ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Lipid.tex
Revision: 2781
Committed: Wed May 31 04:29:44 2006 UTC (18 years, 1 month ago) by tim
Content type: application/x-tex
File size: 17737 byte(s)
Log Message:
begin Liquid Crystal Chapter

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