--- trunk/matt_papers/canidacy_paper/canidacy_paper.tex 2002/08/21 21:25:08 95 +++ trunk/matt_papers/canidacy_paper/canidacy_paper.tex 2003/04/28 16:03:10 507 @@ -1,8 +1,12 @@ \documentclass[11pt]{article} \usepackage{graphicx} +\usepackage{color} +\usepackage{floatflt} \usepackage{amsmath} \usepackage{amssymb} +\usepackage{subfigure} +\usepackage{palatino} \usepackage[ref]{overcite} @@ -19,6 +23,7 @@ \begin{document} + \title{A Mesoscale Model for Phospholipid Simulations} \author{Matthew A. Meineke\\ @@ -31,63 +36,509 @@ Notre Dame, Indiana 46556} \section{Background and Research Goals} +Simulations of phospholipid bilayers are, by necessity, quite +complex. The lipid molecules are large molecules containing many +atoms, and the head group of the lipid will typically contain charge +separated ions which set up a large dipole within the molecule. Adding +to the complexity are the number of water molecules needed to properly +solvate the lipid bilayer, typically 25 water molecules for every +lipid molecule. Because of these factors, many current simulations are +limited in both length and time scale due to to the sheer number of +calculations performed at every time step and the lifetime of the +researcher. A typical +simulation\cite{saiz02,lindahl00,venable00,Marrink01} will have around +64 phospholipids forming a bilayer approximately 40~$\mbox{\AA}$ by +50~$\mbox{\AA}$ with roughly 25 waters for every lipid. This means +there are on the order of 8,000 atoms needed to simulate these systems +and the trajectories are integrated for times up to 10 ns. + +These limitations make it difficult to study certain biologically +interesting phenomena that don't fit within the short time and length +scale requirements. One such phenomenon is the existence of the ripple +phase ($P_{\beta'}$) of the bilayer between the gel phase +($L_{\beta'}$) and the fluid phase ($L_{\alpha}$). The $P_{\beta'}$ +phase has been shown to have a ripple period of +100-200~$\mbox{\AA}$.\cite{katsaras00,sengupta00} A simulation of this +length scale would require approximately 1,300 lipid molecules and +roughly 25 waters for every lipid to fully solvate the bilayer. With +the large number of atoms involved in a simulation of this magnitude, +steps \emph{must} be taken to simplify the system to the point where +the numbers of atoms becomes reasonable. + +Another system of interest would be drug molecule diffusion through +the membrane. Due to the fluid-like properties of a lipid membrane, +not all diffusion takes place at membrane channels. It is of interest +to study certain molecules that may incorporate themselves directly +into the membrane. These molecules may then have an appreciable +waiting time (on the order of nanoseconds) within the +bilayer. Simulation of such a long time scale again requires +simplification of the system in order to lower the number of +calculations needed at each time step or to increase the length of +each time step. + + \section{Methodology} -\subsection{Length Scale Simplifications} +\subsection{Length and Time Scale Simplifications} -The length scale simplifications are aimed at increaseing the number -of molecules simulated without drastically increasing the -computational cost of the system. This is done by a combination of -substituting less expensive interactions for expensive ones and -decreasing the number of interaction sites per molecule. Namely, -charge distributions are replaced with dipoles, and unified atoms are -used in place of water and phospholipid head groups. +The length scale simplifications are aimed at increasing the number of +molecules that can be simulated without drastically increasing the +computational cost of the simulation. This is done through a +combination of substituting less expensive interactions for expensive +ones and decreasing the number of interaction sites per +molecule. Namely, groups of point charges are replaced with single +point-dipoles, and unified atoms are used in place of water, +phospholipid head groups, and alkyl groups. -The replacement of charge distributions with dipoles allows us to -replace an interaction that has a relatively long range, $\frac{1}{r}$ -for the charge charge potential, with that of a relitively short -range, $\frac{1}{r^{3}}$ for dipole - dipole potentials -(Equations~\ref{eq:dipolePot} and \ref{eq:chargePot}). This allows us -to use computaional simplifications algorithms such as Verlet neighbor -lists,\cite{allen87:csl} which gives computaional scaling by $N$. This -is in comparison to the Ewald sum\cite{leach01:mm} needed to compute -the charge - charge interactions which scales at best by $N -\ln N$. +The replacement of charges with dipoles allows us to replace an +interaction that has a relatively long range ($\frac{1}{r}$ for the +coulomb potential) with that of a relatively short range +($\frac{1}{r^{3}}$ for dipole - dipole potentials). Combined with +Verlet neighbor lists,\cite{allen87:csl} this should result in an +algorithm which scales linearly with increasing system size. This is +in comparison to the Ewald sum\cite{leach01:mm} needed to compute +periodic replicas of the coulomb interactions, which scales at +best\cite{darden93:pme} by $N \ln N$. +The second step taken to simplify the number of calculations is to +incorporate unified models for groups of atoms. In the case of water, +we use the soft sticky dipole (SSD) model developed by +Ichiye\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md} +(Section~\ref{sec:ssdModel}). For the phospholipids, a unified head +atom with a dipole will replace the atoms in the head group, while +unified $\text{CH}_2$ and $\text{CH}_3$ atoms will replace the alkyl +units in the tails (Section~\ref{sec:lipidModel}). This model is +similar in practice to that of Lipowsky and Goetz\cite{goetz98}, where +the whole system is reduced to attractive and repulsive Lennard-Jones +spheres. However, our model gives a greater level of detail to each +unified molecule, namely through dipole, bend, and torsion +interactions. + +The time scale simplifications are introduced so that we can take +longer time steps. By increasing the size of the time steps taken by +the simulation, we are able to integrate a given length of time using +fewer calculations. However, care must be taken that any +simplifications used still conserve the total energy of the +simulation. In practice, this means taking steps small enough to +resolve all motion in the system without accidently moving an object +too far along a repulsive energy surface before it feels the effect of +the surface. + +In our simulation we have chosen to constrain all bonds to be of fixed +length. This means the bonds are no longer allowed to vibrate about +their equilibrium positions. Bond vibrations are typically the fastest +periodic motion in a dynamics simulation. By taking this action, we +are able to take time steps of 3 fs while still maintaining constant +energy. This is in contrast to the 1 fs time step typically needed to +conserve energy when bonds lengths are allowed to oscillate. + +\subsection{The Soft Sticky Dipole Water Model} +\label{sec:ssdModel} + +\begin{figure} +\centering +\includegraphics[width=50mm]{ssd.epsi} +\caption{The SSD model with the oxygen and hydrogen atoms drawn in for reference.} +\label{fig:ssdModel} +\end{figure} + +The water model used in our simulations is a modified soft +Stockmayer-sphere model.\cite{stevens95} Like the Stockmayer-sphere, the SSD +model consists of a Lennard-Jones interaction site and a +dipole both located at the water's center of mass (Figure +\ref{fig:ssdModel}). However, the SSD model extends this by adding a +tetrahedral potential to correct for hydrogen bonding. + +The SSD water potential for a pair of water molecules is then given by +the following equation: \begin{equation} -V^{\text{dp}}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i}, +V_{\text{SSD}} = V_{\text{LJ}}(r_{i\!j}) + V_{\text{dp}}(\mathbf{r}_{i\!j}, + \boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j}) + + V_{\text{sp}}(\mathbf{r}_{i\!j},\boldsymbol{\Omega}_{i}, + \boldsymbol{\Omega}_{j}) +\label{eq:ssdTotPot} +\end{equation} +where $\mathbf{r}_{ij}$ is the vector between molecules $i$ and $j$, +and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ are the +Euler angles of molecule $i$ or $j$ respectively. $V_{\text{LJ}}$ is +the Lennard-Jones potential: +\begin{equation} +V_{\text{LJ}} = + 4\epsilon_{ij} \biggl[ + \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12} + - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6} + \biggr] +\label{eq:lennardJonesPot} +\end{equation} +where $\sigma_{ij}$ scales the length of the interaction, and +$\epsilon_{ij}$ scales the energy of the potential. For SSD, +$\sigma_{\text{SSD}} = 3.051 \mbox{ +\AA}$ and $\epsilon_{\text{SSD}} = 0.152\text{ kcal/mol}$. +$V_{\text{dp}}$ is the dipole potential: +\begin{equation} +V_{\text{dp}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i}, \boldsymbol{\Omega}_{j}) = \frac{1}{4\pi\epsilon_{0}} \biggl[ \frac{\boldsymbol{\mu}_{i} \cdot \boldsymbol{\mu}_{j}}{r^{3}_{ij}} - - \frac{3(\boldsymbol{\mu} \cdot \mathbf{r}_{ij}) % - (\boldsymbol{\mu} \cdot \mathbf{r}_{ij}) }{r^{5}_{ij}} \biggr] + \frac{3(\boldsymbol{\mu}_i \cdot \mathbf{r}_{ij}) % + (\boldsymbol{\mu}_j \cdot \mathbf{r}_{ij}) } + {r^{5}_{ij}} \biggr] \label{eq:dipolePot} \end{equation} +where $\boldsymbol{\mu}_i$ is the dipole vector of molecule $i$, +$\boldsymbol{\mu}_i$ points along the z-axis in the body fixed +frame. This frame is then oriented in space by the Euler angles, +$\boldsymbol{\Omega}_i$. The SSD model specifies a dipole magnitude of +2.35~D for water. +The hydrogen bonding is modeled by the $V_{\text{sp}}$ +term of the potential. Its form is as follows: \begin{equation} -V^{\text{ch}}_{ij}(\mathbf{r}_{ij}) = \frac{q_{i}q_{j}}% - {4\pi\epsilon_{0} r_{ij}} -\label{eq:chargePot} -\end{equation} +V_{\text{sp}}(\mathbf{r}_{i\!j},\boldsymbol{\Omega}_{i}, + \boldsymbol{\Omega}_{j}) = + v^{\circ}[s(r_{ij})w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i}, + \boldsymbol{\Omega}_{j}) + + + s'(r_{ij})w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i}, + \boldsymbol{\Omega}_{j})] +\label{eq:spPot} +\end{equation} +Where $v^\circ$ scales strength of the interaction. +$w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$ +and +$w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$ +are responsible for the tetrahedral potential and a correction to the +tetrahedral potential respectively. They are, +\begin{equation} +w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j}) = + \sin\theta_{ij} \sin 2\theta_{ij} \cos 2\phi_{ij} + + \sin \theta_{ji} \sin 2\theta_{ji} \cos 2\phi_{ji} +\label{eq:spPot2} +\end{equation} +and +\begin{equation} +\begin{split} +w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j}) = + &(\cos\theta_{ij}-0.6)^2(\cos\theta_{ij} + 0.8)^2 \\ + &+ (\cos\theta_{ji}-0.6)^2(\cos\theta_{ji} + 0.8)^2 - 2w^{\circ} +\end{split} +\label{eq:spCorrection} +\end{equation} +The angles $\theta_{ij}$ and $\phi_{ij}$ are defined by the spherical +coordinates of the position of molecule $j$ in the reference frame +fixed on molecule $i$ with the z-axis aligned with the dipole moment. +The correction +$w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$ +is needed because +$w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$ +vanishes when $\theta_{ij}$ is $0^\circ$ or $180^\circ$. -The second step taken to simplify the number of calculationsis to -incorporate unified models for groups of atoms. In the case of water, -we use the soft sticky dipole (SSD) model developed by -Ichiye\cite{Liu96} (Section~\ref{sec:ssdModel}). For the phospholipids, a -unified head atom with a dipole will replace the atoms in the head -group, while unified $\text{CH}_2$ and $\text{CH}_3$ atoms will -replace the alkanes in the tails (Section~\ref{sec:lipidModel}). +Finally, the sticky potential is scaled by a switching function, +$s(r_{ij})$, that scales smoothly between 0 and 1. It is represented +by: +\begin{equation} +s(r_{ij}) = + \begin{cases} + 1& \text{if $r_{ij} < r_{L}$}, \\ + \frac{(r_{U} - r_{ij})^2 (r_{U} + 2r_{ij} + - 3r_{L})}{(r_{U}-r_{L})^3}& + \text{if $r_{L} \leq r_{ij} \leq r_{U}$},\\ + 0& \text{if $r_{ij} \geq r_{U}$}. + \end{cases} +\label{eq:spCutoff} +\end{equation} +Despite the apparent complexity of Equation \ref{eq:spPot}, the SSD +model is still computationally inexpensive. This is due to Equation +\ref{eq:spCutoff}. With $r_{L}$ being 2.75~$\mbox{\AA}$ and $r_{U}$ +being equal to either 3.35~$\mbox{\AA}$ for $s(r_{ij})$ or +4.0~$\mbox{\AA}$ for $s'(r_{ij})$, the sticky potential is only active +over an extremely short range, and then only with other SSD +molecules. Therefore, it's predominant interaction is through the +point dipole and the Lennard-Jones sphere. Of these, only the dipole +interaction can be considered ``long-range''. -\subsection{Time Scale Simplifications} - -\subsection{The Soft Sticky Water Model} -\label{sec:ssdModel} - \subsection{The Phospholipid Model} \label{sec:lipidModel} +\begin{figure} +\centering +\includegraphics[angle=-90,width=80mm]{lipidModel.epsi} +\caption{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ is the bend angle, $\mu$ is the dipole moment of the head group, and n is the chain length.} +\label{fig:lipidModel} +\end{figure} -\bibliographystyle{achemso} -\bibliography{canidacy_paper} -\end{document} \ No newline at end of file +The lipid molecules in our simulations are unified atom models. Figure +\ref{fig:lipidModel} shows a schematic for one of our +lipids. The head group of the phospholipid is replaced by a single +Lennard-Jones sphere with a freely oriented dipole placed at it's +center. The magnitude of the dipole moment is 20.6 D, chosen to match +that of DPPC\cite{Cevc87}. The tail atoms are unified $\text{CH}_2$ +and $\text{CH}_3$ atoms and are also modeled as Lennard-Jones +spheres. The total potential for the lipid is represented by Equation +\ref{eq:lipidModelPot}. + +\begin{equation} +V_{\text{lipid}} = + \sum_{i}V_{i}^{\text{internal}} + + \sum_i \sum_{j>i} \sum_{\alpha_i} + \sum_{\beta_j} + V_{\text{LJ}}(r_{\alpha_{i}\beta_{j}}) + +\sum_i\sum_{j>i}V_{\text{dp}}(r_{1_i,1_j},\Omega_{1_i},\Omega_{1_j}) +\label{eq:lipidModelPot} +\end{equation} +where, +\begin{equation} +V_{i}^{\text{internal}} = + \sum_{\text{bends}}V_{\text{bend}}(\theta_{\alpha\beta\gamma}) + + \sum_{\text{torsions}}V_{\text{tors.}}(\phi_{\alpha\beta\gamma\zeta}) + + \sum_{\alpha_i} \sum_{\beta_i > (\alpha_i + 4)}V_{\text{LJ}} + (r_{\alpha_i \beta_i}) +\label{eq:lipidModelPotInternal} +\end{equation} + +The non-bonded interactions, $V_{\text{LJ}}$ and $V_{\text{dp}}$, are +the Lennard-Jones and dipole-dipole interactions respectively. For the +bonded potentials, only the bend and the torsional potentials are +calculated. The bond potential is not calculated, and the bond lengths +are constrained via RATTLE.\cite{leach01:mm} The bend potential is of +the form: +\begin{equation} +V_{\text{bend}}(\theta_{\alpha\beta\gamma}) + = k_{\theta}\frac{(\theta_{\alpha\beta\gamma} - \theta_0)^2}{2} +\label{eq:bendPot} +\end{equation} +Where $k_{\theta}$ sets the stiffness of the bend potential, and $\theta_0$ +sets the equilibrium bend angle. The torsion potential was given by: +\begin{equation} +V_{\text{tors.}}(\phi_{\alpha\beta\gamma\zeta}) + = c_1 [1+\cos\phi_{\alpha\beta\gamma\zeta}] + + c_2 [1 - \cos(2\phi_{\alpha\beta\gamma\zeta})] + + c_3 [1 + \cos(3\phi_{\alpha\beta\gamma\zeta})] +\label{eq:torsPot} +\end{equation} +All parameters for bonded and non-bonded potentials in the tail atoms +were taken from TraPPE.\cite{Siepmann1998} The bonded interactions for +the head atom were also taken from TraPPE, however it's dipole moment +and mass were based on the properties of the phosphatidylcholine head +group. The Lennard-Jones parameter for the head group was chosen such +that it was roughly twice the size of a $\text{CH}_3$ atom, and it's +well depth was set to be approximately equal to that of $\text{CH}_3$. + +\section{Initial Simulation: 25 lipids in water} +\label{sec:5x5} + +\subsection{Starting Configuration and Parameters} +\label{sec:5x5Start} + +\begin{figure} +\centering +\mbox{ + \subfigure[The starting configuration of the 25 lipid system.]{% + \label{fig:5x5Start}% + \includegraphics[width=70mm]{5x5-initial.eps}}\quad + \subfigure[The 25 lipid system at 6.27~ns.]{% + \label{fig:5x5Final}% + \includegraphics[width=70mm]{5x5-6.27ns.eps}} +} +\caption{Snapshots of the 25 lipid system. Carbon tail atoms are drawn in gray, the phospholipid head groups are colored blue, and the waters are scaled down for visibility. A box has been drawn around the periodic image.} +\end{figure} + +Our first simulation is an array of 25 single chain lipids in a sea +of water (Figure \ref{fig:5x5Start}). The total number of water +molecules is 1386, giving a final of water concentration of 70\% +wt. The simulation box measures 34.5~$\mbox{\AA}$ x 39.4~$\mbox{\AA}$ +x 39.4~$\mbox{\AA}$ with periodic boundary conditions imposed. The +system is simulated in the micro-canonical (NVE) ensemble with an +average temperature of 300~K. + +\subsection{Results} +\label{sec:5x5Results} + +\begin{figure} +\centering + \subfigure[The self correlation of the phospholipid head groups. $g(r)$ is on the top, the bottom chart is the $g_\gamma(r)$.]{% + \label{fig:5x5HHCorr}% + \includegraphics[angle=-90,width=80mm]{all5x5-HEAD-HEAD.epsi}% + } + \subfigure[The $g(r)$ for the $\text{CH}_2$ molecules in the chain tails]{% + \label{fig:5x5CCg}% + \includegraphics[angle=-90,width=80mm]{all5x5-CH2-CH2.epsi}} + \subfigure% + [The pair correlations between the head groups and the water]{% + \label{fig:5x5HXCorr}% + \includegraphics[angle=-90,width=80mm]{all5x5-HEAD-X.epsi}} +\caption{The pair correlation functions for the 25 lipid system} +\end{figure} + + +Figure \ref{fig:5x5Final} shows a snapshot of the system at +6.27~ns. Note that the system has spontaneously self assembled into a +bilayer. Discussion of the length scales of the bilayer will follow in +this section. However, it is interesting to note a key qualitative +property of the system revealed by this snapshot, the tail chains are +tilted to the bilayer normal. This is usually indicative of the gel +($L_{\beta'}$) phase. In this system, the box size is probably too +small for the bilayer to relax to the fluid ($P_{\alpha}$) phase. This +demonstrates a need for an isobaric-isothermal ensemble where the box +size may relax or expand to keep the system at 1~atm. + +The simulation was analyzed using the radial distribution function, +$g(r)$, which has the form: +\begin{equation} +g(r) = \frac{V}{N_{\text{pairs}}}\langle \sum_{i} \sum_{j > i} + \delta(|\mathbf{r} - \mathbf{r}_{ij}|) \rangle +\label{eq:gofr} +\end{equation} +Equation \ref{eq:gofr} gives us information about the spacing of two +species as a function of radius. Essentially, if the observer were +located at atom $i$ and were looking out in all directions, $g(r)$ +shows the relative density of atom $j$ at any given radius, $r$, +normalized by the expected density of atom $j$ at $r$. In a +homogeneously mixed fluid, $g(r)$ will approach 1 at large $r$, as a +fluid contains no long range structure to contribute peaks in the +number density. + +For the species containing dipoles, a second pair-wise distribution +function was used, $g_{\gamma}(r)$. It is of the form: +\begin{equation} +g_{\gamma}(r) = \langle \sum_i \sum_{j>i} + (\cos \gamma_{ij}) \delta(| \mathbf{r} - \mathbf{r}_{ij}|) \rangle +\label{eq:gammaofr} +\end{equation} +Where $\gamma_{ij}$ is the angle between the dipole of atom $j$ with +respect to the dipole of atom $i$. This correlation will vary between +$+1$ and $-1$ when the two dipoles are perfectly aligned and +anti-aligned respectively. This then gives us information about how +directional species are aligned with each other as a function of +distance. + +Figure \ref{fig:5x5HHCorr} shows the two self correlation functions +for the Head groups of the lipids. The first peak in the $g(r)$ at +4.03~$\mbox{\AA}$ is the nearest neighbor separation of the heads of +two lipids. This corresponds to a maximum in the $g_{\gamma}(r)$ which +means that the two neighbors on the same leaf have their dipoles +aligned. The broad peak at 6.5~$\mbox{\AA}$ is the inter-surface +spacing. Here, there is a corresponding anti-alignment in the angular +correlation. This means that although the dipoles are aligned on the +same monolayer, the dipoles will orient themselves to be anti-aligned +on a opposite facing monolayer. With this information, the two peaks +at 7.0~$\mbox{\AA}$ and 7.4~$\mbox{\AA}$ are head groups on the same +monolayer, and they are the second nearest neighbors to the head +group. The peak is likely a split peak because of the small statistics +of this system. Finally, the peak at 8.0~$\mbox{\AA}$ is likely the +second nearest neighbor on the opposite monolayer because of the +anti-alignment evident in the angular correlation. + +Figure \ref{fig:5x5CCg} shows the radial distribution function for the +$\text{CH}_2$ unified atoms. The spacing of the atoms along the tail +chains accounts for the regularly spaced sharp peaks, but the broad +underlying peak with its maximum at 4.6~$\mbox{\AA}$ is the +distribution of chain-chain distances between two lipids. The final +Figure, Figure \ref{fig:5x5HXCorr}, includes the correlation functions +between the Head group and the SSD atoms. The peak in $g(r)$ at +3.6~$\mbox{\AA}$ is the distance of closest approach between the two, +and $g_{\gamma}(r)$ shows that the SSD atoms will align their dipoles +with the head groups at close distance. However, as one increases the +distance, the SSD atoms are no longer aligned. + +\section{Second Simulation: 50 randomly oriented lipids in water} +\label{sec:r50} + +\subsection{Starting Configuration and Parameters} +\label{sec:r50Start} + +\begin{figure} +\centering +\mbox{ + \subfigure[The starting configuration of the 50 lipid system.]{% + \label{fig:r50Start}% + \includegraphics[width=70mm]{r50-initial.eps}}\quad + \subfigure[The 50 lipid system at 2.21~ns]{% + \label{fig:r50Final}% + \includegraphics[width=70mm]{r50-2.21ns.eps}} +} +\caption{Snapshots of the 50 lipid system} +\end{figure} + +The second simulation consists of 50 single chained lipid molecules +embedded in a sea of 1384 SSD waters (54\% wt.). The lipids in this +simulation were started with random orientation and location (Figure +\ref{fig:r50Start} ) The simulation box measured 26.6~$\mbox{\AA}$ x +26.6~$\mbox{\AA}$ x 108.4~$\mbox{\AA}$ with periodic boundary conditions +imposed. The simulation was run in the NVE ensemble with an average +temperature of 300~K. + +\subsection{Results} +\label{sec:r50Results} + +\begin{figure} +\centering + \subfigure[The self correlation of the phospholipid head groups.]{% + \label{fig:r50HHCorr}% + \includegraphics[angle=-90,width=80mm]{r50-HEAD-HEAD.epsi}% + } + \subfigure% + [The pair correlations between the head groups and the water]{% + \label{fig:r50HXCorr}% + \includegraphics[angle=-90,width=80mm]{r50-HEAD-X.epsi}% + } + \subfigure[The $g(r)$ for the $\text{CH}_2$ molecules in the chain tails]{% + \label{fig:r50CCg}% + \includegraphics[angle=-90,width=80mm]{r50-CH2-CH2.epsi}} + +\caption{The pair correlation functions for the 50 lipid system} +\end{figure} + +Figure \ref{fig:r50Final} is a snapshot of the system at 2.21~ns. Here +we see that the system has already aggregated into several micelles +and two are even starting to merge. It will be interesting to watch as +this simulation continues what the total time scale for the micelle +aggregation and bilayer formation will be, in Marrink's\cite{Marrink01} +simulation, bilayer aggregation is predicted to occur around 10~ns. + +Figures \ref{fig:r50HHCorr}, \ref{fig:r50HXCorr}, and \ref{fig:r50CCg} are +the same correlation functions for the random 50 simulation as for the +previous simulation of 25 lipids. What is most interesting to note, is +the high degree of similarity between the correlation functions +between systems. Even though the 25 lipid simulation formed a bilayer +and the random 50 simulation is still in the micelle stage, both have +an inter-surface spacing of 6.5~$\mbox{\AA}$ with the same +characteristic anti-alignment between surfaces. Not as surprising, is +the consistency of the closest packing statistics between +systems. Namely, a head-head closest approach distance of +4~$\mbox{\AA}$, and similar findings for the chain-chain and +head-water distributions as in the 25 lipid system. + +\section{Future Directions} + +Current simulations indicate that our model is a feasible one, however +improvements will need to be made to allow the system to be simulated +in the isobaric-isothermal ensemble. This will relax the system to an +equilibrium configuration at room temperature and pressure allowing us +to compare our model to experimental results. Also, we are in the +process of parallelizing the code for an even greater speedup. This +will allow us to simulate large enough systems to examine phenomena +such as the ripple phase and drug molecule diffusion + +Once the work has been completed on the simulation engine, we will +then use it to explore the phase diagram for our model. By +characterizing how our model parameters affect the bilayer properties, +we will tailor our model to more closely match real biological +molecules. With this information, we will then incorporate +biologically relevant molecules into the system and observe their +transport properties across the membrane. + +\section{Acknowledgments} + +I would like to thank Dr.~J.~Daniel Gezelter for his guidance on this +project. I would also like to acknowledge the following for their help +and discussions during this project: Christopher Fennell, Charles +Vardeman, Teng Lin, Megan Sprague, Patrick Conforti, and Dan +Combest. Funding for this project came from the National Science +Foundation. + +\pagebreak +\bibliographystyle{achemso} +\bibliography{canidacy_paper} +\end{document}