ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/matt_papers/canidacy_paper/canidacy_paper.tex
Revision: 108
Committed: Wed Sep 11 15:58:41 2002 UTC (22 years ago) by mmeineke
Content type: application/x-tex
File size: 20290 byte(s)
Log Message:
text is finished YAY!!!

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     atoms,and the head group of the lipid will typically contain charge
40     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     solvate the lipid bilayer. Because of these factors, many current
43     simulations are limited in both length and time scale due to to the
44     sheer number of calculations performed at every time step and the
45     lifetime of the researcher. A typical
46     simulation\cite{saiz02,lindahl00,venable00,Marrink01} will have around
47     64 phospholipids forming a bilayer approximately 40~$\mbox{\AA}$ by
48     50~$\mbox{\AA}$ with roughly 25 waters for every lipid. This means
49     there are on the order of 8,000 atoms needed to simulate these systems
50     and the trajectories in turn are integrated for times up to 10 ns.
51    
52     These limitations make it difficult to study certain biologically
53     interesting phenomena that don't fit within the short time and length
54     scale requirements. One such phenomena is the existence of the ripple
55     phase ($P_{\beta'}$) of the bilayer between the gel phase
56     ($L_{\beta'}$) and the fluid phase ($L_{\alpha}$). The $P_{\beta'}$
57     phase has been shown to have a ripple period of
58     100-200~$\mbox{\AA}$.\cite{katsaras00,sengupta00} A simulation of this
59     length scale would require approximately 1,300 lipid molecules and
60     roughly 25 waters for every lipid to fully solvate the bilayer. With
61     the large number of atoms involved in a simulation of this magnitude,
62     steps \emph{must} be taken to simplify the system to the point where
63     these numbers are reasonable.
64    
65     Another system of interest would be drug molecule diffusion through
66     the membrane. Due to the fluid like properties of a lipid membrane,
67     not all diffusion takes place at membrane channels. It is of interest
68     to study certain molecules that may incorporate themselves directly
69     into the membrane. These molecules may then have an appreciable
70     waiting time (on the order of nanoseconds) within the
71     bilayer. Simulation of such a long time scale again requires
72     simplification of the system in order to lower the number of
73     calculations needed at each time step.
74    
75    
76 mmeineke 95 \section{Methodology}
77    
78 mmeineke 96 \subsection{Length and Time Scale Simplifications}
79 mmeineke 95
80 mmeineke 106 The length scale simplifications are aimed at increasing the number of
81     molecules that can be simulated without drastically increasing the
82 mmeineke 95 computational cost of the system. This is done by a combination of
83     substituting less expensive interactions for expensive ones and
84     decreasing the number of interaction sites per molecule. Namely,
85 mmeineke 106 point charge distributions are replaced with dipoles, and unified atoms are
86     used in place of water, phospholipid head groups, and alkyl groups.
87 mmeineke 95
88     The replacement of charge distributions with dipoles allows us to
89 mmeineke 106 replace an interaction that has a relatively long range,
90     $(\frac{1}{r})$, for the coulomb potential, with that of a relatively
91     short range, $(\frac{1}{r^{3}})$, for dipole - dipole
92     potentials. Combined with a computational simplification algorithm
93     such as a Verlet neighbor list,\cite{allen87:csl} this should give
94     computational scaling by $N$. This is in comparison to the Ewald
95     sum\cite{leach01:mm} needed to compute the coulomb interactions, which
96     scales at best by $N \ln N$.
97 mmeineke 95
98 mmeineke 106 The second step taken to simplify the number of calculations is to
99 mmeineke 95 incorporate unified models for groups of atoms. In the case of water,
100     we use the soft sticky dipole (SSD) model developed by
101     Ichiye\cite{Liu96} (Section~\ref{sec:ssdModel}). For the phospholipids, a
102     unified head atom with a dipole will replace the atoms in the head
103     group, while unified $\text{CH}_2$ and $\text{CH}_3$ atoms will
104 mmeineke 106 replace the alkyl units in the tails (Section~\ref{sec:lipidModel}).
105 mmeineke 95
106 mmeineke 106 The time scale simplifications are introduced so that we can take
107     longer time steps. By increasing the size of the time steps taken by
108     the simulation, we are able to integrate the simulation trajectory
109     with fewer calculations. However, care must be taken that any
110     simplifications used, still conserve the total energy of the
111     simulation. In practice, this means taking steps small enough to
112     resolve all motion in the system without accidently moving an object
113     too far along a repulsive energy surface before it feels the effect of
114     the surface.
115 mmeineke 95
116 mmeineke 96 In our simulation we have chosen to constrain all bonds to be of fixed
117     length. This means the bonds are no longer allowed to vibrate about
118 mmeineke 106 their equilibrium positions. Bond vibrations are typically the fastest
119     periodic motion in a dynamics simulation. By taking this action, we
120     are able to take time steps of 3 fs while still maintaining constant
121     energy. This is in contrast to the 1 fs time step typically needed to
122     conserve energy when bonds lengths are allowed to oscillate.
123 mmeineke 95
124     \subsection{The Soft Sticky Water Model}
125     \label{sec:ssdModel}
126    
127 mmeineke 106 %\begin{floatingfigure}{55mm}
128     %\includegraphics[width=45mm]{ssd.epsi}
129     %\caption{The SSD model with the oxygen and hydrogen atoms drawn in for reference. \vspace{5mm}}
130     %\label{fig:ssdModel}
131     %\end{floatingfigure}
132 mmeineke 96
133 mmeineke 106 The water model used in our simulations is a modified soft
134     Stockmayer-sphere model.\cite{stevens95} Like the Stockmayer-sphere, the SSD
135 mmeineke 98 model\cite{Liu96} consists of a Lennard-Jones interaction site and a
136     dipole both located at the water's center of mass (Figure
137     \ref{fig:ssdModel}). However, the SSD model extends this by adding a
138     tetrahedral potential to correct for hydrogen bonding.
139    
140 mmeineke 106 The SSD water potential is then given by the following equation:
141 mmeineke 96 \begin{equation}
142 mmeineke 106 V_{\text{SSD}} = V_{\text{LJ}}(r_{i\!j}) + V_{\text{dp}}(\mathbf{r}_{i\!j},
143 mmeineke 96 \boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
144     + V_{\text{sp}}(\mathbf{r}_{i\!j},\boldsymbol{\Omega}_{i},
145     \boldsymbol{\Omega}_{j})
146 mmeineke 98 \label{eq:ssdTotPot}
147 mmeineke 96 \end{equation}
148 mmeineke 106 $V_{\text{LJ}}$ is the Lennard-Jones potential:
149     \begin{equation}
150     V_{\text{LJ}} =
151     4\epsilon_{ij} \biggl[
152     \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
153     - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
154     \biggr]
155     \label{eq:lennardJonesPot}
156     \end{equation}
157     where $r_{ij}$ is the distance between two $ij$ pairs, $\sigma_{ij}$
158 mmeineke 108 scales the length of the interaction, and $\epsilon_{ij}$ scales the
159 mmeineke 106 energy of the potential. For SSD, $\sigma_{\text{SSD}} = 3.051 \mbox{
160     \AA}$ and $\epsilon_{\text{SSD}} = 0.152\text{ kcal/mol}$.
161     $V_{\text{dp}}$ is the dipole potential:
162     \begin{equation}
163     V_{\text{dp}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
164     \boldsymbol{\Omega}_{j}) = \frac{1}{4\pi\epsilon_{0}} \biggl[
165     \frac{\boldsymbol{\mu}_{i} \cdot \boldsymbol{\mu}_{j}}{r^{3}_{ij}}
166     -
167     \frac{3(\boldsymbol{\mu}_i \cdot \mathbf{r}_{ij}) %
168     (\boldsymbol{\mu}_j \cdot \mathbf{r}_{ij}) }
169     {r^{5}_{ij}} \biggr]
170     \label{eq:dipolePot}
171     \end{equation}
172     where $\mathbf{r}_{ij}$ is the vector between $i$ and $j$,
173     $\boldsymbol{\Omega}$ is the orientation of the species, and
174     $\boldsymbol{\mu}$ is the dipole vector. The SSD model specifies a dipole
175     magnitude of 2.35~D for water.
176 mmeineke 96
177 mmeineke 101 The hydrogen bonding of the model is governed by the $V_{\text{sp}}$
178 mmeineke 106 term of the potential. Its form is as follows:
179 mmeineke 96 \begin{equation}
180     V_{\text{sp}}(\mathbf{r}_{i\!j},\boldsymbol{\Omega}_{i},
181     \boldsymbol{\Omega}_{j}) =
182     v^{\circ}[s(r_{ij})w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
183     \boldsymbol{\Omega}_{j})
184     +
185     s'(r_{ij})w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
186     \boldsymbol{\Omega}_{j})]
187 mmeineke 98 \label{eq:spPot}
188 mmeineke 96 \end{equation}
189 mmeineke 106 Where $v^\circ$ scales strength of the interaction.
190 mmeineke 98 $w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
191     and
192     $w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
193     are responsible for the tetrahedral potential and a correction to the
194     tetrahedral potential respectively. They are,
195 mmeineke 96 \begin{equation}
196     w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j}) =
197     \sin\theta_{ij} \sin 2\theta_{ij} \cos 2\phi_{ij}
198     + \sin \theta_{ji} \sin 2\theta_{ji} \cos 2\phi_{ji}
199 mmeineke 106 \label{eq:spPot2}
200 mmeineke 96 \end{equation}
201 mmeineke 98 and
202 mmeineke 96 \begin{equation}
203     \begin{split}
204 mmeineke 98 w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j}) =
205     &(\cos\theta_{ij}-0.6)^2(\cos\theta_{ij} + 0.8)^2 \\
206     &+ (\cos\theta_{ji}-0.6)^2(\cos\theta_{ji} + 0.8)^2 - 2w^{\circ}
207 mmeineke 96 \end{split}
208 mmeineke 98 \label{eq:spCorrection}
209 mmeineke 96 \end{equation}
210 mmeineke 106 The angles $\theta_{ij}$ and $\phi_{ij}$ are defined by the spherical
211     polar coordinates of the position of sphere $j$ in the reference frame
212     fixed on sphere $i$ with the z-axis aligned with the dipole moment.
213     The correction
214 mmeineke 98 $w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
215     is needed because
216     $w_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
217 mmeineke 106 vanishes when $\theta_{ij}$ is $0^\circ$ or $180^\circ$.
218 mmeineke 96
219 mmeineke 106 Finally, the sticky potential is scaled by a cutoff function,
220 mmeineke 101 $s(r_{ij})$ that scales smoothly between 0 and 1. It is represented
221     by:
222 mmeineke 96 \begin{equation}
223     s(r_{ij}) =
224     \begin{cases}
225     1& \text{if $r_{ij} < r_{L}$}, \\
226     \frac{(r_{U} - r_{ij})^2 (r_{U} + 2r_{ij}
227     - 3r_{L})}{(r_{U}-r_{L})^3}&
228     \text{if $r_{L} \leq r_{ij} \leq r_{U}$},\\
229     0& \text{if $r_{ij} \geq r_{U}$}.
230     \end{cases}
231 mmeineke 98 \label{eq:spCutoff}
232 mmeineke 96 \end{equation}
233    
234 mmeineke 106 Despite the apparent complexity of Equation \ref{eq:spPot}, the SSD
235 mmeineke 108 model is still computationally inexpensive. This is due to Equation
236 mmeineke 107 \ref{eq:spCutoff}. With $r_{L}$ being 2.75~$\mbox{\AA}$ and $r_{U}$
237 mmeineke 106 being equal to either 3.35~$\mbox{\AA}$ for $s(r_{ij})$ or
238     4.0~$\mbox{\AA}$ for $s'(r_{ij})$, the sticky potential is only active
239 mmeineke 108 over an extremely short range, and then only with other SSD
240 mmeineke 106 molecules. Therefore, it's predominant interaction is through it's
241     point dipole and Lennard-Jones sphere.
242 mmeineke 98
243 mmeineke 95 \subsection{The Phospholipid Model}
244     \label{sec:lipidModel}
245    
246 mmeineke 107 %\begin{floatingfigure}{90mm}
247     %\includegraphics[angle=-90,width=80mm]{lipidModel.epsi}
248     %\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}}
249     %\label{fig:lipidModel}
250     %\end{floatingfigure}
251 mmeineke 95
252 mmeineke 99 The lipid molecules in our simulations are unified atom models. Figure
253     \ref{fig:lipidModel} shows a template drawing for one of our
254     lipids. The Head group of the phospholipid is replaced by a single
255     Lennard-Jones sphere with a freely oriented dipole placed at it's
256     center. The magnitude of it's dipole moment is 20.6 D. The tail atoms
257 mmeineke 106 are unified $\text{CH}_2$ and $\text{CH}_3$ atoms and are also modeled
258 mmeineke 99 as Lennard-Jones spheres. The total potential for the lipid is
259     represented by Equation \ref{eq:lipidModelPot}.
260    
261     \begin{equation}
262     V_{\mbox{lipid}} = \overbrace{%
263     V_{\text{bend}}(\theta_{ijk}) +%
264     V_{\text{tors.}}(\phi_{ijkl})}^{bonded}
265     + \overbrace{%
266     V_{\text{LJ}}(r_{i\!j}) +
267     V_{\text{dp}}(r_{i\!j},\Omega_{i},\Omega_{j})%
268     }^{non-bonded}
269     \label{eq:lipidModelPot}
270     \end{equation}
271    
272 mmeineke 106 The non-bonded interactions, $V_{\text{LJ}}$ and $V_{\text{dp}}$, are
273 mmeineke 99 the Lennard-Jones and dipole-dipole interactions respectively. For the
274     non-bonded potentials, only the bend and the torsional potentials are
275     calculated. The bond potential is not calculated, and the bond lengths
276     are constrained via RATTLE.\cite{leach01:mm} The bend potential is of
277     the form:
278     \begin{equation}
279     V_{\text{bend}}(\theta_{ijk}) = k_{\theta}\frac{(\theta_{ijk} - \theta_0)^2}{2}
280     \label{eq:bendPot}
281     \end{equation}
282 mmeineke 100 Where $k_{\theta}$ sets the stiffness of the bend potential, and $\theta_0$
283     sets the equilibrium bend angle. The torsion potential was given by:
284     \begin{equation}
285 mmeineke 106 V_{\text{tors.}}(\phi_{ijkl})= c_1[1+\cos\phi_{ijkl}]
286     + c_2 [1 - \cos(2\phi_{ijkl})] + c_3[1 + \cos(3\phi_{ijkl})]
287 mmeineke 100 \label{eq:torsPot}
288     \end{equation}
289 mmeineke 102 All parameters for bonded and non-bonded potentials in the tail atoms
290     were taken from TraPPE.\cite{Siepmann1998} The bonded interactions for
291     the head atom were also taken from TraPPE, however it's dipole moment
292 mmeineke 106 and mass were based on the properties of the phosphatidylcholine head
293     group. The Lennard-Jones parameter for the head group was chosen such
294     that it was roughly twice the size of a $\text{CH}_3$ atom, and it's
295     well depth was set to be approximately equal to that of $\text{CH}_3$.
296 mmeineke 99
297 mmeineke 102 \section{Initial Simulation: 25 lipids in water}
298     \label{sec:5x5}
299 mmeineke 101
300 mmeineke 106 \subsection{Starting Configuration and Parameters}
301 mmeineke 102 \label{sec:5x5Start}
302 mmeineke 101
303 mmeineke 108 Our first simulation is an array of 25 single chained lipids in a sea
304 mmeineke 102 of water (Figure \ref{fig:5x5Start}). The total number of water
305 mmeineke 108 molecules is 1386, giving a final of water concentration of 70\%
306     wt. The simulation box measures 34.5~$\mbox{\AA}$ x 39.4~$\mbox{\AA}$
307 mmeineke 107 x 39.4~$\mbox{\AA}$ with periodic boundary conditions imposed. The
308 mmeineke 108 system is simulated in the micro-canonical (NVE) ensemble with an
309 mmeineke 102 average temperature of 300~K.
310    
311     \subsection{Results}
312     \label{sec:5x5Results}
313    
314     Figure \ref{fig:5x5Final} shows a snapshot of the system at
315 mmeineke 107 3.6~ns. Note that the system has spontaneously self assembled into a
316     bilayer. Discussion of the length scales of the bilayer will follow in
317     this section. However, it is interesting to note a key qualitative
318     property of the system revealed by this snapshot, the tail chains are
319     tilted to the bilayer normal. This is usually indicative of the gel
320     ($L_{\beta'}$) phase. In this system, the box size is probably too
321     small for the bilayer to relax to the fluid ($P_{\alpha}$) phase. This
322     demonstrates a need for an isobaric-isothermal ensemble where the box
323     size may relax or expand to keep the system at a 1~atm.
324 mmeineke 102
325 mmeineke 108 The simulation was analyzed using the radial distribution function,
326     $g(r)$, which has the form:
327 mmeineke 106 \begin{equation}
328 mmeineke 107 g(r) = \frac{V}{N_{\text{pairs}}}\langle \sum_{i} \sum_{j > i}
329 mmeineke 106 \delta(|\mathbf{r} - \mathbf{r}_{ij}|) \rangle
330     \label{eq:gofr}
331     \end{equation}
332     Equation \ref{eq:gofr} gives us information about the spacing of two
333     species as a function of radius. Essentially, if the observer were
334     located at atom $i$ and were looking out in all directions, $g(r)$
335     shows the relative density of atom $j$ at any given radius, $r$,
336     normalized by the expected density of atom $j$ at $r$. In a
337     homogeneously mixed fluid, $g(r)$ will approach 1 at large $r$, as a
338     fluid contains no long range structure to contribute peaks in the
339     number density.
340 mmeineke 102
341 mmeineke 106 For the species containing dipoles, a second pair wise distribution
342     function was used, $g_{\gamma}(r)$. It is of the form:
343     \begin{equation}
344     g_{\gamma}(r) = foobar
345     \label{eq:gammaofr}
346     \end{equation}
347     Where $\gamma_{ij}$ is the angle between the dipole of atom $j$ with
348     respect to the dipole of atom $i$. This correlation will vary between
349     $+1$ and $-1$ when the two dipoles are perfectly aligned and
350     anti-aligned respectively. This then gives us information about how
351     directional species are aligned with each other as a function of
352     distance.
353 mmeineke 102
354 mmeineke 106 Figure \ref{fig:5x5HHCorr} shows the two self correlation functions
355 mmeineke 108 for the Head groups of the lipids. The first peak in the $g(r)$ at
356     4.03~$\mbox{\AA}$ is the nearest neighbor separation of the heads of
357     two lipids. This corresponds to a maximum in the $g_{\gamma}(r)$ which
358     means that the two neighbors on the same monolayer have their dipoles
359     aligned. The broad peak at 6.5~$\mbox{\AA}$ is the inter-bilayer
360     spacing. Here, there is a corresponding anti-alignment in the angular
361     correlation. This means that although the dipoles are aligned on the
362     same monolayer, the dipoles will orient themselves to be anti-aligned
363     on a opposite facing monolayer. With this information, the two peaks
364     at 7.0~$\mbox{\AA}$ and 7.4~$\mbox{\AA}$ are head groups on the same
365     monolayer, and they are the second nearest neighbors to the head
366     group. The peak is likely a split peak because of the small statistics
367     of this system. Finally, the peak at 8.0~$\mbox{\AA}$ is likely the
368     second nearest neighbor on the opposite monolayer because of the
369     anti-alignment evident in the angular correlation.
370 mmeineke 102
371 mmeineke 108 Figure \ref{fig:5x5CCg} shows the radial distribution function for the
372     $\text{CH}_2$ unified atoms. The spacing of the atoms along the tail
373     chains accounts for the regularly spaced sharp peaks, but the broad
374     underlying peak with its maximum at 4.6~$\mbox{\AA}$ is the
375     distribution of chain-chain distances between two lipids. The final
376     Figure, Figure \ref{fig:5x5HXCorr}, includes the correlation functions
377     between the Head group and the SSD atoms. The peak in $g(r)$ at
378     3.6~$\mbox{\AA}$ is the distance of closest approach between the two,
379     and $g_{\gamma}(r)$ shows that the SSD atoms will align their dipoles
380     with the head groups at close distance. However, as one increases the
381     distance, the SSD atoms are no longer aligned.
382 mmeineke 102
383 mmeineke 108 \section{Second Simulation: 50 randomly oriented lipids in water}
384     \label{sec:r50}
385 mmeineke 102
386 mmeineke 108 \subsection{Starting Configuration and Parameters}
387     \label{sec:r50Start}
388 mmeineke 102
389 mmeineke 108 The second simulation consists of 50 single chained lipid molecules
390     embedded in a sea of 1384 SSD waters (54\% wt.). The lipids in this
391     simulation were started with random orientation and location (Figure
392     \ref{fig:r50Start} ) The simulation box measured 34.5~$\mbox{\AA}$ x
393     39.4~$\mbox{\AA}$ x 39.4~$\mbox{\AA}$ with periodic boundary conditions
394     imposed. The simulation was run in the NVE ensemble with an average
395     temperature of 300~K.
396 mmeineke 102
397 mmeineke 108 \subsection{Results}
398     \label{sec:r50Results}
399 mmeineke 102
400 mmeineke 108 Figure \ref{fig:r50Final} is a snapshot of the system at 2.0~ns. Here
401     we see that the system has already aggregated into several micelles
402     and two are even starting to merge. It will be interesting to watch as
403     this simulation continues what the total time scale for the micelle
404     aggregation and bilayer formation will be.
405    
406     Figures \ref{fig:r50HHCorr}, \ref{fig:r50CCg}, and \ref{fig:r50} are
407     the same correlation functions for the random 50 simulation as for the
408     previous simulation of 25 lipids. What is most interesting to note, is
409     the high degree of similarity between the correlation functions for
410     each system. Even though the 25 lipid simulation formed a bilayer and
411     the random 50 simulation is still in the micelle stage, both have a
412     inter surface spacing of 6.5~$\mbox{\AA}$ with the same characteristic
413     anti-alignment between surfaces. Not as surprising, is the consistency
414     of the closest packing statistics between systems. Namely, a head-head
415     closest approach distance of 4~$\mbox{\AA}$, and similar findings for
416     the chain-chain and head-water distributions as in the 25 lipid
417     system.
418    
419 mmeineke 101 \section{Future Directions}
420    
421 mmeineke 108 Current simulations indicate that our model is a feasible one, however
422     improvements will need to be made to allow the system to simulate an
423     isobaric-isothermal ensemble. This will allow the system to relax to
424     an equilibrium configuration at room temperature and pressure allowing
425     us to compare our model to experimental results. Also, we plan to
426     parallelize the code for an even greater speedup. This will allow us
427     to simulate the size systems needed to examine phenomena such as the
428     ripple phase and drug molecule diffusion
429 mmeineke 101
430 mmeineke 108 Once the work has completed on the simulation engine, we would then
431     like to use it to explore phase diagram for our model. By
432     characterizing how our model parameters affect the bilayer properties,
433     we hope to tailor our model to more closely match real biological
434     molecules. With this information, we then hope to incorporate
435     biologically relevant molecules into the system and observe their
436     transport properties across the membrane.
437    
438     \section{Acknowledgments}
439    
440     I would like to thank Dr. J.Daniel Gezelter for his guidance on this
441     project. I would also like to acknowledge the following for their help
442     and discussions during this project: Christopher Fennell, Charles
443     Vardeman, Teng Lin, Megan Sprague, Patrick Conforti, and Dan Combest.
444    
445     \pagebreak
446     \bibliographystyle{achemso}
447     \bibliography{canidacy_paper}
448     \end{document}