ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Lipid.tex
Revision: 2807
Committed: Wed Jun 7 01:49:15 2006 UTC (18 years ago) by tim
Content type: application/x-tex
File size: 17938 byte(s)
Log Message:
more reference fixes

File Contents

# Content
1 \chapter{\label{chapt:lipid}LIPID MODELING}
2
3 \section{\label{lipidSection:introduction}Introduction}
4
5 Under biologically relevant conditions, phospholipids are solvated
6 in aqueous solutions at a roughly 25:1 ratio. Solvation can have a
7 tremendous impact on transport phenomena in biological membranes
8 since it can affect the dynamics of ions and molecules that are
9 transferred across membranes. Studies suggest that because of the
10 directional hydrogen bonding ability of the lipid headgroups, a
11 small number of water molecules are strongly held around the
12 different parts of the headgroup and are oriented by them with
13 residence times for the first hydration shell being around 0.5 - 1
14 ns\cite{Ho1992}. In the second solvation shell, some water molecules
15 are weakly bound, but are still essential for determining the
16 properties of the system. Transport of various molecular species
17 into living cells is one of the major functions of membranes. A
18 thorough understanding of the underlying molecular mechanism for
19 solute diffusion is crucial to the further studies of other related
20 biological processes. All transport across cell membranes takes
21 place by one of two fundamental processes: Passive transport is
22 driven by bulk or inter-diffusion of the molecules being transported
23 or by membrane pores which facilitate crossing. Active transport
24 depends upon the expenditure of cellular energy in the form of ATP
25 hydrolysis. As the central processes of membrane assembly,
26 translocation of phospholipids across membrane bilayers requires the
27 hydrophilic head of the phospholipid to pass through the highly
28 hydrophobic interior of the membrane, and for the hydrophobic tails
29 to be exposed to the aqueous environment\cite{Sasaki2004}. A number
30 of studies indicate that the flipping of phospholipids occurs
31 rapidly in the eukaryotic ER and the bacterial cytoplasmic membrane
32 via a bi-directional, facilitated diffusion process requiring no
33 metabolic energy input. Another system of interest would be the
34 distribution of sites occupied by inhaled anesthetics in membrane.
35 Although the physiological effects of anesthetics have been
36 extensively studied, the controversy over their effects on lipid
37 bilayers still continues. Recent deuterium NMR measurements on
38 halothane in POPC lipid bilayers suggest the anesthetics are
39 primarily located at the hydrocarbon chain region\cite{Baber1995}.
40 Infrared spectroscopy experiments suggest that halothane in DMPC
41 lipid bilayers lives near the membrane/water
42 interface\cite{Lieb1982}.
43
44 Molecular dynamics simulations have proven to be a powerful tool for
45 studying the functions of biological systems, providing structural,
46 thermodynamic and dynamical information. Unfortunately, much of
47 biological interest happens on time and length scales well beyond
48 the range of current simulation technologies.
49 %review of coarse-grained modeling
50 Several schemes are proposed in this chapter to overcome these
51 difficulties.
52
53 \section{\label{lipidSection:model}Model}
54
55 \subsection{\label{lipidSection:SSD}The Soft Sticky Dipole Water Model}
56
57 In a typical bilayer simulation, the dominant portion of the
58 computation time will be spent calculating water-water interactions.
59 As an efficient solvent model, the Soft Sticky Dipole (SSD) water
60 model\cite{Chandra1999,Fennell2004} is used as the explicit solvent
61 in this project. Unlike other water models which have partial
62 charges distributed throughout the whole molecule, the SSD water
63 model consists of a single site which is a Lennard-Jones interaction
64 site, as well as a point dipole. A tetrahedral potential is added to
65 correct for hydrogen bonding. The following equation describes the
66 interaction between two water molecules:
67 \[
68 V_{SSD} = V_{LJ} (r_{ij} ) + V_{dp} (r_{ij} ,\Omega _i ,\Omega _j )
69 + V_{sticky} (r_{ij} ,\Omega _i ,\Omega _j )
70 \]
71 where $r_{ij}$ is the vector between molecule $i$ and molecule $j$,
72 $\Omega _i$ and $\Omega _j$ are the orientational degrees of freedom
73 for molecule $i$ and molecule $j$ respectively.
74 \[
75 V_{LJ} (r_{ij} ) = 4\varepsilon _{ij} \left[ {\left( {\frac{{\sigma
76 _{ij} }}{{r_{ij} }}} \right)^{12} - \left( {\frac{{\sigma _{ij}
77 }}{{r_{ij} }}} \right)^6 } \right]
78 \]
79 \[
80 V_{dp} (r_{ij} ,\Omega _i ,\Omega _j ) = \frac{1}{{4\pi \varepsilon
81 _0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{r_{ij}^3 }} -
82 \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
83 r_{ij} } \right)}}{{r_{ij}^5 }}} \right]
84 \]
85 \[
86 V_{sticky} (r_{ij} ,\Omega _i ,\Omega _j ) = v_0 [s(r_{ij} )w(r_{ij}
87 ,\Omega _i ,\Omega _j ) + s'(r_{ij} )w'(r_{ij} ,\Omega _i ,\Omega _j
88 )]
89 \]
90 where $v_0$ is a strength parameter, $s$ and $s'$ are cubic
91 switching functions, while $w$ and $w'$ are responsible for the
92 tetrahedral potential and the short-range correction to the dipolar
93 interaction respectively.
94 \[
95 \begin{array}{l}
96 w(r_{ij} ,\Omega _i ,\Omega _j ) = \sin \theta _{ij} \sin 2\theta _{ij} \cos 2\phi _{ij} \\
97 w'(r_{ij} ,\Omega _i ,\Omega _j ) = (\cos \theta _{ij} - 0.6)^2 (\cos \theta _{ij} + 0.8)^2 - w_0 \\
98 \end{array}
99 \]
100 Although dipole-dipole and sticky interactions are more
101 mathematically complicated than Coulomb interactions, the number of
102 pair interactions is reduced dramatically both because the model
103 only contains a single-point as well as "short range" nature of the
104 higher order interaction.
105
106 \subsection{\label{lipidSection:coarseGrained}The Coarse-Grained Phospholipid Model}
107
108 Fig.~\ref{lipidFigure:coarseGrained} shows a schematic for our
109 coarse-grained phospholipid model. The lipid head group is modeled
110 by a linear rigid body which consists of three Lennard-Jones spheres
111 and a centrally located point-dipole. The backbone atoms in the
112 glycerol motif are modeled by Lennard-Jones spheres with dipoles.
113 Alkyl groups in hydrocarbon chains are replaced with unified
114 $\text{{\sc CH}}_2$ or $\text{{\sc CH}}_3$ atoms.
115
116 \begin{figure}
117 \centering
118 \includegraphics[width=3in]{coarse_grained.eps}
119 \caption[A representation of coarse-grained phospholipid model]{}
120 \label{lipidFigure:coarseGrained}
121 \end{figure}
122
123 Accurate and efficient computation of electrostatics is one of the
124 most difficult tasks in molecular modeling. Traditionally, the
125 electrostatic interaction between two molecular species is
126 calculated as a sum of interactions between pairs of point charges,
127 using Coulomb's law:
128 \[
129 V = \sum\limits_{i = 1}^{N_A } {\sum\limits_{j = 1}^{N_B }
130 {\frac{{q_i q_j }}{{4\pi \varepsilon _0 r_{ij} }}} }
131 \]
132 where $N_A$ and $N_B$ are the number of point charges in the two
133 molecular species. Originally developed to study ionic crystals, the
134 Ewald summation method mathematically transforms this
135 straightforward but conditionally convergent summation into two more
136 complicated but rapidly convergent sums. One summation is carried
137 out in reciprocal space while the other is carried out in real
138 space. An alternative approach is a multipole expansion, which is
139 based on electrostatic moments, such as charge (monopole), dipole,
140 quadruple etc.
141
142 Here we consider a linear molecule which consists of two point
143 charges $\pm q$ located at $ ( \pm \frac{d}{2},0)$. The
144 electrostatic potential at point $P$ is given by:
145 \[
146 \frac{1}{{4\pi \varepsilon _0 }}\left( {\frac{{ - q}}{{r_ - }} +
147 \frac{{ + q}}{{r_ + }}} \right) = \frac{1}{{4\pi \varepsilon _0
148 }}\left( {\frac{{ - q}}{{\sqrt {r^2 + \frac{{d^2 }}{4} + rd\cos
149 \theta } }} + \frac{q}{{\sqrt {r^2 + \frac{{d^2 }}{4} - rd\cos
150 \theta } }}} \right)
151 \]
152
153 \begin{figure}
154 \centering
155 \includegraphics[width=3in]{charge_dipole.eps}
156 \caption[Electrostatic potential due to a linear molecule comprising
157 two point charges]{Electrostatic potential due to a linear molecule
158 comprising two point charges} \label{lipidFigure:chargeDipole}
159 \end{figure}
160
161 The basic assumption of the multipole expansion is $r \gg d$ , thus,
162 $\frac{{d^2 }}{4}$ inside the square root of the denominator is
163 neglected. This is a reasonable approximation in most cases.
164 Unfortunately, in our headgroup model, the distance of charge
165 separation $d$ (4.63 $\AA$ in PC headgroup) may be comparable to
166 $r$. Nevertheless, we could still assume $ \cos \theta \approx 0$
167 in the central region of the headgroup. Using Taylor expansion and
168 associating appropriate terms with electric moments will result in a
169 "split-dipole" approximation:
170 \[
171 V(r) = \frac{1}{{4\pi \varepsilon _0 }}\frac{{r\mu \cos \theta
172 }}{{R^3 }}
173 \]
174 where$R = \sqrt {r^2 + \frac{{d^2 }}{4}}$ Placing a dipole at point
175 $P$ and applying the same strategy, the interaction between two
176 split-dipoles is then given by:
177 \[
178 V_{sd} (r_{ij} ,\Omega _i ,\Omega _j ) = \frac{1}{{4\pi \varepsilon
179 _0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{R_{ij}^3 }} -
180 \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
181 r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
182 \]
183 where $\mu _i$ and $\mu _j$ are the dipole moment of molecule $i$
184 and molecule $j$ respectively, $r_{ij}$ is vector between molecule
185 $i$ and molecule $j$, and $R_{ij}$ is given by,
186 \[
187 R_{ij} = \sqrt {r_{ij}^2 + \frac{{d_i^2 }}{4} + \frac{{d_j^2
188 }}{4}}
189 \]
190 where $d_i$ and $d_j$ are the charge separation distance of dipole
191 and respectively. This approximation to the multipole expansion
192 maintains the fast fall-off of the multipole potentials but lacks
193 the normal divergences when two polar groups get close to one
194 another.
195 %description of the comparsion
196 \begin{figure}
197 \centering
198 \includegraphics[width=\linewidth]{split_dipole.eps}
199 \caption[Comparison between electrostatic approximation]{Electron
200 density profile along the bilayer normal.}
201 \label{lipidFigure:splitDipole}
202 \end{figure}
203
204 %\section{\label{lipidSection:methods}Methods}
205
206 \section{\label{lipidSection:resultDiscussion}Results and Discussion}
207
208 \subsection{One Lipid in Sea of Water Molecules}
209
210 To exclude the inter-headgroup interaction, atomistic models of one
211 lipid (DMPC or DLPE) in sea of water molecules (TIP3P) were built
212 and studied using atomistic molecular dynamics. The simulation was
213 analyzed using a set of radial distribution functions, which give
214 the probability of finding a pair of molecular species a distance
215 apart, relative to the probability expected for a completely random
216 distribution function at the same density.
217
218 \begin{equation}
219 g_{AB} (r) = \frac{1}{{\rho _B }}\frac{1}{{N_A }} < \sum\limits_{i
220 \in A} {\sum\limits_{j \in B} {\delta (r - r_{ij} )} } >
221 \end{equation}
222 \begin{equation}
223 g_{AB} (r,\cos \theta ) = \frac{1}{{\rho _B }}\frac{1}{{N_A }} <
224 \sum\limits_{i \in A} {\sum\limits_{j \in B} {\delta (r - r_{ij} )}
225 } \delta (\cos \theta _{ij} - \cos \theta ) >
226 \end{equation}
227
228 From figure 4(a), we can identify the first solvation shell (3.5
229 $\AA$) and the second solvation shell (5.0 $\AA$) from both plots.
230 However, the corresponding orientations are different. In DLPE,
231 water molecules prefer to sit around -NH3 group due to the hydrogen
232 bonding. In contrast, because of the hydrophobic effect of the
233 -N(CH3)3 group, the preferred position of water molecules in DMPC is
234 around the -PO4 Group. When the water molecules are far from the
235 headgroup, the distribution of the two angles should be uniform. The
236 correlation close to center of the headgroup dipole (< 5 $\AA$) in
237 both plots tell us that in the closely-bound region, the dipoles of
238 the water molecules are preferentially anti-aligned with the dipole
239 of headgroup.
240
241 \begin{figure}
242 \centering
243 \includegraphics[width=\linewidth]{g_atom.eps}
244 \caption[The pair correlation functions for atomistic models]{}
245 \label{lipidFigure:PCFAtom}
246 \end{figure}
247
248 The initial configurations of coarse-grained systems are constructed
249 from the previous atomistic ones. The parameters for the
250 coarse-grained model in Table~\ref{lipidTable:parameter} are
251 estimated and tuned using isothermal-isobaric molecular dynamics.
252 Pair distribution functions calculated from coarse-grained models
253 preserve the basic characteristics of the atomistic simulations. The
254 water density, measured in a head-group-fixed reference frame,
255 surrounding two phospholipid headgroups is shown in
256 Fig.~\ref{lipidFigure:PCFCoarse}. It is clear that the phosphate end
257 in DMPC and the amine end in DMPE are the two most heavily solvated
258 atoms.
259
260 \begin{figure}
261 \centering
262 \includegraphics[width=\linewidth]{g_coarse.eps}
263 \caption[The pair correlation functions for coarse-grained models]{}
264 \label{lipidFigure:PCFCoarse}
265 \end{figure}
266
267 \begin{figure}
268 \centering
269 \includegraphics[width=\linewidth]{EWD_coarse.eps}
270 \caption[Excess water density of coarse-grained phospholipids]{ }
271 \label{lipidFigure:EWDCoarse}
272 \end{figure}
273
274 \begin{table}
275 \caption{The Parameters For Coarse-grained Phospholipids}
276 \label{lipidTable:parameter}
277 \begin{center}
278 \begin{tabular}{|l|c|c|c|c|c|}
279 \hline
280 % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
281 Atom type & Mass & $\sigma$ & $\epsilon$ & charge & Dipole \\
282 $\text{{\sc CH}}_2$ & 14.03 & 3.95 & 0.0914 & 0 & 0 \\
283 $\text{{\sc CH}}_3$ & 15.04 & 3.75 & 0.195 & 0 & 0 \\
284 $\text{{\sc CE}}$ & 28.01 & 3.427& 0.294 & 0 & 1.693 \\
285 $\text{{\sc CK}}$ & 28.01 & 3.592& 0.311 & 0 & 2.478 \\
286 $\text{{\sc PO}}_4$ & 108.995& 3.9 & 1.88 & -1& 0 \\
287 $\text{{\sc HDP}}$ & 14.03 & 4.0 & 0.13 & 0 & 0 \\
288 $\text{{\sc NC}}_4$ & 73.137 & 4.9 & 0.88 & +1& 0 \\
289 $\text{{\sc NH}}_3$ & 17.03 & 3.25 & 0.17 & +1& 0\\
290 \hline
291 \end{tabular}
292 \end{center}
293 \end{table}
294
295 \subsection{Bilayer Simulations Using Coarse-grained Model}
296
297 A bilayer system consisting of 128 DMPC lipids and 3655 water
298 molecules has been constructed from an atomistic coordinate
299 file.[15] The MD simulation is performed at constant temperature, T
300 = 300K, and constant pressure, p = 1 atm, and consisted of an
301 equilibration period of 2 ns. During the equilibration period, the
302 system was initially simulated at constant volume for 1ns. Once the
303 system was equilibrated at constant volume, the cell dimensions of
304 the system was relaxed by performing under NPT conditions using
305 Nos¨¦-Hoover extended system isothermal-isobaric dynamics. After
306 equilibration, different properties were evaluated over a production
307 run of 5 ns.
308
309 \begin{figure}
310 \centering
311 \includegraphics[width=\linewidth]{bilayer.eps}
312 \caption[Image of a coarse-grained bilayer system]{A coarse-grained
313 bilayer system consisting of 128 DMPC lipids and 3655 SSD water
314 molecules.}
315 \label{lipidFigure:bilayer}
316 \end{figure}
317
318 \subsubsection{Electron Density Profile (EDP)}
319
320 Assuming a gaussian distribution of electrons on each atomic center
321 with a variance estimated from the size of the van der Waals radius,
322 the EDPs which are proportional to the density profiles measured
323 along the bilayer normal obtained by x-ray scattering experiment,
324 can be expressed by\cite{Tu1995}
325 \begin{equation}
326 \rho _{x - ray} (z)dz \propto \sum\limits_{i = 1}^N {\frac{{n_i
327 }}{V}\frac{1}{{\sqrt {2\pi \sigma ^2 } }}e^{ - (z - z_i )^2 /2\sigma
328 ^2 } dz},
329 \end{equation}
330 where $\sigma$ is the variance equal to the van der Waals radius,
331 $n_i$ is the atomic number of site $i$ and $V$ is the volume of the
332 slab between $z$ and $z+dz$ . The highest density of total EDP
333 appears at the position of lipid-water interface corresponding to
334 headgroup, glycerol, and carbonyl groups of the lipids and the
335 distribution of water locked near the head groups, while the lowest
336 electron density is in the hydrocarbon region. As a good
337 approximation to the thickness of the bilayer, the headgroup spacing
338 , is defined as the distance between two peaks in the electron
339 density profile, calculated from our simulations to be 34.1 $\AA$.
340 This value is close to the x-ray diffraction experimental value 34.4
341 $\AA$\cite{Petrache1998}.
342
343 \begin{figure}
344 \centering
345 \includegraphics[width=\linewidth]{electron_density.eps}
346 \caption[The density profile of the lipid bilayers]{Electron density
347 profile along the bilayer normal. The water density is shown in red,
348 the density due to the headgroups in green, the glycerol backbone in
349 brown, $\text{{\sc CH}}_2$ in yellow, $\text{{\sc CH}}_3$ in cyan,
350 and total density due to DMPC in blue.}
351 \label{lipidFigure:electronDensity}
352 \end{figure}
353
354 \subsubsection{$\text{S}_{\text{{\sc cd}}}$ Order Parameter}
355
356 Measuring deuterium order parameters by NMR is a useful technique to
357 study the orientation of hydrocarbon chains in phospholipids. The
358 order parameter tensor $S$ is defined by:
359 \begin{equation}
360 S_{ij} = \frac{1}{2} < 3\cos \theta _i \cos \theta _j - \delta
361 _{ij} >
362 \end{equation}
363 where $\theta$ is the angle between the $i$th molecular axis and
364 the bilayer normal ($z$ axis). The brackets denote an average over
365 time and molecules. The molecular axes are defined:
366 \begin{itemize}
367 \item $\mathbf{\hat{z}}$ is the vector from $C_{n-1}$ to $C_{n+1}$.
368 \item $\mathbf{\hat{y}}$ is the vector that is perpendicular to $z$ and
369 in the plane through $C_{n-1}$, $C_{n}$, and $C_{n+1}$.
370 \item $\mathbf{\hat{x}}$ is the vector perpendicular to
371 $\mathbf{\hat{y}}$ and $\mathbf{\hat{z}}$.
372 \end{itemize}
373 In coarse-grained model, although there are no explicit hydrogens,
374 the order parameter can still be written in terms of carbon ordering
375 at each point of the chain\cite{Egberts1988}
376 \begin{equation}
377 S_{ij} = \frac{1}{2} < 3\cos \theta _i \cos \theta _j - \delta
378 _{ij} >.
379 \end{equation}
380
381 Fig.~\ref{lipidFigure:Scd} shows the order parameter profile
382 calculated for our coarse-grained DMPC bilayer system at 300K. Also
383 shown are the experimental data of Tu\cite{Tu1995}. The fact that
384 $\text{S}_{\text{{\sc cd}}}$ order parameters calculated from
385 simulation are higher than the experimental ones is ascribed to the
386 assumption of the locations of implicit hydrogen atoms which are
387 fixed in coarse-grained models at positions relative to the CC
388 vector.
389
390 \begin{figure}
391 \centering
392 \includegraphics[width=\linewidth]{scd.eps}
393 \caption[$\text{S}_{\text{{\sc cd}}}$ order parameter]{A comparison
394 of $|\text{S}_{\text{{\sc cd}}}|$ between coarse-grained model
395 (blue) and DMPC\cite{petrache00} (black) near 300~K.}
396 \label{lipidFigure:Scd}
397 \end{figure}
398
399 %\subsection{Bilayer Simulations Using Gay-Berne Ellipsoid Model}