ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Lipid.tex
Revision: 2778
Committed: Fri May 26 17:56:36 2006 UTC (18 years, 1 month ago) by tim
Content type: application/x-tex
File size: 17119 byte(s)
Log Message:
minor change

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 Figure 1 shows a schematic for our coarse-grained phospholipid
105 model. The lipid head group is modeled by a linear rigid body which
106 consists of three Lennard-Jones spheres and a centrally located
107 point-dipole. The backbone atoms in the glycerol motif are modeled
108 by Lennard-Jones spheres with dipoles. Alkyl groups in hydrocarbon
109 chains are replaced with unified CH2 or CH3 atoms.
110
111 Accurate and efficient computation of electrostatics is one of the
112 most difficult tasks in molecular modeling. Traditionally, the
113 electrostatic interaction between two molecular species is
114 calculated as a sum of interactions between pairs of point charges,
115 using Coulomb's law:
116 \[
117 V = \sum\limits_{i = 1}^{N_A } {\sum\limits_{j = 1}^{N_B }
118 {\frac{{q_i q_j }}{{4\pi \varepsilon _0 r_{ij} }}} }
119 \]
120 where $N_A$ and $N_B$ are the number of point charges in the two
121 molecular species. Originally developed to study ionic crystals, the
122 Ewald summation method mathematically transforms this
123 straightforward but conditionally convergent summation into two more
124 complicated but rapidly convergent sums. One summation is carried
125 out in reciprocal space while the other is carried out in real
126 space. An alternative approach is a multipole expansion, which is
127 based on electrostatic moments, such as charge (monopole), dipole,
128 quadruple etc.
129
130 Here we consider a linear molecule which consists of two point
131 charges $\pm q$ located at $ ( \pm \frac{d}{2},0)$. The
132 electrostatic potential at point $P$ is given by:
133 \[
134 \frac{1}{{4\pi \varepsilon _0 }}\left( {\frac{{ - q}}{{r_ - }} +
135 \frac{{ + q}}{{r_ + }}} \right) = \frac{1}{{4\pi \varepsilon _0
136 }}\left( {\frac{{ - q}}{{\sqrt {r^2 + \frac{{d^2 }}{4} + rd\cos
137 \theta } }} + \frac{q}{{\sqrt {r^2 + \frac{{d^2 }}{4} - rd\cos
138 \theta } }}} \right)
139 \]
140
141 The basic assumption of the multipole expansion is $r \gg d$ , thus,
142 $\frac{{d^2 }}{4}$ inside the square root of the denominator is
143 neglected. This is a reasonable approximation in most cases.
144 Unfortunately, in our headgroup model, the distance of charge
145 separation $d$ (4.63 $\AA$ in PC headgroup) may be comparable to
146 $r$. Nevertheless, we could still assume $ \cos \theta \approx 0$
147 in the central region of the headgroup. Using Taylor expansion and
148 associating appropriate terms with electric moments will result in a
149 "split-dipole" approximation:
150 \[
151 V(r) = \frac{1}{{4\pi \varepsilon _0 }}\frac{{r\mu \cos \theta
152 }}{{R^3 }}
153 \]
154 where$R = \sqrt {r^2 + \frac{{d^2 }}{4}}$ Placing a dipole at point
155 $P$ and applying the same strategy, the interaction between two
156 split-dipoles is then given by:
157 \[
158 V_{sd} (r_{ij} ,\Omega _i ,\Omega _j ) = \frac{1}{{4\pi \varepsilon
159 _0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{R_{ij}^3 }} -
160 \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
161 r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
162 \]
163 where $\mu _i$ and $\mu _j$ are the dipole moment of molecule $i$
164 and molecule $j$ respectively, $r_{ij}$ is vector between molecule
165 $i$ and molecule $j$, and $R_{ij{$ is given by,
166 \[
167 R_{ij} = \sqrt {r_{ij}^2 + \frac{{d_i^2 }}{4} + \frac{{d_j^2
168 }}{4}}
169 \]
170 where $d_i$ and $d_j$ are the charge separation distance of dipole
171 and respectively. This approximation to the multipole expansion
172 maintains the fast fall-off of the multipole potentials but lacks
173 the normal divergences when two polar groups get close to one
174 another.
175
176 \begin{figure}
177 \centering
178 \includegraphics[width=\linewidth]{split_dipole.eps}
179 \caption[Comparison between electrostatic approximation]{Electron
180 density profile along the bilayer normal.}
181 \label{lipidFigure:splitDipole}
182 \end{figure}
183
184 %\section{\label{lipidSection:methods}Methods}
185
186 \section{\label{lipidSection:resultDiscussion}Results and Discussion}
187
188 \subsection{One Lipid in Sea of Water Molecules}
189
190 To exclude the inter-headgroup interaction, atomistic models of one
191 lipid (DMPC or DLPE) in sea of water molecules (TIP3P) were built
192 and studied using atomistic molecular dynamics. The simulation was
193 analyzed using a set of radial distribution functions, which give
194 the probability of finding a pair of molecular species a distance
195 apart, relative to the probability expected for a completely random
196 distribution function at the same density.
197
198 \begin{equation}
199 g_{AB} (r) = \frac{1}{{\rho _B }}\frac{1}{{N_A }} < \sum\limits_{i
200 \in A} {\sum\limits_{j \in B} {\delta (r - r_{ij} )} } >
201 \end{equation}
202 \begin{equation}
203 g_{AB} (r,\cos \theta ) = \frac{1}{{\rho _B }}\frac{1}{{N_A }} <
204 \sum\limits_{i \in A} {\sum\limits_{j \in B} {\delta (r - r_{ij} )}
205 } \delta (\cos \theta _{ij} - \cos \theta ) >
206 \end{equation}
207
208 From figure 4(a), we can identify the first solvation shell (3.5
209 $\AA$) and the second solvation shell (5.0 $\AA$) from both plots.
210 However, the corresponding orientations are different. In DLPE,
211 water molecules prefer to sit around -NH3 group due to the hydrogen
212 bonding. In contrast, because of the hydrophobic effect of the
213 -N(CH3)3 group, the preferred position of water molecules in DMPC is
214 around the -PO4 Group. When the water molecules are far from the
215 headgroup, the distribution of the two angles should be uniform. The
216 correlation close to center of the headgroup dipole (< 5 $\AA$) in
217 both plots tell us that in the closely-bound region, the dipoles of
218 the water molecules are preferentially anti-aligned with the dipole
219 of headgroup.
220
221 \begin{figure}
222 \centering
223 \includegraphics[width=\linewidth]{g_atom.eps}
224 \caption[The pair correlation functions for atomistic models]{}
225 \label{lipidFigure:PCFAtom}
226 \end{figure}
227
228 The initial configurations of coarse-grained systems are constructed
229 from the previous atomistic ones. The parameters for the
230 coarse-grained model in Table~\ref{lipidTable:parameter} are
231 estimated and tuned using isothermal-isobaric molecular dynamics.
232 Pair distribution functions calculated from coarse-grained models
233 preserve the basic characteristics of the atomistic simulations. The
234 water density, measured in a head-group-fixed reference frame,
235 surrounding two phospholipid headgroups is shown in
236 Fig.~\ref{lipidFigure:PCFCoarse}. It is clear that the phosphate end
237 in DMPC and the amine end in DMPE are the two most heavily solvated
238 atoms.
239
240 \begin{figure}
241 \centering
242 \includegraphics[width=\linewidth]{g_coarse.eps}
243 \caption[The pair correlation functions for coarse-grained models]{}
244 \label{lipidFigure:PCFCoarse}
245 \end{figure}
246
247 \begin{figure}
248 \centering
249 \includegraphics[width=\linewidth]{EWD_coarse.eps}
250 \caption[Excess water density of coarse-grained phospholipids]{ }
251 \label{lipidFigure:EWDCoarse}
252 \end{figure}
253
254 \begin{table}
255 \caption{The Parameters For Coarse-grained Phospholipids}
256 \label{lipidTable:parameter}
257 \begin{center}
258 \begin{tabular}{|l|c|c|c|c|c|}
259 \hline
260 % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
261 Atom type & Mass & $\sigma$ & $\epsilon$ & charge & Dipole \\
262
263 $\text{{\sc CH}}_2$ & 14.03 & 3.95 & 0.0914 & 0 & 0 \\
264 $\text{{\sc CH}}_3$ & 15.04 & 3.75 & 0.195 & 0 & 0 \\
265 $\text{{\sc CE}}$ & 28.01 & 3.427& 0.294 & 0 & 1.693
266 $\text{{\sc CK}}$ & 28.01 & 3.592& 0.311 & 0 & 2.478 \\
267 $\text{{\sc PO}}_4$ & 108.995& 3.9 & 1.88 & -1& 0 \\
268 $\text{{\sc HDP}}$ & 14.03 & 4.0 & 0.13 & 0 & 0 \\
269 $\text{{\sc NC}}_4$ & 73.137 & 4.9 & 0.88 & +1& 0 \\
270 $\text{{\sc NH}}_3$ & 17.03 & 3.25 & 0.17 & +1& 0\\
271 \hline
272 \end{tabular}
273 \end{center}
274 \end{table}
275
276 \subsection{Bilayer Simulations Using Coarse-grained Model}
277
278 A bilayer system consisting of 128 DMPC lipids and 3655 water
279 molecules has been constructed from an atomistic coordinate
280 file.[15] The MD simulation is performed at constant temperature, T
281 = 300K, and constant pressure, p = 1 atm, and consisted of an
282 equilibration period of 2 ns. During the equilibration period, the
283 system was initially simulated at constant volume for 1ns. Once the
284 system was equilibrated at constant volume, the cell dimensions of
285 the system was relaxed by performing under NPT conditions using
286 Nos¨¦-Hoover extended system isothermal-isobaric dynamics. After
287 equilibration, different properties were evaluated over a production
288 run of 5 ns.
289
290 \begin{figure}
291 \centering
292 \includegraphics[width=\linewidth]{bilayer.eps}
293 \caption[Image of a coarse-grained bilayer system]{A coarse-grained
294 bilayer system consisting of 128 DMPC lipids and 3655 SSD water
295 molecules.}
296 \label{lipidFigure:bilayer}
297 \end{figure}
298
299 \subsubsection{Electron Density Profile (EDP)}
300
301 Assuming a gaussian distribution of electrons on each atomic center
302 with a variance estimated from the size of the van der Waals radius,
303 the EDPs which are proportional to the density profiles measured
304 along the bilayer normal obtained by x-ray scattering experiment,
305 can be expressed by
306 \begin{equation}
307 \rho _{x - ray} (z)dz \propto \sum\limits_{i = 1}^N {\frac{{n_i
308 }}{V}\frac{1}{{\sqrt {2\pi \sigma ^2 } }}e^{ - (z - z_i )^2 /2\sigma
309 ^2 } dz},
310 \end{equation}
311 where $\sigma$ is the variance equal to the van der Waals radius,
312 $n_i$ is the atomic number of site $i$ and $V$ is the volume of the
313 slab between $z$ and $z+dz$ . The highest density of total EDP
314 appears at the position of lipid-water interface corresponding to
315 headgroup, glycerol, and carbonyl groups of the lipids and the
316 distribution of water locked near the head groups, while the lowest
317 electron density is in the hydrocarbon region. As a good
318 approximation to the thickness of the bilayer, the headgroup spacing
319 , is defined as the distance between two peaks in the electron
320 density profile, calculated from our simulations to be 34.1 $\AA$.
321 This value is close to the x-ray diffraction experimental value 34.4
322 $\AA$.
323
324 \begin{figure}
325 \centering
326 \includegraphics[width=\linewidth]{electron_density.eps}
327 \caption[The density profile of the lipid bilayers]{Electron density
328 profile along the bilayer normal. The water density is shown in red,
329 the density due to the headgroups in green, the glycerol backbone in
330 brown, $\text{{\sc CH}}_2$ in yellow, $\text{{\sc CH}}_3$ in cyan,
331 and total density due to DMPC in blue.}
332 \label{lipidFigure:electronDensity}
333 \end{figure}
334
335 \subsubsection{$\text{S}_{\text{{\sc cd}}}$ Order Parameter}
336
337 Measuring deuterium order parameters by NMR is a useful technique to
338 study the orientation of hydrocarbon chains in phospholipids. The
339 order parameter tensor $S$ is defined by:
340 \begin{equation}
341 S_{ij} = \frac{1}{2} < 3\cos \theta _i \cos \theta _j - \delta
342 _{ij} >
343 \end{equation}
344 where $\theta$ is the angle between the $i$th molecular axis and
345 the bilayer normal ($z$ axis). The brackets denote an average over
346 time and molecules. The molecular axes are defined:
347 \begin{itemize}
348 \item $\mathbf{\hat{z}}$ is the vector from $C_{n-1}$ to $C_{n+1}$.
349 \item $\mathbf{\hat{y}}$ is the vector that is perpendicular to $z$ and
350 in the plane through $C_{n-1}$, $C_{n}$, and $C_{n+1}$.
351 \item $\mathbf{\hat{x}}$ is the vector perpendicular to
352 $\mathbf{\hat{y}}$ and $\mathbf{\hat{z}}$.
353 \end{itemize}
354 In coarse-grained model, although there are no explicit hydrogens,
355 the order parameter can still be written in terms of carbon ordering
356 at each point of the chain
357 \begin{equation}
358 S_{ij} = \frac{1}{2} < 3\cos \theta _i \cos \theta _j - \delta
359 _{ij} >.
360 \end{equation}
361
362 Fig.~\ref{lipidFigure:Scd} shows the order parameter profile
363 calculated for our coarse-grained DMPC bilayer system at 300K. Also
364 shown are the experimental data of Tiburu. The fact that
365 $\text{S}_{\text{{\sc cd}}}$ order parameters calculated from
366 simulation are higher than the experimental ones is ascribed to the
367 assumption of the locations of implicit hydrogen atoms which are
368 fixed in coarse-grained models at positions relative to the CC
369 vector.
370
371 \begin{figure}
372 \centering
373 \includegraphics[width=\linewidth]{scd.eps}
374 \caption[$\text{S}_{\text{{\sc cd}}}$ order parameter]{A comparison
375 of $|\text{S}_{\text{{\sc cd}}}|$ between coarse-grained model
376 (blue) and DMPC\cite{petrache00} (black) near 300~K.}
377 \label{lipidFigure:Scd}
378 \end{figure}