ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
Revision: 3199
Committed: Thu Jul 26 19:47:06 2007 UTC (17 years, 1 month ago) by gezelter
Content type: application/x-tex
File size: 29904 byte(s)
Log Message:
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 gezelter 3199 \usepackage{amsmath}
4     \usepackage{amssymb}
5 xsun 3147 \usepackage{graphicx}
6    
7     \begin{document}
8     \renewcommand{\thefootnote}{\fnsymbol{footnote}}
9     \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
10    
11     %\bibliographystyle{aps}
12    
13 gezelter 3199 \title{Dipolar Ordering in Molecular-scale Models of the Ripple Phase
14     in Lipid Membranes}
15 xsun 3147 \author{Xiuquan Sun and J. Daniel Gezelter}
16     \email[E-mail:]{gezelter@nd.edu}
17     \affiliation{Department of Chemistry and Biochemistry,\\
18 gezelter 3199 University of Notre Dame, \\
19 xsun 3147 Notre Dame, Indiana 46556}
20    
21     \date{\today}
22    
23     \begin{abstract}
24 gezelter 3195 The ripple phase in phosphatidylcholine (PC) bilayers has never been
25     completely explained.
26 xsun 3147 \end{abstract}
27    
28     \pacs{}
29     \maketitle
30    
31 xsun 3174 \section{Introduction}
32     \label{sec:Int}
33 gezelter 3195 Fully hydrated lipids will aggregate spontaneously to form bilayers
34     which exhibit a variety of phases depending on their temperatures and
35     compositions. Among these phases, a periodic rippled phase
36     ($P_{\beta'}$) appears as an intermediate phase between the gel
37     ($L_\beta$) and fluid ($L_{\alpha}$) phases for relatively pure
38     phosphatidylcholine (PC) bilayers. The ripple phase has attracted
39     substantial experimental interest over the past 30 years. Most
40     structural information of the ripple phase has been obtained by the
41     X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
42     microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
43     et al.} used atomic force microscopy (AFM) to observe ripple phase
44     morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
45     experimental results provide strong support for a 2-dimensional
46     hexagonal packing lattice of the lipid molecules within the ripple
47     phase. This is a notable change from the observed lipid packing
48     within the gel phase.~\cite{Cevc87}
49 xsun 3174
50 gezelter 3195 A number of theoretical models have been presented to explain the
51     formation of the ripple phase. Marder {\it et al.} used a
52     curvature-dependent Landau-de Gennes free-energy functional to predict
53     a rippled phase.~\cite{Marder84} This model and other related continuum
54     models predict higher fluidity in convex regions and that concave
55     portions of the membrane correspond to more solid-like regions.
56     Carlson and Sethna used a packing-competition model (in which head
57     groups and chains have competing packing energetics) to predict the
58     formation of a ripple-like phase. Their model predicted that the
59     high-curvature portions have lower-chain packing and correspond to
60     more fluid-like regions. Goldstein and Leibler used a mean-field
61     approach with a planar model for {\em inter-lamellar} interactions to
62     predict rippling in multilamellar phases.~\cite{Goldstein88} McCullough
63     and Scott proposed that the {\em anisotropy of the nearest-neighbor
64     interactions} coupled to hydrophobic constraining forces which
65     restrict height differences between nearest neighbors is the origin of
66     the ripple phase.~\cite{McCullough90} Lubensky and MacKintosh
67     introduced a Landau theory for tilt order and curvature of a single
68     membrane and concluded that {\em coupling of molecular tilt to membrane
69     curvature} is responsible for the production of
70     ripples.~\cite{Lubensky93} Misbah, Duplat and Houchmandzadeh proposed
71     that {\em inter-layer dipolar interactions} can lead to ripple
72     instabilities.~\cite{Misbah98} Heimburg presented a {\em coexistence
73     model} for ripple formation in which he postulates that fluid-phase
74     line defects cause sharp curvature between relatively flat gel-phase
75     regions.~\cite{Heimburg00} Kubica has suggested that a lattice model of
76     polar head groups could be valuable in trying to understand bilayer
77     phase formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations
78     of lamellar stacks of hexagonal lattices to show that large headgroups
79     and molecular tilt with respect to the membrane normal vector can
80     cause bulk rippling.~\cite{Bannerjee02}
81 xsun 3174
82 gezelter 3195 In contrast, few large-scale molecular modelling studies have been
83     done due to the large size of the resulting structures and the time
84     required for the phases of interest to develop. With all-atom (and
85     even unified-atom) simulations, only one period of the ripple can be
86     observed and only for timescales in the range of 10-100 ns. One of
87     the most interesting molecular simulations was carried out by De Vries
88     {\it et al.}~\cite{deVries05}. According to their simulation results,
89     the ripple consists of two domains, one resembling the gel bilayer,
90     while in the other, the two leaves of the bilayer are fully
91     interdigitated. The mechanism for the formation of the ripple phase
92     suggested by their work is a packing competition between the head
93     groups and the tails of the lipid molecules.\cite{Carlson87} Recently,
94 gezelter 3199 the ripple phase has also been studied by Lenz and Schmid using Monte
95 gezelter 3195 Carlo simulations.\cite{Lenz07} Their structures are similar to the De
96     Vries {\it et al.} structures except that the connection between the
97     two leaves of the bilayer is a narrow interdigitated line instead of
98     the fully interdigitated domain. The symmetric ripple phase was also
99     observed by Lenz {\it et al.}, and their work supports other claims
100     that the mismatch between the size of the head group and tail of the
101     lipid molecules is the driving force for the formation of the ripple
102     phase. Ayton and Voth have found significant undulations in
103     zero-surface-tension states of membranes simulated via dissipative
104     particle dynamics, but their results are consistent with purely
105     thermal undulations.~\cite{Ayton02}
106 xsun 3174
107 gezelter 3195 Although the organization of the tails of lipid molecules are
108     addressed by these molecular simulations and the packing competition
109     between headgroups and tails is strongly implicated as the primary
110     driving force for ripple formation, questions about the ordering of
111     the head groups in ripple phase has not been settled.
112 xsun 3174
113 gezelter 3195 In a recent paper, we presented a simple ``web of dipoles'' spin
114     lattice model which provides some physical insight into relationship
115     between dipolar ordering and membrane buckling.\cite{Sun2007} We found
116     that dipolar elastic membranes can spontaneously buckle, forming
117     ripple-like topologies. The driving force for the buckling in dipolar
118     elastic membranes the antiferroelectric ordering of the dipoles, and
119     this was evident in the ordering of the dipole director axis
120     perpendicular to the wave vector of the surface ripples. A similiar
121     phenomenon has also been observed by Tsonchev {\it et al.} in their
122 gezelter 3199 work on the spontaneous formation of dipolar peptide chains into
123     curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
124 gezelter 3195
125     In this paper, we construct a somewhat more realistic molecular-scale
126     lipid model than our previous ``web of dipoles'' and use molecular
127     dynamics simulations to elucidate the role of the head group dipoles
128     in the formation and morphology of the ripple phase. We describe our
129     model and computational methodology in section \ref{sec:method}.
130     Details on the simulations are presented in section
131     \ref{sec:experiment}, with results following in section
132     \ref{sec:results}. A final discussion of the role of dipolar heads in
133     the ripple formation can be found in section
134 xsun 3174 \ref{sec:discussion}.
135    
136 gezelter 3196 \section{Computational Model}
137 xsun 3174 \label{sec:method}
138    
139 gezelter 3199 \begin{figure}[htb]
140     \centering
141     \includegraphics[width=4in]{lipidModels}
142     \caption{Three different representations of DPPC lipid molecules,
143     including the chemical structure, an atomistic model, and the
144     head-body ellipsoidal coarse-grained model used in this
145     work.\label{fig:lipidModels}}
146     \end{figure}
147    
148 gezelter 3195 Our simple molecular-scale lipid model for studying the ripple phase
149     is based on two facts: one is that the most essential feature of lipid
150     molecules is their amphiphilic structure with polar head groups and
151     non-polar tails. Another fact is that the majority of lipid molecules
152     in the ripple phase are relatively rigid (i.e. gel-like) which makes
153     some fraction of the details of the chain dynamics negligible. Figure
154     \ref{fig:lipidModels} shows the molecular strucure of a DPPC
155     molecule, as well as atomistic and molecular-scale representations of
156     a DPPC molecule. The hydrophilic character of the head group is
157     largely due to the separation of charge between the nitrogen and
158     phosphate groups. The zwitterionic nature of the PC headgroups leads
159     to abnormally large dipole moments (as high as 20.6 D), and this
160     strongly polar head group interacts strongly with the solvating water
161     layers immediately surrounding the membrane. The hydrophobic tail
162     consists of fatty acid chains. In our molecular scale model, lipid
163     molecules have been reduced to these essential features; the fatty
164     acid chains are represented by an ellipsoid with a dipolar ball
165     perched on one end to represent the effects of the charge-separated
166     head group. In real PC lipids, the direction of the dipole is
167     nearly perpendicular to the tail, so we have fixed the direction of
168     the point dipole rigidly in this orientation.
169 xsun 3147
170 gezelter 3195 The ellipsoidal portions of the model interact via the Gay-Berne
171     potential which has seen widespread use in the liquid crystal
172 gezelter 3199 community. Ayton and Voth have also used Gay-Berne ellipsoids for
173     modelling large length-scale properties of lipid
174     bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
175     was a single site model for the interactions of rigid ellipsoidal
176 gezelter 3195 molecules.\cite{Gay81} It can be thought of as a modification of the
177     Gaussian overlap model originally described by Berne and
178     Pechukas.\cite{Berne72} The potential is constructed in the familiar
179     form of the Lennard-Jones function using orientation-dependent
180     $\sigma$ and $\epsilon$ parameters,
181 xsun 3174 \begin{eqnarray*}
182     V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
183     r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
184     {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
185     {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
186     -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
187     {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
188 gezelter 3195 \label{eq:gb}
189     \end{eqnarray*}
190    
191     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
192 gezelter 3199 \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
193     \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
194 gezelter 3195 are dependent on the relative orientations of the two molecules (${\bf
195     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
196 gezelter 3199 intermolecular separation (${\bf \hat{r}}_{ij}$). $\sigma$ and
197     $\sigma_0$ are also governed by shape mixing and anisotropy variables,
198 gezelter 3195 \begin {equation}
199     \begin{array}{rcl}
200 gezelter 3199 \sigma_0 & = & \sqrt{d_i^2 + d_j^2} \\ \\
201     \chi & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 -
202     d_j^2 \right)}{\left( l_j^2 + d_i^2 \right) \left(l_i^2 +
203     d_j^2 \right)}\right]^{1/2} \\ \\
204     \alpha^2 & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 +
205     d_i^2 \right)}{\left( l_j^2 - d_j^2 \right) \left(l_i^2 +
206     d_j^2 \right)}\right]^{1/2},
207 gezelter 3195 \end{array}
208     \end{equation}
209 gezelter 3199 where $l$ and $d$ describe the length and width of each uniaxial
210     ellipsoid. These shape anisotropy parameters can then be used to
211     calculate the range function,
212     \begin {equation}
213     \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) = \sigma_{0}
214     \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
215     \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
216     \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
217     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
218     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
219     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
220     \right]^{-1/2}
221     \end{equation}
222    
223     Gay-Berne ellipsoids also have an energy scaling parameter,
224     $\epsilon^s$, which describes the well depth for two identical
225     ellipsoids in a {\it side-by-side} configuration. Additionaly, a well
226     depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
227     the ratio between the well depths in the {\it end-to-end} and
228     side-by-side configurations. As in the range parameter, a set of
229     mixing and anisotropy variables can be used to describe the well
230     depths for dissimilar particles,
231     \begin {eqnarray*}
232     \epsilon_0 & = & \sqrt{\epsilon^s_i * \epsilon^s_j} \\ \\
233     \epsilon^r & = & \sqrt{\epsilon^r_i * \epsilon^r_j} \\ \\
234     \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
235     \\ \\
236     \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
237     \end{eqnarray*}
238     The form of the strength function is somewhat complicated,
239     \begin {eqnarray*}
240     \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
241     \epsilon_{0} \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
242     \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
243     \hat{r}}_{ij}) \\ \\
244     \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
245     \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
246     \hat{u}}_{j})^{2}\right]^{-1/2} \\ \\
247     \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) &
248     = &
249     1 - \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
250     \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
251     \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
252     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
253     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
254     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
255     \end {eqnarray*}
256     although many of the quantities and derivatives are identical with
257     those obtained for the range parameter. Ref. \onlinecite{Luckhurst90}
258     has a particularly good explanation of the choice of the Gay-Berne
259     parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
260     excellent overview of the computational methods that can be used to
261     efficiently compute forces and torques for this potential can be found
262     in Ref. \onlinecite{Golubkov06}
263    
264     The choices of parameters we have used in this study correspond to a
265     shape anisotropy of 3 for the chain portion of the molecule. In
266     principle, this could be varied to allow for modeling of longer or
267     shorter chain lipid molecules. For these prolate ellipsoids, we have:
268 gezelter 3195 \begin{equation}
269     \begin{array}{rcl}
270 gezelter 3199 d & < & l \\
271     \epsilon^{r} & < & 1
272 gezelter 3195 \end{array}
273     \end{equation}
274    
275 gezelter 3199 \begin{figure}[htb]
276     \centering
277     \includegraphics[width=4in]{2lipidModel}
278     \caption{The parameters defining the behavior of the lipid
279     models. $l / d$ is the ratio of the head group to body diameter.
280     Molecular bodies had a fixed aspect ratio of 3.0. The solvent model
281     was a simplified 4-water bead ($\sigma_w \approx d$) that has been
282     used in other coarse-grained (DPD) simulations. The dipolar strength
283     (and the temperature and pressure) were the only other parameters that
284     were varied systematically.\label{fig:lipidModel}}
285     \end{figure}
286 gezelter 3195
287     To take into account the permanent dipolar interactions of the
288     zwitterionic head groups, we place fixed dipole moments $\mu_{i}$ at
289 gezelter 3199 one end of the Gay-Berne particles. The dipoles are oriented at an
290     angle $\theta = \pi / 2$ relative to the major axis. These dipoles
291 gezelter 3195 are protected by a head ``bead'' with a range parameter which we have
292 gezelter 3199 varied between $1.20 d$ and $1.41 d$. The head groups interact with
293     each other using a combination of Lennard-Jones,
294 xsun 3174 \begin{eqnarray*}
295 gezelter 3195 V_{ij} = 4\epsilon \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
296     \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
297 xsun 3174 \end{eqnarray*}
298 gezelter 3199 and dipole-dipole,
299 xsun 3174 \begin{eqnarray*}
300 gezelter 3195 V_{ij} = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
301     \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
302     \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
303 xsun 3174 \end{eqnarray*}
304 gezelter 3195 potentials.
305     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
306     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
307 gezelter 3199 pointing along the inter-dipole vector $\mathbf{r}_{ij}$.
308 gezelter 3195
309     For the interaction between nonequivalent uniaxial ellipsoids (in this
310 gezelter 3199 case, between spheres and ellipsoids), the spheres are treated as
311     ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
312     ratio of 1 ($\epsilon^e = \epsilon^s$). The form of the Gay-Berne
313     potential we are using was generalized by Cleaver {\it et al.} and is
314     appropriate for dissimilar uniaxial ellipsoids.\cite{Cleaver96}
315 xsun 3147
316 gezelter 3199 The solvent model in our simulations is identical to one used by
317     Marrink {\it et al.} in their dissipative particle dynamics (DPD)
318     simulation of lipid bilayers.\cite{Marrink04} This solvent bead is a single
319     site that represents four water molecules (m = 72 amu) and has
320     comparable density and diffusive behavior to liquid water. However,
321     since there are no electrostatic sites on these beads, this solvent
322     model cannot replicate the dielectric properties of water.
323 xsun 3198 \begin{table*}
324     \begin{minipage}{\linewidth}
325     \begin{center}
326 gezelter 3199 \caption{Potential parameters used for molecular-scale coarse-grained
327     lipid simulations}
328     \begin{tabular}{llccc}
329 xsun 3198 \hline
330 gezelter 3199 & & Head & Chain & Solvent \\
331 xsun 3198 \hline
332 gezelter 3199 $d$ (\AA) & & varied & 4.6 & 4.7 \\
333     $l$ (\AA) & & 1 & 3 & 1 \\
334     $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
335     $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 \\
336     $m$ (amu) & & 196 & 760 & 72.06112 \\
337     $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
338     \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
339     \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
340     \multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\
341     $\mu$ (Debye) & & varied & 0 & 0 \\
342 xsun 3198 \end{tabular}
343     \label{tab:parameters}
344     \end{center}
345     \end{minipage}
346     \end{table*}
347 gezelter 3195
348 gezelter 3199 A switching function has been applied to all potentials to smoothly
349     turn off the interactions between a range of $22$ and $25$ \AA.
350 gezelter 3186
351 gezelter 3196 \section{Experimental Methodology}
352 xsun 3174 \label{sec:experiment}
353 xsun 3147
354 gezelter 3196 To create unbiased bilayers, all simulations were started from two
355     perfectly flat monolayers separated by a 20 \AA\ gap between the
356     molecular bodies of the upper and lower leaves. The separated
357     monolayers were evolved in a vaccum with $x-y$ anisotropic pressure
358 xsun 3174 coupling. The length of $z$ axis of the simulations was fixed and a
359     constant surface tension was applied to enable real fluctuations of
360 gezelter 3196 the bilayer. Periodic boundaries were used, and $480-720$ lipid
361     molecules were present in the simulations depending on the size of the
362     head beads. The two monolayers spontaneously collapse into bilayer
363     structures within 100 ps, and following this collapse, all systems
364     were equlibrated for $100$ ns at $300$ K.
365 xsun 3147
366 gezelter 3196 The resulting structures were then solvated at a ratio of $6$ DPD
367     solvent beads (24 water molecules) per lipid. These configurations
368     were then equilibrated for another $30$ ns. All simulations with
369     solvent were carried out at constant pressure ($P=1$ atm) by $3$D
370     anisotropic coupling, and constant surface tension ($\gamma=0.015$
371     UNIT). Given the absence of fast degrees of freedom in this model, a
372     timestep of $50$ fs was utilized. Data collection for structural
373     properties of the bilayers was carried out during a final 5 ns run
374     following the solvent equilibration. All simulations were performed
375     using the OOPSE molecular modeling program.\cite{Meineke05}
376    
377     \section{Results}
378 xsun 3174 \label{sec:results}
379 xsun 3147
380 xsun 3174 Snapshots in Figure \ref{fig:phaseCartoon} show that the membrane is
381     more corrugated increasing size of the head groups. The surface is
382     nearly flat when $\sigma_h=1.20\sigma_0$. With
383     $\sigma_h=1.28\sigma_0$, although the surface is still flat, the
384     bilayer starts to splay inward; the upper leaf of the bilayer is
385     connected to the lower leaf with an interdigitated line defect. Two
386     periodicities with $100$ \AA\ width were observed in the
387     simulation. This structure is very similiar to the structure observed
388     by de Vries and Lenz {\it et al.}. The same basic structure is also
389     observed when $\sigma_h=1.41\sigma_0$, but the wavelength of the
390     surface corrugations depends sensitively on the size of the ``head''
391     beads. From the undulation spectrum, the corrugation is clearly
392     non-thermal.
393     \begin{figure}[htb]
394     \centering
395 gezelter 3199 \includegraphics[width=4in]{phaseCartoon}
396 xsun 3174 \caption{A sketch to discribe the structure of the phases observed in
397     our simulations.\label{fig:phaseCartoon}}
398     \end{figure}
399 xsun 3147
400 xsun 3174 When $\sigma_h=1.35\sigma_0$, we observed another corrugated surface
401     morphology. This structure is different from the asymmetric rippled
402     surface; there is no interdigitation between the upper and lower
403     leaves of the bilayer. Each leaf of the bilayer is broken into several
404     hemicylinderical sections, and opposite leaves are fitted together
405     much like roof tiles. Unlike the surface in which the upper
406     hemicylinder is always interdigitated on the leading or trailing edge
407     of lower hemicylinder, the symmetric ripple has no prefered direction.
408     The corresponding cartoons are shown in Figure
409     \ref{fig:phaseCartoon} for elucidation of the detailed structures of
410     different phases. Figure \ref{fig:phaseCartoon} (a) is the flat phase,
411     (b) is the asymmetric ripple phase corresponding to the lipid
412     organization when $\sigma_h=1.28\sigma_0$ and $\sigma_h=1.41\sigma_0$,
413     and (c) is the symmetric ripple phase observed when
414     $\sigma_h=1.35\sigma_0$. In symmetric ripple phase, the bilayer is
415     continuous everywhere on the whole membrane, however, in asymmetric
416     ripple phase, the bilayer is intermittent domains connected by thin
417     interdigitated monolayer which consists of upper and lower leaves of
418     the bilayer.
419     \begin{table*}
420     \begin{minipage}{\linewidth}
421     \begin{center}
422     \caption{}
423     \begin{tabular}{lccc}
424     \hline
425     $\sigma_h/\sigma_0$ & type of phase & $\lambda/\sigma_0$ & $A/\sigma_0$\\
426     \hline
427     1.20 & flat & N/A & N/A \\
428     1.28 & asymmetric flat & 21.7 & N/A \\
429     1.35 & symmetric ripple & 17.2 & 2.2 \\
430     1.41 & asymmetric ripple & 15.4 & 1.5 \\
431     \end{tabular}
432     \label{tab:property}
433     \end{center}
434     \end{minipage}
435     \end{table*}
436 xsun 3147
437 xsun 3174 The membrane structures and the reduced wavelength $\lambda/\sigma_0$,
438     reduced amplitude $A/\sigma_0$ of the ripples are summerized in Table
439     \ref{tab:property}. The wavelength range is $15~21$ and the amplitude
440     is $1.5$ for asymmetric ripple and $2.2$ for symmetric ripple. These
441     values are consistent to the experimental results. Note, the
442     amplitudes are underestimated without the melted tails in our
443     simulations.
444    
445 gezelter 3195 \begin{figure}[htb]
446     \centering
447 gezelter 3199 \includegraphics[width=4in]{topDown}
448 gezelter 3195 \caption{Top views of the flat (upper), asymmetric ripple (middle),
449     and symmetric ripple (lower) phases. Note that the head-group dipoles
450     have formed head-to-tail chains in all three of these phases, but in
451     the two rippled phases, the dipolar chains are all aligned
452     {\it perpendicular} to the direction of the ripple. The flat membrane
453     has multiple point defects in the dipolar orientational ordering, and
454     the dipolar ordering on the lower leaf of the bilayer can be in a
455     different direction from the upper leaf.\label{fig:topView}}
456     \end{figure}
457    
458 xsun 3174 The $P_2$ order paramters (for molecular bodies and head group
459     dipoles) have been calculated to clarify the ordering in these phases
460     quantatively. $P_2=1$ means a perfectly ordered structure, and $P_2=0$
461     implies orientational randomization. Figure \ref{fig:rP2} shows the
462     $P_2$ order paramter of the dipoles on head group rising with
463     increasing head group size. When the heads of the lipid molecules are
464     small, the membrane is flat. The dipolar ordering is essentially
465 xsun 3189 frustrated on orientational ordering in this circumstance. Figure
466 gezelter 3195 \ref{fig:topView} shows the snapshots of the top view for the flat system
467 xsun 3189 ($\sigma_h=1.20\sigma$) and rippled system
468     ($\sigma_h=1.41\sigma$). The pointing direction of the dipoles on the
469     head groups are represented by two colored half spheres from blue to
470     yellow. For flat surfaces, the system obviously shows frustration on
471     the dipolar ordering, there are kinks on the edge of defferent
472     domains. Another reason is that the lipids can move independently in
473     each monolayer, it is not nessasory for the direction of dipoles on
474     one leaf is consistant to another layer, which makes total order
475     parameter is relatively low. With increasing head group size, the
476     surface is corrugated, and dipoles do not move as freely on the
477 xsun 3174 surface. Therefore, the translational freedom of lipids in one layer
478 xsun 3147 is dependent upon the position of lipids in another layer, as a
479 xsun 3174 result, the symmetry of the dipoles on head group in one layer is tied
480     to the symmetry in the other layer. Furthermore, as the membrane
481     deforms from two to three dimensions due to the corrugation, the
482     symmetry of the ordering for the dipoles embedded on each leaf is
483     broken. The dipoles then self-assemble in a head-tail configuration,
484     and the order parameter increases dramaticaly. However, the total
485     polarization of the system is still close to zero. This is strong
486     evidence that the corrugated structure is an antiferroelectric
487 xsun 3189 state. From the snapshot in Figure \ref{}, the dipoles arrange as
488     arrays along $Y$ axis and fall into head-to-tail configuration in each
489     line, but every $3$ or $4$ lines of dipoles change their direction
490     from neighbour lines. The system shows antiferroelectric
491     charactoristic as a whole. The orientation of the dipolar is always
492     perpendicular to the ripple wave vector. These results are consistent
493     with our previous study on dipolar membranes.
494 xsun 3147
495 xsun 3174 The ordering of the tails is essentially opposite to the ordering of
496     the dipoles on head group. The $P_2$ order parameter decreases with
497     increasing head size. This indicates the surface is more curved with
498     larger head groups. When the surface is flat, all tails are pointing
499     in the same direction; in this case, all tails are parallel to the
500     normal of the surface,(making this structure remindcent of the
501     $L_{\beta}$ phase. Increasing the size of the heads, results in
502     rapidly decreasing $P_2$ ordering for the molecular bodies.
503 gezelter 3199
504 xsun 3174 \begin{figure}[htb]
505     \centering
506     \includegraphics[width=\linewidth]{rP2}
507     \caption{The $P_2$ order parameter as a funtion of the ratio of
508     $\sigma_h$ to $\sigma_0$.\label{fig:rP2}}
509     \end{figure}
510 xsun 3147
511 xsun 3174 We studied the effects of the interactions between head groups on the
512     structure of lipid bilayer by changing the strength of the dipole.
513     Figure \ref{fig:sP2} shows how the $P_2$ order parameter changes with
514     increasing strength of the dipole. Generally the dipoles on the head
515     group are more ordered by increase in the strength of the interaction
516     between heads and are more disordered by decreasing the interaction
517     stength. When the interaction between the heads is weak enough, the
518     bilayer structure does not persist; all lipid molecules are solvated
519     directly in the water. The critial value of the strength of the dipole
520     depends on the head size. The perfectly flat surface melts at $5$
521 xsun 3182 $0.03$ debye, the asymmetric rippled surfaces melt at $8$ $0.04$
522     $0.03$ debye, the symmetric rippled surfaces melt at $10$ $0.04$
523     debye. The ordering of the tails is the same as the ordering of the
524     dipoles except for the flat phase. Since the surface is already
525     perfect flat, the order parameter does not change much until the
526     strength of the dipole is $15$ debye. However, the order parameter
527     decreases quickly when the strength of the dipole is further
528     increased. The head groups of the lipid molecules are brought closer
529     by stronger interactions between them. For a flat surface, a large
530     amount of free volume between the head groups is available, but when
531     the head groups are brought closer, the tails will splay outward,
532     forming an inverse micelle. When $\sigma_h=1.28\sigma_0$, the $P_2$
533     order parameter decreases slightly after the strength of the dipole is
534     increased to $16$ debye. For rippled surfaces, there is less free
535     volume available between the head groups. Therefore there is little
536     effect on the structure of the membrane due to increasing dipolar
537     strength. However, the increase of the $P_2$ order parameter implies
538     the membranes are flatten by the increase of the strength of the
539     dipole. Unlike other systems that melt directly when the interaction
540     is weak enough, for $\sigma_h=1.41\sigma_0$, part of the membrane
541     melts into itself first. The upper leaf of the bilayer becomes totally
542     interdigitated with the lower leaf. This is different behavior than
543     what is exhibited with the interdigitated lines in the rippled phase
544     where only one interdigitated line connects the two leaves of bilayer.
545 xsun 3174 \begin{figure}[htb]
546     \centering
547     \includegraphics[width=\linewidth]{sP2}
548     \caption{The $P_2$ order parameter as a funtion of the strength of the
549     dipole.\label{fig:sP2}}
550     \end{figure}
551 xsun 3147
552 xsun 3174 Figure \ref{fig:tP2} shows the dependence of the order parameter on
553     temperature. The behavior of the $P_2$ order paramter is
554     straightforward. Systems are more ordered at low temperature, and more
555     disordered at high temperatures. When the temperature is high enough,
556 xsun 3182 the membranes are instable. For flat surfaces ($\sigma_h=1.20\sigma_0$
557     and $\sigma_h=1.28\sigma_0$), when the temperature is increased to
558     $310$, the $P_2$ order parameter increases slightly instead of
559     decreases like ripple surface. This is an evidence of the frustration
560     of the dipolar ordering in each leaf of the lipid bilayer, at low
561     temperature, the systems are locked in a local minimum energy state,
562     with increase of the temperature, the system can jump out the local
563     energy well to find the lower energy state which is the longer range
564     orientational ordering. Like the dipolar ordering of the flat
565     surfaces, the ordering of the tails of the lipid molecules for ripple
566     membranes ($\sigma_h=1.35\sigma_0$ and $\sigma_h=1.41\sigma_0$) also
567     show some nonthermal characteristic. With increase of the temperature,
568     the $P_2$ order parameter decreases firstly, and increases afterward
569     when the temperature is greater than $290 K$. The increase of the
570     $P_2$ order parameter indicates a more ordered structure for the tails
571     of the lipid molecules which corresponds to a more flat surface. Since
572     our model lacks the detailed information on lipid tails, we can not
573     simulate the fluid phase with melted fatty acid chains. Moreover, the
574     formation of the tilted $L_{\beta'}$ phase also depends on the
575     organization of fatty groups on tails.
576 xsun 3174 \begin{figure}[htb]
577     \centering
578     \includegraphics[width=\linewidth]{tP2}
579     \caption{The $P_2$ order parameter as a funtion of
580     temperature.\label{fig:tP2}}
581     \end{figure}
582 xsun 3147
583 xsun 3174 \section{Discussion}
584     \label{sec:discussion}
585 xsun 3147
586 gezelter 3199 \newpage
587 xsun 3147 \bibliography{mdripple}
588     \end{document}