ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ripple2/ripple.tex
Revision: 3098
Committed: Thu Dec 28 21:55:59 2006 UTC (17 years, 6 months ago) by gezelter
Content type: application/x-tex
File size: 32038 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 gezelter 3075
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 gezelter 3098 \email[E-mail:]{gezelter@nd.edu}
14 gezelter 3075 \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 gezelter 3098 We present a simple model for dipolar elastic membranes that gives
22 gezelter 3075 lattice-bound point dipoles complete orientational freedom as well as
23     translational freedom along one coordinate (out of the plane of the
24 gezelter 3098 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 gezelter 3075 \end{abstract}
39    
40     \pacs{68.03.Hj, 82.20.Wt}
41     \maketitle
42    
43    
44     \section{Introduction}
45     \label{Int}
46    
47 gezelter 3098 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 gezelter 3075
58 gezelter 3098 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 gezelter 3075
65 gezelter 3098 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 gezelter 3075
93 gezelter 3098 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 gezelter 3075
104     The problem with using atomistic and even coarse-grained approaches to
105 gezelter 3098 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 gezelter 3075 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 gezelter 3098 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 gezelter 3075
130 gezelter 3098 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 gezelter 3075 \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.
136    
137     \section{2-D Dipolar Membrane}
138     \label{sec:model}
139    
140     The point of developing this model was to arrive at the simplest
141     possible theoretical model which could exhibit spontaneous corrugation
142 gezelter 3098 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 ($0 <= \theta
150     <= \pi$, $0 <= \phi < 2
151 gezelter 3075 \pi$). 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 gezelter 3098 \begin{eqnarray}
156     V = \sum_i & & \left( \sum_{j>i} \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[
157 gezelter 3075 {\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 gezelter 3098 \right. \nonumber \\
161     & & \left. + \sum_{j \in NN_i}^6 \frac{k_r}{2}\left(
162     r_{ij}-\sigma \right)^2 \right)
163 gezelter 3075 \label{eq:pot}
164 gezelter 3098 \end{eqnarray}
165 gezelter 3075
166 gezelter 3098
167 gezelter 3075 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
170     potential is governed by three parameters, the dipolar strength
171     ($\mu$), the harmonic spring constant ($k_r$) and the preferred
172     intermolecular spacing ($\sigma$). In practice, we set the value of
173     $\sigma$ to the average inter-molecular spacing from the planar
174     lattice, yielding a potential model that has only two parameters for a
175     particular choice of lattice constants $a$ (along the $x$-axis) and
176     $b$ (along the $y$-axis). We also define a set of reduced parameters
177     based on the length scale ($\sigma$) and the energy of the harmonic
178     potential at a deformation of 2 $\sigma$ ($\epsilon = k_r \sigma^2 /
179     2$). Using these two constants, we perform our calculations using
180     reduced distances, ($r^{*} = r / \sigma$), temperatures ($T^{*} = 2
181     k_B T / k_r \sigma^2$), densities ($\rho^{*} = N \sigma^2 / L_x L_y$),
182     and dipole moments ($\mu^{*} = \mu / \sqrt{4 \pi \epsilon_0 \sigma^5
183     k_r / 2}$).
184    
185     To investigate the phase behavior of this model, we have performed a
186     series of Metropolis Monte Carlo simulations of moderately-sized (34.3
187 xsun 3097 $\sigma$ on a side) patches of membrane hosted on both triangular
188     ($\gamma = a/b = \sqrt{3}$) and distorted ($\gamma \neq \sqrt{3}$)
189 gezelter 3075 lattices. The linear extent of one edge of the monolayer was $20 a$
190     and the system was kept roughly square. The average distance that
191     coplanar dipoles were positioned from their six nearest neighbors was
192 xsun 3097 1 $\sigma$ (on both triangular and distorted lattices). Typical
193     system sizes were 1360 dipoles for the triangular lattices and
194     840-2800 dipoles for the distorted lattices. Two-dimensional periodic
195     boundary conditions were used, and the cutoff for the dipole-dipole
196     interaction was set to 4.3 $\sigma$. Since dipole-dipole interactions
197     decay rapidly with distance, and since the intrinsic three-dimensional
198     periodicity of the Ewald sum can give artifacts in 2-d systems, we
199     have chosen not to use it in these calculations. Although the Ewald
200     sum has been reformulated to handle 2-D
201     systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} these methods
202     are computationally expensive,\cite{Spohr97,Yeh99} and are not
203     necessary in this case. All parameters ($T^{*}$, $\mu^{*}$, and
204     $\gamma$) were varied systematically to study the effects of these
205     parameters on the formation of ripple-like phases.
206 gezelter 3075
207     \section{Results and Analysis}
208     \label{sec:results}
209    
210     \subsection{Dipolar Ordering and Coexistence Temperatures}
211     The principal method for observing the orientational ordering
212     transition in dipolar systems is the $P_2$ order parameter (defined as
213     $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
214     eigenvalue of the matrix,
215     \begin{equation}
216     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
217     \begin{array}{ccc}
218     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
219     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
220     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
221     \end{array} \right).
222     \label{eq:opmatrix}
223     \end{equation}
224     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
225     for dipole $i$. $P_2$ will be $1.0$ for a perfectly-ordered system
226     and near $0$ for a randomized system. Note that this order parameter
227     is {\em not} equal to the polarization of the system. For example,
228     the polarization of the perfect antiferroelectric system is $0$, but
229     $P_2$ for an antiferroelectric system is $1$. The eigenvector of
230     $\mathsf{S}$ corresponding to the largest eigenvalue is familiar as
231     the director axis, which can be used to determine a privileged dipolar
232     axis for dipole-ordered systems. The top panel in Fig. \ref{phase}
233     shows the values of $P_2$ as a function of temperature for both
234 xsun 3097 triangular ($\gamma = 1.732$) and distorted ($\gamma=1.875$)
235 gezelter 3075 lattices.
236    
237 gezelter 3098 \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 gezelter 3075 \end{figure}
246    
247     There is a clear order-disorder transition in evidence from this data.
248 xsun 3097 Both the triangular and distorted lattices have dipolar-ordered
249 gezelter 3075 low-temperature phases, and orientationally-disordered high
250 xsun 3097 temperature phases. The coexistence temperature for the triangular
251     lattice is significantly lower than for the distorted lattices, and
252     the bulk polarization is approximately $0$ for both dipolar ordered
253     and disordered phases. This gives strong evidence that the dipolar
254     ordered phase is antiferroelectric. We have verified that this
255     dipolar ordering transition is not a function of system size by
256     performing identical calculations with systems twice as large. The
257     transition is equally smooth at all system sizes that were studied.
258     Additionally, we have repeated the Monte Carlo simulations over a wide
259     range of lattice ratios ($\gamma$) to generate a dipolar
260     order/disorder phase diagram. The bottom panel in Fig. \ref{phase}
261     shows that the triangular lattice is a low-temperature cusp in the
262     $T^{*}-\gamma$ phase diagram.
263 gezelter 3075
264     This phase diagram is remarkable in that it shows an antiferroelectric
265     phase near $\gamma=1.732$ where one would expect lattice frustration
266     to result in disordered phases at all temperatures. Observations of
267     the configurations in this phase show clearly that the system has
268     accomplished dipolar orderering by forming large ripple-like
269     structures. We have observed antiferroelectric ordering in all three
270 xsun 3097 of the equivalent directions on the triangular lattice, and the dipoles
271 gezelter 3075 have been observed to organize perpendicular to the membrane normal
272     (in the plane of the membrane). It is particularly interesting to
273     note that the ripple-like structures have also been observed to
274     propagate in the three equivalent directions on the lattice, but the
275     {\em direction of ripple propagation is always perpendicular to the
276     dipole director axis}. A snapshot of a typical antiferroelectric
277     rippled structure is shown in Fig. \ref{fig:snapshot}.
278    
279 gezelter 3098 \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 gezelter 3075 \end{figure}
289    
290 xsun 3097 Although the snapshot in Fig. \ref{fig:snapshot} gives the appearance
291     of three-row stair-like structures, these appear to be transient. On
292     average, the corrugation of the membrane is a relatively smooth,
293     long-wavelength phenomenon, with occasional steep drops between
294     adjacent lines of anti-aligned dipoles.
295    
296     The height-dipole correlation function ($C(r, \cos \theta)$) makes the
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 gezelter 3098 intermolecular vector $\vec{r}_{ij}$ and direction of dipole $i$,
301 xsun 3097 \begin{equation}
302 gezelter 3098 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 xsun 3097 \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 gezelter 3098 \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 xsun 3097 Anti-correlation (contours below 0) is only seen when the inter-dipole
320 gezelter 3098 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 xsun 3097 \end{figure}
325    
326 gezelter 3075 \subsection{Discriminating Ripples from Thermal Undulations}
327    
328     In order to be sure that the structures we have observed are actually
329     a rippled phase and not simply thermal undulations, we have computed
330     the undulation spectrum,
331     \begin{equation}
332     h(\vec{q}) = A^{-1/2} \int d\vec{r}
333     h(\vec{r}) e^{-i \vec{q}\cdot\vec{r}}
334     \end{equation}
335     where $h(\vec{r})$ is the height of the membrane at location $\vec{r}
336 xsun 3097 = (x,y)$.~\cite{Safran94,Seifert97} In simple (and more complicated)
337     elastic continuum models, it can shown that in the $NVT$ ensemble, the
338     absolute value of the undulation spectrum can be written,
339 gezelter 3075 \begin{equation}
340 gezelter 3098 \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{k_c q^4 +
341     \gamma q^2},
342 gezelter 3075 \label{eq:fit}
343     \end{equation}
344 gezelter 3098 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 xsun 3097 \begin{equation}
350 gezelter 3098 \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{\gamma q^2},
351 xsun 3097 \label{eq:fit2}
352     \end{equation}
353 gezelter 3075
354     The undulation spectrum is computed by superimposing a rectangular
355     grid on top of the membrane, and by assigning height ($h(\vec{r})$)
356     values to the grid from the average of all dipoles that fall within a
357     given $\vec{r}+d\vec{r}$ grid area. Empty grid pixels are assigned
358     height values by interpolation from the nearest neighbor pixels. A
359     standard 2-d Fourier transform is then used to obtain $\langle |
360 xsun 3097 h(q)|^2 \rangle$. Alternatively, since the dipoles sit on a Bravais
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 gezelter 3098 model (a likely extension), an interpolated grid method for computing
365     undulation spectra will be required.
366 gezelter 3075
367 xsun 3097 As mentioned above, the best fits to our undulation spectra are
368 gezelter 3098 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 gezelter 3075
378 gezelter 3098 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 gezelter 3075
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 gezelter 3098 \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 gezelter 3075 \end{figure}
401    
402     \subsection{Effects of Potential Parameters on Amplitude and Wavelength}
403    
404     We have used two different methods to estimate the amplitude and
405     periodicity of the ripples. The first method requires projection of
406     the ripples onto a one dimensional rippling axis. Since the rippling
407     is always perpendicular to the dipole director axis, we can define a
408     ripple vector as follows. The largest eigenvector, $s_1$, of the
409     $\mathsf{S}$ matrix in Eq. \ref{eq:opmatrix} is projected onto a
410     planar director axis,
411     \begin{equation}
412     \vec{d} = \left(\begin{array}{c}
413     \vec{s}_1 \cdot \hat{i} \\
414     \vec{s}_1 \cdot \hat{j} \\
415     0
416     \end{array} \right).
417     \end{equation}
418     ($\hat{i}$, $\hat{j}$ and $\hat{k}$ are unit vectors along the $x$,
419     $y$, and $z$ axes, respectively.) The rippling axis is in the plane of
420     the membrane and is perpendicular to the planar director axis,
421     \begin{equation}
422     \vec{q}_{\mathrm{rip}} = \vec{d} \times \hat{k}
423     \end{equation}
424     We can then find the height profile of the membrane along the ripple
425     axis by projecting heights of the dipoles to obtain a one-dimensional
426     height profile, $h(q_{\mathrm{rip}})$. Ripple wavelengths can be
427     estimated from the largest non-thermal low-frequency component in the
428     fourier transform of $h(q_{\mathrm{rip}})$. Amplitudes can be
429     estimated by measuring peak-to-trough distances in
430     $h(q_{\mathrm{rip}})$ itself.
431    
432     A second, more accurate, and simpler method for estimating ripple
433     shape is to extract the wavelength and height information directly
434     from the largest non-thermal peak in the undulation spectrum. For
435     large-amplitude ripples, the two methods give similar results. The
436     one-dimensional projection method is more prone to noise (particularly
437 xsun 3097 in the amplitude estimates for the distorted lattices). We report
438 gezelter 3075 amplitudes and wavelengths taken directly from the undulation spectrum
439     below.
440    
441 xsun 3097 In the triangular lattice ($\gamma = \sqrt{3}$), the rippling is
442 gezelter 3075 observed for temperatures ($T^{*}$) from $61-122$. The wavelength of
443     the ripples is remarkably stable at 21.4~$\sigma$ for all but the
444     temperatures closest to the order-disorder transition. At $T^{*} =
445     122$, the wavelength drops to 17.1~$\sigma$.
446    
447     The dependence of the amplitude on temperature is shown in the top
448     panel of Fig. \ref{fig:Amplitude}. The rippled structures shrink
449     smoothly as the temperature rises towards the order-disorder
450     transition. The wavelengths and amplitudes we observe are
451     surprisingly close to the $\Lambda / 2$ phase observed by Kaasgaard
452     {\it et al.} in their work on PC-based lipids.\cite{Kaasgaard03}
453     However, this is coincidental agreement based on a choice of 7~\AA~as
454     the mean spacing between lipids.
455    
456 gezelter 3098 \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 gezelter 3075 \end{figure}
466    
467     The ripples can be made to disappear by increasing the internal
468     surface tension (i.e. by increasing $k_r$ or equivalently, reducing
469     the dipole moment). The amplitude of the ripples depends critically
470     on the strength of the dipole moments ($\mu^{*}$) in Eq. \ref{eq:pot}.
471     If the dipoles are weakened substantially (below $\mu^{*}$ = 20) at a
472     fixed temperature of 94, the membrane loses dipolar ordering
473     and the ripple structures. The ripples reach a peak amplitude of
474     3.7~$\sigma$ at a dipolar strength of 25. We show the dependence
475     of ripple amplitude on the dipolar strength in
476     Fig. \ref{fig:Amplitude}.
477    
478 xsun 3097 \subsection{Distorted lattices}
479 gezelter 3075
480     We have also investigated the effect of the lattice geometry by
481     changing the ratio of lattice constants ($\gamma$) while keeping the
482     average nearest-neighbor spacing constant. The antiferroelectric state
483     is accessible for all $\gamma$ values we have used, although the
484 xsun 3097 distorted triangular lattices prefer a particular director axis due to
485 gezelter 3075 the anisotropy of the lattice.
486    
487 xsun 3097 Our observation of rippling behavior was not limited to the triangular
488     lattices. In distorted lattices the antiferroelectric phase can
489 gezelter 3075 develop nearly instantaneously in the Monte Carlo simulations, and
490     these dipolar-ordered phases tend to be remarkably flat. Whenever
491 xsun 3097 rippling has been observed in these distorted lattices
492 gezelter 3075 (e.g. $\gamma = 1.875$), we see relatively short ripple wavelengths
493     (14 $\sigma$) and amplitudes of 2.4~$\sigma$. These ripples are
494     weakly dependent on dipolar strength (see Fig. \ref{fig:Amplitude}),
495     although below a dipolar strength of $\mu^{*} = 20$, the membrane
496     loses dipolar ordering and displays only thermal undulations.
497    
498     The ripple phase does {\em not} appear at all values of $\gamma$. We
499     have only observed non-thermal undulations in the range $1.625 <
500     \gamma < 1.875$. Outside this range, the order-disorder transition in
501     the dipoles remains, but the ordered dipolar phase has only thermal
502     undulations. This is one of our strongest pieces of evidence that
503 xsun 3097 rippling is a symmetry-breaking phenomenon for triangular and
504     nearly-triangular lattices.
505 gezelter 3075
506     \subsection{Effects of System Size}
507     To evaluate the effect of finite system size, we have performed a
508 xsun 3097 series of simulations on the triangular lattice at a reduced
509 gezelter 3075 temperature of 122, which is just below the order-disorder transition
510     temperature ($T^{*} = 139$). These conditions are in the
511     dipole-ordered and rippled portion of the phase diagram. These are
512     also the conditions that should be most susceptible to system size
513     effects.
514    
515 gezelter 3098 \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 gezelter 3075 \end{figure}
521    
522     There is substantial dependence on system size for small (less than
523     29~$\sigma$) periodic boxes. Notably, there are resonances apparent
524     in the ripple amplitudes at box lengths of 17.3 and 29.5 $\sigma$.
525     For larger systems, the behavior of the ripples appears to have
526     stabilized and is on a trend to slightly smaller amplitudes (and
527     slightly longer wavelengths) than were observed from the 34.3 $\sigma$
528     box sizes that were used for most of the calculations.
529    
530     It is interesting to note that system sizes which are multiples of the
531     default ripple wavelength can enhance the amplitude of the observed
532     ripples, but appears to have only a minor effect on the observed
533     wavelength. It would, of course, be better to use system sizes that
534     were many multiples of the ripple wavelength to be sure that the
535     periodic box is not driving the phenomenon, but at the largest system
536     size studied (70 $\sigma$ $\times$ 70 $\sigma$), the number of dipoles
537     (5440) made long Monte Carlo simulations prohibitively expensive.
538    
539     \section{Discussion}
540     \label{sec:discussion}
541    
542     We have been able to show that a simple dipolar lattice model which
543     contains only molecular packing (from the lattice), anisotropy (in the
544     form of electrostatic dipoles) and a weak surface tension (in the form
545     of a nearest-neighbor harmonic potential) is capable of exhibiting
546     stable long-wavelength non-thermal surface corrugations. The best
547     explanation for this behavior is that the ability of the dipoles to
548     translate out of the plane of the membrane is enough to break the
549 xsun 3097 symmetry of the triangular lattice and allow the energetic benefit from
550 gezelter 3075 the formation of a bulk antiferroelectric phase. Were the weak
551     surface tension absent from our model, it would be possible for the
552     entire lattice to ``tilt'' using $z$-translation. Tilting the lattice
553 xsun 3097 in this way would yield an effectively non-triangular lattice which
554 gezelter 3075 would avoid dipolar frustration altogether. With the surface tension
555     in place, bulk tilt causes a large strain, and the simplest way to
556     release this strain is along line defects. Line defects will result
557     in rippled or sawtooth patterns in the membrane, and allow small
558     ``stripes'' of membrane to form antiferroelectric regions that are
559     tilted relative to the averaged membrane normal.
560    
561     Although the dipole-dipole interaction is the major driving force for
562     the long range orientational ordered state, the formation of the
563     stable, smooth ripples is a result of the competition between the
564     surface tension and the dipole-dipole interactions. This statement is
565     supported by the variation in $\mu^{*}$. Substantially weaker dipoles
566     relative to the surface tension can cause the corrugated phase to
567     disappear.
568    
569 xsun 3097 The packing of the dipoles into a nearly-triangular lattice is clearly
570 gezelter 3075 an important piece of the puzzle. The dipolar head groups of lipid
571     molecules are sterically (as well as electrostatically) anisotropic,
572 xsun 3097 and would not be able to pack in triangular arrangements without the
573     steric interference of adjacent molecular bodies. Since we only see
574     rippled phases in the neighborhood of $\gamma=\sqrt{3}$, this implies
575     that there is a role played by the lipid chains in the organization of
576     the triangularly ordered phases which support ripples in realistic
577     lipid bilayers.
578 gezelter 3075
579     The most important prediction we can make using the results from this
580     simple model is that if dipolar ordering is driving the surface
581     corrugation, the wave vectors for the ripples should always found to
582     be {\it perpendicular} to the dipole director axis. This prediction
583     should suggest experimental designs which test whether this is really
584     true in the phosphatidylcholine $P_{\beta'}$ phases. The dipole
585     director axis should also be easily computable for the all-atom and
586     coarse-grained simulations that have been published in the literature.
587    
588     Our other observation about the ripple and dipolar directionality is
589     that the dipole director axis can be found to be parallel to any of
590 xsun 3097 the three equivalent lattice vectors in the triangular lattice.
591 gezelter 3075 Defects in the ordering of the dipoles can cause the dipole director
592     (and consequently the surface corrugation) of small regions to be
593     rotated relative to each other by 120$^{\circ}$. This is a similar
594     behavior to the domain rotation seen in the AFM studies of Kaasgaard
595     {\it et al.}\cite{Kaasgaard03}
596    
597     Although our model is simple, it exhibits some rich and unexpected
598     behaviors. It would clearly be a closer approximation to the reality
599     if we allowed greater translational freedom to the dipoles and
600     replaced the somewhat artificial lattice packing and the harmonic
601     ``surface tension'' with more realistic molecular modeling
602     potentials. What we have done is to present an extremely simple model
603     which exhibits bulk non-thermal corrugation, and our explanation of
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 gezelter 3098
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 gezelter 3075 \bibliography{ripple}
615     \end{document}