ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
Revision: 3351
Committed: Fri Feb 29 22:02:20 2008 UTC (16 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 49283 byte(s)
Log Message:
final version?

File Contents

# User Rev Content
1 gezelter 3202 \documentclass[12pt]{article}
2 gezelter 3351 \usepackage{graphicx}
3 gezelter 3202 \usepackage{times}
4     \usepackage{mathptm}
5     \usepackage{tabularx}
6     \usepackage{setspace}
7 gezelter 3199 \usepackage{amsmath}
8     \usepackage{amssymb}
9 gezelter 3202 \usepackage[ref]{overcite}
10     \pagestyle{plain}
11     \pagenumbering{arabic}
12     \oddsidemargin 0.0cm \evensidemargin 0.0cm
13     \topmargin -21pt \headsep 10pt
14     \textheight 9.0in \textwidth 6.5in
15     \brokenpenalty=10000
16     \renewcommand{\baselinestretch}{1.2}
17     \renewcommand\citemid{\ } % no comma in optional reference note
18 xsun 3147
19 gezelter 3351
20 xsun 3147 \begin{document}
21 gezelter 3202 %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
22     %\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
23 xsun 3147
24 gezelter 3202 \bibliographystyle{achemso}
25 xsun 3147
26 gezelter 3203 \title{Dipolar ordering in the ripple phases of molecular-scale models
27 gezelter 3269 of lipid membranes}
28 gezelter 3202 \author{Xiuquan Sun and J. Daniel Gezelter \\
29     Department of Chemistry and Biochemistry,\\
30 gezelter 3199 University of Notre Dame, \\
31 xsun 3147 Notre Dame, Indiana 46556}
32    
33 gezelter 3202 %\email[E-mail:]{gezelter@nd.edu}
34    
35 xsun 3147 \date{\today}
36    
37 gezelter 3202 \maketitle
38    
39 xsun 3147 \begin{abstract}
40 gezelter 3203 Symmetric and asymmetric ripple phases have been observed to form in
41     molecular dynamics simulations of a simple molecular-scale lipid
42     model. The lipid model consists of an dipolar head group and an
43     ellipsoidal tail. Within the limits of this model, an explanation for
44     generalized membrane curvature is a simple mismatch in the size of the
45     heads with the width of the molecular bodies. The persistence of a
46     {\it bilayer} structure requires strong attractive forces between the
47     head groups. One feature of this model is that an energetically
48     favorable orientational ordering of the dipoles can be achieved by
49     out-of-plane membrane corrugation. The corrugation of the surface
50 gezelter 3204 stabilizes the long range orientational ordering for the dipoles in the
51     head groups which then adopt a bulk anti-ferroelectric state. We
52 gezelter 3203 observe a common feature of the corrugated dipolar membranes: the wave
53     vectors for the surface ripples are always found to be perpendicular
54     to the dipole director axis.
55 xsun 3147 \end{abstract}
56    
57 gezelter 3202 %\maketitle
58 gezelter 3203 \newpage
59 xsun 3147
60 xsun 3174 \section{Introduction}
61     \label{sec:Int}
62 gezelter 3195 Fully hydrated lipids will aggregate spontaneously to form bilayers
63     which exhibit a variety of phases depending on their temperatures and
64     compositions. Among these phases, a periodic rippled phase
65     ($P_{\beta'}$) appears as an intermediate phase between the gel
66     ($L_\beta$) and fluid ($L_{\alpha}$) phases for relatively pure
67     phosphatidylcholine (PC) bilayers. The ripple phase has attracted
68     substantial experimental interest over the past 30 years. Most
69     structural information of the ripple phase has been obtained by the
70     X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
71     microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
72     et al.} used atomic force microscopy (AFM) to observe ripple phase
73     morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
74     experimental results provide strong support for a 2-dimensional
75     hexagonal packing lattice of the lipid molecules within the ripple
76     phase. This is a notable change from the observed lipid packing
77 gezelter 3269 within the gel phase,~\cite{Cevc87} although Tenchov {\it et al.} have
78     recently observed near-hexagonal packing in some phosphatidylcholine
79 xsun 3270 (PC) gel phases.\cite{Tenchov2001} The X-ray diffraction work by
80 gezelter 3204 Katsaras {\it et al.} showed that a rich phase diagram exhibiting both
81     {\it asymmetric} and {\it symmetric} ripples is possible for lecithin
82     bilayers.\cite{Katsaras00}
83 xsun 3174
84 gezelter 3195 A number of theoretical models have been presented to explain the
85     formation of the ripple phase. Marder {\it et al.} used a
86 gezelter 3204 curvature-dependent Landau-de~Gennes free-energy functional to predict
87 gezelter 3269 a rippled phase.~\cite{Marder84} This model and other related
88     continuum models predict higher fluidity in convex regions and that
89     concave portions of the membrane correspond to more solid-like
90     regions. Carlson and Sethna used a packing-competition model (in
91     which head groups and chains have competing packing energetics) to
92     predict the formation of a ripple-like phase. Their model predicted
93     that the high-curvature portions have lower-chain packing and
94     correspond to more fluid-like regions. Goldstein and Leibler used a
95     mean-field approach with a planar model for {\em inter-lamellar}
96     interactions to predict rippling in multilamellar
97     phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
98     anisotropy of the nearest-neighbor interactions} coupled to
99     hydrophobic constraining forces which restrict height differences
100     between nearest neighbors is the origin of the ripple
101     phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
102     theory for tilt order and curvature of a single membrane and concluded
103     that {\em coupling of molecular tilt to membrane curvature} is
104     responsible for the production of ripples.~\cite{Lubensky93} Misbah,
105     Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
106     interactions} can lead to ripple instabilities.~\cite{Misbah98}
107     Heimburg presented a {\em coexistence model} for ripple formation in
108     which he postulates that fluid-phase line defects cause sharp
109     curvature between relatively flat gel-phase regions.~\cite{Heimburg00}
110     Kubica has suggested that a lattice model of polar head groups could
111     be valuable in trying to understand bilayer phase
112     formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of
113     lamellar stacks of hexagonal lattices to show that large head groups
114 gezelter 3195 and molecular tilt with respect to the membrane normal vector can
115 gezelter 3269 cause bulk rippling.~\cite{Bannerjee02} Recently, Kranenburg and Smit
116     described the formation of symmetric ripple-like structures using a
117     coarse grained solvent-head-tail bead model.\cite{Kranenburg2005}
118     Their lipids consisted of a short chain of head beads tied to the two
119     longer ``chains''.
120 xsun 3174
121 gezelter 3204 In contrast, few large-scale molecular modeling studies have been
122 gezelter 3195 done due to the large size of the resulting structures and the time
123     required for the phases of interest to develop. With all-atom (and
124     even unified-atom) simulations, only one period of the ripple can be
125 gezelter 3204 observed and only for time scales in the range of 10-100 ns. One of
126     the most interesting molecular simulations was carried out by de~Vries
127 gezelter 3195 {\it et al.}~\cite{deVries05}. According to their simulation results,
128     the ripple consists of two domains, one resembling the gel bilayer,
129     while in the other, the two leaves of the bilayer are fully
130     interdigitated. The mechanism for the formation of the ripple phase
131     suggested by their work is a packing competition between the head
132     groups and the tails of the lipid molecules.\cite{Carlson87} Recently,
133 gezelter 3199 the ripple phase has also been studied by Lenz and Schmid using Monte
134 gezelter 3195 Carlo simulations.\cite{Lenz07} Their structures are similar to the De
135     Vries {\it et al.} structures except that the connection between the
136     two leaves of the bilayer is a narrow interdigitated line instead of
137     the fully interdigitated domain. The symmetric ripple phase was also
138     observed by Lenz {\it et al.}, and their work supports other claims
139     that the mismatch between the size of the head group and tail of the
140     lipid molecules is the driving force for the formation of the ripple
141     phase. Ayton and Voth have found significant undulations in
142     zero-surface-tension states of membranes simulated via dissipative
143     particle dynamics, but their results are consistent with purely
144     thermal undulations.~\cite{Ayton02}
145 xsun 3174
146 gezelter 3195 Although the organization of the tails of lipid molecules are
147     addressed by these molecular simulations and the packing competition
148 gezelter 3204 between head groups and tails is strongly implicated as the primary
149 gezelter 3195 driving force for ripple formation, questions about the ordering of
150 gezelter 3203 the head groups in ripple phase have not been settled.
151 xsun 3174
152 gezelter 3195 In a recent paper, we presented a simple ``web of dipoles'' spin
153     lattice model which provides some physical insight into relationship
154     between dipolar ordering and membrane buckling.\cite{Sun2007} We found
155     that dipolar elastic membranes can spontaneously buckle, forming
156 gezelter 3203 ripple-like topologies. The driving force for the buckling of dipolar
157 gezelter 3204 elastic membranes is the anti-ferroelectric ordering of the dipoles.
158 gezelter 3203 This was evident in the ordering of the dipole director axis
159 gezelter 3204 perpendicular to the wave vector of the surface ripples. A similar
160 gezelter 3195 phenomenon has also been observed by Tsonchev {\it et al.} in their
161 gezelter 3199 work on the spontaneous formation of dipolar peptide chains into
162     curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
163 gezelter 3195
164     In this paper, we construct a somewhat more realistic molecular-scale
165     lipid model than our previous ``web of dipoles'' and use molecular
166     dynamics simulations to elucidate the role of the head group dipoles
167     in the formation and morphology of the ripple phase. We describe our
168     model and computational methodology in section \ref{sec:method}.
169     Details on the simulations are presented in section
170     \ref{sec:experiment}, with results following in section
171     \ref{sec:results}. A final discussion of the role of dipolar heads in
172     the ripple formation can be found in section
173 xsun 3174 \ref{sec:discussion}.
174    
175 gezelter 3196 \section{Computational Model}
176 xsun 3174 \label{sec:method}
177    
178 gezelter 3199 \begin{figure}[htb]
179     \centering
180     \includegraphics[width=4in]{lipidModels}
181     \caption{Three different representations of DPPC lipid molecules,
182     including the chemical structure, an atomistic model, and the
183     head-body ellipsoidal coarse-grained model used in this
184     work.\label{fig:lipidModels}}
185     \end{figure}
186    
187 gezelter 3195 Our simple molecular-scale lipid model for studying the ripple phase
188     is based on two facts: one is that the most essential feature of lipid
189     molecules is their amphiphilic structure with polar head groups and
190     non-polar tails. Another fact is that the majority of lipid molecules
191     in the ripple phase are relatively rigid (i.e. gel-like) which makes
192     some fraction of the details of the chain dynamics negligible. Figure
193 gezelter 3204 \ref{fig:lipidModels} shows the molecular structure of a DPPC
194 gezelter 3195 molecule, as well as atomistic and molecular-scale representations of
195     a DPPC molecule. The hydrophilic character of the head group is
196     largely due to the separation of charge between the nitrogen and
197     phosphate groups. The zwitterionic nature of the PC headgroups leads
198     to abnormally large dipole moments (as high as 20.6 D), and this
199     strongly polar head group interacts strongly with the solvating water
200     layers immediately surrounding the membrane. The hydrophobic tail
201     consists of fatty acid chains. In our molecular scale model, lipid
202     molecules have been reduced to these essential features; the fatty
203     acid chains are represented by an ellipsoid with a dipolar ball
204     perched on one end to represent the effects of the charge-separated
205     head group. In real PC lipids, the direction of the dipole is
206     nearly perpendicular to the tail, so we have fixed the direction of
207     the point dipole rigidly in this orientation.
208 xsun 3147
209 gezelter 3195 The ellipsoidal portions of the model interact via the Gay-Berne
210     potential which has seen widespread use in the liquid crystal
211 gezelter 3199 community. Ayton and Voth have also used Gay-Berne ellipsoids for
212 gezelter 3204 modeling large length-scale properties of lipid
213 gezelter 3199 bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
214     was a single site model for the interactions of rigid ellipsoidal
215 gezelter 3195 molecules.\cite{Gay81} It can be thought of as a modification of the
216     Gaussian overlap model originally described by Berne and
217     Pechukas.\cite{Berne72} The potential is constructed in the familiar
218     form of the Lennard-Jones function using orientation-dependent
219     $\sigma$ and $\epsilon$ parameters,
220 gezelter 3202 \begin{equation*}
221 xsun 3174 V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
222     r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
223     {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
224     {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
225     -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
226     {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
227 gezelter 3195 \label{eq:gb}
228 gezelter 3202 \end{equation*}
229 gezelter 3195
230     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
231 gezelter 3199 \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
232     \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
233 gezelter 3195 are dependent on the relative orientations of the two molecules (${\bf
234     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
235 gezelter 3199 intermolecular separation (${\bf \hat{r}}_{ij}$). $\sigma$ and
236     $\sigma_0$ are also governed by shape mixing and anisotropy variables,
237 gezelter 3202 \begin {eqnarray*}
238 gezelter 3199 \sigma_0 & = & \sqrt{d_i^2 + d_j^2} \\ \\
239     \chi & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 -
240     d_j^2 \right)}{\left( l_j^2 + d_i^2 \right) \left(l_i^2 +
241     d_j^2 \right)}\right]^{1/2} \\ \\
242     \alpha^2 & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 +
243     d_i^2 \right)}{\left( l_j^2 - d_j^2 \right) \left(l_i^2 +
244     d_j^2 \right)}\right]^{1/2},
245 gezelter 3202 \end{eqnarray*}
246 gezelter 3199 where $l$ and $d$ describe the length and width of each uniaxial
247     ellipsoid. These shape anisotropy parameters can then be used to
248     calculate the range function,
249 gezelter 3202 \begin{equation*}
250 gezelter 3199 \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) = \sigma_{0}
251     \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
252     \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
253     \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
254     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
255     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
256     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
257     \right]^{-1/2}
258 gezelter 3202 \end{equation*}
259 gezelter 3199
260     Gay-Berne ellipsoids also have an energy scaling parameter,
261     $\epsilon^s$, which describes the well depth for two identical
262 gezelter 3204 ellipsoids in a {\it side-by-side} configuration. Additionally, a well
263 gezelter 3199 depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
264     the ratio between the well depths in the {\it end-to-end} and
265     side-by-side configurations. As in the range parameter, a set of
266     mixing and anisotropy variables can be used to describe the well
267     depths for dissimilar particles,
268     \begin {eqnarray*}
269     \epsilon_0 & = & \sqrt{\epsilon^s_i * \epsilon^s_j} \\ \\
270     \epsilon^r & = & \sqrt{\epsilon^r_i * \epsilon^r_j} \\ \\
271     \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
272     \\ \\
273     \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
274     \end{eqnarray*}
275     The form of the strength function is somewhat complicated,
276     \begin {eqnarray*}
277     \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
278     \epsilon_{0} \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
279     \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
280     \hat{r}}_{ij}) \\ \\
281     \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
282     \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
283     \hat{u}}_{j})^{2}\right]^{-1/2} \\ \\
284     \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) &
285     = &
286     1 - \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
287     \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
288     \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
289     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
290     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
291     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
292     \end {eqnarray*}
293     although many of the quantities and derivatives are identical with
294 gezelter 3202 those obtained for the range parameter. Ref. \citen{Luckhurst90}
295 gezelter 3199 has a particularly good explanation of the choice of the Gay-Berne
296     parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
297     excellent overview of the computational methods that can be used to
298     efficiently compute forces and torques for this potential can be found
299 gezelter 3202 in Ref. \citen{Golubkov06}
300 gezelter 3199
301     The choices of parameters we have used in this study correspond to a
302     shape anisotropy of 3 for the chain portion of the molecule. In
303     principle, this could be varied to allow for modeling of longer or
304     shorter chain lipid molecules. For these prolate ellipsoids, we have:
305 gezelter 3195 \begin{equation}
306     \begin{array}{rcl}
307 gezelter 3199 d & < & l \\
308     \epsilon^{r} & < & 1
309 gezelter 3195 \end{array}
310     \end{equation}
311 gezelter 3200 A sketch of the various structural elements of our molecular-scale
312     lipid / solvent model is shown in figure \ref{fig:lipidModel}. The
313     actual parameters used in our simulations are given in table
314     \ref{tab:parameters}.
315 gezelter 3195
316 gezelter 3199 \begin{figure}[htb]
317     \centering
318     \includegraphics[width=4in]{2lipidModel}
319     \caption{The parameters defining the behavior of the lipid
320 gezelter 3351 models. $\sigma_h / d$ is the ratio of the head group to body diameter.
321 gezelter 3199 Molecular bodies had a fixed aspect ratio of 3.0. The solvent model
322     was a simplified 4-water bead ($\sigma_w \approx d$) that has been
323 gezelter 3351 used in other coarse-grained simulations. The dipolar strength
324 gezelter 3199 (and the temperature and pressure) were the only other parameters that
325     were varied systematically.\label{fig:lipidModel}}
326     \end{figure}
327 gezelter 3195
328     To take into account the permanent dipolar interactions of the
329 gezelter 3203 zwitterionic head groups, we have placed fixed dipole moments $\mu_{i}$ at
330 gezelter 3199 one end of the Gay-Berne particles. The dipoles are oriented at an
331     angle $\theta = \pi / 2$ relative to the major axis. These dipoles
332 gezelter 3203 are protected by a head ``bead'' with a range parameter ($\sigma_h$) which we have
333 gezelter 3199 varied between $1.20 d$ and $1.41 d$. The head groups interact with
334     each other using a combination of Lennard-Jones,
335 gezelter 3202 \begin{equation}
336 gezelter 3200 V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
337 gezelter 3195 \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
338 gezelter 3202 \end{equation}
339 gezelter 3199 and dipole-dipole,
340 gezelter 3202 \begin{equation}
341 gezelter 3200 V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
342     \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
343 gezelter 3195 \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
344     \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
345 gezelter 3202 \end{equation}
346 gezelter 3195 potentials.
347     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
348     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
349 gezelter 3199 pointing along the inter-dipole vector $\mathbf{r}_{ij}$.
350 gezelter 3195
351 gezelter 3269 Since the charge separation distance is so large in zwitterionic head
352     groups (like the PC head groups), it would also be possible to use
353     either point charges or a ``split dipole'' approximation,
354     \begin{equation}
355     V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
356     \hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{R_{ij}^3 }} -
357     \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
358     r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
359     \end{equation}
360     where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and
361     $j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given
362     by,
363     \begin{equation}
364     R_{ij} = \sqrt {r_{ij}^2 + \frac{{d_i^2 }}{4} + \frac{{d_j^2
365     }}{4}}.
366     \end{equation}
367 gezelter 3351 Here, $d_i$ and $d_j$ are charge separation distances associated with
368     each of the two dipolar sites. This approximation to the multipole
369     expansion maintains the fast fall-off of the multipole potentials but
370     lacks the normal divergences when two polar groups get close to one
371     another.
372 gezelter 3269
373 gezelter 3195 For the interaction between nonequivalent uniaxial ellipsoids (in this
374 gezelter 3199 case, between spheres and ellipsoids), the spheres are treated as
375     ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
376 gezelter 3200 ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of
377     the Gay-Berne potential we are using was generalized by Cleaver {\it
378     et al.} and is appropriate for dissimilar uniaxial
379     ellipsoids.\cite{Cleaver96}
380 xsun 3147
381 gezelter 3269 The solvent model in our simulations is similar to the one used by
382     Marrink {\it et al.} in their coarse grained simulations of lipid
383     bilayers.\cite{Marrink04} The solvent bead is a single site that
384     represents four water molecules (m = 72 amu) and has comparable
385     density and diffusive behavior to liquid water. However, since there
386     are no electrostatic sites on these beads, this solvent model cannot
387     replicate the dielectric properties of water. Note that although we
388     are using larger cutoff and switching radii than Marrink {\it et al.},
389     our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the
390 gezelter 3351 solvent diffuses at 0.43 $\AA^2 ps^{-1}$ (only twice as fast as liquid
391     water).
392 gezelter 3203
393 xsun 3198 \begin{table*}
394     \begin{minipage}{\linewidth}
395     \begin{center}
396 gezelter 3199 \caption{Potential parameters used for molecular-scale coarse-grained
397     lipid simulations}
398     \begin{tabular}{llccc}
399 xsun 3198 \hline
400 gezelter 3199 & & Head & Chain & Solvent \\
401 xsun 3198 \hline
402 gezelter 3200 $d$ (\AA) & & varied & 4.6 & 4.7 \\
403     $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
404 gezelter 3199 $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
405     $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 \\
406 gezelter 3200 $m$ (amu) & & 196 & 760 & 72.06 \\
407 gezelter 3199 $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
408     \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
409     \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
410     \multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\
411     $\mu$ (Debye) & & varied & 0 & 0 \\
412 xsun 3198 \end{tabular}
413     \label{tab:parameters}
414     \end{center}
415     \end{minipage}
416     \end{table*}
417 gezelter 3195
418 gezelter 3203 \section{Experimental Methodology}
419     \label{sec:experiment}
420 gezelter 3186
421 gezelter 3200 The parameters that were systematically varied in this study were the
422     size of the head group ($\sigma_h$), the strength of the dipole moment
423     ($\mu$), and the temperature of the system. Values for $\sigma_h$
424 gezelter 3351 ranged from 5.5 \AA\ to 6.5 \AA. If the width of the tails is taken
425 gezelter 3203 to be the unit of length, these head groups correspond to a range from
426     $1.2 d$ to $1.41 d$. Since the solvent beads are nearly identical in
427     diameter to the tail ellipsoids, all distances that follow will be
428     measured relative to this unit of distance. Because the solvent we
429     are using is non-polar and has a dielectric constant of 1, values for
430     $\mu$ are sampled from a range that is somewhat smaller than the 20.6
431 gezelter 3204 Debye dipole moment of the PC head groups.
432 gezelter 3200
433 gezelter 3196 To create unbiased bilayers, all simulations were started from two
434 gezelter 3200 perfectly flat monolayers separated by a 26 \AA\ gap between the
435 gezelter 3196 molecular bodies of the upper and lower leaves. The separated
436 gezelter 3204 monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
437 xsun 3174 coupling. The length of $z$ axis of the simulations was fixed and a
438     constant surface tension was applied to enable real fluctuations of
439 gezelter 3200 the bilayer. Periodic boundary conditions were used, and $480-720$
440     lipid molecules were present in the simulations, depending on the size
441     of the head beads. In all cases, the two monolayers spontaneously
442     collapsed into bilayer structures within 100 ps. Following this
443 gezelter 3204 collapse, all systems were equilibrated for $100$ ns at $300$ K.
444 xsun 3147
445 gezelter 3200 The resulting bilayer structures were then solvated at a ratio of $6$
446 gezelter 3196 solvent beads (24 water molecules) per lipid. These configurations
447 gezelter 3200 were then equilibrated for another $30$ ns. All simulations utilizing
448     the solvent were carried out at constant pressure ($P=1$ atm) with
449 gezelter 3269 $3$D anisotropic coupling, and small constant surface tension
450 gezelter 3203 ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
451 gezelter 3204 this model, a time step of $50$ fs was utilized with excellent energy
452 gezelter 3200 conservation. Data collection for structural properties of the
453     bilayers was carried out during a final 5 ns run following the solvent
454 gezelter 3269 equilibration. Orientational correlation functions and diffusion
455     constants were computed from 30 ns simulations in the microcanonical
456     (NVE) ensemble using the average volume from the end of the constant
457     pressure and surface tension runs. The timestep on these final
458     molecular dynamics runs was 25 fs. No appreciable changes in phase
459     structure were noticed upon switching to a microcanonical ensemble.
460     All simulations were performed using the {\sc oopse} molecular
461     modeling program.\cite{Meineke05}
462 gezelter 3196
463 gezelter 3203 A switching function was applied to all potentials to smoothly turn
464 gezelter 3269 off the interactions between a range of $22$ and $25$ \AA. The
465     switching function was the standard (cubic) function,
466     \begin{equation}
467     s(r) =
468     \begin{cases}
469     1 & \text{if $r \le r_{\text{sw}}$},\\
470     \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
471     {(r_{\text{cut}} - r_{\text{sw}})^3}
472     & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
473     0 & \text{if $r > r_{\text{cut}}$.}
474     \end{cases}
475     \label{eq:dipoleSwitching}
476     \end{equation}
477 gezelter 3203
478 gezelter 3196 \section{Results}
479 xsun 3174 \label{sec:results}
480 xsun 3147
481 gezelter 3203 The membranes in our simulations exhibit a number of interesting
482     bilayer phases. The surface topology of these phases depends most
483     sensitively on the ratio of the size of the head groups to the width
484     of the molecular bodies. With heads only slightly larger than the
485 gezelter 3204 bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
486 gezelter 3203
487     Increasing the head / body size ratio increases the local membrane
488     curvature around each of the lipids. With $\sigma_h=1.28 d$, the
489     surface is still essentially flat, but the bilayer starts to exhibit
490     signs of instability. We have observed occasional defects where a
491     line of lipid molecules on one leaf of the bilayer will dip down to
492     interdigitate with the other leaf. This gives each of the two bilayer
493     leaves some local convexity near the line defect. These structures,
494     once developed in a simulation, are very stable and are spaced
495     approximately 100 \AA\ away from each other.
496    
497     With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
498     resolves into a ``symmetric'' ripple phase. Each leaf of the bilayer
499     is broken into several convex, hemicylinderical sections, and opposite
500     leaves are fitted together much like roof tiles. There is no
501     interdigitation between the upper and lower leaves of the bilayer.
502    
503     For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
504     local curvature is substantially larger, and the resulting bilayer
505     structure resolves into an asymmetric ripple phase. This structure is
506 gezelter 3204 very similar to the structures observed by both de~Vries {\it et al.}
507 gezelter 3203 and Lenz {\it et al.}. For a given ripple wave vector, there are two
508     possible asymmetric ripples, which is not the case for the symmetric
509     phase observed when $\sigma_h = 1.35 d$.
510    
511 xsun 3174 \begin{figure}[htb]
512     \centering
513 gezelter 3199 \includegraphics[width=4in]{phaseCartoon}
514 gezelter 3203 \caption{The role of the ratio between the head group size and the
515     width of the molecular bodies is to increase the local membrane
516     curvature. With strong attractive interactions between the head
517     groups, this local curvature can be maintained in bilayer structures
518     through surface corrugation. Shown above are three phases observed in
519     these simulations. With $\sigma_h = 1.20 d$, the bilayer maintains a
520     flat topology. For larger heads ($\sigma_h = 1.35 d$) the local
521     curvature resolves into a symmetrically rippled phase with little or
522     no interdigitation between the upper and lower leaves of the membrane.
523     The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
524     asymmetric rippled phases with interdigitation between the two
525     leaves.\label{fig:phaseCartoon}}
526 xsun 3174 \end{figure}
527 xsun 3147
528 gezelter 3203 Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
529     ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
530     phases are shown in Figure \ref{fig:phaseCartoon}.
531    
532 gezelter 3204 It is reasonable to ask how well the parameters we used can produce
533     bilayer properties that match experimentally known values for real
534     lipid bilayers. Using a value of $l = 13.8$ \AA for the ellipsoidal
535     tails and the fixed ellipsoidal aspect ratio of 3, our values for the
536     area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
537     entirely on the size of the head bead relative to the molecular body.
538     These values are tabulated in table \ref{tab:property}. Kucera {\it
539     et al.} have measured values for the head group spacings for a number
540 gezelter 3351 of PC lipid bilayers that range from 30.8 \AA\ (DLPC) to 37.8 \AA\ (DPPC).
541 gezelter 3204 They have also measured values for the area per lipid that range from
542     60.6
543     \AA$^2$ (DMPC) to 64.2 \AA$^2$
544     (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
545     largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
546     bilayers (specifically the area per lipid) that resemble real PC
547     bilayers. The smaller head beads we used are perhaps better models
548     for PE head groups.
549    
550 xsun 3174 \begin{table*}
551     \begin{minipage}{\linewidth}
552     \begin{center}
553 gezelter 3204 \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
554     and amplitude observed as a function of the ratio between the head
555     beads and the diameters of the tails. Ripple wavelengths and
556     amplitudes are normalized to the diameter of the tail ellipsoids.}
557     \begin{tabular}{lccccc}
558 xsun 3174 \hline
559 gezelter 3204 $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
560     lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
561 xsun 3174 \hline
562 gezelter 3204 1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
563     1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
564     1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
565     1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
566 xsun 3174 \end{tabular}
567     \label{tab:property}
568     \end{center}
569     \end{minipage}
570     \end{table*}
571 xsun 3147
572 gezelter 3200 The membrane structures and the reduced wavelength $\lambda / d$,
573     reduced amplitude $A / d$ of the ripples are summarized in Table
574 gezelter 3203 \ref{tab:property}. The wavelength range is $15 - 17$ molecular bodies
575 gezelter 3200 and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
576 gezelter 3203 $2.2$ for symmetric ripple. These values are reasonably consistent
577     with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
578     Note, that given the lack of structural freedom in the tails of our
579     model lipids, the amplitudes observed from these simulations are
580     likely to underestimate of the true amplitudes.
581 xsun 3174
582 gezelter 3195 \begin{figure}[htb]
583     \centering
584 gezelter 3199 \includegraphics[width=4in]{topDown}
585 gezelter 3203 \caption{Top views of the flat (upper), symmetric ripple (middle),
586     and asymmetric ripple (lower) phases. Note that the head-group
587     dipoles have formed head-to-tail chains in all three of these phases,
588     but in the two rippled phases, the dipolar chains are all aligned {\it
589     perpendicular} to the direction of the ripple. Note that the flat
590     membrane has multiple vortex defects in the dipolar ordering, and the
591     ordering on the lower leaf of the bilayer can be in an entirely
592     different direction from the upper leaf.\label{fig:topView}}
593 gezelter 3195 \end{figure}
594    
595 gezelter 3202 The principal method for observing orientational ordering in dipolar
596     or liquid crystalline systems is the $P_2$ order parameter (defined
597     as $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
598     eigenvalue of the matrix,
599     \begin{equation}
600     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
601     \begin{array}{ccc}
602     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
603     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
604     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
605     \end{array} \right).
606     \label{eq:opmatrix}
607     \end{equation}
608     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
609     for molecule $i$. (Here, $\hat{\bf u}_i$ can refer either to the
610     principal axis of the molecular body or to the dipole on the head
611     group of the molecule.) $P_2$ will be $1.0$ for a perfectly-ordered
612     system and near $0$ for a randomized system. Note that this order
613     parameter is {\em not} equal to the polarization of the system. For
614     example, the polarization of a perfect anti-ferroelectric arrangement
615     of point dipoles is $0$, but $P_2$ for the same system is $1$. The
616     eigenvector of $\mathsf{S}$ corresponding to the largest eigenvalue is
617     familiar as the director axis, which can be used to determine a
618     privileged axis for an orientationally-ordered system. Since the
619     molecular bodies are perpendicular to the head group dipoles, it is
620     possible for the director axes for the molecular bodies and the head
621     groups to be completely decoupled from each other.
622    
623 gezelter 3200 Figure \ref{fig:topView} shows snapshots of bird's-eye views of the
624 gezelter 3203 flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
625 gezelter 3200 bilayers. The directions of the dipoles on the head groups are
626     represented with two colored half spheres: blue (phosphate) and yellow
627     (amino). For flat bilayers, the system exhibits signs of
628 gezelter 3202 orientational frustration; some disorder in the dipolar head-to-tail
629     chains is evident with kinks visible at the edges between differently
630     ordered domains. The lipids can also move independently of lipids in
631     the opposing leaf, so the ordering of the dipoles on one leaf is not
632     necessarily consistent with the ordering on the other. These two
633 gezelter 3200 factors keep the total dipolar order parameter relatively low for the
634     flat phases.
635 xsun 3147
636 gezelter 3200 With increasing head group size, the surface becomes corrugated, and
637     the dipoles cannot move as freely on the surface. Therefore, the
638     translational freedom of lipids in one layer is dependent upon the
639 gezelter 3202 position of the lipids in the other layer. As a result, the ordering of
640 gezelter 3200 the dipoles on head groups in one leaf is correlated with the ordering
641     in the other leaf. Furthermore, as the membrane deforms due to the
642     corrugation, the symmetry of the allowed dipolar ordering on each leaf
643     is broken. The dipoles then self-assemble in a head-to-tail
644     configuration, and the dipolar order parameter increases dramatically.
645     However, the total polarization of the system is still close to zero.
646     This is strong evidence that the corrugated structure is an
647 gezelter 3204 anti-ferroelectric state. It is also notable that the head-to-tail
648 gezelter 3202 arrangement of the dipoles is always observed in a direction
649     perpendicular to the wave vector for the surface corrugation. This is
650     a similar finding to what we observed in our earlier work on the
651     elastic dipolar membranes.\cite{Sun2007}
652 gezelter 3200
653     The $P_2$ order parameters (for both the molecular bodies and the head
654     group dipoles) have been calculated to quantify the ordering in these
655 gezelter 3202 phases. Figure \ref{fig:rP2} shows that the $P_2$ order parameter for
656     the head-group dipoles increases with increasing head group size. When
657     the heads of the lipid molecules are small, the membrane is nearly
658     flat. Since the in-plane packing is essentially a close packing of the
659     head groups, the head dipoles exhibit frustration in their
660     orientational ordering.
661 gezelter 3200
662 gezelter 3202 The ordering trends for the tails are essentially opposite to the
663     ordering of the head group dipoles. The tail $P_2$ order parameter
664     {\it decreases} with increasing head size. This indicates that the
665     surface is more curved with larger head / tail size ratios. When the
666     surface is flat, all tails are pointing in the same direction (normal
667     to the bilayer surface). This simplified model appears to be
668     exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$
669     phase. We have not observed a smectic C gel phase ($L_{\beta'}$) for
670     this model system. Increasing the size of the heads results in
671     rapidly decreasing $P_2$ ordering for the molecular bodies.
672 gezelter 3199
673 xsun 3174 \begin{figure}[htb]
674     \centering
675     \includegraphics[width=\linewidth]{rP2}
676 gezelter 3202 \caption{The $P_2$ order parameters for head groups (circles) and
677     molecular bodies (squares) as a function of the ratio of head group
678     size ($\sigma_h$) to the width of the molecular bodies ($d$). \label{fig:rP2}}
679 xsun 3174 \end{figure}
680 xsun 3147
681 gezelter 3202 In addition to varying the size of the head groups, we studied the
682     effects of the interactions between head groups on the structure of
683     lipid bilayer by changing the strength of the dipoles. Figure
684     \ref{fig:sP2} shows how the $P_2$ order parameter changes with
685     increasing strength of the dipole. Generally, the dipoles on the head
686     groups become more ordered as the strength of the interaction between
687     heads is increased and become more disordered by decreasing the
688 gezelter 3204 interaction strength. When the interaction between the heads becomes
689 gezelter 3202 too weak, the bilayer structure does not persist; all lipid molecules
690     become dispersed in the solvent (which is non-polar in this
691 gezelter 3204 molecular-scale model). The critical value of the strength of the
692 gezelter 3202 dipole depends on the size of the head groups. The perfectly flat
693     surface becomes unstable below $5$ Debye, while the rippled
694     surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
695    
696     The ordering of the tails mirrors the ordering of the dipoles {\it
697     except for the flat phase}. Since the surface is nearly flat in this
698     phase, the order parameters are only weakly dependent on dipolar
699     strength until it reaches $15$ Debye. Once it reaches this value, the
700     head group interactions are strong enough to pull the head groups
701     close to each other and distort the bilayer structure. For a flat
702     surface, a substantial amount of free volume between the head groups
703     is normally available. When the head groups are brought closer by
704 gezelter 3203 dipolar interactions, the tails are forced to splay outward, first forming
705     curved bilayers, and then inverted micelles.
706 gezelter 3202
707     When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
708 gezelter 3204 when the strength of the dipole is increased above $16$ Debye. For
709 gezelter 3202 rippled bilayers, there is less free volume available between the head
710     groups. Therefore increasing dipolar strength weakly influences the
711     structure of the membrane. However, the increase in the body $P_2$
712     order parameters implies that the membranes are being slightly
713     flattened due to the effects of increasing head-group attraction.
714    
715     A very interesting behavior takes place when the head groups are very
716     large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the
717     dipolar strength is relatively weak ($\mu < 9$ Debye). In this case,
718     the two leaves of the bilayer become totally interdigitated with each
719     other in large patches of the membrane. With higher dipolar
720     strength, the interdigitation is limited to single lines that run
721     through the bilayer in a direction perpendicular to the ripple wave
722     vector.
723    
724 xsun 3174 \begin{figure}[htb]
725     \centering
726     \includegraphics[width=\linewidth]{sP2}
727 gezelter 3202 \caption{The $P_2$ order parameters for head group dipoles (a) and
728     molecular bodies (b) as a function of the strength of the dipoles.
729     These order parameters are shown for four values of the head group /
730     molecular width ratio ($\sigma_h / d$). \label{fig:sP2}}
731 xsun 3174 \end{figure}
732 xsun 3147
733 gezelter 3202 Figure \ref{fig:tP2} shows the dependence of the order parameters on
734     temperature. As expected, systems are more ordered at low
735     temperatures, and more disordered at high temperatures. All of the
736     bilayers we studied can become unstable if the temperature becomes
737     high enough. The only interesting feature of the temperature
738     dependence is in the flat surfaces ($\sigma_h=1.20 d$ and
739     $\sigma_h=1.28 d$). Here, when the temperature is increased above
740     $310$K, there is enough jostling of the head groups to allow the
741     dipolar frustration to resolve into more ordered states. This results
742     in a slight increase in the $P_2$ order parameter above this
743     temperature.
744    
745     For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$),
746     there is a slightly increased orientational ordering in the molecular
747     bodies above $290$K. Since our model lacks the detailed information
748     about the behavior of the lipid tails, this is the closest the model
749     can come to depicting the ripple ($P_{\beta'}$) to fluid
750     ($L_{\alpha}$) phase transition. What we are observing is a
751     flattening of the rippled structures made possible by thermal
752     expansion of the tightly-packed head groups. The lack of detailed
753     chain configurations also makes it impossible for this model to depict
754     the ripple to gel ($L_{\beta'}$) phase transition.
755    
756 xsun 3174 \begin{figure}[htb]
757     \centering
758     \includegraphics[width=\linewidth]{tP2}
759 gezelter 3202 \caption{The $P_2$ order parameters for head group dipoles (a) and
760     molecular bodies (b) as a function of temperature.
761     These order parameters are shown for four values of the head group /
762     molecular width ratio ($\sigma_h / d$).\label{fig:tP2}}
763 xsun 3174 \end{figure}
764 xsun 3147
765 gezelter 3203 Fig. \ref{fig:phaseDiagram} shows a phase diagram for the model as a
766     function of the head group / molecular width ratio ($\sigma_h / d$)
767     and the strength of the head group dipole moment ($\mu$). Note that
768     the specific form of the bilayer phase is governed almost entirely by
769     the head group / molecular width ratio, while the strength of the
770     dipolar interactions between the head groups governs the stability of
771     the bilayer phase. Weaker dipoles result in unstable bilayer phases,
772     while extremely strong dipoles can shift the equilibrium to an
773     inverted micelle phase when the head groups are small. Temperature
774     has little effect on the actual bilayer phase observed, although higher
775     temperatures can cause the unstable region to grow into the higher
776     dipole region of this diagram.
777    
778     \begin{figure}[htb]
779     \centering
780     \includegraphics[width=\linewidth]{phaseDiagram}
781     \caption{Phase diagram for the simple molecular model as a function
782     of the head group / molecular width ratio ($\sigma_h / d$) and the
783     strength of the head group dipole moment
784     ($\mu$).\label{fig:phaseDiagram}}
785     \end{figure}
786    
787 gezelter 3268 We have computed translational diffusion constants for lipid molecules
788     from the mean-square displacement,
789     \begin{equation}
790 gezelter 3351 D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
791 gezelter 3268 \end{equation}
792     of the lipid bodies. Translational diffusion constants for the
793     different head-to-tail size ratios (all at 300 K) are shown in table
794 gezelter 3269 \ref{tab:relaxation}. We have also computed orientational correlation
795     times for the head groups from fits of the second-order Legendre
796     polynomial correlation function,
797     \begin{equation}
798     C_{\ell}(t) = \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
799 gezelter 3351 \mu}_{i}(0) \right) \rangle
800 gezelter 3269 \end{equation}
801     of the head group dipoles. The orientational correlation functions
802     appear to have multiple components in their decay: a fast ($12 \pm 2$
803     ps) decay due to librational motion of the head groups, as well as
804     moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$)
805     components. The fit values for the moderate and slow correlation
806     times are listed in table \ref{tab:relaxation}. Standard deviations
807     in the fit time constants are quite large (on the order of the values
808     themselves).
809 gezelter 3264
810 gezelter 3269 Sparrman and Westlund used a multi-relaxation model for NMR lineshapes
811     observed in gel, fluid, and ripple phases of DPPC and obtained
812 gezelter 3268 estimates of a correlation time for water translational diffusion
813     ($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time
814     corresponds to water bound to small regions of the lipid membrane.
815 gezelter 3269 They further assume that the lipids can explore only a single period
816     of the ripple (essentially moving in a nearly one-dimensional path to
817     do so), and the correlation time can therefore be used to estimate a
818     value for the translational diffusion constant of $2.25 \times
819     10^{-11} m^2 s^{-1}$. The translational diffusion constants we obtain
820     are in reasonable agreement with this experimentally determined
821     value. However, the $T_2$ relaxation times obtained by Sparrman and
822     Westlund are consistent with P-N vector reorientation timescales of
823     2-5 ms. This is substantially slower than even the slowest component
824     we observe in the decay of our orientational correlation functions.
825     Other than the dipole-dipole interactions, our head groups have no
826     shape anisotropy which would force them to move as a unit with
827     neighboring molecules. This would naturally lead to P-N reorientation
828     times that are too fast when compared with experimental measurements.
829 gezelter 3264
830 xsun 3262 \begin{table*}
831     \begin{minipage}{\linewidth}
832     \begin{center}
833 gezelter 3269 \caption{Fit values for the rotational correlation times for the head
834     groups ($\tau^h$) and molecular bodies ($\tau^b$) as well as the
835     translational diffusion constants for the molecule as a function of
836 gezelter 3351 the head-to-body width ratio. All correlation functions and transport
837     coefficients were computed from microcanonical simulations with an
838     average temperture of 300 K. In all of the phases, the head group
839     correlation functions decay with an fast librational contribution ($12
840     \pm 1$ ps). There are additional moderate ($\tau^h_{\rm mid}$) and
841     slow $\tau^h_{\rm slow}$ contributions to orientational decay that
842     depend strongly on the phase exhibited by the lipids. The symmetric
843     ripple phase ($\sigma_h / d = 1.35$) appears to exhibit the slowest
844     molecular reorientation.}
845 gezelter 3269 \begin{tabular}{lcccc}
846 xsun 3262 \hline
847 gezelter 3269 $\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm
848 gezelter 3351 slow} (\mu s)$ & $\tau^b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\
849 xsun 3262 \hline
850 gezelter 3269 1.20 & $0.4$ & $9.6$ & $9.5$ & $0.43(1)$ \\
851     1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\
852     1.35 & $3.2$ & $4.0$ & $0.9$ & $3.42(1)$ \\
853     1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\
854 xsun 3262 \end{tabular}
855     \label{tab:relaxation}
856     \end{center}
857     \end{minipage}
858     \end{table*}
859    
860 xsun 3174 \section{Discussion}
861     \label{sec:discussion}
862 xsun 3147
863 gezelter 3203 Symmetric and asymmetric ripple phases have been observed to form in
864     our molecular dynamics simulations of a simple molecular-scale lipid
865     model. The lipid model consists of an dipolar head group and an
866     ellipsoidal tail. Within the limits of this model, an explanation for
867     generalized membrane curvature is a simple mismatch in the size of the
868     heads with the width of the molecular bodies. With heads
869     substantially larger than the bodies of the molecule, this curvature
870     should be convex nearly everywhere, a requirement which could be
871     resolved either with micellar or cylindrical phases.
872 xsun 3201
873 gezelter 3203 The persistence of a {\it bilayer} structure therefore requires either
874     strong attractive forces between the head groups or exclusionary
875     forces from the solvent phase. To have a persistent bilayer structure
876     with the added requirement of convex membrane curvature appears to
877     result in corrugated structures like the ones pictured in
878     Fig. \ref{fig:phaseCartoon}. In each of the sections of these
879     corrugated phases, the local curvature near a most of the head groups
880     is convex. These structures are held together by the extremely strong
881     and directional interactions between the head groups.
882 xsun 3201
883 gezelter 3269 The attractive forces holding the bilayer together could either be
884     non-directional (as in the work of Kranenburg and
885     Smit),\cite{Kranenburg2005} or directional (as we have utilized in
886     these simulations). The dipolar head groups are key for the
887     maintaining the bilayer structures exhibited by this particular model;
888     reducing the strength of the dipole has the tendency to make the
889     rippled phase disappear. The dipoles are likely to form attractive
890     head-to-tail configurations even in flat configurations, but the
891     temperatures are high enough that vortex defects become prevalent in
892     the flat phase. The flat phase we observed therefore appears to be
893     substantially above the Kosterlitz-Thouless transition temperature for
894     a planar system of dipoles with this set of parameters. For this
895     reason, it would be interesting to observe the thermal behavior of the
896     flat phase at substantially lower temperatures.
897 xsun 3201
898 gezelter 3203 One feature of this model is that an energetically favorable
899     orientational ordering of the dipoles can be achieved by forming
900     ripples. The corrugation of the surface breaks the symmetry of the
901 gezelter 3204 plane, making vortex defects somewhat more expensive, and stabilizing
902 gezelter 3203 the long range orientational ordering for the dipoles in the head
903     groups. Most of the rows of the head-to-tail dipoles are parallel to
904 gezelter 3204 each other and the system adopts a bulk anti-ferroelectric state. We
905 gezelter 3203 believe that this is the first time the organization of the head
906     groups in ripple phases has been addressed.
907    
908     Although the size-mismatch between the heads and molecular bodies
909     appears to be the primary driving force for surface convexity, the
910     persistence of the bilayer through the use of rippled structures is a
911     function of the strong, attractive interactions between the heads.
912     One important prediction we can make using the results from this
913     simple model is that if the dipole-dipole interaction is the leading
914     contributor to the head group attractions, the wave vectors for the
915     ripples should always be found {\it perpendicular} to the dipole
916     director axis. This echoes the prediction we made earlier for simple
917     elastic dipolar membranes, and may suggest experimental designs which
918     will test whether this is really the case in the phosphatidylcholine
919     $P_{\beta'}$ phases. The dipole director axis should also be easily
920     computable for the all-atom and coarse-grained simulations that have
921     been published in the literature.\cite{deVries05}
922    
923 gezelter 3351 Experimental verification of our predictions of dipolar orientation
924     correlating with the ripple direction would require knowing both the
925     local orientation of a rippled region of the membrane (available via
926     AFM studies of supported bilayers) as well as the local ordering of
927     the membrane dipoles. Obtaining information about the local
928     orientations of the membrane dipoles may be available from
929     fluorescence detected linear dichroism (LD). Benninger {\it et al.}
930     have recently used axially-specific chromophores
931     2-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-pentanoyl)-1-hexadecanoyl-sn-glycero-3-phospocholine
932     ($\beta$-BODIPY FL C5-HPC or BODIPY-PC) and 3,3'
933     dioctadecyloxacarbocyanine perchlorate (DiO) in their
934     fluorescence-detected linear dichroism (LD) studies of plasma
935     membranes of living cells.\cite{Benninger:2005qy} The DiO dye aligns
936     its transition moment perpendicular to the membrane normal, while the
937     BODIPY-PC transition dipole is parallel with the membrane normal.
938     Without a doubt, using fluorescence detection of linear dichroism in
939     concert with AFM surface scanning would be difficult experiments to
940     carry out. However, there is some hope of performing experiments to
941     either verify or falsify the predictions of our simulations.
942    
943 xsun 3201 Although our model is simple, it exhibits some rich and unexpected
944 gezelter 3203 behaviors. It would clearly be a closer approximation to reality if
945     we allowed bending motions between the dipoles and the molecular
946     bodies, and if we replaced the rigid ellipsoids with ball-and-chain
947     tails. However, the advantages of this simple model (large system
948 gezelter 3204 sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
949 gezelter 3203 for a wide range of parameters. Our explanation of this rippling
950 xsun 3201 phenomenon will help us design more accurate molecular models for
951 gezelter 3203 corrugated membranes and experiments to test whether or not
952     dipole-dipole interactions exert an influence on membrane rippling.
953 gezelter 3199 \newpage
954 xsun 3147 \bibliography{mdripple}
955     \end{document}