ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/lipid.tex
Revision: 1061
Committed: Fri Feb 20 17:52:57 2004 UTC (20 years, 5 months ago) by mmeineke
Content type: application/x-tex
File size: 14554 byte(s)
Log Message:
cleaned up the references and the tables in the lipid section. also spell checked the thing.

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, computer speeds have allowed for the atomistic
8 simulation of phospholipid bilayers. These simulations have ranged
9 from simulation of the gel phase ($L_{\beta}$) of
10 dipalmitoylphosphatidylcholine (DPPC),\cite{lindahl00} to the
11 spontaneous aggregation of DPPC molecules into fluid phase
12 ($L_{\alpha}$) bilayers.\cite{marrink01} With the exception of a few
13 ambitious
14 simulations,\cite{marrink01:undulation,marrink:2002,lindahl00} most
15 investigations are limited to 64 to 256
16 phospholipids.\cite{lindahl00,sum:2003,venable00,gomez:2003,smondyrev:1999,marrink01}
17 This is due to the expense of the computer calculations involved when
18 performing these simulations. To properly hydrate a bilayer, one
19 typically needs 25 water molecules for every lipid, bringing the total
20 number of atoms simulated to roughly 8,000 for a system of 64 DPPC
21 molecules. Added to the difficulty is the electrostatic nature of the
22 phospholipid head groups and water, requiring the computationally
23 expensive Ewald sum or its slightly faster derivative particle mesh
24 Ewald sum.\cite{nina:2002,norberg:2000,patra:2003} These factors all
25 limit the potential size and time lengths of bilayer simulations.
26
27 Unfortunately, much of biological interest happens on time and length
28 scales infeasible with current simulation. One such example is the
29 observance of a ripple phase ($P_{\beta^{\prime}}$) between the
30 $L_{\beta}$ and $L_{\alpha}$ phases of certain phospholipid
31 bilayers.\cite{katsaras00,sengupta00} These ripples are shown to
32 have periodicity on the order of 100-200~$\mbox{\AA}$. A simulation on
33 this length scale would have approximately 1,300 lipid molecules with
34 an additional 25 water molecules per lipid to fully solvate the
35 bilayer. A simulation of this size is impractical with current
36 atomistic models.
37
38 Another class of simulations to consider, are those dealing with the
39 diffusion of molecules through a bilayer. Due to the fluid-like
40 properties of a lipid membrane, not all diffusion across the membrane
41 happens at pores. Some molecules of interest may incorporate
42 themselves directly into the membrane. Once here, they may possess an
43 appreciable waiting time (on the order of 10's to 100's of
44 nanoseconds) within the bilayer. Such long simulation times are
45 difficulty to obtain when integrating the system with atomistic
46 detail.
47
48 Addressing 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 $\frac{1}{r}$), eliminating the
72 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
82 The following is an outline of this chapter.
83 Sec.~\ref{lipidSec:Methods} is an introduction to the lipid model
84 used in these simulations. As well as clarification about the water
85 model and integration techniques. The various simulation setups
86 explored in this research are outlined in
87 Sec.~\ref{lipidSec:ExpSetup}. Sec.~\ref{lipidSec:Results} and
88 Sec.~\ref{lipidSec:Discussion} give a summary of the results and
89 interpretation of those results respectively. Finally, the
90 conclusions of this chapter are presented in
91 Sec.~\ref{lipidSec:Conclusion}.
92
93 \section{\label{lipidSec:Methods}Methods}
94
95 \subsection{\label{lipidSec:lipidModel}The Lipid Model}
96
97 \begin{figure}
98 \centering
99 \includegraphics[width=\linewidth]{twoChainFig.eps}
100 \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.}
101 \label{lipidFig:lipidModel}
102 \end{figure}
103
104 The phospholipid model used in these simulations is based on the
105 design illustrated in Fig.~\ref{lipidFig:lipidModel}. The head group
106 of the phospholipid is replaced by a single Lennard-Jones sphere of
107 diameter $\sigma_{\text{head}}$, with $\epsilon_{\text{head}}$ scaling
108 the well depth of its van der Walls interaction. This sphere also
109 contains a single dipole of magnitude $|\boldsymbol{\mu}|$, where
110 $|\boldsymbol{\mu}|$ can be varied to mimic the charge separation of a
111 given phospholipid head group. The atoms of the tail region are
112 modeled by unified atom beads. They are free of partial charges or
113 dipoles, containing only Lennard-Jones interaction sites at their
114 centers of mass. As with the head groups, their potentials can be
115 scaled by $\sigma_{\text{tail}}$ and $\epsilon_{\text{tail}}$.
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 \frac{3(\boldsymbol{\hat{u}}_i \cdot \mathbf{r}_{ij}) %
133 (\boldsymbol{\hat{u}}_j \cdot \mathbf{r}_{ij}) }
134 {r^{2}_{ij}} \biggr]
135 \label{lipidEq:dipolePot}
136 \end{equation}
137 Where $V_{\text{LJ}}$ is the Lennard-Jones potential and
138 $V_{\text{dipole}}$ is the dipole-dipole potential. As previously
139 stated, $\sigma_{ij}$ and $\epsilon_{ij}$ are the Lennard-Jones
140 parameters which scale the length and depth of the interaction
141 respectively, and $r_{ij}$ is the distance between beads $i$ and $j$.
142 In $V_{\text{dipole}}$, $\mathbf{r}_{ij}$ is the vector starting at
143 bead$i$ and pointing towards bead $j$. Vectors $\mathbf{\Omega}_i$
144 and $\mathbf{\Omega}_j$ are the orientational degrees of freedom for
145 beads $i$ and $j$. $|\mu_i|$ is the magnitude of the dipole moment of
146 $i$, and $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
147 vector of $\boldsymbol{\Omega}_i$.
148
149 The model also allows for the bonded interactions bends, and torsions.
150 The bond between two beads on a chain is of fixed length, and is
151 maintained according to the {\sc rattle} algorithm.\cite{andersen83}
152 The bends are subject to a harmonic potential:
153 \begin{equation}
154 V_{\text{bend}}(\theta) = k_{\theta}( \theta - \theta_0 )^2
155 \label{lipidEq:bendPot}
156 \end{equation}
157 where $k_{\theta}$ scales the strength of the harmonic well, and
158 $\theta$ is the angle between bond vectors
159 (Fig.~\ref{lipidFig:lipidModel}). In addition, we have placed a
160 ``ghost'' bend on the phospholipid head. The ghost bend adds a
161 potential to keep the dipole pointed along the bilayer surface, where
162 $theta$ is now the angle the dipole makes with respect to the {\sc
163 head}-$\text{{\sc ch}}_2$ bond vector. The torsion potential is given
164 by:
165 \begin{equation}
166 V_{\text{torsion}}(\phi) =
167 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0
168 \label{lipidEq:torsionPot}
169 \end{equation}
170 Here, the parameters $k_0$, $k_1$, $k_2$, and $k_3$ fit the cosine
171 power series to the desired torsion potential surface, and $\phi$ is
172 the angle the two end atoms have rotated about the middle bond
173 (Fig.:\ref{lipidFig:lipidModel}). Long range interactions such as the
174 Lennard-Jones potential are excluded for atom pairs involved in the
175 same bond, bend, or torsion. However, internal interactions not
176 directly involved in a bonded pair are calculated.
177
178 All simulations presented here use a two chained lipid as pictured in
179 Fig.~\ref{lipidFig:lipidModel}. The chains are both eight beads long,
180 and their mass and Lennard Jones parameters are summarized in
181 Table~\ref{lipidTable:tcLJParams}. The magnitude of the dipole moment
182 for the head bead is 10.6~Debye, and the bend and torsion parameters
183 are summarized in Table~\ref{lipidTable:tcBendParams} and
184 \ref{lipidTable:tcTorsionParams}.
185
186 \begin{table}
187 \caption{The Lennard Jones Parameters for the two chain phospholipids.}
188 \label{lipidTable:tcLJParams}
189 \begin{center}
190 \begin{tabular}{|l|c|c|c|}
191 \hline
192 & mass (amu) & $\sigma$($\mbox{\AA}$) & $\epsilon$ (kcal/mol) \\ \hline
193 {\sc head} & 72 & 4.0 & 0.185 \\ \hline
194 {\sc ch}\cite{Siepmann1998} & 13.02 & 4.0 & 0.0189 \\ \hline
195 $\text{{\sc ch}}_2$\cite{Siepmann1998} & 14.03 & 3.95 & 0.18 \\ \hline
196 $\text{{\sc ch}}_3$\cite{Siepmann1998} & 15.04 & 3.75 & 0.25 \\ \hline
197 {\sc ssd}\cite{liu96:new_model} & 18.03 & 3.051 & 0.152 \\ \hline
198 \end{tabular}
199 \end{center}
200 \end{table}
201
202 \begin{table}
203 \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}}
204 \label{lipidTable:tcBendParams}
205 \begin{center}
206 \begin{tabular}{|l|c|c|}
207 \hline
208 & $k_{\theta}$ ( kcal/($\text{mol deg}^2$) ) & $\theta_0$ ( deg ) \\ \hline
209 {\sc ghost}-{\sc head}-$\text{{\sc ch}}_2$ & 0.00177 & 129.78 \\ \hline
210 $x$-{\sc ch}-$y$ & 58.84 & 112.0 \\ \hline
211 $x$-$\text{{\sc ch}}_2$-$y$ & 58.84 & 114.0 \\ \hline
212 \end{tabular}
213 \end{center}
214 \end{table}
215
216 \begin{table}
217 \caption[Torsion Parameters for the two chain phospholipids]{Torsion Parameters for the two chain phospholipids. Alkane parameters based on TraPPE.\cite{Siepmann1998}}
218 \label{lipidTable:tcTorsionParams}
219 \begin{center}
220 \begin{tabular}{|l|c|c|c|c|}
221 \hline
222 All are in kcal/mol $\rightarrow$ & $k_3$ & $k_2$ & $k_1$ & $k_0$ \\ \hline
223 $x$-{\sc ch}-$y$-$z$ & 3.3254 & -0.4215 & -1.686 & 1.1661 \\ \hline
224 $x$-$\text{{\sc ch}}_2$-$\text{{\sc ch}}_2$-$y$ & 5.9602 & -0.568 & -3.802 & 2.1586 \\ \hline
225 \end{tabular}
226 \end{center}
227 \end{table}
228
229
230 \section{\label{lipidSec:furtherMethod}Further Methodology}
231
232 As mentioned previously, the water model used throughout these
233 simulations was the {\sc ssd} model of
234 Ichiye.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md} A
235 discussion of the model can be found in Sec.~\ref{oopseSec:SSD}. As
236 for the integration of the equations of motion, all simulations were
237 performed in an orthorhombic periodic box with a thermostat on
238 velocities, and an independent barostat on each Cartesian axis $x$,
239 $y$, and $z$. This is the $\text{NPT}_{xyz}$. ensemble described in
240 Sec.~\ref{oopseSec:Ensembles}.
241
242
243 \subsection{\label{lipidSec:ExpSetup}Experimental Setup}
244
245 Two main starting configuration classes were used in this research:
246 random and ordered bilayers. The ordered bilayer starting
247 configurations were all started from an equilibrated bilayer at
248 300~K. The original configuration for the first 300~K run was
249 assembled by placing the phospholipids centers of mass on a planar
250 hexagonal lattice. The lipids were oriented with their long axis
251 perpendicular to the plane. The second leaf simply mirrored the first
252 leaf, and the appropriate number of waters were then added above and
253 below the bilayer.
254
255 The random configurations took more work to generate. To begin, a
256 test lipid was placed in a simulation box already containing water at
257 the intended density. The waters were then tested for overlap with
258 the lipid using a 5.0~$\mbox{\AA}$ buffer distance. This gave an
259 estimate for the number of waters each lipid would displace in a
260 simulation box. A target number of waters was then defined which
261 included the number of waters each lipid would displace, the number of
262 waters desired to solvate each lipid, and a fudge factor to pad the
263 initialization.
264
265 Next, a cubic simulation box was created that contained at least the
266 target number of waters in an FCC lattice (the lattice was for ease of
267 placement). What followed was a RSA simulation similar to those of
268 Chapt.~\ref{chapt:RSA}. The lipids were sequentially given a random
269 position and orientation within the box. If a lipid's position caused
270 atomic overlap with any previously adsorbed lipid, its position and
271 orientation were rejected, and a new random adsorption site was
272 attempted. The RSA simulation proceeded until all phospholipids had
273 been adsorbed. After adsorption, all water molecules with locations
274 that overlapped with the atomic coordinates of the lipids were
275 removed.
276
277 Finally, water molecules were removed one by one at random until the
278 desired number of waters per lipid was reached. The typical low final
279 density for these initial configurations was not a problem, as the box
280 would shrink to an appropriate size within the first 50~ps of a
281 simulation in the $\text{NPT}_{xyz}$ ensemble.
282
283 \subsection{\label{lipidSec:Configs}The simulation configurations}
284
285 Table ~\ref{lipidTable:simNames} summarizes the names and important
286 details of the simulations. The B set of simulations were all started
287 in an ordered bilayer and observed over a period of 10~ns. Simulation
288 RL was integrated for approximately 20~ns starting from a random
289 configuration as an example of spontaneous bilayer aggregation.
290 Lastly, simulation RH was also started from a random configuration,
291 but with a lesser water content and higher temperature to show the
292 spontaneous aggregation of an inverted hexagonal lamellar phase.