ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
Revision: 3270
Committed: Fri Oct 26 22:19:44 2007 UTC (16 years, 10 months ago) by xsun
Content type: application/x-tex
File size: 48015 byte(s)
Log Message:
fix citation.

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 gezelter 3269 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 3269 within the gel phase,~\cite{Cevc87} although Tenchov {\it et al.} have
79     recently observed near-hexagonal packing in some phosphatidylcholine
80 xsun 3270 (PC) gel phases.\cite{Tenchov2001} The X-ray diffraction work by
81 gezelter 3204 Katsaras {\it et al.} showed that a rich phase diagram exhibiting both
82     {\it asymmetric} and {\it symmetric} ripples is possible for lecithin
83     bilayers.\cite{Katsaras00}
84 xsun 3174
85 gezelter 3195 A number of theoretical models have been presented to explain the
86     formation of the ripple phase. Marder {\it et al.} used a
87 gezelter 3204 curvature-dependent Landau-de~Gennes free-energy functional to predict
88 gezelter 3269 a rippled phase.~\cite{Marder84} This model and other related
89     continuum models predict higher fluidity in convex regions and that
90     concave portions of the membrane correspond to more solid-like
91     regions. Carlson and Sethna used a packing-competition model (in
92     which head groups and chains have competing packing energetics) to
93     predict the formation of a ripple-like phase. Their model predicted
94     that the high-curvature portions have lower-chain packing and
95     correspond to more fluid-like regions. Goldstein and Leibler used a
96     mean-field approach with a planar model for {\em inter-lamellar}
97     interactions to predict rippling in multilamellar
98     phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
99     anisotropy of the nearest-neighbor interactions} coupled to
100     hydrophobic constraining forces which restrict height differences
101     between nearest neighbors is the origin of the ripple
102     phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
103     theory for tilt order and curvature of a single membrane and concluded
104     that {\em coupling of molecular tilt to membrane curvature} is
105     responsible for the production of ripples.~\cite{Lubensky93} Misbah,
106     Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
107     interactions} can lead to ripple instabilities.~\cite{Misbah98}
108     Heimburg presented a {\em coexistence model} for ripple formation in
109     which he postulates that fluid-phase line defects cause sharp
110     curvature between relatively flat gel-phase regions.~\cite{Heimburg00}
111     Kubica has suggested that a lattice model of polar head groups could
112     be valuable in trying to understand bilayer phase
113     formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of
114     lamellar stacks of hexagonal lattices to show that large head groups
115 gezelter 3195 and molecular tilt with respect to the membrane normal vector can
116 gezelter 3269 cause bulk rippling.~\cite{Bannerjee02} Recently, Kranenburg and Smit
117     described the formation of symmetric ripple-like structures using a
118     coarse grained solvent-head-tail bead model.\cite{Kranenburg2005}
119     Their lipids consisted of a short chain of head beads tied to the two
120     longer ``chains''.
121 xsun 3174
122 gezelter 3204 In contrast, few large-scale molecular modeling studies have been
123 gezelter 3195 done due to the large size of the resulting structures and the time
124     required for the phases of interest to develop. With all-atom (and
125     even unified-atom) simulations, only one period of the ripple can be
126 gezelter 3204 observed and only for time scales in the range of 10-100 ns. One of
127     the most interesting molecular simulations was carried out by de~Vries
128 gezelter 3195 {\it et al.}~\cite{deVries05}. According to their simulation results,
129     the ripple consists of two domains, one resembling the gel bilayer,
130     while in the other, the two leaves of the bilayer are fully
131     interdigitated. The mechanism for the formation of the ripple phase
132     suggested by their work is a packing competition between the head
133     groups and the tails of the lipid molecules.\cite{Carlson87} Recently,
134 gezelter 3199 the ripple phase has also been studied by Lenz and Schmid using Monte
135 gezelter 3195 Carlo simulations.\cite{Lenz07} Their structures are similar to the De
136     Vries {\it et al.} structures except that the connection between the
137     two leaves of the bilayer is a narrow interdigitated line instead of
138     the fully interdigitated domain. The symmetric ripple phase was also
139     observed by Lenz {\it et al.}, and their work supports other claims
140     that the mismatch between the size of the head group and tail of the
141     lipid molecules is the driving force for the formation of the ripple
142     phase. Ayton and Voth have found significant undulations in
143     zero-surface-tension states of membranes simulated via dissipative
144     particle dynamics, but their results are consistent with purely
145     thermal undulations.~\cite{Ayton02}
146 xsun 3174
147 gezelter 3195 Although the organization of the tails of lipid molecules are
148     addressed by these molecular simulations and the packing competition
149 gezelter 3204 between head groups and tails is strongly implicated as the primary
150 gezelter 3195 driving force for ripple formation, questions about the ordering of
151 gezelter 3203 the head groups in ripple phase have not been settled.
152 xsun 3174
153 gezelter 3195 In a recent paper, we presented a simple ``web of dipoles'' spin
154     lattice model which provides some physical insight into relationship
155     between dipolar ordering and membrane buckling.\cite{Sun2007} We found
156     that dipolar elastic membranes can spontaneously buckle, forming
157 gezelter 3203 ripple-like topologies. The driving force for the buckling of dipolar
158 gezelter 3204 elastic membranes is the anti-ferroelectric ordering of the dipoles.
159 gezelter 3203 This was evident in the ordering of the dipole director axis
160 gezelter 3204 perpendicular to the wave vector of the surface ripples. A similar
161 gezelter 3195 phenomenon has also been observed by Tsonchev {\it et al.} in their
162 gezelter 3199 work on the spontaneous formation of dipolar peptide chains into
163     curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
164 gezelter 3195
165     In this paper, we construct a somewhat more realistic molecular-scale
166     lipid model than our previous ``web of dipoles'' and use molecular
167     dynamics simulations to elucidate the role of the head group dipoles
168     in the formation and morphology of the ripple phase. We describe our
169     model and computational methodology in section \ref{sec:method}.
170     Details on the simulations are presented in section
171     \ref{sec:experiment}, with results following in section
172     \ref{sec:results}. A final discussion of the role of dipolar heads in
173     the ripple formation can be found in section
174 xsun 3174 \ref{sec:discussion}.
175    
176 gezelter 3196 \section{Computational Model}
177 xsun 3174 \label{sec:method}
178    
179 gezelter 3199 \begin{figure}[htb]
180     \centering
181     \includegraphics[width=4in]{lipidModels}
182     \caption{Three different representations of DPPC lipid molecules,
183     including the chemical structure, an atomistic model, and the
184     head-body ellipsoidal coarse-grained model used in this
185     work.\label{fig:lipidModels}}
186     \end{figure}
187    
188 gezelter 3195 Our simple molecular-scale lipid model for studying the ripple phase
189     is based on two facts: one is that the most essential feature of lipid
190     molecules is their amphiphilic structure with polar head groups and
191     non-polar tails. Another fact is that the majority of lipid molecules
192     in the ripple phase are relatively rigid (i.e. gel-like) which makes
193     some fraction of the details of the chain dynamics negligible. Figure
194 gezelter 3204 \ref{fig:lipidModels} shows the molecular structure of a DPPC
195 gezelter 3195 molecule, as well as atomistic and molecular-scale representations of
196     a DPPC molecule. The hydrophilic character of the head group is
197     largely due to the separation of charge between the nitrogen and
198     phosphate groups. The zwitterionic nature of the PC headgroups leads
199     to abnormally large dipole moments (as high as 20.6 D), and this
200     strongly polar head group interacts strongly with the solvating water
201     layers immediately surrounding the membrane. The hydrophobic tail
202     consists of fatty acid chains. In our molecular scale model, lipid
203     molecules have been reduced to these essential features; the fatty
204     acid chains are represented by an ellipsoid with a dipolar ball
205     perched on one end to represent the effects of the charge-separated
206     head group. In real PC lipids, the direction of the dipole is
207     nearly perpendicular to the tail, so we have fixed the direction of
208     the point dipole rigidly in this orientation.
209 xsun 3147
210 gezelter 3195 The ellipsoidal portions of the model interact via the Gay-Berne
211     potential which has seen widespread use in the liquid crystal
212 gezelter 3199 community. Ayton and Voth have also used Gay-Berne ellipsoids for
213 gezelter 3204 modeling large length-scale properties of lipid
214 gezelter 3199 bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
215     was a single site model for the interactions of rigid ellipsoidal
216 gezelter 3195 molecules.\cite{Gay81} It can be thought of as a modification of the
217     Gaussian overlap model originally described by Berne and
218     Pechukas.\cite{Berne72} The potential is constructed in the familiar
219     form of the Lennard-Jones function using orientation-dependent
220     $\sigma$ and $\epsilon$ parameters,
221 gezelter 3202 \begin{equation*}
222 xsun 3174 V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
223     r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
224     {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
225     {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
226     -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
227     {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
228 gezelter 3195 \label{eq:gb}
229 gezelter 3202 \end{equation*}
230 gezelter 3195
231     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
232 gezelter 3199 \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
233     \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
234 gezelter 3195 are dependent on the relative orientations of the two molecules (${\bf
235     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
236 gezelter 3199 intermolecular separation (${\bf \hat{r}}_{ij}$). $\sigma$ and
237     $\sigma_0$ are also governed by shape mixing and anisotropy variables,
238 gezelter 3202 \begin {eqnarray*}
239 gezelter 3199 \sigma_0 & = & \sqrt{d_i^2 + d_j^2} \\ \\
240     \chi & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 -
241     d_j^2 \right)}{\left( l_j^2 + d_i^2 \right) \left(l_i^2 +
242     d_j^2 \right)}\right]^{1/2} \\ \\
243     \alpha^2 & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 +
244     d_i^2 \right)}{\left( l_j^2 - d_j^2 \right) \left(l_i^2 +
245     d_j^2 \right)}\right]^{1/2},
246 gezelter 3202 \end{eqnarray*}
247 gezelter 3199 where $l$ and $d$ describe the length and width of each uniaxial
248     ellipsoid. These shape anisotropy parameters can then be used to
249     calculate the range function,
250 gezelter 3202 \begin{equation*}
251 gezelter 3199 \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) = \sigma_{0}
252     \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
253     \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
254     \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
255     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
256     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
257     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
258     \right]^{-1/2}
259 gezelter 3202 \end{equation*}
260 gezelter 3199
261     Gay-Berne ellipsoids also have an energy scaling parameter,
262     $\epsilon^s$, which describes the well depth for two identical
263 gezelter 3204 ellipsoids in a {\it side-by-side} configuration. Additionally, a well
264 gezelter 3199 depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
265     the ratio between the well depths in the {\it end-to-end} and
266     side-by-side configurations. As in the range parameter, a set of
267     mixing and anisotropy variables can be used to describe the well
268     depths for dissimilar particles,
269     \begin {eqnarray*}
270     \epsilon_0 & = & \sqrt{\epsilon^s_i * \epsilon^s_j} \\ \\
271     \epsilon^r & = & \sqrt{\epsilon^r_i * \epsilon^r_j} \\ \\
272     \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
273     \\ \\
274     \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
275     \end{eqnarray*}
276     The form of the strength function is somewhat complicated,
277     \begin {eqnarray*}
278     \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
279     \epsilon_{0} \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
280     \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
281     \hat{r}}_{ij}) \\ \\
282     \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
283     \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
284     \hat{u}}_{j})^{2}\right]^{-1/2} \\ \\
285     \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) &
286     = &
287     1 - \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
288     \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
289     \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
290     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
291     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
292     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
293     \end {eqnarray*}
294     although many of the quantities and derivatives are identical with
295 gezelter 3202 those obtained for the range parameter. Ref. \citen{Luckhurst90}
296 gezelter 3199 has a particularly good explanation of the choice of the Gay-Berne
297     parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
298     excellent overview of the computational methods that can be used to
299     efficiently compute forces and torques for this potential can be found
300 gezelter 3202 in Ref. \citen{Golubkov06}
301 gezelter 3199
302     The choices of parameters we have used in this study correspond to a
303     shape anisotropy of 3 for the chain portion of the molecule. In
304     principle, this could be varied to allow for modeling of longer or
305     shorter chain lipid molecules. For these prolate ellipsoids, we have:
306 gezelter 3195 \begin{equation}
307     \begin{array}{rcl}
308 gezelter 3199 d & < & l \\
309     \epsilon^{r} & < & 1
310 gezelter 3195 \end{array}
311     \end{equation}
312 gezelter 3200 A sketch of the various structural elements of our molecular-scale
313     lipid / solvent model is shown in figure \ref{fig:lipidModel}. The
314     actual parameters used in our simulations are given in table
315     \ref{tab:parameters}.
316 gezelter 3195
317 gezelter 3199 \begin{figure}[htb]
318     \centering
319     \includegraphics[width=4in]{2lipidModel}
320     \caption{The parameters defining the behavior of the lipid
321     models. $l / d$ is the ratio of the head group to body diameter.
322     Molecular bodies had a fixed aspect ratio of 3.0. The solvent model
323     was a simplified 4-water bead ($\sigma_w \approx d$) that has been
324     used in other coarse-grained (DPD) simulations. The dipolar strength
325     (and the temperature and pressure) were the only other parameters that
326     were varied systematically.\label{fig:lipidModel}}
327     \end{figure}
328 gezelter 3195
329     To take into account the permanent dipolar interactions of the
330 gezelter 3203 zwitterionic head groups, we have placed fixed dipole moments $\mu_{i}$ at
331 gezelter 3199 one end of the Gay-Berne particles. The dipoles are oriented at an
332     angle $\theta = \pi / 2$ relative to the major axis. These dipoles
333 gezelter 3203 are protected by a head ``bead'' with a range parameter ($\sigma_h$) which we have
334 gezelter 3199 varied between $1.20 d$ and $1.41 d$. The head groups interact with
335     each other using a combination of Lennard-Jones,
336 gezelter 3202 \begin{equation}
337 gezelter 3200 V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
338 gezelter 3195 \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
339 gezelter 3202 \end{equation}
340 gezelter 3199 and dipole-dipole,
341 gezelter 3202 \begin{equation}
342 gezelter 3200 V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
343     \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
344 gezelter 3195 \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
345     \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
346 gezelter 3202 \end{equation}
347 gezelter 3195 potentials.
348     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
349     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
350 gezelter 3199 pointing along the inter-dipole vector $\mathbf{r}_{ij}$.
351 gezelter 3195
352 gezelter 3269 Since the charge separation distance is so large in zwitterionic head
353     groups (like the PC head groups), it would also be possible to use
354     either point charges or a ``split dipole'' approximation,
355     \begin{equation}
356     V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
357     \hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{R_{ij}^3 }} -
358     \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
359     r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
360     \end{equation}
361     where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and
362     $j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given
363     by,
364     \begin{equation}
365     R_{ij} = \sqrt {r_{ij}^2 + \frac{{d_i^2 }}{4} + \frac{{d_j^2
366     }}{4}}.
367     \end{equation}
368     Here, $d_i$ and $d_j$ are effect charge separation distances
369     associated with each of the two dipolar sites. This approximation to
370     the multipole expansion maintains the fast fall-off of the multipole
371     potentials but lacks the normal divergences when two polar groups get
372     close to one another.
373    
374 gezelter 3195 For the interaction between nonequivalent uniaxial ellipsoids (in this
375 gezelter 3199 case, between spheres and ellipsoids), the spheres are treated as
376     ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
377 gezelter 3200 ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of
378     the Gay-Berne potential we are using was generalized by Cleaver {\it
379     et al.} and is appropriate for dissimilar uniaxial
380     ellipsoids.\cite{Cleaver96}
381 xsun 3147
382 gezelter 3269 The solvent model in our simulations is similar to the one used by
383     Marrink {\it et al.} in their coarse grained simulations of lipid
384     bilayers.\cite{Marrink04} The solvent bead is a single site that
385     represents four water molecules (m = 72 amu) and has comparable
386     density and diffusive behavior to liquid water. However, since there
387     are no electrostatic sites on these beads, this solvent model cannot
388     replicate the dielectric properties of water. Note that although we
389     are using larger cutoff and switching radii than Marrink {\it et al.},
390     our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the
391     solvent diffuses at 0.43 $\AA^2 ps^{-1}$ (roughly twice as fast as
392     liquid water).
393 gezelter 3203
394 xsun 3198 \begin{table*}
395     \begin{minipage}{\linewidth}
396     \begin{center}
397 gezelter 3199 \caption{Potential parameters used for molecular-scale coarse-grained
398     lipid simulations}
399     \begin{tabular}{llccc}
400 xsun 3198 \hline
401 gezelter 3199 & & Head & Chain & Solvent \\
402 xsun 3198 \hline
403 gezelter 3200 $d$ (\AA) & & varied & 4.6 & 4.7 \\
404     $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
405 gezelter 3199 $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
406     $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 \\
407 gezelter 3200 $m$ (amu) & & 196 & 760 & 72.06 \\
408 gezelter 3199 $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
409     \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
410     \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
411     \multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\
412     $\mu$ (Debye) & & varied & 0 & 0 \\
413 xsun 3198 \end{tabular}
414     \label{tab:parameters}
415     \end{center}
416     \end{minipage}
417     \end{table*}
418 gezelter 3195
419 gezelter 3203 \section{Experimental Methodology}
420     \label{sec:experiment}
421 gezelter 3186
422 gezelter 3200 The parameters that were systematically varied in this study were the
423     size of the head group ($\sigma_h$), the strength of the dipole moment
424     ($\mu$), and the temperature of the system. Values for $\sigma_h$
425 gezelter 3203 ranged from 5.5 \AA\ to 6.5 \AA\ . If the width of the tails is taken
426     to be the unit of length, these head groups correspond to a range from
427     $1.2 d$ to $1.41 d$. Since the solvent beads are nearly identical in
428     diameter to the tail ellipsoids, all distances that follow will be
429     measured relative to this unit of distance. Because the solvent we
430     are using is non-polar and has a dielectric constant of 1, values for
431     $\mu$ are sampled from a range that is somewhat smaller than the 20.6
432 gezelter 3204 Debye dipole moment of the PC head groups.
433 gezelter 3200
434 gezelter 3196 To create unbiased bilayers, all simulations were started from two
435 gezelter 3200 perfectly flat monolayers separated by a 26 \AA\ gap between the
436 gezelter 3196 molecular bodies of the upper and lower leaves. The separated
437 gezelter 3204 monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
438 xsun 3174 coupling. The length of $z$ axis of the simulations was fixed and a
439     constant surface tension was applied to enable real fluctuations of
440 gezelter 3200 the bilayer. Periodic boundary conditions were used, and $480-720$
441     lipid molecules were present in the simulations, depending on the size
442     of the head beads. In all cases, the two monolayers spontaneously
443     collapsed into bilayer structures within 100 ps. Following this
444 gezelter 3204 collapse, all systems were equilibrated for $100$ ns at $300$ K.
445 xsun 3147
446 gezelter 3200 The resulting bilayer structures were then solvated at a ratio of $6$
447 gezelter 3196 solvent beads (24 water molecules) per lipid. These configurations
448 gezelter 3200 were then equilibrated for another $30$ ns. All simulations utilizing
449     the solvent were carried out at constant pressure ($P=1$ atm) with
450 gezelter 3269 $3$D anisotropic coupling, and small constant surface tension
451 gezelter 3203 ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
452 gezelter 3204 this model, a time step of $50$ fs was utilized with excellent energy
453 gezelter 3200 conservation. Data collection for structural properties of the
454     bilayers was carried out during a final 5 ns run following the solvent
455 gezelter 3269 equilibration. Orientational correlation functions and diffusion
456     constants were computed from 30 ns simulations in the microcanonical
457     (NVE) ensemble using the average volume from the end of the constant
458     pressure and surface tension runs. The timestep on these final
459     molecular dynamics runs was 25 fs. No appreciable changes in phase
460     structure were noticed upon switching to a microcanonical ensemble.
461     All simulations were performed using the {\sc oopse} molecular
462     modeling program.\cite{Meineke05}
463 gezelter 3196
464 gezelter 3203 A switching function was applied to all potentials to smoothly turn
465 gezelter 3269 off the interactions between a range of $22$ and $25$ \AA. The
466     switching function was the standard (cubic) function,
467     \begin{equation}
468     s(r) =
469     \begin{cases}
470     1 & \text{if $r \le r_{\text{sw}}$},\\
471     \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
472     {(r_{\text{cut}} - r_{\text{sw}})^3}
473     & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
474     0 & \text{if $r > r_{\text{cut}}$.}
475     \end{cases}
476     \label{eq:dipoleSwitching}
477     \end{equation}
478 gezelter 3203
479 gezelter 3196 \section{Results}
480 xsun 3174 \label{sec:results}
481 xsun 3147
482 gezelter 3203 The membranes in our simulations exhibit a number of interesting
483     bilayer phases. The surface topology of these phases depends most
484     sensitively on the ratio of the size of the head groups to the width
485     of the molecular bodies. With heads only slightly larger than the
486 gezelter 3204 bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
487 gezelter 3203
488     Increasing the head / body size ratio increases the local membrane
489     curvature around each of the lipids. With $\sigma_h=1.28 d$, the
490     surface is still essentially flat, but the bilayer starts to exhibit
491     signs of instability. We have observed occasional defects where a
492     line of lipid molecules on one leaf of the bilayer will dip down to
493     interdigitate with the other leaf. This gives each of the two bilayer
494     leaves some local convexity near the line defect. These structures,
495     once developed in a simulation, are very stable and are spaced
496     approximately 100 \AA\ away from each other.
497    
498     With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
499     resolves into a ``symmetric'' ripple phase. Each leaf of the bilayer
500     is broken into several convex, hemicylinderical sections, and opposite
501     leaves are fitted together much like roof tiles. There is no
502     interdigitation between the upper and lower leaves of the bilayer.
503    
504     For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
505     local curvature is substantially larger, and the resulting bilayer
506     structure resolves into an asymmetric ripple phase. This structure is
507 gezelter 3204 very similar to the structures observed by both de~Vries {\it et al.}
508 gezelter 3203 and Lenz {\it et al.}. For a given ripple wave vector, there are two
509     possible asymmetric ripples, which is not the case for the symmetric
510     phase observed when $\sigma_h = 1.35 d$.
511    
512 xsun 3174 \begin{figure}[htb]
513     \centering
514 gezelter 3199 \includegraphics[width=4in]{phaseCartoon}
515 gezelter 3203 \caption{The role of the ratio between the head group size and the
516     width of the molecular bodies is to increase the local membrane
517     curvature. With strong attractive interactions between the head
518     groups, this local curvature can be maintained in bilayer structures
519     through surface corrugation. Shown above are three phases observed in
520     these simulations. With $\sigma_h = 1.20 d$, the bilayer maintains a
521     flat topology. For larger heads ($\sigma_h = 1.35 d$) the local
522     curvature resolves into a symmetrically rippled phase with little or
523     no interdigitation between the upper and lower leaves of the membrane.
524     The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
525     asymmetric rippled phases with interdigitation between the two
526     leaves.\label{fig:phaseCartoon}}
527 xsun 3174 \end{figure}
528 xsun 3147
529 gezelter 3203 Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
530     ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
531     phases are shown in Figure \ref{fig:phaseCartoon}.
532    
533 gezelter 3204 It is reasonable to ask how well the parameters we used can produce
534     bilayer properties that match experimentally known values for real
535     lipid bilayers. Using a value of $l = 13.8$ \AA for the ellipsoidal
536     tails and the fixed ellipsoidal aspect ratio of 3, our values for the
537     area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
538     entirely on the size of the head bead relative to the molecular body.
539     These values are tabulated in table \ref{tab:property}. Kucera {\it
540     et al.} have measured values for the head group spacings for a number
541     of PC lipid bilayers that range from 30.8 \AA (DLPC) to 37.8 (DPPC).
542     They have also measured values for the area per lipid that range from
543     60.6
544     \AA$^2$ (DMPC) to 64.2 \AA$^2$
545     (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
546     largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
547     bilayers (specifically the area per lipid) that resemble real PC
548     bilayers. The smaller head beads we used are perhaps better models
549     for PE head groups.
550    
551 xsun 3174 \begin{table*}
552     \begin{minipage}{\linewidth}
553     \begin{center}
554 gezelter 3204 \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
555     and amplitude observed as a function of the ratio between the head
556     beads and the diameters of the tails. Ripple wavelengths and
557     amplitudes are normalized to the diameter of the tail ellipsoids.}
558     \begin{tabular}{lccccc}
559 xsun 3174 \hline
560 gezelter 3204 $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
561     lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
562 xsun 3174 \hline
563 gezelter 3204 1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
564     1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
565     1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
566     1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
567 xsun 3174 \end{tabular}
568     \label{tab:property}
569     \end{center}
570     \end{minipage}
571     \end{table*}
572 xsun 3147
573 gezelter 3200 The membrane structures and the reduced wavelength $\lambda / d$,
574     reduced amplitude $A / d$ of the ripples are summarized in Table
575 gezelter 3203 \ref{tab:property}. The wavelength range is $15 - 17$ molecular bodies
576 gezelter 3200 and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
577 gezelter 3203 $2.2$ for symmetric ripple. These values are reasonably consistent
578     with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
579     Note, that given the lack of structural freedom in the tails of our
580     model lipids, the amplitudes observed from these simulations are
581     likely to underestimate of the true amplitudes.
582 xsun 3174
583 gezelter 3195 \begin{figure}[htb]
584     \centering
585 gezelter 3199 \includegraphics[width=4in]{topDown}
586 gezelter 3203 \caption{Top views of the flat (upper), symmetric ripple (middle),
587     and asymmetric ripple (lower) phases. Note that the head-group
588     dipoles have formed head-to-tail chains in all three of these phases,
589     but in the two rippled phases, the dipolar chains are all aligned {\it
590     perpendicular} to the direction of the ripple. Note that the flat
591     membrane has multiple vortex defects in the dipolar ordering, and the
592     ordering on the lower leaf of the bilayer can be in an entirely
593     different direction from the upper leaf.\label{fig:topView}}
594 gezelter 3195 \end{figure}
595    
596 gezelter 3202 The principal method for observing orientational ordering in dipolar
597     or liquid crystalline systems is the $P_2$ order parameter (defined
598     as $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
599     eigenvalue of the matrix,
600     \begin{equation}
601     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
602     \begin{array}{ccc}
603     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
604     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
605     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
606     \end{array} \right).
607     \label{eq:opmatrix}
608     \end{equation}
609     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
610     for molecule $i$. (Here, $\hat{\bf u}_i$ can refer either to the
611     principal axis of the molecular body or to the dipole on the head
612     group of the molecule.) $P_2$ will be $1.0$ for a perfectly-ordered
613     system and near $0$ for a randomized system. Note that this order
614     parameter is {\em not} equal to the polarization of the system. For
615     example, the polarization of a perfect anti-ferroelectric arrangement
616     of point dipoles is $0$, but $P_2$ for the same system is $1$. The
617     eigenvector of $\mathsf{S}$ corresponding to the largest eigenvalue is
618     familiar as the director axis, which can be used to determine a
619     privileged axis for an orientationally-ordered system. Since the
620     molecular bodies are perpendicular to the head group dipoles, it is
621     possible for the director axes for the molecular bodies and the head
622     groups to be completely decoupled from each other.
623    
624 gezelter 3200 Figure \ref{fig:topView} shows snapshots of bird's-eye views of the
625 gezelter 3203 flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
626 gezelter 3200 bilayers. The directions of the dipoles on the head groups are
627     represented with two colored half spheres: blue (phosphate) and yellow
628     (amino). For flat bilayers, the system exhibits signs of
629 gezelter 3202 orientational frustration; some disorder in the dipolar head-to-tail
630     chains is evident with kinks visible at the edges between differently
631     ordered domains. The lipids can also move independently of lipids in
632     the opposing leaf, so the ordering of the dipoles on one leaf is not
633     necessarily consistent with the ordering on the other. These two
634 gezelter 3200 factors keep the total dipolar order parameter relatively low for the
635     flat phases.
636 xsun 3147
637 gezelter 3200 With increasing head group size, the surface becomes corrugated, and
638     the dipoles cannot move as freely on the surface. Therefore, the
639     translational freedom of lipids in one layer is dependent upon the
640 gezelter 3202 position of the lipids in the other layer. As a result, the ordering of
641 gezelter 3200 the dipoles on head groups in one leaf is correlated with the ordering
642     in the other leaf. Furthermore, as the membrane deforms due to the
643     corrugation, the symmetry of the allowed dipolar ordering on each leaf
644     is broken. The dipoles then self-assemble in a head-to-tail
645     configuration, and the dipolar order parameter increases dramatically.
646     However, the total polarization of the system is still close to zero.
647     This is strong evidence that the corrugated structure is an
648 gezelter 3204 anti-ferroelectric state. It is also notable that the head-to-tail
649 gezelter 3202 arrangement of the dipoles is always observed in a direction
650     perpendicular to the wave vector for the surface corrugation. This is
651     a similar finding to what we observed in our earlier work on the
652     elastic dipolar membranes.\cite{Sun2007}
653 gezelter 3200
654     The $P_2$ order parameters (for both the molecular bodies and the head
655     group dipoles) have been calculated to quantify the ordering in these
656 gezelter 3202 phases. Figure \ref{fig:rP2} shows that the $P_2$ order parameter for
657     the head-group dipoles increases with increasing head group size. When
658     the heads of the lipid molecules are small, the membrane is nearly
659     flat. Since the in-plane packing is essentially a close packing of the
660     head groups, the head dipoles exhibit frustration in their
661     orientational ordering.
662 gezelter 3200
663 gezelter 3202 The ordering trends for the tails are essentially opposite to the
664     ordering of the head group dipoles. The tail $P_2$ order parameter
665     {\it decreases} with increasing head size. This indicates that the
666     surface is more curved with larger head / tail size ratios. When the
667     surface is flat, all tails are pointing in the same direction (normal
668     to the bilayer surface). This simplified model appears to be
669     exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$
670     phase. We have not observed a smectic C gel phase ($L_{\beta'}$) for
671     this model system. Increasing the size of the heads results in
672     rapidly decreasing $P_2$ ordering for the molecular bodies.
673 gezelter 3199
674 xsun 3174 \begin{figure}[htb]
675     \centering
676     \includegraphics[width=\linewidth]{rP2}
677 gezelter 3202 \caption{The $P_2$ order parameters for head groups (circles) and
678     molecular bodies (squares) as a function of the ratio of head group
679     size ($\sigma_h$) to the width of the molecular bodies ($d$). \label{fig:rP2}}
680 xsun 3174 \end{figure}
681 xsun 3147
682 gezelter 3202 In addition to varying the size of the head groups, we studied the
683     effects of the interactions between head groups on the structure of
684     lipid bilayer by changing the strength of the dipoles. Figure
685     \ref{fig:sP2} shows how the $P_2$ order parameter changes with
686     increasing strength of the dipole. Generally, the dipoles on the head
687     groups become more ordered as the strength of the interaction between
688     heads is increased and become more disordered by decreasing the
689 gezelter 3204 interaction strength. When the interaction between the heads becomes
690 gezelter 3202 too weak, the bilayer structure does not persist; all lipid molecules
691     become dispersed in the solvent (which is non-polar in this
692 gezelter 3204 molecular-scale model). The critical value of the strength of the
693 gezelter 3202 dipole depends on the size of the head groups. The perfectly flat
694     surface becomes unstable below $5$ Debye, while the rippled
695     surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
696    
697     The ordering of the tails mirrors the ordering of the dipoles {\it
698     except for the flat phase}. Since the surface is nearly flat in this
699     phase, the order parameters are only weakly dependent on dipolar
700     strength until it reaches $15$ Debye. Once it reaches this value, the
701     head group interactions are strong enough to pull the head groups
702     close to each other and distort the bilayer structure. For a flat
703     surface, a substantial amount of free volume between the head groups
704     is normally available. When the head groups are brought closer by
705 gezelter 3203 dipolar interactions, the tails are forced to splay outward, first forming
706     curved bilayers, and then inverted micelles.
707 gezelter 3202
708     When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
709 gezelter 3204 when the strength of the dipole is increased above $16$ Debye. For
710 gezelter 3202 rippled bilayers, there is less free volume available between the head
711     groups. Therefore increasing dipolar strength weakly influences the
712     structure of the membrane. However, the increase in the body $P_2$
713     order parameters implies that the membranes are being slightly
714     flattened due to the effects of increasing head-group attraction.
715    
716     A very interesting behavior takes place when the head groups are very
717     large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the
718     dipolar strength is relatively weak ($\mu < 9$ Debye). In this case,
719     the two leaves of the bilayer become totally interdigitated with each
720     other in large patches of the membrane. With higher dipolar
721     strength, the interdigitation is limited to single lines that run
722     through the bilayer in a direction perpendicular to the ripple wave
723     vector.
724    
725 xsun 3174 \begin{figure}[htb]
726     \centering
727     \includegraphics[width=\linewidth]{sP2}
728 gezelter 3202 \caption{The $P_2$ order parameters for head group dipoles (a) and
729     molecular bodies (b) as a function of the strength of the dipoles.
730     These order parameters are shown for four values of the head group /
731     molecular width ratio ($\sigma_h / d$). \label{fig:sP2}}
732 xsun 3174 \end{figure}
733 xsun 3147
734 gezelter 3202 Figure \ref{fig:tP2} shows the dependence of the order parameters on
735     temperature. As expected, systems are more ordered at low
736     temperatures, and more disordered at high temperatures. All of the
737     bilayers we studied can become unstable if the temperature becomes
738     high enough. The only interesting feature of the temperature
739     dependence is in the flat surfaces ($\sigma_h=1.20 d$ and
740     $\sigma_h=1.28 d$). Here, when the temperature is increased above
741     $310$K, there is enough jostling of the head groups to allow the
742     dipolar frustration to resolve into more ordered states. This results
743     in a slight increase in the $P_2$ order parameter above this
744     temperature.
745    
746     For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$),
747     there is a slightly increased orientational ordering in the molecular
748     bodies above $290$K. Since our model lacks the detailed information
749     about the behavior of the lipid tails, this is the closest the model
750     can come to depicting the ripple ($P_{\beta'}$) to fluid
751     ($L_{\alpha}$) phase transition. What we are observing is a
752     flattening of the rippled structures made possible by thermal
753     expansion of the tightly-packed head groups. The lack of detailed
754     chain configurations also makes it impossible for this model to depict
755     the ripple to gel ($L_{\beta'}$) phase transition.
756    
757 xsun 3174 \begin{figure}[htb]
758     \centering
759     \includegraphics[width=\linewidth]{tP2}
760 gezelter 3202 \caption{The $P_2$ order parameters for head group dipoles (a) and
761     molecular bodies (b) as a function of temperature.
762     These order parameters are shown for four values of the head group /
763     molecular width ratio ($\sigma_h / d$).\label{fig:tP2}}
764 xsun 3174 \end{figure}
765 xsun 3147
766 gezelter 3203 Fig. \ref{fig:phaseDiagram} shows a phase diagram for the model as a
767     function of the head group / molecular width ratio ($\sigma_h / d$)
768     and the strength of the head group dipole moment ($\mu$). Note that
769     the specific form of the bilayer phase is governed almost entirely by
770     the head group / molecular width ratio, while the strength of the
771     dipolar interactions between the head groups governs the stability of
772     the bilayer phase. Weaker dipoles result in unstable bilayer phases,
773     while extremely strong dipoles can shift the equilibrium to an
774     inverted micelle phase when the head groups are small. Temperature
775     has little effect on the actual bilayer phase observed, although higher
776     temperatures can cause the unstable region to grow into the higher
777     dipole region of this diagram.
778    
779     \begin{figure}[htb]
780     \centering
781     \includegraphics[width=\linewidth]{phaseDiagram}
782     \caption{Phase diagram for the simple molecular model as a function
783     of the head group / molecular width ratio ($\sigma_h / d$) and the
784     strength of the head group dipole moment
785     ($\mu$).\label{fig:phaseDiagram}}
786     \end{figure}
787    
788 gezelter 3268 We have computed translational diffusion constants for lipid molecules
789     from the mean-square displacement,
790     \begin{equation}
791     D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
792     \end{equation}
793     of the lipid bodies. Translational diffusion constants for the
794     different head-to-tail size ratios (all at 300 K) are shown in table
795 gezelter 3269 \ref{tab:relaxation}. We have also computed orientational correlation
796     times for the head groups from fits of the second-order Legendre
797     polynomial correlation function,
798     \begin{equation}
799     C_{\ell}(t) = \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
800     \mu}_{i}(0) \right)
801     \end{equation}
802     of the head group dipoles. The orientational correlation functions
803     appear to have multiple components in their decay: a fast ($12 \pm 2$
804     ps) decay due to librational motion of the head groups, as well as
805     moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$)
806     components. The fit values for the moderate and slow correlation
807     times are listed in table \ref{tab:relaxation}. Standard deviations
808     in the fit time constants are quite large (on the order of the values
809     themselves).
810 gezelter 3264
811 gezelter 3269 Sparrman and Westlund used a multi-relaxation model for NMR lineshapes
812     observed in gel, fluid, and ripple phases of DPPC and obtained
813 gezelter 3268 estimates of a correlation time for water translational diffusion
814     ($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time
815     corresponds to water bound to small regions of the lipid membrane.
816 gezelter 3269 They further assume that the lipids can explore only a single period
817     of the ripple (essentially moving in a nearly one-dimensional path to
818     do so), and the correlation time can therefore be used to estimate a
819     value for the translational diffusion constant of $2.25 \times
820     10^{-11} m^2 s^{-1}$. The translational diffusion constants we obtain
821     are in reasonable agreement with this experimentally determined
822     value. However, the $T_2$ relaxation times obtained by Sparrman and
823     Westlund are consistent with P-N vector reorientation timescales of
824     2-5 ms. This is substantially slower than even the slowest component
825     we observe in the decay of our orientational correlation functions.
826     Other than the dipole-dipole interactions, our head groups have no
827     shape anisotropy which would force them to move as a unit with
828     neighboring molecules. This would naturally lead to P-N reorientation
829     times that are too fast when compared with experimental measurements.
830 gezelter 3264
831 xsun 3262 \begin{table*}
832     \begin{minipage}{\linewidth}
833     \begin{center}
834 gezelter 3269 \caption{Fit values for the rotational correlation times for the head
835     groups ($\tau^h$) and molecular bodies ($\tau^b$) as well as the
836     translational diffusion constants for the molecule as a function of
837     the head-to-body width ratio (all at 300 K). In all of the phases,
838     the head group correlation functions decay with an fast librational
839     contribution ($12 \pm 1$ ps). There are additional moderate
840     ($\tau^h_{\rm mid}$) and slow $\tau^h_{\rm slow}$ contributions to
841     orientational decay that depend strongly on the phase exhibited by the
842     lipids. The symmetric ripple phase ($\sigma_h / d = 1.35$) appears to
843     exhibit the slowest molecular reorientation.}
844     \begin{tabular}{lcccc}
845 xsun 3262 \hline
846 gezelter 3269 $\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm
847     slow} (\mu s)$ & $\tau_b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\
848 xsun 3262 \hline
849 gezelter 3269 1.20 & $0.4$ & $9.6$ & $9.5$ & $0.43(1)$ \\
850     1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\
851     1.35 & $3.2$ & $4.0$ & $0.9$ & $3.42(1)$ \\
852     1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\
853 xsun 3262 \end{tabular}
854     \label{tab:relaxation}
855     \end{center}
856     \end{minipage}
857     \end{table*}
858    
859 xsun 3174 \section{Discussion}
860     \label{sec:discussion}
861 xsun 3147
862 gezelter 3203 Symmetric and asymmetric ripple phases have been observed to form in
863     our molecular dynamics simulations of a simple molecular-scale lipid
864     model. The lipid model consists of an dipolar head group and an
865     ellipsoidal tail. Within the limits of this model, an explanation for
866     generalized membrane curvature is a simple mismatch in the size of the
867     heads with the width of the molecular bodies. With heads
868     substantially larger than the bodies of the molecule, this curvature
869     should be convex nearly everywhere, a requirement which could be
870     resolved either with micellar or cylindrical phases.
871 xsun 3201
872 gezelter 3203 The persistence of a {\it bilayer} structure therefore requires either
873     strong attractive forces between the head groups or exclusionary
874     forces from the solvent phase. To have a persistent bilayer structure
875     with the added requirement of convex membrane curvature appears to
876     result in corrugated structures like the ones pictured in
877     Fig. \ref{fig:phaseCartoon}. In each of the sections of these
878     corrugated phases, the local curvature near a most of the head groups
879     is convex. These structures are held together by the extremely strong
880     and directional interactions between the head groups.
881 xsun 3201
882 gezelter 3269 The attractive forces holding the bilayer together could either be
883     non-directional (as in the work of Kranenburg and
884     Smit),\cite{Kranenburg2005} or directional (as we have utilized in
885     these simulations). The dipolar head groups are key for the
886     maintaining the bilayer structures exhibited by this particular model;
887     reducing the strength of the dipole has the tendency to make the
888     rippled phase disappear. The dipoles are likely to form attractive
889     head-to-tail configurations even in flat configurations, but the
890     temperatures are high enough that vortex defects become prevalent in
891     the flat phase. The flat phase we observed therefore appears to be
892     substantially above the Kosterlitz-Thouless transition temperature for
893     a planar system of dipoles with this set of parameters. For this
894     reason, it would be interesting to observe the thermal behavior of the
895     flat phase at substantially lower temperatures.
896 xsun 3201
897 gezelter 3203 One feature of this model is that an energetically favorable
898     orientational ordering of the dipoles can be achieved by forming
899     ripples. The corrugation of the surface breaks the symmetry of the
900 gezelter 3204 plane, making vortex defects somewhat more expensive, and stabilizing
901 gezelter 3203 the long range orientational ordering for the dipoles in the head
902     groups. Most of the rows of the head-to-tail dipoles are parallel to
903 gezelter 3204 each other and the system adopts a bulk anti-ferroelectric state. We
904 gezelter 3203 believe that this is the first time the organization of the head
905     groups in ripple phases has been addressed.
906    
907     Although the size-mismatch between the heads and molecular bodies
908     appears to be the primary driving force for surface convexity, the
909     persistence of the bilayer through the use of rippled structures is a
910     function of the strong, attractive interactions between the heads.
911     One important prediction we can make using the results from this
912     simple model is that if the dipole-dipole interaction is the leading
913     contributor to the head group attractions, the wave vectors for the
914     ripples should always be found {\it perpendicular} to the dipole
915     director axis. This echoes the prediction we made earlier for simple
916     elastic dipolar membranes, and may suggest experimental designs which
917     will test whether this is really the case in the phosphatidylcholine
918     $P_{\beta'}$ phases. The dipole director axis should also be easily
919     computable for the all-atom and coarse-grained simulations that have
920     been published in the literature.\cite{deVries05}
921    
922 xsun 3201 Although our model is simple, it exhibits some rich and unexpected
923 gezelter 3203 behaviors. It would clearly be a closer approximation to reality if
924     we allowed bending motions between the dipoles and the molecular
925     bodies, and if we replaced the rigid ellipsoids with ball-and-chain
926     tails. However, the advantages of this simple model (large system
927 gezelter 3204 sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
928 gezelter 3203 for a wide range of parameters. Our explanation of this rippling
929 xsun 3201 phenomenon will help us design more accurate molecular models for
930 gezelter 3203 corrugated membranes and experiments to test whether or not
931     dipole-dipole interactions exert an influence on membrane rippling.
932 gezelter 3199 \newpage
933 xsun 3147 \bibliography{mdripple}
934     \end{document}