ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
Revision: 3200
Committed: Fri Jul 27 21:59:45 2007 UTC (16 years, 11 months ago) by gezelter
Content type: application/x-tex
File size: 31166 byte(s)
Log Message:
friday 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 gezelter 3200 A sketch of the various structural elements of our molecular-scale
275     lipid / solvent model is shown in figure \ref{fig:lipidModel}. The
276     actual parameters used in our simulations are given in table
277     \ref{tab:parameters}.
278 gezelter 3195
279 gezelter 3199 \begin{figure}[htb]
280     \centering
281     \includegraphics[width=4in]{2lipidModel}
282     \caption{The parameters defining the behavior of the lipid
283     models. $l / d$ is the ratio of the head group to body diameter.
284     Molecular bodies had a fixed aspect ratio of 3.0. The solvent model
285     was a simplified 4-water bead ($\sigma_w \approx d$) that has been
286     used in other coarse-grained (DPD) simulations. The dipolar strength
287     (and the temperature and pressure) were the only other parameters that
288     were varied systematically.\label{fig:lipidModel}}
289     \end{figure}
290 gezelter 3195
291     To take into account the permanent dipolar interactions of the
292     zwitterionic head groups, we place fixed dipole moments $\mu_{i}$ at
293 gezelter 3199 one end of the Gay-Berne particles. The dipoles are oriented at an
294     angle $\theta = \pi / 2$ relative to the major axis. These dipoles
295 gezelter 3195 are protected by a head ``bead'' with a range parameter which we have
296 gezelter 3199 varied between $1.20 d$ and $1.41 d$. The head groups interact with
297     each other using a combination of Lennard-Jones,
298 xsun 3174 \begin{eqnarray*}
299 gezelter 3200 V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
300 gezelter 3195 \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
301 xsun 3174 \end{eqnarray*}
302 gezelter 3199 and dipole-dipole,
303 xsun 3174 \begin{eqnarray*}
304 gezelter 3200 V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
305     \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
306 gezelter 3195 \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
307     \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
308 xsun 3174 \end{eqnarray*}
309 gezelter 3195 potentials.
310     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
311     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
312 gezelter 3199 pointing along the inter-dipole vector $\mathbf{r}_{ij}$.
313 gezelter 3195
314     For the interaction between nonequivalent uniaxial ellipsoids (in this
315 gezelter 3199 case, between spheres and ellipsoids), the spheres are treated as
316     ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
317 gezelter 3200 ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of
318     the Gay-Berne potential we are using was generalized by Cleaver {\it
319     et al.} and is appropriate for dissimilar uniaxial
320     ellipsoids.\cite{Cleaver96}
321 xsun 3147
322 gezelter 3199 The solvent model in our simulations is identical to one used by
323     Marrink {\it et al.} in their dissipative particle dynamics (DPD)
324     simulation of lipid bilayers.\cite{Marrink04} This solvent bead is a single
325     site that represents four water molecules (m = 72 amu) and has
326     comparable density and diffusive behavior to liquid water. However,
327     since there are no electrostatic sites on these beads, this solvent
328     model cannot replicate the dielectric properties of water.
329 xsun 3198 \begin{table*}
330     \begin{minipage}{\linewidth}
331     \begin{center}
332 gezelter 3199 \caption{Potential parameters used for molecular-scale coarse-grained
333     lipid simulations}
334     \begin{tabular}{llccc}
335 xsun 3198 \hline
336 gezelter 3199 & & Head & Chain & Solvent \\
337 xsun 3198 \hline
338 gezelter 3200 $d$ (\AA) & & varied & 4.6 & 4.7 \\
339     $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
340 gezelter 3199 $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
341     $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 \\
342 gezelter 3200 $m$ (amu) & & 196 & 760 & 72.06 \\
343 gezelter 3199 $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
344     \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
345     \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
346     \multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\
347     $\mu$ (Debye) & & varied & 0 & 0 \\
348 xsun 3198 \end{tabular}
349     \label{tab:parameters}
350     \end{center}
351     \end{minipage}
352     \end{table*}
353 gezelter 3195
354 gezelter 3199 A switching function has been applied to all potentials to smoothly
355     turn off the interactions between a range of $22$ and $25$ \AA.
356 gezelter 3186
357 gezelter 3200 The parameters that were systematically varied in this study were the
358     size of the head group ($\sigma_h$), the strength of the dipole moment
359     ($\mu$), and the temperature of the system. Values for $\sigma_h$
360     ranged from 5.5 \AA\ to 6.5 \AA\ . If the width of the tails is
361     taken to be the unit of length, these head groups correspond to a
362     range from $1.2 d$ to $1.41 d$. Since the solvent beads are nearly
363     identical in diameter to the tail ellipsoids, all distances that
364     follow will be measured relative to this unit of distance.
365    
366 gezelter 3196 \section{Experimental Methodology}
367 xsun 3174 \label{sec:experiment}
368 xsun 3147
369 gezelter 3196 To create unbiased bilayers, all simulations were started from two
370 gezelter 3200 perfectly flat monolayers separated by a 26 \AA\ gap between the
371 gezelter 3196 molecular bodies of the upper and lower leaves. The separated
372     monolayers were evolved in a vaccum with $x-y$ anisotropic pressure
373 xsun 3174 coupling. The length of $z$ axis of the simulations was fixed and a
374     constant surface tension was applied to enable real fluctuations of
375 gezelter 3200 the bilayer. Periodic boundary conditions were used, and $480-720$
376     lipid molecules were present in the simulations, depending on the size
377     of the head beads. In all cases, the two monolayers spontaneously
378     collapsed into bilayer structures within 100 ps. Following this
379     collapse, all systems were equlibrated for $100$ ns at $300$ K.
380 xsun 3147
381 gezelter 3200 The resulting bilayer structures were then solvated at a ratio of $6$
382 gezelter 3196 solvent beads (24 water molecules) per lipid. These configurations
383 gezelter 3200 were then equilibrated for another $30$ ns. All simulations utilizing
384     the solvent were carried out at constant pressure ($P=1$ atm) with
385     $3$D anisotropic coupling, and constant surface tension
386     ($\gamma=0.015$ UNIT). Given the absence of fast degrees of freedom in
387     this model, a timestep of $50$ fs was utilized with excellent energy
388     conservation. Data collection for structural properties of the
389     bilayers was carried out during a final 5 ns run following the solvent
390     equilibration. All simulations were performed using the OOPSE
391     molecular modeling program.\cite{Meineke05}
392 gezelter 3196
393     \section{Results}
394 xsun 3174 \label{sec:results}
395 xsun 3147
396 xsun 3174 Snapshots in Figure \ref{fig:phaseCartoon} show that the membrane is
397 gezelter 3200 more corrugated with increasing size of the head groups. The surface
398     is nearly flat when $\sigma_h=1.20 d$. With $\sigma_h=1.28 d$,
399     although the surface is still flat, the bilayer starts to splay
400     inward; the upper leaf of the bilayer is connected to the lower leaf
401     with an interdigitated line defect. Two periodicities with $100$ \AA\
402     wavelengths were observed in the simulation. This structure is very
403     similiar to the structure observed by de Vries and Lenz {\it et
404     al.}. The same basic structure is also observed when $\sigma_h=1.41
405     d$, but the wavelength of the surface corrugations depends sensitively
406     on the size of the ``head'' beads. From the undulation spectrum, the
407     corrugation is clearly non-thermal.
408 xsun 3174 \begin{figure}[htb]
409     \centering
410 gezelter 3199 \includegraphics[width=4in]{phaseCartoon}
411 xsun 3174 \caption{A sketch to discribe the structure of the phases observed in
412     our simulations.\label{fig:phaseCartoon}}
413     \end{figure}
414 xsun 3147
415 gezelter 3200 When $\sigma_h=1.35 d$, we observed another corrugated surface
416     morphology. This structure is different from the asymmetric rippled
417 xsun 3174 surface; there is no interdigitation between the upper and lower
418     leaves of the bilayer. Each leaf of the bilayer is broken into several
419     hemicylinderical sections, and opposite leaves are fitted together
420     much like roof tiles. Unlike the surface in which the upper
421     hemicylinder is always interdigitated on the leading or trailing edge
422 gezelter 3200 of lower hemicylinder, this ``symmetric'' ripple has no prefered
423     direction. The corresponding structures are shown in Figure
424 xsun 3174 \ref{fig:phaseCartoon} for elucidation of the detailed structures of
425 gezelter 3200 different phases. The top panel in figure \ref{fig:phaseCartoon} is
426     the flat phase, the middle panel shows the asymmetric ripple phase
427     corresponding to $\sigma_h = 1.41 d$ and the lower panel shows the
428     symmetric ripple phase observed when $\sigma_h=1.35 d$. In the
429     symmetric ripple, the bilayer is continuous over the whole membrane,
430     however, in asymmetric ripple phase, the bilayer domains are connected
431     by thin interdigitated monolayers that share molecules between the
432     upper and lower leaves.
433 xsun 3174 \begin{table*}
434     \begin{minipage}{\linewidth}
435     \begin{center}
436 gezelter 3200 \caption{Phases, ripple wavelengths and amplitudes observed as a
437     function of the ratio between the head beads and the diameters of the
438     tails. All lengths are normalized to the diameter of the tail
439     ellipsoids.}
440 xsun 3174 \begin{tabular}{lccc}
441     \hline
442 gezelter 3200 $\sigma_h / d$ & type of phase & $\lambda / d$ & $A / d$\\
443 xsun 3174 \hline
444     1.20 & flat & N/A & N/A \\
445 gezelter 3200 1.28 & asymmetric ripple or flat & 21.7 & N/A \\
446 xsun 3174 1.35 & symmetric ripple & 17.2 & 2.2 \\
447     1.41 & asymmetric ripple & 15.4 & 1.5 \\
448     \end{tabular}
449     \label{tab:property}
450     \end{center}
451     \end{minipage}
452     \end{table*}
453 xsun 3147
454 gezelter 3200 The membrane structures and the reduced wavelength $\lambda / d$,
455     reduced amplitude $A / d$ of the ripples are summarized in Table
456     \ref{tab:property}. The wavelength range is $15~21$ molecular bodies
457     and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
458     $2.2$ for symmetric ripple. These values are consistent to the
459     experimental results. Note, that given the lack of structural freedom
460     in the tails of our model lipids, the amplitudes observed from these
461     simulations are likely to underestimate of the true amplitudes.
462 xsun 3174
463 gezelter 3195 \begin{figure}[htb]
464     \centering
465 gezelter 3199 \includegraphics[width=4in]{topDown}
466 gezelter 3195 \caption{Top views of the flat (upper), asymmetric ripple (middle),
467     and symmetric ripple (lower) phases. Note that the head-group dipoles
468     have formed head-to-tail chains in all three of these phases, but in
469     the two rippled phases, the dipolar chains are all aligned
470     {\it perpendicular} to the direction of the ripple. The flat membrane
471     has multiple point defects in the dipolar orientational ordering, and
472     the dipolar ordering on the lower leaf of the bilayer can be in a
473     different direction from the upper leaf.\label{fig:topView}}
474     \end{figure}
475    
476 gezelter 3200 Figure \ref{fig:topView} shows snapshots of bird's-eye views of the
477     flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.41, 1.35 d$)
478     bilayers. The directions of the dipoles on the head groups are
479     represented with two colored half spheres: blue (phosphate) and yellow
480     (amino). For flat bilayers, the system exhibits signs of
481     orientational frustration; some disorder in the dipolar chains is
482     evident with kinks visible at the edges between different ordered
483     domains. The lipids can also move independently of lipids in the
484     opposing leaf, so the ordering of the dipoles on one leaf is not
485     necessarily consistant with the ordering on the other. These two
486     factors keep the total dipolar order parameter relatively low for the
487     flat phases.
488 xsun 3147
489 gezelter 3200 With increasing head group size, the surface becomes corrugated, and
490     the dipoles cannot move as freely on the surface. Therefore, the
491     translational freedom of lipids in one layer is dependent upon the
492     position of lipids in the other layer. As a result, the ordering of
493     the dipoles on head groups in one leaf is correlated with the ordering
494     in the other leaf. Furthermore, as the membrane deforms due to the
495     corrugation, the symmetry of the allowed dipolar ordering on each leaf
496     is broken. The dipoles then self-assemble in a head-to-tail
497     configuration, and the dipolar order parameter increases dramatically.
498     However, the total polarization of the system is still close to zero.
499     This is strong evidence that the corrugated structure is an
500     antiferroelectric state. It is also notable that the head-to-tail
501     arrangement of the dipoles is in a direction perpendicular to the wave
502     vector for the surface corrugation. This is a similar finding to what
503     we observed in our earlier work on the elastic dipolar
504     membranes.\cite{Sun07}
505    
506     The $P_2$ order parameters (for both the molecular bodies and the head
507     group dipoles) have been calculated to quantify the ordering in these
508     phases. $P_2 = 1$ implies a perfectly ordered structure, and $P_2 = 0$
509     implies complete orientational randomization. Figure \ref{fig:rP2}
510     shows the $P_2$ order parameter for the head-group dipoles increasing
511     with increasing head group size. When the heads of the lipid molecules
512     are small, the membrane is nearly flat. The dipolar ordering exhibits
513     frustrated orientational ordering in this circumstance.
514    
515 xsun 3174 The ordering of the tails is essentially opposite to the ordering of
516 gezelter 3200 the dipoles on head group. The $P_2$ order parameter {\it decreases}
517     with increasing head size. This indicates that the surface is more
518     curved with larger head / tail size ratios. When the surface is flat,
519     all tails are pointing in the same direction (parallel to the normal
520     of the surface). This simplified model appears to be exhibiting a
521     smectic A fluid phase, similar to the real $L_{\beta}$ phase. We have
522     not observed a smectic C gel phase ($L_{\beta'}$) for this model
523     system. Increasing the size of the heads, results in rapidly
524     decreasing $P_2$ ordering for the molecular bodies.
525 gezelter 3199
526 xsun 3174 \begin{figure}[htb]
527     \centering
528     \includegraphics[width=\linewidth]{rP2}
529 gezelter 3200 \caption{The $P_2$ order parameter as a function of the ratio of
530     $\sigma_h$ to $d$. \label{fig:rP2}}
531 xsun 3174 \end{figure}
532 xsun 3147
533 xsun 3174 We studied the effects of the interactions between head groups on the
534     structure of lipid bilayer by changing the strength of the dipole.
535     Figure \ref{fig:sP2} shows how the $P_2$ order parameter changes with
536     increasing strength of the dipole. Generally the dipoles on the head
537     group are more ordered by increase in the strength of the interaction
538     between heads and are more disordered by decreasing the interaction
539     stength. When the interaction between the heads is weak enough, the
540     bilayer structure does not persist; all lipid molecules are solvated
541     directly in the water. The critial value of the strength of the dipole
542     depends on the head size. The perfectly flat surface melts at $5$
543 xsun 3182 $0.03$ debye, the asymmetric rippled surfaces melt at $8$ $0.04$
544     $0.03$ debye, the symmetric rippled surfaces melt at $10$ $0.04$
545     debye. The ordering of the tails is the same as the ordering of the
546     dipoles except for the flat phase. Since the surface is already
547     perfect flat, the order parameter does not change much until the
548     strength of the dipole is $15$ debye. However, the order parameter
549     decreases quickly when the strength of the dipole is further
550     increased. The head groups of the lipid molecules are brought closer
551     by stronger interactions between them. For a flat surface, a large
552     amount of free volume between the head groups is available, but when
553     the head groups are brought closer, the tails will splay outward,
554     forming an inverse micelle. When $\sigma_h=1.28\sigma_0$, the $P_2$
555     order parameter decreases slightly after the strength of the dipole is
556     increased to $16$ debye. For rippled surfaces, there is less free
557     volume available between the head groups. Therefore there is little
558     effect on the structure of the membrane due to increasing dipolar
559     strength. However, the increase of the $P_2$ order parameter implies
560     the membranes are flatten by the increase of the strength of the
561     dipole. Unlike other systems that melt directly when the interaction
562     is weak enough, for $\sigma_h=1.41\sigma_0$, part of the membrane
563     melts into itself first. The upper leaf of the bilayer becomes totally
564     interdigitated with the lower leaf. This is different behavior than
565     what is exhibited with the interdigitated lines in the rippled phase
566     where only one interdigitated line connects the two leaves of bilayer.
567 xsun 3174 \begin{figure}[htb]
568     \centering
569     \includegraphics[width=\linewidth]{sP2}
570     \caption{The $P_2$ order parameter as a funtion of the strength of the
571     dipole.\label{fig:sP2}}
572     \end{figure}
573 xsun 3147
574 xsun 3174 Figure \ref{fig:tP2} shows the dependence of the order parameter on
575     temperature. The behavior of the $P_2$ order paramter is
576     straightforward. Systems are more ordered at low temperature, and more
577     disordered at high temperatures. When the temperature is high enough,
578 xsun 3182 the membranes are instable. For flat surfaces ($\sigma_h=1.20\sigma_0$
579     and $\sigma_h=1.28\sigma_0$), when the temperature is increased to
580     $310$, the $P_2$ order parameter increases slightly instead of
581     decreases like ripple surface. This is an evidence of the frustration
582     of the dipolar ordering in each leaf of the lipid bilayer, at low
583     temperature, the systems are locked in a local minimum energy state,
584     with increase of the temperature, the system can jump out the local
585     energy well to find the lower energy state which is the longer range
586     orientational ordering. Like the dipolar ordering of the flat
587     surfaces, the ordering of the tails of the lipid molecules for ripple
588     membranes ($\sigma_h=1.35\sigma_0$ and $\sigma_h=1.41\sigma_0$) also
589     show some nonthermal characteristic. With increase of the temperature,
590     the $P_2$ order parameter decreases firstly, and increases afterward
591     when the temperature is greater than $290 K$. The increase of the
592     $P_2$ order parameter indicates a more ordered structure for the tails
593     of the lipid molecules which corresponds to a more flat surface. Since
594     our model lacks the detailed information on lipid tails, we can not
595     simulate the fluid phase with melted fatty acid chains. Moreover, the
596     formation of the tilted $L_{\beta'}$ phase also depends on the
597     organization of fatty groups on tails.
598 xsun 3174 \begin{figure}[htb]
599     \centering
600     \includegraphics[width=\linewidth]{tP2}
601     \caption{The $P_2$ order parameter as a funtion of
602     temperature.\label{fig:tP2}}
603     \end{figure}
604 xsun 3147
605 xsun 3174 \section{Discussion}
606     \label{sec:discussion}
607 xsun 3147
608 gezelter 3199 \newpage
609 xsun 3147 \bibliography{mdripple}
610     \end{document}