ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ripple2/original.tex
Revision: 3098
Committed: Thu Dec 28 21:55:59 2006 UTC (17 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 35321 byte(s)
Log Message:
Getting ready for publication

File Contents

# User Rev Content
1 gezelter 3098 \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}}
7     \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
8    
9     %\bibliographystyle{aps}
10    
11     \title{Spontaneous Corrugation of Dipolar Membranes}
12     \author{Xiuquan Sun and J. Daniel Gezelter}
13     \email[E-mail:]{gezelter@nd.edu}
14     \affiliation{Department of Chemistry and Biochemistry,\\
15     University of Notre Dame, \\
16     Notre Dame, Indiana 46556}
17    
18     \date{\today}
19    
20     \begin{abstract}
21     We present a simple model for dipolar 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.
38     \end{abstract}
39    
40     \pacs{68.03.Hj, 82.20.Wt}
41     \maketitle
42    
43    
44     \section{Introduction}
45     \label{Int}
46     There has been intense recent interest in the phase behavior of
47     dipolar
48     fluids.\cite{Tlusty00,Teixeira00,Tavares02,Duncan04,Holm05,Duncan06}
49     Due to the anisotropic interactions between dipoles, dipolar fluids
50     can present anomalous phase behavior. Examples of condensed-phase
51     dipolar systems include ferrofluids, electro-rheological fluids, and
52     even biological membranes. Computer simulations have provided useful
53     information on the structural features and phase transition of the
54     dipolar fluids. Simulation results indicate that at low densities,
55     these fluids spontaneously organize into head-to-tail dipolar
56     ``chains''.\cite{Teixeira00,Holm05} At low temperatures, these chains
57     and rings prevent the occurrence of a liquid-gas phase transition.
58     However, Tlusty and Safran showed that there is a defect-induced phase
59     separation into a low-density ``chain'' phase and a higher density
60     Y-defect phase.\cite{Tlusty00} Recently, inspired by experimental
61     studies on monolayers of dipolar fluids, theoretical models using
62     two-dimensional dipolar soft spheres have appeared in the literature.
63     Tavares {\it et al.} tested their theory for chain and ring length
64     distributions in two dimensions and carried out Monte Carlo
65     simulations in the low-density phase.\cite{Tavares02} Duncan and Camp
66     performed dynamical simulations on two-dimensional dipolar fluids to
67     study transport and orientational dynamics in these
68     systems.\cite{Duncan04} They have recently revisited two-dimensional
69     systems to study the kinetic conditions for the defect-induced
70     condensation into the Y-defect phase.\cite{Duncan06}
71    
72     Although they are not traditionally classified as 2-dimensional
73     dipolar fluids, hydrated lipids aggregate spontaneously to form
74     bilayers which exhibit a variety of phases depending on their
75     temperatures and compositions. At high temperatures, the fluid
76     ($L_{\alpha}$) phase of Phosphatidylcholine (PC) lipids closely
77     resembles a dipolar fluid. However, at lower temperatures, packing of
78     the molecules becomes important, and the translational freedom of
79     lipid molecules is thought to be substantially restricted. A
80     corrugated or ``rippled'' phase ($P_{\beta'}$) appears as an
81     intermediate phase between the gel ($L_\beta$) and fluid
82     ($L_{\alpha}$) phases for relatively pure phosphatidylcholine (PC)
83     bilayers. The $P_{\beta'}$ phase has attracted substantial
84     experimental interest over the past 30 years. Most structural
85     information of the ripple phase has been obtained by the X-ray
86     diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
87     microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
88     et al.} used atomic force microscopy (AFM) to observe ripple phase
89     morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
90     experimental results provide strong support for a 2-dimensional
91     triangular packing lattice of the lipid molecules within the ripple
92     phase. This is a notable change from the observed lipid packing
93     within the gel phase.~\cite{Cevc87}
94    
95     Although the results of dipolar fluid simulations can not be directly
96     mapped onto the phases of lipid bilayers, the rich behaviors exhibited
97     by simple dipolar models can give us some insight into the corrugation
98     phenomenon of the $P_{\beta'}$ phase. There have been a number of
99     theoretical approaches (and some heroic simulations) undertaken to try
100     to explain this phase, but to date, none have looked specifically at
101     the contribution of the dipolar character of the lipid head groups
102     towards this corrugation. Before we present our simple model, we will
103     briefly survey the previous theoretical work on this topic.
104    
105     The theoretical models that have been put forward to explain the
106     formation of the $P_{\beta'}$ phase have presented a number of
107     conflicting but intriguing explanations. Marder {\it et al.} used a
108     curvature-dependent Landau-de Gennes free-energy functional to predict
109     a rippled phase.~\cite{Marder84} This model and other related
110     continuum models predict higher fluidity in convex regions and that
111     concave portions of the membrane correspond to more solid-like
112     regions. Carlson and Sethna used a packing-competition model (in
113     which head groups and chains have competing packing energetics) to
114     predict the formation of a ripple-like phase. Their model predicted
115     that the high-curvature portions have lower-chain packing and
116     correspond to more fluid-like regions. Goldstein and Leibler used a
117     mean-field approach with a planar model for {\em inter-lamellar}
118     interactions to predict rippling in multilamellar
119     phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
120     anisotropy of the nearest-neighbor interactions} coupled to
121     hydrophobic constraining forces which restrict height differences
122     between nearest neighbors is the origin of the ripple
123     phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
124     theory for tilt order and curvature of a single membrane and concluded
125     that {\em coupling of molecular tilt to membrane curvature} is
126     responsible for the production of ripples.~\cite{Lubensky93} Misbah,
127     Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
128     interactions} can lead to ripple instabilities.~\cite{Misbah98}
129     Heimburg presented a {\em coexistence model} for ripple formation in
130     which he postulates that fluid-phase line defects cause sharp
131     curvature between relatively flat gel-phase regions.~\cite{Heimburg00}
132     Kubica has suggested that a lattice model of polar head groups could
133     be valuable in trying to understand bilayer phase
134     formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of
135     lamellar stacks of triangular lattices to show that large headgroups
136     and molecular tilt with respect to the membrane normal vector can
137     cause bulk rippling.~\cite{Bannerjee02}
138    
139     Large-scale molecular dynamics simulations have also been performed on
140     rippled phases using united atom as well as molecular scale
141     models. De~Vries {\it et al.} studied the structure of lecithin ripple
142     phases via molecular dynamics and their simulations seem to support
143     the coexistence models (i.e. fluid-like chain dynamics was observed in
144     the kink regions).~\cite{deVries05} A similar coarse-grained approach
145     has been used to study the line tension of bilayer
146     edges.\cite{Jiang04,deJoannis06} Ayton and Voth have found significant
147     undulations in zero-surface-tension states of membranes simulated via
148     dissipative particle dynamics, but their results are consistent with
149     purely thermal undulations.~\cite{Ayton02} Brannigan, Tamboli and
150     Brown have used a molecular scale model to elucidate the role of
151     molecular shape on membrane phase behavior and
152     elasticity.~\cite{Brannigan04b} They have also observed a buckled
153     hexatic phase with strong tail and moderate alignment
154     attractions.~\cite{Brannigan04a}
155    
156     The problem with using atomistic and even coarse-grained approaches to
157     study this phenomenon is that only a relatively small number of
158     periods of the corrugation (i.e. one or two) can be realistically
159     simulated given current technology. Also, simulations of lipid
160     bilayers are traditionally carried out with periodic boundary
161     conditions in two or three dimensions and these have the potential to
162     enhance the periodicity of the system at that wavelength. To avoid
163     this pitfall, we are using a model which allows us to have
164     sufficiently large systems so that we are not causing artificial
165     corrugation through the use of periodic boundary conditions.
166    
167     At the other extreme in density from the traditional simulations of
168     dipolar fluids is the behavior of dipoles locked on regular lattices.
169     Ferroelectric states (with long-range dipolar order) can be observed
170     in dipolar systems with non-triangular packings. However, {\em
171     triangularly}-packed 2-D dipolar systems are inherently frustrated and
172     one would expect a dipolar-disordered phase to be the lowest free
173     energy configuration. Therefore, it would seem unlikely that a
174     frustrated lattice in a dipolar-disordered state could exhibit the
175     long-range periodicity in the range of 100-600 \AA (as exhibited in
176     the ripple phases studied by Kaasgard {\it et
177     al.}).~\cite{Kaasgaard03}
178    
179     Is there an intermediate model between the low-density dipolar fluids
180     and the rigid lattice models which has the potential to exhibit the
181     corrugation phenomenon of the $P_{\beta'}$ phase? What we present
182     here is an attempt to find a simple dipolar model which will exhibit
183     this behavior. We are using a modified XYZ lattice model; details of
184     the model can be found in section
185     \ref{sec:model}, results of Monte Carlo simulations using this model
186     are presented in section
187     \ref{sec:results}, and section \ref{sec:discussion} contains our conclusions.
188    
189     \section{2-D Dipolar Membrane}
190     \label{sec:model}
191    
192     The point of developing this model was to arrive at the simplest
193     possible theoretical model which could exhibit spontaneous corrugation
194     of a two-dimensional dipolar medium. Since molecules in the ripple
195     phase have limited translational freedom, we have chosen a lattice to
196     support the dipoles in the x-y plane. The lattice may be either
197     triangular (lattice constants $a/b = \sqrt{3}$) or distorted.
198     However, each dipole has 3 degrees of freedom. They may move freely
199     {\em out} of the x-y plane (along the $z$ axis), and they have
200     complete orientational freedom ($0 <= \theta <= \pi$, $0 <= \phi < 2
201     \pi$). This is essentially a modified X-Y-Z model with translational
202     freedom along the z-axis.
203    
204     The potential energy of the system,
205     \begin{eqnarray}
206     V = \sum_i & & \left( \sum_{j>i} \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[
207     {\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j} -
208     3({\mathbf{\hat u}_i} \cdot {\mathbf{\hat
209     r}_{ij}})({\mathbf{\hat u}_j} \cdot {\mathbf{\hat r}_{ij}})\right]
210     \right. \nonumber \\
211     & & \left. + \sum_{j \in NN_i}^6 \frac{k_r}{2}\left(
212     r_{ij}-\sigma \right)^2 \right)
213     \label{eq:pot}
214     \end{eqnarray}
215    
216    
217     In this potential, $\mathbf{\hat u}_i$ is the unit vector pointing
218     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
219     pointing along the inter-dipole vector $\mathbf{r}_{ij}$. The entire
220     potential is governed by three parameters, the dipolar strength
221     ($\mu$), the harmonic spring constant ($k_r$) and the preferred
222     intermolecular spacing ($\sigma$). In practice, we set the value of
223     $\sigma$ to the average inter-molecular spacing from the planar
224     lattice, yielding a potential model that has only two parameters for a
225     particular choice of lattice constants $a$ (along the $x$-axis) and
226     $b$ (along the $y$-axis). We also define a set of reduced parameters
227     based on the length scale ($\sigma$) and the energy of the harmonic
228     potential at a deformation of 2 $\sigma$ ($\epsilon = k_r \sigma^2 /
229     2$). Using these two constants, we perform our calculations using
230     reduced distances, ($r^{*} = r / \sigma$), temperatures ($T^{*} = 2
231     k_B T / k_r \sigma^2$), densities ($\rho^{*} = N \sigma^2 / L_x L_y$),
232     and dipole moments ($\mu^{*} = \mu / \sqrt{4 \pi \epsilon_0 \sigma^5
233     k_r / 2}$).
234    
235     To investigate the phase behavior of this model, we have performed a
236     series of Metropolis Monte Carlo simulations of moderately-sized (34.3
237     $\sigma$ on a side) patches of membrane hosted on both triangular
238     ($\gamma = a/b = \sqrt{3}$) and distorted ($\gamma \neq \sqrt{3}$)
239     lattices. The linear extent of one edge of the monolayer was $20 a$
240     and the system was kept roughly square. The average distance that
241     coplanar dipoles were positioned from their six nearest neighbors was
242     1 $\sigma$ (on both triangular and distorted lattices). Typical
243     system sizes were 1360 dipoles for the triangular lattices and
244     840-2800 dipoles for the distorted lattices. Two-dimensional periodic
245     boundary conditions were used, and the cutoff for the dipole-dipole
246     interaction was set to 4.3 $\sigma$. Since dipole-dipole interactions
247     decay rapidly with distance, and since the intrinsic three-dimensional
248     periodicity of the Ewald sum can give artifacts in 2-d systems, we
249     have chosen not to use it in these calculations. Although the Ewald
250     sum has been reformulated to handle 2-D
251     systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} these methods
252     are computationally expensive,\cite{Spohr97,Yeh99} and are not
253     necessary in this case. All parameters ($T^{*}$, $\mu^{*}$, and
254     $\gamma$) were varied systematically to study the effects of these
255     parameters on the formation of ripple-like phases.
256    
257     \section{Results and Analysis}
258     \label{sec:results}
259    
260     \subsection{Dipolar Ordering and Coexistence Temperatures}
261     The principal method for observing the orientational ordering
262     transition in dipolar systems is the $P_2$ order parameter (defined as
263     $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
264     eigenvalue of the matrix,
265     \begin{equation}
266     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
267     \begin{array}{ccc}
268     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
269     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
270     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
271     \end{array} \right).
272     \label{eq:opmatrix}
273     \end{equation}
274     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
275     for dipole $i$. $P_2$ will be $1.0$ for a perfectly-ordered system
276     and near $0$ for a randomized system. Note that this order parameter
277     is {\em not} equal to the polarization of the system. For example,
278     the polarization of the perfect antiferroelectric system is $0$, but
279     $P_2$ for an antiferroelectric system is $1$. The eigenvector of
280     $\mathsf{S}$ corresponding to the largest eigenvalue is familiar as
281     the director axis, which can be used to determine a privileged dipolar
282     axis for dipole-ordered systems. The top panel in Fig. \ref{phase}
283     shows the values of $P_2$ as a function of temperature for both
284     triangular ($\gamma = 1.732$) and distorted ($\gamma=1.875$)
285     lattices.
286    
287     \begin{figure}
288     \includegraphics[width=\linewidth]{phase.pdf}
289     \caption{\label{phase} Top panel: The $P_2$ dipolar order parameter as
290     a function of temperature for both triangular ($\gamma = 1.732$) and
291     distorted ($\gamma = 1.875$) lattices. Bottom Panel: The phase
292     diagram for the dipolar membrane model. The line denotes the division
293     between the dipolar ordered (antiferroelectric) and disordered phases.
294     An enlarged view near the triangular lattice is shown inset.}
295     \end{figure}
296    
297     There is a clear order-disorder transition in evidence from this data.
298     Both the triangular and distorted lattices have dipolar-ordered
299     low-temperature phases, and orientationally-disordered high
300     temperature phases. The coexistence temperature for the triangular
301     lattice is significantly lower than for the distorted lattices, and
302     the bulk polarization is approximately $0$ for both dipolar ordered
303     and disordered phases. This gives strong evidence that the dipolar
304     ordered phase is antiferroelectric. We have verified that this
305     dipolar ordering transition is not a function of system size by
306     performing identical calculations with systems twice as large. The
307     transition is equally smooth at all system sizes that were studied.
308     Additionally, we have repeated the Monte Carlo simulations over a wide
309     range of lattice ratios ($\gamma$) to generate a dipolar
310     order/disorder phase diagram. The bottom panel in Fig. \ref{phase}
311     shows that the triangular lattice is a low-temperature cusp in the
312     $T^{*}-\gamma$ phase diagram.
313    
314     This phase diagram is remarkable in that it shows an antiferroelectric
315     phase near $\gamma=1.732$ where one would expect lattice frustration
316     to result in disordered phases at all temperatures. Observations of
317     the configurations in this phase show clearly that the system has
318     accomplished dipolar orderering by forming large ripple-like
319     structures. We have observed antiferroelectric ordering in all three
320     of the equivalent directions on the triangular lattice, and the dipoles
321     have been observed to organize perpendicular to the membrane normal
322     (in the plane of the membrane). It is particularly interesting to
323     note that the ripple-like structures have also been observed to
324     propagate in the three equivalent directions on the lattice, but the
325     {\em direction of ripple propagation is always perpendicular to the
326     dipole director axis}. A snapshot of a typical antiferroelectric
327     rippled structure is shown in Fig. \ref{fig:snapshot}.
328    
329     \begin{figure}
330     \includegraphics[width=\linewidth]{snapshot.pdf}
331     \caption{\label{fig:snapshot} Top and Side views of a representative
332     configuration for the dipolar ordered phase supported on the
333     triangular lattice. Note the antiferroelectric ordering and the long
334     wavelength buckling of the membrane. Dipolar ordering has been
335     observed in all three equivalent directions on the triangular lattice,
336     and the ripple direction is always perpendicular to the director axis
337     for the dipoles.}
338     \end{figure}
339    
340     Although the snapshot in Fig. \ref{fig:snapshot} gives the appearance
341     of three-row stair-like structures, these appear to be transient. On
342     average, the corrugation of the membrane is a relatively smooth,
343     long-wavelength phenomenon, with occasional steep drops between
344     adjacent lines of anti-aligned dipoles.
345    
346     The height-dipole correlation function ($C(r, \cos \theta)$) makes the
347     connection between dipolar ordering and the wave vector of the ripple
348     even more explicit. $C(r, \cos \theta)$ is an angle-dependent pair
349     distribution function. The angle ($\theta$) is defined by the
350     intermolecular vector $\vec{r}_{ij}$ and direction of dipole $i$,
351     \begin{equation}
352     C(r, \cos \theta) = \frac{\langle \sum_{i}
353     \sum_{j} h_i \cdot h_j \delta(r - r_{ij}) \delta(\cos \theta_{ij} -
354     \cos \theta)\rangle} {\langle h^2 \rangle}
355     \end{equation}
356     where $\cos \theta_{ij} = \hat{\mu}_{i} \cdot \hat{r}_{ij}$ and
357     $\hat{r}_{ij} = \vec{r}_{ij} / r_{ij}$. Fig. \ref{fig:CrossCorrelation}
358     shows contours of this correlation function for both anti-ferroelectric, rippled
359     membranes as well as for the dipole-disordered portion of the phase diagram.
360    
361     \begin{figure}
362     \includegraphics[width=\linewidth]{height-dipole-correlation.pdf}
363     \caption{\label{fig:CrossCorrelation} Contours of the height-dipole
364     correlation function as a function of the dot product between the
365     dipole ($\hat{\mu}$) and inter-dipole separation vector ($\hat{r}$)
366     and the distance ($r$) between the dipoles. Perfect height
367     correlation (contours approaching 1) are present in the ordered phase
368     when the two dipoles are in the same head-to-tail line.
369     Anti-correlation (contours below 0) is only seen when the inter-dipole
370     vector is perpendicular to the dipoles. In the dipole-disordered
371     portion of the phase diagram, there is only weak correlation in the
372     dipole direction and this correlation decays rapidly to zero for
373     intermolecular vectors that are not dipole-aligned.}
374     \end{figure}
375    
376     \subsection{Discriminating Ripples from Thermal Undulations}
377    
378     In order to be sure that the structures we have observed are actually
379     a rippled phase and not simply thermal undulations, we have computed
380     the undulation spectrum,
381     \begin{equation}
382     h(\vec{q}) = A^{-1/2} \int d\vec{r}
383     h(\vec{r}) e^{-i \vec{q}\cdot\vec{r}}
384     \end{equation}
385     where $h(\vec{r})$ is the height of the membrane at location $\vec{r}
386     = (x,y)$.~\cite{Safran94,Seifert97} In simple (and more complicated)
387     elastic continuum models, it can shown that in the $NVT$ ensemble, the
388     absolute value of the undulation spectrum can be written,
389     \begin{equation}
390     \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{k_c q^4 +
391     \gamma q^2},
392     \label{eq:fit}
393     \end{equation}
394     where $k_c$ is the bending modulus for the membrane, and $\gamma$ is
395     the mechanical surface tension.~\cite{Safran94} The systems studied in
396     this paper have essentially zero bending moduli ($k_c$) and relatively
397     large mechanical surface tensions ($\gamma$), so a much simpler form
398     can be written,
399     \begin{equation}
400     \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{\gamma q^2},
401     \label{eq:fit2}
402     \end{equation}
403    
404     The undulation spectrum is computed by superimposing a rectangular
405     grid on top of the membrane, and by assigning height ($h(\vec{r})$)
406     values to the grid from the average of all dipoles that fall within a
407     given $\vec{r}+d\vec{r}$ grid area. Empty grid pixels are assigned
408     height values by interpolation from the nearest neighbor pixels. A
409     standard 2-d Fourier transform is then used to obtain $\langle |
410     h(q)|^2 \rangle$. Alternatively, since the dipoles sit on a Bravais
411     lattice, one could use the heights of the lattice points themselves as
412     the grid for the Fourier transform (without interpolating to a square
413     grid). However, if lateral translational freedom is added to this
414     model (a likely extension), an interpolated grid method for computing
415     undulation spectra will be required.
416    
417     As mentioned above, the best fits to our undulation spectra are
418     obtained by setting the value of $k_c$ to 0. In Fig. \ref{fig:fit} we
419     show typical undulation spectra for two different regions of the phase
420     diagram along with their fits from the Landau free energy approach
421     (Eq. \ref{eq:fit2}). In the high-temperature disordered phase, the
422     Landau fits can be nearly perfect, and from these fits we can estimate
423     the tension in the surface. In reduced units, typical values of
424     $\gamma^{*} = \gamma / \epsilon = 2500$ are obtained for the
425     disordered phase ($\gamma^{*} = 2551.7$ in the top panel of
426     Fig. \ref{fig:fit}).
427    
428     Typical values of $\gamma^{*}$ in the dipolar-ordered phase are much
429     higher than in the dipolar-disordered phase ($\gamma^{*} = 73,538$ in
430     the lower panel of Fig. \ref{fig:fit}). For the dipolar-ordered
431     triangular lattice near the coexistence temperature, we also observe
432     long wavelength undulations that are far outliers to the fits. That
433     is, the Landau free energy fits are well within error bars for most of
434     the other points, but can be off by {\em orders of magnitude} for a
435     few low frequency components.
436    
437     We interpret these outliers as evidence that these low frequency modes
438     are {\em non-thermal undulations}. We take this as evidence that we
439     are actually seeing a rippled phase developing in this model system.
440    
441     \begin{figure}
442     \includegraphics[width=\linewidth]{logFit.pdf}
443     \caption{\label{fig:fit} Evidence that the observed ripples are {\em
444     not} thermal undulations is obtained from the 2-d fourier transform
445     $\langle |h^{*}(\vec{q})|^2 \rangle$ of the height profile ($\langle
446     h^{*}(x,y) \rangle$). Rippled samples show low-wavelength peaks that
447     are outliers on the Landau free energy fits by an order of magnitude.
448     Samples exhibiting only thermal undulations fit Eq. \ref{eq:fit}
449     remarkably well.}
450     \end{figure}
451    
452     \subsection{Effects of Potential Parameters on Amplitude and Wavelength}
453    
454     We have used two different methods to estimate the amplitude and
455     periodicity of the ripples. The first method requires projection of
456     the ripples onto a one dimensional rippling axis. Since the rippling
457     is always perpendicular to the dipole director axis, we can define a
458     ripple vector as follows. The largest eigenvector, $s_1$, of the
459     $\mathsf{S}$ matrix in Eq. \ref{eq:opmatrix} is projected onto a
460     planar director axis,
461     \begin{equation}
462     \vec{d} = \left(\begin{array}{c}
463     \vec{s}_1 \cdot \hat{i} \\
464     \vec{s}_1 \cdot \hat{j} \\
465     0
466     \end{array} \right).
467     \end{equation}
468     ($\hat{i}$, $\hat{j}$ and $\hat{k}$ are unit vectors along the $x$,
469     $y$, and $z$ axes, respectively.) The rippling axis is in the plane of
470     the membrane and is perpendicular to the planar director axis,
471     \begin{equation}
472     \vec{q}_{\mathrm{rip}} = \vec{d} \times \hat{k}
473     \end{equation}
474     We can then find the height profile of the membrane along the ripple
475     axis by projecting heights of the dipoles to obtain a one-dimensional
476     height profile, $h(q_{\mathrm{rip}})$. Ripple wavelengths can be
477     estimated from the largest non-thermal low-frequency component in the
478     fourier transform of $h(q_{\mathrm{rip}})$. Amplitudes can be
479     estimated by measuring peak-to-trough distances in
480     $h(q_{\mathrm{rip}})$ itself.
481    
482     A second, more accurate, and simpler method for estimating ripple
483     shape is to extract the wavelength and height information directly
484     from the largest non-thermal peak in the undulation spectrum. For
485     large-amplitude ripples, the two methods give similar results. The
486     one-dimensional projection method is more prone to noise (particularly
487     in the amplitude estimates for the distorted lattices). We report
488     amplitudes and wavelengths taken directly from the undulation spectrum
489     below.
490    
491     In the triangular lattice ($\gamma = \sqrt{3}$), the rippling is
492     observed for temperatures ($T^{*}$) from $61-122$. The wavelength of
493     the ripples is remarkably stable at 21.4~$\sigma$ for all but the
494     temperatures closest to the order-disorder transition. At $T^{*} =
495     122$, the wavelength drops to 17.1~$\sigma$.
496    
497     The dependence of the amplitude on temperature is shown in the top
498     panel of Fig. \ref{fig:Amplitude}. The rippled structures shrink
499     smoothly as the temperature rises towards the order-disorder
500     transition. The wavelengths and amplitudes we observe are
501     surprisingly close to the $\Lambda / 2$ phase observed by Kaasgaard
502     {\it et al.} in their work on PC-based lipids.\cite{Kaasgaard03}
503     However, this is coincidental agreement based on a choice of 7~\AA~as
504     the mean spacing between lipids.
505    
506     \begin{figure}
507     \includegraphics[width=\linewidth]{properties_sq.pdf}
508     \caption{\label{fig:Amplitude} a) The amplitude $A^{*}$ of the ripples
509     vs. temperature for a triangular lattice. b) The amplitude $A^{*}$ of
510     the ripples vs. dipole strength ($\mu^{*}$) for both the triangular
511     lattice (circles) and distorted lattice (squares). The reduced
512     temperatures were kept fixed at $T^{*} = 94$ for the triangular
513     lattice and $T^{*} = 106$ for the distorted lattice (approximately 2/3
514     of the order-disorder transition temperature for each lattice).}
515     \end{figure}
516    
517     The ripples can be made to disappear by increasing the internal
518     surface tension (i.e. by increasing $k_r$ or equivalently, reducing
519     the dipole moment). The amplitude of the ripples depends critically
520     on the strength of the dipole moments ($\mu^{*}$) in Eq. \ref{eq:pot}.
521     If the dipoles are weakened substantially (below $\mu^{*}$ = 20) at a
522     fixed temperature of 94, the membrane loses dipolar ordering
523     and the ripple structures. The ripples reach a peak amplitude of
524     3.7~$\sigma$ at a dipolar strength of 25. We show the dependence
525     of ripple amplitude on the dipolar strength in
526     Fig. \ref{fig:Amplitude}.
527    
528     \subsection{Distorted lattices}
529    
530     We have also investigated the effect of the lattice geometry by
531     changing the ratio of lattice constants ($\gamma$) while keeping the
532     average nearest-neighbor spacing constant. The antiferroelectric state
533     is accessible for all $\gamma$ values we have used, although the
534     distorted triangular lattices prefer a particular director axis due to
535     the anisotropy of the lattice.
536    
537     Our observation of rippling behavior was not limited to the triangular
538     lattices. In distorted lattices the antiferroelectric phase can
539     develop nearly instantaneously in the Monte Carlo simulations, and
540     these dipolar-ordered phases tend to be remarkably flat. Whenever
541     rippling has been observed in these distorted lattices
542     (e.g. $\gamma = 1.875$), we see relatively short ripple wavelengths
543     (14 $\sigma$) and amplitudes of 2.4~$\sigma$. These ripples are
544     weakly dependent on dipolar strength (see Fig. \ref{fig:Amplitude}),
545     although below a dipolar strength of $\mu^{*} = 20$, the membrane
546     loses dipolar ordering and displays only thermal undulations.
547    
548     The ripple phase does {\em not} appear at all values of $\gamma$. We
549     have only observed non-thermal undulations in the range $1.625 <
550     \gamma < 1.875$. Outside this range, the order-disorder transition in
551     the dipoles remains, but the ordered dipolar phase has only thermal
552     undulations. This is one of our strongest pieces of evidence that
553     rippling is a symmetry-breaking phenomenon for triangular and
554     nearly-triangular lattices.
555    
556     \subsection{Effects of System Size}
557     To evaluate the effect of finite system size, we have performed a
558     series of simulations on the triangular lattice at a reduced
559     temperature of 122, which is just below the order-disorder transition
560     temperature ($T^{*} = 139$). These conditions are in the
561     dipole-ordered and rippled portion of the phase diagram. These are
562     also the conditions that should be most susceptible to system size
563     effects.
564    
565     \begin{figure}
566     \includegraphics[width=\linewidth]{SystemSize.pdf}
567     \caption{\label{fig:systemsize} The ripple wavelength (top) and
568     amplitude (bottom) as a function of system size for a triangular
569     lattice ($\gamma=1.732$) at $T^{*} = 122$.}
570     \end{figure}
571    
572     There is substantial dependence on system size for small (less than
573     29~$\sigma$) periodic boxes. Notably, there are resonances apparent
574     in the ripple amplitudes at box lengths of 17.3 and 29.5 $\sigma$.
575     For larger systems, the behavior of the ripples appears to have
576     stabilized and is on a trend to slightly smaller amplitudes (and
577     slightly longer wavelengths) than were observed from the 34.3 $\sigma$
578     box sizes that were used for most of the calculations.
579    
580     It is interesting to note that system sizes which are multiples of the
581     default ripple wavelength can enhance the amplitude of the observed
582     ripples, but appears to have only a minor effect on the observed
583     wavelength. It would, of course, be better to use system sizes that
584     were many multiples of the ripple wavelength to be sure that the
585     periodic box is not driving the phenomenon, but at the largest system
586     size studied (70 $\sigma$ $\times$ 70 $\sigma$), the number of dipoles
587     (5440) made long Monte Carlo simulations prohibitively expensive.
588    
589     \section{Discussion}
590     \label{sec:discussion}
591    
592     We have been able to show that a simple dipolar lattice model which
593     contains only molecular packing (from the lattice), anisotropy (in the
594     form of electrostatic dipoles) and a weak surface tension (in the form
595     of a nearest-neighbor harmonic potential) is capable of exhibiting
596     stable long-wavelength non-thermal surface corrugations. The best
597     explanation for this behavior is that the ability of the dipoles to
598     translate out of the plane of the membrane is enough to break the
599     symmetry of the triangular lattice and allow the energetic benefit from
600     the formation of a bulk antiferroelectric phase. Were the weak
601     surface tension absent from our model, it would be possible for the
602     entire lattice to ``tilt'' using $z$-translation. Tilting the lattice
603     in this way would yield an effectively non-triangular lattice which
604     would avoid dipolar frustration altogether. With the surface tension
605     in place, bulk tilt causes a large strain, and the simplest way to
606     release this strain is along line defects. Line defects will result
607     in rippled or sawtooth patterns in the membrane, and allow small
608     ``stripes'' of membrane to form antiferroelectric regions that are
609     tilted relative to the averaged membrane normal.
610    
611     Although the dipole-dipole interaction is the major driving force for
612     the long range orientational ordered state, the formation of the
613     stable, smooth ripples is a result of the competition between the
614     surface tension and the dipole-dipole interactions. This statement is
615     supported by the variation in $\mu^{*}$. Substantially weaker dipoles
616     relative to the surface tension can cause the corrugated phase to
617     disappear.
618    
619     The packing of the dipoles into a nearly-triangular lattice is clearly
620     an important piece of the puzzle. The dipolar head groups of lipid
621     molecules are sterically (as well as electrostatically) anisotropic,
622     and would not be able to pack in triangular arrangements without the
623     steric interference of adjacent molecular bodies. Since we only see
624     rippled phases in the neighborhood of $\gamma=\sqrt{3}$, this implies
625     that there is a role played by the lipid chains in the organization of
626     the triangularly ordered phases which support ripples in realistic
627     lipid bilayers.
628    
629     The most important prediction we can make using the results from this
630     simple model is that if dipolar ordering is driving the surface
631     corrugation, the wave vectors for the ripples should always found to
632     be {\it perpendicular} to the dipole director axis. This prediction
633     should suggest experimental designs which test whether this is really
634     true in the phosphatidylcholine $P_{\beta'}$ phases. The dipole
635     director axis should also be easily computable for the all-atom and
636     coarse-grained simulations that have been published in the literature.
637    
638     Our other observation about the ripple and dipolar directionality is
639     that the dipole director axis can be found to be parallel to any of
640     the three equivalent lattice vectors in the triangular lattice.
641     Defects in the ordering of the dipoles can cause the dipole director
642     (and consequently the surface corrugation) of small regions to be
643     rotated relative to each other by 120$^{\circ}$. This is a similar
644     behavior to the domain rotation seen in the AFM studies of Kaasgaard
645     {\it et al.}\cite{Kaasgaard03}
646    
647     Although our model is simple, it exhibits some rich and unexpected
648     behaviors. It would clearly be a closer approximation to the reality
649     if we allowed greater translational freedom to the dipoles and
650     replaced the somewhat artificial lattice packing and the harmonic
651     ``surface tension'' with more realistic molecular modeling
652     potentials. What we have done is to present an extremely simple model
653     which exhibits bulk non-thermal corrugation, and our explanation of
654     this rippling phenomenon will help us design more accurate molecular
655     models for corrugated membranes and experiments to test whether
656     rippling is dipole-driven or not.
657    
658     \begin{acknowledgments}
659     Support for this project was provided by the National Science
660     Foundation under grant CHE-0134881. The authors would like to thank
661     the reviewers for helpful comments.
662     \end{acknowledgments}
663    
664     \bibliography{ripple}
665     \end{document}