{equation} | 
| 197 |  |  | 1 = \int d \vec q~^N \int d \vec p~^N \rho, | 
| 198 | xsun | 3354 | \label{Ineq:normalized} | 
| 199 | xsun | 3347 | \end{equation} | 
| 200 |  |  | then the value of $\rho$ gives the probability of finding the system | 
| 201 | xsun | 3372 | in a unit volume in phase space. | 
| 202 | xsun | 3347 |  | 
| 203 |  |  | Liouville's theorem describes the change in density $\rho$ with | 
| 204 | xsun | 3372 | time. The number of representative points at a given volume in phase | 
| 205 | xsun | 3347 | space at any instant can be written as: | 
| 206 |  |  | \begin{equation} | 
| 207 | xsun | 3354 | \label{Ineq:deltaN} | 
| 208 | 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. | 
| 209 |  |  | \end{equation} | 
| 210 | xsun | 3372 | To calculate the change in the number of representative points in this | 
| 211 | xsun | 3347 | volume, let us consider a simple condition: the change in the number | 
| 212 | xsun | 3372 | of representative points along the $q_1$ axis. The rate of the number | 
| 213 |  |  | of the representative points entering the volume at $q_1$ per unit time | 
| 214 |  |  | is: | 
| 215 | xsun | 3347 | \begin{equation} | 
| 216 | xsun | 3354 | \label{Ineq:deltaNatq1} | 
| 217 | xsun | 3347 | \rho~\dot q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N, | 
| 218 |  |  | \end{equation} | 
| 219 | xsun | 3372 | and the rate of the number of representative points leaving the volume | 
| 220 | xsun | 3347 | at another position $q_1 + \delta q_1$ is: | 
| 221 |  |  | \begin{equation} | 
| 222 | xsun | 3354 | \label{Ineq:deltaNatq1plusdeltaq1} | 
| 223 | xsun | 3347 | \left( \rho + \frac{\partial \rho}{\partial q_1} \delta q_1 \right)\left(\dot q_1 + | 
| 224 |  |  | \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. | 
| 225 |  |  | \end{equation} | 
| 226 | xsun | 3372 | Here the higher order differentials are neglected. So the change in | 
| 227 |  |  | the number of representative points is the difference between | 
| 228 |  |  | eq.~\ref{Ineq:deltaNatq1} and eq.~\ref{Ineq:deltaNatq1plusdeltaq1}, | 
| 229 |  |  | which gives us: | 
| 230 | xsun | 3347 | \begin{equation} | 
| 231 | xsun | 3354 | \label{Ineq:deltaNatq1axis} | 
| 232 | 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, | 
| 233 |  |  | \end{equation} | 
| 234 |  |  | where, higher order differetials are neglected. If we sum over all the | 
| 235 | xsun | 3372 | axes in the phase space, we can get the change in the number of | 
| 236 |  |  | representative points in a given volume with time: | 
| 237 | xsun | 3347 | \begin{equation} | 
| 238 | xsun | 3354 | \label{Ineq:deltaNatGivenVolume} | 
| 239 | xsun | 3347 | \frac{d(\delta N)}{dt} = -\sum_{i=1}^N \left[\rho \left(\frac{\partial | 
| 240 |  |  | {\dot q_i}}{\partial q_i} + \frac{\partial | 
| 241 |  |  | {\dot p_i}}{\partial p_i}\right) + \left( \frac{\partial {\rho}}{\partial | 
| 242 |  |  | 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. | 
| 243 |  |  | \end{equation} | 
| 244 | xsun | 3372 | From Hamilton's equations of motion, | 
| 245 | xsun | 3347 | \begin{equation} | 
| 246 |  |  | \frac{\partial {\dot q_i}}{\partial q_i} = - \frac{\partial | 
| 247 |  |  | {\dot p_i}}{\partial p_i}, | 
| 248 | xsun | 3354 | \label{Ineq:canonicalFormOfEquationOfMotion} | 
| 249 | xsun | 3347 | \end{equation} | 
| 250 |  |  | this cancels out the first term on the right side of | 
| 251 | xsun | 3354 | eq.~\ref{Ineq:deltaNatGivenVolume}. If both sides of | 
| 252 |  |  | eq.~\ref{Ineq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2 | 
| 253 | xsun | 3347 | \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N$, then we | 
| 254 | xsun | 3372 | arrive at Liouville's theorem: | 
| 255 | xsun | 3347 | \begin{equation} | 
| 256 |  |  | \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = -\sum_{i} \left( | 
| 257 |  |  | \frac{\partial {\rho}}{\partial | 
| 258 |  |  | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right). | 
| 259 | xsun | 3354 | \label{Ineq:simpleFormofLiouville} | 
| 260 | xsun | 3347 | \end{equation} | 
| 261 |  |  | This is the basis of statistical mechanics. If we move the right | 
| 262 | xsun | 3354 | side of equation~\ref{Ineq:simpleFormofLiouville} to the left, we | 
| 263 | xsun | 3347 | will obtain | 
| 264 |  |  | \begin{equation} | 
| 265 |  |  | \left( \frac{\partial \rho}{\partial t} \right)_{q, p} + \sum_{i} \left( | 
| 266 |  |  | \frac{\partial {\rho}}{\partial | 
| 267 |  |  | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right) | 
| 268 |  |  | = 0. | 
| 269 | xsun | 3354 | \label{Ineq:anotherFormofLiouville} | 
| 270 | xsun | 3347 | \end{equation} | 
| 271 |  |  | It is easy to note that the left side of | 
| 272 | xsun | 3354 | equation~\ref{Ineq:anotherFormofLiouville} is the total derivative of | 
| 273 | xsun | 3372 | $\rho$ with respect to $t$, which means | 
| 274 | xsun | 3347 | \begin{equation} | 
| 275 |  |  | \frac{d \rho}{dt} = 0, | 
| 276 | xsun | 3354 | \label{Ineq:conservationofRho} | 
| 277 | xsun | 3347 | \end{equation} | 
| 278 |  |  | and the rate of density change is zero in the neighborhood of any | 
| 279 | xsun | 3372 | selected moving representative points in the phase space. | 
| 280 | xsun | 3347 |  | 
| 281 | xsun | 3375 | The type of thermodynamic ensemble is determined by the density | 
| 282 | xsun | 3347 | distribution. If we consider the density distribution as only a | 
| 283 |  |  | function of $q$ and $p$, which means the rate of change of the phase | 
| 284 | xsun | 3372 | space density in the neighborhood of all representative points in the | 
| 285 | xsun | 3347 | phase space is zero, | 
| 286 |  |  | \begin{equation} | 
| 287 |  |  | \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = 0. | 
| 288 | xsun | 3354 | \label{Ineq:statEquilibrium} | 
| 289 | xsun | 3347 | \end{equation} | 
| 290 |  |  | We may conclude the ensemble is in {\it statistical equilibrium}. An | 
| 291 | xsun | 3372 | ensemble in statistical equilibrium means the system is also in | 
| 292 | xsun | 3347 | macroscopic equilibrium. If $\left( \frac{\partial \rho}{\partial t} | 
| 293 |  |  | \right)_{q, p} = 0$, then | 
| 294 |  |  | \begin{equation} | 
| 295 |  |  | \sum_{i} \left( | 
| 296 |  |  | \frac{\partial {\rho}}{\partial | 
| 297 |  |  | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right) | 
| 298 |  |  | = 0. | 
| 299 | xsun | 3354 | \label{Ineq:constantofMotion} | 
| 300 | xsun | 3347 | \end{equation} | 
| 301 |  |  | If $\rho$ is a function only of some constant of the motion, $\rho$ is | 
| 302 |  |  | independent of time. For a conservative system, the energy of the | 
| 303 | xsun | 3372 | system is one of the constants of the motion. There are many | 
| 304 |  |  | thermodynamically relevant ensembles: when the density distribution is | 
| 305 |  |  | constant everywhere in the phase space, | 
| 306 | xsun | 3347 | \begin{equation} | 
| 307 |  |  | \rho = \mathrm{const.} | 
| 308 | xsun | 3354 | \label{Ineq:uniformEnsemble} | 
| 309 | xsun | 3347 | \end{equation} | 
| 310 | xsun | 3375 | the ensemble is called {\it uniform ensemble}, but this ensemble has | 
| 311 |  |  | little relevance for physical chemistry. It is an ensemble with | 
| 312 |  |  | essentially infinite temperature. | 
| 313 | xsun | 3372 |  | 
| 314 |  |  | \subsubsection{The Microcanonical Ensemble\label{In:sssec:microcanonical}} | 
| 315 | xsun | 3375 | The most useful ensemble for Molecular Dynamics is the {\it | 
| 316 |  |  | microcanonical ensemble}, for which: | 
| 317 | xsun | 3347 | \begin{equation} | 
| 318 |  |  | \rho = \delta \left( H(q^N, p^N) - E \right) \frac{1}{\Sigma (N, V, E)} | 
| 319 | xsun | 3354 | \label{Ineq:microcanonicalEnsemble} | 
| 320 | xsun | 3347 | \end{equation} | 
| 321 |  |  | where $\Sigma(N, V, E)$ is a normalization constant parameterized by | 
| 322 |  |  | $N$, the total number of particles, $V$, the total physical volume and | 
| 323 |  |  | $E$, the total energy. The physical meaning of $\Sigma(N, V, E)$ is | 
| 324 |  |  | the phase space volume accessible to a microcanonical system with | 
| 325 |  |  | energy $E$ evolving under Hamilton's equations. $H(q^N, p^N)$ is the | 
| 326 |  |  | Hamiltonian of the system. The Gibbs entropy is defined as | 
| 327 |  |  | \begin{equation} | 
| 328 |  |  | S = - k_B \int d \vec q~^N \int d \vec p~^N \rho \ln [C^N \rho], | 
| 329 | xsun | 3354 | \label{Ineq:gibbsEntropy} | 
| 330 | xsun | 3347 | \end{equation} | 
| 331 |  |  | where $k_B$ is the Boltzmann constant and $C^N$ is a number which | 
| 332 | xsun | 3372 | makes the argument of $\ln$ dimensionless. In this case, $C^N$ is the | 
| 333 | xsun | 3375 | total phase space volume of one state which has the same units as | 
| 334 |  |  | $\Sigma(N, V, E)$. The entropy of a microcanonical ensemble is given | 
| 335 |  |  | by | 
| 336 | xsun | 3347 | \begin{equation} | 
| 337 |  |  | S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right). | 
| 338 | xsun | 3354 | \label{Ineq:entropy} | 
| 339 | xsun | 3347 | \end{equation} | 
| 340 | xsun | 3372 |  | 
| 341 |  |  | \subsubsection{The Canonical Ensemble\label{In:sssec:canonical}} | 
| 342 | xsun | 3347 | If the density distribution $\rho$ is given by | 
| 343 |  |  | \begin{equation} | 
| 344 |  |  | \rho = \frac{1}{Z_N}e^{-H(q^N, p^N) / k_B T}, | 
| 345 | xsun | 3354 | \label{Ineq:canonicalEnsemble} | 
| 346 | xsun | 3347 | \end{equation} | 
| 347 |  |  | the ensemble is known as the {\it canonical ensemble}. Here, | 
| 348 |  |  | \begin{equation} | 
| 349 |  |  | Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N  e^{-H(q^N, p^N) / k_B T}, | 
| 350 | xsun | 3354 | \label{Ineq:partitionFunction} | 
| 351 | xsun | 3347 | \end{equation} | 
| 352 | xsun | 3375 | which is also known as the canonical {\it partition | 
| 353 |  |  | function}. $\Gamma$ indicates that the integral is over all phase | 
| 354 |  |  | space. In the canonical ensemble, $N$, the total number of particles, | 
| 355 |  |  | $V$, total volume, and $T$, the temperature, are constants. The | 
| 356 |  |  | systems with the lowest energies hold the largest | 
| 357 |  |  | population. Thermodynamics maximizes the entropy, $S$, implying that | 
| 358 | xsun | 3347 | \begin{equation} | 
| 359 |  |  | \begin{array}{ccc} | 
| 360 |  |  | \delta S = 0 & \mathrm{and} & \delta^2 S < 0. | 
| 361 |  |  | \end{array} | 
| 362 | xsun | 3354 | \label{Ineq:maximumPrinciple} | 
| 363 | xsun | 3347 | \end{equation} | 
| 364 | xsun | 3372 | From Eq.~\ref{Ineq:maximumPrinciple} and two constrains on the | 
| 365 |  |  | canonical ensemble, {\it i.e.}, total probability and average energy | 
| 366 |  |  | must be conserved, the partition function can be shown to be | 
| 367 |  |  | equivalent to | 
| 368 | xsun | 3347 | \begin{equation} | 
| 369 |  |  | Z_N = e^{-A/k_B T}, | 
| 370 | xsun | 3354 | \label{Ineq:partitionFunctionWithFreeEnergy} | 
| 371 | xsun | 3347 | \end{equation} | 
| 372 |  |  | where $A$ is the Helmholtz free energy. The significance of | 
| 373 | xsun | 3354 | Eq.~\ref{Ineq:entropy} and~\ref{Ineq:partitionFunctionWithFreeEnergy} is | 
| 374 | xsun | 3347 | that they serve as a connection between macroscopic properties of the | 
| 375 | xsun | 3372 | system and the distribution of microscopic states. | 
| 376 | xsun | 3347 |  | 
| 377 |  |  | There is an implicit assumption that our arguments are based on so | 
| 378 | xsun | 3375 | far. Tow representative points in phase space are equally likely to be | 
| 379 |  |  | found if they have the same energy. In other words, all energetically | 
| 380 |  |  | accessible states are represented , and the probabilities to find the | 
| 381 |  |  | system in any of the accessible states is equal to that states | 
| 382 |  |  | Boltzmann weight. This is called the principle of equal a {\it priori} | 
| 383 | xsun | 3372 | probabilities. | 
| 384 | xsun | 3347 |  | 
| 385 | xsun | 3372 | \subsection{Statistical Averages\label{In:ssec:average}} | 
| 386 |  |  | Given a density distribution $\rho$ in phase space, the average | 
| 387 | xsun | 3347 | of any quantity ($F(q^N, p^N$)) which depends on the coordinates | 
| 388 |  |  | ($q^N$) and the momenta ($p^N$) for all the systems in the ensemble | 
| 389 |  |  | can be calculated based on the definition shown by | 
| 390 | xsun | 3354 | Eq.~\ref{Ineq:statAverage1} | 
| 391 | xsun | 3347 | \begin{equation} | 
| 392 | xsun | 3372 | \langle F(q^N, p^N) \rangle = \frac{\int d \vec q~^N \int d \vec p~^N | 
| 393 |  |  | F(q^N, p^N) \rho}{\int d \vec q~^N \int d \vec p~^N \rho}. | 
| 394 | xsun | 3354 | \label{Ineq:statAverage1} | 
| 395 | xsun | 3347 | \end{equation} | 
| 396 |  |  | Since the density distribution $\rho$ is normalized to unity, the mean | 
| 397 |  |  | value of $F(q^N, p^N)$ is simplified to | 
| 398 |  |  | \begin{equation} | 
| 399 | xsun | 3372 | \langle F(q^N, p^N) \rangle = \int d \vec q~^N \int d \vec p~^N F(q^N, | 
| 400 |  |  | p^N) \rho, | 
| 401 | xsun | 3354 | \label{Ineq:statAverage2} | 
| 402 | xsun | 3347 | \end{equation} | 
| 403 | xsun | 3372 | called the {\it ensemble average}. However, the quantity is often | 
| 404 |  |  | averaged for a finite time in real experiments, | 
| 405 | xsun | 3347 | \begin{equation} | 
| 406 |  |  | \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty} | 
| 407 | xsun | 3372 | \frac{1}{T} \int_{t_0}^{t_0+T} F[q^N(t), p^N(t)] dt. | 
| 408 | xsun | 3354 | \label{Ineq:timeAverage1} | 
| 409 | xsun | 3347 | \end{equation} | 
| 410 |  |  | Usually this time average is independent of $t_0$ in statistical | 
| 411 | xsun | 3354 | mechanics, so Eq.~\ref{Ineq:timeAverage1} becomes | 
| 412 | xsun | 3347 | \begin{equation} | 
| 413 |  |  | \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty} | 
| 414 | xsun | 3372 | \frac{1}{T} \int_{0}^{T} F[q^N(t), p^N(t)] dt | 
| 415 | xsun | 3354 | \label{Ineq:timeAverage2} | 
| 416 | xsun | 3347 | \end{equation} | 
| 417 | xsun | 3375 | for an finite time interval, $T$. | 
| 418 | xsun | 3347 |  | 
| 419 | xsun | 3372 | \subsubsection{Ergodicity\label{In:sssec:ergodicity}} | 
| 420 |  |  | The {\it ergodic hypothesis}, an important hypothesis governing modern | 
| 421 |  |  | computer simulations states that the system will eventually pass | 
| 422 | xsun | 3347 | arbitrarily close to any point that is energetically accessible in | 
| 423 |  |  | phase space. Mathematically, this leads to | 
| 424 |  |  | \begin{equation} | 
| 425 | xsun | 3372 | \langle F(q^N, p^N) \rangle = \langle F(q^N, p^N) \rangle_t. | 
| 426 | xsun | 3354 | \label{Ineq:ergodicity} | 
| 427 | xsun | 3347 | \end{equation} | 
| 428 | xsun | 3372 | Eq.~\ref{Ineq:ergodicity} validates Molecular Dynamics as a form of | 
| 429 |  |  | averaging for sufficiently ergodic systems. Also Monte Carlo may be | 
| 430 |  |  | used to obtain ensemble averages of a quantity which are related to | 
| 431 |  |  | time averages measured in experiments. | 
| 432 | xsun | 3347 |  | 
| 433 | xsun | 3372 | \subsection{Correlation Functions\label{In:ssec:corr}} | 
| 434 |  |  | Thermodynamic properties can be computed by equilibrium statistical | 
| 435 |  |  | mechanics. On the other hand, {\it Time correlation functions} are a | 
| 436 |  |  | powerful tool to understand the evolution of a dynamical | 
| 437 |  |  | systems. Imagine that property $A(q^N, p^N, t)$ as a function of | 
| 438 |  |  | coordinates $q^N$ and momenta $p^N$ has an intial value at $t_0$, and | 
| 439 |  |  | at a later time $t_0 + \tau$ this value has changed. If $\tau$ is very | 
| 440 |  |  | small, the change of the value is minor, and the later value of | 
| 441 |  |  | $A(q^N, p^N, t_0 + \tau)$ is correlated to its initial value. However, | 
| 442 |  |  | when $\tau$ is large, this correlation is lost. A time correlation | 
| 443 |  |  | function measures this relationship and is defined | 
| 444 |  |  | by~\cite{Berne90} | 
| 445 | xsun | 3347 | \begin{equation} | 
| 446 | xsun | 3372 | C_{AA}(\tau) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarrow | 
| 447 |  |  | \infty} | 
| 448 | xsun | 3347 | \frac{1}{T} \int_{0}^{T} dt A(t) A(t + \tau). | 
| 449 | xsun | 3354 | \label{Ineq:autocorrelationFunction} | 
| 450 | xsun | 3347 | \end{equation} | 
| 451 | xsun | 3372 | Eq.~\ref{Ineq:autocorrelationFunction} is the correlation function of | 
| 452 |  |  | a single variable, called an {\it autocorrelation function}. The | 
| 453 |  |  | definition of the correlation function for two different variables is | 
| 454 |  |  | similar to that of autocorrelation function, which is | 
| 455 | xsun | 3347 | \begin{equation} | 
| 456 | xsun | 3372 | C_{AB}(\tau) = \langle A(0)B(\tau) \rangle = \lim_{T \rightarrow \infty} | 
| 457 | xsun | 3347 | \frac{1}{T} \int_{0}^{T} dt A(t) B(t + \tau), | 
| 458 | xsun | 3354 | \label{Ineq:crosscorrelationFunction} | 
| 459 | xsun | 3347 | \end{equation} | 
| 460 | xsun | 3375 | and is called a {\it cross correlation function}. | 
| 461 | xsun | 3347 |  | 
| 462 | xsun | 3372 | We know from the ergodic hypothesis that there is a relationship | 
| 463 |  |  | between time average and ensemble average. We can put the correlation | 
| 464 | xsun | 3375 | function in a classical mechanical form, | 
| 465 | xsun | 3347 | \begin{equation} | 
| 466 | xsun | 3375 | C_{AA}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N, p^N] | 
| 467 |  |  | A[q^N(\tau), p^N(\tau)] \rho(q^N, p^N) | 
| 468 | xsun | 3354 | \label{Ineq:autocorrelationFunctionCM} | 
| 469 | xsun | 3347 | \end{equation} | 
| 470 | xsun | 3375 | where $q^N(\tau)$, $p^N(\tau)$ is the phase space point that follows | 
| 471 |  |  | classical evolution of the point $q^N$, $p^N$ after a tme $\tau$ has | 
| 472 |  |  | elapsed, and | 
| 473 | xsun | 3347 | \begin{equation} | 
| 474 | xsun | 3375 | C_{AB}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N, p^N] | 
| 475 |  |  | B[q^N(\tau), p^N(\tau)] \rho(q^N, p^N) | 
| 476 | xsun | 3354 | \label{Ineq:crosscorrelationFunctionCM} | 
| 477 | xsun | 3347 | \end{equation} | 
| 478 | xsun | 3374 | as the autocorrelation function and cross correlation functions | 
| 479 | xsun | 3347 | respectively. $\rho(q, p)$ is the density distribution at equillibrium | 
| 480 | xsun | 3374 | in phase space. In many cases, correlation functions decay as a | 
| 481 |  |  | single exponential in time | 
| 482 | xsun | 3347 | \begin{equation} | 
| 483 |  |  | C(t) \sim e^{-t / \tau_r}, | 
| 484 | xsun | 3354 | \label{Ineq:relaxation} | 
| 485 | xsun | 3347 | \end{equation} | 
| 486 | xsun | 3374 | where $\tau_r$ is known as relaxation time which describes the rate of | 
| 487 | xsun | 3347 | the decay. | 
| 488 |  |  |  | 
| 489 | xsun | 3374 | \section{Methodology\label{In:sec:method}} | 
| 490 |  |  | The simulations performed in this dissertation branch into two main | 
| 491 |  |  | categories, Monte Carlo and Molecular Dynamics. There are two main | 
| 492 |  |  | differences between Monte Carlo and Molecular Dynamics | 
| 493 |  |  | simulations. One is that the Monte Carlo simulations are time | 
| 494 |  |  | independent methods of sampling structural features of an ensemble, | 
| 495 |  |  | while Molecular Dynamics simulations provide dynamic | 
| 496 | xsun | 3375 | information. Additionally, Monte Carlo methods are stochastic | 
| 497 |  |  | processes; the future configurations of the system are not determined | 
| 498 | xsun | 3374 | by its past. However, in Molecular Dynamics, the system is propagated | 
| 499 | xsun | 3375 | by Hamilton's equations of motion, and the trajectory of the system | 
| 500 | xsun | 3374 | evolving in phase space is deterministic. Brief introductions of the | 
| 501 |  |  | two algorithms are given in section~\ref{In:ssec:mc} | 
| 502 |  |  | and~\ref{In:ssec:md}. Langevin Dynamics, an extension of the Molecular | 
| 503 |  |  | Dynamics that includes implicit solvent effects, is introduced by | 
| 504 | xsun | 3347 | section~\ref{In:ssec:ld}. | 
| 505 |  |  |  | 
| 506 |  |  | \subsection{Monte Carlo\label{In:ssec:mc}} | 
| 507 | xsun | 3374 | A Monte Carlo integration algorithm was first introduced by Metropolis | 
| 508 |  |  | {\it et al.}~\cite{Metropolis53} The basic Metropolis Monte Carlo | 
| 509 |  |  | algorithm is usually applied to the canonical ensemble, a | 
| 510 | xsun | 3375 | Boltzmann-weighted ensemble, in which $N$, the total number of | 
| 511 |  |  | particles, $V$, the total volume, and $T$, the temperature are | 
| 512 |  |  | constants. An average in this ensemble is given | 
| 513 | xsun | 3347 | \begin{equation} | 
| 514 | xsun | 3374 | \langle A \rangle = \frac{1}{Z_N} \int d \vec q~^N \int d \vec p~^N | 
| 515 |  |  | A(q^N, p^N) e^{-H(q^N, p^N) / k_B T}. | 
| 516 | xsun | 3354 | \label{Ineq:energyofCanonicalEnsemble} | 
| 517 | xsun | 3347 | \end{equation} | 
| 518 | xsun | 3374 | The Hamiltonian is the sum of the kinetic energy, $K(p^N)$, a function | 
| 519 |  |  | of momenta and the potential energy, $U(q^N)$, a function of | 
| 520 |  |  | positions, | 
| 521 | xsun | 3347 | \begin{equation} | 
| 522 |  |  | H(q^N, p^N) = K(p^N) + U(q^N). | 
| 523 | xsun | 3354 | \label{Ineq:hamiltonian} | 
| 524 | xsun | 3347 | \end{equation} | 
| 525 | xsun | 3374 | If the property $A$ is a function only of position ($ A = A(q^N)$), | 
| 526 |  |  | the mean value of $A$ can be reduced to | 
| 527 | xsun | 3347 | \begin{equation} | 
| 528 | xsun | 3374 | \langle A \rangle = \frac{\int d \vec q~^N A e^{-U(q^N) / k_B T}}{\int d \vec q~^N e^{-U(q^N) / k_B T}}, | 
| 529 | xsun | 3354 | \label{Ineq:configurationIntegral} | 
| 530 | xsun | 3347 | \end{equation} | 
| 531 |  |  | The kinetic energy $K(p^N)$ is factored out in | 
| 532 | xsun | 3354 | Eq.~\ref{Ineq:configurationIntegral}. $\langle A | 
| 533 | xsun | 3374 | \rangle$ is now a configuration integral, and | 
| 534 | xsun | 3354 | Eq.~\ref{Ineq:configurationIntegral} is equivalent to | 
| 535 | xsun | 3347 | \begin{equation} | 
| 536 | xsun | 3375 | \langle A \rangle = \int d \vec q~^N A \rho(q^N), | 
| 537 | xsun | 3354 | \label{Ineq:configurationAve} | 
| 538 | xsun | 3347 | \end{equation} | 
| 539 | xsun | 3375 | where $\rho(q^N)$ is a configurational probability | 
| 540 |  |  | \begin{equation} | 
| 541 |  |  | \rho(q^N) = \frac{e^{-U(q^N) / k_B T}}{\int d \vec q~^N e^{-U(q^N) / k_B T}}. | 
| 542 |  |  | \label{Ineq:configurationProb} | 
| 543 |  |  | \end{equation} | 
| 544 | xsun | 3347 |  | 
| 545 | xsun | 3374 | In a Monte Carlo simulation of a system in the canonical ensemble, the | 
| 546 |  |  | probability of the system being in a state $s$ is $\rho_s$, the change | 
| 547 |  |  | of this probability with time is given by | 
| 548 | xsun | 3347 | \begin{equation} | 
| 549 |  |  | \frac{d \rho_s}{dt} = \sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ], | 
| 550 | xsun | 3354 | \label{Ineq:timeChangeofProb} | 
| 551 | xsun | 3347 | \end{equation} | 
| 552 |  |  | where $w_{ss'}$ is the tansition probability of going from state $s$ | 
| 553 |  |  | to state $s'$. Since $\rho_s$ is independent of time at equilibrium, | 
| 554 |  |  | \begin{equation} | 
| 555 |  |  | \frac{d \rho_{s}^{equilibrium}}{dt} = 0, | 
| 556 | xsun | 3354 | \label{Ineq:equiProb} | 
| 557 | xsun | 3347 | \end{equation} | 
| 558 | xsun | 3375 | the sum of transition probabilities $\sum_{s'} [ -w_{ss'}\rho_s + | 
| 559 |  |  | w_{s's}\rho_{s'} ]$ is $0$ for all $s'$. So | 
| 560 | xsun | 3347 | \begin{equation} | 
| 561 |  |  | \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = \frac{w_{s's}}{w_{ss'}}. | 
| 562 | xsun | 3354 | \label{Ineq:relationshipofRhoandW} | 
| 563 | xsun | 3347 | \end{equation} | 
| 564 | xsun | 3374 | If the ratio of state populations | 
| 565 | xsun | 3347 | \begin{equation} | 
| 566 | xsun | 3374 | \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = e^{-(U_s - U_{s'}) / k_B T}. | 
| 567 |  |  | \label{Ineq:satisfyofBoltzmannStatistics} | 
| 568 |  |  | \end{equation} | 
| 569 |  |  | then the ratio of transition probabilities, | 
| 570 |  |  | \begin{equation} | 
| 571 | xsun | 3347 | \frac{w_{s's}}{w_{ss'}} = e^{-(U_s - U_{s'}) / k_B T}, | 
| 572 | xsun | 3354 | \label{Ineq:conditionforBoltzmannStatistics} | 
| 573 | xsun | 3347 | \end{equation} | 
| 574 | xsun | 3375 | An algorithm that indicates how a Monte Carlo simulation generates a | 
| 575 | xsun | 3374 | transition probability governed by | 
| 576 |  |  | \ref{Ineq:conditionforBoltzmannStatistics}, is given schematically as, | 
| 577 | xsun | 3347 | \begin{enumerate} | 
| 578 | xsun | 3374 | \item\label{Initm:oldEnergy} Choose a particle randomly, and calculate | 
| 579 | xsun | 3375 | the energy of the rest of the system due to the current location of | 
| 580 |  |  | the particle. | 
| 581 | xsun | 3374 | \item\label{Initm:newEnergy} Make a random displacement of the particle, | 
| 582 |  |  | calculate the new energy taking the movement of the particle into account. | 
| 583 | xsun | 3347 | \begin{itemize} | 
| 584 | xsun | 3375 | \item If the energy goes down, keep the new configuration. | 
| 585 | xsun | 3374 | \item If the energy goes up, pick a random number between $[0,1]$. | 
| 586 | xsun | 3347 | \begin{itemize} | 
| 587 | xsun | 3374 | \item If the random number smaller than | 
| 588 | xsun | 3375 | $e^{-(U_{new} - U_{old})} / k_B T$, keep the new configuration. | 
| 589 | xsun | 3374 | \item If the random number is larger than | 
| 590 | xsun | 3375 | $e^{-(U_{new} - U_{old})} / k_B T$, keep the old configuration. | 
| 591 | xsun | 3347 | \end{itemize} | 
| 592 |  |  | \end{itemize} | 
| 593 | xsun | 3374 | \item\label{Initm:accumulateAvg} Accumulate the averages based on the | 
| 594 | xsun | 3375 | current configuration. | 
| 595 | xsun | 3374 | \item Go to step~\ref{Initm:oldEnergy}. | 
| 596 | xsun | 3347 | \end{enumerate} | 
| 597 | xsun | 3375 | It is important for sampling accuracy that the old configuration is | 
| 598 | xsun | 3374 | sampled again if it is kept. | 
| 599 | xsun | 3347 |  | 
| 600 |  |  | \subsection{Molecular Dynamics\label{In:ssec:md}} | 
| 601 |  |  | Although some of properites of the system can be calculated from the | 
| 602 | xsun | 3374 | ensemble average in Monte Carlo simulations, due to the absence of the | 
| 603 |  |  | time dependence, it is impossible to gain information on dynamic | 
| 604 |  |  | properties from Monte Carlo simulations. Molecular Dynamics evolves | 
| 605 |  |  | the positions and momenta of the particles in the system. The | 
| 606 |  |  | evolution of the system obeys the laws of classical mechanics, and in | 
| 607 |  |  | most situations, there is no need to account for quantum effects. In a | 
| 608 |  |  | real experiment, the instantaneous positions and momenta of the | 
| 609 | xsun | 3375 | particles in the system are ofter neither important nor measurable, | 
| 610 |  |  | the observable quantities are usually an average value for a finite | 
| 611 |  |  | time interval. These quantities are expressed as a function of | 
| 612 |  |  | positions and momenta in Molecular Dynamics simulations. For example, | 
| 613 | xsun | 3374 | temperature of the system is defined by | 
| 614 | xsun | 3347 | \begin{equation} | 
| 615 | xsun | 3374 | \frac{3}{2} N k_B T = \langle \sum_{i=1}^N \frac{1}{2} m_i v_i \rangle, | 
| 616 | xsun | 3354 | \label{Ineq:temperature} | 
| 617 | xsun | 3347 | \end{equation} | 
| 618 | xsun | 3374 | here $m_i$ is the mass of particle $i$ and $v_i$ is the velocity of | 
| 619 |  |  | particle $i$. The right side of Eq.~\ref{Ineq:temperature} is the | 
| 620 |  |  | average kinetic energy of the system. | 
| 621 | xsun | 3347 |  | 
| 622 | xsun | 3374 | The initial positions of the particles are chosen so that there is no | 
| 623 |  |  | overlap of the particles. The initial velocities at first are | 
| 624 |  |  | distributed randomly to the particles using a Maxwell-Boltzmann | 
| 625 | xsun | 3375 | distribution, and then shifted to make the total linear momentum of | 
| 626 |  |  | the system $0$. | 
| 627 | xsun | 3374 |  | 
| 628 |  |  | The core of Molecular Dynamics simulations is the calculation of | 
| 629 |  |  | forces and the integration algorithm. Calculation of the forces often | 
| 630 |  |  | involves enormous effort. This is the most time consuming step in the | 
| 631 |  |  | Molecular Dynamics scheme. Evaluation of the forces is mathematically | 
| 632 |  |  | simple, | 
| 633 | xsun | 3347 | \begin{equation} | 
| 634 |  |  | f(q) = - \frac{\partial U(q)}{\partial q}, | 
| 635 | xsun | 3354 | \label{Ineq:force} | 
| 636 | xsun | 3347 | \end{equation} | 
| 637 | xsun | 3374 | where $U(q)$ is the potential of the system. However, the numerical | 
| 638 |  |  | details of this computation are often quite complex. Once the forces | 
| 639 |  |  | computed, the positions and velocities are updated by integrating | 
| 640 |  |  | Hamilton's equations of motion, | 
| 641 |  |  | \begin{eqnarray} | 
| 642 |  |  | \dot p_i & = & -\frac{\partial H}{\partial q_i} = -\frac{\partial | 
| 643 |  |  | U(q_i)}{\partial q_i} = f(q_i) \\ | 
| 644 |  |  | \dot q_i & = & p_i | 
| 645 | xsun | 3354 | \label{Ineq:newton} | 
| 646 | xsun | 3374 | \end{eqnarray} | 
| 647 |  |  | The classic example of an integrating algorithm is the Verlet | 
| 648 |  |  | algorithm, which is one of the simplest algorithms for integrating the | 
| 649 |  |  | equations of motion. The Taylor expansion of the position at time $t$ | 
| 650 |  |  | is | 
| 651 | xsun | 3347 | \begin{equation} | 
| 652 | xsun | 3375 | q(t+\Delta t)= q(t) + \frac{p(t)}{m} \Delta t + \frac{f(t)}{2m}\Delta t^2 + | 
| 653 | xsun | 3347 | \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} + | 
| 654 |  |  | \mathcal{O}(\Delta t^4) | 
| 655 | xsun | 3354 | \label{Ineq:verletFuture} | 
| 656 | xsun | 3347 | \end{equation} | 
| 657 |  |  | for a later time $t+\Delta t$, and | 
| 658 |  |  | \begin{equation} | 
| 659 | xsun | 3375 | q(t-\Delta t)= q(t) - \frac{p(t)}{m} \Delta t + \frac{f(t)}{2m}\Delta t^2 - | 
| 660 | xsun | 3347 | \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} + | 
| 661 |  |  | \mathcal{O}(\Delta t^4) , | 
| 662 | xsun | 3354 | \label{Ineq:verletPrevious} | 
| 663 | xsun | 3347 | \end{equation} | 
| 664 | xsun | 3374 | for a previous time $t-\Delta t$. Adding Eq.~\ref{Ineq:verletFuture} | 
| 665 |  |  | and~\ref{Ineq:verletPrevious} gives | 
| 666 | xsun | 3347 | \begin{equation} | 
| 667 |  |  | q(t+\Delta t)+q(t-\Delta t) = | 
| 668 |  |  | 2q(t) + \frac{f(t)}{m}\Delta t^2 + \mathcal{O}(\Delta t^4), | 
| 669 | xsun | 3354 | \label{Ineq:verletSum} | 
| 670 | xsun | 3347 | \end{equation} | 
| 671 |  |  | so, the new position can be expressed as | 
| 672 |  |  | \begin{equation} | 
| 673 |  |  | q(t+\Delta t) \approx | 
| 674 |  |  | 2q(t) - q(t-\Delta t) + \frac{f(t)}{m}\Delta t^2. | 
| 675 | xsun | 3354 | \label{Ineq:newPosition} | 
| 676 | xsun | 3347 | \end{equation} | 
| 677 | xsun | 3374 | The higher order terms in $\Delta t$ are omitted. | 
| 678 | xsun | 3347 |  | 
| 679 | xsun | 3374 | Numerous techniques and tricks have been applied to Molecular Dynamics | 
| 680 |  |  | simulations to gain greater efficiency or accuracy. The engine used in | 
| 681 |  |  | this dissertation for the Molecular Dynamics simulations is {\sc | 
| 682 |  |  | oopse}. More details of the algorithms and techniques used in this | 
| 683 |  |  | code can be found in Ref.~\cite{Meineke2005}. | 
| 684 | xsun | 3347 |  | 
| 685 |  |  | \subsection{Langevin Dynamics\label{In:ssec:ld}} | 
| 686 | xsun | 3375 | In many cases, the properites of the solvent (like the water in the | 
| 687 |  |  | lipid-water system studied in this dissertation) are less interesting | 
| 688 |  |  | to the researchers than the behavior of the solute. However, the major | 
| 689 |  |  | computational expense is ofter the solvent-solvent interactions, this | 
| 690 |  |  | is due to the large number of the solvent molecules when compared to | 
| 691 |  |  | the number of solute molecules. The ratio of the number of lipid | 
| 692 |  |  | molecules to the number of water molecules is $1:25$ in our | 
| 693 |  |  | lipid-water system. The efficiency of the Molecular Dynamics | 
| 694 |  |  | simulations is greatly reduced, with up to 85\% of CPU time spent | 
| 695 |  |  | calculating only water-water interactions. | 
| 696 | xsun | 3347 |  | 
| 697 | xsun | 3374 | As an extension of the Molecular Dynamics methodologies, Langevin | 
| 698 |  |  | Dynamics seeks a way to avoid integrating the equations of motion for | 
| 699 |  |  | solvent particles without losing the solvent effects on the solute | 
| 700 |  |  | particles. One common approximation is to express the coupling of the | 
| 701 |  |  | solute and solvent as a set of harmonic oscillators. The Hamiltonian | 
| 702 |  |  | of such a system is written as | 
| 703 | xsun | 3347 | \begin{equation} | 
| 704 |  |  | H = \frac{p^2}{2m} + U(q) + H_B + \Delta U(q), | 
| 705 | xsun | 3354 | \label{Ineq:hamiltonianofCoupling} | 
| 706 | xsun | 3347 | \end{equation} | 
| 707 | xsun | 3374 | where $H_B$ is the Hamiltonian of the bath which  is a set of N | 
| 708 |  |  | harmonic oscillators | 
| 709 | xsun | 3347 | \begin{equation} | 
| 710 |  |  | H_B = \sum_{\alpha = 1}^{N} \left\{ \frac{p_\alpha^2}{2m_\alpha} + | 
| 711 |  |  | \frac{1}{2} m_\alpha \omega_\alpha^2 q_\alpha^2\right\}, | 
| 712 | xsun | 3354 | \label{Ineq:hamiltonianofBath} | 
| 713 | xsun | 3347 | \end{equation} | 
| 714 | xsun | 3374 | $\alpha$ runs over all the degree of freedoms of the bath, | 
| 715 |  |  | $\omega_\alpha$ is the bath frequency of oscillator $\alpha$, and | 
| 716 |  |  | $\Delta U(q)$ is the bilinear coupling given by | 
| 717 | xsun | 3347 | \begin{equation} | 
| 718 |  |  | \Delta U = -\sum_{\alpha = 1}^{N} g_\alpha q_\alpha q, | 
| 719 | xsun | 3354 | \label{Ineq:systemBathCoupling} | 
| 720 | xsun | 3347 | \end{equation} | 
| 721 | xsun | 3374 | where $g_\alpha$ is the coupling constant for oscillator $\alpha$. By | 
| 722 |  |  | solving the Hamilton's equations of motion, the {\it Generalized | 
| 723 |  |  | Langevin Equation} for this system is derived as | 
| 724 | xsun | 3347 | \begin{equation} | 
| 725 |  |  | m \ddot q = -\frac{\partial W(q)}{\partial q} - \int_0^t \xi(t) \dot q(t-t')dt' + R(t), | 
| 726 | xsun | 3354 | \label{Ineq:gle} | 
| 727 | xsun | 3347 | \end{equation} | 
| 728 |  |  | with mean force, | 
| 729 |  |  | \begin{equation} | 
| 730 |  |  | W(q) = U(q) - \sum_{\alpha = 1}^N \frac{g_\alpha^2}{2m_\alpha | 
| 731 | xsun | 3374 | \omega_\alpha^2}q^2. | 
| 732 | xsun | 3354 | \label{Ineq:meanForce} | 
| 733 | xsun | 3347 | \end{equation} | 
| 734 | xsun | 3374 | The {\it friction kernel}, $\xi(t)$, depends only on the coordinates | 
| 735 |  |  | of the solute particles, | 
| 736 | xsun | 3347 | \begin{equation} | 
| 737 |  |  | \xi(t) = \sum_{\alpha = 1}^N \frac{-g_\alpha}{m_\alpha | 
| 738 |  |  | \omega_\alpha} \cos(\omega_\alpha t), | 
| 739 | xsun | 3354 | \label{Ineq:xiforGLE} | 
| 740 | xsun | 3347 | \end{equation} | 
| 741 | xsun | 3374 | and a ``random'' force, | 
| 742 | xsun | 3347 | \begin{equation} | 
| 743 |  |  | R(t) = \sum_{\alpha = 1}^N \left( g_\alpha q_\alpha(0)-\frac{g_\alpha}{m_\alpha | 
| 744 |  |  | \omega_\alpha^2}q(0)\right) \cos(\omega_\alpha t) + \frac{\dot | 
| 745 |  |  | q_\alpha(0)}{\omega_\alpha} \sin(\omega_\alpha t), | 
| 746 | xsun | 3354 | \label{Ineq:randomForceforGLE} | 
| 747 | xsun | 3347 | \end{equation} | 
| 748 | xsun | 3375 | that depends only on the initial conditions. The relationship of | 
| 749 |  |  | friction kernel $\xi(t)$ and random force $R(t)$ is given by the | 
| 750 |  |  | second fluctuation dissipation theorem, | 
| 751 | xsun | 3347 | \begin{equation} | 
| 752 | xsun | 3374 | \xi(t) = \frac{1}{k_B T} \langle R(t)R(0) \rangle. | 
| 753 | xsun | 3354 | \label{Ineq:relationshipofXiandR} | 
| 754 | xsun | 3347 | \end{equation} | 
| 755 | xsun | 3374 | In the harmonic bath this relation is exact and provable from the | 
| 756 |  |  | definitions of these quantities. In the limit of static friction, | 
| 757 | xsun | 3347 | \begin{equation} | 
| 758 |  |  | \xi(t) = 2 \xi_0 \delta(t). | 
| 759 | xsun | 3354 | \label{Ineq:xiofStaticFriction} | 
| 760 | xsun | 3347 | \end{equation} | 
| 761 | xsun | 3374 | After substituting $\xi(t)$ into Eq.~\ref{Ineq:gle} with | 
| 762 |  |  | Eq.~\ref{Ineq:xiofStaticFriction}, the {\it Langevin Equation} is | 
| 763 |  |  | extracted, | 
| 764 | xsun | 3347 | \begin{equation} | 
| 765 |  |  | m \ddot q = -\frac{\partial U(q)}{\partial q} - \xi \dot q(t) + R(t). | 
| 766 | xsun | 3354 | \label{Ineq:langevinEquation} | 
| 767 | xsun | 3347 | \end{equation} | 
| 768 | xsun | 3374 | Application of the Langevin Equation to dynamic simulations is | 
| 769 |  |  | discussed in Ch.~\ref{chap:ld}. |