ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
Revision: 3196
Committed: Wed Jul 25 21:39:35 2007 UTC (17 years, 1 month ago) by gezelter
Content type: application/x-tex
File size: 27646 byte(s)
Log Message:
more edits

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 gezelter 3195
284 gezelter 3186 \begin{figure}[htb]
285     \centering
286 gezelter 3196 \includegraphics[width=\linewidth]{2lipidModel}
287 gezelter 3186 \caption{The parameters defining the behavior of the lipid
288     models. $\sigma_h / \sigma_0$ is the ratio of the head group to body
289 gezelter 3196 diameter. Molecular bodies had a fixed aspect ratio of 3.0. The
290     solvent model was a simplified 4-water bead ($\sigma_w = 1.02
291     \sigma_0$) that has been used in other coarse-grained (DPD) simulations.
292     The dipolar strength (and the temperature and pressure) were the only
293     other parameters that were varied
294     systematically.\label{fig:lipidModel}}
295 gezelter 3186 \end{figure}
296    
297 gezelter 3196 \section{Experimental Methodology}
298 xsun 3174 \label{sec:experiment}
299 xsun 3147
300 gezelter 3196 To create unbiased bilayers, all simulations were started from two
301     perfectly flat monolayers separated by a 20 \AA\ gap between the
302     molecular bodies of the upper and lower leaves. The separated
303     monolayers were evolved in a vaccum with $x-y$ anisotropic pressure
304 xsun 3174 coupling. The length of $z$ axis of the simulations was fixed and a
305     constant surface tension was applied to enable real fluctuations of
306 gezelter 3196 the bilayer. Periodic boundaries were used, and $480-720$ lipid
307     molecules were present in the simulations depending on the size of the
308     head beads. The two monolayers spontaneously collapse into bilayer
309     structures within 100 ps, and following this collapse, all systems
310     were equlibrated for $100$ ns at $300$ K.
311 xsun 3147
312 gezelter 3196 The resulting structures were then solvated at a ratio of $6$ DPD
313     solvent beads (24 water molecules) per lipid. These configurations
314     were then equilibrated for another $30$ ns. All simulations with
315     solvent were carried out at constant pressure ($P=1$ atm) by $3$D
316     anisotropic coupling, and constant surface tension ($\gamma=0.015$
317     UNIT). Given the absence of fast degrees of freedom in this model, a
318     timestep of $50$ fs was utilized. Data collection for structural
319     properties of the bilayers was carried out during a final 5 ns run
320     following the solvent equilibration. All simulations were performed
321     using the OOPSE molecular modeling program.\cite{Meineke05}
322    
323     \section{Results}
324 xsun 3174 \label{sec:results}
325 xsun 3147
326 xsun 3174 Snapshots in Figure \ref{fig:phaseCartoon} show that the membrane is
327     more corrugated increasing size of the head groups. The surface is
328     nearly flat when $\sigma_h=1.20\sigma_0$. With
329     $\sigma_h=1.28\sigma_0$, although the surface is still flat, the
330     bilayer starts to splay inward; the upper leaf of the bilayer is
331     connected to the lower leaf with an interdigitated line defect. Two
332     periodicities with $100$ \AA\ width were observed in the
333     simulation. This structure is very similiar to the structure observed
334     by de Vries and Lenz {\it et al.}. The same basic structure is also
335     observed when $\sigma_h=1.41\sigma_0$, but the wavelength of the
336     surface corrugations depends sensitively on the size of the ``head''
337     beads. From the undulation spectrum, the corrugation is clearly
338     non-thermal.
339     \begin{figure}[htb]
340     \centering
341     \includegraphics[width=\linewidth]{phaseCartoon}
342     \caption{A sketch to discribe the structure of the phases observed in
343     our simulations.\label{fig:phaseCartoon}}
344     \end{figure}
345 xsun 3147
346 xsun 3174 When $\sigma_h=1.35\sigma_0$, we observed another corrugated surface
347     morphology. This structure is different from the asymmetric rippled
348     surface; there is no interdigitation between the upper and lower
349     leaves of the bilayer. Each leaf of the bilayer is broken into several
350     hemicylinderical sections, and opposite leaves are fitted together
351     much like roof tiles. Unlike the surface in which the upper
352     hemicylinder is always interdigitated on the leading or trailing edge
353     of lower hemicylinder, the symmetric ripple has no prefered direction.
354     The corresponding cartoons are shown in Figure
355     \ref{fig:phaseCartoon} for elucidation of the detailed structures of
356     different phases. Figure \ref{fig:phaseCartoon} (a) is the flat phase,
357     (b) is the asymmetric ripple phase corresponding to the lipid
358     organization when $\sigma_h=1.28\sigma_0$ and $\sigma_h=1.41\sigma_0$,
359     and (c) is the symmetric ripple phase observed when
360     $\sigma_h=1.35\sigma_0$. In symmetric ripple phase, the bilayer is
361     continuous everywhere on the whole membrane, however, in asymmetric
362     ripple phase, the bilayer is intermittent domains connected by thin
363     interdigitated monolayer which consists of upper and lower leaves of
364     the bilayer.
365     \begin{table*}
366     \begin{minipage}{\linewidth}
367     \begin{center}
368     \caption{}
369     \begin{tabular}{lccc}
370     \hline
371     $\sigma_h/\sigma_0$ & type of phase & $\lambda/\sigma_0$ & $A/\sigma_0$\\
372     \hline
373     1.20 & flat & N/A & N/A \\
374     1.28 & asymmetric flat & 21.7 & N/A \\
375     1.35 & symmetric ripple & 17.2 & 2.2 \\
376     1.41 & asymmetric ripple & 15.4 & 1.5 \\
377     \end{tabular}
378     \label{tab:property}
379     \end{center}
380     \end{minipage}
381     \end{table*}
382 xsun 3147
383 xsun 3174 The membrane structures and the reduced wavelength $\lambda/\sigma_0$,
384     reduced amplitude $A/\sigma_0$ of the ripples are summerized in Table
385     \ref{tab:property}. The wavelength range is $15~21$ and the amplitude
386     is $1.5$ for asymmetric ripple and $2.2$ for symmetric ripple. These
387     values are consistent to the experimental results. Note, the
388     amplitudes are underestimated without the melted tails in our
389     simulations.
390    
391 gezelter 3195 \begin{figure}[htb]
392     \centering
393     \includegraphics[width=\linewidth]{topDown}
394     \caption{Top views of the flat (upper), asymmetric ripple (middle),
395     and symmetric ripple (lower) phases. Note that the head-group dipoles
396     have formed head-to-tail chains in all three of these phases, but in
397     the two rippled phases, the dipolar chains are all aligned
398     {\it perpendicular} to the direction of the ripple. The flat membrane
399     has multiple point defects in the dipolar orientational ordering, and
400     the dipolar ordering on the lower leaf of the bilayer can be in a
401     different direction from the upper leaf.\label{fig:topView}}
402     \end{figure}
403    
404 xsun 3174 The $P_2$ order paramters (for molecular bodies and head group
405     dipoles) have been calculated to clarify the ordering in these phases
406     quantatively. $P_2=1$ means a perfectly ordered structure, and $P_2=0$
407     implies orientational randomization. Figure \ref{fig:rP2} shows the
408     $P_2$ order paramter of the dipoles on head group rising with
409     increasing head group size. When the heads of the lipid molecules are
410     small, the membrane is flat. The dipolar ordering is essentially
411 xsun 3189 frustrated on orientational ordering in this circumstance. Figure
412 gezelter 3195 \ref{fig:topView} shows the snapshots of the top view for the flat system
413 xsun 3189 ($\sigma_h=1.20\sigma$) and rippled system
414     ($\sigma_h=1.41\sigma$). The pointing direction of the dipoles on the
415     head groups are represented by two colored half spheres from blue to
416     yellow. For flat surfaces, the system obviously shows frustration on
417     the dipolar ordering, there are kinks on the edge of defferent
418     domains. Another reason is that the lipids can move independently in
419     each monolayer, it is not nessasory for the direction of dipoles on
420     one leaf is consistant to another layer, which makes total order
421     parameter is relatively low. With increasing head group size, the
422     surface is corrugated, and dipoles do not move as freely on the
423 xsun 3174 surface. Therefore, the translational freedom of lipids in one layer
424 xsun 3147 is dependent upon the position of lipids in another layer, as a
425 xsun 3174 result, the symmetry of the dipoles on head group in one layer is tied
426     to the symmetry in the other layer. Furthermore, as the membrane
427     deforms from two to three dimensions due to the corrugation, the
428     symmetry of the ordering for the dipoles embedded on each leaf is
429     broken. The dipoles then self-assemble in a head-tail configuration,
430     and the order parameter increases dramaticaly. However, the total
431     polarization of the system is still close to zero. This is strong
432     evidence that the corrugated structure is an antiferroelectric
433 xsun 3189 state. From the snapshot in Figure \ref{}, the dipoles arrange as
434     arrays along $Y$ axis and fall into head-to-tail configuration in each
435     line, but every $3$ or $4$ lines of dipoles change their direction
436     from neighbour lines. The system shows antiferroelectric
437     charactoristic as a whole. The orientation of the dipolar is always
438     perpendicular to the ripple wave vector. These results are consistent
439     with our previous study on dipolar membranes.
440 xsun 3147
441 xsun 3174 The ordering of the tails is essentially opposite to the ordering of
442     the dipoles on head group. The $P_2$ order parameter decreases with
443     increasing head size. This indicates the surface is more curved with
444     larger head groups. When the surface is flat, all tails are pointing
445     in the same direction; in this case, all tails are parallel to the
446     normal of the surface,(making this structure remindcent of the
447     $L_{\beta}$ phase. Increasing the size of the heads, results in
448     rapidly decreasing $P_2$ ordering for the molecular bodies.
449     \begin{figure}[htb]
450     \centering
451     \includegraphics[width=\linewidth]{rP2}
452     \caption{The $P_2$ order parameter as a funtion of the ratio of
453     $\sigma_h$ to $\sigma_0$.\label{fig:rP2}}
454     \end{figure}
455 xsun 3147
456 xsun 3174 We studied the effects of the interactions between head groups on the
457     structure of lipid bilayer by changing the strength of the dipole.
458     Figure \ref{fig:sP2} shows how the $P_2$ order parameter changes with
459     increasing strength of the dipole. Generally the dipoles on the head
460     group are more ordered by increase in the strength of the interaction
461     between heads and are more disordered by decreasing the interaction
462     stength. When the interaction between the heads is weak enough, the
463     bilayer structure does not persist; all lipid molecules are solvated
464     directly in the water. The critial value of the strength of the dipole
465     depends on the head size. The perfectly flat surface melts at $5$
466 xsun 3182 $0.03$ debye, the asymmetric rippled surfaces melt at $8$ $0.04$
467     $0.03$ debye, the symmetric rippled surfaces melt at $10$ $0.04$
468     debye. The ordering of the tails is the same as the ordering of the
469     dipoles except for the flat phase. Since the surface is already
470     perfect flat, the order parameter does not change much until the
471     strength of the dipole is $15$ debye. However, the order parameter
472     decreases quickly when the strength of the dipole is further
473     increased. The head groups of the lipid molecules are brought closer
474     by stronger interactions between them. For a flat surface, a large
475     amount of free volume between the head groups is available, but when
476     the head groups are brought closer, the tails will splay outward,
477     forming an inverse micelle. When $\sigma_h=1.28\sigma_0$, the $P_2$
478     order parameter decreases slightly after the strength of the dipole is
479     increased to $16$ debye. For rippled surfaces, there is less free
480     volume available between the head groups. Therefore there is little
481     effect on the structure of the membrane due to increasing dipolar
482     strength. However, the increase of the $P_2$ order parameter implies
483     the membranes are flatten by the increase of the strength of the
484     dipole. Unlike other systems that melt directly when the interaction
485     is weak enough, for $\sigma_h=1.41\sigma_0$, part of the membrane
486     melts into itself first. The upper leaf of the bilayer becomes totally
487     interdigitated with the lower leaf. This is different behavior than
488     what is exhibited with the interdigitated lines in the rippled phase
489     where only one interdigitated line connects the two leaves of bilayer.
490 xsun 3174 \begin{figure}[htb]
491     \centering
492     \includegraphics[width=\linewidth]{sP2}
493     \caption{The $P_2$ order parameter as a funtion of the strength of the
494     dipole.\label{fig:sP2}}
495     \end{figure}
496 xsun 3147
497 xsun 3174 Figure \ref{fig:tP2} shows the dependence of the order parameter on
498     temperature. The behavior of the $P_2$ order paramter is
499     straightforward. Systems are more ordered at low temperature, and more
500     disordered at high temperatures. When the temperature is high enough,
501 xsun 3182 the membranes are instable. For flat surfaces ($\sigma_h=1.20\sigma_0$
502     and $\sigma_h=1.28\sigma_0$), when the temperature is increased to
503     $310$, the $P_2$ order parameter increases slightly instead of
504     decreases like ripple surface. This is an evidence of the frustration
505     of the dipolar ordering in each leaf of the lipid bilayer, at low
506     temperature, the systems are locked in a local minimum energy state,
507     with increase of the temperature, the system can jump out the local
508     energy well to find the lower energy state which is the longer range
509     orientational ordering. Like the dipolar ordering of the flat
510     surfaces, the ordering of the tails of the lipid molecules for ripple
511     membranes ($\sigma_h=1.35\sigma_0$ and $\sigma_h=1.41\sigma_0$) also
512     show some nonthermal characteristic. With increase of the temperature,
513     the $P_2$ order parameter decreases firstly, and increases afterward
514     when the temperature is greater than $290 K$. The increase of the
515     $P_2$ order parameter indicates a more ordered structure for the tails
516     of the lipid molecules which corresponds to a more flat surface. Since
517     our model lacks the detailed information on lipid tails, we can not
518     simulate the fluid phase with melted fatty acid chains. Moreover, the
519     formation of the tilted $L_{\beta'}$ phase also depends on the
520     organization of fatty groups on tails.
521 xsun 3174 \begin{figure}[htb]
522     \centering
523     \includegraphics[width=\linewidth]{tP2}
524     \caption{The $P_2$ order parameter as a funtion of
525     temperature.\label{fig:tP2}}
526     \end{figure}
527 xsun 3147
528 xsun 3174 \section{Discussion}
529     \label{sec:discussion}
530 xsun 3147
531     \bibliography{mdripple}
532     \end{document}