ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
Revision: 3198
Committed: Wed Jul 25 22:42:56 2007 UTC (17 years, 1 month ago) by xsun
Content type: application/x-tex
File size: 28199 byte(s)
Log Message:
add the parameter table

File Contents

# User Rev Content
1 xsun 3147 %\documentclass[aps,pre,twocolumn,amssymb,showpacs,floatfix]{revtex4}
2     \documentclass[aps,pre,preprint,amssymb,showpacs]{revtex4}
3     \usepackage{graphicx}
4    
5     \begin{document}
6     \renewcommand{\thefootnote}{\fnsymbol{footnote}}
7     \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
8    
9     %\bibliographystyle{aps}
10    
11 xsun 3174 \title{Dipolar Ordering of the Ripple Phase in Lipid Membranes}
12 xsun 3147 \author{Xiuquan Sun and J. Daniel Gezelter}
13     \email[E-mail:]{gezelter@nd.edu}
14     \affiliation{Department of Chemistry and Biochemistry,\\
15     University of Notre Dame, \\
16     Notre Dame, Indiana 46556}
17    
18     \date{\today}
19    
20     \begin{abstract}
21 gezelter 3195 The ripple phase in phosphatidylcholine (PC) bilayers has never been
22     completely explained.
23 xsun 3147 \end{abstract}
24    
25     \pacs{}
26     \maketitle
27    
28 xsun 3174 \section{Introduction}
29     \label{sec:Int}
30 gezelter 3195 Fully hydrated lipids will aggregate spontaneously to form bilayers
31     which exhibit a variety of phases depending on their temperatures and
32     compositions. Among these phases, a periodic rippled phase
33     ($P_{\beta'}$) appears as an intermediate phase between the gel
34     ($L_\beta$) and fluid ($L_{\alpha}$) phases for relatively pure
35     phosphatidylcholine (PC) bilayers. The ripple phase has attracted
36     substantial experimental interest over the past 30 years. Most
37     structural information of the ripple phase has been obtained by the
38     X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
39     microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
40     et al.} used atomic force microscopy (AFM) to observe ripple phase
41     morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
42     experimental results provide strong support for a 2-dimensional
43     hexagonal packing lattice of the lipid molecules within the ripple
44     phase. This is a notable change from the observed lipid packing
45     within the gel phase.~\cite{Cevc87}
46 xsun 3174
47 gezelter 3195 A number of theoretical models have been presented to explain the
48     formation of the ripple phase. Marder {\it et al.} used a
49     curvature-dependent Landau-de Gennes free-energy functional to predict
50     a rippled phase.~\cite{Marder84} This model and other related continuum
51     models predict higher fluidity in convex regions and that concave
52     portions of the membrane correspond to more solid-like regions.
53     Carlson and Sethna used a packing-competition model (in which head
54     groups and chains have competing packing energetics) to predict the
55     formation of a ripple-like phase. Their model predicted that the
56     high-curvature portions have lower-chain packing and correspond to
57     more fluid-like regions. Goldstein and Leibler used a mean-field
58     approach with a planar model for {\em inter-lamellar} interactions to
59     predict rippling in multilamellar phases.~\cite{Goldstein88} McCullough
60     and Scott proposed that the {\em anisotropy of the nearest-neighbor
61     interactions} coupled to hydrophobic constraining forces which
62     restrict height differences between nearest neighbors is the origin of
63     the ripple phase.~\cite{McCullough90} Lubensky and MacKintosh
64     introduced a Landau theory for tilt order and curvature of a single
65     membrane and concluded that {\em coupling of molecular tilt to membrane
66     curvature} is responsible for the production of
67     ripples.~\cite{Lubensky93} Misbah, Duplat and Houchmandzadeh proposed
68     that {\em inter-layer dipolar interactions} can lead to ripple
69     instabilities.~\cite{Misbah98} Heimburg presented a {\em coexistence
70     model} for ripple formation in which he postulates that fluid-phase
71     line defects cause sharp curvature between relatively flat gel-phase
72     regions.~\cite{Heimburg00} Kubica has suggested that a lattice model of
73     polar head groups could be valuable in trying to understand bilayer
74     phase formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations
75     of lamellar stacks of hexagonal lattices to show that large headgroups
76     and molecular tilt with respect to the membrane normal vector can
77     cause bulk rippling.~\cite{Bannerjee02}
78 xsun 3174
79 gezelter 3195 In contrast, few large-scale molecular modelling studies have been
80     done due to the large size of the resulting structures and the time
81     required for the phases of interest to develop. With all-atom (and
82     even unified-atom) simulations, only one period of the ripple can be
83     observed and only for timescales in the range of 10-100 ns. One of
84     the most interesting molecular simulations was carried out by De Vries
85     {\it et al.}~\cite{deVries05}. According to their simulation results,
86     the ripple consists of two domains, one resembling the gel bilayer,
87     while in the other, the two leaves of the bilayer are fully
88     interdigitated. The mechanism for the formation of the ripple phase
89     suggested by their work is a packing competition between the head
90     groups and the tails of the lipid molecules.\cite{Carlson87} Recently,
91     the ripple phase has also been studied by the XXX group using Monte
92     Carlo simulations.\cite{Lenz07} Their structures are similar to the De
93     Vries {\it et al.} structures except that the connection between the
94     two leaves of the bilayer is a narrow interdigitated line instead of
95     the fully interdigitated domain. The symmetric ripple phase was also
96     observed by Lenz {\it et al.}, and their work supports other claims
97     that the mismatch between the size of the head group and tail of the
98     lipid molecules is the driving force for the formation of the ripple
99     phase. Ayton and Voth have found significant undulations in
100     zero-surface-tension states of membranes simulated via dissipative
101     particle dynamics, but their results are consistent with purely
102     thermal undulations.~\cite{Ayton02}
103 xsun 3174
104 gezelter 3195 Although the organization of the tails of lipid molecules are
105     addressed by these molecular simulations and the packing competition
106     between headgroups and tails is strongly implicated as the primary
107     driving force for ripple formation, questions about the ordering of
108     the head groups in ripple phase has not been settled.
109 xsun 3174
110 gezelter 3195 In a recent paper, we presented a simple ``web of dipoles'' spin
111     lattice model which provides some physical insight into relationship
112     between dipolar ordering and membrane buckling.\cite{Sun2007} We found
113     that dipolar elastic membranes can spontaneously buckle, forming
114     ripple-like topologies. The driving force for the buckling in dipolar
115     elastic membranes the antiferroelectric ordering of the dipoles, and
116     this was evident in the ordering of the dipole director axis
117     perpendicular to the wave vector of the surface ripples. A similiar
118     phenomenon has also been observed by Tsonchev {\it et al.} in their
119     work on the spontaneous formation of dipolar molecules into curved
120     nano-structures.\cite{Tsonchev04}
121    
122     In this paper, we construct a somewhat more realistic molecular-scale
123     lipid model than our previous ``web of dipoles'' and use molecular
124     dynamics simulations to elucidate the role of the head group dipoles
125     in the formation and morphology of the ripple phase. We describe our
126     model and computational methodology in section \ref{sec:method}.
127     Details on the simulations are presented in section
128     \ref{sec:experiment}, with results following in section
129     \ref{sec:results}. A final discussion of the role of dipolar heads in
130     the ripple formation can be found in section
131 xsun 3174 \ref{sec:discussion}.
132    
133 gezelter 3196 \section{Computational Model}
134 xsun 3174 \label{sec:method}
135    
136 gezelter 3195 Our simple molecular-scale lipid model for studying the ripple phase
137     is based on two facts: one is that the most essential feature of lipid
138     molecules is their amphiphilic structure with polar head groups and
139     non-polar tails. Another fact is that the majority of lipid molecules
140     in the ripple phase are relatively rigid (i.e. gel-like) which makes
141     some fraction of the details of the chain dynamics negligible. Figure
142     \ref{fig:lipidModels} shows the molecular strucure of a DPPC
143     molecule, as well as atomistic and molecular-scale representations of
144     a DPPC molecule. The hydrophilic character of the head group is
145     largely due to the separation of charge between the nitrogen and
146     phosphate groups. The zwitterionic nature of the PC headgroups leads
147     to abnormally large dipole moments (as high as 20.6 D), and this
148     strongly polar head group interacts strongly with the solvating water
149     layers immediately surrounding the membrane. The hydrophobic tail
150     consists of fatty acid chains. In our molecular scale model, lipid
151     molecules have been reduced to these essential features; the fatty
152     acid chains are represented by an ellipsoid with a dipolar ball
153     perched on one end to represent the effects of the charge-separated
154     head group. In real PC lipids, the direction of the dipole is
155     nearly perpendicular to the tail, so we have fixed the direction of
156     the point dipole rigidly in this orientation.
157 xsun 3147
158 xsun 3174 \begin{figure}[htb]
159     \centering
160 gezelter 3186 \includegraphics[width=\linewidth]{lipidModels}
161     \caption{Three different representations of DPPC lipid molecules,
162     including the chemical structure, an atomistic model, and the
163     head-body ellipsoidal coarse-grained model used in this
164     work.\label{fig:lipidModels}}
165 xsun 3174 \end{figure}
166 xsun 3147
167 gezelter 3195 The ellipsoidal portions of the model interact via the Gay-Berne
168     potential which has seen widespread use in the liquid crystal
169     community. In its original form, the Gay-Berne potential was a single
170     site model for the interactions of rigid ellipsoidal
171     molecules.\cite{Gay81} It can be thought of as a modification of the
172     Gaussian overlap model originally described by Berne and
173     Pechukas.\cite{Berne72} The potential is constructed in the familiar
174     form of the Lennard-Jones function using orientation-dependent
175     $\sigma$ and $\epsilon$ parameters,
176 xsun 3174 \begin{eqnarray*}
177     V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
178     r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
179     {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
180     {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
181     -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
182     {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
183 gezelter 3195 \label{eq:gb}
184     \end{eqnarray*}
185    
186    
187    
188     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
189     \hat{r}}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
190     \hat{u}}_{j},{\bf \hat{r}}))$ parameters
191     are dependent on the relative orientations of the two molecules (${\bf
192     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
193     intermolecular separation (${\bf \hat{r}}$). The functional forms for
194     $\sigma({\bf
195     \hat{u}}_{i},{\bf
196     \hat{u}}_{j},{\bf \hat{r}})$ and $\epsilon({\bf \hat{u}}_{i},{\bf
197     \hat{u}}_{j},{\bf \hat{r}}))$ are given elsewhere
198     and will not be repeated here. However, $\epsilon$ and $\sigma$ are
199     governed by two anisotropy parameters,
200     \begin {equation}
201     \begin{array}{rcl}
202     \chi & = & \frac
203     {(\sigma_{e}/\sigma_{s})^2-1}{(\sigma_{e}/\sigma_{s})^2+1} \\ \\
204     \chi\prime & = & \frac {1-(\epsilon_{e}/\epsilon_{s})^{1/\mu}}{1+(\epsilon_{e}/
205     \epsilon_{s})^{1/\mu}}
206     \end{array}
207     \end{equation}
208     In these equations, $\sigma$ and $\epsilon$ refer to the point of
209     closest contact and the depth of the well in different orientations of
210     the two molecules. The subscript $s$ refers to the {\it side-by-side}
211     configuration where $\sigma$ has it's smallest value,
212     $\sigma_{s}$, and where the potential well is $\epsilon_{s}$ deep.
213     The subscript $e$ refers to the {\it end-to-end} configuration where
214     $\sigma$ is at it's largest value, $\sigma_{e}$ and where the well
215     depth, $\epsilon_{e}$ is somewhat smaller than in the side-by-side
216     configuration. For the prolate ellipsoids we are using, we have
217     \begin{equation}
218     \begin{array}{rcl}
219     \sigma_{s} & < & \sigma_{e} \\
220     \epsilon_{s} & > & \epsilon_{e}
221     \end{array}
222     \end{equation}
223     Ref. \onlinecite{Luckhurst90} has a particularly good explanation of the
224     choice of the Gay-Berne parameters $\mu$ and $\nu$ for modeling liquid
225     crystal molecules.
226    
227     The breadth and length of tail are $\sigma_0$, $3\sigma_0$,
228     corresponding to a shape anisotropy of 3 for the chain portion of the
229     molecule. In principle, this could be varied to allow for modeling of
230     longer or shorter chain lipid molecules.
231    
232     To take into account the permanent dipolar interactions of the
233     zwitterionic head groups, we place fixed dipole moments $\mu_{i}$ at
234     one end of the Gay-Berne particles. The dipoles will be oriented at
235     an angle $\theta = \pi / 2$ relative to the major axis. These dipoles
236     are protected by a head ``bead'' with a range parameter which we have
237     varied between $1.20\sigma_0$ and $1.41\sigma_0$. The head groups
238     interact with each other using a combination of Lennard-Jones,
239 xsun 3174 \begin{eqnarray*}
240 gezelter 3195 V_{ij} = 4\epsilon \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
241     \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
242 xsun 3174 \end{eqnarray*}
243 gezelter 3195 and dipole,
244 xsun 3174 \begin{eqnarray*}
245 gezelter 3195 V_{ij} = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
246     \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
247     \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
248 xsun 3174 \end{eqnarray*}
249 gezelter 3195 potentials.
250     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
251     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
252     pointing along the inter-dipole vector $\mathbf{r}_{ij}$.
253    
254     For the interaction between nonequivalent uniaxial ellipsoids (in this
255     case, between spheres and ellipsoids), the range parameter is
256     generalized as\cite{Cleaver96}
257 xsun 3174 \begin{eqnarray*}
258     \sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}}) =
259     {\sigma_{0}} \left(1-\frac{1}{2}\chi\left[\frac{(\alpha{\mathbf{\hat r}_{ij}}
260     \cdot {\mathbf{\hat u}_i} + \alpha^{-1}{\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat
261     u}_j})^2}{1+\chi({\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j})} +
262     \frac{(\alpha{\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat u}_i} - \alpha^{-1}{\mathbf{\hat r}_{ij}}
263     \cdot {\mathbf{\hat u}_j})^2}{1-\chi({\mathbf{\hat u}_i} \cdot
264     {\mathbf{\hat u}_j})} \right] \right)^{-\frac{1}{2}}
265     \end{eqnarray*}
266     where $\alpha$ is given by
267     \begin{eqnarray*}
268     \alpha^2 =
269     \left[\frac{\left({\sigma_e}_i^2-{\sigma_s}_i^2\right)\left({\sigma_e}_j^2+{\sigma_s}_i^2\right)}{\left({\sigma_e}_j^2-{\sigma_s}_j^2\right)\left({\sigma_e}_i^2+{\sigma_s}_j^2\right)}
270     \right]^{\frac{1}{2}}
271     \end{eqnarray*}
272 gezelter 3195 the strength parameter has been adjusted as suggested by Cleaver {\it
273     et al.}\cite{Cleaver96} A switching function has been applied to all
274 gezelter 3196 potentials to smoothly turn off the interactions between a range of $22$ and $25$ \AA.
275 xsun 3147
276 gezelter 3196 The solvent model in our simulations is identical to one used by XXX
277     in their dissipative particle dynamics (DPD) simulation of lipid
278     bilayers.]cite{XXX} This solvent bead is a single site that represents
279     four water molecules (m = 72 amu) and has comparable density and
280     diffusive behavior to liquid water. However, since there are no
281     electrostatic sites on these beads, this solvent model cannot
282     replicate the dielectric properties of water.
283 xsun 3198 \begin{table*}
284     \begin{minipage}{\linewidth}
285     \begin{center}
286     \caption{}
287     \begin{tabular}{lccc}
288     \hline
289     N/A & Head & Chain & Solvent \\
290     \hline
291     $\sigma_0$ (\AA) & varied & 4.6 & 4.7 \\
292     l (aspect ratio) & N/A & 3 & N/A \\
293     $\epsilon_0$ (kcal/mol) & 0.185 & 1.0 & 0.8 \\
294     $\epsilon_e$ (aspect ratio) & N/A & 0.2 & N/A \\
295     M (amu) & 196 & 760 & 72.06112 \\
296     $I_{xx}$, $I_{yy}$, $I_{zz}$ (amu \AA$^2$) & 1125, 1125, 0 & 45000, 45000, 9000 & N/A \\
297     $\mu$ (Debye) & varied & N/A & N/A \\
298     \end{tabular}
299     \label{tab:parameters}
300     \end{center}
301     \end{minipage}
302     \end{table*}
303 gezelter 3195
304 gezelter 3186 \begin{figure}[htb]
305     \centering
306 gezelter 3196 \includegraphics[width=\linewidth]{2lipidModel}
307 gezelter 3186 \caption{The parameters defining the behavior of the lipid
308     models. $\sigma_h / \sigma_0$ is the ratio of the head group to body
309 gezelter 3196 diameter. Molecular bodies had a fixed aspect ratio of 3.0. The
310     solvent model was a simplified 4-water bead ($\sigma_w = 1.02
311     \sigma_0$) that has been used in other coarse-grained (DPD) simulations.
312     The dipolar strength (and the temperature and pressure) were the only
313     other parameters that were varied
314     systematically.\label{fig:lipidModel}}
315 gezelter 3186 \end{figure}
316    
317 gezelter 3196 \section{Experimental Methodology}
318 xsun 3174 \label{sec:experiment}
319 xsun 3147
320 gezelter 3196 To create unbiased bilayers, all simulations were started from two
321     perfectly flat monolayers separated by a 20 \AA\ gap between the
322     molecular bodies of the upper and lower leaves. The separated
323     monolayers were evolved in a vaccum with $x-y$ anisotropic pressure
324 xsun 3174 coupling. The length of $z$ axis of the simulations was fixed and a
325     constant surface tension was applied to enable real fluctuations of
326 gezelter 3196 the bilayer. Periodic boundaries were used, and $480-720$ lipid
327     molecules were present in the simulations depending on the size of the
328     head beads. The two monolayers spontaneously collapse into bilayer
329     structures within 100 ps, and following this collapse, all systems
330     were equlibrated for $100$ ns at $300$ K.
331 xsun 3147
332 gezelter 3196 The resulting structures were then solvated at a ratio of $6$ DPD
333     solvent beads (24 water molecules) per lipid. These configurations
334     were then equilibrated for another $30$ ns. All simulations with
335     solvent were carried out at constant pressure ($P=1$ atm) by $3$D
336     anisotropic coupling, and constant surface tension ($\gamma=0.015$
337     UNIT). Given the absence of fast degrees of freedom in this model, a
338     timestep of $50$ fs was utilized. Data collection for structural
339     properties of the bilayers was carried out during a final 5 ns run
340     following the solvent equilibration. All simulations were performed
341     using the OOPSE molecular modeling program.\cite{Meineke05}
342    
343     \section{Results}
344 xsun 3174 \label{sec:results}
345 xsun 3147
346 xsun 3174 Snapshots in Figure \ref{fig:phaseCartoon} show that the membrane is
347     more corrugated increasing size of the head groups. The surface is
348     nearly flat when $\sigma_h=1.20\sigma_0$. With
349     $\sigma_h=1.28\sigma_0$, although the surface is still flat, the
350     bilayer starts to splay inward; the upper leaf of the bilayer is
351     connected to the lower leaf with an interdigitated line defect. Two
352     periodicities with $100$ \AA\ width were observed in the
353     simulation. This structure is very similiar to the structure observed
354     by de Vries and Lenz {\it et al.}. The same basic structure is also
355     observed when $\sigma_h=1.41\sigma_0$, but the wavelength of the
356     surface corrugations depends sensitively on the size of the ``head''
357     beads. From the undulation spectrum, the corrugation is clearly
358     non-thermal.
359     \begin{figure}[htb]
360     \centering
361     \includegraphics[width=\linewidth]{phaseCartoon}
362     \caption{A sketch to discribe the structure of the phases observed in
363     our simulations.\label{fig:phaseCartoon}}
364     \end{figure}
365 xsun 3147
366 xsun 3174 When $\sigma_h=1.35\sigma_0$, we observed another corrugated surface
367     morphology. This structure is different from the asymmetric rippled
368     surface; there is no interdigitation between the upper and lower
369     leaves of the bilayer. Each leaf of the bilayer is broken into several
370     hemicylinderical sections, and opposite leaves are fitted together
371     much like roof tiles. Unlike the surface in which the upper
372     hemicylinder is always interdigitated on the leading or trailing edge
373     of lower hemicylinder, the symmetric ripple has no prefered direction.
374     The corresponding cartoons are shown in Figure
375     \ref{fig:phaseCartoon} for elucidation of the detailed structures of
376     different phases. Figure \ref{fig:phaseCartoon} (a) is the flat phase,
377     (b) is the asymmetric ripple phase corresponding to the lipid
378     organization when $\sigma_h=1.28\sigma_0$ and $\sigma_h=1.41\sigma_0$,
379     and (c) is the symmetric ripple phase observed when
380     $\sigma_h=1.35\sigma_0$. In symmetric ripple phase, the bilayer is
381     continuous everywhere on the whole membrane, however, in asymmetric
382     ripple phase, the bilayer is intermittent domains connected by thin
383     interdigitated monolayer which consists of upper and lower leaves of
384     the bilayer.
385     \begin{table*}
386     \begin{minipage}{\linewidth}
387     \begin{center}
388     \caption{}
389     \begin{tabular}{lccc}
390     \hline
391     $\sigma_h/\sigma_0$ & type of phase & $\lambda/\sigma_0$ & $A/\sigma_0$\\
392     \hline
393     1.20 & flat & N/A & N/A \\
394     1.28 & asymmetric flat & 21.7 & N/A \\
395     1.35 & symmetric ripple & 17.2 & 2.2 \\
396     1.41 & asymmetric ripple & 15.4 & 1.5 \\
397     \end{tabular}
398     \label{tab:property}
399     \end{center}
400     \end{minipage}
401     \end{table*}
402 xsun 3147
403 xsun 3174 The membrane structures and the reduced wavelength $\lambda/\sigma_0$,
404     reduced amplitude $A/\sigma_0$ of the ripples are summerized in Table
405     \ref{tab:property}. The wavelength range is $15~21$ and the amplitude
406     is $1.5$ for asymmetric ripple and $2.2$ for symmetric ripple. These
407     values are consistent to the experimental results. Note, the
408     amplitudes are underestimated without the melted tails in our
409     simulations.
410    
411 gezelter 3195 \begin{figure}[htb]
412     \centering
413     \includegraphics[width=\linewidth]{topDown}
414     \caption{Top views of the flat (upper), asymmetric ripple (middle),
415     and symmetric ripple (lower) phases. Note that the head-group dipoles
416     have formed head-to-tail chains in all three of these phases, but in
417     the two rippled phases, the dipolar chains are all aligned
418     {\it perpendicular} to the direction of the ripple. The flat membrane
419     has multiple point defects in the dipolar orientational ordering, and
420     the dipolar ordering on the lower leaf of the bilayer can be in a
421     different direction from the upper leaf.\label{fig:topView}}
422     \end{figure}
423    
424 xsun 3174 The $P_2$ order paramters (for molecular bodies and head group
425     dipoles) have been calculated to clarify the ordering in these phases
426     quantatively. $P_2=1$ means a perfectly ordered structure, and $P_2=0$
427     implies orientational randomization. Figure \ref{fig:rP2} shows the
428     $P_2$ order paramter of the dipoles on head group rising with
429     increasing head group size. When the heads of the lipid molecules are
430     small, the membrane is flat. The dipolar ordering is essentially
431 xsun 3189 frustrated on orientational ordering in this circumstance. Figure
432 gezelter 3195 \ref{fig:topView} shows the snapshots of the top view for the flat system
433 xsun 3189 ($\sigma_h=1.20\sigma$) and rippled system
434     ($\sigma_h=1.41\sigma$). The pointing direction of the dipoles on the
435     head groups are represented by two colored half spheres from blue to
436     yellow. For flat surfaces, the system obviously shows frustration on
437     the dipolar ordering, there are kinks on the edge of defferent
438     domains. Another reason is that the lipids can move independently in
439     each monolayer, it is not nessasory for the direction of dipoles on
440     one leaf is consistant to another layer, which makes total order
441     parameter is relatively low. With increasing head group size, the
442     surface is corrugated, and dipoles do not move as freely on the
443 xsun 3174 surface. Therefore, the translational freedom of lipids in one layer
444 xsun 3147 is dependent upon the position of lipids in another layer, as a
445 xsun 3174 result, the symmetry of the dipoles on head group in one layer is tied
446     to the symmetry in the other layer. Furthermore, as the membrane
447     deforms from two to three dimensions due to the corrugation, the
448     symmetry of the ordering for the dipoles embedded on each leaf is
449     broken. The dipoles then self-assemble in a head-tail configuration,
450     and the order parameter increases dramaticaly. However, the total
451     polarization of the system is still close to zero. This is strong
452     evidence that the corrugated structure is an antiferroelectric
453 xsun 3189 state. From the snapshot in Figure \ref{}, the dipoles arrange as
454     arrays along $Y$ axis and fall into head-to-tail configuration in each
455     line, but every $3$ or $4$ lines of dipoles change their direction
456     from neighbour lines. The system shows antiferroelectric
457     charactoristic as a whole. The orientation of the dipolar is always
458     perpendicular to the ripple wave vector. These results are consistent
459     with our previous study on dipolar membranes.
460 xsun 3147
461 xsun 3174 The ordering of the tails is essentially opposite to the ordering of
462     the dipoles on head group. The $P_2$ order parameter decreases with
463     increasing head size. This indicates the surface is more curved with
464     larger head groups. When the surface is flat, all tails are pointing
465     in the same direction; in this case, all tails are parallel to the
466     normal of the surface,(making this structure remindcent of the
467     $L_{\beta}$ phase. Increasing the size of the heads, results in
468     rapidly decreasing $P_2$ ordering for the molecular bodies.
469     \begin{figure}[htb]
470     \centering
471     \includegraphics[width=\linewidth]{rP2}
472     \caption{The $P_2$ order parameter as a funtion of the ratio of
473     $\sigma_h$ to $\sigma_0$.\label{fig:rP2}}
474     \end{figure}
475 xsun 3147
476 xsun 3174 We studied the effects of the interactions between head groups on the
477     structure of lipid bilayer by changing the strength of the dipole.
478     Figure \ref{fig:sP2} shows how the $P_2$ order parameter changes with
479     increasing strength of the dipole. Generally the dipoles on the head
480     group are more ordered by increase in the strength of the interaction
481     between heads and are more disordered by decreasing the interaction
482     stength. When the interaction between the heads is weak enough, the
483     bilayer structure does not persist; all lipid molecules are solvated
484     directly in the water. The critial value of the strength of the dipole
485     depends on the head size. The perfectly flat surface melts at $5$
486 xsun 3182 $0.03$ debye, the asymmetric rippled surfaces melt at $8$ $0.04$
487     $0.03$ debye, the symmetric rippled surfaces melt at $10$ $0.04$
488     debye. The ordering of the tails is the same as the ordering of the
489     dipoles except for the flat phase. Since the surface is already
490     perfect flat, the order parameter does not change much until the
491     strength of the dipole is $15$ debye. However, the order parameter
492     decreases quickly when the strength of the dipole is further
493     increased. The head groups of the lipid molecules are brought closer
494     by stronger interactions between them. For a flat surface, a large
495     amount of free volume between the head groups is available, but when
496     the head groups are brought closer, the tails will splay outward,
497     forming an inverse micelle. When $\sigma_h=1.28\sigma_0$, the $P_2$
498     order parameter decreases slightly after the strength of the dipole is
499     increased to $16$ debye. For rippled surfaces, there is less free
500     volume available between the head groups. Therefore there is little
501     effect on the structure of the membrane due to increasing dipolar
502     strength. However, the increase of the $P_2$ order parameter implies
503     the membranes are flatten by the increase of the strength of the
504     dipole. Unlike other systems that melt directly when the interaction
505     is weak enough, for $\sigma_h=1.41\sigma_0$, part of the membrane
506     melts into itself first. The upper leaf of the bilayer becomes totally
507     interdigitated with the lower leaf. This is different behavior than
508     what is exhibited with the interdigitated lines in the rippled phase
509     where only one interdigitated line connects the two leaves of bilayer.
510 xsun 3174 \begin{figure}[htb]
511     \centering
512     \includegraphics[width=\linewidth]{sP2}
513     \caption{The $P_2$ order parameter as a funtion of the strength of the
514     dipole.\label{fig:sP2}}
515     \end{figure}
516 xsun 3147
517 xsun 3174 Figure \ref{fig:tP2} shows the dependence of the order parameter on
518     temperature. The behavior of the $P_2$ order paramter is
519     straightforward. Systems are more ordered at low temperature, and more
520     disordered at high temperatures. When the temperature is high enough,
521 xsun 3182 the membranes are instable. For flat surfaces ($\sigma_h=1.20\sigma_0$
522     and $\sigma_h=1.28\sigma_0$), when the temperature is increased to
523     $310$, the $P_2$ order parameter increases slightly instead of
524     decreases like ripple surface. This is an evidence of the frustration
525     of the dipolar ordering in each leaf of the lipid bilayer, at low
526     temperature, the systems are locked in a local minimum energy state,
527     with increase of the temperature, the system can jump out the local
528     energy well to find the lower energy state which is the longer range
529     orientational ordering. Like the dipolar ordering of the flat
530     surfaces, the ordering of the tails of the lipid molecules for ripple
531     membranes ($\sigma_h=1.35\sigma_0$ and $\sigma_h=1.41\sigma_0$) also
532     show some nonthermal characteristic. With increase of the temperature,
533     the $P_2$ order parameter decreases firstly, and increases afterward
534     when the temperature is greater than $290 K$. The increase of the
535     $P_2$ order parameter indicates a more ordered structure for the tails
536     of the lipid molecules which corresponds to a more flat surface. Since
537     our model lacks the detailed information on lipid tails, we can not
538     simulate the fluid phase with melted fatty acid chains. Moreover, the
539     formation of the tilted $L_{\beta'}$ phase also depends on the
540     organization of fatty groups on tails.
541 xsun 3174 \begin{figure}[htb]
542     \centering
543     \includegraphics[width=\linewidth]{tP2}
544     \caption{The $P_2$ order parameter as a funtion of
545     temperature.\label{fig:tP2}}
546     \end{figure}
547 xsun 3147
548 xsun 3174 \section{Discussion}
549     \label{sec:discussion}
550 xsun 3147
551     \bibliography{mdripple}
552     \end{document}