ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mdRipple/mdripple.tex
(Generate patch)

Comparing trunk/mdRipple/mdripple.tex (file contents):
Revision 3189 by xsun, Fri Jul 20 00:18:15 2007 UTC vs.
Revision 3202 by gezelter, Wed Aug 1 16:07:12 2007 UTC

# Line 1 | Line 1
1   %\documentclass[aps,pre,twocolumn,amssymb,showpacs,floatfix]{revtex4}
2 < \documentclass[aps,pre,preprint,amssymb,showpacs]{revtex4}
2 > %\documentclass[aps,pre,preprint,amssymb]{revtex4}
3 > \documentclass[12pt]{article}
4 > \usepackage{times}
5 > \usepackage{mathptm}
6 > \usepackage{tabularx}
7 > \usepackage{setspace}
8 > \usepackage{amsmath}
9 > \usepackage{amssymb}
10   \usepackage{graphicx}
11 + \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  
21   \begin{document}
22 < \renewcommand{\thefootnote}{\fnsymbol{footnote}}
23 < \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
22 > %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
23 > %\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
24  
25 < %\bibliographystyle{aps}
25 > \bibliographystyle{achemso}
26  
27 < \title{Dipolar Ordering of the Ripple Phase in Lipid Membranes}
28 < \author{Xiuquan Sun and J. Daniel Gezelter}
29 < \email[E-mail:]{gezelter@nd.edu}
30 < \affiliation{Department of Chemistry and Biochemistry,\\
31 < University of Notre Dame, \\
27 > \title{Dipolar Ordering in Molecular-scale Models of the Ripple Phase
28 > in Lipid Membranes}
29 > \author{Xiuquan Sun and J. Daniel Gezelter \\
30 > Department of Chemistry and Biochemistry,\\
31 > University of Notre Dame, \\
32   Notre Dame, Indiana 46556}
33  
34 + %\email[E-mail:]{gezelter@nd.edu}
35 +
36   \date{\today}
37  
38 < \begin{abstract}
38 > \maketitle
39  
40 + \begin{abstract}
41 + The ripple phase in phosphatidylcholine (PC) bilayers has never been
42 + completely explained.
43   \end{abstract}
44  
45 < \pacs{}
25 < \maketitle
45 > %\maketitle
46  
47   \section{Introduction}
48   \label{sec:Int}
49 + Fully hydrated lipids will aggregate spontaneously to form bilayers
50 + which exhibit a variety of phases depending on their temperatures and
51 + compositions. Among these phases, a periodic rippled phase
52 + ($P_{\beta'}$) appears as an intermediate phase between the gel
53 + ($L_\beta$) and fluid ($L_{\alpha}$) phases for relatively pure
54 + phosphatidylcholine (PC) bilayers.  The ripple phase has attracted
55 + substantial experimental interest over the past 30 years. Most
56 + structural information of the ripple phase has been obtained by the
57 + X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
58 + microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
59 + et al.} used atomic force microscopy (AFM) to observe ripple phase
60 + morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
61 + experimental results provide strong support for a 2-dimensional
62 + hexagonal packing lattice of the lipid molecules within the ripple
63 + phase.  This is a notable change from the observed lipid packing
64 + within the gel phase.~\cite{Cevc87}
65  
66 < As one of the most important components in the formation of the
67 < biomembrane, lipid molecules attracted numerous studies in the past
68 < several decades. Due to their amphiphilic structure, when dispersed in
69 < water, lipids can self-assemble to construct a bilayer structure. The
70 < phase behavior of lipid membrane is well understood. The gel-fluid
71 < phase transition is known as main phase transition. However, there is
72 < an intermediate phase between gel and fluid phase for some lipid (like
73 < phosphatidycholine (PC)) membranes. This intermediate phase
74 < distinguish itself from other phases by its corrugated membrane
75 < surface, therefore this phase is termed ``ripple'' ($P_{\beta '}$)
76 < phase. The phase transition between gel-fluid and ripple phase is
77 < called pretransition. Since the pretransition usually occurs in room
78 < temperature, there might be some important biofuntions carried by the
79 < ripple phase for the living organism.
66 > A number of theoretical models have been presented to explain the
67 > formation of the ripple phase. Marder {\it et al.} used a
68 > curvature-dependent Landau-de Gennes free-energy functional to predict
69 > a rippled phase.~\cite{Marder84} This model and other related continuum
70 > models predict higher fluidity in convex regions and that concave
71 > portions of the membrane correspond to more solid-like regions.
72 > Carlson and Sethna used a packing-competition model (in which head
73 > groups and chains have competing packing energetics) to predict the
74 > formation of a ripple-like phase.  Their model predicted that the
75 > high-curvature portions have lower-chain packing and correspond to
76 > more fluid-like regions.  Goldstein and Leibler used a mean-field
77 > approach with a planar model for {\em inter-lamellar} interactions to
78 > predict rippling in multilamellar phases.~\cite{Goldstein88} McCullough
79 > and Scott proposed that the {\em anisotropy of the nearest-neighbor
80 > interactions} coupled to hydrophobic constraining forces which
81 > restrict height differences between nearest neighbors is the origin of
82 > the ripple phase.~\cite{McCullough90} Lubensky and MacKintosh
83 > introduced a Landau theory for tilt order and curvature of a single
84 > membrane and concluded that {\em coupling of molecular tilt to membrane
85 > curvature} is responsible for the production of
86 > ripples.~\cite{Lubensky93} Misbah, Duplat and Houchmandzadeh proposed
87 > that {\em inter-layer dipolar interactions} can lead to ripple
88 > instabilities.~\cite{Misbah98} Heimburg presented a {\em coexistence
89 > model} for ripple formation in which he postulates that fluid-phase
90 > line defects cause sharp curvature between relatively flat gel-phase
91 > regions.~\cite{Heimburg00} Kubica has suggested that a lattice model of
92 > polar head groups could be valuable in trying to understand bilayer
93 > phase formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations
94 > of lamellar stacks of hexagonal lattices to show that large headgroups
95 > and molecular tilt with respect to the membrane normal vector can
96 > cause bulk rippling.~\cite{Bannerjee02}
97  
98 < The ripple phase is observed experimentally by x-ray diffraction
99 < ~\cite{Sun96,Katsaras00}, freeze-fracture electron microscopy
100 < (FFEM)~\cite{Copeland80,Meyer96}, and atomic force microscopy (AFM)
101 < recently~\cite{Kaasgaard03}. The experimental studies suggest two
102 < kinds of ripple structures: asymmetric (sawtooth like) and symmetric
103 < (sinusoidal like) ripple phases. Substantial number of theoretical
104 < explaination applied on the formation of the ripple
105 < phases~\cite{Marder84,Carlson87,Goldstein88,McCullough90,Lubensky93,Misbah98,Heimburg00,Kubica02,Bannerjee02}.
106 < In contrast, few molecular modelling have been done due to the large
107 < size of the resulting structures and the time required for the phases
108 < of interest to develop. One of the interesting molecular simulations
109 < was carried out by De Vries and Marrink {\it et
110 < al.}~\cite{deVries05}. According to their dynamic simulation results,
111 < the ripple consists of two domains, one is gel bilayer, and in the
112 < other domain, the upper and lower leaves of the bilayer are fully
113 < interdigitated. The mechanism of the formation of the ripple phase in
114 < their work suggests the theory that the packing competition between
115 < head group and tail of lipid molecules is the driving force for the
116 < formation of the ripple phase~\cite{Carlson87}. Recently, the ripple
117 < phase is also studied by using monte carlo simulation~\cite{Lenz07},
118 < the ripple structure is similar to the results of Marrink except that
119 < the connection of the upper and lower leaves of the bilayer is an
120 < interdigitated line instead of the fully interdigitated
121 < domain. Furthermore, the symmetric ripple phase was also observed in
69 < their work. They claimed the mismatch between the size of the head
70 < group and tail of the lipid molecules is the driving force for the
71 < formation of the ripple phase.
98 > In contrast, few large-scale molecular modelling studies have been
99 > done due to the large size of the resulting structures and the time
100 > required for the phases of interest to develop.  With all-atom (and
101 > even unified-atom) simulations, only one period of the ripple can be
102 > observed and only for timescales in the range of 10-100 ns.  One of
103 > the most interesting molecular simulations was carried out by De Vries
104 > {\it et al.}~\cite{deVries05}. According to their simulation results,
105 > the ripple consists of two domains, one resembling the gel bilayer,
106 > while in the other, the two leaves of the bilayer are fully
107 > interdigitated.  The mechanism for the formation of the ripple phase
108 > suggested by their work is a packing competition between the head
109 > groups and the tails of the lipid molecules.\cite{Carlson87} Recently,
110 > the ripple phase has also been studied by Lenz and Schmid using Monte
111 > Carlo simulations.\cite{Lenz07} Their structures are similar to the De
112 > Vries {\it et al.} structures except that the connection between the
113 > two leaves of the bilayer is a narrow interdigitated line instead of
114 > the fully interdigitated domain.  The symmetric ripple phase was also
115 > observed by Lenz {\it et al.}, and their work supports other claims
116 > that the mismatch between the size of the head group and tail of the
117 > lipid molecules is the driving force for the formation of the ripple
118 > phase. Ayton and Voth have found significant undulations in
119 > zero-surface-tension states of membranes simulated via dissipative
120 > particle dynamics, but their results are consistent with purely
121 > thermal undulations.~\cite{Ayton02}
122  
123 < Although the organizations of the tails of lipid molecules are
124 < addressed by these molecular simulations, the ordering of the head
125 < group in ripple phase is still not settlement. We developed a simple
126 < ``web of dipoles'' spin lattice model which provides some physical
127 < insight in our previous studies~\cite{Sun2007}, we found the dipoles
78 < on head groups of the lipid molecules are ordered in an
79 < antiferroelectric state. The similiar phenomenon is also observed by
80 < Tsonchev {\it et al.} when they studied the formation of the
81 < nanotube\cite{Tsonchev04}.
123 > Although the organization of the tails of lipid molecules are
124 > addressed by these molecular simulations and the packing competition
125 > between headgroups and tails is strongly implicated as the primary
126 > driving force for ripple formation, questions about the ordering of
127 > the head groups in ripple phase has not been settled.
128  
129 < In this paper, we made a more realistic coarse-grained lipid model to
130 < understand the primary driving force for membrane corrugation and to
131 < elucidate the organization of the anisotropic interacting head group
132 < via molecular dynamics simulation. We will talk about our model and
133 < methodology in section \ref{sec:method}, and details of the simulation
134 < in section \ref{sec:experiment}. The results are shown in section
135 < \ref{sec:results}. At last, we will discuss the results in section
129 > In a recent paper, we presented a simple ``web of dipoles'' spin
130 > lattice model which provides some physical insight into relationship
131 > between dipolar ordering and membrane buckling.\cite{Sun2007} We found
132 > that dipolar elastic membranes can spontaneously buckle, forming
133 > ripple-like topologies.  The driving force for the buckling in dipolar
134 > elastic membranes the antiferroelectric ordering of the dipoles, and
135 > this was evident in the ordering of the dipole director axis
136 > perpendicular to the wave vector of the surface ripples.  A similiar
137 > phenomenon has also been observed by Tsonchev {\it et al.} in their
138 > work on the spontaneous formation of dipolar peptide chains into
139 > curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
140 >
141 > In this paper, we construct a somewhat more realistic molecular-scale
142 > lipid model than our previous ``web of dipoles'' and use molecular
143 > dynamics simulations to elucidate the role of the head group dipoles
144 > in the formation and morphology of the ripple phase.  We describe our
145 > model and computational methodology in section \ref{sec:method}.
146 > Details on the simulations are presented in section
147 > \ref{sec:experiment}, with results following in section
148 > \ref{sec:results}.  A final discussion of the role of dipolar heads in
149 > the ripple formation can be found in section
150   \ref{sec:discussion}.
151  
152 < \section{Methodology and Model}
152 > \section{Computational Model}
153   \label{sec:method}
154  
95 Our idea for developing a simple and reasonable lipid model to study
96 the ripple phase of lipid bilayers is based on two facts: one is that
97 the most essential feature of lipid molecules is their amphiphilic
98 structure with polar head groups and non-polar tails. Another fact is
99 that dominant numbers of lipid molecules are very rigid in ripple
100 phase which allows the details of the lipid molecules neglectable. The
101 lipid model is shown in Figure \ref{fig:lipidMM}. Figure
102 \ref{fig:lipidMM}a shows the molecular strucure of a DPPC molecule. The
103 hydrophilic character of the head group is the effect of the strong
104 dipole composed by a positive charge sitting on the nitrogen and a
105 negative charge on the phosphate. The hydrophobic tail consists of
106 fatty acid chains. In our model shown in Figure \ref{fig:lipidMM}b,
107 lipid molecules are represented by rigid bodies made of one head
108 sphere with a point dipole sitting on it and one ellipsoid tail, the
109 direction of the dipole is fixed to be perpendicular to the tail. The
110 breadth and length of tail are $\sigma_0$, $3\sigma_0$. The diameter
111 of heads varies from $1.20\sigma_0$ to $1.41\sigma_0$.  The model of
112 the solvent in our simulations is inspired by the idea of ``DPD''
113 water. Every four water molecules are reprsented by one sphere.
114
155   \begin{figure}[htb]
156   \centering
157 < \includegraphics[width=\linewidth]{lipidModels}
157 > \includegraphics[width=4in]{lipidModels}
158   \caption{Three different representations of DPPC lipid molecules,
159   including the chemical structure, an atomistic model, and the
160   head-body ellipsoidal coarse-grained model used in this
161   work.\label{fig:lipidModels}}
162   \end{figure}
163  
164 < Spheres interact each other with Lennard-Jones potential
165 < \begin{eqnarray*}
166 < V_{ij} = 4\epsilon \left[\left(\frac{\sigma_0}{r_{ij}}\right)^{12} -
167 < \left(\frac{\sigma_0}{r_{ij}}\right)^6\right]
168 < \end{eqnarray*}
169 < here, $\sigma_0$ is diameter of the spherical particle. $r_{ij}$ is
170 < the distance between two spheres. $\epsilon$ is the well depth.
171 < Dipoles interact each other with typical dipole potential
172 < \begin{eqnarray*}
173 < V_{ij} = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
174 < \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
175 < \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
176 < \end{eqnarray*}
177 < In this potential, $\mathbf{\hat u}_i$ is the unit vector pointing
178 < along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
179 < pointing along the inter-dipole vector $\mathbf{r}_{ij}$. Identical
180 < ellipsoids interact each other with Gay-Berne potential.
181 < \begin{eqnarray*}
164 > Our simple molecular-scale lipid model for studying the ripple phase
165 > is based on two facts: one is that the most essential feature of lipid
166 > molecules is their amphiphilic structure with polar head groups and
167 > non-polar tails. Another fact is that the majority of lipid molecules
168 > in the ripple phase are relatively rigid (i.e. gel-like) which makes
169 > some fraction of the details of the chain dynamics negligible.  Figure
170 > \ref{fig:lipidModels} shows the molecular strucure of a DPPC
171 > molecule, as well as atomistic and molecular-scale representations of
172 > a DPPC molecule.  The hydrophilic character of the head group is
173 > largely due to the separation of charge between the nitrogen and
174 > phosphate groups.  The zwitterionic nature of the PC headgroups leads
175 > to abnormally large dipole moments (as high as 20.6 D), and this
176 > strongly polar head group interacts strongly with the solvating water
177 > layers immediately surrounding the membrane.  The hydrophobic tail
178 > consists of fatty acid chains.  In our molecular scale model, lipid
179 > molecules have been reduced to these essential features; the fatty
180 > acid chains are represented by an ellipsoid with a dipolar ball
181 > perched on one end to represent the effects of the charge-separated
182 > head group.  In real PC lipids, the direction of the dipole is
183 > nearly perpendicular to the tail, so we have fixed the direction of
184 > the point dipole rigidly in this orientation.  
185 >
186 > The ellipsoidal portions of the model interact via the Gay-Berne
187 > potential which has seen widespread use in the liquid crystal
188 > community.  Ayton and Voth have also used Gay-Berne ellipsoids for
189 > modelling large length-scale properties of lipid
190 > bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
191 > was a single site model for the interactions of rigid ellipsoidal
192 > molecules.\cite{Gay81} It can be thought of as a modification of the
193 > Gaussian overlap model originally described by Berne and
194 > Pechukas.\cite{Berne72} The potential is constructed in the familiar
195 > form of the Lennard-Jones function using orientation-dependent
196 > $\sigma$ and $\epsilon$ parameters,
197 > \begin{equation*}
198   V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
199   r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
200   {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
201   {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
202   -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
203   {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
204 < \end{eqnarray*}
205 < where $\sigma_0 = \sqrt 2\sigma_s$, and orietation-dependent range
206 < parameter is given by
207 < \begin{eqnarray*}
208 < \sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}}) =
209 < {\sigma_{0}} \left(1-\frac{1}{2}\chi\left[\frac{({\mathbf{\hat r}_{ij}}
210 < \cdot {\mathbf{\hat u}_i} + {\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat
211 < u}_j})^2}{1+\chi({\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j})} +
212 < \frac{({\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat u}_i} - {\mathbf{\hat r}_{ij}}
213 < \cdot {\mathbf{\hat u}_j})^2}{1-\chi({\mathbf{\hat u}_i} \cdot
214 < {\mathbf{\hat u}_j})} \right] \right)^{-\frac{1}{2}}
215 < \end{eqnarray*}
216 < and the strength anisotropy function is,
217 < \begin{eqnarray*}
218 < \epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}}) =
219 < {\epsilon^\nu}({\mathbf{\hat u}_i}, {\mathbf{\hat
220 < u}_j}){\epsilon'^\mu}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
221 < {\mathbf{\hat r}_{ij}})
166 < \end{eqnarray*}
167 < with $\nu$ and $\mu$ being adjustable exponent, and
168 < $\epsilon({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j})$,
169 < $\epsilon'({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
170 < r}_{ij}})$ defined as
171 < \begin{eqnarray*}
172 < \epsilon({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}) =
173 < \epsilon_0\left[1-\chi^2({\mathbf{\hat u}_i} \cdot {\mathbf{\hat
174 < u}_j})^2\right]^{-\frac{1}{2}}
204 > \label{eq:gb}
205 > \end{equation*}
206 >
207 > The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
208 > \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
209 > \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
210 > are dependent on the relative orientations of the two molecules (${\bf
211 > \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
212 > intermolecular separation (${\bf \hat{r}}_{ij}$).  $\sigma$ and
213 > $\sigma_0$ are also governed by shape mixing and anisotropy variables,
214 > \begin {eqnarray*}
215 > \sigma_0 & = & \sqrt{d_i^2 + d_j^2} \\ \\
216 > \chi & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 -
217 > d_j^2 \right)}{\left( l_j^2 + d_i^2 \right) \left(l_i^2 +
218 > d_j^2 \right)}\right]^{1/2} \\ \\
219 > \alpha^2 & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 +
220 > d_i^2 \right)}{\left( l_j^2 - d_j^2 \right) \left(l_i^2 +
221 > d_j^2 \right)}\right]^{1/2},
222   \end{eqnarray*}
223 < \begin{eqnarray*}
224 < \epsilon'({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}}) =
225 < 1-\frac{1}{2}\chi'\left[\frac{({\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat
226 < u}_i} + {\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat
227 < u}_j})^2}{1+\chi'({\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j})} +
228 < \frac{({\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat u}_i} - {\mathbf{\hat r}_{ij}}
229 < \cdot {\mathbf{\hat u}_j})^2}{1-\chi'({\mathbf{\hat u}_i} \cdot
230 < {\mathbf{\hat u}_j})} \right]
223 > where $l$ and $d$ describe the length and width of each uniaxial
224 > ellipsoid.  These shape anisotropy parameters can then be used to
225 > calculate the range function,
226 > \begin{equation*}
227 > \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) = \sigma_{0}
228 > \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
229 > \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
230 > \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
231 > \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
232 > \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
233 > \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
234 > \right]^{-1/2}
235 > \end{equation*}
236 >
237 > Gay-Berne ellipsoids also have an energy scaling parameter,
238 > $\epsilon^s$, which describes the well depth for two identical
239 > ellipsoids in a {\it side-by-side} configuration.  Additionaly, a well
240 > depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
241 > the ratio between the well depths in the {\it end-to-end} and
242 > side-by-side configurations.  As in the range parameter, a set of
243 > mixing and anisotropy variables can be used to describe the well
244 > depths for dissimilar particles,
245 > \begin {eqnarray*}
246 > \epsilon_0 & = & \sqrt{\epsilon^s_i  * \epsilon^s_j} \\ \\
247 > \epsilon^r & = & \sqrt{\epsilon^r_i  * \epsilon^r_j} \\ \\
248 > \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
249 > \\ \\
250 > \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
251   \end{eqnarray*}
252 < the diameter dependent parameter $\chi$ is given by
253 < \begin{eqnarray*}
254 < \chi = \frac{({\sigma_s}^2 -
255 < {\sigma_e}^2)}{({\sigma_s}^2 + {\sigma_e}^2)}
256 < \end{eqnarray*}
257 < and the well depth dependent parameter $\chi'$ is given by
258 < \begin{eqnarray*}
259 < \chi' = \frac{({\epsilon_s}^{\frac{1}{\mu}} -
260 < {\epsilon_e}^{\frac{1}{\mu}})}{({\epsilon_s}^{\frac{1}{\mu}} +
261 < {\epsilon_e}^{\frac{1}{\mu}})}
262 < \end{eqnarray*}
263 < $\sigma_s$ is the side-by-side width, and $\sigma_e$ is the end-to-end
264 < length. $\epsilon_s$ is the side-by-side well depth, and $\epsilon_e$
265 < is the end-to-end well depth. For the interaction between
266 < nonequivalent uniaxial ellipsoids (in this case, between spheres and
267 < ellipsoids), the range parameter is generalized as\cite{Cleaver96}
268 < \begin{eqnarray*}
269 < \sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}}) =
270 < {\sigma_{0}} \left(1-\frac{1}{2}\chi\left[\frac{(\alpha{\mathbf{\hat r}_{ij}}
271 < \cdot {\mathbf{\hat u}_i} + \alpha^{-1}{\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat
272 < u}_j})^2}{1+\chi({\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j})} +
273 < \frac{(\alpha{\mathbf{\hat r}_{ij}} \cdot {\mathbf{\hat u}_i} - \alpha^{-1}{\mathbf{\hat r}_{ij}}
274 < \cdot {\mathbf{\hat u}_j})^2}{1-\chi({\mathbf{\hat u}_i} \cdot
275 < {\mathbf{\hat u}_j})} \right] \right)^{-\frac{1}{2}}
276 < \end{eqnarray*}
210 < where $\alpha$ is given by
211 < \begin{eqnarray*}
212 < \alpha^2 =
213 < \left[\frac{\left({\sigma_e}_i^2-{\sigma_s}_i^2\right)\left({\sigma_e}_j^2+{\sigma_s}_i^2\right)}{\left({\sigma_e}_j^2-{\sigma_s}_j^2\right)\left({\sigma_e}_i^2+{\sigma_s}_j^2\right)}
214 < \right]^{\frac{1}{2}}
215 < \end{eqnarray*}
216 < the strength parameter is adjusted by the suggestion of
217 < \cite{Cleaver96}. All potentials are truncated at $25$ \AA\ and
218 < shifted at $22$ \AA.
252 > The form of the strength function is somewhat complicated,
253 > \begin {eqnarray*}
254 > \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
255 > \epsilon_{0}  \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
256 > \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
257 > \hat{r}}_{ij}) \\ \\
258 > \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
259 > \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
260 > \hat{u}}_{j})^{2}\right]^{-1/2} \\ \\
261 > \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) &
262 > = &
263 > 1 - \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
264 > \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
265 > \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
266 > \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
267 > \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
268 > \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
269 > \end {eqnarray*}
270 > although many of the quantities and derivatives are identical with
271 > those obtained for the range parameter. Ref. \citen{Luckhurst90}
272 > has a particularly good explanation of the choice of the Gay-Berne
273 > parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
274 > excellent overview of the computational methods that can be used to
275 > efficiently compute forces and torques for this potential can be found
276 > in Ref. \citen{Golubkov06}
277  
278 + The choices of parameters we have used in this study correspond to a
279 + shape anisotropy of 3 for the chain portion of the molecule.  In
280 + principle, this could be varied to allow for modeling of longer or
281 + shorter chain lipid molecules. For these prolate ellipsoids, we have:
282 + \begin{equation}
283 + \begin{array}{rcl}
284 + d & < & l \\
285 + \epsilon^{r} & < & 1
286 + \end{array}
287 + \end{equation}
288 + A sketch of the various structural elements of our molecular-scale
289 + lipid / solvent model is shown in figure \ref{fig:lipidModel}.  The
290 + actual parameters used in our simulations are given in table
291 + \ref{tab:parameters}.
292 +
293   \begin{figure}[htb]
294   \centering
295 < \includegraphics[height=4in]{lipidModel}
295 > \includegraphics[width=4in]{2lipidModel}
296   \caption{The parameters defining the behavior of the lipid
297 < models. $\sigma_h / \sigma_0$ is the ratio of the head group to body
298 < diameter.  Molecular bodies all had an aspect ratio of 3.0.  The
299 < dipolar strength (and the temperature and pressure) wer the only other
300 < parameters that wer varied systematically.\label{fig:lipidModel}}
297 > models. $l / d$ is the ratio of the head group to body diameter.
298 > Molecular bodies had a fixed aspect ratio of 3.0.  The solvent model
299 > was a simplified 4-water bead ($\sigma_w \approx d$) that has been
300 > used in other coarse-grained (DPD) simulations.  The dipolar strength
301 > (and the temperature and pressure) were the only other parameters that
302 > were varied systematically.\label{fig:lipidModel}}
303   \end{figure}
304  
305 < \section{Experiment}
305 > To take into account the permanent dipolar interactions of the
306 > zwitterionic head groups, we place fixed dipole moments $\mu_{i}$ at
307 > one end of the Gay-Berne particles.  The dipoles are oriented at an
308 > angle $\theta = \pi / 2$ relative to the major axis.  These dipoles
309 > are protected by a head ``bead'' with a range parameter which we have
310 > varied between $1.20 d$ and $1.41 d$.  The head groups interact with
311 > each other using a combination of Lennard-Jones,
312 > \begin{equation}
313 > V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
314 > \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
315 > \end{equation}
316 > and dipole-dipole,
317 > \begin{equation}
318 > V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
319 > \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
320 > \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
321 > \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
322 > \end{equation}
323 > potentials.  
324 > In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
325 > along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
326 > pointing along the inter-dipole vector $\mathbf{r}_{ij}$.  
327 >
328 > For the interaction between nonequivalent uniaxial ellipsoids (in this
329 > case, between spheres and ellipsoids), the spheres are treated as
330 > ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
331 > ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$).  The form of
332 > the Gay-Berne potential we are using was generalized by Cleaver {\it
333 > et al.} and is appropriate for dissimilar uniaxial
334 > ellipsoids.\cite{Cleaver96}
335 >
336 > The solvent model in our simulations is identical to one used by
337 > Marrink {\it et al.}  in their dissipative particle dynamics (DPD)
338 > simulation of lipid bilayers.\cite{Marrink04} This solvent bead is a single
339 > site that represents four water molecules (m = 72 amu) and has
340 > comparable density and diffusive behavior to liquid water.  However,
341 > since there are no electrostatic sites on these beads, this solvent
342 > model cannot replicate the dielectric properties of water.
343 > \begin{table*}
344 > \begin{minipage}{\linewidth}
345 > \begin{center}
346 > \caption{Potential parameters used for molecular-scale coarse-grained
347 > lipid simulations}
348 > \begin{tabular}{llccc}
349 > \hline
350 >  & &  Head & Chain & Solvent \\
351 > \hline
352 > $d$ (\AA) & & varied & 4.6  & 4.7 \\
353 > $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
354 > $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
355 > $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 &  1 \\
356 > $m$ (amu) & & 196 & 760 & 72.06 \\
357 > $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
358 > \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
359 > \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
360 > \multicolumn{2}c{$I_{zz}$} &  0 &    9000 & N/A \\
361 > $\mu$ (Debye) & & varied & 0 & 0 \\
362 > \end{tabular}
363 > \label{tab:parameters}
364 > \end{center}
365 > \end{minipage}
366 > \end{table*}
367 >
368 > A switching function has been applied to all potentials to smoothly
369 > turn off the interactions between a range of $22$ and $25$ \AA.
370 >
371 > The parameters that were systematically varied in this study were the
372 > size of the head group ($\sigma_h$), the strength of the dipole moment
373 > ($\mu$), and the temperature of the system.  Values for $\sigma_h$
374 > ranged from 5.5 \AA\  to 6.5 \AA\ .  If the width of the tails is
375 > taken to be the unit of length, these head groups correspond to a
376 > range from $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly
377 > identical in diameter to the tail ellipsoids, all distances that
378 > follow will be measured relative to this unit of distance.
379 >
380 > \section{Experimental Methodology}
381   \label{sec:experiment}
382  
383 < To make the simulations less expensive and to observe long-time
384 < behavior of the lipid membranes, all simulations were started from two
385 < separate monolayers in the vaccum with $x-y$ anisotropic pressure
383 > To create unbiased bilayers, all simulations were started from two
384 > perfectly flat monolayers separated by a 26 \AA\ gap between the
385 > molecular bodies of the upper and lower leaves.  The separated
386 > monolayers were evolved in a vaccum with $x-y$ anisotropic pressure
387   coupling. The length of $z$ axis of the simulations was fixed and a
388   constant surface tension was applied to enable real fluctuations of
389 < the bilayer. Periodic boundaries were used. There were $480-720$ lipid
390 < molecules in the simulations depending on the size of the head
391 < beads. All the simulations were equlibrated for $100$ ns at $300$
392 < K. The resulting structures were solvated in water ($6$ DPD
393 < water/lipid molecule). These configurations were relaxed for another
243 < $30$ ns relaxation. All simulations with water were carried out at
244 < constant pressure ($P=1$ atm) by $3$D anisotropic coupling, and
245 < constant surface tension ($\gamma=0.015$). Given the absence of fast
246 < degrees of freedom in this model, a timestep of $50$ fs was
247 < utilized. Simulations were performed by using OOPSE
248 < package\cite{Meineke05}.
389 > the bilayer. Periodic boundary conditions were used, and $480-720$
390 > lipid molecules were present in the simulations, depending on the size
391 > of the head beads.  In all cases, the two monolayers spontaneously
392 > collapsed into bilayer structures within 100 ps. Following this
393 > collapse, all systems were equlibrated for $100$ ns at $300$ K.
394  
395 < \section{Results and Analysis}
395 > The resulting bilayer structures were then solvated at a ratio of $6$
396 > solvent beads (24 water molecules) per lipid. These configurations
397 > were then equilibrated for another $30$ ns. All simulations utilizing
398 > the solvent were carried out at constant pressure ($P=1$ atm) with
399 > $3$D anisotropic coupling, and constant surface tension
400 > ($\gamma=0.015$ UNIT). Given the absence of fast degrees of freedom in
401 > this model, a timestep of $50$ fs was utilized with excellent energy
402 > conservation.  Data collection for structural properties of the
403 > bilayers was carried out during a final 5 ns run following the solvent
404 > equilibration.  All simulations were performed using the OOPSE
405 > molecular modeling program.\cite{Meineke05}
406 >
407 > \section{Results}
408   \label{sec:results}
409  
410   Snapshots in Figure \ref{fig:phaseCartoon} show that the membrane is
411 < more corrugated increasing size of the head groups. The surface is
412 < nearly flat when $\sigma_h=1.20\sigma_0$. With
413 < $\sigma_h=1.28\sigma_0$, although the surface is still flat, the
414 < bilayer starts to splay inward; the upper leaf of the bilayer is
415 < connected to the lower leaf with an interdigitated line defect. Two
416 < periodicities with $100$ \AA\ width were observed in the
417 < simulation. This structure is very similiar to the structure observed
418 < by de Vries and Lenz {\it et al.}. The same basic structure is also
419 < observed when $\sigma_h=1.41\sigma_0$, but the wavelength of the
420 < surface corrugations depends sensitively on the size of the ``head''
421 < beads. From the undulation spectrum, the corrugation is clearly
265 < non-thermal.
411 > more corrugated with increasing size of the head groups. The surface
412 > is nearly flat when $\sigma_h=1.20 d$. With $\sigma_h=1.28 d$,
413 > although the surface is still flat, the bilayer starts to splay
414 > inward; the upper leaf of the bilayer is connected to the lower leaf
415 > with an interdigitated line defect. Two periodicities with $100$ \AA\
416 > wavelengths were observed in the simulation. This structure is very
417 > similiar to the structure observed by de Vries and Lenz {\it et
418 > al.}. The same basic structure is also observed when $\sigma_h=1.41
419 > d$, but the wavelength of the surface corrugations depends sensitively
420 > on the size of the ``head'' beads. From the undulation spectrum, the
421 > corrugation is clearly non-thermal.
422   \begin{figure}[htb]
423   \centering
424 < \includegraphics[width=\linewidth]{phaseCartoon}
424 > \includegraphics[width=4in]{phaseCartoon}
425   \caption{A sketch to discribe the structure of the phases observed in
426   our simulations.\label{fig:phaseCartoon}}
427   \end{figure}
428  
429 < When $\sigma_h=1.35\sigma_0$, we observed another corrugated surface
430 < morphology. This structure is different from the asymmetric rippled
429 > When $\sigma_h=1.35 d$, we observed another corrugated surface
430 > morphology.  This structure is different from the asymmetric rippled
431   surface; there is no interdigitation between the upper and lower
432   leaves of the bilayer. Each leaf of the bilayer is broken into several
433   hemicylinderical sections, and opposite leaves are fitted together
434   much like roof tiles. Unlike the surface in which the upper
435   hemicylinder is always interdigitated on the leading or trailing edge
436 < of lower hemicylinder, the symmetric ripple has no prefered direction.
437 < The corresponding cartoons are shown in Figure
436 > of lower hemicylinder, this ``symmetric'' ripple has no prefered
437 > direction.  The corresponding structures are shown in Figure
438   \ref{fig:phaseCartoon} for elucidation of the detailed structures of
439 < different phases. Figure \ref{fig:phaseCartoon} (a) is the flat phase,
440 < (b) is the asymmetric ripple phase corresponding to the lipid
441 < organization when $\sigma_h=1.28\sigma_0$ and $\sigma_h=1.41\sigma_0$,
442 < and (c) is the symmetric ripple phase observed when
443 < $\sigma_h=1.35\sigma_0$. In symmetric ripple phase, the bilayer is
444 < continuous everywhere on the whole membrane, however, in asymmetric
445 < ripple phase, the bilayer is intermittent domains connected by thin
446 < interdigitated monolayer which consists of upper and lower leaves of
291 < the bilayer.
439 > different phases.  The top panel in figure \ref{fig:phaseCartoon} is
440 > the flat phase, the middle panel shows the asymmetric ripple phase
441 > corresponding to $\sigma_h = 1.41 d$ and the lower panel shows the
442 > symmetric ripple phase observed when $\sigma_h=1.35 d$. In the
443 > symmetric ripple, the bilayer is continuous over the whole membrane,
444 > however, in asymmetric ripple phase, the bilayer domains are connected
445 > by thin interdigitated monolayers that share molecules between the
446 > upper and lower leaves.
447   \begin{table*}
448   \begin{minipage}{\linewidth}
449   \begin{center}
450 < \caption{}
450 > \caption{Phases, ripple wavelengths and amplitudes observed as a
451 > function of the ratio between the head beads and the diameters of the
452 > tails.  All lengths are normalized to the diameter of the tail
453 > ellipsoids.}
454   \begin{tabular}{lccc}
455   \hline
456 < $\sigma_h/\sigma_0$ & type of phase & $\lambda/\sigma_0$ & $A/\sigma_0$\\
456 > $\sigma_h / d$ & type of phase & $\lambda / d$ & $A / d$\\
457   \hline
458   1.20 & flat & N/A & N/A \\
459 < 1.28 & asymmetric flat & 21.7 & N/A \\
459 > 1.28 & asymmetric ripple or flat & 21.7 & N/A \\
460   1.35 & symmetric ripple & 17.2 & 2.2 \\
461   1.41 & asymmetric ripple & 15.4 & 1.5 \\
462   \end{tabular}
# Line 307 | Line 465 | The membrane structures and the reduced wavelength $\l
465   \end{minipage}
466   \end{table*}
467  
468 < The membrane structures and the reduced wavelength $\lambda/\sigma_0$,
469 < reduced amplitude $A/\sigma_0$ of the ripples are summerized in Table
470 < \ref{tab:property}. The wavelength range is $15~21$ and the amplitude
471 < is $1.5$ for asymmetric ripple and $2.2$ for symmetric ripple. These
472 < values are consistent to the experimental results. Note, the
473 < amplitudes are underestimated without the melted tails in our
474 < simulations.
468 > The membrane structures and the reduced wavelength $\lambda / d$,
469 > reduced amplitude $A / d$ of the ripples are summarized in Table
470 > \ref{tab:property}. The wavelength range is $15~21$ molecular bodies
471 > and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
472 > $2.2$ for symmetric ripple. These values are consistent to the
473 > experimental results.  Note, that given the lack of structural freedom
474 > in the tails of our model lipids, the amplitudes observed from these
475 > simulations are likely to underestimate of the true amplitudes.
476  
477 < The $P_2$ order paramters (for molecular bodies and head group
478 < dipoles) have been calculated to clarify the ordering in these phases
479 < quantatively. $P_2=1$ means a perfectly ordered structure, and $P_2=0$
480 < implies orientational randomization. Figure \ref{fig:rP2} shows the
481 < $P_2$ order paramter of the dipoles on head group rising with
482 < increasing head group size. When the heads of the lipid molecules are
483 < small, the membrane is flat. The dipolar ordering is essentially
484 < frustrated on orientational ordering in this circumstance. Figure
485 < \ref{} shows the snapshots of the top view for the flat system
486 < ($\sigma_h=1.20\sigma$) and rippled system
487 < ($\sigma_h=1.41\sigma$). The pointing direction of the dipoles on the
488 < head groups are represented by two colored half spheres from blue to
330 < yellow. For flat surfaces, the system obviously shows frustration on
331 < the dipolar ordering, there are kinks on the edge of defferent
332 < domains. Another reason is that the lipids can move independently in
333 < each monolayer, it is not nessasory for the direction of dipoles on
334 < one leaf is consistant to another layer, which makes total order
335 < parameter is relatively low. With increasing head group size, the
336 < surface is corrugated, and dipoles do not move as freely on the
337 < surface. Therefore, the translational freedom of lipids in one layer
338 < is dependent upon the position of lipids in another layer, as a
339 < result, the symmetry of the dipoles on head group in one layer is tied
340 < to the symmetry in the other layer. Furthermore, as the membrane
341 < deforms from two to three dimensions due to the corrugation, the
342 < symmetry of the ordering for the dipoles embedded on each leaf is
343 < broken. The dipoles then self-assemble in a head-tail configuration,
344 < and the order parameter increases dramaticaly. However, the total
345 < polarization of the system is still close to zero. This is strong
346 < evidence that the corrugated structure is an antiferroelectric
347 < state. From the snapshot in Figure \ref{}, the dipoles arrange as
348 < arrays along $Y$ axis and fall into head-to-tail configuration in each
349 < line, but every $3$ or $4$ lines of dipoles change their direction
350 < from neighbour lines. The system shows antiferroelectric
351 < charactoristic as a whole. The orientation of the dipolar is always
352 < perpendicular to the ripple wave vector. These results are consistent
353 < with our previous study on dipolar membranes.
477 > \begin{figure}[htb]
478 > \centering
479 > \includegraphics[width=4in]{topDown}
480 > \caption{Top views of the flat (upper), asymmetric ripple (middle),
481 > and symmetric ripple (lower) phases.  Note that the head-group dipoles
482 > have formed head-to-tail chains in all three of these phases, but in
483 > the two rippled phases, the dipolar chains are all aligned
484 > {\it perpendicular} to the direction of the ripple.  The flat membrane
485 > has multiple point defects in the dipolar orientational ordering, and
486 > the dipolar ordering on the lower leaf of the bilayer can be in a
487 > different direction from the upper leaf.\label{fig:topView}}
488 > \end{figure}
489  
490 < The ordering of the tails is essentially opposite to the ordering of
491 < the dipoles on head group. The $P_2$ order parameter decreases with
492 < increasing head size. This indicates the surface is more curved with
493 < larger head groups. When the surface is flat, all tails are pointing
494 < in the same direction; in this case, all tails are parallel to the
495 < normal of the surface,(making this structure remindcent of the
496 < $L_{\beta}$ phase. Increasing the size of the heads, results in
490 > The principal method for observing orientational ordering in dipolar
491 > or liquid crystalline systems is the $P_2$ order parameter (defined
492 > as $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
493 > eigenvalue of the matrix,
494 > \begin{equation}
495 > {\mathsf{S}} = \frac{1}{N} \sum_i \left(
496 > \begin{array}{ccc}
497 >        u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
498 >        u^{y}_i u^{x}_i  & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
499 >        u^{z}_i u^{x}_i & u^{z}_i u^{y}_i  & u^{z}_i u^{z}_i -\frac{1}{3}
500 > \end{array} \right).
501 > \label{eq:opmatrix}
502 > \end{equation}
503 > Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
504 > for molecule $i$.  (Here, $\hat{\bf u}_i$ can refer either to the
505 > principal axis of the molecular body or to the dipole on the head
506 > group of the molecule.)  $P_2$ will be $1.0$ for a perfectly-ordered
507 > system and near $0$ for a randomized system.  Note that this order
508 > parameter is {\em not} equal to the polarization of the system.  For
509 > example, the polarization of a perfect anti-ferroelectric arrangement
510 > of point dipoles is $0$, but $P_2$ for the same system is $1$.  The
511 > eigenvector of $\mathsf{S}$ corresponding to the largest eigenvalue is
512 > familiar as the director axis, which can be used to determine a
513 > privileged axis for an orientationally-ordered system.  Since the
514 > molecular bodies are perpendicular to the head group dipoles, it is
515 > possible for the director axes for the molecular bodies and the head
516 > groups to be completely decoupled from each other.
517 >
518 > Figure \ref{fig:topView} shows snapshots of bird's-eye views of the
519 > flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.41, 1.35 d$)
520 > bilayers.  The directions of the dipoles on the head groups are
521 > represented with two colored half spheres: blue (phosphate) and yellow
522 > (amino).  For flat bilayers, the system exhibits signs of
523 > orientational frustration; some disorder in the dipolar head-to-tail
524 > chains is evident with kinks visible at the edges between differently
525 > ordered domains.  The lipids can also move independently of lipids in
526 > the opposing leaf, so the ordering of the dipoles on one leaf is not
527 > necessarily consistent with the ordering on the other.  These two
528 > factors keep the total dipolar order parameter relatively low for the
529 > flat phases.
530 >
531 > With increasing head group size, the surface becomes corrugated, and
532 > the dipoles cannot move as freely on the surface. Therefore, the
533 > translational freedom of lipids in one layer is dependent upon the
534 > position of the lipids in the other layer.  As a result, the ordering of
535 > the dipoles on head groups in one leaf is correlated with the ordering
536 > in the other leaf.  Furthermore, as the membrane deforms due to the
537 > corrugation, the symmetry of the allowed dipolar ordering on each leaf
538 > is broken. The dipoles then self-assemble in a head-to-tail
539 > configuration, and the dipolar order parameter increases dramatically.
540 > However, the total polarization of the system is still close to zero.
541 > This is strong evidence that the corrugated structure is an
542 > antiferroelectric state.  It is also notable that the head-to-tail
543 > arrangement of the dipoles is always observed in a direction
544 > perpendicular to the wave vector for the surface corrugation.  This is
545 > a similar finding to what we observed in our earlier work on the
546 > elastic dipolar membranes.\cite{Sun2007}
547 >
548 > The $P_2$ order parameters (for both the molecular bodies and the head
549 > group dipoles) have been calculated to quantify the ordering in these
550 > phases.  Figure \ref{fig:rP2} shows that the $P_2$ order parameter for
551 > the head-group dipoles increases with increasing head group size. When
552 > the heads of the lipid molecules are small, the membrane is nearly
553 > flat. Since the in-plane packing is essentially a close packing of the
554 > head groups, the head dipoles exhibit frustration in their
555 > orientational ordering.
556 >
557 > The ordering trends for the tails are essentially opposite to the
558 > ordering of the head group dipoles. The tail $P_2$ order parameter
559 > {\it decreases} with increasing head size. This indicates that the
560 > surface is more curved with larger head / tail size ratios. When the
561 > surface is flat, all tails are pointing in the same direction (normal
562 > to the bilayer surface).  This simplified model appears to be
563 > exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$
564 > phase.  We have not observed a smectic C gel phase ($L_{\beta'}$) for
565 > this model system.  Increasing the size of the heads results in
566   rapidly decreasing $P_2$ ordering for the molecular bodies.
567 +
568   \begin{figure}[htb]
569   \centering
570   \includegraphics[width=\linewidth]{rP2}
571 < \caption{The $P_2$ order parameter as a funtion of the ratio of
572 < $\sigma_h$ to $\sigma_0$.\label{fig:rP2}}
571 > \caption{The $P_2$ order parameters for head groups (circles) and
572 > molecular bodies (squares) as a function of the ratio of head group
573 > size ($\sigma_h$) to the width of the molecular bodies ($d$). \label{fig:rP2}}
574   \end{figure}
575  
576 < We studied the effects of the interactions between head groups on the
577 < structure of lipid bilayer by changing the strength of the dipole.
578 < Figure \ref{fig:sP2} shows how the $P_2$ order parameter changes with
579 < increasing strength of the dipole. Generally the dipoles on the head
580 < group are more ordered by increase in the strength of the interaction
581 < between heads and are more disordered by decreasing the interaction
582 < stength. When the interaction between the heads is weak enough, the
583 < bilayer structure does not persist; all lipid molecules are solvated
584 < directly in the water. The critial value of the strength of the dipole
585 < depends on the head size. The perfectly flat surface melts at $5$
586 < $0.03$ debye, the asymmetric rippled surfaces melt at $8$ $0.04$
587 < $0.03$ debye, the symmetric rippled surfaces melt at $10$ $0.04$
588 < debye. The ordering of the tails is the same as the ordering of the
589 < dipoles except for the flat phase. Since the surface is already
590 < perfect flat, the order parameter does not change much until the
591 < strength of the dipole is $15$ debye. However, the order parameter
592 < decreases quickly when the strength of the dipole is further
593 < increased. The head groups of the lipid molecules are brought closer
594 < by stronger interactions between them. For a flat surface, a large
595 < amount of free volume between the head groups is available, but when
596 < the head groups are brought closer, the tails will splay outward,
597 < forming an inverse micelle. When $\sigma_h=1.28\sigma_0$, the $P_2$
598 < order parameter decreases slightly after the strength of the dipole is
599 < increased to $16$ debye. For rippled surfaces, there is less free
600 < volume available between the head groups. Therefore there is little
601 < effect on the structure of the membrane due to increasing dipolar
602 < strength. However, the increase of the $P_2$ order parameter implies
603 < the membranes are flatten by the increase of the strength of the
604 < dipole. Unlike other systems that melt directly when the interaction
605 < is weak enough, for $\sigma_h=1.41\sigma_0$, part of the membrane
606 < melts into itself first. The upper leaf of the bilayer becomes totally
607 < interdigitated with the lower leaf. This is different behavior than
608 < what is exhibited with the interdigitated lines in the rippled phase
609 < where only one interdigitated line connects the two leaves of bilayer.
576 > In addition to varying the size of the head groups, we studied the
577 > effects of the interactions between head groups on the structure of
578 > lipid bilayer by changing the strength of the dipoles.  Figure
579 > \ref{fig:sP2} shows how the $P_2$ order parameter changes with
580 > increasing strength of the dipole.  Generally, the dipoles on the head
581 > groups become more ordered as the strength of the interaction between
582 > heads is increased and become more disordered by decreasing the
583 > interaction stength.  When the interaction between the heads becomes
584 > too weak, the bilayer structure does not persist; all lipid molecules
585 > become dispersed in the solvent (which is non-polar in this
586 > molecular-scale model).  The critial value of the strength of the
587 > dipole depends on the size of the head groups.  The perfectly flat
588 > surface becomes unstable below $5$ Debye, while the  rippled
589 > surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
590 >
591 > The ordering of the tails mirrors the ordering of the dipoles {\it
592 > except for the flat phase}. Since the surface is nearly flat in this
593 > phase, the order parameters are only weakly dependent on dipolar
594 > strength until it reaches $15$ Debye.  Once it reaches this value, the
595 > head group interactions are strong enough to pull the head groups
596 > close to each other and distort the bilayer structure. For a flat
597 > surface, a substantial amount of free volume between the head groups
598 > is normally available.  When the head groups are brought closer by
599 > dipolar interactions, the tails are forced to splay outward, forming
600 > first curved bilayers, and then inverted micelles.
601 >
602 > When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
603 > when the strength of the dipole is increased above $16$ debye. For
604 > rippled bilayers, there is less free volume available between the head
605 > groups. Therefore increasing dipolar strength weakly influences the
606 > structure of the membrane.  However, the increase in the body $P_2$
607 > order parameters implies that the membranes are being slightly
608 > flattened due to the effects of increasing head-group attraction.
609 >
610 > A very interesting behavior takes place when the head groups are very
611 > large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the
612 > dipolar strength is relatively weak ($\mu < 9$ Debye). In this case,
613 > the two leaves of the bilayer become totally interdigitated with each
614 > other in large patches of the membrane.   With higher dipolar
615 > strength, the interdigitation is limited to single lines that run
616 > through the bilayer in a direction perpendicular to the ripple wave
617 > vector.
618 >
619   \begin{figure}[htb]
620   \centering
621   \includegraphics[width=\linewidth]{sP2}
622 < \caption{The $P_2$ order parameter as a funtion of the strength of the
623 < dipole.\label{fig:sP2}}
622 > \caption{The $P_2$ order parameters for head group dipoles (a) and
623 > molecular bodies (b) as a function of the strength of the dipoles.
624 > These order parameters are shown for four values of the head group /
625 > molecular width ratio ($\sigma_h / d$). \label{fig:sP2}}
626   \end{figure}
627  
628 < Figure \ref{fig:tP2} shows the dependence of the order parameter on
629 < temperature. The behavior of the $P_2$ order paramter is
630 < straightforward. Systems are more ordered at low temperature, and more
631 < disordered at high temperatures. When the temperature is high enough,
632 < the membranes are instable. For flat surfaces ($\sigma_h=1.20\sigma_0$
633 < and $\sigma_h=1.28\sigma_0$), when the temperature is increased to
634 < $310$, the $P_2$ order parameter increases slightly instead of
635 < decreases like ripple surface. This is an evidence of the frustration
636 < of the dipolar ordering in each leaf of the lipid bilayer, at low
637 < temperature, the systems are locked in a local minimum energy state,
638 < with increase of the temperature, the system can jump out the local
639 < energy well to find the lower energy state which is the longer range
640 < orientational ordering. Like the dipolar ordering of the flat
641 < surfaces, the ordering of the tails of the lipid molecules for ripple
642 < membranes ($\sigma_h=1.35\sigma_0$ and $\sigma_h=1.41\sigma_0$) also
643 < show some nonthermal characteristic. With increase of the temperature,
644 < the $P_2$ order parameter decreases firstly, and increases afterward
645 < when the temperature is greater than $290 K$. The increase of the
646 < $P_2$ order parameter indicates a more ordered structure for the tails
647 < of the lipid molecules which corresponds to a more flat surface. Since
648 < our model lacks the detailed information on lipid tails, we can not
649 < simulate the fluid phase with melted fatty acid chains. Moreover, the
650 < formation of the tilted $L_{\beta'}$ phase also depends on the
434 < organization of fatty groups on tails.
628 > Figure \ref{fig:tP2} shows the dependence of the order parameters on
629 > temperature.  As expected, systems are more ordered at low
630 > temperatures, and more disordered at high temperatures.  All of the
631 > bilayers we studied can become unstable if the temperature becomes
632 > high enough.  The only interesting feature of the temperature
633 > dependence is in the flat surfaces ($\sigma_h=1.20 d$ and
634 > $\sigma_h=1.28 d$).  Here, when the temperature is increased above
635 > $310$K, there is enough jostling of the head groups to allow the
636 > dipolar frustration to resolve into more ordered states.  This results
637 > in a slight increase in the $P_2$ order parameter above this
638 > temperature.
639 >
640 > For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$),
641 > there is a slightly increased orientational ordering in the molecular
642 > bodies above $290$K.  Since our model lacks the detailed information
643 > about the behavior of the lipid tails, this is the closest the model
644 > can come to depicting the ripple ($P_{\beta'}$) to fluid
645 > ($L_{\alpha}$) phase transition.  What we are observing is a
646 > flattening of the rippled structures made possible by thermal
647 > expansion of the tightly-packed head groups.  The lack of detailed
648 > chain configurations also makes it impossible for this model to depict
649 > the ripple to gel ($L_{\beta'}$) phase transition.
650 >
651   \begin{figure}[htb]
652   \centering
653   \includegraphics[width=\linewidth]{tP2}
654 < \caption{The $P_2$ order parameter as a funtion of
655 < temperature.\label{fig:tP2}}
654 > \caption{The $P_2$ order parameters for head group dipoles (a) and
655 > molecular bodies (b) as a function of temperature.
656 > These order parameters are shown for four values of the head group /
657 > molecular width ratio ($\sigma_h / d$).\label{fig:tP2}}
658   \end{figure}
659  
660   \section{Discussion}
661   \label{sec:discussion}
662  
663 + The ripple phases have been observed in our molecular dynamic
664 + simulations using a simple molecular lipid model. The lipid model
665 + consists of an anisotropic interacting dipolar head group and an
666 + ellipsoid shape tail. According to our simulations, the explanation of
667 + the formation for the ripples are originated in the size mismatch
668 + between the head groups and the tails. The ripple phases are only
669 + observed in the studies using larger head group lipid models. However,
670 + there is a mismatch betweent the size of the head groups and the size
671 + of the tails in the simulations of the flat surface. This indicates
672 + the competition between the anisotropic dipolar interaction and the
673 + packing of the tails also plays a major role for formation of the
674 + ripple phase. The larger head groups provide more free volume for the
675 + tails, while these hydrophobic ellipsoids trying to be close to each
676 + other, this gives the origin of the spontanous curvature of the
677 + surface, which is believed as the beginning of the ripple phases. The
678 + lager head groups cause the spontanous curvature inward for both of
679 + leaves of the bilayer. This results in a steric strain when the tails
680 + of two leaves too close to each other. The membrane has to be broken
681 + to release this strain. There are two ways to arrange these broken
682 + curvatures: symmetric and asymmetric ripples. Both of the ripple
683 + phases have been observed in our studies. The difference between these
684 + two ripples is that the bilayer is continuum in the symmetric ripple
685 + phase and is disrupt in the asymmetric ripple phase.
686 +
687 + Dipolar head groups are the key elements for the maintaining of the
688 + bilayer structure. The lipids are solvated in water when lowering the
689 + the strength of the dipole on the head groups. The long range
690 + orientational ordering of the dipoles can be achieved by forming the
691 + ripples, although the dipoles are likely to form head-to-tail
692 + configurations even in flat surface, the frustration prevents the
693 + formation of the long range orientational ordering for dipoles. The
694 + corrugation of the surface breaks the frustration and stablizes the
695 + long range oreintational ordering for the dipoles in the head groups
696 + of the lipid molecules. Many rows of the head-to-tail dipoles are
697 + parallel to each other and adopt the antiferroelectric state as a
698 + whole. This is the first time the organization of the head groups in
699 + ripple phases of the lipid bilayer has been addressed.
700 +
701 + The most important prediction we can make using the results from this
702 + simple model is that if dipolar ordering is driving the surface
703 + corrugation, the wave vectors for the ripples should always found to
704 + be {\it perpendicular} to the dipole director axis.  This prediction
705 + should suggest experimental designs which test whether this is really
706 + true in the phosphatidylcholine $P_{\beta'}$ phases.  The dipole
707 + director axis should also be easily computable for the all-atom and
708 + coarse-grained simulations that have been published in the literature.
709 +
710 + Although our model is simple, it exhibits some rich and unexpected
711 + behaviors.  It would clearly be a closer approximation to the reality
712 + if we allowed greater translational freedom to the dipoles and
713 + replaced the somewhat artificial lattice packing and the harmonic
714 + elastic tension with more realistic molecular modeling potentials.
715 + What we have done is to present a simple model which exhibits bulk
716 + non-thermal corrugation, and our explanation of this rippling
717 + phenomenon will help us design more accurate molecular models for
718 + corrugated membranes and experiments to test whether rippling is
719 + dipole-driven or not.
720 +
721 + \newpage
722   \bibliography{mdripple}
723   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines