ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/matt_papers/canidacy_paper/canidacy_paper.tex
Revision: 110
Committed: Mon Sep 16 22:10:09 2002 UTC (21 years, 9 months ago) by mmeineke
Content type: application/x-tex
File size: 21967 byte(s)
Log Message:
made some final revisions, and started to add some figures

File Contents

# User Rev Content
1 mmeineke 95 \documentclass[11pt]{article}
2    
3     \usepackage{graphicx}
4 mmeineke 106 \usepackage{color}
5 mmeineke 98 \usepackage{floatflt}
6 mmeineke 95 \usepackage{amsmath}
7     \usepackage{amssymb}
8     \usepackage[ref]{overcite}
9    
10    
11    
12     \pagestyle{plain}
13     \pagenumbering{arabic}
14     \oddsidemargin 0.0cm \evensidemargin 0.0cm
15     \topmargin -21pt \headsep 10pt
16     \textheight 9.0in \textwidth 6.5in
17     \brokenpenalty=10000
18     \renewcommand{\baselinestretch}{1.2}
19     \renewcommand\citemid{\ } % no comma in optional reference note
20    
21    
22     \begin{document}
23    
24 mmeineke 98
25 mmeineke 95 \title{A Mesoscale Model for Phospholipid Simulations}
26    
27     \author{Matthew A. Meineke\\
28     Department of Chemistry and Biochemistry\\
29     University of Notre Dame\\
30     Notre Dame, Indiana 46556}
31    
32     \date{\today}
33     \maketitle
34    
35     \section{Background and Research Goals}
36    
37 mmeineke 106 Simulations of phospholipid bilayers are, by necessity, quite
38     complex. The lipid molecules are large molecules containing many
39 mmeineke 110 atoms, and the head group of the lipid will typically contain charge
40 mmeineke 106 separated ions which set up a large dipole within the molecule. Adding
41     to the complexity are the number of water molecules needed to properly
42 mmeineke 110 solvate the lipid bilayer, typically 25 water molecules for every
43     lipid molecule. Because of these factors, many current simulations are
44     limited in both length and time scale due to to the sheer number of
45     calculations performed at every time step and the lifetime of the
46     researcher. A typical
47 mmeineke 106 simulation\cite{saiz02,lindahl00,venable00,Marrink01} will have around
48     64 phospholipids forming a bilayer approximately 40~$\mbox{\AA}$ by
49     50~$\mbox{\AA}$ with roughly 25 waters for every lipid. This means
50     there are on the order of 8,000 atoms needed to simulate these systems
51 mmeineke 110 and the trajectories are integrated for times up to 10 ns.
52 mmeineke 106
53     These limitations make it difficult to study certain biologically
54     interesting phenomena that don't fit within the short time and length
55     scale requirements. One such phenomena is the existence of the ripple
56     phase ($P_{\beta'}$) of the bilayer between the gel phase
57     ($L_{\beta'}$) and the fluid phase ($L_{\alpha}$). The $P_{\beta'}$
58     phase has been shown to have a ripple period of
59     100-200~$\mbox{\AA}$.\cite{katsaras00,sengupta00} A simulation of this
60     length scale would require approximately 1,300 lipid molecules and
61     roughly 25 waters for every lipid to fully solvate the bilayer. With
62     the large number of atoms involved in a simulation of this magnitude,
63     steps \emph{must} be taken to simplify the system to the point where
64 mmeineke 110 the numbers of atoms becomes reasonable.
65 mmeineke 106
66     Another system of interest would be drug molecule diffusion through
67 mmeineke 110 the membrane. Due to the fluid-like properties of a lipid membrane,
68 mmeineke 106 not all diffusion takes place at membrane channels. It is of interest
69     to study certain molecules that may incorporate themselves directly
70     into the membrane. These molecules may then have an appreciable
71     waiting time (on the order of nanoseconds) within the
72     bilayer. Simulation of such a long time scale again requires
73     simplification of the system in order to lower the number of
74 mmeineke 110 calculations needed at each time step or to increase the length of
75     each time step.
76 mmeineke 106
77    
78 mmeineke 95 \section{Methodology}
79    
80 mmeineke 96 \subsection{Length and Time Scale Simplifications}
81 mmeineke 95
82 mmeineke 106 The length scale simplifications are aimed at increasing the number of
83     molecules that can be simulated without drastically increasing the
84 mmeineke 110 computational cost of the simulation. This is done through a
85     combination of substituting less expensive interactions for expensive
86     ones and decreasing the number of interaction sites per
87     molecule. Namely, point charge distributions are replaced with
88     dipoles, and unified atoms are used in place of water, phospholipid
89     head groups, and alkyl groups.
90 mmeineke 95
91     The replacement of charge distributions with dipoles allows us to
92 mmeineke 110 replace an interaction that has a relatively long range ($\frac{1}{r}$
93     for the coulomb potential) with that of a relatively short range
94     ($\frac{1}{r^{3}}$ for dipole - dipole potentials). Combined with
95     Verlet neighbor lists,\cite{allen87:csl} this should result in an
96     algorithm wich scales linearly with increasing system size. This is in
97     comparison to the Ewald sum\cite{leach01:mm} needed to compute
98     periodic replicas of the coulomb interactions, which scales at best by
99     $N \ln N$.
100 mmeineke 95
101 mmeineke 106 The second step taken to simplify the number of calculations is to
102 mmeineke 95 incorporate unified models for groups of atoms. In the case of water,
103     we use the soft sticky dipole (SSD) model developed by
104 mmeineke 110 Ichiye\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md}
105     (Section~\ref{sec:ssdModel}). For the phospholipids, a unified head
106     atom with a dipole will replace the atoms in the head group, while
107     unified $\text{CH}_2$ and $\text{CH}_3$ atoms will replace the alkyl
108     units in the tails (Section~\ref{sec:lipidModel}).
109 mmeineke 95
110 mmeineke 106 The time scale simplifications are introduced so that we can take
111     longer time steps. By increasing the size of the time steps taken by
112 mmeineke 110 the simulation, we are able to integrate a given length of time using
113     fewer calculations. However, care must be taken that any
114 mmeineke 106 simplifications used, still conserve the total energy of the
115     simulation. In practice, this means taking steps small enough to
116     resolve all motion in the system without accidently moving an object
117     too far along a repulsive energy surface before it feels the effect of
118     the surface.
119 mmeineke 95
120 mmeineke 96 In our simulation we have chosen to constrain all bonds to be of fixed
121     length. This means the bonds are no longer allowed to vibrate about
122 mmeineke 106 their equilibrium positions. Bond vibrations are typically the fastest
123     periodic motion in a dynamics simulation. By taking this action, we
124     are able to take time steps of 3 fs while still maintaining constant
125     energy. This is in contrast to the 1 fs time step typically needed to
126     conserve energy when bonds lengths are allowed to oscillate.
127 mmeineke 95
128     \subsection{The Soft Sticky Water Model}
129     \label{sec:ssdModel}
130    
131 mmeineke 110 \begin{figure}
132     \begin{center}
133     \includegraphics[width=50mm]{ssd.epsi}
134     \caption{The SSD model with the oxygen and hydrogen atoms drawn in for reference.}
135     \end{center}
136     \label{fig:ssdModel}
137     \end{figure}
138 mmeineke 96
139 mmeineke 106 The water model used in our simulations is a modified soft
140     Stockmayer-sphere model.\cite{stevens95} Like the Stockmayer-sphere, the SSD
141 mmeineke 110 model consists of a Lennard-Jones interaction site and a
142 mmeineke 98 dipole both located at the water's center of mass (Figure
143     \ref{fig:ssdModel}). However, the SSD model extends this by adding a
144     tetrahedral potential to correct for hydrogen bonding.
145    
146 mmeineke 110 The SSD water potential for a pair of water molecules is then given by
147     the following equation:
148 mmeineke 96 \begin{equation}
149 mmeineke 106 V_{\text{SSD}} = V_{\text{LJ}}(r_{i\!j}) + V_{\text{dp}}(\mathbf{r}_{i\!j},
150 mmeineke 96 \boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
151     + V_{\text{sp}}(\mathbf{r}_{i\!j},\boldsymbol{\Omega}_{i},
152     \boldsymbol{\Omega}_{j})
153 mmeineke 98 \label{eq:ssdTotPot}
154 mmeineke 96 \end{equation}
155 mmeineke 110 where $\mathbf{r}_{ij}$ is the vector between molecules $i$ and $j$,
156     and $\boldsymbol{\Omega}$ is the orientation of molecule $i$ or $j$
157     respectively. $V_{\text{LJ}}$ is the Lennard-Jones potential:
158 mmeineke 106 \begin{equation}
159     V_{\text{LJ}} =
160     4\epsilon_{ij} \biggl[
161     \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
162     - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
163     \biggr]
164     \label{eq:lennardJonesPot}
165     \end{equation}
166 mmeineke 110 here $\sigma_{ij}$
167 mmeineke 108 scales the length of the interaction, and $\epsilon_{ij}$ scales the
168 mmeineke 106 energy of the potential. For SSD, $\sigma_{\text{SSD}} = 3.051 \mbox{
169     \AA}$ and $\epsilon_{\text{SSD}} = 0.152\text{ kcal/mol}$.
170     $V_{\text{dp}}$ is the dipole potential:
171     \begin{equation}
172     V_{\text{dp}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
173     \boldsymbol{\Omega}_{j}) = \frac{1}{4\pi\epsilon_{0}} \biggl[
174     \frac{\boldsymbol{\mu}_{i} \cdot \boldsymbol{\mu}_{j}}{r^{3}_{ij}}
175     -
176     \frac{3(\boldsymbol{\mu}_i \cdot \mathbf{r}_{ij}) %
177     (\boldsymbol{\mu}_j \cdot \mathbf{r}_{ij}) }
178     {r^{5}_{ij}} \biggr]
179     \label{eq:dipolePot}
180     \end{equation}
181 mmeineke 110 where $\boldsymbol{\mu}_i$ is the dipole vector of molecule $i$,
182     $\boldsymbol{\mu}_i$ takes its orientation from
183     $\boldsymbol{\Omega}_i$. The SSD model specifies a dipole magnitude of
184     2.35~D for water.
185 mmeineke 96
186 mmeineke 110 The hydrogen bonding is modeled by the $V_{\text{sp}}$
187 mmeineke 106 term of the potential. Its form is as follows:
188 mmeineke 96 \begin{equation}
189     V_{\text{sp}}(\mathbf{r}_{i\!j},\boldsymbol{\Omega}_{i},
190     \boldsymbol{\Omega}_{j}) =
191     v^{\circ}[s(r_{ij})w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
192     \boldsymbol{\Omega}_{j})
193     +
194     s'(r_{ij})w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
195     \boldsymbol{\Omega}_{j})]
196 mmeineke 98 \label{eq:spPot}
197 mmeineke 96 \end{equation}
198 mmeineke 106 Where $v^\circ$ scales strength of the interaction.
199 mmeineke 98 $w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
200     and
201     $w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
202     are responsible for the tetrahedral potential and a correction to the
203     tetrahedral potential respectively. They are,
204 mmeineke 96 \begin{equation}
205     w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j}) =
206     \sin\theta_{ij} \sin 2\theta_{ij} \cos 2\phi_{ij}
207     + \sin \theta_{ji} \sin 2\theta_{ji} \cos 2\phi_{ji}
208 mmeineke 106 \label{eq:spPot2}
209 mmeineke 96 \end{equation}
210 mmeineke 98 and
211 mmeineke 96 \begin{equation}
212     \begin{split}
213 mmeineke 98 w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j}) =
214     &(\cos\theta_{ij}-0.6)^2(\cos\theta_{ij} + 0.8)^2 \\
215     &+ (\cos\theta_{ji}-0.6)^2(\cos\theta_{ji} + 0.8)^2 - 2w^{\circ}
216 mmeineke 96 \end{split}
217 mmeineke 98 \label{eq:spCorrection}
218 mmeineke 96 \end{equation}
219 mmeineke 106 The angles $\theta_{ij}$ and $\phi_{ij}$ are defined by the spherical
220 mmeineke 110 coordinates of the position of molecule $j$ in the reference frame
221     fixed on molecule $i$ with the z-axis aligned with the dipole moment.
222 mmeineke 106 The correction
223 mmeineke 98 $w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
224     is needed because
225     $w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
226 mmeineke 106 vanishes when $\theta_{ij}$ is $0^\circ$ or $180^\circ$.
227 mmeineke 96
228 mmeineke 106 Finally, the sticky potential is scaled by a cutoff function,
229 mmeineke 110 $s(r_{ij})$, that scales smoothly between 0 and 1. It is represented
230 mmeineke 101 by:
231 mmeineke 96 \begin{equation}
232     s(r_{ij}) =
233     \begin{cases}
234     1& \text{if $r_{ij} < r_{L}$}, \\
235     \frac{(r_{U} - r_{ij})^2 (r_{U} + 2r_{ij}
236     - 3r_{L})}{(r_{U}-r_{L})^3}&
237     \text{if $r_{L} \leq r_{ij} \leq r_{U}$},\\
238     0& \text{if $r_{ij} \geq r_{U}$}.
239     \end{cases}
240 mmeineke 98 \label{eq:spCutoff}
241 mmeineke 96 \end{equation}
242    
243 mmeineke 106 Despite the apparent complexity of Equation \ref{eq:spPot}, the SSD
244 mmeineke 108 model is still computationally inexpensive. This is due to Equation
245 mmeineke 107 \ref{eq:spCutoff}. With $r_{L}$ being 2.75~$\mbox{\AA}$ and $r_{U}$
246 mmeineke 106 being equal to either 3.35~$\mbox{\AA}$ for $s(r_{ij})$ or
247     4.0~$\mbox{\AA}$ for $s'(r_{ij})$, the sticky potential is only active
248 mmeineke 108 over an extremely short range, and then only with other SSD
249 mmeineke 110 molecules. Therefore, it's predominant interaction is through the
250     point dipole and the Lennard-Jones sphere. Of these, only the dipole
251     interaction can be considered ``long-range''.
252 mmeineke 98
253 mmeineke 95 \subsection{The Phospholipid Model}
254     \label{sec:lipidModel}
255    
256 mmeineke 110 \begin{figure}
257     \begin{center}
258     \includegraphics[angle=-90,width=80mm]{lipidModel.epsi}
259     \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. \vspace{5mm}}
260     \end{center}
261     \label{fig:lipidModel}
262     \end{figure}
263 mmeineke 95
264 mmeineke 99 The lipid molecules in our simulations are unified atom models. Figure
265 mmeineke 110 \ref{fig:lipidModel} shows a schematic for one of our
266 mmeineke 99 lipids. The Head group of the phospholipid is replaced by a single
267     Lennard-Jones sphere with a freely oriented dipole placed at it's
268 mmeineke 110 center. The magnitude of the dipole moment is 20.6 D, chosen to match
269     that of DPPC\cite{Cevc87}. The tail atoms are unified $\text{CH}_2$
270     and $\text{CH}_3$ atoms and are also modeled as Lennard-Jones
271     spheres. The total potential for the lipid is represented by Equation
272     \ref{eq:lipidModelPot}.
273 mmeineke 99
274     \begin{equation}
275 mmeineke 110 V_{\text{lipid}} =
276     \sum_{i}V_{i}^{\text{internal}}
277     + \sum_i \sum_{j>i} \sum_{\text{$\alpha$ in $i$}}
278     \sum_{\text{$\beta$ in $j$}}
279     V_{\text{LJ}}(r_{\alpha_{i}\beta_{j}})
280     +\sum_i\sum_{j>i}V_{\text{dp}}(r_{1_i,1_j},\Omega_{1_i},\Omega_{1_j})
281 mmeineke 99 \label{eq:lipidModelPot}
282     \end{equation}
283 mmeineke 110 where,
284     \begin{equation}
285     V_{i}^{\text{internal}} =
286     \sum_{\text{bends}}V_{\text{bend}}(\theta_{\alpha\beta\gamma})
287     + \sum_{\text{torsions}}V_{\text{tors.}}(\phi_{\alpha\beta\gamma\zeta})
288     + \sum_{\alpha} \sum_{\beta>\alpha}V_{\text{LJ}}(r_{\alpha \beta})
289     \label{eq:lipidModelPotInternal}
290     \end{equation}
291 mmeineke 99
292 mmeineke 106 The non-bonded interactions, $V_{\text{LJ}}$ and $V_{\text{dp}}$, are
293 mmeineke 99 the Lennard-Jones and dipole-dipole interactions respectively. For the
294 mmeineke 110 bonded potentials, only the bend and the torsional potentials are
295 mmeineke 99 calculated. The bond potential is not calculated, and the bond lengths
296     are constrained via RATTLE.\cite{leach01:mm} The bend potential is of
297     the form:
298     \begin{equation}
299 mmeineke 110 V_{\text{bend}}(\theta_{\alpha\beta\gamma})
300     = k_{\theta}\frac{(\theta_{\alpha\beta\gamma} - \theta_0)^2}{2}
301 mmeineke 99 \label{eq:bendPot}
302     \end{equation}
303 mmeineke 100 Where $k_{\theta}$ sets the stiffness of the bend potential, and $\theta_0$
304     sets the equilibrium bend angle. The torsion potential was given by:
305     \begin{equation}
306 mmeineke 110 V_{\text{tors.}}(\phi_{\alpha\beta\gamma\zeta})
307     = c_1 [1+\cos\phi_{\alpha\beta\gamma\zeta}]
308     + c_2 [1 - \cos(2\phi_{\alpha\beta\gamma\zeta})]
309     + c_3 [1 + \cos(3\phi_{\alpha\beta\gamma\zeta})]
310 mmeineke 100 \label{eq:torsPot}
311     \end{equation}
312 mmeineke 102 All parameters for bonded and non-bonded potentials in the tail atoms
313     were taken from TraPPE.\cite{Siepmann1998} The bonded interactions for
314     the head atom were also taken from TraPPE, however it's dipole moment
315 mmeineke 106 and mass were based on the properties of the phosphatidylcholine head
316     group. The Lennard-Jones parameter for the head group was chosen such
317     that it was roughly twice the size of a $\text{CH}_3$ atom, and it's
318     well depth was set to be approximately equal to that of $\text{CH}_3$.
319 mmeineke 99
320 mmeineke 102 \section{Initial Simulation: 25 lipids in water}
321     \label{sec:5x5}
322 mmeineke 101
323 mmeineke 106 \subsection{Starting Configuration and Parameters}
324 mmeineke 102 \label{sec:5x5Start}
325 mmeineke 101
326 mmeineke 110 \begin{figure}
327     \begin{center}
328     \includegraphics[width=70mm]{5x5-initial.eps}
329     \caption{The starting configuration of the 25 lipid system. A box is drawn around the periodic image.}
330     \end{center}
331     \label{fig:5x5Start}
332     \end{figure}
333    
334     \begin{figure}
335     \begin{center}
336     \includegraphics[width=70mm]{5x5-6.27ns.eps}
337     \caption{The 25 lipid system at 6.27~ns}
338     \end{center}
339     \label{fig:5x5Final}
340     \end{figure}
341    
342     Our first simulation is an array of 25 single chain lipids in a sea
343 mmeineke 102 of water (Figure \ref{fig:5x5Start}). The total number of water
344 mmeineke 108 molecules is 1386, giving a final of water concentration of 70\%
345     wt. The simulation box measures 34.5~$\mbox{\AA}$ x 39.4~$\mbox{\AA}$
346 mmeineke 107 x 39.4~$\mbox{\AA}$ with periodic boundary conditions imposed. The
347 mmeineke 108 system is simulated in the micro-canonical (NVE) ensemble with an
348 mmeineke 102 average temperature of 300~K.
349    
350     \subsection{Results}
351     \label{sec:5x5Results}
352    
353     Figure \ref{fig:5x5Final} shows a snapshot of the system at
354 mmeineke 110 6.27~ns. Note that the system has spontaneously self assembled into a
355 mmeineke 107 bilayer. Discussion of the length scales of the bilayer will follow in
356     this section. However, it is interesting to note a key qualitative
357     property of the system revealed by this snapshot, the tail chains are
358     tilted to the bilayer normal. This is usually indicative of the gel
359     ($L_{\beta'}$) phase. In this system, the box size is probably too
360     small for the bilayer to relax to the fluid ($P_{\alpha}$) phase. This
361     demonstrates a need for an isobaric-isothermal ensemble where the box
362     size may relax or expand to keep the system at a 1~atm.
363 mmeineke 102
364 mmeineke 108 The simulation was analyzed using the radial distribution function,
365     $g(r)$, which has the form:
366 mmeineke 106 \begin{equation}
367 mmeineke 107 g(r) = \frac{V}{N_{\text{pairs}}}\langle \sum_{i} \sum_{j > i}
368 mmeineke 106 \delta(|\mathbf{r} - \mathbf{r}_{ij}|) \rangle
369     \label{eq:gofr}
370     \end{equation}
371     Equation \ref{eq:gofr} gives us information about the spacing of two
372     species as a function of radius. Essentially, if the observer were
373     located at atom $i$ and were looking out in all directions, $g(r)$
374     shows the relative density of atom $j$ at any given radius, $r$,
375     normalized by the expected density of atom $j$ at $r$. In a
376     homogeneously mixed fluid, $g(r)$ will approach 1 at large $r$, as a
377     fluid contains no long range structure to contribute peaks in the
378     number density.
379 mmeineke 102
380 mmeineke 106 For the species containing dipoles, a second pair wise distribution
381     function was used, $g_{\gamma}(r)$. It is of the form:
382     \begin{equation}
383 mmeineke 110 g_{\gamma}(r) = \langle \sum_i \sum_{j>i}
384     (\cos \gamma_{ij}) \delta(| \mathbf{r} - \mathbf{r}_{ij}|) \rangle
385 mmeineke 106 \label{eq:gammaofr}
386     \end{equation}
387     Where $\gamma_{ij}$ is the angle between the dipole of atom $j$ with
388     respect to the dipole of atom $i$. This correlation will vary between
389     $+1$ and $-1$ when the two dipoles are perfectly aligned and
390     anti-aligned respectively. This then gives us information about how
391     directional species are aligned with each other as a function of
392     distance.
393 mmeineke 102
394 mmeineke 106 Figure \ref{fig:5x5HHCorr} shows the two self correlation functions
395 mmeineke 108 for the Head groups of the lipids. The first peak in the $g(r)$ at
396     4.03~$\mbox{\AA}$ is the nearest neighbor separation of the heads of
397     two lipids. This corresponds to a maximum in the $g_{\gamma}(r)$ which
398 mmeineke 110 means that the two neighbors on the same leaf have their dipoles
399     aligned. The broad peak at 6.5~$\mbox{\AA}$ is the inter-surface
400 mmeineke 108 spacing. Here, there is a corresponding anti-alignment in the angular
401     correlation. This means that although the dipoles are aligned on the
402     same monolayer, the dipoles will orient themselves to be anti-aligned
403     on a opposite facing monolayer. With this information, the two peaks
404     at 7.0~$\mbox{\AA}$ and 7.4~$\mbox{\AA}$ are head groups on the same
405     monolayer, and they are the second nearest neighbors to the head
406     group. The peak is likely a split peak because of the small statistics
407     of this system. Finally, the peak at 8.0~$\mbox{\AA}$ is likely the
408     second nearest neighbor on the opposite monolayer because of the
409     anti-alignment evident in the angular correlation.
410 mmeineke 102
411 mmeineke 108 Figure \ref{fig:5x5CCg} shows the radial distribution function for the
412     $\text{CH}_2$ unified atoms. The spacing of the atoms along the tail
413     chains accounts for the regularly spaced sharp peaks, but the broad
414     underlying peak with its maximum at 4.6~$\mbox{\AA}$ is the
415     distribution of chain-chain distances between two lipids. The final
416     Figure, Figure \ref{fig:5x5HXCorr}, includes the correlation functions
417     between the Head group and the SSD atoms. The peak in $g(r)$ at
418     3.6~$\mbox{\AA}$ is the distance of closest approach between the two,
419     and $g_{\gamma}(r)$ shows that the SSD atoms will align their dipoles
420     with the head groups at close distance. However, as one increases the
421     distance, the SSD atoms are no longer aligned.
422 mmeineke 102
423 mmeineke 108 \section{Second Simulation: 50 randomly oriented lipids in water}
424     \label{sec:r50}
425 mmeineke 102
426 mmeineke 108 \subsection{Starting Configuration and Parameters}
427     \label{sec:r50Start}
428 mmeineke 102
429 mmeineke 110 \begin{figure}
430     \begin{center}
431     \includegraphics[width=70mm]{r50-initial.eps}
432     \caption{The starting configuration of the 50 lipid system.}
433     \end{center}
434     \label{fig:r50Start}
435     \end{figure}
436    
437     \begin{figure}
438     \begin{center}
439     \includegraphics[width=70mm]{r50-2.21ns.eps}
440     \caption{The 50 lipid system at 2.21~ns}
441     \end{center}
442     \label{fig:r50Final}
443     \end{figure}
444    
445 mmeineke 108 The second simulation consists of 50 single chained lipid molecules
446     embedded in a sea of 1384 SSD waters (54\% wt.). The lipids in this
447     simulation were started with random orientation and location (Figure
448 mmeineke 110 \ref{fig:r50Start} ) The simulation box measured 26.6~$\mbox{\AA}$ x
449     26.6~$\mbox{\AA}$ x 108.4~$\mbox{\AA}$ with periodic boundary conditions
450 mmeineke 108 imposed. The simulation was run in the NVE ensemble with an average
451     temperature of 300~K.
452 mmeineke 102
453 mmeineke 108 \subsection{Results}
454     \label{sec:r50Results}
455 mmeineke 102
456 mmeineke 108 Figure \ref{fig:r50Final} is a snapshot of the system at 2.0~ns. Here
457     we see that the system has already aggregated into several micelles
458     and two are even starting to merge. It will be interesting to watch as
459     this simulation continues what the total time scale for the micelle
460     aggregation and bilayer formation will be.
461    
462     Figures \ref{fig:r50HHCorr}, \ref{fig:r50CCg}, and \ref{fig:r50} are
463     the same correlation functions for the random 50 simulation as for the
464     previous simulation of 25 lipids. What is most interesting to note, is
465 mmeineke 110 the high degree of similarity between the correlation functions
466     between systems. Even though the 25 lipid simulation formed a bilayer
467     and the random 50 simulation is still in the micelle stage, both have
468     an inter-surface spacing of 6.5~$\mbox{\AA}$ with the same
469     characteristic anti-alignment between surfaces. Not as surprising, is
470     the consistency of the closest packing statistics between
471     systems. Namely, a head-head closest approach distance of
472     4~$\mbox{\AA}$, and similar findings for the chain-chain and
473     head-water distributions as in the 25 lipid system.
474 mmeineke 108
475 mmeineke 101 \section{Future Directions}
476    
477 mmeineke 108 Current simulations indicate that our model is a feasible one, however
478 mmeineke 110 improvements will need to be made to allow the system to be simulated
479     in the isobaric-isothermal ensemble. This will relax the system to an
480     equilibrium configuration at room temperature and pressure allowing us
481     to compare our model to experimental results. Also, we are in the
482     process of parallelizeing the code for an even greater speedup. This
483     will allow us to simulate the size systems needed to examine phenomena
484     such as the ripple phase and drug molecule diffusion
485 mmeineke 101
486 mmeineke 110 Once the work has been completed on the simulation engine, we will
487     then use it to explore the phase diagram for our model. By
488 mmeineke 108 characterizing how our model parameters affect the bilayer properties,
489 mmeineke 110 we will tailor our model to more closely match real biological
490     molecules. With this information, we will then incorporate
491 mmeineke 108 biologically relevant molecules into the system and observe their
492     transport properties across the membrane.
493    
494     \section{Acknowledgments}
495    
496     I would like to thank Dr. J.Daniel Gezelter for his guidance on this
497     project. I would also like to acknowledge the following for their help
498     and discussions during this project: Christopher Fennell, Charles
499 mmeineke 110 Vardeman, Teng Lin, Megan Sprague, Patrick Conforti, and Dan
500     Combest. Funding for this project came from the National Science
501     Foundation.
502 mmeineke 108
503     \pagebreak
504     \bibliographystyle{achemso}
505     \bibliography{canidacy_paper}
506     \end{document}