ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
Revision: 3264
Committed: Thu Oct 18 21:07:03 2007 UTC (16 years, 10 months ago) by gezelter
Content type: application/x-tex
File size: 43345 byte(s)
Log Message:
Added table

File Contents

# User Rev Content
1 xsun 3147 %\documentclass[aps,pre,twocolumn,amssymb,showpacs,floatfix]{revtex4}
2 gezelter 3202 %\documentclass[aps,pre,preprint,amssymb]{revtex4}
3     \documentclass[12pt]{article}
4     \usepackage{times}
5     \usepackage{mathptm}
6     \usepackage{tabularx}
7     \usepackage{setspace}
8 gezelter 3199 \usepackage{amsmath}
9     \usepackage{amssymb}
10 xsun 3147 \usepackage{graphicx}
11 gezelter 3202 \usepackage[ref]{overcite}
12     \pagestyle{plain}
13     \pagenumbering{arabic}
14     \oddsidemargin 0.0cm \evensidemargin 0.0cm
15     \topmargin -21pt \headsep 10pt
16     \textheight 9.0in \textwidth 6.5in
17     \brokenpenalty=10000
18     \renewcommand{\baselinestretch}{1.2}
19     \renewcommand\citemid{\ } % no comma in optional reference note
20 xsun 3147
21     \begin{document}
22 gezelter 3202 %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
23     %\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
24 xsun 3147
25 gezelter 3202 \bibliographystyle{achemso}
26 xsun 3147
27 gezelter 3203 \title{Dipolar ordering in the ripple phases of molecular-scale models
28     of lipid membranes}
29 gezelter 3202 \author{Xiuquan Sun and J. Daniel Gezelter \\
30     Department of Chemistry and Biochemistry,\\
31 gezelter 3199 University of Notre Dame, \\
32 xsun 3147 Notre Dame, Indiana 46556}
33    
34 gezelter 3202 %\email[E-mail:]{gezelter@nd.edu}
35    
36 xsun 3147 \date{\today}
37    
38 gezelter 3202 \maketitle
39    
40 xsun 3147 \begin{abstract}
41 gezelter 3203 Symmetric and asymmetric ripple phases have been observed to form in
42     molecular dynamics simulations of a simple molecular-scale lipid
43     model. The lipid model consists of an dipolar head group and an
44     ellipsoidal tail. Within the limits of this model, an explanation for
45     generalized membrane curvature is a simple mismatch in the size of the
46     heads with the width of the molecular bodies. The persistence of a
47     {\it bilayer} structure requires strong attractive forces between the
48     head groups. One feature of this model is that an energetically
49     favorable orientational ordering of the dipoles can be achieved by
50     out-of-plane membrane corrugation. The corrugation of the surface
51 gezelter 3204 stabilizes the long range orientational ordering for the dipoles in the
52     head groups which then adopt a bulk anti-ferroelectric state. We
53 gezelter 3203 observe a common feature of the corrugated dipolar membranes: the wave
54     vectors for the surface ripples are always found to be perpendicular
55     to the dipole director axis.
56 xsun 3147 \end{abstract}
57    
58 gezelter 3202 %\maketitle
59 gezelter 3203 \newpage
60 xsun 3147
61 xsun 3174 \section{Introduction}
62     \label{sec:Int}
63 gezelter 3195 Fully hydrated lipids will aggregate spontaneously to form bilayers
64     which exhibit a variety of phases depending on their temperatures and
65     compositions. Among these phases, a periodic rippled phase
66     ($P_{\beta'}$) appears as an intermediate phase between the gel
67     ($L_\beta$) and fluid ($L_{\alpha}$) phases for relatively pure
68     phosphatidylcholine (PC) bilayers. The ripple phase has attracted
69     substantial experimental interest over the past 30 years. Most
70     structural information of the ripple phase has been obtained by the
71     X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
72     microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
73     et al.} used atomic force microscopy (AFM) to observe ripple phase
74     morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
75     experimental results provide strong support for a 2-dimensional
76     hexagonal packing lattice of the lipid molecules within the ripple
77     phase. This is a notable change from the observed lipid packing
78 gezelter 3204 within the gel phase.~\cite{Cevc87} The X-ray diffraction work by
79     Katsaras {\it et al.} showed that a rich phase diagram exhibiting both
80     {\it asymmetric} and {\it symmetric} ripples is possible for lecithin
81     bilayers.\cite{Katsaras00}
82 xsun 3174
83 gezelter 3195 A number of theoretical models have been presented to explain the
84     formation of the ripple phase. Marder {\it et al.} used a
85 gezelter 3204 curvature-dependent Landau-de~Gennes free-energy functional to predict
86 gezelter 3195 a rippled phase.~\cite{Marder84} This model and other related continuum
87     models predict higher fluidity in convex regions and that concave
88     portions of the membrane correspond to more solid-like regions.
89     Carlson and Sethna used a packing-competition model (in which head
90     groups and chains have competing packing energetics) to predict the
91     formation of a ripple-like phase. Their model predicted that the
92     high-curvature portions have lower-chain packing and correspond to
93     more fluid-like regions. Goldstein and Leibler used a mean-field
94     approach with a planar model for {\em inter-lamellar} interactions to
95     predict rippling in multilamellar phases.~\cite{Goldstein88} McCullough
96     and Scott proposed that the {\em anisotropy of the nearest-neighbor
97     interactions} coupled to hydrophobic constraining forces which
98     restrict height differences between nearest neighbors is the origin of
99     the ripple phase.~\cite{McCullough90} Lubensky and MacKintosh
100     introduced a Landau theory for tilt order and curvature of a single
101     membrane and concluded that {\em coupling of molecular tilt to membrane
102     curvature} is responsible for the production of
103     ripples.~\cite{Lubensky93} Misbah, Duplat and Houchmandzadeh proposed
104     that {\em inter-layer dipolar interactions} can lead to ripple
105     instabilities.~\cite{Misbah98} Heimburg presented a {\em coexistence
106     model} for ripple formation in which he postulates that fluid-phase
107     line defects cause sharp curvature between relatively flat gel-phase
108     regions.~\cite{Heimburg00} Kubica has suggested that a lattice model of
109     polar head groups could be valuable in trying to understand bilayer
110     phase formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations
111 gezelter 3204 of lamellar stacks of hexagonal lattices to show that large head groups
112 gezelter 3195 and molecular tilt with respect to the membrane normal vector can
113     cause bulk rippling.~\cite{Bannerjee02}
114 xsun 3174
115 gezelter 3204 In contrast, few large-scale molecular modeling studies have been
116 gezelter 3195 done due to the large size of the resulting structures and the time
117     required for the phases of interest to develop. With all-atom (and
118     even unified-atom) simulations, only one period of the ripple can be
119 gezelter 3204 observed and only for time scales in the range of 10-100 ns. One of
120     the most interesting molecular simulations was carried out by de~Vries
121 gezelter 3195 {\it et al.}~\cite{deVries05}. According to their simulation results,
122     the ripple consists of two domains, one resembling the gel bilayer,
123     while in the other, the two leaves of the bilayer are fully
124     interdigitated. The mechanism for the formation of the ripple phase
125     suggested by their work is a packing competition between the head
126     groups and the tails of the lipid molecules.\cite{Carlson87} Recently,
127 gezelter 3199 the ripple phase has also been studied by Lenz and Schmid using Monte
128 gezelter 3195 Carlo simulations.\cite{Lenz07} Their structures are similar to the De
129     Vries {\it et al.} structures except that the connection between the
130     two leaves of the bilayer is a narrow interdigitated line instead of
131     the fully interdigitated domain. The symmetric ripple phase was also
132     observed by Lenz {\it et al.}, and their work supports other claims
133     that the mismatch between the size of the head group and tail of the
134     lipid molecules is the driving force for the formation of the ripple
135     phase. Ayton and Voth have found significant undulations in
136     zero-surface-tension states of membranes simulated via dissipative
137     particle dynamics, but their results are consistent with purely
138     thermal undulations.~\cite{Ayton02}
139 xsun 3174
140 gezelter 3195 Although the organization of the tails of lipid molecules are
141     addressed by these molecular simulations and the packing competition
142 gezelter 3204 between head groups and tails is strongly implicated as the primary
143 gezelter 3195 driving force for ripple formation, questions about the ordering of
144 gezelter 3203 the head groups in ripple phase have not been settled.
145 xsun 3174
146 gezelter 3195 In a recent paper, we presented a simple ``web of dipoles'' spin
147     lattice model which provides some physical insight into relationship
148     between dipolar ordering and membrane buckling.\cite{Sun2007} We found
149     that dipolar elastic membranes can spontaneously buckle, forming
150 gezelter 3203 ripple-like topologies. The driving force for the buckling of dipolar
151 gezelter 3204 elastic membranes is the anti-ferroelectric ordering of the dipoles.
152 gezelter 3203 This was evident in the ordering of the dipole director axis
153 gezelter 3204 perpendicular to the wave vector of the surface ripples. A similar
154 gezelter 3195 phenomenon has also been observed by Tsonchev {\it et al.} in their
155 gezelter 3199 work on the spontaneous formation of dipolar peptide chains into
156     curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
157 gezelter 3195
158     In this paper, we construct a somewhat more realistic molecular-scale
159     lipid model than our previous ``web of dipoles'' and use molecular
160     dynamics simulations to elucidate the role of the head group dipoles
161     in the formation and morphology of the ripple phase. We describe our
162     model and computational methodology in section \ref{sec:method}.
163     Details on the simulations are presented in section
164     \ref{sec:experiment}, with results following in section
165     \ref{sec:results}. A final discussion of the role of dipolar heads in
166     the ripple formation can be found in section
167 xsun 3174 \ref{sec:discussion}.
168    
169 gezelter 3196 \section{Computational Model}
170 xsun 3174 \label{sec:method}
171    
172 gezelter 3199 \begin{figure}[htb]
173     \centering
174     \includegraphics[width=4in]{lipidModels}
175     \caption{Three different representations of DPPC lipid molecules,
176     including the chemical structure, an atomistic model, and the
177     head-body ellipsoidal coarse-grained model used in this
178     work.\label{fig:lipidModels}}
179     \end{figure}
180    
181 gezelter 3195 Our simple molecular-scale lipid model for studying the ripple phase
182     is based on two facts: one is that the most essential feature of lipid
183     molecules is their amphiphilic structure with polar head groups and
184     non-polar tails. Another fact is that the majority of lipid molecules
185     in the ripple phase are relatively rigid (i.e. gel-like) which makes
186     some fraction of the details of the chain dynamics negligible. Figure
187 gezelter 3204 \ref{fig:lipidModels} shows the molecular structure of a DPPC
188 gezelter 3195 molecule, as well as atomistic and molecular-scale representations of
189     a DPPC molecule. The hydrophilic character of the head group is
190     largely due to the separation of charge between the nitrogen and
191     phosphate groups. The zwitterionic nature of the PC headgroups leads
192     to abnormally large dipole moments (as high as 20.6 D), and this
193     strongly polar head group interacts strongly with the solvating water
194     layers immediately surrounding the membrane. The hydrophobic tail
195     consists of fatty acid chains. In our molecular scale model, lipid
196     molecules have been reduced to these essential features; the fatty
197     acid chains are represented by an ellipsoid with a dipolar ball
198     perched on one end to represent the effects of the charge-separated
199     head group. In real PC lipids, the direction of the dipole is
200     nearly perpendicular to the tail, so we have fixed the direction of
201     the point dipole rigidly in this orientation.
202 xsun 3147
203 gezelter 3195 The ellipsoidal portions of the model interact via the Gay-Berne
204     potential which has seen widespread use in the liquid crystal
205 gezelter 3199 community. Ayton and Voth have also used Gay-Berne ellipsoids for
206 gezelter 3204 modeling large length-scale properties of lipid
207 gezelter 3199 bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
208     was a single site model for the interactions of rigid ellipsoidal
209 gezelter 3195 molecules.\cite{Gay81} It can be thought of as a modification of the
210     Gaussian overlap model originally described by Berne and
211     Pechukas.\cite{Berne72} The potential is constructed in the familiar
212     form of the Lennard-Jones function using orientation-dependent
213     $\sigma$ and $\epsilon$ parameters,
214 gezelter 3202 \begin{equation*}
215 xsun 3174 V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
216     r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
217     {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
218     {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
219     -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
220     {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
221 gezelter 3195 \label{eq:gb}
222 gezelter 3202 \end{equation*}
223 gezelter 3195
224     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
225 gezelter 3199 \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
226     \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
227 gezelter 3195 are dependent on the relative orientations of the two molecules (${\bf
228     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
229 gezelter 3199 intermolecular separation (${\bf \hat{r}}_{ij}$). $\sigma$ and
230     $\sigma_0$ are also governed by shape mixing and anisotropy variables,
231 gezelter 3202 \begin {eqnarray*}
232 gezelter 3199 \sigma_0 & = & \sqrt{d_i^2 + d_j^2} \\ \\
233     \chi & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 -
234     d_j^2 \right)}{\left( l_j^2 + d_i^2 \right) \left(l_i^2 +
235     d_j^2 \right)}\right]^{1/2} \\ \\
236     \alpha^2 & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 +
237     d_i^2 \right)}{\left( l_j^2 - d_j^2 \right) \left(l_i^2 +
238     d_j^2 \right)}\right]^{1/2},
239 gezelter 3202 \end{eqnarray*}
240 gezelter 3199 where $l$ and $d$ describe the length and width of each uniaxial
241     ellipsoid. These shape anisotropy parameters can then be used to
242     calculate the range function,
243 gezelter 3202 \begin{equation*}
244 gezelter 3199 \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) = \sigma_{0}
245     \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
246     \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
247     \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
248     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
249     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
250     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
251     \right]^{-1/2}
252 gezelter 3202 \end{equation*}
253 gezelter 3199
254     Gay-Berne ellipsoids also have an energy scaling parameter,
255     $\epsilon^s$, which describes the well depth for two identical
256 gezelter 3204 ellipsoids in a {\it side-by-side} configuration. Additionally, a well
257 gezelter 3199 depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
258     the ratio between the well depths in the {\it end-to-end} and
259     side-by-side configurations. As in the range parameter, a set of
260     mixing and anisotropy variables can be used to describe the well
261     depths for dissimilar particles,
262     \begin {eqnarray*}
263     \epsilon_0 & = & \sqrt{\epsilon^s_i * \epsilon^s_j} \\ \\
264     \epsilon^r & = & \sqrt{\epsilon^r_i * \epsilon^r_j} \\ \\
265     \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
266     \\ \\
267     \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
268     \end{eqnarray*}
269     The form of the strength function is somewhat complicated,
270     \begin {eqnarray*}
271     \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
272     \epsilon_{0} \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
273     \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
274     \hat{r}}_{ij}) \\ \\
275     \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
276     \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
277     \hat{u}}_{j})^{2}\right]^{-1/2} \\ \\
278     \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) &
279     = &
280     1 - \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
281     \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
282     \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
283     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
284     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
285     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
286     \end {eqnarray*}
287     although many of the quantities and derivatives are identical with
288 gezelter 3202 those obtained for the range parameter. Ref. \citen{Luckhurst90}
289 gezelter 3199 has a particularly good explanation of the choice of the Gay-Berne
290     parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
291     excellent overview of the computational methods that can be used to
292     efficiently compute forces and torques for this potential can be found
293 gezelter 3202 in Ref. \citen{Golubkov06}
294 gezelter 3199
295     The choices of parameters we have used in this study correspond to a
296     shape anisotropy of 3 for the chain portion of the molecule. In
297     principle, this could be varied to allow for modeling of longer or
298     shorter chain lipid molecules. For these prolate ellipsoids, we have:
299 gezelter 3195 \begin{equation}
300     \begin{array}{rcl}
301 gezelter 3199 d & < & l \\
302     \epsilon^{r} & < & 1
303 gezelter 3195 \end{array}
304     \end{equation}
305 gezelter 3200 A sketch of the various structural elements of our molecular-scale
306     lipid / solvent model is shown in figure \ref{fig:lipidModel}. The
307     actual parameters used in our simulations are given in table
308     \ref{tab:parameters}.
309 gezelter 3195
310 gezelter 3199 \begin{figure}[htb]
311     \centering
312     \includegraphics[width=4in]{2lipidModel}
313     \caption{The parameters defining the behavior of the lipid
314     models. $l / d$ is the ratio of the head group to body diameter.
315     Molecular bodies had a fixed aspect ratio of 3.0. The solvent model
316     was a simplified 4-water bead ($\sigma_w \approx d$) that has been
317     used in other coarse-grained (DPD) simulations. The dipolar strength
318     (and the temperature and pressure) were the only other parameters that
319     were varied systematically.\label{fig:lipidModel}}
320     \end{figure}
321 gezelter 3195
322     To take into account the permanent dipolar interactions of the
323 gezelter 3203 zwitterionic head groups, we have placed fixed dipole moments $\mu_{i}$ at
324 gezelter 3199 one end of the Gay-Berne particles. The dipoles are oriented at an
325     angle $\theta = \pi / 2$ relative to the major axis. These dipoles
326 gezelter 3203 are protected by a head ``bead'' with a range parameter ($\sigma_h$) which we have
327 gezelter 3199 varied between $1.20 d$ and $1.41 d$. The head groups interact with
328     each other using a combination of Lennard-Jones,
329 gezelter 3202 \begin{equation}
330 gezelter 3200 V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
331 gezelter 3195 \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
332 gezelter 3202 \end{equation}
333 gezelter 3199 and dipole-dipole,
334 gezelter 3202 \begin{equation}
335 gezelter 3200 V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
336     \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
337 gezelter 3195 \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
338     \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
339 gezelter 3202 \end{equation}
340 gezelter 3195 potentials.
341     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
342     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
343 gezelter 3199 pointing along the inter-dipole vector $\mathbf{r}_{ij}$.
344 gezelter 3195
345     For the interaction between nonequivalent uniaxial ellipsoids (in this
346 gezelter 3199 case, between spheres and ellipsoids), the spheres are treated as
347     ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
348 gezelter 3200 ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of
349     the Gay-Berne potential we are using was generalized by Cleaver {\it
350     et al.} and is appropriate for dissimilar uniaxial
351     ellipsoids.\cite{Cleaver96}
352 xsun 3147
353 gezelter 3199 The solvent model in our simulations is identical to one used by
354     Marrink {\it et al.} in their dissipative particle dynamics (DPD)
355 gezelter 3203 simulation of lipid bilayers.\cite{Marrink04} This solvent bead is a
356     single site that represents four water molecules (m = 72 amu) and has
357 gezelter 3199 comparable density and diffusive behavior to liquid water. However,
358     since there are no electrostatic sites on these beads, this solvent
359 gezelter 3203 model cannot replicate the dielectric properties of water.
360    
361 xsun 3198 \begin{table*}
362     \begin{minipage}{\linewidth}
363     \begin{center}
364 gezelter 3199 \caption{Potential parameters used for molecular-scale coarse-grained
365     lipid simulations}
366     \begin{tabular}{llccc}
367 xsun 3198 \hline
368 gezelter 3199 & & Head & Chain & Solvent \\
369 xsun 3198 \hline
370 gezelter 3200 $d$ (\AA) & & varied & 4.6 & 4.7 \\
371     $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
372 gezelter 3199 $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
373     $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 \\
374 gezelter 3200 $m$ (amu) & & 196 & 760 & 72.06 \\
375 gezelter 3199 $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
376     \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
377     \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
378     \multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\
379     $\mu$ (Debye) & & varied & 0 & 0 \\
380 xsun 3198 \end{tabular}
381     \label{tab:parameters}
382     \end{center}
383     \end{minipage}
384     \end{table*}
385 gezelter 3195
386 gezelter 3203 \section{Experimental Methodology}
387     \label{sec:experiment}
388 gezelter 3186
389 gezelter 3200 The parameters that were systematically varied in this study were the
390     size of the head group ($\sigma_h$), the strength of the dipole moment
391     ($\mu$), and the temperature of the system. Values for $\sigma_h$
392 gezelter 3203 ranged from 5.5 \AA\ to 6.5 \AA\ . If the width of the tails is taken
393     to be the unit of length, these head groups correspond to a range from
394     $1.2 d$ to $1.41 d$. Since the solvent beads are nearly identical in
395     diameter to the tail ellipsoids, all distances that follow will be
396     measured relative to this unit of distance. Because the solvent we
397     are using is non-polar and has a dielectric constant of 1, values for
398     $\mu$ are sampled from a range that is somewhat smaller than the 20.6
399 gezelter 3204 Debye dipole moment of the PC head groups.
400 gezelter 3200
401 gezelter 3196 To create unbiased bilayers, all simulations were started from two
402 gezelter 3200 perfectly flat monolayers separated by a 26 \AA\ gap between the
403 gezelter 3196 molecular bodies of the upper and lower leaves. The separated
404 gezelter 3204 monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
405 xsun 3174 coupling. The length of $z$ axis of the simulations was fixed and a
406     constant surface tension was applied to enable real fluctuations of
407 gezelter 3200 the bilayer. Periodic boundary conditions were used, and $480-720$
408     lipid molecules were present in the simulations, depending on the size
409     of the head beads. In all cases, the two monolayers spontaneously
410     collapsed into bilayer structures within 100 ps. Following this
411 gezelter 3204 collapse, all systems were equilibrated for $100$ ns at $300$ K.
412 xsun 3147
413 gezelter 3200 The resulting bilayer structures were then solvated at a ratio of $6$
414 gezelter 3196 solvent beads (24 water molecules) per lipid. These configurations
415 gezelter 3200 were then equilibrated for another $30$ ns. All simulations utilizing
416     the solvent were carried out at constant pressure ($P=1$ atm) with
417     $3$D anisotropic coupling, and constant surface tension
418 gezelter 3203 ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
419 gezelter 3204 this model, a time step of $50$ fs was utilized with excellent energy
420 gezelter 3200 conservation. Data collection for structural properties of the
421     bilayers was carried out during a final 5 ns run following the solvent
422     equilibration. All simulations were performed using the OOPSE
423     molecular modeling program.\cite{Meineke05}
424 gezelter 3196
425 gezelter 3203 A switching function was applied to all potentials to smoothly turn
426     off the interactions between a range of $22$ and $25$ \AA.
427    
428 gezelter 3196 \section{Results}
429 xsun 3174 \label{sec:results}
430 xsun 3147
431 gezelter 3203 The membranes in our simulations exhibit a number of interesting
432     bilayer phases. The surface topology of these phases depends most
433     sensitively on the ratio of the size of the head groups to the width
434     of the molecular bodies. With heads only slightly larger than the
435 gezelter 3204 bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
436 gezelter 3203
437     Increasing the head / body size ratio increases the local membrane
438     curvature around each of the lipids. With $\sigma_h=1.28 d$, the
439     surface is still essentially flat, but the bilayer starts to exhibit
440     signs of instability. We have observed occasional defects where a
441     line of lipid molecules on one leaf of the bilayer will dip down to
442     interdigitate with the other leaf. This gives each of the two bilayer
443     leaves some local convexity near the line defect. These structures,
444     once developed in a simulation, are very stable and are spaced
445     approximately 100 \AA\ away from each other.
446    
447     With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
448     resolves into a ``symmetric'' ripple phase. Each leaf of the bilayer
449     is broken into several convex, hemicylinderical sections, and opposite
450     leaves are fitted together much like roof tiles. There is no
451     interdigitation between the upper and lower leaves of the bilayer.
452    
453     For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
454     local curvature is substantially larger, and the resulting bilayer
455     structure resolves into an asymmetric ripple phase. This structure is
456 gezelter 3204 very similar to the structures observed by both de~Vries {\it et al.}
457 gezelter 3203 and Lenz {\it et al.}. For a given ripple wave vector, there are two
458     possible asymmetric ripples, which is not the case for the symmetric
459     phase observed when $\sigma_h = 1.35 d$.
460    
461 xsun 3174 \begin{figure}[htb]
462     \centering
463 gezelter 3199 \includegraphics[width=4in]{phaseCartoon}
464 gezelter 3203 \caption{The role of the ratio between the head group size and the
465     width of the molecular bodies is to increase the local membrane
466     curvature. With strong attractive interactions between the head
467     groups, this local curvature can be maintained in bilayer structures
468     through surface corrugation. Shown above are three phases observed in
469     these simulations. With $\sigma_h = 1.20 d$, the bilayer maintains a
470     flat topology. For larger heads ($\sigma_h = 1.35 d$) the local
471     curvature resolves into a symmetrically rippled phase with little or
472     no interdigitation between the upper and lower leaves of the membrane.
473     The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
474     asymmetric rippled phases with interdigitation between the two
475     leaves.\label{fig:phaseCartoon}}
476 xsun 3174 \end{figure}
477 xsun 3147
478 gezelter 3203 Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
479     ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
480     phases are shown in Figure \ref{fig:phaseCartoon}.
481    
482 gezelter 3204 It is reasonable to ask how well the parameters we used can produce
483     bilayer properties that match experimentally known values for real
484     lipid bilayers. Using a value of $l = 13.8$ \AA for the ellipsoidal
485     tails and the fixed ellipsoidal aspect ratio of 3, our values for the
486     area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
487     entirely on the size of the head bead relative to the molecular body.
488     These values are tabulated in table \ref{tab:property}. Kucera {\it
489     et al.} have measured values for the head group spacings for a number
490     of PC lipid bilayers that range from 30.8 \AA (DLPC) to 37.8 (DPPC).
491     They have also measured values for the area per lipid that range from
492     60.6
493     \AA$^2$ (DMPC) to 64.2 \AA$^2$
494     (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
495     largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
496     bilayers (specifically the area per lipid) that resemble real PC
497     bilayers. The smaller head beads we used are perhaps better models
498     for PE head groups.
499    
500 xsun 3174 \begin{table*}
501     \begin{minipage}{\linewidth}
502     \begin{center}
503 gezelter 3204 \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
504     and amplitude observed as a function of the ratio between the head
505     beads and the diameters of the tails. Ripple wavelengths and
506     amplitudes are normalized to the diameter of the tail ellipsoids.}
507     \begin{tabular}{lccccc}
508 xsun 3174 \hline
509 gezelter 3204 $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
510     lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
511 xsun 3174 \hline
512 gezelter 3204 1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
513     1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
514     1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
515     1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
516 xsun 3174 \end{tabular}
517     \label{tab:property}
518     \end{center}
519     \end{minipage}
520     \end{table*}
521 xsun 3147
522 gezelter 3200 The membrane structures and the reduced wavelength $\lambda / d$,
523     reduced amplitude $A / d$ of the ripples are summarized in Table
524 gezelter 3203 \ref{tab:property}. The wavelength range is $15 - 17$ molecular bodies
525 gezelter 3200 and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
526 gezelter 3203 $2.2$ for symmetric ripple. These values are reasonably consistent
527     with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
528     Note, that given the lack of structural freedom in the tails of our
529     model lipids, the amplitudes observed from these simulations are
530     likely to underestimate of the true amplitudes.
531 xsun 3174
532 gezelter 3195 \begin{figure}[htb]
533     \centering
534 gezelter 3199 \includegraphics[width=4in]{topDown}
535 gezelter 3203 \caption{Top views of the flat (upper), symmetric ripple (middle),
536     and asymmetric ripple (lower) phases. Note that the head-group
537     dipoles have formed head-to-tail chains in all three of these phases,
538     but in the two rippled phases, the dipolar chains are all aligned {\it
539     perpendicular} to the direction of the ripple. Note that the flat
540     membrane has multiple vortex defects in the dipolar ordering, and the
541     ordering on the lower leaf of the bilayer can be in an entirely
542     different direction from the upper leaf.\label{fig:topView}}
543 gezelter 3195 \end{figure}
544    
545 gezelter 3202 The principal method for observing orientational ordering in dipolar
546     or liquid crystalline systems is the $P_2$ order parameter (defined
547     as $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
548     eigenvalue of the matrix,
549     \begin{equation}
550     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
551     \begin{array}{ccc}
552     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
553     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
554     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
555     \end{array} \right).
556     \label{eq:opmatrix}
557     \end{equation}
558     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
559     for molecule $i$. (Here, $\hat{\bf u}_i$ can refer either to the
560     principal axis of the molecular body or to the dipole on the head
561     group of the molecule.) $P_2$ will be $1.0$ for a perfectly-ordered
562     system and near $0$ for a randomized system. Note that this order
563     parameter is {\em not} equal to the polarization of the system. For
564     example, the polarization of a perfect anti-ferroelectric arrangement
565     of point dipoles is $0$, but $P_2$ for the same system is $1$. The
566     eigenvector of $\mathsf{S}$ corresponding to the largest eigenvalue is
567     familiar as the director axis, which can be used to determine a
568     privileged axis for an orientationally-ordered system. Since the
569     molecular bodies are perpendicular to the head group dipoles, it is
570     possible for the director axes for the molecular bodies and the head
571     groups to be completely decoupled from each other.
572    
573 gezelter 3200 Figure \ref{fig:topView} shows snapshots of bird's-eye views of the
574 gezelter 3203 flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
575 gezelter 3200 bilayers. The directions of the dipoles on the head groups are
576     represented with two colored half spheres: blue (phosphate) and yellow
577     (amino). For flat bilayers, the system exhibits signs of
578 gezelter 3202 orientational frustration; some disorder in the dipolar head-to-tail
579     chains is evident with kinks visible at the edges between differently
580     ordered domains. The lipids can also move independently of lipids in
581     the opposing leaf, so the ordering of the dipoles on one leaf is not
582     necessarily consistent with the ordering on the other. These two
583 gezelter 3200 factors keep the total dipolar order parameter relatively low for the
584     flat phases.
585 xsun 3147
586 gezelter 3200 With increasing head group size, the surface becomes corrugated, and
587     the dipoles cannot move as freely on the surface. Therefore, the
588     translational freedom of lipids in one layer is dependent upon the
589 gezelter 3202 position of the lipids in the other layer. As a result, the ordering of
590 gezelter 3200 the dipoles on head groups in one leaf is correlated with the ordering
591     in the other leaf. Furthermore, as the membrane deforms due to the
592     corrugation, the symmetry of the allowed dipolar ordering on each leaf
593     is broken. The dipoles then self-assemble in a head-to-tail
594     configuration, and the dipolar order parameter increases dramatically.
595     However, the total polarization of the system is still close to zero.
596     This is strong evidence that the corrugated structure is an
597 gezelter 3204 anti-ferroelectric state. It is also notable that the head-to-tail
598 gezelter 3202 arrangement of the dipoles is always observed in a direction
599     perpendicular to the wave vector for the surface corrugation. This is
600     a similar finding to what we observed in our earlier work on the
601     elastic dipolar membranes.\cite{Sun2007}
602 gezelter 3200
603     The $P_2$ order parameters (for both the molecular bodies and the head
604     group dipoles) have been calculated to quantify the ordering in these
605 gezelter 3202 phases. Figure \ref{fig:rP2} shows that the $P_2$ order parameter for
606     the head-group dipoles increases with increasing head group size. When
607     the heads of the lipid molecules are small, the membrane is nearly
608     flat. Since the in-plane packing is essentially a close packing of the
609     head groups, the head dipoles exhibit frustration in their
610     orientational ordering.
611 gezelter 3200
612 gezelter 3202 The ordering trends for the tails are essentially opposite to the
613     ordering of the head group dipoles. The tail $P_2$ order parameter
614     {\it decreases} with increasing head size. This indicates that the
615     surface is more curved with larger head / tail size ratios. When the
616     surface is flat, all tails are pointing in the same direction (normal
617     to the bilayer surface). This simplified model appears to be
618     exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$
619     phase. We have not observed a smectic C gel phase ($L_{\beta'}$) for
620     this model system. Increasing the size of the heads results in
621     rapidly decreasing $P_2$ ordering for the molecular bodies.
622 gezelter 3199
623 xsun 3174 \begin{figure}[htb]
624     \centering
625     \includegraphics[width=\linewidth]{rP2}
626 gezelter 3202 \caption{The $P_2$ order parameters for head groups (circles) and
627     molecular bodies (squares) as a function of the ratio of head group
628     size ($\sigma_h$) to the width of the molecular bodies ($d$). \label{fig:rP2}}
629 xsun 3174 \end{figure}
630 xsun 3147
631 gezelter 3202 In addition to varying the size of the head groups, we studied the
632     effects of the interactions between head groups on the structure of
633     lipid bilayer by changing the strength of the dipoles. Figure
634     \ref{fig:sP2} shows how the $P_2$ order parameter changes with
635     increasing strength of the dipole. Generally, the dipoles on the head
636     groups become more ordered as the strength of the interaction between
637     heads is increased and become more disordered by decreasing the
638 gezelter 3204 interaction strength. When the interaction between the heads becomes
639 gezelter 3202 too weak, the bilayer structure does not persist; all lipid molecules
640     become dispersed in the solvent (which is non-polar in this
641 gezelter 3204 molecular-scale model). The critical value of the strength of the
642 gezelter 3202 dipole depends on the size of the head groups. The perfectly flat
643     surface becomes unstable below $5$ Debye, while the rippled
644     surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
645    
646     The ordering of the tails mirrors the ordering of the dipoles {\it
647     except for the flat phase}. Since the surface is nearly flat in this
648     phase, the order parameters are only weakly dependent on dipolar
649     strength until it reaches $15$ Debye. Once it reaches this value, the
650     head group interactions are strong enough to pull the head groups
651     close to each other and distort the bilayer structure. For a flat
652     surface, a substantial amount of free volume between the head groups
653     is normally available. When the head groups are brought closer by
654 gezelter 3203 dipolar interactions, the tails are forced to splay outward, first forming
655     curved bilayers, and then inverted micelles.
656 gezelter 3202
657     When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
658 gezelter 3204 when the strength of the dipole is increased above $16$ Debye. For
659 gezelter 3202 rippled bilayers, there is less free volume available between the head
660     groups. Therefore increasing dipolar strength weakly influences the
661     structure of the membrane. However, the increase in the body $P_2$
662     order parameters implies that the membranes are being slightly
663     flattened due to the effects of increasing head-group attraction.
664    
665     A very interesting behavior takes place when the head groups are very
666     large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the
667     dipolar strength is relatively weak ($\mu < 9$ Debye). In this case,
668     the two leaves of the bilayer become totally interdigitated with each
669     other in large patches of the membrane. With higher dipolar
670     strength, the interdigitation is limited to single lines that run
671     through the bilayer in a direction perpendicular to the ripple wave
672     vector.
673    
674 xsun 3174 \begin{figure}[htb]
675     \centering
676     \includegraphics[width=\linewidth]{sP2}
677 gezelter 3202 \caption{The $P_2$ order parameters for head group dipoles (a) and
678     molecular bodies (b) as a function of the strength of the dipoles.
679     These order parameters are shown for four values of the head group /
680     molecular width ratio ($\sigma_h / d$). \label{fig:sP2}}
681 xsun 3174 \end{figure}
682 xsun 3147
683 gezelter 3202 Figure \ref{fig:tP2} shows the dependence of the order parameters on
684     temperature. As expected, systems are more ordered at low
685     temperatures, and more disordered at high temperatures. All of the
686     bilayers we studied can become unstable if the temperature becomes
687     high enough. The only interesting feature of the temperature
688     dependence is in the flat surfaces ($\sigma_h=1.20 d$ and
689     $\sigma_h=1.28 d$). Here, when the temperature is increased above
690     $310$K, there is enough jostling of the head groups to allow the
691     dipolar frustration to resolve into more ordered states. This results
692     in a slight increase in the $P_2$ order parameter above this
693     temperature.
694    
695     For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$),
696     there is a slightly increased orientational ordering in the molecular
697     bodies above $290$K. Since our model lacks the detailed information
698     about the behavior of the lipid tails, this is the closest the model
699     can come to depicting the ripple ($P_{\beta'}$) to fluid
700     ($L_{\alpha}$) phase transition. What we are observing is a
701     flattening of the rippled structures made possible by thermal
702     expansion of the tightly-packed head groups. The lack of detailed
703     chain configurations also makes it impossible for this model to depict
704     the ripple to gel ($L_{\beta'}$) phase transition.
705    
706 xsun 3174 \begin{figure}[htb]
707     \centering
708     \includegraphics[width=\linewidth]{tP2}
709 gezelter 3202 \caption{The $P_2$ order parameters for head group dipoles (a) and
710     molecular bodies (b) as a function of temperature.
711     These order parameters are shown for four values of the head group /
712     molecular width ratio ($\sigma_h / d$).\label{fig:tP2}}
713 xsun 3174 \end{figure}
714 xsun 3147
715 gezelter 3203 Fig. \ref{fig:phaseDiagram} shows a phase diagram for the model as a
716     function of the head group / molecular width ratio ($\sigma_h / d$)
717     and the strength of the head group dipole moment ($\mu$). Note that
718     the specific form of the bilayer phase is governed almost entirely by
719     the head group / molecular width ratio, while the strength of the
720     dipolar interactions between the head groups governs the stability of
721     the bilayer phase. Weaker dipoles result in unstable bilayer phases,
722     while extremely strong dipoles can shift the equilibrium to an
723     inverted micelle phase when the head groups are small. Temperature
724     has little effect on the actual bilayer phase observed, although higher
725     temperatures can cause the unstable region to grow into the higher
726     dipole region of this diagram.
727    
728     \begin{figure}[htb]
729     \centering
730     \includegraphics[width=\linewidth]{phaseDiagram}
731     \caption{Phase diagram for the simple molecular model as a function
732     of the head group / molecular width ratio ($\sigma_h / d$) and the
733     strength of the head group dipole moment
734     ($\mu$).\label{fig:phaseDiagram}}
735     \end{figure}
736    
737 gezelter 3264
738     We have also computed orientational diffusion constants for the head
739     groups from the relaxation of the second-order Legendre polynomial
740     correlation function,
741     \begin{eqnarray}
742     C_{\ell}(t) & = & \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
743     \mu}_{i}(0) \right) \rangle \\ \\
744     & \approx & e^{-\ell(\ell + 1) \theta t},
745     \end{eqnarray}
746     of the head group dipoles. In this last line, we have used a simple
747     ``Debye''-like model for the relaxation of the correlation function,
748     specifically in the case when $\ell = 2$. The computed orientational
749     diffusion constants are given in table \ref{tab:relaxation}. The
750     notable feature we observe is that the orientational diffusion
751     constant for the head group exhibits an order of magnitude decrease
752     upon entering the rippled phase. Our orientational correlation times
753     are substantially in excess of those provided by...
754    
755    
756 xsun 3262 \begin{table*}
757     \begin{minipage}{\linewidth}
758     \begin{center}
759 gezelter 3264 \caption{Rotational diffusion constants for the head groups
760     ($\theta_h$) and molecular bodies ($\theta_b$) as a function of the
761     head-to-body width ratio. The orientational mobility of the head
762     groups experiences an {\it order of magnitude decrease} upon entering
763     the rippled phase, which suggests that the rippling is tied to a
764     freezing out of head group orientational freedom. Uncertainties in
765     the last digit are indicated by the values in parentheses.}
766 xsun 3262 \begin{tabular}{lcc}
767     \hline
768 gezelter 3264 $\sigma_h / d$ & $\theta_h (\mu s^{-1})$ & $\theta_b (1/fs)$ \\
769 xsun 3262 \hline
770 gezelter 3264 1.20 & $0.206(1) $ & $0.0175(5) $ \\
771     1.28 & $0.179(2) $ & $0.055(2) $ \\
772     1.35 & $0.025(1) $ & $0.195(3) $ \\
773     1.41 & $0.023(1) $ & $0.024(3) $ \\
774 xsun 3262 \end{tabular}
775     \label{tab:relaxation}
776     \end{center}
777     \end{minipage}
778     \end{table*}
779    
780 xsun 3174 \section{Discussion}
781     \label{sec:discussion}
782 xsun 3147
783 gezelter 3203 Symmetric and asymmetric ripple phases have been observed to form in
784     our molecular dynamics simulations of a simple molecular-scale lipid
785     model. The lipid model consists of an dipolar head group and an
786     ellipsoidal tail. Within the limits of this model, an explanation for
787     generalized membrane curvature is a simple mismatch in the size of the
788     heads with the width of the molecular bodies. With heads
789     substantially larger than the bodies of the molecule, this curvature
790     should be convex nearly everywhere, a requirement which could be
791     resolved either with micellar or cylindrical phases.
792 xsun 3201
793 gezelter 3203 The persistence of a {\it bilayer} structure therefore requires either
794     strong attractive forces between the head groups or exclusionary
795     forces from the solvent phase. To have a persistent bilayer structure
796     with the added requirement of convex membrane curvature appears to
797     result in corrugated structures like the ones pictured in
798     Fig. \ref{fig:phaseCartoon}. In each of the sections of these
799     corrugated phases, the local curvature near a most of the head groups
800     is convex. These structures are held together by the extremely strong
801     and directional interactions between the head groups.
802 xsun 3201
803 gezelter 3203 Dipolar head groups are key for the maintaining the bilayer structures
804     exhibited by this model. The dipoles are likely to form head-to-tail
805     configurations even in flat configurations, but the temperatures are
806     high enough that vortex defects become prevalent in the flat phase.
807     The flat phase we observed therefore appears to be substantially above
808     the Kosterlitz-Thouless transition temperature for a planar system of
809     dipoles with this set of parameters. For this reason, it would be
810     interesting to observe the thermal behavior of the flat phase at
811     substantially lower temperatures.
812 xsun 3201
813 gezelter 3203 One feature of this model is that an energetically favorable
814     orientational ordering of the dipoles can be achieved by forming
815     ripples. The corrugation of the surface breaks the symmetry of the
816 gezelter 3204 plane, making vortex defects somewhat more expensive, and stabilizing
817 gezelter 3203 the long range orientational ordering for the dipoles in the head
818     groups. Most of the rows of the head-to-tail dipoles are parallel to
819 gezelter 3204 each other and the system adopts a bulk anti-ferroelectric state. We
820 gezelter 3203 believe that this is the first time the organization of the head
821     groups in ripple phases has been addressed.
822    
823     Although the size-mismatch between the heads and molecular bodies
824     appears to be the primary driving force for surface convexity, the
825     persistence of the bilayer through the use of rippled structures is a
826     function of the strong, attractive interactions between the heads.
827     One important prediction we can make using the results from this
828     simple model is that if the dipole-dipole interaction is the leading
829     contributor to the head group attractions, the wave vectors for the
830     ripples should always be found {\it perpendicular} to the dipole
831     director axis. This echoes the prediction we made earlier for simple
832     elastic dipolar membranes, and may suggest experimental designs which
833     will test whether this is really the case in the phosphatidylcholine
834     $P_{\beta'}$ phases. The dipole director axis should also be easily
835     computable for the all-atom and coarse-grained simulations that have
836     been published in the literature.\cite{deVries05}
837    
838 xsun 3201 Although our model is simple, it exhibits some rich and unexpected
839 gezelter 3203 behaviors. It would clearly be a closer approximation to reality if
840     we allowed bending motions between the dipoles and the molecular
841     bodies, and if we replaced the rigid ellipsoids with ball-and-chain
842     tails. However, the advantages of this simple model (large system
843 gezelter 3204 sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
844 gezelter 3203 for a wide range of parameters. Our explanation of this rippling
845 xsun 3201 phenomenon will help us design more accurate molecular models for
846 gezelter 3203 corrugated membranes and experiments to test whether or not
847     dipole-dipole interactions exert an influence on membrane rippling.
848 gezelter 3199 \newpage
849 xsun 3147 \bibliography{mdripple}
850     \end{document}