ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ripple2/ripple.tex
(Generate patch)

Comparing trunk/ripple2/ripple.tex (file contents):
Revision 3091 by xsun, Tue Dec 26 21:50:50 2006 UTC vs.
Revision 3101 by gezelter, Fri Dec 29 20:22:32 2006 UTC

# Line 1 | Line 1
1 < \documentclass[aps,pre,endfloats*,preprint,amssymb,showpacs]{revtex4}
2 < \usepackage{epsfig}
1 > %\documentclass[aps,pre,twocolumn,amssymb,showpacs,floatfix]{revtex4}
2 > \documentclass[aps,pre,preprint,amssymb,showpacs]{revtex4}
3 > \usepackage{graphicx}
4  
5   \begin{document}
6   \renewcommand{\thefootnote}{\fnsymbol{footnote}}
# Line 9 | Line 10
10  
11   \title{Spontaneous Corrugation of Dipolar Membranes}
12   \author{Xiuquan Sun and J. Daniel Gezelter}
13 < \email[]{E-mail: gezelter@nd.edu}
13 > \email[E-mail:]{gezelter@nd.edu}
14   \affiliation{Department of Chemistry and Biochemistry,\\
15   University of Notre Dame, \\
16   Notre Dame, Indiana 46556}
# Line 17 | Line 18 | We present a simple model for dipolar membranes that g
18   \date{\today}
19  
20   \begin{abstract}
21 < We present a simple model for dipolar membranes that gives
21 > We present a simple model for dipolar elastic membranes that gives
22   lattice-bound point dipoles complete orientational freedom as well as
23   translational freedom along one coordinate (out of the plane of the
24 < membrane).  There is an additional harmonic surface tension which
25 < binds each of the dipoles to the six nearest neighbors on either
26 < hexagonal or distorted-hexagonal lattices.  The translational freedom
27 < of the dipoles allows hexagonal lattices to find states that break out
28 < of the normal orientational disorder of frustrated configurations and
29 < which are stabilized by long-range antiferroelectric ordering.  In
30 < order to break out of the frustrated states, the dipolar membranes
31 < form corrugated or ``rippled'' phases that make the lattices
32 < effectively non-hexagonal.  We observe three common features of the
33 < corrugated dipolar membranes: 1) the corrugated phases develop easily
34 < when hosted on hexagonal lattices, 2) the wave vectors for the surface
35 < ripples are always found to be perpendicular to the dipole director
36 < axis, and 3) on hexagonal lattices, the dipole director axis is found
37 < to be parallel to any of the three equivalent lattice directions.
24 > membrane).  There is an additional harmonic term which binds each of
25 > the dipoles to the six nearest neighbors on either triangular or
26 > distorted lattices.  The translational freedom of the dipoles allows
27 > triangular lattices to find states that break out of the normal
28 > orientational disorder of frustrated configurations and which are
29 > stabilized by long-range anti-ferroelectric ordering.  In order to
30 > break out of the frustrated states, the dipolar membranes form
31 > corrugated or ``rippled'' phases that make the lattices effectively
32 > non-triangular.  We observe three common features of the corrugated
33 > dipolar membranes: 1) the corrugated phases develop easily when hosted
34 > on triangular lattices, 2) the wave vectors for the surface ripples
35 > are always found to be perpendicular to the dipole director axis, and
36 > 3) on triangular lattices, the dipole director axis is found to be
37 > parallel to any of the three equivalent lattice directions.
38   \end{abstract}
39  
40   \pacs{68.03.Hj, 82.20.Wt}
# Line 42 | Line 43 | There has been intense recent interest in the phase be
43  
44   \section{Introduction}
45   \label{Int}
45 There has been intense recent interest in the phase behavior of
46 dipolar
47 fluids.\cite{Tlusty00,Teixeira00,Tavares02,Duncan04,Holm05,Duncan06}
48 Due to the anisotropic interactions between dipoles, dipolar fluids
49 can present anomalous phase behavior.  Examples of condensed-phase
50 dipolar systems include ferrofluids, electro-rheological fluids, and
51 even biological membranes.  Computer simulations have provided useful
52 information on the structural features and phase transition of the
53 dipolar fluids. Simulation results indicate that at low densities,
54 these fluids spontaneously organize into head-to-tail dipolar
55 ``chains''.\cite{Teixeira00,Holm05} At low temperatures, these chains
56 and rings prevent the occurrence of a liquid-gas phase transition.
57 However, Tlusty and Safran showed that there is a defect-induced phase
58 separation into a low-density ``chain'' phase and a higher density
59 Y-defect phase.\cite{Tlusty00} Recently, inspired by experimental
60 studies on monolayers of dipolar fluids, theoretical models using
61 two-dimensional dipolar soft spheres have appeared in the literature.
62 Tavares {\it et al.} tested their theory for chain and ring length
63 distributions in two dimensions and carried out Monte Carlo
64 simulations in the low-density phase.\cite{Tavares02} Duncan and Camp
65 performed dynamical simulations on two-dimensional dipolar fluids to
66 study transport and orientational dynamics in these
67 systems.\cite{Duncan04} They have recently revisited two-dimensional
68 systems to study the kinetic conditions for the defect-induced
69 condensation into the Y-defect phase.\cite{Duncan06}
46  
47 < Although they are not traditionally classified as 2-dimensional
48 < dipolar fluids, hydrated lipids aggregate spontaneously to form
49 < bilayers which exhibit a variety of phases depending on their
50 < temperatures and compositions.  At high temperatures, the fluid
51 < ($L_{\alpha}$) phase of Phosphatidylcholine (PC) lipids closely
52 < resembles a dipolar fluid.  However, at lower temperatures, packing of
53 < the molecules becomes important, and the translational freedom of
54 < lipid molecules is thought to be substantially restricted.  A
55 < corrugated or ``rippled'' phase ($P_{\beta'}$) appears as an
56 < intermediate phase between the gel ($L_\beta$) and fluid
57 < ($L_{\alpha}$) phases for relatively pure phosphatidylcholine (PC)
58 < bilayers.  The $P_{\beta'}$ phase has attracted substantial
59 < experimental interest over the past 30 years. Most structural
60 < information of the ripple phase has been obtained by the X-ray
61 < diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
47 > The properties of polymeric membranes are known to depend sensitively
48 > on the details of the internal interactions between the constituent
49 > monomers.  A flexible membrane will always have a competition between
50 > the energy of curvature and the in-plane stretching energy and will be
51 > able to buckle in certain limits of surface tension and
52 > temperature.\cite{Safran94} The buckling can be non-specific and
53 > centered at dislocation~\cite{Seung1988} or grain-boundary
54 > defects,\cite{Carraro1993} or it can be directional and cause long
55 > ``roof-tile'' or tube-like structures to appear in
56 > partially-polymerized phospholipid vesicles.\cite{Mutz1991}
57 >
58 > One would expect that anisotropic local interactions could lead to
59 > interesting properties of the buckled membrane.  We report here on the
60 > buckling behavior of a membrane composed of harmonically-bound, but
61 > freely-rotating electrostatic dipoles.  The dipoles have strongly
62 > anisotropic local interactions and the membrane exhibits coupling
63 > between the buckling and the long-range ordering of the dipoles.
64 >
65 > Buckling behavior in liquid crystalline and biological membranes is a
66 > well-known phenomenon.  Relatively pure phosphatidylcholine (PC)
67 > bilayers form a corrugated or ``rippled'' phase ($P_{\beta'}$) which
68 > appears as an intermediate phase between the gel ($L_\beta$) and fluid
69 > ($L_{\alpha}$) phases.  The $P_{\beta'}$ phase has attracted
70 > substantial experimental interest over the past 30 years. Most
71 > structural information of the ripple phase has been obtained by the
72 > X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
73   microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
74   et al.} used atomic force microscopy (AFM) to observe ripple phase
75   morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
76   experimental results provide strong support for a 2-dimensional
77 < hexagonal packing lattice of the lipid molecules within the ripple
77 > triangular packing lattice of the lipid molecules within the ripple
78   phase.  This is a notable change from the observed lipid packing
79 < within the gel phase.~\cite{Cevc87}
79 > within the gel phase.~\cite{Cevc87} There have been a number of
80 > theoretical
81 > approaches~\cite{Marder84,Goldstein88,McCullough90,Lubensky93,Misbah98,Heimburg00,Kubica02,Bannerjee02}
82 > (and some heroic
83 > simulations~\cite{Ayton02,Jiang04,Brannigan04a,deVries05,deJoannis06})
84 > undertaken to try to explain this phase, but to date, none have looked
85 > specifically at the contribution of the dipolar character of the lipid
86 > head groups towards this corrugation.  Lipid chain interdigitation
87 > certainly plays a major role, and the structures of the ripple phase
88 > are highly ordered.  The model we investigate here lacks chain
89 > interdigitation (as well as the chains themselves!) and will not be
90 > detailed enough to rule in favor of (or against) any of these
91 > explanations for the $P_{\beta'}$ phase.
92  
93 < Although the results of dipolar fluid simulations can not be directly
94 < mapped onto the phases of lipid bilayers, the rich behaviors exhibited
95 < by simple dipolar models can give us some insight into the corrugation
96 < phenomenon of the $P_{\beta'}$ phase.  There have been a number of
97 < theoretical approaches (and some heroic simulations) undertaken to try
98 < to explain this phase, but to date, none have looked specifically at
99 < the contribution of the dipolar character of the lipid head groups
100 < towards this corrugation.  Before we present our simple model, we will
101 < briefly survey the previous theoretical work on this topic.
102 <
103 < The theoretical models that have been put forward to explain the
104 < formation of the $P_{\beta'}$ phase have presented a number of
105 < conflicting but intriguing explanations. Marder {\it et al.} used a
107 < curvature-dependent Landau-de Gennes free-energy functional to predict
108 < a rippled phase.~\cite{Marder84} This model and other related
109 < continuum models predict higher fluidity in convex regions and that
110 < concave portions of the membrane correspond to more solid-like
111 < regions.  Carlson and Sethna used a packing-competition model (in
112 < which head groups and chains have competing packing energetics) to
113 < predict the formation of a ripple-like phase.  Their model predicted
114 < that the high-curvature portions have lower-chain packing and
115 < correspond to more fluid-like regions.  Goldstein and Leibler used a
116 < mean-field approach with a planar model for {\em inter-lamellar}
117 < interactions to predict rippling in multilamellar
118 < phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
119 < anisotropy of the nearest-neighbor interactions} coupled to
120 < hydrophobic constraining forces which restrict height differences
121 < between nearest neighbors is the origin of the ripple
122 < phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
123 < theory for tilt order and curvature of a single membrane and concluded
124 < that {\em coupling of molecular tilt to membrane curvature} is
125 < responsible for the production of ripples.~\cite{Lubensky93} Misbah,
126 < Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
127 < interactions} can lead to ripple instabilities.~\cite{Misbah98}
128 < Heimburg presented a {\em coexistence model} for ripple formation in
129 < which he postulates that fluid-phase line defects cause sharp
130 < curvature between relatively flat gel-phase regions.~\cite{Heimburg00}
131 < Kubica has suggested that a lattice model of polar head groups could
132 < be valuable in trying to understand bilayer phase
133 < formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of
134 < lamellar stacks of hexagonal lattices to show that large headgroups
135 < and molecular tilt with respect to the membrane normal vector can
136 < cause bulk rippling.~\cite{Bannerjee02}
137 <
138 < Large-scale molecular dynamics simulations have also been performed on
139 < rippled phases using united atom as well as molecular scale
140 < models. De~Vries {\it et al.} studied the structure of lecithin ripple
141 < phases via molecular dynamics and their simulations seem to support
142 < the coexistence models (i.e. fluid-like chain dynamics was observed in
143 < the kink regions).~\cite{deVries05} A similar coarse-grained approach
144 < has been used to study the line tension of bilayer
145 < edges.\cite{Jiang04,deJoannis06} Ayton and Voth have found significant
146 < undulations in zero-surface-tension states of membranes simulated via
147 < dissipative particle dynamics, but their results are consistent with
148 < purely thermal undulations.~\cite{Ayton02} Brannigan, Tamboli and
149 < Brown have used a molecular scale model to elucidate the role of
150 < molecular shape on membrane phase behavior and
151 < elasticity.~\cite{Brannigan04b} They have also observed a buckled
152 < hexatic phase with strong tail and moderate alignment
153 < attractions.~\cite{Brannigan04a}
93 > Membranes containing electrostatic dipoles can also exhibit the
94 > flexoelectric effect,\cite{Todorova2004,Harden2006,Petrov2006} which
95 > is the ability of mechanical deformations to result in electrostatic
96 > organization of the membrane.  This phenomenon is a curvature-induced
97 > membrane polarization which can lead to potential differences across a
98 > membrane.  Reverse flexoelectric behavior (in which applied currents
99 > effect membrane curvature) has also been observed.  Explanations of
100 > the details of these effects have typically utilized membrane
101 > polarization perpendicular to the face of the
102 > membrane,\cite{Petrov2006} and the effect has been observed in both
103 > biological,\cite{Raphael2000} bent-core liquid
104 > crystalline,\cite{Harden2006} and polymer-dispersed liquid crystalline
105 > membranes.\cite{Todorova2004}
106  
107   The problem with using atomistic and even coarse-grained approaches to
108 < study this phenomenon is that only a relatively small number of
109 < periods of the corrugation (i.e. one or two) can be realistically
110 < simulated given current technology.  Also, simulations of lipid
111 < bilayers are traditionally carried out with periodic boundary
108 > study membrane buckling phenomena is that only a relatively small
109 > number of periods of the corrugation (i.e. one or two) can be
110 > realistically simulated given current technology.  Also, simulations
111 > of lipid bilayers are traditionally carried out with periodic boundary
112   conditions in two or three dimensions and these have the potential to
113   enhance the periodicity of the system at that wavelength.  To avoid
114   this pitfall, we are using a model which allows us to have
115   sufficiently large systems so that we are not causing artificial
116   corrugation through the use of periodic boundary conditions.
117  
118 < At the other extreme in density from the traditional simulations of
119 < dipolar fluids is the behavior of dipoles locked on regular lattices.
120 < Ferroelectric states (with long-range dipolar order) can be observed
121 < in dipolar systems with non-hexagonal packings.  However, {\em
122 < hexagonally}-packed 2-D dipolar systems are inherently frustrated and
123 < one would expect a dipolar-disordered phase to be the lowest free
124 < energy configuration.  Therefore, it would seem unlikely that a
125 < frustrated lattice in a dipolar-disordered state could exhibit the
126 < long-range periodicity in the range of 100-600 \AA (as exhibited in
127 < the ripple phases studied by Kaasgard {\it et
128 < al.}).~\cite{Kaasgaard03}
118 > The simplest dipolar membrane is one in which the dipoles are located
119 > on fixed lattice sites. Ferroelectric states (with long-range dipolar
120 > order) can be observed in dipolar systems with non-triangular
121 > packings.  However, {\em triangularly}-packed 2-D dipolar systems are
122 > inherently frustrated and one would expect a dipolar-disordered phase
123 > to be the lowest free energy
124 > configuration.\cite{Toulouse1977,Marland1979} Dipolar lattices already
125 > have rich phase behavior, but in order to allow the membrane to
126 > buckle, a single degree of freedom (translation normal to the membrane
127 > face) must be added to each of the dipoles.  It would also be possible
128 > to allow complete translational freedom.  This approach
129 > is similar in character to a number of elastic Ising models that have
130 > been developed to explain interesting mechanical properties in
131 > magnetic alloys.\cite{Renard1966,Zhu2005,Zhu2006,Jiang2006}
132  
133 < Is there an intermediate model between the low-density dipolar fluids
134 < and the rigid lattice models which has the potential to exhibit the
135 < corrugation phenomenon of the $P_{\beta'}$ phase?  What we present
181 < here is an attempt to find a simple dipolar model which will exhibit
182 < this behavior.  We are using a modified XYZ lattice model; details of
183 < the model can be found in section
133 > What we present here is an attempt to find the simplest dipolar model
134 > which will exhibit buckling behavior.  We are using a modified XYZ
135 > lattice model; details of the model can be found in section
136   \ref{sec:model}, results of Monte Carlo simulations using this model
137   are presented in section
138   \ref{sec:results}, and section \ref{sec:discussion} contains our conclusions.
# Line 190 | Line 142 | of a two-dimensional dipolar medium.  Since molecules
142  
143   The point of developing this model was to arrive at the simplest
144   possible theoretical model which could exhibit spontaneous corrugation
145 < of a two-dimensional dipolar medium.  Since molecules in the ripple
146 < phase have limited translational freedom, we have chosen a lattice to
147 < support the dipoles in the x-y plane.  The lattice may be either
148 < hexagonal (lattice constants $a/b = \sqrt{3}$) or non-hexagonal.
149 < However, each dipole has 3 degrees of freedom.  They may move freely
150 < {\em out} of the x-y plane (along the $z$ axis), and they have
151 < complete orientational freedom ($0 <= \theta <= \pi$, $0 <= \phi < 2
145 > of a two-dimensional dipolar medium.  Since molecules in polymerized
146 > membranes and in the $P_{\beta'}$ ripple phase have limited
147 > translational freedom, we have chosen a lattice to support the dipoles
148 > in the x-y plane.  The lattice may be either triangular (lattice
149 > constants $a/b =
150 > \sqrt{3}$) or distorted.  However, each dipole has 3 degrees of
151 > freedom.  They may move freely {\em out} of the x-y plane (along the
152 > $z$ axis), and they have complete orientational freedom ($0 <= \theta
153 > <= \pi$, $0 <= \phi < 2
154   \pi$).  This is essentially a modified X-Y-Z model with translational
155   freedom along the z-axis.
156  
157   The potential energy of the system,
158 < \begin{equation}
159 < V = \sum_i \left( \sum_{j \in NN_i}^6
206 < \frac{k_r}{2}\left( r_{ij}-\sigma \right)^2  +  \sum_{j>i} \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[
158 > \begin{eqnarray}
159 > V = \sum_i & & \left( \sum_{j>i} \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[
160   {\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j} -
161   3({\mathbf{\hat u}_i} \cdot {\mathbf{\hat
162   r}_{ij}})({\mathbf{\hat u}_j} \cdot {\mathbf{\hat r}_{ij}})\right]
163 < \right)
163 > \right. \nonumber \\
164 > & & \left. + \sum_{j \in NN_i}^6 \frac{k_r}{2}\left(
165 > r_{ij}-\sigma \right)^2 \right)
166   \label{eq:pot}
167 < \end{equation}
167 > \end{eqnarray}
168  
169   In this potential, $\mathbf{\hat u}_i$ is the unit vector pointing
170   along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
# Line 227 | Line 182 | k_r / 2}$).
182   reduced distances, ($r^{*} = r / \sigma$), temperatures ($T^{*} = 2
183   k_B T / k_r \sigma^2$), densities ($\rho^{*} = N \sigma^2 / L_x L_y$),
184   and dipole moments ($\mu^{*} = \mu / \sqrt{4 \pi \epsilon_0 \sigma^5
185 < k_r / 2}$).
185 > k_r / 2}$).  It should be noted that the density ($\rho^{*}$) depends
186 > only on the mean particle spacing in the $x-y$ plane; the lattice is
187 > fully populated.
188  
189   To investigate the phase behavior of this model, we have performed a
190   series of Metropolis Monte Carlo simulations of moderately-sized (34.3
191 < $\sigma$ on a side) patches of membrane hosted on both hexagonal
192 < ($\gamma = a/b = \sqrt{3}$) and non-hexagonal ($\gamma \neq \sqrt{3}$)
191 > $\sigma$ on a side) patches of membrane hosted on both triangular
192 > ($\gamma = a/b = \sqrt{3}$) and distorted ($\gamma \neq \sqrt{3}$)
193   lattices.  The linear extent of one edge of the monolayer was $20 a$
194   and the system was kept roughly square. The average distance that
195   coplanar dipoles were positioned from their six nearest neighbors was
196 < 1 $\sigma$ (on both hexagonal and non-hexagonal lattices).  Typical
197 < system sizes were 1360 dipoles for the hexagonal lattices and 840-2800
198 < dipoles for the non-hexagonal lattices.  Periodic boundary conditions
199 < were used, and the cutoff for the dipole-dipole interaction was set to
200 < 4.3 $\sigma$.  All parameters ($T^{*}$, $\mu^{*}$, and $\gamma$) were
201 < varied systematically to study the effects of these parameters on the
202 < formation of ripple-like phases.
196 > 1 $\sigma$ (on both triangular and distorted lattices).  Typical
197 > system sizes were 1360 dipoles for the triangular lattices and
198 > 840-2800 dipoles for the distorted lattices.  Two-dimensional periodic
199 > boundary conditions were used, and the cutoff for the dipole-dipole
200 > interaction was set to 4.3 $\sigma$.  This cutoff is roughly 2.5 times
201 > the typical real-space electrostatic cutoff for molecular systems.
202 > Since dipole-dipole interactions decay rapidly with distance, and
203 > since the intrinsic three-dimensional periodicity of the Ewald sum can
204 > give artifacts in 2-d systems, we have chosen not to use it in these
205 > calculations.  Although the Ewald sum has been reformulated to handle
206 > 2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} these
207 > methods are computationally expensive,\cite{Spohr97,Yeh99} and are not
208 > necessary in this case.  All parameters ($T^{*}$, $\mu^{*}$, and
209 > $\gamma$) were varied systematically to study the effects of these
210 > parameters on the formation of ripple-like phases.
211  
212   \section{Results and Analysis}
213   \label{sec:results}
# Line 265 | Line 230 | the polarization of the perfect antiferroelectric syst
230   for dipole $i$.  $P_2$ will be $1.0$ for a perfectly-ordered system
231   and near $0$ for a randomized system.  Note that this order parameter
232   is {\em not} equal to the polarization of the system.  For example,
233 < the polarization of the perfect antiferroelectric system is $0$, but
234 < $P_2$ for an antiferroelectric system is $1$.  The eigenvector of
233 > the polarization of the perfect anti-ferroelectric system is $0$, but
234 > $P_2$ for an anti-ferroelectric system is $1$.  The eigenvector of
235   $\mathsf{S}$ corresponding to the largest eigenvalue is familiar as
236   the director axis, which can be used to determine a privileged dipolar
237   axis for dipole-ordered systems.  The top panel in Fig. \ref{phase}
238   shows the values of $P_2$ as a function of temperature for both
239 < hexagonal ($\gamma = 1.732$) and non-hexagonal ($\gamma=1.875$)
239 > triangular ($\gamma = 1.732$) and distorted ($\gamma=1.875$)
240   lattices.
241  
242 < \begin{figure}[ht]
243 < \centering
244 < \caption{Top panel: The $P_2$ dipolar order parameter as a function of
245 < temperature for both hexagonal ($\gamma = 1.732$) and non-hexagonal
246 < ($\gamma = 1.875$) lattices.  Bottom Panel: The phase diagram for the
247 < dipolar membrane model.  The line denotes the division between the
248 < dipolar ordered (antiferroelectric) and disordered phases.  An
249 < enlarged view near the hexagonal lattice is shown inset.}
285 < \includegraphics[width=\linewidth]{phase.pdf}
286 < \label{phase}
242 > \begin{figure}
243 > \includegraphics[width=\linewidth]{phase}
244 > \caption{\label{phase} Top panel: The $P_2$ dipolar order parameter as
245 > a function of temperature for both triangular ($\gamma = 1.732$) and
246 > distorted ($\gamma = 1.875$) lattices.  Bottom Panel: The phase
247 > diagram for the dipolar membrane model.  The line denotes the division
248 > between the dipolar ordered (anti-ferroelectric) and disordered phases.
249 > An enlarged view near the triangular lattice is shown inset.}
250   \end{figure}
251  
252   There is a clear order-disorder transition in evidence from this data.
253 < Both the hexagonal and non-hexagonal lattices have dipolar-ordered
253 > Both the triangular and distorted lattices have dipolar-ordered
254   low-temperature phases, and orientationally-disordered high
255 < temperature phases.  The coexistence temperature for the hexagonal
256 < lattice is significantly lower than for the non-hexagonal lattices,
257 < and the bulk polarization is approximately $0$ for both dipolar
258 < ordered and disordered phases.  This gives strong evidence that the
259 < dipolar ordered phase is antiferroelectric.  We have repeated the
260 < Monte Carlo simulations over a wide range of lattice ratios ($\gamma$)
261 < to generate a dipolar order/disorder phase diagram.  The bottom panel
262 < in Fig. \ref{phase} shows that the hexagonal lattice is a
263 < low-temperature cusp in the $T^{*}-\gamma$ phase diagram.
255 > temperature phases.  The coexistence temperature for the triangular
256 > lattice is significantly lower than for the distorted lattices, and
257 > the bulk polarization is approximately $0$ for both dipolar ordered
258 > and disordered phases.  This gives strong evidence that the dipolar
259 > ordered phase is anti-ferroelectric.  We have verified that this
260 > dipolar ordering transition is not a function of system size by
261 > performing identical calculations with systems twice as large.  The
262 > transition is equally smooth at all system sizes that were studied.
263 > Additionally, we have repeated the Monte Carlo simulations over a wide
264 > range of lattice ratios ($\gamma$) to generate a dipolar
265 > order/disorder phase diagram.  The bottom panel in Fig. \ref{phase}
266 > shows that the triangular lattice is a low-temperature cusp in the
267 > $T^{*}-\gamma$ phase diagram.
268  
269 < This phase diagram is remarkable in that it shows an antiferroelectric
270 < phase near $\gamma=1.732$ where one would expect lattice frustration
271 < to result in disordered phases at all temperatures.  Observations of
272 < the configurations in this phase show clearly that the system has
273 < accomplished dipolar orderering by forming large ripple-like
274 < structures.  We have observed antiferroelectric ordering in all three
275 < of the equivalent directions on the hexagonal lattice, and the dipoles
276 < have been observed to organize perpendicular to the membrane normal
277 < (in the plane of the membrane).  It is particularly interesting to
278 < note that the ripple-like structures have also been observed to
279 < propagate in the three equivalent directions on the lattice, but the
280 < {\em direction of ripple propagation is always perpendicular to the
281 < dipole director axis}.  A snapshot of a typical antiferroelectric
282 < rippled structure is shown in Fig. \ref{fig:snapshot}.
269 > This phase diagram is remarkable in that it shows an
270 > anti-ferroelectric phase near $\gamma=1.732$ where one would expect
271 > lattice frustration to result in disordered phases at all
272 > temperatures.  Observations of the configurations in this phase show
273 > clearly that the system has accomplished dipolar ordering by forming
274 > large ripple-like structures.  We have observed anti-ferroelectric
275 > ordering in all three of the equivalent directions on the triangular
276 > lattice, and the dipoles have been observed to organize perpendicular
277 > to the membrane normal (in the plane of the membrane).  It is
278 > particularly interesting to note that the ripple-like structures have
279 > also been observed to propagate in the three equivalent directions on
280 > the lattice, but the {\em direction of ripple propagation is always
281 > perpendicular to the dipole director axis}.  A snapshot of a typical
282 > anti-ferroelectric rippled structure is shown in
283 > Fig. \ref{fig:snapshot}.
284  
285 < \begin{figure}[ht]
286 < \centering
287 < \caption{Top and Side views of a representative configuration for the
288 < dipolar ordered phase supported on the hexagonal lattice. Note the
289 < antiferroelectric ordering and the long wavelength buckling of the
290 < membrane.  Dipolar ordering has been observed in all three equivalent
291 < directions on the hexagonal lattice, and the ripple direction is
292 < always perpendicular to the director axis for the dipoles.}
293 < \includegraphics[width=5.5in]{snapshot.pdf}
326 < \label{fig:snapshot}
285 > \begin{figure}
286 > \includegraphics[width=\linewidth]{snapshot}
287 > \caption{\label{fig:snapshot} Top and Side views of a representative
288 > configuration for the dipolar ordered phase supported on the
289 > triangular lattice. Note the anti-ferroelectric ordering and the long
290 > wavelength buckling of the membrane.  Dipolar ordering has been
291 > observed in all three equivalent directions on the triangular lattice,
292 > and the ripple direction is always perpendicular to the director axis
293 > for the dipoles.}
294   \end{figure}
295  
296 + Although the snapshot in Fig. \ref{fig:snapshot} gives the appearance
297 + of three-row stair-like structures, these appear to be transient.  On
298 + average, the corrugation of the membrane is a relatively smooth,
299 + long-wavelength phenomenon, with occasional steep drops between
300 + adjacent lines of anti-aligned dipoles.
301 +
302 + The height-dipole correlation function ($C_{\textrm{hd}}(r, \cos
303 + \theta)$) makes the connection between dipolar ordering and the wave
304 + vector of the ripple even more explicit.  $C_{\textrm{hd}}(r, \cos
305 + \theta)$ is an angle-dependent pair distribution function. The angle
306 + ($\theta$) is the angle between the intermolecular vector
307 + $\vec{r}_{ij}$ and direction of dipole $i$,
308 + \begin{equation}
309 + C_{\textrm{hd}} = \frac{\langle \frac{1}{n(r)} \sum_{i}\sum_{j>i}
310 + h_i \cdot h_j \delta(r - r_{ij}) \delta(\cos \theta_{ij} -
311 + \cos \theta)\rangle} {\langle h^2 \rangle}
312 + \end{equation}
313 + where $\cos \theta_{ij} = \hat{\mu}_{i} \cdot \hat{r}_{ij}$ and
314 + $\hat{r}_{ij} = \vec{r}_{ij} / r_{ij}$.  $n(r)$ is the number of
315 + dipoles found in a cylindrical shell between $r$ and $r+\delta r$ of
316 + the central particle. Fig. \ref{fig:CrossCorrelation} shows contours
317 + of this correlation function for both anti-ferroelectric, rippled
318 + membranes as well as for the dipole-disordered portion of the phase
319 + diagram.
320 +
321 + \begin{figure}
322 + \includegraphics[width=\linewidth]{hdc}
323 + \caption{\label{fig:CrossCorrelation} Contours of the height-dipole
324 + correlation function as a function of the dot product between the
325 + dipole ($\hat{\mu}$) and inter-dipole separation vector ($\hat{r}$)
326 + and the distance ($r$) between the dipoles.  Perfect height
327 + correlation (contours approaching 1) are present in the ordered phase
328 + when the two dipoles are in the same head-to-tail line.
329 + Anti-correlation (contours below 0) is only seen when the inter-dipole
330 + vector is perpendicular to the dipoles.  In the dipole-disordered
331 + portion of the phase diagram, there is only weak correlation in the
332 + dipole direction and this correlation decays rapidly to zero for
333 + intermolecular vectors that are not dipole-aligned.}
334 + \end{figure}
335 +
336 + The height-dipole correlation function gives a map of how the topology
337 + of the membrane surface varies with angular deviation around a given
338 + dipole.  The upper panel of Fig. \ref{fig:CrossCorrelation} shows that
339 + in the anti-ferroelectric phase, the dipole heights are strongly
340 + correlated for dipoles in head-to-tail arrangements, and this
341 + correlation persists for very long distances (up to 15 $\sigma$).  For
342 + portions of the membrane located perpendicular to a given dipole, the
343 + membrane height becomes anti-correlated at distances of 10 $\sigma$.
344 + The correlation function is relatively smooth; there are no steep
345 + jumps or steps, so the stair-like structures in
346 + Fig. \ref{fig:snapshot} are indeed transient and disappear when
347 + averaged over many configurations.  In the dipole-disordered phase,
348 + the height-dipole correlation function is relatively flat (and hovers
349 + near zero).  The only significant height correlations are for axial
350 + dipoles at very short distances ($r \approx
351 + \sigma$).
352 +
353   \subsection{Discriminating Ripples from Thermal Undulations}
354  
355   In order to be sure that the structures we have observed are actually
# Line 336 | Line 360 | where $h(\vec{r})$ is the height of the membrane at lo
360   h(\vec{r}) e^{-i \vec{q}\cdot\vec{r}}
361   \end{equation}
362   where $h(\vec{r})$ is the height of the membrane at location $\vec{r}
363 < = (x,y)$.~\cite{Safran94} In simple (and more complicated) elastic
364 < continuum models, Brannigan {\it et al.} have shown that in the $NVT$
365 < ensemble, the absolute value of the undulation spectrum can be
342 < written,
363 > = (x,y)$.~\cite{Safran94,Seifert97} In simple (and more complicated)
364 > elastic continuum models, it can shown that in the $NVT$ ensemble, the
365 > absolute value of the undulation spectrum can be written,
366   \begin{equation}
367 < \langle | h(q)|^2 \rangle_{NVT} = \frac{k_B T}{k_c |\vec{q}|^4 +
368 < \tilde{\gamma}|\vec{q}|^2},
367 > \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{k_c q^4 +
368 > \gamma q^2},
369   \label{eq:fit}
370   \end{equation}
371 < where $k_c$ is the bending modulus for the membrane, and
372 < $\tilde{\gamma}$ is the mechanical surface
373 < tension.~\cite{Brannigan04b}
371 > where $k_c$ is the bending modulus for the membrane, and $\gamma$ is
372 > the mechanical surface tension.~\cite{Safran94} The systems studied in
373 > this paper have essentially zero bending moduli ($k_c$) and relatively
374 > large mechanical surface tensions ($\gamma$), so a much simpler form
375 > can be written,
376 > \begin{equation}
377 > \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{\gamma q^2},
378 > \label{eq:fit2}
379 > \end{equation}
380  
381   The undulation spectrum is computed by superimposing a rectangular
382   grid on top of the membrane, and by assigning height ($h(\vec{r})$)
# Line 355 | Line 384 | h(q)|^2 \rangle$.
384   given $\vec{r}+d\vec{r}$ grid area.  Empty grid pixels are assigned
385   height values by interpolation from the nearest neighbor pixels.  A
386   standard 2-d Fourier transform is then used to obtain $\langle |
387 < h(q)|^2 \rangle$.
387 > h(q)|^2 \rangle$.  Alternatively, since the dipoles sit on a Bravais
388 > lattice, one could use the heights of the lattice points themselves as
389 > the grid for the Fourier transform (without interpolating to a square
390 > grid).  However, if lateral translational freedom is added to this
391 > model (a likely extension), an interpolated grid method for computing
392 > undulation spectra will be required.
393  
394 < The systems studied in this paper have relatively small bending moduli
395 < ($k_c$) and relatively large mechanical surface tensions
396 < ($\tilde{\gamma}$).  In practice, the best fits to our undulation
397 < spectra are obtained by approximating the value of $k_c$ to 0.  In
398 < Fig. \ref{fig:fit} we show typical undulation spectra for two
399 < different regions of the phase diagram along with their fits from the
400 < Landau free energy approach (Eq. \ref{eq:fit}).  In the
401 < high-temperature disordered phase, the Landau fits can be nearly
402 < perfect, and from these fits we can estimate the bending modulus and
403 < the mechanical surface tension.
394 > As mentioned above, the best fits to our undulation spectra are
395 > obtained by setting the value of $k_c$ to 0.  In Fig. \ref{fig:fit} we
396 > show typical undulation spectra for two different regions of the phase
397 > diagram along with their fits from the Landau free energy approach
398 > (Eq. \ref{eq:fit2}).  In the high-temperature disordered phase, the
399 > Landau fits can be nearly perfect, and from these fits we can estimate
400 > the tension in the surface.  In reduced units, typical values of
401 > $\gamma^{*} = \gamma / \epsilon = 2500$ are obtained for the
402 > disordered phase ($\gamma^{*} = 2551.7$ in the top panel of
403 > Fig. \ref{fig:fit}).
404  
405 < For the dipolar-ordered hexagonal lattice near the coexistence
406 < temperature, however, we observe long wavelength undulations that are
407 < far outliers to the fits.  That is, the Landau free energy fits are
408 < well within error bars for all other points, but can be off by {\em
409 < orders of magnitude} for a few low frequency components.
405 > Typical values of $\gamma^{*}$ in the dipolar-ordered phase are much
406 > higher than in the dipolar-disordered phase ($\gamma^{*} = 73,538$ in
407 > the lower panel of Fig. \ref{fig:fit}).  For the dipolar-ordered
408 > triangular lattice near the coexistence temperature, we also observe
409 > long wavelength undulations that are far outliers to the fits.  That
410 > is, the Landau free energy fits are well within error bars for most of
411 > the other points, but can be off by {\em orders of magnitude} for a
412 > few low frequency components.
413  
414   We interpret these outliers as evidence that these low frequency modes
415   are {\em non-thermal undulations}.  We take this as evidence that we
416   are actually seeing a rippled phase developing in this model system.
417  
418 < \begin{figure}[ht]
419 < \centering
420 < \caption{Evidence that the observed ripples are {\em not} thermal
421 < undulations is obtained from the 2-d fourier transform $\langle
422 < |h^{*}(\vec{q})|^2 \rangle$ of the height profile ($\langle h^{*}(x,y)
423 < \rangle$). Rippled samples show low-wavelength peaks that are
424 < outliers on the Landau free energy fits.  Samples exhibiting only
425 < thermal undulations fit Eq. \ref{eq:fit} remarkably well.}
426 < \includegraphics[width=5.5in]{fit.pdf}
390 < \label{fig:fit}
418 > \begin{figure}
419 > \includegraphics[width=\linewidth]{logFit}
420 > \caption{\label{fig:fit} Evidence that the observed ripples are {\em
421 > not} thermal undulations is obtained from the 2-d Fourier transform
422 > $\langle |h^{*}(\vec{q})|^2 \rangle$ of the height profile ($\langle
423 > h^{*}(x,y) \rangle$). Rippled samples show low-wavelength peaks that
424 > are outliers on the Landau free energy fits by an order of magnitude.
425 > Samples exhibiting only thermal undulations fit Eq. \ref{eq:fit}
426 > remarkably well.}
427   \end{figure}
428  
429   \subsection{Effects of Potential Parameters on Amplitude and Wavelength}
# Line 416 | Line 452 | fourier transform of $h(q_{\mathrm{rip}})$.  Amplitude
452   axis by projecting heights of the dipoles to obtain a one-dimensional
453   height profile, $h(q_{\mathrm{rip}})$. Ripple wavelengths can be
454   estimated from the largest non-thermal low-frequency component in the
455 < fourier transform of $h(q_{\mathrm{rip}})$.  Amplitudes can be
455 > Fourier transform of $h(q_{\mathrm{rip}})$.  Amplitudes can be
456   estimated by measuring peak-to-trough distances in
457   $h(q_{\mathrm{rip}})$ itself.
458  
423 \begin{figure}[ht]
424 \centering
425 \caption{Contours of the height-dipole correlation function as a function
426 of the dot product between the dipole ($\hat{\mu}$) and inter-dipole
427 separation vector ($\hat{r}$) and the distance ($r$) between the dipoles.
428 Perfect height correlation (contours approaching 1) are present in the
429 ordered phase when the two dipoles are in the same head-to-tail line.
430 Anti-correlation (contours below 0) is only seen when the inter-dipole
431 vector is perpendicular to the dipoles. }
432 \includegraphics[width=\linewidth]{height-dipole-correlation.pdf}
433 \label{fig:CrossCorrelation}
434 \end{figure}
435
459   A second, more accurate, and simpler method for estimating ripple
460   shape is to extract the wavelength and height information directly
461   from the largest non-thermal peak in the undulation spectrum.  For
462   large-amplitude ripples, the two methods give similar results.  The
463   one-dimensional projection method is more prone to noise (particularly
464 < in the amplitude estimates for the non-hexagonal lattices).  We report
464 > in the amplitude estimates for the distorted lattices).  We report
465   amplitudes and wavelengths taken directly from the undulation spectrum
466   below.
467  
468 < In the hexagonal lattice ($\gamma = \sqrt{3}$), the rippling is
468 > In the triangular lattice ($\gamma = \sqrt{3}$), the rippling is
469   observed for temperatures ($T^{*}$) from $61-122$.  The wavelength of
470   the ripples is remarkably stable at 21.4~$\sigma$ for all but the
471   temperatures closest to the order-disorder transition.  At $T^{*} =
# Line 457 | Line 480 | the mean spacing between lipids.
480   However, this is coincidental agreement based on a choice of 7~\AA~as
481   the mean spacing between lipids.
482  
483 < \begin{figure}[ht]
484 < \centering
485 < \caption{a) The amplitude $A^{*}$ of the ripples vs. temperature for a
486 < hexagonal lattice. b) The amplitude $A^{*}$ of the ripples vs. dipole
487 < strength ($\mu^{*}$) for both the hexagonal lattice (circles) and
488 < non-hexagonal lattice (squares).  The reduced temperatures were kept
489 < fixed at $T^{*} = 94$ for the hexagonal lattice and $T^{*} = 106$ for
490 < the non-hexagonal lattice (approximately 2/3 of the order-disorder
491 < transition temperature for each lattice).}
469 < \includegraphics[width=\linewidth]{properties_sq.pdf}
470 < \label{fig:Amplitude}
483 > \begin{figure}
484 > \includegraphics[width=\linewidth]{properties_sq}
485 > \caption{\label{fig:Amplitude} a) The amplitude $A^{*}$ of the ripples
486 > vs. temperature for a triangular lattice. b) The amplitude $A^{*}$ of
487 > the ripples vs. dipole strength ($\mu^{*}$) for both the triangular
488 > lattice (circles) and distorted lattice (squares).  The reduced
489 > temperatures were kept fixed at $T^{*} = 94$ for the triangular
490 > lattice and $T^{*} = 106$ for the distorted lattice (approximately 2/3
491 > of the order-disorder transition temperature for each lattice).}
492   \end{figure}
493  
494   The ripples can be made to disappear by increasing the internal
495 < surface tension (i.e. by increasing $k_r$ or equivalently, reducing
495 > elastic tension (i.e. by increasing $k_r$ or equivalently, reducing
496   the dipole moment).  The amplitude of the ripples depends critically
497   on the strength of the dipole moments ($\mu^{*}$) in Eq. \ref{eq:pot}.
498   If the dipoles are weakened substantially (below $\mu^{*}$ = 20) at a
# Line 481 | Line 502 | Fig. \ref{fig:Amplitude}.
502   of ripple amplitude on the dipolar strength in
503   Fig. \ref{fig:Amplitude}.
504  
505 < \subsection{Non-hexagonal lattices}
505 > \subsection{Distorted lattices}
506  
507   We have also investigated the effect of the lattice geometry by
508   changing the ratio of lattice constants ($\gamma$) while keeping the
509 < average nearest-neighbor spacing constant. The antiferroelectric state
509 > average nearest-neighbor spacing constant. The anti-ferroelectric state
510   is accessible for all $\gamma$ values we have used, although the
511 < distorted hexagonal lattices prefer a particular director axis due to
511 > distorted triangular lattices prefer a particular director axis due to
512   the anisotropy of the lattice.
513  
514 < Our observation of rippling behavior was not limited to the hexagonal
515 < lattices.  In non-hexagonal lattices the antiferroelectric phase can
514 > Our observation of rippling behavior was not limited to the triangular
515 > lattices.  In distorted lattices the anti-ferroelectric phase can
516   develop nearly instantaneously in the Monte Carlo simulations, and
517   these dipolar-ordered phases tend to be remarkably flat.  Whenever
518 < rippling has been observed in these non-hexagonal lattices
518 > rippling has been observed in these distorted lattices
519   (e.g. $\gamma = 1.875$), we see relatively short ripple wavelengths
520   (14 $\sigma$) and amplitudes of 2.4~$\sigma$.  These ripples are
521   weakly dependent on dipolar strength (see Fig. \ref{fig:Amplitude}),
# Line 506 | Line 527 | rippling is a symmetry-breaking phenomenon for hexagon
527   \gamma < 1.875$.  Outside this range, the order-disorder transition in
528   the dipoles remains, but the ordered dipolar phase has only thermal
529   undulations.  This is one of our strongest pieces of evidence that
530 < rippling is a symmetry-breaking phenomenon for hexagonal and
531 < nearly-hexagonal lattices.
530 > rippling is a symmetry-breaking phenomenon for triangular and
531 > nearly-triangular lattices.
532  
533   \subsection{Effects of System Size}
534   To evaluate the effect of finite system size, we have performed a
535 < series of simulations on the hexagonal lattice at a reduced
535 > series of simulations on the triangular lattice at a reduced
536   temperature of 122, which is just below the order-disorder transition
537   temperature ($T^{*} = 139$).  These conditions are in the
538   dipole-ordered and rippled portion of the phase diagram.  These are
539   also the conditions that should be most susceptible to system size
540   effects.
541  
542 < \begin{figure}[ht]
543 < \centering
544 < \caption{The ripple wavelength (top) and amplitude (bottom) as a
545 < function of system size for a hexagonal lattice ($\gamma=1.732$) at $T^{*} =
546 < 122$.}
526 < \includegraphics[width=\linewidth]{SystemSize.pdf}
527 < \label{fig:systemsize}
542 > \begin{figure}
543 > \includegraphics[width=\linewidth]{SystemSize}
544 > \caption{\label{fig:systemsize} The ripple wavelength (top) and
545 > amplitude (bottom) as a function of system size for a triangular
546 > lattice ($\gamma=1.732$) at $T^{*} = 122$.}
547   \end{figure}
548  
549   There is substantial dependence on system size for small (less than
# Line 549 | Line 568 | form of electrostatic dipoles) and a weak surface tens
568  
569   We have been able to show that a simple dipolar lattice model which
570   contains only molecular packing (from the lattice), anisotropy (in the
571 < form of electrostatic dipoles) and a weak surface tension (in the form
571 > form of electrostatic dipoles) and a weak elastic tension (in the form
572   of a nearest-neighbor harmonic potential) is capable of exhibiting
573   stable long-wavelength non-thermal surface corrugations.  The best
574   explanation for this behavior is that the ability of the dipoles to
575   translate out of the plane of the membrane is enough to break the
576 < symmetry of the hexagonal lattice and allow the energetic benefit from
577 < the formation of a bulk antiferroelectric phase.  Were the weak
578 < surface tension absent from our model, it would be possible for the
576 > symmetry of the triangular lattice and allow the energetic benefit
577 > from the formation of a bulk anti-ferroelectric phase.  Were the weak
578 > elastic tension absent from our model, it would be possible for the
579   entire lattice to ``tilt'' using $z$-translation.  Tilting the lattice
580 < in this way would yield an effectively non-hexagonal lattice which
581 < would avoid dipolar frustration altogether.  With the surface tension
582 < in place, bulk tilt causes a large strain, and the simplest way to
583 < release this strain is along line defects.  Line defects will result
584 < in rippled or sawtooth patterns in the membrane, and allow small
585 < ``stripes'' of membrane to form antiferroelectric regions that are
586 < tilted relative to the averaged membrane normal.
580 > in this way would yield an effectively non-triangular lattice which
581 > would avoid dipolar frustration altogether.  With the elastic tension
582 > in place, bulk tilt causes a large strain, and the least costly way to
583 > release this strain is between two rows of anti-aligned dipoles.
584 > These ``breaks'' will result in rippled or sawtooth patterns in the
585 > membrane, and allow small stripes of membrane to form
586 > anti-ferroelectric regions that are tilted relative to the averaged
587 > membrane normal.
588  
589   Although the dipole-dipole interaction is the major driving force for
590   the long range orientational ordered state, the formation of the
591   stable, smooth ripples is a result of the competition between the
592 < surface tension and the dipole-dipole interactions.  This statement is
592 > elastic tension and the dipole-dipole interactions.  This statement is
593   supported by the variation in $\mu^{*}$.  Substantially weaker dipoles
594   relative to the surface tension can cause the corrugated phase to
595   disappear.
596  
597 < The packing of the dipoles into a nearly-hexagonal lattice is clearly
597 > The packing of the dipoles into a nearly-triangular lattice is clearly
598   an important piece of the puzzle.  The dipolar head groups of lipid
599   molecules are sterically (as well as electrostatically) anisotropic,
600 < and would not be able to pack hexagonally without the steric
600 > and would not pack in triangular arrangements without the steric
601   interference of adjacent molecular bodies.  Since we only see rippled
602   phases in the neighborhood of $\gamma=\sqrt{3}$, this implies that
603 < there is a role played by the lipid chains in the organization of the
604 < hexagonally ordered phases which support ripples in realistic lipid
605 < bilayers.
603 > even if this dipolar mechanism is the correct explanation for the
604 > ripple phase in realistic bilayers, there would still be a role played
605 > by the lipid chains in the in-plane organization of the triangularly
606 > ordered phases which could support ripples.  The present model is
607 > certainly not detailed enough to answer exactly what drives the
608 > formation of the $P_{\beta'}$ phase in real lipids, but suggests some
609 > avenues for further experiments.
610  
611   The most important prediction we can make using the results from this
612   simple model is that if dipolar ordering is driving the surface
# Line 595 | Line 619 | the three equivalent lattice vectors in the hexagonal
619  
620   Our other observation about the ripple and dipolar directionality is
621   that the dipole director axis can be found to be parallel to any of
622 < the three equivalent lattice vectors in the hexagonal lattice.
622 > the three equivalent lattice vectors in the triangular lattice.
623   Defects in the ordering of the dipoles can cause the dipole director
624   (and consequently the surface corrugation) of small regions to be
625   rotated relative to each other by 120$^{\circ}$.  This is a similar
# Line 606 | Line 630 | replaced the somewhat artificial lattice packing and t
630   behaviors.  It would clearly be a closer approximation to the reality
631   if we allowed greater translational freedom to the dipoles and
632   replaced the somewhat artificial lattice packing and the harmonic
633 < ``surface tension'' with more realistic molecular modeling
634 < potentials.  What we have done is to present an extremely simple model
635 < which exhibits bulk non-thermal corrugation, and our explanation of
636 < this rippling phenomenon will help us design more accurate molecular
637 < models for corrugated membranes and experiments to test whether
638 < rippling is dipole-driven or not.
639 < \clearpage
633 > elastic tension with more realistic molecular modeling potentials.
634 > What we have done is to present a simple model which exhibits bulk
635 > non-thermal corrugation, and our explanation of this rippling
636 > phenomenon will help us design more accurate molecular models for
637 > corrugated membranes and experiments to test whether rippling is
638 > dipole-driven or not.
639 >
640 > \begin{acknowledgments}
641 > Support for this project was provided by the National Science
642 > Foundation under grant CHE-0134881.  The authors would like to thank
643 > the reviewers for helpful comments.
644 > \end{acknowledgments}
645 >
646   \bibliography{ripple}
617 \printfigures
647   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines