ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/lipid.tex
Revision: 1115
Committed: Thu Apr 15 19:19:06 2004 UTC (20 years, 3 months ago) by mmeineke
Content type: application/x-tex
File size: 34393 byte(s)
Log Message:
final copies sent to the printer

File Contents

# User Rev Content
1 mmeineke 971
2    
3 mmeineke 1105 \chapter{\label{chapt:lipid}PHOSPHOLIPID SIMULATIONS}
4 mmeineke 971
5     \section{\label{lipidSec:Intro}Introduction}
6    
7 mmeineke 1083 In the past 10 years, increasing computer speeds have allowed for the
8     atomistic simulation of phospholipid bilayers for increasingly
9 mmeineke 1087 relevant lengths of time. These simulations have ranged from
10 mmeineke 1089 simulation of the gel ($L_{\beta}$) phase of
11 mmeineke 1061 dipalmitoylphosphatidylcholine (DPPC),\cite{lindahl00} to the
12 mmeineke 1001 spontaneous aggregation of DPPC molecules into fluid phase
13 mmeineke 1115 ($L_{\alpha}$) bilayers.\cite{Marrink01} With the exception of a few
14 mmeineke 1061 ambitious
15 mmeineke 1089 simulations, \cite{marrink01:undulation,marrink:2002,lindahl00} most
16 mmeineke 1083 investigations are limited to a range of 64 to 256
17 mmeineke 1115 phospholipids.\cite{lindahl00,sum:2003,venable00,gomez:2003,smondyrev:1999,Marrink01}
18 mmeineke 1083 The expense of the force calculations involved when performing these
19     simulations limits the system size. To properly hydrate a bilayer, one
20 mmeineke 1001 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 mmeineke 1061 molecules. Added to the difficulty is the electrostatic nature of the
23 mmeineke 1083 phospholipid head groups and water, requiring either the
24 mmeineke 1089 computationally expensive, direct Ewald sum or the slightly faster particle
25     mesh Ewald (PME) sum.\cite{nina:2002,norberg:2000,patra:2003} These factors
26     all limit the system size and time scales of bilayer simulations.
27 mmeineke 1001
28     Unfortunately, much of biological interest happens on time and length
29 mmeineke 1089 scales well beyond the range of current simulation technologies. One
30     such example is the observance of a ripple ($P_{\beta^{\prime}}$)
31     phase which appears between the $L_{\beta}$ and $L_{\alpha}$ phases of
32     certain phospholipid bilayers
33     (Fig.~\ref{lipidFig:phaseDiag}).\cite{katsaras00,sengupta00} These
34     ripples are known from x-ray diffraction data to have periodicities on
35     the order of 100-200~$\mbox{\AA}$.\cite{katsaras00} A simulation on
36     this length scale would have approximately 1,300 lipid molecules with
37     an additional 25 water molecules per lipid to fully solvate the
38     bilayer. A simulation of this size is impractical with current
39     atomistic models.
40 mmeineke 1001
41 mmeineke 1089 \begin{figure}
42     \centering
43     \includegraphics[width=\linewidth]{ripple.eps}
44     \caption[Diagram of the bilayer gel to fluid phase transition]{Diagram showing the $P_{\beta^{\prime}}$ phase as a transition between the $L_{\beta}$ and $L_{\alpha}$ phases}
45     \label{lipidFig:phaseDiag}
46     \end{figure}
47    
48 mmeineke 1083 The time and length scale limitations are most striking in transport
49 mmeineke 1089 phenomena. Due to the fluid-like properties of lipid membranes, not
50     all small molecule diffusion across the membranes happens at pores.
51     Some molecules of interest may incorporate themselves directly into
52     the membrane. Once there, they may exhibit appreciable waiting times
53     (on the order of 10's to 100's of nanoseconds) within the
54     bilayer. Such long simulation times are nearly impossible to obtain
55     when integrating the system with atomistic detail.
56 mmeineke 1001
57 mmeineke 1083 To address these issues, several schemes have been proposed. One
58 mmeineke 1089 approach by Goetz and Lipowsky\cite{goetz98} is to model the entire
59 mmeineke 1001 system as Lennard-Jones spheres. Phospholipids are represented by
60     chains of beads with the top most beads identified as the head
61     atoms. Polar and non-polar interactions are mimicked through
62 mmeineke 1089 attractive and soft-repulsive potentials respectively. A model
63     proposed by Marrinck \emph{et. al}.\cite{marrink04}~uses a similar
64     technique for modeling polar and non-polar interactions with
65 mmeineke 1001 Lennard-Jones spheres. However, they also include charges on the head
66     group spheres to mimic the electrostatic interactions of the
67 mmeineke 1089 bilayer. The solvent spheres are kept charge-neutral and
68 mmeineke 1001 interact with the bilayer solely through an attractive Lennard-Jones
69     potential.
70    
71     The model used in this investigation adds more information to the
72 mmeineke 1026 interactions than the previous two models, while still balancing the
73 mmeineke 1089 need for simplification of atomistic detail. The model uses
74     unified-atom Lennard-Jones spheres for the head and tail groups of the
75 mmeineke 1061 phospholipids, allowing for the ability to scale the parameters to
76 mmeineke 1026 reflect various sized chain configurations while keeping the number of
77     interactions small. What sets this model apart, however, is the use
78 mmeineke 1061 of dipoles to represent the electrostatic nature of the
79 mmeineke 1026 phospholipids. The dipole electrostatic interaction is shorter range
80 mmeineke 1089 than Coulombic ($\frac{1}{r^3}$ versus $\frac{1}{r}$), and therefore
81     eliminates the need for the costly Ewald sum.
82 mmeineke 1001
83 mmeineke 1089 Another key feature of this model is the use of a dipolar water model
84 mmeineke 1061 to represent the solvent. The soft sticky dipole ({\sc ssd}) water
85     \cite{liu96:new_model} relies on the dipole for long range electrostatic
86     effects, but also contains a short range correction for hydrogen
87 mmeineke 1089 bonding. In this way the simulated systems in this research mimic the
88     entropic contribution to the hydrophobic effect due to hydrogen-bond
89     network deformation around a non-polar entity, \emph{i.e.}~the
90     phospholipid molecules. This effect has been missing from previous
91     reduced models.
92 mmeineke 1026
93     The following is an outline of this chapter.
94 mmeineke 1083 Sec.~\ref{lipidSec:Methods} is an introduction to the lipid model used
95     in these simulations, as well as clarification about the water model
96     and integration techniques. The various simulations explored in this
97     research are outlined in
98     Sec.~\ref{lipidSec:ExpSetup}. Sec.~\ref{lipidSec:resultsDis} gives a
99     summary and interpretation of the results. Finally, the conclusions
100     of this chapter are presented in Sec.~\ref{lipidSec:Conclusion}.
101 mmeineke 1026
102 mmeineke 971 \section{\label{lipidSec:Methods}Methods}
103    
104 mmeineke 1026 \subsection{\label{lipidSec:lipidModel}The Lipid Model}
105    
106 mmeineke 971 \begin{figure}
107 mmeineke 1061 \centering
108     \includegraphics[width=\linewidth]{twoChainFig.eps}
109     \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.}
110 mmeineke 971 \label{lipidFig:lipidModel}
111     \end{figure}
112    
113     The phospholipid model used in these simulations is based on the
114     design illustrated in Fig.~\ref{lipidFig:lipidModel}. The head group
115     of the phospholipid is replaced by a single Lennard-Jones sphere of
116 mmeineke 1061 diameter $\sigma_{\text{head}}$, with $\epsilon_{\text{head}}$ scaling
117     the well depth of its van der Walls interaction. This sphere also
118     contains a single dipole of magnitude $|\boldsymbol{\mu}|$, where
119     $|\boldsymbol{\mu}|$ can be varied to mimic the charge separation of a
120 mmeineke 971 given phospholipid head group. The atoms of the tail region are
121 mmeineke 1083 modeled by beads representing multiple methyl groups. They are free
122     of partial charges or dipoles, and contain only Lennard-Jones
123     interaction sites at their centers of mass. As with the head groups,
124     their potentials can be scaled by $\sigma_{\text{tail}}$ and
125     $\epsilon_{\text{tail}}$.
126 mmeineke 971
127 mmeineke 1089 The possible long range interactions between atomic groups in the
128     lipids are given by the following:
129 mmeineke 971 \begin{equation}
130 mmeineke 1061 V_{\text{LJ}}(r_{ij}) =
131     4\epsilon_{ij} \biggl[
132     \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
133     - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
134 mmeineke 1114 \biggr],
135 mmeineke 971 \label{lipidEq:LJpot}
136     \end{equation}
137     and
138     \begin{equation}
139 mmeineke 1061 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
140     \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
141     \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
142     -
143 mmeineke 1083 3(\boldsymbol{\hat{u}}_i \cdot \mathbf{\hat{r}}_{ij}) %
144 mmeineke 1114 (\boldsymbol{\hat{u}}_j \cdot \mathbf{\hat{r}}_{ij} )\biggr].
145 mmeineke 971 \label{lipidEq:dipolePot}
146     \end{equation}
147 mmeineke 1114 Here $V_{\text{LJ}}$ is the Lennard-Jones potential and
148 mmeineke 971 $V_{\text{dipole}}$ is the dipole-dipole potential. As previously
149     stated, $\sigma_{ij}$ and $\epsilon_{ij}$ are the Lennard-Jones
150     parameters which scale the length and depth of the interaction
151     respectively, and $r_{ij}$ is the distance between beads $i$ and $j$.
152     In $V_{\text{dipole}}$, $\mathbf{r}_{ij}$ is the vector starting at
153 mmeineke 1083 bead $i$ and pointing towards bead $j$. Vectors $\mathbf{\Omega}_i$
154 mmeineke 971 and $\mathbf{\Omega}_j$ are the orientational degrees of freedom for
155     beads $i$ and $j$. $|\mu_i|$ is the magnitude of the dipole moment of
156     $i$, and $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
157 mmeineke 1087 vector rotated with Euler angles: $\boldsymbol{\Omega}_i$.
158 mmeineke 971
159 mmeineke 1089 The model also allows for the intra-molecular bend and torsion
160     interactions. The bond between two beads on a chain is of fixed
161     length, and is maintained using the {\sc rattle}
162     algorithm.\cite{andersen83} The bends are subject to a harmonic
163     potential:
164 mmeineke 971 \begin{equation}
165 mmeineke 1114 V_{\text{bend}}(\theta) = k_{\theta}( \theta - \theta_0 )^2,
166 mmeineke 971 \label{lipidEq:bendPot}
167     \end{equation}
168 mmeineke 1061 where $k_{\theta}$ scales the strength of the harmonic well, and
169     $\theta$ is the angle between bond vectors
170     (Fig.~\ref{lipidFig:lipidModel}). In addition, we have placed a
171 mmeineke 1089 ``ghost'' bend on the phospholipid head. The ghost bend is a bend
172     potential which keeps the dipole roughly perpendicular to the
173     molecular body, where $\theta$ is now the angle the dipole makes with
174     respect to the {\sc head}-$\text{{\sc ch}}_2$ bond vector
175     (Fig.~\ref{lipidFig:ghostBend}). This bend mimics the hinge between
176     the phosphatidyl part of the PC head group and the remainder of the
177     molecule. The torsion potential is given by:
178 mmeineke 971 \begin{equation}
179 mmeineke 1061 V_{\text{torsion}}(\phi) =
180 mmeineke 1114 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0.
181 mmeineke 971 \label{lipidEq:torsionPot}
182     \end{equation}
183     Here, the parameters $k_0$, $k_1$, $k_2$, and $k_3$ fit the cosine
184     power series to the desired torsion potential surface, and $\phi$ is
185 mmeineke 1061 the angle the two end atoms have rotated about the middle bond
186     (Fig.:\ref{lipidFig:lipidModel}). Long range interactions such as the
187     Lennard-Jones potential are excluded for atom pairs involved in the
188 mmeineke 1089 same bond, bend, or torsion. However, long-range interactions for
189     pairs of atoms not directly involved in a bond, bend, or torsion are
190     calculated.
191 mmeineke 971
192 mmeineke 1083 \begin{figure}
193     \centering
194 mmeineke 1089 \includegraphics[width=0.5\linewidth]{ghostBendFig.eps}
195     \caption[Depiction of the ``ghost'' bend]{The ``ghost'' bend is a bend potential added to restrain 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.}
196 mmeineke 1083 \label{lipidFig:ghostBend}
197     \end{figure}
198    
199 mmeineke 1089 All simulations presented here use a two-chain lipid as pictured in
200 mmeineke 1061 Fig.~\ref{lipidFig:lipidModel}. The chains are both eight beads long,
201 mmeineke 1026 and their mass and Lennard Jones parameters are summarized in
202     Table~\ref{lipidTable:tcLJParams}. The magnitude of the dipole moment
203 mmeineke 1089 for the head bead is 10.6~Debye (approximately half the magnitude of
204     the dipole on the PC head group\cite{Cevc87}), and the bend and
205     torsion parameters are summarized in
206     Table~\ref{lipidTable:tcBendParams} and
207 mmeineke 1061 \ref{lipidTable:tcTorsionParams}.
208 mmeineke 971
209 mmeineke 1061 \begin{table}
210 mmeineke 1115 \caption{THE LENNARD JONES PARAMETERS FOR THE TWO CHAIN PHOSPHOLIPIDS}
211 mmeineke 1061 \label{lipidTable:tcLJParams}
212     \begin{center}
213 mmeineke 1089 \begin{tabular}{|l|c|c|c|c|}
214 mmeineke 1061 \hline
215 mmeineke 1089 & mass (amu) & $\sigma$($\mbox{\AA}$) & $\epsilon$ (kcal/mol) %
216     & $|\mathbf{\hat{\mu}}|$ (Debye) \\ \hline
217     {\sc head} & 72 & 4.0 & 0.185 & 10.6 \\ \hline
218     {\sc ch}\cite{Siepmann1998} & 13.02 & 4.0 & 0.0189 & 0.0 \\ \hline
219     $\text{{\sc ch}}_2$\cite{Siepmann1998} & 14.03 & 3.95 & 0.18 & 0.0 \\ \hline
220     $\text{{\sc ch}}_3$\cite{Siepmann1998} & 15.04 & 3.75 & 0.25 & 0.0 \\ \hline
221     {\sc ssd}\cite{liu96:new_model} & 18.03 & 3.051 & 0.152 & 0.0 \\ \hline
222 mmeineke 1061 \end{tabular}
223     \end{center}
224     \end{table}
225 mmeineke 1026
226 mmeineke 1061 \begin{table}
227 mmeineke 1115 \caption{BEND PARAMETERS FOR THE TWO CHAIN PHOSPHOLIPIDS}
228 mmeineke 1061 \label{lipidTable:tcBendParams}
229     \begin{center}
230     \begin{tabular}{|l|c|c|}
231     \hline
232     & $k_{\theta}$ ( kcal/($\text{mol deg}^2$) ) & $\theta_0$ ( deg ) \\ \hline
233     {\sc ghost}-{\sc head}-$\text{{\sc ch}}_2$ & 0.00177 & 129.78 \\ \hline
234     $x$-{\sc ch}-$y$ & 58.84 & 112.0 \\ \hline
235     $x$-$\text{{\sc ch}}_2$-$y$ & 58.84 & 114.0 \\ \hline
236     \end{tabular}
237 mmeineke 1089 \begin{minipage}{\linewidth}
238     \begin{center}
239     \vspace{2mm}
240     All alkane parameters are based off of those from TraPPE.\cite{Siepmann1998}
241 mmeineke 1061 \end{center}
242 mmeineke 1089 \end{minipage}
243     \end{center}
244 mmeineke 1061 \end{table}
245    
246     \begin{table}
247 mmeineke 1115 \caption{TORSION PARAMETERS FOR THE TWO CHAIN PHOSPHOLIPIDS}
248 mmeineke 1061 \label{lipidTable:tcTorsionParams}
249     \begin{center}
250     \begin{tabular}{|l|c|c|c|c|}
251     \hline
252     All are in kcal/mol $\rightarrow$ & $k_3$ & $k_2$ & $k_1$ & $k_0$ \\ \hline
253     $x$-{\sc ch}-$y$-$z$ & 3.3254 & -0.4215 & -1.686 & 1.1661 \\ \hline
254     $x$-$\text{{\sc ch}}_2$-$\text{{\sc ch}}_2$-$y$ & 5.9602 & -0.568 & -3.802 & 2.1586 \\ \hline
255     \end{tabular}
256 mmeineke 1089 \begin{minipage}{\linewidth}
257     \begin{center}
258     \vspace{2mm}
259     All alkane parameters are based off of those from TraPPE.\cite{Siepmann1998}
260 mmeineke 1061 \end{center}
261 mmeineke 1089 \end{minipage}
262     \end{center}
263 mmeineke 1061 \end{table}
264    
265    
266     \section{\label{lipidSec:furtherMethod}Further Methodology}
267    
268 mmeineke 1026 As mentioned previously, the water model used throughout these
269 mmeineke 1089 simulations was the {\sc ssd/e} model of Fennell and
270     Gezelter,\cite{fennell04} earlier forms of this model can be found in
271     Ichiye \emph{et
272     al}.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md} A
273 mmeineke 1061 discussion of the model can be found in Sec.~\ref{oopseSec:SSD}. As
274     for the integration of the equations of motion, all simulations were
275     performed in an orthorhombic periodic box with a thermostat on
276     velocities, and an independent barostat on each Cartesian axis $x$,
277 mmeineke 1089 $y$, and $z$. This is the $\text{NPT}_{xyz}$. integrator described in
278     Sec.~\ref{oopseSec:integrate}. With $\tau_B = 1.5$~ps and $\tau_T =
279     1.2$~ps, the box volume stabilizes after 50~ps, and fluctuates about
280     its equilibrium value by $\sim 0.6\%$, temperature fluctuations are
281     about $\sim 1.4\%$ of their set value, and pressure fluctuations are
282     the largest, varying as much as $\pm 250$~atm. However, such large
283 mmeineke 1114 fluctuations in pressure are typical for liquid state simulations.\cite{leach01:mm}
284 mmeineke 1026
285    
286     \subsection{\label{lipidSec:ExpSetup}Experimental Setup}
287    
288 mmeineke 1089 Two main classes of starting configurations were used in this research:
289     random and ordered bilayers. The ordered bilayer simulations were all
290     started from an equilibrated bilayer configuration at 300~K. The original
291     configuration for the first 300~K run was assembled by placing the
292     phospholipids centers of mass on a planar hexagonal lattice. The
293     lipids were oriented with their principal axis perpendicular to the plane.
294     The bottom leaf simply mirrored the top leaf, and the appropriate
295     number of water molecules were then added above and below the bilayer.
296 mmeineke 1026
297     The random configurations took more work to generate. To begin, a
298     test lipid was placed in a simulation box already containing water at
299 mmeineke 1089 the intended density. The water molecules were then tested against
300     the lipid using a 5.0~$\mbox{\AA}$ overlap test with any atom in the
301     lipid. This gave an estimate for the number of water molecules each
302     lipid would displace in a simulation box. A target number of water
303     molecules was then defined which included the number of water
304     molecules each lipid would displace, the number of water molecules
305     desired to solvate each lipid, and a factor to pad the initial box
306     with a few extra water molecules.
307 mmeineke 1026
308     Next, a cubic simulation box was created that contained at least the
309 mmeineke 1089 target number of water molecules in an FCC lattice (the lattice was for ease of
310 mmeineke 1026 placement). What followed was a RSA simulation similar to those of
311     Chapt.~\ref{chapt:RSA}. The lipids were sequentially given a random
312     position and orientation within the box. If a lipid's position caused
313 mmeineke 1083 atomic overlap with any previously placed lipid, its position and
314     orientation were rejected, and a new random placement site was
315 mmeineke 1026 attempted. The RSA simulation proceeded until all phospholipids had
316 mmeineke 1089 been placed. After placement of all lipid molecules, water
317 mmeineke 1083 molecules with locations that overlapped with the atomic coordinates
318     of the lipids were removed.
319 mmeineke 1026
320 mmeineke 1089 Finally, water molecules were removed at random until the desired water
321     to lipid ratio was achieved. The typical low final density for these
322     initial configurations was not a problem, as the box shrinks to an
323     appropriate size within the first 50~ps of a simulation under the
324     NPTxyz integrator.
325 mmeineke 1026
326 mmeineke 1083 \subsection{\label{lipidSec:Configs}Configurations}
327 mmeineke 1026
328 mmeineke 1089 The first class of simulations were started from ordered bilayers. All
329     configurations consisted of 60 lipid molecules with 30 lipids on each
330     leaf, and were hydrated with 1620 {\sc ssd/e} molecules. The original
331     configuration was assembled according to Sec.~\ref{lipidSec:ExpSetup}
332     and simulated for a length of 10~ns at 300~K. The other temperature
333     runs were started from a configuration 7~ns in to the 300~K
334     simulation. Their temperatures were modified with the thermostatting
335     algorithm in the NPTxyz integrator. All of the temperature variants
336     were also run for 10~ns, with only the last 5~ns being used for
337     accumulation of statistics.
338 mmeineke 1083
339     The second class of simulations were two configurations started from
340     randomly dispersed lipids in a ``gas'' of water. The first
341     ($\text{R}_{\text{I}}$) was a simulation containing 72 lipids with
342 mmeineke 1089 1800 {\sc ssd/e} molecules simulated at 300~K. The second
343     ($\text{R}_{\text{II}}$) was 90 lipids with 1350 {\sc ssd/e} molecules
344 mmeineke 1083 simulated at 350~K. Both simulations were integrated for more than
345 mmeineke 1089 20~ns to observe whether our model is capable of spontaneous
346     aggregation into known phospholipid macro-structures:
347     $\text{R}_{\text{I}}$ into a bilayer, and $\text{R}_{\text{II}}$ into
348     a inverted rod.
349 mmeineke 1083
350     \section{\label{lipidSec:resultsDis}Results and Discussion}
351    
352 mmeineke 1084 \subsection{\label{lipidSec:densProf}Density Profile}
353    
354     Fig.~\ref{lipidFig:densityProfile} illustrates the densities of the
355 mmeineke 1087 atoms in the bilayer systems normalized by the bulk density as a
356 mmeineke 1084 function of distance from the center of the box. The profile is taken
357 mmeineke 1115 along the bilayer normal (in this case the $z$ axis). The first interesting point to note is the penetration of water into the membrane. Water penetrates about 5~$\mbox{\AA}$ into the bilayer, completely solvating the head groups. This is common in atomistic and some coarse grain simulations of phospholipid bilayers.\cite{Marrink01,marrink04,klein01} It is an indication that the water molecules are very attracted to the polar head region, yet there is still enough of a hydrophobic effect to exclude water from the inside of the bilayer.
358 mmeineke 1084
359 mmeineke 1114 Another interesting point is the fluidity of the chains. Although the ends of the tails, the $\text{{\sc ch}}_3$ atoms, are mostly concentrated at the centers of the bilayers, they have a significant density around the head regions. This indicates that there is much freedom of movement in the chains of our model. Typical atomistic simulations of DPPC show the terminal groups concentrated at the center of the bilayer.\cite{marrink03:vesicles} This is most likely an indication that our chain lengths are too small, and given longer chains, the tail groups would stay more deeply buried in the bilayer.
360    
361 mmeineke 1115 The last point to consider is the splitting in the density peak of the {\sc head} atom at 270~K. This implies that there is some freezing of structure at this temperature. By 280~K, this feature is smoothed out, demonstrating a more fluid phase in the bilayer. Within the time scale of the simulation, the gel phase has not formed at 270~K, so this splitting in the peak is likely a glassy transition in the head groups, and could possibly indicate that we are simulating in a super cooled region of our phospholipid model.
362 mmeineke 1114
363 mmeineke 1084 \begin{figure}
364     \centering
365     \includegraphics[width=\linewidth]{densityProfile.eps}
366     \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.}
367     \label{lipidFig:densityProfile}
368     \end{figure}
369    
370    
371     \subsection{\label{lipidSec:scd}$\text{S}_{\text{{\sc cd}}}$ Order Parameters}
372    
373 mmeineke 1083 The $\text{S}_{\text{{\sc cd}}}$ order parameter is often reported in
374 mmeineke 1087 the experimental characterizations of phospholipids. It is obtained
375 mmeineke 1083 through deuterium NMR, and measures the ordering of the carbon
376     deuterium bond in relation to the bilayer normal at various points
377 mmeineke 1114 along the chains. The order parameter has a range of $[1,-\frac{1}{2}]$. A value of 1
378     implies full order aligned to the bilayer axis, 0 implies full
379     disorder, and $-\frac{1}{2}$ implies full order perpendicular to the
380     bilayer axis. The {\sc cd} bond vector for carbons near the head group
381     are usually ordered perpendicular to the bilayer normal, with tails
382     farther away tending toward disorder. This makes the order parameter
383     negative for most carbons, and as such $|S_{\text{{\sc cd}}}|$ is more
384     commonly reported than $S_{\text{{\sc cd}}}$.
385    
386     In our model, there are no explicit hydrogens, but
387 mmeineke 1083 the order parameter can be written in terms of the carbon ordering at
388     each point in the chain:\cite{egberts88}
389     \begin{equation}
390 mmeineke 1114 S_{\text{{\sc cd}}} = \frac{2}{3}S_{xx} + \frac{1}{3}S_{yy},
391 mmeineke 1083 \label{lipidEq:scd1}
392     \end{equation}
393 mmeineke 1114 where $S_{ij}$ is given by:
394 mmeineke 1083 \begin{equation}
395 mmeineke 1084 S_{ij} = \frac{1}{2}\Bigl\langle(3\cos\Theta_i\cos\Theta_j
396 mmeineke 1114 - \delta_{ij})\Bigr\rangle.
397 mmeineke 1083 \label{lipidEq:scd2}
398     \end{equation}
399 mmeineke 1084 Here, $\Theta_i$ is the angle the $i$th axis in the reference frame of
400     the carbon atom makes with the bilayer normal. The brackets denote an
401     average over time and molecules. The carbon atom axes are defined:
402 mmeineke 1089 \begin{itemize}
403 mmeineke 1114 \item $\mathbf{\hat{z}}$ is the vector from $C_{n-1}$ to $C_{n+1}$.
404     \item $\mathbf{\hat{y}}$ is the vector that is perpendicular to $z$ and
405     in the plane through $C_{n-1}$, $C_{n}$, and $C_{n+1}$.
406     \item $\mathbf{\hat{x}}$ is the vector perpendicular to
407 mmeineke 1083 $\mathbf{\hat{y}}$ and $\mathbf{\hat{z}}$.
408 mmeineke 1089 \end{itemize}
409     This assumes that the hydrogen atoms are always in a plane
410     perpendicular to the $C_{n-1}-C_{n}-C_{n+1}$ plane.
411 mmeineke 1083
412 mmeineke 1084 Fig.~\ref{lipidFig:scdFig} shows the $S_{\text{{\sc cd}}}$ order
413     parameters for the bilayer system at 300~K. There is no appreciable
414     difference in the plots for the various temperatures, however, there
415 mmeineke 1089 is a larger difference between our model's ordering, and the
416     experimentally observed ordering of DMPC. As our values are closer to
417     $-\frac{1}{2}$, this implies more ordering perpendicular to the normal
418     than in a real system. This is due to the model having only one carbon
419     group separating the chains from the top of the lipid. In DMPC, with
420     the flexibility inherent in a multiple atom head group, as well as a
421     glycerol linkage between the head group and the acyl chains, there is
422 mmeineke 1115 more loss of ordering by the point when the chains start. Also, there is more ordering in the model due to the our assumptions about the locations of the hydrogen atoms. Our method assumes a rigid location for each hydrogen atom based on the carbon positions. This does not allow for any small fluctuations in their positions that would be inherent in an atomistic simulation or in experiments. These small fluctuations would serve to lower the ordering measured in the $S_{\text{{\sc cd}}}$.
423 mmeineke 1083
424     \begin{figure}
425     \centering
426     \includegraphics[width=\linewidth]{scdFig.eps}
427     \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.}
428     \label{lipidFig:scdFig}
429     \end{figure}
430    
431 mmeineke 1087 \subsection{\label{lipidSec:p2Order}$P_2$ Order Parameter}
432 mmeineke 1083
433 mmeineke 1087 The $P_2$ order parameter allows us to measure the amount of
434 mmeineke 1089 directional ordering that exists in the bodies of the molecules making
435     up the bilayer. Each lipid molecule can be thought of as a cylindrical
436     rod with the head group at the top. If all of the rods are perfectly
437     aligned, the $P_2$ order parameter will be $1.0$. If the rods are
438     completely disordered, the $P_2$ order parameter will be 0. For a
439     collection of unit vectors pointing along the principal axes of the
440     rods, the $P_2$ order parameter can be solved via the following
441 mmeineke 1087 method.\cite{zannoni94}
442 mmeineke 1083
443 mmeineke 1089 Define an ordering tensor $\overleftrightarrow{\mathsf{Q}}$, such that,
444 mmeineke 1087 \begin{equation}
445 mmeineke 1089 \overleftrightarrow{\mathsf{Q}} = \frac{1}{N}\sum_i^N %
446 mmeineke 1087 \begin{pmatrix} %
447     u_{ix}u_{ix}-\frac{1}{3} & u_{ix}u_{iy} & u_{ix}u_{iz} \\
448     u_{iy}u_{ix} & u_{iy}u_{iy}-\frac{1}{3} & u_{iy}u_{iz} \\
449     u_{iz}u_{ix} & u_{iz}u_{iy} & u_{iz}u_{iz}-\frac{1}{3} %
450 mmeineke 1114 \end{pmatrix},
451 mmeineke 1087 \label{lipidEq:po1}
452     \end{equation}
453 mmeineke 1114 where the $u_{i\alpha}$ is the $\alpha$ element of the unit vector
454 mmeineke 1087 $\mathbf{\hat{u}}_i$, and the sum over $i$ averages over the whole
455 mmeineke 1089 collection of unit vectors. This allows the tensor to be written:
456 mmeineke 1087 \begin{equation}
457 mmeineke 1089 \overleftrightarrow{\mathsf{Q}} = \frac{1}{N}\sum_i^N \biggl[
458     \mathbf{\hat{u}}_i \otimes \mathbf{\hat{u}}_i
459 mmeineke 1114 - \frac{1}{3} \cdot \mathsf{1} \biggr].
460 mmeineke 1087 \label{lipidEq:po2}
461     \end{equation}
462 mmeineke 1083
463 mmeineke 1089 After constructing the tensor, diagonalizing
464     $\overleftrightarrow{\mathsf{Q}}$ yields three eigenvalues and
465     eigenvectors. The eigenvector associated with the largest eigenvalue,
466     $\lambda_{\text{max}}$, is the director axis for the system of unit
467     vectors. The director axis is the average direction all of the unit vectors
468     are pointing. The $P_2$ order parameter is then simply
469 mmeineke 1087 \begin{equation}
470 mmeineke 1114 \langle P_2 \rangle = \frac{3}{2}\lambda_{\text{max}}.
471 mmeineke 1087 \label{lipidEq:po3}
472     \end{equation}
473 mmeineke 1083
474 mmeineke 1087 Table~\ref{lipidTab:blSummary} summarizes the $P_2$ values for the
475 mmeineke 1115 bilayers, as well as for the dipole orientations. The unit vector for the
476 mmeineke 1087 lipid molecules was defined by finding the moment of inertia for each
477     lipid, then setting $\mathbf{\hat{u}}$ to point along the axis of
478 mmeineke 1114 minimum inertia (the long axis). For the {\sc head} atoms, the unit vector simply
479 mmeineke 1087 pointed in the same direction as the dipole moment. For the lipid
480     molecules, the ordering was consistent across all temperatures, with
481     the director pointed along the $z$ axis of the box. More
482     interestingly, is the high degree of ordering the dipoles impose on
483 mmeineke 1089 the {\sc head} atoms. The directors for the dipoles themselves
484 mmeineke 1115 consistently pointed along the plane of the bilayer, with head groups lining up in rows of alternating alignment. The ordering implies that the dipole interaction is a little too strong, or that perhaps the dipoles are allowed to approach each other a bit too closely. A possible change in future models would alter the size or shape of the head group to discourage too rigid ordering of the dipoles.
485 mmeineke 1087
486 mmeineke 1083 \begin{table}
487 mmeineke 1115 \caption{BILAYER STRUCTURAL PROPERTIES AS A FUNCTION OF TEMPERATURE}
488 mmeineke 1087 \label{lipidTab:blSummary}
489 mmeineke 1083 \begin{center}
490     \begin{tabular}{|c|c|c|c|c|}
491     \hline
492 mmeineke 1084 Temperature (K) & $\langle L_{\perp}\rangle$ ($\mbox{\AA}$) & %
493     $\langle A_{\parallel}\rangle$ ($\mbox{\AA}^2$) & %
494     $\langle P_2\rangle_{\text{Lipid}}$ & %
495     $\langle P_2\rangle_{\text{{\sc head}}}$ \\ \hline
496 mmeineke 1083 270 & 18.1 & 58.1 & 0.253 & 0.494 \\ \hline
497     275 & 17.2 & 56.7 & 0.295 & 0.514 \\ \hline
498     277 & 16.9 & 58.0 & 0.301 & 0.541 \\ \hline
499     280 & 17.4 & 58.0 & 0.274 & 0.488 \\ \hline
500     285 & 16.9 & 57.6 & 0.270 & 0.616 \\ \hline
501     290 & 17.0 & 57.6 & 0.263 & 0.534 \\ \hline
502     293 & 17.5 & 58.0 & 0.227 & 0.643 \\ \hline
503     300 & 16.9 & 57.6 & 0.315 & 0.536 \\ \hline
504     \end{tabular}
505     \end{center}
506     \end{table}
507 mmeineke 1087
508 mmeineke 1089 \subsection{\label{lipidSec:miscData}Further Structural Data}
509 mmeineke 1087
510     Also summarized in Table~\ref{lipidTab:blSummary}, are the bilayer
511 mmeineke 1089 thickness ($\langle L_{\perp}\rangle$) and area per lipid ($\langle
512     A_{\parallel}\rangle$). The bilayer thickness was measured from the
513     peak to peak {\sc head} atom distance in the density profiles. The
514 mmeineke 1087 area per lipid data compares favorably with values typically seen for
515 mmeineke 1089 DMPC (60.0~$\mbox{\AA}^2$ at 303~K)\cite{petrache00}. Although our
516 mmeineke 1087 values are lower this is most likely due to the shorter chain length
517     of our model (8 versus 14 for DMPC).
518    
519 mmeineke 1089 \subsection{\label{lipidSec:diffusion}Lateral Diffusion Constants}
520    
521     The lateral diffusion constant, $D_L$, is the constant characterizing
522     the diffusive motion of the lipid molecules within the plane of the bilayer. It
523     is given by the following Einstein relation:\cite{allen87:csl}
524     \begin{equation}
525     D_L = \lim_{t\rightarrow\infty}\frac{1}{4t}\langle |\mathbf{r}(t)
526 mmeineke 1114 - \mathbf{r}(0)|^2\rangle,
527 mmeineke 1089 \end{equation}
528 mmeineke 1114 where $\mathbf{r}(t)$ is the $xy$ position of the lipid at time $t$
529 mmeineke 1115 (assuming the $z$-axis is parallel to the bilayer normal). Calculating the $D_L$ involves first plotting the mean square displacement, $\langle |\mathbf{r}(t) - \mathbf{r}(0)|^2\rangle$, finding the slope at long times, and dividing the slope by 4 to give the diffusion constant (Fig.~\ref{lipidFig:msdFig}). When finding the slope only the 1~ns to 3~ns times are considered. Points at the longer times are not included due to the lack of good statistics at long time intervals.
530 mmeineke 1089
531     Fig.~\ref{lipidFig:diffusionFig} shows the lateral diffusion constants
532     as a function of temperature. There is a definite increase in the
533     lateral diffusion with higher temperatures, which is exactly what one
534     would expect with greater fluidity of the chains. However, the
535 mmeineke 1114 diffusion constants are two orders of magnitude larger than those
536     typical of DPPC ($\sim10^{-9}\text{cm}^2/\text{s}$ over this temperature range).\cite{Cevc87} This is what one would expect as the DPPC
537     molecule is sterically larger and heavier than our model, indicating that further modifications to the model should increase the lengths of the tail chains, and perhaps explore larger, more massive head groups. In contrast, the diffusion constant of
538 mmeineke 1089 the {\sc ssd} water, $9.84\times 10^{-6}\,\text{cm}^2/\text{s}$, is
539     reasonably close to the bulk water diffusion constant ($2.2999\times
540     10^{-5}\,\text{cm}^2/\text{s}$).\cite{Holz00}
541    
542     \begin{figure}
543     \centering
544 mmeineke 1112 \includegraphics[width=\linewidth]{msdFig.eps}
545 mmeineke 1115 \caption[Lateral mean square displacement for the phospholipid at 300~K]{This is a representative lateral mean square displacement for the center of mass motion of the phospholipid model. This particular example is from the 300~K simulation. The box is drawn about the region used in the calculation of the diffusion constant.}
546 mmeineke 1112 \label{lipidFig:msdFig}
547     \end{figure}
548    
549     \begin{figure}
550     \centering
551 mmeineke 1089 \includegraphics[width=\linewidth]{diffusionFig.eps}
552     \caption[The lateral diffusion constants versus temperature]{The lateral diffusion constants for the bilayers as a function of temperature.}
553     \label{lipidFig:diffusionFig}
554     \end{figure}
555    
556 mmeineke 1087 \subsection{\label{lipidSec:randBilayer}Bilayer Aggregation}
557    
558     A very important accomplishment for our model is its ability to
559     spontaneously form bilayers from a randomly dispersed starting
560     configuration. Fig.~\ref{lipidFig:blImage} shows an image sequence for
561 mmeineke 1115 the bilayer aggregation. After 1.0~ns, bulk aggregation has occured. By 5.0~ns, the basic bilayer aggregation can be seen, however there is a vertical lipid bridge connecting the periodic image of the bilayer to itself. At 15.0~ns, the lipid bridge has finally broken up, and the lipid molecules are starting to re-incorporate themselves into the bilayer. A water pore is still present through the membrane. In the last frame at 42.0~ns, the water pore is still present, although does show some signs of rejoining the bulk water section. These behaviors are typical for coarse grain model simulations, which can have lipid bridge lifetimes of up to 20~ns, and water pores typically lasting 3 to 25~ns.\cite{marrink04}
562 mmeineke 1087
563     \begin{figure}
564     \centering
565     \includegraphics[width=\linewidth]{bLayerImage.eps}
566     \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.}
567     \label{lipidFig:blImage}
568     \end{figure}
569    
570     \subsection{\label{lipidSec:randIrod}Inverted Rod Aggregation}
571    
572     Fig.~\ref{lipidFig:iRimage} shows a second aggregation sequence
573     simulated in this research. Here the fraction of water had been
574     significantly decreased to observe how the model would respond. After
575     1.5~ns, The main body of water in the system has already collected
576     into a central water channel. By 10.0~ns, the channel has widened
577 mmeineke 1089 slightly, but there are still many water molecules permeating the
578 mmeineke 1114 lipid macro-structure. At 35.0~ns, the central water channel has
579 mmeineke 1089 stabilized and several smaller water channels have been absorbed by
580     the main one. However, there is still an appreciable water
581     concentration throughout the lipid structure.
582 mmeineke 1087
583     \begin{figure}
584     \centering
585     \includegraphics[width=\linewidth]{iRodImage.eps}
586     \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}.}
587     \label{lipidFig:iRimage}
588     \end{figure}
589    
590     \section{\label{lipidSec:Conclusion}Conclusion}
591    
592 mmeineke 1089 We have presented a simple unified-atom phospholipid model capable of
593     spontaneous aggregation into a bilayer and an inverted rod
594     structure. The time scales of the macro-molecular aggregations are
595 mmeineke 1114 approximately 24~ns, with water permeation of the structures persisting for times longer than the scope of both aggregations. In addition the model's properties have been
596 mmeineke 1089 explored over a range of temperatures through prefabricated
597     bilayers. No freezing transition is seen in the temperature range of
598     our current simulations. However, structural information from 270~K
599     may imply that a freezing event is on a much longer time scale than
600     that explored in this current research. Further studies of this system
601     could extend the time length of the simulations at the low
602     temperatures to observe whether lipid crystallization can occur within
603     the framework of this model.
604    
605     Potential problems that may be obstacles in further research, is the
606     lack of detail in the head region. As the chains are almost directly
607     attached to the {\sc head} atom, there is no buffer between the
608     actions of the head group and the tails. Another disadvantage of the
609     model is the dipole approximation will alter results when details
610     concerning a charged solute's interactions with the bilayer. However,
611     it is important to keep in mind that the dipole approximation can be
612     kept an advantage by examining solutes that do not require point
613     charges, or at the least, require only dipole approximations
614     themselves. Other advantages of the model include the ability to alter
615     the size of the unified-atoms so that the size of the lipid can be
616     increased without adding to the number of interactions in the
617     system. However, what sets our model apart from other current
618     simplified models,\cite{goetz98,marrink04} is the information gained
619     by observing the ordering of the head groups dipole's in relation to
620     each other and the solvent without the need for point charges and the
621     Ewald sum.