ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
Revision: 3203
Committed: Thu Aug 2 15:57:11 2007 UTC (17 years, 1 month ago) by gezelter
Content type: application/x-tex
File size: 40424 byte(s)
Log Message:
Nearing completion

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