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 3097 by xsun, Wed Dec 27 22:13:09 2006 UTC vs.
Revision 3098 by gezelter, Thu Dec 28 21:55:59 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]{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 < triangular or distorted lattices.  The translational freedom
27 < of the dipoles allows triangular 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-triangular.  We observe three common features of the
33 < corrugated dipolar membranes: 1) the corrugated phases develop easily
34 < when hosted on triangular 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 triangular 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 antiferroelectric 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
81 < ($L_{\alpha}$) phases for relatively pure phosphatidylcholine (PC)
82 < bilayers.  The $P_{\beta'}$ phase has attracted substantial
83 < experimental interest over the past 30 years. Most structural
84 < information of the ripple phase has been obtained by the X-ray
85 < diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
86 < microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
87 < et al.} used atomic force microscopy (AFM) to observe ripple phase
88 < morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
89 < experimental results provide strong support for a 2-dimensional
90 < triangular packing lattice of the lipid molecules within the ripple
91 < phase.  This is a notable change from the observed lipid packing
92 < within the gel phase.~\cite{Cevc87}
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 < Although the results of dipolar fluid simulations can not be directly
59 < mapped onto the phases of lipid bilayers, the rich behaviors exhibited
60 < by simple dipolar models can give us some insight into the corrugation
61 < phenomenon of the $P_{\beta'}$ phase.  There have been a number of
62 < theoretical approaches (and some heroic simulations) undertaken to try
63 < to explain this phase, but to date, none have looked specifically at
100 < the contribution of the dipolar character of the lipid head groups
101 < towards this corrugation.  Before we present our simple model, we will
102 < briefly survey the previous theoretical work on this topic.
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 < The theoretical models that have been put forward to explain the
66 < formation of the $P_{\beta'}$ phase have presented a number of
67 < conflicting but intriguing explanations. 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
70 < continuum models predict higher fluidity in convex regions and that
71 < concave portions of the membrane correspond to more solid-like
72 < regions.  Carlson and Sethna used a packing-competition model (in
73 < which head groups and chains have competing packing energetics) to
74 < predict the formation of a ripple-like phase.  Their model predicted
75 < that the high-curvature portions have lower-chain packing and
76 < correspond to more fluid-like regions.  Goldstein and Leibler used a
77 < mean-field approach with a planar model for {\em inter-lamellar}
78 < interactions to predict rippling in multilamellar
79 < phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
80 < anisotropy of the nearest-neighbor interactions} coupled to
81 < hydrophobic constraining forces which restrict height differences
82 < between nearest neighbors is the origin of the ripple
83 < phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
84 < theory for tilt order and curvature of a single membrane and concluded
85 < that {\em coupling of molecular tilt to membrane curvature} is
86 < responsible for the production of ripples.~\cite{Lubensky93} Misbah,
87 < Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
88 < interactions} can lead to ripple instabilities.~\cite{Misbah98}
89 < Heimburg presented a {\em coexistence model} for ripple formation in
90 < which he postulates that fluid-phase line defects cause sharp
91 < 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 triangular lattices to show that large headgroups
135 < and molecular tilt with respect to the membrane normal vector can
136 < cause bulk rippling.~\cite{Bannerjee02}
65 > Buckling behavior in liquid crystalline and biological membranes is a
66 > well-known phenomenon.  Relatively pure phosphatidylcholine (PC)
67 > bilayers are known to form a corrugated or ``rippled'' phase
68 > ($P_{\beta'}$) which appears as an intermediate phase between the gel
69 > ($L_\beta$) and fluid ($L_{\alpha}$) phases.  The $P_{\beta'}$ phase
70 > has attracted substantial experimental interest over the past 30
71 > years. Most structural information of the ripple phase has been
72 > obtained by the X-ray diffraction~\cite{Sun96,Katsaras00} and
73 > freeze-fracture electron microscopy (FFEM).~\cite{Copeland80,Meyer96}
74 > Recently, Kaasgaard {\it et al.} used atomic force microscopy (AFM) to
75 > observe ripple phase morphology in bilayers supported on
76 > mica.~\cite{Kaasgaard03} The experimental results provide strong
77 > support for a 2-dimensional triangular packing lattice of the lipid
78 > molecules within the ripple phase.  This is a notable change from the
79 > observed lipid packing within the gel phase.~\cite{Cevc87} There have
80 > been a number of 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 < Large-scale molecular dynamics simulations have also been performed on
94 < rippled phases using united atom as well as molecular scale
95 < models. De~Vries {\it et al.} studied the structure of lecithin ripple
96 < phases via molecular dynamics and their simulations seem to support
97 < the coexistence models (i.e. fluid-like chain dynamics was observed in
98 < the kink regions).~\cite{deVries05} A similar coarse-grained approach
99 < has been used to study the line tension of bilayer
100 < edges.\cite{Jiang04,deJoannis06} Ayton and Voth have found significant
101 < undulations in zero-surface-tension states of membranes simulated via
102 < 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 > Another interesting properties of elastic membranes containing
94 > electrostatic dipoles is the phenomenon of flexoelectricity,\cite{}
95 > which is the ability of mechanical deformations of the membrane to
96 > result in electrostatic organization of the membrane.  This phenomenon
97 > is a curvature-induced membrane polarization which can lead to
98 > potential differences across a membrane.  Reverse flexoelectric
99 > behavior (in which applied alternating currents affect membrane
100 > curvature) has also been observed.  Explanations of the details of
101 > these effects have typically utilized membrane polarization parallel
102 > to the membrane normal.\cite{}
103  
104   The problem with using atomistic and even coarse-grained approaches to
105 < study this phenomenon is that only a relatively small number of
106 < periods of the corrugation (i.e. one or two) can be realistically
107 < simulated given current technology.  Also, simulations of lipid
108 < bilayers are traditionally carried out with periodic boundary
105 > study membrane buckling phenomena is that only a relatively small
106 > number of periods of the corrugation (i.e. one or two) can be
107 > realistically simulated given current technology.  Also, simulations
108 > of lipid bilayers are traditionally carried out with periodic boundary
109   conditions in two or three dimensions and these have the potential to
110   enhance the periodicity of the system at that wavelength.  To avoid
111   this pitfall, we are using a model which allows us to have
112   sufficiently large systems so that we are not causing artificial
113   corrugation through the use of periodic boundary conditions.
114  
115 < At the other extreme in density from the traditional simulations of
116 < dipolar fluids is the behavior of dipoles locked on regular lattices.
117 < Ferroelectric states (with long-range dipolar order) can be observed
118 < in dipolar systems with non-triangular packings.  However, {\em
119 < triangularly}-packed 2-D dipolar systems are inherently frustrated and
120 < one would expect a dipolar-disordered phase to be the lowest free
121 < energy configuration.  Therefore, it would seem unlikely that a
122 < frustrated lattice in a dipolar-disordered state could exhibit the
123 < long-range periodicity in the range of 100-600 \AA (as exhibited in
124 < the ripple phases studied by Kaasgard {\it et
125 < al.}).~\cite{Kaasgaard03}
115 > The simplest dipolar membrane is one in which the dipoles are located
116 > on fixed lattice sites. Ferroelectric states (with long-range dipolar
117 > order) can be observed in dipolar systems with non-triangular
118 > packings.  However, {\em triangularly}-packed 2-D dipolar systems are
119 > inherently frustrated and one would expect a dipolar-disordered phase
120 > to be the lowest free energy
121 > configuration.\cite{Toulouse1977,Marland1979} Dipolar lattices already
122 > have rich phase behavior, but in order to allow the membrane to
123 > buckle, a single degree of freedom (translation normal to the membrane
124 > face) must be added to each of the dipoles.  It would also be possible
125 > to allow complete translational freedom.  This approach
126 > is similar in character to a number of elastic Ising models that have
127 > been developed to explain interesting mechanical properties in
128 > magnetic alloys.\cite{Renard1966,Zhu2005,Zhu2006,Jiang2006}
129  
130 < Is there an intermediate model between the low-density dipolar fluids
131 < and the rigid lattice models which has the potential to exhibit the
132 < 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
130 > What we present here is an attempt to find the simplest dipolar model
131 > which will exhibit buckling behavior.  We are using a modified XYZ
132 > lattice model; details of the model can be found in section
133   \ref{sec:model}, results of Monte Carlo simulations using this model
134   are presented in section
135   \ref{sec:results}, and section \ref{sec:discussion} contains our conclusions.
# Line 190 | Line 139 | of a two-dimensional dipolar medium.  Since molecules
139  
140   The point of developing this model was to arrive at the simplest
141   possible theoretical model which could exhibit spontaneous corrugation
142 < of a two-dimensional dipolar medium.  Since molecules in the ripple
143 < phase have limited translational freedom, we have chosen a lattice to
144 < support the dipoles in the x-y plane.  The lattice may be either
145 < triangular (lattice constants $a/b = \sqrt{3}$) or distorted.
146 < However, each dipole has 3 degrees of freedom.  They may move freely
147 < {\em out} of the x-y plane (along the $z$ axis), and they have
148 < complete orientational freedom ($0 <= \theta <= \pi$, $0 <= \phi < 2
142 > of a two-dimensional dipolar medium.  Since molecules in polymerized
143 > membranes and in in the $P_{\beta'}$ ripple phase have limited
144 > translational freedom, we have chosen a lattice to support the dipoles
145 > in the x-y plane.  The lattice may be either triangular (lattice
146 > constants $a/b =
147 > \sqrt{3}$) or distorted.  However, each dipole has 3 degrees of
148 > freedom.  They may move freely {\em out} of the x-y plane (along the
149 > $z$ axis), and they have complete orientational freedom ($0 <= \theta
150 > <= \pi$, $0 <= \phi < 2
151   \pi$).  This is essentially a modified X-Y-Z model with translational
152   freedom along the z-axis.
153  
154   The potential energy of the system,
155 < \begin{equation}
156 < 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[
155 > \begin{eqnarray}
156 > V = \sum_i & & \left( \sum_{j>i} \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[
157   {\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j} -
158   3({\mathbf{\hat u}_i} \cdot {\mathbf{\hat
159   r}_{ij}})({\mathbf{\hat u}_j} \cdot {\mathbf{\hat r}_{ij}})\right]
160 < \right)
160 > \right. \nonumber \\
161 > & & \left. + \sum_{j \in NN_i}^6 \frac{k_r}{2}\left(
162 > r_{ij}-\sigma \right)^2 \right)
163   \label{eq:pot}
164 < \end{equation}
164 > \end{eqnarray}
165  
166 +
167   In this potential, $\mathbf{\hat u}_i$ is the unit vector pointing
168   along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
169   pointing along the inter-dipole vector $\mathbf{r}_{ij}$.  The entire
# Line 281 | Line 234 | lattices.
234   triangular ($\gamma = 1.732$) and distorted ($\gamma=1.875$)
235   lattices.
236  
237 < \begin{figure}[ht]
238 < \centering
239 < \caption{Top panel: The $P_2$ dipolar order parameter as a function of
240 < temperature for both triangular ($\gamma = 1.732$) and distorted
241 < ($\gamma = 1.875$) lattices.  Bottom Panel: The phase diagram for the
242 < dipolar membrane model.  The line denotes the division between the
243 < dipolar ordered (antiferroelectric) and disordered phases.  An
244 < enlarged view near the triangular lattice is shown inset.}
292 < \includegraphics[width=\linewidth]{phase.pdf}
293 < \label{phase}
237 > \begin{figure}
238 > \includegraphics[width=\linewidth]{phase}
239 > \caption{\label{phase} Top panel: The $P_2$ dipolar order parameter as
240 > a function of temperature for both triangular ($\gamma = 1.732$) and
241 > distorted ($\gamma = 1.875$) lattices.  Bottom Panel: The phase
242 > diagram for the dipolar membrane model.  The line denotes the division
243 > between the dipolar ordered (antiferroelectric) and disordered phases.
244 > An enlarged view near the triangular lattice is shown inset.}
245   \end{figure}
246  
247   There is a clear order-disorder transition in evidence from this data.
# Line 325 | Line 276 | rippled structure is shown in Fig. \ref{fig:snapshot}.
276   dipole director axis}.  A snapshot of a typical antiferroelectric
277   rippled structure is shown in Fig. \ref{fig:snapshot}.
278  
279 < \begin{figure}[ht]
280 < \centering
281 < \caption{Top and Side views of a representative configuration for the
282 < dipolar ordered phase supported on the triangular lattice. Note the
283 < antiferroelectric ordering and the long wavelength buckling of the
284 < membrane.  Dipolar ordering has been observed in all three equivalent
285 < directions on the triangular lattice, and the ripple direction is
286 < always perpendicular to the director axis for the dipoles.}
287 < \includegraphics[width=5.5in]{snapshot.pdf}
337 < \label{fig:snapshot}
279 > \begin{figure}
280 > \includegraphics[width=\linewidth]{snapshot}
281 > \caption{\label{fig:snapshot} Top and Side views of a representative
282 > configuration for the dipolar ordered phase supported on the
283 > triangular lattice. Note the antiferroelectric ordering and the long
284 > wavelength buckling of the membrane.  Dipolar ordering has been
285 > observed in all three equivalent directions on the triangular lattice,
286 > and the ripple direction is always perpendicular to the director axis
287 > for the dipoles.}
288   \end{figure}
289  
290   Although the snapshot in Fig. \ref{fig:snapshot} gives the appearance
# Line 347 | Line 297 | intermolecular vector $\vec{r}_{ij}$ and dipolar-axis
297   connection between dipolar ordering and the wave vector of the ripple
298   even more explicit.  $C(r, \cos \theta)$ is an angle-dependent pair
299   distribution function. The angle ($\theta$) is defined by the
300 < intermolecular vector $\vec{r}_{ij}$ and dipolar-axis of atom $i$,
300 > intermolecular vector $\vec{r}_{ij}$ and direction of dipole $i$,
301   \begin{equation}
302 < C(r, \cos \theta) = \langle \sum_{i}
303 < \sum_{j} h_i \cdot h_j \delta(r - r_{ij}) \delta(\cos \theta_{ij} - \cos \theta)\rangle / \langle h^2 \rangle
302 > C(r, \cos \theta) = \frac{\langle \sum_{i}
303 > \sum_{j} h_i \cdot h_j \delta(r - r_{ij}) \delta(\cos \theta_{ij} -
304 > \cos \theta)\rangle} {\langle h^2 \rangle}
305   \end{equation}
306   where $\cos \theta_{ij} = \hat{\mu}_{i} \cdot \hat{r}_{ij}$ and
307   $\hat{r}_{ij} = \vec{r}_{ij} / r_{ij}$.   Fig. \ref{fig:CrossCorrelation}
308   shows contours of this correlation function for both anti-ferroelectric, rippled
309   membranes as well as for the dipole-disordered portion of the phase diagram.  
310  
311 < \begin{figure}[ht]
312 < \centering
313 < \caption{Contours of the height-dipole correlation function as a function
314 < of the dot product between the dipole ($\hat{\mu}$) and inter-dipole
315 < separation vector ($\hat{r}$) and the distance ($r$) between the dipoles.
316 < Perfect height correlation (contours approaching 1) are present in the
317 < ordered phase when the two dipoles are in the same head-to-tail line.
311 > \begin{figure}
312 > \includegraphics[width=\linewidth]{hdc}
313 > \caption{\label{fig:CrossCorrelation} Contours of the height-dipole
314 > correlation function as a function of the dot product between the
315 > dipole ($\hat{\mu}$) and inter-dipole separation vector ($\hat{r}$)
316 > and the distance ($r$) between the dipoles.  Perfect height
317 > correlation (contours approaching 1) are present in the ordered phase
318 > when the two dipoles are in the same head-to-tail line.
319   Anti-correlation (contours below 0) is only seen when the inter-dipole
320 < vector is perpendicular to the dipoles.  In the dipole-disordered portion
321 < of the phase diagram, there is only weak correlation in the dipole direction
322 < and this correlation decays rapidly to zero for intermolecular vectors that are
323 < not dipole-aligned.}
372 < \includegraphics[width=\linewidth]{height-dipole-correlation.pdf}
373 < \label{fig:CrossCorrelation}
320 > vector is perpendicular to the dipoles.  In the dipole-disordered
321 > portion of the phase diagram, there is only weak correlation in the
322 > dipole direction and this correlation decays rapidly to zero for
323 > intermolecular vectors that are not dipole-aligned.}
324   \end{figure}
325  
326   \subsection{Discriminating Ripples from Thermal Undulations}
# Line 387 | Line 337 | absolute value of the undulation spectrum can be writt
337   elastic continuum models, it can shown that in the $NVT$ ensemble, the
338   absolute value of the undulation spectrum can be written,
339   \begin{equation}
340 < \langle | h(q)|^2 \rangle_{NVT} = \frac{k_B T}{k_c |\vec{q}|^4 +
341 < \tilde{\gamma}|\vec{q}|^2},
340 > \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{k_c q^4 +
341 > \gamma q^2},
342   \label{eq:fit}
343   \end{equation}
344 < where $k_c$ is the bending modulus for the membrane, and
345 < $\tilde{\gamma}$ is the mechanical surface tension.~\cite{Safran94}
346 < The systems studied in this paper have essentially zero bending moduli
347 < ($k_c$) and relatively large mechanical surface tensions
348 < ($\tilde{\gamma}$), so a much simpler form can be written,
344 > where $k_c$ is the bending modulus for the membrane, and $\gamma$ is
345 > the mechanical surface tension.~\cite{Safran94} The systems studied in
346 > this paper have essentially zero bending moduli ($k_c$) and relatively
347 > large mechanical surface tensions ($\gamma$), so a much simpler form
348 > can be written,
349   \begin{equation}
350 < \langle | h(q)|^2 \rangle_{NVT} = \frac{k_B T}{\tilde{\gamma}|\vec{q}|^2},
350 > \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{\gamma q^2},
351   \label{eq:fit2}
352   \end{equation}
353  
# Line 411 | Line 361 | model, an interpolated method for computing undulation
361   lattice, one could use the heights of the lattice points themselves as
362   the grid for the Fourier transform (without interpolating to a square
363   grid).  However, if lateral translational freedom is added to this
364 < model, an interpolated method for computing undulation spectra will be
365 < required.
364 > model (a likely extension), an interpolated grid method for computing
365 > undulation spectra will be required.
366  
367   As mentioned above, the best fits to our undulation spectra are
368 < obtained by approximating the value of $k_c$ to 0.  In
369 < Fig. \ref{fig:fit} we show typical undulation spectra for two
370 < different regions of the phase diagram along with their fits from the
371 < Landau free energy approach (Eq. \ref{eq:fit2}).  In the
372 < high-temperature disordered phase, the Landau fits can be nearly
373 < perfect, and from these fits we can estimate the tension in the
374 < surface.
368 > obtained by setting the value of $k_c$ to 0.  In Fig. \ref{fig:fit} we
369 > show typical undulation spectra for two different regions of the phase
370 > diagram along with their fits from the Landau free energy approach
371 > (Eq. \ref{eq:fit2}).  In the high-temperature disordered phase, the
372 > Landau fits can be nearly perfect, and from these fits we can estimate
373 > the tension in the surface.  In reduced units, typical values of
374 > $\gamma^{*} = \gamma / \epsilon = 2500$ are obtained for the
375 > disordered phase ($\gamma^{*} = 2551.7$ in the top panel of
376 > Fig. \ref{fig:fit}).
377  
378 < For the dipolar-ordered triangular lattice near the coexistence
379 < temperature, however, we observe long wavelength undulations that are
380 < far outliers to the fits.  That is, the Landau free energy fits are
381 < well within error bars for most of the other points, but can be off by
382 < {\em orders of magnitude} for a few low frequency components.
378 > Typical values of $\gamma^{*}$ in the dipolar-ordered phase are much
379 > higher than in the dipolar-disordered phase ($\gamma^{*} = 73,538$ in
380 > the lower panel of Fig. \ref{fig:fit}).  For the dipolar-ordered
381 > triangular lattice near the coexistence temperature, we also observe
382 > long wavelength undulations that are far outliers to the fits.  That
383 > is, the Landau free energy fits are well within error bars for most of
384 > the other points, but can be off by {\em orders of magnitude} for a
385 > few low frequency components.
386  
387   We interpret these outliers as evidence that these low frequency modes
388   are {\em non-thermal undulations}.  We take this as evidence that we
389   are actually seeing a rippled phase developing in this model system.
390  
391 < \begin{figure}[ht]
392 < \centering
393 < \caption{Evidence that the observed ripples are {\em not} thermal
394 < undulations is obtained from the 2-d fourier transform $\langle
395 < |h^{*}(\vec{q})|^2 \rangle$ of the height profile ($\langle h^{*}(x,y)
396 < \rangle$). Rippled samples show low-wavelength peaks that are
397 < outliers on the Landau free energy fits.  Samples exhibiting only
398 < thermal undulations fit Eq. \ref{eq:fit} remarkably well.}
399 < \includegraphics[width=5.5in]{logFit.pdf}
445 < \label{fig:fit}
391 > \begin{figure}
392 > \includegraphics[width=\linewidth]{logFit}
393 > \caption{\label{fig:fit} Evidence that the observed ripples are {\em
394 > not} thermal undulations is obtained from the 2-d fourier transform
395 > $\langle |h^{*}(\vec{q})|^2 \rangle$ of the height profile ($\langle
396 > h^{*}(x,y) \rangle$). Rippled samples show low-wavelength peaks that
397 > are outliers on the Landau free energy fits by an order of magnitude.
398 > Samples exhibiting only thermal undulations fit Eq. \ref{eq:fit}
399 > remarkably well.}
400   \end{figure}
401  
402   \subsection{Effects of Potential Parameters on Amplitude and Wavelength}
# Line 499 | Line 453 | the mean spacing between lipids.
453   However, this is coincidental agreement based on a choice of 7~\AA~as
454   the mean spacing between lipids.
455  
456 < \begin{figure}[ht]
457 < \centering
458 < \caption{a) The amplitude $A^{*}$ of the ripples vs. temperature for a
459 < triangular lattice. b) The amplitude $A^{*}$ of the ripples vs. dipole
460 < strength ($\mu^{*}$) for both the triangular lattice (circles) and
461 < distorted lattice (squares).  The reduced temperatures were kept
462 < fixed at $T^{*} = 94$ for the triangular lattice and $T^{*} = 106$ for
463 < the distorted lattice (approximately 2/3 of the order-disorder
464 < transition temperature for each lattice).}
511 < \includegraphics[width=\linewidth]{properties_sq.pdf}
512 < \label{fig:Amplitude}
456 > \begin{figure}
457 > \includegraphics[width=\linewidth]{properties_sq}
458 > \caption{\label{fig:Amplitude} a) The amplitude $A^{*}$ of the ripples
459 > vs. temperature for a triangular lattice. b) The amplitude $A^{*}$ of
460 > the ripples vs. dipole strength ($\mu^{*}$) for both the triangular
461 > lattice (circles) and distorted lattice (squares).  The reduced
462 > temperatures were kept fixed at $T^{*} = 94$ for the triangular
463 > lattice and $T^{*} = 106$ for the distorted lattice (approximately 2/3
464 > of the order-disorder transition temperature for each lattice).}
465   \end{figure}
466  
467   The ripples can be made to disappear by increasing the internal
# Line 560 | Line 512 | effects.
512   also the conditions that should be most susceptible to system size
513   effects.
514  
515 < \begin{figure}[ht]
516 < \centering
517 < \caption{The ripple wavelength (top) and amplitude (bottom) as a
518 < function of system size for a triangular lattice ($\gamma=1.732$) at $T^{*} =
519 < 122$.}
568 < \includegraphics[width=\linewidth]{SystemSize.pdf}
569 < \label{fig:systemsize}
515 > \begin{figure}
516 > \includegraphics[width=\linewidth]{SystemSize}
517 > \caption{\label{fig:systemsize} The ripple wavelength (top) and
518 > amplitude (bottom) as a function of system size for a triangular
519 > lattice ($\gamma=1.732$) at $T^{*} = 122$.}
520   \end{figure}
521  
522   There is substantial dependence on system size for small (less than
# Line 654 | Line 604 | rippling is dipole-driven or not.
604   this rippling phenomenon will help us design more accurate molecular
605   models for corrugated membranes and experiments to test whether
606   rippling is dipole-driven or not.
607 < \clearpage
607 >
608 > \begin{acknowledgments}
609 > Support for this project was provided by the National Science
610 > Foundation under grant CHE-0134881.  The authors would like to thank
611 > the reviewers for helpful comments.
612 > \end{acknowledgments}
613 >
614   \bibliography{ripple}
659 \printfigures
615   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines