ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Lipid.tex
Revision: 2847
Committed: Fri Jun 9 19:25:07 2006 UTC (18 years ago) by tim
Content type: application/x-tex
File size: 20860 byte(s)
Log Message:
minor fix

File Contents

# User Rev Content
1 tim 2685 \chapter{\label{chapt:lipid}LIPID MODELING}
2    
3     \section{\label{lipidSection:introduction}Introduction}
4    
5 tim 2731 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 tim 2786 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 tim 2776 Infrared spectroscopy experiments suggest that halothane in DMPC
41 tim 2786 lipid bilayers lives near the membrane/water
42     interface\cite{Lieb1982}.
43 tim 2731
44 tim 2776 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 tim 2844 the range of current simulation technologies. Several schemes are
49     proposed in this chapter to overcome these difficulties.
50 tim 2731
51 tim 2844 \section{\label{lipidSection:model}Model and Methodology}
52 tim 2685
53 tim 2776 \subsection{\label{lipidSection:SSD}The Soft Sticky Dipole Water Model}
54 tim 2685
55 tim 2776 In a typical bilayer simulation, the dominant portion of the
56     computation time will be spent calculating water-water interactions.
57     As an efficient solvent model, the Soft Sticky Dipole (SSD) water
58 tim 2807 model\cite{Chandra1999,Fennell2004} is used as the explicit solvent
59 tim 2786 in this project. Unlike other water models which have partial
60     charges distributed throughout the whole molecule, the SSD water
61     model consists of a single site which is a Lennard-Jones interaction
62     site, as well as a point dipole. A tetrahedral potential is added to
63     correct for hydrogen bonding. The following equation describes the
64     interaction between two water molecules:
65 tim 2776 \[
66     V_{SSD} = V_{LJ} (r_{ij} ) + V_{dp} (r_{ij} ,\Omega _i ,\Omega _j )
67     + V_{sticky} (r_{ij} ,\Omega _i ,\Omega _j )
68     \]
69     where $r_{ij}$ is the vector between molecule $i$ and molecule $j$,
70     $\Omega _i$ and $\Omega _j$ are the orientational degrees of freedom
71     for molecule $i$ and molecule $j$ respectively.
72     \[
73     V_{LJ} (r_{ij} ) = 4\varepsilon _{ij} \left[ {\left( {\frac{{\sigma
74     _{ij} }}{{r_{ij} }}} \right)^{12} - \left( {\frac{{\sigma _{ij}
75     }}{{r_{ij} }}} \right)^6 } \right]
76     \]
77     \[
78     V_{dp} (r_{ij} ,\Omega _i ,\Omega _j ) = \frac{1}{{4\pi \varepsilon
79     _0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{r_{ij}^3 }} -
80     \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
81     r_{ij} } \right)}}{{r_{ij}^5 }}} \right]
82     \]
83     \[
84     V_{sticky} (r_{ij} ,\Omega _i ,\Omega _j ) = v_0 [s(r_{ij} )w(r_{ij}
85     ,\Omega _i ,\Omega _j ) + s'(r_{ij} )w'(r_{ij} ,\Omega _i ,\Omega _j
86     )]
87     \]
88     where $v_0$ is a strength parameter, $s$ and $s'$ are cubic
89     switching functions, while $w$ and $w'$ are responsible for the
90     tetrahedral potential and the short-range correction to the dipolar
91     interaction respectively.
92     \[
93     \begin{array}{l}
94     w(r_{ij} ,\Omega _i ,\Omega _j ) = \sin \theta _{ij} \sin 2\theta _{ij} \cos 2\phi _{ij} \\
95     w'(r_{ij} ,\Omega _i ,\Omega _j ) = (\cos \theta _{ij} - 0.6)^2 (\cos \theta _{ij} + 0.8)^2 - w_0 \\
96     \end{array}
97     \]
98     Although dipole-dipole and sticky interactions are more
99     mathematically complicated than Coulomb interactions, the number of
100     pair interactions is reduced dramatically both because the model
101     only contains a single-point as well as "short range" nature of the
102     higher order interaction.
103    
104     \subsection{\label{lipidSection:coarseGrained}The Coarse-Grained Phospholipid Model}
105    
106 tim 2781 Fig.~\ref{lipidFigure:coarseGrained} shows a schematic for our
107     coarse-grained phospholipid model. The lipid head group is modeled
108     by a linear rigid body which consists of three Lennard-Jones spheres
109     and a centrally located point-dipole. The backbone atoms in the
110     glycerol motif are modeled by Lennard-Jones spheres with dipoles.
111     Alkyl groups in hydrocarbon chains are replaced with unified
112     $\text{{\sc CH}}_2$ or $\text{{\sc CH}}_3$ atoms.
113 tim 2776
114 tim 2781 \begin{figure}
115     \centering
116 tim 2806 \includegraphics[width=3in]{coarse_grained.eps}
117 tim 2844 \caption[A representation of coarse-grained phospholipid model]{A
118     representation of coarse-grained phospholipid model. The lipid
119     headgroup consists of $\text{{\sc PO}}_4$ group (yellow),
120     $\text{{\sc NC}}_4$ group (blue) and a united $\text{{\sc C}} atom
121     (gray) $ with a dipole, while the glycerol backbone includes dipolar
122     $\text{{\sc CE}}$ (read) and $\text{{\sc CK}}$ (pink) groups. Alkyl
123     groups in hydrocarbon chains are simply represented by gray united
124     atoms.} \label{lipidFigure:coarseGrained}
125 tim 2781 \end{figure}
126    
127 tim 2776 Accurate and efficient computation of electrostatics is one of the
128     most difficult tasks in molecular modeling. Traditionally, the
129     electrostatic interaction between two molecular species is
130     calculated as a sum of interactions between pairs of point charges,
131     using Coulomb's law:
132     \[
133     V = \sum\limits_{i = 1}^{N_A } {\sum\limits_{j = 1}^{N_B }
134     {\frac{{q_i q_j }}{{4\pi \varepsilon _0 r_{ij} }}} }
135     \]
136     where $N_A$ and $N_B$ are the number of point charges in the two
137     molecular species. Originally developed to study ionic crystals, the
138     Ewald summation method mathematically transforms this
139     straightforward but conditionally convergent summation into two more
140     complicated but rapidly convergent sums. One summation is carried
141     out in reciprocal space while the other is carried out in real
142     space. An alternative approach is a multipole expansion, which is
143     based on electrostatic moments, such as charge (monopole), dipole,
144     quadruple etc.
145    
146     Here we consider a linear molecule which consists of two point
147     charges $\pm q$ located at $ ( \pm \frac{d}{2},0)$. The
148     electrostatic potential at point $P$ is given by:
149     \[
150     \frac{1}{{4\pi \varepsilon _0 }}\left( {\frac{{ - q}}{{r_ - }} +
151     \frac{{ + q}}{{r_ + }}} \right) = \frac{1}{{4\pi \varepsilon _0
152     }}\left( {\frac{{ - q}}{{\sqrt {r^2 + \frac{{d^2 }}{4} + rd\cos
153     \theta } }} + \frac{q}{{\sqrt {r^2 + \frac{{d^2 }}{4} - rd\cos
154     \theta } }}} \right)
155     \]
156    
157 tim 2781 \begin{figure}
158     \centering
159 tim 2806 \includegraphics[width=3in]{charge_dipole.eps}
160 tim 2844 \caption[An illustration of split-dipole
161     approximation]{Electrostatic potential due to a linear molecule
162     comprising two point charges with opposite charges. }
163     \label{lipidFigure:chargeDipole}
164 tim 2781 \end{figure}
165    
166 tim 2776 The basic assumption of the multipole expansion is $r \gg d$ , thus,
167     $\frac{{d^2 }}{4}$ inside the square root of the denominator is
168     neglected. This is a reasonable approximation in most cases.
169     Unfortunately, in our headgroup model, the distance of charge
170     separation $d$ (4.63 $\AA$ in PC headgroup) may be comparable to
171     $r$. Nevertheless, we could still assume $ \cos \theta \approx 0$
172     in the central region of the headgroup. Using Taylor expansion and
173     associating appropriate terms with electric moments will result in a
174     "split-dipole" approximation:
175     \[
176     V(r) = \frac{1}{{4\pi \varepsilon _0 }}\frac{{r\mu \cos \theta
177     }}{{R^3 }}
178     \]
179     where$R = \sqrt {r^2 + \frac{{d^2 }}{4}}$ Placing a dipole at point
180     $P$ and applying the same strategy, the interaction between two
181     split-dipoles is then given by:
182     \[
183     V_{sd} (r_{ij} ,\Omega _i ,\Omega _j ) = \frac{1}{{4\pi \varepsilon
184     _0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{R_{ij}^3 }} -
185     \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
186     r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
187     \]
188     where $\mu _i$ and $\mu _j$ are the dipole moment of molecule $i$
189     and molecule $j$ respectively, $r_{ij}$ is vector between molecule
190 tim 2800 $i$ and molecule $j$, and $R_{ij}$ is given by,
191 tim 2776 \[
192     R_{ij} = \sqrt {r_{ij}^2 + \frac{{d_i^2 }}{4} + \frac{{d_j^2
193     }}{4}}
194     \]
195     where $d_i$ and $d_j$ are the charge separation distance of dipole
196     and respectively. This approximation to the multipole expansion
197     maintains the fast fall-off of the multipole potentials but lacks
198     the normal divergences when two polar groups get close to one
199 tim 2844 another. The comparision between different electrostatic
200     approximation is shown in \ref{lipidFigure:splitDipole}. Due to the
201     divergence at the central region of the headgroup introduced by
202     dipole-dipole approximation, we discover that water molecules are
203     locked into the central region in the simulation. This artifact can
204     be corrected using split-dipole approximation or other accurate
205     methods.
206 tim 2776 \begin{figure}
207     \centering
208 tim 2847 \includegraphics[width=3in]{split_dipole.eps}
209 tim 2844 \caption[Comparison between electrostatic
210     approximation]{Electrostatic potential map for two pairs of charges
211     with different alignments: (a) illustration of different alignments;
212     (b) charge-charge interaction; (c) dipole-dipole approximation; (d)
213     split-dipole approximation.} \label{lipidFigure:splitDipole}
214 tim 2776 \end{figure}
215    
216     %\section{\label{lipidSection:methods}Methods}
217    
218 tim 2685 \section{\label{lipidSection:resultDiscussion}Results and Discussion}
219 tim 2776
220     \subsection{One Lipid in Sea of Water Molecules}
221    
222     To exclude the inter-headgroup interaction, atomistic models of one
223     lipid (DMPC or DLPE) in sea of water molecules (TIP3P) were built
224     and studied using atomistic molecular dynamics. The simulation was
225     analyzed using a set of radial distribution functions, which give
226     the probability of finding a pair of molecular species a distance
227     apart, relative to the probability expected for a completely random
228     distribution function at the same density.
229    
230     \begin{equation}
231     g_{AB} (r) = \frac{1}{{\rho _B }}\frac{1}{{N_A }} < \sum\limits_{i
232     \in A} {\sum\limits_{j \in B} {\delta (r - r_{ij} )} } >
233     \end{equation}
234     \begin{equation}
235     g_{AB} (r,\cos \theta ) = \frac{1}{{\rho _B }}\frac{1}{{N_A }} <
236     \sum\limits_{i \in A} {\sum\limits_{j \in B} {\delta (r - r_{ij} )}
237     } \delta (\cos \theta _{ij} - \cos \theta ) >
238     \end{equation}
239    
240 tim 2844 From Fig.~\ref{lipidFigure:PCFAtom}, we can identify the first
241     solvation shell (3.5 $\AA$) and the second solvation shell (5.0
242     $\AA$) from both plots. However, the corresponding orientations are
243     different. In DLPE, water molecules prefer to sit around $\text{{\sc
244     NH}}_3$ group due to the hydrogen bonding. In contrast, because of
245     the hydrophobic effect of the $ {\rm{N(CH}}_{\rm{3}}
246     {\rm{)}}_{\rm{3}} $ group, the preferred position of water molecules
247     in DMPC is around the $\text{{\sc PO}}_4$ Group. When the water
248     molecules are far from the headgroup, the distribution of the two
249     angles should be uniform. The correlation close to center of the
250     headgroup dipole in both plots tell us that in the closely-bound
251     region, the dipoles of the water molecules are preferentially
252     anti-aligned with the dipole of headgroup. When the water molecules
253     are far from the headgroup, the distribution of the two angles
254     should be uniform. The correlation close to center of the headgroup
255     dipole in both plots tell us that in the closely-bound region, the
256     dipoles of the water molecules are preferentially aligned with the
257     dipole of headgroup.
258 tim 2776
259     \begin{figure}
260     \centering
261     \includegraphics[width=\linewidth]{g_atom.eps}
262 tim 2844 \caption[The pair correlation functions for atomistic models]{The
263     pair correlation functions for atomistic models: (a)$g(r,\cos \theta
264     )$ for DMPC; (b) $g(r,\cos \theta )$ for DLPE; (c)$g(r,\cos \omega
265     )$ for DMPC; (d)$g(r,\cos \omega )$ for DLPE; (e)$g(\cos \theta,\cos
266     \omega)$ for DMPC; (f)$g(\cos \theta,\cos \omega)$ for DMLPE.}
267 tim 2776 \label{lipidFigure:PCFAtom}
268     \end{figure}
269    
270     The initial configurations of coarse-grained systems are constructed
271     from the previous atomistic ones. The parameters for the
272     coarse-grained model in Table~\ref{lipidTable:parameter} are
273     estimated and tuned using isothermal-isobaric molecular dynamics.
274     Pair distribution functions calculated from coarse-grained models
275     preserve the basic characteristics of the atomistic simulations. The
276     water density, measured in a head-group-fixed reference frame,
277     surrounding two phospholipid headgroups is shown in
278     Fig.~\ref{lipidFigure:PCFCoarse}. It is clear that the phosphate end
279     in DMPC and the amine end in DMPE are the two most heavily solvated
280     atoms.
281    
282     \begin{figure}
283     \centering
284     \includegraphics[width=\linewidth]{g_coarse.eps}
285 tim 2844 \caption[The pair correlation functions for coarse-grained
286     models]{The pair correlation functions for coarse-grained models:
287     (a)$g(r,\cos \theta )$ for DMPC; (b) $g(r,\cos \theta )$ for DLPE.}
288 tim 2776 \label{lipidFigure:PCFCoarse}
289     \end{figure}
290    
291     \begin{figure}
292     \centering
293     \includegraphics[width=\linewidth]{EWD_coarse.eps}
294 tim 2847 \caption[Excess water density of coarse-grained
295     phospholipids]{Excess water density of coarse-grained
296     phospholipids.} \label{lipidFigure:EWDCoarse}
297 tim 2776 \end{figure}
298    
299     \begin{table}
300     \caption{The Parameters For Coarse-grained Phospholipids}
301     \label{lipidTable:parameter}
302     \begin{center}
303     \begin{tabular}{|l|c|c|c|c|c|}
304     \hline
305     % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
306     Atom type & Mass & $\sigma$ & $\epsilon$ & charge & Dipole \\
307     $\text{{\sc CH}}_2$ & 14.03 & 3.95 & 0.0914 & 0 & 0 \\
308     $\text{{\sc CH}}_3$ & 15.04 & 3.75 & 0.195 & 0 & 0 \\
309 tim 2781 $\text{{\sc CE}}$ & 28.01 & 3.427& 0.294 & 0 & 1.693 \\
310 tim 2776 $\text{{\sc CK}}$ & 28.01 & 3.592& 0.311 & 0 & 2.478 \\
311     $\text{{\sc PO}}_4$ & 108.995& 3.9 & 1.88 & -1& 0 \\
312     $\text{{\sc HDP}}$ & 14.03 & 4.0 & 0.13 & 0 & 0 \\
313     $\text{{\sc NC}}_4$ & 73.137 & 4.9 & 0.88 & +1& 0 \\
314     $\text{{\sc NH}}_3$ & 17.03 & 3.25 & 0.17 & +1& 0\\
315     \hline
316     \end{tabular}
317     \end{center}
318     \end{table}
319    
320     \subsection{Bilayer Simulations Using Coarse-grained Model}
321    
322     A bilayer system consisting of 128 DMPC lipids and 3655 water
323 tim 2844 molecules has been constructed from an atomistic coordinate file.
324     The MD simulation is performed at constant temperature, T = 300K,
325     and constant pressure, p = 1 atm, and consisted of an equilibration
326     period of 2 ns. During the equilibration period, the system was
327     initially simulated at constant volume for 1 ns. Once the system was
328     equilibrated at constant volume, the cell dimensions of the system
329     was relaxed by performing under NPT conditions using Nos¨¦-Hoover
330     extended system isothermal-isobaric dynamics. After equilibration,
331     different properties were evaluated over a production run of 5 ns.
332 tim 2776
333     \begin{figure}
334     \centering
335     \includegraphics[width=\linewidth]{bilayer.eps}
336     \caption[Image of a coarse-grained bilayer system]{A coarse-grained
337     bilayer system consisting of 128 DMPC lipids and 3655 SSD water
338     molecules.}
339     \label{lipidFigure:bilayer}
340     \end{figure}
341    
342 tim 2819 \subsubsection{\textbf{Electron Density Profile (EDP)}}
343 tim 2776
344     Assuming a gaussian distribution of electrons on each atomic center
345     with a variance estimated from the size of the van der Waals radius,
346     the EDPs which are proportional to the density profiles measured
347     along the bilayer normal obtained by x-ray scattering experiment,
348 tim 2786 can be expressed by\cite{Tu1995}
349 tim 2776 \begin{equation}
350     \rho _{x - ray} (z)dz \propto \sum\limits_{i = 1}^N {\frac{{n_i
351     }}{V}\frac{1}{{\sqrt {2\pi \sigma ^2 } }}e^{ - (z - z_i )^2 /2\sigma
352     ^2 } dz},
353     \end{equation}
354     where $\sigma$ is the variance equal to the van der Waals radius,
355     $n_i$ is the atomic number of site $i$ and $V$ is the volume of the
356     slab between $z$ and $z+dz$ . The highest density of total EDP
357     appears at the position of lipid-water interface corresponding to
358     headgroup, glycerol, and carbonyl groups of the lipids and the
359     distribution of water locked near the head groups, while the lowest
360     electron density is in the hydrocarbon region. As a good
361     approximation to the thickness of the bilayer, the headgroup spacing
362     , is defined as the distance between two peaks in the electron
363     density profile, calculated from our simulations to be 34.1 $\AA$.
364     This value is close to the x-ray diffraction experimental value 34.4
365 tim 2786 $\AA$\cite{Petrache1998}.
366 tim 2776
367     \begin{figure}
368     \centering
369     \includegraphics[width=\linewidth]{electron_density.eps}
370     \caption[The density profile of the lipid bilayers]{Electron density
371     profile along the bilayer normal. The water density is shown in red,
372     the density due to the headgroups in green, the glycerol backbone in
373     brown, $\text{{\sc CH}}_2$ in yellow, $\text{{\sc CH}}_3$ in cyan,
374     and total density due to DMPC in blue.}
375     \label{lipidFigure:electronDensity}
376     \end{figure}
377    
378 tim 2819 \subsubsection{\textbf{$\text{S}_{\text{{\sc cd}}}$ Order Parameter}}
379 tim 2776
380     Measuring deuterium order parameters by NMR is a useful technique to
381     study the orientation of hydrocarbon chains in phospholipids. The
382     order parameter tensor $S$ is defined by:
383     \begin{equation}
384     S_{ij} = \frac{1}{2} < 3\cos \theta _i \cos \theta _j - \delta
385     _{ij} >
386     \end{equation}
387     where $\theta$ is the angle between the $i$th molecular axis and
388     the bilayer normal ($z$ axis). The brackets denote an average over
389     time and molecules. The molecular axes are defined:
390     \begin{itemize}
391     \item $\mathbf{\hat{z}}$ is the vector from $C_{n-1}$ to $C_{n+1}$.
392     \item $\mathbf{\hat{y}}$ is the vector that is perpendicular to $z$ and
393     in the plane through $C_{n-1}$, $C_{n}$, and $C_{n+1}$.
394     \item $\mathbf{\hat{x}}$ is the vector perpendicular to
395     $\mathbf{\hat{y}}$ and $\mathbf{\hat{z}}$.
396     \end{itemize}
397     In coarse-grained model, although there are no explicit hydrogens,
398     the order parameter can still be written in terms of carbon ordering
399 tim 2786 at each point of the chain\cite{Egberts1988}
400 tim 2776 \begin{equation}
401     S_{ij} = \frac{1}{2} < 3\cos \theta _i \cos \theta _j - \delta
402     _{ij} >.
403     \end{equation}
404    
405     Fig.~\ref{lipidFigure:Scd} shows the order parameter profile
406     calculated for our coarse-grained DMPC bilayer system at 300K. Also
407 tim 2786 shown are the experimental data of Tu\cite{Tu1995}. The fact that
408 tim 2776 $\text{S}_{\text{{\sc cd}}}$ order parameters calculated from
409     simulation are higher than the experimental ones is ascribed to the
410     assumption of the locations of implicit hydrogen atoms which are
411     fixed in coarse-grained models at positions relative to the CC
412     vector.
413    
414     \begin{figure}
415     \centering
416     \includegraphics[width=\linewidth]{scd.eps}
417     \caption[$\text{S}_{\text{{\sc cd}}}$ order parameter]{A comparison
418     of $|\text{S}_{\text{{\sc cd}}}|$ between coarse-grained model
419 tim 2845 (blue) and DMPC\cite{Petrache2000} (black) near 300~K.}
420 tim 2776 \label{lipidFigure:Scd}
421     \end{figure}
422 tim 2781
423     %\subsection{Bilayer Simulations Using Gay-Berne Ellipsoid Model}
424 tim 2844
425     \section{\label{lipidSection:Conclusion}Conclusion}
426    
427     Atomistic simulations are used in this study to determine the
428     preferred orientation and location of water molecules relative to
429     the location and orientation of the PC and PE lipid headgroups.
430     Based on the result from all-atom simulations, we developed a simple
431     coarse-grained model capturing essential features of the headgroup
432     solvation which is crucial to transport process in membrane system.
433     In addition, the model has been explored in a bilayer system which
434     is shown to have reasonable electron density profile,
435     $\text{S}_{\text{{\sc cd}}}$ order parameter and other structural
436     properties. The accuracy of this model is achieved by matching
437     atomistic result. It is also easy to represent different
438     phosphorlipids by changing a few parameters of the model. Another
439     important characteristic of this model distinguishing itself from
440     other models\cite{Goetz1998,Marrink2004} is the computational speed
441     gaining by introducing short range electrostatic approximation.
442     Further studies of this system using z-constraint method could
443     extend the time length of the simulations to study transport
444     phenomena in large-scale membrane systems.