ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/lipid.tex
Revision: 1087
Committed: Fri Mar 5 22:16:34 2004 UTC (20 years, 4 months ago) by mmeineke
Content type: application/x-tex
File size: 28993 byte(s)
Log Message:
semi-final draft

File Contents

# Content
1
2
3 \chapter{\label{chapt:lipid}Phospholipid Simulations}
4
5 \section{\label{lipidSec:Intro}Introduction}
6
7 In the past 10 years, increasing computer speeds have allowed for the
8 atomistic simulation of phospholipid bilayers for increasingly
9 relevant lengths of time. These simulations have ranged from
10 simulation of the gel phase ($L_{\beta}$) of
11 dipalmitoylphosphatidylcholine (DPPC),\cite{lindahl00} to the
12 spontaneous aggregation of DPPC molecules into fluid phase
13 ($L_{\alpha}$) bilayers.\cite{marrink01} With the exception of a few
14 ambitious
15 simulations,\cite{marrink01:undulation,marrink:2002,lindahl00} most
16 investigations are limited to a range of 64 to 256
17 phospholipids.\cite{lindahl00,sum:2003,venable00,gomez:2003,smondyrev:1999,marrink01}
18 The expense of the force calculations involved when performing these
19 simulations limits the system size. To properly hydrate a bilayer, one
20 typically needs 25 water molecules for every lipid, bringing the total
21 number of atoms simulated to roughly 8,000 for a system of 64 DPPC
22 molecules. Added to the difficulty is the electrostatic nature of the
23 phospholipid head groups and water, requiring either the
24 computationally expensive Ewald sum or the faster, particle mesh Ewald
25 sum.\cite{nina:2002,norberg:2000,patra:2003} These factors all limit
26 the system size and time scales of bilayer simulations.
27
28 Unfortunately, much of biological interest happens on time and length
29 scales well beyond the range of current simulation technology. One
30 such example is the observance of a ripple phase
31 ($P_{\beta^{\prime}}$) between the $L_{\beta}$ and $L_{\alpha}$ phases
32 of certain phospholipid bilayers.\cite{katsaras00,sengupta00} These
33 ripples are shown to have periodicity on the order of
34 100-200~$\mbox{\AA}$. A simulation on this length scale would have
35 approximately 1,300 lipid molecules with an additional 25 water
36 molecules per lipid to fully solvate the bilayer. A simulation of this
37 size is impractical with current atomistic models.
38
39 The time and length scale limitations are most striking in transport
40 phenomena. Due to the fluid-like properties of a lipid membrane, not
41 all diffusion across the membrane happens at pores. Some molecules of
42 interest may incorporate themselves directly into the membrane. Once
43 here, they may possess an appreciable waiting time (on the order of
44 10's to 100's of nanoseconds) within the bilayer. Such long simulation
45 times are nearly impossible to obtain when integrating the system with
46 atomistic detail.
47
48 To address these issues, several schemes have been proposed. One
49 approach by Goetz and Liposky\cite{goetz98} is to model the entire
50 system as Lennard-Jones spheres. Phospholipids are represented by
51 chains of beads with the top most beads identified as the head
52 atoms. Polar and non-polar interactions are mimicked through
53 attractive and soft-repulsive potentials respectively. A similar
54 model proposed by Marrinck \emph{et. al}.\cite{marrink04}~uses a
55 similar technique for modeling polar and non-polar interactions with
56 Lennard-Jones spheres. However, they also include charges on the head
57 group spheres to mimic the electrostatic interactions of the
58 bilayer. While the solvent spheres are kept charge-neutral and
59 interact with the bilayer solely through an attractive Lennard-Jones
60 potential.
61
62 The model used in this investigation adds more information to the
63 interactions than the previous two models, while still balancing the
64 need for simplifications over atomistic detail. The model uses
65 Lennard-Jones spheres for the head and tail groups of the
66 phospholipids, allowing for the ability to scale the parameters to
67 reflect various sized chain configurations while keeping the number of
68 interactions small. What sets this model apart, however, is the use
69 of dipoles to represent the electrostatic nature of the
70 phospholipids. The dipole electrostatic interaction is shorter range
71 than Coulombic ($\frac{1}{r^3}$ versus $\frac{1}{r}$), and eliminates
72 the need for a costly Ewald sum.
73
74 Another key feature of this model, is the use of a dipolar water model
75 to represent the solvent. The soft sticky dipole ({\sc ssd}) water
76 \cite{liu96:new_model} relies on the dipole for long range electrostatic
77 effects, but also contains a short range correction for hydrogen
78 bonding. In this way the systems in this research mimic the entropic
79 contribution to the hydrophobic effect due to hydrogen-bond network
80 deformation around a non-polar entity, \emph{i.e.}~the phospholipid
81 molecules.
82
83 The following is an outline of this chapter.
84 Sec.~\ref{lipidSec:Methods} is an introduction to the lipid model used
85 in these simulations, as well as clarification about the water model
86 and integration techniques. The various simulations explored in this
87 research are outlined in
88 Sec.~\ref{lipidSec:ExpSetup}. Sec.~\ref{lipidSec:resultsDis} gives a
89 summary and interpretation of the results. Finally, the conclusions
90 of this chapter are presented in Sec.~\ref{lipidSec:Conclusion}.
91
92 \section{\label{lipidSec:Methods}Methods}
93
94 \subsection{\label{lipidSec:lipidModel}The Lipid Model}
95
96 \begin{figure}
97 \centering
98 \includegraphics[width=\linewidth]{twoChainFig.eps}
99 \caption[The two chained lipid model]{Schematic diagram of the double chain phospholipid model. The head group (in red) has a point dipole, $\boldsymbol{\mu}$, located at its center of mass. The two chains are eight methylene groups in length.}
100 \label{lipidFig:lipidModel}
101 \end{figure}
102
103 The phospholipid model used in these simulations is based on the
104 design illustrated in Fig.~\ref{lipidFig:lipidModel}. The head group
105 of the phospholipid is replaced by a single Lennard-Jones sphere of
106 diameter $\sigma_{\text{head}}$, with $\epsilon_{\text{head}}$ scaling
107 the well depth of its van der Walls interaction. This sphere also
108 contains a single dipole of magnitude $|\boldsymbol{\mu}|$, where
109 $|\boldsymbol{\mu}|$ can be varied to mimic the charge separation of a
110 given phospholipid head group. The atoms of the tail region are
111 modeled by beads representing multiple methyl groups. They are free
112 of partial charges or dipoles, and contain only Lennard-Jones
113 interaction sites at their centers of mass. As with the head groups,
114 their potentials can be scaled by $\sigma_{\text{tail}}$ and
115 $\epsilon_{\text{tail}}$.
116
117 The long range interactions between lipids are given by the following:
118 \begin{equation}
119 V_{\text{LJ}}(r_{ij}) =
120 4\epsilon_{ij} \biggl[
121 \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
122 - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
123 \biggr]
124 \label{lipidEq:LJpot}
125 \end{equation}
126 and
127 \begin{equation}
128 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
129 \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
130 \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
131 -
132 3(\boldsymbol{\hat{u}}_i \cdot \mathbf{\hat{r}}_{ij}) %
133 (\boldsymbol{\hat{u}}_j \cdot \mathbf{\hat{r}}_{ij} \biggr]
134 \label{lipidEq:dipolePot}
135 \end{equation}
136 Where $V_{\text{LJ}}$ is the Lennard-Jones potential and
137 $V_{\text{dipole}}$ is the dipole-dipole potential. As previously
138 stated, $\sigma_{ij}$ and $\epsilon_{ij}$ are the Lennard-Jones
139 parameters which scale the length and depth of the interaction
140 respectively, and $r_{ij}$ is the distance between beads $i$ and $j$.
141 In $V_{\text{dipole}}$, $\mathbf{r}_{ij}$ is the vector starting at
142 bead $i$ and pointing towards bead $j$. Vectors $\mathbf{\Omega}_i$
143 and $\mathbf{\Omega}_j$ are the orientational degrees of freedom for
144 beads $i$ and $j$. $|\mu_i|$ is the magnitude of the dipole moment of
145 $i$, and $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
146 vector rotated with Euler angles: $\boldsymbol{\Omega}_i$.
147
148 The model also allows for the bonded interactions bends, and torsions.
149 The bond between two beads on a chain is of fixed length, and is
150 maintained according to the {\sc rattle} algorithm.\cite{andersen83}
151 The bends are subject to a harmonic potential:
152 \begin{equation}
153 V_{\text{bend}}(\theta) = k_{\theta}( \theta - \theta_0 )^2
154 \label{lipidEq:bendPot}
155 \end{equation}
156 where $k_{\theta}$ scales the strength of the harmonic well, and
157 $\theta$ is the angle between bond vectors
158 (Fig.~\ref{lipidFig:lipidModel}). In addition, we have placed a
159 ``ghost'' bend on the phospholipid head. The ghost bend adds a
160 potential to keep the dipole pointed along the bilayer surface, where
161 $\theta$ is now the angle the dipole makes with respect to the {\sc
162 head}-$\text{{\sc ch}}_2$ bond vector
163 (Fig.~\ref{lipidFig:ghostBend}). The torsion potential is given by:
164 \begin{equation}
165 V_{\text{torsion}}(\phi) =
166 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0
167 \label{lipidEq:torsionPot}
168 \end{equation}
169 Here, the parameters $k_0$, $k_1$, $k_2$, and $k_3$ fit the cosine
170 power series to the desired torsion potential surface, and $\phi$ is
171 the angle the two end atoms have rotated about the middle bond
172 (Fig.:\ref{lipidFig:lipidModel}). Long range interactions such as the
173 Lennard-Jones potential are excluded for atom pairs involved in the
174 same bond, bend, or torsion. However, internal interactions not
175 directly involved in a bonded pair are calculated.
176
177 \begin{figure}
178 \centering
179 \includegraphics[width=\linewidth]{ghostBendFig.eps}
180 \caption[Depiction of the ``ghost'' bend]{The ``ghost'' bend is a bend potential added to constrain the motion of the dipole on the {\sc head} group. The potential follows Eq.~\ref{lipidEq:bendPot} where $\theta$ is now the angle that the dipole makes with the {\sc head}-$\text{{\sc ch}}_2$ bond vector.}
181 \label{lipidFig:ghostBend}
182 \end{figure}
183
184 All simulations presented here use a two chained lipid as pictured in
185 Fig.~\ref{lipidFig:lipidModel}. The chains are both eight beads long,
186 and their mass and Lennard Jones parameters are summarized in
187 Table~\ref{lipidTable:tcLJParams}. The magnitude of the dipole moment
188 for the head bead is 10.6~Debye, and the bend and torsion parameters
189 are summarized in Table~\ref{lipidTable:tcBendParams} and
190 \ref{lipidTable:tcTorsionParams}.
191
192 \begin{table}
193 \caption{The Lennard Jones Parameters for the two chain phospholipids.}
194 \label{lipidTable:tcLJParams}
195 \begin{center}
196 \begin{tabular}{|l|c|c|c|}
197 \hline
198 & mass (amu) & $\sigma$($\mbox{\AA}$) & $\epsilon$ (kcal/mol) \\ \hline
199 {\sc head} & 72 & 4.0 & 0.185 \\ \hline
200 {\sc ch}\cite{Siepmann1998} & 13.02 & 4.0 & 0.0189 \\ \hline
201 $\text{{\sc ch}}_2$\cite{Siepmann1998} & 14.03 & 3.95 & 0.18 \\ \hline
202 $\text{{\sc ch}}_3$\cite{Siepmann1998} & 15.04 & 3.75 & 0.25 \\ \hline
203 {\sc ssd}\cite{liu96:new_model} & 18.03 & 3.051 & 0.152 \\ \hline
204 \end{tabular}
205 \end{center}
206 \end{table}
207
208 \begin{table}
209 \caption[Bend Parameters for the two chain phospholipids]{Bend Parameters for the two chain phospholipids. All alkane parameters are based off of those from TraPPE.\cite{Siepmann1998}}
210 \label{lipidTable:tcBendParams}
211 \begin{center}
212 \begin{tabular}{|l|c|c|}
213 \hline
214 & $k_{\theta}$ ( kcal/($\text{mol deg}^2$) ) & $\theta_0$ ( deg ) \\ \hline
215 {\sc ghost}-{\sc head}-$\text{{\sc ch}}_2$ & 0.00177 & 129.78 \\ \hline
216 $x$-{\sc ch}-$y$ & 58.84 & 112.0 \\ \hline
217 $x$-$\text{{\sc ch}}_2$-$y$ & 58.84 & 114.0 \\ \hline
218 \end{tabular}
219 \end{center}
220 \end{table}
221
222 \begin{table}
223 \caption[Torsion Parameters for the two chain phospholipids]{Torsion Parameters for the two chain phospholipids. Alkane parameters based on TraPPE.\cite{Siepmann1998}}
224 \label{lipidTable:tcTorsionParams}
225 \begin{center}
226 \begin{tabular}{|l|c|c|c|c|}
227 \hline
228 All are in kcal/mol $\rightarrow$ & $k_3$ & $k_2$ & $k_1$ & $k_0$ \\ \hline
229 $x$-{\sc ch}-$y$-$z$ & 3.3254 & -0.4215 & -1.686 & 1.1661 \\ \hline
230 $x$-$\text{{\sc ch}}_2$-$\text{{\sc ch}}_2$-$y$ & 5.9602 & -0.568 & -3.802 & 2.1586 \\ \hline
231 \end{tabular}
232 \end{center}
233 \end{table}
234
235
236 \section{\label{lipidSec:furtherMethod}Further Methodology}
237
238 As mentioned previously, the water model used throughout these
239 simulations was the {\sc ssd} model of
240 Ichiye.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md} A
241 discussion of the model can be found in Sec.~\ref{oopseSec:SSD}. As
242 for the integration of the equations of motion, all simulations were
243 performed in an orthorhombic periodic box with a thermostat on
244 velocities, and an independent barostat on each Cartesian axis $x$,
245 $y$, and $z$. This is the $\text{NPT}_{xyz}$. ensemble described in
246 Sec.~\ref{oopseSec:integrate}.
247
248
249 \subsection{\label{lipidSec:ExpSetup}Experimental Setup}
250
251 Two main starting configuration classes were used in this research:
252 random and ordered bilayers. The ordered bilayer starting
253 configurations were all started from an equilibrated bilayer at
254 300~K. The original configuration for the first 300~K run was
255 assembled by placing the phospholipids centers of mass on a planar
256 hexagonal lattice. The lipids were oriented with their long axis
257 perpendicular to the plane. The second leaf simply mirrored the first
258 leaf, and the appropriate number of waters were then added above and
259 below the bilayer.
260
261 The random configurations took more work to generate. To begin, a
262 test lipid was placed in a simulation box already containing water at
263 the intended density. The waters were then tested for overlap with
264 the lipid using a 5.0~$\mbox{\AA}$ buffer distance. This gave an
265 estimate for the number of waters each lipid would displace in a
266 simulation box. A target number of waters was then defined which
267 included the number of waters each lipid would displace, the number of
268 waters desired to solvate each lipid, and a factor to pad the
269 initial box with a few extra water molecules.
270
271 Next, a cubic simulation box was created that contained at least the
272 target number of waters in an FCC lattice (the lattice was for ease of
273 placement). What followed was a RSA simulation similar to those of
274 Chapt.~\ref{chapt:RSA}. The lipids were sequentially given a random
275 position and orientation within the box. If a lipid's position caused
276 atomic overlap with any previously placed lipid, its position and
277 orientation were rejected, and a new random placement site was
278 attempted. The RSA simulation proceeded until all phospholipids had
279 been adsorbed. After placement of all lipid molecules, water
280 molecules with locations that overlapped with the atomic coordinates
281 of the lipids were removed.
282
283 Finally, water molecules were removed at random until the desired
284 number of waters per lipid was reached. The typical low final density
285 for these initial configurations was not a problem, as the box shrinks
286 to an appropriate size within the first 50~ps of a simulation in the
287 $\text{NPT}_{xyz}$ ensemble.
288
289 \subsection{\label{lipidSec:Configs}Configurations}
290
291 The first class of simulations were started from ordered
292 bilayers. They were all configurations consisting of 60 lipid
293 molecules with 30 lipids on each leaf, and were hydrated with 1620
294 {\sc ssd} molecules. The original configuration was assembled
295 according to Sec.~\ref{lipidSec:ExpSetup} and simulated for a length
296 of 10~ns at 300~K. The other temperature runs were started from a
297 frame 7~ns into the 300~K simulation. Their temperatures were reset
298 with the thermostating algorithm in the $\text{NPT}_{xyz}$
299 integrator. All of the temperature variants were also run for 10~ns,
300 with only the last 5~ns being used for accumulation of statistics.
301
302 The second class of simulations were two configurations started from
303 randomly dispersed lipids in a ``gas'' of water. The first
304 ($\text{R}_{\text{I}}$) was a simulation containing 72 lipids with
305 1800 {\sc ssd} molecules simulated at 300~K. The second
306 ($\text{R}_{\text{II}}$) was 90 lipids with 1350 {\sc ssd} molecules
307 simulated at 350~K. Both simulations were integrated for more than
308 20~ns, and illustrate the spontaneous aggregation of the lipid model
309 into phospholipid macro-structures: $\text{R}_{\text{I}}$ into a
310 bilayer, and $\text{R}_{\text{II}}$ into a inverted rod.
311
312 \section{\label{lipidSec:resultsDis}Results and Discussion}
313
314 \subsection{\label{lipidSec:diffusion}Lateral Diffusion Constants}
315
316 The lateral diffusion constant, $D_L$, is the constant characterizing
317 the diffusive motion of the lipid within the plane of the bilayer. It
318 is given by the following Einstein relation valid at long
319 times:\cite{allen87:csl}
320 \begin{equation}
321 2tD_L = \frac{1}{2}\langle |\mathbf{r}(t) - \mathbf{r}(0)|^2\rangle
322 \end{equation}
323 Where $\mathbf{r}(t)$ is the position of the lipid at time $t$, and is
324 constrained to lie within a plane. For the bilayer simulations the
325 plane of constrained motion was that perpendicular to the bilayer
326 normal, namely the $xy$-plane.
327
328 Fig.~\ref{lipidFig:diffusionFig} shows the lateral diffusion constants
329 as a function of temperature. There is a definite increase in the
330 lateral diffusion with higher temperatures, which is exactly what one
331 would expect with greater fluidity of the chains. However, the
332 diffusion constants are all two orders of magnitude smaller than those
333 typical of DPPC.\cite{Cevc87} This is counter-intuitive as the DPPC
334 molecule is sterically larger and heavier than our model. This could
335 be an indication that our model's chains are too interwoven and hinder
336 the motion of the lipid, or that a simplification in the model's
337 forces has led to a slowing of diffusive behavior within the
338 bilayer. In contrast, the diffusion constant of the {\sc ssd} water,
339 $9.84\times 10^{-6}\,\text{cm}^2/\text{s}$, compares favorably with
340 that of bulk water ($2.2999\times
341 10^{-5}\,\text{cm}^2/\text{s}$\cite{Holz00}).
342
343 \begin{figure}
344 \centering
345 \includegraphics[width=\linewidth]{diffusionFig.eps}
346 \caption[The lateral diffusion constants versus temperature]{The lateral diffusion constants for the bilayers as a function of temperature.}
347 \label{lipidFig:diffusionFig}
348 \end{figure}
349
350 \subsection{\label{lipidSec:densProf}Density Profile}
351
352 Fig.~\ref{lipidFig:densityProfile} illustrates the densities of the
353 atoms in the bilayer systems normalized by the bulk density as a
354 function of distance from the center of the box. The profile is taken
355 along the bilayer normal, in this case the $z$ axis. The profile at
356 270~K shows several structural features that are largely smoothed out
357 by 300~K. The left peak for the {\sc head} atoms is split at 270~K,
358 implying that some freezing of the structure might already be occurring
359 at this temperature. From the dynamics, the tails at this temperature
360 are very much fluid, but the profile could indicate that a phase
361 transition may simply be beyond the length scale of the current
362 simulation. In all profiles, the water penetrates almost
363 5~$\mbox{\AA}$ into the bilayer, completely solvating the {\sc head}
364 atoms. The $\text{{\sc ch}}_3$ atoms although mainly centered at the
365 middle of the bilayer, show appreciable penetration into the head
366 group region. This indicates that the chains have enough mobility to
367 bend back upward to allow the ends to explore areas around the {\sc
368 head} atoms. It is unlikely that this is penetration from a lipid of
369 the opposite face, as the lipids are only 12~$\mbox{\AA}$ in length,
370 and the typical leaf spacing as measured from the {\sc head-head}
371 spacing in the profile is 17.5~$\mbox{\AA}$.
372
373 \begin{figure}
374 \centering
375 \includegraphics[width=\linewidth]{densityProfile.eps}
376 \caption[The density profile of the lipid bilayers]{The density profile of the lipid bilayers along the bilayer normal. The black lines are the {\sc head} atoms, red lines are the {\sc ch} atoms, green lines are the $\text{{\sc ch}}_2$ atoms, blue lines are the $\text{{\sc ch}}_3$ atoms, and the magenta lines are the {\sc ssd} atoms.}
377 \label{lipidFig:densityProfile}
378 \end{figure}
379
380
381 \subsection{\label{lipidSec:scd}$\text{S}_{\text{{\sc cd}}}$ Order Parameters}
382
383 The $\text{S}_{\text{{\sc cd}}}$ order parameter is often reported in
384 the experimental characterizations of phospholipids. It is obtained
385 through deuterium NMR, and measures the ordering of the carbon
386 deuterium bond in relation to the bilayer normal at various points
387 along the chains. In our model, there are no explicit hydrogens, but
388 the order parameter can be written in terms of the carbon ordering at
389 each point in the chain:\cite{egberts88}
390 \begin{equation}
391 S_{\text{{\sc cd}}} = \frac{2}{3}S_{xx} + \frac{1}{3}S_{yy}
392 \label{lipidEq:scd1}
393 \end{equation}
394 Where $S_{ij}$ is given by:
395 \begin{equation}
396 S_{ij} = \frac{1}{2}\Bigl\langle(3\cos\Theta_i\cos\Theta_j
397 - \delta_{ij})\Bigr\rangle
398 \label{lipidEq:scd2}
399 \end{equation}
400 Here, $\Theta_i$ is the angle the $i$th axis in the reference frame of
401 the carbon atom makes with the bilayer normal. The brackets denote an
402 average over time and molecules. The carbon atom axes are defined:
403 $\mathbf{\hat{z}}\rightarrow$ vector from $C_{n-1}$ to $C_{n+1}$;
404 $\mathbf{\hat{y}}\rightarrow$ vector that is perpendicular to $z$ and
405 in the plane through $C_{n-1}$, $C_{n}$, and $C_{n+1}$;
406 $\mathbf{\hat{x}}\rightarrow$ vector perpendicular to
407 $\mathbf{\hat{y}}$ and $\mathbf{\hat{z}}$.
408
409 The order parameter has a range of $[1,-\frac{1}{2}]$. A value of 1
410 implies full order aligned to the bilayer axis, 0 implies full
411 disorder, and $-\frac{1}{2}$ implies full order perpendicular to the
412 bilayer axis. The {\sc cd} bond vector for carbons near the head group
413 are usually ordered perpendicular to the bilayer normal, with tails
414 farther away tending toward disorder. This makes the order parameter
415 negative for most carbons, and as such $|S_{\text{{\sc cd}}}|$ is more
416 commonly reported than $S_{\text{{\sc cd}}}$.
417
418 Fig.~\ref{lipidFig:scdFig} shows the $S_{\text{{\sc cd}}}$ order
419 parameters for the bilayer system at 300~K. There is no appreciable
420 difference in the plots for the various temperatures, however, there
421 is a larger difference between our models ordering, and that of
422 DMPC. As our values are closer to $-\frac{1}{2}$, this implies more
423 ordering perpendicular to the normal than in a real system. This is
424 due to the model having only one carbon group separating the chains
425 from the top of the lipid. In DMPC, with the flexibility inherent in a
426 multiple atom head group, as well as a glycerol linkage between the
427 head group and the acyl chains, there is more loss of ordering by the
428 point when the chains start.
429
430 \begin{figure}
431 \centering
432 \includegraphics[width=\linewidth]{scdFig.eps}
433 \caption[$\text{S}_{\text{{\sc cd}}}$ order parameter for our model]{A comparison of $|\text{S}_{\text{{\sc cd}}}|$ between our model (blue) and DMPC\cite{petrache00} (black) near 300~K.}
434 \label{lipidFig:scdFig}
435 \end{figure}
436
437 \subsection{\label{lipidSec:p2Order}$P_2$ Order Parameter}
438
439 The $P_2$ order parameter allows us to measure the amount of
440 directional ordering that exists in the bilayer. Each lipid molecule
441 can be thought of as a cylindrical tube with the head group at the
442 top. If all of the cylinders are perfectly aligned, the $P_2$ order
443 parameter will be $1.0$. If the cylinders are completely dispersed,
444 the $P_2$ order parameter will be 0. For a collection of unit vectors,
445 the $P_2$ order parameter can be solved via the following
446 method.\cite{zannoni94}
447
448 Define an ordering matrix $\mathbf{Q}$, such that,
449 \begin{equation}
450 \mathbf{Q} = \frac{1}{N}\sum_i^N %
451 \begin{pmatrix} %
452 u_{ix}u_{ix}-\frac{1}{3} & u_{ix}u_{iy} & u_{ix}u_{iz} \\
453 u_{iy}u_{ix} & u_{iy}u_{iy}-\frac{1}{3} & u_{iy}u_{iz} \\
454 u_{iz}u_{ix} & u_{iz}u_{iy} & u_{iz}u_{iz}-\frac{1}{3} %
455 \end{pmatrix}
456 \label{lipidEq:po1}
457 \end{equation}
458 Where the $u_{i\alpha}$ is the $\alpha$ element of the unit vector
459 $\mathbf{\hat{u}}_i$, and the sum over $i$ averages over the whole
460 collection of unit vectors. This allows the matrix element
461 $Q_{\alpha\beta}$ to be written:
462 \begin{equation}
463 Q_{\alpha\beta} = \langle u_{\alpha}u_{\beta} -
464 \frac{1}{3}\delta_{\alpha\beta} \rangle
465 \label{lipidEq:po2}
466 \end{equation}
467
468 Having constructed the matrix, diagonalizing $\mathbf{Q}$ yields three
469 eigenvalues and eigenvectors. The eigenvector associated with the
470 largest eigenvalue, $\lambda_{\text{max}}$, is the director for the
471 system of unit vectors. The director is the average direction all of
472 the unit vectors are pointing. The $P_2$ order parameter is then
473 simply
474 \begin{equation}
475 \langle P_2 \rangle = \frac{3}{2}\lambda_{\text{max}}
476 \label{lipidEq:po3}
477 \end{equation}
478
479 Table~\ref{lipidTab:blSummary} summarizes the $P_2$ values for the
480 bilayers, as well as the dipole orientations. The unit vector for the
481 lipid molecules was defined by finding the moment of inertia for each
482 lipid, then setting $\mathbf{\hat{u}}$ to point along the axis of
483 minimum inertia. For the {\sc head} atoms, the unit vector simply
484 pointed in the same direction as the dipole moment. For the lipid
485 molecules, the ordering was consistent across all temperatures, with
486 the director pointed along the $z$ axis of the box. More
487 interestingly, is the high degree of ordering the dipoles impose on
488 the {\sc head} atoms. The directors for the dipoles consistently
489 pointed along the plane of the bilayer, with the directors
490 anti-aligned on the top and bottom leaf.
491
492 \begin{table}
493 \caption[Structural properties of the bilayers]{Bilayer Structural properties as a function of temperature.}
494 \label{lipidTab:blSummary}
495 \begin{center}
496 \begin{tabular}{|c|c|c|c|c|}
497 \hline
498 Temperature (K) & $\langle L_{\perp}\rangle$ ($\mbox{\AA}$) & %
499 $\langle A_{\parallel}\rangle$ ($\mbox{\AA}^2$) & %
500 $\langle P_2\rangle_{\text{Lipid}}$ & %
501 $\langle P_2\rangle_{\text{{\sc head}}}$ \\ \hline
502 270 & 18.1 & 58.1 & 0.253 & 0.494 \\ \hline
503 275 & 17.2 & 56.7 & 0.295 & 0.514 \\ \hline
504 277 & 16.9 & 58.0 & 0.301 & 0.541 \\ \hline
505 280 & 17.4 & 58.0 & 0.274 & 0.488 \\ \hline
506 285 & 16.9 & 57.6 & 0.270 & 0.616 \\ \hline
507 290 & 17.0 & 57.6 & 0.263 & 0.534 \\ \hline
508 293 & 17.5 & 58.0 & 0.227 & 0.643 \\ \hline
509 300 & 16.9 & 57.6 & 0.315 & 0.536 \\ \hline
510 \end{tabular}
511 \end{center}
512 \end{table}
513
514 \subsection{\label{lipidSec:miscData}Further Bilayer Data}
515
516 Also summarized in Table~\ref{lipidTab:blSummary}, are the bilayer
517 thickness and area per lipid. The bilayer thickness was measured from
518 the peak to peak {\sc head} atom distance in the density profiles. The
519 area per lipid data compares favorably with values typically seen for
520 DMPC (60.0~$\mbox{\AA}^2$ at 303~K)\cite{petrache00}. Although are
521 values are lower this is most likely due to the shorter chain length
522 of our model (8 versus 14 for DMPC).
523
524 \subsection{\label{lipidSec:randBilayer}Bilayer Aggregation}
525
526 A very important accomplishment for our model is its ability to
527 spontaneously form bilayers from a randomly dispersed starting
528 configuration. Fig.~\ref{lipidFig:blImage} shows an image sequence for
529 the bilayer aggregation. After 3.0~ns, the basic form of the bilayer
530 can already be seen. By 7.0~ns, the bilayer has a lipid bridge
531 stretched across the simulation box to itself that will turn out to be
532 very long lived ($\sim$20~ns), as well as a water pore, that will
533 persist for the length of the current simulation. At 24~ns, the lipid
534 bridge is dispersed, and the bilayer is still integrating the lipid
535 molecules from the bridge into itself, and has still been unable to
536 disperse the water pore.
537
538 \begin{figure}
539 \centering
540 \includegraphics[width=\linewidth]{bLayerImage.eps}
541 \caption[Image sequence of the bilayer aggregation]{Image sequence of the bilayer aggregation. The blue beads are the {\sc head} atoms the grey beads are the chains, and the red and white bead are the water molecules. A box has been drawn around the periodic image.}
542 \label{lipidFig:blImage}
543 \end{figure}
544
545 \subsection{\label{lipidSec:randIrod}Inverted Rod Aggregation}
546
547 Fig.~\ref{lipidFig:iRimage} shows a second aggregation sequence
548 simulated in this research. Here the fraction of water had been
549 significantly decreased to observe how the model would respond. After
550 1.5~ns, The main body of water in the system has already collected
551 into a central water channel. By 10.0~ns, the channel has widened
552 slightly, but there are still many sub channels permeating the lipid
553 macro-structure. At 23.0~ns, the central water channel has stabilized
554 and several smaller water channels have been absorbed to main
555 one. However, there are still several other channels that persist
556 through the lipid structure.
557
558 \begin{figure}
559 \centering
560 \includegraphics[width=\linewidth]{iRodImage.eps}
561 \caption[Image sequence of the inverted rod aggregation]{Image sequence of the inverted rod aggregation. color scheme is the same as in Fig.~\ref{lipidFig:blImage}.}
562 \label{lipidFig:iRimage}
563 \end{figure}
564
565 \section{\label{lipidSec:Conclusion}Conclusion}
566
567 We have presented a phospholipid model capable of spontaneous
568 aggregation into a bilayer and an inverted rod structure. The time
569 scales of the macro-molecular aggregations are in excess of 24~ns. In
570 addition the model's bilayer properties have been explored over a
571 range of temperatures through prefabricated bilayers. No freezing
572 transition is seen in the temperature range of our current
573 simulations. However, structural information from the lowest
574 temperature may imply that a freezing event is on a much longer time
575 scale than that explored in this current research. Further studies of
576 this system could extend the time length of the simulations at the low
577 temperatures to observe whether lipid crystallization occurs within the
578 framework of this model.